Você está na página 1de 848

BLADDER CANCER

EDITORS

MARK SOLOWAY - SAAD KHOURY

2nd International Consultation on Bladder Cancer - Vienna

Second Edition
2012
1

ICUD

Acknowledgement :
We wish to express our sincere appreciation to Adrienne Carmack for her amazing editorial
assistance in reading each chapter and modifying syntax where appropriate. Despite her dual
role as a mother and urologist she was able to find the time to help us once again with the
Bladder Cancer Recommendations.

For information

Distributor : EDITIONS 21

5 Boulevard Flandrin - 75016 Paris - FRANCE


E-mail : consulturo@aol.com

ICUD-EAU 2012

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted in
any form or by any means, electronic, mechanical, photocopying, recording, or otherwise, without prior permission of the publisher.
Accurate indications, adverse reactions, and dosage schedules for drugs are provided in this book, but it is possible
that they may change. The reader is urged to review the package information data of the manufacturers
of the medications mentioned.
The Publishers have made every effort to trace the copyright holders for borrowed material. If they have inadvertently overlooked any, they will be pleased to make the necessary arrangements at the first opportunity.

The great tragedy of science :


The slaying of a beautiful hypothesis by an ugly fact
Thomas Huxley (1825-1895)

ISBN: 978-9953-493-20-6

Contents
- Contents
3
- ICUD Evidence Based Medicine Overview
4
- TNM
6
- Faculty
7
- Editorial
13
Committee 1: Screening, Diagnosis, and Evaluation
23
Dr. Paul K. Hegarty, Dr. Ashish M. Kamat, Mohan Arianayagam,
Peter Black, Sam S. Chang, Peter E. Clark, Judson D. Davies, Jinhai Fan,
Jason R. Gee, Nicholas J. Hegarty, Murugesan Manoharan, Tim S. OBrien,
Amit Patel, SudhirRawal, Rafael Sanchez-Salas, Robert S. Svatek, William Turner
Committee 2: Pathology Consensus Guidelines by the Pathology
63
of Bladder Cancer Work Group
Mahul B. Amin, Victor E. Reuter, Jonathan I. Epstein, David J. Grignon,
Donna E. Hansel, Oscar Lin, Jesse K. McKenney, Rodolfo Montironi,
Gladell P. Paner, Mark Soloway, and Members of the Pathology of Bladder Cancer
Work Group
Committee 3 A: Molecular Markers for Bladder Cancer Screening,
169
Early Diagnosis, and Surveillance
Shahrokh F. Shariat, Pierre I. Karakiewicz, Bernd J. Schmitz-Drger,
Michael Droller, Yair Lotan, Vinata B. Lokeshwar, Bas W. van Rhijn,
George P. Hemstreet, PerUno Malmstrom, Yves Fradet, MLiss Hudson,
Osamu Ogawa, Michael J. Marberger
Committee 3 B: Molecular Markers for Bladder Cancer Staging and Prognosis
207
Shahrokh F. Shariat, Pierre I. Karakiewicz, Michael J. Droller, Yves Fradet,
George P. Hemstreet, MLiss Hudson, Yair Lotan, Peruno Malmstrom,
Michael J. Marberger, Osamu Ogawa, Bernd J. Schmitz-Drger, Christian Seitz,
Maxine Sun, Bas W. Van Rhijn, Vinata B. Lokeshwar
Committee 4: Low Grade Ta Urothelial Carcinoma of the Bladder
231
Badrinath Konety, Willem Oosterlinck, Sam Chang, Sigurdur Gudjonsson,
Raj Pruthi, Mark Soloway, Eduardo Solsona, Paul Sved
Committee 5: High grade Ta, CIS, and T1 Urothelial Carcinoma
247
Maximilian Burger, Fred Witjes, Marko Babjuk, Maurizio Brausi,
Chris Cheng, Eva Comperat, Colin Dinney, Wolfgang Jager, Wolfgang Otto,
Jay Shah, Joachim Throf
Committee 6: Muscle-invasive, Presumably Regional, Tumor
269
Jason Efstathiou, David I. Quinn, Arnulf Stenzl, Joaquim Bellmunt,
Michael S. Cookson, Georgios Gakis, Khurshid A. Guru, Axel Heidenreich,
Patrick Keegan, Seth P. Lerner, Edward M. Messing, Arthur I. Sagalowsky,
Mark P. Schoenberg, William U. Shipley, Arlene O. Siefker-Radtke, Cora Sternberg
Committee 7: Urinary Diversion
343
Richard E. Hautmann, Bjoern G. Volkmer, Hassan Abol-Enein,
Thomas Davidsson, Sigurdur Gudjonsson, Stefan H. Hautmann, Henriette V. Holm,
Cheryl T. Lee, Fredrik Liedberg, Stephan Madersbacher, Murugesan Manoharan,
Wiking Mansson, Robert D. Mills, David F. Penson, Eila C. Skinner, Raimund Stein,
Urs E. Studer, Joachim W. Thueroff, William H. Turner
Committee 8: Urothelial Carcinoma of the Prostate
379
Joan Palou, David Wood, H. Al-Ahmadie, B. Bochner, Henk van der Poel,
Ofer Yossepowitch
Committee 9: Chemotherapy for Metastatic Urothelial Carcinoma
393
CN. Sternberg, D. Bajorin, R. Dreicer, M. Galsky, D. George, M. Milowsky,
D. Quinn, G. Sonpavde, WM. Stadler, D. Theodorescu, D. Vaughn
Committee 10: Non-urothelial Cancer of the Urinary Bladder
409
Bruce R Kava, Hassan Abol-Enein, Jack Baniel, Adrienne Carmack,
Alexandra Coquhoun, William Turner
3

EVIDENCE BASED MEDICINE OVERVIEW OF THE MAIN STEPS FOR


DEVELOPING AND GRADING GUIDELINE RECOMMENDATIONS.
INTRODUCTION

2.2 How papers are analysed?

The International Consultation on Urological Diseases (ICUD)


is a non-governmental organization registered with the World
Health Organisation (WHO). In the last ten years Consultations
have been organised on BPH, Prostate Cancer, Urinary Stone
Disease, Nosocomial Infections, Erectile Dysfunction and
Urinary Incontinence. These consultations have looked at
published evidence and produced recommendations at four
levels; highly recommended, recommended, optional and not
recommended. This method has been useful but the ICUD
believes that there should be more explicit statements of the
levels of evidence that generate the subsequent grades of
recommendations.

Papers published in peer reviewed journals have differing


quality and level of evidence.
Each committee will rate the included papers according to
levels of evidence (see below).
The level (strength) of evidence provided by an individual
study depends on the ability of the study design to minimise
the possibility of bias and to maximise attribution.
is influenced by:
the type of study
The hierarchy of study types are:

The Agency for Health Care Policy and Research


(AHCPR) have used specified evidence levels to justify
recommendations for the investigation and treatment of a
variety of conditions. The Oxford Centre for Evidence Based
Medicine have produced a widely accepted adaptation of the
work of AHCPR. (June 5th 2001 http://minerva.minervation.
com/cebm/docs/levels.html). The ICUD has examined the
Oxford guidelines and discussed with the Oxford group
their applicability to the Consultations organised by ICUD. It
is highly desirable that the recommendations made by the
Consultations follow an accepted grading system supported
by explicit levels of evidence.

- Systematic reviews and meta-analysis of randomised


controlled trials
- Randomised controlled trials
- Non-randomised cohort studies
- Case control studies
- Case series
- Expert opinion
how well the study was designed and carried out
Failure to give due attention to key aspects of study
methodology increase the risk of bias or confounding factors,
and thus reduces the studys reliability.

The ICUD proposes that future consultations should use a


modified version of the Oxford system which can be directly
mapped onto the Oxford system.

The use of standard check lists is recommended to insure


that all relevant aspects are considered and that a consistent
approach is used in the methodological assessment of the
evidence.

1. 1st Step: Define the specific questions or statements


that the recommendations are supposed to address.
2. 2nd Step: Analyse and rate (level of evidence) the
relevant papers published in the literature.

The objective of the check list is to give a quality rating for


individual studies.

The analysis of the literature is an important step in preparing


recommendations and their guarantee of quality.

how well the study was reported

2.1 What papers should be included in the analysis?

The ICUD has adopted the CONSORT statement and its


widely accepted check list. The CONSORT statement and
the checklist are available at

Papers published, or accepted for publication in the peer


reviewed issues of journals.
T
 he committee should do its best to search for papers
accepted for publication by the peer reviewed journals in
the relevant field but not yet published.

http: //www.consort-statement.org
2.3 How papers are rated?
Papers are rated following a Level of Evidence scale.

A
 bstracts published in peer review journals should be
identified. If of sufficient interest the author(s) should
be asked for full details of methodology and results. The
relevant committee members can then peer review the
data, and if the data confirms the details in the abstract,
then that abstract may be included, with an explanatory
footnote. This is a complex issue it may actually increase
publication bias as uninteresting abstracts commonly do
not progress to full publication.

ICUD has modified the Oxford Center for Evidence-Based


Medicine levels of evidence.
The levels of evidence scales vary between types of studies
(ie therapy, diagnosis, differential diagnosis/symptom
prevalence study).
the Oxford Center for Evidence-Based Medicine Website:
http://minerva.minervation.com/cebm/docs/ levels.html
3. 3rd Step: Synthesis of the evidence

Papers published in non peer reviewed supplements will


not be included.

After the selection of the papers and the rating of the level
of evidence of each study, the next step is to compile a
summary of the individual studies and the overall direction of
the evidence in an Evidence Table.

An exhaustive list should be obtained through:


I. the major databases covering the last ten years (e.g.
Medline, Embase, Cochrane Library, Biosis, Science
Citation Index)

4. 4
 th Step: Considered judgment
individual clinical expertise)

II. the table of contents of the major journals of urology and


other relevant journals, for the last three months, to take
into account the possible delay in the indexation of the
published papers in the databases.

(integration

of

Having completed a rigorous and objective synthesis of the


evidence base, the committee must then make a judgement
as to the grade of the recommendation on the basis of this
evidence. This requires the exercise of judgement based on
clinical experience as well as knowledge of the evidence and
the methods used to generate it. Evidence based medicine
requires the integration of individual clinical expertise with best

It is expected that the highly experienced and expert


committee members provide additional assurance that no
important study would be missed using this review process.

available external clinical evidence from systematic research.


Without the former, practice quickly becomes tyrannised
by evidence, for even excellent external evidence may be
inapplicable to, or inappropriate for, an individual patient:
without current bestevidence, practice quickly becomes
out of date. Although it is not practical to lay our rules for
exercising judgement, guideline development groups are
asked to consider theevidence in terms of quantity, quality,
and consistency; applicability; generalisability; and clinical
impact.

6.2 Grades of Recommendation


The ICUD will use the four grades from the Oxford system.
As with levels of evidence the grades of evidence may apply
either positively (do the procedure) or negatively (dont do the
procedure). Where there is disparity of evidence, for example
if there were three well conducted RCTs indicating that Drug
A was superior to placebo, but one RCT whose results show
no difference, then there has to be an individual judgement
as to the grade of recommendation given and the rationale
explained.

5. 5th Step: Final Grading

Grade A recommendation usually depends on consistent


level 1 evidence and often means that the recommendation is effectively mandatory and placed within a clinical care
pathway. However, there will be occasions where excellent
evidence (level 1) does not lead to a Grade A recommendation, for example, if the therapy is prohibitively expensive,
dangerous or unethical. Grade A recommendation can
follow from Level 2 evidence. However, a Grade A recommendation needs a greater body of evidence if based on
anything except Level 1 evidence

The grading of the recommendation is intended to strike an


appropriate balance between incorporating the complexity of
type and quality of the evidence and maintaining clarity for
guideline users.
The recommendations for grading follow the Oxford Centre
for Evidence-Based Medicine.
The levels of evidence shown below have again been
modified in the light of previous consultations. There are now
4 levels of evidence instead of 5.

Grade B recommendation usually depends on consistent


level 2 and or 3 studies, or majority evidence from RCTs.

The grades of recommendation have not been reduced and a


no recommendation possible grade has been added.
6. Levels of Evidence and Grades of Recommendation
Therapeutic Interventions

Grade C recommendation usually depends on level 4 studies or majority evidence from level 2/3 studies or Dephi
processed expert opinion.

All interventions should be judged by the body of evidence


for their efficacy, tolerability, safety, clinical effectiveness and
cost effectiveness. It is accepted that at present little data
exists on cost effectiveness for most interventions.

Grade D No recommendation possible would be used


where the evidence is inadequate or conflicting and when
expert opinion is delivered without a formal analytical process, such as by Dephi.

6.1 Levels of Evidence

7. Levels of Evidence and Grades of Recommendation


for Methods of Assessment and Investigation

Firstly, it should be stated that any level of evidence may be


positive (the therapy works) or negative (the therapy doesnt
work). A level of evidence is given to each individual study.

From initial discussions with the Oxford group it is clear that


application of levels of evidence/grades of recommendation
for diagnostic techniques is much more complex than for
interventions.

L
 evel 1 evidence (incorporates Oxford 1a, 1b) usually
involves meta-anaylsis of trials (RCTs) or a good quality
randomised controlled trial, or all or none studies in which
no treatment is not an option, for example in vesicovaginal
fistula.

The ICUD recommend, that, as a minimum, any test should


be subjected to three questions:
1. d
 oes the test have good technical performance, for
example, do three aliquots of the same urine sample give
the same result when subjected to stix testing?

Level 2 evidence (incorporates Oxford 2a, 2b and 2c)


includes low quality RCT (e.g. < 80% follow up) or
metaanalysis (with homogeneity) of good quality prospective
cohort studies. These may include a single group when
individuals who develop the condition are compared with
others from within the original cohort group. There can be
parallel cohorts, where those with the condition in the first
group are compared with those in the second group.

2. D
 oes the test have good diagnostic performance, ideally
against a gold standard measure?
3. D
 oes the test have good therapeutic performance, that is,
does the use of the test alter clinical management, does
the use of the test improve outcome?
For the third component (therapeutic performance) the same
approach can be used as for section 6.

L
 evel 3 evidence (incorporates Oxford 3a, 3b and 4)
includes:

8. L
 evels of Evidence and Grades of Recommendation
for Basic Science and Epidemiology Studies

good quality retrospective case-control studies where


a group of patients who have a condition are matched
appropriately (e.g. for age, sex etc) with control individuals
who do not have the condition.

The proposed ICUD system does not easily fit into these
areas of science. Further research needs to be carried out,
in order to develop explicit levels of evidence that can lead
to recommendations as to the soundness of data in these
important aspects of medicine.

good quality case series where a complete group of patients


all, with the same condition/disease/therapeutic intervention,
are described, without a comparison control group.

CONCLUSION
The ICUD believes that its consultations should follow
the ICUD system of levels of evidence and grades of
recommendation, where possible. This system can be
mapped to the Oxford system.

L
 evel 4 evidence (incorporates Oxford 4) includes expert
opinion were the pinion is based not on evidence but on
first principles (e.g. physiological or anatomical) or bench
research. The Delphi process can be used to give expert
opinion greater authority. In the Delphi process a series of
questions are posed to a panel; the answers are collected
into a series of options; the options are serially ranked;
if a 75% agreement is reached then a Delphi consensus
statement can be made.

There are aspects to the ICUD system that require further


research and development, particularly diagnostic
performance and cost effectiveness, and also factors
such as patient preference.
P. Abrams, S Khoury, A. Grant 19/1/04

TNM staging

T (Primary tumour)
TX Primary tumour cannot be assessed
T0 No evidence of primary tumour
Ta Non-invasive papillary carcinoma
Tis Carcinoma in situ (flat tumour)
T1 Tumour invades subepithelial connective tissue
T2 Tumour invades muscle
T2a Tumour invades superficial muscle (inner half)
T2b Tumour invades deep muscle (outer half)
T3 Tumour invades perivesical tissue:
T3a Microscopically
T3b Macroscopically (extravesical mass)
T4 Tumour invades any of the following: prostate, uterus, vagina, pelvic wall, abdominal
wall
T4a Tumour invades prostate, uterus or vagina
T4b Tumour invades pelvic wall or abdominal wall
N (Lymph nodes)
NX Regional lymph nodes cannot be assessed
N0 No regional lymph node metastasis
N1 Metastasis in a single lymph node 2 cm or less in greatest dimension
N2 Metastasis in a single lymph node more than 2 cm but not more than 5 cm in greatest
dimension, or multiple lymph nodes, none more than 5 cm in greatest dimension
N3 Metastasis in a lymph node more than 5 cm in greatest dimension
M (Distant metastasis)
MX Distant metastasis cannot be assessed
M0 No distant metastasis
M1 Distant metastasis

Faculty

(Chairs are underligned)

Editors
Mark S. Soloway,

Department of Urology
University of Miami,
Miami, USA

Saad Khoury,

Departement of Urology
Hopital de la Piti
Paris, France

Committe 1
Screening, Diagnosis, and
Evaluation
Mohan Arianayagam,
Urology department
University of Miami
Miami, FL 33136
USA

Jinhai Fan,

First Affiliated Hospital of Xian,


Jiaotong University College of
Medicine Xian 710061,
China

Jason R Gee,

Lahey Clinic Medical Center


Associate Professor of Urology
Tufts University School of Medicine
Burlington, Massachusetts
USA

Nicholas Hegarty,

Consultant Urologist,
Mater Private Hospital,
Dublin (Ireland).

Paul K. Hegarty,

Consultant Urologic Surgeon


Guys and St Thomas Hospital,
London, UK

Ashish M. Kamat,

William Bill Turner,


Addenbrookes Hospital
Cambridge University
Hospital NHS
Foundation Trust
Cambridge
UK

Committee 2
Pathology Consensus
Guidelines by the Pathology
of Bladder Cancer Work
Group
Mahul B. Amin,

Department of Pathology &


Laboratory Medicine,CedarsSinaiMedicalCenter,Los
Angeles,CA,USA

Director of Urologic Oncology


Followship
Dept. of Urology / Unit 1373 UT
MD Anderson Cancer Center
Houston, USA

Jonathan I. Epstein,

University of British Columbia,


Dept. of Urologic Sciences
Vancouver Campus
Vancouver, B.C.,
Canada

Murugesan Manoharan,

David J. Grignon,

University of Miami
Department of urology
Miami, FL 33136
USA

Department of Pathology &


Laboratory Medicine, Indiana
University Purdue University
Indianapolis, Indianapolis, IN, USA

Sam S. Chang,

Tim O Brien,

Donna E. Hansel,

Peter Black,

Vanderbilt University Medical


Center
Dept. of Urologic Surgery &
Vanderbilt-Ingram Comprehensive
Cancer Center
Nashville,
USA

Peter Clark,

Vanderbilt University Medical


Center
Dept. of Urologic Surgery &
Vanderbilt-Ingram Comprehensive
Cancer Center
Nashville,
USA

Judson D Davies,

Vanderbilt University Medical


Center, Dept. of Urologic
Surgery & Vanderbilt-Ingram
Comprehensive Cancer Center
Nashville
USA

Guys and St Thomas Hospital,


London, UK

Amit Patel

Guys and St Thomas Hospital,


London, UK

Rafael Sanchez-Salas,
Institut Montsouris
Paris, France

Robert Svatek,

UT Health Science Center San


Antonio
Dept. of Urology, Division of
Urologic Oncology
San Antonio, USA

Sudhir Rawal,

Director of Surgical Oncology


Chief Urogenital oncology
Rajiv Gandhi Cancer Institute &
research centre
Sector 5 Rohini, New Delhi ,
India

Department of Pathology,
JohnsHopkinsMedical
Institutions,Baltimore,MD,USA

Department of Anatomic
Pathology, Cleveland Clinic,
Cleveland, OH, USA

Oscar Lin,

Department of Pathology, Memorial Sloan-Kettering Cancer Center,


New York, NY, USA

Jesse K. McKenney,

Department of Pathology, Stanford University School of Medicine,


Stanford, CA, USA

Rodolfo Montironi,

Section of Pathological Anatomy,


Polytechnic University of the Marche
Region, School of Medicine, United
Hospitals, Ancona, Italy

Gladell P. Paner,

Department of Pathology, University of Chicago, Chicago, IL, USA

Victor E. Reuter,

Pierre I. Karakiewicz,

Mark Soloway,

Bernd J. Schmitz-Drger,

Department of Pathology, Memorial Sloan-Kettering Cancer Center,


New York
Department of Urology, University
of Miami, Miami, FL, USA

committee 3A
Molecular Markers for
Bladder Cancer Screening,
Early Diagnosis, and
Surveillance
Michael J. Droller,

Department of Urology, Mount Sinai


School of Medicine, New York,
NY, USA

Yves Fradet,

Department of Urology,
Laval University, Laval,
Quebec, Canada

George P. Hemstreet,

Department of Urology, University


of Nebraska Medical Center and
Omaha Veterans Affairs Hospital,
Omaha, Nebraska,
USA

MLiss Hudson,

Division of Urologic Surgery,


Washington University,
St. Louis, Missouri, USA

Vinata B. Lokeshwar,

Department of Urology,
University of Miami, Miami,
USA

Yair Lotan,

Department of Urology,
UT Southwestern Medical Center,
Dallas, TX, USA

PerUno Malmstrom,

Department of Urology,
University Hospital, Uppsala,
Sweden

Michael J. Marberger,

Department of Urology, Medical


University of Vienna, Vienna,
Austria

Osamu Ogawa,

Department of Urology,
Graduate School of Medicine,
Kyoto University, Kyoto,
Japan

Department of Urology, University


of Montreal, Montreal,
Quebec, Canada
Urologie, EuromedClinic,
Europa-Allee 1, 90763 Frth,
Germany

Shahrokh F. Shariat,

Department of Urology and Division


of Medical Oncology, Weill Cornell
Medical College,
New York-Presbyterian Hospital,
New York, NY, USA

Bas W. van Rhijn,

Department of Urology, Antoni van


Leeuwenhoek Hospital/Netherlands
Cancer Institute, Amsterdam, The
Netherlands

Commette 3 B
Molecular markers for
bladder cancer staging and
prognosis
Michael Droller,

Department of Urology, Mount Sinai


School of Medicine, New York, NY,
USA

Yves Fradet,

Department of Urology, Laval


University, Laval, Quebec, Canada

George P. Hemstreet,

Department of Urology, University


of Nebraska Medical Center and
Omaha Veterans Affairs Hospital,
Omaha, Nebraska

MLiss Hudson,

Division of Urologic Surgery,


Washington University, St. Louis,
Missouri, USA

Pierre I. Karakiewicz,

Department of Urology, University


of Montreal, Montreal, Quebec,
Canada

Vinata B. Lokeshwar,

Departments of Urology and Cell


Biology and Anatomy, University of
Miami Miller School of Medicine,
Miami, Florida, USA

Yair Lotan,

Department
of
Urology,
UT
Southwestern Medical Center,
Dallas, TX, USA

Peruno Malmstrom,

Department of Urology, University


Hospital, Uppsala, Sweden

Michael J. Marberger,

Department of Urology, Medical


University of Vienna, Vienna,
Austria

Osamu Ogawa,

Department of Urology, Graduate


School of Medicine, Kyoto
University, Kyoto, Japan

Bernd J. Schmitz-Drger,

Urologie, EuromedClinic, EuropaAllee 1, Frth, Germany

Christian Seitz,

Department of Urology, St John of


God Hospital, Vienna, Austria

Shahrokh F. Shariat,

Department of Urology and Division


of Medical Oncology, Weill Cornell
Medical College, New YorkPresbyterian Hospital, New York,
NY, USA

Maxine Sun,

Department of Urology, University


of Montreal, Montreal, Quebec,
Canada

Bas W. van Rhijn,

Department of Urology,
Netherlands Cancer Institute,
Antoni van Leeuwenhoek Hospital,
Amsterdam, the Netherlands.

Committee 4
Low Grade Ta Urothelial
Carcinoma of the Bladder
Sam S. Chang,

Department of Urologic Surgery


Vanderbilt University School of
Medicine
Nashville, Tennessee, USA

Sigurdur Gudjonsson,

Department of Clinical Sciences


Urology, Skne University Hospital
Malm, Sweden

Badrinath Konety,

Department of Urology
University of Minnesota
Minneapolis, MN USA

Willem Oosterlinck,

Ghent University Hospital


Ghent, Belgium

Raj Pruthi,

Division of Urologic Surgery


UNC Department of Surgery
Chapel Hill, NC 27599-7235
USA

Mark S. Soloway,

Department of Urology
University of Miami,
USA

Eduaro Solsona,

Service of Urology
Valenciano Oncology Institute
Valencia, Spain.

Paul Sved,

University of Sydney
Sydney, Australia

Committee 5
High grade Ta, CIS and T1
Urothelial carcinoma
Marko Babjuk,

Department of Urology
Hospital Motol
2nd Faculty of Medicine
Charles University in Praha
Czech Republic

Wolfgang Jager,

Department of Urology
University Medical Center
Johannes Gutenberg University
Gebude 604
Langenbeckstrae 1
55131 Mainz, Germany

Wolfgang Otto,

Dept. of Urology
Caritas St. Josef Medical Centre,
University of Regensburg
Landshuterstr. Regensburg,
Germany

Jay Shah,

Department of Urology
UT MD Anderson Cancer Center
Houston, TX
USA

Joachim W. Throff,

Department of Urology
University Medical Center
Johannes Gutenberg University
Gebude 604
Langenbeckstrae 1
55131 Mainz, Germany

J. Alfred Witjes,

Dept of Urology
Radboud University Nijmegen
Medical Centre
The Netherlands

Maurizio A Brausi,

Ausl Modena
Dept of Urology
B Ramazzini Hospital Modena
Italy

Max Burger,

Dept. of Urology
Caritas St. Josef Medical Centre,
University of Regensburg
Landshuterstr. Regensburg,
Germany

Christopher Cheng,

Department of Urology
Singapore General Hospital
Singapore

Eva Comprat,

Service dAnatomie Pathologique


Hpital La Piti Salptrire,
UPMC Paris VI
Paris, France

Colin Dinney,

Department of Urology
UT MD Anderson Cancer Center
1515 Holcombe Blvd., Unit 1373
Houston,
USA

Khurshid A. Guru,

Department of Urology, Roswell


Park Cancer Institute, Buffalo, NY,
USA

Axel Heidenreich,

Department of Urology, University


Hospital, RWTH Aachen
University, Aachen, Germany

Patrick Keegan,

Department of Urology, Vanderbilt


Medical Center, Nashville, TN,
USA

Seth P. Lerner,

Scott Department of Urology,


Bayor College of Medicine,
Houston, TX, USA

Edward M. Messing,

Department of Urology, University


of Rochester, Medical Center,
Rochester, NY, USA

David I. Quinn,

Keck School of Medicine,


University of Southern California,
Los Angeles, CA, USA

Arthur I. Sagalowsky,

Department of Urology, UT
Southwestern Medical Center,
Dallas, TX, USA

Mark P. Schoenberg,

Committee 6
Muscle-invasive,
Presumably Regional,
Tumor
Joaquim Bellmunt,

Medical Oncology, University


Hospital del Mar, Barcelona, Spain

Department of Urology, James


Buchanan Brady Urological
Institute, Johns Hopkins Medical
Institutions, Baltimore, MD, USA

William U. Shipley,

Department of Radiation Oncology,


Harvard Medical School and
Massachusetts General Hospital,
Boston, MA, USA,

Arlene O. Siefker-Radtke,

Department of Urology, Vanderbilt


Medical Center, Nashville, TN,
USA

Department of Genitourinary
Medical Oncology, Division of
Cancer Medicine, The University
of Texas MD Anderson Cancer
Center, Houston, TX, USA

Jason Efstathiou,

Arnulf Stenzl,

Michael S. Cookson,

Department of Radiation Oncology,


Harvard Medical School and
Massachusetts General Hospital,
Boston, MA, USA,

Georgios Gakis,

Department of Urology, University


Hospital, Eberhard-Karls
University, Tbingen, Germany

Department of Urology, University


Hospital Tbingen, Eberhard-Karls
University Tbingen, Germany

Cora Sternberg,

Department of Medical Oncology,


San Camillo and Forlanini
Hospital,
Rome, Italy

Committee 7
Urinary Diversion

David F. Penson,

David Wood,

Eila C. Skinner,

Offer Yossepowitch,

Vanderbilt University Medical


Center, Nashville, TN,
USA

Department of Urology,
Mansoura University, Mansoura,
Egypt

Keck USC School of Medicine,


University of Southern
California, Los Angeles, CA,
USA

Thomas Davidsson,

Raimund Stein,

Hassan Abol-Enein,

Department of Urology,
Kristianstad County Hospital,
Kristianstad,
Sweden

Sigurdur Gudjonsson,

Department of Urology,
Skane University Hospital, Malm,
Sweden

Department of Urology,
Johannes Gutenberg University,
Mainz, Germany

Urs E. Studer,

Department of Urology,
Inselspital Bern,
Switzerland

Joachim W. Thueroff,

University of Ulm,
Germany

Department of Urology,
Johannes Gutenberg University,
Mainz, Germany

Stefan H. Hautmann,

William H. Turner,

Richard E. Hautmann,

Department of Urology,
Mrkische Kliniken Ldenscheid,
Germany

Henriette V. Holm,

Department of Urology,
Oslo University Hospital, Oslo,
Norway

Cheryl T. Lee,

Department of Urology,
Addenbrookes Hospital,
Cambridge University,
UK

Bjoern G. Volkmer,

Department of Urology,
Klinikum Kassel,
Kassel, Germany

Committe 8

Comprehensive Cancer Center,


University of Michigan,
Ann Arbor, MI,
USA

Urothelial Carcinoma of the


Prostate

Fredrik Liedberg,

Hikmat Al-Ahmadie,

Department of Urology,
Skane University Hospital, Malm,
Sweden

Stephan Madersbacher,
Department of Urology,
Donauspital, Vienna,
Austria

Murugesan Manoharan,

Department of Urology,
University of Miami, Miami, FL,
USA

Wiking Mansson,

InstituteofUrology
RabinMedicalCenter
TelAvivUniversity
Tel Aviv
Israel

Committee 9
Chemotherapy for
Metastatic Urothelial
Carcinoma
DF Bajorin,

Memorial Sloan-Kettering Cancer


Center
Weil Medical College of Cornell
University
New York, USA

Robert Dreicer,

Department of Solid Tumor


Oncology,
Taussig Cancer Institute,
Cleveland Clinic, Cleveland, OH,
USA

Matthew D. Galsky,

Genitourinary Medical Oncology


Tisch Cancer Institute
Mount Sinai School of Medicine
New York, USA

Daniel George,

Department of Pathology
Memorial Sloan-Kettering Cancer
Center
New York, NY 10044, USA

Divisions of Medical Oncology and


Urology
Duke University Medical Center
Durham, NC 27705
USA

Bernard H. Bochner,

MI Milowsky,

Memorial Sloan Kettering Cancer


Center
Department of Urology
Weill Medical College of Cornell
University
USA

Joan Palou,

Department of Urology,
Skane University Hospital, Malm,
Sweden

Urological Oncology Dept.


Fundaci Puigvert
Barcelona
Spain

Robert D. Mills,

Henk g van der Poel,

The Norfolk and Norwich


University Hospital,
Norwich,
UK

Department of Urology
University of Michigan
USA

Dept urology
Netherlands cancer institute
Amsterdam
The Netherlands

10

Memorial Sloan-Kettering Cancer


Center
New york, USA

DI Quinn,

Kenneth J. Norris Jr.


Comprehensive Cancer Center
University of Southern California,
Los Angeles, CA 90033
USA

Guru Sonpavde,

Texas Oncology and the Veterans


Affairs Medical Center
Baylor College of Medicine
Webster, Texas,
USA

Walter M. Stadler,

Sections Hematology/Oncology &


Urology
University of Chicago
Chicago, IL
USA

Cora N. Sternberg,

William H Turner,

Department of Urology
Addenbrookes Hospital
Cambridge University Hospitals
NHS Foundation Trust
Cambridge
United Kingdom

Department of Medical Oncology


San Camillo and Forlanini
Hospital
Rome - Italy

Dan Theodorescu,

University of Colorado Cancer


Center
University of Colorado
Denver, CO,
USA

David J. Vaughn,

Abramson Cancer Center


University of Pennsylvania
Philadelphia, PA
USA

Committee 10
Non-urothelial Cancer of
the Urinary Bladder
Hassan Abol-Enein,

Urology and Nephrology Center


Mansoura University
Mansoura, Egypt

Jack Baniel,

Urology Department
Rabin Medical Center, Beilinson
Campus
Petach Tikva,
Israel

Adrienne J. K. Carmack,
Brenham Urology
Brenham, Texas,
USA

Alexandra J Colquhoun,

Cambridge University Hospitals


NHS Foundation Trust
Cambridge,
United Kingdom

Bruce R Kava,

Department of Urology
University of Miami Miller School of
Medicine
Department of Veterans Affairs
Medical Center, Miami
Miami, Florida,
USA

11

12

Editorial

Mark S. Soloway

It has been my privilege to chair the recently completed International Consultation on Urologic
Diseases (ICUD) Recommendations on Bladder Cancer (BC). After over one year of intense
interactive and enthusiastic collaboration with over 100 experts in various aspects of BC the
recommendations from the 10 committees were presented at the 2011 European Association
of Urology Congress in Vienna on March 18, 2011. Each committee had a chair and co chair
and they selected a broad based group of men and women with expertise in bladder cancer.
This editorial will provide only some of the highlights of the consultation. The entire proceedings will be found in this book . The comprehensive nature of this text will serve as a superb
resource for many years. Of note the first author of each chapter is not necessarily the same
as the chair or co chair of the committee.
Bladder cancer is the second most common malignancy of the genitourinary tract. Its protracted course and requirement for frequent monitoring is responsible for the huge costs involved with the care of patients with BC. It has been stated that it is one of the most expensive
cancers. It is a heterogeneous neoplasm and even within some categories there are significant differences in behavior among different histologic variants. These nuances are critical to
proper management. Unlike prostate cancer the level of awareness for this neoplasm is quite
low among not only the public but primary care physicians as well. This is one of the primary
problems that leads to a delay in diagnosis which may adversely impact survival in men and
women with high grade BC. Thus one of our challenges is to educate the public and members
of the medical profession about the association of risk factors such as cigarette smoke and
pelvic radiation with the development of bladder cancer as well as the need for prompt investigation of hematuria or persistent irritative bladder symptoms.
13

their overall medical history and their prior history


of urothelial cancer. In a discussion of some of the
new techniques to enhance endoscopy and TURBT
they highlight the fact that residual or new tumors
are found in up to 50% of patients at their first post
diagnosis follow up or at a reTUR. Many if not most of
the tumors identified at a repeat endoscopy/TURBT
performed within 4 to 6 weeks of the prior procedure
are due to missed or incompletely resected tumor.
This fact highlights the difficulty in performing this
procedure as the goal is always to remove all evident
tumor at the initial TURBT.

Committee 1: Diagnosis and


staging
Committee 1 was charged with reviewing the
diagnosis and staging of BC. The co chairs are
Ashish Kamat and Paul Hegarty. The sections are
repleat with valuable information about all aspects
of the diagnosis, endoscopic management, and
imaging of BC.
The authors stress the importance of prompt
investigation of men and women with gross or
microscopic hematuria to determine whether a
malignancy is the cause of the bleeding. Although
the liklihood of identifying cancer as the source of
microhematuria is low, upper urinary tract imaging
as well as cystoscopy are indicated particularly
if there are any risk factors or the individual is
over 50. There is no absolute recommendation on
the method of imaging the upper urinary tract. In
most centers the CTurogram is the most common
technique to provide excellent anatomy of the
kidneys and ureters as well as the urinary bladder.
The excretory urogram (IVP) is gradually going the
way of the rotary dial telephone. Nonetheless CT
requires intravenous contrast, subjects the patient
to significant radiation and is expensive. Urinary
cytology should be considered part of the hematuria
work up. Urine cytology is specific and sensitive for
the detection of high grade BC. Although cytology
is not very sensitive for the detection of low grade
tumors this is not a reason to shun its use as
there is no necessity to detect low grade tumors
early. Although cytology and other urine based
markers are recommended for the investigation of
individuals with hematuria the committee concluded
that urinary markers including cytology have not
been shown to be helpful in detecting BC in an
asymptomatic high risk group, e. g. occupational
exposure, smokers. This is largely related to the low
prevalence of bladder cancer even in a so called
high risk population. Among the various urine based
markers the committee indicated that the FISH test
is expensive and there are false positive tests. As
a result a patient may be subjected to a thorough
urologic investigation following a positive FISH test
yet no urothelial cancer is detected. Its use might be
beneficial when there is an atypical cytology reading.
Thus the routine use of urinary markers other than
cytology have yet to find a place in routine urology
practice.

Several new methods are nicely described that


might improve the urologists ability to resect all
the tumor, i.e. perform a complete TURBT. These
technical innovations include photodynamic
diagnostic cystoscopy (PDD), narrow band
imaging (NBI), Ramon spectroscopy, and optical
coherence tomography. These techniques are
designed either to aid in the detection of tumors
not seen by standard white light endoscopy or offer
the potential to provide in situ tumor staging. I have
had experience with PDD and NBI and they do
provide added benefit in some patients, particularly
in identifying papillary tumors and distinguishing
CIS. There are a number of trials demonstrating
that PDD can detect more tumors and this leads
to a lower recurrence rate presumably as a result
of a more complete TURBT. The relative reduction
in recurrent tumors was 16% in one prospective
randomized multicenter trial. This technique requires
the intravesical instillation of a solution containing
hexaminolevulinic acid 1 hour before the endoscopy/
TURBT. NBI has the advantage of not requiring
instillation of any exogenous material. It is initiated
by touching a button on the camera or light source
which places a filter in the optical system. Papillary
tumors in particular are identified by a characteristic
blue appearance.
Upon identifying a bladder tumor the urologist must
decide upon the depth of resection required to
properly stage the tumor. The TURBT provides the
tissue i.e. tumor, which is the foundation for the
decision making and treatment of bladder cancer.
If the endoscopic appearance suggests a possible
T1 or T2 tumor (lamina propria or muscularis propria
invasion) the resection must contain muscle from the
muscularis propria in order to determine the depth
of penetration. In the case of an invasive tumor the
specimen should be analyzed for lymphovascular
invasion as this is an important prognostic factor
along with grade and depth of invasion.

The committee thoroughly reviewed lower urinary


tract endoscopy and discussed in detail some of
the more recent approaches to enhance endoscopy
and endoscopic tumor resection (TURBT). A
TURBT is an under estimated surgical procedure
and requires skill and judgement. The procedure
begins with a thorough history including a review
of the patients unique background which includes

I was particularly pleased to note that the authors


reiterated that bladder perforation should be
discouraged. Many years ago urologists thought
it was acceptable to resect through the wall of the
bladder in an effort to remove a bladder tumor.
The risks clearly outweigh the potential benefit of
eradicating a T3 tumor by endoscopic resection.

14

All pathology reports will be accompanied by the


pertinent histology. Viewing histology via the internet
should be similar to what most of us currently rely
on for viewing images of the upper urinary tract.
We would not make a decision to remove a kidney,
for example, without reviewing the actual CT/MRI.
Why should we not have the same opportunity with
biopsy material?

Committee 2: Pathology
Committee 2 was devoted to the pathology of
BC. The co chairs are M. Amin and Victor Reuter.
They collaborated with a virtual whos who of GU
pathologists to write an outstanding chapter. The
text is accompanied by superb photomicrographs.
The chapter provides an overview of both urothelial
(UC) and non urothelial cancers of the bladder. The
authors carefully illustrate the differences between
dysplasia (not cancer) and carcinoma in situ (CIS).
CIS is a flat, high grade cancer confined to the
urothelium. Dysplasia may require monitoring while
CIS is treated with intravesical therapy, e. g. BCG.

This extensive review of the pathology of bladder


cancer will serve as a superb reference for anyone
interested in this subject. This is emphasized by the
number of references 761!

Committee 3: Molecular
Markers for Bladder Cancer

There are many important variants of urothelial


cancer (UC) and they have important implications
for treatment beyond grade and stage. For example
both micropapillary and the nested variant of UC
have a distinct histologic appearance and behave
in amore aggressive manner than other high grade
UCs. The implication is therefore that a diagnosis
of nested or micropapillary UC at stage T1 should
include a discussion of total cystectomy at the time
of diagnosis.

Committee 3, authored by S. Shariat et al, details


the various molecular markers which might be
useful in BC. As it is a heterogeneous neoplasm
there are numerous molecular signature patterns
which potentially may be immensely helpful for
directing treatment. Some of these markers include
p53, p21, pRB, and p27 among others. Despite a
host of studies that have analyzed the addition or
deletion of these and other genes or gene products
the information derived from these studies have yet
to impact on the decisions we make in the clinic.
The authors discuss a number of reasons for this
which include the differences in the antibodies used
to identify the marker in question and the statistical
methodology in specific studies. For those interested
in the state of the art of this subject this chapter is
a terrific resource.

Another tumor type that has important treatment


implications is small cell carcinoma (SCC) of the
bladder. Like SCC at other sites (lung, prostate) these
tumors metastasize early and the initial treatment
should be systemic chemotherapy, not cystectomy.
This chapter has a superb section detailing the use
of immunohistochemistry which may aid in the
diagnosis of different subtypes of bladder cancers.
There are some markers which may provide
prognostic information as well. Most urologists like
myself are not terribly conversant with this area but
should attempt to become conversant about the
potential role this plays in our ability to decide the
best treatment approach. There is also a section
on urinary markers including cytology and the
chromosome addition/deletion FISH test from the
pathologists perspective.

The second part of this chapter reviews the current


status of urinary markers for screening, diagnosis,
and surveillance of BC. The authors stress the
importance of urinary markers and the opportunity
they might have to play a larger role in the diagnosis
and monitoring of BC. Unfortunately the potential
of urinary cytology and other urine based markers
has yet to be realized. Although these markers
have some efficacy in detecting patients with BC
the specificity and sensitivity is such that they have
not found wide utilization by most urologists. The
discussion of cytology and its lack of sensitivity in
low grade BC must be countered with the fact that
an early diagnosis of the benign acting low grade
tumor is not critical however we have abundant
evidence that there is a substantial delay in the
diagnosis of high grade potentially lethal BC. Thus
the authors highlight the potential role of cytology
or other markers for the early detection of bladder
cancer in a high risk population, i. e. hematuria,
smokers, etc.

The authors detail the necessary information


that constitutes an adequate report from the
pathologist to the clinician. Although there are
many occasions when it would be useful for the
urologist or other oncologist to look at the slides with
the pathologist this is often not feasible so the report
is the primary communication and must contain the
critical information the urologist/oncologist requires
to make treatment decisions. This includes not only
the tumor grade and stage but more precisely the
depth of lamina propria invasion, presence of LVI,
presence or lack thereof of muscularis propria, and
whether there is CIS. In the not too distant future
I believe we will have a readily available method
to access our patients histology on the internet.

The authors comment on why most of the urine based


markers have not been integrated into a routine work

15

up for patients with hematuria. This is partly related to


the relatively high rate of false positive tests and thus
their lack of specificity. In contrast when one has a
report of malignant cells identified in a urine cytology
the clinician has the obligation to search for a tumor
as the chance of a false positive is very low.

BCG is not recommended as first line intravesical


therapy in low risk patients as the side effect profile
is higher than with chemotherapy and the efficacy is
equivalent. BCG should be considered for patients
who develop a high grade Ta or T1 tumor as well
as those who have multiple recurrences despite
intravesical chemotherapy.

Committee 4: Low Grade Ta


Urothelial Carcinoma of the
Bladder

Committee 5: High grade Ta,


CIS, and T1 Urothelial
Carcinoma

Committee 4 concentrated on low grade Ta bladder


tumors. The chairs are B. J. Konety and Wim
Oosterlinck. Similar to the conclusions noted above
as related to cytology and urine based markers
these authors make it clear that cytology and urine
based markers do not have a role in the monitoring
of patients with low grade tumors. The cost and
problems with false positives and/or low sensitivity
in the context of an already highly sensitive and
specific although minimally invasive test to diagnosis
and treat these tumors i. e. cystoscopy have
relegated these tests to a minor role. The authors
emphasize that the risk of a subsequent upper tract
tumor following initial diagnosis and treatment of a
low grade Ta tumor is low and therefore there is no
indication for regular monitoring of the upper tract,
e. g. IVP, CT. This recommendation, if followed, will
save money as well as lessen the risk of radiation
for patients who have LG Ta tumors. Obviously any
change in grade, unexplained hematuria or positive
cytology would prompt consideration for upper tract
imaging.

Committee 5 chaired by Fred Witjes and Max Berger


provided recommendations on high grade Ta, CIS,
and T1 UC. They indicate that the risk of developing
an upper tract tumor over the course of a patients
life is sufficiently high to warrant periodic upper tract
surveillance. Unfortunately there is a paucity of data
on how frequently to perform upper tract imaging
and importantly whether any schedule identifies a life
threatening upper tract UC in a timely fashion. Most
guidelines suggest performing an imaging study
every 1- 3 years following resection of a high grade
Ta/T1 but a high grade upper tract urothelial tumor
is likely to grow and invade or metastasize within
this time frame. If one is really concerned about
detecting a high grade upper tract tumor promptly
then frequent cytology possibly supplemented by
ultrasonography might be a practical and potentially
useful alternative.
Cystoscopy is recommended every 3-4 months for
BC surveillance following the diagnosis of a high
grade UC because there is concern about waiting
too long before detecting and treating a high grade
recurrence. In contrast to the patient with low grade
Ta tumors the risk of recurrence and progression is
substantial. Tumors of the urinary bladder which are
high grade and Ta are initially treated by complete
endoscopic resection which may be facilitated by
the use of photodynamic diagnostic techniques
using blue light technology. If the surgeon is not
confidant that he/she has performed a complete
resection then a re-resection should be performed
within 4 to 6 weeks. Once a complete resection has
been assured intravesical BCG should be initiated
2 weeks after the TURBT. The accepted instillation
is weekly for six weeks with maintenance therapy for
a minimum of one year. There is no data on what
is the optimal maintenance schedule although
most studies have used the schema initiated by the
SOG group. There is evidence that an initial single
instillation of chemotherapy provides a longer
tumor free interval. The precise role of post TURBT
intravesical chemotherapy in these patients is
controversial.

There are a variety of ways the urologist can


manage the patient with a low grade Ta tumor:
cold cup resection/fulguration in the operating
room, office fulguration, formal TURBT, and
initial observation for recurrent very small tumors
which appear to be Ta. The goal is to minimize
the risks of therapy in a tumor that rarely, if ever,
progresses in stage. Because patients with a low
grade Ta tumor infrequently develop a high grade
BC the committee recommends less vigorous
endoscopic surveillance. Thus if the initial three
month cystoscopy is negative the next cystoscopy
may be at 1 year.
Following the removal of one or more low grade
Ta tumor(s) all current guidelines recommend the
instillation of at least one dose of intravesical
chemotherapy to be instilled into the bladder
within hours of the resection. This should not be
prescribed if there has been perforation of the
bladder. Although this recommendation is consistent
among most guideline committees (EAU, AUA, SIU)
data from insurance providers in the US indicate it
is infrequently practiced. This may in part be related
to the obstacles of giving chemotherapy in the
postop recovery room. The committee indicated that

Carcinoma in situ is usually multifocal and often


seen in conjunction with other tumors in the bladder.
Its presence signifies multifocal urothelial cancer.

16

When identified as the only tumor it is treated with


intravesical BCG.

reviewing the treatment alternatives for muscle


invasive UC (MIBC). This is a wonderful and
comprehensive overview of the surgical approaches
to MIBC which include partial and total cystectomy,
and TURBT alone. The initial part of the chapter
deals with the procedure of total cystectomy
which in men traditionally includes removal of the
prostate and in women removal of the uterus and
the anterior vaginal wall. The precise extent of the
procedure may vary depending on the extent of the
tumor, patient age, etc. They discuss the variations
of this procedure and note there may be occasions
when the prostate and uterus and vagina may be
spared. The authors caution that the surgeon must
weigh any potential quality of life benefit from a
less extensive procedure with the risk of leaving
an aggressive cancer in situ. A particular example
would be the woman who is having a cystectomy for
high grade Ta/T1/CIS UC and wishes an orthotopic
neobladder. As long as the staging is accurate it may
be wise to preserve the anterior vaginal wall and
uterus if no prior hysterectomy to act as a support
for the pouch.

Patients with a high grade T1 tumor(s) are a


particular challenge as the risk of initial under staging
and progression is substantial. Some will advocate
cystectomy at initial diagnosis in patients which
have particularly adverse prognostic factors such as
multifocal T1 tumor, associated CIS, lymphovascular
invasion (LVI) or tumors that cannot be completely
resected. There is a strong recommendation that
following the resection of a high grade T1 urothelial
carcinoma with or without muscularis propria in the
specimen a reTURBT should be performed to ensure
complete removal of all tumor before commencing
BCG.
It is important to emphasize that 10 to 20% of
patients with an initial high grade T1 UC who are
treated by a TURBT followed by BCG ultimately die
of metastatic urothelial carcinoma. Thus there should
be some consideration and discussion regarding
initial cystectomy.
One of the important subjects discussed in this
chapter are the definitions of BCG failure. Patients
who are refractory to BCG never achieve a disease
free status with one or two six-week courses of BCG.
Patients who have an initial complete response to
TURBT followed by BCG and develop another tumor
after a year without therapy can be considered for
re-treatment with BCG after complete resection of
their tumor. Patients who have an initial complete
responseto TURBT followed by BCG and develop
another tumor after a year without therapy can be
considered for re-treatment with BCG after complete
resection of their tumor. Lastly there are patients
who are intolerant of BCG. For patients with high
grade urothelial carcinoma that recur promptly
despite BCG there are several alternatives but
cystectomy should be considered the initial option
with alternative intravesical therapies reserved for
those unfit for cystectomy.

The literature outlining the extent of the pelvic


lymph node dissection is detailed. There is
substantial data from a number of sources stressing
the benefit of extending the node dissection superior
to the external iliac vessels and removing lymph
nodes along the common iliac vessels and possibly
even to the aortic bifurcation. Although removing the
lymph nodes superior to the external iliac vessels
yields more negative and positive nodes it remains
unclear that the additional dissection results in a
longer disease free survival. It is clear however that
the quality of the cystectomy and node dissection
is a factor in the prognosis of patients undergoing
this procedure. The percentage of patients with
one or more positive lymph nodes is approximately
25% from large RC series and further highlights the
aggressive nature of high grade MIBC and the need
for some early detection program to improve upon
the survival rate from this disease.

Patients who receive BCG after resection of a


high grade Ta/T1 tumor and have not achieved a
complete response at three months are given the
alternatives of another 6 weeks of BCG or proceed
to cystectomy. Although the authors stress there are
insufficient data to direct the timing of abandoning
bladder preservation there is a growing consensus
that if the tumors are T1 a cystectomy in medically
fit patients is appropriate. The chance that the tumor
is understaged is substantial and delay in removing
the bladder can be lethal.

The complication rate from radical cystectomy


ranges from 30 to 50%. There is a long list of medical
and surgical problems that can occur after a RC/
urinary diversion and some can be addressed with
appropriate attention to detail and risk factors. Once
again a plea to have these procedures concentrated
in centers where there are surgeons and medical and
surgical intensivists accustomed to this operation as
well as the perioperative care of high risk patients.
Thanks to technical improvements to aid the surgeon
such as vascular staplers, bipolar cautery devices as
well as an improved understanding of pelvic anatomy
particularly related to the prostate the blood loss
from this surgery has declined. The perioperative
mortality remains relatively stable and ranges from
1 to 5% depending on a number of factors such as
patient comorbidity, age, prior surgery, etc.

Committee 6: Muscle-invasive,
Presumably Regional, Tumor
Committee 6 chaired by Jason Efstathiou, David
Quinn and Arnulf Stenzl aided by 13 urologists,
radiation or medical oncologists had the task of

17

Another alternative for the management of muscle


invasive bladder cancer is bladder preservation.
This approach is for selected patients who meet
a short list of inclusion criteria which includes a
normal upper urinary tract (no hydronephrosis), an
adequate bladder capacity, and they must have
had acomplete TURBT. It is preferable they do not
have CIS. This program requires the interest and
expertise of three disciplines: urology, radiation and
medical oncology.

and cisplatin. Randomized studies have shown


equivalence between MVAC and GC in metastatic
urothelial cancer with less toxicity with the two drug
regimen
One of the important conclusions is that patients are
more likely to receive the intended chemotherapy
if given before cystectomy, i. e. induction rather
than as an adjuvant. This is largely related to delays
or failure to receive chemotherapy as a result of
the frequent complications following this extensive
operation. This is a disease of elderly patients and
many simply refuse or cannot tolerate our optimal
chemotherapy regimen after a cystectomy and
diversion. These factors are less operative when
considering induction chemotherapy.

The treatment plan calls for an initial complete


endoscopic tumor resection followed by 40 Gy
radiation therapy with concomitant cisplatin based
chemotherapy. At this point a repeat cystoscopy
is required to determine whether there has been
a complete response. If not the patient is advised
to have their bladder removed. If there has been a
complete response the initial treatment is consolidated
with additional radiation and chemotherapy with
careful endoscopic monitoring. Over the last 20
years the protocol has been changed from a
neoadjuvant chemotherapy approach with cisplatin
based combination chemotherapy to be followed by
radiation therapy and cisplatin to the current approach
of twice a day radiation and concurrent cisplatin
and a taxane. This is then followed by four cycles
of gemcitabine and cisplatin. The results from this
bladder sparing approach in a highly selected group
of patients is similar in disease free survival to initial
radical cystectomy. A high percentage of the patients
have adequate bladder function.

This chapter has excellent sections discussing the


role of RC as initial therapy for several variants
of high grade UC. Specifically they discuss the
nested variant, micropapillary, and UC with adeno
and squamous elements. All of these tumors are
associated with a poorer prognosis and thus a
RC should be seriously considered when they are
diagnosed at stage T1.
Another section discusses the role of monitoring
the upper tract following a diagnosis of MIBC
of the bladder. Some of the key points relate to a
risk related strategy for monitoring the upper tract
since there are no optimal recommendations for
imaging the upper tract. No strategy to date has
demonstrated that frequent CT or other imaging
modality performed at one or two year intervals will
detect tumors in a timely fashion. Thus patients with
tumor at the ureteral margin or with bladder CIS
should have more frequent monitoring. Cytology
might be a reasonable adjunct to imaging as it can
be performed frequently at relatively little cost and
does not add the risk of radiation associated with CT
scanning.

The important subject of perioperative systemic


chemotherapy is addressed by a group of
outstanding medical oncologists. The disease free
survival following radical cystectomy has plateaued
over the last decade. In order to improve the
cure rate we will require either effective systemic
therapy or a better method to identify and treat
patients with high grade UC before metastases
occur. We have systemic chemotherapy that can
produce major responses in locally advanced and
metastatic UC although complete responses in the
latter category are no more than 10%. It is thus a
reasonably chemotherapy sensitive tumor with the
use of cisplatin based combination chemotherapy.
Prospective randomized trials have examined the
benefit of induction (neoadjuvant) and adjuvant
chemotherapy. These results are thoroughly
reviewed in this chapter. There have been two
prospective randomized trials that indicate at least
a 5% survival advantage in favor of patients who
received induction or neoadjuvant chemotherapy
before cystectomy or radiation. Meta analysis of
these neoadjuvant trials support the benefit of
cisplatin based combination chemotherapy as
induction treatment. The authors indicate that at
this time the four drug combination identified as
MVAC remains the regimen of choice although in
practical terms most oncologists use gemcitabine

The committee dealing with muscle invasive bladder


cancer also stressed the importance of initial
chemotherapy using cisplatin and etoposide for
small cell carcinoma. Most of these patients have
metastatic disease at diagnosis despite negative
staging studies and therefore it is a recommendation
to initiate treatment with chemotherapy. Once
a complete response is achieved consolidation is
performed with surgery or radiation.

Committee 7: Urinary Diversion


Committee 7 deals with urinary diversion. This
committee is chaired by Richard Hautmann and
Bjoern Volkmer. The 19 member committee stresses
that a radical cystectomy and urinary diversion
may be the most difficult operation in urologic
surgery. The procedures are often performed in

18

elderly patients with significant co morbidity usually


related to years of cigarette smoking. The operation,
whether by a standard open approach or with
laparoscopy, is long and requires that the surgery
be performed with attention to detail. At this time
most of the urinary diversions are being performed
with an open technique even if the RC is performed
laparoscopically.

for a neobladder. Many patients more than 70


years old wish to avoid an abdominal stoma. The
motivation of the patient is the most important factor
in the decision to perform a bladder substitute. The
urethral anastomotic stricture rate is low. With
procedures performed at centers of excellence the
stricture or stenosis at the uretero-ileal junction is
less than 5%. Orthotopic diversion in women is
becoming increasingly utilized and the motivation of
the patient is critical. One would hesitate to perform
a neobladder if the carcinoma invades the bladder
neck or vaginal wall.

The committee points out that there are few studies


which adequately address the complications
related to radical cystectomy. An ideal analysis
should include operative time, length of stay,
estimated blood loss, time to return to work, as well
as operative mortality. Few reports detail readmission
rates, reoperation rates, time in intensive care, and
additional interventional radiologic procedures both
in the short-term and long-term. Since there are no
randomized studies all recommendations are level
III or IV and mostly expert opinion.

The authors discuss the role of renal function both


in deciding the type of diversion and the need to
monitor renal function after the diversion. In order to
perform an orthotopic diversion the patient must have
a creatinine clearance of >50 cc/min. Few patients
have a decline in renal function following a diversion.
One should monitor the upper tract with periodic
creatinine levels as well as by renal ultrasound. Early
diagnosis of stricture at the ureteroileal anastomosis
is important.

This committee surveyed the members of the


working group and note that approximately 50%
of men and women who have a urinary diversion
at these centers of excellence had an orthotopic
neobladder, most of the remaining patients had an
ileal conduit. A continent diversion to the skin is
uncommon. Of interest in population based surveys
of the type of diversion only 15% of patients undergo
an orthotopic diversion however in centers devoted
to the care of these patients up to 70% have an
orthotopic neobladder. The committee suggests that
a high volume center should perform at least 25 RCs
per year. At many centers, including ours, one of us
performs the RC and another the diversion. This
limits the time in the operating room to under 2 - 3
hours for each and I believe has some benefits.

The authors have an excellent section discussing


perioperative complications of urinary diversion.
They stress some practical points such as the
preservation of the anterior vaginal wall, when
possible, and the use of an omental interposition flap
to decrease the chance of a urethral vaginal fistula.

Committee 8: Urothelial
Carcinoma of the Prostate
Committee 8, authored by Juan Palou, David
Wood, and four other urologists reviewed the
topic of urothelial carcinoma of the prostate. Many
years ago this was thought to be a rare entity. With
improvements in endoscopy and TURBT along
with intravesical therapy urothelial carcinoma of the
prostate is more frequent. With control of tumors in
the bladder we have changed the natural history of
UC of the lower urinary tract. The urologist managing
patients with high grade Ta/T1/CIS of the bladder
must always consider the status of the prostatic
urethra and guard against a non visible area of
high grade cancer in the prostate. A positive urinary
cytology in the absence of a visible source in the
bladder requires a biopsy of the prostatic urethra,
which should be a TUR resection biopsy.

The committee highlights the fact that the extent of


the cancer in the pelvis is not always a barrier to
perform an orthotopic neobladder. Positive lymph
nodes, for example, should not exclude a patient
from a neobladder. There is no consensus on the
need for preoperative biopsy of the prostate to
determine the acceptability for an orthotopic urinary
diversion as the frozen section of the urethral margin
will serve as the determining factor of whether an
anastomosis can be made to the urethra. Some
surgeons like to have information on the status of
the prostatic urethra in hand to discuss the type
of diversion. I rely on the intraoperative frozen
section.

It is important to accurately stage the extent of high


grade urothelial carcinoma of the prostatic urethra.
The stages are: carcinoma in situ (CIS), involvement
of the prostatic ducts, and invasion of the prostatic
stroma. The latter portends a particularly adverse
prognostic factor. The 5 year survival is 60% if there
is no stromal invasion and 30% if there is. Stromal
invasion can either occur by direct extension of a
muscle invasive tumor from the bladder or by growth

Nerve sparing can be utilized depending on the


extent of the tumor and will allow preservation of
erectile function in some young men who undergo
a total cystoprostatectomy. The committee indicates
that preserving the seminal vesicles and the
nerves surrounding them may be beneficial
in preserving more nerves involved in erectile
function and continence. There is no age cut off

19

from the prostate urothelium down the ducts into the


stroma. When stromal invasion occurs by extension
from the bladder the survival is particularly poor
despite cystoprostatectomy.

first line chemotherapy. Unlike other cancers in which


there are effective second and third line programs
there is a lack of effective salvage therapy in UC.

Treatment for CIS of the prostate is endoscopic


resection followed by BCG. High grade tumor
after an initial course of BCG should prompt
consideration for a RC. Tumor in the ducts can
be initially treated similarly although some advocate
cystoprostatectomy at initial diagnosis of ductal
involvement. The prognosis for patients with stromal
invasion is poor despite a cystoprostatectomy.
Induction chemotherapy should be considered in
such cases.

Chapter 10: Non-urothelial


carcinoma of the urinary
bladder
Chapter 10 authored by Bruce Kava, Alex
Colquhoun, Jack Baniel et al reviews non-urothelial
carcinoma of the urinary bladder. This category
includes squamous cell, adenocarcinoma, and
small cell carcinoma of the urinary bladder. Other
very rare histologic types are also mentioned.
There is a paucity of information based upon any
randomized trial to determine the optimal approach
for the first two of these tumor types. Since there is
no effective systemic therapy for squamous cell or
adenocarcinoma there is a consensus that the best
option for the majority of these tumors is RC.

Papillary low-grade tumors of the prostate


have a good prognosis and can be managed with
endoscopic resection with or without intravesical
chemotherapy.

Committee 9: Chemotherapy

Squamous cell carcinoma is relatively infrequent


outside of Egypt or other countries in which
schistosomiasis is endemic. Squamous cell
carcinoma of the bladder may be caused by
chronic inflammation of the bladder. Patients with
an indwelling catheter, bladder calculi, spinal cord
injured patients who have a neurogenic bladder are
at an increased risk for squamous cell carcinoma of
the bladder. The local failure rate for squamous cell
carcinoma appears to be higher than for UC of the
bladder.

Committee 9 is authored by Cora Sternberg with


the help of 9 outstanding medical oncologists and
one urologist. Their chapter discusses the role of
systemic chemotherapy for advanced BC. There are
several prospective, randomized trials which have
examined the activity of systemic chemotherapy
using the four drug combination referred to as MVAC
or the combination of cisplatin and gemcitabine.
An important 400 patient prospective, randomized
trial compared these two regimens and demonstrated
equivalent efficacy and less toxicity with the two drug
combination. Thus although some oncologists feel
that in the healthy patient MVAC might be superior
the majority of oncologists use the two drug regimen
for the majority of patients for first line chemotherapy
in locally advanced or metastatic UC. The committee
also discussed the role of high-dose MVAC with
G-CSF rescue.

Adenocarcinoma of the bladder has several


subtypes which include urachal and non urachal
tumors. The former are usually treated by extensive
partial cystectomy ensuring negative margins at
frozen section.
Despite apparent local control the incidence of local
and systemic failure is high. Unfortunately there is
no effective adjuvant chemotherapy although in
some instances radiation might be considered.

The authors discuss the common problem of


cisplatin ineligible patients. Cisplatin is the most
active single agent for UC however many patients
with advanced urothelial cancer are elderly and/or
have reduced renal or cardiac function. Fourty per
cent of patients are ineligible to receive cisplatin
due to renal impairment. Most studies which have
given chemotherapy in these patients have used
carboplatin in place of cisplatin. One of the problems
of evaluating these reports is the variability in the
definition of unfit patients. The traditional exclusion
criteria are a performance status less than 60% and
a creatinine clearance less than 30 ML per minute.
Patients also require a cardiac ejection fraction better
than 40% in order to receive cisplatin.

Small cell carcinoma has been discussed in


part in the sections on pathology and MIBC. The
key here is to make the diagnosis after the initial
diagnostic TURBT. After thorough staging the initial
treatment is cisplatin and etoposide systemic
chemotherapy. The initial response is usually
excellent. Following chemotherapy there are no data
driven recommendations on whether the optimal
approach is RC, radiation, or observation. I have
treated patients using each of these with success.

The chapter discusses the many agents that have


been used for patients who have progressed despite

20

BLADDER CANCER

Editors

Mark S. Soloway and Saad Khoury,

21

22

Committee 1

Screening, Diagnosis,
and Evaluation
Chairs:
Dr. Paul K. Hegarty,
Dr. Ashish M. Kamat
Members:
Mohan Arianayagam,
Peter Black,
Sam S. Chang,
Peter E. Clark,
Judson D. Davies,
Jinhai Fan,
Jason R. Gee,
Nicholas J. Hegarty,
Murugesan Manoharan,
Tim S. OBrien,
Amit Patel,
SudhirRawal,
Rafael Sanchez-Salas,
Robert S. Svatek,
William Turner

23

Table of Contents
A. SCREENING AND DIAGNOSIS:
EVALUATION OF PATIENTS WITH
HEMATURIA

D. Imaging and Bladder Cancer

I. SCREENING

I. Methods

II. PRESENTATION

II. Virtual Cystoscopy

III. INVESTIGATION

III. C
 T in Localized Bladder Cancer Staging

IV. URINE CYTOLOGY

IV. MRI in Localized Bladder


Cancer Staging

V. URINARY BIOMARKERS

V. Detection of Nodal Disease


and Metastasis

B. Endoscopic Examination of
the Lower Urinary Tract

VI. Imaging after Bladder Cancer Treatment

I. White Light Cystoscopy


II. FLUORESCENCE CYSTOSCOPY

VII. G
 eneral Comments

III. New OPTICAL IMAGING Techniques for Bladder Cancer


DIAGNOSIS

Summary of the Recommendations

C. Transurethral Resection
of Bladder Tumors
I. Preoperative assessment
II. Technique
II. Technique
III. T
 URBT in
stances

Specific

Circum-

IV. S
 ample Transportand Pathologic Processing: Or How
best to help your pathologist give you the most useful pathology report

24

Screening, Diagnosis, and Evaluation


Dr. Paul K. Hegarty, Dr. Ashish M. Kamat
Mohan Arianayagam, Peter Black, Sam S. Chang,
Peter E. Clark, Judson D. Davies, Jinhai Fan, Jason R. Gee,
Nicholas J. Hegarty, PaulK. Hegarty, Murugesan Manoharan,
Tim S. OBrien, Amit Patel, SudhirRawal, Rafael Sanchez-Salas
Robert S. Svatek, William Turner
cancer on subsequent investigation is 5% [level of
evidence 2; Britton et al, 1989]. It is unknown if early
detection of bladder cancer following investigation
of asymptomatic microscopic hematuria leads to an
improved outcome. Findings on physical examination
are unremarkable in patients with noninvasive
bladder cancer.

A. SCREENING AND DIAGNOSIS:


EVALUATION OF PATIENTS WITH
HEMATURIA

III. INVESTIGATION

I. SCREENING

Cystoscopy is the cornerstone investigation in


patients with suspected bladder cancer. Flexible
cystoscopy can be performed in the office with local
anesthetic lubricant. The office-based procedure can
be omitted in patients in whom bladder cancer has
already been identified on imaging and who proceed
directly to transurethral resection.

The object of screening is to identify disease at an


earlier stage when it is more amenable to treatment.
An evaluation of the quality of studies concluded that
there were no randomized controlled trials or highquality controlled observational studies comparing
screening to not screening for bladder cancer [Chou
2010]. No study was identified that randomized
screen-detected cancers to treatment versus nontreatment. Thus the recommendations on screening
remain unchanged from the 2006 ICUD guidelines.

In patients with hematuria, imaging is an important


part of the evaluation; imaging is used primarily to
visualize the upper urinary tract as cystoscopy allows
inspection of the lower urinary tract. The incidence
of upper urinary tract malignancy among patients
undergoing evaluation because of hematuria is 0.20.7% [level of evidence 2; Khadra et al, 2000, Edwards
et al, 2006]. Options for imaging are ultrasonography,
IVU, CT urography, MRI, or a combination of these.
Ultrasonography and IVU appear to be equally
sensitive in diagnosing upper tract disease [level of
evidence 2; Khadra et al, 2000]. Using CT as the gold
standard, ultrasonography has a positive predictive
value of 100% in detecting renal cancer [level of
evidence 2; Datta et al, 2002]. Improvements in the
technical quality of ultrasonography have resulted
in improved performance of this imaging modality
in detecting bladder cancer; reported sensitivities
are 63-98%, and reported specificity is 99% [level
of evidence 2; Datta et al, 2002]. Thus, diagnostic
flexible cystoscopy can be omitted in cases in
which ultrasonography reveals bladder cancer. CT

II. PRESENTATION
Hematuria, typically painless, is the cardinal
presenting symptom of noninvasive bladder cancer.
In patients with CIS, hematuria may be accompanied
by irritative voiding symptoms. The incidence of
bladder cancer is 17-18.9% in patients presenting
with macroscopic hematuria and 4.8-6% in patients
presenting with microscopic hematuria [level of
evidence 2; Edwards et al, 2006, Datta et al, 2002,
Mishriki et al, 2008]. The prevalence of asymptomatic
hematuria ranges from 0.19% to 21% depending
on the population being studied, and prevalence
appears to increase with age [Rodgers et al, 2006].
The prevalence of microscopic hematuria is 23% in
men over 60 years of age, and in men older than 60
years with microscopic hematuria, the risk of bladder

25

urography has reported sensitivities of 88-100%


and specificities of 93-100% in diagnosing urothelial
carcinoma. A meta-analysis of 5 studies of CT
urography yielded a pooled sensitivity of 96% and a
pooled specificity of 99% in diagnosing upper urinary
tract malignancy [level of evidence 2; Chlapoutakis
et al, 2010]. The disadvantages of CT urography are
relatively high radiation dose, need for intravenous
contrast medium, limited availability, and relatively
high cost.

positive only when tumor cells are identified, the


positive predictive value of this test is very high
[Bastacky et al, 1999]. The diagnostic yield of urine
cytology is increased if more than 3 samples are
analyzed [Planz et al, 2005].
Sampling error and tumor load are 2 parameters that
have to be considered in the interpretation of urine
cytology. Voided urine specimens have a lesser
diagnostic yield than bladder washings. There have
been instances in which findings on cytology are
positive but findings on cystoscopy or imaging are
negative.

Edwards et al (2010) tested a protocol in which


patients with hematuria underwent flexible
cystoscopy, ultrasonography, and plain radiography
as first-line procedures and IVU if there were
abnormal findings on first-line procedures or there
was persistent hematuria in a patient over 40 years
of age. With a medium follow-up time of 4 years, the
risk of missing a tumor with this protocol in a patient
with microscopic hematuria was less than 1%. For
macroscopic hematuria, the risk of missing a tumor
was 2% for men over 30 years of age and women
over 50 years of age, and the risk increased a further
1% for each decade over 50 years of age for both
sexes [level of evidence 2]. These so-called missed
tumors probably represent new tumors or tumors
not visible on initial cystoscopy [level of evidence 3;
Messing et al, 2005].

The role of urine cytology in the diagnosis of


bladder cancer has been universally accepted and
extensively analyzed. However, a critical review
of the literature reveals a paucity of randomized
controlled trials of urine cytology. Most of the earlier
studies concentrated on the sensitivity, specificity,
and positive predictive value. A pooled analysis
of 36 studies reported a sensitivity of 44% and a
specificity of 96% for urine cytology [Mowatt et al,
2010]. Once urinary biomarkers were introduced,
the focus of research shifted to comparison of
cytology and biomarkers and subsequently to the
utility of biomarkers as an adjunct to cytology. In
a meta-analysis comparing fluorescence in situ
hybridization (FISH) and cytology, even though the
sensitivity and specificity of FISH were superior to
those of urine cytology, the sensitivity of FISH was
lower than that of cytology for higher-stage tumors
[Yang et al, 2009].

IV. URINE CYTOLOGY


Urine cytology is not a laboratory test; rather, it is
a pathologists interpretation of the morphologic
features of urothelial cells. The combination of urine
cytology and cystoscopy has been the gold standard
for the diagnosis and surveillance of bladder
cancer.

The American Urological Association best practice


policy recommends voided urine cytology for all
patients with risk factors for urothelial carcinoma. For
patients with asymptomatic microscopic hematuria
without risk factors for urothelial carcinoma, either
urine cytology or cystoscopy can be used [Grossfeld
et al, 2001].

Cytology is a simple, noninvasive, cost-effective


procedure and the the combination of urine cytology
and cystoscopy is superior to cystoscopy alone in
the detection of high-grade flat lesionsi.e., CIS and
dysplasia; in the detection of high-grade urothelial
cancer; and in the detection of upper urinary tract
tumors. Urine cytology is especially valuable for
differentiating high-grade urothelial carcinomas from
low-grade urothelial neoplasms, a feat that urinary
biomarkers [see the next section] have not yet
achieved [Lokeshwar et al, 2005].

Because of its high specificity and sensitivity in the


detection of pTis disease, urine cytology fulfills the
requirements for an adjunct to cystoscopy in the
surveillance of patients with pTa-T1 bladder urothelial
carcinoma [Babjuk et al, 2008]. A multicenter study
performed in the United States to compare the
performance of cytology, the ImmunoCyt tumor
marker test, and the combination of cytology and
ImmunoCyt in patients monitored for recurrence
of bladder cancer showed that ImmunoCyt
enhanced the sensitivity of cytology [Messing et
al, 2005]. About 50-70% of patients with superficial
bladder cancer develop recurrence within 5 years,
and 25% eventually develop invasive disease.
Lifelong surveillance is therefore mandatory in
the management of non muscle invasive cancer
[Fritsche et al, 2010]. Urine cytology remains the
gold standard for surveillance of patients with non
muscle invasive cancer.

Urine cytology in the diagnosis of bladder cancer has


a moderate sensitivity overall that is complemented
by high specificity. Cytology has a high sensitivity
and high negative predictive value in detecting
high-grade bladder tumors; it has a low sensitivity
and low negative predictive value in detecting lowgrade bladder neoplasms. The specificity of cytology
is superior to that of most of the currently available
bladder tumor markers [see the next section].
Because findings on urine cytology are considered

26

Other recent studies regarding FISH in the diagnosis


of bladder cancer call into question the role of this
assay in clinical practice. In a study by Ferra et al
(2009), patients with suspicious findings on urine
cytology were also evaluated by FISH. When
findings on cystoscopy and biopsy were used as the
gold standard, the sensitivity of FISH was 68.3% and
the specificity was 39.7%, indicating that in patients
with suspicious findings on urine cytology and
negative findings on cystoscopy, positive findings
on FISH could not justify aggressive evaluation
[Ferra et al, 2009]. Youssef et al (2010) performed
a review of 142 patients undergoing cystoscopic
surveillance and confirmed that a cancer diagnosis
in patients with negative findings on both cystoscopy
and cytology was rare (1/111 patients). Furthermore,
cancer in patients with equivocal or positive findings
on cystoscopy was relatively common, such that
the authors concluded that FISH was not useful in
these cases either and that a cost analysis should
be done before adoption of widespread use of FISH
in patients with negative findings on urine cytology
[Youssef et al, 2010].

V. URINARY BIOMARKERS
Urinary biomarkers for the diagnosis of bladder
cancer have been evaluated extensively, but to date,
there have been relatively few completed randomized
trials. Glas et al (2003) performed a meta-analysis
of 42 case series (n=5,706) that compared urine
cytology versus urinary biomarkersincluding
NMP22, telomerase, BTA, BTA-Trak, and BTAStatin the diagnosis of primary bladder cancer.
None of the urinary biomarkers exhibited sensitivity
that would justify the use of biomarker measurement
as a substitute for cystoscopy. Furthermore, when
all studies were combined, urine cytology had
a specificity of 94%, which was higher than the
specificity of any of the biomarker tests evaluated.
Twenty-two of the 42 studies included in the metaanalysis used a case-control design, which can be
associated with bias [Glas et al, 2003]. In a more
recent, cross-sectional study involving 668 patients,
Grossman et al (2006) found that the sensitivity of the
NMP22 test as a point-of-care test for the detection
of recurrent bladder cancer was 49.5%, versus only
12.2% for urine cytology. While these results cannot
justify the use of NMP22 in lieu of cystoscopy, the
addition of the NMP22 assay did improve sensitivity
compared with cystoscopy alone (99% vs. 91%),
whereas the addition of voided urine cytology did not
significantly increase the sensitivity compared with
cystoscopy alone (94% vs. 91%) [Grossman et al,
2006].

Combinations of urinary biomarkers have also been


explored. In a study by Horstmann et al (2009), 221
patients undergoing cystoscopic surveillance for
non-muscle-invasive bladder cancer were evaluated
with urine cytology, NMP22 testing, FISH, and
ImmunoCyt. Sensitivity was increased to over 90%
and negative predictive value was increased to
over 80% with combinations of 2 or 3 biomarkers,
although specificity was reduced to an average of
44% with 2 biomarkers and 35% with 3 biomarkers
[Horstmann 2009].

Results for urinary biomarker tests vary widely. For


instance, for NMP22, the sensitivity has been found
in cohort studies to be as low as 44% [Chahal et al,
2001] and as high as 91% [Miyanaga et al, 1999]. The
specificity of NMP22 in the Chahal et al and Miyanaga
et al studies was 91% and 76%, respectively, and
in another series, the specificity was only 40%
[Tritschler et al, 2007]. Significant variability has also
been noted for other urinary biomarkers, including
BTA (sensitivity of 28% [Johnston et al, 1997] versus
80% [Leyh et al, 1997]]; telomerase (specificity of
24% [Cassel et al, 2001] versus 100% [Kinoshita et
al, 1997]]; and cytokeratin 8/18 (UBC) (sensitivity of
12.1% [Babjuk et al, 2008] versus 70.5% [Hakenberg
et al, 2004]].

Currently, there is no consensus regarding the use of


urinary biomarkers for the diagnosis of bladder cancer.
While results of many carefully conducted clinical
studies have been reported, randomized clinical
trials are needed to determine the appropriate use
of urinary biomarkers in diagnosis and surveillance
of bladder cancer.

REFERENCES
1. Babjuk M, Soukup V, Pesl M, Kostirova M, Drncova E,
Smolova H, et al. Urinary cytology and quantitative BTA and
UBC tests in surveillance of patients with pTapT1 bladder
urothelial carcinoma. Urology 2008; 71: 718-7. Level of
Evidence 2

Results for FISH appear more consistent. In a recent


meta-analysis of 14 studies involving 2,477 FISH
tests, for all studies combined, the sensitivity of FISH
was 72% (69-75%), and the specificity was 83% (8285%). In comparison, for all studies combined, the
sensitivity of cytology was 42% (38-45%), and the
specificity was 96% (95-97%). However, when cases
of non muscle invasive cancer were excluded from
the analysis, only a marginal difference was noted
in the performance of FISH and cytology [Hajdinjak,
2008].

2. Babjuk M, Soukup V, Pesl M, Kostrov M, Drncov E,


Smolov H, Szakacsov M, Getzenberg R, Pavlk I,
Dvorcek J. Urinary cytology and quantitative BTA and
UBC tests in surveillance of patients with pTapT1 bladder
urothelial carcinoma. Urology 2008; 71: 718-22. Level of
Evidence 2
3. Bastacky S, Ibrahim S, Wilczynski SP, et al. The accuracy
of urinary cytology in daily practice. Cancer 1999 ; 87: 11828. Level of Evidence 2
4. Britton JP, Dowell AC, Whelan P. Dipstick haematuria and
bladder cancer in men over 60: results of a community
study. BMJ 1989; 299: 10102. Level of Evidence 2

27

5. Cassel A, Rahat MA, Lahat N, Lindenfeld N, Mecz Y, Stein


A. Telomerase activity and cytokeratin 20 as markers for
the detection and followup of transitional cell carcinoma:
an unfulfilled promise. J Urol 2001; 166: 841-4. Level of
Evidence 3

20. Johnston B, Morales A, Emerson L, Lundie M. Rapid


detection of bladder cancer: a comparative study of point of
care tests. J Urol 1997; 158: 2098-101. Level of Evidence
3
21. Khadra MH, Pickard RS, Charlton M, Powell PH, Neal DE.
A prospective analysis of 1,930 patients with hematuria to
evaluate current diagnostic practice. J Urol 2000;163:5247.
Level of Evidence 2

6. Chahal R, Darshane A, Browning AJ, Sundaram SK.


Evaluation of the clinical value of urinary NMP22 as a
marker in the screening and surveillance of transitional cell
carcinoma of the urinary bladder. Eur Urol 2001; 40: 41520. Level of Evidence 2

22. Kinoshita H, Ogawa O, Kakehi Y, Mishina M, Mitsumori K,


Itoh N, et al. Detection of telomerase activity in exfoliated
cells in urine from patients with bladder cancer. J Natl
Cancer Inst 1997; 89: 724-30. Level of Evidence 3

7. Chlapoutakis K, Theocharopoulos N, Yarmenitis S, Damilakis


J. Performance of computed tomographic urography in
diagnosis of upper urinary tract urothelial carcinoma, in
patients presenting with haematuria: systematic review
and meta-analysis. Eur J Radiol 2010;73:3348. Level of
Evidence 2

23. Leyh H, Hall R, Mazeman E, Blumenstein BA. Comparison


of the Bard BTA test with voided urine and bladder wash
cytology in the diagnosis and management of cancer of the
bladder. Urology 1997; 50: 49-53. Level of Evidence 2

8. Chou R, Dana T. Screening adults for bladder cancer:


Update of the 2004 evidence review for the US Preventive
Services Task Force. Rockville (MD): agency for Healthcare
Research and Quality (US): 2010 Oct.

24. Lokeshwar VB, Habuchi T, Grossman HB, Murphy WM,


Hautmann SH,et al. Bladder tumor markers beyond
cytology: International Consensus Panel on Bladder Tumor
Markers. Urology 2005; 66(6 Suppl 1): 35-63. Level of
Evidence 2

9. Datta SN, Allen GM, Evans R, Vaughton KC, Lucas


MG. Urinary tract ultrasonography in the evaluation of
haematuria: a report of over 1000 cases. Ann R Coll Surg
Engl 2002;84:20-5. Level of Evidence 2

25. Messing EM, Teot L, Korman H, Underhill E, Barker E, Stork


B, et al. Performance of urine test inpatients monitored
for recurrence of bladder cancer: a multicenter study in
the United States. J Urol 2005; 174: 123841. Level of
Evidence 1

10. Edwards TJ, Dickinson AJ, Natale S, Gosling J, McGrath


JS. A prospective analysis of the diagnostic yield resulting
from the attendance of 4020 patients at a protocol-driven
haematuria clinic. BJU Int 2006;9 7:3015. Level of
Evidence 2

26. Mishriki SF, Nabi G, Cohen NP. Diagnosis of urologic


malignancies in patients with asymptomatic dipstick
hematuria: prospective study with 13 years follow-up.
Urology 2008;71:136. Level of Evidence 2

11. Edwards TJ, Dickinson AJ, Gosling J, McInerney PD,


Natale S, McGrath JS. Patient-specific risk of undetected
malignant disease after investigation for haematuria, based
on a 4-year follow-up. BJU Int ePub Aug 2010. Level of
Evidence 2b

27. Miyanaga N, Akaza H, Tsukamoto T, Ishikawa S, Noguchi


R, Ohtani M, et al. Urinary nuclear matrix protein 22 as a
new marker for the screening of urothelial cancer in patients
with microscopic hematuria. Int J Urol 1999; 6: 173-7. Level
of Evidence 2

12. Ferra S, Denley R, Herr H, Dalbagni G, Jhanwar S, Lin


O. Reflex UroVysion testing in suspicious urine cytology
cases. Cancer 2009; 117: 7-14. Level of Evidence 3

28. Mowatt G,Zhu S,Kilonzo M,Boachie C,Fraser C,Griffiths


TR,NDow J,Nabi G,Cook J,Vale L. Systematic review
of the clinical effectiveness and cost-effectiveness of
photodynamic diagnosis and urine biomarkers (FISH,
ImmunoCyt, NMP22) and cytology for the detection and
follow-up of bladder cancer. Health Technol Assess.2010;
14: 1-331, iii-iv. Level of Evidence 2

13. Fritsche HA, Grossman HB, Lerner SP, Sawczuk I. National


Academy of Clinical Biochemistry Guidelines for the Use
of Tumor Markers in Bladder Cancer. Clin Chem 2010; 56:
e1-48Level of Evidence 1
14. Glas AS, Roos D, Deutekom M, Zwinderman AH, Bossuyt
PM, Kurth KH. Tumor markers in the diagnosis of primary
bladder cancer. A systematic review. J Urol 2003; 169:
197582. Level of Evidence 2a
15. Grossfeld GD, Litwin MS, Wolf JS Jr, Hricak H, Shuler
CL, Agerter DC, Caroll PR. Evaluation of asymptomatic
microscopic hematuria in adults: the American Urological
Association best practice policy--part II: patient evaluation,
cytology, voided markers, imaging, cystoscopy, nephrology
evaluation, and follow-up. Urology 2001; 57: 604-10 Level
of Evidence 2
16. Grossman HB, Soloway M, Messing E, et al. Surveillance
for recurrent bladder cancer using a point-of-care proteomic
assay. JAMA 2006; 295: 299305. Level of Evidence 2
17. Hajdinjak T. UroVysion FISH test for detecting urothelial
cancers: meta-analysis of diagnostic accuracy and
comparison with urinary cytology testing. Urol Oncol 2008;
26: 646-51. Level of Evidence 2a
18. Hakenberg OW, Fuessel S, Richter K, Froehner M,
Oehlschlaeger S, Rathert P, Meye A, Wirth MP. Qualitative
and quantitative assessment of urinary cytokeratin 8 and
18 fragments compared with voided urine cytology in
diagnosis of bladder carcinoma. Urology 2004; 64: 1121-6.
Level of Evidence 2
19. Horstmann M, Patschan O, Hennenlotter J, Senger E, Feil
G, Stenzl A. Combinations of urine-based tumour markers
in bladder cancer surveillance. Scand J Urol Nephrol 2009;
43: 461-6. Level of Evidence 2

29. Planz B, Jochims E, Deix T, Caspers HP, Jakse G,


Boecking A. The role of urinary cytology for detection of
bladder cancer. Eur J Surg Oncol 2005; 31: 3048. Level of
Evidence 2
30. Rodgers MA, Hempel S, Aho T, Kelly JD, Kleijnen J,
Westwood M. Diagnostic tests used in the investigation
of adult haematuria: a systematic review. BJU Int 2006;
98:1154-60. Level of Evidence 2
31. Tritschler S, Scharf S, Karl A, Tilki D, Knuechel R, Hartmann
A, Stief CG, Zaak D. Validation of the diagnostic value of
NMP22 BladderChek test as a marker for bladder cancer
by photodynamic diagnosis. Eur Urol 2007; 51: 403-7.
Level of Evidence 2
32. Tritschler S, Sommer ML, Straub J, Hocaoglu Y, Tilki
D, Strittmatter F, Zaak D, et al. Urinary cytology in era
of fluorescence endoscopy: redefining the role of an
established method with a new reference standard. Urology
2010; 76: 677-80. Level of Evidence 3
33. Yang MG, Zhao XK, Xiao N. Meta analysis of fluorescence
in situ hybridization and cytology for the diagnosis of
bladder cancer. Chin J Cancer 2009; 28: 655-62. Level of
Evidence 2
34. Youssef RF, Schlomer BJ, Ho R, Sagalowsky AI, Ashfaq R,
Lotan Y. Role of fluorescence in situ hybridization in bladder
cancer surveillance of patients with negative cytology. Urol
Oncol 2010 May 5. Urol Oncol 2010 May 5 Epub ahead of
print Level of Evidence 2

28

The large bore of rigid cystoscopes typically allows


for excellent irrigation flow and visualization, even
when there is mildly to moderately bloody urine or
debris, and provides a port that can accommodate
a variety of instruments. However, because of their
large size and rigid nature, rigid cystoscopes typically
cannot be used effectively in the office; most often,
particularly in men, rigid cystoscopes are used in
an operating room with the patient under general or
some form of regional anesthesia. In addition, for a
rigid cystoscope to be used effectively, the patient
must be placed in a dorsal lithotomy position, and in
the occasional patient this may pose difficulties.

B. Endoscopic Examination of
the Lower Urinary Tract
Sam S. Chang, Peter E. Clark, Judson D. Davies,
Murugesan Manoharan, Tim S. OBrien, Amit Patel,
Jinhai Fan
One of the cornerstones of the diagnosis and
management of urothelial carcinoma is cystoscopic
examination of the lower urinary tract. In this
section, we will provide an overview of white light
cystoscopy (WLC); the techniques of lower urinary
tract endoscopy; and newer optical technologies
that can enhance endoscopy, including fluorescence
cystoscopy, narrow band imaging (NBI), Raman
spectroscopy (RS), and optical coherence
tomography (OCT). The evidence presented in
this section is level 3 unless stated otherwise, and
all recommendations presented in this section are
grade C unless stated otherwise.

2. Flexible Cystoscopy
Rigid cystoscopy was the standard of care in urology
for many years, but starting in the early 1970s, the
advent of better fiberoptic technologies permitted
the development of flexible instruments that could
be used more easily than rigid cystoscopes in the
office. The first published report of use of a flexible
fibercystoscope for examination of the bladder
neck was by Tsuchida and Sugawara in 1973.[1] This
report was followed over a decade later, in 1984,
by the development of the first commercial flexible
instrument built specifically for cystoscopy. Since
that time, the use of flexible fiberoptic cystoscopy
has grown rapidly, and flexible cystoscopy is now a
standard method for diagnosis and surveillance of
a variety of lower urinary tract disorders, including
urothelial carcinoma. The fact that flexible cystoscopy
can generally be performed without anesthesia has
led to widespread acceptance of this procedure; it
is now the most common office-based procedure
performed by urologists in developed countries such
as the United States. Over time, manufacturers have
developed improved optics that permit smaller-caliber
scopes while maintaining high image quality, and the
development of a working channel allowed for the
use of flexible instruments. This, coupled with the
advent of active deflection, has permitted a range of
office-based procedures that can potentially be done
without the need for general anesthesia. In addition
to decreased discomfort, the other advantage of
flexible cystoscopy is the use of active deflection to
improve visualization of the anterior bladder neck.

I. White Light Cystoscopy


White light endoscopic examination of both the
urethra and the bladder (white light cystoscopy;
WLC) remains the gold standard for screening and
diagnosis for multiple diseases of the lower urinary
tract, including urothelial carcinoma. Cystoscopy not
only permits visualization of the bladder urothelium
but also affords access to the ureteral orifices to
facilitate assessment and treatment of the upper
urinary tract.
Cystoscopy can be performed utilizing either rigid
or flexible endoscopes, depending on the clinical
circumstances.The standard caliber measurement
for all endoscopes is based on the French (Fr) scale,
in which 1 Fr equals 1/3 mm (for example, a 12-Fr
endoscope has a diameter of 4 mm). Endoscopes
from 8-12 Fr are typically used for pediatric patients,
and endoscopes from 16-28 Fr are typically used for
adult patients.

1. Rigid Cystoscopy
For screening and diagnostic work, most often
sheaths of 20-22 Fr are used for adult patients,
and smaller sizes are used for pediatric patients.
For suspected urethral neoplasms, a 0-degree lens
is useful. For examination of the prostatic urethra,
the bladder trigone, and the bladder wall except
the part immediately adjacent to the bladder neck,
a 30-degree lens is generally used. For more
acute or difficult angles, particularly at the anterior
bladder neck, a 70-degree or 90-degree lens may be
employed, although in such cases, use of adjunctive
implements, such as catheters or biopsy forceps, is
typically not possible without specialized deflecting
bridge adaptors.

Despite their advantages and widespread use,


flexible cystoscopes have some disadvantages
compared to rigid endoscopes, such as the small
irrigation port and lack of a separate working sheath,
which limits the ability to irrigate and use instruments
simultaneously. Typically only the smallest of tumors
can be ablated or biopsied using only flexible
endoscopy. Finally, flexible endoscopes are typically
more costly and more prone to damage than rigid
endoscopes if not handled properly, including during
cleaning and sterilization between cases.
The flexible fiberoptic cystoscopes commonly in

29

use today also have some optical disadvantages


compared to rigid endoscopes. These disadvantages
are due to the inherent imaging limitations of fiberoptic
technology. The diameter of the glass fibers carrying
the image is finite; this results in a pixelated or
screen door image appearance.[2] The shaft of a
flexible cystoscope typically has 3 fiberoptic bundles.
Two carry the light from the generating source, and
the third carries the image back to the eye piece.
As a consequence, the image obtained is actually
a composite matrix of each of the individual fibers
in the bundle. This creates an image that is in fact
the merging of the individual dots in the matrix into
1 image, analogous to what is seen in newspaper
images. This is an inherent limitation of fiberoptic
technology, but digital endoscopy, as discussed
below, overcomes these limitations.

First introduced commercially in 2005, digital


endoscopes have been studied to a limited degree
in the clinical setting. Quayle et al[6] found that
digital-sensor-based flexible scopes had better
optics than fiberoptic cystoscopes. Okhunov et al
conducted a prospective clinical trial comparing
digital cystoscopes to fiberoptic cystoscopes for
office-based flexible cystoscopy in over 1,000
patients.[7] These authors reported that operating
surgeons found the digital cystoscopes to be lighter
and easier to maneuver. There was better deflection
with instruments placed in the working channels
of digital endoscopes without apparent loss of
instrument functionality. Both types of endoscope
were found to be relatively durable; both had a 0.2%
repair rate over the course of the study. It should be
emphasized that digital endoscopes are often more
expensive, so whether they will ultimately prove to be
cost-effective remains to be seen. The development
of digital endoscopes has, however, permitted the
development of other promising technologies to be
discussed later in this chapter, such as NBI.

3. Advances in White Light


Cystoscopy (WLC)
Video cameras as an adjunct to endoscopy were
first introduced by French researchers in 1956,[2]
and this concept has led to improved ergonomics,
improved safety through avoidance of contact with
body fluids, enhanced patient and resident education,
and improved documentation of findings and sharing
of information among physicians through the use
of digital cameras and recording devices.[3] One
study showed that male patients who were able to
monitor their cystoscopy by watching the procedure
on a video monitor experienced up to 40% less
pain and discomfort during endoscopy compared
to patients who did not watch their procedure on a
video monitor.[4]

4. Conclusion
WLC, whether performed in the office utilizing flexible
instruments or performed in the operating room
using rigid instruments, is the standard approach
to diagnosis and management of lower urinary tract
disease, including urothelial carcinoma, and is the
gold standard against which other approaches must
be compared.

II. FLUORESCENCE CYSTOSCOPY

Perhaps the biggest recent advance in WLC has


been the development of digital endoscopy. With
this method, light from the generating source still
travels through the traditional fiberoptic imaging
bundle. However, the image is no longer carried
back along fiberoptic bundles to the eye piece.
Instead, there is a digital sensor at the very tip of the
endoscope that is based on a charge-coupled device
and complementary metal oxide semiconductor
chips. The image is acquired through the digital
sensor through millions of photodiodes that convert
the photons of light into an electric current that
is subsequently transformed into voltage, then
amplified and converted to a digital format.[2] The
semiconductor chips transfer the information to a
receiver that presents the image on a monitor and/
or stores it digitally.

1. P
 hotodynamic Diagnosis of
Bladder Tumors
Non-muscle-invasive bladder cancer is associated
with a significant risk of recurrence; thus, lifelong
patient surveillance is required. Recurrence of nonmuscle-invasive bladder cancer has been postulated
to result either from failure to identify abnormal
mucosal areas or from incomplete resection under
conventional WLC. After WLC and resection,
reported rates of residual or recurrent disease
at follow-up WLC can be as high as 55%.[9-11]
Fluorescence cystoscopy (photodynamic diagnosis;
PDD) can improve the visualization of bladder tumors
compared to standard WLC. Improved visualization
has led in turn to reduced rates of residual tumor at
first-check cystoscopy and in some series to reduced
rates of recurrence.[9-11]

Digital endoscopy offers the promise of improved


optical resolution, contrast, and color differentiation
and may be more durable than traditional flexible
endoscopy. Ex vivo studies have suggested that
distal endoscopes are superior to more traditional
fiberoptic cystoscopes in terms of resolution, contrast
discrimination, and red color differentiation.[5]

However, many of the studies to date on PDD have


shortcomings, including inclusion in the studies of
both patients with newly presenting and patients
with recurrent tumors; failure to use the best current

30

standard of carefor example, transurethral resection


plus immediate single instillation of intravesical
mitomycin Cin the control arm of studies; use of
surrogate endpoints such as histology rather than
the more clinically relevant endpoint of recurrence;
and, in older studies, use of the photosensitizer
5-aminolevulinate (5-ALA) rather than the currently
licensed product hexylaminolevulinate (Hexvix,
Cysview; see the next section).

At the molecular level, a photon of light interacts with


the electron shell of a single PPIX molecule, causing
excitation. Fluorescence is emitted as the electron
returns to its ground state. When tissues with high
levels of PPIX are illuminated by a blue light, they
fluoresce red (Figure 1).

b) Practical Considerations
Fluorescence cystoscopy uses a normal rigid
cystoscope combined with an excitation light source
called a D-light. The D-light consists of a shortarc xenon lamp and an integrated observation
filter to filter out all but the fluorescence emission
wavelength that is of interest and enough excitation
blue light back-scattered to suppress nonspecific red
fluorescence. Urine fluoresces yellow-green under
fluorescence (blue light) cystoscopy; therefore,
the bladder should be emptied before blue light
illumination.

a) Principle of PDD
PDD relies on an endogenous or exogenous
photosensitizer that causes abnormal cells to
fluoresce under a specific wavelength of light.
Currently, the most frequently used exogenous
photosensitizers are derived from the heme
biosynthetic pathway. Nonactive molecules are
converted in a series of enzymatic reactions to
the active fluorescent molecule protoporphyrin
IX (PPIX), which under normal conditions is then
rapidly converted to heme. This metabolic reaction
takes place in all nucleated cells.

Lowering of the intensity of the bright blue light to


the tone of fluorescence results in images 10 times
darker than those produced by WLC. As a result of
the darker image, the video photography is done
at 15 rather than 50 frames per second, and the
cystoscope needs to be moved slowly to reduce
strobe effect. It is recommended that the bulk of
tumor resection be carried out under WLC with blue
light reserved for checking the resection margins.

The starting point of the heme biosynthetic pathway


is 5-ALA. Administration of exogenous 5-ALA causes
rapid rises in the levels of PPIX. Fluorescence
microscopy following administration of 5-ALA has
shown tumor selectivity of PPIX accumulation,
which is at least 5 times as great in tumors as in
normal urothelium.[1] Significant differences in heme
metabolism have been demonstrated between
tumor and normal cells, and these differences
promote the selective accumulation of PPIX.[1]
Higher concentrations of intracellular 5-ALA are
from the reduced diffusion barrier to the absorption
of exogenous 5-ALA. This is in part due to fewer
intercellular tight junctions and recruitment of
transmembrane channel proteins, which provide
active transport of 5-ALA into the cytoplasm.[2]

The intensity of PPIX fluorescence diminishes with


time, a process referred to as photobleaching.
Illumination causes the degradation of the tetrapyrole
ring of PPIX and return of the delocalized electrons
to ground state and a reduction in fluorescence. This
effect is greatest at the blue-light wavelength but
also occurs under white-light conditions, although
at a less rapid rate. Reducing the viewing distance
between the cystoscope and the urothelium also
increases the rate of photobleaching. Fluorescence
is greatly reduced when the viewing distance is
halved.[3]

Fluorescence of PPIX is achieved by the presence


of pyrole rings, which have photo-optical properties
and fluoresce when excited by light of certain
wavelengths. PPIX absorbs light at around 400 nm
(violet blue) and emits fluorescence at 635 nm (red).

Artifact fluorescence is defined as an increase


in fluorescence intensity not due to tumor. False

Figure 1. Cystoscopy of papillary bladder tumours under white (left) and blue (right) cystoscopy

31

fluorescence is usually a less-bright pink color and


less discrete than the fluorescence in tumor cells
and occurs at sites of thickened urothelium. Folds
of the urothelium will cause such artifact, and
therefore the bladder should be sufficiently full to
minimize folds. Other common sites where artifact
fluorescence is seen are along blood vessels, on the
trigone, and around the ureteric orifices. Squamous
metaplasia at the bladder neck and the trigone
produces nonspecific fluorescence, which is used
as a control to confirm the adequate absorption
of the fluorophore when there is no fluorescence
elsewhere in the bladder. The viewing angle
between the cystoscope and the bladder can also
give rise to false fluorescence due to increasing the
thickness of the urothelium with tangential viewing
angles. This problem can be overcome by using
a range of cystoscopes to distinguish pathological
from tangential fluorescence.

underwent WLC resection and 11% of those who


underwent PDD resection [20]. If validated, these
results would indicate improved visualization of the
tumor with PDD and thus more complete resection.
Improved detection and resection of bladder tumors
with PDD has been reported to improve recurrencefree survival compared to WLC. Stenzl et al
performed a multicenter randomized controlled trial
and over a 9-month follow-up period found that the
tumor recurrence rate was lower in the PDD group
(47%; 128 of 271) than in the white light group (56%;
157 of 280) (p=0.026); the relative reduction in the
recurrence rate was 16% [21]. Trials with a longer
follow-up duration but using 5-ALA have shown that
the reduction in recurrence extends up to 8 years of
follow-up.

3. E
 xamination of Patients with
Positive Findings on Voided Urine
Cytology but Negative Findings on
WLC (Grade C)

Recommendations for Use of PDD


PDD is currently recommended for treatment
of patients with newly presenting non-muscleinvasive bladder tumors; examination of patients
with positive findings on voided urine cytology but
negative findings on WLC; treatment of multifocal
recurrent bladder tumors; and teaching of urological
surgeons in training. Each of these situations is
discussed below.

PDD has been shown to detect new bladder tumors


in approximately 30% of patients with positive
findings on urine cytology and negative findings
on urinary tract investigations [23]. This rate can
rise to approximately 66% in patients previously
treated for bladder cancer. Patients with negative
findings on PDD and normalized cytology usually
have a clear urinary tract for at least 2 years[14]. If
findings on cytology remain positive, an abnormality
is likely to be detected within 1 year. Improved
detection of bladder lesions can reduce the need
for bilateral ureterorenoscopy and stents and their
associated morbidity. Tritschler et al evaluated the
use of fluorescence endoscopy and found that the
sensitivity and specificity of bladder wash cytology
was greater with fluorescence cystoscopy than with
white light endoscopy. However, 3.4% of tumors
were diagnosed by bladder wash cytology only, and
most of these were high-grade tumors [Tritschler et
al, 2010].

2. Evaluation and Treatment of


Patients with Newly Presenting
Non-muscle-invasive Bladder
Tumors (Grade C)
The advantage of WLC plus PDD over WLC alone
in the visualization and staging of bladder tumors is
widely recognized.[12-17] The tumor detection rate is
73-96% with WLC alone, compared to 90-96% with
WLC plus PDD.[12-17] PDD is particularly helpful
in the detection of carcinoma in situ (CIS); rates of
detection are 23-68% with WLC alone, compared to
91-97% with WLC plus PDD. While PDD improves
sensitivity in the diagnosis of non-muscle-invasive
bladder tumors, there remain issues surrounding its
specificity. Early studies suggested a slightly higher
false-positive rate of 3-8%[13-15] due to nonspecific
inflammation that can occur after treatment of
a urinary tract infection, after TURBT, or after
intravesical therapy with bacillus Calmette-Gurin
(BCG). It is therefore recommended that PDD be
delayed for 9-12 weeks after any activity that could
result in bladder inflammation [18]. Intravesical
mitomycin C is not associated with a higher falsepositive rate [19].

4. T
 reatment of Multifocal
Recurrent Bladder Tumors
(Grade C)
Patients with multifocal bladder tumors are at high
risk for the development of recurrence. Holmang
et al reported that 88% of patients with 3 or more
tumors experienced recurrence within 5 years after
diagnosis [24]. Ray et al showed that PDD detected
additional disease in 8 (57%) of 14 patients with
already confirmed recurrence [18]. Four patients in
whom additional disease was detected were found
to have CIS, and thus had treatment with BCG
initiated. Only the results from a phase II trial as
to whether PDD in patients with multifocal disease
affects recurrence rates on follow-up, the results of
phase III trials are awaited.

PDD may also be able to improve the quality of


resection compared to WLC. Geavlete et al reported
at repeat TURBT 6 weeks after initial resection that
residual tumor was found in 31% of patients who

32

5. Teaching Urological Surgeons in


Training (Grade D)

follow-up of non-muscle-invasive bladder cancer.


However, these techniques have a low sensitivity
and other limitations. First, CIS may easily be
missed by standard WLC. Second, no information
on histopathologic information is provided during
WLC, which often makes discrimination between
inflammatory lesions and CIS difficult because both
can appear as mucosal red spots. Finally, estimates
of grade or stage of bladder cancer made during
cystoscopy are often not accurate even when
they are made by an experienced urologist [26].
New optical imaging modalities for bladder cancer
diagnosis have been developed to overcome the
limitations of cystoscopy and cytology. These new
techniques include PDD, NBI, RS, and OCT. We
herein present our review of these new techniques
for detecting bladder cancer (except for PDD, which
was discussed earlier in the chapter) and speculate
regarding their potential future applications.

PDD offers endoscopists better visualization of tumors


and their resection margins, thereby facilitating
improved transurethral resection of bladder tumor
(TURBT) technique. Thus, PDD is a useful tool for
teaching urological surgeons in training.

6. Additional Considerations
Whether early repeat cystoscopy is required after
initial tumor resection with PDD is not yet known.
PDD cystoscopy could potentially help rationalize
the use of early resection, optimize case selection,
improve outpatient management modalities, and
permit more precise tumor resection. PDD is not
currently recommended in patients for whom
cystectomy is indicated or in the outpatient setting
with flexible cystoscopes. Although PDD flexible
cystoscopy has been shown to be feasible and as
effective as PDD rigid cystoscopy [25], the need for
a second cystoscopy under a general anesthetic for
biopsy has not made this approach cost-effective.

1. Narrow Band Imaging


a) Principle of NBI
For NBI, modified optical filters are used in the
light source of a video endoscope system to
narrow the bandwidth of spectral transmittance.
NBI enhances the differences in penetration depth
between wavelengths. Light penetration depth within
tissue is highly dependent on wavelength: shorter
wavelengths produce only superficial penetration
and longer wavelengths produce deeper penetration.
Blue light, therefore, penetrates supercially, while
red light penetrates deeply [27]. NBI narrows the
bandwidth of light output from the endoscope
system to between 415 nm and 540 nm. The relative
intensities of blue and green light are increased,
while the intensity of red light is decreased to a
minimum. This narrow bandwidth of green and blue
light is strongly absorbed by hemoglobin, so NBI
enhances visibility of surface capillaries and blood
vessels in the submucosa without the use of dyes
(Figure 2). Systems that have integrated NBI and
WLC are already commercially available. With the

Conclusions
Non-muscle-invasive bladder cancer is associated
with a significant risk of recurrence; thus, lifelong
patient surveillance is required. PDD can improve
detection of disease, especially CIS; improve
completeness of resection; and, most importantly,
reduce rates of recurrence of non-muscle-invasive
bladder cancer. Improved initial resection, such as
with PDD, has been suggested to improve recurrence
rates after intravesical chemotherapy, but this needs
to be further evaluated in clinical trials.

III. New OPTICAL IMAGING


Techniques for Bladder
Cancer DIAGNOSIS
Cystoscopy and cytology are considered to
be the current gold standard for detection and

Figure 2. Representation of the narrowed bandwidth of light (a), its penetration of tissue (b)

33

push of a button, the NBI mode is activated by


mechanical insertion of a narrow-band filter in front
of the white light source.

Possible limitations of the currently published studies


are their monocentric nature and the possible
observer bias since WLC and NBI were performed
sequentially and observed by the same urologist. Ye
et al conducted a study designed to compare the rate
of detection of non-muscle-invasive bladder cancer by
NBI cystoscopy versus WLC in a multicenter setting
with a randomized sequence of the 2 procedures.
One hundred three consecutive patients from 4
academic centers in China were included in this
prospective, multicenter, phase III trial presented in
abstract form in 2010 [34]. Eighty-seven patients had
confirmed bladder tumors. In 29 of these 87 patients,
tumors were detected by NBI cystoscopy only, and in
4 patients, tumors were identified by WLC only. The
mean number of detected tumors per patient was
1.45 with WLC and 1.85 with NBI cystoscopy.

b) NBI Cystoscopy for Detection of Nonmuscle-invasive Bladder Cancer


NBI cystoscopy is easy for surgeons to adopt and
has a high sensitivity for detecting small papillary and
otherwise undetectable cancers, especially CIS.
NBI was initially shown to be useful in gastrointestinal
disease, particularly for detection of adenomas
at colonoscopy and for follow-up of Barretts
esophagus [30]. The first report of NBI in bladder
cancer was published by Bryan et al in 2008 [31].
These authors performed WLC and subsequent NBI
flexible cystoscopy to detect bladder cancer in 29
patients with recurrent non-muscle-invasive bladder
cancer [31]. NBI flexible cystoscopy provided
much better visualization of bladder cancer than
did conventional flexible WLC. NBI cystoscopy
revealed 15 additional tumors (in 12 patients) not
detected by WLC. However, these additional tumors
were not confirmed histopathologically because all
tumors were treated by diathermy ablation following
cystoscopic examination. With NBI cystoscopy, the
vasculature appears dark green or black against the
almost white normal urothelium, whereas with WLC,
tumors appear red in a background of normal pink
urothelium (Figure 3).

c) NBI Cystoscopy for Detection of Residual


Cancer after Treatment
Two trials to detect residual cancer by NBI cystoscopy
were conducted in patients with high-risk nonmuscle-invasive bladder cancer. In a series of 47
patients, all patients were examined with WLC and
NBI cystoscopy and underwent a second TURBT
procedure approximately 1 month after an initial
TURBT. Overall, 16 of 47 patients were discovered to
have residual or recurrent cancer, including 6 patients
with high-grade cancers detected only by NBI [35]. In
another series, 61 patients were evaluated with both
WLC and NBI cystoscopy 3 months after beginning
induction therapy with BCG. NBI correctly identied
21 of 22 cases of residual cancer [36].

Sensitivity and specicity rates for NBI cystoscopy and


WLC in patients with non-muscle-invasive bladder
cancer and CIS are presented in Table 1. Herr et al
described a series of 427 patients who underwent
follow-up with WLC and NBI cystoscopy, 103 of
whom had tumor recurrence [32]. Ninety patients
(sensitivity 87%) had their disease diagnosed by
WLC, and the other 13 (sensitivity 100%) had their
disease detected only by NBI cystoscopy, including
8 patients with CIS, 4 patients with Ta disease, and 1
patient with T1 disease. NBI cystoscopy performed
better than WLC in demarcating margins of CIS
lesions from surrounding normal-appearing mucosa.
In another series, Tatsugami et al performed WLC
and then NBI cystoscopy in 104 consecutive patients
[27]. In 39 (26.9%) of 161 sites suspicious for tumor,
bladder tumors were identied only by NBI. These
tumors were CIS in 25 patients, Ta tumors in 12
patients, and T1 tumors in 2 patients. In this entire
series, 14 of 30 patients with CIS had their disease
detected only by NBI. However, the aforementioned
studies might be flawed owing to observer bias since
WLC and NBI cystoscopy were performed by the
same urologist. To avoid observer bias, Cauberg et al
conducted a trial in which WLC and NBI cystoscopy
were performed by different surgeons in 95 patients
[33]. Seventy-eight patients had histopathologically
conrmed non-muscle-invasive bladder cancer. NBI
identied additional tumors in 28 of those 78 patients.
In contrast, WLC identied additional tumors in only
3 of those 78 patients.

d) NBI Cystoscopy as a Follow-Up Tool


Another trial compared the usefulness of WLC and
NBI cystoscopy as follow-up tools. Patients were
followed up with WLC for 3 years and with NBI
cystoscopy for the next 3 years. The trial revealed
fewer tumor recurrences, smaller numbers of
recurrent tumors, and a longer recurrence-free
survival time with NBI cystoscopy. Small solitary
papillary tumors and clusters of papillary tumors
were more efficiently treated with NBI cystoscopy
than with WLC [37]. However, only patients with
frequently recurrent tumors were included in this
study. Hence, the authors were unable to discriminate
the natural history of disease from an influence of
NBI cystoscopy on outcomes. With longer follow-up
periods, tumors generally recur less frequently [38].
More investigation appears to be indicated to confirm
this trials conclusions.

e) WLC versus NBI Cystoscopy for TURBT


NBI technology is available for rigid cystoscopy and
bladder tumor resection. One feasibility report of
TURBT performed with NBI cystoscopy indicated
shorter operative times, shorter time to catheter

34

Figure 3. Examples of the differences between tumors visualized with WLC (left) and NBI cystoscopy (right).
(Photos were cited from Case study of NBI (Specific Wavelength Light) Vol.1 which was published by Olympus
medical systems Corporation.)

35

Table 1. Comparison of sensitivity and specicity between NBI and WLC for NMIBC
Reference

Number of patients

Tatsugami et al
Herr et al

27

32

Cauberg et al

33

Ye ZQ et al34

NBI
Sensitivity

WLC

Specicity

Sensitivity

Specicity

104

92.7

70.9

57.3

86.2

427

100.0

82.0

87.0

85.0

95

94.7

68.4

79.2

75.5

103

96.4

60.9

75.4

58.6

removal, and shorter time to hospital discharge


with NBI cystoscopy than with WLC, although the
differences were not statistically significant [39]. The
authors speculated that use of NBI cystoscopy for
TURBT would substantially reduce the recurrence
rate of bladder tumors. However, the observer in
this trial was allowed to take a second look with
the alternative form of imaging, which may have
introduced a bias.

associated with inflammatory mucosal appearances


mimicking urothelial tumors (see Figure 3). In view of
its low specificity (60-85%), use of NBI in combination
with other modalities such as urine cytology, RS, or
OCT that allow histopathologic diagnosis may well
be advised.

f) Learning Curve for NBI Cystoscopy

Raman
spectroscopy
measures
molecular
components of tissue both qualitatively and
quantitatively. Photons interact with laser light
to contribute to or derive energy from a tissues
intramolecular bonds. The result is a change in the
bonds vibrational state and emission of a different
wavelength of scattered light [43]. Each molecule has
unique vibrational energy levels and corresponding
wavelength shifts. When a tissue is exposed to laser
light, altered wavelengths from different molecules
combine to form the Raman spectrum, which
depends on the molecular composition of the tissue
being investigated. Using RS, a pseudocolor map of
examined tissue is created. Tissue areas with similar
molecular composition, and therefore with similar
spectra, are depicted in the same color. Such images
are comparable to histopathology and images of
stained tissues. Molecular composition changes
if pathologic transformations occur. Thus, RS can
provide an objective prediction of the pathologic
diagnosis [26].

2. Raman Spectroscopy
a) Principle of RS

Herr et al demonstrated that both experienced and


relatively inexperienced urologists quickly learned
NBI cystoscopy as a supplement to WLC [40]. That
is, use of NBI cystoscopy to detect and characterize
bladder tumors is not associated with a steep learning
curve. Bryan et al confirmed these findings [41].

g) NBI Cystoscopy Controversies


By permitting better visualization of the margins
of non-muscle-invasive papillary and at bladder
lesions, NBI cystoscopy facilitates more thorough
excision of tumor. Fluorescence cystoscopy (PDD)
using 5-ALA or its hexyl ester has been recognized
to improve the detection of non-muscle-invasive
papillary and at bladder lesions compared to WLC
and to reduce the recurrence rate of non-muscleinvasive bladder cancer [17]. However, whether the
better visualization and high sensitivity associated
with NBI cystoscopy compared to WLC likewise
translate into lower tumor recurrence rates cannot
be confirmed. Prospective randomized trials with
long-term follow-up are ongoing to compare the
efficacy of TURBT performed with WLC and with NBI
cystoscopy [39]. Results of these trials are eagerly
awaited.

b) RS for Detection of Non-muscle-invasive


Bladder Cancer
Several trials were performed to compare the
diagnostic accuracy of RS and histology in ex vivo
bladder samples. A study of 15 tumor and nontumor
bladder tissue samples demonstrated that RS had
a sensitivity of 92% and a specificity of 94% in the
diagnosis of bladder cancer [44]. In another series,
of 75 bladder tissue samples, Crow et al showed
that RS could differentiate normal, inflammatory, and
malignant tissue with a sensitivity of 90-95% and a
specificity of 95-98% [46]. Discrimination between
grade 1 or 2 and grade 3 malignant samples was
achieved with sensitivity and specificity of greater
than 93%. The ability to discriminate between CIS,

Because NBI cystoscopy detects ulcers of bladder


mucosa and areas of angiogenesis more readily
than does WLC, the value of NBI cystoscopy in
the diagnosis of interstitial cystitis/painful bladder
syndrome is apparent [42]. One notable deficiency
of NBI cystoscopy for bladder cancer detection is
its low specificity (high false-positive rate), which
leads to unnecessary resection of noncancerous
tissue. Intravesical agents (immunotherapy or
chemotherapy), cystitis, and hematuria may be

36

inflammation, and the changes associated with


intravesical therapy, all of which can appear as red
spots in the bladder mucosa, is an important finding.
Draga et al were the first to perform RS in vivo [47].The
microber optic probes were utilized endoscopically
through a working channel. These authors found
that RS distinguished normal bladder from cancer
with a sensitivity of 85% and a specificity of 79%.
Combining the new diagnostic modalities of RS
and 5-ALA, Grimbergen et al showed that applying
5-ALA affected the Raman spectra of bladder tissues
[48]. Whether combining fluorescence with RS could
improve its diagnostic capability is unknown.
Shapiro et al used RS to detect tumor epithelial cells
in urine samples and in a broader sense to diagnose
cancer in isolated cell samples [49]. Baseline spectral
features were obtained from bladder cancer tissue
for high-grade (T2), low-grade (Ta), CIS, and normal
bladder tissue (Figure 4). The signal obtained from
epithelial cells in urine samples correctly diagnosed
bladder cancer with a sensitivity of 96% (100% of
the high-grade tumors) and a specificity of 90%
(Figure 5). Furthermore, the authors delineated the
thresholds for normal tissue, low-grade tumors, and
high-grade tumors, which makes the use of RS for
bladder cancer diagnosis feasible.

c) Future of RS for Detection of Bladder


Cancer
To date, only a few small-scale clinical trials have
 conducted of RS for detection of non-musclebeen
invasive
bladder cancer. Small, flexible fiberoptic


probes compatible with the working channel of a rigid


or flexible cystoscope have been developed [45]. In
other fields of medicine, findings have been reported
for use of RS to check intraoperative margins during
partial mastectomy and for use of RS to characterize
atherosclerotic plaques during vascular surgery.
In bladder cancer, RS must be targeted at visually
suspicious lesions or lesions identified by other
methods, such as PDD or NBI cystoscopy. The
potential advantage of RS is its ability to provide a
noninvasive, real-time, and objective prediction of
the pathologic diagnosis [46].
Application of RS to urine samples presents an
exciting approach to the diagnosis of bladder cancer,
one with sensitivity and specificity similar to that of
cystoscopy. Maybe one day RS can be used instead
of urine cytology as the preliminary screening study
for bladder cancer. Further investigations are needed
to verify the performance of RS in urine samples
and also in other cellular systems to determine the
applicability of RS to other malignancies.

3. Optical Coherence Tomography


a) Principle of OCT
OCT produces high-resolution, cross-sectional
images of tissue. OCT is similar to B-mode
ultrasound imaging except that OCT measures
reflected infrared light rather than acoustic waves.
The intensity of back-reflected light from structures
within tissue is displayed as a function of depth [52].
Because of its high resolution (approximately 10
micrometers), superior dynamic range (approximate

Figure
4. Raman spectrum obtained from high-grade urothelial carcinoma (UC; dashed) tissue, low-grade
Figure4.Ramanspectrumobtainedfromhighgradeurothelialcarcinoma(UC;dashed)
UC tissue (dotted), and normal tissue (solid). The 1584-cm-1wave is higher in high-grade tumors but is also
present in low-grade tissue (The figure was cited from Shapiro A 49).

tissue,lowgradeUCtissue(dotted),andnormaltissue(solid).The1584cm1waveishigher

inhighgradetumorsbutisalsopresentinlowgradetissue(ThefigurewascitedfromShapir
37
A49).

Page 45: TEXT: The signal obtained from epithelial cells in urine samples correctly diagnosed bladder cancer
with a sensitivity of 96% (100% of the high-grade tumors) and a specificity of 90% (Figure 4).

Raman shift (-cm-1)


Figure 5. Raman spectra of cumulative single, exemplary cells from cases of high-grade (solid bold), lowgrade (dotted), and normal cells (solid) (The figure was cited from Shapiro A 49).

140 dB), and optimal depth of penetration (up to


2-3 mm), OCT is potentially useful for noninvasive
screening for superficial lesions in bladder mucosa
[53]. The 3 anatomic layers of the normal bladder
wall (the urothelium, lamina propria, and muscularis
propria) can be clearly distinguished from one
another with OCT. This is possible because of the
different scattering properties of each distinct layer
[52].

WLC; 97.5% and 78.6%, respectively, for PDD; and


97.5% and 97.9%, respectively for PDD combined
with OCT (Figure 6 and 7). Thus, combining PDD
with targeted OCT increased specicity substantially.
To enhance the accuracy of detection of CIS, Ren
et al used 3-dimensional OCT [60]. Sensitivity and
specificity were 56.5% and 61.5%, respectively,
with 2-dimensional OCT and 95.7% and 92.3%,
respectively, with 3-dimensional OCT.

b) OCT for Detection of Bladder Cancer

c) Limitations of OCT

OCT allows noninvasive acquisition of real-time,


high-resolution images of the bladder and provides
information about the depth of tumor growth. As of
this writing, results of 5 studies on the diagnostic
accuracy of OCT have been published [54-58]. Goh
et al and Sengottayan et al found similar results:
OCT in the diagnosis of muscle-invasive bladder
cancer had a sensitivity of 100% and a specificity
of 90%. However, these studies were flawed by the
inclusion of only a few muscle-invasive tumors.

Some authors point out that most false-positive


findings on OCT are caused by inflammatory lesions
of the bladder. Whether or not previous intravesical
instillations or radiotherapy, both of which can
cause inflammation, influence the accuracy of OCT
is unknown. In addition, OCT may have limited
reliability in the measurements of muscle-invasive
tumors because of insufficient imaging depth.
Furthermore, OCT is not ideal for screening the
entire bladder because, in the absence of visually
suspect lesions, OCT must be used in combination
with other diagnostic methods (e.g., NBI or PDD) to
identify the region of interest [26].

Schmidbauer et al performed WLC followed by


fluorescence cystoscopy (PDD) and OCT scanning
[59]. These investigators found sensitivity and
specicity rates of 69.3% and 83.7%, respectively, for

In contrast to ultrasonography, OCT can be

38

Figure 6. (a) Optical coherence tomography image of normal bladder wall: urothelium (U), lamina propria
(LP), and muscularis layer (MP); (b) OCT probe placed on the bladder wall, as seen during cystoscopy (The
figure was cited from Schmidbauer J [59]).

Figure 7. Pathologic findings on optical coherence tomography imaging: (a)dysplasia; (b) carcinoma in situ;
(c)papillary Ta lesion; (d)T1 lesion; (e)muscle-invasive urothelial cell carcinoma( The figure was cited from
Schmidbauer J [59]).

39

performed at other sites within the urinary tract


through optical fibers without the need for a distal
transducer or transducing medium [52]. The
interpretation of OCT scans requires some training
and is operator dependent. To address this problem,
Lingley-Papadopoulos et al presented an automated
algorithm that uses texture analysis to discriminate
bladder cancer in OCT images [61]. In differentiating
benign from malignant tissue, this algorithm had a
sensitivity of 92% and a specificity of 62%.

performed in vivo and have shown impressive


preliminary results. RS may be utilized to detect
tumor epithelial cells in urine samples, which may
represent a new approach for preliminarily screening
of patients who present with hematuria. However,
more research must be conducted to assess the
advantages of these new techniques.
Since RS and OCT have a limited field of view,
screening the entire bladder with either of these
techniques would be excessively time-consuming.
On the other hand, while NBI cystoscopy and PDD
allow the operator to quickly assess the entire
bladder wall, these techniques suffer from low
specificity. Therefore, a strategy in which RS or OCT
measurements are targeted at visually suspicious
lesions or lesions identified by PDD and/or NBI
cystoscopy may overcome the drawbacks of the
various modalities to increase both the sensitivity
and the specificity of bladder cancer detection.
Encouragingly, the first steps with this approach
have been taken, and promising preliminary data
indicate that RS and/or OCT combined with PDD
and/or NBI likely will lead to better characterization of
non-muscle-invasive bladder cancer. We speculate
that RS for detection of tumor epithelial cells in
urine samples might be the alternative to cytology
for preliminarily screening patients with hematuria
or a history of non-muscle-invasive bladder cancer.
NBI cystoscopy or PDD combined with either RS
or OCT could produce more sensitive and specific
detection, more complete endoscopic resection,
and lower recurrence rates for non-muscle-invasive
bladder cancer. However, whether the combinations
predict a better prognosis and fewer inspections by
cystoscopy is unknown. Also, we must consider the
efficacy of inspection against the cost. More research
has to be conducted before these techniques can
be implemented in the management of non-muscleinvasive bladder cancer.

4. Summary
Compared to WLC, NBI, the newest optical diagnostic
technique, permits an improved evaluation of nonmuscle-invasive bladder cancer, particularly at
the margins of non-muscle-invasive papillary and
at bladder lesions. Furthermore, NBI cystoscopy
significantly enhances tumor detection compared
with WLC. NBI cystoscopy plus WLC appears to
be superior to WLC alone, and NBI cystoscopy is
an easy-to-handle complement to WLC. However,
whether NBI cystoscopy alone is as good as WLC
alone has yet to be established in larger series.
No trials have confirmed the superiority of TURBT
performed with NBI cystoscopy to TURBT performed
with WLC after long-term follow-up. However,
ongoing trials might generate more robust data
concerning these applications of NBI. In future
studies comparing NBI cystoscopy and WLC, we
should avoid the observation bias introduced by
permitting a second look.
PDD has been applied extensively to enhance the
rate of detection of non-muscle-invasive bladder
cancer. However, PDD is costly and time-consuming:
5-ALA is expensive, and the time between instillation
of 5-ALA and fluorescence endoscopy is at least
1 hour. To date, PDD has been shown to have
sensitivity and specificity similar to that of NBI
in the detection of non-muscle-invasive bladder
cancer. It is difficult to compare these 2 modalities
because of different protocols, and no 3-arm study
is yet available. Nevertheless, 3-arm studies for
direct comparison of NBI cystoscopy versus PDD
versus WLC in terms of tumor detection rates, falsepositivity rates, recurrence rates, mortality rates,
and cost would be helpful. However, because of the
different principles underlying NBI cystoscopy and
PDD, the additional tumors identified by them may
not be the same ones.

References
1. Tsuchida, S., Sugawara, H.: A new flexible fibercystoscope
for visualization of the bladder neck. J Urol, 109: 830,
1973
2. Natalin, R. A., Landman, J.: Where next for the endoscope?
Nat Rev Urol, 6: 622, 2009
3. Campbell, M. F., Wein, A. J., Kavoussi, L. R.: CampbellWalsh urology / editor-in-chief, Alan J. Wein ; editors, Louis
R. Kavoussi ... [et al.], 9th ed. Philadelphia: W.B. Saunders,
pp. 4 v. (xlii, 3945, cxv p.), 2007

RS and OCT are 2 new optical diagnostic modalities.


Both of these involve access to the bladder obtained
through the working channel of a cystoscope and are
used to predict histopathologic diagnosis objectively
and in real time. RS can measure the molecular
components of tissue in a qualitative and quantitative
way, while OCT can produce high-resolution,
cross-sectional images of tissue comparable to
histopathology images. Both techniques have been

4. Cornel, E. B., Oosterwijk, E., Kiemeney, L. A.: The effect on


pain experienced by male patients of watching their officebased flexible cystoscopy. BJU Int, 102: 1445, 2008
5. Borin, J. F., Abdelshehid, C. S., Clayman, R. V.: Comparison
of resolution, contrast, and color differentiation among
fiberoptic and digital flexible cystoscopes. J Endourol, 20:
54, 2006
6. Quayle, S. S., Ames, C. D., Lieber, D. et al.: Comparison
of optical resolution with digital and standard fiberoptic
cystoscopes in an in vitro model. Urology, 66: 489, 2005

40

7. Okhunov, Z., Hruby, G. W., Mirabile, G. et al.: Prospective


comparison of flexible fiberoptic and digital cystoscopes.
Urology, 74: 427, 2009

23. Ray, E. R., Chatterton, K., Khan, M. S. et al.:


Hexylaminolaevulinate blue light fluorescence cystoscopy
in the investigation of clinically unconfirmed positive urine
cytology. BJU Int, 103: 1363, 2009

8. Schulze, M., Stotz, N., Rassweiler, J.: Retrospective


analysis of transurethral resection, second-look resection,
and long-term chemo-metaphylaxis for superficial bladder
cancer: indications and efficacy of a differentiated approach.
J Endourol, 21: 1533, 2007

24. Holmang, S., Hedelin, H., Anderstrom, C. et al.: Recurrence


and progression in low grade papillary urothelial tumors. J
Urol, 162: 702, 1999

9. Riedl, C. R., Daniltchenko, D., Koenig, F. et al.: Fluorescence


endoscopy with 5-aminolevulinic acid reduces early
recurrence rate in superficial bladder cancer. J Urol, 165:
1121, 2001

25. Witjes, J. A., Moonen, P. M., van der Heijden, A. G.:


Comparison of hexaminolevulinate based flexible and rigid
fluorescence cystoscopy with rigid white light cystoscopy in
bladder cancer: results of a prospective Phase II study. Eur
Urol, 47: 319, 2005

10. Kriegmair, M., Zaak, D., Rothenberger, K. H. et al.:


Transurethral resection for bladder cancer using
5-aminolevulinic acid induced fluorescence endoscopy
versus white light endoscopy. J Urol, 168: 475, 2002

26. Cauberg, E. C., de Bruin, D. M., Faber, D. J. et al.: A


new generation of optical diagnostics for bladder cancer:
technology, diagnostic accuracy, and future applications.
Eur Urol, 56: 287, 2009

11. Filbeck, T., Pichlmeier, U., Knuechel, R. et al.: Clinically


relevant improvement of recurrence-free survival with
5-aminolevulinic acid induced fluorescence diagnosis in
patients with superficial bladder tumors. J Urol, 168: 67,
2002

27. Tatsugami, K., Kuroiwa, K., Kamoto, T. et al.: Evaluation of


narrow-band imaging as a complementary method for the
detection of bladder cancer. J Endourol, 24: 1807, 2010
28. Kuznetsov, K., Lambert, R., Rey, J. F.: Narrow-band
imaging: potential and limitations. Endoscopy, 38: 76,
2006

12. Schmidbauer, J., Witjes, F., Schmeller, N. et al.:


Improved detection of urothelial carcinoma in situ with
hexaminolevulinate fluorescence cystoscopy. J Urol, 171:
135, 2004

29. Song, L. M., Adler, D. G., Conway, J. D. et al.: Narrow band


imaging and multiband imaging. Gastrointest Endosc, 67:
581, 2008

13. Grossman, H. B., Gomella, L., Fradet, Y. et al.: A phase III,


multicenter comparison of hexaminolevulinate fluorescence
cystoscopy and white light cystoscopy for the detection of
superficial papillary lesions in patients with bladder cancer.
J Urol, 178: 62, 2007

30. East, J. E., Suzuki, N., Stavrinidis, M. et al.: Narrow band


imaging for colonoscopic surveillance in hereditary nonpolyposis colorectal cancer. Gut, 57: 65, 2008
31. Bryan, R. T., Billingham, L. J., Wallace, D. M.: Narrow-band
imaging flexible cystoscopy in the detection of recurrent
urothelial cancer of the bladder. BJU Int, 101: 702, 2008

14. Fradet, Y., Grossman, H. B., Gomella, L. et al.: A comparison


of hexaminolevulinate fluorescence cystoscopy and white
light cystoscopy for the detection of carcinoma in situ in
patients with bladder cancer: a phase III, multicenter study.
J Urol, 178: 68, 2007

32. Herr, H. W., Donat, S. M.: A comparison of white-light


cystoscopy and narrow-band imaging cystoscopy to detect
bladder tumour recurrences. BJU Int, 102: 1111, 2008
33. Cauberg, E. C., Kloen, S., Visser, M. et al.: Narrow band
imaging cystoscopy improves the detection of non-muscleinvasive bladder cancer. Urology, 76: 658, 2010

15. Jocham, D., Witjes, F., Wagner, S. et al.: Improved detection


and treatment of bladder cancer using hexaminolevulinate
imaging: a prospective, phase III multicenter study. J Urol,
174: 862, 2005

34. Ye, Z. Q., Fan, J. H.: Prospective within-patient multicentre phase III trial to compare white-light cystoscope
with narrow band imaging in the detection of non muscleinvasive bladder cancer. J Urol: 1355A, 2010
35. Naselli, A., Introini, C., Bertolotto, F. et al.: Narrow band
imaging for detecting residual/recurrent cancerous tissue
during second transurethral resection of newly diagnosed
non-muscle-invasive high-grade bladder cancer. BJU Int,
105: 208, 2010
36. Herr, H. W.: Narrow-band imaging cystoscopy to evaluate
the response to bacille Calmette-Guerin therapy: preliminary
results. BJU Int, 105: 314, 2010
37. Herr, H. W., Donat, S. M.: Reduced bladder tumour
recurrence rate associated with narrow-band imaging
surveillance cystoscopy. BJU Int, 2010
38. Holmang, S., Hedelin, H., Anderstrom, C. et al.: The
relationship among multiple recurrences, progression and
prognosis of patients with stages Ta and T1 transitional cell
cancer of the bladder followed for at least 20 years. J Urol,
153: 1823, 1995
39. Naselli, A., Introini, C., Bertolotto, F. et al.: Feasibility of
transurethral resection of bladder lesion performed entirely
by means of narrow-band imaging. J Endourol, 24: 1131,
2010
40. Herr, H., Donat, M., Dalbagni, G. et al.: Narrow-band
imaging cystoscopy to evaluate bladder tumours--individual
surgeon variability. BJU Int, 106: 53, 2010
41. Bryan, R. T., Shah, Z. H., Collins, S. I. et al.: Narrow-band
imaging flexible cystoscopy: a new users experience. J
Endourol, 24: 1339, 2010

16. Jichlinski, P., Guillou, L., Karlsen, S. J. et al.: Hexyl


aminolevulinate fluorescence cystoscopy: new diagnostic
tool for photodiagnosis of superficial bladder cancer--a
multicenter study. J Urol, 170: 226, 2003
17. Denzinger, S., Burger, M., Walter, B. et al.: Clinically
relevant reduction in risk of recurrence of superficial bladder
cancer using 5-aminolevulinic acid-induced fluorescence
diagnosis: 8-year results of prospective randomized study.
Urology, 69: 675, 2007
18. Ray, E. R., Chatterton, K., Thomas, K. et al.:
Hexylaminolevulinate photodynamic diagnosis for multifocal
recurrent nonmuscle invasive bladder cancer. J Endourol,
23: 983, 2009
19. Karl, A., Tritschler, S., Stanislaus, P. et al.: Positive urine
cytology but negative white-light cystoscopy: an indication
for fluorescence cystoscopy? BJU Int, 103: 484, 2009
20. Geavlete, B., Jecu, M., Multescu, R. et al.: HAL blue-light
cystoscopy in high-risk nonmuscle-invasive bladder cancer-re-TURBT recurrence rates in a prospective, randomized
study. Urology, 76: 664, 2010
21. Stenzl A, Burger M, Fradet Y. et al.: Hexaminolevulinate
guided fluorescence cystoscopy reduces recurrence in
patients with nonmuscle invasive bladder cancer. J Urol.
2010 ; 184: 1907-13
22. Babjuk, M., Soukup, V., Petrik, R. et al.: 5-aminolaevulinic
acid-induced fluorescence cystoscopy during transurethral
resection reduces the risk of recurrence in stage Ta/T1
bladder cancer. BJU Int, 96: 798, 2005

41

42. Ueda, T., Nakagawa, M., Okamura, M. et al.: New


cystoscopic diagnosis for interstitial cystitis/painful bladder
syndrome using narrow-band imaging system. Int J Urol,
15: 1039, 2008

61. Lingley-Papadopoulos, C. A., Loew, M. H., Manyak, M. J. et


al.: Computer recognition of cancer in the urinary bladder
using optical coherence tomography and texture analysis.
J Biomed Opt, 13: 024003, 2008

43. Crow, P., Stone, N., Kendall, C. A. et al.: The use of


Raman spectroscopy to identify and grade prostatic
adenocarcinoma in vitro. Br J Cancer, 89: 106, 2003

62. Kobayashi, Y., Hayashino, Y., Jackson, J. L. et al.:


Diagnostic performance of chromoendoscopy and narrow
band imaging for colonic neoplasms: a Meta-analysis.
Colorectal Dis, 2010

44. de Jong, B. W., Schut, T. C., Maquelin, K. et al.:


Discrimination between nontumor bladder tissue and tumor
by Raman spectroscopy. Anal Chem, 78: 7761, 2006

C. Transurethral Resection
of Bladder Tumors

45. Koljenovic, S., Bakker Schut, T. C., Wolthuis, R. et al.:


Tissue characterization using high wave number Raman
spectroscopy. J Biomed Opt, 10: 031116, 2005
46. Crow, P., Uff, J. S., Farmer, J. A. et al.: The use of Raman
spectroscopy to identify and characterize transitional cell
carcinoma in vitro. BJU Int, 93: 1232, 2004

I. Preoperative assessment

47. Draga, R. O., Grimbergen, M. C., Vijverberg, P. L. et al.:


In vivo bladder cancer diagnosis by high-volume Raman
spectroscopy. Anal Chem, 82: 5993, 2010
48. Grimbergen, M. C., van Swol, C. F., van Moorselaar, R. J. et
al.: Raman spectroscopy of bladder tissue in the presence
of 5-aminolevulinic acid. J Photochem Photobiol B, 95: 170,
2009

1. Imaging
Prior to TURBT, imaging studies are generally
recommended. Imaging studies should also be
performed in the investigation of hematuria, and hence
they may already have been done prior to TURBT. In
general, imaging is required to evaluate the upper
tracts. Ultrasonography is primarily used to evaluate
the renal parenchyma. Intravenous urography (IVU)
can be used to evaluate the urothelium, and lesions
may be apparent as filling defects on the pyelogram
for the upper tracts. Filling defects within the bladder
are also seen. Increasingly, CT is being performed
rather than intravenous pyelography.[1] CT is faster
and more informative than IVU but is associated with
significantly more radiation exposure (Level 4).

49. Shapiro, A., Gofrit, O. N., Pizov, G. et al.: Raman Molecular


Imaging: A Novel Spectroscopic Technique for Diagnosis of
Bladder Cancer in Urine Specimens. Eur Urol, 2010
50. Haka, A. S., Volynskaya, Z., Gardecki, J. A. et al.: In vivo
margin assessment during partial mastectomy breast
surgery using raman spectroscopy. Cancer Res, 66: 3317,
2006
51. Motz, J. T., Fitzmaurice, M., Miller, A. et al.: In vivo Raman
spectral pathology of human atherosclerosis and vulnerable
plaque. J Biomed Opt, 11: 021003, 2006
52. Tearney, G. J., Brezinski, M. E., Southern, J. F. et al.: Optical
biopsy in human urologic tissue using optical coherence
tomography. J Urol, 157: 1915, 1997
53. Pan, Y., Lavelle, J. P., Bastacky, S. I. et al.: Detection of
tumorigenesis in rat bladders with optical coherence
tomography. Med Phys, 28: 2432, 2001

CT urography permits evaluation of both the renal


parenchyma and the urothelium and hence is the
imaging modality of choice in patients scheduled
to undergo TURBT. Critical components of CT
urography include the use of multidetector CT, thin
slice imaging, administration of contrast medium,
and imaging in the excretory phase.[2] CT urography
also allows evaluation of other pathologic conditions
within the genitourinary system, especially
urinary calculi, which may be a benign cause of
hematuria. Multiplanar reformatted images and
3-dimensional volume rendering are also useful
adjuncts to CT urography, yielding 3-dimensional
images that clinicians are familiar with and allowing
accurate evaluation of the upper tracts.[3] Magnetic
resonance urography provides information similar to
that provided by CT urography; however, magnetic
resonance urography is less accurate for calculi and
takes significantly longer to perform.[4] (Level 2B)

54. Hermes, B., Spoler, F., Naami, A. et al.: Visualization of


the basement membrane zone of the bladder by optical
coherence tomography: feasibility of noninvasive evaluation
of tumor invasion. Urology, 72: 677, 2008
55. Manyak, M. J., Gladkova, N. D., Makari, J. H. et al.:
Evaluation of superficial bladder transitional-cell carcinoma
by optical coherence tomography. J Endourol, 19: 570,
2005
56. Goh, A. C., Tresser, N. J., Shen, S. S. et al.: Optical
coherence tomography as an adjunct to white light
cystoscopy for intravesical real-time imaging and staging of
bladder cancer. Urology, 72: 133, 2008
57. Karl, A., Stepp, H., Willmann, E. et al.: Optical coherence
tomography for bladder cancer -- ready as a surrogate for
optical biopsy? Results of a prospective mono-centre study.
Eur J Med Res, 15: 131, 2010
58. Sengottayan, V. K., Vasudeva, P., Dalela, D.: Intravesical
real-time imaging and staging of bladder cancer: Use
of optical coherence tomography. Indian J Urol, 24: 592,
2008
59. Schmidbauer, J., Remzi, M., Klatte, T. et al.: Fluorescence
cystoscopy with high-resolution optical coherence
tomography imaging as an adjunct reduces false-positive
findings in the diagnosis of urothelial carcinoma of the
bladder. Eur Urol, 56: 914, 2009
60. Ren, H., Yuan, Z., Waltzer, W. et al.: Enhancing detection
of bladder carcinoma in situ by 3-dimensional optical
coherence tomography. J Urol, 184: 1499, 2010

In imaging of the bladder, ultrasonography and CT


urography have similar specificities in the diagnosis
of bladder cancer, namely, 94.7% and 96.5%,
respectively. However, CT urography has higher
sensitivity (89.7% versus 69%).[5] In addition, CT
urography demonstrates the size of lesions within

42

the bladder, extravesical spread, and pelvic lymph


node status whereas ultrasonography may not.[4]
Magnetic resonance urography is similarly effective.
[4] (level 2B)

hypotension during regional anesthesia. Diuretics


may cause hypokalemia, which can alter the effect
of muscle relaxants. (level 4)
Diabetic patients need to have their medications
reviewed and need to be given instructions regarding
the use of oral hypoglycemics and insulin while
fasting prior to TURBT.

2. Laboratory Investigations
Only basic laboratory investigations are required prior
to TURBT. Urine should be cultured, and infections
should be treated. In patients with infections, it is
unclear if a subsequent negative culture is required
before TURBT. A complete blood cell count and
measurement of electrolytes and creatinine are
also required as anemia or elevated creatinine
or potassium levels may have implications for
anesthesia. A coagulation profile should be obtained
if indicated. (Level 4)

Finally, patients need to be evaluated to ensure


that they are able to tolerate the surgical position.
Particular care is needed in placing patients with hip
and knee replacements in the lithotomy position.
Previous spinal surgery may make it difficult to use
spinal or epidural anesthesia.

II. Technique

3. Medical Clearance and Anesthesia


Review

1. Anesthesia

Bladder cancer is usually a disease of the elderly,


and hence preoperative medical clearance or
anesthesia review is beneficial and may lead to a
lower rate of cancellation at the time of surgery.[6]
Patients should be evaluated at a preanesthesia
clinic for any issues pertaining to the airway and
respiratory and circulatory systems, for medications,
and for anemia and diabetes. Referral for further
cardiac or respiratory evaluation can then be initiated
as needed.

For TURBT, the goal of anesthesia is to enable


safe resection of the bladder tumor with appropriate
analgesia and relaxation. General or regional
anesthesia or a combination of the two can be used.
Regional anesthesia in the form of spinal or epidural
blockade is effective. It also provides the advantage
of a conscious patient, which means that any
intraperitoneal bladder perforation can be identified
by abdominal or shoulder-tip pain.
General anesthesia with or without paralysis can
also be used. The obturator reflex can be a technical
challenge during TURBT, and full relaxation is
recommended when tumors on the lateral bladder
walls are resected. Alternatively, transvesical
obturator nerve block can be performed using
lidocaine. (Level 1B)

The major components of the clinical evaluation


are the history, physical examination, and
electrocardiography, the goal of which is to identify
any coronary artery disease, previous myocardial
infarction, angina pectoris, or other serious cardiac
disorders like heart failure or disorders requiring
pacemakers.[7] It should be determined whether
the patient has chronic obstructive or restrictive
pulmonary disease, diabetes mellitus, peripheral
vascular disease, or another major comorbid
condition and whether the patient has previously
undergone major vascular, thoracic, or abdominal
surgery. Renal dysfunction is common in elderly
patients, and creatinine level, potassium level,
and estimated glomerular filtration rate are useful
in calculating the dose of antibiotics as well as
anesthetic agents, particularly muscle relaxants.
(Level 4)

2. Antibiotic Prophylaxis
The goal of antibiotic prophylaxis is to prevent
infectious complications from TURBT. As stated
earlier, a preoperative urine culture should be
performed, and any infection should be treated
preoperatively. However, antibiotic prophylaxis is still
recommended at the time of surgery. Oral agents
with good bioavailability can be given several hours
prior to TURBT, but common practice is to administer
an intravenous antibiotic at the time of anesthesia to
ensure that peak plasma concentrations have been
achieved at the time of instrumentation. (Level 4)

A thorough medication history is critical. Antiplatelet


agents may be discontinued 7-10 days prior to
surgery. Warfarin should be discontinued and the
international normalized ratio (INR) should be
checked on the day of surgery. While the INR is
subtherapeutic, low-molecular-weight heparin may
be required. Spinal or epidural anesthesia should
not be performed if the INR is greater than 1.5.[8]
(level 4) Beta-blockers should not be discontinued
before TURBT; however, alpha blockers may need
to be documented as they may result in marked

There is a paucity of data on the role of antibiotics


in TURBT[11]; however, without prophylaxis, the
risk of subsequent urinary tract infection may be
as high as 39%.[12] Two randomized trials have
been performed comparing antibiotic prophylaxis to
placebo following TURBT. MacDermott et al found
a significantly reduced rate of bacteriuria in patients
who received antibiotics compared to patients who
did not receive antimicrobials.[13] Most of the data

43

on antibiotic use in genitourinary surgery is related


to transurethral resection of the prostate (TURP), in
which the benefits of antibiotic prophylaxis are clear.
Systematic reviews by Berry and Barratt and by
Qiang et al demonstrated that prophylactic antibiotics
reduced the incidence of bacteriuria and sepsis
following TURP. Hence, it is recommended that prior
to TURBT, urine should be sterilized if possible, and
prophylactic antibiotics should be administered.
There is no evidence to support prolonged use of
antibiotics after TURBT in the absence of specific
indications like persistent bacteriuria from calculi or
anatomic abnormalities.

The cystoscope is inserted, and the entire urethra is


examined. It is possible to obtain a urine sample by
barbotage immediately after the scope is inserted into
the bladder using gentle irrigation with normal saline
through a catheter, the cystoscope sheath, or the
resectoscope sheath. However, it is recommended
that urine be collected after complete inspection of
the bladder because during barbotage the bladder
wall can be drawn against the sheath, causing
urothelial trauma that may mimic the appearance of
CIS.
The bladder should be examined with both a 30degree and a 70-degree lens. Visualization of
the anterior bladder neck may be enhanced by
using a 120-degree scope or by using retroflexion
with a flexible cystoscope. The dome and anterior
bladder are examined by applying gentle suprapubic
pressure to move the bladder mucosa down and
closer to the lens. It is critical to maintain optimal
bladder distention during cystoscopy and TURBT.
Underdistention prevents complete examination of
the mucosal surface and can result in inadvertent
bladder perforation during TURBT. Overdistention,
on the other hand, can result in thinning of the
bladder mucosa, increasing the likelihood of bladder
perforation and inducing mucosal hemorrhage.
Cystoscopic findings recommended to be reported
are listed in Table 2.

3. Indications for Hospital Admission


Following TURBT, the majority of patients can be
discharged home with a Foley catheter, which can
be removed 1-2 days after the procedure. Indications
for admission to the hospital with continuous bladder
irrigation are very large resections, significant
hematuria, and bladder perforation.

4. Surgical Technique
TURBT in conjunction with cystoscopy and
examination under anesthesia serves both a
diagnostic and a therapeutic purpose. The diagnostic
purpose of TURBT is to obtain a specimen sufficient
to permit proper estimation of the natural history of
the tumor, which is currently based largely on stage,
grade, histological subtype, and the presence of
lymphovascular invasion. The therapeutic purpose
of TURBT is to remove all tumor and provide the best
milieu for intravesical therapy to be most effective.

The smallest resectoscope necessary should be


introduced in an atraumatic fashion. Continuous
irrigation is usually recommended to improve
visualization and prevention overdistention of the
bladder. Delicate movements of the sheath should
be used, along with delicate movement of the loop
itself to adjust the depth of resection.

The surgical technique of TURBT is largely based on


surgeon experience. The description herein expands
on the review provided by Committee 1 of the First
International Consultation on Bladder Tumors [17].

There are 2 basic approaches to performing TURBT:


staged resection and en bloc resection. Both are
described below.

The appropriate anesthesia to use during TURBT is


individualized according to the patient and to surgeon
and anesthesiologist preference. General anesthesia
with endotracheal intubation or a laryngeal mask
airway, spinal anesthesia, or epidural anesthesia
may be appropriate. Resection of tumors along the
lateral wall may elicit an obturator nerve reflex[17].
For such procedures, neuromuscular blockade is
recommended; the use of an obturator nerve block
is optional and is unnecessary in the majority of
patients (Level 4).

a) Staged Resection
Staged TURBT is done in several phases (Level
3). First, the most superficial portion of the tumor
protruding into the bladder is resected, and the
surgeon starts medially and works laterally if
possible. Next, the next layer of tissue is resected in
a similar fashion. Layers of tissue are resected in this
manner until the base of the tumor is reached. Finally,
the base of the tumor is resected. For resection of
the base of the tumor, some authors recommend
starting just adjacent to the bladder tumor to include
normal adjacent mucosa to help guide the depth of
the subsequent resection, progressing towards the
tumor and towards the adjacent normal mucosa on
the other side. Other authors advocate removing
the bulk of the tumor base first and then performing
a third phase in which the tissue surrounding the
tumor base is removed. The resected tissue may be
sent together for collective analysis, or tissue from

The patient is positioned in a dorsal lithotomy


position with the buttocks at the edge of the table and
all pressure points padded. Special care should be
taken to pad the lateral proximal fibula to avoid foot
drop due to neurapraxia of the common peroneal
nerve. An examination under anesthesia prior to
tumor resection is optional. An examination under
anesthesia should always be performed, however,
after tumor resection (see the section Examination
Under Anesthesia below).

44

Table 2. Recommended documentation at the time of TURBT

Cystoscopy
Number of tumors
Description of gross tumor architecture
Size (cm)*
Tumor location(s)

Examples or comments
2
Flat papillary or sessile
3-4cm
1 Right lateral wall

Suspected depth of invasion


Adjacent mucosa (trabeculated, normal)
Presence of diverticula
Ureteral orifices

2- trigone
Appears Non-invasive
Normal adjacent mucosa
No diverticula seen
Normal location, clear efflux of urine

TURBT
Approximate Depth of resection

Superficial, into muscle, into perivesical fat depth

Completeness of resection
Ureteral orifice(s)
Complications

of resection
Complete resection of tumor
Resected or intact ureteral orifice
No complications

EUA (following resection)


Prostate
Bladder
Bladder mass (estimate 3 dimensions when present)

Smooth, symmetric, no nodules, 40grams


Freely mobile
3cm x 3cm x 4cm (mobile)

* Use loop (1cm) as reference

d) Examination under Anesthesia

each stage may be sent separately for a differential


analysis[17-19].

Examination under anesthesia should be performed


following resection of the bladder tumor. The objective
of the examination under anesthesia is to properly
stage the disease and to help guide appropriate
therapy. A properly performed examination under
anesthesia includes a bimanual examination
following drainage of the bladder and without a
catheter in place so that the bladder is completely
empty. One hand is placed in the suprapubic region,
and 1 or 2 fingers of the other hand are placed
into the anus (males) or vagina (females)[17]. The
hand on the abdomen displaces structures, and
the posterior finger or fingers are used to note the
anatomic features. The bladder, prostate, and
other pelvic structures are palpated between the
2 hands[17]. If clinical stage T3 disease (palpable
mass on examination under anesthesia) is present,
this should be recorded, along with an estimate of
the 3-dimensional size of the mass. If a mass or the
bladder is fixed, the location of the fixation should
be noted (Table 2). Involvement of adjacent organs
should be assessed. In addition, the prostate should
be examined.

b) En Bloc Resection
En bloc resection may be used for small tumors,
generally those less than 3 cm in greatest
dimension[20-22] (Level 3). Reported advantages of
en bloc resection include more accurate pathologic
assessment because of decreased cautery artifact,
avoidance of tumor fragmentation, and preservation
of the spatial orientation of the tumor relative to
the bladder wall. There have been no comparative
studies of en bloc TURBT versus staged TURBT.

c) Hemostasis
Following resection, it is important to achieve
hemostasis. Hemostasis may be achieved by
selective cautery of obviously bleeding vessels,
to avoid unnecessary bladder contracture due to
excessive cautery. An alternative approach is to
cauterize the entire base of the resection site to
prevent subsequent bleeding. Cautery near the
ureteral orifices should be avoided if possible to
avoid ureteral stricturing[23-25].

45

tomy, or radical cystectomy should be considered


(Level 4). Small, low-grade lesions can be managed
with careful resection or fulguration (Level 4). In addition, intravesical therapy should be considered,
especially if the completeness of excision is in doubt
(Level 4).

III. TURBT in Specific


Circumstances
1. Diffuse CIS
For patients with diffuse bladder lesions and/or
suspected diffuse CIS, excessive bladder biopsies
and cauterization are not recommended. TURBT
is not therapeutic for most cases of CIS, and there
is clinical evidence indicating that the appropriate
treatment of diffuse CIS is intravesical therapy[26]
(Level 1).

5. Muscle-invasive Tumors
When a muscle-invasive tumor is suspected, the
depth of resection largely depends on the clinical
context. In general, sufficient tissue should be
resected to reveal whether or not tumor is present in
the muscularis propria. Intraoperative frozen tissue
assessment may be used to ensure that sufficient
muscle is available for proper evaluation. In addition,
proper staging of muscle-invasive bladder cancer
requires an examination under anesthesia performed
following complete transurethral resection of
intravesical tumor[33]. Among patients treated with
neoadjuvant chemotherapy, patients with pathologic
T0 disease on radical cystectomy performed after
systemic chemotherapy have a significantly better
prognosis than patients with residual intravesical
disease[34]. However, it is unclear if the completeness
of TURBT influences the response to neoadjuvant
chemotherapy. In general, it is advised that as much
tumor be removed as is safe to decrease the local
tumor burden and possibly optimize the response to
chemotherapy prior to radical cystectomy (Level 5).
For patients who are deemed not candidates for or
who refuse a radical cystectomy, TURBT should be
as thorough as possible to provide the patient with
the best opportunity for cancer control. In such cases,
re-resection to include deep tissue, sometimes into
the perivesical fat, may be warranted[20] (Level 3).

2. Tumors at the Ureteral Orifice


Cautery should be used sparingly near the ureteral
orifice. When cutting current is used to remove
tumors at or near the ureteral orifice, ureteral
stenosis is uncommon (Level 3). A ureteral stent
may be used to prevent temporary obstruction due
to postoperative edema, but whether ureteral stents
prevent ureteral stenosis is unknown [17] (Level 3).
A functional study, such as IVU or a renal scan, is
generally recommended 3-6 weeks after resection
over the ureteral orifice[27] (Level 4). Resection of
the ureteral orifice may lead to vesicoureteral reflux,
which may increase the risk of upper tract recurrence
(Level 3). Therefore, more frequent surveillance may
be needed in patients with reflux (Level 3).

3. Tumors on the Anterior Surface


and Lateral Wall
The anterior surface of the bladder is one of the
most challenging locations for TURBT. Resection of
tumors on the anterior surface can be made easier
by applying counterpressure to the bladder in the
suprapubic region[17] (Level 4). Resection of tumors
on the lateral wall may stimulate the obturator
nerve and result in a sudden adduction of the leg
and subsequent bladder perforation. Maneuvers
to decrease the risk of eliciting an obturator nerve
reflex include use of neuromuscular blockade,
use of obturator block, avoiding overdistention of
the bladder, using intermittent cutting current, and
lowering the resection current (Level 4).

6. Repeat TURBT
Current management guidelines for bladder
cancer are based on the depth of invasion, and
therefore it is critical that the urologist obtain the
most accurate staging assessment. Several cohort
series indicate that the depth of tumor invasion is
often underestimated on an initial TURBT. Herr 101
reported on 150 patients with bladder cancer who
underwent a repeat TURBT; 28 (29%) of the 96
patients with non-muscle-invasive bladder tumors
were reclassified as having muscle-invasive disease
at repeat biopsy. In addition, of the 23 patients
with T1 lesions without muscle in the specimen on
primary resection, 49% had their disease reclassified
as T2 at the second TURBT. However, the patients
in that series represented a unique populationthey
were all referred to a tertiary care center, and the
completeness of the primary TURBT in that setting
may have been limited because of the expectation
of subsequent referral to the tertiary care center.
Nevertheless, others have reported similar findings.
Dutta et al reported understaging in 64% of patients
with T1 disease when muscle was not present in the
specimen versus 30% of patients with T1 disease

4. Tumors in Bladder Diverticula


Diverticula of the bladder lack the muscularis propria, and thus diagnosis and treatment of tumors
in diverticula is challenging. The lack of a muscular
component precludes proper assessment for muscle
invasion. On the other hand, deep biopsies of bladder tumors in diverticula should be avoided because
of the high risk of perforation due to a thin bladder
wall. No prospective comparative trials have been
published on different therapeutic approaches for diverticular tumors. However, general guidelines are
as follows: For large or high-grade tumors arising
from diverticula, a diverticulectomy, partial cystec-

46

when muscle was present (Level 3)102. Zurkirchen


et al reported that 37% of patients with T1 bladder
tumors at initial diagnosis had persistent tumor on
repeat resection[35]. Grimm et al reported finding
residual tumor in 22% of patients, including 53%
of those with bladder tumors initially diagnosed as
T1 (Level 3)[36]. Brauers et al reported that 24% of
patients with non-muscle-invasive bladder cancer
had their disease reclassified as T2 or CIS on repeat
TURBT (Level 3)[37]. A prospective cohort evaluation
by Schips et al revealed residual disease in at least
50% of patients at repeat TURBT, and patients with
multifocality and high-grade disease were more likely
to have persistent disease at repeat resection[38].
Kln et al [39] identified residual tumor at repeat
TURBT in 20 (43.5%) of 46 patients and noted that
patients who underwent a staged resection (see
the section Staged Resection above) had a lower
incidence of tumor at repeat TURBT. Recently, Shen
et al reported results of a large observational study
that included 125 patients with non-muscle-invasive
disease who underwent a repeat TURBT[40]. Of
the 125 patients,43 (34.4%) had residual tumor on
repeat TURBT, and the presence of muscle tissue
in the initial TURBT specimen was significantly
associated with a decreased likelihood of having
residual disease (Level 3)[40].

The percentage of patients with tumor on random


bladder biopsies ranges from 1.5-14.5%[42-45]. In a
European Organisation for Research and Treatment
of Cancer study, 393 patients with solitary Ta and
T1 tumors underwent a biopsy of normal-appearing
mucosa, and CIS was detected in 1.5%[44]. Taguchi
et al reported that CIS was detected in12 (14.5%) of
83 patients who underwent random bladder biopsies
[45]. Fujimato et al performed random bladder
biopsies in 100 consecutive patients with non muscle
invasive cancer, and in 8 of them, biopsy revealed
additional tumor[43]. May et al [46], in a large series,
reported finding urothelial tumor on random bladder
biopsies in 12.4% of patients, including 4% of those
with CIS (Level 3). Moreover, these authors reported
that random bladder biopsy findings influenced the
management in 7% of patients (Level 3)[46].
The risk of tumor implantation and subsequent
tumor recurrence due to random bladder biopsy is
controversial. Levi et al reported that bladder biopsy
using a cold-cup biopsy grasper increased the risk
of tumor recurrence in the presence of CIS[47]. Mufti
and Singh [42], on the other hand, did not find tumor
implantation to be a risk factor for subsequent tumor
development.
A consensus opinion is that routine random bladder
biopsies do not change the clinical treatment in most
cases and are generally not recommended (Level
5). Random bladder biopsies are recommended in
cases of discordance between urine cytology findings
and the gross appearance of the bladder mucosa
and in bladder mapping prior to consideration of a
partial cystectomy. Finally, although there is little
evidence to support recommendations regarding the
location and number of random bladder biopsies, it
is common practice to obtain 5 biopsy specimens,
of the lateral walls, trigone, posterior bladder, and
dome.

Recently, a prospective, randomized trial was


conducted to address the value of repeat TURBT
in patients undergoing intravesical therapy with
mitomycin C. Divrik et al randomly assigned patients
with newly diagnosed T1 bladder cancer to a second
TURBT with adjuvant intravesical mitomycin C (74
patients) or adjuvant mitomycin C following the
initial resection without repeat TURBT (68 patients).
Patients with CIS, incomplete resection, or muscleinvasive disease were excluded from the study. At
a mean follow-up time of 31.5 months, the authors
observed a significantly higher recurrence rate in the
group that did not undergo a second TURBT (63.2%)
than in the group that did undergo a second TURBT
(25.7%). In addition, the progression rates were
higher in the group that did not undergo a second
TURBT (11.8% vs. 4.05%; p=0.097). In summary,
this trial provides evidence that a second TURBT
decreases recurrence rates for patients with T1
bladder cancer (Level 2)[41].

IV. Sample Transport and


Pathologic Processing:
Or How best to help your
pathologist give you the most
useful pathology report

7. Random Bladder Biopsies During


Routine TURBT

The pathologic evaluation of bladder tumor specimens


acquired by TURBT differs markedly from the usual
pathologic evaluation of tumor specimens. This
is due in large part to the nature of the procedure,
which cuts the bladder tumor into multiple fragments
with electrocautery. The pathologist receives a
specimen without any anatomic orientation, has to
process multiple fragments of the specimen, and has
to deal with potentially extensive cautery artifacts.
The pathologists interpretation of this specimen
will guide decisions about subsequent treatment
repeat TURBT, observation, intravesical therapy, or

Random bladder biopsies are performed to identify


the presence of CIS or dysplasia, which may not
be discernible by gross inspection but could alter
management recommendations. Unlike site-selected
biopsies, random bladder biopsies are performed in
completely normal appearing urothelium. The role
of random bladder biopsies during routine TURBT
has not been resolvedthere is a lack of convincing
evidence either for or against routine random bladder
biopsies.

47

radical cystectomy or radiotherapy. The most critical


information gained from pathologic interpretation
of TURBT specimens is information about whether
there is invasion of the lamina propria or muscularis
propria. Detection of lamina propria invasion usually
leads to re-resection and intravesical therapy, if not
immediate cystectomy, and detection of muscularis
propria invasion leads to radical cystectomy,
sometimes preceded by neoadjuvant chemotherapy.
It is imperative that TURBT specimens be handled
and processed optimally to help the pathologist
make the most accurate assessment.

tumor should be sent as a separate specimen to allow


optimal assessment of depth of invasion. Separate
evaluation of each tumor should facilitate detection
of lamina propria invasion in any 1 tumor, reveal
whether the surgeon should perform repeat resection
of specific tumor beds because of inadequate depth
of resection on initial TURBT, and reveal whether the
presence of lamina propria invasion requires second
resection (level 4).
When repeat TURBT is required for more certain
staging of high-grade and/or T1 bladder tumors, the
surgeon should resect the resection bed 4-6 weeks
after initial TURBT. Specimens from repeat TURBT
are generally flat and are sent to the pathologist as
a single specimen per tumor site, without a separate
sample from the tumor base. Interpretation of
specimens from repeat TURBT can be complicated
by inflammatory changes due to the preceding
resection.

1. Tissue Acquisition
The specimen delivered to the pathologist by the
urologist depends on the type of resection. Bladder
biopsies are often performed to evaluate suspiciousappearing areas of bladder mucosa or performed
randomly to rule out CIS. Bladder biopsies are
typically performed with cold-cup biopsy forceps
and include mucosal tissue with a varying depth of
subepithelial tissue, sometimes including muscularis
propria. The specimens are small and friable. They
must be minimally handled to prevent distortion[4850]. Specimens are generally placed immediately
on nonadherent gauze (e.g., Telfa, Kendall) and
transferred directly into formalin (level 4). Orientation
of cold-cup biopsy specimens is often straightforward
because the lesions are generally small and this
facilitates the identification of early invasion[51].
Cautery artifacts are not an important issue in coldcup biopsy specimens, but denudation of mucosa
is a significant risk. Denudation is seen commonly
with CIS, and hence a denuded biopsy specimen is
concerning but nondiagnostic.

Laser TURBT is a novel procedure that has not found


widespread application. In principle, use of laser
energy for resection does not change the handling
and processing of tissue specimens[54]. The
principal limitation of laser TURBT is the potential
for tissue alteration and destruction that prevents
adequate pathologic investigation[55].
En bloc TURBT is a novel approach that is designed
to allow better pathologic evaluation of tumors. Like
laser TURBT, however, en bloc TURBT is not widely
practiced. With en bloc TURBT, the surgeon uses a
flat loop electrode, laser, or knife to resect tumors up
to 3 cm in diameter as a single specimen. This gives
the pathologist more accurate anatomic orientation
and potentially enables better assessment of depth
of invasion.

TURBT of a unifocal tumor is the most common


clinical scenario. If the tumor is small, removal with
cold cup biopsy is often preferable to resection with
a cautery loop to minimize the cautery artifact. If the
tumor is larger, the final specimen will be a collection
of tissue chips, each measuring approximately 6
mm in diameter. These will have varying length and
depth and will be derived from different portions
of the tumor (superficial, middle, and base with
interface to bladder wall). Larger tumors should be
sent as 2 separate samples: 1 from the bulk of the
tumor and 1 from the tumor base[50] (level 4). This
enables the pathologist to make the most accurate
assessment possible of whether there is invasion
into the lamina propria or muscularis propria and
decreases the likelihood that such invasion could be
missed because of sampling errors.

2. Urine Cytology
Urine samples sent for cytologic evaluation also
require special handling and processing [58].
Cytologic evaluation of a urine specimen involves
a certain degree of subjectivity. Interpretation errors
will occur if specimens are collected, handled, and
processed incorrectly[59]. Optimal diagnosis can
only be obtained by following best practices.
Urine for cytologic analysis can be voided, obtained
during catheterization, or taken at the time of
instrumentation. The quantity of cellular material and
the quality of the specimen are best for instrumented
specimens, intermediate for specimens obtained
during catheterization, and worst for voided
specimens[60-62].

TURBT of multifocal tumor requires some additional


considerations. To map the bladder adequately,
each tumor should be sent as a separate specimen.
This practice requires that the surgeon document
pictorially the location of the tumors in the bladder in
the written operative note. The base of each larger

Instrumented urine samples can be obtained by


barbotage, which involves irrigating in and out
to release more cells and thereby enhance the
diagnostic quality of the specimen. It is important
to use normal saline, and not water or glycine for
barbotage, in order to preserve cellular morphology.

48

It is equally critical to obtain specimens prior to


instillation of contrast dye or any other agents.
Ideally, the cystoscope is inserted into the bladder,
and barbotage is performed with saline through the
cystoscope (level 4). As is the case for cytology in
general, these maneuvers are best for detecting
high-grade carcinoma, and both the sensitivity
and specificity are low for low-grade lesions. The
pathologist must be wary of the risk of overdiagnosing
low-grade urothelial neoplasms, especially in the
case of instrumented urine specimens. These
samples often contain benign cellular clumps that
can resemble urothelial carcinoma but usually have
more defined borders and a more densely staining
cytoplasmic collar[63-65].

filtration (Magna MCE nitrocellulose filter, BioBlock


Scientific, Illkirch, France); and Monoprep (Monogen,
Lincolnshire, IL )[66]. These methods deliver a more
uniform, single-layer preparation with more cells,
better preserved morphology, and less background
debris[66-69] (level 3). Urine samples are placed
immediately into proprietary fixatives, which provide
immediate and superior preservation.
Recently, immunocytochemistry for vimentin (but
not high-molecular-weight cytokeratin) has been
reported to better differentiate reactive renal tubular
cells from low-grade urothelial carcinoma cells in
voided urine. Thus this technique eliminates one
specific confounding factor in the evaluation of urine
cytology.

A urine specimen for cytology should not be taken


from the first voided urine of the day. The premise
for this recommendation is that the first voided urine
of the day has likely dwelled in the bladder for a
relatively long time and will have poorly preserved
cellular architecture. A common recommendation is
to have the patient empty the bladder in the morning
and collect the next specimen after 2-3 hours of
adequate hydration (level 4).

3. Clinical Information
Adequate clinical information allows the pathologist
to initiate any specific tests required to guide
clinical management and to avoid unnecessary
investigations[72]. Clinical information may accelerate
the diagnostic process. For example, the pathologist
may request additional immunohistochemical
analyses based on clinical data. Conversely, the
written pathology report is the principal line of
communication from the pathologist back to the
treating clinician, and this report can be tailored to
the clinical need, thereby enhancing communication.
The essential clinical information is highlighted in
Table 3.

The ideal urine specimen volume is poorly defined,


but 25 mL can be considered a minimal volume. A
larger volume may come at the expense of longer
dwell time in the bladder, which can be detrimental.
It is not advisable to collect urine for a defined period
of time, such as 12 or 24 hours, because this leads
to poor specimen quality due to specimen storage
(level 4).

Table 3. Recommended essential clinical information


for pathologic specimen submission

There are no well-defined timelines for processing


urine for cytologic examination, but it is clear that the
degree of cellular deterioration increases with time.
The diagnostic accuracy therefore increases with
the expediency of specimen processing. The time
from specimen collection to processing should not
exceed 8 hours [58] (level 4). Urine is best examined
as a fresh specimen for cytology but, if expeditious
examination of a fresh specimen is not feasible, the
specimen can be preserved with various solutions,
including 70% ethyl alcohol (2:1 dilution by volume),
polyethylene glycol, or the solution in kits used for
liquid-based Papanicolaou testing (level 3).

Tumor characteristics:
Flat
Papillary
Nodular
Ulcerated
Location in the bladder
Clinical history:
P
 rior bladder tumors (grade/stage/histology/
location)
P
 rior intravesical chemotherapy/

Techniques for processing urine specimens for


cytologic analysis in the laboratory vary. Conventional
methods include using filters, cytocentrifugation
(cytospin), or concentrated smears. Newer techniques
include various commercial kits using liquidbased methods: ThinPrep (Hologic, Inc., Bedford,
Massachusetts); SurePath (Becton Dickenson
Company, Franklin Lakes, New Jersey); Millipore
filter preparations (Millipore Corporation, Billerica,
Massachusetts); AutoCyte PREP (BD Diagnostics
Tripath, Burlington, NC); Shandon Cytospin (Thermo
Electron, Waltham, MA); nitrocellulose membrane

immunotherapy
Prior systemic chemotherapy
Prior radiation therapy to the bladder/pelvis
Prior instrumentation
H
 istory of stones, infection, other urologic
conditions
P
 rior non-bladder malignancy (type and
location)

49

Documentation of the site of tissue acquisition is


essential for pathologic identification and evaluation.
In the context of TURBT, the primary information
required is whether the specimen was taken from
the bladder or the urethra. Within the bladder,
documentation of the location from which tissue was
obtained may be important for mapping of tumors
for example, should there be a need for re-resection
of 1 or more sites. Location on the dome may be
relevant for diagnosing urachal carcinoma.

may lead to understaging because of missed focal


invasion into the lamina propria or muscularis propria.
The presence of muscularis propria in the specimen
is considered a measure of quality control[79], and
it is therefore important to obtain a biopsy sample
that includes muscularis propria. If not all chips are
sampled, the pathology report should indicate what
proportion of the chips was sampled[49] (level 4).
The College of American Pathologists has developed
a Protocol for the Examination of Specimens from
Patients with Carcinoma of the Urinary Bladder
(http://www.cap.org/apps/docs/committees/cancer/
cancer_protocols/2009/UrinaryBladder_09protocol.
pdf)[80]. The panel drawing up this protocol was
made up primarily of United States contributors but
also included some members from other countries.
This document focuses mostly on standard
reporting for TURBT specimens but also includes
recommendations for tissue processing.

The pathologist should also be given information


about the patients relevant medical history. A prior
history of kidney stones or urinary tract infection may
affect the interpretation of bladder biopsies. A prior
history of BCG instillation may aid in the diagnosis
of granulomatous changes. Use of photodynamic
agents such as hexaminolevulinate at the time of
TURBT does not appear to affect the pathologic
specimen.
For cytopathology requisitions, it must be noted
whether a patient has an intact bladder or has
undergone prior urinary diversion. Inflammatory
changes associated with urinary diversion may lead
the cytopathologist to be concerned about atypia if
the prior urinary diversion is not mentioned. Previous
BCG therapy and previous stone disease are also
relevant as both can cause artifact.

The College of American Pathologists protocol calls


for examination of 1 microscopic section for every
1 cm of tumor diameter up to 10 sections. If this
evaluation reveals that tumor invades into muscle,
sampling can be considered adequate. If, however,
there is evidence of invasion into the lamina propria
but not muscle, further sampling is necessary to
exclude muscularis propria invasion. Similarly, if
there is no evidence of invasion on the initial analysis,
more tissue must be sampled. Depending on the
size of the tumor, it may be necessary to examine
the entire specimen (level 4).

4. Pathologic Processing
a) Tissue Fixation
In the operating room, all tissue must be placed
immediately in 10% neutral buffered formalin in a
10:1 ratio by volume of fixative to tissue to ensure
adequate fixation. Filling a specimen cup with
tumor chips without adequate fixative will lead to
poor fixation and increased tissue degradation.
Alternatively, 4% paraformaldehyde may be used
instead of neutral buffered formalin. In fact, 4%
paraformaldehyde may even be a superior fixative
for immunohistochemistry as it tends to yield reduced
background staining. However, it must be freshly
made just before use. The duration of fixation can be
12-24 hours for TURBT specimens since each tissue
fragment tends to be small (level 4).

The European Society of Uropathology and the


Uropathology Working Group have also described
standardized methods for handling and reporting
bladder tumor specimens. Their report stems from
a Uropathology Workshop held in Sesto Fiorentino,
Italy, in 2003 and represents a statement of expert
opinion. The standardized methods outlined in the
report incorporate recommendations made by the
College of American Pathologists and the Association
of Directors of Anatomic Pathology and are in line
with the recommendations for tissue sampling
described above. In addition, the European Society
of Uropathology and the Uropathology Working
Group emphasize that the number, size (aggregate
dimensions), and weight of each sample must be
recorded (level 4).

In the pathology laboratory, it is important to avoid


overfilling the specimen cassettes, as this is more
likely to lead to undersampling of some tissue
fragments (level 4).

Assessment of lamina propria invasion is one of the


most significant challenges in bladder pathology.
There is a high degree of interobserver variability in
the assessment of lamina propria invasion, and the
clinical relevance of such assessment is substantial,
as patients with T1 lesions are considered at high
risk for recurrence and progression when bladder
preservation is attempted. The status of the lamina
propria can determine the need for repeat resection
and intravesical BCG or, in patients who have had
prior treatments, the need for cystectomy.

b) Specimen Examination
Every sample sent from the operating room should
be processed individually and reported separately.
Expert pathologists recommend that all tissue
fragments from any TURBT sample be processed.
The degree to which these samples are sectioned
and analyzed is, however, not clearly established, and
practice in this area is not uniform. Undersampling

50

One of the most significant obstacles to accurate


assessment of invasion is poor orientation of the
specimen due to the piecemeal fashion in which it is
delivered to the pathologist and tangential sectioning.
There is a risk that papillary tumors can be sectioned
in such a way that isolated clusters of noninvasive
tumor cells are visualized with surrounded stroma,
giving the appearance of invasion. The pathologist
must deduce from smooth and regular contours
that this is an artifact of sectioning, while irregular
contours with haphazard arrangement favor true
invasion. If necessary, re-embedding and further
sectioning can be helpful[50].

criteria would demand specific immunohistochemical


staining with endothelial markers such as factor-VIIIrelated antigen, CD31, or CD34[48]. In some cases,
staining will not resolve the problem of differentiating
lymphatic from artifactual space entrapment by
tumor cells. Retraction artifact is also prominent in
the micropapillary variant of urothelial carcinoma
which has been well described by Kamat et al.
Cautery artifact can be one of the principal factors
limiting pathologic analysis of TURBT specimens.
Cautery can cause severe morphologic distortion
that prevents adequate interpretation. Lopez-Beltran
et al suggest 2 methods to help overcome cautery
artifact[51]: deeper sections can uncover regions
of tissue that are better preserved, and immunohistochemistry with anti-cytokeratin antibodies can delineate tumor tissue. If adequate interpretation is not
possible despite these methods, that fact needs to
be documented in the pathology report.

Accurate identification of the presence of muscularis


propria in the TURBT specimen is essential for
accurate staging and determination of the appropriate
subsequent clinical management [81]. As mentioned
above, the presence of muscularis propria in the
TURBT specimen is also evolving as an important
quality indicator of the TURBT itself[79].

5. Pathology Reporting

The pathologist must be careful to distinguish


muscularis mucosa from muscularis propria and
sometimes even myofibroblasts in desmoplastic
stromal reaction associated with the carcinoma.
Muscularis mucosa consists of thin muscle bundles
in the lamina propria that are classically described
as being arranged in a single interrupted, dispersed,
or continuous layer[82]. In practice, however,
the distinction may be less obvious, especially
if the muscularis mucosa is hyperplastic, and
immunohistochemical staining with anti-smoothmuscle-specific actin[53] or smoothelin can help in
selected cases (level 3). Smoothelin is a marker of
terminally differentiated smooth muscle cells and
has recently been shown to stain muscularis propria
with relative specificity. Smoothelin stains muscularis
mucosa either weakly or not at all and does not stain
myofibroblasts in desmoplastic stroma.

Careful harvesting, handling, and processing of


TURBT specimens is ultimately of little benefit if the
findings are not reported clearly[92-94]. The presence
of key findings such as focal invasion or concomitant
CIS can be missed by the treating physician if these
findings are not clearly documented in the pathology
report[92]. The goal of the pathologic investigation
and the final report is to provide the urologist or
other treating physician with the most accurate
information possible so that he or she can make
the most appropriate management decisions for the
patient from whom the specimen was obtained[95].
An important component of pathology reporting
is specific documentation of known risk factors for
disease recurrence and progression as well as
predictive factors for response to therapy.
The College of American Pathologists Protocol
for the Examination of Specimens from Patients
with Carcinoma of the Urinary Bladder includes a
component on the optimal reporting of pathologic
findings in bladder tumor specimens (http://www.
cap.org/apps/docs/committees/cancer/cancer_
protocols/2009/UrinaryBladder_09protocol.pdf).
This protocol was first published in 1996[96] and
was updated in 2003[80]. These standards were
subsequently adopted by the Commission on
Cancer of the American College of Surgeons as a
mandated checklist in its Cancer Program Standards
for Approved Cancer Programs (2004). The checklist
is reproduced with permission in Figure 8 (level 4).

After
routine
hematoxylin-eosin
staining,
immunohistochemical analysis of tissue samples
with uncertain histology may be necessary to confirm
or rule out the diagnosis of urothelial carcinoma. The
most common distinction that needs to be made
is the distinction between urothelial carcinoma
and high-grade prostate cancer [86]. Urothelial
carcinoma typically stains positive for uroplakin III,
thrombomodulin, cytokeratin 7, and cytokeratin
20. Negative immunostaining for prostate-specific
antigen and prostatic acid phosphatase is an
important indicator of urothelial carcinoma as
opposed to prostate cancer (level 3).

The Association of Directors of Anatomic Pathology


has generated a Final Anatomic Diagnosis Checklist
for Urinary Bladder Neoplasms, but these guidelines
are for cystectomy specimens only.

Lymphovascular invasion (LVI) is an established


biomarker used for risk stratification in cystectomy
specimens[88-90] as well as in TURBT samples
[91] and hence LVI is often used clinically in bladder
cancer care decision-making. It is therefore relevant
for the pathologist to inspect the sections for LVI and
comment correspondingly in the report (level 3). Strict

It is important that the pathologist state not only


what is present in, but also what is absent from, the
TURBT specimen. It is vital, for example, to state the

51

FIGURE 8 Checklist for pathologic reporng


.

Figure 8. Checklist for pathologic reporting

52

Figure 8. Checklist for pathologic reporting (ctd)

53

10

absence of LVI or the absence of muscularis propria if


the pathologist has examined the specimen carefully
and not identified these features. In patients with
urothelial malignancy, a report of denuded biopsy
(most probably related to CIS) is significantly different
from one stating that no tumor is seen [52]. Tumor
in fat should not be interpreted as perivesical fat
invasion, because the bladder wall can contain fat.

8. Cherian V. Anesthetic implications of urology surgery.


Indian journal of urology 2006;22:194-200.

Just as the pathologic requisition is an important


document for communicating the clinical scenario to
the pathologist, the pathology report is the principal
means by which the pathologist communicates back
to the urologist and to subsequent treating physicians.
In complex cases, face-to-face communication
between the pathologist and the urologist can be
essential for accurately conveying the findings. Such
communication can involve the surgeons delivering
the specimen to the pathologist with specific details
on orientation of the specimen or the pathologists
demonstrating the pathology under the microscope
to the urologist. A more intimate working relationship
is likely to enhance patient care[95]. Nuances are
conveyed better on the telephone or in person than
in the pathologic report.

11. Bootsma AM, Laguna Pes MP, Geerlings SE, Goossens A.


Antibiotic prophylaxis in urologic procedures: a systematic
review. Eur Urol 2008;54:1270-86.

9. Khorrami MH, Javid A, Saryazdi H, Javid M. Transvesical


blockade of the obturator nerve to prevent adductor
contraction in transurethral bladder surgery. J Endourol
2010;24:1651-4.
10. Tatlisen A, Sofikerim M. Obturator nerve block and
transurethral surgery for bladder cancer. Minerva Urol
Nefrol 2007;59:137-41.

12. Trinchieri A, Mangiarotti B, Lizzano R. [Use of levofloxacin


in the antibiotic prophylaxis for diagnostic procedures in
urology]. Arch Ital Urol Androl 2002;74:33-9.
13. MacDermott JP, Ewing RE, Somerville JF, Gray BK.
Cephradine prophylaxis in transurethral procedures for
carcinoma of the bladder. Br J Urol 1988;62:136-9.
14. Delavierre D, Huiban B, Fournier G, Le Gall G, Tande
D, Mangin P. [The value of antibiotic prophylaxis in
transurethral resection of bladder tumors. Apropos of 61
cases]. Prog Urol 1993;3:577-82.
15. Berry A, Barratt A. Prophylactic antibiotic use in
transurethral prostatic resection: a meta-analysis. J Urol
2002;167:571-7.
16. Qiang W, Jianchen W, MacDonald R, Monga M, Wilt TJ.
Antibiotic prophylaxis for transurethral prostatic resection in
men with preoperative urine containing less than 100,000
bacteria per ml: a systematic review. J Urol 2005;173:117581.

When there is diagnostic uncertainty, the pathologist


should make this clear but still provide a likely
diagnosis. The opinion of additional pathologists or
a reference pathologist may be helpful. Finally, the
pathology report should not include management
recommendations because many possible clinical
parameters could make these recommendations
inappropriate[95].

17. Kirkali Z, Chan T, Manoharan M, et al. Bladder Cancer:


Epidemiology, Staging and Grading, and Diagnosis. In:
Soloway MS, Carmack A, Khoury S, eds. 1st International
Consultation on Bladder Tumors. Paris: Health Publication
Ltd; 2004:13-64.
18. Milner WA. Transurethral biopsy; an accurate method of
determining the true malignancy of bladder carcinoma. J
Urol 1949;61:917-23; Disc , 29-9.

References

19. Soloway MS, Patel J. Surgical techniques for endoscopic


resection of bladder cancer. Urol Clin North Am
1992;19:467-71.

1. OConnor O, Fitzgerald E, Maher M. Imaging of hematuria.


American Journal of Roentgenology 2010;195:w263-7.

20. Solsona E, Iborra I, Ricos JV, et al. Recurrence of superficial


bladder tumors in prostatic urethra. Eur Urol 1991;19:8992.

2. Van Der Molen AJ, Cowan NC, Mueller-Lisse UG, NolteErnsting CC, Takahashi S, Cohan RH. CT urography:
definition, indications and techniques. A guideline for
clinical practice. Eur Radiol 2008;18:4-17.
3. Anderson EM, Murphy R, Rennie AT, Cowan NC.
Multidetector computed tomography urography (MDCTU)
for diagnosing urothelial malignancy. Clin Radiol
2007;62:324-32.
4. Kim JK, Park SY, Ahn HJ, Kim CS, Cho KS. Bladder cancer:
analysis of multi-detector row helical CT enhancement
pattern and accuracy in tumor detection and perivesical
staging. Radiology 2004;231:725-31.
5. Knox MK, Cowan NC, Rivers-Bowerman MD, Turney
BW. Evaluation of multidetector computed tomography
urography and ultrasonography for diagnosing bladder
cancer. Clin Radiol 2008;63:1317-25.
6. Cousins J, Howard J, Borra P. Principles of anaesthesia in
urological surgery. BJU Int 2005;96:223-9.
7. Eagle KA, Berger PB, Calkins H, et al. ACC/AHA guideline
update for perioperative cardiovascular evaluation for
noncardiac surgery--executive summary: a report of the
American College of Cardiology/American Heart Association
Task Force on Practice Guidelines (Committee to Update
the 1996 Guidelines on Perioperative Cardiovascular
Evaluation for Noncardiac Surgery). J Am Coll Cardiol
2002;39:542-53.

21. Sonpavde G, Goldman BH, Speights VO, et al. Quality of


pathologic response and surgery correlate with survival
for patients with completely resected bladder cancer after
neoadjuvant chemotherapy. Cancer 2009;115:4104-9.
22. Ukai R, Kawashita E, Ikeda H. A new technique for
transurethral resection of superficial bladder tumor in 1
piece. J Urol 2000;163:878-9.
23. Davis JP. Ureteral injury by trans-urethral electroresection
and coagulation. J Urol 1952;68:168-77.
24. Gottfries A, Nilsson S, Sundin T, Viklund LG. Late effects
of transurethral resection of bladder tumours at the ureteric
orifice. Scand J Urol Nephrol 1975;9:32-5.
25. Posta B, Streit B, Schmauzer J. Transurethral resection
of the carcinomatous ureteral orifice (Analysis of 27
operations). Int Urol Nephrol 1980;12:23-35.
26. Sylvester RJ, van der Meijden AP, Witjes JA, Kurth K.
Bacillus calmette-guerin versus chemotherapy for the
intravesical treatment of patients with carcinoma in situ
of the bladder: a meta-analysis of the published results
of randomized clinical trials. J Urol 2005;174:86-91;
discussion -2.

54

27. Amar AD. Ureterovesical Junction Obstruction Following


Transurethral Resection. Br J Urol 1965;37:307-13.

45. Taguchi I, Gohji K, Hara I, et al. Clinical evaluation of


random biopsy of urinary bladder in patients with superficial
bladder cancer. Int J Urol 1998;5:30-4.

28. Rees RW. The effect of transurethral resection of the


intravesical ureter during the removal of bladder tumours.
Br J Urol 1969;41:2-5.

46. May F, Treiber U, Hartung R, Schwaibold H. Significance of


random bladder biopsies in superficial bladder cancer. Eur
Urol 2003;44:47-50.

29. Hobika JH, Clarke BG. Use of neuromuscular blocking


drugs to counteract thigh-adductor spasm induced by
electrical shocks of obturator nerve during transurethral
resection of bladder tumors. J Urol 1961;85:295-6.

47. Levi AW, Potter SR, Schoenberg MP, Epstein JI. Clinical
significance of denuded urothelium in bladder biopsy. J
Urol 2001;166:457-60.

30. Prentiss RJ, Harvey GW, Bethard WF, Boatwright DE,


Pennington RD. Massive Adductor Muscle Contraction in
Transurethral Surgery: Cause and Prevention; Development
of Electrical Circuitry. J Urol 1965;93:263-71.

48. Lopez-Beltran A, Bassi PF, Pavone-Macaluso M, Montironi


R. Handling and pathology reporting of specimens with
carcinoma of the urinary bladder, ureter, and renal pelvis.
A joint proposal of the European Society of Uropathology
and the Uropathology Working Group. Virchows Arch
2004;445:103-10.

31. Golijanin D, Yossepowitch O, Beck SD, Sogani P, Dalbagni


G. Carcinoma in a bladder diverticulum: presentation and
treatment outcome. J Urol 2003;170:1761-4.

49. Lopez-Beltran A, Bollito E, Luque RJ, Montironi R. A practical


approach to bladder sampling and diagnostic reporting of
pathological findings. Pathologica 2001;93:688-92.

32. Melekos MD, Asbach HW, Barbalias GA. Vesical diverticula:


etiology, diagnosis, tumorigenesis, and treatment. Analysis
of 74 cases. Urology 1987;30:453-7.

50. Cheng L, Montironi R, Davidson DD, Lopez-Beltran A.


Staging and reporting of urothelial carcinoma of the urinary
bladder. Mod Pathol 2009;22 Suppl 2:S70-95.

33. Greene F, Page D, Flemming I. American Joint Committee


on Cancer Staging Manual. New York: Springer-Verlag;
2002.

51. Lopez-Beltran A, Cheng L. Stage pT1 bladder carcinoma:


diagnostic criteria, pitfalls and prognostic significance.
Pathology 2003;35:484-91.

34. Grossman HB, Natale RB, Tangen CM, et al. Neoadjuvant


chemotherapy plus cystectomy compared with cystectomy
alone for locally advanced bladder cancer. N Engl J Med
2003;349:859-66.

52. Lopez-Beltran A, Bassi P, Pavone-Macaluso M, Montironi


R. Handling and pathology reporting of specimens with
carcinoma of the urinary bladder, ureter, and renal pelvis.
Eur Urol 2004;45:257-66.

35. Zurkirchen MA, Sulser T, Gaspert A, Hauri D. Second


transurethral resection of superficial transitional cell
carcinoma of the bladder: a must even for experienced
urologists. Urol Int 2004;72:99-102.

53. Jimenez RE, Keane TE, Hardy HT, Amin MB. pT1 urothelial
carcinoma of the bladder: criteria for diagnosis, pitfalls, and
clinical implications. Adv Anat Pathol 2000;7:13-25.

36. Grimm MO, Steinhoff C, Simon X, Spiegelhalder P,


Ackermann R, Vogeli TA. Effect of routine repeat
transurethral resection for superficial bladder cancer: a
long-term observational study. J Urol 2003;170:433-7.

54. Smith N, Thomas K. Lasers for Bladder Tumors. In:


Dasgupta P, Fitzpatrick J, Kirby R, Gill IS, eds. New
Technologies in Urology. 1 ed. London: Springer Verlag;
2010:71-8.

37. Brauers A, Buettner R, Jakse G. Second resection and


prognosis of primary high risk superficial bladder cancer: is
cystectomy often too early? J Urol 2001;165:808-10.

55. Lopez-Beltran A, Luque RJ, Mazzucchelli R, Scarpelli


M, Montironi R. Changes produced in the urothelium by
traditional and newer therapeutic procedures for bladder
cancer. J Clin Pathol 2002;55:641-7.

38. Schips L, Augustin H, Zigeuner RE, et al. Is repeated


transurethral resection justified in patients with newly
diagnosed
superficial
bladder
cancer?
Urology
2002;59:220-3.

56. Lodde M, Lusuardi L, Palermo S, et al. En bloc transurethral


resection of bladder tumors: use and limits. Urology
2003;62:1089-91.

39. Kln R, Loy V, Huland H. Residual tumor discovered in


routine second transurethral resection in patients with
stage T1 transitional cell carcinoma of the bladder. J Urol
1991;146:316-8.

57. Saito S. Transurethral en bloc resection of bladder tumors.


J Urol 2001;166:2148-50.
58. Rosenthal DL, Raab S, eds. Cytologic Detection of
Urothelial Lesions. New York, NY: Springer-Verlag; 2006.

40. Shen YJ, Ye DW, Yao XD, et al. [Repeat transurethral


resection for non-muscle invasive bladder cancer].
Zhonghua Wai Ke Za Zhi 2009;47:725-7.

59. Raab SS, Grzybicki DM, Vrbin CM, Geisinger KR. Urine
cytology discrepancies: frequency, causes, and outcomes.
Am J Clin Pathol 2007;127:946-53.

41. Divrik RT, Yildirim U, Zorlu F, Ozen H. The effect of repeat


transurethral resection on recurrence and progression
rates in patients with T1 tumors of the bladder who received
intravesical mitomycin: a prospective, randomized clinical
trial. J Urol 2006;175:1641-4.

60. Zein T, Wajsman Z, Englander LS, et al. Evaluation of


bladder washings and urine cytology in the diagnosis of
bladder cancer and its correlation with selected biopsies of
the bladder mucosa. J Urol 1984;132:670-1.

42. Mufti GR, Singh M. Value of random mucosal biopsies in


the management of superficial bladder cancer. Eur Urol
1992;22:288-93.

61. Matzkin H, Moinuddin SM, Soloway MS. Value of urine


cytology versus bladder washing in bladder cancer. Urology
1992;39:201-3.

43. Fujimoto N, Harada S, Terado M, Sato H, Matsumoto T.


Multiple biopsies of normal-looking urothelium in patients
with superficial bladder cancer: Are they necessary? Int J
Urol 2003;10:631-5.

62. Trott PA, Edwards L. Comparison of bladder washings and


urine cytology in the diagnosis of bladder cancer. J Urol
1973;110:664-6.
63. Kannan V, Bose S. Low grade transitional cell carcinoma
and instrument artifact. A challenge in urinary cytology. Acta
Cytol 1993;37:899-902.

44. van der Meijden A, Oosterlinck W, Brausi M, Kurth KH,


Sylvester R, de Balincourt C. Significance of bladder
biopsies in Ta,T1 bladder tumors: a report from the EORTC
Genito-Urinary Tract Cancer Cooperative Group. EORTCGU Group Superficial Bladder Committee. Eur Urol
1999;35:267-71.

64. Chu YC, Han JY, Han HS, Kim JM, Suh JK. Cytologic
evaluation of low grade transitional cell carcinoma and
instrument artifact in bladder washings. Acta Cytol
2002;46:341-8.

55

65. Sack MJ, Artymyshyn RL, Tomaszewski JE, Gupta PK.


Diagnostic value of bladder wash cytology, with special
reference to low grade urothelial neoplasms. Acta Cytol
1995;39:187-94.

83. Vakar-Lopez F, Shen SS, Zhang S, Tamboli P, Ayala AG,


Ro JY. Muscularis mucosae of the urinary bladder revisited
with emphasis on its hyperplastic patterns: a study of a
large series of cystectomy specimens. Ann Diagn Pathol
2007;11:395-401.

66. Laucirica R, Bentz JS, Souers RJ, et al. Do liquid-based


preparations of urinary cytology perform differently than
classically prepared cases? Observations from the College
of American Pathologists Interlaboratory Comparison
Program in Nongynecologic Cytology. Arch Pathol Lab Med
2010;134:19-22.

84. Paner GP, Brown JG, Lapetino S, et al. Diagnostic use


of antibody to smoothelin in the recognition of muscularis
propria in transurethral resection of urinary bladder tumor
(TURBT) specimens. Am J Surg Pathol 2010;34:792-9.
85. Council L, Hameed O. Differential expression of
immunohistochemical markers in bladder smooth muscle
and myofibroblasts, and the potential utility of desmin,
smoothelin, and vimentin in staging of bladder carcinoma.
Mod Pathol 2009;22:639-50.

67. Voss JS, Kipp BR, Krueger AK, et al. Changes in specimen
preparation method may impact urine cytologic evaluation.
Am J Clin Pathol 2008;130:428-33.
68. Nassar H, Ali-Fehmi R, Madan S. Use of ThinPrep
monolayer technique and cytospin preparation in urine
cytology: a comparative analysis. Diagn Cytopathol
2003;28:115-8.

86. Mhawech P, Uchida T, Pelte MF. Immunohistochemical


profile of high-grade urothelial bladder carcinoma and
prostate adenocarcinoma. Hum Pathol 2002;33:1136-40.
87. Mhawech P, Iselin C, Pelte MF. Value of
immunohistochemistry in staging T1 urothelial bladder
carcinoma. Eur Urol 2002;42:459-63.

69. Norimatsu Y, Kawanishi N, Shigematsu Y, Kawabe T,


Ohsaki H, Kobayashi TK. Use of liquid-based preparations
in urine cytology: An evaluation of Liqui-PREP and BD
SurePath. Diagn Cytopathol 2010;38:702-4.

88. Quek ML, Stein JP, Nichols PW, et al. Prognostic


significance of lymphovascular invasion of bladder cancer
treated with radical cystectomy. J Urol 2005;174:103-6.

70. Ohsaki H, Hirakawa E, Nakamura M, Norimatsu Y, Kiyomoto


H, Haba R. Expression of vimentin and high-molecularweight cytokeratin (clone 34ssE12) in differentiating reactive
renal tubular cells from low-grade urothelial carcinoma cells
in voided urine. Cytopathology 2010.

89. Shariat SF, Svatek RS, Tilki D, et al. International validation


of the prognostic value of lymphovascular invasion
in patients treated with radical cystectomy. BJU Int
2010;105:1402-12.

71. Ohsaki H, Hirakawa E, Kushida Y, et al. Can cytological


features differentiate reactive renal tubular cells from
low-grade urothelial carcinoma cells? Cytopathology
2010;21:326-33.

90. Lotan Y, Gupta A, Shariat SF, et al. Lymphovascular


invasion is independently associated with overall survival,
cause-specific survival, and local and distant recurrence in
patients with negative lymph nodes at radical cystectomy. J
Clin Oncol 2005;23:6533-9.

72. Burton JL, Stephenson TJ. Are clinicians failing to supply


adequate information when requesting a histopathological
investigation? J Clin Pathol 2001;54:806-8.

91. Kunju LP, You L, Zhang Y, Daignault S, Montie JE,


Lee CT. Lymphovascular invasion of urothelial cancer
in matched transurethral bladder tumor resection and
radical cystectomy specimens. J Urol 2008;180:1928-32;
discussion 32.

73. Anagnostopoulou I, Rammou-Kinia R, Likourinas M. Urine


cytology evaluation in cases of uretero-ileal cutaneous
diversion. Cytopathology 1995;6:268-72.
74. Wolinska WH, Melamed MR. Urinary conduit cytology.
Cancer 1973;32:1000-6.

92. Powsner SM, Costa J, Homer RJ. Clinicians are from Mars
and pathologists are from Venus. Arch Pathol Lab Med
2000;124:1040-6.

75. Betz SA, See WA, Cohen MB. Granulomatous inflammation


in bladder wash specimens after intravesical bacillus
Calmette-Guerin therapy for transitional cell carcinoma of
the bladder. Am J Clin Pathol 1993;99:244-8.

93. Gephardt GN, Baker PB. Interinstitutional comparison of


bladder carcinoma surgical pathology report adequacy. A
College of American Pathologists Q-Probes Study of 7234
bladder biopsies and curettings in 268 institutions. Arch
Pathol Lab Med 1995;119:681-5.

76. Bhan R, Pisharodi LR, Gudlaugsson E, Bedrossian


C. Cytological, histological, and clinical correlations in
intravesical Bacillus Calmette-Guerin immunotherapy. Ann
Diagn Pathol 1998;2:55-60.

94. Murphy WM. ASCP survey on anatomic pathology


examination of the urinary bladder. Am J Clin Pathol
1994;102:715-23.

77. Kannan V, Gupta D. Calculus artifact. A challenge in urinary


cytology. Acta Cytol 1999;43:794-800.

95. Thompson JF, Scolyer RA. Cooperation between


surgical oncologists and pathologists: a key element of
multidisciplinary care for patients with cancer. Pathology
2004;36:496-503.
96. Hammond EH, Henson DE. Practice protocol for the
examination of specimens removed from patients with
carcinoma of the urinary bladder, ureter, renal pelvis, and
urethra. Arch Pathol Lab Med 1996;120:1103-10.
97. Recommendations for the reporting of urinary bladder
specimens containing bladder neoplasms. Association of
Directors of Anatomic and Surgical Pathology. Am J Clin
Pathol 1996;106:568-70.
98. Murphy WM, Crissman JD, Johansson SL, Ayala AG.
Recommendations for the reporting of urinary bladder
specimens that contain bladder neoplasms. Mod Pathol
1996;9:796-8.
99. Philip AT, Amin MB, Tamboli P, Lee TJ, Hill CE, Ro JY.
Intravesical adipose tissue: a quantitative study of its
presence and location with implications for therapy and
prognosis. Am J Surg Pathol 2000;24:1286-90.

78. Highman W, Wilson E. Urine cytology in patients with


calculi. J Clin Pathol 1982;35:350-6.
79. Mariappan P, Zachou A, Grigor KM. Detrusor muscle in the
first, apparently complete transurethral resection of bladder
tumour specimen is a surrogate marker of resection quality,
predicts risk of early recurrence, and is dependent on
operator experience. Eur Urol 2010;57:843-9.
80. Amin MB, Srigley JR, Grignon DJ, et al. Updated protocol
for the examination of specimens from patients with
carcinoma of the urinary bladder, ureter, and renal pelvis.
Arch Pathol Lab Med 2003;127:1263-79.
81. Ro JY, Ayala AG, el-Naggar A. Muscularis mucosa of
urinary bladder. Importance for staging and treatment. Am
J Surg Pathol 1987;11:668-73.
82. Paner GP, Ro JY, Wojcik EM, Venkataraman G, Datta MW,
Amin MB. Further characterization of the muscle layers and
lamina propria of the urinary bladder by systematic histologic
mapping: implications for pathologic staging of invasive
urothelial carcinoma. Am J Surg Pathol 2007;31:1420-9.

56

or ultrasonography [Qu 2010]. However, the risks


associated with exposure to radiation limits the use
of repeated CT virtual cystoscopy in bladder cancer
surveillance.

100. Coblentz TR, Mills SE, Theodorescu D. Impact of second


opinion pathology in the definitive management of patients
with bladder carcinoma. Cancer 2001;91:1284-90.
101. Herr HW. The value of a second transurethral resection in
evaluating patients with bladder tumors. J Urol 1999; 162:
74-6.

Recommendation

102. Dutta SC, Smith JA Jr, Shappell SB, Coffey CS, Chang
SS, Cookson MS. Clinical under staging of high risk
nonmuscle invasive urothelial carcinoma treated with
radical cystectomy. J Urol 2001; 166: 490-3.

It is difficult to assess the role of virtual cystoscopy


in clinical practice. Certainly catheterless forms of
virtual cystoscopy may allow patients to avoid the
discomfort associated with standard cystoscopy.
Cystoscopy does, however, offer the advantages
of being able to detect tumors of all sizes, being
able to detect areas of mucosal irregularity or CIS,
and allowing biopsy or definitive treatment to be
performed. The broad availability of cystoscopy
and familiarity with its use mean that it remains the
gold standard. Grade C

D. Imaging and Bladder Cancer


I. Methods
Medline, Embase, the Cochrane Library, Biosis, and
Science Citation Index were searched using the
terms bladder cancer, IVU, US, CT, MRI, and PET.
In addition, a direct search was done of the last 3
months issues of the major urology and radiology
journals.

III. CT in Localized Bladder


Cancer Staging
Cross-sectional imaging has a very real role in
assessing the local extent of bladder cancer. The
National Comprehensive Cancer Network guidelines
for Bladder Cancer [Version 1. 2011] state that if the
cystoscopic appearance of a tumor is solid, is high
grade, or suggests invasion into muscle, a CT scan
of the abdomen and pelvis is recommended before
TURBT [Level 4]. CT can be used on its own or in
combination with positron emission tomography
scanning in the detection of both local and distant
disease. The reported sensitivity of CT for detecting
and staging bladder cancer ranges from 62-91%, and
the reported specificity ranges from 63-95% [Baltaci
2008, Blanco Diez 2003, Jaume 2003, Kim 2004,
Koplay 2010]. Studies with multidetector CT showed
sensitivity of 89-91% and specificity of 92-95% [Kim
2004, Koplay 2010]. Protocols for CT for localized
bladder cancer staging vary, as do CT machines
themselves. It has been suggested that acquisition
60 seconds after the administration of contrast yields
the best results [Kim 2004]. In staging patients who
have undergone TURBT, delaying the scan until at
least 1 week after surgery facilitates interpretation
of images and reduces the number of false-positive
results associated with postoperative inflammatory
changes [Kim 2004].

Overall, 3,586 articles were retrieved. The decision


was made to limit the review to articles published
within the last 10 years, which left 1,814 articles.
From these articles, case reports, articles that
pertained to benign disease, and articles that did not
refer to bladder cancer per se were excluded. A total
of 186 articles remained.
These 186 articles were then analyzed as to the
level and quality of evidence and recommendations
made according to the International Consultation on
Urological Diseases guideline [Abrams 2010]. All
evidence quoted is Level 3 unless otherwise stated.

II. Virtual Cystoscopy


Both the American Urological Association and
the European Association of Urology recommend
WLC as the gold standard for the assessment of
suspected bladder tumors [Babjuk 2008, Hall 2007].
However, discomfort or pain is reported by more than
one-third of people undergoing flexible cystoscopy
[Van der Aa 2007]. In an effort to reduce discomfort
while providing information equivalent to or perhaps
superior to the information provided by flexible
cystoscopy, virtual cystoscopy was developed.
Initially, CT virtual cystoscopy was described;
subsequently, MRI virtual cystoscopy was described.
The place of virtual cystoscopy in the diagnostic
pathway is discussed elsewhere in this document.
Virtual cystoscopy has reported sensitivities and
specificities of the order of 90-95%. A meta-analysis
of 26 virtual cystoscopy studies involving 3,084
patients concluded that CT virtual cystoscopy has
higher diagnostic value than MRI virtual cystoscopy

RecommendationS
CT cannot reliably differentiate non-muscleinvasive bladder cancer from T2 disease. Thus,
TURBT with inclusion of muscle layer remains the
preferred diagnostic option. Grade C
CT and in particular multidetector CT is of use in
detecting extravesical spread. Grade B
CT should be delayed for at least 7 days in patients
who have undergone TURBT. Grade B

57

technique, including sensitivity of 94% and specificity


of 95% [Deserno 2004], other studies have shown
little difference compared to contrast-enhanced MRI
[Thoeny 2009, Harisinghani 2005].

IV. MRI in Localized Bladder


Cancer Staging
MRI is performed using T1-weighted, T2-weighted
imaging, with or without gadolinium enhancement
and diffusion weighting. Studies with diffusionweighted MRI show that this method can detect 98100% of bladder tumors [El-Assmy 2009, Matsuki
2007]. Staging sensitivity ranges from 68-80%
and specificity ranges from 90-93% [Tillou 2008,
Watanabe 2009]. Overall, MRI is probably not more
sensitive in defining the T classification than is
TURBT, but in cases in which there is discordance
between findings, close follow-up with early secondlook surgery is advised because as many as 60% of
cases may be understaged on initial TURBT [Tillou
2008]. There is some evidence that MRI can predict
tumor grade [Takeuchi 2009] and other biological
processes such as angiogenesis [Tuncbilek 2009].

2. PET Scanning
A number of tracers have been explored in positron
emission tomography (PET) scanning for bladder
cancer, including 11C-methionine [Ahlstrom 1996)
and 11C-choline [Picchio 2006, de Jong 2002). The
utility of PET for local staging of bladder cancer
has been similar to that of CT; the utility of PET in
evaluation for lymph node metastases has been
superior to that of CT.
F-fluorodeoxyglucose (FDG) is the contrast agent
most widely studied in PET imaging for bladder
cancer. Excretion of FDG in the urine can lead to
misdiagnosis: focal FDG accumulation may mimic
tumor uptake of FDG, whereas diffuse FDG activity
may obscure an FDG-avid pelvic lesion. Repeated
bladder irrigation and retrograde filling and proneposition imaging are useful techniques to ascertain
the nature of any observed FDG accumulation
[Anjos 2007, Lin 2009). Diuretic protocols have been
described that can significantly lower bladder FDG
activity and potentially improve image quality [Nijjar
2010].
18

Comparative studies suggest that MRI is superior


to CT in preoperative staging of bladder cancer,
detecting 78-90% of tumors compared to 67-85%
with CT [Akmangit 2003, Fernandez Mena 2001].
MRI also has the advantage of considerably less
artifact in patients with hip prostheses [Charnley
2005].

Recommendations

Studies using FDG-PET/CT in patients with


muscle-invasive bladder carcinoma cancer show a
sensitivity of 57-81% and a specificity of 88-100%
in the detection of pelvic lymph node metastases
[Apolo 2010, Drieskens 2005, Lodde 2010, Kibel
2010]. Similar results are seen in the assessment of
recurrent disease [Jadvar 2008].

Diffusion-weighted MRI has poor sensitivity in


differentiating between Ta, T1, and T2 bladder
tumors. Grade B
Diffusion-weighted MRI is, however, very useful in
the detection of T3 and T4 disease. Grade B

While comparative studies generally indicate


superiority of PET scanning over CT and MRI in
the detection of lymph node metastases [Apolo
2010, Lodde 2010], at least 1 study has reported
equivalence of PET and CT [Swinnen 2009].

V. Detection of Nodal Disease


and Metastasis
1. CT and MRI

Recommendations

Lymph node metastases in bladder cancer are


detected by CT with a sensitivity of 31-50% and
specificity of 68-100% [Baltaci 2008, Lodde 2010,
Picchio 2006]. While MRI can detect more lymph
nodes overall than CT and is in particular better
than CT in the detection of nodes smaller than
5 mm [Saokar 2010], the ability of MRI to identify
tumor in normal-sized or slightly enlarged nodes
is poor. Efforts to image tumor in normal-sized
nodes with ultrasmall superparamagnetic iron
oxide particles have met with mixed results. These
particles are taken up by macrophages, which
allows the architecture of individual, even normalsized, nodes to be identified. Areas not taking up
ultrasmall superparamagnetic iron oxide particles
are presumed to contain metastases. After early
studies reported very encouraging results with this

CT, MRI, and PET scanning are all of use in the


detection of lymph node metastasis. Grade B
Of the 3 modalities, PET scanning appears to have
the greatest accuracy. Grade C
Experience with ultrasmall superparamagnetic iron
oxide particles is not broad enough to recommend
their routine use. Grade D

VI. Imaging after Bladder


Cancer Treatment
CT after TURBT has a sensitivity of 89% and a
specificity of 95% for the detection of extravesical
spread [Kim 2004]. PET has accuracy similar to

58

that of CT in detecting residual bladder cancer after


TURBT and outperforms CT in detecting lymph node
enlargement [Picchio 2006].

7. Baltaci, S., B. Resorlu, et al. (2008). Computerized


tomography for detecting perivesical infiltration and lymph
node metastases in invasive bladder carcinoma. Urol Int
81:399-402

MRI can differentiate between responders and


nonresponders to chemotherapy with a sensitivity
of 91% and specificity of 93% [Schrier 2006]. In
patients receiving chemotherapy and radiotherapy
prior to surgery, MRI has 57% sensitivity and 92%
specificity in predicting the final histology, and
diffusion-weighted MRI demonstrates significantly
greater accuracy than T1- and T2-weighted MRI
[Yoshida 2010].

8. Blanco Diez, A., L. Alvarez Castelo, et al. (2003). Staging of


infiltrating bladder cancer. Role of CT scan. Arch Esp Urol
56(1):23-29.
9. Charnley, N., A. Morgan, et al. (2005). The use of CTMR image registration to define target volumes in pelvic
radiotherapy in the presence of bilateral hip replacements.
Br J Radiol 78(931): 634-636.
10. de Jong, I.J., J. Pruim, et al. (2002). Visualisation of bladder
cancer using (11)C-choline PET: first clinical experience.
Eur J Nucl Med Mol Imaging. 2002; (10):1283-8.
11. Deserno, W.M, M.G. Harlsinghani, et al. (2004) Urinary
bladder cancer: Preoperative nodal staging with
ferumoxtran-10-enhanced MR imaging. Radiology 233:449456

Chemotherapy results in reduced cellular activity in


viable cancer cells. Thus, the sensitivity of PET for
the detection of metastatic disease drops from 77%
in untreated patients to 50% in patients treated with
chemotherapy [Liu 2006].

12. Drieskens, O., R. Oyen, et al. (2005). FDG-PET for


preoperative staging of bladder cancer. Eur J Nucl Med Mol
Imaging 32(12):1412-7.

VII. General Comments

13. El-Assmy, A., M. E. Abou-El-Ghar, et al. (2009). Bladder


tumour staging: comparison of diffusion- and T2-weighted
MR imaging. Eur Radiol 19(7): 1575-1581.

There is a paucity of randomized controlled trials


of imaging for bladder cancer, which is somewhat
surprising as bladder cancer would seem to lend
itself to such studies. Many of the studies that have
been published were underpowered and quote
differences in outcomes that are frequently based
on trends rather than proven differences. The only
meta-analysis identified in the search was that by
Qu et al. The ICUD has suggested that different
considerations should be applied in the assessment
of diagnostic and investigative tools. One of the tests
that they have suggested is to ask whether the tool
has an effect in clinical practice. By this test, virtual
cystoscopy is disappointingwhile virtual cystoscopy
has been the subject of numerous articles over the
last decade, few centers have adopted this technique
into routine clinical practice.

14. Fernandez Mena, F. J. and Y. C. Moreno-Torres (2001).


[Bladder cancer]. Arch Esp Urol 54(6): 493-510.
15. Hall MC, Chang SS, Dalbagni G, et al. (2007) Guideline
for the management of nonmuscle invasive bladder cancer
(stages Ta, T1, Tis): 2007 update. J Urol 178:2314-2330
16. Harisinghani, M.G., M.A. Saksena, et al. (2005).
Ferumoxtran-10-enhanced MR lymphangiography: Does
contrast-enhanced imaging alone suffice for accurate
lymph node characterization? AJR 186(1):144-148
17. Jadvar, H., V. Quan, et al. (2008). [F-18]-Fluorodeoxyglucose
PET and PET-CT in diagnostic imaging evaluation of
locally recurrent and metastatic bladder transitional cell
carcinoma. Int J Clin Oncol 13(1): 42-47.
18. Jaume, S., Ferrant, M., et al (2003). Tumor detection in the
bladder wall with a measurement of abnormal thickness in
CT scans. IEEE Trans Biomed 50(3): 383-390
19. Kibel, A.S., F. Dehdashti, et al.(2009) Prospective study of
[18F]fluorodeoxyglucose positron emission tomography/
computed tomography for staging of muscle-invasive
bladder carcinoma. J Clin Oncol. 27(26):4314-20.
20. Kilickesmez, O., T. Cimilli, et al. (2009). Diffusion-weighted
MRI of urinary bladder and prostate cancers. Diagn Interv
Radiol 15(2): 104-110.
21. Kim, J.K., P. Soo-Youn, et al. (2004). Bladder cancer:
Analysis of multi-detector row helical CT enhancement
pattern and accuracy in tumor detection and perivesical
staging. Radiology 231(3): 725-731
22. Koplay, M., M. Kantarci, et al. (2010). Diagnostic efficiency
of multidetector computed tomography with multiplanar
reformatted imaging and virtual cystoscopy in the
assessment of bladder tumors after transurethral resection.
J Comput Assist Tomogr 34:121-126.
23. Lin, W.Y., K.B. Wang, et al. (2009). Unexpected accumulation
of F-18 FDG in the urinary bladder after bladder irrigation
and retrograde filling with sterile saline: a possible pitfall in
PET examination. Clin Nucl Med. 34(9):560-3.
24. Liu, I. J., Y. H. Lai, et al. (2006). Evaluation of
fluorodeoxyglucose positron emission tomography imaging
in metastatic transitional cell carcinoma with and without
prior chemotherapy. Urol Int 77(1): 69-75.
25. Lodde, M., L. Lacombe, et al. (2010) Evaluation of
fluorodeoxyglucose positron-emission tomography with
computed tomography for staging of urothelial carcinoma.
BJU Int. epub Feb 11.

References
1. Abrams, P. and S. Khoury. (2010). International Consultation
on Urological Diseases: Evidence-based medicine overview
of the main steps for developing and grading guideline
recommendations. Neurourol Urodyn 29:116-118
2. Ahlstrom, H., P.U. Malmstrom, et al. (1996). Positron
emission tomography in the diagnosis and staging of
urinary bladder cancer. Acta Radiol 37: 1805
3. Akmangit, I., H. Lakadamyali, et al. (2003). [Staging of
urinary bladder tumors with CT and MRI]. Tani Girisim
Radyol 9(1): 63-69.
4. Anjos, D.A., E.C. Etchebehere, et al. (2007). 18F-FDG
PET/CT delayed images after diuretic for restaging invasive
bladder cancer. J Nucl Med. 48(5):764-70.
5. Apolo, A.B., J. Riches, et al. (2010). Clinical value of
fluorine-18 2-fluoro-2-deoxy-D-glucose positron emission
tomography/computed tomography in bladder cancer. J
Clin Oncol. 28(25):3973-8.
6. Babjuk, M., W. Oosterlinck, et al. (2008) EAU guidelines
on nonmuscle invasive urothelial carcinoma of the bladder.
Eur Urol 54:303-314

59

Summary of the
Recommendations

26. Matsuki, M., Y. Inada, et al. (2007). Diffusion-weighted MR


imaging for urinary bladder carcinoma: initial results. Eur
Radiol 17(1): 201-204.
27. Nijjar, S., J. Patterson, et al. (2010). The effect of
furosemide dose timing on bladder activity in oncology
imaging with 18F-fluorodeoxyglucose PET/CT. Nucl Med
Commun.31(2):167-72.

1. Screening and Diagnosis:


a. Currently, the available evidence indicates
that bladder cancer screening is not helpful in
improving survival. Further studies to evaluate
the utility of bladder cancer screening are
warranted (Grade C).

28. Nishimura, K., C. Fujiyama, et al. (2009). The effects of


neoadjuvant chemotherapy and chemo-radiation therapy
on MRI staging in invasive bladder cancer: comparative
study based on the pathological examination of whole layer
bladder wall. Int Urol Nephrol 41(4): 869-875

b. Bladder cancer screening should probably be


confined to patients at high risk for bladder
cancer (Grade C).

29. Picchio, M., U. Treiber, et al. (2006). Value of 11C-choline


PET and contrast-enhanced CT for staging of bladder
cancer: correlation with histopathologic findings. J Nucl
Med. 47(6):938-44

c. Screening in high-risk patients should consist


of yearly cytologic examination of urine (urine
cytology) and dipstick urinalysis to evaluate
for hematuria (Grade C).

30. Qu, X., X. Huang, et al. (2010). Comparison of virtual


cystoscopy and ultrasonography for bladder cancer
detection: A meta-analysis. Eur J Radiol 2010 May 7 Epub
31. Saokar, A., T. Islam, et al. (2010). Detection of lymph
nodes in pelvic malignancies with Computed Tomography
and Magnetic Resonance Imaging. Clin Imaging 34(5):
361-366.

d. T
 here is no correlation between the number of
red blood cells per high-power field seen on
urine microscopy and the diagnosis of bladder
cancer (Grade B).

32. Schrier, B. P., M. Peters, et al. (2006). Evaluation of


chemotherapy with magnetic resonance imaging in patients
with regionally metastatic or unresectable bladder cancer.
Eur Urol 49(4): 698-703

e. Nearly all patients with bladder cancer


diagnosed on cystoscopy have some form of
microscopic or macroscopic hematuria. Hence,
patients with microscopic or macroscopic
hematuria require further evaluation, such as
with flexible cystoscopy (Grade D).

33. Swinnen, G., A. Maes, et al. (2009) FDG-PET/CT for the


Preoperative Lymph Node Staging of Invasive Bladder
Cancer. Eur Urol. 2009
34. Takeuchi, M., S. Sasaki, et al. (2009). Urinary bladder
cancer: diffusion-weighted MR imaging-accuracy for
diagnosing T stage and estimating histologic grade.
Radiology 251(1): 112-121.

f. M
 icroscopic hematuria in patients with bladder
cancer varies in intensity and is intermittent;
hence, lack of hematuria on a single urinalysis
does not exclude bladder cancer (Grade C).

35. Thoeny, H.C., M. Triantafyllou, et al. (2009). Combined


ultrasmall superparamagnetic particles of iron oxideenhanced and diffusion-weighted magnetic resonance
imaging reliably detect pelvic lymph node metastases
in normal-sized nodes of bladder and prostate cancer
patients. Eur Urol 55: 761-769.

g. In patients with irritative voiding symptoms


such as dysuria, frequency, and urgency,
bladder cancer, particularly carcinoma in situ
(CIS), must be ruled out (Grade C).

36. Tillou, X., E. Grardel, et al. (2008). [Can MRI be used to


distinguish between superficial and invasive transitional
cell bladder cancer?]. Prog Urol 18(7): 440-444.
37. Tuncbilek, N., M. Kaplan, et al. (2009). Value of dynamic
contrast-enhanced MRI and correlation with tumor
angiogenesis in bladder cancer. AJR Am J Roentgenol
192(4): 949-955.

h. For patients with asymptomatic microscopic


hematuria without risk factors for urothelial
carcinoma, urine cytology or cystoscopy can
be used (Grade B).

38. Van der Aa, M.N., E.W. Steyerberg, et al. (2008). Patients
perceived burden of cystoscopic and urinary surveillance
of bladder cancer: a randomized comparison. BJU
Int101(9):1106-1110

i. For initial work-up of patients with hematuria,


imaging of the upper tracts must be performed
(Grade B).

39. Watanabe, H., M. Kanematsu, et al. (2009). Preoperative T


staging of urinary bladder cancer: does diffusion-weighted
MRI have supplementary value? AJR Am J Roentgenol
192(5): 1361-1366.

2. Cystoscopy and Urinary Markers


a. Cystoscopy alone is the most cost effective
method to detect recurrence of bladder cancer.
Among urinary markers, urine cytology is the
gold standard for surveillance of patients with
non muscle invasive cancer (Grade B).

40. Yoshida, S., F. Koga, et al. (2010). Initial experience of


diffusion-weighted magnetic resonance imaging to assess
therapeutic response to induction chemoradiotherapy
against muscle-invasive bladder cancer. Urology 75(2):
387-391.

b. C
 ytopathologists
should
nomenclature (Grade C).

use

uniform

c. Bladder wash cytology provides a better


diagnostic yield than voided urine cytology
(Grade B).

60

d. When cystoscopy is performed, thorough


cystoscopy with minimal manipulation should
be performed, the residual urine should be
collected, and then a formal bladder lavage
should be performed. Both the residual urine
and the bladder lavage specimen should be
sent for cytology (Grade D).

e. For patients with invasive bladder tumors,


evaluation for metastatic disease should
include chest radiography, liver function tests,
and alkaline phosphatase measurement
(Grade C).
f. B
 one scan is not necessary in all cases but
should be performed in patients with bone pain
or elevated alkaline phosphatase concentration
(Grade B).

e. In the follow-up of patients with urothelial


neoplasms, urine cytology is most useful for
diagnosis of high-grade tumor recurrence
(Grade B). Urine cytology is especially
valuable
for
differentiating
high-grade
urothelial carcinomas from low-grade urothelial
neoplasms (Grade B).

5. Technique of TURBT
a. A
 t present, there is insufficient information
to support the recommendation of a specific
technique for TURBT (Grade D).

f. T
 here is no consensus on the application of
urinary markers in the diagnosis of bladder
cancer (Grade D).

b. Patients undergoing TURBT should be given


appropriate prophylactic antibiotics (Grade
B).

g. T
 he role of urinary markers in particular FISH
(fluorescence in situ hybridization) appears
to be most useful in the setting of a negative
cystoscopy and atypical cytology. (Grade C).

c. T
 he shape, size, and location of the
tumor should be documented explicitly as
these details provide important prognostic
information (Grade C).

3. Newer Modalities for Endoscopy

d. T
 he appearance of the base of the tumor,
whether sessile or pedunculated, should be
documented as the appearance often predicts
the invasiveness of the tumor (Grade C).

a. White light cystoscopy (WLC) is the gold


standard for evaluation of the lower urinary
tract (Grade B).

e. Separate tumor base and margin biopsies


should be performed during TURBT (Grade
C).

b. Fluorescence cystoscopy improves the rate of


detection of carcinoma in situ (CIS) (Grade B).
Fluorescence-guided transurethral resection
of bladder tumor (TURBT) decreases the
likelihood of leaving behind residual tumor
(Grade B).

f. During TURBT, complete tumor resection


should be attempted except in patients with
diffuse CIS (Grade C).
g. CIS may manifest as velvety erythematous
patches. Endoscopists should specifically look
for such changes, and all suspicious lesions
should be biopsied (Grade C).

c. Fluorescence cystoscopy can be used in the


examination of patients with positive findings
on voided urine cytology but negative findings
on WLC (Grade C).

h. During TURBT, bladder perforation should be


avoided (Grade C).

4. Imaging
a. For initial work-up of patients with hematuria,
imaging of the upper tracts must be performed
(Grade B).

i. When the ureteral orifice is resected, cutting


current should be used; coagulation of the
ureteral orifice should be avoided. Three to 6
weeks after resection of the ureteral orifice, a
functional study should be obtained to check
for stenosis (Grade C).

b. While computed tomography (CT) urography


is becoming the de facto standard imaging
modality for patients with bladder cancer,
intravenous urography (IVU), CT of the
abdomen and pelvis, ultrasound, and magnetic
resonance imaging (MRI) are options (Grade
C).

j. A
 ggressive resection of a tumor in a bladder
diverticulum can lead to perforation. Low-grade,
noninvasive tumors in a bladder diverticulum
may be treated with transurethral resection or
fulguration with or without intravesical therapy
(Grade C).

c. Imaging for staging should be obtained prior to


TURBT or delayed for 1-2 weeks after TURBT
to avoid artifacts (Grade C).

k. A
 second TURBT should be performed in all
patients with a high-grade Ta lesion or any T1
lesion (Grade B). The optimal timing of repeat
TURBT is within 1-4 weeks after the first
resection (Grade C).

d. A
 bdominal and pelvic imaging (MRI or CT)
is not accurate for staging of primary bladder
tumors but may be useful for confirming or
ruling out metastatic disease (Grade B).

61

l. Routine random bladder biopsies are not


recommended (Grade C). Patients with positive
findings on urine cytology and normal findings
on cystoscopy should undergo random bladder
biopsies (Grade B). Patients undergoing
partial cystectomy should undergo random
bladder biopsies (Grade C).
m. Routine prostatic urethral biopsy is not
recommended at initial evaluation. Prostatic
urethral biopsy is indicated in cases of
multifocal urothelial carcinoma of the bladder,
CIS, and visible abnormalities of the prostatic
urothelium (Grade B).
n. Prostatic urethral biopsy should be performed
using electrocautery loop resection including
the 5 oclock and 7 oclock positions of the
verumontanum (Grade B).

62

Committee 2

Pathology Consensus Guidelines


by the Pathology of Bladder
Cancer Work Group
Chairs:
Mahul B. Amin,
Victor E. Reuter
Members:
Jonathan I. Epstein, David J. Grignon, Donna E. Hansel,
Oscar Lin, Jesse K. McKenney, Rodolfo Montironi,
Gladell P. Paner, Mark Soloway,
and

Members of the Pathology of Bladder Cancer Work Group


Participants:
Hikmat A. Al-Ahmadie, MD (New York, NY, USA)
Ferran Algaba, MD (Barcelona, Spain)
Syed Ali, MD (Baltimore, MD)
Isabel Alvardo-Caberrera, MD (Mexico City, Mexico)
Lukas Bubendorf, MD, PhD (Basel, Switzerland)
Liang Cheng, MD (Indiana, IN, USA)
John C. Cheville, MD (Rochester, MN, USA)
Glenn Kristiansen, MD (Zurich, Switzerland)
Richard J. Cote, MD (Miami, FL, USA)
Brett Delahunt (Wellington, New Zealand)
John N. Eble, MD (Indianapolis, IN, USA)
Mahmoud Elbaz, MD (Cairo, Egypt)
Elizabeth Genega, MD (Boston, MA, USA)
Christian Gulmann, MD (Dublin, Ireland)
Arndt Hartmann, MD (Erlangen, Germany)

Cord Langer, MD (Graz, Austria)


Antonio Lopez-Beltran, MD (Cordoba, Spain)
Cristina Magi-Galluzzi, MD, PhD (Cleveland, OH, USA)
Jorda Merce, MD, PhD (Miami, FL, USA)
George J. Netto (Baltimore, MD, USA)
Esther Oliva, MD (Boston, MA, USA)
Priya Rao, MD (Houston, TX, USA)
Jae Y. Ro, MD, PhD (Seoul, S. Korea)
John R. Srigley, MD (Toronto, ON, Canada)
Satish K. Tickoo, MD (New York, NY, USA)
Toyonori Tsuzuk, MD, PhD (Nagoya, Japan)
Saleem A. Umar, MD (Miami, FL, USA)
Theo Van der Kwast, MD (Toronto, ON, Canada)
Robert H. Young, MD (Boston, MA, USA)

63

Table of Contents
Introduction

III. Prognostic and Predictive Biomarkers


in Urothelial Carcinoma

A. Urinary Bladder Histoanatomy and


Impact on Cancer Diagnosis and Staging

IV. The Utility of Immunohistochemistry


In The Differential Diagnosis Of Primary
Versus Metastatic Tumors To The Urinary
Bladder

I. Epithelial Changes
II. Lamina Propria (LP)

V. The Use of Immunohistochemistry in


Flat Urothelial Lesions with Atypia

III. Muscularis Mucosae


IV. Indeterminate Muscle type
V. Muscularis Propria

F. RECOMMENDED NOMENCLATURE FOR URINE


CYTOLOGY

VI. Adipose Tissue

I. Nomenclature and Reporting

VII. Regional Variations in Bladder Wall

II. Role of urine cytology in screening


and monitoring patients with bladder
cancer

B. Classification and Grading of Noninvasive Urothelial Neoplasms

III. Role of FISH in screening and


monitoring of bladder carcinoma

C. UPDATES ON GRADING OF NON-INVASIVE


AND INVASIVE UROTHELIAL CARCINOMA OF THE
URINARY BLADDER

G. UPDATED PROTOCOL FOR THE EXAMINATION


AND HANDLING OF SPECIMENS FROM PATIENTS
WITH URINARY BLADDER CARCINOMA

I. Grading Non-invasive Papillary


Urothelial Neoplasms: Tumor
Distribution and Impact to Outcome of
WHO (2004) Versus the Prior 1973 WHO
Grading System

I. Introduction
II. Relevant Clinical History
III. Urologists Handling of Specimens
IV. Pathologist Handling of Biopsy
Specimens

II. Inter- and Intra-Observer


Reproducibility Studies on Grading

V. Pathologist Handling of
Transurethral resections

III. Grading Papillary Urothelial


Neoplasms with Histologic Heterogeneity

VI. Pathologist Handling of Cystectomy


Specimens

IV. Grading of Invasive Urothelial


Carcinomas (pT1 and pT2+ tumors)

VII. Reporting of Histologic Grade

D. Histologic Types of Bladder Cancer


Including Variants of Urothelial
Carcinoma

VIII. Reporting Variant Urothelial


Histology

I. UROTHELIAL CARCINOMA VARIANTS


Carcinoma

IX. Lymphovascular Invasion (LVI)

II. SQUAMOUS CELL CARCINOMA

X. Reporting on Lymph Nodes

III. ADENOCARCINOMA VARIANTS

XI. Reporting of Prostate Involvement

IV. NEUROENDOCRINE NEOPLASMS

XII. Frozen sections

E. ROLE OF IMMUNOHISTOCHEMISTRY AND


ANCILLARY DIAGNOSTICS IN THE DIAGNOSIS,
STAGING, PROGNOSTICATION AND PREDICTION
OF BLADDER CANCER IN CONTEMPORARY
CLINICAL PRACTICE
I. Immunohistochemical Markers for
Staging of Bladder Cancer
II. Summary of molecular/genetic
alterations in urothelial carcinoma

64

Pathology Consensus Guidelines


by the Pathology of Bladder Cancer
Work Group
Mahul B. Amin, Victor E. Reuter,
Jonathan I. Epstein, David J. Grignon, Donna E. Hansel, Oscar Lin,
Jesse K. McKenney, Rodolfo Montironi, Gladell P. Paner,
Mark Soloway,
and

Members of the Pathology of Bladder Cancer Work Group


by the pathologist in order to provide optimal pathologic reporting of bladder cancer, 4) on nomenclature
for inverted lesions of the bladder, 5) the histologic
types of bladder cancer and the variants of urothelial cancer, some relatively recently described, with
a focus on definition for diagnosis and their implied
clinico-pathologic significance, 6) role of immunohistochemistry and molecular studies in contemporary
routine practice diagnosis, prognostication or predicition of bladder cancer, 7) the reporting of bladder
cancer in transurethral resection of bladder tumors
(TURBT) and cystectomy specimens, and 8) recommended nomenclature for urine cytology and role of
ancillary testing in urine specimens.

Introduction
There have been several forums in the last two
decades in which an international consensus
dialogue pertaining to the pathology of the bladder
cancer have been conducted and promulgated.
In this document we provide new consensus
guidelines built on past such endeavors, supported
by comprehensive literature review of the recent
literature where possible, and based on the best
practices of an expert group of 38 international
urologic pathologists from 13 countries. While strong
evidence is commonly lacking to make many of the
recommendations herein, the proposals made are in
the interest to move the field forward, acknowledging
where data is lacking, but at the same time providing
consensus guidelines to our clinician colleagues to
provide optimal patient care. It is hoped that when
the guidelines made here are based on the absence
of a high level of evidence in the literature, they will
prompt future studies using these guidelines as a
common platform of nomenclature and the state of
understanding of the pathology of bladder cancer
as it pertains to contemporary patient care in the
year 2011. Where possible, levels of evidence and
grades of recommendation (Table 1) are provided.
It must be noted, however, that the criteria for
levels and grades of recommendation are primarily
promulgated for clinical studies and have been used
by our Pathology of Bladder Cancer Work Group as
a framework to support their guidelines.

Table 1. Levels of evidence and grades of


recommendation*

Levels of Evidence

Level 1: Metanalysis of randomized trials of a good


quality randomized trial
Level 2: Low quality randomized trial or metanalysis
of good quality prospective cohort studies
Level 3: Good quality retrospective case control
studies or case series
Level 4: Expert opinion based on first principles or
bench research, not on evidence
Grades of Recommendation
Grade A: Usually consistent level 1 evidence
Grade B: Consistent level 2 or 3 evidence or
majority evidence from randomized trials
Grade C: Level 4 evidence, majority evidence from
level 2-3 studies, expert opinion
Grade D: No recommendation possible because of
inadequate or conflicting evidence

In this international consultation, we focus on a select important unique aspects and provide an update
on: 1) the knowledge of the histoanatomy of the
bladder as it pertains to the diagnosis and staging of
bladder cancer, 2) the prognostic significance of histologic grading by the WHO (2004)/ISUP system, its
contributions, shortfalls and opportunities for refinement, 3) the information needed from the clinicians

*Adapted from Evidence-based medicine: Overview


of the main steps for developing and grading
guideline recommendations, by P Abrams, A Grant,
and S Khoury, January 2004

65

2. Urothelial Hyperplasia

A. Urinary Bladder
Histoanatomy and Impact on
Cancer Diagnosis and Staging

Urothelial hyperplasia is defined as markedly


thickened urothelium in the absence of cytologic
atypia, and can be flat or papillary. Flat hyperplasia
may be seen adjacent to low-grade papillary tumors
[4]. Papillary urothelial hyperplasia is characterized
by presence of undulating hyperplastic urothelium
arranged into thin mucosal papillary folds of varying
heights in a non inflamed setting [4]. The lack of
cytological atypia and true fibrovascular cores
distinguishes this lesion from papillary neoplasia.
Papillary hyperplasia (Figure 3) is often seen in
patients with prior and subsequent diagnosis of lowgrade papillary urothelial lesions (Level 3), and for
a subset of patients is suggested to be a precursor
lesion for mainly the low-grade lesions [4,5]. This
hypothesis is further supported by the presence of
some genetic alterations found in bladder cancer that
also occur in urothelial hyperplasia [6-8]. Hyperplastic
urothelium may also harbor a varying degree of
atypia, and is suggested to be associated with CIS
and high-grade papillary urothelial carcinoma (Level
3) [9].

I. Epithelial Changes
1. Urothelial Denudation
Prior instrumentation and prior intravesical therapy
are frequent contributors to denudation at cystoscopy
or in bladder specimens. Urothelial carcinoma in situ
(CIS) is also often associated with prominent cellular
discohesion and exfoliation of neoplastic cells in
the urine. The presence of extensive denudation
in bladder biopsy samples has been shown to be
associated with CIS on prior or subsequent bladder
biopsies[1]. (Figure 1) Tissue specimens obtained
by hot wire loop may show urothelial denudation
which can be attributed mainly to thermal effects of
the proceedure, and can be seen even in low risk
patients for CIS. In contrast, substantially denuded
biopsy samples obtained by cold cup biopsy are in
most cases related to CIS. Levi et al [1] showed that
31% of patients with denudation on cold cup biopsy
developed CIS within 24 months, and that CIS was
seen in 75% and 29% of patients with and without
history of CIS, respectively. This demonstrates a
higher risk for subsequent CIS in denuded biopsy
samples in the setting of both cold cup procedure
and with prior CIS (Level 3). In cold cup biopsies with
denudation cystitis, 54% of patients had concurrent
positive urine cytology, stressing the importance of
performing concurrent bladder wash, which has
higher sensitivity for detecting urothelial carcinoma
(Level 3) [2]. Denudation may also have some
implications in papillary urothelial lesions. Papillary
urothelial lesions with extensive urothelial denudation
are more often high-grade carcinomas[3]. (Figure 2)
In denuded papillary lesions, pathologists should do
careful search for residual high-grade cells in order
to avoid under diagnosis (Grade C). Denuded flat
and/or papillary urothelial neoplasms may occur
with prominent cautery artifact or in an anatomically
confined area implicating iatrogenic or mechanical
contributing factors for the epithelial exfoliation.

Figure 1. Urothelial
denudation.

Extensive or complete urothelial denudation in


bladder biopsy samples should be reported (Grade
C). Correlation of denuded biopsy samples with
concurrent cytology results may yield a positive
diagnosis of malignancy in these patients (Grade
C). Urothelial denudation in cold cup biopsies of
cystoscopically abnormal areas in patients with prior
CIS and without any recent intravesical interventions
should be dealt with caution by urologists.

carcinoma

in

situ

with

Figure 2. High-grade papillary urothelial carcinoma


with denudation.

66

From a practical viewpoint there is difficulty and


controversy in diagnosing lesions which do not have
fully established features of papillary neoplasia and
also have accompanying cytologic atypia exceeding
reactive changes. This type of setting may occur in
the seeting of prior treatment and in surveillance
biopsies. As an approach, first, strict criteria including
the absence of true fibrovascular cores are necessary
for the diagnosis of papillary urothelial hyperplasia
versus papillary urothelial neoplasia. Second, the
degree of urothelial atypia when present in the
hyperplastic urothelium should be reported using the
2004 WHO/ISUP terminology for urothelial neoplasia
(e.g. dysplasia or CIS) (Grade C). Third, correlation
with cystoscopy would be valuable to ascertain if
clinically the lesions were deemed to be papillary.
In the absence of true papillae histologically and
in the absence of a clinical papillary presentation,
terminology such as dysplasia with early papillary
formations or CIS with early papillary formations
may be used when criteria for papillary neoplasia
are not fulfilled but there is a background of cytologic
atypia (Grade C). These terms are only descriptive
diagnoses as outcome studies using this terminology
are unavailable. Until substantial evidence for
urothelial hyperplasia as precursor lesion to papillary
urothelial neoplasia becomes available, reporting of
papillary hyperplasia in the absence of dysplasia or
CIS is optional (Grade D).

cell carcinoma, and thus, must be closely followed


(Grade C) [13]. Keratinizing squamous metaplasia
is also related to vesical carcinoma and associated
complications, such as bladder contracture and
ureteral obstruction [14]. Squamous metaplasia is
not uncommon in patients with spinal cord injury or
paraplegia with prolonged catheterization or chronically infected bladders [15,16], and higher risk for
squamous cell carcinoma is reported in these patients (Level 3) [17].
The presence of squamous metaplasia in bladder
biopsies should be reported by pathologists with
specification whether keratinizing or not and focal
or multifocal (Grade C). The word squamoid
should be avoided and unequivocal criteria for
squamous differentiation i.e. intracellular bridges
or evidence of keratinization should be present.
Current evidence does not support any form of
intervention including surveillance cystoscopy for
patients with focal changes, although multifocal and/
or extensive lesions often require intervention due
to intractable symptoms or because they precede
or may be associated concurrent with neoplasia
(Grade C). Spinal cord injury or paraplegic patients
with squamous metaplasia must be followed and
evaluated regularly because of the increased risk for
squamous cell carcinoma (Grade C).

Figure 4. Squamous metaplasia, non-keratinizing.

Figure 3. Papillary urothelial hyperplasia.

4. Glandular Metaplasia

3. Squamous Metaplasia

Glandular metaplasia (cystitis glandularis) (Figure 5)


is a benign proliferative change that may occasionally
show presence of intestinal-type goblet cells (cystitis
glandularis, intestinal type or intestinal metaplasia)
(Figure 6). Glandular metaplasia is supposed
to relate the risk of neoplastic transformation
because intestinal metaplasia often coexists with
adenocarcinoma of the urinary bladder [18,19].
Recent studies however showed that glandular
metaplasia of the urinary bladder is not a strong risk
factor for adenocarcinoma or urothelial cancer (Level
3) [20,21]. The significance of glandular metaplasia as
a pre-neoplastic lesion, particularly in the absence of
intestinal metaplsia, is still controversial, however, its

There are two type of squamous metaplasia in the


urothelial tract, non-keratinizing and keratinizing.
Non-keratinizing squamous epithelium is present in
the trigone in 7586% of women and is considered
to be a normal finding (Figure 4) (Level 2) [10-12]; its
occurrence outside this region is considered metaplastic. Keratinizing type of squamous metaplasia
usually occurs in response to chronic irritative stimuli to bladder such as cystitis, lithiasis, diverticula,
or schistosomiasis. Keratinizing type of squamous
metaplasia can be associated with subsequent
or concurrent in situ or invasive carcinomas with
squamous differentiation (Level 3) and has been
suggested to be a precursor lesion for squamous

67

presence should be reported as it is often associated


and correlates with a cystoscopic abnormality.
Multifocal disease with intestinal metaplasia or that
associated with dysplastic changes may warrant
close clinical follow-up (Grade C). Rare cases of
florid intestinal metaplasia and extensive mucin
extravasation (exuberant cystitis cystica) may result
in a pseudotumorous mass and histologically may
mimic adenocarcinoma [22].

have conditions that may explain the presence of an


environment of ischemia or injury, such as cardiac
and vascular diseases. Viral infection related cellular
changes, for example polyoma (BK) virus-type cellular changes can also mimic urothelial carcinoma [27].
Knowledge of prior therapy procedures is invaluable
for pathologists and clinicians should provide this information as some reactive changes may mimic and
complicate the diagnosis of neoplastic conditions.

Figure 5. Cystitis cystica and glandularis.


Figure 7. Urothelial atypia due to radiation.

Figure 6. Cystitis
metaplasia.

glandularis

with

intestinal
Figure 8. Urothelial atypia after BCG therapy. Note
the presence of granuloma.

5. Reactive Changes
Therapeutic procedures for bladder cancer, such
as chemotherapy, immunotherapy, radiotherapy,
photodynamic and laser treatment, previous resection and gene therapy, may produce alterations in
the urothelial mucosa, and some of these changes
may be mistaken for carcinoma [23]. (Figures 7-9)
Intravesical stones, prolonged indwelling catheter,
trauma, infection, etc. also cause epithelial changes
that can mimic epithelial malignancy. Radiation or
chemotherapy cystitis may show epithelial proliferations that may be confused with invasive urothelial
carcinomas even when the therapy was remotely
performed [24,25]. Pseudocarcinomatous epithelial
hyperplasia occurs most commonly after radiation
or chemotherapy but may also occur in relation to
localized ischemia or injury to urothelium [26]. Most
of these patients without radiation or chemotherapy

Figure 9. Urothelial atypia after mitomycin therapy.

68

basement membrane, should ideally be employed


(Grade C). Studies are needed to establish a
clinically significant definition of microinvasive
urothelial carcinoma. Future studies should also
take into account the LP regional variations such as
the trigone where LP is the thinnest (0.46-1.58 mm)
and muscularis propria boundary is not well defined
[37]. Until there is understanding of the definition, it
is recommended that only stringent criteria (similar
to those proposed above) be used or the term
microinvasive carcinoma not be used (Grade C).

II. Lamina Propria (LP)


1. Interpretation Variability of
(Early) Lamina Propria Invasion
(pT1)
The morphologic features of stromal invasion
include retraction artifact, individual tumor cells,
irregular cords, and strands, or isolated single cells,
cytoplasmic eosinophilia (paradoxical maturation)
and desmoplasia [28-30]. When many or most
of these features are appreciable, diagnosis of
invasion is more definite and reproducible. However,
there are issues with interobserver reproducibility in
the diagnosis of early lamina propria (LP) invasion
[31,32]. In one study, 35% of pT1 carcinomas were
downstaged to pTa, and the reviewers staging
allowed a better estimate of risk for subsequent
progression than the initial staging [32]. Another
study showed agreement for pTa versus pT1 in 80%
among 15 pathologists and in 88% upon re-review,
and consensus and original diagnosis agreed in 68%
of cases [33]. Re-staged pT1 by consensus manner
had better prognostic significance compared to its
original non-pT1 diagnosis [33]. In an interobserver
reproducibility study that focused on early invasion,
Shen et al. showed diagnostic agreement in at
least 15 of 19 genitourinary pathologists in 17 of 26
(65%) of cases and highlighted certain problems
when evaluating early invasive cancer (Level
3). von Brunns nest involvement by urothelial
carcinoma and in the background of inflammation
may be mistaken for invasion and has no impact for
disease progression [34]. Because of the diagnostic
difficulty, when early invasive urothelial carcinoma
is suspected, diagnosis through examination of
additional levels, or in a consensus manner with
a pathology colleague or thru quality assurance
meeting is encouraged (Grade C).

3. Substratification or Substaging of
LP Invasion (pT1)
The usefulness of subclassification or substaging
of LP invasion is still controversial. Subjective and
objective methods have been proposed in the
literature. Subjective methods rely on the presence
of muscularis mucosae (MM) and the vascular
plexus in the LP, which are both inconsistent
histoanatomic landmarks (Level 3) [37] (Figure 10).
Depth of invasion may be established either by twotiered or three-tiered systems. Two-tiered system is
by identifying LP invasion up to (pT1a) or beyond
the MM or vascular plexus (pT1b) and three-tired
system is by identifying submucosal tumor invasion
up to (pT1a), in (pT1b), or beyond (pT1c) the MM or
vascular plexus (Level 3) [38-44]. Since the anatomic
landmarks essential for substatification are not always
present, substaging is not currently recommended in
routine surgical pathology sign out. More objective
methods propose measuring the depth of invasion
with thresholds of 0.5 mm [45] or 1.5 mm [46,47].
Problems precluding use of the depth of invasion
criteria include lack of consistent orientation with
the basement membrane, specimen fragmentation,
issues with reproducibility in measurement, and
lack of established clinical significance. Further, this
method may not be practical for general pathologists,
as it may be laborious particularly when multiple
minute invasion foci are to be measured.

2. Microinvasive Urothelial
Carcinoma

Platz et al. evaluated the reproducibility of pT1


substaging and showed little concordance between
participating pathologists (kappa = 0.22; 95%
C.I. = 0.08-0.36) [48]. Non-biased specimens
obtained from various institutions were used which
is important because handling and processing of
TURBT materials varied among institutions [48].
They concluded that substaging was technically
difficult and has little merit, at least for the general
pathologists and urologists. Other than this study, no
other study has been done to evaluate the usefulness
of pT1 substaging among general pathologists.

The concept of microinvasive urothelial carcinoma


is still controversial and criteria for diagnosis had
varied. Initially, microinvasion was defined as LP
invasion by carcinoma to a depth of 5 mm or less
from the urothelial basement membrane [35]. Others
defined microinvasion as infiltration to a depth of
within 2 mm or less in the LP [30,36]. Lopez-Beltran
et al most recently proposed that the microinvasive
tumor into the LP should be composed of no more
than 20 invading cells measured from the stromaepithelial interface [29].

Substaging of pT1 (Figure 11) is currently not


recommended by the WHO/ISUP consensus
guidelines and this committee due to the lack of
widely accepted and reproducible criteria although
we acknowledge there is much need for future
studies (Grade C). It is however recommended for

Currently, there are no standardized criteria or


prospective study for microinvasive urothelial
carcinoma (Level 4). To diagnose early invasion,
stringent criteria such as: only focal invasion, less
than 1 high power field, or 0.5 mm from the nearest

69

pathologists to provide some form of estimate of


the LP invasion in pT1 tumors with respect to depth
and/or quantity of invasion (e.g. focal, multifocal,
extensive, etc.) (Grade C).

these variations from MP. Weaver et al. [52] have


shown that MM was more often seen in female than
in male bladder LP.
For staging of invasive bladder cancer, involvement
of MM is equivalent to LP invasion or pT1 and
involvement of MP is staged as pT2. Distinction is
very important in initial tumor resection or biopsy
as a vital determinant for subsequent therapeutic
decisions. MM when hyperplastic may be difficult
to differentiate from MP [37,53]. Hyperplastic MM
is composed of more than 3 layers of muscle fibers
thick appearing either as fibers parallel to the surface
mucosa or as small rounded bundles when cut in
cross sections [37,53]. Although hyperplastic MM is
present in the LP of different topographic regions of
the entire bladder specimen, it is most often seen in
the dome [37,53]. There are two described patterns:
(a) aggregates of hyperplastic MM with haphazard
(Figure 12) orientation and irregular outlines
morphologically distinct from that of MP and (b)
hyperplastic compact MM with small parallel muscle
fibers and regular outlines arranged singly or in small
groups (Figure 13). The latter may strongly resemble
MP muscle, and the most reliable distinguishing
feature from MP is on the basis of location in LP [37],
which is not always possible in limited specimens
such as in TURBT (Level 3). Attention to quantity
(usually rare and isolated), location, and comparison
(size being much smaller) with more characteristic
MP, if present, would be helpful; isolated bundles
close to urothelium with haphazard fiber orientation
and irregular outlines favor MM over MP [37]. Other
mimickers of MP are cauterized vascular wall and/or
collagen fibers (Level 3) (Figure 14).

Figure 10. Lamina propria vessels are usually seen


at the level of the muscularis mucosae muscle, but
the location may also vary as shown here.

To avoid confusion in interpretation of the pathology


report, documentation of MM-only invasion by
carcinoma is not required, and should be reported
as urothelial carcinoma with LP invasion (pathologic
stage atleast pT1) when MP is not present and
urothelial carcinoma with LP invasion ( pathologic
stage pT1) when MP is present (Grade C).
Involvement of MM may be included in a comment
to provide information on depth/extent of invasion. It
is important to document whether MP is present in a
bladder specimen with at least high-grade pTa and
all invasive urothelial neoplasia (Grade B). Based on
institutional and clinical preference, some centers do
not report the presence or absence of uninvolved MP
in TURBT specimens containing low-grade papillary
tumors and mention the presence of uninvolved MP
in TURBT specimens containing high-grade tumors,
irrespective of invasion only. Although formal studies
are lacking, information on the amount of MP included
in the specimen (isolated and rare muscle bundles
versus established groups of muscle bundles); or MP
in the area of tumor or not in the vicinity of the tumor,
may yield useful information in the management of
invasive bladder cancer. At this point, reporting this
information is optional and may be communicated in
multidisciplinary conferences (Grade D).

Figure 11. Invasive urothelial carcinoma involving


muscularis mucosae muscle (pT1).

III. Muscularis Mucosae


The muscularis mucosae (MM) is an important
histoanatomic structure to differentiate from
muscularis propria (MP) or deep muscle when
staging invasive urothelial carcinoma in the bladder.
MM muscle bundles are typically thin slender
bundles arranged in single layer of interrupted,
dispersed, or continuous muscle [37,49]. The
bladder MM was initially described by Dixon et al.
[50], and Ro et al. [49] subsequently emphasized
the importance of histologic recognition of MM and
its implications in pathologic staging. Keep et al.
[51] pointed out the importance of recognition of
MM in TURBT specimens to prevent overstaging
of invasive urothelial carcinoma. Paner et al. [37]
further described morphologic variations of MM and
as it relates to the different topographical regions
of the bladder and the challenges in distinguishing

70

IV. Indeterminate Muscle type

Figure 12. Hyperplastic


haphazard type.

muscularis

It may be difficult to distinguish between MM and


MP muscle bundles in some TURBT samples.
Hyperplastic MM, desmoplasia, prior procedures,
or fracturing of muscle bundles by invasive
carcinoma often compound the diagnostic dilemma
in reliably distinguishing MM and MP. Identification
as hyperplastic MM may become challenging
and involvement is equivalent to pT1 and not pT2
invasive carcinoma. On the other hand, MP whose
muscle fibers are splayed by invasive carcinoma
and mimic MM, when involved is staged as pT2.
When indeterminate muscle bundles are observed,
the presence of vascular network with thick walled
vessels may aid in recognition as MM (Level 3)
[37,42,49,54,55]. Also, relationship or distance
from the urothelial basement membrane, quantity
and muscle fiber bundle morphology (haphazard
or in comparison to characteristic MP if present)
may provide additional clues for MM (Level 3)
[37]. Muscle bundles indeterminate between MM
and MP should be reported with terminology such
as invasive urothelial carcinoma with invasion of
muscle, indeterminate type to prompt the urologist
for a restaging biopsy procedure (Grade C). A
second biopsy may provide more definitive staging
information or resolve this problematic situation
(Grade C) [56].

mucosae,

V. Muscularis Propria
Documentation of the presence or absence of MP
and its status of involvement by invasive carcinoma is
required in reporting of biopsy or TURBT specimens
with at least high grade pTa and all pT1+ tumors.
Pathologic staging is considered not adequate
without detailing the presence of uninvolved MP
in pTa or pT1 tumors and its because of the risk
of unsampled MP-invasive disease. Current major
guidelines including those by the American Urological
Association, European Association of Urology
and National Comprehensive Cancer Network
recommend reporting whether MP component
is sampled and involved by tumor in TURBT
specimen. If MP is not present, repeat resection is
recommended for verification of pT stage before
subsequent management decision is made (Grade
B). The recent Japanese bladder cancer guideline
under the auspices of the Japanese Urological
Association and Japanese Society of Pathology also
recommends this approach.

Figure 13. Hyperplastic muscularis mucosae,


compact type which resembles muscularis propria
muscle bundles.

Several terminologies have been used to report MP


muscle such as deep muscle, muscle proper or
detrusor muscle. MM is sometimes reported as
superficial muscle. Use of different terminologies
for bladder muscles may generate confusion in

Figure 14. Dense fibrosis associated with invasive


carcinoma.

71

staging interpretation and specimen adequacy.


These terms are not recommended and use of
standardized nomenclature of MP muscle bundles
is recommended in reporting (Grade C). Obtaining
a separate deep or post tumor resection biopsy
is also recommended to ensure sampling of the
muscularis propria.

or TURBT specimens should not be interpreted as


perivesical fat involvement or pT3 disease. Involvement of adipose tissue by carcinoma in TURBT or
biopsy without MP suggests tumor involving at least
the deeper aspects of LP (Level 3).

The MP is composed of three smooth muscle layers


inner and outer longitudinal layer and a central
circular layer. MP invasion has been divided into
two categories, superficial (pT2a) and deep (pT2b)
based on the depth of tumor invasion. Substaging
of pT2 disease (Figure 15) and recognition of
invasion beyond MP (pT3 disease) is not tenable in
TURBT specimens, where only portion of MP can be
present. In cystectomy specimens, pT2 substaging
is complicated by the lack of a reliable anatomical
landmark to aid in their distinction (Level 3). Currently,
there are no data regarding the reproducibility of
pT2a/b distinction among pathologists (Level 4).
Some outcome studies have shown the lack of
prognostic impact and are against the use of pT2
substaging [57-59]. Although some reports showed
prognostic value of pT2a/b subcategories [60,61].
These varied results may be influenced by the
difficult distinction between pT2a/b, and thus, a more
objective and reproducible anatomical or extent of
disease criterion is needed between pT2a and pT2b
subcategories in cystectomy specimens.

Figure 16. Adipose tissue when present in the lamina


propria is typically situated at the deeper aspect of
lamina propria.

Figure 17. Lamina propria adipose tissue in


transurethral resection tissue fragment.

Because of the consistent presence of adipose


tissue (Figure 19) in the MP, the boundary of
perivesical adipose tissue from the deep (outer)
MP may not be easily delineated (Level 3) [62]
(Figure 20). Distinction between pT2b and pT3a
in a cystectomy is not usually straightforward and
requires the use of low power assessment to identify
tumor extending beyond the outermost boundary of
MP muscle (Level 3). The prognostic significance
of pT3a versus pT2 has also been questioned as it
has been shown to have no significant difference in
recurrence and mortality rate and incidence of lymph
node involvement [63].

Figure 15. Urothelial carcinoma invading muscularis


propria muscle (pT2).

VI. Adipose Tissue


Adipose tissue is frequently present in the LP and
MP of the urinary bladder, usually scant in the former
and abundant in the latter [62]. When fat is present
in the LP, it is more typically in the deeper aspects
or in the vicinity of the MP [37] (Figures 16, 17).
Pathologic staging (pT1 or pT2 Figure 18) should
be performed based on the type of muscle and not
based on adipose tissue. It is necessary to stress
that involvement of adipose tissue by tumor in biopsy

Similarly, there is limited reliability of pT3a vs.


pT3b subcategorization as gross involvement of
perivesical fat (extravesical fat) may not always be
readily recognizable (Level 3). Distinction between
pT3a and pT3b relies solely on gross findings, which
can be ambiguous compared to the microscopic

72

Figure 20. The typically ill-defined boundary between


muscularis propria muscle and perivesical adipose
tissue. Low power is important in distinguishing
pT2b vs. pT3a.

Figure 18. Unequivocal invasion of A) lamina propria


with wispy fascicles of muscularis mucosae (pT1)
and B) thick bundles of muscularis propria (pT2).

Figure 21. Urothelial carcinoma with microscopic


invasion of the perivesical fat (pT3a).

Figure 19. Urothelial carcinoma invading adipose


tissue within the muscularis propria (pT2).

assessment. Perivesical tissue reactions to the


tumor or prior biopsy may be over interpreted as
extravesical fat involvement. In addition, residents
or physician assistants who are less experienced
sometimes perform the bladder gross examination
and may not document subtleties at the macroscopic
level. The prognostic significance of pT3 subcategory
is controversial; some reports did not show survival
differences [64-66] while some did show differences
between pT3a and pT3b [67,68] (Figures 21, 22)
(Level 4).

Figure 22. Urothelial carcinoma with macroscopic


invasion of the perivesical fat (pT3b).

73

B. Classification and Grading


of Non-invasive Urothelial
Neoplasms

VII. Regional Variations in


Bladder Wall
There are variations in the LP among the different
topographical regions of the bladder. The trigonal
area has a relatively flatter surface and LP is
attenuated where the depth ranges from 0.46 mm to
1.58 mm [37]. In contrast, the LP is deepest at the
dome where it ranges from 0.98 to 3.07 mm [37].
The differences of the depth of the LP in the different
bladder regions need to be taken into account for
future definitions of microinvasive carcinoma or
criteria for substaging pT1 disease (e.g. 2 mm depth
can already be pT2 at the trigone). The distribution
and morphologic features of MM also varies at the
different topographical sites of the bladder [37].
The LP MM muscle bundles are not well defined
in the trigone and the MP muscle bundles in this
region become progressively smaller in caliber
more superficially as they almost reach the surface
(Figure 23A, B). Further, smooth muscle bundles
of the bladder base and the MP of the intravesical
segment of the ureter may simulate vesical MP
muscle in samples from this region. Hyperplastic
MM is most commonly encountered in the dome.
In 22%, isolated or small groups of compact regular
hyperplastic MM muscle bundles are seen in deep LP
situated between the more typical slender MM layer
and the MP. The more superficial location of the MP
and rarity of MM in the trigone, relative abundance
of hyperplastic MM in dome, and presence of the
more superficial ureteral MP at its insertion site in
the bladder complicate the traditional pT stage
evaluation of invasive cancer in these regions (Level
3) [37]. There is no data to indicate that differences in
histoanatomy of the bladder influence the prognosis
of muscle invasive urothelial carcinoma (Grade D),
although knowledge of these variations is essential
for accurate staging.

1. Introduction
The non-invasive urothelial neoplasms may be flat,
papillary (exophytic) and inverted (endophytic) depending on their growth pattern relationship with the
surface of the surrounding urothelial mucosa [69].
All 3 growth patterns may be seen concomitantly in
a single tumor (Figure 24). Extensive clinicopathological studies have been published related to the
first two growth patterns, whereas the clinical significance of the third has been dealt with only at a
limited extent.
The World Health Organization (W.H.O.) (2004) /
International Society of Urologic Pathologists (ISUP)
classification system is the recommended system for
classification and grading of urothelial neoplasms. It
has also been endorsed by the W.H.O. 2004 Blue
Book, the 4th Series Armed Forces Institutes of
Pathology Fascicle on Bladder, the 7th edition AJCC
Cancer Staging Manual and this consensus group.
W.H.O. (2004)/ISUP grading system has prognostic
(recurrence and progression potential) significance
between different grades for papillary lesions;
although risk prediction at an individual patient level
is still not possible. There is need for incorporation
of the W.H.O. (2004)/ISUP grade along with other
clinical and pathologic parameters in nomograms
and clinical decision-making algorithms to provide
personalized risk stratification in patients with noninvasive urothelial carcinoma. The aggregate data
in the literature on the W.H.O. (2004)/ISUP system
has shown the following benefits: a) establishment

Figure 23. A) Histology of the bladder at the trigone. The muscularis propria reaches to an almost suburothelial
location. B) Muscularis propria muscle bundles at trigone become progressively smaller in caliber and are
more superficial as they almost reach the surface

74

Figure 24. Examples of the spectrum of tumors in a cystectomy (A), Whole mount section of a bladder with
a preoperative diagnosis of muscle invasive high-grade papillary carcinoma; B, Urothelial dysplasia; from
the specimen shown in A), a papillary neoplasm (C, The neoplasm grows exophytically into the lumen of the
urinary system with a papillary configuration; D, Low-grade papillary carcinoma; from the specimen shown
in C), and an inverted neoplasm (E, The neoplasm, even though a polypoid appearance, is characterized
by an epithelium that shows a non-infiltrative growth into the subepithelial connective tissue; F, Low-grade
inverted urothelial carcinoma; from the specimen shown in E).

75

b) Urothelial Dysplasia

of uniform terminology and common definitions for


papillary neoplasms; b) establishment of detailed
criteria of various preneoplastic conditions and
various grades of tumor; c) correlation with urine
cytology terminology, facilitating cyto-histologic
correlation and making it easier for urologists
to manage patients; d) creation of a category of
tumor that identifies a tumor with a negligible risk
of progression (PUNLMP), whereby patients avoid
the label of having cancer which has psychosocial
and financial implications, neither is a benign lesion
(papilloma) diagnosed in these patients, so they
may still be followed up closely; e) approximately
half the cases occurring in young patients (patients
less than 20 years of age) are now diagnosed as
papillary urothelial neoplasms of low malignant
potential instead of transitional cell carcinoma grade
1 obviating a carcinoma diagnosis in patients with
biologically indolent tumors and in keeping with their
molecular profile; f) identification of a distinct group of
patients (high-grade papillary urothelial carcinoma)
who would be ideal candidates for intravesical
therapy; g) identification of a larger group of patients
at high risk for progression for urologists to follow up
more closely; h) removal of ambiguity in diagnostic
categories in WHO 1973 system (e.g., TCC grade
I-II, TCC grade II-III); h) stratification of bladder
tumors into prognostically significant categories; and
i) emergence of molecular correlates and signatures
for high grade tumors and tumors at the low-grade
end of the spectrum which may help provide ancillary
grading tools, facilitate patient management, and
possibly guide in future refinements of the current
grading system.

Urothelial dysplasia is characterized by architectural


distortion and a variable degree of atypia. The
thickness is usually normal ( but may be attenuated
or hyperplastic) and cytological changes are often
restricted to the intermediate and basal cells.
The nuclei are irregularly enlarged with loss of
polarity. Nuclear outlies are irregular and chromatin
abnormalities, although subtle, are present.
Nucleoli are variable, absent but not conspicuously
present. Mitotic activity is scant, usually involving
only the basal and intermediate cell layers. Overall
features are those of a neoplastic atypia (clear cut
nuclear atypia) but which fall short of the criteria
for CIS; dysplasia is not further graded. Low-grade
intraurothelial lesion is a synonym. There is some
evidence, largely genetic, that dysplasia shares
some abnormalities with CIS and therefore likely
represents a precursor lesion. Few studies, most
dated, indicate a 5-19% risk of developing cancer
[72]. The diagnosis of de novo dysplasia (i.e. in a
patient without history of bladder neoplasia) should
not be made or should be made with great caution
as the vast majority of patients, in the limited studies,
do not progress to cancer (Level 3). While dysplasia
likely represents a marker of urothelial instability, the
diagnosis should not by itself invoke any therapy;
continued surveillance is recommended (Grade C).

c) Urothelial Carcinoma In Situ (CIS)


CIS is characterized by architectural disorder and
distinct high-grade nuclear atypia (definitional feature)
including marked irregularities of the nuclear outlines
and chromatin distribution, nuclear pleomorphism,
prominent nucleoli throughout (if nucleoli are present),
nucleolar pleomorphism, and atypical mitoses on the
surface. There is usually loss of nuclear polarity and
the cells show a high degree of atypia. Mitoses are
generally frequent and may be seen at any level of
the epithelium. There is a spectrum of nuclear and
architectural atypia. CIS defined by WHO (2004)
/ ISUP includes cases diagnosed previously as
severe dysplasia and even some cases previously
diagnosed as moderate dysplasia. The classification
system provides detailed criteria for pathologists, as
CIS may be both over- and under-diagnosed. The
development of invasion is seen in 20 to 30% of the
cases (Level 2) [73-78]. Urothelial carcinoma in situ
of the urethra may extend into underlying prostatic
glands (Figure 25).

2. Flat Lesions and Neoplasms


The 2004 WHO/ISUP classification of the flat lesions
includes flat hyperplasia, dysplasia and carcinoma in
situ. In addition, this classification also lists reactive
atypia, secondary to inflammation, and atypia of
uncertain significance.

a) Flat Urothelial Hyperplasia


Urothelial hyperplasia is characterized by markedly
thickened urothelium with an increase in the number
of cell layers, usually 10 or more, usually with
increased cellularity (increased cells per unit area).
The cells do not show cytological abnormalities,
although slight nuclear enlargement may be focally
present. There is cellular order and morphologic
evidence of maturation from base to surface
epithelium is evident. Mitoses are generally absent. It
has been described in association with inflammatory
disorders as well as adjacent to low-grade papillary
tumours [28]. Molecular analyses have shown that
this lesion may be clonally related to the papillary
tumors in patients with bladder cancer and suggest
a role in the pathogenesis of low-grade papillary
urothelial carcinoma (Level 3) [70,71].

3. Papillary (Exophytic) Neoplasms


The lesions and neoplasms of this group grow
exophytically into the lumen of the urinary system
with a papillary configuration. According to the 2004
WHO classification, this group includes urothelial
papilloma, papillary urothelial neoplasm of low
malignant potential (PUNLMP), low-grade papillary
carcinoma and high-grade papillary carcinoma.

76

grade 1 TCC and approximately 70% of grade 2


TCC in the old 1973 WHO system. WHO grade 1
neoplasms showing mild but distinct cytological
atypia and mitoses, are diagnosed in the 2004 WHO
system as low-grade papillary urothelial carcinomas.
At low magnification the cells are relatively uniform
but enlarged in size with frequent loss of cellular
polarity (loss of linear perpendicular orientation
to basement membrane). There is no significant
pleomorphism. The nuclei tend to be uniformly
enlarged with occasional to consitent irregularities
of nuclear contours. The chromatin is relatively fine
to mildly abnormal with small nucleoli. Mitoses may
be present but are few and remain basally located.
These tumors have a significantly higher recurrence
rate than for PUNLMP (Level 2). They also have
a significantly higher rate of stage progression
than PUNLMP but significantly lower than for highgrade papillary carcinoma (Level 2). A review of the
literature reveals a mean recurrence rate of 50% and
mean stage progression rate of 10%.

Figure 25. Extension of urethral urothelial carcinoma


as urothelial carcinoma in situ involving prostatic
glands.

a) Urothelial Papilloma
Urothelial papilloma is characterized by a few
fine papillary fronds lined by normal-appearing
urothelium. Tumors contain slender minimally
branching papillae and the superficial umbrella cells
are often prominent which may show vacuolization.
Urothelial papilloma shows a very low recurrence
rate and minimal to absence progression rate (Level
2) [79-81].

d) High-grade Papillary Carcinoma


High-grade papillary carcinoma is characterized by
cells with high-grade cytologic atypia (resembling
CIS in a flat lesion) on a fibrovascular core. Cases
exhibit a spectrum with some tumors (approximately
30%) considered grade 2 TCC and most cases
of TCC grade 3 in the old 1973 WHO system. On
low power magnification, the tumor often exhibits
complex papillary architecture and fusion. Individual
cells are haphazardly arranged within the epithelium
and have a generally discohesive nature. The tumor
may show extensive denudation. There is a range
of nuclear atypia, some have obvious pleomorphism
while some have monomorphic nuclei with rounding,
nucleomegaly and irregular clumped chromatin. The
chromatin is dense, irregularly distributed and often
clumped. Nucleoli may be single or multiple and
are often prominent. Mitoses are generally frequent
and may be seen at any level of the epithelium. It
is often associated with invasive disease at the
time of diagnosis [82-89]. These tumors not only
have a risk of invasion but also have a significant
risk of recurrence and progression; association
with multifocality and CIS is common (Level 2). The
overall progression rate (to invasive carcinoma)
ranges from 15% to 40%; more than half the cases
are prone to recurrences.

b) Papillary Urothelial Neoplasm of Low


Malignant Potential (PUNLMP)
PUNLMP is characterized by cytologically bland /
normal appearing cells with hyperplasia (increased
cells per unit area and increased cell layers) on
a fibrovascular core. PUNLMP largely, though
not completely, corresponds to grade 1 papillary
transitional cell carcinoma (TCC) of the old 1973
WHO system. The lesion consists mostly of delicate
papillae with little or no fusion. The covering
urothelium shows monotony with normal size and
preserved polarity characterized by upward streaming
of cells perpendicular to basement membrane. The
umbrella cell layer is often preserved and basal cells
may show palisading. Nuclei lack significant nuclear
hyperchromasia or pleomorphism. The chromatin
is fine and nucleoli are inconspicuous. Mitoses
are infrequent and basally located. This tumor has
a significantly lower rate of recurrence than either
low- or high-grade papillary carcinomas and a very
low rate of grade and stage progression (Level 2).
In a review of published studies, a mean tumor
recurrence rate to be 36% and stage progression
rate to be 3.7% [71].

4. Inverted (Endophytic) Neoplasms


Papillary tumors may be associated with a variable
degree of inverted growth patterns. Although
focal areas of inverted growth are not uncommon,
prominent or exclusive inverted growth is much
rarer and when encountered may pose problems
related to grading or assessment of invasion. The
neoplasms with inverted growth are basically

c) Low-grade Papillary Carcinoma


Low-grade papillary urothelial carcinoma shows
cells with low-grade cytologic atypia (resembling
urothelial dysplasia in a flat lesion) on a fibrovascular
core. Cases would span some tumors considered

77

characterized by an epithelium that shows a noninfiltrative growth into the subepithelial connective
tissue, even though, cystoscopically they might show
a polypoid or dome shaped appearance. Although a
formal grading system is not proposed specifically
for lesions with a predominant inverted growth, nor is
one universally used, we recommend that for tumors
with predominalty invereted growth patterns criteria
which essentially parallels the W.H.O. (2004) / ISUP
system for exophytic tumors and flat lesions be used
(Table 2).

Inverted urothelial papilloma consists of thin


anastomosing trabeculae of urothelial cells within
the subepithelial connective tissue and covered by
a normal or attenuated urothelium without cytologic
atypia [90-92]. Inverted papilloma is associated with a
low risk of recurrence (<5%) (Level 2). In comparison
with inverted urothelial papilloma, the architectural
features favoring a diagnosis of urothelial neoplasms
with inverted growth pattern include thick columns
with irregularity in their width and transition into more
solid areas. The characteristic orderly maturation,
spindling, and peripheral palisading seen in inverted
papilloma are generally absent or inconspicuous.

Although prospective formal studies with uniform


criteria or assessing the clinical significance of the
classification terminology are not yet performed,
standardized nomenclature proposed herein may
facilitate such much needed data (Level 4).

Inverted papillary neoplasms of low malignant


potential show thicker or irregular cords, solid or
nodular growth lacking cytologic atypia; when
cytologic atypia is present tumors are graded as lowgrade or high-grade based on the degree of atypia.
Reports of non-invasive urothelial carcinomas with
inverted growth pattern are found in the literature
using variable nomenclature [28,70,93,94] with two
patterns - inverted papilloma-like pattern), or broadfront pattern (Table 3). Minor degrees of inverted

The classification system includes urothelial


papilloma, inverted papillary urothelial neoplasm of
low malignant potential, inverted low-grade urothelial
carcinoma and inverted high-grade urothelial
carcinoma; the last category may be non-invasive or
invasive.
Table 2. Analogy in nomenclature for inverted lesions

Flat Lesions

Papillary Lesions*

Predominant or Exclusive Inverted


Lesions*
Normal
Papilloma
Inverted papilloma
Urothelial hyperplasia Papillary neoplasm of low malignant
Inverted papillary neoplasm of low
potential
malignant potential
Urothelial dysplasia
Papillary urothelial carcinoma, low grade, Inverted papillary urothelial carcinoma,
non-invasive
low grade, non-invasive
Urothelial CIS
Papillary urothelial carcinoma, high grade, Inverted papillary urothelial carcinoma,
non-invasive
high grade, non-invasive
_
Papillary urothelial carcinoma, high grade, Inverted papillary urothelial carcinoma,
invasive
high grade, invasive
From Epstein JI, Reuter VE, Amin MB. Biopsy Interpretation of Bladder, 2nd Ed.
*Lesions may be both exophytic and endophytic
Table 3. Morphologic, immunologic and molecular genetic features of inverted urothelial papilloma and
urothelial carcinoma with inverted pattern
Characteristic

Inverted urothelial papilloma

Surface

Smooth, domed shaped, usually intact


cytologically normal
Endophytic, expansive, sharply
delineated, anastomosing cords and
trabeculae
Orderly polarized cells, some with
spindling and palisading at the periphery.
No significant atypia, mitoses rare
Low p53 expression and Ki-67
proliferation index
Rare deletions at chromosome 9 or 17,
rare FGFR3 mutations, low rate of LOH**
Benign, rare recurrences*

Growth pattern

Cytologic features

Immunohistochemistry
Molecular analysis
Biological potential

*Rare recurrences related to incomplete excision


**Loss of heterozygosity

78

Urothelial carcinoma with inverted


growth
Usually exophytic papillary lesions present
Endophytic, lesional circumscription
variable
Variable, nuclear pleomorphism and atypia
present
Variable, usually high p53 and Ki-67
proliferation index
Frequent FGFR3 mutation, chromosome 9
and 17 deletions
Recurrences and progression may occur

pattern should not be acknowledged in the pathologic


diagnosis. Prominent of predominant patterns may
be diagnosed using terminology such as papillary
urothelial neoplasm of low malignant potential with
prominent inverted growth pattern, no evidence of
invasion or papillary urothelial carcinoma, lowgrade, non invasive, with prominent areas of inverted
growth pattern.

low-grade papillary carcinoma represented the


largest grade category by 2004 WHO/ISUP system
which comprised 36-74% of the papillary tumors,
and G2 (31-84%) the largest when 1973 WHO
system was applied. High-grade papillary carcinoma
(2004 system) generally had greater proportion
of tumors than G3 (1973 system). In 6 of 7 series,
high-grade papillary carcinoma constituted > 10%
of tumors, whereas, G3 was comparatively smaller
and constituted > 5% of tumors in only 1 out of the
7 studies. In most studies, PUNLMP generally had
lower number of tumors than G1 tumors. In summary,
studies have attempted to correlate the 1973 and the
2004 system with variable success as there is not
a direct correlation between the 2 systems due to
differing criteria[70,79,83-85,90,91].

High grade invereted tumors may be non invasive


or invasive. The diagnosis of invasion requires the
unquestionable presence within the lamina propria of
irregularly shaped nests or single cells that may have
evoked a desmoplastic or inflammatory response.
When a stromal response is absent, irregularity
of the contours of the invasive nests, architectural
complexity and recognition of single-cell invasion
are helpful (Table 4).

Since its introduction, several studies had compared


the impact to disease outcome of the 1998 ISUP/
WHO (2004 WHO) system to the 1973 WHO system
in same patient cohort of non-invasive papillary
urothelial neoplasm (Table 6). To eliminate the
influence of stage, studies that combined noninvasive with pT1 tumors or deep muscle-invasive
tumors were not included in this review. Some studies
have shown advantage for the 2004 WHO grading
system over the 1973 WHO grading system.

C. UPDATES ON GRADING OF
NON-INVASIVE AND INVASIVE
UROTHELIAL CARCINOMA OF THE
URINARY BLADDER
Since its introduction in 1998, the different
components of 2004 WHO (1998 ISUP/WHO)
grading system for urothelial neoplasms have been
the subject of several clinicopathological validation
studies, observer variability studies, and comparative
analysis with the older 1973 WHO grading system.
This review focuses on critical diagnostic pathologyrelated issues of the 2004 WHO/ISUP grading
system, with appropriate comparison with the 1973
WHO grading system based on published studies.

Cao et al [95] reviewed 172 pTa (out of 269 non


muscleinvasive tumors) with long follow-up (up to
10 years) and demonstrated significant differences
between low-grade papillary carcinoma and highgrade papillary carcinoma in recurrence-free survival
(p= 0.05, log-rank test) and progression free-survival
(p= 0.01, log-rank test). No PUNLMP was identified
in the study cohort. The 1973 WHO system in
contrast showed significant difference only in
progression free-survival (p=0.03, log-rank test) and
not in recurrence-free survival. Both WHO grading
systems were unable to predict the overall survival.

I. Grading Non-invasive Papillary Urothelial Neoplasms:


Tumor Distribution and Impact
to Outcome of WHO (2004) Versus the Prior 1973 WHO Grading System

Burger et al [96] reviewed 104 pTa tumors that


included 51 early onset (< 45 years old patients)
(median follow-up of 48 months) and 63 regular
onset (> 45 years old patients) (median follow-up
of 43 months) tumors. In early onset pTa tumors,
both WHO grading systems were not significantly
related to recurrence-free survival or to progression
to muscle invasive disease. In regular onset, both
2004 WHO system (p=0.021, log-rank test) and
1973 WHO system (p=0.035, log-rank test) were
predictive of recurrence-free survival, however only

Table 5 shows a comparative distribution of noninvasive papillary urothelial neoplasms according


to the 2004 WHO/ISUP and 1973 WHO grading
systems in same patient cohorts. In most studies,

Table 4. Urothelial carcinoma with inverted pattern. Criteria for invasion.

Features
Contours of neoplastic nests/
cords
Size and shape of nests
Inflammatory and desmoplastic
stroma

Non-invasive
Regular

Invasive
Irregular

Similar, rounded edges


Absent

Variable, irregular and jagged edges


Present

79

the 2004 WHO system was related to progression


(p=0.002).

Schned et al [98] reviewed 504 non-invasive tumors


(mean follow-up 7.2 years) and similarly showed a
gradient of progressive lower survival times from the
lowest to the highest grade tumors in both 2004 and
1973 WHO grading systems. Compared with survival
times for PUNLMP, the hazard ratio for low-grade
papillary carcinoma was 1.9 (95% CI 1.0-3.4) and for
high-grade papillary carcinoma was 3.0 (95% CI 1.56.0). For the 1973 WHO system, compared to G1
tumors, the hazard ratio for G2 was 1.8 (95% CI 1.13.1) and for G3 was 2.4 (95% CI 1.2-4.7). The study
however did not provide information on recurrence,
progression and disease-specific survival of these
cases.

Yin and Leong [97] reviewed 84 pTa tumors and


showed significant differences in recurrence within
36 months between PUNLMP (17% recurrence) and
low-grade papillary carcinoma (45% recurrence) and
between low-grade papillary carcinoma and highgrade papillary carcinoma (74% recurrence) (p<0.5).
In contrast, no significant difference in recurrence
was observed between the 1973 WHO grades
(urothelial papilloma, G1, G2 and G3 with recurrence
of 0%, 41%, 54% and 67%, respectively).
Some studies did not show a clear-cut advantage
in terms of impact to outcome for the 2004 WHO
system over the 1973 WHO grading system.

Samaratunga et al [99] investigated 134 patients with


pTa tumors and showed no statistical significance
in recurrence rates among the 1998 ISUP/WHO
grades. The only statistically significant difference in
the 1973 WHO system was that G3 had increased
recurrences per year compared with papilloma
(p=0.02), G1 (p=0.0001), and G2 (p=0.0001).
Separate analyses showed both 2004 WHO system
and 1973 WHO system independently predicted
progression (p=0.003 and p=0.002, respectively)
together with tumor size.

May et al [76] compared the prognostic implications


of both WHO grading systems in 200 non-invasive
tumors and with mean follow-up of 72 months
using consensus diagnosis of 4 genitourinary
pathologists. No PUNLMP was identified in the
series. By 2004 WHO system, low-grade papillary
carcinoma and high-grade papillary carcinoma had
5-year recurrence free survival of 69.1% and 46.1%,
respectively (p=0.007) and 5-year progression free
survival of 93.5% and 80.2%, respectively (p=0.084).
By 1973 WHO system, G1, G2 and G3 had 5-year
recurrence free survival of 64.5%, 68.1% and 11.1%,
respectively (p<0.001) and 5-year progression free
survival of 97.2%, 86.9% and 64.8%, respectively
(p=0.002).

Oosterhuis et al [100] reclassified 322 pTa tumors


to 1998 ISUP/WHO and log rank test for five
year progression free survival (p=0.06) and five
year recurrence free survival (p=0.13) were not
significantly different between the grade groups as
a whole. The only significant difference between

Table 5. Distribution of non-invasive papillary urothelial neoplasms according to 2004 WHO (1998 ISUP/
WHO) and 1973 WHO grading systems.*

Study

Total

Cao et al, 2010

172

May et al, 2010

200

Burger et al,
2007
Schned et al,
2007
Yin and Leong,
2004
Samaratunga et
al, 2002
Oosterhuis et
al, 2002
Range

114
504
84
134
322

2004 WHO (1998 ISUP/WHO) system


1973 WHO system
Papilloma PUNLMP
LG
HG
Papilloma
G1
G2 (%)
G3
(%)
(%)
PUC
PUC
(%)
(%)
(%)
(%)
(%)
0
65 (36)
117
44
121 (70) 7 (4)
(64)
(26)
1 (0.5)
149 50 (25)
82
109 (54) 9 (5)
(74)
(41)
6 (5)
77 (71) 26 (24)
58
46 (42)
5 (5)
(53)
179 (38)
214 73 (16)
295 154 (31)
55
(46)
(58)
(11)
3 (4)
32 (38) 46 (55) 3 (4)
0
12
53 (63)
19
(14)
(23)
3 (2)
29 (22) 73 (54) 29 (22)
7 (5)
42
79 (59)
6 (4)
(31)
18 (5)
116 (36)
141 45 (14)
2 (1)
31 (9)
286
1 (1)
(44)
(84)**
2-5%
0-38%
364-64%
0-5%
9-58% 31-84% 1-23%
74%

* Includes only studies that accounted the purely non-invasive tumors; studies combining non-invasive and invasive
tumors and/or without the use of the two grading systems were not included; ** Modified 1973 WHO 2a (214, 66%)
and 2b (72, 22%).

80

Table 6. Outcome of non-invasive papillary urothelial neoplasms by 2004 WHO (1998 WHO/ISUP) and 1973 WHO grading systems.
Study
May et al, 2010

Grade
2004 WHO

1973 WHO

Burger et al, 2007

2004 WHO

1973 WHO

2004 WHO

81

1973 WHO

Yin and Leong,


2004

1998 ISUP/WHO

1973 WHO

Samaratunga et
al, 2002

1998 ISUP/WHO

1973 WHO

PUNLMP
LG PUC
HG PUC
G1
G2
G3
PUNLMP (< 45 years)
LG PUC (< 45 years)
HG PUC (< 45 years)
G1 (< 45 years)
G2 (< 45 years)
G3 (< 45 years)
PUNLMP (> 45 years)
LG PUC (> 45 years)
HG PUC (> 45 years)
G1 (> 45 years)
G2 (> 45 years)
G3 (> 45 years)
PUNLMP
LG PUC
HG PUC
Papilloma
G1
G2
G3
Papilloma
PUNLMP
LG PUC
HG PUC
Papilloma
G1
G2
G3

Total
1
149
50
82
109
9
0
34
12
30
15
1
6
43
14
28
31
4
12
53
19
3
32
46
3
3
29
73
29
7
42
79
6

Recurrence
0 (0%)
55(37%)
29 (58%)
36 (44%)
40 (37%)
8 (89%)
2 (17)
24 (45)
14 (74)
0
13 (41)
25 (54)
2 (67)
0.0/year
0.025/year
0.045/year
0.047/year
0.02/year
0.027/year
0.059/year
0.3/year

p-value
<0.001

<0.001

<0.05

<0.05

NS

G3 vs. papilloma, 0.02


G3 vs. G1, 0.00001
G3 vs. G2, 0.00001

Progression
0
10 (67%)
8 (16%)
4 (5%)
11 (10%)
3 (30%)
2 (6%)
3 (25%)
1 (3%)
3 (20%)
1 (100%)
0 (0%)
0 (0%)
5 (36%)
2 (71%)
2 (65%)
1 (25%)
0%
8%
13%
51%
0%
11%
24%
60%

p-value
0.084

0.002

0.085
G1 vs. G2/3, 0.0105

0.002

G1 vs. G2/3, 0.310

Table 6. Outcome of non-invasive papillary urothelial neoplasms by 2004 WHO (1998 WHO/ISUP) and 1973 WHO grading systems. (continued)

PUNLMP vs. LG PUC, 0.0.36


HG PUC vs. PUNLMP, 0.04
HG PUC vs. LG PUC, 0.16
NS
NS

100% PFS
97% PFS
96% PFS
92% PFS
88 PFS
92 PFS
99 PFS
PUNLMP vs. LG PUC, 0.35
HG PUC vs. PUNLMP, 0.1
HG PUC vs. LG PUC, 0.42
86% RFS
68% RFS
66% RFS
57% RFS
80% RFS
68% RFS
55% RFS
18
116
141
45
31
214
72
1
1973 WHO

Papilloma
PUNLMP
LG PUC
HG PUC
G1
G2a
G2b
G3
1998 ISUP/WHO
Oosterhuis et al,
2002

* NS, not significant; RFS, recurrence free survival; PFS, progression free survival

82

two groups was for five-year progression free


survival between PUNLMP and high-grade papillary
carcinoma (p=0.04, log rank test). By 1973 WHO
system, log rank test for five year progression free
survival and five year recurrence free survival were
not significantly different between the grade groups.
The study by Hlmang et al [101] showed usefulness
of 1998 ISUP/WHO grading system, but also
highlighted an advantage for the 1973 WHO grading
system. 363 primary pTa tumors with follow-up of at
least 5 years were classified according to the 1998
ISUP/WHO grading system. There was recurrence
in 35% PUNLMP compared to 71% low-grade
papillary carcinoma and 73% of high-grade papillary
carcinoma (p<0.0001). Progression was 0%, 4% and
23% for PUNLMP, low-grade papillary carcinoma
and high-grade papillary carcinoma, respectively
(p<0.0001). However, when the 108 high-grade
papillary carcinoma where further subdivided by
1973 WHO system into G2 (95) and G3 (13), there
was a significant difference in progression of 20%
and 45%, respectively (p<0.0022).
In terms of tumor distribution, the grading categories
of the 2004 WHO system did not directly translate
from the 1973 WHO system categories (Level 2).
An advantage for the 2004 WHO system was it
allowed identification of a larger number of patients
with non-invasive papillary urothelial carcinoma that
were at higher risk of recurrence and progression,
as comparatively a higher proportion of tumors were
categorized as high-grade papillary carcinoma than
G3 of the 1973 WHO system (Level 2).
In terms of impact to disease outcome, grade overall
provided important prognostic information in most
of the published studies of non-invasive papillary
urothelial neoplasms, with the 2004 WHO system
demonstrating either comparable results or with
some advantages in predicting recurrence and/or
progression than the 1973 WHO system (Level 3).
However, there was evidence suggesting that highgrade papillary carcinoma is heterogeneous in terms
of outcome and subdivision into G2 and G3 by 1973
WHO system may provide additional prognostic
information (Level 3). This data may be helpful in
future refinements of the WHO grading system
to increase the predicitive power of high grade
tumors focusing on morphologic and/ or molecular
subcategorization. Accumulated evidences thus far
support for the continued application of the 2004
WHO system in grading non-invasive papillary
urothelial carcinoma (Grade B).

II. Inter- and Intra-Observer


Reproducibility Studies on
Grading
The percentages of the grading categories for both

WHO grading systems in non-invasive papillary


urothelial neoplasms had varied significantly in
previous studies (Table 5). The percentages of 2004
WHO PUNLMP, low-grade papillary carcinoma and
high-grade papillary carcinoma were 0-38%, 36-74%
and 4-64%, respectively. The percentages of 1973
WHO G1, G2 and G3 were 9-58%, 31-84% and
1-23%, respectively. Differences in patient cohort
geography may partly explain the broad differences,
but the influence of interpretation variability among
pathologists as well cannot be discounted.

Thus, the 2004 WHO system did not show better


intra or inter-observer reproducibility than the 1973
WHO system. The group suggested that the degree
of variability between pathologists could be quantified
using a mean grade, and such can be a potential tool
for quality assurance in pathology.
In the study by Gnl et al [104] of 258 papillary
urothelial carcinoma, the overall degree of
reproducibility between 2 pathologists was higher in
the 2004 WHO system ( 0.59) than the 1973 WHO
system ( 0.41), although both were still moderate
in agreement. PUNLMP showed the lowest degree
of agreement and if excluded from the analysis,
interobserver concordance of the 1998 WHO/ISUP
increased significantly ( 0.84).

Table 7 summarizes several published grading


observer variability studies in papillary urothelial
neoplasms, and includes some that compares the
2004 WHO system with the 1973 WHO grading
system.

In the study by Mamoon et al [105], the agreement


on 80 urothelial neoplasms between 2 pathologists
for the 2004 WHO system was excellent ( 0.91),
whereas agreement for 1973 WHO system was good
( 0.68). By 2004 WHO system, the kappa value for
low grade and high grade tumors showed excellent
agreement while the kappa value for papilloma and
PUNLMP showed fair to good agreement.

In the study by May et al [76], 200 papillary urothelial


carcinomas were reviewed independently by 4
genitourinary pathologists and graded according to
both 2004 WHO and 1973 WHO grading systems.
Owing to the rare PUNLMPs in the cohort, the WHO
2004 approached a two-tiered system (low and high
grades) and showed less interobserver variability
than the 1973 classification ( 0.30-0.52 vs. 0-0.37,
respectively). In comparing the power of both
classifications to separate indolent from aggressive
papillary carcinoma, striking pathologist-dependent
differences were observed. Only 2 of 4 pathologists
provided a grading under the 2004 WHO system that
was significantly prognostic regarding recurrence
and progression.

Campbell et al [106] showed moderate agreement in


grading 49 non-invasive papillary urothelial neoplasm
using WHO 2004 system whether by one ( 0.45) or
two ( 0.60) pathologists.
Yorukoglu et al [107] in a review of 30 papillary
urothelial neoplasms with 6 genitourinary pathologists
showed for both 1998 ISUP/WHO system and 1973
WHO system moderate interobserver agreement
( 0.56 and 0.48, respectively). For both grading
system the mean rate of agreement was lowest for
the lowest grades (PUNLMP and G1) and highest for
highest grades (high grade papillary carcinoma and
G3). There was substantial intraobserver agreement
for both 2004 WHO system and 1973 WHO system
( 0.67 and 0.66, respectively).

In the review by Miyamoto et al [102 ] of a single


institution cohort of 112 low-grade urothelial carcinomas, 8 of 55 cases (14.5%) originally diagnosed
as low grade by general surgical pathologists were
reclassified as high-grade urothelial carcinomas after review by 2 genitourinary pathologists. The main
reason for the upgrade was the presence of previously undetected focal high-grade areas amidst the
predominantly low-grade urothelial carcinoma.

In the study by Murphy et al [108], 247 biopsies


were reviewed by 3 pathologists and by the
1998 ISUP/WHO system, there were substantial
interobserver agreement for high grade papillary
carcinoma and CIS ( 0.75 to 0.82) compared with
slight to moderate agreement for PUNLMP and lowgrade papillary carcinoma ( 0.12 to 0.50). When
discriminating PUNLMP vs. low-grade papillary
carcinoma, discrepancies were 50% after educating
the pathologists compared to 39% before education.
In discriminating low-grade vs. high-grade papillary
carcinoma and PUNLMP vs. high-grade papillary
carcinoma, discrepancies were 15% and no
discrepancy after education.

Van Rhijn et al [103] compared the 2004 WHO system


and 1973 WHO system for observer variability and
prognosis in 173 non-muscle invasive bladder
cancers. For 2004 WHO system, the intraobserver
and interobserver agreements were 71% to 88% (
0.55-0.81) and 39% to 74% ( 0.17-0.58 in first round
and 0.14-0.41 in second review), respectively. If
PUNLMP and low-grade papillary carcinoma were
condensed, the intraobserver agreement increased
to 85-97% ( 0.68-0.91) and interobserver agreement
varied from 68% to 87% ( 0.68-0.91). For WHO
1973, intraobserver and interobserver agreements
(for 2 pathologists) were 80% to 81% ( 0.67-0.69)
and 39% to 64% ( 0.15-0.41). If G1 and G2 were
condensed to one grade, intraobserver agreement
increased to 91-95% ( 0.64-0.88) and interobserver
agreement ranged from 81% to 91% ( 0.47-0.67).

Most published observer variability studies showed


that both 2004 and 1973 WHO grading systems for
bladder urothelial neoplasms suffered from suboptimal
observer agreement among pathologists, with most

83

Table 7. Inter and intra observer variability studies in grading of papillary urothelial neoplasm.
Study
May et al, 2010

Cases

Grading system

Grades (No.)

200 papillary urothelial


neoplasm

2004 WHO

PUNLMP (1)
LG PUC (149)
HG PUC (50)
G1 (82)
G2 (109)
G3 (9)
LG PUC (13)
HG PUC (297)
G1 (0)
G2 (112)
G3 (198)
PUNLMP
LG PUC
HG PUC

1973 WHO

Otto et al, 2010

310 pT1 papillary


urothelial carcinoma

2004 WHO
1973 WHO

Van Rhijn, 2009

173 non-muscle
invasive papillary
urothelial neoplasm

2004 WHO

No. of
Interobserver agreement
pathologists
4
71.5% to 82.5% ( 0.30-52, fair
genitourinary
to moderate)
pathologists
38.0% to 73.0% ( 0-0.37 fair
to moderate)
3

90.3%

39% to 74% (first round


0.17-0.58 and second round
0.14-0.41)

Gnl et al, 2007

258 papillary urothelial


neoplasm

2004 WHO

1973 WHO

G1 (68)
G2 (79)
G3 (26)
PUNLMP
LG PUC
HG PUC
G1
G2
G3

Intraobserver
agreement
-

83..2%

71% to 88% ( 0.550.81)

39% to 64% ( 0.15-0.41)

4 (2
pathologists),
at least 6
months
interval
2 (2
pathologists),
at least 6
months
interval
4

75% ( 0.59, moderate)

84
1973 WHO

No. of times
reviewed
1

62% ( 0.41, moderate)

80% to 81% ( 0.670.69)

Table 7. Inter and intra observer variability studies in grading of papillary urothelial neoplasm.
Mamoon et al,
2006

100 papillary urothelial


neoplasm

2004 WHO

1973 WHO

Overall
Papilloma
PUNLMP
LG PUC
HG PUC
Overall
Papilloma
G1
G2
G3
PUNLMP (12)
LG PUC (28)
HG PUC (9)

6
genitourinary
pathologists

49 non-invasive
papillary urothelial
neoplasm

1998 WHO/ISUP

Yorukoglu et al,
2003

30 non-invasive
papillary urothelial
neoplasm

1998 WHO/ISUP

Overall

1973 WHO

PUNLMP (10)
LG PUC (10)
HG PUC (10)
Overall

85

Campbell et al,
2004

Murphy et al,
2002

247 biopsies

1998 WHO/ISUP

G1
G2
G3
PUNLMP
LG PUC
HG PUC
In situ

95% ( 0.91, excellent)


0.67
0.75
0.92
0.96
80% ( 0.68, good)
0.25
0.79
0.82
0.82
52% ( 0.45, moderate) with 1
pathologist
71% ( 0.60, moderate) with 2
pathologists
70.9% ( 0.56, moderate)

2, at least 4
weeks interval

78.3% ( 0.67,
substantial)
77.8% ( 0.66,
substantial)
-

48%
72.7%
92%
65.1% ( 0.48, moderate)
56%
67.3%
72%
0.12 - 0.50, slight to
moderate
0.7 - 0.82, substantial

1 (divided into
learning and
study sets)

studies showed only moderate agreement (Level 2).


When compared, the 2004 WHO system showed
relatively better reproducibility than the 1973 WHO
system (Level 3). Among the WHO 2004 grades,
PUNLMP and papilloma had the lowest interobserver
reproducibility and distinction of PUNLMP from
low-grade papillary carcinoma appeared to be the
most difficult. Condensing PUNLMP and low-grade
papillary carcinoma improved grading reproducibility
of the 2004 WHO system. Evidence showed that
education of pathologists might help improve
interobserver reproducibility of PUNLMP versus lowgrade papillary carcinoma. For high-grade papillary
carcinoma, a reason for interobserver variability was
tumor heterogeneity wherein a focus of high grade
was not accounted for amidst a predominantly lowgrade papillary tumor.

secondary (p=0.001) and combined (p=0.0001) were


all significant in predicting progression. Significant
difference in progression free survival between
patients with a combined score of 5 (low grade +
high grade) and those with a score of 6 (pure high
grade papillary carcinoma) (p=0.02) was observed.
Of note among mixed tumors with predominant lowgrade papillary carcinoma, more minor component
of high-grade papillary carcinoma (21%, score of 5)
than PUNLMP (6%, score of 3) was observed, the
former by conventional WHO 2004 grading will be
designated as high-grade papillary carcinoma.
Billis et al [109] applied a similar approach to Cheng
et al [114] in 293 bladder tumors but instead used
the 1999 WHO grading system and correlated
with stage. The study was able to stratify G2 into
subgroups (mixed [G1 + G2] and pure [G2+G2]),
which were statistically different when considering
stage. In G3, there was also a trend for statistical
difference between mixed (G2 + G3) and pure
(G3 + G3) tumors. The study however provided no
information on disease outcome.

The few studies that reviewed intraobserver


reproducibility among pathologists showed no
differences between the 2004 and 1973 grading
systems with both showed moderate to substantial
agreement (Level 3).

Bircan et al [110] applied a similar approach to


Cheng et al [111] and Billis et al [109] in 87 bladder
carcinomas and used 1998 ISUP/WHO, 1973 WHO
and 1999 WHO grading systems and correlated
with stage. Mixed histology (odd number scores)
constituted 18%, 29% and 33% of tumors by 1998
ISUP/WHO, 1973 WHO and 1999 WHO grading
systems, respectively. In the 1998 ISUP/WHO
system, there was significant difference between
low-grade and high-grade carcinomas (p=0.000) and
between mixed low-grade and high-grade (p=0.011)
when correlated to stage. The 1998 ISUP/WHO
system positively correlated with stage, but the 1998
combined scoring did not. The 1973 WHO system
positively correlated with stage, and there was a
weak association between the 1973 combined
scoring and stage. The study also did not provide
information on disease outcome.

III. Grading Papillary


Urothelial Neoplasms with
Histologic Heterogeneity
Grade represents a morphological spectrum, and
variation in the degree and distribution of atypia within
one tumor is well-recognized by pathologists (Figure
26). Admixture of at least two different grades in a
papillary urothelial neoplasms is not uncommonly
encountered, and is reported in 3% to 43% of
tumors [76,109-112]. This histologic heterogeneity
contributed to the lumping of grades (e.g. G1-G2,
G2-G3 or G2-G3) in the older grading systems, and
avoidance of these non-definitive lumped grades
was one impetus for modifications toward a more
defined grading system. The 1998 ISUP consensus,
cognizant of urothelial tumors with variable histology,
suggested that grade should be reported according
to the highest grade present in heterogenous tumors,
but stressed the need of studies to determine how
significant a minor component must be in order to
have an impact on overall prognosis [113].
Cheng et al [111] took into account the heterogeneity
in papillary urothelial neoplasms and proposed a
modified grading approach by combining the most and
second most common 1998 ISUP/WHO grades with
corresponding number designations (e.g. PUNLMP =
1, etc.) to come-up with a score, in a similar manner
to Gleason scoring. The volume of a grade should
be >5% to be considered for grading. The combined
scores resulted in a scale of grades 2 to 6, wherein
the odd numbers 3 and 5 representing mixed grades
accounted for 32% of tumors. Histologic grade based
on the worst grade (p=0.0009), primary (p=0.0004),

Figure 26. Grade heterogeneity in papillary urothelial


carcinoma. Low (right) and high grade (left) areas

86

Krger et al [112] examined the prognostic


significance of grade taking into account the presence
tumor heterogeneity in 151 muscle-invasive bladder
carcinomas. Using the 1998 ISUP/WHO grading
system, 43% of tumors had mixed low-grade and
high-grade papillary carcinoma components, which
resulted in a three-tiered grading of low-grade, mixed
and high-grade papillary urothelial carcinomas. By
Kaplan-Meier analysis, both the 1998 ISUP/WHO
grading system and the three-tiered grading failed to
reach the level of statistical significance. However,
when low-grade were combined with mixed versus
high-grade, this approach allowed stratification of
patients showing a significant difference in diseaserelated survival (p=0.0404). In multivariate analysis
however, this two-tiered grading approach did not
reach the level of significance.

(2004) / ISUP grade to manage patients and to determine intervention. Sharing such difficult cases
in a consensus manner either with other pathology
colleagues or through quality assurance meeting is
encouraged (Grade B). While currently there are no
reliable or acceptable immunohistochemical or molecular markers to assist in grading of borderline lesions, there are promising markers under investigation that may complement histological grading in the
future, especially for difficult cases.
From the few studies available, there is evidence to
suggest that pure high-grade papillary carcinoma has
a different biologic behavior or has higher disease
progression than mixed high-grade and low-grade
tumors (Level 3). Additional studies with long-term
follow-up are needed to determine the prognostic
impact and define the allowable extent or percentage
cut-off of high-grade focus in heterogeneous papillary
urothelial neoplasm. More studies are needed to
determine whether the broader scale summation
grading provides advantage in outcome information
and observer agreement than the conventional WHO
2004 grading system.

May et al [76] identified 3% in 269 non-muscle


invasive tumors that contained elements of both lowgrade papillary carcinoma and high-grade papillary
carcinoma and were able to classify these tumors
into low-grade or high-grade grades by separating
high-grade papillary carcinoma based on 5% cut-off
for high grade areas.

IV. Grading of Invasive


Urothelial Carcinomas (pT1
and pT2+ tumors)

Grade heterogeneity is not uncommonly encountered


in papillary urothelial neoplasms. The 2004 WHO
(1998 ISUP/WHO) system recommends grading of
heterogeneous tumor to be based on the highest
grade present in a tumor (Grade C). There is no
current widely acceptable definition or criteria to
provide quantitative estimate of size of smallest focus
required to upgrade a lesion. While arbitrary criteria
of ignoring less than 5% have been proposed, the
prognostic impact of this criterion in unknown (Level
4). Studies are needed to establish quantitative/
semi-quantitative criteria that need to be present
to alter assignment of grade in tumors with grade
heterogeneity.

Unlike non-invasive papillary urothelial neoplasms


(pTa), the role of histologic grade in pT1 and higher
stage tumors has been suggested to be of only
relative importance. The 1998 ISUP consensus
proposed that invasive tumors should be graded
as low or high grade similar to the scheme used
for grading non-invasive lesions [113]. Herein, we
review the impact of 2004 WHO (1998 ISUP/WHO)
system and 1973 WHO system in grading pT1
tumors. Distribution of pT1 tumors from previous
studies according to the 2004 WHO and 1973 WHO
system is shown in Table 8. In this summary, studies
combining pT1 with pTa tumors are not included, as
it is known that grading has impact on non-invasive
papillary tumors.

For high-grade papillary carcinoma, a reason for


interobserver reproducibility is tumor heterogeneity wherein a focus of high grade is not accounted
amidst a predominantly low-grade carcinoma. Some
authors use the term high-grade urothelial carcinoma arising in a background of low-grade urothelial
carcinoma when unequivocal areas of high-grade
carcinoma are seen. While the distinction between
PUNLMP and low-grade papillary carcinoma is not
that critical in an individual patient, distinction of low
grade from high-grade carcinoma is very important.
When assigning the tumor grade in tumors with
borderline grade histology, other tumor parameters
such as multifocality, previous grade of the tumor,
size of lesions, frequency of recurrence, presence
of concurrent CIS, positive cytology may be factored
in the grading; and the presence of a few or many
of these parameters may influence upgrading
to a high grade lesion (Grade C). Urologists often
use these tumor parameters in addition to W.H.O.

In the study of Cao et al [95], 41 of 42 (98%) pT1


tumors were classified as high grade based on the
2004 WHO system. By 1973 WHO system, these
high grade tumors were classified into 27 (64%) G2
and 15 (36%) G3. No significant differences were
noted in recurrence-free survival, progression-free
survival, or overall survival between G2 and G3
pT1 tumors. Analysis cannot be performed with the
2004 WHO system because of the few numbers of
invasive low-grade tumors.
In the study of Otto et al [115], 310 pT1 tumors were
mostly graded as high grade (96%) and only few
were low grade (4%) by 2004 WHO system. By 1973
WHO system, these invasive tumors were classified

87

Table 8. Distribution of pT1 urothelial carcinoma according to 2004 WHO (1998 ISUP/WHO) and 1973 WHO
grading systems.
Study

Total

2004 WHO system


LG PUC (%) HG PUC (%)
Cao et al, 2010
42
1 (2)
41 (98)
Otto et al, 2010
310
13 (4)
297 (96)
Van Rhijn et al, 2009
164
37 (23)
127 (77)
Kamel et al, 2006
105 high grade tumors
105

into G2 (36%) and G3 (64%) with no G1 tumors


identified. The 10-year recurrencefree survival and
overall survival did not differ significantly between
high grade and low-grade tumors (2004 WHO) or
between G2 and G3 tumors (1973 WHO), but the
cancer-specific survival between pT1G2 and pT1G3
by 1973 WHO system was shown to be highly
significant (p<0.001). This is in contrast to the low
grade versus high-grade tumors by 2004 WHO
system that did not significantly differ in cancerspecific survival.

G1 (%)
0
0
0
-

1973 WHO system


G2 (%)
G3 (%)
25 (64)
15 (36)
112 (36%)
198 (64)
74 (45)
90 (55)
61(58)
44 (42)

high-grade G3) (Level 3); such grade assignments


are difficult in routine practice, and criteria for grading
in invasive settings or impact on management based
on grade are not well defined.
It is recommended therefore that when tumors
are invasive, irrespective of depth, they should be
graded as high grade (Grade C). This rule is also
applicable to the general surgical and urologic
pathology community as deceptively bland variants
such as the nested or small tubular variants that
histologically appear low-grade, tend to behave like
invasive high-grade tumors of similar stage (Grade
C). If for institutional disease management teams or
clinical trials (e.g. for some institutions and protocols
in Europe), assigning grade G2 and G3 is important,
the practice should continue based on respective
protocols or institutional guidelines.

In the study by van Rhijn et al [103] of 164 pT1 tumors,


the 1973 WHO system (p=0.032) and not 2004 WHO
system was significant in univariate analyses for
tumor progression. But both WHO grading systems
were not significant in the multivariate analysis, in
contrast to the reviewed stage and presence of CIS.
Kamel et al [116] reviewed 105 pT1 high-grade
tumors by 2004 WHO system with median follow-up
of 4 years and divided these tumors into G2 (61) and
G3 (44) by 1973 WHO grading system. In terms of
actuarial disease specific survival, 20 of 44 (45%)
patients with high-grade/G2 tumors were alive versus
22 of 61 (36%) patients with high-grade/G3 tumors
(p=0.04). In terms of tumor progression, 13 of 44
(30%) of high-grade/G3 tumors progressed versus
12 of 61 (20%) high-grade/G2 tumors (p=0.02).
There was a trend for high-grade/G3 tumors to have
worse progression free survival rates than high
grade/G2 tumors after controlling treatment modality
chosen, and was not statistically significant owing
to the smaller numbers. In multivariate analysis
high grade/G3 was a significant predictor of tumor
progression (p=0.05) and marginally non-significant
predictor of poor patient survival (p=0.056).

D. Histologic Types of
Bladder Cancer Including
Variants of Urothelial
Carcinoma
Bladder carcinoma may exhibit a diverse array of
morphologies and may mimic many non-epithelial
and non-bladder tumorigenic processes. Greater
than 95% of bladder carcinoma consists of urothelial
carcinoma (formerly, transitional cell carcinoma)
and numerous variants thereof. Other major
subclassifications include squamous cell carcinoma,
adenocarcinoma and neuroendocrine carcinoma.
Although virtually all bladder carcinomas arise from
the associated urothelium, the pathogenesis of most
forms of carcinoma remains unknown with the
exception that tobacco smoking and male gender
play a signficant role.

Both W.H.O. (2004) / ISUP and the older 1973


WHO grading systems when applied to pT1 tumors
are confronted by the fact that the vast majority of
invasive tumors are high-grade (high grade by the
W.H.O. (2004) / ISUP system or G2 and G3 by 1973
WHO system), with only few tumors classified as lowgrade. Because of the predominance of high-grade
tumors, no study has shown the value of W.H.O.
(2004) / ISUP system in pure pT1 tumors (without
including pTa tumors). The older WHO (1973) system
has shown the ability in some studies to provide
grade dichotomy (or divisions) in the invasive pT1
tumors (i.e. G2 versus G3 or high-grade G2 versus

Of the less common major subdivision of bladder


carcinoma, squamous cell carcinoma occurs in
<5% of bladder cancer patients in the Western
hemisphere. Higher rates (although declining) are
seen in parts of the Middle East and Africa, where
infection with Schistosomal species plays a causative
role. Adenocarcinoma affects approximately 2% of
bladder cancer patients and may arise either from
the bladder urothelium or the urachus. Clear cell
adenocarcinoma a distinct form of adenocarcinoma
at this site may arise in a background of

88

endometriosis and resemble its counterparts in


the female genital tract. Finally, neuroendocrine
carcinoma may arise in pure form or be associated
with other forms of bladder carcinoma.

c)

Clinical: The presenting symptoms are


nonspecific and typical for urothelial neoplasms
in general with hematuria as the most common
manifestation[125-131].

Level I evidence is limited for most bladder carcinoma


variants, given their relative rarity. Especially
aggressive forms include micropapillary urothelial
carcinoma, sarcomatoid carcinoma, undifferentiated
carcinoma NOS, signet ring adenocarcinoma and
small cell carcinoma, among others. Variants that are
easily missed due to challenging histopathological
features may result in delayed diagnosis and
treatment; a sampling of such variants includes
nested urothelial carcinoma, large nested variant of
urothelial carcinoma, urothelial carcinoma with small
tubules and plasmacytoid urothelial carcinoma.

d) Gross Appearance: No information available.


e) Microscopic Features: Approximately 10%

of urothelial carcinomas contain foci of glandular


(Figure 27) and approximately 60% show variable
amounts of squamous (Figure 28) differentiation.
This may not reflect the actual frequency as this information is not consistently or accurately recorded
by pathologists. Glandular differentiation is usually in
the form of small tubular formations in conventional
urothelial carcinoma or as a histology similar to enteric adenocarcinoma (Figure 27). Rarely, a coexistent signet ring cell or mucinous component may be
present [117-124].

In general, treatment for muscle-invasive disease


involves complete surgical resection (generally radical
cystectomy) and lymph node dissection, regardless
of morphological subtype. Neoadjuvant or adjuvant
chemotherapy may have a role although treatment
regimens may vary for some of the histologic types or
between institutions. Earlier interventions have been
recently described to improve survival and outcomes
for patients with micropapillary urothelial carcinoma.
As our knowledge of these variants expands, it is
likely that more tailored, personalized therapies
may evolve depending on pathological diagnosis.

To designate squamous differentiation, one must


see clear-cut evidence of squamous production
(intracellular keratin, intercellular bridges or keratin
pearls), and the degree of squamous differentiation when present - usually parallels the grade of urothelial
carcinoma (Figure 28). Therefore, the presence of
squamous or glandular differentiation in a poorly
differentiated neoplasm, particularly at metastatic
sites and in the context of a bladder primary, should
suggest the possibility of urothelial differentiation.
To designate a bladder tumor as squamous cell or
adenocarcinoma, a pure or almost pure histology
of squamous cell carcinoma or adenocarcinoma is
required. From a practical viewpoint, we do make note
of prominent squamous or glandular differentiation
in urothelial carcinoma using diagnostic terminology
such as invasive urothelial carcinoma with prominent
squamous (40%) and glandular (25%) differentiation
[117-124]. In the case of metastatic tumors from the
bladder, knowledge of the presence and percentage
of divergent differentiation may be useful to facilitate
comparison with the bladder primary. The diagnosis
of adenocarcinoma is reserved for pure tumors
[117-124].

I. UROTHELIAL CARCINOMA
VARIANTS
1. Urothelial Carcinoma with
Divergent Differentiation
a) Definition: Urothelial carcinoma has propensity

for divergent differentiation, with the most common


being squamous differentiation followed by glandular
differentiation [117-124]. Virtually the whole spectrum
of bladder cancer variants may be seen in variable
proportions accompanying otherwise conventional
urothelial carcinomas, mostly in cases of high-grade
and high-stage disease. The clinical outcome of
some of these variants differs from typical urothelial
carcinoma, therefore recognition of these variants is
important [125-131]. When divergent differentiation
is seen together with conventional urothelial carcinoma, the pathologist should use the terminology
of urothelial carcinoma with ____ differentiation,
inserting the type of differentiation observed.

f) Differential Diagnosis: The main differential

diagnostic considerations include squamous cell


carcinoma and adenocarcinoma. In the urinary
bladder, the term squamous cell carcinoma should
be reserved for tumors with an exclusive squamous
cell component. Basaloid or clear cell-type squamous
differentiation can be observed rarely, and the
pathologist should be aware of this before making
a diagnosis of urothelial carcinoma with a mixed
squamous cell component [117-124]. The diagnosis
of adenocarcinoma of the bladder is reserved for
pure tumors.

b) Incidence:The actual frequency varies, as


this information is not consistently recorded by
pathologists. Some form of divergent differentiation
may be seen in 7% to 27% depending on the type
of specimen (cystectomy vs. transurethral resection)
[125-131].

g) Ancillary Diagnostic Tests: A recent study


found 74% of urothelial carcinomas with squamous
differentiation showed expression of urothelial and
squamous associated markers (S100P 83%, GATA3

89

35%, uroplakin III 13%, CK14 87% and desmoglein-3


70%)[132]; although reactivity for individual markers
within some tumors did not always correspond with
morphologic differentiation. Of the remaining 26%, 4
cases showed an overall squamous immunoprofile
while 2 cases showed an urothelial immunoprofile
[132]. This particular study showed that a panel of
five antibodies reliably distinguishes carcinomas
of the urothelial tract with squamous and urothelial
differentiation suggesting potential utility due to
management implications. Immunostaining in tumors
with both urothelial and squamous differentiation
showed a mixed profile. The expression of MUC5ACapomucin may be useful as immunohistochemical
marker of glandular differentiation in urothelial
tumors [133].

Figure 27. Urothelial carcinoma with glandular


differentiation.

h) Prognosis: The prognostic significance of squa-

mous or glandular differentiation is unclear, although


some studies have suggested an adverse outcome
[117-124]. This difference in prognosis may not be
apparent when corrected for stage of the disease.
Some studies have shown poor response to chemotherapy or radiation therapy, but it is unclear whether
the poor response is a reflection of histology or the
advanced stage of the disease [125-131]. Some
studies have suggested that these variants may be
more resistant to chemotherapy or radiation therapy
than pure urothelial carcinoma, but this has not been
confirmed [125-131]. A recent meta-analysis of reported data sugested that all of the variant histologies portend a worse prognosis than pure urothelial
carcinoma [126]. Classic studies suggest that squamous differentiation is associated with high-grade
disease and advanced stage and predicts a poor response to chemotherapy, radiotherapy, and surgery
[125-131]. In the context of radical cystectomy, a
recent series determined that squamous differentiation within urothelial carcinoma was an independent
predictor of local recurrence on multivariate analysis
[125]. Also, there is no indication that radiotherapy
would offer any benefit to mixed urothelial carcinoma
with squamous differentiation [125-131]. The presence of mixed histologic features at transurethral
resection could indicate locally aggressive disease.
Of 91 patients with metastatic carcinoma, cancer
progressed despite intense chemotherapy in 83%
with mixed adenocarcinoma and 46% with mixed
squamous cell carcinoma, whereas it progressed in
<30% of patients with pure urothelial carcinoma[125131].

Figure 28. Urothelial carcinoma with squamous


differentiation.

2. M
 icropapillary Variant of
Urothelial Carcinoma
a) Definition: Micropapillary carcinoma (MPC) of the

urinary tract is a special type of urothelial carcinoma,


which has papillary structures reminiscent of those
seen in ovarian papillary serous neoplasms and
typically lacks central fibrovascular cores seen within
usual papillary urothelial carcinoma [134]. Most cases
of MPC have been reported in the urinary bladder,
and MPCs of the upper urinary tract including renal
pelvis, ureter and ureteropelvic junction have rarely
been reported [135,136].
The majority of tumors present with advanced
stages and are muscle invasive (> T2) at the time
of presentation. Histologically, the MP component
is encountered in the (a) non-invasive surface
component, (b) invasive component and (c) in
metastasis. The MP pattern may be pure, but is
frequently associated with a high grade conventional
urothelial carcinoma. There is no specified criterion
required to designate a case as MPC, but most
series in the literature have included cases even
with <10% to almost pure MP histology [137,138].
The percentage of MP component has been shown

The presence of squamous or glandular differentiation


in locally advanced urothelial carcinoma of the
bladder does not confer resistance to MVAC and in
fact may be an indication for the use of neoadjuvant
chemotherapy before radical cystectomy [131].
Low-grade urothelial carcinoma with focal squamous
differentiation has a higher recurrence rate [123].

90

to be a significant adverse prognostic factor with the


more MPC component, the worse prognosis[139].
However, since any amount of MPC impacts the
prognosis, no percent limitation is required for the
diagnosis of MPC, but the percentage of the MPC
component should be reported.

invasive component and in all metastatic sites, the


tumor cells are arranged in small tight nests or balls
(Figure 29). The tumor cells in the invasive and
metastatic components are aggregated in lacunae
or stromal retraction spaces, which mimic vascular
invasion. This feature is characteristic of the invasive
MPC. The spaces may be lined focally by flattened
spindled cells or may be devoid of any lining. In most
instances, there is no host response to the tumor
cells that merely seem to occupy hollow spaces
at various random intervals within the tumor. This
pattern of lacunae containing neoplastic cells is also
seen in the metastatic sites. Multiple small cell nests
in the same lacunar spaces are characteristically and
frequently seen. MPC always shows a high nuclear
grade (high grade by the WHO/ISUP classification)
with peripherally oriented nuclei and inverted pattern,
although some areas within a neoplasm may parallel
low-grade urothelial carcinoma. Lymphovascular
invasion (LVI) is present in most of invasive MPCs.
A high frequency has been confirmed by Factor
VIIIR-Ag and Ulex europaeus agglutinin I lectin[1]
as well as CD 31, CD34 and D2-40 [143]. However,
awareness that lacunae of MPC may mimic LVI is
important so as not to overdiagnose the presence of
this feature as LVI . Psammoma bodies, a feature
of ovarian papillary serous neoplasia, are either not
present or exquisitely rare.

b) Incidence: Although it is difficult to estimate

the exact number of reported cases of MPC, at


least 500 cases of this variant have been reported
since the first description of a series of cases in
the urinary bladder in 1994 by Amin et al [134].
This rare histological variant comprises 0.68.2%
of urothelial carcinomas and shows a definite male
predominance (male to female ratio, 5:110:1), which
is higher than in conventional urothelial carcinoma
(3:1) [134,137,138,140]. As mentioned above, since
the amount of MPC component necessary to make
the diagnosis has not been specified, the reported
incidence has varied depending on the amount of
MPC component deemed necessary for diagnosis.
In early studies, the reported incidence of MPC was
0.61.0% of UC. Tumors in these series had at least
10% or 20% with most cases displaying greater
than 50% of MPC. In a recent series, the incidence
was found to be 6-8.2% [140]. In the largest series
reported to date, urothelial carcinoma with any
amount of a MP component was considered as
MPC [137,138]. It is likely that the recently increased
incidence of MPC is due to reporting cases with any
MPC as well as the increased awareness of MPC
by pathologists. Another factor for the different
incidence may be the diagnostic criteria of MPC. A
recent consensus opinion diagnosis result showed
only moderate overall diagnostic reproducibility (k:
0.54). Depending on pathologists, MPC may be
overdiagnosed or underdiagnosed [140].

Cytologic features of MPC include the presence


of singly scattered tumor cells with high nuclear to
cytoplasmic ratio and pleomorphic nuclei, clustered
cells devoid of fibrovascular core (micropapillae),
3-dimensional
cell
aggregates,
cytoplasmic
vacuoles, and micropapillae exhibiting some features
of low-grade urothelial neoplasms[144]. In addition,
urothelial carcinoma in situ is demonstrable in >1/3 of
the cases and concurrent glandular or villoglandular
differentiation is known to occur; rare cases of
mixed trophoblastic differentiation, carcinosarcoma
(sarcomatoid carcinoma), pleomorphic giant cell
carcinoma, lipid cell variant or plasmacytoid UC
have also been reported [145-151].

c) Gross Appearance: There are no specific

gross features to distinguish from other variants


of urothelial carcinoma. MPC shows both surface
papillary and deep invasive component. MPC may
be grossly papillary, polypoid, nodular, ulcerative
and or solid infiltrative. MPC is usually solitary,
but may be multifocal. The size of the tumor is
variable, ranging from a few millimeters to about 10
cm in greatest dimensions. The appearance of the
underlying bladder wall significantly depends on the
extent of invasion, which is commonly extensive in
MPC [141].
The uninvolved mucosa is usually normal, but
sometimes erythematous which may represent
microscopic areas of CIS. No currently well defined
imaging techniques can reliably diagnose MPC
[142].

d) Microscopic Features: The MPC histology

has two distinct patterns; on the surface, it forms


slender, delicate filiform processes with no or rarely
with a fibrovascular core. When cut in cross-sections,
these papillae appear as glomeruloid bodies. In the

Figure 29. Micropapillary variant of urothelial


carcinoma.

91

e) Differential Diagnosis: The main differential

and 4) micropapillae with elongated and slender


nests with average width <4.5 nuclei. Whereas nonclassic MPC with retraction usually showed large
to medium sized nests with >4.5 nuclei and nest
anastomosis and confluence or branching. This
study demonstrated that the overall reproducibility
for the diagnosis of MPC in invasive UCs showing
stromal retraction was only moderate. The diagnosis
of MPC was more reproducible when very restrictive
criteria were applied, but it is not known whether
the use of this diagnostic threshold correlates best
with the clinical outcome. These findings clearly
suggest that more studies are needed to identify the
individual pathological features that might potentially
correlate with an aggressive outcome and lack of
response to intravesical therapy. Finally, caution
should be exercised when considering the use of a
reported MPC diagnosis as the sole determinant in
selecting therapeutic options for individual patients
with invasive urothelial carcinoma.

includes: 1) distinguishing MPC of the bladder from


that of the different sites; and 2) how to separate
classic MPC from conventional UC with retraction
artifact [140,152]. In women, a metastatic papillary
serous carcinoma of the ovary or peritoneal primary
is the main differential diagnosis from MPC of
the bladder, especially if the tumor is originally
discovered in the peritoneum, abdominal lymph
nodes and mesentery or as a carcinoma of unknown
origin. Clinical/radiographic correlation is usually
required, but the possibility of a bladder primary may
be suggested if there is no obvious primary tumor at
another anatomic site. Identification of an admixed
urothelial carcinoma of more typical morphology or
immunohistochemical support (CK 7, CK 20 and
uroplakin III positivity) would be helpful. In a bladder
tumor, pure MP histology may raise concern for a
primary adenocarcinoma of the bladder; however,
the MP architecture is due to neoplastic urothelial
cells in a MP configuration and not due to true
glandular differentiation as also supported by
immunohistochemistry. In primary adenocarcinoma
of the bladder, there is a greater variability of gland
shape and size and of the range of differentiation in
contrast to the typically uniform appearance of MP
component of UC. Carcinomas with MP histology
have also been reported in the lung, breast,
pancreas, colon, stomach and salivary glands. The
best markers to identify urothelial MPC are uroplakin
and CK20, whereas p63, high molecular weight
cytokeratin, and thrombomodulin are less sensitive
and specific. Lung MPC is uniformly TTF-1 positive.
Breast MPC is ER and mammaglobin positive,
while ovarian MPC is ER positive and PAX8/WT-1
positive. No reliable markers for pancreas, stomach
and salivary glands have been reported, but CK20
and CDX2 positivity may be useful to make a
diagnosis of colon primary. In the metastatic setting,
or when MPC occurs without an associated in situ
or conventional carcinoma component for the certain
organ, staining for uroplakin, CK20, TTF-1, ER and
WT-1, and/or PAX8, and mammaglobin is the best
panel for accurately classifying the likely primary site
of MPC [152].

f) Ancillary diagnostic tests: To make a MP


morphology diagnosis, MUC1 expression along
the membrane segment facing the stroma [153]
has been reported to be a feature of MPCs in other
organ systems. Similarly immunohistochemical
expression of CA125, HER2/neu and KL-6 [154,155]
has been reported as distinctive markers of MPC.
However, Sangoi et al stained immunohistochemical
expression of MUC1, CA125 and HER2/neu and
reported that these markers are not entirely specific
for MPC when compared specifically with invasive
urothelial carcinoma having prominent stromal
retraction. As such, these markers lack reliable
diagnostic utility in challenging cases, further casecontrolled studies are needed to compare specificity
versus invasive urothelial carcinoma with prominent
stromal retraction. At present, there are no specific
markers for MPC and no published molecular
comparative data to address this specific diagnostic
issue.

As mentioned in the differential diagnosis, to make


a diagnosis of bladder MPC and to differentiate
from other sites of primary tumor, the most sensitive
marker is reported to be pan-uroplakin, which shows
membranous and/or cytoplasmic staining in most
of bladder cases with negative for breast, lung and
ovary, making it a specific marker for urothelial MPC.
Other markers including EMA, CK20, high molecular
weight cytokeratin, p63, and thrombomodulin are
also positive in urothelial MPC, but less sensitive
and specific compared to pan-uroplakin. Other
markers including CA125, B72.3, Leu M1, BerEp4,
placental alkaline phosphatase have been reported
to be positive. Samaratunga and Khoo also reported
immunohistochemical findings of MPC similar
to conventional urothelial carcinoma with CK7,
CK20, EMA, CEA and high molecular cytokeratin.
However, CA125 staining was seen only in MPC.

Another diagnostic challenge is the differentiation of


classic MPC from conventional UC with retraction
(non-classic). A recent consensus study conducted
by Sangoi et al [140] found that interobserver
reproducibility for a diagnosis of classic and nonclassic MPCs in the 30 study cases was only
moderate (k: 0.54). Although classification as MPC
among the 10 classic MPC cases was relatively
uniform (93% agreement), the classification in the
subset of 20 invasive UCs with extensive retraction
and varying sized tumor nests (non-classic) was
more variable. Classic MPC features included 1)
multiple nests within the same lacunar space, 2)
intracytoplasmic vacuolization, 3) epithelial ring forms,

92

Based on the morphology and immunohistochemical


profile of the MPC, Samaratunga and Khoo [156]
suggest that MPC may be a form of glandular
differentiation in urothelial carcinoma. MPC displays
overexpression of p53 and MIB-1, Aurora-A, but
no statistically significant difference could be made
with conventional urothelial carcinoma except for
Aurora-A. Aurora-A overexpression in MPC may
play a role in aggressive clinical behavior of MPC
[157]. E-cadherin is positive in MPC as seen in usual
urothelial carcinoma, whereas plasmacytoid and
signet ring cell differentiation in UC shows loss of
E-cadherin [158]. The immunoreactivity to Her2Neu
and PTEN has been reported in MPC and might be
relevant in terms of future targeted therapy [159].

percentage of the MP component of the carcinoma.


Radical surgery is mandatory in muscle-invasive
disease, even though patients with lymph node
involvement die from the disease. However, in
non-muscle-invasive disease and in the absence
of associated carcinoma in situ, intravesical BCG
treatment may be offered when the MP component
of the carcinoma component is a small percentage.
The prognosis is uniformly unfavorable; the 5-year
and 10-year survival in the largest study was 51
and 24%, respectively. Samaratunga and Khoo
[156] reported that the prognosis is related to the
proportion and location of the MPC. Cases with
moderate or extensive (> 10%) MPC are at high risk
of being advanced at presentation. Cases with focal
(<10%) MPC and surface MPC have a high chance
of detection at an early stage.

g) Prognosis: There are several important reasons

for recognizing MP variant of urothelial carcinoma.


Most are muscle invasive at presentation and
frequent lymph node and distant metastases
with common sites being lung, liver and bone. (1)
There is an ample evidence to suggest that this
unique MP configuration of UC connotes a more
aggressive clone of neoplastic cells(a) these
tumors are invariably high grade and usually of
high stage and are frequently associated with LVI;
(b) the amount of MP component is correlated
inversely with prognosis, (c) the MP component has
a higher DNA index than did conventional urothelial
carcinoma; and (d) metastatic sites of tumors have a
predominant MP component. (2) Initial presentation
with MP histology in metastatic sites should prompt
consideration of the possibility of UC, especially if the
MP configuration is encountered in the peritoneum,
abdominal lymph nodes or mesentery of a male
patient with an unknown primary or in a female with
normal appearing ovaries. (3) The high association
of MPC with muscle-invasive disease should alert
the pathologists to recommend a clinician to get a
deep biopsy, if the initial biopsy is superficial and
lacks muscularis propria. (4) Recent data suggest
that because of limited response to intravesical
BCG, one group has advocated for early cystectomy
for pTa and pT1 tumors with MP histology, as these
tumors are unlikely to respond to chemotherapy
when used as a secondary treatment modality.
Kamat et al [138] from MD Anderson Cancer Center
reported BCG treatment in 27 of 44 non-muscle
invasive MPCs of bladder (5Ta, 4 CIS and 35T1),
and of these 27 patients, 67% had progression (T2 or
greater), including 22% in whom metastatic disease
developed. Based on their findings, they reported
that the optimal treatment strategy for nonmuscle
invasive MPC is radical cystectomy performed before
progression. However, this recommendation is not
universally accepted as there needs to be confirming
studies from other institutions and also due to the
interobserver variability in the diagnosis of MPC.

3. Nested Variant of Urothelial


Carcinoma
a) Definition: A variant of invasive urothelial

carcinoma consisting of small irregularly distributed


nests with bland cytology which may simulate benign
processes such as Von Brunn nests [143,160-164].

b) Incidence: Rare with less than 100 cases


reported in literature between 1979 and 2007 the
estimated incidences 0.3% [143].

c) Clinical: The tumor typically affects men in later

adulthood (after 50 years of age) [143,162]. The


most frequent clinical manifestations are hematuria,
urgency and signs of ureteral obstruction. At
endoscopy, the tumors may be slightly raised and
erythematous or nodular and infiltrative.

d) Gross Appearance: Tumors typically involve

urinary bladder most commonly involving a trigone or


ureteric orifice. Involvement of ureter or renal pelvis
may also be noted. The size range is from less than
1 cm to 8 cm [143].
e) Microscopic Features: The nested variant of
urothelial carcinoma typically consists of nests of
bland appearing urothelial cells irregularly distributed
within lamina propria and often showing invasion
of muscularis propria (Figures 30, 31). Overlying
papillary tumor or in-situ disease is generally not
present. This tumor has different appearances in
the superficial and deep aspects. In the superficial
zone, the nests are tightly packed but somewhat
haphazardly arranged and often confluent with
little intervening fibromuscular stroma. Sometimes
abortive tubules are noted and this pattern overlaps
with urothelial carcinoma with small tubules (see later
section). The nests generally have bland cytologic
appearance but occasional ones may demonstrate
more atypia with pleomorphic nuclei and enlarged
nucleoli (Figure 31) [165,166]. In the deeper
portions, the tumor cells have a greater degree
of cytologic atypia and the cytoplasm tends to be

Gaya et al [139] recently reported that tumor stage


and patient outcome of MPC may be related to the

93

g) Ancillary Studies: The nested variant of

more eosinophilic than the nests in the superficial


portions. Commonly the nested variant is associated
with deep infiltration of muscularis propria.

urothelial carcinoma typically displays positivity for


high molecular weight keratin (34BE12), cytokeratins
7 and 20 and p63, similar to usual urothelial carcinoma
[143,170]. In comparison to benign mimickers such
as Von Brunn nests, nested urothelial carcinoma
typically shows higher MIB-1 labelling although
examples of nested variant of urothelial carcinoma
showing low MIB-1 expression have been reported
[160,168]. The nested variant of urothelial carcinoma
has also been shown to high positivity for p53 in
comparison to Von Brunn nests. Loss of p21 and
p27 expression has also been observed [160].

h) Prognosis: The nested variant of urothelial

carcinoma usually presents with muscle invasion


and tends to be aggressive [143]. The reported
series are relatively small and it is difficult to
extrapolate specific prognostic information. Some
authors suggest that this tumor is more aggressive
than usual invasive urothelial carcinoma while
others suggest that the prognosis and clinical course
is similar to high grade urothelial carcinoma stage
for stage [143,162,171,172]. The optimum treatment
for nested urothelial carcinoma has not been
established due to the rarity of the tumor. Specifically,
no significant benefit of adjuvant chemotherapy or
radiation has been established [143,162,172].

Figure 30. Nested variant of urothelial carcinoma,


low power.

4. Large Nested Variant of


Urothelial Carcinoma
a) Definition: A variant of nested urothelial
carcinoma consisting of large, irregular to regular
nests with bland cytology [164].

b) Incidence: Rare with the only series reporting 23


consult cases over a 9 year interval.

Figure 31. Nested variant of urothelial carcinoma,


high power.

c) Clinical: Mean patient age is 63.7 years (3989) and 86% are male.
hematuria.

f) Differential Diagnosis: The main differential

diagnosis is a proliferation of Von Brunn nests [167169]. The latter are generally superficial and closely
related to the overlying urothelium. Proliferating Von
Brunn nests tend to have a rounded interface with
lamina propria and exhibit both architectural and
cytological uniformity. In contrast the nested variant of
urothelial carcinoma is more architecturally complex
with irregularly distributed nests within lamina propria.
Confluence and anastomosis of nests is seen and
the lesion is often deeply invasive into a muscularis
propria. Nephrogenic adenoma may occasionally
enter the differential diagnosis although that lesion
does not often consist entirely of small nests and
cords. It usually has tubular, papillary, cystic or
signet-ring features [167,169]. Other neoplasms
including paraganglioma, carcinoid tumor, secondary
prostatic adenocarcinoma and melanoma may enter
the broad differential diagnosis [143,170]. In these
cases the selective use of immunohistochemistry
can help to resolve the diagnostic problem[170].

Typically present with

d) Gross Appearance: No information available.


e) Microscopic Features: Cases are composed of

medium to large sized nests with either rounded or


irregularly shaped borders, composed of cytologically
bland urothelial cells (Figure 32). In a minority of
cases the invasive component has a broad, pushing
invasive front, similar to that identified in verrucous
squamous cell carcinomas. Unlike the nested
variant of urothelial carcinoma, the large nests are
often separated by broad areas of fibrous tissue
and/or smooth muscle with little to no back-to-back
nests present. Also unlike usual nested carcinoma a
surface component is present in most cases, typically
low grade papillary urothelial carcinoma. At the time
of diagnosis, the large nests are invasive into the
muscularis propria in almost all cases. In a minority
of cases, focal areas of conventional patterns of
invasive urothelial may be identified [167,169,173].

94

urothelial-type cells which may be quite attenuated.


True cuboidal or columnar cells are not seen. The
tumor shows a diffuse pattern of infiltration with
haphazard arrangement of nests and tubules. The
cytologic appearance is low grade but at least focally,
some degree of nuclear pleomorphism may be seen.
Invasion of muscularis propria is often seen.

Figure 32. Large nested variant of urothelial


carcinoma.

f) Differential Diagnosis: The major diagnostic

pitfall is in distinguishing large nested variant from


either a proliferation of von Brunn nests or from an
inverted growth pattern of non-invasive papillary
urothelial carcinoma.

Figure 33. Urothelial carcinoma with small tubules.

g) Ancillary Diagnostic Tests: None available.


h) Prognosis: Follow-up was available for 17 of

f) Differential Diagnosis: The differential diagnosis

the 23 reported cases with an average interval of


43 months (range 5 months - 9 years). In 11 of 17
patients, there was no evidence of disease, while 3
patients were alive with disease and 3 had died of
their disease. In two cases, metastases to the lung
were present. While distinguishing the presence of
invasion from either an inverted growth pattern of
non-invasive papillary urothelial carcinoma or a florid
proliferation of von Brunn nests is difficult, recognition
of large nested variant of urothelial carcinoma is
imperative, as they have the ability to metastasize
and result in death [167,169,173].

includes cystitis glandularis and nephrogenic


adenoma. The glands of cystitis glandularis are
architecturally uniform and are lined by cuboidal to
columnar lining cells in contrast to the urothelialtype cells of urothelial carcinoma with small tubules.
Nephrogenic adenoma typically displays a tubular
pattern but the tubules are often tiny and associated
with other typical patterns of nephrogenic adenoma
such as microcystic, signet-ring and cord-like areas.

a) Definition: A variant of nested urothelial

carcinoma having a major component of small to


medium sized tubules which may be confused
with benign or malignant glandular processes
[164,167,169].

Adenocarcinoma of the urinary bladder displays true


glandular structures lined by cuboidal or columnar
cells. Another important differential diagnosis is
secondary adenocarcinoma of prostate [173].
Marker studies for PSA, PAP, high molecular weight
keration, cytokeratins 7 and 20 and p63 are useful to
distinguish between secondary prostatic carcinoma
and urothelial carcinoma with tubular differentiation
[173].

b) Incidence: Extremely rare with less than a dozen

g) Ancillary Studies: Urothelial carcinoma with

5. U
 rothelial Carcinoma with Small
Tubules

reported cases.

small tubules typically shows the marker profile of


urothelial carcinoma which includes positivity for
high molecular weight keratin (34BE12, CK7, CK20)
[173].

c) Clinical: Similar to a nested variant of urothelial


carcinoma.

d) Gross Appearance: Similar to a nested variant

h) Prognosis: Little is known about the prognosis

of urothelial carcinoma.

of urothelial carcinoma with small tubules. Since


urothelial carcinoma with small tubules is often
admixed with solid nests and likely represents part
of the spectrum of nested urothelial carcinoma, an
adverse outcome may occur [174].

e)

Microscopic Features: This deceptively


bland variant of nested urothelial carcinoma is
characterized by irregularly infiltrative small to
medium sized tubules usually admixed with small
solid nests (Figure 33). The tubules are lined by
95

6. Microcystic Urothelial Carcinoma

adenoma sometimes has a microcystic appearance


and enters the differential diagnosis with microcystic
urothelial
carcinoma,
however,
nephrogenic
adenoma typically displays other patterns as well.
In problematic cases, a racemase and Pax 2 or 8
stains maybe useful since it strongly decorates
nephrogenic adenoma.

a) Definition: A variant of urothelial carcinoma

characterized by dilated tubules with microcysts and


macrocyts [169,175-177]. It has been suggested
that the tumor should have at least 25% microcytic
change in order to be designated as microcystic
urothelial carcinoma [143].

Additionally, primary vesicular adenocarcinoma and


secondary adenocarcinoma is sometimes in the
differential diagnosis. Immunohistochemical studies
may be required to resolve the differential diagnosis
[170].

b) Incidence: Rare with less than 20 reported


cases [143].

c) Clinical: No specific clinical features.


d) Gross Appearance: Gross features similar
to nested and small tubular variants of urothelial
carcinoma occasional 1-2 mm macrocysts may be
identified.

g)

Ancillary Studies: Microcystic urothelial


carcinomas stain in a similar fashion to other
urothelial carcinomas [143,177-179].

e) Microscopic Features: The tumor consists of

h) Prognosis: Due to the limited number of cases,

very little is known about the prognosis of this variant,


however, in reported cases, the prognosis has been
similar to usual invasive urothelial carcinoma [143].

tubular structures which are cystically dilated forming


microcysts and occasionally macrocysts (Figure
34). Eosinophilic luminal material and necrotic debris
are often seen. Luminal secretions may display a
targetoid appearance. The luminal secretions stain
positively for periodic acid-Schiff and alcian blue
stains. Calcifications may be present within the cysts.
Urothelial cells, many of which are flattened, line
the cystic structures. Focal mucinous differentiation
may be noted but cuboidal or columnar cells are
not seen. The nuclei are typically bland although
focal atypia may be identified. The tumor has an
infiltrative appearance with involvement of lamina
propria and muscularis propria. A stromal reaction
of varying degree may be identified. The microcystic
pattern may be associated with more typical nested
and small tubular patterns of urothelial carcinoma as
described in prior sections of this publication.

7. Lymphoepithelioma-Like Carcinoma
a) Definition: Lymphoepithelioma-like carcinoma
of the urinary bladder is a variant of urothelial
carcinoma in which there is a dense inflammatory
infiltrate surrounding nests of poorly differentiated
carcinoma cells, imparting a close resemblance to
the well-known lymphoepitheliomas of the head and
neck.

b) Incidence: Although lymphoepithelioma-like

carcinoma was first recognized two decades ago


in 1991, the largest series is 28 cases and the next
three largest series encompass only 13, 11, and 9
cases so lymphoepithelioma-like carcinoma of the
bladder appears rare [180-184].

f) Differential Diagnosis: The most common

differential diagnosis is nephrogenic adenoma


and cystitis cystica glandularis. The latter lesion is
superficial and shows very regular with rounded
epithelial-stromal borders. The glands of cystitis
cystica are often lined by attenuated cuboidal
cells rather than urothelial-type cells. Nephrogenic

c) Clinical: In the 4 largest series, 70% of the

patients have been men. Ages ranged from 44 to


90 years with a median age of 70 years. The age
distribution for women appears to be the same as
for men. The age distribution for patients with pure
lymphoepithelioma-like carcinoma does not appear
to be different from that for patients with mixed
tumors. Macroscopic hematuria is the most common
presenting complaint. Urgency, burning sensation
at micturition, increased frequency of micturition,
and difficulty in micturition have also been reported,
although the last two may be symptoms of
concurrent prostatic hyperplasia. Little information
is available about the cystoscopic appearances of
these tumors.

d) Microscopic Features: As the name indicates,

the essence of the histology of these tumors is


a resemblance to the lymphoepitheliomas of the
nasopahrynx: syncytial-looking nests sheets and
cords of large poorly differentiated carcinoma
cells in a stroma heavily infiltrated by lymphocytes

Figure 34. Microcystic urothelial carcionoma.

96

f) Prognosis: For 28 patients from the major series

(Figure 35). The carcinoma cells have large vesicular


nuclei with prominent nucleoli and mitotic figures
are usually numerous. Rarely, the nuclei may be
dramatically pleomorphic [185]. The lymphoid infiltrate
may includes populations of plasma cells, histiocytes,
and granulocytes. While some lymphoepitheliomalike carcinomas are tumors composed entirely of this
element, many times the diagnosis is applied when
areas of lymphoepithelioma-like carcinoma are
found in tumors which also contain garden-variety
urothelial carcinoma, with or without other variants of
urothelial carcinoma. With no well-defined lower limit
for the amount of lymphoepithelioma-like carcinoma
needed to support the diagnosis, it is possible
that this variant is underdiagnosed. Amin et al
stratified their cases into those composed purely of
lymphoepithelioma-like carcinoma, those composed
predominantly of lymphoepithelioma-like carcinoma,
and those with only focal lymphoepithelioma-like
carcinoma [180]. Tamas et al simplified this approach
by stratifying their cases into those composed purely
of lymphoepithelioma-like carcinoma and those in
which other elements were present. This distinction
appears to have prognostic implications (vide infra)
[183].

with pure lymphoepithelioma-like carcinoma of the


urinary bladder with follow up of 4 months to 216
months (median 29 months) only one is recorded as
having died of the carcinoma and only 2 more are
recorded as having spread or recurrence but had not
died at the time of the report. The outlook for those
with mixtures of other elements may be less favorable.
For 10 patients with focal lymphoepithelioma-like
carcinoma of the urinary bladder, 9 died of the
carcinoma and 1 died shortly after diagnosis of
other causes. For 29 patients with predominant
or mixed lymphoepithelioma-like carcinoma of
the urinary bladder, 10 died of the carcinoma and
3 had major progression of the carcinoma at the
time of the report. While the numbers are small,
the differences in reported outcomes between
pure lymphoepithelioma-like carcinoma and tumors
with other components seems likely to matter. In
pathology reports, it would be appropriate to state
whether or not the tumor is pure lymphoepitheliomalike carcinoma.

8. Lipoid-Rich Variant of Urothelial


Carcinoma
a) Definition: Lipid-rich (lipoid) variant is defined by

the presence of large carcinoma cells with multiple


large, clear cytoplasmic vacuoles, morphologically
similar to lipoblasts [188].

b) Incidence: This is a rare variant of urothelial

carcinoma. Following the initial description [189],


two published series included a total of 32 patients
[148,190].

c) Clinical: Patients have typical symptoms of

urothelial cancer; all studied cases had hematuria


but other symptoms included pain, irritation and
obstruction. There is a striking (>90%) predominance
of males with most patients in the 7th decade of life.

d) Gross Appearance: Cystoscopically and grossly

Figure 35. Lymphoepithelioma-like carcinoma

these tumors show no distinctive features. Most


cases of the lipid-rich variant of urothelial carcinoma
arise in the bladder.

e) Ancillary Diagnostic Tests: In most cases,

no ancillary tests should be required to make the


diagnosis. However, if lymphoma is a consideration
on routine sections, immunohistochemistry to detect
markers of epithelial differentiation will quickly clarify
the situation because the carcinoma cells mark
strongly with antibodies to a variety of markers of
epithelial differentiation. Cytokeratin AE1/3, epithelial
membrane antigen, cytokeratin 7, and cytokeratin 8
are some of the most frequently detected epitopes.
Tumor protein p53 is also frequently detectable in
lymphoepithelioma-like carcinoma of the urinary
bladder. Epstein-Barr virus is found in nasopharyngeal
carcinomas but is not found in lympho-epitheliomalike carcinoma of the urinary bladder [186,187].

e)

Microscopic Features: The low power


appearance is that of high-grade urothelial carcinoma
admixed with large, multi-vacuolated cells that
constitute 10- 50% of the lesion [148]. At high-power,
the diagnostic cells have optically clear vacuoles
which often indent an eccentric nucleus (Figure
36). Some cases may also exhibit other subtypes of
urothelial carcinoma [148,191,192]. Histochemistry
and electron microscopy indicate that the vacuoles
contain lipid rather than mucin [117,148]. The
lipidized cells are carcinoma cells and have similar
immunoprofile to urothelial carcinoma and stain
with EMA, cytokeratins (including cytokeratin 7 in
all cases and cytokeratin 20 and 34BE12 in most
97

b) Incidence: Rare with less than 20 documented

cases) and more specific urothelial markers such as


thrombomodulin. They are negative for S100 [148].

cases in the English literature [198-203] and has so


far not detailed in any large-scale series.

c) Clinical: Patients present with typical symptoms


of urothelial carcinoma, mainly hematuria. Mean age
is 64 (41-82) years.

d) Gross Appearance: Most cases occur in bladder

but ureter/renal pelvis and urethral origin are also


described [198,200,201]. There is no information
available to suggest that gross appearance of tumors
differ from usual type.

e) Microscopic Features: Tumors are typically

high-grade and invasive with papillary and/or solid


growth pattern similar to usual type urothelial
carcinoma. In the majority of cases, more typical
urothelial carcinoma is recognized elsewhere [117]
and tumors may also display other variant types
[199]. Tumor cells are large with abundant clear
cytoplasm and well defined cell membranes and
centrally placed pleomorphic nuclei (Figure 37).

Figure 36. Lipoid-rich variant of urothelial carcinoma.


Solid sheets of high-grade urothelial carcinoma with
abundant vacuolated clear cells, many of which
have a multivacuolated, lipoblast-like appearance
(arrows).

f) Differential Diagnosis: The main differential

diagnosis includes sarcomatoid carcinoma with


heterologous liposarcomatous differentiation and
signet ring cell carcinoma. As both tumors can be
either primary or secondary [193,194], this may be
of importance.

g) Ancillary Diagnostic Tests: Lipid rich cells are

negative for mucin histochemical stains and S100


immunostain. As these cells are immunopositive
for cytokeratin, a panel of mucin stains, S100
immunostain and cytokeratin immunostain may
be useful to rule out sarcomatoid and signet ring
cell carcinoma [121,148,195]. Tumors retain the
immunoprofile of usual type urothelial carcinoma
(positive for cytokeratin 7, and focally positive for
thrombomodulin, cytokeratin 20 and high-molecular
weight cytokeratins) [121,170].

Figure 37. Clear cell (glycogen-rich) variant of


urothelial carcinoma. Expansile, invasive nests
sheets of high-grade urothelial carcinoma showing
clear cytoplasm.

h) Prognosis: The outcome for this variant is poor.

In two series [148,190] that included 32 cases, 27


(84%) were muscle-invasive or beyond (and all of
those not diagnosed as muscle-invasive were staged
based on transurethral tumor resection). The vast
majority of patients (94%) had progressive (local or
metastatic) disease and death occurred in almost
two thirds of patients.

f) Differential Diagnosis: It is important to

9. Clear Cell (Glycogen Rich) Variant


of Urothelial Carcinoma

recognize that clear cell variant urothelial carcinoma


can be confused histologically with other clear cell
neoplasms. Clear cell adenocarcinoma [204-206] and
metastatic renal cell carcinoma [207-211] both enter
into the differential diagnosis as may on occasion
prostatic adenocarcinoma and paraganglioma [212].
Finally, urothelial carcinomas of usual type frequently
exhibit artifactual cytoplasmic clearing, especially in
TURBT specimens where diathermy was used.

a) Definition: Cytoplasmic clearing is focally

g) Ancillary Tests: Immunohistochemistry is similar

present in many, otherwise typical, urothelial


carcinomas [124] as well as in occasional cases of
urothelial dysplasia [196]. In rare instances, clear
cell change is the predominant pattern in which case
the term clear cell variant of urothelial carcinoma is
used [197].

to that of usual type urothelial carcinoma, i.e. positive


for high-molecular weight cytokeratins, cytokeratin 7,
p63, GATA-3, thrombomodulin and variably positive
for cytokeratin 20 [117,170,202,203,213]. Negativity
for renal, prostatic and neuroendocrine markers may
be of help in above differentials.

98

h) Prognosis: The prognostic and therapeutic

(mean 68.7 years). In the pediatric cases, all patients


have been girls; at presentation, one patient was 2
years old, two were 4 years of age and one was
six years old. Hematuria appears to be the most
common presenting sign.

implication of clear cell variant histology remains


uncertain given lack of large series. Tumors are
typically muscle invasive [198-203] and have been
associated with a high (50%) frequency of nodal
metastases [198-203]. See Table 9.

d) Gross: These tumors do not have a specific

Table 9. Clear cell variant. Small series and case


reports. Gender, age and stage
Reference
Kotliar et al, 1995
Braslis et al, 1997

No. Stage
2 N/A
T4
2 T3+

Perez-Montiel et al,
2006

T3+
T3

Yamashita et al,
2006
Rotellini et al, 2010

T4
T3+

Domanowska et al,
2001

gross appearance; some have been described as a


polypoid or submucosal mass [216,217].

Outcome
Not stated
Not stated
6 months
AWOT
Not stated
Not stated

e) Microscopic Features: Tumors with rhabdoid

differentiation have a very distinct microscopic appearance. They are characterized by large discohesive tumor cells that have distinct cell borders,
abundant cytoplasm with a perinuclear eosinophilic
inclusion, and a large nucleus, sometimes eccentrically located within the cell, with a prominent nucleolus (Figure 38). The tumor cells may be arranged
singly, in small clusters or in more diffuse sheets.
Ultrastructural studies have shown that the cytoplasmic inclusion consists of whorls of intermediate filaments.

Not stated
7 months
AWOT
T3+
12 months
AWOT
T3+ N0 Not stated
T3+ N0
T3+ N0
T3+ N1
T3+ N1
T3+ N1

Not stated
Not stated
Not stated
Not stated
Not stated

AWOT, alive without evidence of tumor.

10. Urothelial Carcinoma with


Rhabdoid Features
a) Definition: Urothelial carcinoma with rhabdoid

differentiation is a rare, aggressive tumor. Typically,


in adults, rhabdoid features are seen within an otherwise poorly differentiated conventional urothelial
carcinoma [214]. Pure rhabdoid tumors of the bladder (extra-renal malignant rhabdoid tumor) are even
rarer, and those reported have occurred in pediatric
patients [215-217].

Figure 38. Urothelial carcinoma with rhabdoid


features.

f) Differential Diagnosis: Given that most bladder

tumors with rhabdoid features, at least in adults,


are poorly differentiated urothelial carcinomas
morphologically, diagnoses should not be too
problematic. If, however, the entire tumor consists of
rhabdoid cells, the main differential diagnosis would be
an extra-renal malignant rhabdoid tumor. Much less
likely would be other tumors with dense eosinophilic
cytoplasm, including adrenal cortical carcinoma,
hepatocellular carcinoma, rhabdomyosarcoma, and
epithelioid angiomyolipoma.

b) Incidence: A specific incidence is unknown,

given the rarity of these tumors in the bladder.


Approximately 6 cases have been reported in adults,
all associated with other morphologies including
conventional urothelial carcinoma, malignant fibrous
histiocytoma, and small cell, sarcomatoid, and
myxoid histologies [218-220]. One case of urothelial
carcinoma with sarcomatoid growth and rhabdoid
differentiation in a 2 year old girl has also been
reported [221]. Pure rhabdoid tumors of the urinary
bladder are exceedingly rare with only 3 possible
cases reported to have occurred in pediatric patients
[215-217].

g) Ancillary Diagnostic Tests: Malignant rhabdoid

tumors have a deletion or mutation of the hSNF5/


INI1 gene on chromosome 22q11.2 [222,223]. For
bladder tumors that consist entirely of rhabdoid
cells, some advocate that molecular analysis should
be performed to determine if the tumor is a primary
extra-renal malignant rhabdoid tumor or a urothelial
carcinoma with rhabdoid differentiation [214]. Loss

c) Clinical: In adults, most of the tumors have

occurred in men with an age range of 53-86 years

99

of nuclear INI1 immunohistochemical expression


has been demonstrated to be useful in distinguishing
malignant rhabdoid tumors from other tumors that
may demonstrate rhabdoid features [224] and,
therefore, this test could also be performed.

degree of nuclear pleomorphism is low to moderate.


Binucleation may be present [150,233]. Nucleoli may
be present. The tumors cells are often in a stroma
that appears edematous or myxoid and can be arranged in cords, small nests, sheet-like or scattered
discohesive cells [214,233,235]. The plasmacytoid
tumor cells usually co-exist with conventional urothelial carcinoma, sarcomatoid carcinoma, nested or
micropapillary urothelial carcinoma, glandular morphology, urothelial carcinoma in situ or high grade
papillary urothelial carcinoma [150,228,233,235].

h) Prognosis: In general, bladder urothelial

carcinomas with rhabdoid differentiation are


aggressive high grade, high stage neoplasms and
the patient prognosis is poor [214,219,220]. For the
pediatric patients diagnosed with pure rhabdoid
bladder tumors, two were alive without evidence
of disease at follow-up of 2 years and 9 years
[216,217]. Given the behavior of these tumors, at
least in adults, radical resection seems prudent.
Some pediatric patients received chemotherapy
and/or partial cystectomy or transurethral resection.

11. Plasmacytoid Urothelial


Carcinoma
a) Definition: Plasmacytoid urothelial carcinoma is
a rare variant of urothelial carcinoma characterized
by tumor cells that morphologically resemble plasma
cells. This malignant tumor has been recognized
by the most recent World Health Organization
classification as a distinct variant of urothelial
carcinoma [225].

Figure 39. Plasmacytoid urothelial carcinoma.

b) Incidence: This tumor is not common; the

f) Differential Diagnosis: At cystectomy, plasma-

literature contains individual cases reports and


relatively small series [150,184,192,226-239]. To
date, less than 100 cases have been reported. In
one series of 720 high grade urothelial carcinomas
of the bladder, the incidence of the plasmacytoid
urothelial carcinoma was less than 1% [228], and in
another series of 260 invasive urothelial carcinomas,
the plasmacytoid variant accounted for 2.7% of
cases [231].

cytoid carcinoma is usually associated with conventional urothelial carcinoma. In the setting of a small
bladder biopsy or metastatic tumor containing only
plasmacytoid cells and the absence of a precursor
urothelial neoplasm (in situ carcinoma or papillary
urothelial carcinoma) in a bladder biopsy, the differential diagnosis of plasmacytoid urothelial carcinoma is broad and includes lymphoma/plasmacytoma,
malignant melanoma, rhabdomyosarcoma, lobular
breast carcinoma, gastric adenocarcinoma, neuroendocrine carcinoma, paraganglioma [214,235].

c) Clinical: The majority of patients with plasmacytoid

urothelial carcinoma are men [150,184,192,226239]. The age range at presentation is broad (46-89
years). In some of the larger series, the mean age at
presentation was 64.3 years [235], 66.2 years [227],
67 years[150] and 70 years [233]. Most patients
present with hematuria, with fewer complaints
related to irritative voiding symptoms [150,192,226,
227,230,231,233,235,238,239].

g) Ancillary Diagnostic Tests: Plasmacytoid

urothelial carcinomas are positive for pancytokeratin,


CK7, CK20 and uroplakin III [150,233,235,240].
CD138 is a marker expressed in normal and malignant
plasma cells and is also expressed in plasmacytoid
urothelial carcinoma [150,232-234,240]; therefore,
it cannot be used alone to diagnose plasmacytoid
urothelial carcinoma. In one of the largest and
most recent studies, 78% of the tumor cells were
positive for CD138 [240]. In this same study, 64.5%
of the tumors had loss of membranous E-cadherin
expression and the authors propose that loss of this
expression is a feature of plasmacytoid urothelial
carcinoma. Depending on the specific differential
diagnosis, other immunohistochemical markers can
be performed as necessary.

d) Gross Appearance: The gross findings are

non-specific and little information is provided in the


literature. Only a few reports provide a description of
the tumors appearance which includes sessile [192]
and ulcerated, firm mass or masses [227].

e) Microscopic Features: The tumor is character-

ized by discohesive round to oval tumor cells with


dense eosinophilic cytoplasm which may occasionally contain small vacuoles [150,235]. The nuclei are
round to oval and hyperchromatic and are located
eccentrically within the cell (Figure 39). Typically the

h) Prognosis: Plasmacytoid urothelial carcinoma

typically presents at an advanced stage and, in

100

f) Differential Diagnosis: The differential diagno-

general, the prognosis is poor [150,214,233,235,240].


Treatment options are variable. Some patients
receive neoadjuvant or adjuvant chemotherapy and
resection may consist of a transurethral resection or
cystectomy.

sis includes benign conditions such as pseudosarcomatous myofibroblastic proliferations (including


postoperative spindle cell nodules and inflammatory
myofibroblastic tumors) or more aggressive lesions
such as urothelial carcinoma with chondroid or osseous metaplasia, and leiomyosarcoma [252,253].

12. Sarcomatoid Carcinoma of The


Urinary Bladder
a)

g) Ancillary Diagnostic Tests: The immunohis-

Definition: A malignant urothelial-derived

tochemical profile of sarcomatoid carcinoma includes


immunoreactivity for pancytokeratin, high molecular
weight cytokeratin, CK5/6 and p63. Smooth muscle
actin is only variably positive at best and Alk-1 is
negative these markers are frequently expressed
in inflammatory myofibroblastic tumors/pseudosarcomatous myofibroblastic proliferations and leiomyosarcomas [254,255].

carcinoma that displays both spindle cell and


epithelial elements (formerly carcinosarcoma).

b) Incidence: Sarcomatoid carcinoma accounts


for approximately 0.6% of all bladder cancers
[241,242].

c) Clinical: The mean patient age is 66 years and

the male to female ratio is 3:1. Similar to other bladder


carcinomas, patients with sarcomatoid carcinoma
typically present with hematuria [243-251].

h) Prognosis: An estimated 70% of patients


die within 2 years of diagnosis. Compared with
patients with pure urothelial carcinoma, patients with
sarcomatoid carcinoma are at a greater risk for death
even after adjusting for the stage at presentation.

d) Gross Appearance: The majority of lesions


appear as a solitary, polypoid mass that may involve
any portion of the bladder.

13. Undifferentiated Urothelial


Carcinoma with Trophoblastic
Giant Cells

e) Microscopic Features: The spindle cell com-

ponent usually shows high-grade morphology with


nondescript architecture that often resembles malignant fibrous histiocytoma. (Figure 40) Heterologous
differentiation is often present that may contain - in
decreasing order of frequency - areas that resemble osteosarcoma, chondrosarcoma, rhabdomyosarcoma, liposarcoma, angiosarcoma or a mixture
of sarcomatoid histologies [245,247]. Despite the
morphologic resemblance to sarcoma, it is currently
recognized that these spindle cell areas are derived
from an underlying epithelial malignancy, hence the
revised terminology of sarcomatoid carcinoma. Often associated with areas of sarcomatoid morphology may be foci of overlying flat urothelial carcinoma
in situ or adjacent invasive high grade urothelial carcinoma, squamous cell carcinoma, adenocarcinoma,
or small cell carcinoma.

a) Definition: Trophoblastic differentiation may be

seen in several settings in urothelial carcinoma: 1)


Serologically and/or immunohistochemically in otherwise typical urothelial carcinoma; or 2) Morphological evidence of trophoblastic differentiation either
with scattered trophoblastic giant cells in urothelial
carcinoma or with a biphasic tumor indistinguishable
from choriocarcinoma [256-260]. It is considered
that this phenomenon represents either morphologic
and/or functional trophoblastic differentiation from
reactivation of embryonal cellular programs in tumor
cells.

b) Incidence: 1) Human chorionic gonadotropin

(HCG) and to a lesser extent human placental lactogen (HPL) and pregnancy-specific beta-1-glycoprotein (SP-1) immunoreactive cells have been found
in a minority of urothelial carcinomas, typically high
grade. Although different grading systems have been
used, the incidence for high grade tumors has been
19%, 21%, 40%, and 55% in various studies. Most
studies have found no immunohistochemical HCG
production in low grade tumors, although one study
found 26% [229,256,261-263]. Serum HCG levels
have also been shown to be elevated in some of the
cases with immunohistochemical evidence of HCG
expression.

c) Clinical: In patients with usual urothelial carcinoma expressing HCG immunohistochemically or


by the presence of scattered syncytiotrophoblastic
giant cells, there is no endocrinologic manifestations
related to HCG. Tumors with pure choriocarcinoma
or choriocarcinoma associated with poorly differentiated urothelial carcinoma have in some cases been
associated with gynecomastia.

Figure 40. Sarcomatoid urothelial carcinoma.

101

d) Gross Appearance: High grade tumors

g) Ancillary Diagnostic Tests: Comparative ge-

associated immunohistochemical evidence of HCG


or those with choriocarcinoma usually have been
described as large, exophytic, or fungating, often
with hemorrhage and necrosis.

nomic hybridization showed that both the urothelial


carcinoma and choriocarcinoma shared losses of
chromosomes 9 and 17, similar to the genetic findings previously reported for urothelial carcinoma
[264]. The less well-differentiated choriocarcinomatous component showed more genetic alterations
than the papillary urothelial carcinoma component.
In a case of micropapillary urothelial carcinoma and
choriocarcinoma, the trophoblastic component was
positive for high molecular weight cytokeratin, which
typically is not seen in trophoblasts yet can be seen
in urothelial carcinoma. One case of pure choriocarcinoma has been shown to have an isochromosome
12p, as seen in germ cell tumors [265]. All of the
above findings support that choriocarcinoma is of
urothelial origin and acquires varying degrees of
trophoblastic differentiation.

e) Microscopic Features: Areas of high grade

pleomorphic urothelial carcinoma, whether non-invasive papillary, CIS, or invasive are the most likely
to immunohistochemically express HCG. The HCG
positive cells morphologically look indistinguishable
to adjacent HCG negative cancer cells. Morphological evidence of trophoblastic differentiation shows
either scattered syncytiotrophoblastic giant cells or
choriocarcinoma with both syncytiotrophoblastic
giant cells admixed with cytotrophoblasts indistinguishable from choriocarcinoma. Scattered syncytiotrophoblastic giant cells can be seen in a wide
spectrum of urothelial carcinomas including noninvasive papillary low and high grade urothelial carcinoma, as well as invasive carcinoma. Many cases
reported as choriocarcinoma are not classic choriocarcinomas and when their photomicrographs are
evaluated they represent undifferentiated urothelial
carcinoma with scattered tumor giant cells. Cases
which appear histologically identical to choriocarcinoma typically are admixed with high grade urothelial carcinoma. Even rarer cases appear to be typical
choriocarcinoma from inception without associated
urothelial carcinoma (Figure 41).

h)

Prognosis: HCG immunoreactivity in


pretreatment in some studies has been a strong
predictor of treatment response (22 of 29 patients
with HCG-positive specimens did not respond to
treatment versus 29 of 71 patients with HCG-negative
specimens. Of the few patients with choriocarcinoma,
most have died of disease. Although there has
been anecdotal treatment specifically directed at
trophoblastic tissue, it has failed to influence the
course of the disease. Falling levels of HCG is
not invariably associated with tumor response to
chemotherapy.

f) Differential Diagnosis: Urothelial carcinoma

14. Undifferentiated Urothelial


Carcinoma with Osteoclastic
Giant Cells

with trophoblastic differentiation can be confused


with choriocarcinoma, whereby inappropriate
chemotherapy may be utilized or there could be
a delay in treatment while looking for a primary
elsewhere. In the setting of a primary choriocarcinoma
of the bladder, spread should be excluded from a
gynecological or testicular primary, in women and
men, respectively.

a) Definition: Tumors closely recapitulating the

morphology of osteoclastic giant-cell tumors of


bone or soft parts, with a biphasic tumor composed
of both mononuclear cells and osteoclast-like giant
cells. Tumors are considered a variant of urothelial
carcinoma based on their association with usual
urothelial carcinoma and various ancillary studies
(see below).

b) Incidence: Rare with the largest series composed

of 6 cases from a consult service [266-271]. As of


2006, there were 20 reported cases with a few case
reports reported subsequently.

c) Clinical: Approximately two-thirds of patients are

male with most in their 7th decade. The presenting


symptoms are nonspecific and typical for urothelial
neoplasms in general with gross hematuria as the
most common manifestation. Other complaints
reported in the literature are flank pain, renal colic,
and dysuria.

d) Gross Appearance: Approximately 50% are

Figure 41. Urothelial carcinoma with trophoblastic


giant cells. Pure primary choriocarcinoma of
the bladder with syncytiotrophoblasts and
cytotrophoblasts intimately admixed.

located in the renal pelvis and involve the kidney.

e) Microscopic Features: Neoplasms closely recapitulate giant-cell tumors of bone and soft tissues.

102

h) Prognosis: Renal pelvis tumors are often high

They are composed of a mixture of oval to plump


mononuclear cells and multinucleated osteoclastlike giant cells, which form sheets and nodules
(Figure 42). The stroma is richly vascularized with
frequent areas of erythrocyte extravasation, occasionally forming large blood filled lakes. Large areas
of necrosis are seen in a minority of cases. Osteoclast-like giant cells exhibit phagocytic features with
erythrocytes, hemosiderin, mononuclear cells, or debris readily found within their cytoplasm. Neoplasms
are unencapsulated and demonstrated a tongue-like
infiltrative growth pattern. Lymphovascular invasion
is present in many cases. The vast majority of cases
are associated with either CIS or non-invasive high
grade papillary urothelial carcinoma.

stage at presentation (pT3) with invasion of the renal


parenchyma, hilar adipose tissue, and thick walled
veins. Bladder tumors have been of variable stage
from T1 to T3. In many cases, the disease-free
follow-up is too short to be useful. Of cases with
more meaningful yet still short follow-up, about 50%
of patients have been without evidence of disease
for 1.5-3.5 years. The remaining approximately 50%
of patients have died of disease within a short time
with pulmonary metastases the most common site
of distant disease. Death typically occurs less than
1 year following radical surgery. Extensive surgical
excision appears to be the common consensus
for primary or recurrent lesions until the treatment
benefits of adjuvant chemotherapy or radiotherapy
can be established.

15. U
 ndifferentiated Carcinoma
(Including Giant Cell Carcinoma)
a) Definition: Carcinomas of the urinary bladder

with a phenotype composed of sheets or isolated


undifferentiated cells that do not fit into urothelial,
squamous, adenocarcinoma or any other recognized
category of bladder carcinoma [117,118,120,121,14
9,188,199,272,273]. Some authorities would classify
these cases as large-cell undifferentiated carcinoma
[120,149,273]. Others may prefer to denote
these lesions as high-grade urothelial carcinoma,
given that this is the most common cancer within
the bladder and that this pattern of tumor most
likely represents the most poorly differentiated
manifestation of conventional urothelial carcinoma
(Figure 43) [120,149,273]. Giant cell carcinoma
is an undifferentiated carcinoma of the bladder
that occasionally may be composed predominantly
or purely of poorly differentiated, pleomorphic,
multinucleated, and multinucleated anaplastic cells
with abundant eosinophilic or amphophilic cytoplasm
[117,118,120,121,149,273]. The tumor cells often
contain multiple nucleoli and are similar to giant cell
carcinomas seen in the lung and in other parts of the
body (Figure 44) [274].

Figure 42. Undifferentiated urothelial carcinoma


with numerous osteoclastic giant cells.

f) Differential Diagnosis: One of the differential

diagnoses would be a sarcomatoid urothelial carcinoma, if associated with a conventional urothelial


carcinoma component. If conventional urothelial carcinoma was not identified, the tumor might not be
recognized as a variant of urothelial carcinoma. It
may be misdiagnosed as a sarcoma, either primary
in the bladder or spread to the bladder from an occult primary.

g) Ancillary Diagnostic Tests: In most but not all

cases, the mononuclear cells immunohistochemically demonstrate epithelial differentiation with focal
positivity for various keratins or epithelial membrane
antigen. Ultrastructurally, the mononuclear cells display few desmosomes, microvilli, and intracytoplasmic keratin. The multinucleated cells have identical
morphological and immunohistochemical properties
of osteoclasts; positive for CD-68, LCA, CD51 and
CD54, and primarily negative for cytokeratins and
EMA. Varying percentages of mononuclear cells
express alpha-smooth muscle actin, desmin, S-100,
and CD68. The mononuclear cells have a relatively
high proliferation index of between 20-50%. Based
on these findings the mononuclear cells are regarded as undifferentiated urothelial carcinoma with the
osteoclast-like giant cells representing reactive histiocytes.

b) Incidence: Rare with the largest series composed


of 8 cases of large cell undifferentiated carcinoma,
five of them seen in consultation, and an additional
series of 8 cases of giant cell carcinoma [149,273].

c)

Clinical: The presenting symptoms are


nonspecific and typical for urothelial neoplasms in
general with gross hematuria as the most common
manifestation. Approximately two-thirds of patients
are male with most in their 6th or 7th decade. Large
cell undifferentiated and giant cell carcinomas
present at an advanced stage [149,273].
d) Gross Appearance: No information available.
e) Microscopic Features: Neoplasms closely
recapitulate large-cell undifferentiated or giant-

103

cell carcinoma seen in the lung and in other parts


of the body [274]. They are composed of sheets
of large polygonal or round cells with moderate
to abundant cytoplasm and distinct cell borders
[117,118,120,121,149,188,199,272-274]. The architectural pattern of the tumor varies from infiltrating tumor to solid expansile nests of undifferentiad
cells. Pleomorphic giant cells are prominent in the
giant cell variant. Typical or atypical mitotic figures
are common. Rarely, the stroma is hypocellular and
desmoplastic. The vast majority of cases are associated with invasive high grade conventional urothelial
carcinoma [149,273]. Lymphovascular invasion is
present in many cases.

patient had no evidence of disease at 74 months


[149]. Extensive surgical excision appears to be the
common consensus for primary or recurrent lesions
until the treatment benefits of ad juvant chemotherapy
or radiotherapy can be established.

f)

Differential Diagnosis: If present in a


metastatic site, the histology of undifferentiated
carcinoma would not suggest an urothelial primary
[117,118,120,121,149,188,199,272-274]. The high
degree of nuclear anaplasia helps differentiate
giant cell carcinoma from the osteoclast-type giant
cells that may be seen in urothelial carcinoma
[267]. Other giant cells that may be seen with
urothelial carcinoma include syncytiotrophoblastic
giant cells or foreign body giant cells secondary to
previous biopsy or resection. In limited samples,
undifferentiated carcinoma may be misdiagnosed
as secondary carcinoma or sarcoma, a pitfall of
paramount importance for its clinical management
[117,118,120,121,149,188,199,267,272-274].

Figure 43. Giant cell carcinoma of urinary bladder.

g) Ancillary diagnostic tests: Immunohistochemical staining demonstrated that large cell undifferentiated carcinoma cases were positive for cytokeratins AE1/AE3 and 7; CAM 5.2, CK20, thrombomodulin and Uroplakin III were positive in 6, 3, 3 and
2 cases, respectively.2 Both pleomorphic giant cell
and associated conventional urothelial carcinoma
were positive for cytokeratins 7, CAM 5.2 and AE1/
AE3, and epithelial membrane antigen; P63, thrombomodulin and Uroplakin III were positive in 6, 3 and
2 cases, respectively. 1 Other immunohistochemical
markers performed in the differential diagnosis context included -Feto-protein, HCG, PSA, Vimentin,
synaptophisin and chromogranin and all were negative [149].

Figure 44. Giant cell carcinoma of urinary bladder.

II. SQUAMOUS CELL CARCINOMA


1. Squamous Cell Carcinoma, NonSchistosomal
a) Definition: Urinary bladder squamous cell

carcinoma is a primary malignancy of urothelial


origin that consists almost entirely of keratin-forming
squamous carcinoma.

h) Prognosis: Large-cell undifferentiated bladder

carcinoma is an aggressive variant of urothelial


carcinoma [117,118,120,121,149,188,199,267, 272274]. All patients had advanced stage cancer (>
pT3); and 87.5 % had lymph node metastases. 2On
follow up 75% of patients died of disease from 5 to
26 months and 2 patients were alive with metastases
at 6 and 14 months. The prognosis of large-cell
undifferentiated bladder carcinoma was compared
with conventional urothelial carcinoma of similar
stages showing survival differences (P=0.0004).
Five (62%) of patients with giant cell carcinoma died
of disease from 6 to 17 months and 2 patients were
alive with metastases at 11 and 19 months. One

b) Incidence: Squamous cell carcinoma compris-

es approximately 5% of all bladder cancers in the


Western world [275]. Excluded from this category
are urothelial carcinomas with extensive squamous
differentiation (urothelial carcinoma with divergent
differentiation), although the distinction of these two
entities on histological review may be challenging,
especially on small tissue samples. Risk factors are
associated with long standing inflammation and include long-term indwelling catheters, chronic infection, chronic bladder neck obstruction and bladder
calculi [13,14,276,277]. Human papillomavirus (HPV)

104

g) Differential Diagnosis: The differential diag-

is generally not associated with squamous cell carcinoma except in extremely rare cases [13,278-281].

nosis of bladder squamous cell carcinoma includes


urothelial carcinoma with extensive squamous differentiation (divergent differentiation) or squamous cell
carcinoma secondarily involving the bladder from
sites such as the cervix or anus.

c) Clinical Features: Squamous cell carcinoma

occurs most frequently in the seventh decade [275].


The male to female ratio is approximately 2:1 and
African-Americans are twice as likely to be affected
[281]. Clinical presentation is similar to patients with
urothelial carcinoma and includes hematuria and
irritative voiding symptoms.

h) Prognosis: There are few studies available on

pure non-Schistosomal squamous cell carcinoma


of the bladder. Five-year disease-free survival following radical cystectomy ranges from 43 to 57%
[130,279,282]. The poor outcomes associated with
this disease may be attributed to the high stage at
initial presentation and relative frequency of regional
lymph node metastases in 24% of patients at the
time of cystectomy [279]. However, when compared
to urothelial carcinoma on a stage-by-stage basis,
outcomes appear similar, with a 5-year disease specific survival rate of 57% [130,283]. Squamous cell
carcinoma has been associated with limited responsiveness to chemotherapy and radiation treatment.

d) Gross Appearance: At cystoscopy, squamous

cell carcinoma is frequently sessile, solid and often


demonstrates a white, flaky surface secondary to
keratin formation. Ulceration may occur in up to half
of all carcinomas [279]. Lesions are predominantly
solitary, may involve any aspect of the bladder wall,
although the posterior and lateral walls are most
commonly affected, and range in size from 0.8-6.4
cm (average 3.8 cm) [279].

e) Histologic Features: Squamous cell carcinoma

shows nests of invasive squamous carcinoma cells


with glassy, pink cells and variable keratin formation, often described as keratin pearls. (Figure 45)
Other classic histologic features include intercellular
bridges and sloughed keratin debris. All squamous
cell carcinomas are considered to be high-grade.
Associated mucosal findings include keratinizing squamous metaplasia (62%), nonkeratinizing
squamous metaplasia (44%), squamous cell carcinoma in situ (36%), flat urothelial carcinoma in situ
(16%), squamous metaplasia with dysplasia (9%),
verrucous squamous hyperplasia (7%) and rarely
condyloma accuminata (2%) [279].

A key challenge is in the management of patients


who present with squamous cell carcinoma on initial
biopsy or transurethral resection. Due to sampling,
the exclusion of a high-grade urothelial carcinoma
with squamous differentiation is challenging and
this distinction can be made primarily on radical
cystectomy when the entire lesion is available
for histologic analysis. A second challenge is the
management of in situ squamous lesions found
on biopsy. The most concerning of all associated
mucosal lesions is squamous cell carcinoma in situ,
which has a high likelihood of progression to invasive
carcinoma in patients and one study that followed up
on these in situ lesions found that 5/7 patients with
squamous cell carcinoma in situ on initial sampling
developed subsequent invasive carcinoma [13].

f) Ancillary Diagnostic Tests: Immunohistochemical staining is similar to squamous cell carcinoma arising at other locations and includes positive immunoreactivity to p63, high-molecular weight
cytokeratin and pancytokeratin. Recent studies are
evaluating markers that may distinguish between
squamous cell carcinoma and urothelial carcinoma
with extensive squamous differentiation, which will
be of high utility on biopsy or transurethral resection
specimens.

2. Verrucous Carcinoma
a) Definition: Verrucous carcinoma, a tumor much

more common in the upper aerodigestive tract, anus


and genitalia than the urinary bladder, is a rare subtype of squamous cell carcinoma that is most often
related to urinary schistosomiasis infection [284287], but may also rarely occur in its absence [288294]. It is a well-differentiated invasive squamous
neoplasm with a broad deep border that pushes
into the underlying stroma rather than demonstrating frank stromal invasion typical of a conventional
squamous cell carcinoma.

b) Incidence: In areas with endemic urinary schis-

tosomiasis, verrucous carcinoma accounts for up to


3.4% of all bladder carcinomas [285-287] and 4.66.5% of squamous cell carcinomas [286,287].

c) Clinical: The tumor occurs more frequently in

men than women [285-287], and in regions with endemic urinary schistosomiasis, tends to occur at a
younger age (mean age 44-46 years) in those with

Figure 45. Squamous cell carcinoma.

105

the parasitic infection than in those without the infection (mean age 53 years) and those with bladder cancer in general, who typically present in the
sixth and seventh decades of life [284-287]. In nonendemic regions of the world, a few cases of verrucous carcinoma have been reported to be related
to condyloma acuminatum [288,293], chronic urinary
tract infection [288,289,294] and bladder diverticula
[290]. In some patients, the tumors are not associated with other genitourinary diseases [291,292].
Patients usually have irritative bladder symptoms
at presentation and urine analysis may reveal white
cellular debris, consistent with keratin [284,290,292].
Patients have also reported seeing white material in
their urine [286,287,292]. Hematuria is not common
[286,287].

section, can be very difficult to make. Superficial biopsy samples may not include the base of the lesion,
and may consist only of thickened, well-differentiated, almost benign-appearing, squamous epithelium.
Transurethral resections may not be particularly useful either, as they may be maloriented and muscularis propria, useful to evaluate for invasion, may not
be present. The main differential diagnosis includes:
verrucous squamous hyperplasia, squamous papilloma, pseudoepitheliomatous hyperplasia, condyloma
acuminatum and a conventional invasive squamous
cell carcinoma that contains foci morphologically resembling a verrucous carcinoma.

d) Gross Appearance: Most tumors are solitary

ated with urinary schistosomiasis, the patient follow-up is relatively short (3 months-3 years) [288290,292,294]. Nevertheless, while some of the patients experienced tumor recurrence in the bladder
due to lack of radical resection initially [290,292,294],
all did well after total cystectomy. In larger series
of verrucous carcinoma that occurred in areas with
endemic schistosomiasis, specific outcomes are
not provided [285,286]. However, if these tumors
are diagnosed with strict adherence to histologic
criteria, given that these neoplasms are low grade
neoplasms without a propensity for lymph node metastases, the implication from these studies is that
patients do well. If, however, an invasive component
typical of a conventional squamous cell carcinoma is
present, the possibility for more aggressive behavior
exists. Therefore, in the latter situation, the World
Health Organization [225] has recommended that
such tumors not be diagnosed as verrucous carcinoma. Due to the tendency for tumor recurrence
and potential to be associated with a conventional
squamous cell carcinoma, the treatment of choice
is complete resection, usually by total rather than
partial cystectomy. Radiation treatment is not advocated because of the possibility of development of
anaplasia within the tumor and risk of more aggressive behavior, as documented in these tumors that
occur at other sites [286,289,296-298].

g) Ancillary Diagnostic Tests: None available.


h) Prognosis: For tumors that are not associ-

[287,294]. While the tumor may occur anywhere


within the urinary bladder, one series reported that
80% occurred on the posterior wall or dome [287].
The tumor is typically exophytic, firm, white, and
wart-like.

e) Microscopic Features: Verrucous carcinoma is

a well-differentiated neoplasm, with both exophytic


and endophytic growth, characterized by hyperplastic, parakeratotic squamous epithelium that has minimal cytologic atypia and few mitoses (Figure 46).
The surface of the neoplasm projects as epithelial
papillae above the bladder mucosa. The thickened
squamous epithelium invaginates into the underlying stroma; the deep border consists of a round,
wide base of epithelium that pushes, rather than
frankly invades, into the stroma. A chronic inflammatory infiltrate may be present in the lamina propria,
adjacent to the base of the tumor. Parasite ova may
be present in the bladder tissue. Within the bladder,
verrucous carcinoma arises from areas of squamous
metaplasia; the metaplastic change develops secondary to chronic irritation most often caused by the
parasite [284,295].

f) Differential Diagnosis: A definitive diagnosis of

verrucous carcinoma, without its complete intact re-

3. Squamous Cell Carcinoma,


Schistosomal
a) Definition: Urinary bladder squamous cell carcinoma that consists almost entirely of keratin-forming
squamous carcinoma and is associated with a precedent or concurrent infection with schistosomal species.

b) Incidence: This form of squamous cell carcinoma


occurs most frequently in the Middle East and Africa
and - until recently - represented the most common
form of bladder cancer in Egypt [299-301]. Risk factors
include infection with Schistosoma hematobium [276,
277], exposure to nitrate-based fertilizers, pesticides,
and smoking. Schistosomal eggs deposited within

Figure 46. Verrucous carcinoma.

106

the bladder result in inflammation of the bladder wall


and subsequent carcinoma development.

rarity of these tumors. In larger series, adenocarcinomas constituted 0.55% to 2.6% of all urinary bladder malignancies [302-306], although these data are
hampered by the inclusion of cases of urachal carcinoma. Lack of histological review is also likely to
have a confounding effect, as in one series 58.7%
reported cases of bladder adenocarcinoma were
found, on review, to be other forms of malignancy
[307]. In particular this included urachal adenocarcinomas, urothelial carcinomas with focal glandular
differentiation and metastatic extravesical adenocarcinomas.

c) Clinical Features: The male to female ratio is


slightly higher than non-Schistosomal squamous
carcinoma at 4-5:1, which has been attributed to
the increased exposure of men to infested water
while laboring outdoors. In these endemic areas,
squamous carcinoma often develops decades earlier
than in Western countries.

d) Gross Appearance: Similar to its non-Schis-

tosomal counterpart, squamous cell carcinoma is


frequently solitary and fungating with a white, flaky
surface secondary to keratin formation. Ten to 20
percent of cases may be multifocal.

There is an association between schistosomiasis


and bladder adenocarcinoma and in cases from
Egypt, 5.2% to 11.4% of bladder tumors were found
to be adenocarcinomas [308-310]. An increased
incidence of these tumors is also seen in individuals
with bladder extrophy [310,311].

e) Histologic Features: Schistosomal-associated

squamous cell carcinoma is morphologically similar


to non-Schistosomal disease, with well- to moderately-differentiated carcinoma comprising the majority
of cases. In many instances, calcified Schistosomal
eggs may be found within the bladder wall and show
an associated granulomatous reaction. Superficial
changes include squamous metaplasia in 80% of
cases and squamous carcinoma in situ in 50% of
cases [295].

c) Clinical Features:

Adenocarcinoma of the
bladder occurs most frequently in the fifth to seventh decades, although rare pediatric cases have
been reported [312]. In registry-based studies that
focus on adults and in larger clinical series, patients
ranged in age from 15 to 93 years with mean ages
of 60 to 64 years [303-305,307]. In these series the
male:female ratio was 1.7:1 to 3.2:1. In schistosoma
endemic areas, patients present at a younger age
(mean age 46 to 50.4 years) with a male to female
ratio of 2.18:1 to 2.25:1[308-310,313].

f) Ancillary Diagnostic Tests: Ancillary diagnostic testing is identical to non-Schistosomal disease.

g) Differential Diagnosis: The differential diagnosis of bladder squamous cell carcinoma is similar to
non-Schistosomal disease and includes urothelial
carcinoma with extensive squamous differentiation
(divergent differentiation) or squamous cell carcinoma secondarily involving the bladder from sites such
as the cervix or anus.

Adenocarcinoma is the most common malignancy


associated with bladder extrophy and occurs in
approximately 10% of cases [314,315]. In these
patients tumor development is usually a late
phenomenon and may occur in individuals who
have undergone surgical repair at a young age
[311,316,317].Although adenocarcinomas have
been reported in association with neurogenic bladder
and chronic in-dwelling catheters, these conditions
do not appear to be predisposing factors for these
tumors [318]. There is evidence that non-urothelial
bladder malignancies occur more frequently in
bladder diverticula although, due to small sample
size, it remains uncertain if this is specific for primary
adenocarcinoma [319].

h) Prognosis: The five-year overall survival for

Schistosomal-associated squamous cell carcinoma


is approximately 50%, which reflects recent improvements in early diagnosis and peri-operative
outcomes. Tumor stage, grade and lymph node
involvement are most significantly associated with
survival outcomes. Treatment is primarily surgical in
nature with little benefit from chemotherapy or radiation therapy.

Hematuria is the most common presenting symptom


for bladder adenocarcinoma with other presenting
features being irritative bladder symptoms, flank
pain secondary to outflow obstruction, suprapubic
pain, urinary frequency, dysuria and mucosuria
[302,304,305,307,315,320,321].

III. ADENOCARCINOMA VARIANTS


1. Primnary Adenocarcinoma of
Bladder

d) Gross Appearance: At cystoscopy primary

a) Definition: Urinary bladder adenocarcinoma is a

bladder adenocarcinoma is more frequently solid or


sessile than papillary, and there is usually extensive
ulceration [312,315,322]. Tumors may be mucinous
and hemorrhage is frequently seen. There may be
associated inflammation with the tumor forming an
ill-defined mass [315]. Adenocarcinomas are solitary

primary malignancy that shows glandular differentiation, with features of enteric, mucinous or signet ring
cell carcinoma.

b) Incidence: Population based studies on the incidence of bladder adenocarcinoma are limited by the

107

in >50% of cases and while the trigone is the most


common primary site, all parts of the bladder may be
involved [302,305,307,312,315,322,323]. Infiltration
of the bladder wall and beyond is frequently detected
by physical examination at presentation and at cystoscopy.

intestinal metaplasia [20,325]. Although both


metaplastic conditions result from chronic urothelial
irritation/inflammation, more recent studies have
suggested that these have no associated risk for
the development of adenocarcinoma [20,21]. It has,
however, been noted that the epithelium adjacent to
foci of adenocarcinoma may show atypia or in situ
carcinoma, and in one report this was present in
17% of cases [323,325].

Signet ring cell carcinomas are almost always


invasive at diagnosis. Features of linitis plastica with
diffuse thickening of the bladder wall are commonly
present and the mucosa is frequently ulcerated
[312,320,324].

e) Histologic Features: Urinary bladder adeno-

carcinomas show a morphologic spectrum similar


to that of colonic adenocarcinoma (Figure 47) and
five specific morphotypes have been described
[307,325]. 1. Enteric tumors are acinar, cribriform,
villous or solid. Occasionally these tumors may contain cells that show neuroendocrine differentiation
[325]. 2. Mucinous or colloid carcinomas consist of
individual cells and cell nests within lakes of mucin.
These tumors may also contain mucin filled acini and
occasional signet ring cells may also be seen[320].
3. Signet ring cell carcinoma (Figure 48) is restricted
to those tumors that show a diffuse signet ring morphology with varying amounts of mucin and without
foci of glandular, urothelial or squamous differentiation [195,320,323,326]. 4. Adenocarcinoma not otherwise specified (NOS) is applied to tumors where
the architectural pattern does not conform to the preceding three morphotypes. 5. Mixed forms showing
two or more morphologic patterns. (Figure 49) It has
been suggested that to be included in this category,
a tumor should not exhibit >75% of one tumor morphotype [325] Rare hepatoid forms have also been
described [327].

Figure 47. Adenocarcinoma with enteric features.

In some series clear cell adenocarcinoma of


apparent Mllerian origin are included as variants of
bladder adenocarcinoma, while in other series they
are considered to be a separate entity [307]. The
features of these tumors are considered elsewhere
in this work.

Figure 48.
variant.

Adenocarcinoma,

signet-ring

cell

Of the five morphological variants of bladder


adenocarcinoma, enteric and colloid carcinoma
are the most commonly encountered morphotypes,
although in one series 42% of tumors were
classified as adenocarcinomas NOS, while only
25% of tumors had the features of enteric/mucinous
adenocarcinoma. Signet ring cell carcinomas are
rare and with less than 100 cases reported to date
these tumors constitute approximately 5% of cases
of bladder adenocarcinoma [307,323].
Various grading systems, including that of the World
Health Organization, have been employed for these
tumors. In general signet ring cell adenocarcinomas
are considered to be high grade [322,326].

Figure 49. Adenocarcinoma, enteric and mucinous


features.

Bladder adenocarcinoma is often found in


association with cystitis glandularis and urothelial

108

f) Ancillary Diagnostic Tests: Immunohistochem-

likely to have less of a mucinous component when


compared to tumors of gastrointestinal origin.

ical staining may assist in the diagnosis of primary


urinary bladder adenocarcinoma of non-urachal origin [170,328-331]. These tumors show variable expression of cytokeratin 20, E48, carcinoembryonic
antigen, LeuM1 and catenin. Positive staining for
cytokeratin 7, OC125 and Her-2/neu may occasionally be present [320,330,331]. Tumors are usually
negative for thrombomodulin, monoclonal prostate
specific antigen, prostate specific acid phosphatase
and vimentin, although occasional tumors show
positive staining for prostate antigen P501S [329],
prostate specific membrane antigen[329] and polyclonal prostate specific antigen. In bladder adenocarcinomas, bcl-2 expression has been shown to
decrease, while p53 expression increased, with increasing grade[332], and this may have prognostic
significance.

Signet ring cell adenocarcinoma may occasionally


mimic lobular carcinoma of the breast; however,
immunostaining for ER, PR and GCDFP-15 should
lead to the correct diagnosis.

h) Prognosis: Most studies relating to the outcome

of primary adenocarcinomas of the urinary bladder


are hampered by small sample size. Further confounding factors are that in many series urachal and
non-urachal carcinomas, are admixed [304-306]. In
further series adenocarcinomas, which contained
areas of typical urothelial carcinoma were included
[307]. Bladder adenocarcinomas are often of advanced stage at presentation. On imaging studies
75% of primary non-urachal adenocarcinoma were
found to have diffuse thickening of the bladder wall on
computerized tomography, while 88% had stranding
of perivesical fat. Lymphadenopathy was visible in
25% of cases and in 25% there was direct invasion of
the rectus muscle [334]. These results are reflected
in cystoscopic and gross pathological findings, with
most tumors showing infiltration into the bladder wall
at diagnosis[302,307,312,322,335]. Metastases are
present in approximately 25% of cases with secondary sites in descending order of frequency being liver,
bone, regional lymph nodes, adrenals, peritoneum
and skin [302,307,312,322,335]. Bladder adenocarcinoma has a poor prognosis with 5-year survivals of
11% to 55% being reported [242]. In various studies tumor stage at diagnosis has been shown to be
an important prognostic feature [307,315,322,325].
In a recent registry-based study of 306 patients with
bladder adenocarcinomas [303], the distribution of
cases according to pT staging category was pTis/a
0.3%, pT1 7.8%, pT2 30.4%, pT3 28.4% and pT4
33%, and cancer specific 5-year survival rates for organ-confined and extravesical disease were 82.5%
and 43.3% respectively. Comparison showed that
adenocarcinomas had the same natural history as
urothelial carcinoma following radical cystectomy,
but that patients with adenocarcinoma were usually
of a more advanced stage at the time of treatment.
In a separate study the median survival for patients
with localized tumors was 64 months, while median
survivals for patients with tumors showing regional
and distant spread were 29 and 8 months respectively [242].

g) Differential Diagnosis: The differential diagnosis of bladder adenocarcinoma includes other forms
of bladder neoplasia and secondary adenocarcinoma of the stomach, appendix, large bowel, breast,
endometrium and prostate gland.
Rare villous adenomas of the bladder may resemble
enteric type adenocarcinomas, although they lack
the degree of atypia seen in carcinoma and do
not infiltrate the bladder wall. Villous adenomas
show membranous expression of prostatic specific
membrane antigen and while this may also be seen
in mucinous and signet ring cell adenocarcinomas,
enteric adenocarcinomas usually show cytoplasmic
staining [329]. Urothelial carcinomas may show cystic
change with gland-like lumens, however, in these
tumors more typical areas of urothelial carcinoma
may be identified.
Infiltration of the bladder by prostatic adenocarcinoma may be differentiated from primary bladder
adenocarcinoma by immunohistochemistry as bladder tumors are negative for monoclonal prostate
specific antigen and prostate specific acid phosphatase [333]. Colonic adenocarcinoma infiltrating
or metastatic to the bladder may mimic bladder adenocarcinoma although careful examination will confirm that colonic metastases do not have any in situ
component. Immunohistochemical expression may
also be helpful as positive staining for cytokeratin 7
and negative staining for cytokeratin 20 is seen in
41% of bladder tumors but is not a feature of colonic
adenocarcinoma [170]. Additionally bladder adenocarcinomas often show positivity for thrombomodulin, while colon adenocarcinomas are negative and
unlike bladder tumors, often show nuclear staining
for -catenin [331].

Morphology has been associated with outcome,


with 5-year survival rates being 64% for tumors
showing mixed morphology, 55% for mucinous
adenocarcinomas, 51% for adenocarcinoma NOS,
and 27% for signet ring cell adenocarcinomas,
although these data did not achieve statistical
significance[326]. Signet ring cell carcinoma are
usually of an advanced stage at diagnosis, with mean
survivals of 9 to 12 months, and a 5-year survival of
<11% being reported [195,326,336].

Signet ring cell adenocarcinoma of the bladder may


resemble metastatic gastrointestinal tumors. These
tumors have an immunoexpression similar to enterictype bladder adenocarcinoma and further they are

109

Other features also shown to be of prognostic


significance are tumor size, histological grade and
method of treatment. It is, however, evident that
the prognostic significance of treatment modality
is stage dependent as this determines treatment
selection [322,325].

presenting signs and symptoms are abdominal


mass or pain, suprapubic mass, umbilical mass or
discharge, recurrent urinary tract infections, irritative
voiding symptoms, or urinary outflow obstruction,
while mucinuria is seen in 25% of urachal
adenocarcinomas [338,340,347].

It has been suggested that for localized tumor


observed poor outcomes following cystectomy relate
to occult spread of tumor [322,325]. This may explain
the significantly improved 5-year survival rates (58%
versus 44%) for patients treated with neoadjuvant
radiotherapy, four to ten weeks post-operatively
[337].

d) Gross Appearances: During fetal development,

the urachus contains the allantois and connects the


umbilicus to dome of the bladder. Following birth the
urachus undergoes involution to form the median
umbilical ligament [350]. In one third of individuals the
epithelial-lined urachal lumen persists as urothelium
or mucinous columnar epithelium, resulting in patent
urachus, cysts or tubular canals of varying size and
complexity within the median umbilical ligament
or bladder wall [339]. The embryogenesis of the
urachus accounts for the observation that urachal
carcinomas are seen as a mass at the bladder
dome or on the anterior surface of the bladder.
There is usually infiltration into the bladder wall
although occasionally this may be absent, with the
tumor having a pushing margin and being clearly
demarcated from adjacent bladder tissue. Tumors
are usually bulky at diagnosis with extension beyond
the bladder within the space of Retzius or through
to the anterior abdominal wall in the midline [323].
Less frequently urachal carcinomas infiltrate the
bladder mucosa [340] and may appear grossly as
ulcerated, mucinous or papillary masses [338].

2. Urachal Carcinoma
a) Definition: Urachal carcinomas are malignant

tumors arising in urachal remnants. While the great


majority of these tumors are adenocarcinomas
originating from vestigial urachal epithelium,
squamous cell carcinoma, urothelial carcinoma,
neuroendocrine carcinoma and sarcomas have also
been reported, albeit rarely [338].

b) Incidence: Although urachal remnants are

seen in up to one third of post mortems [339],


urachal malignancies are rare tumors with many
reports consisting of single cases and small
series[340,341]. In population based studies from
the Swedish Cancer Registry, the annual incidence
was 1 in 5 x 106 [342], while in Massachusetts
these tumors comprised 0.01% of all malignant
tumors[343]. Urachal carcinomas have been shown
to constitute 0.07% - 0.7% of bladder carcinomas
in North America and Europe and 0.55% - 1.2% of
bladder carcinomas in Japan [344]. In a separate
study these tumors were shown to comprise 0.17%
- 0.34% of bladder neoplasms and 20% - 39% of
all bladder adenocarcinomas [345]. More recently
urachal cancer was found to account for <0.1% of
all bladder cancers reported to the Ontario Cancer
Registry [346].

Imaging studies suggest a diagnosis of urachal


carcinoma when tumors are seen in the bladder
dome or extending to the anterior abdominal
wall in the midline [349,351]. Calcification is
frequently present within the tumor and is seen on
computerized tomography in approximately 70% of
cases [349,352].

e)

Histological Features: Because of the


morphologic overlap of vesical and urachal
adenocarcinoma, several diagnostic criteria for
urachal tumors have been proposed (Table 10). It
is considered that the Wheeler and Hill [353] and
Mostofi [323] criteria are too restrictive and likely to
exclude cases of true urachal adenocarcinoma. For
this reason the criteria of Gopalan et al. (2009) [340]
are preferred.

c) Clinical Features: There is a male predominance

for all types of urachal cancers with male to female


ratios of 1.8:1 to 3:1 being reported [338,341,347].
In various series, patients ranged in age from 16 to
87 years and the mean age ranged from 50 to 60.7
year[338,340,341,345,347-349]. Approximately 70%
of patients were aged between 41 and 70 years at
diagnosis [338,345].

The
majority
of
urachal
carcinoma
are
adenocarcinoma [325,341,347] and closely resemble
colonic adenocarcinoma. Many of the tumors have
a pronounced mucinous component and this has
been considered as a separate subtype consisting
of single tumor cells or cell nests within lakes of
extracellular mucin [325]. Less frequently the tumor
has a signet ring cell morphology [325,341,347].

For urachal adenocarcinomas the male to female


ratio ranged from 1.35:1 to 3:1[338,346]. Greater
than 70% of patients were aged 41 to 70 years,
while 20% of patients were less than 40 years of
age at diagnosis[338,345]. In a further series 20% of
patients who presented with urachal adenocarcinoma
were greater than 70 years of age[346].

Rare carcinomas have features other than


those of typical adenocarcinoma and cases of
squamous cell carcinoma [341,347,354,355],
urothelial carcinoma[347,355], lymphoepithelial-like

Hematuria is the most common presenting feature


in 55% to 80% of cases [340,344,345,347]. Other

110

carcinoma[341], clear cell carcinoma[341] and small


cell/neuroendocrine carcinoma[347,356,357] have
been described. On occasion more typical urachal
adenocarcinomas may contain focal areas of these
rare forms of carcinoma. Several cases of urachal
sarcoma with features resembling leiomyosarcoma,
rhabdomyosarcoma and hemangiopericytoma have
also been reported [338,358-360].

Extravesical genital tract carcinomas may be


differentiated from urachal adenocarcinomas by
prostate specific antigen, vimentin and CA125
immunohistochemistry [330,362]. Vesical urothelial
carcinomas are usually negative for carcinoembryonic
antigen.
There is often an overlap in the morphologic features
of vesical, colonic and urachal adenocarcinoma and
clinicopathologic correlation, with application of the
defining criteria for urachal carcinoma (Table 10),
usually leads to the correct diagnosis. Each of
these primary tumors may show positive staining
with carcinoembryonic antigen, however, diffuse
expression favors a urachal origin [330]. Unlike
urachal adenocarcinoma, colonic tumors are likely
to show diffuse cytokeratin 20 staining and are
cytokeratin 7 negative.

In larger series tumor grading is reported, although


no formal grading system for these tumors has been
established. In these series tumors are subjectively
graded as well differentiated/poorly differentiated
[347] or well/moderately/poorly differentiated [346].

f) Ancillary Diagnostic Tests:

Studies on
the immunohistochemical expression of urachal
carcinomas are limited [361]. These tumors are
usually positive for carcinoembryonic antigen,
7E12H12 IgM isotype, CDX2 and Leu-M1, while
focal positivity has been reported for cytokeratin 7,
cytokeratin 20, 34E12 and E48 antibody. Stains
for vimentin, Her-2/neu and prostate specific
antigen are negative. Cytoplasmic -catenin and
CA19-9 positivity is seen in the majority of urachal
adenocarcinomas, while CA125 positivity is only
rarely present [330,340,362].

h) Prognosis: In earlier series the prognosis of


urachal carcinoma was found to be poor with overall
survivals of 22.7% and 31%, and a 5 year survival
of 61% being reported [344,363]. In more recent
studies mean overall survival was 122 months[9]
and a meta-analysis of 312 patients showed 5 year
post-operative cancer free survivals to range from
43% to 70%[325,341,345-348,363-367].

A staging system for urachal carcinoma was


established by Sheldon et al [345]. This was
hampered by a relatively large number of categories
and a simplified staging classification has been
proposed [347,365] (Table 11). Advanced tumor

g)

Differential Diagnosis: The differential


diagnosis includes primary adenocarcinoma of the
bladder and metastatic colonic, ovarian and prostatic
carcinomas.

Table 10. Criteria for diagnosis of urachal adenocarcinoma


Wheeler and Hill, 1954

1.
2.
3.

Tumor in dome of bladder.


Absence of cystitis cystica or cystitis glandularis.
Predominant involvement of muscularis rather than submucosa, with
intact or ulcerative bladder epithelium.

4.
5.

Urachal remnant connected to neoplasm.


Suprapubic neoplastic mass.

1.
2.

Criteria additional to Wheeler and Hill


Location in anterior wall of bladder.
Sharp demarcation of tumor and surface epithelium.

3.

Extension to space of Retzius, to umbilicus or anterior abdominal wall.

1.
2.

Tumor in bladder dome.


Sharp demarcation between tumor and bladder epithelium.

3.

Exclusion of another primary (non-urachal) adenocarcinoma.

1.
2.
3.

Tumor in dome/anterior wall.


Epicenter of carcinoma in bladder.
Absence of widespread cystitis cystica/glandularis beyond the dome/
anterior bladder.
No known primary elsewhere.

Mostofi et al, 1955

Johnson et al, 1985

Gopalan et al, 2009

4.

111

3. Clear Cell Adenocarcinoma

stage is associated with a less favorable outcome,


with reported 5 year survivals of 93% for patients
with tumor confined to the urachus and bladder,
69% for extra-vesical and peri-urachal tumors and
0% for tumors within the peritoneal cavity [341,347].
In a further series all patients who developed local
recurrence or metastases were Sheldon stage T3 or
higher [340]. Other features associated with poor
survival are high tumor grade and positive surgical
margins [341,346,347].

a) Definition: This designation should be reserved

for tumors resembling to a significant degree clear cell


adenocarcinoma of mllerian type as encountered
in the female genital tract and should have one or
more of the typical tubulocystic, solid and papillary
patterns as well as one or more of the typical cell
types including clear cell, non-specific cuboidal and
hobnail [205,377].

b) Incidence: These are rare neoplasms and are of

Successful management of urachal carcinoma


is dependent on complete surgical excision of
tumor. This involves partial or radical cystectomy,
removal of the medial umbilical ligament/urachus
and resection of the umbilicus, and in recent
years laparoscopic surgical techniques have been
described [368-370]. Recurrence are more likely
following partial cystectomy [347,371,372] and
common metastatic sites are liver, lung, bone and
peritoneum [340,341,343-347], while metastasis to
the brain and skin are less common [347,373,374].
For patients who developed local recurrence or
metastases, time to relapse is usually short, ranging
from 10 to 22 months [340,341]. Salvage surgery
has been shown to result in a long term cure for 50%
of patients with local recurrence [347].

two fundamental types. Those that arise in females


from endometriosis and accordingly histogenetically
as well as morphologically are identical to mllerian
clear cell carcinoma. The second group includes
neoplasms without a proven association with
endometriosis but with morphologic similarity to
tumors having such an association. Even when both
subsets are combined this is amongst the rarest of
all bladder cancers accounting for no more than
0.01% of cases.

c) Clinical: The great majority (approximately 80%)

of these tumors have occurred in females. There is


a wide age distribution in adult life from the early
twenties to late years but the majority occur in a
somewhat old population (mean 57 years). There is
no unique aspect to the clinical presentation amongst
the reported cases although it is conceivable that
an association in a female of symptoms with the
menstrual cycle could indicate an endometriosis
associated carcinoma.

Adjuvant chemotherapy currently has a limited


efficacy for urachal adenocarcinoma. Although tumor
regression has been reported following chemotherapy
for a few cases [339,340,346,351,375], it has been
noted that this does not alter survival interval for
patients with metastatic disease [347]. Despite this,
it has recently been suggested that 5-fluorouracilbased chemotherapy may have some efficacy
[376].

d) Gross Appearance: Most of the reported

tumors have been polypoid to papillary. An entirely


mural neoplasm, although not reported, could be
encountered due to the origin of some cases in
endometriosis within the muscularis propria of the

Table 11. Staging systems for urachal carcinoma.


1.

2.

Sheldon staging classification


Stage I
Stage II
Stage III
IIIA
IIIB
IIIC
IIID
Stage IV
IVA
IV

Confined to urachal mucosa.


Confined to urachus.
Extension to bladder.
Peri-urachal and vesical fat invasion.
Extension to peritoneum.
Extension to viscera other than bladder
Metastases to regional lymph nodes.
Metastases to other organs.

Mayo staging classification


Stage I
Stage II
Stage III
Stage IV

Confined to urachus and bladder.


Beyond muscular layer of urachus and/or bladder.
Infiltration of regional lymph nodes.
Infiltration of non-regional lymph nodes or distant sites.

112

bladder. The reported tumors have ranged up to 7


cm in greatest dimension.

epithelium may rarely be difficult to distinguish from


dilated glands of associated endometriosis should
the specimen be from a female patient, particularly
if the patient is postmenopausal and atrophy of
the stromal component of endometriosis would be
expected.

e) Microscopic Features: The typical triad is

that of tubulocystic, papillary and diffuse (solid)


arrangements. The commonest is tubulocystic,
the lumens of these formations often containing
basophilic or eosinophilic secretion that may be
mucin positive (Figure 50). The papillae are generally
small and rounded but may be large and delicately
filiform. Rarely they are hyalinized. In tumors with a
diffuse growth the cells may have conspicuous clear
cytoplasm and appreciable clear cytoplasm may
be a feature of the cells lining tubules and cysts or
forming papillae although generally the cells of these
structures have non-specific lightly eosinophilic
cytoplasm of modest amount. A variety of other
rare patterns may be encountered including solid
tubular formations. Some of the cells lining tubules
and cysts may have the features of hobnail cells but
they are generally predominantly non-specific and
cuboidal in nature [205,377].

The differential of clear cell carcinoma also includes


a urothelial (transitional) carcinoma with glandular
differentiation in which the latter resembles to
a degree the pattern of clear cell carcinoma.
Thorough sampling can be crucial in this distinction
and in biopsy or curettage material it may not be
possible to make the distinction but it rarely has
therapeutic implications. Spread to the bladder from
a clear cell carcinoma of the genital tract in females
is excluded primarily by clinical findings. There is
little resemblance between clear cell carcinoma and
metastatic renal cell carcinoma despite the shared
population of clear cells in many cases.

g) Ancillary Diagnostic Tests: These are

rarely needed to establish the diagnosis but


some information potentially of benefit has
accrued, albeit overall knowledge on this aspect is
limited. Most of the reported tumors have stained
immunohistochemically for CA-125 and CK7. CK20
has usually been only focally and weakly positive
although it was strongly positive in one tumor.

h) Prognosis: The prognostic parameters are

similar to those of conventional carcinoma of the


bladder and standard parameters such as invasion
of the lamina propria, and its extent, and even more
so invasion of the muscularis propria are guides to
both therapy and prognosis. There is no evidence
that the behavior, stage for stage, is different from
usual bladder cancer.

Figure 50. Clear cell variant urothelial carcinoma,


tubulocystic pattern. This neoplasm, from a female,
was associated with endometriosis.

IV. NEUROENDOCRINE NEOPLASMS

f) Differential Diagnosis: The major differential

1. Neuroendocrine Carcinoma of The


Bladder

diagnosis is with nephrogenic adenoma because the


tubular and papillary patterns of clear cell carcinoma
may occasionally be reminiscent of similar patterns
of nephrogenic adenoma. It is helpful in many cases
that the gross characteristics are highly suspicious for
a neoplasm although it should be noted that a mass
lesion may be seen in some cases of nephrogenic
adenoma. The typical tiny tubules of nephrogenic
adenoma usually are of much smaller caliber than
the great majority of tubular formations in clear cell
carcinoma. Treacherously, hobnail cells may be
seen in both the benign and malignant processes.
Clear cells are only rarely a feature of nephrogenic
adenoma. There is generally much greater cytologic
atypia and in particular more conspicuous mitotic
activity in cases of clear cell carcinoma than ever
seen in nephrogenic adenoma.

a) Definition: A group of tumors comprising small


cell carcinoma and large cell neuroendocrine
carcinoma and mixed patterns [378].

b) Incidence: Large cell neuroendocrine carcinoma

of the urinary bladder is extremely rare with less than


ten cases reported in the literature [379]. Small cell
carcinoma of the urinary bladder, although relatively
much more common than both carcinoid and large
cell neuroendocrine carcinoma, is an extremely rare
primary bladder malignancy, accounting for less than
1% of urinary bladder cancers [120,380-382].

c) Clinical: Although there are no more than single

cases of primary large cell neuroendocrine carcinoma


of the urinary bladder, the age at diagnosis ranges
from 32 to 82 years, and patients may present with

Dilated cysts lined by compressed neoplastic

113

symptoms similar to those of carcinoid tumor [383385]. Small cell carcinomas mostly occur in the sixth
to seventh decade of life, the mean age being 66
years at the time of diagnosis (range 36 to 85 years),
and have a male preponderance [386]. Patients
usually present with hematuria, which may be
accompanied by dysuria, frequency, nocturia, urinary
obstructive symptoms, or localized abdominal/
pelvic pain [380,386]. Small cell carcinomas arise
most frequently in the lateral walls and dome of the
bladder, and rarely in bladder diverticula [119,120].

be mistaken for malignant lymphoma, poorly differentiated urothelial carcinoma, and even exuberant
inflammation in the background of crush or electrocautery thermal artifact, especially in a superficial or
scant specimen. Pulmonary metastasis or extension
from adjacent viscera must be ruled out by clinicoradiologic correlation, as immunohistochemistry offers
limited assistance in this setting. An additional rare
but significant consideration is alveolar rhabdomyosarcoma, a subset of which can impart a primitive
round cell appearance, imparting a histology reminiscent of small cell carcinoma [120].

d) Gross Appearance: The few reported cases of

g) Ancillary Diagnostic Tests: Tumor cells of

large cell neuroendocrine carcinomas have not been


associated with any one single modality of growth.
Small cell carcinoma may manifest as a single large,
solid, polypoid, sessile or ulcerated mass, which may
be infiltrative at presentation [119,120,387].

large cell neuroendocrine carcinoma are immunopositive for neuroendocrine markers chromogranin A,
synaptophysin, CD56, and neuron-specific enolase,
and for cytokeratins CAM 5.2 and AE1/AE3, and
epithelial membrane antigen. Small cell carcinoma
tumor cells are immunopositive for chromogranin A
and synaptophysin (greater than 60% of cases) and
frequently demonstrate a dot-like immunopositivity
for pan-cytokeratin. TTF-1 may be immunopositive in
up to 40% of cases [388]. Unlike neuroendocrine tumors, lymphomas are immunopositive for leukocyte
common antigen and negative for keratin and neuroendocrine markers. Poorly differentiated urothelial
carcinoma does not express neuroendocrine markers such as chromogranin or synaptophysin, thereby
differentiating it from small cell carcinoma. Prostatic
adenocarcinoma, unlike small cell carcinoma, would
be immunopositive for prostate-specific markers,
prostate specific antigen and prostatic acid phosphatase. Lastly, alveolar rhabdomyosarcoma with
small cell differentiation would be immunopositive
for desmin, myogenin and Myo D1 and negative for
keratin. Interestingly, this tumor is associated with
immunopositivity for the neuroendocrine marker synaptophysin, necessitating the employment of a panel
of immunohistochemical markers.

e) Microscopic Features: Large cell neuroendo-

crine carcinoma is poorly differentiated and highgrade. At low magnification, it demonstrates a noticeable histologic pattern of growth in the form of nests
and trabeculae. Tumor cells have irregular angulated nuclear outlines and prominent nucleoli. Pure
small cell carcinomas have a patternless pattern of
growth, meaning at low magnification, the tumor cells
grow in a diffuse fashion with no clear organization.
As an exception however, one may find focal nesting
pattern. Tumor cells have very little cytoplasm; as a
result, they show nuclear crowding and molding as
though being pushed in to one another (Figure 51).
Nucleoli are inconspicuous and the chromatin is
evenly dispersed. Small cell carcinomas carry a high
proliferation rate, as evidenced by frequent mitoses,
crush artifact, geographic necrosis, and the subsequent Azzopardi effect, whereby blood vessel walls
undergo encrustation by excess liberated DNA.

f) Differential Diagnosis: Large cell neuroen-

docrine carcinoma, also due to its exquisite rarity,


ought not be confused with a metastatic deposit of
its pulmonary counterpart. Small cell carcinoma may

h) Prognosis: Large cell neuroendocrine carcino-

ma appears to behave by way of small cell carcinoma, being aggressive with high metastatic potential, with most reported cases having a fatal outcome
[383,384]. As with primary pure carcinoid tumors of
the urinary bladder, large studies have not been possible due to the limited number of reported cases.
Small cell carcinoma of the urinary bladder often
presents at an advanced stage, with up to 94% of
cases having muscularis propria invasion or extravesical extension. Metastasis at time of presentation is not uncommon, to sites that include regional
lymph nodes, liver, bone, and lung. Treatment recommendations for small cell carcinoma of the urinary
bladder have been based on case reports and small
retrospective series. Cisplatin and etoposide-based
chemotherapy in conjunction with surgery has been
associated with significantly improved survival, including reports of long-term survivors. Various clini-

Figure 51. Small cell carcinoma with squamous


differentiation.

114

f) Differential Diagnosis: Carcinoid tumors can be

cal studies have addressed patient management by


way of different combinations and modalities of surgery, chemotherapy and radiation therapy, making it
exceedingly difficult to correlate outcome between
the different series. Mean survival ranges from 6 to
34.9 months and the reported 5-year survival rate
ranges from 8 to 40%. Organ-confined disease is
more amenable to therapy and is associated with
more favorable patient survival. In sum, prognosis
is influenced by disease extent at diagnosis, employment of chemotherapy, and the patients performance status [386,387,389].

histologically confused with nested variant of urothelial carcinoma, inverted urothelial papilloma, and
metastatic tumors arising in the prostate, gastrointestinal tract and lung. The rarity of carcinoid tumors
of the urinary bladder coupled with the prognostic
significance of its differentials makes it paramount
to employ ancillary studies and clinicoradiologic correlation as part of case work-up.

g) Ancillary Diagnostic Tests: Carcinoid tumor

cells are immunopositive for neuroendocrine markers chromogranin, synaptophysin, serotonin, and
neuron-specific enolase, and for cytokeratin AE1/
AE3.

2. Carcinoid Tumor
a) Definition: A well-differentiated neuroendocrine

h) Prognosis: In the largest series to date, all 6

tumor similar to that seen in other organs which is


part of a very broad spectrum neuroendocrine tumors in the bladder [378].

pure carcinoid tumors with similar morphology had


excellent prognosis [390]. However, the number of
cases reported is still relatively small and the tumors
likely had been completely removed by biopsies.
Pure primary carcinoid tumor of the bladder and
prostatic urethra, when presenting as small polypoid nodules in the bladder is associated with a very
favorable outcome. However, there are reports of
larger more invasive lesion with regional lymph node
or distant metastases [119,392]. Some of the early
reported cases with aggressive behavior were found
to be mixed tumors with associated small cell carcinoma or adenocarcinoma components.

b) Incidence: Pure carcinoid tumor of the urinary


bladder are extremely rare. By way of criteria applied
to their more common pulmonary counterparts, less
than ten cases have been reported in the literature
[390,391].

c) Clinical: Carcinoid tumors of the urinary bladder

have been reported to occur solely in adults ranging from 29 to 75 years of age, with a slight male
preponderance. These patients may present with
hematuria and irritative voiding symptoms.

d) Gross Appearance: Carcinoid tumors usually

E. ROLE OF
IMMUNOHISTOCHEMISTRY
AND ANCILLARY DIAGNOSTICS
IN HE DIAGNOSIS, STAGING,
PROGNOSTICATION AND
PREDICTION OF BLADDER CANCER
IN CONTEMPORARY CLINICAL
PRACTICE

present as small polypoid masses at the bladder


neck/trigone[390].

e) Microscopic Features: Microscopically, carcinoid tumors are submucosal and confined within
the lamina propria, often associated with adjacent
cystitis cystica et glandularis. The tumors are composed of uniform, cuboidal, or columnar cells with
finely stippled chromatin and inconspicuous nucleoli
in a prominent pseudoglandular pattern composed
of acinar and cribriform structures (Figure 52).

I. Immunohistochemical
Markers for Staging of
Bladder Cancer
1. Introduction
Pathologic staging of bladder cancer identified
in transurethral resections or biopsies is the
clinically most powerful determinant with regard
to treatment decision, in addition to grading[393].
Although grading is the strongest prognosticator
for progression of bladder cancer in the subset of
non-invasive (pTa) bladder cancers, grading tends
to lose its independent prognostic value in invasive
bladder cancers[115]. Pathologic staging depends
on the morphologic recognition of invasion beyond
the basement membrane into the lamina propria,

Figure 52. Carcinoid tumor involving the bladder.

115

while invasion of the detrusor muscle serves as a


landmark for T2 bladder cancer.[393] The diagnosis
of muscularis propria invasion generally leads to the
clinical decision to perform a major surgery, including
cystectomy or nephroureterectomy. Histopathological
assessment of invasion of the lamina propria may be
hampered by several factors, which include artefacts
related to the procurement of the tissue and the
morphologic variation of bladder cancers.[393]
Indeed, sometimes it is impossible for a pathologist
to determine whether a bladder cancer is invasive.
Artifactual confounders are the occurrence of severe
cautery and crush effects related to the urological
procedure and the lack of spatial orientation due to the
random embedding of the bladder tissue chips. If the
biopsy or transurethral resection is very superficial,
underlying stroma may be too scarce or even
absent, again precluding the staging of the tumour
sample. Biological tumor-associated factors that
may preclude a proper assessment of pathological
stage are the presence of an inverted growth
pattern, which is more commonly associated with a
trabecular or even pushing invasive growth pattern,
and the occurrence of a florid desmoplastic reaction,
which in particular could hamper the identification
of the detrusor muscle. Further, the variation in
thickness of the muscularis mucosae, which are
normally represented by thin wisps of smooth
muscle fibers may lead to an erroneous assignment
of T2 bladder cancer, particularly when this layer
becomes hypertrophic in appearance. At various
sites in the bladder (e.g. the trigone, the bladder
neck, or the renal calyces), the detrusor muscle may
be located superficial, and as a consequence, its
invasion may spuriously be deemed as invasion of
the muscularis mucosae, resulting in under-staging
of the bladder cancer. Finally, in resected cancers
of bladder diverticula (in fact pseudo-diverticula) it
is generally impossible to distinguish pT1 from pT2
or even pT3 stages as this site generally lacks the
landmark detrusor muscle.

cytokeratin antibodies can facilitate the diagnosis


of invasion in cases with prominent inflammation
obscuring the interface between epithelium and
stroma. The most commonly employed antibodies
for this purpose are pan-keratins, like AE1 / AE3 and
antibodies against cytokeratins 8 and 18[396]. (Table
13) Also, in case of cautery artefacts or as a sequel
to a previous urological procedure, cytokeratin
staining may allow the identification of carcinoma
cells infiltrating the (cauterized) stroma[393,397].
A potential pitfall, particularly in the setting of a
previous urological procedure such as a transurethral
resection, is the positive staining for cytokeratins of
myofibroblasts and smooth muscle cells[397]. Both
myofibroblasts and smooth muscle cells generally
express only low-molecular weight cytokertins,
while epithelial cells will show immunorecativity
to high molecular weight forms. Most importantly,
a correlation of the cytologic appearance of the
cytokeratin positive cells is essential in order to
obtain a correct diagnostic interpretation.
Table 12. Interobserver variation for staging of
bladder cancers.

Author
(reference)

Down staging pT1 Upstaging


-- pTa
pT1 pT2

Abel, 1988

13%

1%

Van der
Meijden, 2000

53%

10%

Bol, 2003

56%

13%

Tosoni et al,
2000

35%

3%

Experimental studies have shown that tumour


invasion is associated with dysregulation of both
adhesion molecules, involved in cell-cell interactions
and cell-basement membrane structures like
hemidesmosomes, and endopeptidases with the
capacity to de-grade extracellular matrix proteins
[398]. Cadherins are a family of cell-cell adhesion
molecules whose expression is altered during
the epithelial mesenchymal transition which is
characteristic of tumour invasion [399,400]. Whereas
carcinoma in situ of the urothelium expresses high
levels of E-cadherin, its expression is reduced, and
a switch to N-cadherin and in particular P-cadherin
expression has been reported in invasive bladder
cancers [399]. Negative markers such as E-cadherin
are obviously not suitable for diagnostic use to
detect invasion, but markers which are selectively
up-regulated in invasive cells may hold diagnostic
promise. P-cadherin expression was shown to be
elevated in particular in the invading edge, but this
molecule is also predominantly expressed in the

A few studies have demonstrated a substantial level


of inter-observer variation among pathologists for
staging of bladder cancers (Table 12)[33,394,395].
For this reason, several attempts have been made
to identify easily applicable histological markers that
would increase the accuracy of pathological staging
of cancers of the urinary tract.

2. Determination of invasiveness of
bladder cancer
The identification of invasive bladder cancer cells
can be challenging, especially when the number of
invading cells is small or in the presence of cautery
or mechanical artefacts. In addition, some tumours
may show pushing growth deep in the lamina propria
without an obvious invasive pattern. Although no
formal studies have been reported in the literature,
it is well-established that immunostaining with anti-

116

Table 13. Most commonly employed immunohistochemical markers for staging of bladder cancers

Pathological issue

Immunohistochemical marker

Caveats

Detection of micro-invasion
into cauterized or damaged
tissue

Pan-keratin, keratin 8, 18

Keratin positive myofibroblasts

Residual cancer cells postTUR

Pan-keratin

Keratin positive myofibroblasts

pT1 substaging

Desmin, pan-keratin

pT1 versus pT2 staging

Smoothelin, vimentin

basal layers of the non-invasive bladder cancers


[398], precluding its use as a diagnostic adjunct for
invasion.

heterogeneous, particularly with regard to risk of


progression and death of disease during followup. Several attempts have been made to substage
T1 bladder cancers. The most often employed
substaging system employs the muscularis mucosae
and the associated plexus venosus in the lamina
propria as reference point, distinguishing a substage
pT1a, pT1b and pT1c [44]. pT1a cancers are the
most superficial bladder cancers, with invasion above
the muscularis mucosae, while pT1c cancers would
invade beyond the plexus venosus. A condensed and
probably more robust substaging system aggregates
pT1a and pT1b cancers in the substage of minimally
invasive bladder cancers [40]. Several studies
reported a considerable inter-observer variation for
this substaging system, while it was also noted that
in a variable proportion of cases pathologists were
not able to assign a substage due to absence of
the muscularis mucosae / plexus venosus and poor
orientation. As a consequence, pT1 substaging has
not been included in the WHO 2010 classification
of bladder cancers. One attempt was made to use
immunohistochemistry for improvement in the
accuracy of pT1 substaging (18), using antibodies
against desmin and pan-keratin; in a series of 93
consecutive pT1 bladder cancers, they observed
that in a small proportion of cases immunostaining
could eliminate the disagreement with regard to pT1
substage among two pathologists when reviewing
H&E stained slides [406]. Thus, it was concluded that
the use of desmin and cytokeratin immunostaining
could be of help to improve the precision of
substaging pT1 bladder cancers; however, there is
no data to show that using immunohistochemistry
for substaging improves the prediction of recurrent
or progressive disease and therefore it is not
recommended for routine application.

Among the extracellular matrix degrading endopeptidases, collagenase-3 (matrix metalloproteinase-13)


and stromelysin-3 have been studied in relatively
small series of bladder cancers for their diagnostic
and prognostic use [401-403]. MMP-13 was expressed predominantly at the invading edge and
invasive nests of about 60% of the invasive carcinomas. Further, a weak expression was also noted
in the stromal fibroblasts of some of the invading
cancers [403]. Given the low sensitivity of MMP-13,
this marker does not seem useful in diagnostics of
bladder cancer invasiveness. Stromelysin expression was particularly noted in the stromal fibroblasts
adjacent to the invasive carcinoma nests of about
70% of carcinomas, but not in the carcinoma cells.
However, a small proportion of non-invasive bladder
cancers were also positive for this marker, diminishing its diagnostic significance [401,402]. Of interest,
stromelysin expression showed a statistically significant correlation with lymphatic vessel invasion.
Further, some studies addressed the potential of
immunostaining for basement membrane proteins as
an adjunct to the morphologic detection of invasive
cells in bladder cancer. Laminins are glycoproteins
which represent a major component of the basement
membrane and they consist of three subunits, i.e. an
, and chain. The LN-5 chain is a specific target
of MMP-2, and this cleavage is critical for tumor
invasion and tissue remodeling. It was reported
that expression of LN-5 2 is increased in the
carcinoma cells at the invasive edge of the bladder
cancers [404,405]. Although LN-5 2 was strongly
and independently associated with unfavourable
prognosis, only about one third of the invasive
carcinomas expressed this marker as compared
to 7% of the non-invasive bladder cancers, which
renders LN-5 2 unsuitable as a diagnostic marker.

4. Immunohistochemical markers for


identification of detrusor muscle
Particularly in transurethral resections of the bladder,
the distinction between T1 and T2 bladder cancer may
be challenging. For this reason immunohistochemical
markers have been employed to help discriminate
muscularis mucosae from muscularis propria.

3. Immunohistochemical markers for


substaging pT1 bladder cancers
Stage

pT1

bladder

cancers

are

Overlap in staining intensity of


muscularis mucosae and detrusor
muscle

clinically

117

Initially, the utility of desmin and/or smooth muscle


type actin in the identification of the detrusor muscle
was evaluated, but these markers do not allow a
differential staining of the muscularis mucosae and
muscularis propria. Histochemical stains such as
hematoxylin-phloxine-saffron or Mallory trichrome
have been used to identify muscle fibers in areas
of desmoplastic stroma or cautery artifact. Although
many pathologists do utilize such stains, there
are no published series reporting on use of any of
these markers in a routine pathology setting. There
is no good data on the value or the validity of this
approach; therefore, these cannot be recommended
for routine use.

and vimentin could help distinguish detrusor


muscle from muscularis mucosae. Two subsequent
studies addressed the question of whether
immunohistochemistry with smoothelin would help
resolve pathologic staging difficulty in problematic
transurethral resection specimens [412,413]. Both
studies confirmed the diagnostic utility of this marker.
In the first study, 4 cases with equivocal muscularis
propria on standard H&E stains were examined and
in the latter a series of 34 specimens submitted
for consultation because of questions relating to
pathological stage were stained for smoothelin. This
publication mentioned that the staining intensity of
blood vessel wall was similar to that of the detrusor
muscle, implying that the vessel wall musculature
could be used as an internal reference [412]. The
latter publication noted occasional overlap in
intensity of smoothelin staining for the two muscle
layers, which might lead to spurious pathological
staging results [413]. Since all studies to date have
demonstrated that immunostaining for smoothelin
with clone R4A monoclonal antibody can help
differentiate between muscularis mucosae and
detrusor muscle, this marker may be employed in
controversial cases as a diagnostic tool. Since a
minority of samples displayed a similar intensity of
staining in both muscle types, the immunostaining
should always be used in conjunction with the H&E
findings. Further, vimentin immunostaining may be of
additional help in distinguishing muscularis mucosae
from detrusor muscle [411].

Smoothelin was identified in 1996 as a marker


for smooth muscle tissue, and two isoforms,
smoothelin-A (59kDa) and smoothelin-B (110-kDa)
have been described [407]. Their function is related to
the contractility of smooth muscle and its expression
is absent in non-contractile or proliferative smooth
muscle cells. (Table 13) Smoothelins are actinbinding proteins that are expressed abundantly in
terminally differentiated visceral (smoothelin-A) and
vascular (smoothelin-B) smooth muscle and both
types can be visualized by monoclonal antibody R4A.
The latter antibody is the most commonly employed
antibody for the immunohistochemical detection of
smoothelin in a diagnostic setting. Recently, Paner
et al. reported the differential staining of detrusor
muscle and muscularis mucosae by this antismoothelin antibody to help distinguish muscularis
mucosae versus detrusor muscle invasion of bladder
cancer [408]. Similarly, in the gastro-intestinal tract
smoothelin is now being studied to distinguish
between pT1 and pT2 adenocarcinomas of Barrett
esophagus [409]. After a second confirmatory
study by Paner et al[410] on a larger series of 70
TUR samples, a subsequent study by Council et al.
compared the diagnostic value of smoothelin with
that of smooth muscle actin, desmin, caldesmon,
and vimentin on a series of 15 cystectomies [411].
In contrast to the similar intense expression levels
of muscularis mucosae and detrusor muscle for
smooth muscle actin, desmin, and the variable
expression levels of caldesmon, smoothelin and
vimentin displayed a differential expression pattern.
Smoothelin was intensely positive for all detrusor
muscle samples and negative or weakly positive for
muscularis mucosae, whereas vimentin was rarely
expressed in the detrusor muscle and positive in 9 of
11 specimens with muscularis mucosae. Council et
al. also noted that antibodies specific for caldesmon
and desmin and to a lesser extent - smoothelin
were able to selectively stain smooth muscle cells
as compared to reactive myofibroblasts [411]. These
authors recommended the use of caldesmon and
desmin immunohistochemistry to distinguish reactive
myofibroblasts, e.g. in desmoplasia from smooth
muscle cells, while a combination of smoothelin

5. Conclusions
Although these markers have been shown to have
promise, the studies are limited and from only a
few institutions. Substantially more data is needed
before these could be recommended for routine
application.
Level of evidence: 4
Recommendation: C

II. Summary of molecular/


genetic alterations in
urothelial carcinoma
1. Introduction
Despite the extensive new information and
knowledge on bladder cancer genetics, epigenetics
and gene expression, much is needed to satisfy
several unanswered questions.
Much of the work is following, and pointing toward,
two distinct pathways that correspond to two main
groups of tumors; superficial (papillary and flat) and
muscularis propria-invasive urothelial carcinoma.
A major question is whether the existing molecular
information can explain tumorigenesis along the

118

Table 14. Common genetic and epigenetic aberrations in urothelial carcinoma


Gene

Location

Deletion/Mutation

12q14.3-q15

Tumor
Suppressor
Oncogene

Clinicopathological Diagnostic Prognostic


Association
Role
Role
HGUC
Y/N
Y/N

Amplification

HGUC

HRAS*

11p15

Oncogene

FGFR3

4p16.3

Oncogene

E2F3

6p22

CCND1

TP53

17p13.1

MDM2

Role

Abnormality

Y/N

Oncogene

Activating Mutations All grades and


stages, up to 15%
Activating Mutations Predominantly lowgrade/stage
Amplification
Invasive HGUC

11q13

Oncogene

Amplification

CDKN2A

9p21

Tumor
Suppressor

Deletion/Mutation/
Methylation

RB1

13q14.2

Deletion/Mutation

Y/N

ERBB2

17q21-q22

Tumor
Suppressor
Oncogene

All grades and


stages, up to 20%
All stages and
grades, primarily
HGUC
Invasive HGUC

Amplification

Invasive HGUC

MYC

8q24.21

Oncogene

Amplification

Invasive HGUC

PTEN

10q23.3

Deletion/Mutation

HGUC, all stages

RASSF1A

3p21.3

Tumor
Suppressor
Tumor
Suppressor

FHIT

3p14.2

Y/N

PTCH

9q22.3

DBC1

9q32-33

TSC1

9q34

Tumor
Suppressor
Tumor
Suppressor
Tumor
Suppressor
Tumor
Suppressor

Methylation

Invasive HGUC,
associated with
progression
Deletion/Methylation HGUC, associated
with progression
Deletion/Mutation
All grades and
stages
Deletion/Methylation All grades and
stages
Deletion/Mutation
All grades and
stages

proposed two major tumor groups in regard to the


molecular events and of the biochemical signaling
pathways involved. Another equally important
question is whether currently available information,
or future knowledge, can predict which superficial
bladder tumors will later progress to become
invasive. Similarly, will there be any marker that
can have prognostic significance within any given
stage of the tumor. Despite this being a goal of many
molecular studies, to date molecular information has
failed to provide robust or reliable predictive markers.
Possibly, key markers still remain to be identified.

chromosome 9 by far the most common cytogenetic


finding (refer to tables 15 and 16 for more details).

2. Molecular alterations in
urothelial carcinoma

There seems to be distinct signaling pathways in


two groups of urothelial carcinoma. The association
of FGFR3 or HRAS mutations in the majority of
superficial tumors suggests that they share changes
in pathway activation, whereas, the common
inactivation of RB1 and TP53 pathways in invasive
tumors may indicate a distinct signaling status.

In invasive urothelial carcinoma, many genetic


alterations have been reported in addition to frequent
chromosome, which involve dysregulations of several
oncogenes and tumor suppressor genes. It has been
recently shown that tumors with aberrations in p53/
MDM2, RB1 and E2F3 are associated with more
genomic instability.

3. Signalling pathways in urothelial


carcinoma

Molecular, genetic and epigenetic aberrations


commonly involved in urothelial carcinoma are
listed in Tables 15 and 16. Briefly, most studies of
low-grade papillary urothelial carcinoma show few
molecular alterations apart from deletions involving
chromosome 9 and mutations of FGFR3 and HRAS.
These tumors are often near-diploid with loss of

Activation of the PI3K/Akt pathway in urothelial


carcinoma could occur via the known mutation

119

of PTEN and TSC1. A number of therapeutic


applications targeting molecules in these pathways
have been developed and are being tested for
potential roles in treating urothelial carcinoma.

provide support for the nuclear shape and also


involved in DNA replication, in RNA transcription
and in regulation of gene expression. This protein
is reported to have a concentration as high as 25
times in urothelial carcinoma as compared to normal
urothelial cells. This assay is approved for both
the detection of new cancers and the follow-up of
patients with a prior history of urothelial carcinoma.
The reported sensitivity ranges are 34.6%100%,
and 49.5%65.0%, but false positive results have
been reported.

4. Biomarkers with diagnostic/


prognostic values in urine
samples
a) BTA tests
This test is based on the detection of the human
complement factor H-related protein, which is
reported to be expressed only in bladder tumor cells.
The sensitivity of the BTA stat is reported to be 50%
for low grade urothelial carcinomas, which is higher
than cytology. Conversely, the specificity of BTA stat
is reportedly lower than cytology.

c) B
 ladder Cancer Immunofluorescence Assay
This is an immunofluorescence assay designed
to improve the sensitivity of urine cytology. It
employs a cocktail of three monoclonal antibodies;
M344, LDQ10 and 19A211. The first two detect a
mucin-like antigen, while the third one recognizes
a high molecular weight glycosylated form of
carcinoembryonic antigen in exfoliated tumour
cells. This assay is approved only for use as a

b) Nuclear Matrix Protein 22 (NMP22)


This test is based on the detection of NMP22,
which is a member of a family of proteins that is
part of the structural framework of the nucleus and

Table 15. Common aberrations in urothelial carcinoma as detected by loss of heterozygosity (LOH) and
comparative genomic hybridization (CGH) analyses
LOH

CGH

Location Clinical Association

Location
9p, 9q, 10q, 11p,Y

Type of
Aberration
Loss

Clinicopathological
Association
non-invasive UC

3p

high grade/high stage

4p

high grade/high stage

1q, 17, 20q

Gain

non-invasive UC

4q

high grade/high stage

11q

Amplification

non-invasive UC

8p

high grade/high stage

Invasive UC

9q

all stages/grades

2q, 4p, 4q, 5q, 6q, 8p, 9p, 9q, 10q, Loss
11p, 11q, 13q, 15q, 17p, 18p, 18q,
Y
1q, 3p, 3q, 5p. 6p, 7p, 8q, 10p, 17q, Gain
19p, 19q, 20p, 20q, Xq

11p

high grade/high stage

Invasive UC

11q

recurrence

1q2224, 3p2425, 6p22, 8p12, Amplification


8q22, 10p1214, 10q2223, 11q13,
12q1221, 17q21, 20q13

14q

carcinoma in situ

Invasive UC

Table 16. Established clinicopathologic prognostic parameters in superficial and muscle invasive urothelial
carcinoma of bladder.
Clinicopathologic Prognostic Parameters
Superficial urothelial carcinoma
Muscle invasive urothelial carcinoma
WHO/ISUP Grade
pT stage
Presence of associated CIS/Dysplasia
Disease duration
Time to and frequency of recurrences
Multifocality
Tumor Size (>3 cm)
Failure of prior BCG Rx
Presence of LVI
Depth of Lamina propria Invasion

pTNM
LVI
Resistance to neoadjuvant chemotherapy
Divergent Histology:
Micropapillary
Osteoclast rich
Undifferentiated/Giant cell
Plasmacytoid

120

surveillance test if used in conjunction with cytology.


The overall sensitivity of the combined Bladder
Cancer Immunofluorescence Assay and cytology is
approximately 84%, which is better than either test
alone and it performs better at the detection of low
grade urothelial carcinoma.

types in regards to biologic behavior and prognosis.


Increasingly, molecular evidence supporting two
divergent pathways of pathogenesis for superficial
and invasive disease is accumulating. Superficial
BC is thought to originate from benign urothelium
through hyperplasia with only a small contribution
(10-15%) to the pool of high grade non-invasive and
subsequently invasive BC. The majority of invasive
tumors appear to originate through progression from
dysplasia to flat CIS and high grade non invasive BC
where genetic instability lead to accumulation of genetic alterations promoting progression to invasive
lesions [414-417].

d) UroVysion
This is a fluorescent in situ hybridization (FISH)
probe set with FDA approval for use in monitoring
tumor recurrence and primary detection of urothelial
carcinoma in voided urine specimens from patients
with gross or microscopic hematuria, but no previous history of urothelial carcinoma. The UroVysion
test probe set contains a mixture of four fluorescent
labeled DNA probes; a locus specific probe to the
9p21 band on chromosome 9 and to the centromere
of chromosomes 3, 7 and 17. The individual sensitivity of the centromeric probes for chromosome 3,
7, 17 is reported to be 73.7%, 76.2% and 61.9%,
respectively, while the sensitivity of homozygous
9p21 deletion for urothelial carcinoma has been reported as 28.6%. The UroVysion test is based on
combination of these probes and the sensitivity and
specificity has been reported to be 72% and 83%,
respectively. This test, however, is not free of false
positive and false negative results.

Clinically, a significant proportion of superficial tumors (pTa and pT1) are deemed to recur following
transurethral resection (TURB) with only a relative
minority of cases enduring progression to highergrade, deeply invasive aggressive tumors.
The superficial BC pathogenesis pathway appears
to be primarily based on alterations in the tyrosine
kinase receptor FGFR-3 and H-RAS[418] and PIK3CA-Akt in a subset of cases [414,419,420]. The
RAS-MAPK and PI3K-Akt pathways are potentially
the most important pathways promoting cell growth
in urothelial neoplasia. RASgenes activating mutations lead to activation of Mitogen-activated protein
kinases (MAPK) and PI3K pathways. Activating mutations in upstream tyrosine kinase receptor FGFR3
seems to be mutually exclusive withRASmutations
given that both signal through a common downstream pathway in urothelial oncogenesis. On the
other hand,PIK3CAmutations generally co-occur
withFGFR3mutations suggesting an additive oncogenic effect ofPIK3CAtoFGFR3mutations.

Level of evidence: 4
Recommendation: C

III. Prognostic and Predictive


Biomarkers in Urothelial
Carcinoma

The pathogenic pathway for muscle invasive BC


primarily involves alterations in tumor suppressor
genes p53, p16 and Rb [416,417,421].

1. Introduction
Molecular diagnostics applications are now an
integral part of the management algorithms of
several solid tumors such as breast, colon, and lung.
In stark contrast, the current clinical management of
urologic malignancies is lagging behind. Clinically
robust molecular tests that can identify patients
that are more likely to respond to a given targeted
agent or even those in need of a more aggressive
treatment based on well-validated molecular
prognosticators are still lacking. Table 17 and 18
outline traditional and potential molecular markers
in bladder cancer. Several promising biomarkers for
detection, prognosis, and targeted therapeutics are
now under evaluation. The following is a discussion
of some candidate biomarkers that may soon
make their transition to clinical assays in urothelial
carcinoma patients.

As illustrated in figure 53, subsequent p53 and Rb


alterations are also needed for the progression of
a subset of superficial lesions into a higher grade
muscle invasive BC.

3. P
 rognostic Biomarkers in
Superficial Non-Muscle Invasive
(NMI-BC) and Muscle Invasive
Urothelial Carcinoma (MI-BC)
The currently established clinicopathologic prognostic
parameters in superficial lesions include: pTNM
stage, WHO/ISUP grade, size of tumor, disease
multifocality, presence of CIS and frequency and rate
of prior recurrences [422]. Prognostic parameters
that can accurately predict the subset of superficial
tumors that will progress are actively sought in order
to identify patients in need for vigilant surveillance
and aggressive treatment plan [423,424]. Equally
needed are markers that will improve prognostication
in muscle invasive disease given the current poor

2. Divergent Molecular Pathogenesis


in Urothelial Carcinoma
Superficial and muscle invasive bladder urothelial
carcinoma (BC) display two distinct clinical pheno-

121

Figure 53. Divergent molecular pathways of oncogenesis in superficial and muscle invasive urothelial
carcinoma of urinary bladder. Genetic alterations are depicted in key stages of disease progression.
Table 17. Potential molecular prognostic parameters in superficial and muscle invasive urothelial
carcinoma of bladder.
Emerging Molecular Prognostic Markers
Superficial Non Muscle Invasive Urothelail Carcinoma (NMIBC)
**Proliferation index (Ki67, MIB1, S phase)
**FGFR3 Mutation/Overexpression (protective)
**mG (FGFR#/MIB1)
**p53 inactivation/accumulation
**DNA Ploidy Status
Multi Traget FISH
HRAS
ERBB3, ERBB4 overexpression (protective)
Loss of E Cadherin
Cell Cycle Control:
Downregulation of Rb expression
Downregulation of p21 expression
Downregulation of p27 expression
Cyclin D3 overexpression
Cyclin D1 overexpression

Muscle Invasive Urothelial Carcinoma (MI-BC)


p53 inactivation/accumulation
Alterations of Rb expression
Loss of p21 expression
Alteration of p16 expression
Loss of E Cadherin
RTK:
EGFR overexpression
HER2 overexpression/amplification
Angiogenesis Markers:
VEGF overexpression
HIF 1A overexpression
TSP1 Overexpression
mTOR-Akt Pathway:
mTOR
Phos S6 expression (protective)

**Multimarker Immunexpression Analysis:


(p53,p27,ki67,Rb,p21)

Genomic and Gene Expression Array Panels

Angiogenesis Markers:
VEGF overexpression
HIF 1A overexpression
TSP1 Overexpression

Epigenetic Alterations:
RASSF1 promotor hypermethylation
E Cad promotor hypermethylation
EDNRB promotor hypermethylation

Genomic and Gene Expression Array Panels


Epigenetic Alterations:
RASSF1 promotor hypermethylation
DAPK promotor hypermethylation
APC promotor hypermethylation
E Cad promotor hypermethylation
EDNRB promotor hypermethylation

122

outcome (60% or less overall survival rate) in this


group of patients [425-427].

prospective study should support a additive role for


the inclusion of the new biomarker in existing management algorithm(s) [445,446]. It is the lack of the
latter crucial steps in biomarkers development that
had hindered the streamlining of clinical utilization
of several promising markers in BC patient management.

The current increasingly detailed understanding of


the molecular pathways involved in urothelial bladder cancer (BC) development and progression
has fueled the field of molecular prognostication,
theranostics and targeted therapy in this disease
[73,424,428-444]. Incorporating molecular biomarkers in clinical management can only be done after
rigorous and extensive validation process. Initial retrospective discovery studies need to be confirmed
and validated in large independent cohorts. The
crucial subsequent step in developing a prognostic
or predictive biomarker is validating the robustnest
of the proposed biomarker in well controlled multiinstitutional randomized prospective study. Such

a) Chromosomal Numerical Alteration


Chromosome 9 alterations are the earliest genetic
alterations in both of the above described divergent
pathways of BC development. They are responsible
for providing the necessary milieu of genetic instability that in turn allows for the accumulation of subsequent genetic defects. Several additional structural/
numerical somatic chromosomal alterations are also
a common occurrence in BC. Among these, gains of
chromosomes 3q, 7p, and 17q and 9p21 deletions
(p16 locus) are of special interest given their potential
diagnostic and prognostic value [447,448]. A multitarget interphase FISH based urine cytogenetic assay
was developed (39) based on the above numerical
chromosomal alterations and is now commercially
available and commonly used in clinical management. Initially FDA approved for surveillance of recurrence in previously diagnosed BC patients, the
test subsequently became approved for screening in
high risk (smoking exposure) patients with hematuria. The multicolor FISH assay appears to enhance
the sensitivity of routine urine cytology analysis and
can be used in combination with routine cytology
as a reflex testing in cases with atypical cytology. A
sensitivity range of 69-87% and a specificity range
of 89-96% have been reported with the multitarget
interphase FISH assay [449]. With the exception of
one study [450], the multitarget FISH urine assay has
been shown to be more sensitive than routine cytology. An additional advantage of urine based FISH
testing could be the anticipatory positive category of
patients identified by such assay. This refers to patients where FISH assay detects molecular alteration
of BC in urine cells several months prior to cancer
detection by cystoscopy or routine cytology. In the
study by Yoder et al [451], two thirds of the 27% of
patients categorized as anticipatory positive developed BC that was detected by cystoscopy up to 29
months later. Such encouraging results point to the
great potential of molecular testing in early detection
and allocation of vigorous/frequent follow-up cystoscopy in at risk patients [452-455].

Table 18. Recommended Nomenclature for Urine


Cytology

I. Adequacy Statement (Optional)


Satisfactory for evaluation
List any quality factors affecting specimen
Unsatisfactory for evaluation (give reason)
II. General Categorization
Negative for epithelial cell abnormality (see
Descriptive Diagnosis)
Epithelial cell abnormality present (see Descriptive
diagnosis)
III. Descriptive Diagnosis
Negative for epithelial cell abnormality
Infectious agents
Bacterial organisms
Fungal organisms
Viral changes (CMV, herpes, adenovirus,
polyomavirus)
Nonspecific inflammatory changes
Acute inflammation
Chronic inflammation
Changes consistent with xanthogranulomatous
pyelonephritis
Cellular changes associated with:
Chemotherapeutic agents
Radiation
Epithelial Cell Abnormalities
Atypical urothelial cells (*see comment)
Low-grade urothelial carcinoma
High-grade urothelial carcinoma (invasive
carcinoma vs. carcinoma in situ)
Squamous cell carcinoma
Adenocarcinoma
Other malignant neoplasms (specify type)

Finally, several recent studies have pointed to


the potential prognostic role for multitarget FISH
analysis [447,448,456-458]. Maffezini et al. were
able to demonstrate that low-risk FISH-positive
patients, defined as 9p21 loss/Ch3 abnormalities,
had a higher rate of recurrence as compared to
FISH-negative patients [456]. The recurrence rate
was even greater in patients with a high-risk positive
FISH (Ch7/Ch17 abnormality). Both Kawauchi et
al., using bladder washings and Kruger et al., using

IV. Other
A comment section should be available and used
at the discretion of the cytopathologist in which
additional findings can be listed and/or further
clarification of findings within one or more of the
above diagnostic categories made. At the discretion
of the cytopathologist, ancillary studies may be
included in the report.

123

formalin fixed paraffin embedded transurethral


biopsy samples, independently found loss of 9p21 to
predict recurrence but not progression in superficial
BC [447,448]. Furthermore, both Savic et al [457].
and Whitson et al [458]. found urine cytology and
FISH in post-BCG bladder washings to be predictive
of failure to BCG therapy in patients with non muscle
invasive disease. Such promising prognostic role
for multitarget FISH awaits prospective randomized
trial before clinical integration into practice
algorithm. Clear guidelines for interpretation and test
performance parameters in terms of interobserver
reproducibility is also needed [459].

and compared it to the European Organization for


Research and Treatment of Cancer (EORTC)NMIBCcalculator[472] (weighted score of six variables
including WHO 1973 grade , stage, presence of CIS,
multiplicity, size, and prior recurrence rate). The mG
(89%) was more reproducible than the pathologic
grade (41-74%). FGFR3 mutations significantly correlated with favorable disease parameters, whereas
increased MIB-1 was frequently seen with pT1, high
grade, and high EORTC risk scores. EORTC risk
score remained significant in multivariable analyses
for recurrence and progression. Importantly, mG also
maintained independent significance for progression
and disease-specific survival and the addition of mG
to the multivariable model for progression increased
the predictive accuracy from 74.9% to 81.7%.

b) Receptor Tyrosine Kinases


Recent studies have pointed to the potential
prognostic value of evaluating the expression of
receptor tyrosine kinases (RTK) such as FGFR-3,
EGFR and other ERB family members (HER2/neu
and ERBB3) [89,417,429,460-467] in superficial and
muscle invasive bladder cancer disease.

Several studies have suggested a negative prognostic


role for HER2/neu amplification/overexpression
in MI-BC [446,473-475]. Most recently, Bolenz et
al. found HER2/neu positive MI-BC patients to be
at twice increased risk for recurrence and cancer
specific mortality on multivariable analyses adjusted
for pathological stage, grade, LVI, lymph node
metastasis and adjuvant chemotherapy [461].

FGFR3 mutations are a common occurrence in


NMIB and can theoretically be used alone or combined with RAS and PIK3CA oncogenes as markers
of early recurrence during surveillance. Both Zuiverloon et al [468] and Miyake et al [469] independently
developed sensitive PCR assays for detectingFGFR3 mutations in voided urine. A positive urine sample by the assay developed by Zuiverloon group was
associated with concomitant or future recurrence in
81% of NMI-BC cases [468]. An even higher positive
predictive value of 90% was achieved in patients
with consecutive FGR3 positive urine samples. Similarly, Miyake et al. were able to detect FGR3 mutation in 53% of their 45 patients and found their assay
to be superior to cytology (78% vs 0) in detecting
post TURB recurrence in NMI-BC harboringFGFR3
mutations in primary tumors [469].

c) P53, Cell Cycle Regulators and Proliferation


Index Markers:
Early studies by Sarkis et al revealed p53 alterations
to be a strong independent predictor of disease
progression in BC (superficial, muscle invasive
as well as CIS) [476-478]. P53 has also been
shown to be predictive of increased sensitivity to
chemotherapeutic agents that damage DNA [479481]. Recent studies have further supported the
prognostic role of p53 in pT1-pT2 patients following
cystectomy showing an independent role for p53
alteration in predicting disease free survival ( DFS)
and disease specific survival (DSS) [482].
Among other G1-S phase cell cycle regulators,
Cyclin D3 and Cyclin D1, p16, p21 and p27 have
also been evaluated as prognosticators in NMI-BC
[480,483-488]. Lopez-Beltran et al [485], confirmed
their initial finding of the independent prognostic
role of cyclin D3 and cyclin D1 overexpression in
predicting progression in pTa and pT1 tumors [480].
Their findings however are in contrast to subsequent
findings by Shariat et al. emphasizing the need for
further validation in muti-institutional large cohorts of
patients [487].

Kompier et al. were recently able to develop multiplex PCR assay for mutational analysis detecting the
most frequent mutations hot spots of HRAS, KRAS,
NRAS, FGFR3 and PIK3CA in formalin fixed paraffin embedded (FFPE) TURB samples [419]. They
demonstrated evidence of at least one mutation in
up to 88% of low grade NMI-BC samples. Hernandez et al revealed that FGFR3 mutations were more
common among low malignant potential neoplasms
(LMPN; 77%) and TaG1/TaG2 tumors (61%/58%)
than among TaG3 tumors (34%) and T1G3 tumors
(17%) [470]. On multivariable analysis, mutations
were associated with increased risk of recurrence in
NMI-BC.

A synergistic prognostic role for combining p53


evaluation with other cell cycle control elements such
as pRb, cyclin E1, p21 and p27 is emerging both in
NMI-BC and MI-BC [424,435,479,489,490]. In a study
by Shariat et al [424], NMI-BC patients with TURB
demonstrating synchronous immunohistochemical
alterations in all four tested markers (p53, p21, pRb
and p27) were at significantly lower likelihood of
sustaining disease free survival (DFS) compared

Van Rhijn et al previously proposed a molecular


grade parameter (mG) based on a combination
of FGFR3 gene mutation status and MIB-1 index
as an alternative to pathologic grade in NMI-BC
[471]. Recently, the same group elegantly validated their previously proposed mG parameter[89]

124

to patients with only three markers. The negative


predictive effect was decreased witthe number of
altered markers (3 vs 2 vs 1). Similarly, the same
group later found that combining p53, p27 and Ki67
assessment in pT1 radical cystecomy specimens
improved the prediction of DFS and DSS [489].

using genes differentially expressed in superficial Vs


muscle invasive tumors. Accuracies of 82% (entire
cohort) and 90% (MI-BC) were also obtained for
predicting overall survival. A genetic profile consisting
of 174 probes was identified in patients with positive
lymph nodes and poor survival.

A similar synergistic prognostic role of assessment


of immunoexpression of multiple molecular markers
(p53, pRb and p21) was demonstrated by Chatterjee
et al in patients undergoing cystectomy for MI-BC
[435].

Recently, Birkhahn et al attempted to identify genes


predictive for recurrence and progression in Ta using
a quantitative pathway-specific approach in a set of
24 key genes by real-time PCR in tumor biopsies at
initial presentation 73.They found CCND3 expression to be highly sensitive and specific for recurrence
(97% and 63%, respectively). While HRAS, E2F1,
BIRC5/Survivin and VEGFR2 were predictive for
progression by univariate analysis, on multivariable
analysis the combination of HRAS, VEGFR2, and
VEGF expression status predicted progression with
an impressive 81% sensitivity and 94% specificity.

The superiority of multimarkers approach compared


to prior single marker approach certainly merits further assessment [476-478,491]. Such multimarker
approach of prognostication could soon be integrated in the standard of care in BC management
once additional multi-institutional, prospective, trials
confirm the above promising findings.

In a recent landmark study, Lindgren et al suggested


that a combined molecular and histopathologic
classification of BC may prove more powerful in
predicting outcome and stratifying treatment [497].
The authors combined gene expression analysis,
whole genome array-CGH analysis and mutational
analysis of FGFR3, PIK3CA, KRAS, HRAS, NRAS,
TP53, CDKN2A, and TSC1 to identify two intrinsic
molecular signatures (MS1 and MS2). Genomic
instability was the most distinguishing genomic feature
of MS2 signature, independent of TP53/MDM2
alterations. Their genetic signatures were validated
in two independent data sets that successfully
classified urothelial carcinomas into low-grade
and high-grade tumors as well NMI-BC and MI-BC
with high precisions and sensitivities. Furthermore,
a gene expression signature that independently
predicts metastasis and disease-specific survival
was also defined. This clearly supports the role of
molecular grading as a complement to standard
pathologic grading.

Tumor proliferation index measured immunohistochemically by either Ki-67 or MIB-1 has been
consistently shown to be a prognosticator in bladder cancer [89,471,480,484,492-495]. As mentioned
above, tumor proliferation index (MIB1) in NMI-BC
plays a prognostic role as one of the elements of the
mG paramater forwarded by Van Rihjn et al [471].
The independent prognostic role of proliferation index measured by Ki-67 has also been shown. In the
study by Quintero et al. Ki67 index in NMI-BC TURB
biopsy was predictive of PFS and DSS [494].
A similar role for proliferation index assessment as
prognosticator is established in MI-BC. Building on
initial findings of significance in an organ confined
subset of MI-BC by Margulis et al a recent report
of the bladder consortium multi-institutional trial (7
institutions and 713 patients) confirmed the role
of proliferation index, measured in cystectomy
specimens [492]. In the later study, Ki-67 improved
prediction of both PFS and DSS when added to
standard prediction models supporting a role for
proliferation index assessment in stratifying patients
for peri-operative systemic chemotherapy. This has
certainly taken Ki-67 assessment a step closer to
clinical applicability in MI-BC.

Mengual et al performed gene expression analysis


in 341 urine samples from NMI-BC and MI-BC
patients and 235 controls by TaqMan Arrays [444].
A 12+2 gene expression signature demonstarted
a staggering 98% sensitivity and 99% specificity
in discriminating between BC and control and 79%
sensitivity and 92% specificity in predicting tumor
aggressiveness (NMI-BC vs MI-BC). The signature
was then validated in voided urine samples and
maintained accuracy. In an integrated genetic/
epigenetic approach, Serizawa et al prospectively
performed mutational screen of a set of 6 genes
(FGFR3, PIK3CA, TP53, HRAS, NRAS and KRAS)
and quantitatively assessed promoter methylation
status of 11 additional genes (APC, ARF, DBC1,
INK4A, RARB, RASSF1A, SFRP1, SFRP2, SFRP4,
SFRP5 and WIF1) in NMI-BC tumor biopsies and
corresponding urine samples from 118 patients and
33 controls.501 A total of 95 oncogenic mutations

d) Gene Expression and Genomic Analysis


Several recent gene expression studies have
highlighted sets of differentially expressed genes
that may play a role in diagnosis and in predicting
recurrence and progression in BC [73,102,416,430433,443,444,496-501].
In a landmark study by Sanchez- Carbayo et al
oligonucleotide arrays were used to analyze transcript
profiles of 105 cases of NMI-BC and MI-BC.433
Hierarchical clustering and supervised algorithms
were used to stratify bladder tumors based on stage,
nodal metastases, and overall survival. Predictive
algorithms were 89% accurate for tumor staging

125

and 189 hypermethylation events were detected.


The total panel of markers provided a sensitivity
of 93% and 70% in biopsies and urine samples
respectively. FGFR3 mutations in combination with
three methylation markers (APC, RASSF1A and
SFRP2) provided a sensitivity of 90% in tumors and
62% in urine with 100% specificity.

dependent probe amplification assay (MS-MLAP), to


analyze 25 tumor suppressor genes (TSG) that have
been thought to play a role in BC oncogenes.502.
The TSG included PTEN, , CD44, WT1, GSTP1,
BRCA2, RB1, TP53, BRCA1, TP73, RARB, VHL,
ESR1, PAX5A, CDKN2A, and PAX6. The authors
found BRCA1, WT1, and RARB to be the most
frequently methylated TSGs with receiver operating
characteristic curve analyses revealing significant
diagnostic accuracies in two additional validation
sets.

With the impending cost and turn around time


advantages of next generation sequencing
technology, the power of genomic approach in
providing a non-invasive diagnostic and predictive
tool should be heavily pursued in a prospective large
cohort.

Finally, assessment of promoter methylation is giving


additional insights on BC oncogenesis. Promoter
methylation of CpG Islands and shores controlling
miRNA expression is one such example [505].

e) Epigenetic Alterations
Epigenetic analysis is also gaining momentum in
BC as a non invasive diagnostic tool for screening
and surveillance. As a prognostic tool, epigenetic
analysis has similarly shown promising potential in
BC patients [501-514].

f) Ploidy and Morphometric Analysis


Several studies have pointed to the indepedent
prognostic role of ploidy and S phase analysis in
NMI-BC [495,515-521]. Ploidy analysis can be
performed by flow cytometry or automated image
cytometry (ICM) and is applicable to urine cytology
specimens as well as biopsy supernatant[ 521] and
disaggregated TURB FFPE specimens [516].

In an early study by Catto et al hypermethylation


anlaysis at 11 CpG promoters islands was performed
by MSP PCR in 116 bladder and 164 upper-tract
tumors [503]. Promoter methylation was found in 86%
of all tumors and the incidence was relatively higher
in upper tract tumors compared to BC. Methylation
was associated with advanced tumor stage and
higher tumor progression and mortality rates. Most
importantly, on multivariate analysis methylation
at the RASSF1A and DAPK genes promoters was
associated with disease progression independent of
tumor stage and grade.

In one of the largest studies assessing DNA ploidy in


NMI-BC (377 test set; 156 validation set); Ali-Et-Dein
et al. found stage, DNA ploidy, tumor multiplicity,
history of recurrence, tumor configuration and type of
adjuvant therapy to independently predict recurrence
[515]. Recurrence at 3-month, grade and DNA ploidy
were the only predictors of progression to muscleinvasion. The constructed Predictive Index model
successfully stratified patients in a second validation
set into three risk groups. Likewise, Baak et al were
able to show ploidy status and S phase measured
by an automated image cytometry platform (ICM) to
be strong independent predictiors of recurrence and
progression in pTa and pT1 patients [516].

The same group, using quantitative methylationspecific PCR (MSP) at 17 candidate gene promoters, found five loci to be associated with progression (RASSF1a, E-cadherin, TNFSR25, EDNRB,
andAPC) [513]. Multivariate analysis revealed that
the overall degree of methylation was more significantly associated with subsequent progression and
death than tumor stage. An epigenetic predictive
models developed using artificial intelligence techniques identified likelihood and timing of progression
with 97% specificity and 75% sensitivity.

Despite all the above encouraging data, ploidy


analysis still await a prospective randomized trial to
bring ICM or flowcytometry technique into current
standard management algorithms for NMI-BC.

g) Emerging Potential Biomarkers:

Among the studies on the diagnostic role of promoter


methylation, the study by Lin et al used MSP assay
in 4 genes (E-cadherin, p16, p14, and RASSF1A)
in primary tumor DNA and urine sediment DNA
from 57 bladder cancer patients 508. MSP detected
hypermethylation in the urine of 80% of tested
patients. Hypermethylation analysis of E-cadherin,
p14 or RASSF1A in urine sediment DNA detected
85% of superficial and low grade BC and 79% of
high grade and 75% of invasive bladder cancers.
The study highlighted the great potential of such
test in detecting NMI-BC. Similar diagnostic role
was also found by Cabello et al. using a novel
technology, methylation-specific multiplex ligation-

Other biomarkers with encouraging but less robust


data on their potential prognostic role in BC include
tumor microenvironement markers such as cell
adhesion E Cadherin and N Cadherin [439,522]
and angiogensis modulators such as HIF 1a, HIF 2,
VEGF, CA IX and Thrombospondin-1 [438,523-528].
In addition, our group and others have demonstrated
a potential prognostic role for mTOR-Akt pathway
markers [439,522,529-532]. Other markers such as
Aurora-A have also been investigated in this setting
[533,534]. Finally, miRNA profile alterations will
certainly be a new area of heavy investigation as a
non invasive diagnostic and as a prognostic tool in
BC patients [505,535-537].

126

The current clinicopathologic based prognostic


approach to predicting progression in superficial BC
[102,494,538-540] will potentially be supplemented
in the not so distant future by a molecular guided
approach based on markers among those listed in
table 16 and 17, [422,427,435,503,513,524,525,
53,532,541-544].

solitary lesions, often (more than half) are located in


the neck or trigone, and the vast majority correspond
to adenocarcinomas [193]. It is important to separate
metastases from primary urinary bladder tumors
as they are associated with different managerial
approaches and ominous prognoses. However, this
distinction may represent an important challenge for
urologists and pathologists. In this section we will be
discussing specific immunohistochemical markers
used in distinguishing primary from secondary
tumors of the urinary bladder.

Finally, from a targeted therapy perspective,


RTK-HRAS-MAPK (superficial disease) and
p53-pRb (muscle-invasive disease) pathogenic
pathways as well as angiogenesis pathway of the
tumor microenvironment offer tremendous new
opportunities for future management of BC. Phase
II trials evaluating the role of tyrosine receptor kinase
inhibitors (TKI) targeting EGFR with small molecules
such as Gefitinib and Sorafinib or monoclonal
antibodies (MoAb) such as Cetuximab are underway.
Other Phase II trials are addressing the role of
Herceptin (anti HER2 MoAb) and Bevacizumab
(anti VEGF MoAb) in BC. Phase I trials testing
the safety of p53 or Rb intravesical gene therapy
are being evaluated. One strategy example is the
intravesical introduction of a wild type p53 loaded
replication deficient Adenovirus (Ad5CMV-TP53) in
an ambitious attempt to compensate for the loss of
p53 function in invasive BC [545-549].

2. Prostatic Adenocarcinoma
Direct invasion is the most frequent mechanism of
bladder involvement by prostatic adenocarcinoma.
The latter can be morphologically indistinguishable
from poorly differentiated urothelial carcinoma,
especially when glandular differentiation or clear
cell features are exhibited by urothelial carcinomas.
This distinction is critical given the staging,
therapeutic, and prognostic implications. Several
immunohistochemical markers have been reported
to reliably differentiate poorly differentiated urothelial
carcinoma from prostatic adenocarcinoma. However,
none of them is sufficiently sensitive and/or specific to
be used alone, and therefore immunohistochemical
panels are recommended. Kunju et al [556] evaluated
the joint utility of prostatic specific antigen (PSA),
prostate acid phosphatase (PAP), high-molecular
weight cytokeratin (HMWCK) (clone 34E12), CK7,
CK20, p63 and -methylacylcoenzyme (P504s) in
this differential diagnosis. They found that 95% of
documented prostatic adenocarcinoma demonstrated
a PSA+, HMWCK/p63- immunohistochemical profile
whereas 97% of urothelial carcinomas tested were
PSA-, HMWCK/ p63+. Negative PSA associated with
HMWCK and/or p63 positivity may also establish
a diagnosis of urothelial carcinoma. Even though
PSA and PAP are reliable markers of prostatic
adenocarcinoma [556-561], their expression can be
absent or only focally and weakly seen in high-grade
tumors. PSA expression is frequently heterogeneous
and consequently, immunostains may need to be
performed on multiple blocks containing tumor
[562]. Additionally, PAP is slightly less specific for
prostatic origin as a minority of urothelial carcinomas
(approximately 10%), may be focally positive for
this marker [562]. Although the majority of prostatic
adenocarcinomas are CK7-/CK20-, while urothelial
carcinomas are CK7+/CK20+ or CK7+/CK 20[556,557,559,563-565], results frequently overlap
making these markers less sensitive and specific
when used in isolation. However, the coordinate
expression patterns of CK7 and CK20 in conjunction
with PAP may be very helpful in differentiating
prostatic carcinoma (PAP+/CK7/CK20) from
urothelial carcinoma (PAP/CK7+/CK20+ or PAP/
CK7+/CK20), especially when the results of the
preliminary panel of PSA, HMWCK, and p63 are
all negative. Furthermore, in the occasion where

h) Conclusion:
At present, there are no prognostic or predictive
biomarkers for urothelial carcinoma that significantly
improve on pathologic grade and stage information;
therefore, none of these evaluations are currently
recommended for routine clinical practice.
Level of evidence: 4
Recommendation: C

IV. The Utility of


Immunohistochemistry In The
Differential Diagnosis Of
Primary Versus Metastatic
Tumors To The Urinary
Bladder
1. Introduction
Tumors metastatic to the urinary bladder are
relatively uncommon accounting for 2 to 14% of all
bladder malignancies [193,194,550,551]. Secondary
neoplasms may affect the urinary bladder by either
direct extension from adjacent anatomic sites or, less
frequently, by lymphovascular spread [552]. Tumors
that may extend directly to the urinary bladder include
colon (21%), prostate (19%), rectum (12%), and
uterine cervix (11%). Distant sites include stomach,
skin, lung, breast, kidney, and testis [193,550,553555]. These tumor deposits are almost always

127

the primary panel demonstrates negative results,


addition of uroplakin III and thrombomodulin may help
as both markers are negative in adenocarcinomas
of prostatic origin. Alphamethylacyl-CoA-racemase
(P504s) is not a useful marker to distinguish prostatic
from urothelial carcinomas since it is also expressed
in approximately one third of urothelial carcinomas
[556,566-568].

rectal or colonic carcinoma, which may secondarily


invade the urinary bladder, may have significant
morphologic overlap with high-grade urothelial
carcinoma with glandular differentiation (either
with mucinous, enteric, or signet ring cell features)
or pure urinary bladder adenocarcinoma. Even
though clinical history is essential and the findings
of cystitis glandularis of intestinal type, carcinoma
in situ, or tumor centered in the inner half of the
bladder wall would favor a primary urothelial origin,
in small specimens these findings may not be seen
and immunohistochemistry may be needed in
elucidating the origin of the tumor [193]. Urothelial
carcinoma with glandular differentiation may be
separated from secondary involvement by colonic
carcinoma as it is positive for CK7 and CK20, but
negative for villin. Colonic adenocarcinoma is
typically CK7 negative and CK20/villin positive
[579,580]. However, it is important to keep in mind
that colonic carcinomas, although uncommonly,
can express CK7 along with CK20 [581]. Although
bladder carcinomas generally lack of villin and CDX2
staining, primary vesical carcinomas with gland
formation may take on an enteric immunophenotype
[579,582,583]. Other markers that may be used as
part of a panel include thrombomodulin, uroplakin
III and -catenin. Thrombomodulin is expressed in
urothelial carcinomas, but it is negative in colorectal
carcinomas, and it appears to be a more sensitive
marker than CEA in this differential diagnosis [584].
Up to date, uroplakin III has only been demonstrated
in urothelial carcinomas [557,585,586]. -catenin
is highly expressed in colonic carcinoma cells
as a result of mutated adenomatous polyposis
coli gene [331], but its expression is minimal to
absent in tumors of urothelial origin including
adenocarcinomas [331,587]. CEA and P504s are not
helpful in either differential diagnosis [193,582]. The
latter is observed in approximately 30% of urothelial
carcinomas [566,588] and up to 65% of urinary
bladder adenocarcinomas [582]. In difficult cases,
colonoscopy should serve as the gold-standard in
this distinction.

Newer prostate restricted markers such as prostate


specific membrane antigen (PSMA) and prostein
(P501S) have been proposed to be a useful adjunct
in the diagnosis of prostate adenocarcinoma as they
have a high sensitivity [569-576] and can be used
in the differential diagnosis between prostatic and
urothelial carcinoma.
A novel marker NKX3.1, an androgen regulated homeodomain gene whose expression in prostate epithelium has been extensively studied, has been established as part of the immunohistochemical panel
to distinguish metastatic prostatic adenocarcinoma
and high grade urothelial carcinoma [333,577]. The
sensitivity of NKX3.1 in prostatic adenocarcinoma
(Gleason scores 8 to 10) ranges from 92% to 94%.
Specificity for prostate versus urothelial origin has
been reported to be 100%, as all urothelial carcinomas are NKX3 negative [577]. The predominant
NKX3 nuclear staining in primary and metastatic prostatic adenocarcinomas is useful when compared to
other prostate specific markers (e.g. PSA, PAP, and
P501s) that exhibit cytoplasmic staining. This staining pattern can be of additional diagnostic benefit
when there is only weak and focal cytoplasmic staining of other prostate markers.
Glutamate decarboxylase 1 (GAD1) is also a
sensitive and specific for prostate tissue, whereas
bladder neoplasms are entirely negative [578]. When
compared to the sensitivity and specificity profile
of PSA and PSMA, GAD1 has been found to as
reliable as PSA and to show higher specificity than
PSMA. GAD1 shows consistent strong expression
in both benign and malignant prostate tissues, while
cancers from the urinary bladder, rectum, and lung
are almost entirely negative [578].

4. Cervical/Vaginal Carcinoma
Squamous cell carcinoma of the cervix or vagina can
secondarily involve the urinary bladder and mimic
a primary squamous cell or high-grade urothelial
carcinoma with or without squamous differentiation
[118]. Likewise, adenocarcinomas of the cervix with
or without intestinal differentiation and transitional cell
carcinoma of the cervix or corpus (the latter very rare)
or vagina, may simulate a primary adenocarcinoma
or urothelial carcinoma of the urinary bladder [589595].

In an attempt to find better markers for urothelial


carcinoma, Higgins et al identified S100P and GATA3
as two distinct genes showing consistently high and
uniform expression in urothelial carcinoma when
compared to prostatic adenocarcinoma[213] In an
immunohistochemical analysis on tissue microarrays,
polyclonal antiserum against S100P protein stained
86% of urothelial carcinomas but only 3% of prostatic
adenocarcinomas and monoclonal antibody against
GATA3 stained 67% urothelial carcinomas but none
of the prostate carcinomas.

In general, all squamous cell carcinomas,


independent of site of origin, are CK7, CK5/6,
cytokeratin 34E12, and p63 positive; therefore,
these markers are not helpful in establishing the site

3. Colorectal carcinoma
The morphologic appearance of poorly differentiated

128

of origin [596-598]. Even though p16, a surrogate


marker of human papilloma virus (HPV), is widely
expressed in squamous cell carcinomas of the cervix
and vagina,[599,600] it has also been reported in
one third of squamous cell carcinomas of the urinary
bladder in both males and female patients [601] and
in high-grade urothelial carcinomas [313]. The use
of in situ hybridization for HPV DNA can be helpful
as this is positive in most squamous cell carcinomas
of the cervix, but not of the bladder (Figure 54A,B)
[602].

and 35 adenocarcinomas of the endometrium [605].


Two relatively new markers proposed by some
investigators to be highly expressed in urothelial
carcinomas of the bladder, namely S100p and
GATA3, have been shown to be also positive in
transitional cell proliferations/tumors of the ovary
[213], but no experience is reported in transitional
cell proliferations of the lower female genital tract.
Cervical adenocarcinomas of the usual or intestinal
type, or even those with signet ring cells, are typically
CK7 positive. Intestinal type adenocarcinomas are
also often positive for CK20 and CDX2, in contrast
to usual cervical adenocarcinomas, which rarely
stain for these markers [590]. However, cervical
carcinomas with signet ring morphology have been
reported to be CK20 and CDX2 negative [590].
Thus, these markers are not useful in distinguishing
secondary involvement by cervical adenocarcinoma
from urinary bladder adenocarcinoma, especially
when the tumor has intestinal morphology. CEA is
not helpful in this differential diagnosis either, while
thrombomodulin expression in urinary bladder
adenocarcinomas is much lower (approximately
17%) when compared to conventional urothelial
carcinomas. Thus, if positive, thrombomodulin is
helpful in favoring a urinary bladder origin of the
adenocarcinoma [584]. Uroplakin expression has
been reported in 50 to 60% of all urothelial carcinomas
[557], but results on urothelial carcinomas with
glandular differentiation or pure adenocarcinomas
are lacking. p16, as discussed in the differential
diagnosis of squamous cell carcinomas, is typically
positive in cervical adenocarcinomas (related to
HPV infection); however, it has not been studied in
primary adenocarcinomas of the urinary bladder and
should be used with caution as it has been shown
to be expressed in endometrial and colorectal
carcinomas [606,607].

Urothelial carcinomas of the urinary bladder are


typically positive for CK7, CK20, uroplakin, and
thrombomodulin with the latter two having 100%
and 96% specificity and 57% and 69% sensitivity,
respectively [585,603]. They are also typically p63
positive [255]. Squamotransitional cell carcinomas
of the uterus have been shown to be CK7 and p63
positive and uroplakin negative, but experience is
very limited [604]. Transitional cell carcinomas of
the uterus and vagina as well as those arising from
the urinary bladder are both positive for CA19.9 and
CEA [603]. No experience has been reported with
thrombomodulin in squamotransitional carcinomas
of the uterus, although Ordoez reported negative
staining on 7 squamous cell carcinomas of the cervix

Clear cell carcinoma can arise in the uterine cervix


or vagina and may secondary involve the urinary
bladder. In such cases, the differential diagnosis
includes a clear cell carcinoma of the urinary bladder
or a urothelial carcinoma with clear cytoplasm
[205,608]. A prior clinical history of diethylstilbestrol
or finding areas of typical urothelial neoplasia would
favor a primary cervical or urinary bladder origin,
respectively. Clear cell carcinomas of the urinary
bladder express low-molecular weight keratins,
CK7 and CK20, but are negative for estrogen and
progesterone receptors [114,204,609-611], an
immunohistochemical profile that overlaps with
that seen in clear cell tumors of the female genital
tract. Primary urinary clear cell carcinomas are also
typically positive for CD10, P504s, and they can
be positive for PAX2 [608]. In contrast, clear cell
carcinoma of the cervix and vagina has been reported
to be CD10 negative, although experience with this
marker is limited [612]. Up to date, PAX2 and P504s
have not been studied in clear cell carcinomas of the

Figure 54. A) Cervical squamous cell carcinoma


involving the urinary bladder shows B) positivity by
HPV in situ hybridiization.

129

female genital tract. PAX8, another positive marker


of clear cell carcinoma of the urinary bladder [613] is
expressed in tumors arising from the ovary [614], but
no experience is reported in clear cell carcinomas of
the uterus or vagina.

setting [618]. Thyroid transcription factor-1 (TTF-1)


plays a role in the development and physiology of the
thyroid gland and lungs, and it is used as a marker of
carcinomas originating in these organs. Commercially
available clones of TTF-1 monoclonal antibodies,
8G7G3/1 and SPT24, have been reported to have
different sensitivities for the detection of neoplasms
of different origins. SPT24 clone detects a significant
number of non-pulmonary primary tumors, especially
invasive urothelial carcinoma (5%) [619]. Therefore,
if SPT24 is used, results should be interpreted with
caution.

5. Gastric carcinoma
Advanced gastric carcinoma, especially if poorly
differentiated or signet ring cell type, may secondarily
involve the urinary bladder. The morphologic
appearance may overlap with that seen in poorly
differentiated or plasmacytoid urothelial carcinoma.
These tumors typically show a CK7+/CK20+
immunohistochemical profile that overlaps with
that observed in urothelial carcinoma. Uroplakin
may be helpful in separating a poorly differentiated
urothelial carcinoma from a poorly differentiated
gastric carcinoma, but is not helpful in the differential
diagnosis with plasmacytoid carcinoma, as the latter
is frequently negative or shows minimal positivity
[233]. Thrombomodulin, another useful marker in
the diagnosis of urothelial carcinoma, has not been
reported in gastric carcinoma, but in some pancreatic
carcinomas [170].

Even though small cell carcinoma of the urinary


bladder accounts for less than 1% of all bladder tumors, the urinary bladder is the most frequent location of extrapulmonary small cell carcinoma [620].
Morphologically, this tumor is identical to its lung
counterpart. The finding of associated conventional
urothelial neoplasia or its variants would favor a primary origin from the urinary bladder, as one third of
small cell carcinomas are admixed with conventional
urothelial carcinoma. Both pulmonary and urinary
bladder small cell carcinomas share a dot-like positive staining with pan-keratin and have variable expression of neuroendocrine markers (chromogranin,
synaptophysin and CD56). Although TTF-1 was initially proposed as a useful marker to confirm the lung
as the origin of a metastatic adenocarcinoma, its expression has been increasingly reported in tumors
from other origins including small cell carcinomas of
the urinary bladder [621].

6. Malignant Melanoma
Melanoma can be primary [615,616] or involve
secondarily (more commonly) the urinary bladder
[194]. In most cases a prior history of melanoma, the
finding of typical histologic patterns and the presence
of melanin pigment are helpful in the diagnosis of
melanoma. Specifically, the finding of an in situ
component may be helpful in establishing a primary
origin in the urinary bladder. However, it is well
known that malignant melanoma in the urinary tract
may closely mimic the appearance of a urothelial
carcinoma, as it can have a papillary, nested,
sarcomatoid, plasmacytoid and even pagetoid
growths as described in urothelial carcinomas
[617]. When separating melanoma from urothelial
carcinoma, it is important to remember that the latter
is frequently positive for S-100[213] and melanomas
may rarely be positive for pankeratins [150], and thus
a panel of antibodies including HMB45, melan-A and
cytokeratins should be used to establish the nature
of the tumor.

When squamous differentiation is the predominant


pattern in a urinary bladder neoplasm in a patient
with a known lung squamous cell carcinoma,
the differential diagnosis between a primary and
a secondary neoplasm may be difficult. A panel
including TTF-1, napsin A and p16 may be helpful,
with TTF-1 and/or napsin A positivity favoring a lung
primary, while p16 positivity typically would favor an
extrapulmonary site [597].

8. Breast Carcinoma
Breast carcinoma metastatic to the urinary bladder
is a rare event [622]. For the most part, the
morphologic appearance of typical ductal carcinoma
(which metastasizes to the urinary bladder less
commonly than lobular carcinoma) does not create
problems in differential diagnosis; however, when
tumors have a micropapillary architecture or are
composed of signet ring cells (either ductal or
lobular), the histologic appearance overlaps with
primary urothelial neoplasms [233,623,624]. When
using immunohistochemistry, it is important to keep
in mind that CK7 expression is seen in both breast
and urothelial carcinomas,[563] although primary
and metastatic ductal carcinoma show important
differences in CK7 expression [625]. Cytokeratin 5/6
positivity can be detected in about half of urothelial

7. Lung Carcinoma
Lung adenocarcinoma is the most frequent type
of pulmonary neoplasm and therefore is the most
frequent type found in a metastatic setting, followed
by small cell and squamous cell carcinoma. When
glandular morphology is the predominant pattern in
a bladder neoplasm in a patient with a known lung
carcinoma or lung mass, the differential diagnosis
between a primary and a secondary neoplasm
should be entertained. The use of a limited
immunohistochemical panel including napsin A,
GATA3, and S100P may be extremely helpful in this

130

carcinomas, but can also be seen in one third of


breast ductal but not lobular carcinomas [626].
Estrogen receptor (ER) expression is typically
seen in breast carcinomas [627], but urothelial
carcinomas can also express ER, especially when
invasive, higher grade and high stage [628-630]. It
is primarily form that is expressed with expression
of the form being much less frequent (95). CK20
is rarely expressed in ductal and lobular carcinoma
in contrast to 30% of transitional carcinomas [563].
Thus, a CK7+/CK20- profile would favor a breast
carcinoma while a CK7+/CK20+ profile would
support the diagnosis of urothelial carcinoma [564].
However, unusual subtypes of breast carcinomas
including mucinous and papillary can be CK20
positive [627,631]. Gross cystic disease protein
15 (GCDP-15), a fairly specific marker for breast
carcinoma, is only expressed in approximately 60%
of these tumors. Thus, a positive stain supports
a diagnosis of metastatic breast carcinoma, but
a negative stain does not exclude it [632]. Other
markers including EMA, CEA are not helpful in this
differential diagnosis as both tumors may stain for
these markers [558,633]. Thrombomodulin and
CK34BE12, reported to be positive in up to 70% and
100% of urothelial carcinomas, can be expressed
also in breast carcinoma [170].

CD10 membranous immunoreactivity is frequently


used in the diagnosis of renal cell carcinoma,
clear cell type [638]. However, there is a growing
list of neoplasms that can be CD10 positive,
including urothelial carcinomas. Bahadir et al [639]
demonstrated that this marker is positive in more
than 40% of urothelial carcinomas and its expression
is strongly correlated with high tumor grade and
stage. Therefore, in a setting where metastatic
renal cell carcinoma is in the differential diagnosis,
CD10 should be interpreted with caution. CK7 and
P504s are sensitive markers for papillary renal
cell carcinoma [613]; however, they also can be
expressed in carcinomas of urothelial origin [556].
Although PAX2 and PAX8 are sensitive and relatively
specific markers for renal neoplasms, regardless
of subtype [638], these two markers have recently
been also found in clear cell adenocarcinoma of the
lower urinary tract limiting its value in this setting
as discussed earlier [613]. C-kit is expressed in
most chromophobe renal cell carcinomas and
oncocytomas, however, clear and papillary renal
cell carcinomas are negative for this marker as are
urothelial carcinomas, except small cell carcinomas
of the urinary bladder (approximately one third are
immunoreactive) [640]; therefore, this marker may
have limited utility in the differential diagnosis of
renal versus urothelial carcinoma. In conclusion,
it is suggested that a panel which includes PAX2
or PAX8 and CD10, supplemented by a urothelial
marker would reliably differentiate the majority of
metastatic renal cell carcinomas from urothelial
carcinoma [638].

9) Renal Cell Carcinoma


Renal cell carcinoma is an uncommon source of
bladder metastasis with fewer than forty reported
cases [208,554,634-637]. When metastatic renal cell
carcinoma is encountered, immunohistochemistry
may be very helpful as: a) either metastatic renal
cell carcinoma may precede the primary tumor or in
patients with a known prior history of renal carcinoma,
metastatic foci may develop after a prolonged period
of time and slides from the primary neoplasm may
not be available for review; and b) metastatic renal
cell carcinoma may look morphologically different
from the primary tumor. Furthermore, in patients
with a prior history of renal carcinoma and a new
bladder tumor, it is extremely important to know if the
tumor is a new primary versus metastases. Typically,
patients with renal cell carcinoma metastatic to the
bladder present with gross hematuria demonstrating
sessile bladder lesions [554]. Histological evaluation
is usually straightforward, however, on occasions,
the differential diagnosis between a metastatic renal
cell carcinoma and urothelial carcinoma is difficult
as both may show overlapping morphologies,
particularly if urothelial carcinoma has clear [202] or
lipid-type cells [148]. A large variety of markers have
been proposed for the diagnosis of primary renal cell
carcinoma, however, studies on the immunoprofile
of metastatic tumors are scarce. Metastatic renal
cell carcinoma usually retains an immunoprofile that
is similar to the one seen in primary neoplasms;
however, intensity and proportion of staining tends
to decrease.

When renal carcinoma originates from the collecting


ducts (collecting duct carcinoma) and metastasizes
to urinary bladder, the differential diagnosis with
urothelial carcinoma with glandular differentiation
can be extremely challenging. Highmolecularweight cytokeratin is expressed in most collecting
duct and urothelial carcinomas [638,641]. Adding
urothelial markers like p63 can be useful in this
differential diagnosis [638,641]. Finally, renal pelvis
urothelial carcinomas are generally morphologically
and immunophenotypically indistinguishable for
their bladder counterpart; however, strong PAX8
expression is seen in a subset of renal pelvic
urothelial carcinomas (unpublished data).

10. Testicular Neoplasms


Metastatic testicular germ cell tumor to the urinary
bladder is extremely rare with few reported cases
in the literature [553,642]. Still, distinguishing
metastatic germ cell from a non germ cell tumor has
important clinical managerial implications. Markers
used in the diagnosis of testicular germ cell tumors
have included placental-like alkaline phosphatase
(PLAP), CD30, CD117(C-KIT), and a-fetoprotein
(AFP). Although these markers are useful, they

131

show only moderate sensitivity/specificity and


significant immunophenotypic overlap with somatic
malignancies that may show PLAP and CD30
expression [643]. Furthermore, as their expression
can be lost in metastatic foci, other markers have
been recently investigated. Glypican 3 has been
proposed as a new marker for yolk sac tumor, but
its staining is typically focal and is not highly specific
[644]; it has been shown to be positive in urothelial
carcinomas with myxoid stroma [645]. Also, novel
embryonic stem cell markers have been identified
as being sensitive and specific for testicular germ
cell tumors. Cheng et al established the sensitivity
and specificity of OCT4 in detecting metastatic germ
cell tumors in retroperitoneal lymph nodes and found
that all embryonal carcinomas and seminomas
showed diffuse and strong OCT4 nuclear staining. In
contrast, yolk sac tumors, choriocarcinoma, mature
teratomas, primitive neuroectodermal tumors,
malignant lymphomas, melanomas and carcinomas
from other sites including prostate, urinary bladder,
pancreas, lung, colon, and kidney, were negative for
this marker [646]. SALL4 has also been reported as
a marker of seminoma and embryonal carcinoma,
but unlike OCT4, SALL4 is also strongly expressed
in yolk sac tumors [647,648].

urothelial dysplasia, which would require clinical outcome studies to be of any value to clinical management. Studies of adjunctive methods of classification in the urothelial atypia of uncertain significance/
urothelial dysplasia histologic spectrum have underscored the problems with immunohistochemistry in
that specific diagnostic setting [649]. We discuss the
use of adjunctive immunohistochemistry to distinguish between urothelial carcinoma in situ and benign reactive lesions in routine diagnostic practice.

2. p53
Urothelial carcinoma in situ may show strong,
diffuse nuclear immunoreactivity for p53 [650-654];
however, the sensitivity has been reported to be as
low as 57%. It is important to acknowledge that the
evaluation of p53 by immunohistochemistry can be
difficult [655]. The intensity of the antibody staining
may fluctuate significantly with small variations in
titer and inter-laboratory variation is not uncommon.
Normal or reactive urothelium, therefore, may show
numerous cells with weak to moderate nuclear
staining. It is critical that one use a high threshold for
a positive result when evaluating p53 in flat urothelial
lesions. Only diffuse 3+ nuclear immunoreactivity in
the neoplastic cell population should be regarded as
positive.

11. Conclusions

3. Cytokeratin 20

Immunohistochemistry is commonly employed, and


sometimes essential, in routine diagnostic practice
for the distinction between a primary bladder
carcinoma and secondary involvement form another
anatomic site. The exact immunohistochemical
panel required, if any, depends on the histologic
features of a given case.

CK20 expression has been studied by numerous


groups that have all verified the typically strong and
diffuse cytoplasmic immunoreactivity in urothelial
carcinoma in situ cells (Figure 55 A,B) [649651,656-658]. The reported sensitivity of CK20 for
carcinoma in situ has ranged from 72-89%, while
the specificity is over 90%. The extent of staining
depends on the pattern of the urothelial carcinoma
in situ being evaluated. Carcinoma in situ shows
full thickness staining when the neoplastic cells
comprise the full thickness of the urothelium;
however, in cases of pagetoid carcinoma in situ, this
staining is restricted to the individual neoplastic cells
without reactivity in the surrounding benign urothelial
cells. In normal urothelium and in reactive urothelial
atypia, cytokeratin 20 expression is restricted to the
superficial umbrella cell layer. In our experience, the
CK20 expression in carcinoma in situ is preserved
after both intravesical and radiation therapy
(unpublished data).

Level of evidence: 4
Recommendation: C

V. The Use of
Immunohistochemistry in Flat
Urothelial Lesions with Atypia
1. Introduction
The histologic classification of flat urothelial lesions
with atypia (i.e. reactive urothelial atypia, urothelial
atypia of uncertain significance, urothelial dysplasia,
and urothelial carcinoma in situ) can be one of the
most difficult diagnostic challenges in genitourinary
pathology. Numerous studies have investigated the
potential utility of adjunctive immunohistochemistry
in the diagnosis of these flat urothelial lesions.

4. CD44s
The CD44s antibody has been suggested as a useful
comparative immunostain to use in conjunction with
CK20 because it has the reverse staining pattern
[651]. In normal urothelium, the basal cell layer shows
membranous reactivity for CD44s, occasionally with
some extension to the intermediate cell levels. In
reactive urothelial atypia, the full thickness of the
urothelium often shows this basal-like membranous

The ability of a given immunophenotype to predict


disease progression, independent of histology, has
not been adequately studied. We, therefore, do
not endorse using immunohistochemistry to distinguish urothelial atypia of uncertain significance from

132

7. Conclusions

staining pattern. In contrast, the neoplastic cells of


urothelial carcinoma in situ do not express CD44s,
and the normal basal cell population is frequently
absent.

Despite the reported prototypical immunostaining


profiles of flat urothelial lesions with atypia, the
authors of this document have anecdotally seen
cases of both morphologically obvious carcinoma
in situ and benign reactive urothelium with
unexpected immunophenotypes. Any adjunctive
immunohistochemistry must be carefully correlated
with the morphologic features of the urothelial
lesion being evaluated. Currently, routine
morphologic evaluation should remain as the goldstandard for the diagnosis of flat urothelial lesions
with atypia. As stated, we do not endorse the
use of immunohistochemistry in the distinction of
dysplasia and atypia of uncertain significance. Of the
immunostains that might be used as confirmatory
adjuncts for carcinoma in situ or reactive atypia
in select difficult cases, CK20, CD44s, and p53
currently offer the most potential utility.

5. p16ink4
A few groups have evaluated the staining pattern
of p16 in urothelial carcinoma in situ [488,649].
Strong and diffuse cytoplasmic expression of p16
is reported in urothelial carcinoma in-situ, while
normal and reactive urothelium both show uniform
weak expression. As with the use of p16 in cervical
dysplasia, the interpretation of varying levels of
intensity may present problems, and one must
maintain a high threshold for a positive result. There
are very few studies with this antibody, and we have
not had similar success with its use.

6. Ki-67
Proliferation markers have also been evaluated
as adjuncts to the diagnosis of carcinoma in situ.
[650,652,657,658] The proliferation rate of carcinoma
in situ, as defined by Ki-67 immunohistochemistry,
has significant overlap with florid reactive atypia,
particularly when inflammatory infiltrates are present.
[657] Therefore, proliferation markers have little role
as diagnostic markers in this setting.

Level of evidence: 4
Recommendation: C

F. RECOMMENDED NOMENCLATURE FOR URINE CYTOLOGY

55A

I. Nomenclature and
Reporting
The Papanicolaou Society of Cytopathology
Practice and Guidelines Task Force recommends
a diagnostic format that mimics the Bethesda 2001
System for reporting cervical cytology format. (Table
18), The diagnostic section for urinary cytology
should include a statement as to the site of origin of
the urinary tract specimen (urethra, urinary bladder,
or ureter/renal pelvis).and also state the technique
used forobtaining the specimen when known (voided
urine, washing, brushing, etc.) [659].
This classification nomenclature raises few issues in
urine cytology diagnosis, particularly in the atypical
and low grade urothelial carcinoma categories. The
diagnosis of low grade urothelial carcinoma is rarely
used due to the lack of specific criteria; therefore,
most low grade urothelial carcinomas are included
in the atypical category. Regarding the atypical
urothelial cells terminology, there is little consensus
in the literature on the criteria for inclusion in this
category Nonetheless, it has been suggested
that this atypical category can be subdivided
into 2 subcategories, atypical urothelial cells of
undetermined significance (AUC-US) and atypical
urothelial cell, cannot rule out high grade urothelial
carcinoma or favor neoplasm. The rationale for this
subclassification is that specimens diagnosed in the
latter category should be referred to cystoscopy while

Figure 55. Urothelial carcinoma in situ on A) H&E


showing B) diffuse cytoplasmic immunoreactivity
for cytokeratin 20.

133

the ones diagnosed as AUC-US could be followed


with repeat urines [660-662].

72 and 83%[679], UroVysion FISH has a sensitivity of 90100% for the detection of invasive bladder
cancer (pT14) and a carcinoma specificity of 95%
while for low-grade non-invasive the sensitivity increases from 25 to 6075% when compared to cytology [680]. Use of Urovysion has been advocated
in equivocal urine specimens, either suspicious or
atypical [455,457,681-684] due to its Increased sensitivity. Skacel et al [685] retrospectively evaluated
the UroVysion FISH assay in 120 urine specimens
and found a sensitivity of 100 and 89% in patients
with suspicious and atypical urine cytology specimens, respectively, while the overall specificity was
97%. A negative FISH result in case of a negative or
atypical cytology does not exclude low-grade urothelial neoplasia, since up to 30% of these tumors are
negative with the FISH assay.
A positive Urovysion test (Figure 57) has been
associated with higher rate of recurrence in patients
treated with intravesical bacillus Calmette-Guerin
(BCG) for high-grade non-muscle invasive urothelial
carcinoma [453,457,458,686,687]. A higher rate
of recurrence has also been noted in patients with
positive Urovysion test and negative cystoscopy and
cytology [455,678,688,689] although this was not
confirmed in all studies [454,690]. Pitfalls of Urovysion
have been associated with overinterpretation of
benign tetraploid cells [691], or chromosomal
aberrations following pelvic irradiation [692], except
deletion of 9p21.

Level of evidence: 4
Recommendation: C

II. Role of urine cytology in


screening and monitoring
patients with bladder cancer
Urinary cytology is a simple, noninvasive, and
relatively inexpensive method for detecting UC [663].
It has consistently shown to have high specificity,
especially in detecting carcinoma in situ and highgrade, flat lesions (Fig 56) [664,665]. However, it is
also accepted that urine cytology has low sensitivity
that varies particularly for low grade urothelial
carcinoma [449,450,666-674]. Factors leading to
the low sensitivity of urine cytology include minimal
atypia seen in low grade UC, paucity of tumor cells in
an inflammatory background, instrumentation effect,
lithiasis, reactive changes, post BCG treatment,
degenerative changes and viral cytopathic effect
[675,676]. Diagnostic accuracy is also dependent on
diagnostic expertise [677].
Level of evidence: 3C
Recommendation: B

The wide application of UroVysion in routine practice


is still controversial due to its high cost [693-695].
Although there is higher accuracy in the detection of
high-grade lesions, the benefit achieved from better
diagnosis of clinically harmless low-grade lesions by
Urovyion is minor and it is achieved at a much higher
cost [661]. The low cost efficiency is also cited as
a problem when Urovysion is used in a population
of patients with hematuria but no increased risk for
urothelial carcinoma [680,693,694].
Level of evidence: 3C
Recommendation: B
Figure 56. High grade urothelial carcinoma in urine
cytology

III. Role of FISH in screening


and monitoring of bladder
carcinoma
Urovysion has been approved by the US Food
and Drug Administration for screening of urothelial
carcinoma in patients with hematuria and monitoring
for tumor recurrence in patients with a prior history of
urothelial carcinoma [449,678].
The overall sensitivity and specificity of Urovysion is

Figure 57. Positive UroVysion in a urine specimen

134

With some updates in select topics, this proposed


guideline aim at standardizing descriptions of
diagnosis and reporting of urothelial carcinoma of
the bladder and to help optimize uniformity between
individual pathology practices and institutions (Grade
C). This protocol may assist pathologists in providing
clinically useful and relevant reporting information in
a multidisciplinary care setting.

G. UPDATED PROTOCOL FOR THE


EXAMINATION AND HANDLING OF
SPECIMENS FROM PATIENTS WITH
URINARY BLADDER CARCINOMA
I. Introduction

II. Relevant Clinical History

Providing the best management for patients with


bladder neoplasia relies on close cooperation and
teamwork among urologists, oncologists, radiologists
and pathologists. The pathologist is involved in
the diagnosis and assessment of prognostic and
therapeutic variables in surgical specimens such
as bladder biopsies, transurethral resection (TUR)
and cystectomy specimens. Pathologists must
report accurately and with minimal variability the key
pathologic parameters using terminologies that are
well understood by clinicians [696].

The ability to have an open communication is


essential for pathologists, urologists and oncologists
in the decision-making process for diagnosis
and management. Adequate clinical information
is important to pathologists in deciding the best
approach in handling and processing the surgical
specimens and directing the most essential
information which to be reported. The pathology
specimen requisition form must be filled with
relevant and sufficient information the clinicians
can provide. With the high degree of accuracy
of cystoscopy, the pathologists should also take
extrameasures of reviewing cystoscopic impressions
before diagnosing a tumor in biopsy or TUR samples
[700,701]. With differing implications for non-invasive
papillary versus flat (carcinoma in situ) tumors and
advent of enhanced cystoscopic (e.g. fluoroscopic)
visualization, urologists impression provides helpful
information to the pathologist in difficult situations,
such as when the entire lesion architecture is not
obvious on the slides (e.g. fragmented or superficial
sections). Indications of the surgical procedure,
such as for surveillance or post intravesical therapy
biopsies, initial or second TUR (e.g. for deep muscle
assessment or completion of resection), early or
salvage cystectomy, among others, are important
for pathologists in specimen handling and reporting.
Other relevant information includes history of
bladder cancer and prior intravesical or neoadjuvant
chemotherapy. Presence of calculi, catheterization
or infection may explain reactive changes (versus
dysplasia) when under consideration.

The most common tissue specimens encountered are


biopsy and TUR, which provide the main information
essential in tailoring the patient subsequent therapy
that includes more aggressive treatment options.
Additional information on curative-intent resections
(e.g. cystectomy) determines the adequacy of
therapy and appropriate surveillance or further
need for surgery or adjuvant chemotherapy and/
or radiotherapy [697]. The recent literature is not
short of satisfactory checklists and commentaries
for handling and reporting of genitourinary pathology
specimens [174,393,698,699]. Organizations such
as the College of American Pathologists (CAP)
recommend reporting in a checklist/synoptic format
and have outlined essential elements for reporting
specimens resected for invasive bladder cancer
[698]. However changes remain in flux as recent
advancements, shift to multidisciplinary care, and
regular introductions of updated clinical guidelines
in bladder cancer diagnosis and management
diversify the nature of specimens and impact the
manner of handling and assessment. For example,
other than the usual cystectomy for muscle invasive
cancer, contemporary cystectomies include early
(non-muscle invasive tumors), post neo-adjuvant
chemotherapy, salvage, or palliative resections.
Eventual pT0 or pTis cystectomies that contain no
invasive cancer are becoming much more common.
Segmental cystectomy for solitary lesions without
pTis lesions amenable to resection with adequate
margin is now considered a feasible surgical option.
Second TUR procedures can be received for prior
inadequate initial TUR staging or deep muscle
sampling, prior incomplete tumor resection, or
bladder-preservation treatment. The differences in
indications and characteristics of tumor content in
these specimens evoke re-appraisal instead of being
uniformly processed in the traditional manner.

III. Urologists Handling of


Specimens
Tissue preservation is required and samples should
be handled the minimum way possible. Ideally, tissue
container with fixatives for biopsy or TUR specimens
should be readily available in the operating rooms. In
situations where patients are part of study protocols
or clinical trials requiring fresh tissue preservation,
the specimen should be sent without any delay to
pathology gross room for immediate processing.
In TUR, a deeper bite for assessment of MP
involvement is preferably submitted separately in
another container especially in situations of larger

135

VI. Pathologist Handling of


Cystectomy Specimens

tumors. This will facilitate search for muscle invasive


component of the tumor. Biopsy or TUR of multifocal
tumors should also be submitted separately according
to site. Cold cup mapping biopsies of visually normal
mucosa and prostatic urethra are needed to exclude
flat lesions and should also be submitted separately
per site. This provides a better picture of the extent
of the lesions in the bladder. Any erythematous or
velvety area should be sampled. In cystectomy,
ureter margins submitted separately should have
proper labelling of the distal end (true margin). In
partial or segmental cystectomy, the bladder wall
resection margins should be marked with identifiers
such as suture to facilitate identification.

The bladder must be opened and after adequate


fixation samples must be taken from representative
areas of the tumor, normal-appearing bladder, the
prostate, seminal vesicles, or other organs included
in the surgical specimen. Minimum sections of the
bladder should include at least the tumor/s including
its deepest penetration, and representative sections
of the dome, trigone, anterior, posterior and lateral
walls. In potential cystectomy with no residual tumor
(pT0), prior surgical sites or mucosal ulcerations
should be submitted entirely for any possible residual
tumor. Adequate mapping of different regions of
the bladder should be performed, as well as in
early cystectomy or in setting of post-neoadjuvant
chemotherapy with less visible or no gross residual
tumors. Any suspicious-looking areas of reddening
and mucosal induration (for CIS) need to be
sampled. Grossly visible tumour in perivesical fat
(pT3b) or at resection margins (R2) must be stated
in the gross examination. For tumor grossly invading
the bladder wall, multiple full thickness sections
including the area grossly suspicious for deepest
penetration should be included in the sections to
assess pT2a/b vs. pT3a. The underlying soft tissue
margin should be examined, inked and sampled.
Unless submitted separately for frozen section,
ureter and urethral margins should be sampled,
preferably taken as shave cross sections so that
the entire mucosal circumference can be visualized.
In case of segmental cystectomy, bladder wall
margins of resection should be identified for tumor
involvement. In cystoprostatectomy specimens,
right and left prostate should be inked differently
and representative sections from both sides are to
be taken including perpendicular sections of the
apical margins, to secure assessment of important
variables in case of concomitant prostate cancer,
which is not uncommon to occur. Sections of the
prostate should include the prostatic urethra to
identify possible tumor invasion originating from
a separate urethral carcinoma (pT2). Sections of
possible deep penetration to prostate from bladder
by the tumor should also be taken to document
higher stage (pT4) bladder cancer. Likewise, tumor
mass along the distal ureter/ureteral orifice should
be closely examined whether or not tumor is arising
from the distal ureter, for appropriate staging. Lymph
nodes must be dissected from the lymphadenectomy
specimens and all grossly and tentatively identified
lymph nodes in the specimen should be submitted.
(Table 20)

IV. Pathologist Handling of


Biopsy Specimens
All biopsy-sampled tissues should be submitted
for processing. Effort should be made to identify
the mucosal aspect of tissue fragments to allow
embedding on edge and better visualization of the
surface urothelium, although this may not always be
possible. For histologic evaluation, at least 3 levels of
tissue sections for each biopsy should be prepared.
Deeper levels into the block are recommended if the
surface urothelium is deemed not entirely visible or
in situations of denuded samples, particularly with
known risk for CIS, to exclude the possibility of
denuded CIS (clinging type) (Grade C).

V. Pathologist Handling of
Transurethral resections
The conventional way of processing TUR specimens
is to submit the tissue entirely. In large tumors
exceeding 10 cm, one approach is to submit 1 section
per cm of tumor diameter (up to 10 cassettes), and
additional selective sampling until MP invasion is
identified, or submitting the rest of the sample if no
MP invasion is recognized after the first selective
sampling. The second TURs for deep muscle
staging are typically low volume samples, which
can be processed entirely. Large bulk completion
TUR for incompletely resected tumors may have an
option to submit less if the initial TUR has already
established the diagnosis and attributed risk (e.g
pT2 stage). Follow-up TUR for bladder-preservation
treatment after chemotherapy and/or radiotherapy
however should be processed the conventional way.
(Table 19)

136

system represents the gold standard for prediction


of recurrence after radical cystectomy in patients
with invasive bladder cancer. Thus, accurate staging
provides broadly applicable prognostic information
and is the basis of all authoritative patient
management decisions [706].

VII. Reporting of Histologic


Grade
Use of the WHO (2004) / ISUP classification system
is recommended (Grade B) [113,275]. Based on
institutional / clinician preference, other systems may
be used along side. Grading should also take cancer
heterogeneity into consideration; approximately onethird of patients with pTa tumor had cancer containing
a different histologic grade. Currently, grading of
papillary urothelial tumors is typically based on the
worst grade present (Grade C).

In addition to advanced tumour stage, however, other


pathological factors have been evaluated as possible
risk factors for recurrence and poor survival. Among
them, lymphovascular invasion (LVI) marks the initial
step of the metastatic process. Its significance in
predicting outcome of affected patients, however,
remains controversial.

VIII. Reporting Variant


Urothelial Histology

According to literature data, LVI can be observed in


about 30-50% of cystectomy specimens (Table 21)
[60,707-717]. Prevalence of LVI correlates significantly with increasing T classification [708,710,712714, 717-721], high tumour grade [712-714,717,718],
and presence of lymph node metastasis [710,712714,717,721,722]. However, in a considerable
number of patients with histologically proven nodal
disease LVI is not detected within the bladder wall.
In the comprehensive study by Lotan et al. LVI was
observed in 9.0% of T1, 23.0% of T2, 50% of T3, and
78% of T4 tumours, respectively [712]. Likewise, the
authors detected LVI in 151 out of 581 (26%) nodenegative and 122 out of 169 (72%) node-positive
cancers. In a recent international validation study
collecting 4,257 patients from 12 centres, Shariat et
al. rendered similar data: LVI was present in 11.0%
of T1, 31.3% of T2, 52.3% of T3, and 60.9% of T4
tumours, respectively, and its presence was significantly associated with high tumour grade (5.1% G1,
38.2% G2, 33.5% G3) and presence of lymph node
metastasis (22.5% N0, 64.7% N1-2) [717].

Urothelial cancer is known to show variant histologic


features otherwise known as divergent differentiation,
with estimates ranging from 7% to 81% in series
specifically reporting differentiation patterns of
urothelial carcinomas [117,702]. Current evidence
suggests that urothelial cancer with divergent
differentiation has a worse prognosis when compared
with pure urothelial cancer, although stage matched
cohorts show limited differences (Level 3). Billis and
colleagues [703] in a study of 165 TURBT noted that
7% of tumors showed squamous and/or glandular
differentiation. Those patients with divergent
differentiation had higher clinical stage at presentation
compared to conventional urothelial carcinoma [703].
Wasco et al [704] in a study of 448 consecutive
TURBTs and 295 subsequent cystectomy, noted
that urothelial cancers with divergent differentiation
were more likely to be muscularis propria invasive at
TURBT and extravesical fat invasive at cystectomy
compared to pure urothelial cancers. Jozwicki et
al [705] performed mapping in 38 cystectomies
specimens, and then correlated the mapping studies
with survival time, the presence of greater than 80%
pure urothelial cancer within a specimen was a
favorable prognostic factor, and increasing numbers
of histologic subtypes (increasingly divergent
differentiation) led to a less favorable prognosis.

LVI has repeatedly been associated with poor


outcome of patients undergoing radical cystectomy
for invasive bladder cancer. This effect is primarily
observed in patients with node-negative cases
disease [708,712-714,717,718,723]. In detail, LVI
proved to be a predictor of local, distant, and overall
tumour recurrence as well as a predictor of adverse
overall and disease-specific survival, independent of
tumour stage and grade [711,712,715-718,724,725].
In some studies, however, discordant results were
obtained. Thus, in the study by Canter et al., LVI
independently predicted overall and disease-specific
survival, whereas no independent influence on
recurrence-free survival was noted [708]. In other
studies, some of them involving rather low numbers
of patients, LVI was identified as an adverse
prognostic variable only in univariate analysis, yet
lost prognostic significance in multivariate analysis
[709,710,713,726-728]. Finally, in two studies, both
involving less than 100 patients, the presence of LVI
was not found to be significantly related to outcome
[729,730].

Divergent morphology must be documented because


of its prognostic implications (Grade B). Further,
reporting facilitate correlation with subsequent
metastasis, were it to occur at remote sites. When
multiple divergent (aberrant) morphologies are
present, one should provide relative percentages for
each of the pattens e.g. invasive urothelial carcinoma
(75%) with squamous (20%) and glandular (5%)
differentiation.

IX. Lymphovascular Invasion


(LVI)
Tumour stage reflected by the AJCC/UICC TNM

137

____ Dysplasia
___ Carcinoma in situ
___ Cannot be determined
Histologic Grade
__ Not applicable
__ Cannot be determined
Urothelial Carcinoma (WHO 2004/ISUP)
___ Papilloma
___ PUNLMP
___ Low-grade
___ High-grade
___ Primary Grade
___ Secondary grade
___ Other (specify): _________________________
Urothelial Carcinoma ( WHO 1973)
__Papilloma
__ Grade 1
__Grade 2
__ Grade 3
__ Primary Grade
__ Secondary Grade

Table 19. URINARY BLADDER: Biopsy and


Transurethral Resection of Bladder Tumor (TURBT)

Note: Use of checklist for biopsy specimens is


optional
Select a single response unless otherwise
indicated.

Table 19. URINARY BLADDER: Biopsy and


Transurethral Resection of Bladder Tumor (TURBT)

Note: Use of checklist for biopsy specimens is


optional
Select a single response unless otherwise
indicated.
Relevant Clinical History
___ Cystoscopic impression of the mucosa
___ Indication of procedure
___ Previous history of cancer in bladder or
elsewhere in the genitourinary tract
___ Prior therapy
Procedure
__ Biopsy
__ TURBT
__ Other (specify): _________________________
__ Not specified
Histologic Type
__ Urothelial (transitional cell) carcinoma
__ Urothelial (transitional cell) carcinoma with % of
squamous differentiation
__ Urothelial (transitional cell) carcinoma with % of
glandular differentiation
__ Urothelial (transitional cell) carcinoma with % of
micropapillary component
__Urothelial (transitional cell) carcinoma with
variant histology (specify): ___________________
__ Squamous cell carcinoma, typical
__ Squamous cell carcinoma, variant histology
(specify): _______________________
__ Adenocarcinoma, typical
__ Adenocarcinoma, variant histology (specify):
___________________________
__ Small cell carcinoma
__Undifferentiated carcinoma (specify):
___________________________
__ Mixed cell type (specify):
___________________________
__ Other (specify):
___________________________
__ Carcinoma, type cannot be determined
Associated Epithelial Lesions (select all that
apply)
__ None identified
__ Urothelial (transitional cell) papilloma (World
Health Organization [WHO] 2004/ International
Society of Urologic Pathology [ISUP])
__ Urothelial (transitional cell) papilloma, inverted
type
__ Papillary urothelial (transitional cell) neoplasm,
low malignant potential
(WHO 2004/ISUP)
____ Hyperplasia

Adenocarcinoma and Squamous Cell Carcinoma

__ GX: Cannot be assessed


__ G1: Well differentiated
__ G2: Moderately differentiated
__ G3: Poorly differentiated
__ Other (specify): __________________________
Tumor Configuration (select all that apply)
__ Papillary
__ Solid/nodule
__ Flat
__ Ulcerated
__ Indeterminate
__ Other (specify): __________________________
pT1 Tumors
___ Muscularis Mucosae Present
___ Thin-Walled blood vessels present
___ Depth of Lamina propria invasion (micrometer)
Type of tumor invasion
___ Single cells
___ Nodular
___ Trabeculae
___ Infiltrative
Stromal Reaction
___ Stromal retraction
___ Stromal edema
___ Inflammation
___ Fibroblastic proliferation
___ Fibrosis
Adequacy of Material for Determining Muscularis
Propria Invasion
___ Muscularis propria (detrusor muscle) not
identified
___ Muscularis propria (detrusor muscle) present
___ Presence of muscularis propria indeterminate
Lymph-Vascular Invasion
___ Not identified
___ Present
___ Indeterminate

138

Microscopic Extent of Tumor (select all that


apply)
___ Cannot be assessed
___ Noninvasive papillary carcinoma
___ Flat carcinoma in situ
___ Tumor invades subepithelial connective tissue
(lamina propria)
___ Tumor invades muscularis propria (detrusor
muscle)
___ Urothelial carcinoma in situ involving prostatic
urethra in prostatic chips sampled by TURBT
___ Urothelial carcinoma in situ involving prostatic
ducts and acini in prostatic chips sampled by
TURBT
___ Urothelial carcinoma invasive into prostatic
stroma in prostatic chips sampled by TURBT
Additional Pathologic Findings (select all that apply)
___ Urothelial dysplasia (low-grade intraurothelial
neoplasia)

was independently associated with cancer-specific


survival, yet not with recurrence-free survival [731].
In all, six studies have so far evaluated the
significance of LVI in TURB material [721,732-735].
The reported prevalence of LVI is 6-28%, which not
unexpectedly is lower than in radical cystectomy
specimens [721,732-735]. According to a recent
investigation involving matched TURB and radical
cystectomy material from 75 patients, concordance
between LVI diagnoses of TURB and cystectomy
material is good in muscle-invasive cancers, while
it is only low in superficial tumours [721]. Overall
sensitivity and specificity of LVI detection in TURB
samples were 37% and 87%, and positive predictive
values were 65% and 67%, respectively. Similar to
the situation in cystectomy specimens, presence of
LVI in TURB material proved to be an independent
predictor of tumour recurrence [733] and progression
[732,733] as well as patients overall [734] and
disease-specific [732] survival, particularly in clinical
stage I or II [723].

Non-neoplastic changes
___ Inflammation/regenerative changes
___ Therapy-related changes
___ Cautery artifact
___ Cystitis cystica glandularis
___ Keratinizing squamous metaplasia
___ Intestinal metaplasia
___ Denuded urothelium
___ Other (specify): _________________________
Comment(s)

Whether or not lymphatic invasion and venous


invasion should be considered separately has
not been adequately addressed because most
studies have not attempted to differentiate between
venous and lymphatic invasion [60,709,712715,717,718,724,730,731,734,735]. Lotan et al
refer to difficulty and lack of reproducibility when
using routine light microscopic examination of
haematoxylin and eosin stained slides [712].
Some studies have, however, differentiated between
lymphatic and venous invasion and defined venous
invasion as tumour present in vessels with a thick
vascular wall and blood cells within the lumen
[720,726,728,736,737]. Using this definition, the
prevalence of venous invasion was generally lower
than that of lymphatic invasion [710,719,726728,736,737]. With respect to prediction of outcome,
venous invasion proved to be superior to lymphatic
invasion in two of these studies [720,737], whereas
a stronger prognostic effect for lymphatic invasion
was noted in one study [736].

Adding LVI, together with other morphological


(presence of carcinoma in situ in the cystectomy
specimen) and clinical (neoadjuvant chemotherapy,
adjuvant chemotherapy, adjuvant radiotherapy)
parameters to multivariate nomograms assessing
risk for disease recurrence or survival significantly
improved accuracy of prediction compared with the
AJCC/UICC staging system [711,716]. Likewise,
adding LVI alone to a base model including tumour
stage and grade, surgical margin status, number of
lymph nodes removed, and adjuvant chemotherapy
significantly improved predictive accuracy for
disease recurrence and cancer-specific survival in
node-negative cases, yet only marginally in nodepositive cases [717].

Despite unanimity regarding the relationship between


presence of LVI and bladder cancer aggressiveness,
there are major issues of concern that have hindered
its integration in treatment guidelines for bladder
cancer and clinical decision making. In fact, the major
factor is poor diagnostic reproducibility: reported
prevalence and levels of prognostic significance vary
considerably, differing from one study to another.
The lack of reproducibility when using routine light
microscopic examination of haematoxylin and eosin
stained slides is well-known [712,738].

The significance of LVI has additionally been


addressed in patients undergoing partial cystectomy
or transurethral tumour resection for bladder cancer
(TURB). Thus, Zhang et al. recently presented 100
patients with muscle-invasive bladder cancer who
underwent bladder-conserving therapy, i.e. partial
cystectomy. LVI was present in 16% of cases and

These differences have mainly been attributed


to lack of standardized assessment of LVI, as
demonstrated by divergent, often poorly described
histological criteria used by pathologists to identify

139

Associated Epithelial Lesions (select all that


apply)
___ None identified
___ Urothelial (transitional cell) papilloma (World
Health Organization [WHO] 2004/ International
Society of Urologic Pathology [ISUP])
___ Urothelial (transitional cell) papilloma, inverted
type
___ Papillary urothelial (transitional cell) neoplasm,
low malignant potential
(WHO 2004/ISUP)
___ Hyperplasia
___ Dysplasia
___ Carcinoma in situ
___ Cannot be determined

Table 20. URINARY BLADDER: Cystectomy, Partial,


Total, or Radical; Anterior Exenteration
Select a single response unless otherwise
indicated.

Relevant Clinical History


___ Diagnosis of previous material
___ Chemotherapy
___ Radiotherapy
___ Preoperative Examination
Specimen
___ Bladder
___ Other (specify): _________________________
___ Not specified
Procedure
___ Partial cystectomy
___ Total cystectomy
___ Radical cystectomy
___ Radical cystoprostatectomy
___ Anterior exenteration
___ Other (specify): _________________________
___ Not specified

Histologic Grade
___ Not applicable
___ Cannot be determined
Urothelial Carcinoma (WHO 2004/ISUP)
___ PUNLMP
___ Low-grade
___ High-grade
___ Primary Grade
___ Secondary Grade
___ Other (specify): _________________________

Tumor Site (select all that apply)


___ Trigone
___ Right lateral wall
___ Left lateral wall
___ Anterior wall
___ Posterior wall
___ Dome
___ Other (specify): _________________________
___ Not specified

Urothelial Carcinoma (WHO 1973)


Papilloma ___
Grade 1 ____
Grade 2 ____
Grade 3 ____
___ Primary pattern
___ Secondary pattern

Tumor Size
Greatest dimension: ____ cm
Additional dimensions: ___x____ cm
___ Cannot be determined (see Comment)

Adenocarcinoma and Squamous Cell Carcinoma


___ GX: Cannot be assessed
___ G1: Well differentiated
___ G2: Moderately differentiated
___ G3: Poorly differentiated
___ Other (specify): _________________________

Histologic Type
___ Urothelial (transitional cell) carcinoma
___ Urothelial (transitional cell) carcinoma with % of
squamous differentiation
___ Urothelial (transitional cell) carcinoma with % of
glandular differentiation
___ Urothelial carcinoma with % of Micropapillary
component
___ Urothelial (transitional cell) carcinoma with
variant histology (specify): ____________________
___ Squamous cell carcinoma, typical
___ Squamous cell carcinoma, variant histology
(specify): ________________________
___ Adenocarcinoma, typical
___ Adenocarcinoma, variant histology (specify):
____________________________
___ Small cell carcinoma
___ Undifferentiated carcinoma (specify):
____________________________
___ Mixed cell type (specify):
____________________________
___ Other (specify): _________________________
___ Carcinoma, type cannot be determined

Tumor Configuration (select all that apply)


___ Papillary
___ Solid/nodule
___ Flat
___ Ulcerated
___ Indeterminate
___ Other (specify): _________________________

140

vagina, pelvic wall, abdominalwall


Tumor invades prostatic stroma or
uterus or vagina
___ pT4b:
Tumor invades pelvic wall or
abdominal wall
Regional Lymph Nodes (pN)
___ pNX:
Lymph nodes cannot be assessed
___ pN0:
No lymph node metastasis
___ pN1:
Single regional lymph node metastasis in the true pelvis (hypogastric,
obturator, external iliac or presacral
lymph node)
___ pN2:
Multiple
regional
lymph
node
metastases in the true pelvis
(hypogastric, obrutrator, external iliac
or presacral lymph node metastasis)
___ pN3:
Lymph node metastasis to the
common iliac lymph nodes
____ Number of nodes examined
____ Number with metastases
____ Size of largest metastases
____ Number of nodes with micrometastasis or
isolated tumor cells
____ Extracapsular extension of metastasis

Microscopic Tumor Extension (select all that


apply)
___ None identified
___ Perivesical fat
___ Rectum
Prostate
___ Prostatic stroma
___ Prostatic urethra
___ Focal involvement
___ Extensive involvement
___ Seminal vesicle (specify laterality):
_______________________
___ Vagina
___ Uterus and adnexae
___ Pelvic sidewall (specify laterality):
_______________________
___ Ureter (specify laterality):
_______________________
___ Other (specify): _________________________

___ pT4a:

Margins (select all that apply)


___ Cannot be assessed
___ Margins uninvolved by invasive carcinoma
Distance of invasive carcinoma from closest
margin: ___mm
Specify margin: ____________________________
___ Margin(s) involved by invasive carcinoma
Specify margin(s): __________________________
___ Margin(s) uninvolved by carcinoma in situ
___ Margin(s) involved by carcinoma in situ
Specify margin(s): __________________________

Distant Metastasis (pM)


___ Not applicable
___ pM1:
Distant metastasis
Specify site(s), if known: _____________________
Additional Pathologic Findings (select all that
apply)
___ Adenocarcinoma of prostate (use protocol for
carcinoma of prostate)
___ Urothelial (transitional cell) carcinoma involving
urethra, prostatic ducts and acini
with or without stromal invasion (use protocol for
carcinoma of urethra)
___ Urothelial dysplasia (low-grade intraurothelial
neoplasia)
___ Inflammation/regenerative changes
___ Therapy-related changes
___ Cystitis cystica glandularis
___ Keratinizing squamous metaplasia
___ Intestinal metaplasia
___ Other (specify): _________________________
___ Tumor Bank

Lymph-Vascular Invasion
___ Not identified
___ Present
___ Indeterminate
Pathologic Staging (pTNM)
TNM Descriptors (required only if applicable) (select
all that apply)
___ m (multiple primary tumors)
___ r (recurrent)
___ y (post-treatment)
Primary Tumor (pT)
___ pTX:
___ pT0:
___ pTa:
___ pTis:
___ pT1:

Primary tumor cannot be assessed


No evidence of primary tumor
Noninvasive papillary carcinoma
Carcinoma in situ: flat tumor
Tumor invades subepithelial connective tissue (lamina propria)
pT2: Tumor invades muscularis propria (detrusor
muscle)
___ pT2a:
Tumor invades superficial muscularis
propria (inner half)
___ pT2b:
Tumor invades deep muscularis
propria (outer half)
pT3: Tumor invades perivesical tissue
___ pT3a:
Microscopically
___ pT3b:
Macroscopically
(extravesicular
mass)
pT4: Tumor invades any of the following: prostatic
stroma, seminal vesicles, uterus,

Comment(s)

141

TABLE 21. Literature review of significance of LVI in bladder cancer (studies with >100 patients)
Author

Surgery
Year

Type of
Surgery

No. Pts

pT
stage

+LN
(%)

+ LVI BI LI
(%) (%) (%)

Followup
(months)

Evolution
(survival
/progression)

Statistical
Significance
(MVA)

Abdel-Latif
et al, 2004

97-99

RC, LA

418

T1-T4

26

32

n/a

n/a

36.6
(mean)

+LVI: 27*
-LVI:56*
(3-ys survival
+LN)

Independent
(+LN)

Bassi et al,
1999

82-94

RC, LA

369

T1-T4

21

n/a

30

48.5
(median)

+BI: 29%
-BI: 56%
+LI: 39%
-LI: 61%
(5-ys survival)

Not
independent

Bolenz et
al, 2010

85-08

RC, LA

1099

T1-T4

None

27

n/a

n/a

n/a

+LVI: 69%
-LVI:74%

Independent
(RFS, CSS,
OS)

Carter et
al,
2008

88-06

RC, LA

356

T1-T4

28

32

n/a

n/a

45.4
(mean)

+LVI: 52 mo
OS
-LVI: 97 mo OS

Independent
(CSS, OS)

Cho et al,
2009

01-07

TURBT

118

T1

n/a

28

n/a

n/a

35
(median)

+LVI: 30%
-LVI:11%

Independent
(progression,
metastasis)

Hara et al,
2001

85-00

RC, LA

154

T1-T4

20

50

n/a

n/a

23
(mean)

+LVI: 42%
-LVI:83%

Not
independent

Harada et
al, 2005

89-03

RC, LA

114

T1-T4

18

n/a

33

48

n/a

+BI: 49%
-BI: 67%
+LI: 53%
-LI: 68%

Not
independent

Harada et
al, 2006

89-04

RC, LA

124

T1-T4

18

n/a

28

47

n/a

+BI: 37%R
-BI: 12%R
+LI: 34%R
-LI: 6%R

Not
independent

Herrmann
et al, 2008

86-05

RC, LA

614

T1-T4

28

n/a

n/a

n/a

44
(mean)

+LVI vs LVI:
1.7 RR

Independent
(CSS, OS)

Hong et al,
2005

91-02

RC, LA

125

T1-T4

None

n/a

21

41
(median)

+BI: HR 4.88
-BI: HR 1
+LI: HR 1.2
-LI: HR 1

Independent,
BI only
(CSS)

Leissner et
al, 2003

87-97

RC, LA

283

T1-T4

31

n/a

13

54

32.8
(mean)

+BI vs BI: RR
1.82

Independent,
BI only
(RFS)

Lotan et al,
2005

84-03

RC, LA

750

T1-T4

22

36

n/a

n/a

15
(mean)

-LN, +LVI:
86.9%
-LN, -LVI:
72.1%

Independent
in -LN,
(RFS, CSS,
OS)

Manoharan
et al, 2010

92-08

RC, LA

357

T1-T4

20

29

n/a

n/a

n/a

+LVI: 42%
-LVI:67%
(10-ys CSS)

Not
independent

Palmieri et
al, 2010

95-07

RC, LA

265

T1-T4

23

29

n/a

n/a

108
(median)

+LVI: 29%
-LVI:53%
(10-ys CSS)

Independent
(CSS)

Quek et al,
2005

71-04

RC, LA

702

T1-T4

22

35

n/a

n/a

132
(median)

-LVI: 74%
+LVI: 42%
(10-ys RFS)

Independent
(RFS, OS)

Shariat et
al, 2010

79-08

RC, LA

4257

T0-T4

25

33

n/a

n/a

43
(median)

+LVI: 39%
-LVI:72%
(10-ys CSS)

Independent
(RFS, CSS)

142

TABLE 21. Literature review of significance of LVI in bladder cancer (studies with >100 patients)
(continued)
Author

Surgery
Year

Type of
Surgery

No. Pts

pT
stage

+LN
(%)

+ LVI BI LI
(%) (%) (%)

Sonpavde
et al, 2010

71-08

RC, LA

707

T2

None

15

n/a

Zhang et
al, 2010

02-07

PC

100

T2-T4

n/a

16

n/a

Followup
(months)

Evolution
(survival
/progression)

Statistical
Significance
(MVA)

n/a

60.9
(median)

+LVI vs LVI:
HR 2.2

Independent
(RFS)

n/a

31.5
(median)

+LVI: 32%
-LVI:72%
(5-ys CSS)

Independent
(CSS)

pTN stage = pathological tumor stage


LVI = lymphovascular invasion; BI = blood vessel invasion; LI = lymphatic invasion
+LN = lymph node metastasis; +LM = loco-regional metastasis
RC = radical cystaectomy, LA = bilateral pelvic lymphadenectomy, PC = partial cystectomy; TURBT = transurethral
resection of bladder tumor
RFS = recurrence free survival, CSS = cancer-specific survival; OS = overall survival;
HR = hazard ratio, RR = relative risk

LVI [707,738]. If histological criteria are mentioned,


most studies refer to presence of tumour cells
within endothelium-lined spaces without underlying
muscular wall [60,712-714,717,721,725]. In fact,
identification of a clear-cut endothelial lining is
considered crucial in differentiating LVI from stromal
retraction, which a frequent finding around tumour
cell nests, particularly at the leading edge of invasion
(Fig 58) [739]. Stromal retraction is commonly overdiagnosed as vascular invasion; however, at least in
the breast cancer literature, there is strong evidence
that stromal retraction is not a processing artifact, has
independent prognostic significance in node negative
disease on its own merit, and may possibly be early
phase of vascular invasion [740]. Finally, varying
levels of experience of pathologists evaluating LVI,
and knowledge of the potential problems and mimics,
may influence accuracy of diagnosis, illustrating the
need for strict morphological criteria to diminish
the problem of interobserver variability [707,738].
Another related problem that is not commonly
discussed is that the general approach to assessing
LVI in diagnostic surgical pathology does not
recognize the varied biologic mechanisms involved
tumor vascularization, some of which are not readily
visible at the light microscopic level [741].

network inherent to an abnormal stroma. Moreover,


in pseudoinvasion due to tissue retraction, the
pseudothrombus usually displays a surface with
a blurred outline and shreds of cytoplasm may be
present between the pseudothrombus and the
supposed vessel wall.

In 2006, Algaba
suggested the following
morphological criteria for the diagnosis of LVI on
haematoxylin and eosin stained slides: morphological
features in favour of LVI are tumour cell thrombi within
isolated spaces with an unequivocal endothelial
lining, preferably located next to arterioles and with
a normal surrounding stromal tissue (Figure 58
A,B) [707]. The tumour thrombus is usually floating,
completely free, within the vessel lumen and may
show some fibrin precipitate and/or blood cells around
it. The tumour cells are usually tightly cohesive and
display a smooth border, the cells in the periphery
of the thrombus having a shell-like morphology. In
contrast, morphological features arguing against LVI
are tumour cell clusters, particularly with invasion of
multiple spaces, surrounded by an ectatic capillary

Figure 58. A) Unequivocal presence of urothelial


carcinoma cells within endothelial-lined vessels.
B) Invasive urothelial carcinoma with retraction
artifact.

143

The role of ancillary techniques, such as


immunohistochemistry, in the differentiation of LVI
from pseudoinvasion still remains to be defined.
Almost all published reports on bladder cancer
rely solely on haematoxylin and eosin stained
slides, and immunohistochemistry; if used at all,
immunohistochemistry seems to be reserved for
equivocal cases. However, data obtained from other
types of cancer clearly demonstrate that additional
histochemical and/or immunohistochemical staining
may be help to identify LVI and increase the accuracy
of identification with LVI by proving true vascular
invasion and/or avoiding false-positive reporting due
to overinterpretation of stromal retraction artefacts.
Thus, immunostaining using monoclonal antibodies
directed against endothelial markers, such as
CD31 and podoplanin (D2-40), the latter specific
for lymphatic endothelial cells, have been shown
to increase the detection rate of LVI compared to
conventional haematoxylin and eosin staining and
to render clinically significant information in breast
[742,743], colorectal [744], endometrial [745],
gastric [746], and oral squamous cell carcinoma
[747,748].

survival. On multivariate analysis, blood vessel


invasion detected by CD31 immunostaining proved
to be the only independent prognostic factor. Finally,
another recent study demonstrated a significant
association between lymphatic invasion, assessed
by D2-40 immunohistochemistry, and presence of
lymph node metastasis [752].
LVI is an important prognostic marker in the
assessment of bladder cancer, including both
cystectomy and TURB specimens, and should
routinely be reported upon in the pathological report
(Grade B). LVI can predict tumor behavior and
guide treatment decisions when detected in TURBT
specimens (Level 2) [721,723,733,734]. LVI was
shown as predictor of poorer outcome including
progression and metastasis in patients with T1
bladder cancer in TUR specimens [723,733,734].
When LVI is identified in TURBT it will be present in
65% of matched cystectomies and associated with
nodal metastases in 41% of cases [721]. LVI was
also shown to be an independent prognostic variable
in bladder cancer patients treated with radical
cystectomy (Level 2) [712,714,715]. In a retrospective
study of 750 patients with bladder cancer, LVI was
shown to be an independent predictor of recurrence
and decreased cause-specific and overall survival
in node-negative invasive bladder cancer treated
with cystectomy [712]. Recent multicenter studies
clearly proved LVI to be an independent prognostic
variable, significantly improving predictive accuracy
of standard prognostic models. This effect was,
however, mainly observed in node-negative cases.
Standardization of assessment is urgently needed,
and we would recommend that the evaluation of LVI
on haematoxylin and eosin stained slides should be
performed following the criteria presented by Algaba
to ensure reproducibility and reliability of pathological
diagnoses [707]. The use of endothelial cell markers,
such as CD31 or D2-40, is an interesting tool
which may facilitate detection of LVI and improve
diagnostic accuracy in equivocal cases. The general
use of immunohistochemistry in the routine setting,
however, cannot be recommended, since performing
two immunostains on even selected paraffin blocks
with bladder cancer would be extremely time
consuming and cost intensive. In addition, the
clinical consequences of an LVI diagnosis, either
in transurethral or cystectomy specimens, currently
remains controversial due to inconsistency of studies
on adjuvant chemotherapy after radical cystectomy.
The over-diagnosis of stromal retraction as LVI,
however, does remains to be a major diagnostic
problem. Adherence to strict morphological criteria
embedded in a setting of continuous medical
education and quality control is essential.

In bladder cancer, two studies, published already in


the early nineties, indicated the potential value of
these ancillary techniques: Angioinvasion diagnosed
on haematoxylin and eosin stained slides was
confirmed by immunohistochemistry in only 5 out
of 36 [749] or 2 out of 5 [750] cases, respectively.
Both studies are fairly old and did use antibodies
directed against Ulex euopaeus agglutinin I (UEA
I), von Willebrand factor, and QBEND/10, which are
antibodies that have generally been replaced with
more specific markers in current practice. Therefore,
we cannot exclude that some technical aspects
inherent to the immunohistochemical staining
evaluation may have caused the high rate of putative
false-positive reporting of LVI on haematoxylin and
eosin stained slides. In 2009, Afonso et al. published
the first systematic study investigating the potential
value of CD31 and D2-40 immunostaining in bladder
cancer including 83 patients with radical cystectomy.
[751] Using haematoxylin and eosin stained slides,
blood and lymphatic vessel invasion were diagnosed
in 19 (23%) and 18 (22%) cases, respectively.
Immunohistochemistry significantly improved the
recognition of vascular invasion for both lymphatic
and blood vessels, regarding particularly for
identifying single intravascular single tumour cells
(40% for CD31, 37% for D2-40), yet not compared
regarding malignant to tumor emboli (13% for CD31,
19% for D2-40). Overall agreement between the
three different methods used to identify LVI was
42.2%. Univariate analysis rendered vascular
invasion diagnosed on haematoxylin and eosin
stained slides as well as blood and lymphatic vessel
invasion diagnosed by immunohistochemistry as
significant prognostic variables regarding for overall

Level of evidence: 4
Recommendation: C

144

with prostatic urothelial carcinoma involving


prostatic stroma have a significantly poorer 5 year
disease specific survival than patients with urothelial
carcinoma of the prostatic urethra. Shen et al [759]
showed that patients with prostatic CIS or urethral
lamina propria invasion had a similar, but higher
incidence of lymph node metastasis and lower longterm and 5-year survival than those patients without
prostatic involvement. Similarly, prostatic stromal
invasion and periprostatic/seminal vesical invasion
had a similar, but much higher nodal metastasis
and worse survival than patients with only prostatic
CIS or urethral lamina propria invasion [759]. It is
apparent that an important etiology of the wide range
of detection of prostatic stromal invasion in various
series is likely the manner of examination of the
prostate [760].

X. Reporting on Lymph Nodes


The minimum number of lymph nodes required in
cystectomy specimen is not established. Recent
studies have emphasized the importance of lymph
node density (LND) for nodal staging, LND is defined by the ratio of positive lymph nodes to the total number of lymph nodes sampled, but again, the
minimum number of lymph node in the specimen for
optimal LND estimation yet to be established [753].
Prior studies suggested that the number of lymph
node, size of tumor metastasis and presence of
extranodal extension have prognostic significance
(Level 3) [707,712,754]. Single-institutions cystectomy series have shown for both node-negative and
node-positive patients with invasive bladder cancer
that overall survival improves with an increasing
number of lymph nodes examined [754]. In a study
by Fang et al [755], minimum number of lymph
nodes submission for radical cystectomy and pelvic
lymphadenectomy was required, having to submit
more tissue, including fat, if minimum number of 16
lymph nodes was not reached. In a span of 4 years,
the median number of lymph node increased from to
20 from 15, and this policy was found to decrease
mortality risk by 48% and significant in multivariate analysis [712]. Stephenson et al [756], showed
the prognostic importance in measuring aggregate
lymph node metastasis diameter. Fleischmann et al
[757] have recently shown that extracapsular extension of lymph node metastases defines a subgroup
with a very poor prognosis.

XII. Frozen sections


Frozen section is not an optimal method for a primary
diagnosis of invasive urothelial carcinoma or to
perform pathologic staging prior to a cystectomy [761].
The criteria for indicating open partial cystectomy
are quite strict. In these cases the urologist may
send the pathologist samples of the margins in order
to determine their status. The cystectomy specimen
should be oriented or marked in such a way that the
pathologist is able to recognize the margins. These
should be suitably inked and the selection of the
margins to be frozen should be made by means of a
perpendicular section of the full thickness of the wall.
In radical cystectomy, the ureteral margins are very
rarely affected by the tumor. If the urologist suspects
this may be the case, frozen section is indicated.
Segment of ureter submitted separately should have
proper labelling of the true margin. The situation is
completely different if the macroscopic appearance is
normal, as the intraepithelial lesions are uncommon
and only the unequivocal cases of CIS should be
diagnosed. Assessment for CIS at the margin in
frozen section may be difficult. In most cases the
status of the urethral mucosa is known through
preoperative biopsies. Some institutions determine
the status of the urethral margin intraoperatively.
Lymph node frozen section evaluation is indicated
when the possibility of not performing cystectomy is
considered.

XI. Reporting of Prostate


Involvement
In surveillance biopsies of prostate/ prostatic urethra
it is important to report the status of urethral mucosa
(CIS or not), if prostatic glands are represented in
the biopsy specimen and if so, are they involved
by urothelial neoplasia or not. Involvement of the
prostatic ducts and acini without stromal invasion
is staged as pTis using the staging system for
Urothelial carcinoma of the Prostate. Involvement
of prostate may be through direct invasion from a
bladder primary (pT4, staging system for Urothelial
carcinoma of the Bladder) or urothelial carcinoma
arising from the prostatic urethra with secondary
prostatic stromal involvement or prostatic stromal
invasion secondary to urothelial carcinoma involving
prostatic ducts and acini. In the two latter situations,
the staging system for Urothelial carcinoma of the
Prostate is used. Patients with urothelial carcinoma
involving prostatic stroma have a significantly worse
5 year disease specific survival than patients with
urothelial carcinoma of the prostatic urethra (Level
3). Cheville et al [758] demonstrated that patients

REFERENCES
1. Levi AW, Potter SR, Schoenberg MP, Epstein JI. Clinical
significance of denuded urothelium in bladder biopsy. J
Urol 2001;166:457-60.
2. Parwani AV, Levi AW, Epstein JI, Ali SZ. Urinary bladder
biopsy with denuded mucosa: denuding cystitiscytopathologic correlates. Diagnostic cytopathology
2004;30:297-300.
3. Owens CL, Epstein JI. Significance of denuded urothelium
in papillary urothelial lesions. The American journal of
surgical pathology 2007;31:298-303.

145

4. Taylor DC, Bhagavan BS, Larsen MP, Cox JA, Epstein JI.
Papillary urothelial hyperplasia. A precursor to papillary
neoplasms. The American journal of surgical pathology
1996;20:1481-8.

23. Lopez-Beltran A, Luque RJ, Mazzucchelli R, Scarpelli


M, Montironi R. Changes produced in the urothelium by
traditional and newer therapeutic procedures for bladder
cancer. Journal of clinical pathology 2002;55:641-7.

5. Readal N, Epstein JI. Papillary urothelial hyperplasia:


relationship
to
urothelial
neoplasms.
Pathology
2010;42:360-3.

24. Baker
PM,
Young
RH.
Radiation-induced
pseudocarcinomatous proliferations of the urinary bladder:
a report of 4 cases. Human pathology 2000;31:678-83.

6. Chow NH, Cairns P, Eisenberger CF, et al. Papillary


urothelial hyperplasia is a clonal precursor to papillary
transitional cell bladder cancer. International journal of
cancer Journal international du cancer 2000;89:514-8.

25. Chan TY, Epstein JI. Radiation or chemotherapy cystitis


with pseudocarcinomatous features. The American
journal of surgical pathology 2004;28:909-13.
26. Lane Z, Epstein JI. Pseudocarcinomatous epithelial
hyperplasia in the bladder unassociated with prior
irradiation or chemotherapy. The American journal of
surgical pathology 2008;32:92-7.

7. Hartmann A, Moser K, Kriegmair M, Hofstetter A,


Hofstaedter F, Knuechel R. Frequent genetic alterations
in simple urothelial hyperplasias of the bladder in
patients with papillary urothelial carcinoma. Am J Pathol
1999;154:721-7.

27. Herawi M, Parwani AV, Chan T, Ali SZ, Epstein JI. Polyoma
virus-associated cellular changes in the urine and bladder
biopsy samples: a cytohistologic correlation. The American
journal of surgical pathology 2006;30:345-50.

8. Obermann EC, Junker K, Stoehr R, et al. Frequent genetic


alterations in flat urothelial hyperplasias and concomitant
papillary bladder cancer as detected by CGH, LOH, and
FISH analyses. J Pathol 2003;199:50-7.

28. Amin MB, Gomez JA, Young RH. Urothelial transitional cell
carcinoma with endophytic growth patterns: a discussion
of patterns of invasion and problems associated with
assessment of invasion in 18 cases. The American journal
of surgical pathology 1997;21:1057-68.

9. Swierczynski SL, Epstein JI. Prognostic significance of


atypical papillary urothelial hyperplasia. Human pathology
2002;33:512-7.
10. Cifuentes L. Epithelium of vaginal type in the female trigone;
the clinical problem of trigonitis. J Urol 1947;57:1028-37.

29. Lopez-Beltran A, Cheng L, Andersson L, et al. Preneoplastic


non-papillary lesions and conditions of the urinary bladder:
an update based on the Ancona International Consultation.
Virchows Archiv : an international journal of pathology
2002;440:3-11.

11. Long ED, Shepherd RT. The incidence and significance of


vaginal metaplasia of the bladder trigone in adult women.
British journal of urology 1983;55:189-94.
12. Ozbey I, Aksoy Y, Polat O, Bicgi O, Demirel A. Squamous
metaplasia of the bladder: findings in 14 patients and review
of the literature. International urology and nephrology
1999;31:457-61.

30. McKenney JK, Gomez JA, Desai S, Lee MW, Amin MB.
Morphologic expressions of urothelial carcinoma in situ: a
detailed evaluation of its histologic patterns with emphasis
on carcinoma in situ with microinvasion. The American
journal of surgical pathology 2001;25:356-62.

13. Guo CC, Fine SW, Epstein JI. Noninvasive squamous


lesions in the urinary bladder: a clinicopathologic analysis
of 29 cases. The American journal of surgical pathology
2006;30:883-91.

31. Shen SS, Guo, C.C., Ro, J.Y., et al. Common problems in
interpretation of early invasive bladder cancer: A consensus
study. Mod Pathol 2011;24:224A.
32. Tosoni I, Wagner U, Sauter G, et al. Clinical significance
of interobserver differences in the staging and grading of
superficial bladder cancer. BJU international 2000;85:4853.
33. Bol MG, Baak JP, Buhr-Wildhagen S, et al. Reproducibility
and prognostic variability of grade and lamina propria
invasion in stages Ta, T1 urothelial carcinoma of the
bladder. J Urol 2003;169:1291-4.
34. Dinney CP, Ramirez EI, Swanson DA, Ro JY, Babaian
RJ, von Eschenbach AC. Management of transitional
cell carcinoma involving von Brunns nests. J Urol
1995;153:944-9.
35. Farrow GM, DC, U. Observations on microinvasive
transitional cell carcioma of the urinary bladder. Clin Oncol
1982;1:609-15.
36. Amin MB, Young RH. Intraepithelial lesions of the urinary
bladder with a discussion of the histogenesis of urothelial
neoplasia. Seminars in diagnostic pathology 1997;14:8497.
37. Paner GP, Ro JY, Wojcik EM, Venkataraman G, Datta MW,
Amin MB. Further characterization of the muscle layers
and lamina propria of the urinary bladder by systematic
histologic mapping: implications for pathologic staging of
invasive urothelial carcinoma. The American journal of
surgical pathology 2007;31:1420-9.
38. Angulo JC, Lopez JI, Grignon DJ, Sanchez-Chapado M.
Muscularis mucosa differentiates two populations with
different prognosis in stage T1 bladder cancer. Urology
1995;45:47-53.

14. Khan MS, Thornhill JA, Gaffney E, Loftus B, Butler MR.


Keratinising squamous metaplasia of the bladder: natural
history and rationalization of management based on review
of 54 years experience. European urology 2002;42:46974.
15. Broecker BH, Klein FA, Hackler RH. Cancer of the bladder
in spinal cord injury patients. J Urol 1981;125:196-7.
16. Hess MJ, Zhan EH, Foo DK, Yalla SV. Bladder cancer
in patients with spinal cord injury. J Spinal Cord Med
2003;26:335-8.
17. Kaufman JM, Fam B, Jacobs SC, et al. Bladder cancer and
squamous metaplasia in spinal cord injury patients. J Urol
1977;118:967-71.
18. Bullock PS, Thoni DE, Murphy WM. The significance of
colonic mucosa (intestinal metaplasia) involving the urinary
tract. Cancer 1987;59:2086-90.
19. Lin JI, Yong HS, Tseng CH, Marsidi PS, Choy C,
Pilloff B. Diffuse cystitis glandularis. Associated with
adenocarcinomatous change. Urology 1980;15:411-5.
20. Corica FA, Husmann DA, Churchill BM, et al. Intestinal
metaplasia is not a strong risk factor for bladder cancer:
study of 53 cases with long-term follow-up. Urology
1997;50:427-31.
21. Smith AK, Hansel DE, Jones JS. Role of cystitis cystica
et glandularis and intestinal metaplasia in development of
bladder carcinoma. Urology 2008;71:915-8.
22. Young RH, Bostwick DG. Florid cystitis glandularis
of intestinal type with mucin extravasation: a mimic
of adenocarcinoma. The American journal of surgical
pathology 1996;20:1462-8.

39. Hasui Y, Osada Y, Kitada S, Nishi S. Significance of invasion


to the muscularis mucosae on the progression of superficial
bladder cancer. Urology 1994;43:782-6.

146

40. Holmang S, Hedelin H, Anderstrom C, Holmberg E,


Johansson SL. The importance of the depth of invasion in
stage T1 bladder carcinoma: a prospective cohort study. J
Urol 1997;157:800-3; discussion 4.

57. Girgin C, Sezer A, Delibas M, Sahin O, Oder M, Dincel C.


Impact of the level of muscle invasion in organ-confined
bladder cancer. Urologia internationalis 2007;78:145-9.
58. Tokgoz H, Turkolmez K, Resorlu B, Kose K, Tulunay O,
Beduk Y. Pathological staging of muscle invasive bladder
cancer. Is substaging of pT2 tumors really necessary?
International braz j urol : official journal of the Brazilian
Society of Urology 2007;33:777-83; discussion 83-4.

41. Orsola A, Trias I, Raventos CX, et al. Initial high-grade


T1 urothelial cell carcinoma: feasibility and prognostic
significance of lamina propria invasion microstaging
(T1a/b/c) in BCG-treated and BCG-non-treated patients.
European urology 2005;48:231-8; discussion 8.

59. Yu RJ, Stein JP, Cai J, Miranda G, Groshen S, Skinner


DG. Superficial (pT2a) and deep (pT2b) muscle invasion
in pathological staging of bladder cancer following radical
cystectomy. J Urol 2006;176:493-8; discussion 8-9.

42. Smits G, Schaafsma E, Kiemeney L, Caris C, Debruyne F,


Witjes JA. Microstaging of pT1 transitional cell carcinoma of
the bladder: identification of subgroups with distinct risks of
progression. Urology 1998;52:1009-13; discussion 13-4.

60. Sonpavde G, Khan MM, Svatek RS, et al. Prognostic risk


stratification of pathological stage T2N0 bladder cancer
after radical cystectomy. BJU international 2010.

43. Sozen S, Akbal C, Sokmensuer C, Ekici S, Ozen H.


Microstaging of pT1 transitional cell carcinoma of the
bladder. Does it really differentiate two populations
with different prognoses? (pT1 subcategory). Urologia
internationalis 2002;69:200-6.

61. Tilki D, Reich O, Karakiewicz PI, et al. Validation of the


AJCC TNM substaging of pT2 bladder cancer: deep muscle
invasion is associated with significantly worse outcome.
European urology 2010;58:112-7.

44. Younes M, Sussman J, True LD. The usefulness of the


level of the muscularis mucosae in the staging of invasive
transitional cell carcinoma of the urinary bladder. Cancer
1990;66:543-8.

62. Philip AT, Amin MB, Tamboli P, Lee TJ, Hill CE, Ro JY.
Intravesical adipose tissue: a quantitative study of its
presence and location with implications for therapy and
prognosis. The American journal of surgical pathology
2000;24:1286-90.

45. van der Aa MN, van Leenders GJ, Steyerberg EW, et al. A
new system for substaging pT1 papillary bladder cancer: a
prognostic evaluation. Human pathology 2005;36:981-6.

63. Bastian PJ, Hutterer GC, Shariat SF, et al. Macroscopic,


but not microscopic, perivesical fat invasion at radical
cystectomy is an adverse predictor of recurrence and
survival. BJU international 2008;101:450-4.

46. Cheng L, Neumann RM, Weaver AL, Spotts BE, Bostwick


DG. Predicting cancer progression in patients with stage T1
bladder carcinoma. J Clin Oncol 1999;17:3182-7.
47. Cheng L, Weaver AL, Neumann RM, Scherer BG, Bostwick
DG. Substaging of T1 bladder carcinoma based on the
depth of invasion as measured by micrometer: A new
proposal. Cancer 1999;86:1035-43.

64. Boudreaux KJ, Jr., Chang SS, Lowrance WT, et al.


Comparison of American Joint Committee on Cancer
pathologic stage T3a versus T3b urothelial carcinoma:
analysis of patient outcomes. Cancer 2009;115:770-5.

48. Platz CE, Cohen MB, Jones MP, Olson DB, Lynch CF. Is
microstaging of early invasive cancer of the urinary bladder
possible or useful? Mod Pathol 1996;9:1035-9.

65. Quek ML, Stein JP, Clark PE, et al. Microscopic and gross
extravesical extension in pathological staging of bladder
cancer. J Urol 2004;171:640-5.

49. Ro JY, Ayala AG, el-Naggar A. Muscularis mucosa of


urinary bladder. Importance for staging and treatment. The
American journal of surgical pathology 1987;11:668-73.

66. Quek ML, Stein JP, Clark PE, et al. Natural history of
surgically treated bladder carcinoma with extravesical
tumor extension. Cancer 2003;98:955-61.

50. Dixon JS, Gosling JA. Histology and fine structure of the
muscularis mucosae of the human urinary bladder. Journal
of anatomy 1983;136:265-71.

67. Scosyrev E, Yao J, Messing E. Microscopic invasion of


perivesical fat by urothelial carcinoma: implications for
prognosis and pathology practice. Urology 2010;76:90813; discussion 14.

51. Keep JC, Piehl M, Miller A, Oyasu R. Invasive carcinomas


of the urinary bladder. Evaluation of tunica muscularis
mucosae involvement. American journal of clinical
pathology 1989;91:575-9.

68. Tilki D, Svatek RS, Karakiewicz PI, et al. pT3 Substaging


is a prognostic indicator for lymph node negative urothelial
carcinoma of the bladder. J Urol 2010;184:470-4.

52. Weaver MG, Abdul-Karim FW. The prevalence and


character of the muscularis mucosae of the human urinary
bladder. Histopathology 1990;17:563-6.

69. Montironi R, Mazzucchelli R, Scarpelli M, Lopez-Beltran A,


Cheng L. Morphological diagnosis of urothelial neoplasms.
Journal of clinical pathology 2008;61:3-10.

53. Vakar-Lopez F, Shen SS, Zhang S, Tamboli P, Ayala AG,


Ro JY. Muscularis mucosae of the urinary bladder revisited
with emphasis on its hyperplastic patterns: a study of a
large series of cystectomy specimens. Annals of diagnostic
pathology 2007;11:395-401.

70. Hodges KB, Lopez-Beltran A, Davidson DD, Montironi R,


Cheng L. Urothelial dysplasia and other flat lesions of the
urinary bladder: clinicopathologic and molecular features.
Human pathology 2010;41:155-62.
71. Montironi R, Lopez-Beltran A, Scarpelli M, Mazzucchelli
R, Cheng L. Morphological classification and definition of
benign, preneoplastic and non-invasive neoplastic lesions
of the urinary bladder. Histopathology 2008;53:621-33.

54. Aydin A, Ucak R, Karakok M, Guldur ME, Kocer NE.


Vascular plexus is a differentation criterion for muscularis
mucosa from muscularis propria in small biopsies and
transurethral resection materials of urinary bladder?
International urology and nephrology 2002;34:315-9.

72. Cheng L, Cheville JC, Neumann RM, Bostwick DG. Natural


history of urothelial dysplasia of the bladder. The American
journal of surgical pathology 1999;23:443-7.

55. Engel P, Anagnostaki L, Braendstrup O. The muscularis


mucosae of the human urinary bladder. Implications for
tumor staging on biopsies. Scandinavian journal of urology
and nephrology 1992;26:249-52.

73. Birkhahn M, Mitra AP, Williams AJ, et al. Predicting


recurrence and progression of noninvasive papillary
bladder cancer at initial presentation based on quantitative
gene expression profiles. European urology 2010;57:1220.

56. Miyamoto H, Epstein JI. Transurethral resection specimens


of the bladder: outcome of invasive urothelial cancer
involving muscle bundles indeterminate between muscularis
mucosae and muscularis propria. Urology 2010;76:600-2.

74. Burger M. Editorial comment on: prognostic accuracy of


individual uropathologists in noninvasive urinary bladder

147

carcinoma: a multicentre study comparing the 1973 and


2004 World Health Organisation classifications. European
urology 2010;57:858.

90. Cheng L, Bostwick DG. Overdiagnosis of bladder


carcinoma. Anal Quant Cytol Histol 2008;30:261-4.
91. Lott S, Wang M, Zhang S, et al. FGFR3 and TP53 mutation
analysis in inverted urothelial papilloma: incidence and
etiological considerations. Mod Pathol 2009;22:627-32.

75. Guey LT, Garcia-Closas M, Murta-Nascimento C, et


al. Genetic susceptibility to distinct bladder cancer
subphenotypes. European urology 2010;57:283-92.

92. Sung MT, Maclennan GT, Lopez-Beltran A, Montironi R,


Cheng L. Natural history of urothelial inverted papilloma.
Cancer 2006;107:2622-7.

76. May M, Brookman-Amissah S, Roigas J, et al. Prognostic


accuracy of individual uropathologists in noninvasive urinary
bladder carcinoma: a multicentre study comparing the
1973 and 2004 World Health Organisation classifications.
European urology 2010;57:850-8.

93. Cheng L, Lopez-Beltran A, MacLennan, Montironi R,


Bostwick DG. Neoplasms of the urinary bladder. In:
Bostwick DG, Cheng L., ed. Urologic Surgical Pathology. 2
ed. Philadelphia: Mosby/Elsevier; 2008:289-92.

77. van Rhijn BW, Burger M, Lotan Y, et al. Recurrence and


progression of disease in non-muscle-invasive bladder
cancer: from epidemiology to treatment strategy. European
urology 2009;56:430-42.

94. Jones TD, Zhang S, Lopez-Beltran A, et al. Urothelial


carcinoma with an inverted growth pattern can be
distinguished from inverted papilloma by fluorescence in
situ hybridization, immunohistochemistry, and morphologic
analysis. The American journal of surgical pathology
2007;31:1861-7.

78. Yorukoglu K, Tuna B, Kirkali Z. Re: Matthias May, Sabine


Brookman-Amissah, Jan Roigas, et al. Prognostic accuracy
of individual uropathologists in noninvasive urinary bladder
carcinoma: a multicentre study comparing the 1973 and
2004 World Health Organisation classifications. Eur Urol
2010;57:850-8. European urology 2010;58:e7; author reply
e8.

95. Cao D, Vollmer RT, Luly J, et al. Comparison of 2004 and


1973 World Health Organization grading systems and their
relationship to pathologic staging for predicting long-term
prognosis in patients with urothelial carcinoma. Urology
2010;76:593-9.

79. Cheng L, Darson M, Cheville JC, et al. Urothelial papilloma


of the bladder. Clinical and biologic implications. Cancer
1999;86:2098-101.

96. Burger M, Denzinger S, Wieland WF, Stief CG, Hartmann


A, Zaak D. Does the current World Health Organization
classification predict the outcome better in patients with
noninvasive bladder cancer of early or regular onset? BJU
international 2008;102:194-7.

80. Magi-Galluzzi C, Epstein JI. Urothelial papilloma of the


bladder: a review of 34 de novo cases. The American
journal of surgical pathology 2004;28:1615-20.
81. McKenney JK, Amin MB, Young RH. Urothelial (transitional
cell) papilloma of the urinary bladder: a clinicopathologic
study of 26 cases. Mod Pathol 2003;16:623-9.
82. Babjuk M. Second resection for non-muscle-invasive
bladder carcinoma: current role and future perspectives.
European urology 2010;58:191-2.
83. Dalbagni G, Vora K, Kaag M, et al. Clinical outcome in a
contemporary series of restaged patients with clinical T1
bladder cancer. European urology 2009;56:903-10.
84. Divrik RT, Sahin AF, Ergor G. Reply from authors re: Marko
Babjuk. Second resection for non-muscle-invasive bladder
carcinoma: current role and future perspectives. Eur Urol
2010;58:191-2 and Giacomo Novara, Vincenzo Ficarra.
Does routine second transurethral resection affect the
long-term outcome of patients with T1 bladder cancer?
Why a flawed randomized controlled trial cannot address
the issue. Eur Urol 2010;58:193-4. European urology
2010;58:195-6.
85. Divrik RT, Sahin AF, Yildirim U, Altok M, Zorlu F. Impact of
routine second transurethral resection on the long-term
outcome of patients with newly diagnosed pT1 urothelial
carcinoma with respect to recurrence, progression rate,
and disease-specific survival: a prospective randomised
clinical trial. European urology 2010;58:185-90.
86. Fritsche HM, Burger M, Svatek RS, et al. Characteristics
and outcomes of patients with clinical T1 grade 3 urothelial
carcinoma treated with radical cystectomy: results from an
international cohort. European urology 2010;57:300-9.
87. Kulkarni GS, Hakenberg OW, Gschwend JE, et al. An
updated critical analysis of the treatment strategy for
newly diagnosed high-grade T1 (previously T1G3) bladder
cancer. European urology 2010;57:60-70.
88. Novara G, Ficarra V. Does routine second transurethral
resection affect the long-term outcome of patients with
T1 bladder cancer? Why a flawed randomized controlled
trial cannot address the issue. European urology
2010;58:193-4.
89. van Rhijn BW, Zuiverloon TC, Vis AN, et al. Molecular grade
(FGFR3/MIB-1) and EORTC risk scores are predictive in
primary non-muscle-invasive bladder cancer. European
urology 2010;58:433-41.

97. Yin H, Leong AS. Histologic grading of noninvasive papillary


urothelial tumors: validation of the 1998 WHO/ISUP system
by immunophenotyping and follow-up. American journal of
clinical pathology 2004;121:679-87.
98. Schned AR, Andrew AS, Marsit CJ, Zens MS, Kelsey
KT, Karagas MR. Survival following the diagnosis of
noninvasive bladder cancer: WHO/International Society of
Urological Pathology versus WHO classification systems. J
Urol 2007;178:1196-200; discussion 200.
99. Samaratunga H, Makarov DV, Epstein JI. Comparison
of WHO/ISUP and WHO classification of noninvasive
papillary urothelial neoplasms for risk of progression.
Urology 2002;60:315-9.
100. Oosterhuis JW, Schapers RF, Janssen-Heijnen ML,
Pauwels RP, Newling DW, ten Kate F. Histological grading
of papillary urothelial carcinoma of the bladder: prognostic
value of the 1998 WHO/ISUP classification system and
comparison with conventional grading systems. Journal of
clinical pathology 2002;55:900-5.
101. Holmang S, Andius P, Hedelin H, Wester K, Busch C,
Johansson SL. Stage progression in Ta papillary urothelial
tumors: relationship to grade, immunohistochemical
expression of tumor markers, mitotic frequency and DNA
ploidy. J Urol 2001;165:1124-8; discussion 8-30.
102. Miyamoto H, Brimo F, Schultz L, et al. Low-grade
papillary urothelial carcinoma of the urinary bladder:
a clinicopathologic analysis of a post-World Health
Organization/International Society of Urological Pathology
classification cohort from a single academic center. Arch
Pathol Lab Med 2010;134:1160-3.
103. van Rhijn BW, van Leenders GJ, Ooms BC, et al. The
pathologists mean grade is constant and individualizes
the prognostic value of bladder cancer grading. European
urology 2010;57:1052-7.
104. Gonul, II, Poyraz A, Unsal C, Acar C, Alkibay T. Comparison
of 1998 WHO/ISUP and 1973 WHO classifications for
interobserver variability in grading of papillary urothelial
neoplasms of the bladder. Pathological evaluation of 258
cases. Urologia internationalis 2007;78:338-44.

148

105. Mamoon N, Iqbal MA, Jamal S, Luqman M. Urothelial


neoplasia of the urinary bladder--comparison of
interobserver variability for WHO Classification 1972 with
WHO/ISUP Consensus Classification 1998. J Ayub Med
Coll Abbottabad 2006;18:4-8.

122. Lopez-Beltran A, Martin J, Garcia J, Toro M. Squamous and


glandular differentiation in urothelial bladder carcinomas.
Histopathology, histochemistry and immunohistochemical
expression of carcinoembryonic antigen. Histology and
histopathology 1988;3:63-8.

106. Campbell PA, Conrad RJ, Campbell CM, Nicol DL,


MacTaggart P. Papillary urothelial neoplasm of low
malignant potential: reliability of diagnosis and outcome.
BJU international 2004;93:1228-31.

123. Lopez-Beltran A, Requena MJ, Alvarez-Kindelan J,


Quintero A, Blanca A, Montironi R. Squamous differentiation
in primary urothelial carcinoma of the urinary tract as seen
by MAC387 immunohistochemistry. Journal of clinical
pathology 2007;60:332-5.

107. Yorukoglu K, Tuna B, Dikicioglu E, et al. Reproducibility of


the 1998 World Health Organization/International Society
of Urologic Pathology classification of papillary urothelial
neoplasms of the urinary bladder. Virchows Archiv : an
international journal of pathology 2003;443:734-40.

124. Lopez-Beltran A, Requena MJ, Cheng L, Montironi R.


Pathological variants of invasive bladder cancer according
to their suggested clinical significance. BJU international
2008;101:275-81.

108. Murphy WM, Takezawa K, Maruniak NA. Interobserver


discrepancy using the 1998 World Health Organization/
International Society of Urologic Pathology classification
of urothelial neoplasms: practical choices for patient care.
J Urol 2002;168:968-72.

125. Antunes AA, Nesrallah LJ, DallOglio MF, et al. The role of
squamous differentiation in patients with transitional cell
carcinoma of the bladder treated with radical cystectomy.
International braz j urol : official journal of the Brazilian
Society of Urology 2007;33:339-45; discussion 46.

109. Billis A, Carvalho RB, Mattos AC, et al. Tumor grade


heterogeneity in urothelial bladder carcinoma--proposal of
a system using combined numbers. Scandinavian journal
of urology and nephrology 2001;35:275-9.

126. Black PC, Brown GA, Dinney CP. The impact of variant
histology on the outcome of bladder cancer treated with
curative intent. Urologic oncology 2009;27:3-7.
127. Dalbagni G, Genega E, Hashibe M, et al. Cystectomy
for bladder cancer: a contemporary series. J Urol
2001;165:1111-6.
128. Logothetis CJ, Dexeus FH, Chong C, et al. Cisplatin,
cyclophosphamide and doxorubicin chemotherapy
for unresectable urothelial tumors: the M.D. Anderson
experience. J Urol 1989;141:33-7.
129. Martin JE, Jenkins BJ, Zuk RJ, Blandy JP, Baithun SI.
Clinical importance of squamous metaplasia in invasive
transitional cell carcinoma of the bladder. Journal of
clinical pathology 1989;42:250-3.
130. Rogers CG, Palapattu GS, Shariat SF, et al. Clinical
outcomes following radical cystectomy for primary
nontransitional cell carcinoma of the bladder compared
to transitional cell carcinoma of the bladder. J Urol
2006;175:2048-53; discussion 53.
131. Scosyrev E, Ely BW, Messing EM, et al. Do mixed
histological features affect survival benefit from neoadjuvant
platinum-based combination chemotherapy in patients with
locally advanced bladder cancer? A secondary analysis of
Southwest Oncology Group-Directed Intergroup Study
(S8710). BJU international 2010.
132. Gulmann C, Paner, G.P., Amin, M.B., et al.
Immunohistochemical Profile to Distinguish Urothelial
from Squamous Differentiation in Carcinomas of Urothelial
Tract. Submitted 2010.
133. Kunze E, Francksen B, Schulz H. Expression of MUC5AC
apomucin in transitional cell carcinomas of the urinary
bladder and its possible role in the development of
mucus-secreting adenocarcinomas. Virchows Archiv : an
international journal of pathology 2001;439:609-15.
134. Amin MB, Ro JY, el-Sharkawy T, et al. Micropapillary
variant of transitional cell carcinoma of the urinary bladder.
Histologic pattern resembling ovarian papillary serous
carcinoma. The American journal of surgical pathology
1994;18:1224-32.
135. Guo CC, Tamboli P, Czerniak B. Micropapillary variant
of urothelial carcinoma in the upper urinary tract: a
clinicopathologic study of 11 cases. Arch Pathol Lab Med
2009;133:62-6.
136. Oh YL, Kim KR. Micropapillary variant of transitional
cell carcinoma of the ureter. Pathology international
2000;50:52-6.
137. Comperat E, Roupret M, Yaxley J, et al. Micropapillary
urothelial carcinoma of the urinary bladder: a
clinicopathological analysis of 72 cases. Pathology
2010;42:650-4.

110. Bircan S, Candir O, Serel TA. Comparison of WHO 1973,


WHO/ISUP 1998, WHO 1999 grade and combined scoring
systems in evaluation of bladder carcinoma. Urologia
internationalis 2004;73:201-8.
111. Cheng L, Neumann RM, Nehra A, Spotts BE, Weaver
AL, Bostwick DG. Cancer heterogeneity and its biologic
implications in the grading of urothelial carcinoma. Cancer
2000;88:1663-70.
112. Kruger S, Thorns C, Bohle A, Feller AC. Prognostic
significance of a grading system considering tumor
heterogeneity in muscle-invasive urothelial carcinoma of
the urinary bladder. International urology and nephrology
2003;35:169-73.
113. Epstein JI, Amin MB, Reuter VR, Mostofi FK. The World
Health Organization/International Society of Urological
Pathology consensus classification of urothelial
(transitional cell) neoplasms of the urinary bladder. Bladder
Consensus Conference Committee. The American journal
of surgical pathology 1998;22:1435-48.
114. Cheng L, Cheville JC, Sebo TJ, Eble JN, Bostwick DG.
Atypical nephrogenic metaplasia of the urinary tract: a
precursor lesion? Cancer 2000;88:853-61.
115. Otto W, Denzinger S, Fritsche HM, et al. The WHO
classification of 1973 is more suitable than the WHO
classification of 2004 for predicting survival in pT1 urothelial
bladder cancer. BJU international 2011;107:404-8.
116. Kamel MH, Daly PJ, Khan MF, Kay EW, OKelly P, Hickey
DP. Survival and progression in high grade tumour subset
of G2 and G3 pT1 bladder transitional cell carcinoma. Eur
J Surg Oncol 2006;32:1139-43.
117. Amin MB. Histological variants of urothelial carcinoma:
diagnostic, therapeutic and prognostic implications. Mod
Pathol 2009;22 Suppl 2:S96-S118.
118. Eble JN, Young RH. Carcinoma of the urinary bladder: a
review of its diverse morphology. Seminars in diagnostic
pathology 1997;14:98-108.
119. Eble JN SG, Epstein JI, Sesterhenn IA. Pathology &
Genetics. Tumours of the Urinary System and Male
Genital Organs. Albany, NY: IARC Press; 2004.
120. Epstein JI, Reuter, V.E., Amin, M.B. Biopsy Interpretation of
the Bladder. 2nd ed. Philadelphia, PA: Lippincott Williams
and Wilkins; 2010.
121. Lopez-Beltran A, Cheng L. Histologic variants of urothelial
carcinoma: differential diagnosis and clinical implications.
Human pathology 2006;37:1371-88.

149

138. Kamat AM, Dinney CP, Gee JR, et al. Micropapillary


bladder cancer: a review of the University of Texas M. D.
Anderson Cancer Center experience with 100 consecutive
patients. Cancer 2007;110:62-7.

MUC1, CA125, and Her2Neu in invasive micropapillary


carcinoma of the urinary tract and typical invasive
urothelial carcinoma with retraction artifact. Mod Pathol
2009;22:660-7.

139. Gaya JM, Palou J, Algaba F, Arce J, Rodriguez-Faba O,


Villavicencio H. The case for conservative management
in the treatment of patients with non-muscle-invasive
micropapillary bladder carcinoma without carcinoma in
situ. The Canadian journal of urology 2010;17:5370-6.

156. Samaratunga H, Khoo K. Micropapillary variant of urothelial


carcinoma of the urinary bladder; a clinicopathological and
immunohistochemical study. Histopathology 2004;45:5564.
157. Comperat E, Roupret M, Conort P, et al. Aurora-A/STK-15
is differentially expressed in the micropapillary variant of
bladder cancer. Urologia internationalis 2009;82:312-7.

140. Sangoi AR, Beck AH, Amin MB, et al. Interobserver


reproducibility in the diagnosis of invasive micropapillary
carcinoma of the urinary tract among urologic pathologists.
The American journal of surgical pathology 2010;34:136776.

158. Lim MG, Adsay NV, Grignon DJ, Osunkoya AO.


E-cadherin expression in plasmacytoid, signet ring cell and
micropapillary variants of urothelial carcinoma: comparison
with usual-type high-grade urothelial carcinoma. Mod
Pathol 2011;24:241-7.

141. Perepletchikov AM, Parwani AV. Micropapillary urothelial


carcinoma: clinico-pathologic review. Pathology, research
and practice 2009;205:807-10.

159. Lopez-Beltran A, Montironi R, Blanca A, Cheng L. Invasive


micropapillary urothelial carcinoma of the bladder. Human
pathology 2010;41:1159-64.

142. Nishizawa K, Kobayashi T, Mitsumori K, Ide Y, Watanabe


J, Ogura K. Micropapillary bladder cancer. International
journal of urology : official journal of the Japanese
Urological Association 2005;12:506-8.

160. Lin O, Cardillo M, Dalbagni G, Linkov I, Hutchinson B,


Reuter VE. Nested variant of urothelial carcinoma: a
clinicopathologic and immunohistochemical study of 12
cases. Mod Pathol 2003;16:1289-98.

143. Zhai QJ, Black J, Ayala AG, Ro JY. Histologic variants


of infiltrating urothelial carcinoma. Arch Pathol Lab Med
2007;131:1244-56.

161. Murphy WM, Deana DG. The nested variant of transitional


cell carcinoma: a neoplasm resembling proliferation of
Brunns nests. Mod Pathol 1992;5:240-3.

144. Nicolas MM, Jagirdar JS, Arisco AM, Valente PT.


Micropapillary carcinoma of the urinary bladder: Report
of a case and review of its cytologic features. Diagnostic
cytopathology 2010.

162. Nese N, Kandiloglu AR, Atesci YZ. Nested variant of


transitional cell carcinoma with osseous metaplasia of
the urinary bladder: a case report and review of published
reports. International journal of urology : official journal of
the Japanese Urological Association 2007;14:365-7.

145. Baschinsky DY, Chen JH, Vadmal MS, Lucas JG,


Bahnson RR, Niemann TH. Carcinosarcoma of the urinary
bladder--an aggressive tumor with diverse histogenesis.
A clinicopathologic study of 4 cases and review of the
literature. Arch Pathol Lab Med 2000;124:1172-8.

163. Stern JB. Unusual benign bladder tumor of Brunn nest


origin. Urology 1979;14:288-9.
164. Talbert ML, Young RH. Carcinomas of the urinary bladder
with deceptively benign-appearing foci. A report of three
cases. The American journal of surgical pathology
1989;13:374-81.

146. Demirovic A, Marusic Z, Lenicek T, et al. CD138-positive


plasmacytoid urothelial carcinoma of urinary bladder with
focal micropapillary features. Tumori 2010;96:358-60.
147. Lim M, Adsay NV, Grignon D, Osunkoya AO. Urothelial
carcinoma with villoglandular differentiation: a study of 14
cases. Mod Pathol 2009;22:1280-6.

165. Cardillo M, Reuter VE, Lin O. Cytologic features of the


nested variant of urothelial carcinoma: a study of seven
cases. Cancer 2003;99:23-7.

148. Lopez-Beltran A, Amin MB, Oliveira PS, et al. Urothelial


carcinoma of the bladder, lipid cell variant: clinicopathologic
findings and LOH analysis. The American journal of
surgical pathology 2010;34:371-6.

166. Drew PA, Furman J, Civantos F, Murphy WM. The nested


variant of transitional cell carcinoma: an aggressive
neoplasm with innocuous histology. Mod Pathol
1996;9:989-94.

149. Lopez-Beltran A, Blanca A, Montironi R, Cheng L,


Regueiro JC. Pleomorphic giant cell carcinoma of the
urinary bladder. Human pathology 2009;40:1461-6.

167. Ro JY, Lapham, R., Amin, M. Deceptively bland


transitional cell carcinoma of the urinary bladder-further
characterization of subtle and diagnostically tracherous
patterns of invasion in urothelial neoplasia. Adv Anat
Pathol 1997;4:244-51.

150. Lopez-Beltran A, Requena MJ, Montironi R, Blanca


A, Cheng L. Plasmacytoid urothelial carcinoma of the
bladder. Human pathology 2009;40:1023-8.
151. Regalado JJ. Mixed micropapillary and trophoblastic
carcinoma of bladder: report of a first case with new
immunohistochemical evidence of urothelial origin. Human
pathology 2004;35:382-4.

168. Volmar KE, Chan TY, De Marzo AM, Epstein JI. Florid
von Brunn nests mimicking urothelial carcinoma: a
morphologic and immunohistochemical comparison to
the nested variant of urothelial carcinoma. The American
journal of surgical pathology 2003;27:1243-52.

152. Lotan TL, Ye H, Melamed J, Wu XR, Shih Ie M, Epstein JI.


Immunohistochemical panel to identify the primary site of
invasive micropapillary carcinoma. The American journal
of surgical pathology 2009;33:1037-41.

169. Young RH, Oliva E. Transitional cell carcinomas of the


urinary bladder that may be underdiagnosed. A report of
four invasive cases exemplifying the homology between
neoplastic and non-neoplastic transitional cell lesions. The
American journal of surgical pathology 1996;20:1448-54.

153. Nassar H, Pansare V, Zhang H, et al. Pathogenesis


of invasive micropapillary carcinoma: role of MUC1
glycoprotein. Mod Pathol 2004;17:1045-50.

170. McKenney JK, Amin MB. The role of immunohistochemistry


in the diagnosis of urinary bladder neoplasms. Seminars
in diagnostic pathology 2005;22:69-87.

154. Ohtsuki Y, Kuroda N, Umeoka T, et al. KL-6 is another


useful marker in assessing a micropapillary pattern in
carcinomas of the breast and urinary bladder, but not the
colon. Medical molecular morphology 2009;42:123-7.

171. Holmang S, Johansson SL. The nested variant of


transitional cell carcinoma--a rare neoplasm with poor
prognosis. Scandinavian journal of urology and nephrology
2001;35:102-5.

155. Sangoi AR, Higgins JP, Rouse RV, Schneider AG,


McKenney JK. Immunohistochemical comparison of

150

172. Liedberg F, Chebil G, Davidsson T, Gadaleanu V, Grabe


M, Mansson W. The nested variant of urothelial carcinoma:
a rare but important bladder neoplasm with aggressive
behavior. Three case reports and a review of the literature.
Urologic oncology 2003;21:7-9.

Organization International Histological Classification of


Tumours. Histological Typing of Urinary Bladder Tumours.
2nd ed. Berlin, Heidelberg: Springer Verlag; 1999.
190. Leroy X, Gonzalez S, Zini L, Aubert S. Lipoid-cell
variant of urothelial carcinoma: a clinicopathologic and
immunohistochemical study of five cases. The American
journal of surgical pathology 2007;31:770-3.

173. Huang Q, Chu PG, Lau SK, Weiss LM. Urothelial carcinoma
of the urinary bladder with a component of acinar/tubular
type differentiation simulating prostatic adenocarcinoma.
Human pathology 2004;35:769-73.

191. Shimada K, Nakamura M, Konishi N. A case of


urothelial carcinoma with triple variants featuring nested,
plasmacytoid, and lipoid cell morphology. Diagnostic
cytopathology 2009;37:272-6.

174. Association of Directors of anatomic and surgical


pathology., 2003. (Accessed at http://www.adasp.org/
Checklists/checklists.htm.)

192. Soylu A, Aydin NE, Yilmaz U, Kutlu R, Gunes A. Urothelial


carcinoma featuring lipid cell and plasmacytoid morphology
with poor prognostic outcome. Urology 2005;65:797.

175. Paz A, Rath-Wolfson L, Lask D, et al. The clinical and


histological features of transitional cell carcinoma of the
bladder with microcysts: analysis of 12 cases. British
journal of urology 1997;79:722-5.

193. Bates AW, Baithun SI. Secondary neoplasms of the


bladder are histological mimics of nontransitional cell
primary tumours: clinicopathological and histological
features of 282 cases. Histopathology 2000;36:32-40.

176. Sari A, Uyaroglu MA, Ermete M, Oder M, Girgin C, Dincer


C. Microcystic urothelial carcinoma of the urinary bladder
metastatic to the penis. Pathology oncology research :
POR 2007;13:170-3.

194. Morichetti D, Mazzucchelli R, Lopez-Beltran A, et al.


Secondary neoplasms of the urinary system and male
genital organs. BJU international 2009;104:770-6.

177. Young RH, Zukerberg LR. Microcystic transitional cell


carcinomas of the urinary bladder. A report of four cases.
American journal of clinical pathology 1991;96:635-9.

195. Grignon DJ, Ro JY, Ayala AG, Johnson DE. Primary


signet-ring cell carcinoma of the urinary bladder. American
journal of clinical pathology 1991;95:13-20.

178. Leroy X, Leteurtre E, De La Taille A, Augusto D, Biserte


J, Gosselin B. Microcystic transitional cell carcinoma: a
report of 2 cases arising in the renal pelvis. Arch Pathol
Lab Med 2002;126:859-61.

196. Berman JJ, Seidman, J.D., Yetter, R., Moore, G.W.


Clear cell dysplasia of the bladder. Report of a case
with flow cytometric analysis. Anal Quant Cytol Histol
1991;13:391-4.

179. Mai KT, Elmontaser G, Perkins DG, Yazdi HM, Stinson WA,
Thijssen A. Histopathological and immunohistochemical
study of papillary urothelial neoplasms of low malignant
potential and grade associated with extensive invasive
low-grade urothelial carcinoma. BJU international
2004;94:544-7.

197. Lopez Beltran A, Sauter, G., Gasser, T. et al. Infiltrating


urothelial carcinoma. Pathology and Genetics of Tumours
of the Urinary System and Male Genital Organs. Lyon,
France: IARC Press; 2004.

180. Amin MB, Ro JY, Lee KM, et al. Lymphoepithelioma-like


carcinoma of the urinary bladder. The American journal of
surgical pathology 1994;18:466-73.

198. Braslis KG, Jones A, Murphy D. Clear-cell transitional cell


carcinoma. The Australian and New Zealand journal of
surgery 1997;67:906-8.

181. Holmang S, Borghede G, Johansson SL. Bladder


carcinoma with lymphoepithelioma-like differentiation: a
report of 9 cases. J Urol 1998;159:779-82.

199. Domanowska E, Jozwicki W, Domaniewski J, et al. Muscleinvasive urothelial cell carcinoma of the human bladder:
multidirectional differentiation and ability to metastasize.
Human pathology 2007;38:741-6.

182. Lopez-Beltran A, Luque RJ, Vicioso L, et al.


Lymphoepithelioma-like carcinoma of the urinary bladder:
a clinicopathologic study of 13 cases. Virchows Archiv : an
international journal of pathology 2001;438:552-7.

200. Kotliar SN, Wood, C.G., Schaeffer, A.J. Transitional cell


carcinoma exhibiting clear cell features. A differential
diagnosis for clear cell adenocarcinoma of the urinary
tract. Arch Pathol Lab Med 1995;119:79-81.

183. Tamas EF, Nielsen ME, Schoenberg MP, Epstein JI.


Lymphoepithelioma-like carcinoma of the urinary tract: a
clinicopathological study of 30 pure and mixed cases. Mod
Pathol 2007;20:828-34.

201. Perez-Montiel D, Wakely PE, Hes O, Michal M, Suster


S. High-grade urothelial carcinoma of the renal pelvis:
clinicopathologic study of 108 cases with emphasis on
unusual morphologic variants. Mod Pathol 2006;19:494503.

184. Zukerberg LR, Harris NL, Young RH. Carcinomas of the


urinary bladder simulating malignant lymphoma. A report
of five cases. The American journal of surgical pathology
1991;15:569-76.

202. Rotellini M, Fondi C, Paglierani M, Stomaci N, Raspollini


MR. Clear cell carcinoma of the bladder in a patient with
a earlier clear cell renal cell carcinoma: a case report with
morphologic, immunohistochemical, and cytogenetical
analysis. Applied immunohistochemistry & molecular
morphology : AIMM / official publication of the Society for
Applied Immunohistochemistry 2010;18:396-9.

185. Fadare O, Renshaw IL, Rubin C. Pleomorphic


lymphoepithelioma-like carcinoma of the urinary bladder.
International journal of clinical and experimental pathology
2009;2:194-9.
186. Gulley ML, Amin MB, Nicholls JM, et al. Epstein-Barr virus
is detected in undifferentiated nasopharyngeal carcinoma
but not in lymphoepithelioma-like carcinoma of the urinary
bladder. Human pathology 1995;26:1207-14.

203. Yamashita R, Yamaguchi R, Yuen K, Niwakawa M, Tobisu


K. Urothelial carcinoma (clear cell variant) diagnosed with
useful immunohistochemistry stain. International journal
of urology : official journal of the Japanese Urological
Association 2006;13:1448-50.

187. Izquierdo-Garcia FM, Garcia-Diez F, Fernandez I, et al.


Lymphoepithelioma-like carcinoma of the bladder: three
cases with clinicopathological and p53 protein expression
study. Virchows Archiv : an international journal of
pathology 2004;444:420-5.

204. Gilcrease MZ, Delgado R, Vuitch F, Albores-Saavedra J.


Clear cell adenocarcinoma and nephrogenic adenoma
of the urethra and urinary bladder: a histopathologic and
immunohistochemical comparison. Human pathology
1998;29:1451-6.

188. Lopez-Beltran A, Requena, M.J., Cheng, L., et al. Infiltrating


urothelial carcinoma. Lyon, France: IARC Press; 2004.

205. Oliva E, Amin MB, Jimenez R, Young RH. Clear cell


carcinoma of the urinary bladder: a report and comparison

189. Mostofi FK, Davis, C.J., Sesterhenn, I.A. World Health

151

of four tumors of mullerian origin and nine of probable


urothelial origin with discussion of histogenesis and
diagnostic problems. The American journal of surgical
pathology 2002;26:190-7.

222. Biegel JA, Kalpana G, Knudsen ES, et al. The role of


INI1 and the SWI/SNF complex in the development of
rhabdoid tumors: meeting summary from the workshop
on childhood atypical teratoid/rhabdoid tumors. Cancer
research 2002;62:323-8.

206. Sung MT, Zhang S, MacLennan GT, et al. Histogenesis of


clear cell adenocarcinoma in the urinary tract: evidence
of urothelial origin. Clinical cancer research : an official
journal of the American Association for Cancer Research
2008;14:1947-55.

223. Versteege I, Sevenet N, Lange J, et al. Truncating


mutations of hSNF5/INI1 in aggressive paediatric cancer.
Nature 1998;394:203-6.
224. Hoot AC, Russo P, Judkins AR, Perlman EJ, Biegel
JA. Immunohistochemical analysis of hSNF5/INI1
distinguishes renal and extra-renal malignant rhabdoid
tumors from other pediatric soft tissue tumors. The
American journal of surgical pathology 2004;28:1485-91.

207. Chinegwundoh FI, Khor T, Leedham PW. Bladder


metastasis from renal cell carcinoma. British journal of
urology 1997;79:650-1.
208. Gelister JS, Falzon M, Crawford R, Chapple CR, Hendry
WF. Urinary tract metastasis from renal carcinoma. British
journal of urology 1992;69:250-2.

225. Grignon D, El-Bolkainy, M.N. Infiltrating urothelial


carcinoma: lymphoma-like and plasmacytoid variants.
Lyon, France: IARC Press; 2004.

209. Matsuo M, Koga S, Nishikido M, et al. Renal cell carcinoma


with solitary metachronous metastasis to the urinary
bladder. Urology 2002;60:911-2.

226. Aldousari S, Sircar K, Kassouf W. Plasmacytoid urothelial


carcinoma of the bladder: a case report. Cases journal
2009;2:6647.

210. Shiraishi K, Mohri J, Inoue R, Kamiryo Y. Metastatic


renal cell carcinoma to the bladder 12 years after
radical nephrectomy. International journal of urology :
official journal of the Japanese Urological Association
2003;10:453-5.

227. Fritsche HM, Burger M, Denzinger S, Legal W, Goebell


PJ, Hartmann A. Plasmacytoid urothelial carcinoma of the
bladder: histological and clinical features of 5 cases. J Urol
2008;180:1923-7.
228. Gaafar A, Garmendia M, de Miguel E, et al. [Plasmacytoid
urothelial carcinoma of the urinary bladder. A study of 7
cases]. Actas urologicas espanolas 2008;32:806-10.

211. Sim SJ, Ro JY, Ordonez NG, Park YW, Kee KH, Ayala
AG. Metastatic renal cell carcinoma to the bladder: a
clinicopathologic and immunohistochemical study. Mod
Pathol 1999;12:351-5.

229. Grammatico D, Grignon DJ, Eberwein P, Shepherd RR,


Hearn SA, Walton JC. Transitional cell carcinoma of the
renal pelvis with choriocarcinomatous differentiation.
Immunohistochemical and immunoelectron microscopic
assessment of human chorionic gonadotropin production
by transitional cell carcinoma of the urinary bladder.
Cancer 1993;71:1835-41.

212. Zhou M, Epstein JI, Young RH. Paraganglioma of the


urinary bladder: a lesion that may be misdiagnosed as
urothelial carcinoma in transurethral resection specimens.
The American journal of surgical pathology 2004;28:94100.

230. Kohno T, Kitamura M, Akai H, Takaha M, Kawahara K,


Oka T. Plasmacytoid urothelial carcinoma of the bladder.
International journal of urology : official journal of the
Japanese Urological Association 2006;13:485-6.

213. Higgins JP, Kaygusuz G, Wang L, et al. Placental S100


(S100P) and GATA3: markers for transitional epithelium
and urothelial carcinoma discovered by complementary
DNA microarray. The American journal of surgical
pathology 2007;31:673-80.

231. Mai KT, Park PC, Yazdi HM, et al. Plasmacytoid urothelial
carcinoma of the urinary bladder report of seven new
cases. European urology 2006;50:1111-4.

214. Epstein JI, Amin, M.B., Reuter, V.E. Histologic variants of


urothelial carcinoma. Philadelphia, PA: Lippincott Williams
and Wilkins; 2010.

232. Mitsogiannis IC, Ioannou MG, Sinani CD, Melekos MD.


Plasmacytoid transitional cell carcinoma of the urinary
bladder. Urology 2005;66:194.
233. Nigwekar P, Tamboli P, Amin MB, Osunkoya AO, Ben-Dor
D. Plasmacytoid urothelial carcinoma: detailed analysis of
morphology with clinicopathologic correlation in 17 cases.
The American journal of surgical pathology 2009;33:41724.
234. Patriarca C, Di Pasquale M, Giunta P, Bergamaschi F.
CD138-positive plasmacytoid urothelial carcinoma of
the bladder. International journal of surgical pathology
2008;16:215-7.
235. Ro JY, Shen SS, Lee HI, et al. Plasmacytoid transitional
cell carcinoma of urinary bladder: a clinicopathologic study
of 9 cases. The American journal of surgical pathology
2008;32:752-7.
236. Sahin AA, Myhre M, Ro JY, Sneige N, Dekmezian RH, Ayala
AG. Plasmacytoid transitional cell carcinoma. Report of a
case with initial presentation mimicking multiple myeloma.
Acta cytologica 1991;35:277-80.
237. Sato K, Ueda Y, Kawamura K, Aihara K, Katsuda S.
Plasmacytoid urothelial carcinoma of the urinary bladder:
a case report and immunohistochemical study. Pathology,
research and practice 2009;205:189-94.
238. Shimada K, Nakamura M, Ishida E, Konishi N. Urothelial
carcinoma with plasmacytoid variants producing both
human chorionic gonadotropin and carbohydrate antigen
19-9. Urology 2006;68:891 e7-10.

215. Carter RL, McCarthy KP, al-Sam SZ, Monaghan P, Agrawal


M, McElwain TJ. Malignant rhabdoid tumour of the bladder
with immunohistochemical and ultrastructural evidence
suggesting histiocytic origin. Histopathology 1989;14:17990.
216. Chang JH, Dikranian AH, Johnston WH, Storch SK,
Hurwitz RS. Malignant extrarenal rhabdoid tumor of the
bladder: 9-year survival after chemotherapy and partial
cystectomy. J Urol 2004;171:820-1.
217. Duvdevani M, Nass D, Neumann Y, Leibovitch I, Ramon
J, Mor Y. Pure rhabdoid tumor of the bladder. J Urol
2001;166:2337.
218. Egawa S, Uchida T, Koshiba K, Kagata Y, Iwabuchi K.
Malignant fibrous histiocytoma of the bladder with focal
rhabdoid tumor differentiation. J Urol 1994;151:154-6.
219. Harris M, Eyden BP, Joglekar VM. Rhabdoid tumour
of the bladder: a histological, ultrastructural and
immunohistochemical study. Histopathology 1987;11:108392.
220. Parwani AV, Herawi M, Volmar K, Tsay SH, Epstein JI.
Urothelial carcinoma with rhabdoid features: report of 6
cases. Human pathology 2006;37:168-72.
221. Inagaki T, Nagata M, Kaneko M, Amagai T, Iwakawa M,
Watanabe T. Carcinosarcoma with rhabdoid features of the
urinary bladder in a 2-year-old girl: possible histogenesis
of stem cell origin. Pathology international 2000;50:973-8.

152

239. Zhang XM, Elhosseiny A, Melamed MR. Plasmacytoid


urothelial carcinoma of the bladder. A case report and
the first description of urinary cytology. Acta cytologica
2002;46:412-6.

256. Dirnhofer S, Koessler P, Ensinger C, Feichtinger H,


Madersbacher S, Berger P. Production of trophoblastic
hormones by transitional cell carcinoma of the bladder:
association to tumor stage and grade. Human pathology
1998;29:377-82.

240. Keck B, Stoehr R, Wach S, et al. The plasmacytoid


carcinoma of the bladder-rare variant of aggressive
urothelial carcinoma. International journal of cancer
Journal international du cancer 2010.

257. Iles RK, Chard T. Human chorionic gonadotropin


expression by bladder cancers: biology and clinical
potential. J Urol 1991;145:453-8.

241. Wang J, Wang FW, Lagrange CA, Hemstreet Iii GP,


Kessinger A. Clinical features of sarcomatoid carcinoma
(carcinosarcoma) of the urinary bladder: analysis of 221
cases. Sarcoma 2010;2010.

258. Iles RK, Jenkins BJ, Oliver RT, Blandy JP, Chard T. Beta
human chorionic gonadotrophin in serum and urine. A
marker for metastatic urothelial cancer. British journal of
urology 1989;64:241-4.

242. Wright JL, Black PC, Brown GA, et al. Differences in


survival among patients with sarcomatoid carcinoma,
carcinosarcoma and urothelial carcinoma of the bladder.
J Urol 2007;178:2302-6; discussion 7.

259. Martin JE, Jenkins BJ, Zuk RJ, Oliver RT, Baithun SI. Human
chorionic gonadotrophin expression and histological
findings as predictors of response to radiotherapy in
carcinoma of the bladder. Virchows Archiv A, Pathological
anatomy and histopathology 1989;414:273-7.

243. Holtz F, Fox JE, Abell MR. Carcinosarcoma of the urinary


bladder. Cancer 1972;29:294-304.

260. Sievert K, Weber EA, Herwig R, Schmid H, Roos S,


Eickenberg HU. Pure primary choriocarcinoma of
the urinary bladder with long-term survival. Urology
2000;56:856.

244. Jones EC, Young RH. Myxoid and sclerosing sarcomatoid


transitional cell carcinoma of the urinary bladder: a
clinicopathologic and immunohistochemical study of 25
cases. Mod Pathol 1997;10:908-16.

261. Campo E, Algaba F, Palacin A, Germa R, Sole-Balcells


FJ, Cardesa A. Placental proteins in high-grade urothelial
neoplasms. An immunohistochemical study of human
chorionic gonadotropin, human placental lactogen,
and pregnancy-specific beta-1-glycoprotein. Cancer
1989;63:2497-504.

245. Lopez-Beltran A, Pacelli A, Rothenberg HJ, et al.


Carcinosarcoma and sarcomatoid carcinoma of the
bladder: clinicopathological study of 41 cases. J Urol
1998;159:1497-503.
246. Nimeh T, Kuang W, Levin HS, Klein EA. Sarcomatoid
transitional cell carcinoma of bladder managed with
transurethral resection alone. J Urol 2002;167:641-2.
247. Perret L, Chaubert P, Hessler D, Guillou L. Primary heterologous carcinosarcoma (metaplastic carcinoma) of the
urinary bladder: a clinicopathologic, immunohistochemical, and ultrastructural analysis of eight cases and a review of the literature. Cancer 1998;82:1535-49.
248. Torenbeek R, Blomjous CE, de Bruin PC, Newling
DW, Meijer CJ. Sarcomatoid carcinoma of the urinary
bladder. Clinicopathologic analysis of 18 cases with
immunohistochemical
and
electron
microscopic
findings. The American journal of surgical pathology
1994;18:241-9.
249. Wang X, MacLennan GT, Zhang S, et al. Sarcomatoid
carcinoma of the upper urinary tract: clinical outcome
and molecular characterization. Human pathology
2009;40:211-7.
250. Young RH. Carcinosarcoma of the urinary bladder. Cancer
1987;59:1333-9.
251. Young RH, Wick MR, Mills SE. Sarcomatoid carcinoma
of the urinary bladder. A clinicopathologic analysis of 12
cases and review of the literature. American journal of
clinical pathology 1988;90:653-61.
252. Lott S, Lopez-Beltran A, Maclennan GT, Montironi R,
Cheng L. Soft tissue tumors of the urinary bladder, Part
I: myofibroblastic proliferations, benign neoplasms, and
tumors of uncertain malignant potential. Human pathology
2007;38:807-23.
253. Lott S, Lopez-Beltran A, Montironi R, MacLennan GT,
Cheng L. Soft tissue tumors of the urinary bladder Part II:
malignant neoplasms. Human pathology 2007;38:963-77.
254. Hodges KB, Lopez-Beltran A, Emerson RE, Montironi
R, Cheng L. Clinical utility of immunohistochemistry
in the diagnoses of urinary bladder neoplasia. Applied
immunohistochemistry & molecular morphology :
AIMM / official publication of the Society for Applied
Immunohistochemistry 2010;18:401-10.
255. Westfall DE, Folpe AL, Paner GP, et al. Utility of a
comprehensive immunohistochemical panel in the
differential diagnosis of spindle cell lesions of the urinary
bladder. The American journal of surgical pathology
2009;33:99-105.

262. Iles RK, Purkis PE, Whitehead PC, Oliver RT, Leigh I, Chard
T. Expression of beta human chorionic gonadotrophin by
non-trophoblastic non-endocrine normal and malignant
epithelial cells. British journal of cancer 1990;61:663-6.
263. Shah VM, Newman J, Crocker J, et al. Ectopic beta-human
chorionic gonadotropin production by bladder urothelial
neoplasia. Arch Pathol Lab Med 1986;110:107-11.
264. Zettl A, Konrad MA, Polzin S, et al. Urothelial carcinoma
of the renal pelvis with choriocarcinomatous features:
genetic evidence of clonal evolution. Human pathology
2002;33:1234-7.
265. Hanna NH, Ulbright TM, Einhorn LH. Primary
choriocarcinoma of the bladder with the detection of
isochromosome 12p. J Urol 2002;167:1781.
266. Amir G, Rosenmann E. Osteoclast-like giant cell tumour of
the urinary bladder. Histopathology 1990;17:413-8.
267. Baydar D, Amin MB, Epstein JI. Osteoclast-rich
undifferentiated carcinomas of the urinary tract. Mod
Pathol 2006;19:161-71.
268. Garcia Garcia F, Garcia Ligero J, Martinez Diaz F, et al.
[Bladder carcinoma with osteoclast-type giant cells. A
case with a rare presentation. Review of the literature].
Actas urologicas espanolas 2003;27:317-20.
269. Kruger S, Johannisson R, Kausch I, Feller AC. Papillary
urothelial bladder carcinoma associated with osteoclastlike giant cells. International urology and nephrology
2005;37:61-4.
270. OConnor RC, Hollowell CM, Laven BA, Yang XJ,
Steinberg GD, Zagaja GP. Recurrent giant cell carcinoma
of the bladder. J Urol 2002;167:1784.
271. Zukerberg LR, Armin AR, Pisharodi L, Young RH.
Transitional cell carcinoma of the urinary bladder with
osteoclast-type giant cells: a report of two cases and
review of the literature. Histopathology 1990;17:407-11.
272. Churg A. The fine structure of large cell undifferentiated
carcinoma of the lung. Evidence for its relation to
squamous cell carcinomas and adenocarcinomas. Human
pathology 1978;9:143-56.

153

273. Lopez-Beltran A, Cheng L, Comperat E, et al. Large


cell undifferentiated carcinoma of the urinary bladder.
Pathology 2010;42:364-8.

293. Walther M, OBrien DP, 3rd, Birch HW. Condylomata


acuminata and verrucous carcinoma of the bladder: case
report and literature review. J Urol 1986;135:362-5.

274. Pelosi G, Fraggetta F, Nappi O, et al. Pleomorphic


carcinomas of the lung show a selective distribution of gene
products involved in cell differentiation, cell cycle control,
tumor growth, and tumor cell motility: a clinicopathologic
and immunohistochemical study of 31 cases. The
American journal of surgical pathology 2003;27:1203-15.

294. Wyatt JK, Craig I. Verrucous carcinoma of urinary bladder.


Urology 1980;16:97-9.
295. Khafagy MM, el-Bolkainy MN, Mansour MA. Carcinoma
of the bilharzial urinary bladder. A study of the associated
mucosal lesions in 86 cases. Cancer 1972;30:150-9.
296. Demian SD, Bushkin FL, Echevarria RA. Perineural
invasion and anaplastic transformation of verrucous
carcinoma. Cancer 1973;32:395-401.

275. Eble JN, Sauter, G., Epstein, J.I. Pathology & Genetics of
Tumors of the Urinary System and Male Genital Organs.
In. Lyon, France: IARC Press; 2004.

297. Kraus FT, Perezmesa C. Verrucous carcinoma. Clinical


and pathologic study of 105 cases involving oral cavity,
larynx and genitalia. Cancer 1966;19:26-38.

276. Mostafa MH, Sheweita SA, OConnor PJ. Relationship


between schistosomiasis and bladder cancer. Clinical
microbiology reviews 1999;12:97-111.

298. Perez CA, Kraus FT, Evans JC, Powers WE. Anaplastic
transformation in verrucous carcinoma of the oral cavity
after radiation therapy. Radiology 1966;86:108-15.

277. Shokeir AA. Squamous cell carcinoma of the bladder:


pathology, diagnosis and treatment. BJU international
2004;93:216-20.
278. Del Mistro A, Koss LG, Braunstein J, Bennett B,
Saccomano G, Simons KM. Condylomata acuminata
of the urinary bladder. Natural history, viral typing, and
DNA content. The American journal of surgical pathology
1988;12:205-15.

299. Godwin JT, Hanash K. Pathology of bilharzial bladder


cancer. Progress in clinical and biological research
1984;162A:95-143.

279. Lagwinski N, Thomas A, Stephenson AJ, et al. Squamous


cell carcinoma of the bladder: a clinicopathologic analysis
of 45 cases. The American journal of surgical pathology
2007;31:1777-87.

301. Zahran MM, Kamel M, Mooro H, Issa A. Bilharziasis of


urinary bladder and ureter: comparative histopathologic
study. Urology 1976;8:73-9.

300. Tannenbaum M. Inflammatory proliferative lesion of urinary


bladder: squamous metaplasia. Urology 1976;7:428-9.

302. Jacobo E, Loening S, Schmidt JD, Culp DA. Primary


adenocarcinoma of the bladder: a retrospective study of
20 patients. J Urol 1977;117:54-6.

280. Mvula M, Iwasaka T, Iguchi A, Nakamura S, Masaki


Z, Sugimori H. Do human papillomaviruses have a
role in the pathogenesis of bladder carcinoma? J Urol
1996;155:471-4.

303. Lughezzani G, Sun M, Jeldres C, et al. Adenocarcinoma


versus urothelial carcinoma of the urinary bladder:
comparison between pathologic stage at radical cystectomy
and cancer-specific mortality. Urology 2010;75:376-81.

281. Porter MP, Voigt LF, Penson DF, Weiss NS. Racial
variation in the incidence of squamous cell carcinoma of
the bladder in the United States. J Urol 2002;168:1960-3.
282. Swanson DA, Liles A, Zagars GK. Preoperative irradiation
and radical cystectomy for stages T2 and T3 squamous
cell carcinoma of the bladder. J Urol 1990;143:37-40.
283. Newman DM, Brown JR, Jay AC, Pontius EE. Squamous
cell carcinoma of the bladder. J Urol 1968;100:470-3.
284. el Bolkainy MN, Ghoneim, N.A., Mansour, M.A. Carcinoma
of the bilharzial bladder in Egypt. Br J Urolol 1972;44:56170.
285. El-Bolkainy MN, Mokhtar NM, Ghoneim MA, Hussein MH.
The impact of schistosomiasis on the pathology of bladder
carcinoma. Cancer 1981;48:2643-8.
286. el-Sebai I, Sherif M, el-Bolkainy MN, Mansour MA,
Ghoneim MA. Verrucose squamous carcinoma of bladder.
Urology 1974;4:407-10.
287. Mahran MR, el-Baz M. Verrucous carcinoma of the bilharzial
bladder. Impact of invasiveness on survival. Scandinavian
journal of urology and nephrology 1993;27:189-92.
288. Batta AG, Engen DE, Reiman HM, Winkelmann RK.
Intravesical condyloma acuminatum with progression to
verrucous carcinoma. Urology 1990;36:457-64.
289. Ellsworth PI, Schned AR, Heaney JA, Snyder PM.
Surgical treatment of verrucous carcinoma of the bladder
unassociated with bilharzial cystitis: case report and
literature review. J Urol 1995;153:411-4.
290. Holck S, Jorgensen L. Verrucous carcinoma of urinary
bladder. Urology 1983;22:435-7.
291. Horner SA, Fisher HA, Barada JH, Eastman AY, Migliozzi
J, Ross JS. Verrucous carcinoma of the bladder. J Urol
1991;145:1261-3.
292. Pierangeli T, Grifoni R, Marchi P, Montironi R, Stefano S.
Verrucous carcinoma in situ of the bladder, not associated
with urinary schistosomiasis. International urology and
nephrology 1989;21:597-602.

304. Malek RS, Rosen JS, ODea MJ. Adenocarcinoma of


bladder. Urology 1983;21:357-9.
305. Nocks BN, Heney NM, Daly JJ. Primary adenocarcinoma
of urinary bladder. Urology 1983;21:26-9.
306. Wilson TG, Pritchett TR, Lieskovsky G, Warner NE,
Skinner DG. Primary adenocarcinoma of bladder. Urology
1991;38:223-6.
307. Anderstrom C, Johansson SL, von Schultz L. Primary
adenocarcinoma of the urinary bladder. A clinicopathologic
and prognostic study. Cancer 1983;52:1273-80.
308. el-Mekresh MM, el-Baz MA, Abol-Enein H, Ghoneim MA.
Primary adenocarcinoma of the urinary bladder: a report
of 185 cases. British journal of urology 1998;82:206-12.
309. Ghoneim MA, Abdel-Latif M, el-Mekresh M, et al.
Radical cystectomy for carcinoma of the bladder: 2,720
consecutive cases 5 years later. J Urol 2008;180:121-7.
310. Zoghloul MS, Nouh, A., Nazmy, M., et al. Long-term
results of primary adenocarcinoma of the urinary bladder.
Urologic oncology 2006;24:13-20.
311. Paulhac P, Maisonnette F, Bourg S, Dumas JP, Colombeau
P. Adenocarcinoma in the exstrophic bladder. Urology
1999;54:744.
312. Kamat MR, Kulkarni JN, Tongaonkar HB. Adenocarcinoma
of the bladder: study of 14 cases and review of the
literature. British journal of urology 1991;68:254-7.
313. Buza N, Cohen PJ, Pei H, Parkash V. Inverse p16 and p63
expression in small cell carcinoma and high-grade urothelial
cell carcinoma of the urinary bladder. International journal
of surgical pathology 2010;18:94-102.
314. Goyanna R, Emmett JL, Mc DJ. Exstrophy of the bladder
complicated by adenocarcinoma. J Urol 1951;65:391400.

154

315. Jones WA, Gibbons RP, Correa RJ, Jr., Cummings KB,
Mason JT. Primary adenocarcinoma of bladder. Urology
1980;15:119-22.

335. Ploeg M, Aben KK, Hulsbergen-van de Kaa CA,


Schoenberg MP, Witjes JA, Kiemeney LA. Clinical
epidemiology of nonurothelial bladder cancer: analysis of
the Netherlands Cancer Registry. J Urol 2010;183:91520.

316. Abeshouse BS. Exstrophy of the bladder complicated by


adenocarcinoma of the bladder and renal calculi. A report
of a case and review of the literature. J Urol 1943;49:25989.

336. Busto Martin LA, Janeiro Pais, P.M., Gonzalez Dacal, J.,
Chantada Abal, V., Busto Castanon, L. Signet-ring cell
adenocarcinoma of the bladder: case series between
1990-2009. Arch Espan Urol 2010;63:150-3.

317. Culp DA. The Histology of the Exstrophied Bladder. J Urol


1964;91:538-48.

337. Zaghloul MS, El Baradie M, Nouh, Abdel-Fatah S, Taher A,


Shalaan M. Prognostic index for primary adenocarcinoma
of the urinary bladder. The gulf journal of oncology
2007:47-54.

318. Bitar M, Mandel E, Kirschenbaum AM, Unger PD. Urinary


bladder adenocarcinoma arising in a spina bifida patient.
Annals of diagnostic pathology 2007;11:453-6.
319. Tamas EF, Stephenson AJ, Campbell SC, Montague
DK, Trusty DC, Hansel DE. Histopathologic features and
clinical outcomes in 71 cases of bladder diverticula. Arch
Pathol Lab Med 2009;133:791-6.

338. Eble JN. Abnormalities of the Urachus. New York, NY:


Churchill-Livingstone; 1989.
339. Schubert GE, Pavkovic MB, Bethke-Bedurftig BA. Tubular
urachal remnants in adult bladders. J Urol 1982;127:40-2.

320. Choi H, Lamb S, Pintar K, Jacobs SC. Primary signetring cell carcinoma of the urinary bladder. Cancer
1984;53:1985-90.

340. Gopalan A, Sharp DS, Fine SW, et al. Urachal carcinoma:


a clinicopathologic analysis of 24 cases with outcome
correlation. The American journal of surgical pathology
2009;33:659-68.

321. Wattenberg CA BJ, Tornmey AR. Exstrophy of the urinary


bladder complicated by adenocarcinoma. J Urol 1956.

341. Herr HW, Bochner BH, Sharp D, Dalbagni G, Reuter VE.


Urachal carcinoma: contemporary surgical outcomes. J
Urol 2007;178:74-8; discussion 8.

322. Kramer SA, Bredael J, Croker BP, Paulson DF, Glenn JF.
Primary non-urachal adenocarcinoma of the bladder. J
Urol 1979;121:278-81.
323. Mostofi FK, Thomson RV, Dean AL, Jr. Mucous
adenocarcinoma of the urinary bladder. Cancer
1955;8:741-58.
324. Braun EV, Ali M, Fayemi AO, Beaugard E. Primary signetring cell carcinoma of the urinary bladder: review of the
literature and report of a case. Cancer 1981;47:1430-5.
325. Grignon DJ, Ro JY, Ayala AG, Johnson DE, Ordonez
NG. Primary adenocarcinoma of the urinary bladder.
A clinicopathologic analysis of 72 cases. Cancer
1991;67:2165-72.
326. Blute ML, Engen DE, Travis WD, Kvols LK. Primary
signet ring cell adenocarcinoma of the bladder. J Urol
1989;141:17-21.
327. Burgues O, Ferrer J, Navarro S, Ramos D, Botella E,
Llombart-Bosch A. Hepatoid adenocarcinoma of the
urinary bladder. An unusual neoplasm. Virchows Archiv :
an international journal of pathology 1999;435:71-5.
328. Abenoza P, Manivel C, Fraley EE. Primary adenocarcinoma
of urinary bladder. Clinicopathologic study of 16 cases.
Urology 1987;29:9-14.
329. Lane Z, Hansel DE, Epstein JI. Immunohistochemical
expression of prostatic antigens in adenocarcinoma and
villous adenoma of the urinary bladder. The American
journal of surgical pathology 2008;32:1322-6.
330. Torenbeek R, Lagendijk JH, Van Diest PJ, Bril H, van de
Molengraft FJ, Meijer CJ. Value of a panel of antibodies to
identify the primary origin of adenocarcinomas presenting
as bladder carcinoma. Histopathology 1998;32:20-7.
331. Wang HL, Lu DW, Yerian LM, et al. Immunohistochemical
distinction between primary adenocarcinoma of the
bladder and secondary colorectal adenocarcinoma. The
American journal of surgical pathology 2001;25:1380-7.

342. Ohman U vGB, Moberg A. Carcinoma of the urachus,


review of literature and report of two cases. Scandinavian
journal of urology and nephrology 1971.
343. Cornil C, Reynolds CT, Kickham CJ. Carcinoma of the
urachus. J Urol 1967;98:93-5.
344. Ghazizadeh M, Yamamoto S, Kurokawa K. Clinical
features of urachal carcinoma in Japan: review of 157
patients. Urol Res 1983;11:235-8.
345. Sheldon CA, Clayman RV, Gonzalez R, Williams RD, Fraley
EE. Malignant urachal lesions. J Urol 1984;131:1-8.
346. Pinthus JH, Haddad R, Trachtenberg J, et al. Population
based survival data on urachal tumors. J Urol
2006;175:2042-7; discussion 7.
347. Ashley RA, Inman BA, Sebo TJ, et al. Urachal carcinoma:
clinicopathologic features and long-term outcomes of an
aggressive malignancy. Cancer 2006;107:712-20.
348. Siefker-Radtke AO, Gee J, Shen Y, et al. Multimodality
management of urachal carcinoma: the M. D. Anderson
Cancer Center experience. J Urol 2003;169:1295-8.
349. Thali-Schwab CM, Woodward PJ, Wagner BJ. Computed
tomographic appearance of urachal adenocarcinomas:
review of 25 cases. European radiology 2005;15:79-84.
350. Begg RC. The Urachus: its Anatomy, Histology and
Development. Journal of anatomy 1930;64:170-83.
351. Chen ZF, Wang F, Qin ZK, et al. [Clinical analysis of
14 cases of urachal carcinoma]. Ai zheng = Aizheng =
Chinese journal of cancer 2008;27:966-9.
352. Dyer RB, Chen MY, Zagoria RJ. Abnormal calcifications in
the urinary tract. Radiographics : a review publication of the
Radiological Society of North America, Inc 1998;18:140524.
353. Wheeler JD, Hill WT. Adenocarcinoma involving the urinary
bladder. Cancer 1954;7:119-35.
354. Kuramoto T, Kikkawa K, Nishihata M, et al. Squamous cell
carcinoma of the urachus producing granulocyte colonystimulating factor. Urology 2009;73:442 e5-7.
355. Rubin JP, Kasznica JM, Davis CA, 3rd, Carpinito GA,
Hirsch EF. Transitional cell carcinoma in a urachal cyst. J
Urol 1999;162:1687-8.
356. Abenoza P, Manivel C, Sibley RK. Adenocarcinoma with
neuroendocrine differentiation of the urinary bladder.
Clinicopathologic, immunohistochemical, and ultrastructural
study. Arch Pathol Lab Med 1986;110:1062-6.

332. Lu QL, Laniado M, Abel PD, Stamp GW, Lalani EN.


Expression of bcl-2 in bladder neoplasms is a cell lineage
associated and p53-independent event. Molecular
pathology : MP 1997;50:28-33.
333. Chuang AY, DeMarzo AM, Veltri RW, Sharma RB, Bieberich
CJ, Epstein JI. Immunohistochemical differentiation of
high-grade prostate carcinoma from urothelial carcinoma.
The American journal of surgical pathology 2007;31:124655.
334. Hughes MJ, Fisher C, Sohaib SA. Imaging features of
primary nonurachal adenocarcinoma of the bladder. AJR
American journal of roentgenology 2004;183:1397-401.

155

357. Satake T, Takeda A, Matsuyama M. Argyrophil cells in the


urachal epithelium and urachal adenocarcinoma. Acta
pathologica japonica 1984;34:1193-9.

378. Vakar-Lopez F, True LD. How wide is the spectrum


of neuroendocrine carcinoma of the urinary bladder?
American journal of clinical pathology 2007;128:723-5.

358. Baglio CM, Crowson CN. Hemangiopericytoma of


Urachus: Report of a Case. J Urol 1964;91:660-2.

379. Alijo Serrano F, Sanchez-Mora N, Angel Arranz J,


Hernandez C, Alvarez-Fernandez E. Large cell and small cell
neuroendocrine bladder carcinoma: immunohistochemical
and outcome study in a single institution. American journal
of clinical pathology 2007;128:733-9.

359. Hama Y, Okizuka H, Kusano S. Pleomorphic sarcoma of


the adult urinary bladder: sonographic findings. Journal of
clinical ultrasound : JCU 2004;32:215-7.

380. Blomjous CE, Vos W, De Voogt HJ, Van der Valk P,


Meijer CJ. Small cell carcinoma of the urinary bladder. A
clinicopathologic, morphometric, immunohistochemical,
and ultrastructural study of 18 cases. Cancer 1989;64:134757.

360. Noyes D, Vinson RK. Urachal leiomyosarcoma. Urology


1981;17:279-80.
361. Pantuck AJ, Bancila E, Das KM, et al. Adenocarcinoma
of the urachus and bladder expresses a unique colonic
epithelial epitope: an immunohistochemical study. J Urol
1997;158:1722-7.

381. Holmang S, Borghede G, Johansson SL. Primary small


cell carcinoma of the bladder: a report of 25 cases. J Urol
1995;153:1820-2.

362. Kajita Y, Habuchi T, Kamoto T, et al. [Long-term clinical


results of 5 cases of urachal carcinoma]. Hinyokika kiyo
Acta urologica Japonica 2000;46:711-4.

382. Quek ML, Nichols PW, Yamzon J, et al. Radical


cystectomy for primary neuroendocrine tumors of the
bladder: the university of southern california experience.
J Urol 2005;174:93-6.

363. Whitehead ED, Tessler AN. Carcinoma of the urachus.


British journal of urology 1971;43:468-76.
364. Dandekar NP, Dalal AV, Tongaonkar HB, Kamat
MR. Adenocarcinoma of bladder. Eur J Surg Oncol
1997;23:157-60.

383. Dundr P, Pesl M, Povysil C, Vitkova I, Dvoracek J. Large


cell neuroendocrine carcinoma of the urinary bladder with
lymphoepithelioma-like features. Pathology, research and
practice 2003;199:559-63.

365. Henly DR, Farrow GM, Zincke H. Urachal cancer: role of


conservative surgery. Urology 1993;42:635-9.

384. Evans AJ, Al-Maghrabi J, Tsihlias J, Lajoie G, Sweet


JM, Chapman WB. Primary large cell neuroendocrine
carcinoma of the urinary bladder. Arch Pathol Lab Med
2002;126:1229-32.

366. Shou J, Ma J, Xu B. [Adenocarcinoma of the urinary


bladder: a report of 27 cases]. Zhonghua zhong liu za zhi
[Chinese journal of oncology] 1999;21:461-3.
367. Wright JL, Porter MP, Li CI, Lange PH, Lin DW. Differences
in survival among patients with urachal and nonurachal
adenocarcinomas of the bladder. Cancer 2006;107:721-8.
368. Chan ES, Ng CF, Chui KL, Lo KL, Hou SM, Yip SK. Novel
approach of laparoscopic transperitoneal en bloc resection
of urachal tumor and umbilectomy with a comparison
of various techniques. Journal of laparoendoscopic &
advanced surgical techniques Part A 2009;19:423-6.
369. Milhoua PM, Knoll A, Bleustein CB, Ghavamian R.
Laparoscopic partial cystectomy for treatment of
adenocarcinoma of the urachus. Urology 2006;67:423
e15- e17.
370. Rabah DM. Robot-assisted partial cystectomy for the
treatment of urachal carcinoma. The Canadian journal of
urology 2007;14:3640-2.
371. Gill HS, Dhillon HK, Woodhouse CR. Adenocarcinoma of
the urinary bladder. British journal of urology 1989;64:13842.
372. Xiaoxu L, Jianhong L, Jinfeng W, Klotz LH. Bladder
adenocarcinoma: 31 reported cases. The Canadian
journal of urology 2001;8:1380-3.
373. Kaido T, Uemura H, Hirao Y, Uranishi R, Nishi N, Sakaki
T. Brain metastases from urachal carcinoma. Journal of
clinical neuroscience : official journal of the Neurosurgical
Society of Australasia 2003;10:703-5.
374. McClelland S, 3rd, Garcia RE, Monaco SE, et al.
Carcinomatous meningitis from urachal carcinoma: the first
reported case. Journal of neuro-oncology 2006;76:171-4.
375. Kojima Y, Yamada Y, Kamisawa H, Sasaki S, Hayashi
Y, Kohri K. Complete response of a recurrent advanced
urachal carcinoma treated by S-1/cisplatin combination
chemotherapy. International journal of urology :
official journal of the Japanese Urological Association
2006;13:1123-5.
376. Siefker-Radtke A. Urachal carcinoma: surgical and
chemotherapeutic options. Expert review of anticancer
therapy 2006;6:1715-21.
377. Chor PJ, Gaum LD, Young RH. Clear cell adenocarcinoma
of the urinary bladder: report of a case of probable
mullerian origin. Mod Pathol 1993;6:225-8.

385. Lee KH, Ryu SB, Lee MC, Park CS, Juhng SW, Choi C.
Primary large cell neuroendocrine carcinoma of the urinary
bladder. Pathology international 2006;56:688-93.
386. Cheng L, Pan CX, Yang XJ, et al. Small cell carcinoma
of the urinary bladder: a clinicopathologic analysis of 64
patients. Cancer 2004;101:957-62.
387. Wang X, MacLennan GT, Lopez-Beltran A, Cheng L.
Small cell carcinoma of the urinary bladder--histogenesis,
genetics, diagnosis, biomarkers, treatment, and prognosis.
Applied immunohistochemistry & molecular morphology
: AIMM / official publication of the Society for Applied
Immunohistochemistry 2007;15:8-18.
388. Agoff SN, Lamps LW, Philip AT, et al. Thyroid transcription
factor-1 is expressed in extrapulmonary small cell
carcinomas but not in other extrapulmonary neuroendocrine
tumors. Mod Pathol 2000;13:238-42.
389. Bex A, Nieuwenhuijzen JA, Kerst M, et al. Small cell
carcinoma of bladder: a single-center prospective study
of 25 cases treated in analogy to small cell lung cancer.
Urology 2005;65:295-9.
390. Chen YB, Epstein JI. Primary carcinoid tumors of the
urinary bladder and prostatic urethra: a clinicopathologic
study of 6 cases. The American journal of surgical
pathology 2011;35:442-6.
391. Murali R, Kneale K, Lalak N, Delprado W. Carcinoid
tumors of the urinary tract and prostate. Arch Pathol Lab
Med 2006;130:1693-706.
392. Sugihara A, Kajio K, Yoshimoto T, et al. Primary carcinoid
tumor of the urinary bladder. International urology and
nephrology 2002;33:53-7.
393. Cheng L, Montironi R, Davidson DD, Lopez-Beltran A.
Staging and reporting of urothelial carcinoma of the urinary
bladder. Mod Pathol 2009;22 Suppl 2:S70-95.
394. Abel PD, Hall RR, Williams G. Should pT1 transitional cell
cancers of the bladder still be classified as superficial?
British journal of urology 1988;62:235-9.
395. Van Der Meijden A, Sylvester R, Collette L, Bono A,
Ten Kate F. The role and impact of pathology review on

156

stage and grade assessment of stages Ta and T1 bladder


tumors: a combined analysis of 5 European Organization
for Research and Treatment of Cancer Trials. J Urol
2000;164:1533-7.

410. Paner GP, Brown JG, Lapetino S, et al. Diagnostic use


of antibody to smoothelin in the recognition of muscularis
propria in transurethral resection of urinary bladder tumor
(TURBT) specimens. The American journal of surgical
pathology 2010;34:792-9.

396. Schaafsma HE, Ramaekers FC, van Muijen GN, et


al. Distribution of cytokeratin polypeptides in human
transitional cell carcinomas, with special emphasis on
changing expression patterns during tumor progression.
Am J Pathol 1990;136:329-43.

411. Council L, Hameed O. Differential expression of


immunohistochemical markers in bladder smooth muscle
and myofibroblasts, and the potential utility of desmin,
smoothelin, and vimentin in staging of bladder carcinoma.
Mod Pathol 2009;22:639-50.

397. Tamas EF, Epstein JI. Detection of residual tumor cells


in bladder biopsy specimens: pitfalls in the interpretation
of cytokeratin stains. The American journal of surgical
pathology 2007;31:390-7.

412. Bovio IM, Al-Quran SZ, Rosser CJ, Algood CB, Drew PA,
Allan RW. Smoothelin immunohistochemistry is a useful
adjunct for assessing muscularis propria invasion in
bladder carcinoma. Histopathology 2010;56:951-6.

398. Bryan RT, Atherfold PA, Yeo Y, et al. Cadherin switching


dictates the biology of transitional cell carcinoma of the
bladder: ex vivo and in vitro studies. J Pathol 2008;215:18494.

413. Miyamoto H, Sharma RB, Illei PB, Epstein JI. Pitfalls in the
use of smoothelin to identify muscularis propria invasion
by urothelial carcinoma. The American journal of surgical
pathology 2010;34:418-22.

399. Bringuier PP, Giroldi LA, Umbas R, Shimazui T, Schalken


JA. Mechanisms associated with abnormal E-cadherin
immunoreactivity in human bladder tumors. International
journal of cancer Journal international du cancer
1999;83:591-5.

414. Mitra AP, Cote RJ. Molecular pathogenesis and diagnostics


of bladder cancer. Annu Rev Pathol 2009;4:251-85.
415. Mitra AP, Cote RJ. Molecular screening for bladder cancer:
progress and potential. Nat Rev Urol 2010;7:11-20.

400. Giroldi LA, Bringuier PP, Shimazui T, Jansen K, Schalken


JA. Changes in cadherin-catenin complexes in the
progression of human bladder carcinoma. International
journal of cancer Journal international du cancer
1999;82:70-6.

416. Mitra AP, Datar RH, Cote RJ. Molecular pathways in


invasive bladder cancer: new insights into mechanisms,
progression, and target identification. J Clin Oncol
2006;24:5552-64.
417. Wu XR. Urothelial tumorigenesis: a tale of divergent
pathways. Nat Rev Cancer 2005;5:713-25.

401. Bostrom PJ, Ravanti L, Reunanen N, et al. Expression of


collagenase-3 (matrix metalloproteinase-13) in transitionalcell carcinoma of the urinary bladder. International journal
of cancer Journal international du cancer 2000;88:41723.

418. Oxford G, Theodorescu D. The role of Ras superfamily


proteins in bladder cancer progression. J Urol
2003;170:1987-93.
419. Kompier LC, Lurkin I, van der Aa MN, van Rhijn BW, van
der Kwast TH, Zwarthoff EC. FGFR3, HRAS, KRAS,
NRAS and PIK3CA mutations in bladder cancer and their
potential as biomarkers for surveillance and therapy. PLoS
One 2010;5:e13821.

402. Mueller J, Steiner C, Hofler H. Stromelysin-3 expression


in noninvasive and invasive neoplasms of the urinary
bladder. Human pathology 2000;31:860-5.
403. Wallard MJ, Pennington CJ, Veerakumarasivam A, et al.
Comprehensive profiling and localisation of the matrix
metalloproteinases in urothelial carcinoma. British journal
of cancer 2006;94:569-77.

420. Lopez-Knowles E, Hernandez S, Malats N, et al. PIK3CA


mutations are an early genetic alteration associated with
FGFR3 mutations in superficial papillary bladder tumors.
Cancer research 2006;66:7401-4.

404. Hindermann W, Berndt A, Haas KM, Wunderlich H,


Katenkamp D, Kosmehl H. Immunohistochemical
demonstration of the gamma2 chain of laminin-5 in urinary
bladder urothelial carcinoma. Impact for diagnosis and
prognosis. Cancer Detect Prev 2003;27:109-15.

421. Kubota Y, Miyamoto H, Noguchi S, et al. The loss of


retinoblastoma gene in association with c-myc and
transforming growth factor-beta 1 gene expression in
human bladder cancer. J Urol 1995;154:371-4.

405. Kiyoshima K, Oda Y, Kinukawa N, Naito S, Tsuneyoshi


M. Overexpression of laminin-5 gamma2 chain and its
prognostic significance in urothelial carcinoma of urinary
bladder: association with expression of cyclooxygenase 2,
epidermal growth factor receptor [corrected] and human
epidermal growth factor receptor [corrected] 2. Human
pathology 2005;36:522-30.

422. ODonnell MA. Advances in the management of superficial


bladder cancer. Semin Oncol 2007;34:85-97.
423. Burger M, van der Aa MN, van Oers JM, et al. Prediction
of progression of non-muscle-invasive bladder cancer
by WHO 1973 and 2004 grading and by FGFR3
mutation status: a prospective study. European urology
2008;54:835-43.

406. Mhawech P, Iselin C, Pelte MF. Value of


immunohistochemistry in staging T1 urothelial bladder
carcinoma. European urology 2002;42:459-63.

424. Shariat SF, Ashfaq R, Sagalowsky AI, Lotan Y. Predictive


value of cell cycle biomarkers in nonmuscle invasive
bladder transitional cell carcinoma. J Urol 2007;177:481-7;
discussion 7.

407. van der Loop FT, Schaart G, Timmer ED, Ramaekers FC,
van Eys GJ. Smoothelin, a novel cytoskeletal protein specific for smooth muscle cells. J Cell Biol 1996;134:401-11.

425. Shariat SF, Chade DC, Karakiewicz PI, et al. Combination


of multiple molecular markers can improve prognostication
in patients with locally advanced and lymph node positive
bladder cancer. J Urol 2010;183:68-75.

408. Paner GP, Shen SS, Lapetino S, et al. Diagnostic utility


of antibody to smoothelin in the distinction of muscularis
propria from muscularis mucosae of the urinary bladder: a
potential ancillary tool in the pathologic staging of invasive
urothelial carcinoma. The American journal of surgical
pathology 2009;33:91-8.

426. Shariat SF, Karakiewicz PI, Palapattu GS, et al. Outcomes


of radical cystectomy for transitional cell carcinoma of
the bladder: a contemporary series from the Bladder
Cancer Research Consortium. J Urol 2006;176:2414-22;
discussion 22.

409. Faragalla HF, Marcon NE, Yousef GM, Streutker CJ.


Immunohistochemical staining for smoothelin in the
duplicated versus the true muscularis mucosae of Barrett
esophagus. The American journal of surgical pathology
2011;35:55-9.

427. Stein JP, Lieskovsky G, Cote R, et al. Radical cystectomy


in the treatment of invasive bladder cancer: long-term
results in 1,054 patients. J Clin Oncol 2001;19:666-75.

157

428. Rabbani F, Koppie TM, Charytonowicz E, Drobnjak M,


Bochner BH, Cordon-Cardo C. Prognostic significance of
p27Kip1 expression in bladder cancer. BJU international
2007;100:259-63.

aggressiveness of bladder urothelial carcinoma. Clinical


cancer research : an official journal of the American
Association for Cancer Research 2010;16:2624-33.
445. Bensalah K, Montorsi F, Shariat SF. Challenges of cancer
biomarker profiling. European urology 2007;52:1601-9.

429. Rotterud R, Nesland JM, Berner A, Fossa SD. Expression


of the epidermal growth factor receptor family in normal
and malignant urothelium. BJU international 2005;95:134450.

446. Bolenz C, Lotan Y. Translational research in bladder


cancer: from molecular pathogenesis to useful tissue
biomarkers. Cancer Biol Ther 2010;10:407-15.

430. Sanchez-Carbayo M, Cordon-Cardo C. Applications of


array technology: identification of molecular targets in
bladder cancer. British journal of cancer 2003;89:2172-7.

447. Kawauchi S, Sakai H, Ikemoto K, et al. 9p21 index as


estimated by dual-color fluorescence in situ hybridization is
useful to predict urothelial carcinoma recurrence in bladder
washing cytology. Human pathology 2009;40:1783-9.

431. Sanchez-Carbayo M, Cordon-Cardo C. Molecular


alterations associated with bladder cancer progression.
Semin Oncol 2007;34:75-84.

448. Kruger S, Mess F, Bohle A, Feller AC. Numerical


aberrations of chromosome 17 and the 9p21 locus are
independent predictors of tumor recurrence in noninvasive transitional cell carcinoma of the urinary bladder.
Int J Oncol 2003;23:41-8.

432. Sanchez-Carbayo M, Socci ND, Charytonowicz E, et


al. Molecular profiling of bladder cancer using cDNA
microarrays: defining histogenesis and biological
phenotypes. Cancer research 2002;62:6973-80.

449. Sarosdy MF, Kahn PR, Ziffer MD, et al. Use of a multitarget
fluorescence in situ hybridization assay to diagnose bladder
cancer in patients with hematuria. J Urol 2006;176:44-7.
450. Moonen PM, Merkx GF, Peelen P, Karthaus HF, Smeets
DF, Witjes JA. UroVysion compared with cytology and
quantitative cytology in the surveillance of non-muscleinvasive bladder cancer. European urology 2007;51:127580; discussion 80.
451. Yoder BJ, Skacel M, Hedgepeth R, et al. Reflex UroVysion
testing of bladder cancer surveillance patients with
equivocal or negative urine cytology: a prospective
study with focus on the natural history of anticipatory
positive findings. American journal of clinical pathology
2007;127:295-301.
452. Ferra S, Denley R, Herr H, Dalbagni G, Jhanwar S, Lin
O. Reflex UroVysion testing in suspicious urine cytology
cases. Cancer 2009;117:7-14.

433. Sanchez-Carbayo M, Socci ND, Lozano J, Saint F,


Cordon-Cardo C. Defining molecular profiles of poor
outcome in patients with invasive bladder cancer using
oligonucleotide microarrays. J Clin Oncol 2006;24:77889.
434. Beekman KW, Bradley D, Hussain M. New molecular
targets and novel agents in the treatment of advanced
urothelial cancer. Semin Oncol 2007;34:154-64.
435. Chatterjee SJ, Datar R, Youssefzadeh D, et al. Combined
effects of p53, p21, and pRb expression in the progression
of bladder transitional cell carcinoma. J Clin Oncol
2004;22:1007-13.
436. Clairotte A, Lascombe I, Fauconnet S, et al. Expression of
E-cadherin and alpha-, beta-, gamma-catenins in patients
with bladder cancer: identification of gamma-catenin as
a new prognostic marker of neoplastic progression in T1
superficial urothelial tumors. American journal of clinical
pathology 2006;125:119-26.

453. Fritsche HM, Burger M, Dietmaier W, et al. Multicolor FISH


(UroVysion) facilitates follow-up of patients with high-grade
urothelial carcinoma of the bladder. American journal of
clinical pathology 2010;134:597-603.

437. Highshaw RA, McConkey DJ, Dinney CP. Integrating basic


science and clinical research in bladder cancer: update
from the first bladder Specialized Program of Research
Excellence (SPORE). Curr Opin Urol 2004;14:295-300.

454. Karnwal A, Venegas R, Shuch B, Bassett J, Rajfer J,


Reznichek R. The role of fluorescence in situ hybridization
assay for surveillance of non-muscle invasive bladder
cancer. The Canadian journal of urology 2010;17:507781.
455. Schlomer BJ, Ho R, Sagalowsky A, Ashfaq R, Lotan Y.
Prospective validation of the clinical usefulness of reflex
fluorescence in situ hybridization assay in patients with
atypical cytology for the detection of urothelial carcinoma
of the bladder. J Urol 2010;183:62-7.
456. Maffezzini M, Campodonico F, Capponi G, et al. Prognostic
significance of fluorescent in situ hybridisation in the followup of non-muscle-invasive bladder cancer. Anticancer Res
2010;30:4761-5.
457. Savic S, Zlobec I, Thalmann GN, et al. The prognostic
value of cytology and fluorescence in situ hybridization
in the follow-up of nonmuscle-invasive bladder cancer
after intravesical Bacillus Calmette-Guerin therapy.
International journal of cancer Journal international du
cancer 2009;124:2899-904.
458. Whitson J, Berry A, Carroll P, Konety B. A multicolour
fluorescence in situ hybridization test predicts recurrence
in patients with high-risk superficial bladder tumours
undergoing intravesical therapy. BJU international
2009;104:336-9.
459. Renshaw AA. UroVysion, urine cytology, and the College
of American Pathologists: where should we go from here?
Arch Pathol Lab Med 2010;134:1106-7.
460. Billerey C, Chopin D, Aubriot-Lorton MH, et al. Frequent
FGFR3 mutations in papillary non-invasive bladder (pTa)
tumors. Am J Pathol 2001;158:1955-9.

438. Ioachim E, Michael MC, Salmas M, et al. Thrombospondin-1


expression in urothelial carcinoma: prognostic significance
and association with p53 alterations, tumour angiogenesis
and extracellular matrix components. BMC Cancer
2006;6:140.
439. Lascombe I, Clairotte A, Fauconnet S, et al. N-cadherin
as a novel prognostic marker of progression in superficial
urothelial tumors. Clinical cancer research : an official
journal of the American Association for Cancer Research
2006;12:2780-7.
440. Miyamoto H, Kubota Y, Fujinami K, et al. Infrequent
somatic mutations of the p16 and p15 genes in human
bladder cancer: p16 mutations occur only in low-grade
and superficial bladder cancers. Oncol Res 1995;7:32730.
441. Miyamoto H, Kubota Y, Shuin T, et al. Analyses of p53
gene mutations in primary human bladder cancer. Oncol
Res 1993;5:245-9.
442. Miyamoto H, Shuin T, Torigoe S, Iwasaki Y, Kubota Y.
Retinoblastoma gene mutations in primary human bladder
cancer. British journal of cancer 1995;71:831-5.
443. Cheng L, Zhang S, MacLennan GT, Williamson SR,
Lopez-Beltran A, Montironi R. Bladder cancer: translating
molecular genetic insights into clinical practice. Human
pathology 2011;42:455-81.
444. Mengual L, Burset M, Ribal MJ, et al. Gene expression
signature in urine for diagnosing and assessing

158

461. Bolenz C, Shariat SF, Karakiewicz PI, et al. Human


epidermal growth factor receptor 2 expression status
provides independent prognostic information in patients
with urothelial carcinoma of the urinary bladder. BJU
international 2010;106:1216-22.

475. Ravery V, Grignon D, Angulo J, et al. Evaluation of


epidermal growth factor receptor, transforming growth
factor alpha, epidermal growth factor and c-erbB2 in
the progression of invasive bladder cancer. Urol Res
1997;25:9-17.
476. Sarkis AS, Bajorin DF, Reuter VE, et al. Prognostic value
of p53 nuclear overexpression in patients with invasive
bladder cancer treated with neoadjuvant MVAC. J Clin
Oncol 1995;13:1384-90.
477. Sarkis AS, Dalbagni G, Cordon-Cardo C, et al. Association
of P53 nuclear overexpression and tumor progression in
carcinoma in situ of the bladder. J Urol 1994;152:388-92.

462. Eissa S, Ali HS, Al Tonsi AH, Zaglol A, El Ahmady O.


HER2/neu expression in bladder cancer: relationship to
cell cycle kinetics. Clin Biochem 2005;38:142-8.
463. Gandour-Edwards R, Lara PN, Jr., Folkins AK, et al. Does
HER2/neu expression provide prognostic information
in patients with advanced urothelial carcinoma? Cancer
2002;95:1009-15.
464. Latif Z, Watters AD, Dunn I, Grigor K, Underwood MA,
Bartlett JM. HER2/neu gene amplification and protein
overexpression in G3 pT2 transitional cell carcinoma of
the bladder: a role for anti-HER2 therapy? Eur J Cancer
2004;40:56-63.

478. Sarkis AS, Dalbagni G, Cordon-Cardo C, et al. Nuclear


overexpression of p53 protein in transitional cell bladder
carcinoma: a marker for disease progression. J Natl
Cancer Inst 1993;85:53-9.
479. Garcia del Muro X, Condom E, Vigues F, et al. p53 and p21
Expression levels predict organ preservation and survival
in invasive bladder carcinoma treated with a combinedmodality approach. Cancer 2004;100:1859-67.

465. Leibl S, Zigeuner R, Hutterer G, Chromecki T, Rehak P,


Langner C. EGFR expression in urothelial carcinoma of the
upper urinary tract is associated with disease progression
and metaplastic morphology. APMIS 2008;116:27-32.
466. Mason RA, Morlock EV, Karagas MR, et al. EGFR
pathway polymorphisms and bladder cancer susceptibility
and prognosis. Carcinogenesis 2009;30:1155-60.

480. Lopez-Beltran A, Luque RJ, Alvarez-Kindelan J, et al.


Prognostic factors in stage T1 grade 3 bladder cancer
survival: the role of G1-S modulators (p53, p21Waf1,
p27kip1, Cyclin D1, and Cyclin D3) and proliferation index
(ki67-MIB1). European urology 2004;45:606-12.

467. Simonetti S, Russo R, Ciancia G, Altieri V, De Rosa


G, Insabato L. Role of polysomy 17 in transitional cell
carcinoma of the bladder: immunohistochemical study of
HER2/neu expression and fish analysis of c-erbB-2 gene
and chromosome 17. International journal of surgical
pathology 2009;17:198-205.

481. Tzai TS, Tsai YS, Chow NH. The prevalence and
clinicopathologic correlate of p16INK4a, retinoblastoma
and p53 immunoreactivity in locally advanced urinary
bladder cancer. Urologic oncology 2004;22:112-8.

468. Zuiverloon TC, van der Aa MN, van der Kwast TH, et al.
Fibroblast growth factor receptor 3 mutation analysis on
voided urine for surveillance of patients with low-grade
non-muscle-invasive bladder cancer. Clinical cancer
research : an official journal of the American Association
for Cancer Research 2010;16:3011-8.
469. Miyake M, Sugano K, Sugino H, et al. Fibroblast growth
factor receptor 3 mutation in voided urine is a useful
diagnostic marker and significant indicator of tumor
recurrence in non-muscle invasive bladder cancer. Cancer
Sci 2010;101:250-8.
470. Hernandez S, Lopez-Knowles E, Lloreta J, et al.
Prospective study of FGFR3 mutations as a prognostic
factor in nonmuscle invasive urothelial bladder carcinomas.
J Clin Oncol 2006;24:3664-71.
471. van Rhijn BW, Vis AN, van der Kwast TH, et al. Molecular
grading of urothelial cell carcinoma with fibroblast growth
factor receptor 3 and MIB-1 is superior to pathologic
grade for the prediction of clinical outcome. J Clin Oncol
2003;21:1912-21.
472. Sylvester RJ, van der Meijden AP, Oosterlinck W, et
al. Predicting recurrence and progression in individual
patients with stage Ta T1 bladder cancer using EORTC
risk tables: a combined analysis of 2596 patients from
seven EORTC trials. European urology 2006;49:466-5;
discussion 75-7.
473. Chakravarti A, Winter K, Wu CL, et al. Expression of the
epidermal growth factor receptor and Her-2 are predictors
of favorable outcome and reduced complete response
rates, respectively, in patients with muscle-invading
bladder cancers treated by concurrent radiation and
cisplatin-based chemotherapy: a report from the Radiation
Therapy Oncology Group. Int J Radiat Oncol Biol Phys
2005;62:309-17.
474. Jimenez RE, Hussain M, Bianco FJ, Jr., et al. Her-2/neu
overexpression in muscle-invasive urothelial carcinoma
of the bladder: prognostic significance and comparative
analysis in primary and metastatic tumors. Clinical cancer
research : an official journal of the American Association
for Cancer Research 2001;7:2440-7.

482. Shariat SF, Bolenz C, Karakiewicz PI, et al. p53 expression


in patients with advanced urothelial cancer of the urinary
bladder. BJU international 2010;105:489-95.
483. Fu TY, Tu MS, Tseng HH, Wang JS. Overexpression
of p27kip1 in urinary bladder urothelial carcinoma.
International journal of urology : official journal of the
Japanese Urological Association 2007;14:1084-7.
484. Kruger S, Mahnken A, Kausch I, Feller AC. P16
immunoreactivity is an independent predictor of tumor
progression in minimally invasive urothelial bladder
carcinoma. European urology 2005;47:463-7.
485. Lopez-Beltran A, Requena MJ, Luque RJ, et al. Cyclin
D3 expression in primary Ta/T1 bladder cancer. J Pathol
2006;209:106-13.
486. Sgambato A, Migaldi M, Faraglia B, et al. Loss of P27Kip1
expression correlates with tumor grade and with reduced
disease-free survival in primary superficial bladder
cancers. Cancer research 1999;59:3245-50.
487. Shariat SF, Ashfaq R, Sagalowsky AI, Lotan Y. Association
of cyclin D1 and E1 expression with disease progression
and biomarkers in patients with nonmuscle-invasive
urothelial cell carcinoma of the bladder. Urologic oncology
2007;25:468-75.
488. Yin M, Bastacky S, Parwani AV, McHale T, Dhir R. p16ink4
immunoreactivity is a reliable marker for urothelial
carcinoma in situ. Human pathology 2008;39:527-35.
489. Shariat SF, Bolenz C, Godoy G, et al. Predictive value of
combined immunohistochemical markers in patients with
pT1 urothelial carcinoma at radical cystectomy. J Urol
2009;182:78-84; discussion
490. Shariat SF, Karakiewicz PI, Ashfaq R, et al. Multiple
biomarkers improve prediction of bladder cancer recurrence
and mortality in patients undergoing cystectomy. Cancer
2008;112:315-25.
491. Dalbagni G, Presti JC, Jr., Reuter VE, et al. Molecular
genetic alterations of chromosome 17 and p53 nuclear
overexpression in human bladder cancer. Diagn Mol
Pathol 1993;2:4-13.

159

492. Margulis V, Lotan Y, Karakiewicz PI, et al. Multi-institutional


validation of the predictive value of Ki-67 labeling index in
patients with urinary bladder cancer. J Natl Cancer Inst
2009;101:114-9.
493. Margulis V, Shariat SF, Ashfaq R, Sagalowsky AI, Lotan
Y. Ki-67 is an independent predictor of bladder cancer
outcome in patients treated with radical cystectomy for
organ-confined disease. Clinical cancer research : an
official journal of the American Association for Cancer
Research 2006;12:7369-73.
494. Quintero A, Alvarez-Kindelan J, Luque RJ, et al. Ki-67
MIB1 labelling index and the prognosis of primary TaT1
urothelial cell carcinoma of the bladder. Journal of clinical
pathology 2006;59:83-8.
495. Ramos D, Ruiz A, Morell L, et al. Prognostic value of
morphometry in low grade papillary urothelial bladder
neoplasms. Anal Quant Cytol Histol 2004;26:285-94.

508. Lin HH, Ke HL, Huang SP, Wu WJ, Chen YK, Chang
LL. Increase sensitivity in detecting superficial, low
grade bladder cancer by combination analysis of
hypermethylation of E-cadherin, p16, p14, RASSF1A
genes in urine. Urologic oncology 2010;28:597-602.
509. Nishiyama N, Arai E, Chihara Y, et al. Genome-wide DNA
methylation profiles in urothelial carcinomas and urothelia
at the precancerous stage. Cancer Sci 2010;101:231-40.
510. Vallot C, Stransky N, Bernard-Pierrot I, et al. A novel
epigenetic phenotype associated with the most aggressive
pathway of bladder tumor progression. J Natl Cancer Inst
2011;103:47-60.
511. Vinci S, Giannarini G, Selli C, et al. Quantitative methylation
analysis of BCL2, hTERT, and DAPK promoters in
urine sediment for the detection of non-muscle-invasive
urothelial carcinoma of the bladder: a prospective, twocenter validation study. Urologic oncology 2011;29:150-6.

496. Heidenblad M, Lindgren D, Jonson T, et al. Tiling


resolution array CGH and high density expression profiling
of urothelial carcinomas delineate genomic amplicons and
candidate target genes specific for advanced tumors.
BMC Med Genomics 2008;1:3.

512. Wiklund ED, Bramsen JB, Hulf T, et al. Coordinated


epigenetic repression of the miR-200 family and miR-205
in invasive bladder cancer. International journal of cancer
Journal international du cancer 2011;128:1327-34.
513. Yates DR, Rehman I, Abbod MF, et al. Promoter
hypermethylation identifies progression risk in bladder
cancer. Clinical cancer research : an official journal of the
American Association for Cancer Research 2007;13:204653.

497. Lindgren D, Frigyesi A, Gudjonsson S, et al. Combined


gene expression and genomic profiling define two intrinsic
molecular subtypes of urothelial carcinoma and gene
signatures for molecular grading and outcome. Cancer
research 2010;70:3463-72.
498. Lindgren D, Liedberg F, Andersson A, et al. Molecular
characterization of early-stage bladder carcinomas by
expression profiles, FGFR3 mutation status, and loss of
9q. Oncogene 2006;25:2685-96.
499. Mitra AP, Pagliarulo V, Yang D, et al. Generation of a
concise gene panel for outcome prediction in urinary
bladder cancer. J Clin Oncol 2009;27:3929-37.
500. Rothman N, Garcia-Closas M, Chatterjee N, et al. A
multi-stage genome-wide association study of bladder
cancer identifies multiple susceptibility loci. Nat Genet
2010;42:978-84.
501. Serizawa RR, Ralfkiaer U, Steven K, et al. Integrated
genetic and epigenetic analysis of bladder cancer reveals
an additive diagnostic value of FGFR3 mutations and
hypermethylation events. International journal of cancer
Journal international du cancer 2011;129:78-87.
502. Cabello MJ, Grau L, Franco N, et al. Multiplexed
methylation profiles of tumor suppressor genes in bladder
cancer. J Mol Diagn 2011;13:29-40.
503. Catto JW, Azzouzi AR, Rehman I, et al. Promoter
hypermethylation is associated with tumor location, stage,
and subsequent progression in transitional cell carcinoma.
J Clin Oncol 2005;23:2903-10.
504. Chan MW, Chan LW, Tang NL, et al. Hypermethylation
of multiple genes in tumor tissues and voided urine in
urinary bladder cancer patients. Clinical cancer research
: an official journal of the American Association for Cancer
Research 2002;8:464-70.
505. Dudziec E, Miah S, Choudhry HM, et al. Hypermethylation
of CpG islands and shores around specific microRNAs and
mirtrons is associated with the phenotype and presence
of bladder cancer. Clinical cancer research : an official
journal of the American Association for Cancer Research
2011;17:1287-96.
506. Friedrich MG, Weisenberger DJ, Cheng JC, et al.
Detection of methylated apoptosis-associated genes in
urine sediments of bladder cancer patients. Clinical cancer
research : an official journal of the American Association
for Cancer Research 2004;10:7457-65.
507. Hoque MO, Begum S, Topaloglu O, et al. Quantitation of
promoter methylation of multiple genes in urine DNA and
bladder cancer detection. J Natl Cancer Inst 2006;98:9961004.

514. Yates DR, Rehman I, Meuth M, Cross SS, Hamdy FC,


Catto JW. Methylational urinalysis: a prospective study of
bladder cancer patients and age stratified benign controls.
Oncogene 2006;25:1984-8.
515. Ali-El-Dein B, Sarhan O, Hinev A, Ibrahiem el HI, Nabeeh
A, Ghoneim MA. Superficial bladder tumours: analysis of
prognostic factors and construction of a predictive index.
BJU international 2003;92:393-9.
516. Baak JP, Bol MG, van Diermen B, et al. DNA cytometric
features in biopsies of TaT1 urothelial cell cancer predict
recurrence and stage progression more accurately than
stage, grade, or treatment modality. Urology 2003;61:126672.
517. Bellaoui H, Chefchaouni MC, Lazrak N, Khalfaoui LC,
Yassine F, Elhamany Z. [Flow cytometric DNA analysis
and cytology in diagnosis and prognosis of bladder
tumors: preliminary results of a comparative study of
bladder lavage]. Ann Urol (Paris) 2002;36:45-52.
518. Bol MG, Baak JP, Diermen B, Janssen EA, Buhr-Wildhagen
SB, Kjellevold KH. Correlation of grade of urothelial cell
carcinomas and DNA histogram features assessed by
flow cytometry and automated image cytometry. Anal Cell
Pathol 2003;25:147-53.
519. Caraway NP, Khanna A, Payne L, Kamat AM, Katz RL.
Combination of cytologic evaluation and quantitative digital
cytometry is reliable in detecting recurrent disease in
patients with urinary diversions. Cancer 2007;111:323-9.
520. Falkman K, Tribukait
U. S-phase fraction
of the bladder--a
study. Scandinavian
2004;38:278-84.

B, Nyman CR, Larsson P, Norming


in superficial urothelial carcinoma
prospective, long-term, follow-up
journal of urology and nephrology

521. Loughman NT, Lin BP, Dent OF, Newland RC. DNA
ploidy of bladder cancer using bladder biopsy supernate
specimens. Anal Quant Cytol Histol 2003;25:146-58.
522. Lin J, Dinney CP, Grossman HB, et al. E-cadherin
promoter polymorphism (C-160A) and risk of recurrence
in patients with superficial bladder cancer. Clin Genet
2006;70:240-5.
523. Chai CY, Chen WT, Hung WC, et al. Hypoxia-inducible
factor-1alpha expression correlates with focal macrophage

160

infiltration, angiogenesis and unfavourable prognosis


in urothelial carcinoma. Journal of clinical pathology
2008;61:658-64.

539. Miyamoto H, Miller JS, Fajardo DA, Lee TK, Netto GJ,
Epstein JI. Non-invasive papillary urothelial neoplasms:
the 2004 WHO/ISUP classification system. Pathology
international 2010;60:1-8.

524. Crew JP, OBrien T, Bicknell R, Fuggle S, Cranston D,


Harris AL. Urinary vascular endothelial growth factor and
its correlation with bladder cancer recurrence rates. J Urol
1999;161:799-804.

540. Soloway MS, Sofer M, Vaidya A. Contemporary


management of stage T1 transitional cell carcinoma of the
bladder. J Urol 2002;167:1573-83.

525. Crew JP, OBrien T, Bradburn M, et al. Vascular


endothelial growth factor is a predictor of relapse and
stage progression in superficial bladder cancer. Cancer
research 1997;57:5281-5.

541. Malekzadeh K, Sobti RC, Nikbakht M, et al. Methylation


patterns of Rb1 and Casp-8 promoters and their impact
on their expression in bladder cancer. Cancer Invest
2009;27:70-80.

526. Ioachim E, Michael M, Salmas M, Michael MM,


Stavropoulos NE, Malamou-Mitsi V. Hypoxia-inducible
factors HIF-1alpha and HIF-2alpha expression in bladder
cancer and their associations with other angiogenesisrelated proteins. Urologia internationalis 2006;77:255-63.

542. Shariat SF, Lotan Y, Karakiewicz PI, et al. p53 predictive


value for pT1-2 N0 disease at radical cystectomy. J Urol
2009;182:907-13.
543. van der Kwast TH, Bapat B. Predicting favourable
prognosis of urothelial carcinoma: gene expression and
genome profiling. Curr Opin Urol 2009;19:516-21.

527. Palit V, Phillips RM, Puri R, Shah T, Bibby MC. Expression


of HIF-1alpha and Glut-1 in human bladder cancer. Oncol
Rep 2005;14:909-13.

544. Wilhelm-Benartzi CS, Koestler DC, Houseman EA, et al.


DNA methylation profiles delineate etiologic heterogeneity
and clinically important subgroups of bladder cancer.
Carcinogenesis 2010;31:1972-6.

528. Turner KJ, Crew JP, Wykoff CC, et al. The hypoxiainducible genes VEGF and CA9 are differentially regulated
in superficial vs invasive bladder cancer. British journal of
cancer 2002;86:1276-82.

545. Bellmunt J, Hussain M, Dinney CP. Novel approaches with


targeted therapies in bladder cancer. Therapy of bladder
cancer by blockade of the epidermal growth factor receptor
family. Crit Rev Oncol Hematol 2003;46 Suppl:S85-104.
546. Black PC, Agarwal PK, Dinney CP. Targeted therapies
in bladder cancer--an update. Urologic oncology
2007;25:433-8.
547. Black PC, Dinney CP. Bladder cancer angiogenesis and
metastasis--translation from murine model to clinical trial.
Cancer Metastasis Rev 2007;26:623-34.
548. Wallerand H, Reiter RR, Ravaud A. Molecular targeting
in the treatment of either advanced or metastatic bladder
cancer or both according to the signalling pathways. Curr
Opin Urol 2008;18:524-32.
549. Wallerand H, Robert G, Bernhard JC, Ravaud A, Ferriere
JM. [Targeted therapy for locally advanced and/or
metastatic bladder cancer]. Prog Urol 2008;18:407-17.
550. Helpap B. In: Eble JN, Sauter G, Epstein JI, Sesterhenn
I, eds. The World Health Organization Classification of
Tumours of the Urinary System and Male Genital System.
Lyon, France: IARC Press, 2004; 2004:148911.
551. Velcheti V, Govindan R. Metastatic cancer involving
bladder: a review. The Canadian journal of urology
2007;14:3443-8.
552. Chow J, Unger P, Xiao G-Q. Metastatic Tumors to the
Urinary Bladder. Mod Pathol 2011;24:185A.
553. Alsharif M, Aslan DL, Jessurun J, Gulbahce HE,
Pambuccian SE. Cytologic diagnosis of metastatic
seminoma to the prostate and urinary bladder: a case
report. Diagnostic cytopathology 2008;36:734-8.
554. McAchran SE, Williams DH, MacLennan GT. Renal
cell carcinoma metastasis to the bladder. J Urol
2010;184:726-7.
555. Bates AW, Baithun SI. The significance of secondary
neoplasms of the urinary and male genital tract. Virchows
Arch 2002;440:640-7.
556. Kunju LP, Mehra R, Snyder M, Shah RB. Prostatespecific antigen, high-molecular-weight cytokeratin (clone
34betaE12), and/or p63: an optimal immunohistochemical
panel to distinguish poorly differentiated prostate
adenocarcinoma from urothelial carcinoma. American
journal of clinical pathology 2006;125:675-81.
557. Mhawech P, Uchida T, Pelte MF. Immunohistochemical
profile of high-grade urothelial bladder carcinoma
and prostate adenocarcinoma. Human pathology
2002;33:1136-40.

529. Platt FM, Hurst CD, Taylor CF, Gregory WM, Harnden P,
Knowles MA. Spectrum of phosphatidylinositol 3-kinase
pathway gene alterations in bladder cancer. Clinical cancer
research : an official journal of the American Association
for Cancer Research 2009;15:6008-17.
530. Schultz L, Albadine R, Hicks J, et al. Expression status
and prognostic significance of mammalian target of
rapamycin pathway members in urothelial carcinoma of
urinary bladder after cystectomy. Cancer 2010;116:551726.
531. Theodoropoulos VE, Lazaris A, Sofras F, et al. Hypoxiainducible factor 1 alpha expression correlates with
angiogenesis and unfavorable prognosis in bladder
cancer. European urology 2004;46:200-8.
532. Tickoo SK, Milowsky MI, Dhar N, et al. Hypoxia-inducible
factor and mammalian target of rapamycin pathway markers
in urothelial carcinoma of the bladder: possible therapeutic
implications. BJU international 2011;107:844-9.
533. Comperat E, Camparo P, Haus R, et al. Aurora-A/STK15 is a predictive factor for recurrent behaviour in noninvasive bladder carcinoma: a study of 128 cases of noninvasive neoplasms. Virchows Archiv : an international
journal of pathology 2007;450:419-24.
534. Mhawech-Fauceglia P, Fischer G, Beck A, Cheney RT,
Herrmann FR. Raf1, Aurora-A/STK15 and E-cadherin
biomarkers expression in patients with pTa/pT1 urothelial
bladder carcinoma; a retrospective TMA study of 246
patients with long-term follow-up. Eur J Surg Oncol
2006;32:439-44.
535. Catto JW, Miah S, Owen HC, et al. Distinct microRNA
alterations characterize high- and low-grade bladder
cancer. Cancer research 2009;69:8472-81.
536. Veerla S, Panagopoulos I, Jin Y, Lindgren D, Hoglund
M. Promoter analysis of epigenetically controlled
genes in bladder cancer. Genes Chromosomes Cancer
2008;47:368-78.
537. Wszolek MF, Rieger-Christ KM, Kenney PA, et al. A
MicroRNA expression profile defining the invasive bladder
tumor phenotype. Urologic oncology 2009.
538. Millan-Rodriguez F, Chechile-Toniolo G, Salvador-Bayarri
J, Palou J, Vicente-Rodriguez J. Multivariate analysis
of the prognostic factors of primary superficial bladder
cancer. J Urol 2000;163:73-8.

161

558. Genega EM, Hutchinson B, Reuter VE, Gaudin PB.


Immunophenotype of high-grade prostatic adenocarcinoma
and urothelial carcinoma. Mod Pathol 2000;13:1186-91.

575. Yin M, Dhir R, Parwani AV. Diagnostic utility of p501s


(prostein) in comparison to prostate specific antigen (PSA)
for the detection of metastatic prostatic adenocarcinoma.
Diagn Pathol 2007;2:41.

559. Bassily NH, Vallorosi CJ, Akdas G, Montie JE, Rubin MA.
Coordinate expression of cytokeratins 7 and 20 in prostate
adenocarcinoma and bladder urothelial carcinoma.
American journal of clinical pathology 2000;113:383-8.

576. Sweat SD, Pacelli A, Murphy GP, Bostwick DG. Prostatespecific membrane antigen expression is greatest in
prostate adenocarcinoma and lymph node metastases.
Urology 1998;52:637-40.

560. Goldstein NS. Immunophenotypic characterization of


225 prostate adenocarcinomas with intermediate or high
Gleason scores. American journal of clinical pathology
2002;117:471-7.

577. Gurel B, Ali TZ, Montgomery EA, et al. NKX3.1 as a marker


of prostatic origin in metastatic tumors. The American
journal of surgical pathology 2010;34:1097-105.

561. Epstein JI. PSA and PAP as immunohistochemical markers


in prostate cancer. Urol Clin North Am 1993;20:757-70.

578. Jaraj S, Augsten M, Haaggarth L, et al. GAD1 is a biomarker


for benign and malignant prostatic tissue. Scandinavian
Journal of Urology and Nephrology 2011;45:39-45.

562. Shah RB, Mehra R, Chinnaiyan AM, et al. Androgenindependent prostate cancer is a heterogeneous group of
diseases: lessons from a rapid autopsy program. Cancer
research 2004;64:9209-16.

579. Tamboli P, Mohsin SK, Hailemariam S, Amin MB.


Colonic adenocarcinoma metastatic to the urinary tract
versus primary tumors of the urinary tract with glandular
differentiation: a report of 7 cases and investigation using
a limited immunohistochemical panel. Arch Pathol Lab
Med 2002;126:1057-63.

563. Chu P, Wu E, Weiss LM. Cytokeratin 7 and cytokeratin 20


expression in epithelial neoplasms: a survey of 435 cases.
Mod Pathol 2000;13:962-72.

580. Werling RW, Yaziji H, Bacchi CE, Gown AM. CDX2, a


highly sensitive and specific marker of adenocarcinomas
of intestinal origin: an immunohistochemical survey of 476
primary and metastatic carcinomas. The American journal
of surgical pathology 2003;27:303-10.

564. Wang NP, Zee S, Zarbo RJ, Bacchi CE, Gown AM.
Coordinate Expression of Cytokeratins 7 and 20
Defines Unique Subsets of Carcinomas. . Applied
Immunohistochemistry 1995;3:99-107.
565. Lindeman N, Weidner N. Immunohistochemical Profile
of Prostatic and Urothelial Carcinoma: Impact of Heatinduced Epitope Retrieval and Presentation of Tumors with
Intermediate Features. Applied Immunohistochemistry
1996;4:264-75.

581. Chu PG, Weiss LM. Keratin expression in human tissues


and neoplasms. Histopathology 2002;40:403-39.
582. Suh N, Yang XJ, Tretiakova MS, Humphrey PA, Wang HL.
Value of CDX2, villin, and alpha-methylacyl coenzyme A
racemase immunostains in the distinction between primary
adenocarcinoma of the bladder and secondary colorectal
adenocarcinoma. Mod Pathol 2005;18:1217-22.

566. Jiang Z, Fanger GR, Woda BA, et al. Expression of alphamethylacyl-CoA racemase (P504s) in various malignant
neoplasms and normal tissues: astudy of 761 cases.
Human pathology 2003;34:792-6.
567. Beach R, Gown AM, De Peralta-Venturina MN, et al.
P504S immunohistochemical detection in 405 prostatic
specimens including 376 18-gauge needle biopsies. The
American journal of surgical pathology 2002;26:1588-96.

583. Paner GP, McKenney JK, Barkan GA, et al.


Immunohistochemical analysis in a morphologic spectrum
of urachal epithelial neoplasms: diagnostic implications
and pitfalls. The American journal of surgical pathology
2011;35:787-98.

568. Tretiakova MS, Sahoo S, Takahashi M, et al. Expression


of alpha-methylacyl-CoA racemase in papillary renal cell
carcinoma. The American journal of surgical pathology
2004;28:69-76.

584. Ordonez NG. Value of thrombomodulin immunostaining in


the diagnosis of transitional cell carcinoma: a comparative
study with carcinoembryonic antigen. Histopathology
1997;31:517-24.

569. Sheridan T, Herawi M, Epstein JI, Illei PB. The role of P501S
and PSA in the diagnosis of metastatic adenocarcinoma of
the prostate. Am J Surg Pathol 2007;31:1351-5.

585. Parker DC, Folpe AL, Bell J, et al. Potential utility of


uroplakin III, thrombomodulin, high molecular weight
cytokeratin, and cytokeratin 20 in noninvasive, invasive,
and metastatic urothelial (transitional cell) carcinomas.
The American journal of surgical pathology 2003;27:1-10.

570. Mhawech-Fauceglia P, Zhang S, Terracciano L, et al.


Prostate-specific membrane antigen (PSMA) protein
expression in normal and neoplastic tissues and its
sensitivity and specificity in prostate adenocarcinoma: an
immunohistochemical study using mutiple tumour tissue
microarray technique. Histopathology 2007;50:472-83.

586. Kaufmann O, Volmerig J, Dietel M. Uroplakin III is a highly


specific and moderately sensitive immunohistochemical
marker for primary and metastatic urothelial carcinomas.
American journal of clinical pathology 2000;113:683-7.

571. Kalos M, Askaa J, Hylander BL, et al. Prostein expression


is highly restricted to normal and malignant prostate
tissues. Prostate 2004;60:246-56.

587. Choi C, Amin MB, Tamboli P, Eble JN, Young RH. Glandular
Neoplasms of the Urachus: Clinicopathologic and
Immunohistochemical Analysis of 43 Cases with Special
Emphasis on Low-Grade Mucinous Cystic Tumors. Mod
Pathol 2005;18:134A.

572. Chang SS, Reuter VE, Heston WD, Gaudin PB.


Comparison of anti-prostate-specific membrane antigen
antibodies and other immunomarkers in metastatic
prostate carcinoma. Urology 2001;57:1179-83.

588. Zhou M, Chinnaiyan AM, Kleer CG, Lucas PC, Rubin


MA. Alpha-Methylacyl-CoA racemase: a novel tumor
marker over-expressed in several human cancers and
their precursor lesions. The American journal of surgical
pathology 2002;26:926-31.

573. Silver DA, Pellicer I, Fair WR, Heston WD, Cordon-Cardo


C. Prostate-specific membrane antigen expression in
normal and malignant human tissues. Clinical cancer
research : an official journal of the American Association
for Cancer Research 1997;3:81-5.

589. Park KJ, Bramlage MP, Ellenson LH, Pirog EC.


Immunoprofile of adenocarcinomas of the endometrium,
endocervix, and ovary with mucinous differentiation.
Applied immunohistochemistry & molecular morphology:
AIMM / official publication of the Society for Applied
Immunohistochemistry 2009;17:8-11.

574. Tanaka M, Komuro I, Inagaki H, Jenkins NA, Copeland


NG, Izumo S. Nkx3.1, a murine homolog of Ddrosophila
bagpipe, regulates epithelial ductal branching and
proliferation of the prostate and palatine glands. Dev Dyn
2000;219:248-60.

162

590. McCluggage WG, Shah R, Connolly LE, McBride HA.


Intestinal-type cervical adenocarcinoma in situ and
adenocarcinoma exhibit a partial enteric immunophenotype
with consistent expression of CDX2. Int J Gynecol Pathol
2008;27:92-100.

606. Huh JW, Lee JH, Kim HR. Expression of p16, p53, and Ki67 in colorectal adenocarcinoma: a study of 356 surgically
resected cases. Hepatogastroenterology 2010;57:734-40.
607. Chew I, Post MD, Silvestro CG. p16 expression in
squamous and trophoblastic lesions of the upper female
genital tract. . Int J Gynecol Pathol 2010;in press.

591. Robinson CE, Sarode VR, Albores-Saavedra J. Mixed


papillary transitional cell carcinoma and adenocarcinoma
of the uterine cervix: a clinicopathologic study of three
cases. Int J Gynecol Pathol 2003;22:220-5.

608. Herawi M, Drew PA, Pan CC, Epstein JI. Clear cell
adenocarcinoma of the bladder and urethra: cases
diffusely mimicking nephrogenic adenoma. Human
pathology 2010;41:594-601.

592. Al-Nafussi AI, Al-Yusif R. Papillary squamotransitional cell


carcinoma of the uterine cervix: an advanced stage disease
despite superficial location: report of two cases and review
of the literature. Eur J Gynaecol Oncol 1998;19:455-7.

609. Oliva E, Young RH. Clear cell adenocarcinoma of the


urethra: a clinicopathologic analysis of 19 cases. Mod
Pathol 1996;9:513-20.

593. Rose PG, Stoler MH, Abdul-Karim FW. Papillary


squamotransitional cell carcinoma of the vagina. Int J
Gynecol Pathol 1998;17:372-5.

610. Sun K, Huan Y, Unger PD. Clear cell adenocarcinoma


of urinary bladder and urethra: another urinary tract
lesion immunoreactive for P504S. Arch Pathol Lab Med
2008;132:1417-22.

594. Marino-Enriquez A, Gonzalez-Rocha T, Burgos E, et


al. Transitional cell carcinoma of the endometrium and
endometrial carcinoma with transitional cell differentiation:
a clinicopathologic study of 5 cases and review of the
literature. Human pathology 2008;39:1606-13.

611. Kawano K, Yano M, Kitahara S, Yasuda K. Clear cell


adenocarcinoma of the female urethra showing strong
immunostaining for prostate-specific antigen. BJU
international 2001;87:412-3.

595. Patrelli TS, Silini EM, Berretta R, et al. Squamotransitional


cell carcinoma of the vagina: diagnosis and clinical
management : a literature review starting from a rare case
report. Pathol Oncol Res 2011;17:149-53.

612. Ordi J, Romagosa C, Tavassoli FA, et al. CD10 expression


in epithelial tissues and tumors of the gynecologic
tract: a useful marker in the diagnosis of mesonephric,
trophoblastic, and clear cell tumors. The American journal
of surgical pathology 2003;27:178-86.

596. Kaufmann O, Fietze E, Mengs J, Dietel M. Value of p63


and cytokeratin 5/6 as immunohistochemical markers
for the differential diagnosis of poorly differentiated and
undifferentiated carcinomas. American journal of clinical
pathology 2001;116:823-30.

613. Tong GX, Weeden EM, Hamele-Bena D, et al. Expression


of PAX8 in nephrogenic adenoma and clear cell
adenocarcinoma of the lower urinary tract: evidence of
related histogenesis? The American journal of surgical
pathology 2008;32:1380-7.

597. Pereira TC, Share SM, Magalhaes AV, Silverman JF.


Can we tell the site of origin of metastatic squamous cell
carcinoma? An immunohistochemical tissue microarray
study of 194 cases. Applied immunohistochemistry &
molecular morphology : AIMM / official publication of the
Society for Applied Immunohistochemistry 2011;19:10-4.

614. Nonaka D, Chiriboga L, Soslow RA. Expression of pax8 as


a useful marker in distinguishing ovarian carcinomas from
mammary carcinomas. The American journal of surgical
pathology 2008;32:1566-71.

598. Houghton O, McCluggage WG. The expression and


diagnostic utility of p63 in the female genital tract. Adv
Anat Pathol 2009;16:316-21.

615. De Torres I, Fortuno MA, Raventos A, Tarragona J, Banus


JM, Vidal MT. Primary malignant melanoma of the bladder:
immunohistochemical study of a new case and review of
the literature. J Urol 1995;154:525-7.

599. Focchi GR, Silva ID, Nogueira-de-Souza NC, Dobo C,


Oshima CT, Stavale JN. Immunohistochemical expression
of p16(INK4A) in normal uterine cervix, nonneoplastic
epithelial lesions, and low-grade squamous intraepithelial
lesions. J Low Genit Tract Dis 2007;11:98-104.

616. Lange-Welker U, Papadopoulos I, Wacker HH. Primary


malignant melanoma of the bladder. A case report and
literature review. Urologia internationalis 1993;50:226-8.
617. Oliva E, Quinn TR, Amin MB, et al. Primary malignant
melanoma of the urethra: a clinicopathologic analysis
of 15 cases. The American journal of surgical pathology
2000;24:785-96.

600. Klaes R, Benner A, Friedrich T, et al. p16INK4a


immunohistochemistry improves interobserver agreement
in the diagnosis of cervical intraepithelial neoplasia. The
American journal of surgical pathology 2002;26:1389-99.

618. Raspollini MR, Comin CE, Crisci A, Chilosi M. The use


of placental S100 (S100P), GATA3 and napsin A in the
differential diagnosis of primary adenocarcinoma of the
bladder and bladder metastasis from adenocarcinoma of
the lung. Pathologica 2010;102:33-5.
619. Matoso A, Singh K, Jacob R, et al. Comparison of
thyroid transcription factor-1 expression by 2 monoclonal
antibodies in pulmonary and nonpulmonary primary
tumors. Applied immunohistochemistry & molecular
morphology : AIMM / official publication of the Society for
Applied Immunohistochemistry 2010;18:142-9.
620. Jones TD, Kernek KM, Yang XJ, et al. Thyroid transcription
factor 1 expression in small cell carcinoma of the urinary
bladder: an immunohistochemical profile of 44 cases.
Human pathology 2005;36:718-23.
621. Thompson S, Cioffi-Lavina M, Gomez C. Distinction
of High-Grade Neuroendocrine Carcinoma/Small Cell
Carcinoma from Conventional Urothelial Carcinoma of
Urinary Bladder: An Immunohistochemical Approach. .
Appl Immunohistochem Mol Morphol 2011 (accepted for
publication).

601. Cioffi-Lavina M, Chapman-Fredricks J, Gomez-Fernandez


C, Ganjei-Azar P, Manoharan M, Jorda M. P16 expression
in squamous cell carcinomas of cervix and bladder.
Applied immunohistochemistry & molecular morphology
: AIMM / official publication of the Society for Applied
Immunohistochemistry 2010;18:344-7.
602. Westenend PJ, Stoop JA, Hendriks JG. Human
papillomaviruses 6/11, 16/18 and 31/33/51 are not
associated with squamous cell carcinoma of the urinary
bladder. BJU international 2001;88:198-201.
603. Soslow RA, Rouse RV, Hendrickson MR, Silva EG,
Longacre TA. Transitional cell neoplasms of the ovary
and urinary bladder: a comparative immunohistochemical
analysis. Int J Gynecol Pathol 1996;15:257-65.
604. Drew PA, Hong B, Massoll NA, Ripley DL. Characterization
of papillary squamotransitional cell carcinoma of the
cervix. J Low Genit Tract Dis 2005;9:149-53.
605. Ordonez NG. Thrombomodulin expression in transitional
cell carcinoma. American journal of clinical pathology
1998;110:385-90.

163

622. Feldman PA, Madeb R, Naroditsky I, Halachmi S,


Nativ O. Metastatic breast cancer to the bladder: a
diagnostic challenge and review of the literature. Urology
2002;59:138.

642. Viddeleer AC, Lycklama a Nijeholt GA, Beekhuis-Brussee


JA. A late manifestation of testicular seminoma in the
bladder in a renal transplant recipient: a case report. J Urol
1992;148:401-2.

623. Ramalingam P, Middleton LP, Tamboli P, Troncoso P, Silva


EG, Ayala AG. Invasive micropapillary carcinoma of the
breast metastatic to the urinary bladder and endometrium:
diagnostic pitfalls and review of the literature of tumors with
micropapillary features. Annals of diagnostic pathology
2003;7:112-9.

643. Wick MR, Swanson PE, Manivel JC. Placental-like


alkaline phosphatase reactivity in human tumors: an
immunohistochemical study of 520 cases. Human
pathology 1987;18:946-54.
644. Preda O, Nicolae A, Aneiros-Fernandez J, Borda
A, Nogales FF. Glypican 3 is a sensitive, but not a
specific, marker for the diagnosis of yolk sac tumours.
Histopathology 2011;58:312-4; author reply 4-5.

624. Ryan PD, Harisinghani M, Lerwill MF, Kaufman DS. Case


records of the Massachusetts General Hospital. Case
6-2006. A 71-year-old woman with urinary incontinence and
a mass in the bladder. N Engl J Med 2006;354:850-6.

645. Cox RM, Schneider AG, Sangoi AR, Clingan WJ,


Gokden N, McKenney JK. Invasive urothelial carcinoma
with chordoid features: a report of 12 distinct cases
characterized by prominent myxoid stroma and cordlike
epithelial architecture. The American journal of surgical
pathology 2009;33:1213-9.

625. Tot T. Cytokeratins 20 and 7 as biomarkers: usefulness in


discriminating primary from metastatic adenocarcinoma.
Eur J Cancer 2002;38:758-63.
626. Chu PG, Weiss LM. Expression of cytokeratin 5/6 in
epithelial neoplasms: an immunohistochemical study of
509 cases. Mod Pathol 2002;15:6-10.
627. Tot T. The role of cytokeratins 20 and 7 and estrogen
receptor analysis in separation of metastatic lobular
carcinoma of the breast and metastatic signet ring
cell carcinoma of the gastrointestinal tract. APMIS
2000;108:467-72.

646. Cheng L. Establishing a germ cell origin for metastatic


tumors using OCT4 immunohistochemistry. Cancer
2004;101:2006-10.
647. Cao D, Humphrey PA, Allan RW. SALL4 is a novel sensitive
and specific marker for metastatic germ cell tumors, with
particular utility in detection of metastatic yolk sac tumors.
Cancer 2009;115:2640-51.

628. Croft PR, Lathrop SL, Feddersen RM, Joste NE. Estrogen
receptor expression in papillary urothelial carcinoma of
the bladder and ovarian transitional cell carcinoma. Arch
Pathol Lab Med 2005;129:194-9.

648. Wang F, Liu A, Peng Y, et al. Diagnostic utility of SALL4


in extragonadal yolk sac tumors: an immunohistochemical
study of 59 cases with comparison to placental-like
alkaline phosphatase, alpha-fetoprotein, and glypican-3.
The American journal of surgical pathology 2009;33:152939.

629. Shen SS, Smith CL, Hsieh JT, et al. Expression of estrogen
receptors-alpha and -beta in bladder cancer cell lines and
human bladder tumor tissue. Cancer 2006;106:2610-6.
630. Kaufmann O, Baume H, Dietel M. Detection of oestrogen
receptors in non-invasive and invasive transitional cell
carcinomas of the urinary bladder using both conventional
immunohistochemistry and the tyramide staining
amplification (TSA) technique. J Pathol 1998;186:165-8.

649. Murata S, Iseki M, Kinjo M, et al. Molecular and


immunohistologic analyses cannot reliably solve diagnostic
variation of flat intraepithelial lesions of the urinary bladder.
American journal of clinical pathology 2010;134:862-72.
650. Mallofre C, Castillo M, Morente V, Sole M.
Immunohistochemical expression of CK20, p53, and Ki-67
as objective markers of urothelial dysplasia. Mod Pathol
2003;16:187-91.

631. Tot T. Patterns of distribution of cytokeratins 20 and 7 in


special types of invasive breast carcinoma: a study of 123
cases. Annals of diagnostic pathology 1999;3:350-6.
632. Lerwill MF. Current practical applications of diagnostic
immunohistochemistry in breast pathology. The American
journal of surgical pathology 2004;28:1076-91.

651. McKenney JK, Desai S, Cohen C, Amin MB. Discriminatory


immunohistochemical staining of urothelial carcinoma
in situ and non-neoplastic urothelium: an analysis of
cytokeratin 20, p53, and CD44 antigens. The American
journal of surgical pathology 2001;25:1074-8.

633. Tavora F, Epstein JI. Urothelial carcinoma with abundant


myxoid stroma. Human pathology 2009;40:1391-8.

652. Sun W, Zhang PL, Herrera GA. p53 protein and Ki67 overexpression in urothelial dysplasia of bladder.
Applied immunohistochemistry & molecular morphology
: AIMM / official publication of the Society for Applied
Immunohistochemistry 2002;10:327-31.

634. Saitoh H. Distant metastasis of renal adenocarcinoma.


Cancer 1981;48:1487-91.
635. Remis RE, Halverstadt DB. Metastatic renal cell carcinoma
to the bladder: case report and review of the literature. J
Urol 1986;136:1294-6.

653. Wagner U, Sauter G, Moch H, et al. Patterns of p53,


erbB-2, and EGF-r expression in premalignant lesions of
the urinary bladder. Human pathology 1995;26:970-8.

636. Gulati M, Gore JL, Pantuck AJ, Kim Y, Barajas L, Rajfer


J. Ureteral tumor thrombus from renal cell carcinoma
extending into bladder. Urologic oncology 2007;25:393-5.

654. Yildiz IZ, Recavarren R, Armah HB, Bastacky S, Dhir


R, Parwani AV. Utility of a dual immunostain cocktail
comprising of p53 and CK20 to aid in the diagnosis of nonneoplastic and neoplastic bladder biopsies. Diagn Pathol
2009;4:35.

637. Raviv S, Eggener SE, Williams DH, Garnett JE, Pins MR,
Smith ND. Long-term survival after drop metastases of
renal cell carcinoma to the bladder. Urology 2002;60:697.
638. Truong LD, Shen SS. Immunohistochemical diagnosis of
renal neoplasms. Arch Pathol Lab Med 2011;135:92-109.
639. Bahadir B, Behzatoglu K, Bektas S, Bozkurt ER, Ozdamar
SO. CD10 expression in urothelial carcinoma of the
bladder. Diagn Pathol 2009;4:38.

655. Kay EW, Barry Walsh CJ, Whelan D, OGrady A, Leader


MB. Inter-observer variation of p53 immunohistochemistry-an assessment of a practical problem and comparison
with other studies. Br J Biomed Sci 1996;53:101-7.

640. Pan CX, Yang XJ, Lopez-Beltran A, et al. c-kit Expression


in small cell carcinoma of the urinary bladder: prognostic
and therapeutic implications. Mod Pathol 2005;18:320-3.

656. Harnden P, Eardley I, Joyce AD, Southgate J. Cytokeratin


20 as an objective marker of urothelial dysplasia. British
journal of urology 1996;78:870-5.

641. Jorda M, Manoharam M. Collecting Duct Carcinoma of


Kidney: Differential Diagnosis of Neoplasms Involving the
Renal Medulla. . Pathology Case Reviews 2006;11:191-6.

657. Kunju LP, Lee CT, Montie J, Shah RB. Utility of cytokeratin
20 and Ki-67 as markers of urothelial dysplasia. Pathology
international 2005;55:248-54.

164

658. Yin H, He Q, Li T, Leong AS. Cytokeratin 20 and Ki-67


to distinguish carcinoma in situ from flat non-neoplastic
urothelium. Applied immunohistochemistry & molecular
morphology : AIMM / official publication of the Society for
Applied Immunohistochemistry 2006;14:260-5.
659. Layfield LJ, Elsheikh TM, Fili A, Nayar R, Shidham V.
Review of the state of the art and recommendations of
the Papanicolaou Society of Cytopathology for urinary
cytology procedures and reporting : the Papanicolaou
Society of Cytopathology Practice Guidelines Task Force.
Diagnostic cytopathology 2004;30:24-30.
660. Brimo F, Vollmer RT, Case B, Aprikian A, Kassouf W, Auger
M. Accuracy of urine cytology and the significance of an
atypical category. American journal of clinical pathology
2009;132:785-93.

677. Bastacky S, Ibrahim S, Wilczynski SP, Murphy WM. The


accuracy of urinary cytology in daily practice. Cancer
1999;87:118-28.
678. Sarosdy MF, Schellhammer P, Bokinsky G, et al.
Clinical evaluation of a multi-target fluorescent in situ
hybridization assay for detection of bladder cancer. J Urol
2002;168:1950-4.
679. Hajdinjak T. UroVysion FISH test for detecting urothelial
cancers: meta-analysis of diagnostic accuracy and
comparison with urinary cytology testing. Urologic
oncology 2008;26:646-51.
680. Halling KC, Kipp BR. Bladder cancer detection using FISH
(UroVysion assay). Adv Anat Pathol 2008;15:279-86.
681. Glatz K, Willi N, Glatz D, et al. An international telecytologic
quiz on urinary cytology reveals educational deficits
and absence of a commonly used classification system.
American journal of clinical pathology 2006;126:294-301.

661. Bubendorf L. Multiprobe fluorescence in situ hybridization


(UroVysion) for the detection of urothelial carcinoma FISHing for the right catch. Acta cytologica;55:113-9.
662. Renshaw AA. Subclassifying atypical urinary cytology
specimens. Cancer 2000;90:222-9.

682. Kipp BR, Halling KC, Campion MB, et al. Assessing the
value of reflex fluorescence in situ hybridization testing
in the diagnosis of bladder cancer when routine urine
cytological examination is equivocal. J Urol 2008;179:1296301; discussion 301.

663. Caraway NP, Katz RL. A review on the current state of


urine cytology emphasizing the role of fluorescence in
situ hybridization as an adjunct to diagnosis. Cancer
Cytopathol;118:175-83.
664. Hudson MA, Herr HW. Carcinoma in situ of the bladder. J
Urol 1995;153:564-72.

683. Kipp BR, Tanasescu M, Else TA, et al. Quantitative


fluorescence in situ hybridization and its ability to predict
bladder cancer recurrence and progression to muscleinvasive bladder cancer. J Mol Diagn 2009;11:148-54.

665. Wiener HG, Mian C, Haitel A, Pycha A, Schatzl G,


Marberger M. Can urine bound diagnostic tests replace
cystoscopy in the management of bladder cancer? J Urol
1998;159:1876-80.

684. Lotan Y, Bensalah K, Ruddell T, Shariat SF, Sagalowsky AI,


Ashfaq R. Prospective evaluation of the clinical usefulness
of reflex fluorescence in situ hybridization assay in patients
with atypical cytology for the detection of urothelial
carcinoma of the bladder. J Urol 2008;179:2164-9.

666. Halling KC, King W, Sokolova IA, et al. A comparison of


cytology and fluorescence in situ hybridization for the
detection of urothelial carcinoma. J Urol 2000;164:176875.

685. Skacel M, Fahmy M, Brainard JA, et al. Multitarget


fluorescence in situ hybridization assay detects transitional
cell carcinoma in the majority of patients with bladder
cancer and atypical or negative urine cytology. J Urol
2003;169:2101-5.

667. Krause FS, Rauch A, Schrott KM, Engehausen DG. Clinical


decisions for treatment of different staged bladder cancer
based on multitarget fluorescence in situ hybridization
assays? World J Urol 2006;24:418-22.

686. Kipp BR, Fritcher EG, del Rosario KM, Stevens CL, Sebo
TJ, Halling KC. A systematic approach to identifying
urothelial cells likely to be polysomic by fluorescence in
situ hybridization. Anal Quant Cytol Histol 2005;27:31722.

668. Laudadio J, Keane TE, Reeves HM, et al. Fluorescence in


situ hybridization for detecting transitional cell carcinoma:
implications for clinical practice. BJU international
2005;96:1280-5.

687. Mengual L, Marin-Aguilera M, Ribal MJ, et al. Clinical utility


of fluorescent in situ hybridization for the surveillance of
bladder cancer patients treated with bacillus CalmetteGuerin therapy. European urology 2007;52:752-9.

669. Lodde M, Mian C, Comploj E, et al. uCyt+ test: alternative


to cystoscopy for less-invasive follow-up of patients with
low risk of urothelial carcinoma. Urology 2006;67:950-4.
670. Mian C, Lodde M, Comploj E, et al. Multiprobe
fluorescence in situ hybridisation: prognostic perspectives
in superficial bladder cancer. Journal of clinical pathology
2006;59:984-7.

688. Gudjonsson S, Isfoss BL, Hansson K, et al. The value of the


UroVysion assay for surveillance of non-muscle-invasive
bladder cancer. European urology 2008;54:402-8.
689. Zellweger T, Benz G, Cathomas G, et al. Multi-target
fluorescence in situ hybridization in bladder washings
for prediction of recurrent bladder cancer. International
journal of cancer Journal international du cancer
2006;119:1660-5.

671. Schroeder GL, Lorenzo-Gomez MF, Hautmann SH, et al.


A side by side comparison of cytology and biomarkers for
bladder cancer detection. J Urol 2004;172:1123-6.
672. Sun Y, He DL, Ma Q, et al. Comparison of seven screening
methods in the diagnosis of bladder cancer. Chin Med J
(Engl) 2006;119:1763-71.
673. Tetu B, Tiguert R, Harel F, Fradet Y. ImmunoCyt/uCyt+
improves the sensitivity of urine cytology in patients followed
for urothelial carcinoma. Mod Pathol 2005;18:83-9.
674. Toma MI, Friedrich MG, Hautmann SH, et al. Comparison
of the ImmunoCyt test and urinary cytology with other
urine tests in the detection and surveillance of bladder
cancer. World J Urol 2004;22:145-9.
675. Mack D, Frick J. Diagnostic problems of urine cytology
on initial follow-up after intravesical immunotherapy with
Calmette-Guerin bacillus for superficial bladder cancer.
Urologia internationalis 1994;52:204-7.
676. Rosenthal DL, Raab, S.S. Cytologic Detection of Urothelial
Lesions. In. Pittsburgh, PA: Springer; 2005.

690. Youssef RF, Schlomer BJ, Ho R, Sagalowsky AI, Ashfaq


R, Lotan Y. Role of fluorescence in situ hybridization in
bladder cancer surveillance of patients with negative
cytology. Urologic oncology.
691. Wojcik EM, Brownlie RJ, Bassler TJ, Miller MC. Superficial
urothelial (umbrella) cells. A potential cause of abnormal
DNA ploidy results in urine specimens. Anal Quant Cytol
Histol 2000;22:411-5.
692. Chrouser K, Leibovich B, Bergstralh E, Zincke H, Blute M.
Bladder cancer risk following primary and adjuvant external
beam radiation for prostate cancer. J Urol 2005;174:10710; discussion 10-1.
693. Murphy WM. Whats the trouble with cytology? J Urol
2006;176:2343-6.

165

694. Nieder AM, Soloway MS, Herr HW. Should we abandon


the FISH test? European urology 2007;51:1469-71.

712. Lotan Y, Gupta A, Shariat SF, et al. Lymphovascular


invasion is independently associated with overall survival,
cause-specific survival, and local and distant recurrence in
patients with negative lymph nodes at radical cystectomy.
J Clin Oncol 2005;23:6533-9.

695. Renshaw AA. UroVysion, urine cytology, and the College


of American Pathologists: where should we go from here?
Arch Pathol Lab Med;134:1106-7.

713. Manoharan M, Katkoori D, Kishore TA, Jorda M, Luongo


T, Soloway MS. Lymphovascular invasion in radical
cystectomy specimen: is it an independent prognostic
factor in patients without lymph node metastases? World J
Urol 2010;28:233-7.

696. Chandra A, Griffiths D, McWilliam LJ. Best practice:


gross examination and sampling of surgical specimens
from the urinary bladder. Journal of clinical pathology
2010;63:475-9.
697. Humphrey PA. Urinary bladder pathology 2004: an update.
Annals of diagnostic pathology 2004;8:380-9.

714. Palmieri F, Brunocilla E, Bertaccini A, et al. Prognostic


value of lymphovascular invasion in bladder cancer in
patients treated with radical cystectomy. Anticancer Res
2010;30:2973-6.

698. Amin MB. Protocol for the examination of specimens from


patients with carcinoma of the urinary bladder. In: College
of American Pathologists; 2011.

715. Quek ML, Stein JP, Nichols PW, et al. Prognostic


significance of lymphovascular invasion of bladder cancer
treated with radical cystectomy. J Urol 2005;174:103-6.

699. Lopez-Beltran A, Bassi P, Pavone-Macaluso M, Montironi


R. Handling and pathology reporting of specimens with
carcinoma of the urinary bladder, ureter, and renal pelvis.
European urology 2004;45:257-66.

716. Shariat SF, Karakiewicz PI, Palapattu GS, et al.


Nomograms provide improved accuracy for predicting
survival after radical cystectomy. Clinical cancer research:
an official journal of the American Association for Cancer
Research 2006;12:6663-76.

700. Cina SJ, Epstein JI, Endrizzi JM, Harmon WJ, Seay TM,
Schoenberg MP. Correlation of cystoscopic impression
with histologic diagnosis of biopsy specimens of the
bladder. Human pathology 2001;32:630-7.

717. Shariat SF, Svatek RS, Tilki D, et al. International validation


of the prognostic value of lymphovascular invasion in
patients treated with radical cystectomy. BJU international
2010;105:1402-12.

701. Smith JA, Jr., Labasky RF, Cockett AT, Fracchia JA, Montie
JE, Rowland RG. Bladder cancer clinical guidelines
panel summary report on the management of nonmuscle
invasive bladder cancer (stages Ta, T1 and TIS). The
American Urological Association. J Urol 1999;162:1697701.

718. Bolenz C, Herrmann E, Bastian PJ, et al. Lymphovascular


invasion is an independent predictor of oncological
outcomes in patients with lymph node-negative urothelial
bladder cancer treated by radical cystectomy: a multicentre
validation trial. BJU international 2010;106:493-9.

702. Chalasani V, Chin JL, Izawa JI. Histologic variants of


urothelial bladder cancer and nonurothelial histology in
bladder cancer. Can Urol Assoc J 2009;3:S193-8.

719. Hong B, Park S, Hong JH, Kim CS, Ro JY, Ahn H. Prognostic
value of lymphovascular invasion in transitional cell
carcinoma of upper urinary tract. Urology 2005;65:692-6.

703. Billis A, Schenka AA, Ramos CC, Carneiro LT, Araujo V.


Squamous and/or glandular differentiation in urothelial
carcinoma: prevalence and significance in transurethral
resections of the bladder. International urology and
nephrology 2001;33:631-3.

720. Hong SK, Kwak C, Jeon HG, Lee E, Lee SE. Do vascular,
lymphatic, and perineural invasion have prognostic
implications for bladder cancer after radical cystectomy?
Urology 2005;65:697-702.

704. Wasco MJ, Daignault S, Zhang Y, et al. Urothelial


carcinoma with divergent histologic differentiation (mixed
histologic features) predicts the presence of locally
advanced bladder cancer when detected at transurethral
resection. Urology 2007;70:69-74.

721. Kunju LP, You L, Zhang Y, Daignault S, Montie JE, Lee


CT. Lymphovascular invasion of urothelial cancer in
matched transurethral bladder tumor resection and
radical cystectomy specimens. J Urol 2008;180:1928-32;
discussion 32.

705. Jozwicki W, Domaniewski J, Skok Z, Wolski Z,


Domanowska E, Jozwicka G. Usefulness of histologic
homogeneity estimation of muscle-invasive urinary
bladder cancer in an individual prognosis: a mapping
study. Urology 2005;66:1122-6.
706. Margulis V, Lotan Y, Montorsi F, Shariat SF. Predicting
survival after radical cystectomy for bladder cancer. BJU
international 2008;102:15-22.
707. Algaba F. Lymphovascular invasion as a prognostic tool
for advanced bladder cancer. Curr Opin Urol 2006;16:36771.
708. Canter D, Guzzo T, Resnick M, et al. The presence of
lymphovascular invasion in radical cystectomy specimens
from patients with urothelial carcinoma portends a poor
clinical prognosis. BJU international 2008;102:952-7.
709. Hara S, Miyake H, Fujisawa M, et al. Prognostic variables
in patients who have undergone radical cystectomy for
transitional cell carcinoma of the bladder. Jpn J Clin Oncol
2001;31:399-402.
710. Harada K, Sakai I, Kurahashi T, et al. Clinicopathological
features of recurrence after radical cystectomy for patients
with transitional cell carcinoma of the bladder. International
urology and nephrology 2006;38:49-55.
711. Karakiewicz PI, Shariat SF, Palapattu GS, et al. Nomogram
for predicting disease recurrence after radical cystectomy
for transitional cell carcinoma of the bladder. J Urol
2006;176:1354-61; discussion 61-2.

722. Abdel-Latif M, Abol-Enein H, El-Baz M, Ghoneim MA.


Nodal involvement in bladder cancer cases treated with
radical cystectomy: incidence and prognosis. J Urol
2004;172:85-9.
723. Streeper NM, Simons CM, Konety BR, et al. The
significance of lymphovascular invasion in transurethral
resection of bladder tumour and cystectomy specimens
on the survival of patients with urothelial bladder cancer.
BJU international 2009;103:475-9.
724. Herrmann E, Stoter E, van Ophoven A, et al. The
prognostic impact of pelvic lymph node metastasis and
lymphovascular invasion on bladder cancer. International
journal of urology : official journal of the Japanese
Urological Association 2008;15:607-11.
725. Turkolmez K, Tokgoz H, Resorlu B, Kose K, Beduk Y.
Muscle-invasive bladder cancer: predictive factors and
prognostic difference between primary and progressive
tumors. Urology 2007;70:477-81.
726. Bassi P, Ferrante GD, Piazza N, et al. Prognostic factors
of outcome after radical cystectomy for bladder cancer: a
retrospective study of a homogeneous patient cohort. J
Urol 1999;161:1494-7.
727. Furukawa J, Miyake H, Hara I, Takenaka A, Fujisawa
M. Clinical outcome of radical cystectomy for patients

166

with pT4 bladder cancer. International journal of urology:


official journal of the Japanese Urological Association
2008;15:58-61.

invasion demonstrate that it is the predominant method


of vascular invasion in breast cancer and has important
clinical consequences. The American journal of surgical
pathology 2007;31:1825-33.

728. Harada K, Sakai I, Hara I, Eto H, Miyake H. Prognostic


significance of vascular invasion in patients with bladder
cancer who underwent radical cystectomy. International
journal of urology : official journal of the Japanese
Urological Association 2005;12:250-5.

744. Harris EI, Lewin DN, Wang HL, et al. Lymphovascular


invasion in colorectal cancer: an interobserver variability
study. The American journal of surgical pathology
2008;32:1816-21.

729. Cheng L, Neumann RM, Scherer BG, et al. Tumor size


predicts the survival of patients with pathologic stage T2
bladder carcinoma: a critical evaluation of the depth of
muscle invasion. Cancer 1999;85:2638-47.

745. Alexander-Sefre F, Singh N, Ayhan A, Salveson


HB, Wilbanks G, Jacobs IJ. Detection of tumour
lymphovascular space invasion using dual cytokeratin and
CD31 immunohistochemistry. Journal of clinical pathology
2003;56:786-8.

730. Ennis RD, Petrylak DP, Singh P, et al. The effect of


cystectomy, and perioperative methotrexate, vinblastine,
doxorubicin and cisplatin chemotherapy on the risk and
pattern of relapse in patients with muscle invasive bladder
cancer. J Urol 2000;163:1413-8.

746. Kim JH, Park SS, Park SH, et al. Clinical significance of
immunohistochemically-identified lymphatic and/or blood
vessel tumor invasion in gastric cancer. J Surg Res
2010;162:177-83.
747. Kurtz KA, Hoffman HT, Zimmerman MB, Robinson RA.
Perineural and vascular invasion in oral cavity squamous
carcinoma: increased incidence on re-review of slides and
by using immunohistochemical enhancement. Arch Pathol
Lab Med 2005;129:354-9.

731. Zhang M, Tao R, Zhang C, Shen Z. Lymphovascular


invasion and the presence of more than three tumors are
associated with poor outcomes of muscle-invasive bladder
cancer after bladder-conserving therapies. Urology
2010;76:902-7.
732. Andius P, Johansson SL, Holmang S. Prognostic factors in
stage T1 bladder cancer: tumor pattern (solid or papillary)
and vascular invasion more important than depth of
invasion. Urology 2007;70:758-62.

748. ODonnell RK, Feldman M, Mick R, Muschel RJ.


Immunohistochemical method identifies lymphovascular
invasion in a majority of oral squamous cell carcinomas
and discriminates between blood and lymphatic vessel
invasion. J Histochem Cytochem 2008;56:803-10.

733. Cho KS, Seo HK, Joung JY, et al. Lymphovascular


invasion in transurethral resection specimens as
predictor of progression and metastasis in patients with
newly diagnosed T1 bladder urothelial cancer. J Urol
2009;182:2625-30.

749. Larsen MP, Steinberg GD, Brendler CB, Epstein JI. Use
of Ulex europaeus agglutinin I (UEAI) to distinguish
vascular and pseudovascular invasion in transitional cell
carcinoma of bladder with lamina propria invasion. Mod
Pathol 1990;3:83-8.

734. Lopez JI, Angulo JC. The prognostic significance of vascular


invasion in stage T1 bladder cancer. Histopathology
1995;27:27-33.

750. Ramani P, Birch BR, Harland SJ, Parkinson MC. Evaluation


of endothelial markers in detecting blood and lymphatic
channel invasion in pT1 transitional carcinoma of bladder.
Histopathology 1991;19:551-4.

735. Park J, Song C, Hong JH, et al. Prognostic significance of


non-papillary tumor morphology as a predictor of cancer
progression and survival in patients with primary T1G3
bladder cancer. World J Urol 2009;27:277-83.

751. Afonso J, Santos LL, Amaro T, Lobo F, Longatto-Filho A.


The aggressiveness of urothelial carcinoma depends to a
large extent on lymphovascular invasion--the prognostic
contribution of related molecular markers. Histopathology
2009;55:514-24.

736. Horikawa Y, Kumazawa T, Narita S, et al. Lymphatic


invasion is a prognostic factor for bladder cancer treated
with radical cystectomy. Int J Clin Oncol 2007;12:131-6.

752. Ma Y, Hou Y, Liu B, Li X, Yang S, Ma J. Intratumoral


lymphatics and lymphatic vessel invasion detected
by D2-40 are essential for lymph node metastasis in
bladder transitional cell carcinoma. Anat Rec (Hoboken)
2010;293:1847-54.

737. Leissner J, Koeppen C, Wolf HK. Prognostic significance


of vascular and perineural invasion in urothelial
bladder cancer treated with radical cystectomy. J Urol
2003;169:955-60.
738. Reuter VE. Lymphovascular invasion as an independent
predictor of recurrence and survival in node-negative
bladder cancer remains to be proven. J Clin Oncol
2005;23:6450-1.

753. Kassouf W, Leibovici D, Munsell MF, Dinney CP, Grossman


HB, Kamat AM. Evaluation of the relevance of lymph node
density in a contemporary series of patients undergoing
radical cystectomy. J Urol 2006;176:53-7; discussion 7.

739. Lapham RL, Ro JY, Staerkel GA, Ayala AG. Pathology of


transitional cell carcinoma of the bladder and its clinical
implications. Semin Surg Oncol 1997;13:307-18.

754. Herr HW, Faulkner JR, Grossman HB, Crawford ED.


Pathologic evaluation of radical cystectomy specimens: a
cooperative group report. Cancer 2004;100:2470-5.

740. Acs G, Dumoff KL, Solin LJ, Pasha T, Xu X, Zhang PJ.


Extensive retraction artifact correlates with lymphatic
invasion and nodal metastasis and predicts poor outcome
in early stage breast carcinoma. The American journal of
surgical pathology 2007;31:129-40.

755. Fang AC, Ahmad AE, Whitson JM, Ferrell LD, Carroll PR,
Konety BR. Effect of a minimum lymph node policy in
radical cystectomy and pelvic lymphadenectomy on lymph
node yields, lymph node positivity rates, lymph node
density, and survivorship in patients with bladder cancer.
Cancer 2010;116:1901-8.

741. Auguste P, Lemiere S, Larrieu-Lahargue F, Bikfalvi A.


Molecular mechanisms of tumor vascularization. Crit Rev
Oncol Hematol 2005;54:53-61.

756. Stephenson AJ, Gong MC, Campbell SC, Fergany AF,


Hansel DE. Aggregate lymph node metastasis diameter
and survival after radical cystectomy for invasive bladder
cancer. Urology 2010;75:382-6.

742. Marinho VF, Metze K, Sanches FS, Rocha GF, Gobbi H.


Lymph vascular invasion in invasive mammary carcinomas
identified by the endothelial lymphatic marker D2-40 is
associated with other indicators of poor prognosis. BMC
Cancer 2008;8:64.

757. Fleischmann A, Thalmann GN, Markwalder R, Studer UE.


Extracapsular extension of pelvic lymph node metastases
from urothelial carcinoma of the bladder is an independent
prognostic factor. J Clin Oncol 2005;23:2358-65.

743. Mohammed RA, Martin SG, Gill MS, Green AR, Paish EC,
Ellis IO. Improved methods of detection of lymphovascular

167

758. Cheville JC, Dundore PA, Bostwick DG, et al. Transitional


cell carcinoma of the prostate: clinicopathologic study of
50 cases. Cancer 1998;82:703-7.
759. Shen SS, Lerner SP, Muezzinoglu B, Truong LD, Amiel
G, Wheeler TM. Prostatic involvement by transitional
cell carcinoma in patients with bladder cancer and its
prognostic significance. Human pathology 2006;37:72634.
760. Oliva IV, Smith SL, Chen Z, Osunkoya AO. Urothelial
carcinoma of the bladder with transmural and direct
prostatic stromal invasion: does extent of stromal invasion
significantly impact patient outcome? Human pathology
2011;42:51-6.
761. Algaba F, Arce Y, Lopez-Beltran A, Montironi R, Mikuz
G, Bono AV. Intraoperative frozen section diagnosis in
urological oncology. European urology 2005;47:129-36.

168

Committee 3 A

Molecular Markers for Bladder


Cancer Screening, Early
Diagnosis, and Surveillance
Chairs:
Shahrokh F. Shariat,
Pierre I. Karakiewicz

Members:
Bernd J. Schmitz-Drger, Michael Droller, Yair Lotan,
Vinata B. Lokeshwar, Bas W. van Rhijn, George P. Hemstreet,
PerUno Malmstrom, Yves Fradet, MLiss Hudson, Osamu Ogawa,
Michael J. Marberger

*Address all correspondence to:


S.F. SHARIAT
Weill Medical College of Cornell University
New York Presbyterian Hospital,
525 East 68th Street
New York, NY 10021
Email: sfshariat@gmail.com
Telephone: 469 363 8500

169

Table of Contents
I. Introduction

III. Marker performance


1. Diagnosis and surveillance

1. Why did we fail in the past?

2. Screening

2. Potential indications for marker


use

3. Problems in marker comparison


4. Recommendations
5. Marker performance

II. Materials and Methods

6. Problems in the assessment of


marker trials
7. Key questions

1. Data collection
2. Criteria for assessment of
reporting, marker status and LoE

IV. Appendix

Abstract:
Due to the lack of disease-specific symptoms diagnosis and follow-up of bladder cancer has remained a
challenge to the urologic community. Cystoscopy, commonly accepted as a gold standard for the detection of
bladder cancer, is invasive and relatively expensive while urine cytology is of limited value specifically in low
grade disease. Through the last decades numerous molecular assays for the diagnosis of urothelial cancer
have been developed and were investigated with regard to their clinical use. However, although all of these
assays have been shown to have superior sensitivity as compared to urine cytology, none of them has been
included in clinical guidelines. The key reason for this situation is that none of the assays has been included
into clinical decision making so far. We reviewed the current status and performance of modern molecular
urine tests following systematic analysis of the value and limitations of commercially available assays. Despite
considerable advances through recent years the authors feel that at this stage the added value of molecular
markers for the diagnosis of urothelial tumors has not yet been identified. Current data suggest that some
of these markers may have the potential to play a role in screening and surveillance of bladder cancer. Well
designed protocols and prospective, controlled trials will be needed to provide the basis to determine whether
integration of molecular markers into clinical decision-making will be of value in the future.

170

Molecular Markers for Bladder


Cancer Screening, Early Diagnosis,
and Surveillance
Bernd J. Schmitz-Drger, Michael Droller, Yair Lotan,
Vinata B. Lokeshwar, Bas W. van Rhijn, George P. Hemstreet,
PerUno Malmstrom, Yves Fradet, MLiss Hudson, Osamu Ogawa,
Michael J. Marberger, Pierre I. Karakiewicz, Shahrokh F. Shariat
such as fluorescence or narrow-band imaging are
emerging, the invasiveness and added costs of these
procedures underscores the need for better, simpler
and cheaper diagnostic tests in the management of
bladder cancer patients.[5-7]

I. Introduction
Bladder cancer represents an important cause of
cancer morbidity and mortality. In 2010, once again,
it was the second most common genitourinary
cancer in the United States, with a projected 70,530
new cases and 14,680 deaths.[1] Currently, it is
estimated that there are more than 1,000,000 men
and women alive in the United States and Europe
who have a history of bladder cancer. At the time of
initial diagnosis, approximately 70% of patients have
cancers confined to the epithelium or the subepithelial
connective tissue. These cancers usually are
managed with endoscopic resection and the selective
use of intravesical therapy. The recurrence rate for
these tumors ranges from 50% to 70%, and between
10% and 15% progress to muscle-invasive disease
over 5 years.[2,3] Disease recurrence may occur in
the bladder or in the upper urinary tract even after
several years, necessitating life-long surveillance.
Management of this population is costly because of
the extended surveillance and need for repeated use
of endoscopic and intravesical therapies. Patients
with muscle-invasive and metastatic disease have
a much more precarious survival outcome and also
contribute greatly to the cost of bladder cancer care
because of the expense of radical cystectomy and
systemic chemotherapy. As a result, bladder cancer
is the most expensive cancer to treat on a per patient
basis in the United States.[4]

Voided urine cytology is a highly specific,


noninvasive adjunct to cystoscopy. It has good
sensitivity for detecting high-grade urothelial cancer,
but sensitivity for detection of low-grade tumors
ranges from only 4% to 31%.[8] Furthermore the
accuracy of cytology is dependent upon the level of
expertise of the pathologist and, in consequence,
not readily available in all places. Thus, specifically
in the surveillance of papillary low grade tumors a
noninvasive, highly sensitive and specific bladder
cancer marker could decrease the frequency of
cystoscopies, thereby improving patient quality of
life, and potentially decrease costs by substituting
a less expensive, noninvasive test for the more
expensive endoscopic procedure. In high grade
disease increased sensitivity of markers might lead
to earlier detection of tumor recurrence and, in
consequence improve patient survival.
The potential to decrease costs associated with
cystoscopy would depend upon the cost of the noninvasive biomarker and its specificity such that a false
positive result and the unnecessary need for further
evaluation and associated patient anxiety would not
be a frequent occurrence. Correspondingly, the use
of urinary cytology in monitoring for persistence or
recurrence of high grade disease would also allow
for a decrease in the frequency of cystoscopies if
its specificity were sufficiently high and pathologists
were adequately trained so that they could deliver
a reliable negative reading. Use of other urinary
markers would depend upon their having a similarly
high specificity for high risk disease in this context.

Due to the lack of disease-specific symptoms


diagnosis and follow-up of bladder cancer has
remained a challenge to the urologic community.
Cystoscopy, commonly accepted as a gold standard
for the detection of bladder cancer, is invasive and
relatively expensive thus limiting the frequency of
its use. Although new cystoscopic technologies

The requirements for an ideal marker have been

171

defined using the terms easier, better, faster,


cheaper.[9] Easier in this definition refers to the
assays analytical performance and robustness. For
an assay to be clinically applicable, it should be able
to be performed easily and promptly in a clinical
environment. Better is by the far the most important
challenge that has to be addressed. Demonstrating
information equal to current clinically available
variables is not enough. Any newly discovered
marker should provide additional information that
is helpful to the clinician for the management of the
disease thus providing an added value to the current
situation. Faster means that a new marker should
be able to make the information available in an
efficient and timely manner. Even if the marker has
been proven to offer valuable information regarding
the disease, an unreasonable period of time for its
delivery could considerably decrease its practical
utility. Cheaper is essential for a marker to be costeffective. With health care expenditures reaching
record levels, medical decision making is increasingly
affected by economic concerns. Nevertheless, many
parameters must be considered when assessing
economic impact of a marker: in addition to the
mere costs of the assay, potential clinical benefits
(avoidance of further diagnostic interventions or
ineffective therapy, or benefit from targeted therapy)
need to be considered.

prostate cancer screening where the diagnosis is


usually being sought in asymptomatic individuals
who may themselves request a screening test.
It seems obvious that new tests for the initial
diagnosis of bladder cancer should be investigated in
patients with symptoms and/or signs associated with
this disease. This will pertain largely to patients who
have gross hematuria, those who may have irritative
voiding symptoms without urinary tract infection, and
those found on routine urinalysis to have microscopic
hematuria. However, an investigation of the literature
shows that this approach is often neglected. In
contrast, the vast majority of studies are case-control
trials comparing artificially composed study cohorts,
in which the prevalence of the disease frequently
exceeds 50%. High disease prevalence is usually not
seen in urological practice and such an evaluation
is likely to result in an optimistic assessment of the
positive predictive value. Therefore, this study design
may be misleading when the findings are translated
into routine clinical use.
In conclusion, an insufficient evaluation process is
one of the reasons for the lack of incorporation of
modern bladder cancer tests into clinical decision
making. We still lack good clinical practice
guidelines for the evaluation of diagnostic markers.
The different phases for development and validation
of diagnostic markers in clinical practise have been
defined.[12] However, these four phases, defined
in analogy to the classification used for therapeutic
trials, still provide only a framework for the detailed
assessment of a new diagnostic marker.[7,13]

A significant amount of laboratory and clinical


investigations have developed numerous new urine
markers for the diagnosis of bladder cancer. Many
of them exhibit sensitivity considerably superior to
that of standard urine cytology, particularly in low to
moderate grade diatheses, and are frequently used.
However, despite FDA approval of some of these
assays, none of them has achieved acceptance as a
standard diagnostic procedure in clinical guidelines.
[10,11]

2. Potential indications for marker


use
The following putative indications for the use
of diagnostic bladder cancer markers can be
delineated:

In this report we try to delineate the reasons for


this situation. Furthermore, we will focus in this
manuscript on characterizing any added value and
corresponding clinical utility of modern diagnostic
markers. To this end, we define added value as
either improvement of diagnostic accuracy, reduction
(not necessarily replacement) of other diagnostic
(occasionally invasive) measures, improvement
in quality of life, and/or reduction of costs. Each of
these will be assessed in the context of requirement
for an ideal marker as described above.

Screening
Voiding symptoms, hematuria, risk populations
(occupational exposure/life style)
Reflex testing
Follow-up of patients with bladder cancer

a) Screening
Bladder cancer screening could be an indication for
the use of a non-invasive diagnostic test. Although
the mortality/incidence ratio is higher for bladder
than prostate cancer, the low prevalence of bladder
cancer in the general population along with the low
mortality from bladder cancer due to a large number
of cases with non-fatal tumors has been an obstacle
to develop effective screening strategies for bladder
cancer. Nevertheless, data from a few screening
trials and theoretical considerations on costeffectiveness issues recently have revitalized this

1. Why did we fail in the past?


Although non-invasive tests are labelled to diagnose
bladder cancer, it remains unclear how they can
effectively be integrated into clinical decision
making, particularly when making an initial diagnosis
because the presenting signs and symptoms may
be caused by a number of different diseases and
conditions. This situation is different from that in

172

discussion.[14] Screening of well-defined high risk


populations with a disease prevalence comparable
to other tumor entities that have been accepted for
screening (e.g. breast cancer or colorectal cancer)
may offer a solution to this problem.[15]

Since hematuria assessment is not a screening


approach in its original sense the respective studies
were grouped below among the diagnostic studies.

1. Voiding symptoms

Exposure to specific chemical carcinogens is a wellestablished risk factor for urothelial cancer. Exposure
to aromatic amines, in chemical industry workplaces
involving use of these chemical substances and in
the rubber industry, is associated with bladder cancer.
An excess risk was also reported for dyers in textile
industries, painters, varnishers and hairdressers.
Other agents having been associated with UBC are
polycyclic aromatic hydrocarbons (PAHs), used in
aluminum production, coal gasification, coal tars,
roofing and carbon black manufacture [25]

3. Risk populations

Irritative voiding symptoms may occur in patients


with bladder cancer. This may occur in association
with gross hematuria in which the bleeding itself may
produce irritation, or the involvement of the trigone,
bladder neck, or prostatic urethra by carcinoma in
situ. However, the prevalence of bladder cancer
in patients with irritative voiding symptoms barely
exceeds that of age-matched controls because of so
many other conditions that cause these symptoms.
For example the close association of voiding
symptoms due to prostate enlargement common
in the male population with bladder cancer is a
significant confounding factor and might prevent
adequate patient selection. An effective selection of
patients at risk of having bladder cancer based upon
any degree of voiding symptoms appears impossible.
In consequence, despite a good correlation between
irritative voiding symptoms and bladder cancer,
this condition alone is currently not suited for the
identification of a patient cohort that should undergo
further assessment for bladder cancer.

While these professions are rarer today and safety


measures reducing exposure were introduced at
least in Western plants an excess risk of bladder
cancer still exists among workers exposed to diesel
engine exhaust, such as drivers. In a meta-analysis,
the relative risk among truck-drivers was 1.17
(95%CI 1.06-1.29) and among bus drivers, it was
1.33 (95%CI 1.22-1.45) [26]
Today, cigarette smoking is the most important
risk factor for UBC, accounting for 50% of cases in
men and 35% in women. A meta-analysis reported
that current cigarette smokers have a risk of 2.57
(95%CI 2.20-3.00) compared with non-smokers.[27]
A positive dose-response relationship is found with
both number of cigarettes smoked daily and number
of years of smoking.

2. Hematuria
We do have increased bladder cancer prevalence
rates in gross hematuria justifying a complete
clinical work-up of these patients.[15-19] This is in
contrast to microhematuria, a frequent condition in
the general population. Although the prevalence
of bladder cancer [20,21] is lower in patients with
microscopic hematuria, complete urological workup remains a matter of discussion.[22] This dilemma
has resulted in the discrimination between high risk
and low risk populations with a focus of diagnostic
efforts on those patients at higher risk. However,
apart from bladder cancer, there may be other
conditions correlated with hematuria that may
require urological intervention. However, information
on these additional conditions is rare. Finally, there is
concern that primary care physicians do not always
refer patients with hematuria for evaluation thus
introducing a further selection bias.[23,24]

Screening for bladder cancer requires definition


of cohorts with adequate disease prevalence.
The prevalence data provided above suggest
that professional exposure alone may not justify
screening; however, professional exposure in
combination with other risk factors may yield
populations with disease prevalence suitable for a
screening approach. However, this will be further
modulated by performance characteristics of the
marker used and cost/benefit considerations.

b) Reflex testing
Use of molecular markers for so-called Reflex
testing has gained some interest. The idea behind
this strategy is to improve the accuracy of a previous
test (mostly cytology) but also to minimize expenses
for molecular assays. In most cases bladder cancer
patients with a negative cytology test subsequently
undergo reflex testing with a more sensitive assay.
This procedure makes use of the high specificity of
urine cytology on the one hand and aims at improving
sensitivity of non-invasive diagnosis. Several studies
on reflex testing have been published using the
UroVysion assay.[28,29] As with most findings, it is
important to validate the clinical utility of markers. One
study prospectively validated the role of Urovysion in

The current pathways for the assessment of


patients with hematuria have disadvantages. While
endoscopy remains invasive and costly, it is still
required because of the low sensitivity of urine
cytology. In addition, the sensitivity of imaging for the
detection of upper urinary tract tumors is insufficient;
further confounding the problem of reaching an
accurate diagnosis of urothelial cancer. As a result,
assessment of patients with hematuria could be an
area where new diagnostic markers could be clinical
helpful.

173

patients with atypical cytology and noted cystoscopic


findings had an important effect on the performance
of the marker.[30] Nevertheless, any added value
of this approach in the context of what we have
described above requires validation.

2. C
 riteria for assessment of
reporting, marker status and LoE

c) Follow-up

T
 he Level of Evidence (LoE) for diagnostic
procedures (Oxford classification) [36].

It was a specific challenge of this assessment to


classify the different markers and trials according to

Surveillance of patients with a history of bladder


cancer is a key area for the use of new diagnostic
markers, because the prevalence of the disease is
high in this group and new urinary tests will therefore
have a better positive predictive value than urine
cytology. These tests can detect bladder cancer
before they are visually evident. [28,31] However,
this causes a significant problem in defining negative
tests. Currently, there is no easy way of separating
false positive tests from true positive tests when
patients do not have clinically evident tumor.

T
 he accuracy of data reporting according to the
STARD criteria [37,38] .
T
 he status of the marker with regard to clinical
implementation (IBCN classification) [12].
By this procedure we intended that readers should
be informed on the quality of marker studies and
the status of a given marker with regard to clinical
implementation in a structured and reproducible
way. Furthermore, use of the respective tools were
to be made in order to

In general, two different directions for a use of urine


tests are conceivable:

encourage readers, reviewers and publishers


to use a structured and transparent process for
manuscript assessment, but also

Surveillance of patients with low risk tumors aimed


at a reduction of the frequency of diagnostic
cystoscopies.

to identify deficiencies of the tools available and


produce proposals for their improvement.

Follow-up of patients with high risk tumors with


the intention to recognize tumor recurrence and
progression as early as possible.

All statements and recommendations were


discussed within the group. Recommendations were
provided and categorized according to the criteria
of the Agency for Health Care Policy and Research
(AHCPR) [39] and required consensus of the group.

Some studies suggest that a use of non-invasive


diagnostic markers in follow-up of bladder cancer
may be helpful; [8,32-34] however, prospective
analyses to define the consequences from a negative
or positive test result are still lacking.

III. Marker performance


In this part of the assessment, information on the
performance of commercially available molecular
diagnostic markers is provided. This information
is based upon a critical evaluation of the currently
available literature.

II. Materials and methods


1. Data collection

As of March 2010, six urinary tumor marker tests


have been cleared by the U.S. Food and Drug
Administration (FDA) and are in clinical use. These
tests are:

This review was restricted to commercially available


assays (Tab. 1). The assessment was based upon
a systematic literature search in medical databases
(PubMed). All studies on the diagnostic use of the
respective markers were screened and reviews
as well as repeated publications of the same data
were identified and excluded. Studies investigating
prognosis were not considered for this analysis but
are part of a second paper. [35] For some markers,
well-executed meta-analyses were used as a basis
for assessment [33,34] while for other markers
detailed analysis of studies that had been published
in English through January 2011 was performed.
Sensitivity was assessed based upon histopathologic
results only. Studies on non-urothelial tumors or
trials not comprising information required for a basic
assessment (e.g. stage, grade) were excluded. If
deemed necessary additional publications in other
languages were considered.

T
 he quantitative BTA TRAK and the qualitative
point-of-care BTA (bladder tumor antigen) stat
test (both by Polymedco Inc., Cortlandt Manor, NY,
USA).
T
 he quantitative immunoassay NMP22 and
the qualitative, point-of-care test NMP22
BladderChek (Matritech Inc.,Newton, MA, USA).
T
 he UroVysion Bladder Cancer Kit (Vysis Inc.,
Downers Grove, IL, USA), a multiple marker
fluorescence in situ hybridization (FISH) test.
T
 he uCytTM test, an immunocytochemical assay
(Scimedx Inc., Denville, NJ, USA).
With the exception of the uCyt test, which is only

174

Table 1. Commercially available bladder tumor markers (basic information)


Test/
marker
Cytology

Marker
detected/ Specimen
Marker type

Tumor cells

Hematuria A: Hemoglobin
detection
B: Red blood cells

Assay type

F
D
A Manufacturer
approval

Voided urine Microscopy


Barbotage
specimen
Exfoliated cells
A:
Voided A: Dipstick
urine

n.a.

B:
urine

Voided B: Interference-contrast
microscopy or
Red blood cell analyzer
BTA-Stat
Complement
factor Voided urine
Dipstick immunoassay Diagnosis,
H-related protein (and
follow-up
also
Complement
factor H)
BTA-TRAK Complement
factor Voided urine
Sandwich ELISA
Diagnosis,
H-related protein (and
follow-up
also
Complement
factor H)
NMP-22
Nuclear
mitotic Voided urine
Sandwich ELISA
follow-up
apparatus protein

A: Bayer Corp.
B: -

B a r d / B i o n
Diagnostics

B a r d / B i o n
Diagnostics

Matritech, Inc.

NMP-22

Nuclear
mitotic Voided urine
apparatus protein

Point-of-care device

BLCA-4

Nuclear matrix protein

Voided urine

ELISA (using a rabbit polyclonal antibody)

E i c h r o m
Technologies

Survivin

A member of inhibitors Voided urine


of apoptosis gene
family

F u j i r e b i o
Diagnostics, Inc.

UBC

Cytokeratin 8 and 18 Voided urine


(cytoskeletal proteins)

Bio-dot test
(dot blot assay using
a
rabbit
polyclonal
antibody
Sandwich ELISA or a point of care test

C Y F R A Cytokeratin
19
(a Voided urine
21-1
cytoskeletal protein)

Bio
International;
Roche Diagnostics

DD23

Immunoradiometric assay
or
electrochemiluminescent immuno assay
185-kDa
tumor Exfoliated cells Immunocytochemistry
associated Antigen

UroCor
Labs,
Oklahoma City, OK,
USA

Carcinoembryonic
antigen, 2 bladder
tumor cell-associated
mucins
Alterations
in
chromosomes 3, 7, 17
and 9p21

Scimedx, Inc.

uCyt+

UroVysion

Voided urine, Immunocytochemistry


Exfoliated cells

Voided urine, Multi-colored,


Exfoliated cells probe FISH

175

Diagnosis Matritech, Inc.


high risk,
follow-up

Follow-up

IDL Biotech.

multi- Diagnosis, Abbott, Vysis


follow-up

cleared for monitoring bladder cancer recurrence,


all tests are FDA-approved as adjunctive tests for
use in the initial diagnosis of bladder cancer and
surveillance of bladder cancer patients, in conjunction
with standard procedures.

tests. A negative predictive value represents the


percent of negative tests in patients without disease
versus all negative tests. Tests with high positive
predictive value will accurately identify patients with
disease by a positive test result. Tests with high
negative predictive value will accurately identify
patients without disease by negative test results.
Both depend on the prevalence of disease in the
test population because a greater proportion of test
results will then more accurately reflect the true
status of disease in that population. Thus, according
to the contingency table the resultant numerator will
be greater and the denominator less, each leading to
a higher predictive value.

The performance of biomarkers depends upon their


sensitivity (positivity of a marker in the presence
of disease), specificity (negativity of a marker in
the absence of disease), positive predictive value
(probability of disease if a marker is positive), and
negative predictive value (probability of no disease
if a marker is negative). A threshold can be set for
interpreting a test result as positive or negative, this
in turn influencing a markers sensitivity or specificity
(Fig. 1). A markers predictive value will be influenced
by the prevalence of a condition in a test population,
thereby affecting calculations of the probability
of the presence or absence of the disease in that
population on the basis of a positive or negative test
result (Fig. 2).

Of relevance to each of the biomarkers discussed


in this survey is the concern that decisions may be
based on an arbitrary threshold. This will dichotomize
a test that biologically is actually a continuous
variable. Thus, although thresholds may be set to
determine the likelihood of detecting true versus
false positives and true versus false negatives, this
may be misleading in interpreting test results and
their use. This can be important in both low risk and
high risk disease in influencing how a marker may
be applied in screening, surveillance, or determining
efficacy of treatment.

In applying these contingency tables, sensitivity


represents the performance of a test in patients
withd isease (true positives versus all patients
with disease) (Fig. 1). Specificity represents the
performance of a test in patients without disease
(true negatives versus all patients without disease).
Tests with high sensitivity will not overlook patients
with disease. This may allow earlier diagnosis and/
or prevent a delay in diagnosis. Tests with high
specificity may not subject patients who do not have
disease to unnecessary procedures.

1. Diagnosis and surveillance


Performance characteristics may be obtained
by assessment of trials claiming to investigate a
diagnostic use of non-invasive molecular markers.
These trials are inhomogeneous since they are
composed from case control studies with a high
prevalence of cases and from trials targeting
frequently poorly characterized cohorts (usually
designed as cases suspicious for bladder cancer,
hematuria in some cases) with a lower prevalence
of bladder cancer. Therefore, the reported range of
sensitivity is usually wider as compared to that in
follow up studies.

In bladder cancer, the value of biomarkers according


to their sensitivity and/or specificity will depend
upon the particular objectives in screening versus
monitoring in the context of low risk versus high risk
disease and the definition of risk as the disposition to
recurrence versus progression.
Thresholds can be set to determine the likelihood of
detecting true versus false positive and true versus
false negative test results. The resultant increased
or decreased sensitivities or specificities can
determine the usefulness of a biomarker meeting
particular objectives in screening versus monitoring
for disease recurrence. Accordingly, high thresholds
will increase specificity while decreasing sensitivity
because of fewer false positives and more false
negatives (Fig. 1). Correspondingly, low thresholds
will increase sensitivity while decreasing specificity
because of fewer false negatives and more false
positives.

Few trials may be classified as true screening trials


investigating predefined cohorts of asymptomatic
individuals (e.g. smokers, professionally exposed
individuals, cohorts randomly invited for screening)
rendering marker positive individuals for urological
evaluation. These studies are addressed separately.

a) Urine-based Markers
1. NMP22
Nuclear matrix proteins (NMPs) are part of the
structural framework of the nucleus and provide
support for the nuclear shape. These proteins have
also been attributed roles in DNA replication, in
ribonucleic acid transcription and in the regulation
of gene expression. One member of this family,
nuclear mitotic apparatus protein (NMP22), is much
more prevalent in malignant urothelial cells than in
their normal counterparts. Apoptosis is accompanied

Predictive values in bladder cancer can also be


placed in the context of low risk versus high risk
disease and the role that biomarkers play depending
upon the threshold that is set for positivity or
negativity of a particular test value (Fig. 2). A positive
predictive value represents the percent of positive
tests in patients with disease versus all positive

176

Contingency Analysis of
Sensitivity and Specificity
Disease Status
Absent (-)
Present (+)

Pos (+)

Test Result
Neg (-)

Sensitivity = true positives a / (true positives a + false negatives)


Sensitivity = true negatives d / (true negatives d + false positives)
Figure 1. Contingency analysis of sensitivity and specificity

Contingency Analysis
of Predictive Values
Disease Status
Absent (-)
Present (+)

Pos (+)

Test Result
Neg (-)

Positive Predictive Value = true positives a / (true pos a + false positive)


Negative Predictive Value = true negatives d / (true neg d + false negative)
Figure 2. Contingency analysis of predictive values

177

with a release of NMP22 into the urine, and patients


with bladder cancer have a significantly elevated
concentration of NMP22. Both a laboratory-based
quantitative microplate enzyme immunoassay and
a qualitative point-of-care test (BladderChek Test;
Matritech Inc., Massachusetts, USA) are available
and are FDA-approved for use in bladder cancer
surveillance. The latter is also approved for detection
of bladder cancer in high-risk patients.

significant overlap in studies comparing markers and


cytology for high grade cancer, high stage cancers
and patients with CIS (Tab. 2 and 3). Nevertheless,
in general urinary bladder markers also perform
better in patients with higher stage disease (Tab. 4)
and higher biologic aggressiveness (Tab. 5).
Boman and colleagues evaluated NMP22 (cut-off
> 4 U/ml) sensitivities in 250 patients and found
sensitivities of 46% (41/90), 58% (21/36), 76%
(22/29) in detecting tumors < 10 mm, 1120 mm
and > 21 mm, respectively.[40] Sanchez-Carbayo
and colleagues evaluated 187 patients with NMP22
(cut-off >14.6 U/ml) and reported sensitivities of 83%
(19/23), 81% (34/42) and 93% (38/41) in detecting
tumors < 5 mm, 530 mm and > 30 mm, respectively.
[41]

Sensitivity
There have been several meta-analyses that have
evaluated the sensitivity of commonly used markers
(Tab. 2). When compared with cytology, NMP22 as
well as other markers generally have a significantly
higher sensitivity for detecting bladder cancer. This
improvement in sensitivity is primarily in detection
of low grade and low stage bladder cancers with

The data on the impact of tumor number on sensitivity

Table 2. Marker Sensitivity and Specificity of Cytology and Commercially Available Markers (data from Metaanalyses)
Marker

Median Sensitivity (range) Median Specificity (range) Total Number of Patients

Cytology
Glas (51)

55 (48-62)*

94 (90-96)*

3,444

Lotan (8)

34 (20-53)

99 (83-99)

2,767

Van Rhijn (32)

35 (13-75)

94 (85-100)

5,545

Mowatt (52)

44 (38-51)*

96 (94-98)*

14,260

BTA stat

Glas (51)

70 (66-74)*

75 (64-84)*

1,160

Lotan (8)

71 (57-82)

73 (61-82)

2,534

Van Rhijn (32)

58 (29-74)

73 (56-86)

3,461

67 (60-73)*

78 (72-83)*

2,290

Lotan (8)

73 (47-87)

80 (58-91)

2,413

Van Rhijn (32)

71 (47-100)

73 (55-98)

2,041

Mowatt (52) pooled

68 (62-74)*

79 (74-84)*

10,119

Mowatt (52) BladderChek

65 (50-85)

81 (40-87)

2,426

Van Rhijn (32)

67 (52-100)

75 (62-82)

959

Mowatt (52)

84 (77-91)*

75 (68-83)*

3,041

This assessment

81 (42-100)

75 (62-95)

4,899

NMP22 (assay/bladder check)


Glas (51)

uCyt+/Immunocyt

FISH (Urovysion)

Hajdinjak (31)

72 (69-75)*

83 (82-85)*

2,477

Mowatt (52)

76 (65-84)*

85 (78-92)*

3,101

This assessment

72 (23-100)

80 (40-100)

2,852

*95% CI

178

Table 3. Sensitivity of Cytology and Commercially Available Markers (data from Meta-analyses) Based on
Tumor Grade
Marker

# studies

Grade 1

Grade 2

Grade 3

Lotan (8)

.12 (.04-.31)

.26 (.17-.37)

.64 (.38-.84)

Van Rhijn (32)

0.17

0.34

0.58

Lotan (8)

.47 (.38-.56)

.73 (.59-.83)

.94 (.55-.99)

Van Rhijn (32)

0.45

0.6

0.75

Van Rhijn (32)

0.41

0.53

0.8

Lotan (8)

.61(.35-.81)

.71 (.41-.90)

.79 (.63-.89)

Van Rhijn (32)

0.78

0.9

This assessment

19

0.75

084

0.84

Van Rhijn (32)

0.56

0.78

0.95

This assessment

21

0.53

0.81

0.79

Cytology

BTA stat

NMP22

Immunocyt

FISH (Urovysion)

Table 4. Association of Cytology and Commercially Available Markers (data from Meta-analyses) with Tumor
Stage [Lotan (8)]
Marker

# studies

Ta

T1

>T2

TIS

Cytology

.15 (.09-.25)

.46 (.34-.59)

.55 (.35-.73)

.63 (.29-.87)

BTA stat

.57 (.47-.67)

.82 (.63-.92)

.91 (.74-.97)

.66 (.42-.83)

NMP 22

.60 (.42-.76)

.85 (.27-.97)

.89 (.50-.98)

.73 (.54-.86)

Table 5. Sensitivity of Cytology and Commercially Available Markers (data from Meta-analyses) and
Association with Tumor Aggressiveness [Mowatt (52)]
Less aggressive/lower
risk (pTa, G1, G2)
median% (range)

More aggressive
(pT1, G3, CIS)
median% (range)

CIS median% (range)

Number of Patients
(total)

Cytology

27 (0-93)

69 (0-100)

78 (0-100)

12,566

NMP22

50 (0-86)

83 (0-100)

83 (0-100)

7556

FISH (Urovysion)

65 (32-100)

95 (50-100)

100 (50-100)

2164

Immunocyt

81 (55-90)

90 (67-100)

100 (67-100)

2502

is still controversial. Poulakis et al. evaluated 739


patients using NMP22 (cut-off 8.25 U/ml) and
found sensitivities of 79% (165/208), 90% (83/92)
and 97% (96/99) in patients with one, two to three,
and more than three tumors, respectively.[42] On
the other hand, Sanchez-Carbayo and colleagues
evaluated 187 patients using NMP22 (cut-off 14.6
U/ml) and found sensitivities of 72% (18/25) and 75%
(61/81) in patients with single and multiple tumors,
respectively.[41] This discrepancy may relate to
the level of NMP22 reaching threshold based upon

the amount of apoptotic cell debris (the basis of a


positive test) shed into the urine. Tumor volume may
reflect either size or number of lesions in contributing
to a positive test result.
There is also a possible impact of marker sensitivity
based on whether the marker is used for detection or
surveillance. However, this may be related to the fact
that tumors are larger at diagnosis or have a more
advanced stage than during surveillance. Boman
found that NMP22 has higher sensitivity for new as
compared to recurrent tumors; this appeared due to

179

higher stage and grade at presentation and larger


tumor size.[40]

quality of reporting according to the STARD criteria


is moderate to poor, in part due to the fact that the
majority of trials were conducted earlier.[37,38]
We conclude that the level of evidence of studies
on NMP22 is LoE III and in some studies LoE IIb
according to the Oxford classification.[36] Phase
III IBCN trials are lacking.[12,51] This translates
into a maximal LoE Grade IIa for meta-analyses.
[8,33,34,52,53]

In two large prospective multicenter studies using


the NMP22 BladderChek test, the NMP22 assay was
positive in 44 of 79 patients with cancer (sensitivity
55.7%; 95% confidence interval [CI], 44.1%-66.7%)
in the detection setting and sensitivity was 49.5%
(51/103; 95% CI, 39.5%-59.5%) in the surveillance
setting.[32,43] While the sensitivity for surveillance
was lower, the overlapping confidence intervals
suggest no significant difference in performance of
NMP22 in detection and surveillance setting.

2. BTA stat, BTA trak


Among the non-invasive tests developed to detect
urothelial carcinoma, those derived from basement
membrane fragments found in urine from bladder
cancer patients included a series called BTA assays.
The original BTA test was a colorimetric three-step,
latex-agglutination process that qualitatively detected
bladder tumor-associated analytes (BTA) using
human immunoglobulin G.[54-56] The BTA analytes
were polypeptides ranging from 16 to 65 kDa thought
to be released into the urine by proteases released
from tumor cells during stromal invasion.[54-56]
This test was supplanted by 2 newer versions, the
BTA stat and the BTA TRAK which detect different
protein(s) than the original.[54-57] Extrapolation
of sensitivity or specificity results from studies of
the original BTA test to BTA stat or BTA TRAK are,
therefore, not valid. The original BTA test is no longer
commercially available.

Specificity
The main disadvantage of current markers is their
lower specificity compared with cytology (Table 1).
NMP22 is a protein that localizes with the spindle
poles during mitosis and thus regulates chromatid and
daughter cell separation.[44] There is a substantially
higher level of NMP22 in the urine of patients with
bladder cancer. However, because this protein
is released from dead and dying urothelial cells,
many benign conditions of the urinary tract, such as
stones, infection, inflammation, and hematuria may
carry these proteins as well and cystoscopy can
cause a false-positive reading. In a study of NMP22
and BTA stat in 278 symptomatic patients who
presented to a urology clinic, Sharma et al found that
greater than 80% of the false-positive results were
clinically categorized as benign inflammatory or
infectious conditions, renal or bladder calculi, recent
history of a foreign body in the urinary tract, bowel
interposition segment, another genitourinary cancer
or an instrumented urinary sample.[45] History of
ureteral stents or any bowel interposition segment
had a 100% false-positive rate. Exclusion of all
6 clinical categories improved the specificity and
positive predictive value of NMP22 (95.6%, 87.5%)
and BTA stat (91.5%, 69.7%), and was similar to
urinary cytology.

Both BTA stat and TRAK detect human complement


factor H related protein (hCFHrp) and complement
factor H.[57] hCFHrp is thought to interrupt the
complement cascade and confer a selective growth
advantage to cancer cells by allowing them to evade
the host immune system. The dissociation of C3
activating convertase and degradation of C3b are
thought to be enhanced by complement factor H.[5558] Both tests are non-invasive and approved by the
Food and Drug Administration (FDA) as adjuncts
to cystoscopy in the detection of urothelial cancer,
not as primary diagnostic tools.[59,60] BTA stat is a
qualitative test, while BTA TRAK is quantitative.[1,6]
Both have been performed on fresh, refrigerated,
or frozen urine obtained as voided or catheterized
specimens.[40,42,45,49,55,59,61-74]

One consideration that is often raised is the


possibility that a urine based marker may become
positive prior to visualization of a tumor. This has
been termed an anticipatory positive result. There
are several studies that found a greater likelihood
of recurrence in patients with a positive FISH assay
compared to those with negative assays in the
absence of a visualized tumor.[46-48] This has also
been reported for NMP22 and Immunocyt/uCyt,
albeit in a small number of patients.[16,49,50] Thus
the lack of specificity is the major limitation of these
urine markers. Strategies to manage patients with
a positive marker,[12,51] without cystoscopically
visible tumor is crucial to the future applicability of
markers.

Data quality

The BTA stat is an inexpensive, office-based, singlestep, immunochromatographic assay usually performed on voided fresh or refrigerated urine samples
producing results in 5 minutes with minimal training
of personnel.[57] Five drops of urine are placed into
a well containing a colloidal gold conjugate antibody
and allowed to react. Only if tumor antigen is present does a visible colored line form when the antigen
conjugate complexes are captured by a second antibody. BTA stat has been used in the detection of initial, recurrent, and upper tract urothelial carcinoma
(Tab. 6).[40,42,45,49,62-74]

As for other markers discussed in this assessment,

The BTA TRAK is a sandwich immunoassay method

180

requiring trained laboratory technologists several


hours to complete.[75] In this assay, antihuman
complement factor H related protein monoclonal
antibody coated onto 96-well microtiter plate captures
its target in urine. A second alkaline phosphatase
labeled reporter antibody and a substrate solution
are used to detect hCFHrp. Human complement
factor H is also detected. Comparison to a calibration
curve created from kit standards is used to determine
the amount of hCFHrp present. The cut-off limit
recommended by the manufacturer is 14 units/ml,
where 1 unit is 4.7 ng of hCFHrp (Tab. 6).[58,59,61]

Level 2a evidence as identified by a meta-analysis


of data on BTA Stat and TRAK testing was provided
in these review articles, with the highest number of
subjects reported in the Glas and van Rhijn articles.
[34,52] Each included many of the same source
documents, and thus each was dependent on the
quality of these sources. As described by Glas[52],
the quality of the literature is weak and we concur,
based on our evaluation using the STARD checklist.
[38] No study met all 25 STARD items.
There are several clinical scenarios in which either
of the BTA tests could prove useful. The first is as a
diagnostic tool for the detection of primary urothelial
carcinoma in subjects with signs and symptoms of
bladder cancer or at screening of risk populations.
The FDA has not approved either BTA stat or TRAK
for this indication.[6] Tumor markers in the diagnosis
of primary bladder cancer. A systematic review by
Glas would appear to address this situation.[52] In
this meta-analysis, of 1160 subjects providing urine
samples for BTA stat testing, the sensitivity was 70%
(95% CI: 66-74I) and specificity 75% (95% CI: 64-

Using a PubMed search for BTA, we identified 7


review articles in English on bladder tumor markers
in use that included BTA stat or TRAK testing.
[34,52,54,56,58-60] Sample source documents
from these were selected based on frequency of
citation and to include global urologic participants.
With the exception of those studies performed on
archived urine specimens from prior studies,[55,61]
the majority of studies discussed are International
Bladder Cancer Network (IBCN) Phase II studies.

Table 6. BARD stat and TRAK assay: Patient demographics


Author

Gender

Sarosdy 1997 54

M
443
M
141
Partial record
Partial record

F
332
F
77

M
151
M
145
M
485
NS
M
199
NS
M
127
M
172
M
207
M
152
NS
M
208
M
80

F
28
F
23
F
254

Babjuk 2002 70
Ellis 1997 60
Tsui 2007 72
Serretta 2000 68
Giannopoulos 2000 47
Poukalis 2001 40
Nasuti 1998 66
Wiener 1998 61
Sharma 1999 43
Sozen 1999 64
Leyh 1999 62
Pode 1999 63
Ramakumar 1999 65
Raitanen 2001 69
Boman 2002 38
Schroeder 2004 71

F
92
F
13
F
68
F
43
F
44
F
96
F
35

NS = Not stated

181

Race

# of Sites

Urine sample
type
Voided Archived
frozen
Voided, NS

Age

NS

14+

NS,

NS
NS

14
NS

NS
NS

1
1

Voided, frozen
Voided
fresh
Voided Fresh &
frozen
Voided
fresh
Voided
fresh
NS
Voided Fresh &
frozen
Voided NS
Voided or cath

NS

NS

NS

NS
NS

1
1

NS
NS
98% white

Voided or cath

Yes

NS

Voided fresh

NS

NS

Voided fresh

Yes

NS
NS

18
1

Voided fresh
NS

NS
Yes

NS

Unknown Collection Yes


frozen

Yes
Yes

Yes
Yes
Yes
NS
Yes
NS
Yes

84). The sensitivity and specificity of the BTA TRAK


test was 66% (95% CI: 62-71), and 65% (95% CI:
45-81) on data collected from 829 subjects who
provided urine samples for this same meta-analysis.
Thus, level 2a evidence does not support the use of
either BTA test alone for the detection of urothelial
carcinoma (Tab 7 and 8).

than stat (71% vs. 58%, respectively). The median


specificity was higher (73%) for 2084 BTA stat-tested
subjects than for 195 BTA TRAK-tested subjects
(66%). Subset analysis of recurrent tumor stratified
by grade showed lower sensitivities for grade 1
and 2 tumors for both BTA stat and TRAK (grade 1
= 45% and 55%, respectively; grade 2 = 60% and
59%, respectively) as compared to grade 3 tumors
(75% and 74%, respectively). A trend of increasing
sensitivity and specificity for overall tumor detection
has also been noted with increasing tumor stages.
[58] Furthermore, the BTA stat test has been shown
to have a lower sensitivity for detecting recurrent as
opposed to primary tumors, possibly related to the
smaller size of recurrent tumors BTA TRAK showed
increasing sensitivity and specificity with higher
tumor grades and stages (Tab. 7 and 8).[56]

Glas[52] further contributes to our understanding of


the literature by noting how study design influenced
results. With regard to the BTA stat test, sensitivity
was estimated to be significantly lower in case control
studies (66%, 95% CI: 60-71) when compared to
cohort studies (77%; 95%CI: 71-82). Specificity of the
BTA stat test was overestimated when interpretation
of results occurred in a manner which was not
blinded. Glas[52] described the studies available
for this meta-analysis as weak as most were not a
consecutive series of subjects suspected of having a
bladder tumor with independent assessment of the
marker test and reference standard.

Unfortunately, both current forms of the BTA test


are limited by conditions producing false positive
results. Because complement factor H is present
at high concentrations in blood, a positive BTA
stat or TRAK test will occur when hematuria is
present, regardless of the presence or absence of
urothelial tumor.[59,60] Greater than 80% of false
positives to either form of BTA test occur in subjects
with hematuria, dysuria, incontinence, a history of
intravesical therapy, ureteral stents or nephrostomy
tubes, renal or bladder calculi, benign inflammatory
disease (urinary tract infections or prostatitis), bowel
interpositions or other genitourinary cancers (renal or
prostate).[45,59,60,63,67] While use of exclusionary

The second clinical situation, monitoring of subjects


with a prior history of bladder cancer for recurrence,
is the indication for which the FDA has approved the
BTA tests as an adjunct to cystoscopy.[60] Urine
markers for bladder cancer surveillance: A systematic
review by van Rhijn[34] appears to address this
scenario. These authors reported a total of 1377
subjects on whom the BTA stat was performed and
360 subjects on whom the BTA TRAK was performed
in the setting of surveillance for recurrent disease.
The median sensitivity was higher for BTA TRAK

Table 7. BARD stat assay: Individual Analyses for Overall Sensitivity and Specificity
Author
Sarosdy 54
Wiener 61
Pode* 63
Irani 114
Babjuk 2002 70
Sozen 1999 64
Leyh 1997 62
Sharma 43
Ramakumar 654
Giannopouloulos 47
Nasuti 66
Heicappell 67
Raitanen 69
Boman 38
Schroeder 71
Halling 45
Serretta 68
Poulakis 40

%
Sensitivity
327^220 58
291
57
250
83
81
65
88
87
140
69
231^107 73
278
68
196
74
168
72
100
100
354
63
445
53
250+
64
115
53
2806
78
179
57
739
70

%
Specificity
72
68
69
72
74
68
52
82
73
57
84
93
86
64
77
74
62
67

N True
Positives
147
62
106
32

N True
Negatives
75
NS
NS
23

28
40
23
48*
50
3
105
63
96
31
16
279

NS = Not stated
+ = 289 samples from 250 patients
6 = 280 samples from 250 patients

182

N False
Negatives
73
NS
22
17

%
PPV
82
56

%
NPV
51
70

68
26
201
101
28
81
174
246
91
59

N False
Positives
32
64
NS
9
18
32
45
43
38
28
16
13
81
17
18

12
15
11
15
18
0
62
55
55
28

70

67

35
54
70
16

83
87
58
*

44
85
63

82
62
68

24
223

40
110

12
120

72

65

Table 8. BARD stat and TRAK assay: Tumor Characteristics

Author
Grade 1 Grade 2 Grade 3 pTa
Sarosdy 1997 54
57
56
95
111
Babjuk 2002 70
27
27
24
29
Ellis 1997 60
53
56
96
108
Tsui 2007 72
16
14
15
NS
Serretta 2000 68
7
19
29
13
Giannopoulos 2000 47 26
39
34
49
Poulakis 2001 40
130
167
70
209
Raitanen 2001 69
48
35
3
54
Nasuti 1998 66
0
2
1
NS*
Wiener 1998 61
23
38
30
47
Sharma 1999 43
NS
NS
NS
NS
Sozen 1999 64
14
16
10
16
Leyh 1999 62
26
45
36
58
Pode 1999 63
25
58
45
72
Ramakumar 1999 65
9
26
13
25
Boman 2002 38
54
70
35
117
Schroeder 2004 71
8
26
23
28
NA = Not applicable
NS = Not stated
*included more than one group together
^grade and stage unknown in 32, 33 subjects, respectively

criteria improve the performance of both BTA tests,


the signs and symptoms of benign inflammatory
conditions overlap those seen in subjects with
urothelial carcinoma. This limits the usefulness of
these tests for discriminating between malignant
and non-malignant states.[45,67] In particular, false
positives for up to 2 years after intravesical bacillus
Calmette-Guerin (BCG) therapy limits the BTA
tests usefulness in monitoring for recurrent tumor.
[64] False positives are more commonly seen with
the BTA stat as opposed to the BTA TRAK test.[60]
False positives are seen in < 5% of subjects with no
known urinary pathology.[45]

pT1
38
26
38
NS
27
26
47
20
NS*
25
NS
10
27
29
NS*
24
13

pTIS
18
3
18
NS
3
5
29
2
0
NS
NS
NS
5
NS
11
NS
5

>pT2
50
20
50
NS
12
16
112
3
1
19
NS
14
17
27
NS*
14
11

%Sensitivities recorded
BTA stat
BTA TRAK
Yes
NA
Yes
Yes
NA
Yes
NA
Yes
Yes
NA
Yes
NA
Yes
NA
Yes
NA
Yes
NA
Yes
NA
Yes
NA
Yes
NA
Yes
NA
Yes
NA
Yes
NA
Yes
NA
Yes
NA

controls for office-based laboratory procedures such


as the BTA stat test and declining re-imbursement
for such point-of-contact testing by Medicare and
private healthcare insurance companies have likely
reduced use of these tests by many urologists. The
cost to perform BTA TRAK testing and its limited
use in only bladder cancer subjects may also have
discouraged clinical laboratories from performing this
test, particularly in comparison to cytology, which is
routinely performed on multiple fluids from multiple
organ sites.
Suggested future trials for use of the BTA stat and
TRAK tests.

Although the value of this and other biomarkers in


detecting high risk (high grade) disease has been
suggested in numerous studies, their high rate of
false positive results and consequent low specificity
(through the influence of inflammatory conditions and
prior intravesical therapies either with chemotherapy
of BCG) somewhat limits their practical usefulness
in monitoring for disease persistence or recurrence.
In this regard, urinary cytology may be of value
because of its high specificity in the setting of high
grade disease.

At this time, we have not found evidence to endorse


use of either BTA test in screening for bladder
cancer. Use of BTA stat in subjects with a history of
urothelial cancer and a normal urinalysis could be
prospectively studied to determine if this combination
of tests (which might fail to detect small, low grade
recurrences) could safely reduce the frequency of
surveillance cystoscopies without compromising
cancer control. The suggestion by Blumenstein[76]
that serial measurements of BTA TRAK tests could
be useful in predicting recurrence in the individual
patient requires confirmation in a large, prospective,
multi-center trial. As multiple candidate antibodies
were obtained during the development of the BTA
tests, it is possible that an additional antibody that
does not cross-react with substances in the urine
of those with benign inflammatory conditions and/

Relatively few recent studies have been published


on the BTA tests, and most date from 1999 to 2001.
This may in part be explained by the decreasing
levels of specificity reported for the BTA stat test
between 1997 and 2001, and thus, lower enthusiasm
for their use. Additionally, the increase in regulatory

183

or hematuria could be isolated and used to further


refine BTA testing.

two other studies, the overall sensitivity was less than


50%.[71,72] For UBC-ELISA, different studies have
used different cut-off limits with a range between
0.16 g/L and 15 g/L. In one study, the cut-off limit
was called an Index value, which was calculated
by dividing the value during follow-up divided by the
value before the first transurethral resection.[74] In
some studies, the UBC values were normalized to
creatinine, whereas in other studies they were not
normalized; the manufacturer does not recommend
such normalization. For these reasons, a metaanalysis of UBC-ELISA results from different studies
cannot be and should not be performed.

3. UBC tests
UBC-Rapid and UBC-ELISA tests are immunological
assays available from IDL, IDL Biotech, (Borlange,
Sweden). Both assays detect cytokeratin 8 and 18
fragments in urine. Cytokeratins are intermediate
filament type cytoskeletal proteins specific for
epithelial cell origin. In human cells, a total of 20
cytokeratins have been identified and the expression
of cytokeratins 8, 18, 19, and 20 at the protein or
mRNA level has been evaluated as bladder cancer
markers.[77] Since cytokeratins are intracellular
proteins, the detection of these proteins in urine
is possible only when they are released in urine
following cell death. The UBC-Rapid assay is a
qualitative point-of-care assay wherein cytokeratin 8
and 18 fragments present in urine react with goldlabeled antibodies forming a complex. This complex
migrates to the reaction zone where it reacts with
a specific antibody to produce a dark line (positive
result). Any excess of gold-labeled antibodies
continues to migrate into the second capture zone to
form a second dark line (test control).[78]

4. Survivin
Survivin is an anti-apoptotic protein that is a
member of the inhibition of apoptosis protein (IAP)
gene family. Survivin levels are elevated in bladder
cancer, and therefore, survivin has been suggested
as a promising biomarker for bladder cancer.[80-82]
The commercially available bio-dot assay (Fujirebio
Diagnostics, Inc.) for Survivin is a technique that
is routinely performed in any research laboratory
to detect a variety of proteins. In Survivin bio-dot
assay (a dot-blot assay) urine samples are blotted
onto a nitrocellulose or ImmobilonP membrane,
along with recombinant survivin protein blotted
at various concentrations. Following sequential
incubations with a rabbit polyclonal anti-survivin
antibody and a horseradish peroxidise-conjugated
secondary antibody, the dots are visualized by
chemiluminenscence. The intensity of the specimens
and the standard-dots is measured by densitometry
and the amount of survivin in specimens is determined
from a standard curve. This assay however has not
been used recently and the current assays reported
in various articles are either quantitative reverse
transcription polymerase chain reaction (Q-PCR)
or qualitative reverse transcription PCR (RT-PCR)
assays. In the latter survivin mRNA expression is
detected by agarose gel electrophoresis, followed
by ethidium bromide staining.

UBC-ELISA is a solid-phase 2-step sandwich assay.


Specimens, standards, and controls are incubated
with microtiter wells coated with a mouse monoclonal
anti-UBC antibody. Following incubation, unbound
samples are washed off and the wells are incubated
with a horseradish peroxidase second antibody.
Following washing, the wells are developed using
tetramethylbenzidine substrate and the color
intensity is measured at 450 nm. Levels of UBC in
the urine specimens are calculated from a standard
curve. The manufacturer-suggested cutoff limit for
UBC-ELISA is 12 g/L. UBC-ELISA requires sending
samples to specialized laboratories, where trained
personnel can conduct the ELISA.
Pubmed search of the term UBC and bladder
cancer resulted in 73 hits. After examining the title
and the abstract of each article, nineteen articles
were found to be on UBC tests. In these nineteen
studies, 623 subjects have been assayed by the
UBC-Rapid test and 3,102 individuals have been
assayed by UBC-ELISA.

Pubmed search of the term Survivin and bladder


cancer resulted in 126 hits, which included 12
reviews. After examining the title and the abstract
of each article, ten articles were found to have
evaluated the efficacy of survivin as a urine marker
using the bio-dot, Q-PCR or RT-PCR assays. One
of these studies was a prospective cohort screening
study but it did not include any bladder cancer cases.
[83]

According to the STARD criteria the quality of many


articles was moderate to good with a few articles
displaying excellent quality of reporting. The majority
of studies provided LoE Grade III and IV evidence.
Three cohort studies were classified as LoE IIb
and one study by Hedelin et al was a prospective
screening study.[79]

According to the STARD criteria the quality of many


articles was moderate to good, with a few articles
displaying excellent quality of reporting. The majority
of studies provided LoE Grade III and IV evidence.
Three cohort studies were classified as LoE IIb and
one study by Davies et al had both a retrospective
blinded cohort and an prospective cohort.[83]

Meta-analysis of UBC-Rapid in three studies


reporting 623 patients (but UBC-Rapid assay was
performed on 515 patients) showed an overall
sensitivity of 59.3% with 86.1% specificity. However,
it is noteworthy that barring the initial study[78], in

184

Since the dot-blot assay detects survivin protein


and the PCR assays detect mRNA expression, the
results reported in studies using the dot blot assays
and PCR assays cannot be and should not be used
to perform a meta-analysis of the Survivin marker.
Furthermore, in each study the PCR primers used
for Q-PCR or RT-PCR were different, and therefore,
no two PCR studies are alike. In addition, Q-PCR is
quantitative and RT-PCR is qualitative, and therefore,
the results of these PCR studies also should not
be used to perform a meta-analysis. Davies et al
reported that urinary survivin levels decrease after
radical prostatectomy; this study also reported an
unusually high false positive rate, at 50%. Given
that each study has used different techniques to
assay survivin expression, this marker is not ready
for diagnosis and/or surveillance of bladder cancer
patients.

three articles which evaluated the efficacy of BLCA-4


by ELISA was good. Since these were case-control
studies, the evidence provided was classified as
Grade IV. Since these three studies either used
the same or similar patient populations [86,87] or
different assays, a meta-analysis cannot be and
should not be performed.

6. Cyfra 21-1
CYFRA21-1 is a Cytokeratin (CK) based assay.
CKs are intermediate filament proteins specific
for epithelial cells. A given epithelium can be
characterized by a chain-specific CK expression
pattern. In general, overexpression of a particular
chain-specific CK is associated with bladder.
CYFRA21-1 is an ELISA that detects fragments of
CK19 with the help of two monoclonal antibodies
(BM19.21 and KS19.1) in urine. Urinary stones,
infection and previous intravesical treatment with
BCG caused false positive results.[91] In 3 studies
from 2 institutes analyzing CYFRA21.1 in patients
under surveillance, sensitivity was 85% in 156 cancer
positive patients. Specificity was 82% in 323 patients
with no tumor at cystoscopy. Sanchez et al. [41,92]
reported similar results for NMP22 and CYFRA21-1
which is not what one would expect from a urine
marker with such a high potential. Moreover, the
number of studies with CYFRA21.1 is relatively low,
cutoff values have not yet been defined properly and
the additional value over NMP22 is not obvious.

5. BLCA-4
BLCA-4 assay is a sandwich ELISA commercially
available from Eichrom Technologies. BLCA-4 is a
nuclear matrix protein and has homology to ELK3
gene, a member of the ETS family of transcription
factors.[84] BLCA-4 is differentially up-regulated in
bladder cancer cells and tissues and was identified
by two dimensional gel electrophoresis of the nuclear
matrix components from normal and tumor tissues.
[85] The initial assay for BLCA-4 measurement in
urine was an indirect ELISA.[86,87] In those studies,
the cut-off limit to detect bladder cancer was set
at 13 g/mg protein. Later a sandwich ELISA that
utilizes 2 monoclonal antibodies was developed
and a cut of limit of absorbance set at 0.04 units at
630 nm was used for calculating the sensitivity and
specificity.[88] Therefore unlike the previous indirect
ELISA, neither urine precipitation nor normalization
of the BLCA-4 levels to urinary protein concentration
is carried out when using the sandwich ELISA for
BLCA-4 measurement.

The body of evidence for CYFRA21-1 is limited.


Only few reports on marker performance have
been published. Reporting quality is moderate to
poor, the level of evidence provided ranges from 4
to 3B, marker status according to the IBCN criteria
is considered to be level I. In summary, CK based
assays, particularly CYFRA21-1, are promising;
however, current information at this stage is
insufficient for any definite statements on the clinical
use in bladder cancer detection and follow-up.

Pubmed search for the term BLCA-4 and bladder


cancer resulted in 14 hits, which included six
reviews on bladder tumor markers. After examining
the title and the abstract of each article, three articles
were found to evaluate the efficacy of the BLCA-4
marker for the detection of bladder cancer. All of
these studies were case-control and from a single
institution [86-88]. Another study also found that
BLCA-4 levels were significantly elevated among
patients who have suffered from schistosomiasis
and have developed bladder abnormalities as
evaluated by ultrasound when compared to those
without detectable bladder abnormalities [89].
Recently, in suppressive subtractive hybridization
(SSH) and cDNA microarray, conducted on voided
urine specimens showed that BLCA-4 expression
has 90% sensitivity and 85% specificity to detect
bladder cancer.[90]

b) Cell-based assays
1. DD23
DD23 is a murine monoclonal antibody that was
evaluated in 1996 with quantitative fluorescence
image analysis in exfoliated urothelial cells.[93]
When used as a quantitative marker to detect
bladder cancer, sensitivity was 85% (41 cases) and
specificity in asymptomatic age-matched controls
was 95% (41 subjects).[93] The DD23 assay test
was subsequently developed using an avidinbiotin alkaline phosphatase immunocytochemical
procedure.[94,95] A single positive cell was
considered a positive urine test. In 308 cases
under surveillance for non-muscle invasive bladder
cancer, sensitivity was 81% and specificity 60%.
[94] In another study from the same authors in 81
patients analysing 151 samples sensitivity was 70%
and specificity 60%.[95] The authors concluded

According to the STARD criteria, the quality of the

185

that DD23 was able to enhance the sensitivity of


cytology, in particular for low grade tumors.[94,95]
The first results in patients under surveillance are
characterized by a low specificity which implies that
DD23 is not an ideal marker to lower the cystoscopy
frequency in these patients.

studies was classified as LoE 2b.


One remarkable feature of the uCytTM assay is
a reproducibly high sensitivity specifically in low
grade lesions. On average, detection rate for low
grade tumors was 75%, sensitivity for G2 and high
grade tumors was approximately 85% (Tab. 9).
Overall specificity was 75%. Discriminating between
diagnostic and follow-up trials, sensitivity appears to
be lower specifically in low and intermediate grade
lesions in diagnostic studies as compared to followup trials. However, this conclusion is based upon a
small number of cases.

The body of evidence for DD23 is limited. Only few


reports on marker performance have been published.
Reporting quality is moderate to poor, the level of
evidence provided ranges from 4 to 2B, marker
status according to the IBCN criteria is considered
to be level I. In summary, current data do not permit
definite conclusions on a clinical use of DD23.

The uCytTM assay has been reported to be confounded


by a variety of different urological conditions (BPE,
hematuria, urolithiasis or inflammatory conditions).
However, studies in hematuria populations suggest
that the impact of these conditions on test specificity
is limited.[17,97]

2. uCyt+TM /ImmunocytTM
The uCyt+TM assay, formerly ImmunocytTM, is a
commercially available immunocytological assay
based upon microscopical detection of tumorassociated cellular antigens in urine-derived
urothelial cells by immunofluorescence (Scimedx
Inc., Denville, NJ, USA). For tumor cell detection
an antibody cocktail containing fluorescein-labeled
monoclonal antibodies M344 and LDQ10 directed
against sulfated mucin glycoproteins and Texasred linked antibody 19A211 against glycosylated
forms of high molecular carcinoembryonic antigens
(hmCEA) is used. After staining the samples are
studied for immunofluorescence examining more
than 500 nuclei. In most studies specimens with
>1 green or red urothelial cell are considered
immunocytologically positive.

In general, the uCytTM assay appears more sensitive


but less specific as compared with urine cytology.
Side-by-side comparisons with other assays have
been reported. However, interpretation is difficult
since adequate experience of the investigators is a
prerequisite to obtain meaningful results for the uCytTM
assay but is not disclosed in these publications.
There was one prospective trial on marker-guided
follow-up providing information that may be
classified as LoE grade 1b [130,131] according to
the Oxford classification for diagnostic procedures.
[36] The very same trial was classified as a phase
III trial concerning the IBCN classification on marker
development,[12] while all remaining studies were
considered phase II.

The uCytTM test is a cell based assay. Assay costs


and requirements concerning lab equipment, time
for specimen processing and reading as well as
experience necessary for adequate interpretation
of the staining must be considered high. These
properties restrict the use of this test to more
specialized laboratories. Reproducibility, i.e. interobserver variability, is reasonable provided that
reading is performed by trained staff with high
experience.[96]

Mian and coworkers [130] reported on a prospective


marker-based follow-up trial in 942 patients: after
an initial 3 month follow-up including cystoscopy,
cytology and uCytTM test patients were divided into
3 categories according to their risk profile. Low risk
patients were followed using uCytTM and cytology;
cystoscopy was performed if one of the tests was
positive or after 1 year. Intermediate risk tumors
were followed by cytology, uCytTM and UrovysionTM.
Negative patients or patients with a positive p16/
CEP3 FISH test were further followed as low risk
patients, while patients with positive cytology, uCytTM
and CEP7/17 positive UrovysionTM test underwent
cystoscopy. High risk patients were followed by
cystoscopy every 3 months. The authors reported
that in low risk patients the number of cystoscopies
could be safely reduced and significant costs were
saved. [131] However, precise data required for
understanding of the authors conclusions are
lacking. In addition, analysis of the benefit or added
value in comparing fewer cystoscopies in low
risk disease patients, reliability of cytology in high
risk patients, and costs of uCyt and Urovysion are
required in considering the results in the context of
the type of cancer diathesis being monitored.

A literature search on the terms immunocytology,


immunocyt, uCyt, and bladder cancer yielded
49 hits. After removal of reviews, meta-analyses and
redundant trials, 20 studies assessable for criteria
concerning assay performance and comprising
more than 5,000 individuals were identified forming
the basis for this assessment of assay performance.
[17,97-129]
Accuracy of reporting according to the STARD
criteria [37,38] was mostly moderate or poor, only few
papers displaying good reporting quality. Specifically,
information on the training and experience of
investigators a parameter highly affecting uCytTM
results was not provided and information on the
blinding of investigators towards clinical observations
was rare. The majority of trials provided LoE Grade 3
and 4 evidence; however, information from 8 cohort

186

Table 9. Performance characteristics for uCyt+TM


low grade tumors acc. to the 2004 classification were included in the G1 category acc. to the 1973/1998 classification, high grade tumors acc. to the 2004 classification and
CIS were included in the G3 category;
inf. = informative; UUT = upper urinary tract tumors; # = no. of test (not of patients; therefore not considered for specificity calculation)
cohort study: consecutive patients, no healthy controls included; marker guided prospective trial: clinical decision making based upon marker result;
LoE: case control studies were considered LoE grade IV, studies including diagnostic and follow up patients were considered LoE grade IIIb, studies including clearly defined
patient cohorts, consecutive cases were considered LoE grade IIb, results from a marker guided prospective trial were considered LoE grade Ib;
* Marker status according to IBCN classification 2008; ** no. of requirements met acc. to STARD recommendations
Note: Li data for specificity but not considered for sensitivity analysis
Study type

%
Sensitivity

% Specificity

187

Author (year) Reference

S t u d y D i a g n o s i s / No. of Grade I/
design
Follow-up
patients LG

Grade II

Grade III/ HG/


CIS

Remarks

LoE

IBCN*/ STARD**
Status

Fradet (1997) 115

C a s e mixed

300

23/27

41/43

24/25

79/102

272/300 inf.

Mian (1999) 116

control
Cohort

mixed

264

85.2
21/25

93
22/25

96
28/29

77.3
135/170

249/264 inf.

3b

16/25
II

Olsson (2001) 117

Cohort

mixed

121

84
8/8

88
14/14

97
8/8

79.4
57/83

114/121 inf.

3b

18/25
II

Lodde (2001) 118

Cohort

diagnostic

37

100
1/3

100
6/6

100
4/5

68.7
20/21

2b

15/25
II

Mian (2003) 119

(UUT!)
Cohort

mixed

181

33
25/31

100
21/24

80
23/25

95
71/101

173/181 inf.

3b

15/25
II

Feil G (2003) 120

Cohort

mixed

92

80.6
1/7

87.5
4/9

92
6/10

71
73/87

113/121 inf.

3b

16/25
II

Piaton (2003) 121

Cohort

diagnostic

236

14.3
4/10

42.9
15/17

60
23/30

83.9
130/151

231/236 inf.

2b

14/25
II

Pfister (2003) 122


Piaton (2003) 121

Cohort

Follow up

458

40
13/21

88.2
16/24

76.7
30/39

83.3
286/342

451/458 inf.

2b

19/25
II

Pfister (2003) 122


Hautmann (2004) 123

C a s e diagnostic

94

61.9
3/4

66.7
7/15

76.9
9/11

81.9
48/64

PPV 54.5

19/25
II

75

46.7

81.8

75

NPV 81.3

control

14/25

Table 9. Performance characteristics for uCyt+TM (continued)


Toma (2004) 124

Cohort

mixed

126

6/7

17/23

10/12

60/82

Cohort

Follow-up

904

85.7
48/64

73.9

Tetu (2005) 125

83.3
34/40

72.5
453/734

870/904 inf.

85

62

PPV 26

75 (LMP+LG)
Messing (2005) 126

Cohort

Follow up

Mian (2005) 127

Cohort

mixed

Mian (2006) 96

(CIS!)
M a r k e r - Follow up

341

188

prospec.
Cohort

4/6

206/274

79

90

67

75

35/35

12/17

942

96/121

74/88

100
82/89

70.6
1152/1588#

79.3

84.1

92.1

72.5

41

8/13

11/11

10/16

Schmitz-Drger (2007/2008) Cohort


diagnostic
129,49
(microhematuria)
Soyuer (2009) 130
C a s e mixed
control
Horstmann (2009) 131
Cohort
Follow up

222

62
4/6
66

91
4/4
100

63
170/201
85

Schmitz-Drger (2008/2010) Cohort

103

24/31
77.4
20/32
62
7/8

21/23
91.3
20/28
72
11/13
85

132,49

Li (2010) 133

Follow-up

9/10

35

guided
Sullivan (2007) 128

22/28

diagnostic

90
221

( g r o s s
hematuria)
C a s e mixed

87

1881/1991 inf.

II

2b

13/25
II
17/25

2b

II
19/25

3b

II

1b

15/25
III
17/25

PPV 43
NPV 88

2b

II

2b

17/25
II
18/25

31/36
86.1
78/108
72
64/78

211/222 inf.
PPV 25.8
NPV 99
PPV 90
NPV 79.5
PPV 72
NPV 74
100/103 inf.

82

PPV 57.6

4
2b
2b

II
15/25
II
17/25
II
19/25

NPV 94.4
191

control
Total

44/53
82

NPV 93
327/341 inf.
PPV 72
NPV 74

3b

4899

334/446
74.9

76/93

85/98

81.6

86.7

290/344
84.3

387/462
83.8

2.068/2.745
75.3

II
15/25

4.992/5.242
95.2

3. Urovysion

and low grade lesions. [33] This is paralleled by a


high specificity of appr. 80%, however, again with a
broad range from 43 100% (Tab. 10).

The UroVysion multicolor fluorescence in situ


hybridization (FISH) test (Vysis, Abbott Laboratories,
Des Plaines, IL, USA) is a cell based assay containing
probes to the centromeres of chromosomes 3, 7, and
17 and to the 9p21 locus. The assay was approved
by the U.S. Food and Drug Administration (FDA) for
surveillance of patients with previous bladder cancer
but also for diagnosis in hematuria. A minimum
of 25 morphologically abnormal cells is viewed.
Detection of 4 or more cells that have gains in 2 or
more of chromosomes 3, 7, and 17 in the same cell
or at least 12 cells without a signal for P16 tumor
suppressor gene locus 9p21 are mostly classified as
a pathologic result. However, a variety of different
definitions and cut-off levels are being used.[33]

In general, the UroVysion assay appears more


sensitive (even in high grade disease) but less specific
as compared with urine cytology. A recent metaanalysis of FISH showed that overall performance of
FISH was better than that of cytology (area under the
curve: 87% vs 63%).[33] This difference, however,
was almost entirely attributable to the difference in
overall performance in diagnosing pTa patients; the
higher performance disappeared when pTa patients
were excluded from the analysis (area under the
curve: 94% vs 91%). Side-by-side comparisons
with other assays have been reported; however,
interpretation is difficult since adequate experience of
the investigators is prerequisite to obtain meaningful
results for the UroVysion assay but not disclosed in
these publications.

Assay costs and requirements concerning lab


equipment, time for specimen processing and
reading as well as experience necessary for adequate
interpretation of the staining must be considered
high. These properties restrict the use of this test to
more specialized laboratories but may also explain
the great ranges in sensitivity and specificity reported
for this assay. Reproducibility has been reported to
be good provided that the reading is performed by
experienced laboratory staff. The UroVysion assay
has been reported to be confounded by a variety
of different urological conditions (other tumors,
urolithiasis or inflammatory conditions). Another
limitation is that the rate of non-informative cases of
appr. 10% must be anticipated (Tab. 10).

Mian and coworkers [130] reported on a prospective


marker-based follow-up trial in 942 patients: after
an initial 3 month follow-up including cystoscopy,
cytology and uCytTM test patients were divided into 3
categories according to their risk profile. Intermediate
risk tumors were followed by cytology, uCytTM and
UrovysionTM. Negative patients or patients with a
positive p16/CEP3 FISH test were further followed
as low risk patients, while patients with positive
cytology, uCytTM and CEP7/17 positive UrovysionTM
test underwent cystoscopy. Information on the
results of the UroVysion data, however, has not been
published yet.

A literature search on the terms FISH, UroVysion,


and bladder cancer yielded 331 hits. After removal
of reviews, metaanalyses and redundant trials, 21
studies assessable for criteria concerning assay
performance and comprising 2852 individuals were
identified forming the basis for this assessment of
assay performance.[28,46-48,99,101-106,110,132141]

Although there is a relatively high rate of falsepositive results translating into a relatively low PPV
of the test, findings from several studies suggest that
the low specificity in follow-up trials may be explained
in part as an anticipatory positive result, in which a
premalignant change precedes the discovery of a
recurrent malignancy.[28,132,142] One study [132]
found that 89% of the patients that had a falsepositive test had a positive bladder biopsy within 12
months of the test, while another found that FISH
preceded tumor recurrence in 85% of patients.
[142] Nonetheless, the real role of an anticipatory
positive result is still unclear as many patients with
non-muscleinvasive bladder cancer eventually
experience disease recurrence.

Accuracy of reporting according to the STARD criteria


[37,38] was mostly moderate or poor, few mostly
more recent - papers displaying good reporting
quality. Specifically, information on the training and
experience of investigators a parameter highly
affecting UroVysion results is not provided and
information on the blinding of investigators towards
clinical observations is rare. Ten trials provided LoE
Grade 3 and 4 evidence; however, information from
11 cohort studies was classified as LoE 2b.

In considering the observations and conclusions


reported in these studies, it also becomes important
to consider the cost of these tests, especially if
sufficient information is otherwise available through
less costly standard examinations (cystoscopy,
cytology) or other approved biomarkers. Because
of the importance of determining any added
value in the use of a particular test, costs, difficulty
in performance, confusion of interpretation in a
particular clinical setting, and the emotional stress
encountered by both patient and physician in

The broad range for sensitivity and specificity of


UroVysion FISH reported in different papers is
notable and may not only reflect patient selection,
study design or tumor prevalence but also technical
aspects such as cut-off definitions and experience of
laboratory staff. However, in systematic reviews and
meta-analyses sensitivity has been found to exceed
70% and even approach 80% when omitting small

189

Table 10. Performance characteristics for UroVysionTM


low grade tumors acc. to the 2004 classification were included in the G1 category acc. to the 1973/1998 classification, high grade tumors acc. to the 2004 classification and
CIS were included in the G3 category;
inf. = informative; UUT = upper urinary tract tumors; # = no. of test (not of patients; therefore not considered for specificity calculation)
cohort study: consecutive patients, no healthy controls included;
LoE: case control studies were considered LoE grade IV, studies including diagnostic and follow up patients were considered LoE grade IIIb, studies including clearly defined
patient cohorts, consecutive cases were considered LoE grade IIb, results from marker guided prospective trials were considered LoE grade Ib;
* Marker status according to IBCN classification 2008; ** no. of requirements met acc. to STARD recommendations
Study type

%
Sensitivity

Author (year) Reference

Study design D i a g n o s i s / No. of Grade I/


Follow-up
patients LG

Bubendorf (2001) 46

Case control

Placer (2002)

15/21 (71) 25/29 (86)

190

8/15

438

(53.3)
(83.3)
12/22 (55) 7/9 (78)

Mian (2003) 119

C o h o r t / Follow-up
Case control
(separate)
Cohort
Mixed

57

7/8 (87)

Skacel (2003) 99

Cohort

111

19/23 (83) 28/35 (80)

Sarosdy (2002) 44

mixed

91

Grade II

86

134

Case control

mixed

% Specificity

diagnostic

10/12

19/19 (100)

Grade III/
HG/CIS
16/17 (94)

Remarks

26/27 (96.3)

19/19

29/34

(100)
17/18 (94)

(85.3)
75/114 (65.8)
260/275 (94.5)
controls

1/1 (100)

11/24 (46.4)

23/24 (96)

28/29 (97)

LoE

Concordance 4
voided/ barbotage 85%
4
2b

IBCN*/ STARD**
Status
II
15/25
II
18/25
II
19/25

5/57 n. inf.

3b

II

2b

16/25
II

Reflex

19/25

Veeramachaneni (2003) 135

(neg. Cytol)
n.r.

mixed

121

Krause (2004) 136

Case control

mixed

84

Varella-Garcia (2004) 137

Cohort

Follow-up

19

Pycha (2004) 138

Cohort

Follow-up

49

1/3

12/14

0/1

n.r.

36/121

(33)
10/14
(71)
2/2
(100)

(84)
10/11
(91)
3/3
(100)

(0)
44/44
(100)
1/2
(50)
1 2 / 3 5
(34.3)

25/35
(71)
12/12
(100)
12/14 (85.7)

4/106

4
2b
2b

II
15/25
II
17/25
II
17/25
II
17/25

Table 10. Performance characteristics for UroVysionTM (continued)


Kipp (2005) 139

Follow-up

37

Laudadio (2005) 140

Cohort
Prospec-tive
Cohort

mixed

300

14/25 (56)

12/25
(48)
18/19 (95)

Junker (2006) 141

Cohort

diagnostic

121

n.r.
(37)

n.r.
(91.7)

Bergmann (2007) 142

41

Moonen (2007) 143

C o h o r t Follow-up
retrospec-tive
Cohort
Follow-up

Yoder (2007) 27

C o h o r t Follow-up

249

n.r.
(65.4)
30/39#
(77)
6/27
7/19
(21.4)
(36.8)
6 / 1 9 15/20 (75)
(31.6)

(90.5)

105

Reflex

12/12
(100)
167/256 (65)

23/28
(82.6)
80/86#
(93)
12/18
37/41
(66.7)
(89.7)
3 8 / 4 2 1 4 7 / 1 6 8
(87.5)

(neg. Cytol)

191

Riesz (2007) 144

Case control

Frigerio (2007) 145

Case control mixed


retrospec-tive
Cohort
n.r.

Ferra (2009) 113

diagnostic

50

4/9 (44.4)

56

10/18 (55)

140

9/19

(100)
21/24
(87.5)
45/60

(47.4)

(75)

Reflex

16/16 (100)

1 4 / 1 4 11/11 (100)

After intravesi- 2b
cal prophylaxis
16/141 n. inf.
3b
20/141 n. inf.

2b

16/162 n. inf.

2b

10/113 n. inf.

2b

35/56
pat. 2b
(62.5%)
UC
neg.,
FISH
pos. develop
tumor
5/55 n. inf.
4

n.r.
27/68

4
10/161 n. inf.

3b

(39.7)

II
17/25
II
16/25
II
16/25
II
17/25
II
18/25
II
20/25

I
14/25
I
(14/25)
II
17/25

Susp./pos.
Caraway (2010) 146

Cytol.
C o h o r t mixed

Youssef (2010) 147

retrospec-tive
Cohort
F o l l o w - u p 142

600

n e g a t i v e
Mian (2010) 148

Cohort

170/263
(64.6)

540/632#

65/1006 n. inf.

3b

II

19/142 n. inf.

2b

16/25
II

1/7

3/10

(85.4)
100/106

(14.3)

(30)

(94.3)

cytology!
D i a g n o s t i c 55

24/24 (100)

34/38 (89.5)

18/25
1/68 n. inf.

(UUT)
Total

2b

II
17/25

2852

124/232

152/187

2 9 6 / 3 7 3 1 0 3 6 / 1 2 9 3 206/2263 (9.1)

(53.4)

(81.3)

(79.3)

(80.1)

assessing the reliability of a test result should all


be considered in the application of any marker for
routine clinical use.

smokers with >25 RBC/l on dipstick testing might


be a scenario to be considered.
A key problem of this concept is the high prevalence
of hematuria in the general population, along with
its low specificity, raising unnecessary anxiety in
screened subjects and requiring urologic workup in a high number of individuals without bladder
cancer. Hedelin and coworkers tried to correct for
this parameter by increasing the cut-off level for
hematuria but the efficacy of this measure need
to be confirmed. On the other hand, detection of
additional diseases requiring intervention is frequent
in hematuria patients and needs also to be taken into
account when considering this approach.[19,147]

2. Screening
Screening for bladder cancer, i.e. investigation of
an asymptomatic population, represents a specific
diagnostic challenge. While screening for breast
cancer and colorectal cancer has gained social
acceptance, bladder cancer screening has not been
considered a reasonable approach mainly due to
the low prevalence of the disease in an unselected
population. Nevertheless, few studies have been
published reporting on screening for bladder cancer
in populations with an increased risk of developing
bladder cancer.

b) Smoking:
Steiner et al. invited 183 subjects identified as
smoking 40+ pack-years to join a bladder cancer
screening program including urinary dipstick test,
urine cytology, NMP 22 BladderChek and UroVysion.
[14] 75 subjects with at least one positive test result
were offered urologic work-up. 5 urothelial cancers
(3 bladder tumors 1x pTa LG, 2x carcinoma in situ
and two upper urinary tract tumors (pTaG1 and
pTxN2G3)) were detected. While this study found a
higher incidence of cancer, another study of 1502
subjects with greater than 10 years of smoking
screened for bladder cancer using BladderChek
found only 2 cancers and one patient with atypia.
[15]

a) Hematuria:
Messing et al. invited 1575 men aged 50 years
and older to test their urine repetitively with a
chemical reagent strip for hemoglobin.[21,143-145]
Participants with positive test results underwent
standard urologic evaluation. Bladder cancer stages
and grades as well as the outcomes of men with
screen-detected tumors were compared with the
grades, stages, and outcomes of an age-matched
cohort of men with newly diagnosed BC who
were reported to the Wisconsin Tumor Registry in
1988 (n=509). Two hundred fifty-eight screening
participants (16.4%) were evaluated for hematuria,
and 21 participants (8.1%) were diagnosed with
bladder cancer. Proportions of low-grade (Grade 1
and 2) superficial (Stage Ta and T1) versus highgrade (Grade 3) superficial or invasive (Stage < T2)
cancers in screened men (52.4% vs. 47.7%) and
in men from the tumor registry (60.3% vs. 39.7%)
were similar (p=0.50). The proportion of high-grade
superficial or invasive bladder tumors were lower in
screened men (10%) than in unscreened men (60%;
p=0.002). At 14 years of follow-up, cancer specific
survival in screen-detected patients was 100%,
whereas 20.4% of unscreened men had died of
bladder cancer (p=0.02).

c) Professional exposure:
In a prospective study, Hemstreet et al. assessed
the risk for the development of bladder cancer in a
group of 1788 Chinese workers who were exposed
to benzidine using a biomarker profile over a period
of 6 years.[100] This biomarker profile included
the analysis of DNA ploidy, G-actin, and tumorassociated antigen P-300. Although the biomarker
profile placed only 21% of the exposed workers in a
high or moderate risk group, 87% of the 28 bladder
cancer cases in the entire cohort were found in this
group, and all of the tumors were clinically organconfined. Interestingly, a positive biomarker profile
occurred 15 to 33 months before the clinical detection
of bladder cancer.

Hedelin et al. investigated 2000 randomly selected


men, age 60-70 years, invited to participate in
a screening program based upon dipstick for
hematuria and the UBC assay.[146] Men with 5-10
red blood cells (RBC)/l and an International Prostate
Symptom Score (IPSS) of >10 and all men with >25
RBC/l and/or elevated UBC levels underwent both
white-light and fluorescence cystoscopy. In 14% of
the responding 1096 men microhematuria with 5-10
RBC/l was observed. One tumor was detected in
the 62 men with 5-10 RBC/l and an IPSS of >10.
Among the 10% of men (n=112) with >25 RBC/l,
four bladder tumors were detected. Another two
tumors were detected in men without hematuria but
with a positive UBC test. The authors concluded
that hematuria-based screening among older male

Giberti et al. investigated 171 workers at an Italian


Coke plant with long-term exposure to PAHs using
dip stick testing for hematuria, cytology and the uCyt+
assay.[148] Although uCyt+ was positive in 12% of
the screened subjects subsequent urological work-up
yielded no urothelial cancers in this cohort. While the
relatively young age of the screened subjects (mean
53 years) may have affected disease prevalence,
a low cut-off value for the uCyt+ assay could be
responsible for the low specificity observed.
Several other studies investigating professionally
exposed risk populations (Fire fighters, chemical

192

workers, workers in alloy smelters [147,149,150]


using NMP22, uCyt+ or a mix of different molecular
markers, demonstrated good sensitivity and
specificity for the markers. However, due to the low
prevalence of disease (< 1%) in the cohorts studied,
the positive predictive value of the assays remains
unsatisfactory.

Several investigators have demonstrated a


correlation between sensitivity and specificity for a
variety of biomarkers. As for PSA in prostate cancer
for example, an increased cut-off level will both,
increase specificity and decrease sensitivity (and
vice versa for a decreased cut-off). Since different
threshold definitions are in use for several assays
(e.g. UBC, NMP22, UroVysion), the choice of a cutoff level will automatically have significant impact on
the performance of a given assay (see above).

Davies et al. targeted another risk group screening


457 patients with spinal cord injury for 5 years using
urine cytology, the BTA stat test and the survivin
assay.[83] A total of 1075 urine specimens from
457 patients were analyzed. Of the 1073 BTA stat
tests, 119 showed positive reactions (specificity
88.9%) and 954 were negative. In the survivin
assays, 47 samples had a score of one, 38 a score
of two, and 9 a score of three (specificity 91.2%). No
cytology specimens were noted to have malignant
cells (specificity 100%). None of the three patients
diagnosed with bladder cancer had a positive test
result.

While investigator bias may not play a role in pointof-care assays, this is of particular relevance for cellbased tests (e.g. UroVysion or uCyt+). The precision
of these tests is clearly correlated with training status
and experience of the laboratory staff.[96] Since the
experience of investigators is not reported in the
literature despite specific recommendations to do
this (STARD), it is impossible to estimate the impact
of investigator bias in the comparison of different
assays.

In summary, despite a limited number of studies


there is evidence that screening for bladder cancer
in general is feasible and screened subjects may
benefit from early cancer detection. However, cost
calculations based upon the results from published
trials suggest costs between $ 25.000 50.000/
cancer detected.[6] This finding clearly points at
a careful selection of high risk populations and
demonstrates the necessity of an effective design of
future protocols.

Furthermore, it is the scientific norm to report


innovations, promising results, and initially positive
observations, all of which contribute to an enthusiasm
for an early clinical application. However, such initial
reports may be limited in their study design, length of
follow-up, and numbers needed to provide statistical
power in order to validate results. These together
with misapplication of observations to different
clinical scenarios may account for the commonly
observed failure to validate initially promising albeit
preliminary reports.

3. Problems of marker comparison

With decreasing financial resources of health care


systems worldwide cost-benefit considerations
have gained considerable importance. This aspect,
however, has rarely been addressed and included in
a side-by-side comparison of different assays.

There is a variety of reasons for why a comparison


of markers, aiming at the identification of the best
marker, is of limited value:
Different performance profile

Although these problems and limitations discussed


above currently prevent a sufficient comparison
between different markers and urine cytology, it
is evident that there is an urgent need to identify
the optimal diagnostic armamentarium for the
different clinical scenarios. Therefore, well designed
prospective studies are needed to confirm the
significance of urine cytology and identify an potential
added value of the markers.

Threshold definitions
Technical aspects
Cost-benefit considerations
This assessment has demonstrated that markers
have different performance profiles. While some
of these markers may have a similar sensitivity
through all tumor grades, others may have a higher
sensitivity in high grade tumors. Similarly, specificity
particularly in urine-based markers and, to a lesser
extent of cell-based assays, is highly dependent upon
the composition of the tumor-negative cohort. While
these markers in general may have good specificity
in a healthy control population, they uniformly have
a lower specificity in cohorts comprising patients
suffering from non cancer-related urological diseases
(e.g. inflammatory conditions, stone disease,
hematuria, or BPE). As a consequence, a different
composition of a study population will directly affect
the results of a given study.

4. Recommendations
As with other previous reviews and meta-analyses,
this assessment confirms the overall superiority
of molecular markers over conventional urine
cytology concerning test sensitivity. This improved
sensitivity is more pronounced in low grade tumors
but is also seen in high grade bladder cancer. In
contrast, urine cytology is clearly superior to that
of the molecular markers with regard to specificity,
and its role in bladder cancer diagnosis needs to
be considered in this context.

193

With the exception of NMP22, only limited additional


information has been gained for other urine-based
markers (e.g. BARD stat and TRAK, Cyfra21-1,
UBC) through the last decade. In contrast, many
reports concerning cell-based markers (UroVysion,
uCyt+) were published only after 2001. Despite
numerous publications, progress with regard
to an introduction of these markers into clinical
decision-making is limited. The first prospective
trials on a marker-guided follow-up in bladder
cancer patients are currently underway. In addition,
several screening studies could demonstrate the
feasibility of a use of molecular markers for this
indication. Results of these and other such studies
will determine their usefulness and role in clinical
practice.

Furthermore, reimbursement policy in different


countries will have additional influence on the use of
molecular markers.
In order to obtain a better idea on the performance
of a given assay, marker assessment needs to
follow a standardized and transparent evaluation
process. It remains one of the great challenges in
marker development to define a standard procedure
and, finally, introduce this standard into the scientific
community.

6. Problems in the assessment of


marker trials
The question of marker performance has been
addressed within a number of meta-analyses.
[6,33,53,151]. However, the problems of these
analyses are significant for several reasons and,
in consequence, conclusions derived from these
analyses are heavily biased. One of the key problems
is the highly differing quality of the trials, which hardly
permits common analysis. Furthermore, different
study design, patient selection, tumor prevalence,
distribution of tumor grade and stage, study
endpoints and several other parameters will further
confound the results of any combined analysis.

5. Marker performance
There is no marker that meets all of the postulates
of a so called ideal marker,[9] but markers have
been described with a high overall sensitivity, a
high sensitivity for low or high grade disease, high
specificity, reasonable expense and point-of-care
capabilities. However, it is obvious that urologists
will have to select markers that meet specific clinical
needs. In a screening scenario high specificity of a
marker is mandatory since otherwise the number of
patients undergoing a marker-initiated evaluation
will be inappropriately high. This contrasts to the
requirements in a follow-up setting, when sensitivity
of an assay is of key importance in order not to
miss bladder cancer persistence or recurrence. In
addition, the diagnostic strategy might also affect
selection of a marker: several investigators may
favor markers with good performance in low grade
disease since appr. 70% of all bladder tumors are
low grade. Other urologists may prefer markers with
a high sensitivity in high grade disease, arguing that it
may be appropriate to delay detection of a low grade
tumor that does not pose life-threatening risk but that
high grade tumors should be reliably detected.

In order to standardize the evaluation of molecular


markers several tools for assessment of diagnostic
markers were used in this analysis. These tools
included for the analysis a questionnaire for the
quality of reporting (STARD), the definition of the
level of evidence according to the Oxford criteria,
and the classification of the marker status.[12,3638]
Concerning the reporting quality, no single study
included all 25 STARD items.[37,38] Certain items,
such as use of Mesh headings to identify sensitivity,
specificity or diagnostic accuracy, or reporting
of adverse events associated with testing were
uniformly missing, although these items may have
less importance to data quality than others. While
the majority of studies included some population
demographics, race, for example, was recorded in
only one of the sampled sources. All authors provided
information on test performance techniques, and
most described collection and handling of specimens;
however, information on the reproducibility of test
results, the training and qualifications of those
performing the assays, whether testers were blinded
from other results, and handling of indeterminate,
out-lying, or missing results were often lacking or
ambiguous.[70] Many studies stratified subjects
according to grade [WHO 1998 and 2004] and
stage of tumors [TMN 1997] and provided subset
analyses of sensitivity and specificity of the tests
for these subsets. Due to mostly low numbers of
subjects in some groups, the validity of drawing
conclusions from this data is uncertain. Authors
are generally to be commended for a reasonable

Importantly, there are additional obstacles that make


it difficult to identify the most accurate test system in
any clinical situation. This finding impedes a simple
comparison of tests based upon performance data
reported in case-control studies. Several parameters
have been shown to affect the performance of a
given test.
Patient selection
Prevalence of the disease
Study design
Study endpoints
Sample size
Technique/standardization/Cut-off
Experience

194

description of statistical methods used, including


confidence intervals on reported data. Clearly,
implementation of standardized reporting of studies
that adhere to consistent guidelines such as STARD
recommendations would improve our understanding
of tumor markers.

Data obtained in high-risk groups undergoing urinary


dipstick screening for bladder cancer suggest that
the bladder tumors discovered when evaluating all
patients with asymptomatic microscopic hematuria
may be more amenable to curative treatment than
those normally encountered, thereby reducing
morbidity and mortality associated with bladder
cancer in these patients. [20,21,143-145] Since
improved survival of screened patients was not
demonstrated in a randomized fashion but only
in comparison with a cancer register, this study
presents interesting information; however, it cannot
serve to provide a final decision on the benefit of
hematuria screening.

It can be argued, that STARD guidelines are


yet imperfect. According to experiences made
throughout this assessment, some items are
differently interpreted by observers and opinions on
the necessity of including all items, and the relative
importance of certain items, was not universally
agreed upon. However, currently STARD provides
us with a starting point for collecting comparable
data among studies.

Meanwhile, further studies targeting risk populations


such as smokers and professionally exposed
individuals could demonstrate that screening for
bladder cancer using molecular markers is feasible.
However, despite selecting risk populations in most
of these trials tumor prevalence was still too low
to make bladder cancer screening a cost-effective
procedure.

Recently, a new definition for diagnostic trials was


developed for the Oxford Classification on the Levels
of Evidence.[36] This classification has been used in
this assessment but appears more difficult to apply
if compared to the recommendations for therapeutic
trials. This was partly due to deficits in reporting.
But for some studies, application of certain criteria
was not possible. Nevertheless, the new Oxford
classification on diagnostic trials is promising but
may need minor modification.

Since identification of high risk populations suited


for a screening scenario remains the key problem
for bladder cancer screening the development and
validation of respective risk calculators (risk-adapted
screening) might be an option for the future.

For definition of the stage of implementing new


markers into clinical decision making the IBCN
classification has been developed and was used
in this assessment.[12] However, using this
classification it became evident that more precise
definitions on the requirements for allocating a given
study to a certain stage are mandatory and that this
classification requires revision.

Question: (How) can molecular markers support


screening of patients at risk of having or developing
bladder cancer?
Statement:
Feasibility of bladder cancer screening has been
demonstrated in several prospective trials. The results
from one study using dip-stick testing for hematuria
suggests a survival benefit of individuals undergoing
hematuria screening. Because of weak controls in this
report validation of the results and improved definition
of risk populations suited for screening is required.

7. Key questions:
a) (How) can molecular markers support
screening of patients at risk of having or
developing bladder cancer?
Considering bladder cancer screening the key
question to be answered is if early detection of
bladder cancer may have any impact on cure rates
and, subsequently, on patient survival. Through the
last decades a growing body of evidence has been
accumulated suggesting that early detection and
treatment of bladder cancer may indeed reduce
cancer specific mortality [127-129] thus providing
arguments for this procedure.

References: [14,21,143,144]
Recommendation:
Bladder cancer screening using urine for testing is
promising but cannot be recommended at present.
LoE: 1b
Grade: B
Agreement: 92%

However, due to the low prevalence of bladder


cancer in the general population (0.001%) and in
people above the age of 50 (0.67% - 1.13%), mass
screening for bladder cancer, with the possibility
of detecting a significant number of false positives
requiring an unnecessary work-up, would certainly
not be cost-effective. [6] As a consequence of these
considerations those few trials addressing screening
for bladder cancer targeted high risk populations.

b) (How) can molecular markers be used in


reflex testing for bladder cancer?
Reflex testing e.g. in the follow-up of patients with
bladder cancer with atypical cytology finding is a
logical approach. However, experience with this
procedure at present is very limited and does not
permit a definite statement. In consequence, this
strategy should be exploited in more detail within
prospective controlled studies.

195

urine cytology with regard to test sensitivity even


in high grade bladder cancer. It remains unclear if
these results are based upon a systematic deficiency
of urine cytology or if performance quality has
decreased in the last decades due to changes in the
training of pathologists.

Question: (How) can molecular markers be used in


reflex testing for bladder cancer?
Statement:
At present experience with reflex testing is very limited
(mostly restricted to FISH technique) and therefore
does not permit a definite statement. Reflex testing
should be explored in more detail within prospective
controlled studies

The lower specificity of molecular markers as


compared to urine cytology appears not worrisome
in this population since in the surveillance of patients
with high grade tumors sensitivity appears to be
of the utmost importance. In addition, it may be
questioned if at least a part of false positive results
may be explained as an anticipatory positive finding,
predicting tumor recurrence.[28,132]

References: [28,132,142,152]
Recommendation:
Reflex testing is considered experimental at present
and should not be used within a clinical setting.
LoE: 2b
Grade: B
Agreement: 92%

However, prospective studies demonstrating an


added value of molecular markers in the follow-up of
patients with high grade bladder cancer are missing,
thus not supporting their use in clinical practice.

c) (How) can molecular markers support


follow-up of patients with superficial low risk
bladder cancer?

Question: (How) can molecular markers support


follow-up of patients with superficial high risk bladder
cancer?
Statement:
Molecular markers detect high grade bladder cancer
with high sensitivity. At this stage it remains unclear
how molecular markers can support surveillance of
patients with high grade bladder cancer.
Recommendation:
A use of molecular markers in surveillance of
patients with high grade bladder cancer cannot be
recommended.
LoE: 2b
Grade: B
Agreement: 92%

There is clear evidence that modern molecular


markers outperform urine cytology concerning
sensitivity in the diagnosis of patients with noninvasive low grade tumors. In addition, due to the
low risk of tumor progression a marker-guided
surveillance could significantly reduce the number
of control cystoscopies without placing patients
at significant risk. However, until today only one
prospective trial using a marker-guided surveillance
protocol has been performed. [130] Information from
this study, however, is still preliminary and does not
yet permit recommendation of this procedure for
clinical routine use.
Question: (How) can molecular markers support
follow-up of patients with superficial low risk bladder
cancer?
Statement:
Marker-guided follow-up of patients with non-muscleinvasive low risk tumors appears feasible. However,
studies proving the efficacy of this concept and
demonstrating an added value for patients or the
health system are lacking.
Recommendation:
Marker-guided follow-up of patients with superficial
low grade bladder cancer appears attractive; however,
based upon current levels of evidence this procedure
cannot be recommended at present
LoE: 1b
Grade: B
Agreement: 92%

8. Outlook
Although molecular bladder cancer assays have
been shown to have superior sensitivity as compared
to urine cytology, none of them has been included in
clinical guidelines. The key reason for this situation is
that none of the assays have been incorporated into
clinical decision making so far. In consequence, an
added value of molecular markers for the diagnosis
of urothelial tumors has not yet been identified.
However, the current data suggest that some of
these markers do have the potential to play a role
in screening and surveillance of bladder cancer in
the future. Current screening protocols however are
hampered by a low disease prevalence thus inhibiting
an acceptable cost/benefit ratio. The introduction of
risk calculators into screening protocols could make
up for the deficits of a mass screening approach.

d) (How) can molecular markers support


follow-up of patients with superficial high risk
bladder cancer?

Furthermore, the introduction of molecular markers


in the follow-up of patients with low risk bladder
cancer might also represent a scenario that should

The assessment of marker performance suggests


that several molecular markers may outperform

196

be further investigated. Preliminary reports suggest


that this procedure is feasible. However, detailed
information for a definite judgment is lacking.

Question: (How) can molecular markers support


follow-up of patients with superficial low risk bladder
cancer?
Statement:
Marker-guided follow-up of patients with non-muscleinvasive low risk tumors appears feasible. However,
studies proving the efficacy of this concept and
demonstrating an added value for patients or the
health system are lacking.
Recommendation:
Marker-guided follow-up of patients with superficial
low grade bladder cancer appears attractive; however,
based upon current levels of evidence this procedure

The scientific community is urged to develop


protocols and conduct prospective trials to provide
the basis for an integration of molecular markers into
clinical decision making in the future.[7,84,153]

IV. Appendix
Questions
Question: (How) can molecular markers support

cannot be recommended at present


LoE: 1b
Grade: B
Agreement: 92%

screening of patients at risk of having or developing


bladder cancer?
Statement:
Feasibility of bladder cancer screening has been
demonstrated in several prospective trials. The results
from one study using dip-stick testing for hematuria
suggests a survival benefit of individuals undergoing
hematuria screening. Because of weak controls in this
report validation of the results and improved definition
of risk populations suited for screening is required.

Question: (How) can molecular markers support


follow-up of patients with superficial high risk bladder
cancer?
Statement:
Molecular markers detect high grade bladder cancer
with high sensitivity. At this stage it remains unclear
how molecular markers can support surveillance of
patients with high grade bladder cancer.
Recommendation:
A use of molecular markers in surveillance of
patients with high grade bladder cancer cannot be
recommended.
LoE: 2b
Grade: B
Agreement: 92%

References: [14,15,21,143,144]
Recommendation:
Bladder cancer screening using urine for testing is
promising but cannot be recommended at present.
LoE: 1b
Grade: B
Agreement: 92%

Question: (How) can molecular markers be used in


reflex testing for bladder cancer?
Statement:
At present experience with reflex testing is very limited
(mostly restricted to FISH technique) and therefore
does not permit a definite statement. Reflex testing
should be explored in more detail within prospective
controlled studies
References: [28,132,142,152]
Recommendation:
Reflex testing is considered experimental at present
and should not be used within a clinical setting.
LoE: 2b
Grade: B
Agreement: 92%

197

Levels of Evidence[34]
Oxford Centre for Evidence-based Medicine Levels of Evidence (May 2001)
Level Prognosis
1a

1b

Systematic review (SR) (with


homogeneity*) of inception cohort
studies; CDR validated in
different populations
Individual inception cohort study
with > 80% follow-up; CDR
validated in a single population

1c

All or none case-series

2a

2c
3a

SR (with homogeneity*) of either


retrospective cohort studies or
untreated control groups in RCTs
Retrospective cohort study or
follow-up of untreated control
patients in an RCT; Derivation
of CDR or validated on splitsample only
Outcomes Research
-

3b

Case-series (and poor quality


prognostic cohort studies***)
Expert opinion without explicit
critical appraisal, or based on
physiology, bench research or
first principles

2b

Diagnosis

Economic and decision


analyses
SR (with homogeneity*) of Level SR (with homogeneity*) of Level 1
1 diagnostic studies; CDR with economic studies
1b studies from different clinical
centres
Validating** cohort study with Analysis based on clinically
good reference standards; sensible costs or alternatives;
or CDR tested within one clinical systematic review(s) of the
centre
evidence; and including multi-way
sensitivity analyses
Absolute SpPins and SnNouts Absolute better-value or worsevalue analyses
SR (with homogeneity*) of Level SR (with homogeneity*) of Level
>2 diagnostic studies
>2 economic studies
Exploratory** cohort study with
good reference standards;
CDR after derivation, or
validated only on split-sample
or databases
SR (with homogeneity*) of 3b and
better studies
Non-consecutive study; or without
consistently applied reference
standards

Case-control study, poor or nonindependent reference standard


Expert opinion without explicit
critical appraisal, or based on
physiology, bench research or
first principles

Analysis based on clinically


sensible costs or alternatives;
limited review(s) of the evidence,
or single studies; and including
multi-way sensitivity analyses
Audit or outcomes research
SR (with homogeneity*) of 3b and
better studies
Analysis
based
on
limited
alternatives or costs, poor quality
estimates of data, but including
sensitivity analyses incorporating
clinically sensible variations.
Analysis with no sensitivity
analysis
Expert opinion without explicit
critical appraisal, or based
on economic theory or first
principles

Produced by Bob Phillips, Chris Ball, Dave Sackett, Doug Badenoch, Sharon Straus, Brian Haynes, Martin
Dawes since November 1998. Updated by Jeremy Howick March 2009.

Notes

worrisome, and not all worrisome heterogeneity


need be statistically significant. As noted above,
studies displaying worrisome heterogeneity should
be tagged with a - at the end of their designated
level.

Users can add a minus-sign - to denote the level of


that fails to provide a conclusive answer because:
EITHER a single result with a wide Confidence
Interval

Clinical Decision Rule. (These are algorithms


or scoring systems that lead to a prognostic
estimation or a diagnostic category.)

OR a Systematic Review with troublesome


heterogeneity.

See note above for advice on how to understand,


rate and use trials or other studies with wide
confidence intervals.

Such evidence is inconclusive, and therefore can


only generate Grade D recommendations.
* By homogeneity we mean a systematic review
that is free of worrisome variations (heterogeneity)
in the directions and degrees of results between
individual studies. Not all systematic reviews with
statistically significant heterogeneity need be

Met when all patients died before the Rx became


available, but some now survive on it; or when some
patients died before the Rx became available, but
none now die on it.

198

By poor quality cohort study we mean one that


failed to clearly define comparison groups and/
or failed to measure exposures and outcomes in
the same (preferably blinded), objective way in
both exposed and non-exposed individuals and/
or failed to identify or appropriately control known
confounders and/or failed to carry out a sufficiently
long and complete follow-up of patients. By poor
quality case-control study we mean one that
failed to clearly define comparison groups and/
or failed to measure exposures and outcomes in
the same (preferably blinded), objective way in
both cases and controls and/or failed to identify
or appropriately control known confounders.

applied to all patients. Poor reference standards


are haphazardly applied, but still independent
of the test. Use of a non-independent reference
standard (where the test is included in the
reference, or where the testing affects the
reference) implies a level 4 study.
B
 etter-value treatments are clearly as good
but cheaper, or better at the same or reduced
cost. Worse-value treatments are as good and
more expensive, or worse and the equally or
more expensive.
** V
 alidating studies test the quality of a specific
diagnostic test, based on prior evidence. An
exploratory study collects information and trawls
the data (e.g. using a regression analysis) to find
which factors are significant.

Split-sample validation is achieved by collecting


all the information in a single tranche, then
artificially dividing this into derivation and
validation samples.

*** B
 y poor quality prognostic cohort study we mean
one in which sampling was biased in favour of
patients who already had the target outcome, or
the measurement of outcomes was accomplished
in <80% of study patients, or outcomes were
determined in an unblinded, non-objective way, or
there was no correction for confounding factors.

A
 n Absolute SpPin is a diagnostic finding
whose Specificity is so high that a Positive result
rules-in the diagnosis. An Absolute SnNout is
a diagnostic finding whose Sensitivity is so high
that a Negative result rules-out the diagnosis.
Good, better, bad and worse refer to the
comparisons between treatments in terms of
their clinical risks and benefits.

**** G
 ood follow-up in a differential diagnosis study
is >80%, with adequate time for alternative
diagnoses to emerge (for example 1-6 months
acute, 1 - 5 years chronic)

G
 ood reference standards are independent of
the test, and applied blindly or objectively to

199

STARD checklist for reporting of studies of diagnostic accuracy


(version January 2003)

Section and Topic

Item
On
#
page #
TITLE/ABSTRACT/
1 Identify the article as a study of diagnostic accuracy (recommend MeSH
KEYWORDS
heading <sensitivity and specificity>).
INTRODUCTION
2 State the research questions or study aims, such as estimating diagnostic
accuracy or comparing accuracy between tests or across participant groups.
METHODS
Participants
3 The study population: The inclusion and exclusion criteria, setting and
locations where data were collected.
4 Participant recruitment: Was recruitment based on presenting symptoms,
results from previous tests, or the fact that the participants had received the
index tests or the reference standard?
5 Participant sampling: Was the study population a consecutive series of
participants defined by the selection criteria in item 3 and 4? If not, specify
how participants were further selected.
6 Data collection: Was data collection planned before the index test and reference
standard were performed (prospective study) or after (retrospective study)?
Test methods
7 The reference standard and its rationale.
8 Technical specifications of material and methods involved including how and
when measurements were taken, and/or cite references for index tests and
reference standard.
9 Definition of and rationale for the units, cut-offs and/or categories of the
results of the index tests and the reference standard.
10 The number, training and expertise of the persons executing and reading the
index tests and the reference standard.
11 Whether or not the readers of the index tests and reference standard were
blind (masked) to the results of the other test and describe any other clinical
information available to the readers.
Statistical methods 12 Methods for calculating or comparing measures of diagnostic accuracy, and
the statistical methods used to quantify uncertainty (e.g. 95% confidence
intervals).
13 Methods for calculating test reproducibility, if done.
RESULTS
Participants
14 When study was performed, including beginning and end dates of recruitment.
15 Clinical and demographic characteristics of the study population (at least
information on age, gender, spectrum of presenting symptoms).
16 The number of participants satisfying the criteria for inclusion who did or
did not undergo the index tests and/or the reference standard; describe
why participants failed to undergo either test (a flow diagram is strongly
recommended).
Test results
17 Time-interval between the index tests and the reference standard, and any
treatment administered in between.
18 Distribution of severity of disease (define criteria) in those with the target
condition; other diagnoses in participants without the target condition.
19 A cross tabulation of the results of the index tests (including indeterminate
and missing results) by the results of the reference standard; for continuous
results, the distribution of the test results by the results of the reference
standard.
20 Any adverse events from performing the index tests or the reference standard.
Estimates
21 Estimates of diagnostic accuracy and measures of statistical uncertainty
(e.g. 95% confidence intervals).
22 How indeterminate results, missing data and outliers of the index tests were
handled.
23 Estimates of variability of diagnostic accuracy between subgroups of
participants, readers or centers, if done.
24 Estimates of test reproducibility, if done.
DISCUSSION
25 Discuss the clinical applicability of the study findings.

200

References

19. Grossfeld GD, Carroll PR: Evaluation of asymptomatic


microscopic hematuria. Urol Clin North Am 25:661-76,
1998

1. Jemal A, Siegel R, Xu J, et al: Cancer statistics, 2010. CA


Cancer J Clin 60:277-300, 2010

20. Madeb R, Golijanin D, Knopf J, et al: Long-term outcome


of patients with a negative work-up for asymptomatic
microhematuria. Urology 75:20-5
21. Madeb R, Messing EM: Long-term outcome of home
dipstick testing for hematuria. World J Urol 26:19-24, 2008

2. Prout GR, Barton BA, Griffin PP, et al: Treated history


of noninvasive grade 1 transitional cell carcinoma. The
National Bladder Cancer Group. J Urol 148:1413-9, 1992

22. Malmstrom PU: Time to abandon testing for microscopic


haematuria in adults? Bmj 326:813-5, 2003

3. Pagano F, Bassi P, Galetti TP, et al: Results of


contemporary radical cystectomy for invasive bladder
cancer: a clinicopathological study with an emphasis
on the inadequacy of the tumor, nodes and metastases
classification. J Urol 145:45-50, 1991

23. Nieder AM, Lotan Y, Nuss GR, et al: Are patients with
hematuria appropriately referred to Urology? A multiinstitutional questionnaire based survey. Urol Oncol, 2008
24. Johnson EK, Daignault S, Zhang Y, et al: Patterns of
hematuria referral to urologists: does a gender disparity
exist? Urology 72:498-502; discussion 502-3, 2008

4. Botteman MF, Pashos CL, Redaelli A, et al: The health


economics of bladder cancer: a comprehensive review of
the published literature. Pharmacoeconomics 21:1315-30,
2003

25. Murta-Nascimento C, Schmitz-Drger B, Zeegers M, et


al: Epidemiology of urinary bladder cancer: from tumor
development to patients death. World Journal of Urology
25:285-295, 2007

5. Lotan Y, Kamat AM, Porter MP, et al: Key concerns about


the current state of bladder cancer: a position paper
from the Bladder Cancer Think Tank, the Bladder Cancer
Advocacy Network, and the Society of Urologic Oncology.
Cancer 115:4096-103, 2009

26. Boffetta P, Silverman DT: A Meta-Analysis of Bladder


Cancer and Diesel Exhaust Exposure. Epidemiology
12:125-130, 2001

6. Lotan Y, Svatek RS, Malats N: Screening for bladder


cancer: a perspective. World J Urol 26:13-8, 2008

27. Zeegers MP, Tan FE, Dorant E, et al: The impact of


characteristics of cigarette smoking on urinary tract cancer
risk: a meta-analysis of epidemiologic studies. Cancer
89:630-9, 2000

7. Shariat SF, Lotan Y, Vickers A, et al: Statistical consideration


for clinical biomarker research in bladder cancer. Urol Oncol
28:389-400

28. Yoder BJ, Skacel M, Hedgepeth R, et al: Reflex UroVysion


testing of bladder cancer surveillance patients with
equivocal or negative urine cytology: a prospective study
with focus on the natural history of anticipatory positive
findings. Am J Clin Pathol 127:295-301, 2007

8. Lotan Y, Roehrborn CG: Sensitivity and specificity of


commonly available bladder tumor markers versus
cytology: results of a comprehensive literature review and
meta-analyses. Urology 61:109-18; discussion 118, 2003
9. Bensalah K, Montorsi F, Shariat SF: Challenges of cancer
biomarker profiling. Eur Urol 52:1601-9, 2007

29. Lotan Y, Bensalah K, Ruddell T, et al: Prospective


evaluation of the clinical usefulness of reflex fluorescence
in situ hybridization assay in patients with atypical cytology
for the detection of urothelial carcinoma of the bladder. J
Urol 179:2164-9, 2008

10. Babjuk M, Oosterlinck W, Sylvester R, et al: EAU Guidelines


on Non-Muscle-Invasive Urothelial Carcinoma of the
Bladder, the 2011 Update. Eur Urol 59:997-1008

30. Schlomer BJ, Ho R, Sagalowsky A, et al: Prospective


validation of the clinical usefulness of reflex fluorescence in
situ hybridization assay in patients with atypical cytology for
the detection of urothelial carcinoma of the bladder. J Urol
183:62-7

11. Hall MC, Chang SS, Dalbagni G, et al: Guideline for the
management of nonmuscle invasive bladder cancer (stages
Ta, T1, and Tis): 2007 update. J Urol 178:2314-30, 2007
12. Goebell PJ, Groshen SL, Schmitz-Drager BJ: Guidelines
for development of diagnostic markers in bladder cancer.
World J Urol 26:5-11, 2008

31. Schmitz-Drger BJ, Fradet Y, Grossman HB: Bladder


cancer markers in patient management: the current
perspective. World J Urol 26:1-3, 2008

13. Shariat SF, Canto EI, Kattan MW, et al: Beyond prostatespecific antigen: new serologic biomarkers for improved
diagnosis and management of prostate cancer. Rev Urol
6:58-72, 2004

32. Grossman HB, Soloway M, Messing E, et al: Surveillance


for recurrent bladder cancer using a point-of-care proteomic
assay. JAMA 295:299-305, 2006

14. Steiner H, Bergmeister M, Verdorfer I, et al: Early results of


bladder-cancer screening in a high-risk population of heavy
smokers. BJU Int 102:291-6, 2008
15. Lotan Y, Elias K, Svatek RS, et al: Bladder cancer screening
in a high risk asymptomatic population using a point of
care urine based protein tumor marker. J Urol 182:52-7;
discussion 58, 2009

33. Hajdinjak T: UroVysion FISH test for detecting urothelial


cancers: meta-analysis of diagnostic accuracy and
comparison with urinary cytology testing. Urol Oncol
26:646-51, 2008
34. van Rhijn BWG, van der Poel HG, van der Kwast TH: Urine
Markers for Bladder Cancer Surveillance: A Systematic
Review. European Urology 47:736-748, 2005

16. Schmitz-Drager BJ, Tirsar LA, Schmitz-Drager C, et


al: Immunocytology in the assessment of patients with
asymptomatic hematuria. World J Urol 26:31-7, 2008
17. Schmitz-Drager BJ, Beiche B, Tirsar LA, et al:
Immunocytology in the assessment of patients with
asymptomatic microhaematuria. Eur Urol 51:1582-8;
discussion 1588, 2007
18. Grossfeld GD, Litwin MS, Wolf JS, Jr., et al: Evaluation
of asymptomatic microscopic hematuria in adults: the
American Urological Association best practice policy--part
II: patient evaluation, cytology, voided markers, imaging,
cystoscopy, nephrology evaluation, and follow-up. Urology
57:604-10, 2001

35. Shariat S, Lotan L, Schmitz-Drger B, et al: Molecular


Markers in the Management of Patients with Bladder
Cancer: I. Prognostic molecular markers for bladder cancer,
2010
36. Oxford: Oxford Centre for Evidence-based Medicine Levels
of Evidence, May 2001
37. Bossuyt PM, Reitsma JB, Bruns DE, et al: Towards
complete and accurate reporting of studies of diagnostic
accuracy: the STARD initiative. Standards for Reporting of
Diagnostic Accuracy. Clin Chem 49:1-6, 2003
38. Bossuyt PM, Reitsma JB, Bruns DE, et al: The STARD

201

statement for reporting studies of diagnostic accuracy:


explanation and elaboration. Clin Chem 49:7-18, 2003

56. Konety BR, Getzenberg RH: Urine based markers of


urological malignancy. J Urol 165:600-11, 2001

39. McCormick KA, Fleming B: Clinical practice guidelines. The


Agency for Health Care Policy and Research fosters the
development of evidence-based guidelines. Health Prog
73:30-4, 1992

57. Kinders R, Jones T, Root R, et al: Complement factor H or a


related protein is a marker for transitional cell cancer of the
bladder. Clin Cancer Res 4:2511-20, 1998
58. Malkowicz SB: The application of human complement
factor H-related protein (BTA TRAK) in monitoring patients
with bladder cancer. Urol Clin North Am 27:63-73, ix, 2000

40. Boman H, Hedelin H, Holmng S: Four bladder tumor


markers have a disappointingly low sensitivity for small size
and low grade recurrence. J Urol 167:80-3, 2002

59. Lokeshwar VB, Soloway MS: Current bladder tumor tests:


does their projected utility fulfill clinical necessity? J Urol
165:1067-77, 2001

41. Snchez-Carbayo M, Herrero E, Megas J, et al:


Comparative sensitivity of urinary CYFRA 21-1, urinary
bladder cancer antigen, tissue polypeptide antigen, tissue
polypeptide antigen and NMP22 to detect bladder cancer. J
Urol 162:1951-6, 1999

60. Lokeshwar VB, Habuchi T, Grossman HB, et al: Bladder


tumor markers beyond cytology: International Consensus
Panel on bladder tumor markers. Urology 66:35-63, 2005

42. Poulakis V, Witzsch U, De Vries R, et al: A comparison


of urinary nuclear matrix protein-22 and bladder tumour
antigen tests with voided urinary cytology in detecting and
following bladder cancer: the prognostic value of falsepositive results. BJU Int 88:692-701, 2001

61. Ellis WJ, Blumenstein BA, Ishak LM, et al: Clinical


evaluation of the BTA TRAK assay and comparison to
voided urine cytology and the Bard BTA test in patients with
recurrent bladder tumors. The Multi Center Study Group.
Urology 50:882-7, 1997

43. Grossman HB, Messing E, Soloway M, et al: Detection


of bladder cancer using a point-of-care proteomic assay.
JAMA 293:810-6, 2005

62. Wiener HG, Mian C, Haitel A, et al: Can urine bound


diagnostic tests replace cystoscopy in the management of
bladder cancer? J Urol 159:1876-80, 1998

44. Shariat SF, Karam JA, Lotan Y, et al: Critical evaluation


of urinary markers for bladder cancer detection and
monitoring. Rev Urol 10:120-35, 2008

63. Leyh H, Marberger M, Conort P, et al: Comparison of the


BTA stat test with voided urine cytology and bladder wash
cytology in the diagnosis and monitoring of bladder cancer.
Eur Urol 35:52-6, 1999

45. Sharma S, Zippe CD, Pandrangi L, et al: Exclusion criteria


enhance the specificity and positive predictive value of
NMP22 and BTA stat. J Urol 162:53-7, 1999

64. Pode D, Shapiro A, Wald M, et al: Noninvasive detection


of bladder cancer with the BTA stat test. J Urol 161:443-6,
1999

46. Sarosdy MF, Schellhammer P, Bokinsky G, et al: Clinical


evaluation of a multi-target fluorescent in situ hybridization
assay for detection of bladder cancer. J Urol 168:1950-4,
2002

65. Szen S, Biri H, Sinik Z, et al: Comparison of the nuclear


matrix protein 22 with voided urine cytology and BTA stat
test in the diagnosis of transitional cell carcinoma of the
bladder. Eur Urol 36:225-9, 1999

47. Halling KC, King W, Sokolova IA, et al: A comparison of


cytology and fluorescence in situ hybridization for the
detection of urothelial carcinoma. J Urol 164:1768-75,
2000

66. Ramakumar S, Bhuiyan J, Besse JA, et al: Comparison of


screening methods in the detection of bladder cancer. J
Urol 161:388-94, 1999

48. Bubendorf L, Grilli B, Sauter G, et al: Multiprobe FISH


for enhanced detection of bladder cancer in voided urine
specimens and bladder washings. Am J Clin Pathol 116:7986, 2001

67. Nasuti JF, Gomella LG, Ismial M, et al: Utility of the BTA
stat test kit for bladder cancer screening. Diagn Cytopathol
21:27-9, 1999
68. Heicappell R, Wettig IC, Schostak M, et al: Quantitative
detection of human complement factor H-related protein in
transitional cell carcinoma of the urinary bladder. Eur Urol
35:81-7, 1999

49. Giannopoulos A, Manousakas T, Mitropoulos D, et al:


Comparative evaluation of the BTAstat test, NMP22,
and voided urine cytology in the detection of primary and
recurrent bladder tumors. Urology 55:871-5, 2000

69. Serretta V, Pomara G, Rizzo I, et al: Urinary BTA-stat,


BTA-trak and NMP22 in surveillance after TUR of recurrent
superficial transitional cell carcinoma of the bladder. Eur
Urol 38:419-25, 2000
70. Raitanen MP, Marttila T, Nurmi M, et al: Human complement
factor H related protein test for monitoring bladder cancer. J
Urol 165:374-7, 2001
71. Babjuk M, Kostrov M, Mudra K, et al: Qualitative and
quantitative detection of urinary human complement factor
H-related protein (BTA stat and BTA TRAK) and fragments
of cytokeratins 8, 18 (UBC rapid and UBC IRMA) as
markers for transitional cell carcinoma of the bladder. Eur
Urol 41:34-9, 2002
72. Schroeder GL, Lorenzo-Gomez MF, Hautmann SH, et al:
A side by side comparison of cytology and biomarkers for
bladder cancer detection. J Urol 172:1123-6, 2004
73. Tsui KH, Chen SM, Wang TM, et al: Comparisons of voided
urine cytology, nuclear matrix protein-22 and bladder tumor
associated antigen tests for bladder cancer of geriatric male
patients in Taiwan, China. Asian J Androl 9:711-5, 2007
74. Babjuk M, Soukup V, Pesl M, et al: Urinary cytology and
quantitative BTA and UBC tests in surveillance of patients
with pTapT1 bladder urothelial carcinoma. Urology 71:71822, 2008

50. Moonen PM, Merkx GF, Peelen P, et al: UroVysion


compared with cytology and quantitative cytology in the
surveillance of non-muscle-invasive bladder cancer. Eur
Urol 51:1275-80; discussion 1280, 2007
51. Goebell PJ, Groshen S, Schmitz-Drager BJ, et al: The
International Bladder Cancer Bank: proposal for a new
study concept. Urol Oncol 22:277-84, 2004
52. Glas AS, Roos D, Deutekom M, et al: Tumor markers in the
diagnosis of primary bladder cancer. A systematic review. J
Urol 169:1975-82, 2003
53. Mowatt G, Zhu S, Kilonzo M, et al: Systematic review
of the clinical effectiveness and cost-effectiveness of
photodynamic diagnosis and urine biomarkers (FISH,
ImmunoCyt, NMP22) and cytology for the detection and
follow-up of bladder cancer. Health Technol Assess 14:1331, iii-iv, 2010
54. Villicana P, Whiting B, Goodison S, et al: Urine-based
assays for the detection of bladder cancer. Biomark Med
3:265, 2009
55. Sarosdy MF, Hudson MA, Ellis WJ, et al: Improved detection
of recurrent bladder cancer using the Bard BTA stat Test.
Urology 50:349-53, 1997

202

75. Eissa S, Zohny S, Zekri A, et al: Diagnostic value of


fibronectin and mutant p53 in the urine of patients with
bladder cancer: impact on clinicopathological features and
disease recurrence. Med Oncol, 2009

93. Bonner RB, Liebert M, Hurst RE, et al: Characterization


of the DD23 tumor-associated antigen for bladder cancer
detection and recurrence monitoring. Marker Network
for Bladder Cancer. Cancer Epidemiol Biomarkers Prev
5:971-8, 1996

76. Blumenstein BA, Ellis WJ, Ishak LM: The relationship


between serial measurements of the level of a bladder
tumor associated antigen and the potential for recurrence.
J Urol 161:57-60; discussion 60-1, 1999

94. Sawczuk IS, Pickens CL, Vasa UR, et al: DD23 Biomarker:
a prospective clinical assessment in routine urinary cytology
specimens from patients being monitored for TCC. Urol
Oncol 7:185-90, 2002

77. Lokeshwar V, Habuchi T, Grossman H, et al: Cytology and


tumor markers: Tumor markers beyond cytology, Bladder
tumors(Soloway MS, Carmack A, Khoury S eds0Health
Publication Ltd, 2006

95. Gilbert SM, Veltri RW, Sawczuk A, et al: Evaluation of


DD23 as a marker for detection of recurrent transitional
cell carcinoma of the bladder in patients with a history of
bladder cancer. Urology 61:539-43, 2003

78. Snchez-Carbayo M, Herrero E, Megas J, et al: Initial


evaluation of the new urinary bladder cancer rapid test in
the detection of transitional cell carcinoma of the bladder.
Urology 54:656-61, 1999

96. Beiche B, Ebert T, Schmitz-Drger B: Immunzytologie in der


Diagnostik des Urothelkarzinoms ein reproduzierbares
Testverfahren?, Urologe A, 2002, pp 45

79. Hedelin H, Jonsson K, Salomonsson K, et al: Screening for


bladder tumours in men aged 60-70 years with a bladder
tumour marker (UBC) and dipstick-detected haematuria
using both white-light and fluorescence cystoscopy. Scand
J Urol Nephrol 40:26-30, 2006

97. Schmitz-Drager B, Tirsar LA, Schmitz-Drager C, et al:


Immunocytology in the assessment of patients with painless
gross haematuria. BJU Int 101:455-8, 2008
98. Hemstreet GP, 3rd, Yin S, Ma Z, et al: Biomarker risk
assessment and bladder cancer detection in a cohort
exposed to benzidine. J Natl Cancer Inst 93:427-36, 2001

80. Sharp JD, Hausladen DA, Maher MG, et al: Bladder cancer
detection with urinary survivin, an inhibitor of apoptosis.
Front Biosci 7:e36-41, 2002

99. Veeramachaneni R, Nordberg ML, Shi R, et al: Evaluation


of fluorescence in situ hybridization as an ancillary tool to
urine cytology in diagnosing urothelial carcinoma. Diagn
Cytopathol 28:301-7, 2003

81. Schultz IJ, Witjes JA, Swinkels DW, et al: Bladder cancer
diagnosis and recurrence prognosis: comparison of
markers with emphasis on survivin. Clin Chim Acta 368:2032, 2006

100. Hemstreet GP, Yin S, Ma Z, et al: Biomarker risk


assessment and bladder cancer detection in a cohort
exposed to benzidine. J Natl Cancer Inst 93:427-36, 2001

82. Margulis V, Lotan Y, Shariat SF: Survivin: a promising


biomarker for detection and prognosis of bladder cancer.
World J Urol 26:59-65, 2008

101. Caraway NP, Khanna A, Fernandez RL, et al: Fluorescence


in situ hybridization for detecting urothelial carcinoma: a
clinicopathologic study. Cancer Cytopathol 118:259-68

83. Davies B, Chen JJ, McMurry T, et al: Efficacy of BTA stat,


cytology, and survivin in bladder cancer surveillance over 5
years in patients with spinal cord injury. Urology 66:908-11,
2005

102. Frigerio S, Padberg BC, Strebel RT, et al: Improved


detection of bladder carcinoma cells in voided urine by
standardized microsatellite analysis. Int J Cancer 121:32938, 2007

84. Shariat SF, Karam JA, Lerner SP: Molecular markers in


bladder cancer. Curr Opin Urol 18:1-8, 2008

103. Mian C, Mazzoleni G, Vikoler S, et al: Fluorescence in


situ hybridisation in the diagnosis of upper urinary tract
tumours. Eur Urol 58:288-92

85. Getzenberg RH, Konety BR, Oeler TA, et al: Bladder


cancer-associated nuclear matrix proteins. Cancer Res
56:1690-4, 1996
86. Konety BR, Nguyen TS, Brenes G, et al: Clinical usefulness
of the novel marker BLCA-4 for the detection of bladder
cancer. J Urol 164:634-9, 2000

104. Moonen PM, Merkx GF, Peelen P, et al: UroVysion


compared with cytology and quantitative cytology in the
surveillance of non-muscle-invasive bladder cancer. Eur
Urol 51:1275-80; discussion 1280, 2007

87. Konety BR, Nguyen TS, Dhir R, et al: Detection of bladder


cancer using a novel nuclear matrix protein, BLCA-4. Clin
Cancer Res 6:2618-25, 2000

105. Riesz P, Lotz G, Paska C, et al: Detection of bladder cancer


from the urine using fluorescence in situ hybridization
technique. Pathol Oncol Res 13:187-94, 2007

88. Van Le TS, Miller R, Barder T, et al: Highly specific urinebased marker of bladder cancer. Urology 66:1256-60,
2005

106. Youssef RF, Schlomer BJ, Ho R, et al: Role of fluorescence


in situ hybridization in bladder cancer surveillance of
patients with negative cytology. Urol Oncol

89. Shiff C, Veltri R, Naples J, et al: Ultrasound verification


of bladder damage is associated with known biomarkers
of bladder cancer in adults chronically infected with
Schistosoma haematobium in Ghana. Trans R Soc Trop
Med Hyg 100:847-54, 2006

107. Mian C, Pycha A, Wiener H, et al: Immunocyt: a new tool


for detecting transitional cell cancer of the urinary tract. J
Urol 161:1486-9, 1999
108. Olsson H, Zackrisson B: ImmunoCyt a useful method in
the follow-up protocol for patients with urinary bladder
carcinoma. Scand J Urol Nephrol 35:280-2, 2001

90. Guo B, Che T, Shi B, et al: Screening and identification


of specific markers for bladder transitional cell carcinoma
from urine urothelial cells with suppressive subtractive
hybridization and cDNA microarray. Can Urol Assoc J,
2011

109. Lodde M, Mian C, Wiener H, et al: Detection of upper


urinary tract transitional cell carcinoma with ImmunoCyt:
a preliminary report. Urology 58:362-6, 2001

91. Nisman B, Barak V, Shapiro A, et al: Evaluation of urine


CYFRA 21-1 for the detection of primary and recurrent
bladder carcinoma. Cancer 94:2914-22, 2002

110. Mian C, Lodde M, Comploj E, et al: Liquid-based cytology


as a tool for the performance of uCyt+ and Urovysion
Multicolour-FISH in the detection of urothelial carcinoma.
Cytopathology 14:338-42, 2003

92. Snchez-Carbayo M, Urrutia M, Gonzlez de Buitrago JM,


et al: Utility of serial urinary tumor markers to individualize
intervals between cystoscopies in the monitoring of patients
with bladder carcinoma. Cancer 92:2820-8, 2001

111. Feil G, Zumbragel A, Paulgen-Nelde HJ, et al: Accuracy


of the ImmunoCyt assay in the diagnosis of transitional
cell carcinoma of the urinary bladder. Anticancer Res
23:963-7, 2003

203

112. Piaton E, Daniel L, Verriele V, et al: Improved


detection of urothelial carcinomas with fluorescence
immunocytochemistry (uCyt+ assay) and urinary cytology:
results of a French Prospective Multicenter Study. Lab
Invest 83:845-52, 2003

130. Mian C, Maier K, Comploj E, et al: uCyt+/ImmunoCyt in


the detection of recurrent urothelial carcinoma: an update
on 1991 analyses. Cancer 108:60-5, 2006
131. Lodde M, Mian C, Comploj E, et al: uCyt+ test: alternative
to cystoscopy for less-invasive follow-up of patients with
low risk of urothelial carcinoma. Urology 67:950-4, 2006

113. Pfister C, Chautard D, Devonec M, et al: Immunocyt test


improves the diagnostic accuracy of urinary cytology:
results of a French multicenter study. J Urol 169:921-4,
2003

132. Skacel M, Fahmy M, Brainard JA, et al: Multitarget


fluorescence in situ hybridization assay detects transitional
cell carcinoma in the majority of patients with bladder
cancer and atypical or negative urine cytology. J Urol
169:2101-5, 2003

114. Hautmann S, Toma M, Lorenzo Gomez MF, et al:


Immunocyt and the HA-HAase urine tests for the detection
of bladder cancer: a side-by-side comparison. Eur Urol
46:466-71, 2004

133. Ferra S, Denley R, Herr H, et al: Reflex UroVysion testing


in suspicious urine cytology cases. Cancer 117:7-14,
2009

115. Toma MI, Friedrich MG, Hautmann SH, et al: Comparison


of the ImmunoCyt test and urinary cytology with other
urine tests in the detection and surveillance of bladder
cancer. World J Urol 22:145-9, 2004

134. Pycha A, Lodde M, Comploj E, et al: Intermediate-risk


urothelial carcinoma: an unresolved problem? Urology
63:472-5, 2004

116. Tetu B, Tiguert R, Harel F, et al: ImmunoCyt/uCyt+


improves the sensitivity of urine cytology in patients
followed for urothelial carcinoma. Mod Pathol 18:83-9,
2005

135. Bergman J, Reznichek RC, Rajfer J: Surveillance of


patients with bladder carcinoma using fluorescent in-situ
hybridization on bladder washings. BJU Int 101:26-9,
2008

117. Messing EM, Teot L, Korman H, et al: Performance of


urine test in patients monitored for recurrence of bladder
cancer: a multicenter study in the United States. J Urol
174:1238-41, 2005

136. Junker K, Fritsch T, Hartmann A, et al: Multicolor


fluorescence in situ hybridization (M-FISH) on cells from
urine for the detection of bladder cancer. Cytogenet
Genome Res 114:279-83, 2006

118. Lodde M, Mian C, Comploj E, et al: uCyt+ test: alternative


to cystoscopy for less-invasive follow-up of patients with
low risk of urothelial carcinoma. Urology 67:950-4, 2006

137. Kipp BR, Karnes RJ, Brankley SM, et al: Monitoring


intravesical therapy for superficial bladder cancer using
fluorescence in situ hybridization. J Urol 173:401-4, 2005

119. Mian C, Maier K, Comploj E, et al: uCyt+/ImmunoCyt in


the detection of recurrent urothelial carcinoma: an update
on 1991 analyses. Cancer 108:60-5, 2006

138. Krause FS, Rauch A, Schrott KM, et al: Clinical decisions


for treatment of different staged bladder cancer based
on multitarget fluorescence in situ hybridization assays?
World J Urol 24:418-22, 2006

120. Mian C, Lodde M, Comploj E, et al: The value of the


ImmunoCyt/uCyt+ test in the detection and follow-up of
carcinoma in situ of the urinary bladder. Anticancer Res
25:3641-4, 2005

139. Laudadio J, Keane TE, Reeves HM, et al: Fluorescence in


situ hybridization for detecting transitional cell carcinoma:
implications for clinical practice. BJU Int 96:1280-5, 2005

121. Sullivan PS, Nooraie F, Sanchez H, et al: Comparison of


ImmunoCyt, UroVysion, and urine cytology in detection
of recurrent urothelial carcinoma: a split-sample study.
Cancer 117:167-73, 2009

140. Placer J, Espinet B, Salido M, et al: Clinical utility of a


multiprobe FISH assay in voided urine specimens for the
detection of bladder cancer and its recurrences, compared
with urinary cytology. Eur Urol 42:547-52, 2002

122. Soyuer I, Sofikerim M, Tokat F, et al: Which urine marker


test provides more diagnostic value in conjunction with
standard cytology- ImmunoCyt/uCyt+ or Cytokeratin 20
expression. Diagn Pathol 4:20, 2009

141. Varella-Garcia M, Akduman B, Sunpaweravong P, et al:


The UroVysion fluorescence in situ hybridization assay
is an effective tool for monitoring recurrence of bladder
cancer. Urol Oncol 22:16-9, 2004

123. Horstmann M, Patschan O, Hennenlotter J, et al:


Combinations of urine-based tumour markers in bladder
cancer surveillance. Scand J Urol Nephrol 43:461-6,
2009

142. Gofrit ON, Zorn KC, Silvestre J, et al: The predictive value
of multi-targeted fluorescent in-situ hybridization in patients
with history of bladder cancer. Urol Oncol 26:246-9, 2008
143. Messing EM, Young TB, Hunt VB, et al: Comparison of
bladder cancer outcome in men undergoing hematuria
home screening versus those with standard clinical
presentations. Urology 45:387-96; discussion 396-7,
1995

124. Schmitz-Drager BJ, Tirsar LA, Schmitz-Drager C, et al:


[Analyses of the role of immunocytology in the differential
diagnosis of patients with asymptomatic microhematuria].
Urologe A 47:190-4, 2008
125. Schmitz-Drager BJ, Tirsar LA, Schmitz-Drager C, et al:
[Role of immunocytology in the evaluation of patients with
painless gross hematuria]. Urologe A 49:741-6

144. Messing EM, Young TB, Hunt VB, et al: Hematuria home
screening: repeat testing results. J Urol 154:57-61, 1995

126. Li HX, Li M, Li CL, et al: ImmunoCyt and cytokeratin


20 immunocytochemistry as adjunct markers for urine
cytologic detection of bladder cancer: a prospective study.
Anal Quant Cytol Histol 32:45-52

145. Messing EM, Madeb R, Young T, et al: Long-term outcome


of hematuria home screening for bladder cancer in men.
Cancer 107:2173-9, 2006
146. Hedelin H, Jonsson K, Salomonsson K, et al: Screening for
bladder tumours in men aged 60-70 years with a bladder
tumour marker (UBC) and dipstick-detected haematuria
using both white-light and fluorescence cystoscopy. Scand
J Urol Nephrol 40:26-30, 2006

127. Gore JL, Lai J, Setodji CM, et al: Mortality increases when
radical cystectomy is delayed more than 12 weeks: results
from a Surveillance, Epidemiology, and End ResultsMedicare analysis. Cancer 115:988-96, 2009
128. Wallace DM, Bryan RT, Dunn JA, et al: Delay and survival
in bladder cancer. BJU Int 89:868-78, 2002

147. Schmitz-Drger B, L-A Schwentner, CHennenlotter,


JStenzl, AMian, CMartini, TLodde, MCha, EKShariat,
SF: What is behind asymptomatic microhematuria?
Comparison of 3 contemporary cohorts, 2010

129. Hollenbeck BK, Dunn RL, Ye Z, et al: Delays in diagnosis


and bladder cancer mortality. Cancer 116:5235-42, 2010

204

148. Giberti C, Gallo F, Schenone M, et al: Early results of


urothelial carcinoma screening in a risk population of
coke workers: urothelial carcinoma among coke workers.
Biomed Environ Sci 23:300-4, 2010
149. Feil G, Sievert K, Nasterlack M, et al: Early diagnosis of
bladder cancer in high-risk populations with urine-based
tumor marker tests Interim data of the prospective study
UROSCREEN, Annual Meeting of the American Urological
Association 2010, 2010
150. Pesch B, Nasterlack M, Eberle F, et al: The role of
haematuria in bladder cancer screening among men with
former occupational exposure to aromatic amines. BJU
Int, 2011
151. Greene KL, Berry A, Konety BR: Diagnostic Utility of
the ImmunoCyt/uCyt+ Test in Bladder Cancer. Rev Urol
8:190-7, 2006
152. Ferra S, Denley R, Herr H, et al: Reflex UroVysion testing
in suspicious urine cytology cases. Cancer 117:7-14,
2009
153. Shariat SF, Margulis V, Lotan Y, et al: Nomograms for
bladder cancer. Eur Urol 54:41-53, 2008

205

206

Committee 3 B

Molecular Markers for Bladder


Cancer Staging and Prognosis
Chairs:
Shahrokh F. Shariat,
Pierre I. Karakiewicz

Members:
Michael J. Droller, Yves Fradet, George P. Hemstreet,
MLiss Hudson, Yair Lotan, Peruno Malmstrom,
Michael J. Marberger, Osamu Ogawa, Bernd J. Schmitz-Drger,
Christian Seitz, Maxine Sun, Bas W. Van Rhijn, Vinata B. Lokeshwar

Address all correspondence to:


S. F. SHARIAT
Weill Medical College of Cornell University
New York Presbyterian Hospital,
525 East 68th Street
New York, NY 10021
Email: sfshariat@gmail.com
Telephone: 469 363 8500

207

Table of Contents
I. Introduction

IV. Promising Blood-based


Prognostic Molecular
Markers

II. Evidence Acquisition

1. P
 lasma-insulin Growth Factor
Binding Protein (IGFBP)-3

III. Evidence Synthesis

2. T
 ransforming Growth Factor
(TGF)-1

1. Challenges and Statistical


Considerations for Bladder
Cancer Prognostic Molecular
Marker Research

3. Matrix Metalloproteinase (MMP)


4. Urokinase Plasminogen Activation
(uPA)

2. Promising Tissue-based Prognostic


Molecular Markers

5. Interleukin-6 (IL-6) and IL-6 Soluble


Receptor (IL-6sR)

3. Apoptosis

6. Soluble E-cadherin (sE-Cadherin)

4. Angiogenesis

V. High-throughput Tumor
Profiling

5. Signaling Proteins
6. Hormone Receptors
7. Combination of Tissue-Based
Molecular Markers

208

Molecular Markers for Bladder


Cancer Staging and Prognosis
Shahrokh F. Shariat
Michael J. Droller, Yves Fradet, George P. Hemstreet,
MLiss Hudson, Pierre I. Karakiewicz, Yair Lotan,
Peruno Malmstrom, Michael J. Marberger, Osamu Ogawa,
Bernd J. Schmitz-Drger, Christian Seitz, Maxine Sun,
Bas W. Van Rhijn, Vinata B. Lokeshwar
gov). A MEDLINE search was performed with
emphasis on urothelial malignancies and prognostic
markers using combinations of the following terms:
urinary tract cancer, bladder carcinomas, urothelial
carcinomas,
detection,
bladder,
recurrence,
and prognostic markers. Basically, articles were
considered between 2000 and 2011. No evidence
level 1 information from prospective randomized
trials was available. Articles were selected for
this review with regards to the following criteria:
evolution of concepts, development and refinement
of techniques, quality of the study, and relevance.
Older studies were included selectively if historically
relevant or in case of scanty data in more recent
publications.

I. Introduction
Intensive molecular research over the last decade
has provided great insight to the biology of bladder
cancer and is beginning to shape clinical practice.
Technologies such as high-throughput transcript
profiling, microarrays, and proteomics have facilitated
the identification and understanding of molecular
pathways and molecules that are active in bladder
cancer. Molecular profiling of bladder cancer is
beginning to offer additional means to predict tumor
behavior and thereby potentially improve treatment
outcomes. In addition, new markers can conceivably
serve as targets for novel therapeutic approaches
and allow improved selection of therapeutic options.

III. Evidence Synthesis

It is now clear that different forms of bladder cancer


develop along independent but possibly interrelated
molecular pathways. Each of the steps associated
with the development of malignancy, including
tumor initiation, progression, and metastasis involve
multiple genetic and epigenetic events indicating
that few tumors are genetically and/or phenotypically
identical. This is reflected by the heterogeneity in
clinical and histologic presentation, intrinsic biologic
potential of the various diatheses, and treatment
outcomes. Current methods of risk stratification do not
fully predict these distinctions. In this review, we will
discuss the challenges and statistical considerations
of molecular marker research and than explore
recent advances regarding the biology and potential
clinical utility and consequent integration of molecular
markers for bladder cancer staging, prognostication,
and therapeutic development.

1. C
 hallenges and Statistical
Considerations for Bladder
Cancer Prognostic Molecular
Marker Research
Despite a plethora of molecular markers reported to
be clinically promising, there is currently no single
prognostic molecular marker in routine clinical use.
Challenges for the application of molecular markers
in cancer care include analytical and regulatory
barriers.[3-8] In addition, the absent use of bladder
cancer molecular markers in clinical practice is a
result of poor application of statistics and study
design by which such markers might have been
validated. Molecular marker research is usually
done in the context of standard clinical care, rather
than clinical trials. This has been largely guided by
intuition and anecdotal experience rather that wellstructured analyses. As a result, most molecular
marker findings are not reproducible. Furthermore,
most molecular markers that appear biomedically

II. Evidence Acquisition


The literature was reviewed using the National
Library of Medicine database (http://www.pubmed.

209

prominent journals (Figure 2).[19-21] The goal


of these guidelines is to encourage transparent
and complete reporting and to help readers judge
the data and understand the context in which the
conclusions apply. Indeed, it provides a detailed
description as to the minimum amount of information
that should be given in the reporting of results from
molecular marker studies. The REMARK lists 20
items that investigators should attempt to report in
any molecular marker study. Another tool aimed
at standardizing the quality of molecular marker
research is the Tumor Marker Utility Grading System
(Figure 3).[22] This is a scale of levels of evidence,
designed to help place molecular marker studies into
a context of validity.

and statistically significant at one center are not


confirmed by studies at other centers.[9-14] This
prompted the development of guidelines intended
to ensure that molecular marker studies conform to
some basic standards in design and reporting.[4-8]

In 2002, the National Cancer Institutes Early


Detection Research Network developed a 5 phase
approach to systematic discovery and evaluation
of molecular markers.[15-17] These phases of
research are generally ordered in a sequence from
discovery to validation and ultimately to assessment
of benefit according to the strength of evidence
that each provides. This approach is not only an
intellectual process but also provides a clear scale
by which researchers, patients, reviewers, and
investors can evaluate the status of a molecular
An issue that has received less attention is the degree
marker in its development process. The phases of
to which research on molecular markers has made
research are generally ordered according to the
sufficient use of clinically relevant statistics, such
strength of evidence that each provides in favor of
as the assessment of predictive accuracy, decision
the biomarker. The results from earlier phases are
analysis, and/or experimental methodology. The
generally necessary to design later phases. This
fundamental idea behind the concept of personalized
classification of studies into a sequence of phases,
medicine is that it is possible to identify patterns of
from discovery to validation and assessment of
demographic, clinical, genomic/proteomic, and other
benefit has been adapted by several bladder cancer
types of biological data that can be used together
groups (Figure 1).
to benefit individual patients. Most molecular
markers do not provide sufficient information to be
In an effort to address the pervasive problem of
used independent of other information. However,
inadequacies
in the design, analysis, and reporting
1
the optimaltouse
molecular
markers would be in
2
Figure
Modificationstudies,
of the structured
the of
systematic
discovery,
of molecular
marker1.prognostic
a set ofphase-approach
3 recommendations such as
evaluation,
and validation
biomarkers in a model that includes standard
the of
incorporation
reporting
the REMARK
4
clinical data together with relevant additional data
has been
developed and adopted by many
5

Phase

Goals/aims

Experimentation

Sample details

Preclinical

Exploratory; nominate

Preclinical study

Possible bias: small size and

Testing

and rank candidate

for hypothesis generation

convenience sampling

biomarker profiles
0

Develop an assay with

Reproducibility and

clinically reproducible results

robustness of assay; No
assessment of benefit

II

Test on small sample to

Marker optimization, establish

Sample population assay

determine benefit

prediction rules, determine

developed from candidate

cut-offs

biomarker profile

Determine operating

Retrospective design

Sample population should be

characteristics & internal

the target population

validation
III

External validation

Retrospective or prospective,

Multi-institutional, large study

Generaliziability, Impact on
clinical decision-making
IV

Assess whether biomarker


reduces the burden of disease

Post-approval reporting and


testing for other disease
processes or disease stages

Figure 1. Modification of the structured phase-approach to the systematic discovery, evaluation, and
validation of biomarkers

210

Item
INTRODUCTION
1 State the marker examined, the study objectives, and any pre-specified hypotheses.
MATERIALS AND METHODS Patients
Describe the characteristics (e.g., disease stage or co-morbidities) of the study patients, including their
source and inclusion and exclusion criteria.
3 Describe treatments received and how chosen (e.g., randomized or rule-based).
Specimen characteristics
Describe type of biological material used (including control samples) and methods of preservation and
4
storage.
Assay methods
Specify the assay method used and provide (or reference) a detailed protocol, including specific
reagents or kits used, quality control procedures, reproducibility assessments, quantitation methods,
5
and scoring and reporting protocols. Specify whether and how assays were performed blinded to the
study endpoint.
Study design
State the method of case selection, including whether prospective or retrospective and whether
6a
stratification or matching (e.g., by stage of disease or age) was used.
Specify the time period from which cases were taken, the end of the follow-up period, and the median
6b
follow-up time.
7 Precisely define all clinical endpoints examined.
8 List all candidate variables initially examined or considered for inclusion in models.
Give rationale for sample size; if the study was designed to detect a specified effect size, give the
9
target power and effect size.
Statistical analysis methods
Specify all statistical methods, including details of any variable selection procedures and other model10
building issues, how model assumptions were verified, and how missing data were handled.
Clarify how marker values were handled in the analyses; if relevant, describe methods used for
11
cutpoint determination.
2

RESULTS
Data
Describe the flow of patients through the study, including the number of patients included in each stage
12 of the analysis (a diagram may be helpful) and reasons for dropout. Specifically, both overall and for
each subgroup extensively examined report the numbers of patients and the number of events.
Report distributions of basic demographic characteristics (at least age and sex), standard (disease13
specific) prognostic variables, and tumor marker, including numbers of missing values.
Analysis and presentation
14 Show the relation of the marker to standard prognostic variables.
Present univariate analyses showing the relation between the marker and outcome, with the estimated
effect (e.g., hazard ratio and survival probability). Preferably provide similar analyses for all other
15
variables being analyzed. For the effect of a tumor marker on a time-to-event outcome, a Kaplan-Meier
plot is recommended.
For key multivariable analyses, report estimated effects (e.g., hazard ratio) with confidence intervals
16
for the marker and, at least for the final model, all other variables in the model.
Among reported results, provide estimated effects with confidence intervals from an analysis in which
17
the marker and standard prognostic variables are included, regardless of their statistical significance.
If done, report results of further investigations, such as checking assumptions, sensitivity analyses, and
18
internal validation.
DISCUSSION
Interpret the results in the context of the pre-specified hypotheses and other relevant studies; include a
19
discussion of limitations of the study.
20 Discuss implications for future research and clinical value.
Figure 2. Reporting recommendations for tumor marker prognostic studies (REMARK) NCI-EORTC (with
permission from McShane et al.[21])

211

Figure 3. Levels of evidence for grading clinical utility of biomarkers: Tumor Marker Utility Grading System
(TMUGS) (reproduced with permission of Hayes et al. [22])

with regards to assay performance (i.e., medications


or nutrition that may interfere with an assay), so that
the model could then be used to provide individual
patient care.

impractical for most molecular markers. However,


decision analytic techniques can often be used
instead of trials. The key point of decision analysis is
that the consequences of clinical decisions would be
incorporated in these analyses. This decision analytic
evaluation should be performed during later stages
of research and before clinical implementation of a
tool. Furthermore, external validation needs to be
performed prior to utilization of the model. Wider
adoption of decision analytic methods will provide
better insights as to whether any of the plethora
of new molecular markers improves assessment,
treatment outcomes, and ultimate health.

To determine the value of a new molecular marker, it


is not sufficient to show that it is significantly related
to prognosis or outcome, statistically significant in
a multivariable model that includes the standard
clinical and pathologic factors, or more significant
than standard clinical and pathologic factors. Rather,
for a molecular marker to be clinically useful, it is
necessary to show that adding the molecular marker
to an existing model based on the most important
clinical and pathologic factors substantively
improves the predictive accuracy (discrimination
and calibration) of the model. The next step is to
assess whether the increase in accuracy and risk
assessment translates into improved individualized,
evidence-based treatment recommendations with
eventual superior patient outcomes. Indeed, an
improvement only in predictive accuracy, although
necessary, is not in itself sufficient to assess whether
using the marker in practice would actually benefit
patients.

2. P
 romising Tissue-based Prognostic
Molecular Markers
a) Cell Cycle Markers
The cell cycle is a series of carefully coordinated
and regulated steps that govern cellular proliferation.
Progression through the cell cycle is mediated in part
by the buildup of cyclins, proteins that activate cyclindependent kinases (cdks). Cdks then phosphorylate
the retinoblastoma (Rb) protein, a classic tumor
suppressor and key component of the G1/S
checkpoint. This allows DNA replication to proceed.
Cdk inhibitors of both the inosin phosphokinase
(INK) family, such as p16INK4A and the KIP family,
such as p21 and p27, act as brakes on cell cycle
progression (Figure 4). The p53 protein serves as
the guardian of the genome by inducing multiple
mechanisms of cell cycle arrest after cell stress.[29]

Establishing clinical relevance of a molecular


marker test to guide therapeutic decisions requires
demonstrating that it can classify patients into
distinct subgroups with different recommended
managements. While this would be ideally assessed
in a randomized clinical trial, the substantial
human and financial resources necessary for such
randomized trials would make such an evaluation

Mutations of cell cycle regulatory genes are the

212

42

6
7
8
9

the bottom arrow. Abbreviations: FGFR3, fibroblast growth factor receptor 3 gene; mt, mutation;
, elevated expression (Ki-67, p53); CIS, carcinoma in situ.

MOLECULAR AND CLINICAL PATHWAYS OF BLADDER CANCER

FGFR3 mutation

pTa/1 G1/2

pT2 G2/3

Normal
Urothelium

Metastasis
p53
Ki-67

pTa/1 G2/3
CIS

pT2 G3

INCREASING NUMBER OF GENETIC EVENTS


10
11
Figure
12 4. Two-pathway model of the disease pathogenesis of Bladder cancer (BC). This figure shows the
combination
of molecular and pathological data. Arrow thickness is indicative for the percentage of tumors.
13
The
14 FGFR3 mutation is largely responsible for the favorable molecular pathway. Among many others, P53
and
15 Ki-67 over-expression are examples of unfavorable NMI-BC. Molecular alterations, not included in the
figure in the interest of clarity, are represented by the bottom arrow. Abbreviations: FGFR3, fibroblast growth
factor receptor 3 gene; mt, mutation; , elevated expression (Ki-67, p53); CIS, carcinoma in situ.

most common genetic alterations found in human


cancer, including bladder cancer. Alterations in
the p53 and retinoblastoma (Rb) tumor suppressor
genes are predominant components in the
development of bladder cancer, but the downstream
effectors and other pathways that contribute to
urothelial transformation remain to be more fully
defined. p21WAF1/CIP1 and p27Kip1 are cyclindependent kinase inhibitors that can arrest the cell
cycle by blocking G1/S progression. p21 is a wellcharacterized p53-inducible gene. As we will discuss,
alterations of p53, pRB, p21, and p27 expression are
markers that have prognostic significance in bladder
cancer patients.

1. p53
The tumor suppressor p53 is the most commonly
found mutated gene in human cancer. Activated
by various types of cellular stress, including DNA
damage and oncogene activation, p53 acts as a
sequence specific transcription factor and appears
to initiate a cascade of molecular events that result
in cell cycle arrest, DNA repair, and apoptosis,
thereby preventing the generation of genetically
altered cells. Mutation or loss of chromosome 17p
and consequent functional inactivation of p53 leads
to the loss of ability to affect the important cell
process and consequent uncontrolled proliferation
and growth of a malignancy.
Many studies have demonstrated that the majority of
invasive bladder cancers display allelic loss of the p53
gene. While most of these support that p53 nuclear
accumulation is predictive of outcome, particularly
for patients treated with radical cystectomy, there
is evidence for and against almost every aspect
of the role of p53 in cancer biology and clinical

outcomes or application.[9] These discrepancies


may be related to the choice of antibody used in p53
assays, variability in interpretation and stratification
criteria, and inconsistencies in specimen handling
and other technical procedures. A meta-analysis
of the role of p53 in bladder cancer, found 117
studies including 10,026 patients, with sample sizes
ranging from 12 to 270 patients (mean 86, SD 53).[9]
Immunohistochemistry was the most commonly (96%
of the 117 studies) used approach for assessment of
p53 status, with molecular analysis being used in the
other 5 studies.
In a tissue microarray study containing specimens
from over 400 bladder cancer patients, the proportion
of altered p53 progressively increased in specimens
from normal urothelium to non-muscle-invasive
bladder cancer to carcinoma in situ and then to muscleinvasive bladder cancer, being highest in metastatic
lymph node specimens. In accordance with all large
previous studies, the degree of p53 expression
(reflecting chromosome 17p abnormalities) was
significantly associated with pathologic tumor stage,
43lymphovascular invasion, lymph node metastases,
pathologic tumor grade, disease recurrence, and
bladder cancer-specific death. Interestingly, the
p53 phenotype was a stronger predictor of bladder
cancer outcomes (disease recurrence and survival)
in patients treated with radical cystectomy than
chromosomal alterations and degree of expression
of p21, pRB, p27, p16, cyclin E1, and cyclin D1. In a
multi-institutional tissue microarray study comprising
over 1200 patients treated with radical cystectomy
for bladder cancer, p53 improved the accuracy of
predictive models that included standard clinical
and pathologic features for prediction of disease
recurrence and bladder cancer-specific survival.
Interestingly, in patients with advanced bladder

213

cancer (pT3-4 N0 or pT, any N+), adding p53 to a


base model of clinical and pathologic characteristics
did not improve overall predictive accuracy.[56]
However, p53 improved prediction of recurrencefree and cancer-specific survival in patients with
pT1-2N0 disease by a statistically and prognostically
significant margin.[47], Taken together, these studies
suggest that p53 can improve bladder cancer
prognostication in high risk disease; however, the
prognostic value of p53 decreases significantly in
advanced disease.

trial found no clinical utility for p53 expression in


predicting survival in high-risk patients with T1
tumors, but this may have been affected by the low
event rate related to treatment efficacy.
Finally, in a recent study of 80 consecutive patients
with pT1N0 bladder cancer, expression of p53
was altered in 25% of patients and p53 was
independently associated with bladder cancer
recurrence (HR 3.66; p=0.033) and cancer-specific
mortality (HR 5.25; p=0.016).[57] The concordance
index of a base model that included tumor grade,
lymph node status, lymphovascular invasion, and
concomitant carcinoma in situ for bladder cancer
recurrence and bladder cancer-specific mortality
was 54.7% and 64.3%, respectively. Addition of
p53 increased the concordance indices of the base
model for bladder cancer recurrence and cancerspecific mortality to 60% and 67.9%, respectively.
The authors concluded that p53 can help stratify the
heterogeneous population of pT1 patients into risk
groups that can be used to guide clinical decisionmaking regarding observation versus adjuvant
therapy. These findings are in line with data from the
ISBC analysis combining immunohistochemical p53
staining results from 23 different studies.[50] This
combined analysis of 929 patients with T1 bladder
cancer demonstrated an independent prognostic
value of p53 immunohistochemistry in multivariable
analysis regarding tumor progression but not overall
survival. However, these observations need to be
reproduced by other groups.

Although p53 nuclear accumulation is associated


with mutations in the p53 gene of chromosome 17,
significant discordance exists. Lamy et al showed
inactive p53 variants in up to 44% of bladder
cancers, with the highest proportion in high grade
tumors.[58] Stern et al investigated the frequency of
p53 mutations in bladder cancer, taking into account
DNA repair genotypes.[59] They found that tumor
progression pathways may differ among subjects
with different DNA repair polymorphisms, possibly
explaining the inter-individual variation observed for
the type and frequency of p53 gene mutations.
In addition to genetic factors, environmental
exposures are likely to play an important role in
bladder carcinogenesis. Based on this concept, Ryk
et al suggested that the occurrence of p53 mutations
may be due to genetic polymorphisms resulting in
altered metabolism and repair.[60] This may result
from individual modulation of mutagenic DNA
adduct levels, and may then be of importance for an
individuals risk of developing bladder cancer.

At this time, the use of p53 is still not established despite more than 100 studies evaluating utility in bladder cancer patients. In the meta-analysis mentioned
above, p53 independently predicted recurrence,
progression, and mortality in only 27% (9 of 34),
50% (12 of 24), and 29% (10 of 35), respectively.9 In
the studies that used Cox regression modeling, the
overall risk of recurrence was 1.6 (95% CI 1.2-2.1),
progression was 3.1 (1.9-4.9), and mortality was 1.4
(1.2-1.7). Nevertheless, given that most larger, welldone studies have shown a prognostic value to p53
in patients with advanced urothelial carcinoma of the
bladder, a large multicenter adjuvant chemotherapy
study was initiated with the goal of randomizing patients with a nuclear p53 accumulation (p53+) to adjuvant chemotherapy or observation within 10 weeks
following radical cystectomy (for details see: https://
www.uscnorris.com/p53/). This clinical trial might
shed more light on the role of p53 in identifying patients at high risk for disease recurrence as well as
response to adjuvant chemotherapy using MVAC.

Accurately sub-stratifying patients with high grade


urothelial carcinoma involving the lamina propria
(stage T1) regarding risk and potential treatment
outcomes represents a particular challenge for
urologists. There is conflicting data on whether
p53 immunohistochemistry can reliably categorize
patients with T1 tumors into different risk strata
with regards to clinical outcomes.[33,48] A
preliminary report from the EPICURO showed that
p53 overexpression correlated with higher grade
and stage of the disease in 995 patients with nonmuscle-invasive bladder cancer.[61] However,
the prognostic significance of p53 dropped out in
multivariate analyses. In a follow-up study, using
both immunohistochemistry and gene sequencing
of 119 tumors, these authors found that the p53
pathway is inactivated in most T1 grade 3 tumors,
but failed to identify any prognostic value for p53
in these patients.[61] Moonen et al also reported
no additional value for p53 mutation analysis over
known prognostic factors using specimens from 105
patients with high-risk non-muscle-invasive bladder
cancer.[62] Another group reported that p53 nuclear
accumulation was not predictive of response to
intravesical BCG, cancer recurrence, progression,
or cancer-specific survival.[63] A recent prospective

2. pRB
The retinoblastoma protein (pRb) is a pivotal
cell cycle regulator that plays a role in stem cell
maintenance, tissue regeneration, differentiation,
and developmental programs. Interference with

214

the function of pRb or components of its pathway is


a major mechanism by which cancer cells escape
normal growth control mechanisms. Specifically,
pRb plays a key role in coordinating cell-cycle
control and G1/S phase progression. Rb regulated
protein-to-protein activation of E2Fs and other
transcriptional regulators leads to phosphorylation
of cyclin-dependent kinases (cdk4/cdk6 and cdk2)
and release of E2F factors, which initiate the DNA
duplication machinery and G1/S transition. Over
the past decade, it has become clear that the
genomic or epigenetic loss of pRb function leads to
uncontrolled cell growth. This is not only an initiating
event in tumorigenesis, but also a step associated
with malignant progression and aggressive outcome.
However, recent studies suggest that the predictive
power of pRB may be inferior to that of other cell
cycle regulators both in non-muscle-invasive and
muscle-invasive bladder cancer. This may be the
result of the complex redundant pathway that pRb
coordinates.

an independent role in predicting outcome but this


varies by disease stage. In patients with muscleinvasive bladder cancer, p21 was an independent
predictor of both disease recurrence and cancerspecific mortality. In patients with organ-confined
disease, p21 remained independently associated
with disease recurrence and bladder cancer-specific
death. In addition, the authors found that p21 has
cooperative/synergistic action with p53, p27, and
pRB.49 Conversely, in a study of 80 patients with
pT1N0 bladder cancer who underwent cystectomy,
p21 was not an independent predictor of disease
recurrence or cancer-specific mortality.[57] Similarly,
p21 was not an independent predictor in patients
with advanced disease (T4, N positive).[45,55] p21
may be most useful in patients with T2, T3, nodenegative disease especially in combination with
other markers.[49]

4. p27
p27Kip1 is a member of the Cip/Kip family of Cdk
inhibitors, which binds to Cdk2 (and to other Cdks)
and potently inhibits Cdk2 kinase activity,[66]
potentiating cell-cycle arrest in the G1 phase. In
quiescent cells (G0 phase), p27 levels are high
concomitant with low Cdk activity. As cells progress
through the cell cycle and enter S phase, p27 levels
decrease, allowing the kinase activity of Cdks to
increase.

3. p21
The product of the p21 gene binds to and inhibits
the activity of cyclin-dependent kinase 2 or cyclindependent kinase 4 complexes, and thus functions
as a regulator of cell cycle progression at G1.[65] In
a large immunohistochemical study of p21 and p53
expression in 242 patients, a multivariate analysis
showed that p21 status was an independent
predictor of tumor recurrence and survival after
radical cystectomy.[36] Moreover, patients with p53altered/p21-altered tumors demonstrated a higher
rate of disease relapse and worse cancer-specific
survival, compared with those with p53-altered/p21normal tumors.[36]

Two studies have found limited predictive value


for p27 in patients with non-muscle-invasive
bladder cancer. However, in patients with muscleinvasive disease treated with radical cystectomy,
p27 significantly improved prediction of bladder
cancer recurrence and survival. In patients treated
with radical cystectomy, p27 is a key component
of a panel of immunohistochemical markers
which improved predictive value of a nomogram
based on standard clinical and pathological
characteristics both in patients with broad stage
characteristics (pT1-3N0M0) and in a more select
population of pT1 patients. Furthermore, p27 was
independently associated with disease recurrence
and progression.

The value of p21 expression depends on the stage


and management of bladder cancer patients. In
a study of 49 patients with carcinoma in situ only,
altered p21 expression was independently associated
with bladder cancer recurrence and progression.
[46] Patients with carcinoma in situ and altered
expression for both p53 and p21 are at the greatest
risk of bladder cancer recurrence, progression, and,
most importantly, mortality, suggesting a potential
rationale for early definitive therapy in these patients.
On the other hand, an intact pathway at the level
of p21 seems to abrogate the detrimental effects
of altered p53 immunoreactivity on the outcome of
bladder carcinoma in situ. In another study of 74
patients with non-muscle-invasive papillary bladder
cancer, p21 was independently associated with tumor
progression. Combination of p21 with p53, pRB, and
p27 stratified patients into statistically significantly
different risk groups for disease recurrence and
progression in patients with non-invasive disease.
[43]

5. Cyclins
The deregulation of the G1/S transition is a hallmark
of cancer, allowing uncontrolled cell proliferation.
Cyclins D and E are responsible for the initial and
terminal phases of G1, respectively. Cyclin E1
is the predominant regulatory protein active at
the G1/S transition and its alteration significantly
affects carcinogenesis. Normal bladder urothelium
expresses cyclin E1 only in the umbrella cells, while
cancer cells already expressed cyclin E1 in the
earliest stages of the disease process.[53] Cyclin E1
expression was significantly decreased in patients
with features of biologically and clinically aggressive
bladder cancer; decreased expression of cyclin E1

In patients who undergo cystectomy, p21 may play

215

was significantly associated with advanced pathologic


stage, lymphovascular invasion, metastases to
regional lymph nodes, and bladder cancer-specific
mortality after controlling for the effects of standard
pathologic features in both transurethral resection
and radical cystectomy specimens.

effector by cleaving multiple cellular components.


[74] Giannopolou et al studied caspase-3 expression
in 53 patients with bladder cancer and did not find
a correlation between caspase-3 expression and
tumor grade or stage.[75] Burton et al[76] evaluated
caspase-3 expression in 34 patients with carcinoma
in situ, of whom 41% developed invasive bladder
cancer. They reported that activated caspase-3
overexpression was associated with higher rates of
disease invasiveness. Conversely, a more recent
study by Karam et al[54] involving 226 consecutive
patients treated with radical cystectomy reported
that 49% of the patients had loss of caspase-3
expression, which was associated with higher
pathologic grade and stage, and presence of lymph
node metastasis. Moreover, loss of caspase-3 was
an independent predictor of bladder cancer-specific
survival after radical cystectomy.[54]

Although the same authors reported that cyclin D1


immunoreactivity was elevated in bladder cancer
patients compared to controls, it was not associated
with clinical or pathologic characteristics among
bladder cancer patients.[53] While loss of cyclin
D1 expression was associated with an increased
probability of disease recurrence and bladder
cancer-specific mortality in univariate analyses,
this association was not significant in a multivariate
analysis that adjusted for the effects of standard
pathologic features. These findings are consistent
with previous reports showing that alteration in cyclin
D1 is an early event in bladder tumorigenesis, but it
does not add prognostic significance.

c) Bcl-2
Bcl-2 is an anti-apoptotic protein present in
intracellular membranes, and controls cytochrome c
location, caspase status, and ion channels involved
in apoptosis.[77] Overexpression of bcl-2 was
found in 32% of radical cystectomy specimens,
and correlated with higher pathologic stage,
disease recurrence, and cancer-specific mortality.
[54] In agreement with these findings, two other
groups reported that over-expression of bcl-2 was
associated with worse all-cause survival and lower
response rates to chemotherapy.[79]

3. Apoptosis
Apoptosis, or programmed cell death, is a complex
and highly regulated process comprised of a series of
coordinated steps resulting in cell death. Alterations
in apoptosis pathways are important in tumorigenesis
and cancer progression, as they allow cancerous
cells to survive, resist a variety of stresses, and
express more invasiveness.[70] Bladder cancer has
been shown to resist programmed cell death with
alteration of both pro- and anti-apoptotic signaling.

a) Fas (CD95)

d) Survivin

Fas is a death receptor; when activated by Fas


ligand (FasL), Fas activates downstream signals
that induce apoptosis.[71] Fas can cause apoptosis
in immune cells and prevent inflammatory killing
of tumor cells. Svatek et al reported that the
soluble form of Fas was present in the cell lysate
and supernatant of high-grade bladder cancer cell
lines, suggesting that Fas is likely to be produced
and released by bladder cancer cells.[72] Urine
soluble Fas was an independent predictor of the
presence of bladder cancer and invasiveness in
patients with a past history of non-muscle-invasive
disease. Immunohistochemical analysis in 123
bladder cancer specimens and the mucosa of 30
benign bladders revealed that Fas was expressed
in 90% of benign specimens compared to 58% of
cancer specimens.[73] In addition, decreased Fas
expression was associated with higher tumor stage,
grade, and disease-specific mortality in univariate
analysis.[73] These observations suggest that Fas
may play a role in the local immunosuppression in
bladder cancer associated with tumor development,
biologic aggressiveness, and progression.

Survivin is a member of the Inhibitor of Apoptosis


(IAP) family, and inhibits apoptosis, at least partly, by
blocking downstream caspase activity. Survivin also
controls mitotic progression and induces changes
in gene expression that are associated with tumor
cell invasion.Survivin mRNA is selectively expressed
during embryonic and fetal development,[84]
becomes undetectable or expressed at low levels
in most differentiated normal adult tissues, and is
overexpressed in human cancers.[83] In bladder
cancer, survivin expression, both at the protein
and the mRNA level, is associated with cancer
presence, higher tumor grade, and advanced
pathologic stage.[85-87] Survivin overexpression
was present in 63% of bladder cancer specimens,
and was associated with higher pathologic stage,
presence of lymphovascular invasion, lymph node
metastasis, bladder cancer recurrence, and bladder
cancer-specific survival in 219 patients treated with
radical cystectomy.[54] In addition, the proportion of
specimens with survivin overexpression increased
gradually from non-muscle-invasive bladder cancer
to advanced bladder cancer and to metastatic lymph
node tissue.[88] In a large multicenter validation
study, addition of survivin improved the accuracy of
standard clinicopathologic features for prediction of

b) Caspase-3
Caspase-3 is a protease that acts as an apoptosis

216

disease recurrence and cancer-specific survival in a


subgroup of patients with pT1-3N0M0 disease.[89]
Survivin is a highly promising marker that deserves
further investigation as a prognostic/predictive
marker as well as a target for therapy.[90]

these 2 molecular markers were strongly correlated.


These findings were independently confirmed in a
study of 204 patients with muscle-invasive disease
treated with radical cystectomy.[99] Decreased
thrombospondin-1 was independently associated
with disease recurrence and cancer-specific survival.
Moreover, loss of thrombospondin-1 expression
was associated with alterations in other cell cycle
regulators such as p21 and p27.

4. Angiogenesis
Angiogenesis, the process of new blood vessel
formation, is a critical event in the initiation and
progression of solid malignancies. Angiogenesis
has been traditionally quantified by measuring
microvessel density. However, a variety of molecular
markers are also associated with angiogenesisrelated events.

c) Vascular Endothelial Growth Factor


Of the various angiogenic factors, VEGF has been
identified as a crucial regulator of normal and pathologic angiogenesis. VEGF produces a number of
important biological effects such as endothelial mitogenesis and migration, extracellular matrix remodeling via induction of proteases, increased vascular
permeability, and maintenance of newly formed vasculature. Several members of the VEGF family have
been characterized. In a small study of 45 patients,
VEGF-C expression was localized to the cytoplasm
of bladder cancer cells, with minimal expression in
normal bladder epithelium.[100] VEGF-C expression
was significantly associated with tumor size, pathologic stage and grade, lymphovascular invasion,
and pelvic lymph node metastasis. In multivariable
analysis, VEGF-C expression was an independent
predictor of pelvic lymph node metastasis.[100] In
addition, patients with high VEGF-C expression
had a worse prognosis compared to those with low
VEGF-C expression. In a separate study of 126 patients treated with transurethral resection of bladder
cancer, higher expression of VEGF-A, VEGF-C, and
VEGF-D were all associated with increasing tumor
stage and grade.[101] VEGF-C expression was
also associated with higher microvessel density. In
a study of 204 patients treated with radical cystectomy and bilateral pelvic lymphadenectomy, VEGF
was over-expressed in 86% of patients, supporting
its role in bladder tumorigenesis and identifying it as
a potential target for therapy.[94]

a) Microvessel density
Bochner et al[91] evaluated 164 bladder cancer
specimens and reported that patients with high
microvessel density (>100 microvessels per hpf)
were at highest risk of disease recurrence (68%
at 5 years) and cancer-specific mortality (66% at 5
years) compared to patients with lower microvessel
density (MVD). In addition, MVD was an independent
prognostic factor when adjusted for the effect of
tumor stage, grade, and lymph node status. However,
other investigators could not confirm the prognostic
value of MVD.[92-94] In 204 patients treated with
radical cystectomy, Shariat et al failed to detect an
association between MVD and prognosis, but they
found that MVD was higher in patients with lymph
node metastasis.[94] Jaeger et al examined 41
muscle-invasive tumors following radical cystectomy,
demonstrating higher MVD in the primary tumor if
LN metastases were present.[95] Explanations for
these discrepancies in prognostic value of MVD
include differences in the choice of antibody used
for immunohistochemistry, as well as staining and
scoring protocols. For example, the endothelium of
tumor and normal tissues is heterogeneous. Thus,
pan-endothelial antibodies such as CD31 may not
stain tumor vessels to the same degree and may
generally react better with larger vessels.[96]

In addition, VEGF overexpression was associated


with pathologic features, and altered expression
of p21, p27, pRB, cyclin E1 and Ki-67, suggesting
complex interactions between different bladder
cancer-associated
molecules.
While
these
associations are important, VEGF expression had
no independent prognostic value. Increased VEGF
levels can result in increased vascular permeability
and interstitial fluid pressure, impairing chemotherapy
delivery. Thus, adding anti-VEGF to chemotherapy
regimens for bladder cancer might lead to improved
responses.

b) Thrombospondin-1
Thrombospondin-1, an important component of
the extracellular matrix, has been implicated in
the regulation of cell growth and proliferation, cell
motility, cytoskeletal organization, inflammation
and wound healing, and the development and
differentiation of cell types. Recently, it has been
shown to be a potent inhibitor of angiogenesis in vitro
and in vivo, and its expression is inversely related to
microvessel density.[97] Grossfeld et al previously
reported that altered thrombospondin-1 expression
was independently associated with an increased
risk of disease recurrence and all cause mortality
in 163 patients treated with radical cystectomy.[98]
However, the prognostic value of thrombospondin-1
was not independent of p53 expression status since

d) Basic Fibroblast Growth Factor


Similar to VEGF, bFGF is a potent pro-angiogenic
growth factor. It has been hypothesized that during
wound healing and tumor development bFGF is
activated, thereby mediating the formation of new

217

blood vessels. bFGF may behave as a transforming/


oncogenic factor inducing cell proliferation and
motility. The interaction with specific receptors may
lead to unchecked proliferation via the Ras-MAPK
pathway. The most common oncogenes and some
of the tumor suppressor genes relevant to bladder
cancer are components of this pathway.

not found in bladder cancer [108].


The effect of inhibition of EGFR signaling in urothelial cells is cystostatic.[109] Therefore, although
combination therapies with conventional cytotoxic
agents are underway, the requirement for cell division to allow some agents to exert their effects may
be problematic and timing of delivery may be critical.
Inhibitors of EGFR are available and have potential
clinical efficacy. Overexpression of EGFR was associated with disease progression of pT1 high grade
urothelial carcinoma of the bladder and decreased
survival.[110] Consequently, gefitinib, which targets
EGFR, was studied in preclinical in vitro and subcutaneous in vivo models of urothelial carcinoma of the
bladder with promising results.[111] Unfortunately,
a clinical phase II trial (CALGB 90102) investigating
the treatment of measurable metastatic urothelial
carcinoma of the bladder found a high toxicity and
no survival benefit when gefitinib was added to gemcitabine and cisplatin chemotherapy.[112]

bFGF has been evaluated as a potential urinary


molecular marker in bladder cancer. Higher levels of
bFGF were found in urine samples from patients with
bladder cancer compared with those with benign
disease or a history of urothelial carcinoma but no
current tumor.[104] Others have measured urinary
bFGF in patients with various grades and stages
of urothelial carcinoma, and noted a correlation of
increased bFGF with advanced disease.[105] In a
tissue microarray study of 204 radical cystectomy
patients, bFGF was associated with established
features of biologically aggressive bladder cancer
such as pathologic stage, lymphovascular invasion,
lymph node metastasis, molecular markers
commonly altered in bladder cancer (i.e., p27, pRb,
and Ki-67) and disease recurrence. In addition,
altered bFGF expression has been associated with
resistance to cisplatin in human bladder cancer
cell lines.[106] Since bFGF may also contribute
to bladder cancer progression via its extracellular
matrix involvement, bFGF could be a candidate for
therapeutic targeting.

Patient response is clearly correlated to the presence


of activating mutations in the kinase domain or a
variant of EGFR (known as EGFRvIII). Importantly,
several studies have demonstrated that mutations
within the tyrosine kinase domain of EGFR and
expression of EGFRvIII are rare events in bladder
cancer, these likely limiting the utility of EGFR as a
prognostic marker or therapeutic target.

5. Signaling Proteins

b) RAS

Many receptor tyrosine kinases and their downstream


effectors and regulators have genetic alteration and/
or altered expression in bladder cancer. Several of
these are oncogenes that have a dominant effect on
cell phenotype and these may represent particularly
good therapeutic targets.

Approximately 13% of bladder cancers of all grades


and stages contain a mutation in one of the RAS
genes (HRAS, NRAS, KRAS2).[114] This implicates
the RAS-MAP kinase (RAS-MAPK) pathway and
potentially the PI3K pathway in these tumors.
Although the RAS proteins themselves have proven
difficult to target, there is currently much interest in
developing small molecule inhibitors of other key
proteins in these pathways that may be applicable to
RAS-mutant bladder cancers. Further confirmation
of the activation status and examination of the
relationship to RAS and other mutations, including
FGFR3 and PIK3CA mutations (see below), will be
required to allow rational selection of patients for
these types of therapy.

a) ERBB Family Receptors


Receptor tyrosine kinases are attractive therapeutic
targets and currently there is much effort to target
these either with antibodies or small molecules.
Epidermal growth factor receptor (EGFR) is a receptor
tyrosine kinase implicated in the pathogenesis of
a variety of human cancers. Overexpression of
EGFR in bladder cancer was identified more than
20 years ago [107]. The mechanism for upregulation
is unknown but is not related to gene amplification.
There are several approved drugs (antibodies and
small molecules) that interfere with signaling via this
receptor and there are now several ongoing clinical
studies in bladder cancer. However, large trials in
non-small cell carcinoma of the lung have not shown
significant benefit. And although there is some
success in specific organs, it is difficult to predict
response as this is not related directly to expression
levels. In non-small cell carcinoma of the lung,
activating mutation of the receptor has been shown
to be predictive of response. But such mutations are

c) PI3K Pathway Genes


Several genes in the PI3K pathway are implicated
in bladder cancer. PTEN, a lipid phosphatase that
removes phosphate from phosphatidylinositol 3,4,5trisphosphate produced by PI3K, is commonly deleted in muscle-invasive tumors. In some cases, biallelic inactivation by mutation in the retained allele
or homozygous deletion has been found.[115-117]
Recent studies have identified activating point mutations in the alpha catalytic subunit of PI3 kinase
p110alpha (PIK3CA) in bladder cancer, with highest
frequency in non-invasive papillary tumors.[118] This

218

mutational activation of PI3K results in phosphorylation of AKT, which in turn regulates many cellular
activities including cell proliferation, stimulation of
protein synthesis via the mTOR pathway, effects on
metabolism and evasion of apoptosis [119].

situ. This molecular evidence for the clinical twopathway model in bladder cancer (Figure 5) is
further substantiated by the absence of FGFR3
mutations in 20 cases of CIS[123] and the mutually
exclusive occurrence of FGFR3 mutations and
altered expression of Ki-67, p53, and p27.

TSC1 acts in the mTOR pathway downstream of


AKT and this pathway is activated in bladder cancer
via mutational inactivation of TSC1. Inhibitors of
mTOR (rapamycin and analogues) are approved for
clinical use in kidney cancer and other malignancies
but not yet tested in bladder cancer. Inhibitors of
other pathway molecules are in development and
some show promising activity. For example, there
are reports of successful inhibition of tumor growth
in preclinical mouse models by a dual inhibitor of
p110alpha and mTOR. For rational selection of
appropriate inhibitors in the future, it will be important
to determine the relationship of different molecular
lesions in the pathway to tumor phenotype and
clinical outcome and to test the relative sensitivity of
cells with specific pathway alterations in preclinical
models.

FGFR3 mutations were found to protect against


progression of non-muscle-invasive bladder cancer.
However, a combination of FGFR3 with another
molecular marker associated with worse prognosis
if positive, like Ki-67, seemed to confer a more
accurate prediction of progression and diseasespecific survival than FGFR3 or Ki-67 alone.
Molecular grade (mG) based on FGFR3 mutation
status and Ki-67 expression was proposed as an
alternative to pathologic grade.[127] Using these
2 molecular markers, a 3-tier grading system was
developed covering all cases: mG1 (mutation; normal
expression), favorable prognosis; mG2 (no mutation;
normal expression or mutation; high expression),
intermediate prognosis; mG3 (no mutation; high
expression), poor prognosis.[124] A recent study
compared the reproducibility of pathological grading
and mG on the same series of patients.[127] The
reproducibility of mG was almost perfect (kappa;
0.76), whereas reproducibility for pathologic grade
was only fair to substantial (kappa: 0.17-0.58). In
addition, combining the EORTC risk scores and mG
led to a more accurate prediction of progression
in patients with intermediate and high-risk EORTC
scores.[127] In this recent study with long-term followup, EORTC high-risk combined with mG3 showed
50% (15/30) progression as opposed to 20% (6/29)
progression in EORTC high-risk combined with
mG1-2. Taken together, mG (composed of 3 mGs)
proved more reproducible than pathological grade
assessment, making it a reliable and robust tool to
assess progression in non-muscle-invasive bladder
cancer. An independent prospective analysis of mG
has to be done to confirm the promising retrospective
data described in this paragraph.

d) Fibroblast Growth Factor Receptor 3


Fibroblast growth factor receptor 3 (FGFR3)
belongs to a tyrosine kinase receptor family and
regulates diverse cellular processes including
growth, differentiation, and angiogenesis. It is
activated by a point mutation in bladder cancer.
[122] Activating FGFR3 mutations are present in
up to 60% of bladder tumors, and are characteristic
of papillary, non-muscle-invasive bladder cancer.
[122] Surprisingly, FGFR3 mutations were found to
be related to favorable disease with 84% of pTaG1
tumors having a mutation as compared to 7% of
pT2G3 or higher stage tumors.[123] In general, the
FGFR3 mutation represents a subset of low grade,
low stage tumors that rarely progress and hence
have a good prognosis.[122-124]
Lindgren et al used expression profiling, mutation
analysis, and loss of heterozygosity to molecularly
characterize a cohort of 75 Ta and T1 bladder
cancers.[125] They confirmed that low grade, low
stage bladder cancer is characterized by FGFR3
receptor activity, high protein synthesis, and low
cell-cycle activity. Using screening for FGFR3
mutations by direct bidirectional sequencing and for
genome-wide molecular changes with microarray
technology of 35 low grade Ta and 50 T1 or grade
3 tumors, Zieger et al found that FGFR3-mutated
non-muscle-invasive tumors can progress and retain
their mutation and chromosomal instability patterns
during progression.[126] Their results suggest that
the decreasing frequency of FGFR3 mutations in
later stages of tumor development is caused by the
emergence of tumors following a different molecular
pathway with no FGFR3 mutations. Tumors of this
latter pathway were associated with carcinoma in

It is clear that the FGFR3 gene plays a major role in


the development of low-grade non-muscle-invasive
bladder tumors; however, there is little evidence
that it plays any role in the development of invasive
cancer or in the progression of non-muscle--invasive
to invasive disease. One of the most important
clinical questions remains whether FGFR3 can help
risk stratify the heterogeneous group of T1 grade 3
bladder cancers. As illustrated above, a combination
of FGFR3 with another molecular marker associated
with worse clinical features may be the answer.

6. Hormone Receptors
a) Human Epidermal Growth Factor Receptor 2
Human epidermal growth factor receptor 2
(HER-2) is a tyrosine kinase in the EGFR family
that can promote oncogenesis. HER-2 has been

219

best characterized in breast cancer, in which


overexpression of HER-2 is a poor prognostic factor
and inhibition of HER-2 function has proven to be an
effective approach for the treatment of breast cancer
cells that overexpress HER-2. Conflicting data exists
regarding the expression and significance of HER-2
in bladder cancer.

used tissue microarrays and different antibodies,


correlated in a blinded fashion with results observed
in two centers and included automated and manual
scoring systems. Results were further validated
and confirmed on a small representative sample
population analyzed in a previous positive study[139]
to show similar findings, despite antibody epitope,
reagents, and laboratories.

A number of studies have shown limited or no


prognostic value of examining HER-2 expression.
[128-133] For instance, Kassouf etal reported no
significant correlation between HER-2 expression
status and clinical outcome in 184 patients with
primary bladder cancer, and another study found that
HER-2/neu overexpression in primary or metastatic
tumors did not predict survival in a cohort of 80
consecutive patients with muscle-invasive bladder
cancers. However, other studies have reported that
higher HER-2 expression levels are associated with
poor prognosis[134-136]. Recently, a study of 198
patients observed on multivariable analyses that
HER-2 positive patients were at increased risk for
both bladder cancer recurrence and cancer-specific
mortality compared with patients with negative
HER-2 expression.[137]

c) E
 strogen
Receptor

Receptor

and

Progesterone

Like AR, the estrogen receptor (ER) is a member


of the nuclear receptor superfamily, and acts a
transcription factor after binding to estrogens and
translocation to the nucleus. Two subtypes of ER
exist, with varying expression and functional profiles:
estrogen receptor alpha (ERalpha) and estrogen
receptor beta (ERbeta). ER is expressed in a subset
of bladder tumors; however, the prognostic and
functional significance remains unclear. ERalpha
is rarely expressed by bladder cancer specimens.
[141] It has been reported that expression of ERbeta
is associated with pathologic characteristics of
bladder cancer (stage and grade).[142] However,
several studies have shown limited or no impact of
ER expression on prognosis of patients with bladder
cancer.

These conflicting data may be partially explained by


differences in patient populations and techniques
of HER-2 analysis and immunostaining among the
various studies. Clearly, additional data is necessary
from prospective studies using standardized
analytical techniques for HER-2 expression. Clinical
trials have been initiated to investigate the efficacy
of anti-HER-2 therapy in patients with advanced
bladder cancer, and these will provide further insight
to the role of HER-2 in disease progression.

The progesterone receptor (PR) is another


intracellular steroid receptor that acts by a
mechanism similar to that of AR and ER; however, a
recent report indicates it is not expressed in bladder
cancer.[141]

7. C
 ombination of Tissue-Based
Molecular Markers

b) Androgen Receptor

Given the complexity of the molecular abnormalities


associated with bladder cancer, it is improbable that
a single marker can accurately segregate tumors
of similar clinicopathologic phenotypes into distinct
prognostic categories. A marker may reflect disruption of a biochemical pathway by a particular mechanism but not by another. Therefore, as previously
shown, combinations of independent, complementary markers may provide a more accurate prediction
of outcome compared to a single marker.

The androgen receptor (AR) is a nuclear receptor


and ligand-dependent transcription factor that
mediates the biologic effects of androgens. Boorjian
et al reported that expression of androgen receptor
was inversely correlated with pathological stage in
a series of 49 patients with bladder cancer; 75%
of non-muscle-invasive tumors expressed the AR
compared with 21.4% of muscle-invasive tumors.
[138] Similarly, another study confirmed that loss
of AR expression was associated with higher grade
and invasive tumors; however, no association was
found with patient outcomes.[139]

Karam and coworkers found that p53, Bcl-2,


caspase-3, and survivin display distinct association
patterns with tumor stage and grade, lymphovascular
invasion, and carcinoma in situ.[54] The number of
simultaneously altered apoptosis markers was an
important prognostic indicator for disease recurrence
and bladder cancer-specific survival in patients
treated by radical cystectomy. Changes in expression
of markers were frequent in individuals with bladder
cancer, with only 22 patients (10%) showing normal
expression of all 4 markers. Alteration of fewer than
4 apoptosis markers did not provide independent
prognostic information after controlling for the

Although several groups have argued for a role for AR


in pathogenesis of bladder cancer, evidence for using
androgen receptor as either a prognostic marker or
therapeutic target in bladder cancer remains limited.
In a study comprising 472 patients with urothelial
bladder carcinoma from two centers, the authors
did not observe any association between AR protein
expression and bladder cancer stage, grade, or
outcomes.[140] Further data did not suggest that
loss of AR expression is gender-related. The authors

220

effects of standard pathologic features; only when


all 4 markers were altered was significance reached.
This is explained by the fact that Bcl-2, caspase-3,
p53, and survivin have not only interrelated but also
independent roles in the apoptotic pathway.

and proven in large multi-institutional prospective


controlled trials.

Analysis of any combination of cell cycle regulators


(i.e., p53, pRB, p21, p27, cyclin D1, and/or cyclin
E1) provides additional prognostic information
beyond that obtained from any single molecular
marker or combination of 2 or 3 molecular markers.
This makes sense because it investigates multiple
pathways including downstream effectors rather than
a single crossroad in a pathway. Higher total number
of altered molecular markers was associated with
a progressive, proportional increase in the risk of
advanced pathologic tumor stage, lymphovascular
invasion, lymph node metastases, disease recurrence,
and death from bladder cancer. These molecular
markers had a cooperative or synergistic role in the
biologic behavior of bladder cancer. Furthermore,
examining how many markers are altered and adding
this information to a nomogram of clinicopathologic
criteria significantly improved predictive accuracy for
disease recurrence and cancer-specific mortality.
[45] Other studies have confirmed this principle in
selected patient populations; addition of a number
of altered molecular markers to a base predictive
model significantly improved predictive accuracy
for 80 patients with T1 disease,[57] 324 patients
with organ-confined bladder cancer,[145] and 692
patients with advanced bladder cancer.[55]

Insulin growth factor (IGFs) and IGFBPs have a


central role in the regulation of growth, cellular proliferation and transformation, and apoptosis. IGFBP-3
belongs to a family of high affinity IGFBPs, which
directly mediate IGF-independent actions, such as
regulation of cell growth and induction of apoptosis,
through binding to its own putative receptor. However, evidence also suggests that IGFBP-3 has its
own pro-apoptotic effects, independent of its ability
to modulate IGF bioavailability.[148] The activities of
IGFBP-3 are regulated in a complex manner, such
as IGFBP-3 proteases like kallikreins, cathepsin,
and matrix metalloproteinases, which release IGFs
from IGFBP-3. [149] Equally noteworthy is that several studies have found that IGFBP-3 is capable of
mediating the antiproliferative effects of tumor suppressor protein p53, transforming growth factor-beta
(TGF-1), retinoic acid, vitamin D, antiestrogens,
and tumor necrosis factor-alpha. Previous investigators collected preoperative plasma levels of IGFBP-3
measured using DSL-enzyme-linked immunosorbent assays in a group of 51 patients who underwent
radical cystectomy and compared it to the levels in
blood from 44 healthy controls.[150] The association
between IGFBP-3 and IGF-I was also tested, and
authors found that both levels were strongly correlated with one and another among patients with bladder cancer (r=0.568, p<0.001). Preoperative plasma
IGFBP-3 concentration level was lower in patients
with regional lymph node metastases (p=0.047) and
was a significant predictor of lymph node involvement, independent of clinical stage and grade (HR:
0.9, p=0.047). This confirms the speculations that
a relationship between circulatory IGFBP-3 levels
and metastasis may exist. Moreover, authors of that
study revealed that low levels of IGFBP-3 portended
an increased risk of disease recurrence (HR: 0.9,
p=0.049) and cancer-specific mortality (HR: 0.9,
p=0.04) in the preoperative setting after adjusting
for the effects of clinical stage and grade. A better
understanding of the biologic mechanisms and the
role of IGFBP-3 in bladder cancer is worthy of investigation.

1. P
 lasma-insulin Growth Factor
Binding Protein (IGFBP)-3

Because tumorigenesis and progression is a


process involving multiple genetic defects, it is likely
that alterations in other pathways will also affect
bladder cancer progression and metastasis. These
data, together with the current paradigm that bladder
cancer develops along multiple molecular pathways,
suggest that including multiple molecular markers in
models designed to predict outcomes could enhance
their predictive power. Therefore, use of multiple
molecular markers likely represents the future of risk
stratification and should be used to guide patient
counseling and management decisions such as
neoadjuvant and adjuvant therapy.

IV. Promising Blood-based


Prognostic Molecular
Markers

2. T
 ransforming Growth Factor
(TGF)-1

The use of biomarkers in the blood offers several


advantages over tissue samples, such as that
of higher sample homogeneity and its minimally
invasive nature. Furthermore, blood samples
provide information that can be known at any
disease stage, offering an alternative to surgical and
clinical decision-making. However, whether further
value exists in regards to the clinical applicability of
these blood-based biomarkers remains to be tested

TGF-1 inhibits cell growth in cellular proliferation,


chemotaxis,
cellular
differentiation,
immune
response, and angiogenesis. Loss of response to
the antiproliferative effects of TGF-1 is associated
with the progressive stages of carcinogenesis.[151]
The effects of TGF-1 are mediated by membranebound serine-threonine kinase receptors, otherwise
known as TGF-1 receptor types I and II (TGF-

221

1-R1 and TGF-1-RII). Alteration of TGF-1 is


common in bladder cancer and overexpression of
TGF-1 is associated with the loss of expression of
its receptors.[151] An increased expression of TGF1 has also been linked with features of advanced
bladder cancer such as tumor grade, stage, and
lymphovascular invasion.[151-153] Elevated plasma
TGF-1 levels (n=51) have been associated with an
increased probability of regional and distant lymph
node metastasis even after adjusting for clinical
stage and grade (odds ratio [OR]: 12.8, p=0.03).
[153] However, others could not confirm this finding.
[152] Elevated levels of TGF-1 were associated
with an increased risk of disease recurrence and
cancer-specific mortality both in the preoperative
(HR: 1.6, p=0.009 and HR: 2.9, p=0.02, respectively)
and postoperative settings (HR: 3.1, p=0.01 and HR:
3.0, p=0.02, respectively).[153]

originate at least partly from the primary bladder


cancer cells. Nonetheless, further studies with larger
sample sizes are needed to better understand the
biologic and prognostic role of MMP-7 in bladder
cancer.
Limited data exist on other circulating MMPs
in patients with bladder cancer. Some reported
elevated concentrations of MMP-2 and MMP-3 in
patients with advanced bladder cancer (pathological
T24, N+, M+) compared to those in patients with
non-muscle-invasive bladder cancer (Ta-T1, N0,
M0). Lower levels of plasma MMP-2 (p=0.03)
have been associated with more advanced tumor
stage and grade.[163] Moreover, it appears that
the tissue inhibitor of metalloproteinases-(TIMP) 2,
which inhibits protease activity of MMP-2, could also
suppress invasion, metastasis, neovascularization,
and growth in some human tumors, and that the ratio
of MMP-2 and TIMP-2 is considered an independent
predictor of cancer recurrence.[167] Increased
MMP-9 serum levels have also been found in patients
with advanced clinical stage and grade.[168]

3. Matrix Metalloproteinase (MMP)


MMP is a family of zinc-dependent proteases involved
in the breakdown and remodeling of extracellular
matrix. A degradation of extracellular matrix is vital
in normal physiological processes, as well as in
pathological processes such as tumor progression
and metastasis. [154] In addition, MMPs can induce
several molecular processes implicated in tumor
progression through their cleaving abilities. MMPs
can mobilize pro-angiogenic factors, but are also
capable of generating angiogenic inhibitors, such
as endostatin and angiostatin. [155-157] In bladder
cancer, MMP expression has been associated with
more advanced stage and grade.[158-162] Plasma
MMP concentrations (-2, -3, -7, -8, and -9) collected
from 135 patients with high grade T1 or higher stage
bladder cancer[163] revealed that MMP-7 was
elevated in patients with advanced clinical stage
(p=0.02). Moreover, patients with MMP-7 levels
above the median (300 pg/mL) were at increased
risk of cancer-specific mortality compared to patients
with MMP-7 below the median (5-year survival
estimates: 48% vs. 74%, p=0.01). After adjusting
for the effects of clinical grade, MMP-7 remained an
independent predictor of cancer-specific mortality
(HR: 2.2, p=0.02).

4. Urokinase Plasminogen Activation


(uPA)
uPA is a serine protease involved in the formation and
regulation of blood vessels and clots, bone modeling,
and activation of metalloproteinases and growth
factors. It holds an important role in tumor invasion
and metastasis by accelerating the conversion of
plasminogen to plasmin. Subsequently, the plasmin
cleaves to components of the basement membrane
and extracellular matrix to allow tumor cells to
access lymph vessels and vasculature. The inactive
precursor of uPA is activated by binding to a specific
membrane-bound or soluble cell surface receptor
(uPAR). This hastens the conversion of plasminogen
into plasmin.
In bladder cancer cells, the presence of both uPA
and uPAR were necessary for in vitro invasion,
and invasive potential was reduced when their
interaction was interrupted. [171] Plasma levels of
uPA and uPAR were significantly higher in bladder
cancer patients (n=51) compared to healthy subjects
(n=44, p<0.001). [172] Higher uPA concentration was
associated with lymphovascular invasion (p=0.02)
and regional lymph node metastases (p=0.02),
while higher uPAR concentration was associated
with distant lymph node metastasis (p=0.04). In the
preoperative multivariable model, only uPA emerged
as an independent predictor of disease recurrence
(HR: 3.7, p=0.03) and cancer-specific mortality
(HR: 3.7, p=0.04). After validation, these markers
may be useful in selecting patients most likely to
benefit from intensified follow-up and therapy. These
markers should also be considered for integration in
prediction tools.

Szarvas et al confirmed these results in 79 serum


samples from bladder cancer patients and 19 healthy
individuals.[164] Higher levels of serum MMP-7 were
associated with bladder cancer presence (p<0.001),
lymph node metastasis (p=0.002), overall (p=0.008),
and cancer-specific mortality (p=0.006). However,
MMP-7 was not associated with tumor stage
(p=0.1) or grade (p=0.2). In multivariable analyses,
higher serum MMP-7 remained associated with
metastatic cancer (HR: 2.5, p=0.048) and cancerspecific mortality (HR: 2.0, p=0.04). Interestingly,
serum levels of MMP-7 decreased significantly after
surgery in 75% of cases suggesting that the higher
circulating MMP-7 in bladder cancer patients might

222

5. Interleukin-6 (IL-6) and IL-6 Soluble


Receptor (IL-6sR)

appropriate database and bioinformatic support, will


be effective tools to screen for novel markers and
to fingerprint each cancer. Indeed, the advent of
high-throughput methods of molecular analysis has
allowed the comprehensive survey of the genetic
profiles characteristic of distinct tumor types and
identification of targets and pathways that may
underlie particular clinical behavior. Several groups
also have used gene expression profiling of bladder
cancer tissues to identify signature genes that
distinguish bladder cancer subclasses and genetic
pathways underlying bladder cancer progression.
Such signature genes ideally would provide a
molecular basis for classification and also yield
insight into the molecular events underlying different
bladder cancer clinical phenotypes.

IL-6 is a cytokine that plays a role in the regulation of


the immune system via the proliferation and activation
of cytotoxic T cells, proliferation and differentiation of
B cells, and production of acute phase proteins. IL-6
signaling is initiated when IL-6 binds to the ligand
specific non-signaling receptor IL-6R, which also
exists in soluble form as IL-6sR. The latter arises
from 1 or 2 independent cellular processes, namely
differential mRNA splicing or proteolytic cleavage
of membrane bound IL-6 receptors. Previous data
support the hypothesis that bladder cancer cell lines
produce more IL-6 than normal urothelial cells.[173]
It seems that IL-6 confers a selective advantage for
certain disseminated cells to develop into metastatic
tumors. In a prospective study of 51 bladder
cancer patients, elevated IL-6 and IL-6sR levels
were associated with more advanced pathological
stage, lymphovascular invasion, and regional and
distant lymph node metastases.[174] Only IL-6
was correlated with clinical stage and tumor grade.
Both IL-6 and IL-6sR were independent predictors
of lymph node metastasis, disease recurrence,
and cancer-specific mortality after adjusting for the
effects of clinical stage and grade.

Microarray-based approaches have been used


extensively to look for expression profiles that could
sub-classify bladder cancers. Dyrskjot et al used
high-density oligonucleotide microarrays to identify
a 45-gene signature of disease progression in a
training set of 29 bladder cancer patients.[126] This
signature was then examined using an independent
test set (74 superficial tumor samples) revealing a
significant correlation between gene expression
classifications and clinical outcome.
Blaveri et al used cDNA microarrays (10,368 genes)
to identify differentially expressed genes along
the course of disease progression in 80 bladder
tumors, 9 bladder cancer cell lines, and 3 normal
bladder samples.[187] Unsupervised hierarchical
clustering successfully separated the samples into
2 subgroups containing non-muscle-invasive versus
muscle-invasive tumors, supported by a 90.5%
success rate. Tumors could also be classified
into transitional versus squamous subtypes (89%
success rate) and good versus bad prognosis
(78% success rate). Sanchez-Carbayo et al used
oligonucleotide arrays to identify genetic signatures
characteristic of aggressive clinical behavior in
advanced bladder tumors by transcript profiling of 52
normal urothelium, 33 superficial, and 72 invasive
tumors.[188] Unsupervised clustering classified them
with 82.2% accuracy, while predictive algorithms
rendered an 89% correct rate for tumor staging using
genes differentially expressed between superficial
and invasive tumors. Accuracies of 82% and 90%
were obtained for predicting overall survival when
considering all patients with bladder cancer or only
patients with invasive disease, respectively.

6. Soluble E-cadherin (sE-Cadherin)


E-cadherin is a member of the family of transmembrane
glycoproteins involved in calcium dependent
intercellular adhesion. [175] Normal E-cadherin
expression suppresses tumor progression. When
altered, it allows cancer cells to detach, invade,
and access the lymphatic and vascular system. In
bladder cancer, as in other malignancies, a decrease
in E-cadherin expression has been linked with an
increased risk of metastases, tumor progression,
and cancer-specific mortality. [177-182] sE-cadherin
is the degradation product of cellular E-cadherin,
generated by Ca2+ ion dependent proteolytic activity.
Several authors have found increased serum levels
of sE-cadherin in bladder cancer patients compared
to non-cancer subjects. Moreover, sE-cadherin was
significantly associated with regional lymph node
metastasis as well as disease recurrence. However,
sE-cadherin failed to achieve statistical significance
when cancer-specific mortality was the endpoint of
interest.

V. High-throughput Tumor
Profiling

Target validation of synuclein by immunohistochemistry on tissue arrays (N=294) sustained its association with tumor staging and outcome. Lindgren
et al successfully used this approach to distinguish
specific molecular subtypes of bladder cancer and
define gene signatures that can classify bladder
cancer by grade as well as muscle-invasion status.
Furthermore, they reported a gene expression sig-

As our understanding of the individuality of cancers


evolves, we will rely on molecular fingerprints to
tailor management and therapeutic approaches
to the clinical, morphological, and molecular
features of each patient. High-throughput genomic,
transcriptomic, and proteomic assays, with

223

nature that is associated with both metastasis and


survival.[189] Takata et al examined expression profiles of 27,648 genes in 27 cases of bladder cancer
to predict response to a neoadjuvant methotrexate,
vinblastine, doxorubicin, and cisplatin (M-VAC) chemotherapy regimen.[190] They identified 14 genes
with a correct prediction in 8 out of 9 patients. These
studies suggest that defining gene expression signatures may be a relevant complement to standard
clinical and pathological risk stratification. However,
current costs associated with examination of such a
large number of genes make transition to the clinical
realm difficult.

9. Malats N, Bustos A, Nascimento CM, et al: P53 as a


prognostic marker for bladder cancer: a meta-analysis and
review. Lancet Oncol 6:678-86, 2005

Conclusions

14. Diamandis EP: Proteomic patterns to identify ovarian


cancer: 3 years on. Expert Rev Mol Diagn 4:575-7, 2004

10. Dalbagni G, Parekh DJ, Ben-Porat L, et al: Prospective


evaluation of p53 as a prognostic marker in T1 transitional
cell carcinoma of the bladder. BJU Int 99:281-5, 2007
11. Svatek RS, Jeldres C, Karakiewicz PI, et al: Pretreatment biomarker levels improve the accuracy of postprostatectomy nomogram for prediction of biochemical
recurrence. Prostate 69:886-94, 2009
12. Bensalah K, Lotan Y, Karam JA, et al: New circulating
biomarkers for prostate cancer. Prostate Cancer Prostatic
Dis, 2007
13. Check E: Proteomics and cancer: running before we can
walk? Nature 429:496-7, 2004

Bladder cancer is an especially complex and


heterogeneous disease with a broad spectrum of
histologic findings. Molecular medicine holds the
promise that clinical outcomes will be improved by
directing therapy toward the mechanisms and targets
associated with the growth of an individual patients
tumor. The advent of high-throughput technologies
is allowing comprehensive identification of molecular
targets and molecular markers specific for bladder
cancer. Such identification processes may provide a
better understanding of the biology associated with
tumorigenesis and tumor progression. However,
further research is warranted in the field to translate
the identification of these molecular targets into
potential predictive molecular markers. The challenge
remains to optimize measurement of these targets,
evaluate the impact of such targets for therapeutic
drug development, and translate molecular markers
into improved clinical management of bladder cancer
patients.

15. Verma M, Srivastava S: New cancer biomarkers deriving


from NCI early detection research. Recent Results Cancer
Res 163:72-84; discussion 264-6, 2003
16. Winget MD, Baron JA, Spitz MR, et al: Development
of common data elements: the experience of and
recommendations from the early detection research
network. Int J Med Inform 70:41-8, 2003
17. Goebell PJ, Groshen SL, Schmitz-Drager BJ: Guidelines
for development of diagnostic markers in bladder cancer.
World J Urol 26:5-11, 2008
18. Lokeshwar VB, Habuchi T, Grossman HB, et al: Bladder
tumor markers beyond cytology: International Consensus
Panel on bladder tumor markers. Urology 66:35-63, 2005
19. McShane LM, Altman DG, Sauerbrei W, et al: REporting
recommendations for tumor MARKer prognostic studies
(REMARK). Nat Clin Pract Urol 2:416-22, 2005
20. Mallett S, Timmer A, Sauerbrei W, et al: Reporting of
prognostic studies of tumour markers: a review of published
articles in relation to REMARK guidelines. Br J Cancer,
2009
21. McShane LM, Altman DG, Sauerbrei W, et al: Reporting
recommendations for tumor marker prognostic studies
(REMARK). J Natl Cancer Inst 97:1180-4, 2005
22. Hayes DF, Bast RC, Desch CE, et al: Tumor marker utility
grading system: a framework to evaluate clinical utility of
tumor markers. J Natl Cancer Inst 88:1456-66, 1996

REFERENCES

23. Shariat SF, Karakiewicz PI: Perspectives on prostate


cancer biomarkers. Eur Urol 54:8-10, 2008

1. Shariat S, Karam J, Lerner S: Molecular markers in bladder


cancer. Curr Opin Urol 18:1-8, 2008

24. Shariat SF, Karakiewicz PI, Roehrborn CG, et al: An


updated catalog of prostate cancer predictive tools. Cancer
113:3075-99, 2008

2. Karam JA, Shariat SF, Hsieh JT, et al: Genomics: a preview


of genomic medicine. BJU Int 102:1221-7, 2008
3. Bensalah K, Montorsi F, Shariat SF: Challenges of cancer
biomarker profiling. Eur Urol 52:1601-9, 2007

25. Shariat SF, Karam JA, Margulis V, et al: New blood-based


biomarkers for the diagnosis, staging and prognosis of
prostate cancer. BJU Int 101:675-83, 2008

4. Lotan Y: Role of biomarkers to predict outcomes and


response to therapy. Urologic Oncology: Seminars and
Original Investigations 28:97-101, 2010

26. Shariat SF, Karam JA, Walz J, et al: Improved prediction of


disease relapse after radical prostatectomy through a panel
of preoperative blood-based biomarkers. Clin Cancer Res
14:3785-91, 2008

5. Shariat SF, Canto EI, Kattan MW, et al: Beyond prostatespecific antigen: new serologic biomarkers for improved
diagnosis and management of prostate cancer. Rev Urol
6:58-72, 2004

27. Shariat SF, Lotan Y, Vickers A, et al: Statistical consideration


for clinical biomarker research in bladder cancer. Urol Oncol
28:389-400, 2010

6. Shariat SF, Karam JA, Roehrborn CG: Blood biomarkers


for prostate cancer detection and prognosis. Future Oncol
3:449-61, 2007

28. Vickers AJ, Elkin EB: Decision curve analysis: a novel


method for evaluating prediction models. Med Decis
Making 26:565-74, 2006

7. Shariat SF, Kattan MW, Vickers AJ, et al: Critical review of


prostate cancer predictive tools. Future Oncol 5:1555-84,
2009

29. Vogelstein B, Lane D, Levine AJ: Surfing the p53 network.


Nature 408:307-10, 2000

8. Shariat SF, Lotan Y, Vickers A, et al: Statistical consideration


for clinical biomarker research in bladder cancer. Urol Oncol
28:389-400

30. Ruas M, Peters G: The p16INK4a/CDKN2A tumor


suppressor and its relatives. Biochim Biophys Acta
1378:F115-77, 1998

224

31. Esrig D, Spruck CHd, Nichols PW, et al: p53 nuclear protein
accumulation correlates with mutations in the p53 gene,
tumor grade, and stage in bladder cancer. Am J Pathol
143:1389-97, 1993

49. Shariat SF, Zlotta AR, Ashfaq R, et al: Cooperative effect


of cell-cycle regulators expression on bladder cancer
development and biologic aggressiveness. Mod Pathol
20:445-59, 2007

32. Cote RJ, Dunn MD, Chatterjee SJ, et al: Elevated and
absent pRb expression is associated with bladder cancer
progression and has cooperative effects with p53. Cancer
Res 58:1090-4, 1998

50. Dalbagni G, Presti J, Reuter V, et al: Genetic alterations in


bladder cancer. Lancet 342:469-71, 1993

33. Grossman HB, Liebert M, Antelo M, et al: p53 and RB


expression predict progression in T1 bladder cancer. Clin
Cancer Res 4:829-34, 1998

52. Goebell PJ, Groshen SG, Schmitz-Drger BJ, et al: p53


immunohistochemistry in bladder cancer--a new approach
to an old question. Urol Oncol 28:377-88, 2010

51. Knowles MA, Elder PA, Williamson M, et al: Allelotype of


human bladder cancer. Cancer Res 54:531-8, 1994

34. Sarkis AS, Dalbagni G, Cordon-Cardo C, et al: Nuclear


overexpression of p53 protein in transitional cell bladder
carcinoma: a marker for disease progression. J Natl Cancer
Inst 85:53-9, 1993

53. Shariat SF, Ashfaq R, Sagalowsky AI, et al: Correlation of


cyclin D1 and E1 expression with bladder cancer presence,
invasion, progression, and metastasis. Hum Pathol, 2006
54. Karam JA, Lotan Y, Karakiewicz PI, et al: Use of combined
apoptosis biomarkers for prediction of bladder cancer
recurrence and mortality after radical cystectomy. Lancet
Oncol 8:128-36, 2007

35. Esrig D, Elmajian D, Groshen S, et al: Accumulation of


nuclear p53 and tumor progression in bladder cancer. N
Engl J Med 331:1259-64, 1994
36. Stein JP, Ginsberg DA, Grossfeld GD, et al: Effect of
p21WAF1/CIP1 expression on tumor progression in bladder
cancer. J Natl Cancer Inst 90:1072-9, 1998

55. Shariat SF, Chade DC, Karakiewicz PI, et al: Combination


of multiple molecular markers can improve prognostication
in patients with locally advanced and lymph node positive
bladder cancer. J Urol 183:68-75

37. Tokunaga H, Shariat SF, Green AE, et al: Correlation


of immunohistochemical molecular staging of bladder
biopsies and radical cystectomy specimens. Int J Radiat
Oncol Biol Phys 51:16-22, 2001

56. Shariat SF, Bolenz C, Karakiewicz PI, et al: p53 expression


in patients with advanced urothelial cancer of the urinary
bladder. BJU Int 105:489-95

38. Benedict WF, Lerner SP, Zhou J, et al: Level of retinoblastoma


protein expression correlates with p16 (MTS-1/INK4A/
CDKN2) status in bladder cancer. Oncogene 18:1197-203,
1999

57. Shariat SF, Bolenz C, Godoy G, et al: Predictive value of


combined immunohistochemical markers in patients with
pT1 urothelial carcinoma at radical cystectomy. J Urol
182:78-84; discussion 84, 2009

39. Cordon-Cardo C, Wartinger D, Petrylak D, et al: Altered


expression of the retinoblastoma gene product: prognostic
indicator in bladder cancer. J Natl Cancer Inst 84:1251-6,
1992

58. Lamy A, Gobet F, Laurent M, et al: Molecular profiling of


bladder tumors based on the detection of FGFR3 and TP53
mutations. J Urol 176:2686-9, 2006
59. Stern MC, Conway K, Li Y, et al: DNA repair gene
polymorphisms and probability of p53 mutation in bladder
cancer. Mol Carcinog 45:715-9, 2006

40. Shariat SF, Tokunaga H, Zhou J, et al: p53, p21, pRB, and
p16 expression predict clinical outcome in cystectomy with
bladder cancer. J Clin Oncol 22:1014-24, 2004

60. Ryk C, Kumar R, Sanyal S, et al: Influence of polymorphism


in DNA repair and defence genes on p53 mutations in
bladder tumours. Cancer Lett 241:142-9, 2006

41. Chatterjee SJ, Datar R, Youssefzadeh D, et al: Combined


effects of p53, p21, and pRb expression in the progression
of bladder transitional cell carcinoma. J Clin Oncol 22:100713, 2004

61. Lopez-Knowles E, Hernandez S, Kogevinas M, et al: The


p53 pathway and outcome among patients with T1G3
bladder tumors. Clin Cancer Res 12:6029-36, 2006

42. Shariat SF, Ashfaq R, Sagalowsky AI, et al: Association of


cyclin D1 and E1 expression with disease progression and
biomarkers in patients with nonmuscle-invasive urothelial
cell carcinoma of the bladder. Urol Oncol 25:468-75, 2007

62. Moonen PM, van Balken-Ory B, Kiemeney LA, et al:


Prognostic value of p53 for high risk superficial bladder
cancer with long-term followup. J Urol 177:80-3, 2007

43. Shariat SF, Ashfaq R, Sagalowsky AI, et al: Predictive value


of cell cycle biomarkers in nonmuscle invasive bladder
transitional cell carcinoma. J Urol 177:481-7; discussion
487, 2007

63. Esuvaranathan K, Chiong E, Thamboo TP, et al: Predictive


value of p53 and pRb expression in superficial bladder
cancer patients treated with BCG and interferon-alpha.
Cancer 109:1097-105, 2007

44. Bensalah K, Traxer O, Shariat SF, et al: [What are the


differences between American and French urologists?].
Prog Urol 17:1367-71, 2007

64. Yurakh AO, Ramos D, Calabuig-Farinas S, et al: Molecular


and immunohistochemical analysis of the prognostic value
of cell-cycle regulators in urothelial neoplasms of the
bladder. Eur Urol 50:506-15; discussion 515, 2006

45. Shariat SF, Karakiewicz PI, Ashfaq R, et al: Multiple


biomarkers improve prediction of bladder cancer recurrence
and mortality in patients undergoing cystectomy. Cancer
112:315-25, 2008

65. Harada K, Ogden GR: An overview of the cell cycle arrest


protein, p21(WAF1). Oral Oncol 36:3-7, 2000
66. Sherr CJ, Roberts JM: CDK inhibitors: positive and negative
regulators of G1-phase progression. Genes Dev 13:150112, 1999

46. Shariat SF, Kim J, Raptidis G, et al: Association of p53


and p21 expression with clinical outcome in patients with
carcinoma in situ of the urinary bladder. Urology 61:1140-5,
2003

67. Schrier BP, Vriesema JL, Witjes JA, et al: The predictive
value of p53, p27(kip1), and alpha-catenin for progression
in superficial bladder carcinoma. Eur Urol 50:76-82, 2006

47. Shariat SF, Lotan Y, Karakiewicz PI, et al: p53 Predictive


value for pT1-2 N0 disease at radical cystectomy. J Urol
182:907-13, 2009

68. Richter J, Wagner U, Kononen J, et al: High-throughput


tissue microarray analysis of cyclin E gene amplification
and overexpression in urinary bladder cancer. Am J Pathol
157:787-94, 2000

48. Shariat SF, Weizer AZ, Green A, et al: Prognostic value


of P53 nuclear accumulation and histopathologic features
in T1 transitional cell carcinoma of the urinary bladder.
Urology 56:735-40, 2000

69. Del Pizzo JJ, Borkowski A, Jacobs SC, et al: Loss of cell
cycle regulators p27(Kip1) and cyclin E in transitional cell

225

carcinoma of the bladder correlates with tumor grade and


patient survival. Am J Pathol 155:1129-36, 1999

biomarker for detection and prognosis of bladder cancer.


World J Urol 26:59-65, 2008

70. Reed JC: Dysregulation of apoptosis in cancer. J Clin Oncol


17:2941-53, 1999

91. Bochner BH, Cote RJ, Weidner N, et al: Angiogenesis in


bladder cancer: relationship between microvessel density
and tumor prognosis. J Natl Cancer Inst 87:1603-12, 1995

71. Nagata S: Apoptosis by death factor. Cell 88:355-65, 1997

92. Inoue K, Slaton JW, Karashima T, et al: The prognostic


value of angiogenesis factor expression for predicting
recurrence and metastasis of bladder cancer after
neoadjuvant chemotherapy and radical cystectomy. Clin
Cancer Res 6:4866-73, 2000

72. Svatek RS, Herman MP, Lotan Y, et al: Soluble Fas-A


promising novel urinary marker for the detection of recurrent
superficial bladder cancer. Cancer 106:1701-7, 2006
73. Yamana K, Bilim V, Hara N, et al: Prognostic impact of FAS/
CD95/APO-1 in urothelial cancers: decreased expression
of Fas is associated with disease progression. Br J Cancer
93:544-51, 2005

93. Streeter EH, Harris AL: Angiogenesis in bladder cancer-prognostic marker and target for future therapy. Surg Oncol
11:85-100, 2002

74. Arai M, Sasaki A, Saito N, et al: Immunohistochemical


analysis of cleaved caspase-3 detects high level of
apoptosis frequently in diffuse large B-cell lymphomas of
the central nervous system. Pathol Int 55:122-9, 2005

94. Shariat SF, Youssef RF, Gupta A, et al: Association


of angiogenesis related markers with bladder cancer
outcomes and other molecular markers. J Urol 183:174450, 2010

75. Giannopoulou I, Nakopoulou L, Zervas A, et al:


Immunohistochemical study of pro-apoptotic factors Bax,
Fas and CPP32 in urinary bladder cancer: prognostic
implications. Urol Res 30:342-5, 2002

95. Jaeger TM, Weidner N, Chew K, et al: Tumor angiogenesis


correlates with lymph node metastases in invasive bladder
cancer. J Urol 154:69-71, 1995
96. Wang JM, Kumar S, Pye D, et al: Breast carcinoma:
comparative study of tumor vasculature using two
endothelial cell markers. J Natl Cancer Inst 86:386-8,
1994

76. Burton PB, Anderson CJ, Corbishly CM: Caspase 3 and


p27 as predictors of invasive bladder cancer. N Engl J Med
343:1418-20, 2000
77. Kluck RM, Bossy-Wetzel E, Green DR, et al: The release
of cytochrome c from mitochondria: a primary site for Bcl-2
regulation of apoptosis. Science 275:1132-6, 1997

97. Dameron KM, Volpert OV, Tainsky MA, et al: Control


of angiogenesis in fibroblasts by p53 regulation of
thrombospondin-1. Science 265:1582-4, 1994

78. Ong F, Moonen LM, Gallee MP, et al: Prognostic factors in


transitional cell cancer of the bladder: an emerging role for
Bcl-2 and p53. Radiother Oncol 61:169-75, 2001

98. Grossfeld GD, Ginsberg DA, Stein JP, et al:


Thrombospondin-1 expression in bladder cancer:
association with p53 alterations, tumor angiogenesis, and
tumor progression. J Natl Cancer Inst 89:219-27, 1997

79. Kong G, Shin KY, Oh YH, et al: Bcl-2 and p53 expressions
in invasive bladder cancers. Acta Oncol 37:715-20, 1998

99. Shariat SF, Youssef RF, Gupta A, et al: Association


of angiogenesis related markers with bladder cancer
outcomes and other molecular markers. J Urol 183:174450

80. Ambrosini G, Adida C, Altieri DC: A novel anti-apoptosis


gene, survivin, expressed in cancer and lymphoma. Nat
Med 3:917-21, 1997
81. Akhtar M, Gallagher L, Rohan S: Survivin: role in diagnosis,
prognosis, and treatment of bladder cancer. Adv Anat
Pathol 13:122-6, 2006

100. Zu X, Tang Z, Li Y, et al: Vascular endothelial growth


factor-C expression in bladder transitional cell cancer
and its relationship to lymph node metastasis. BJU Int
98:1090-3, 2006

82. Salz W, Eisenberg D, Plescia J, et al: A survivin gene


signature predicts aggressive tumor behavior. Cancer Res
65:3531-4, 2005

101. Miyata Y, Kanda S, Ohba K, et al: Lymphangiogenesis and


angiogenesis in bladder cancer: prognostic implications
and regulation by vascular endothelial growth factors-A,
-C, and -D. Clin Cancer Res 12:800-6, 2006

83. Altieri DC: Survivin, versatile modulation of cell division and


apoptosis in cancer. Oncogene 22:8581-9, 2003

102. Elfiky AA, Rosenberg JE: Targeting angiogenesis in


bladder cancer. Curr Oncol Rep 11:244-9, 2009

84. Adida C, Crotty PL, McGrath J, et al: Developmentally


regulated expression of the novel cancer anti-apoptosis
gene survivin in human and mouse differentiation. Am J
Pathol 152:43-9, 1998

103. Xia G, Kumar SR, Hawes D, et al: Expression and


significance of vascular endothelial growth factor receptor
2 in bladder cancer. J Urol 175:1245-52, 2006

85. Smith SD, Wheeler MA, Plescia J, et al: Urine detection of


survivin and diagnosis of bladder cancer. Jama 285:324328, 2001

104. Chodak GW, Hospelhorn V, Judge SM, et al: Increased


levels of fibroblast growth factor-like activity in urine
from patients with bladder or kidney cancer. Cancer Res
48:2083-8, 1988

86. Shariat SF, Casella R, Khoddami SM, et al: Urine detection


of survivin is a sensitive marker for the noninvasive
diagnosis of bladder cancer. J Urol 171:626-30, 2004

105. Watanabe H, Hori A, Seno M, et al: A sensitive enzyme


immunoassay for human basic fibroblast growth factor.
Biochem Biophys Res Commun 175:229-35, 1991

87. Schultz IJ, Kiemeney LA, Karthaus HF, et al: Survivin


mRNA copy number in bladder washings predicts tumor
recurrence in patients with superficial urothelial cell
carcinomas. Clin Chem 50:1425-8, 2004

106. Miyake H, Hara I, Gohji K, et al: Expression of basic


fibroblast growth factor is associated with resistance to
cisplatin in a human bladder cancer cell line. Cancer Lett
123:121-6, 1998

88. Shariat SF, Ashfaq R, Karakiewicz PI, et al: Survivin


expression is associated with bladder cancer presence,
stage, progression, and mortality. Cancer 109:1106-13,
2007

107. Neal DE, Marsh C, Bennett MK, et al: Epidermal-growthfactor receptors in human bladder cancer: comparison of
invasive and superficial tumours. Lancet 1:366-8, 1985

89. Shariat SF, Karakiewicz PI, Godoy G, et al: Survivin as a


prognostic marker for urothelial carcinoma of the bladder:
a multicenter external validation study. Clin Cancer Res
15:7012-9, 2009

108. Villares GJ, Zigler M, Blehm K, et al: Targeting EGFR in


bladder cancer. World J Urol 25:573-9, 2007
109. Maclaine NJ, Wood MD, Holder JC, et al: Sensitivity of
normal, paramalignant, and malignant human urothelial

90. Margulis V, Lotan Y, Shariat SF: Survivin: a promising

226

cells to inhibitors of the epidermal growth factor receptor


signaling pathway. Mol Cancer Res 6:53-63, 2008

pattern of ErbB family receptors signifies an aggressive


variant of bladder cancer. J Urol 179:353-8, 2008

110. Mellon K, Wright C, Kelly P, et al: Long-term outcome


related to epidermal growth factor receptor status in
bladder cancer. J Urol 153:919-25, 1995

129. Jimenez RE, Hussain M, Bianco FJ, Jr., et al: Her-2/neu


overexpression in muscle-invasive urothelial carcinoma
of the bladder: prognostic significance and comparative
analysis in primary and metastatic tumors. Clin Cancer
Res 7:2440-7, 2001

111. Kassouf W, Luongo T, Brown G, et al: Schedule dependent


efficacy of gefitinib and docetaxel for bladder cancer. J
Urol 176:787-92, 2006

130. Mellon JK, Lunec J, Wright C, et al: C-erbB-2 in bladder


cancer: molecular biology, correlation with epidermal
growth factor receptors and prognostic value. J Urol
155:321-6, 1996

112. Philips GK, Halabi S, Sanford BL, et al: A phase II trial


of cisplatin, fixed dose-rate gemcitabine and gefitinib for
advanced urothelial tract carcinoma: results of the Cancer
and Leukaemia Group B 90102. BJU Int 101:20-5, 2008

131. Chow NH, Chan SH, Tzai TS, et al: Expression profiles of
ErbB family receptors and prognosis in primary transitional
cell carcinoma of the urinary bladder. Clin Cancer Res
7:1957-62, 2001

113. Blehm KN, Spiess PE, Bondaruk JE, et al: Mutations


within the kinase domain and truncations of the epidermal
growth factor receptor are rare events in bladder cancer:
implications for therapy. Clin Cancer Res 12:4671-7,
2006

132. Liedberg F, Anderson H, Chebil G, et al: Tissue microarray


based analysis of prognostic markers in invasive bladder
cancer: much effort to no avail? Urol Oncol 26:17-24,
2008

114. Jebar AH, Hurst CD, Tomlinson DC, et al: FGFR3 and Ras
gene mutations are mutually exclusive genetic events in
urothelial cell carcinoma. Oncogene 24:5218-5225, 2005

133. Gandour-Edwards R, Lara PN, Jr., Folkins AK, et al: Does


HER2/neu expression provide prognostic information
in patients with advanced urothelial carcinoma? Cancer
95:1009-15, 2002

115. Aveyard JS, Skilleter A, Habuchi T, et al: Somatic mutation


of PTEN in bladder carcinoma. Br J Cancer 80:904-8,
1999

134. Kruger S, Weitsch G, Buttner H, et al: HER2 overexpression


in muscle-invasive urothelial carcinoma of the bladder:
prognostic implications. Int J Cancer 102:514-8, 2002

116. Cairns P, Evron E, Okami K, et al: Point mutation and


homozygous deletion of PTEN/MMAC1 in primary bladder
cancers. Oncogene 16:3215-8, 1998

135. Sato K, Moriyama M, Mori S, et al: An immunohistologic


evaluation of C-erbB-2 gene product in patients with
urinary bladder carcinoma. Cancer 70:2493-8, 1992

117. Wang DS, Rieger-Christ K, Latini JM, et al: Molecular


analysis of PTEN and MXI1 in primary bladder carcinoma.
Int J Cancer 88:620-5, 2000

136. Kolla SB, Seth A, Singh MK, et al: Prognostic significance of


Her2/neu overexpression in patients with muscle invasive
urinary bladder cancer treated with radical cystectomy. Int
Urol Nephrol 40:321-7, 2008

118. Lopez-Knowles E, Hernandez S, Malats N, et al: PIK3CA


mutations are an early genetic alteration associated with
FGFR3 mutations in superficial papillary bladder tumors.
Cancer Res 66:7401-4, 2006

137. Bolenz C, Shariat SF, Karakiewicz PI, et al: Human


epidermal growth factor receptor 2 expression status
provides independent prognostic information in patients
with urothelial carcinoma of the urinary bladder. BJU Int

119. Engelman JA, Luo J, Cantley LC: The evolution of


phosphatidylinositol 3-kinases as regulators of growth and
metabolism. Nat Rev Genet 7:606-19, 2006
120. Fan QW, Knight ZA, Goldenberg DD, et al: A dual PI3
kinase/mTOR inhibitor reveals emergent efficacy in
glioma. Cancer Cell 9:341-9, 2006

138. Boorjian S, Ugras S, Mongan NP, et al: Androgen receptor


expression is inversely correlated with pathologic tumor
stage in bladder cancer. Urology 64:383-8, 2004

121. Raynaud FI, Eccles S, Clarke PA, et al: Pharmacologic


characterization of a potent inhibitor of class I
phosphatidylinositide 3-kinases. Cancer Res 67:5840-50,
2007

139. Tuygun C, Kankaya D, Imamoglu A, et al: Sex-specific


hormone receptors in urothelial carcinomas of the
human urinary bladder: A comparative analysis of
clinicopathological features and survival outcomes
according to receptor expression. Urol Oncol, 2009

122. Knowles MA: Role of FGFR3 in urothelial cell carcinoma:


biomarker and potential therapeutic target. World J Urol
25:581-93, 2007

140. Mir C, Shariat SF, Van Der Kwast TH, et al: Loss of
androgen receptor expression is not associated with
pathological stage, grade, gender or outcome in bladder
cancer: a large multi-institutional study. BJU Int, 2010

123. Billerey C, Chopin D, Aubriot-Lorton MH, et al: Frequent


FGFR3 mutations in papillary non-invasive bladder (pTa)
tumors. Am J Pathol 158:1955-9, 2001

141. Bolenz C, Lotan Y, Ashfaq R, et al: Estrogen and


Progesterone Hormonal Receptor Expression in Urothelial
Carcinoma of the Bladder. Eur Urol, 2009

124. van Rhijn BW, Vis AN, van der Kwast TH, et al: Molecular
grading of urothelial cell carcinoma with fibroblast growth
factor receptor 3 and MIB-1 is superior to pathologic grade
for the prediction of clinical outcome. J Clin Oncol 21:191221, 2003

142. Shen SS, Smith CL, Hsieh JT, et al: Expression of estrogen
receptors-alpha and -beta in bladder cancer cell lines and
human bladder tumor tissue. Cancer 106:2610-6, 2006

125. Lindgren D, Liedberg F, Andersson A, et al: Molecular


characterization of early-stage bladder carcinomas by
expression profiles, FGFR3 mutation status, and loss of
9q. Oncogene 25:2685-96, 2006

143. Basakci A, Kirkali Z, Tuzel E, et al: Prognostic significance


of estrogen receptor expression in superficial transitional
cell carcinoma of the urinary bladder. Eur Urol 41:342-5,
2002

126. Dyrskjot L, Zieger K, Kruhoffer M, et al: A molecular


signature in superficial bladder carcinoma predicts clinical
outcome. Clin Cancer Res 11:4029-36, 2005
127. van Rhijn BW, Zuiverloon TC, Vis AN, et al: Molecular
grade (FGFR3/MIB-1) and EORTC risk scores are
predictive in primary non-muscle-invasive bladder cancer.
Eur Urol 58:433-41, 2010

144. Cordon-Cardo C, Zhang ZF, Dalbagni G, et al: Cooperative


effects of p53 and pRB alterations in primary superficial
bladder tumors. Cancer Res 57:1217-21, 1997
145. Shariat SF, Karakiewicz PI, Ashfaq R, et al: Combination
of cell cycle regulating bio-markers improves prognosis
in patients with organ-confined urothelial carcinoma at
radical cystectomy. J Urol 179:578, 2008

128. Kassouf W, Black PC, Tuziak T, et al: Distinctive expression

146. Rajah R, Valentinis B, Cohen P: Insulin-like growth

227

factor (IGF)-binding protein-3 induces apoptosis and


mediates the effects of transforming growth factor-beta1
on programmed cell death through a p53- and IGFindependent mechanism. J Biol Chem 272:12181-8, 1997

predictor of poor survival in bladder cancer. Cancer


Science 101:1300-8, 2010
165. Gohji K, Fujimoto N, Komiyama T, et al: Elevation of
serum levels of matrix metalloproteinase-2 and -3 as
new predictors of recurrence in patients with urothelial
carcinoma. Cancer 78:2379-87, 1996

147. Hoeflich A, Fettscher O, Lahm H, et al: Overexpression


of insulin-like growth factor-binding protein-2 results in
increased tumorigenic potential in Y-1 adrenocortical
tumor cells. Cancer Res 60:834-8, 2000

166. Staack A, Badendieck S, Schnorr D, et al: Combined


determination of plasma MMP2, MMP9, and TIMP1
improves the non-invasive detection of transitional cell
carcinoma of the bladder. BMC Urol 6:19, 2006

148. Butt AJ, Firth SM, Baxter RC: The IGF axis and programmed
cell death. Immunol Cell Biol 77:256-62, 1999
149. Rajah R, Katz L, Nunn S, et al: Insulin-like growth factor
binding protein (IGFBP) proteases: functional regulators
of cell growth. Prog Growth Factor Res 6:273-84, 1995

167. Gohji K, Fujimoto N, Fujii A, et al: Prognostic significance


of circulating matrix metalloproteinase-2 to tissue inhibitor
of metalloproteinases-2 ratio in recurrence of urothelial
cancer after complete resection. Cancer Res 56:3196-8,
1996

150. Shariat SF, Kim J, Nguyen C, et al: Correlation of


preoperative levels of IGF-I and IGFBP-3 with pathologic
parameters and clinical outcome in patients with bladder
cancer. Urology 61:359-64, 2003

168. Guan K-P, Ye H-Y, Yan Z, et al: Serum levels of endostatin


and matrix metalloproteinase-9 associated with high
stage and grade primary transitional cell carcinoma of the
bladder. Urology 61:719-23, 2003

151. Kim JH, Shariat SF, Kim IY, et al: Predictive value of
expression of transforming growth factor-beta(1) and
its receptors in transitional cell carcinoma of the urinary
bladder. Cancer 92:1475-83, 2001

169. Wang Y: The role and regulation of urokinase-type


plasminogen activator receptor gene expression in cancer
invasion and metastasis. Med Res Rev 21:146-70., 2001

152. Eder IE, Stenzl A, Hobisch A, et al: Transforming growth


factors-beta 1 and beta 2 in serum and urine from patients
with bladder carcinoma. The Journal of Urology 156:953-7,
1996

170. Andreasen PA, Egelund R, Petersen HH: The plasminogen


activation system in tumor growth, invasion, and
metastasis. Cell Mol Life Sci 57:25-40, 2000

153. Shariat SF, Kim JH, Andrews B, et al: Preoperative plasma


levels of transforming growth factor beta(1) strongly predict
clinical outcome in patients with bladder carcinoma.
Cancer 92:2985-92, 2001

171. Hudson MA, McReynolds LM: Urokinase and the urokinase


receptor: association with in vitro invasiveness of human
bladder cancer cell lines. J Natl Cancer Inst 89:709-17,
1997

154. Nagase H, Woessner JF: Matrix metalloproteinases. J Biol


Chem 274:21491-4, 1999

172. Shariat SF, Monoski MA, Andrews B, et al: Association


of plasma urokinase-type plasminogen activator and
its receptor with clinical outcome in patients undergoing
radical cystectomy for transitional cell carcinoma of the
bladder. Urology 61:1053-8, 2003

155. Ito T-K, Ishii G, Saito S, et al: Degradation of soluble VEGF


receptor-1 by MMP-7 allows VEGF access to endothelial
cells. Blood 113:2363-9, 2009

173. Okamoto M, Hattori K, Oyasu R: Interleukin-6 functions as


an autocrine growth factor in human bladder carcinoma
cell lines in vitro. Int J Cancer 72:149-54, 1997

156. Cornelius LA, Nehring LC, Harding E, et al: Matrix


metalloproteinases generate angiostatin: effects on
neovascularization. J Immunol 161:6845-52, 1998

174. Andrews B, Shariat SF, Kim J-H, et al: Preoperative plasma


levels of interleukin-6 and its soluble receptor predict
disease recurrence and survival of patients with bladder
cancer. The Journal of Urology 167:1475-81, 2002

157. Ferreras M, Felbor U, Lenhard T, et al: Generation and


degradation of human endostatin proteins by various
proteinases. FEBS Lett 486:247-51, 2000
158. Shiomi T, Okada Y: MT1-MMP and MMP-7 in invasion and
metastasis of human cancers. Cancer Metastasis Rev
22:145-52, 2003

175. Takeichi M: Cadherin cell adhesion receptors as a


morphogenetic regulator. Science 251:1451-5, 1991
176. Matsuura K, Kawanishi J, Fujii S, et al: Altered expression
of E-cadherin in gastric cancer tissues and carcinomatous
fluid. British Journal of Cancer 66:1122-30, 1992

159. Vasala K, Kuvaja P, Turpeenniemi-Hujanen T: Low


circulating levels of ProMMP-2 are associated with
adverse prognosis in bladder cancer. Tumour Biol 29:27986, 2008

177. Popov Z, Gil-Diez de Medina S, Lefrere-Belda MA, et


al: Low E-cadherin expression in bladder cancer at the
transcriptional and protein level provides prognostic
information. British Journal of Cancer 83:209-14, 2000

160. Vasala K, Pkko P, Turpeenniemi-Hujanen T: Matrix


metalloproteinase-9 (MMP-9) immunoreactive protein in
urinary bladder cancer: a marker of favorable prognosis.
Anticancer Res 28:1757-61, 2008

178. Byrne RR, Shariat SF, Brown R, et al: E-cadherin


immunostaining of bladder transitional cell carcinoma,
carcinoma in situ and lymph node metastases with longterm followup. The Journal of Urology 165:1473-9, 2001

161. Vasala K, Pkk P, Turpeenniemi-Hujanen T: Matrix


metalloproteinase-2 immunoreactive protein as a
prognostic marker in bladder cancer. Urology 62:952-7,
2003

179. Jones JL, Royall JE, Walker RA: E-cadherin relates to


EGFR expression and lymph node metastasis in primary
breast carcinoma. British Journal of Cancer 74:1237-41,
1996

162. Wallard MJ, Pennington CJ, Veerakumarasivam A, et al:


Comprehensive profiling and localisation of the matrix
metalloproteinases in urothelial carcinoma. British Journal
of Cancer 94:569-77, 2006

180. Bremnes RM, Veve R, Gabrielson E, et al: High-throughput


tissue microarray analysis used to evaluate biology and
prognostic significance of the E-cadherin pathway in nonsmall-cell lung cancer. J Clin Oncol 20:2417-28, 2002

163. Svatek RS, Shah JB, Xing J, et al: A multiplexed, particlebased flow cytometric assay identified plasma matrix
metalloproteinase-7 to be associated with cancer-related
death among patients with bladder cancer. Cancer
116:4513-9, 2010

181. Shariat SF, Pahlavan S, Baseman AG, et al: E-cadherin


expression predicts clinical outcome in carcinoma in situ
of the urinary bladder. Urology 57:60-5, 2001

164. Szarvas T, Becker M, Vom Dorp F, et al: Matrix


metalloproteinase-7 as a marker of metastasis and

182. Shariat SF, Matsumoto K, Casella R, et al: Urinary levels

228

of soluble e-cadherin in the detection of transitional cell


carcinoma of the urinary bladder. Eur Urol 48:69-76, 2005
183. Damsky CH, Richa J, Solter D, et al: Identification and
purification of a cell surface glycoprotein mediating
intercellular adhesion in embryonic and adult tissue. Cell
34:455-66, 1983
184. Wheelock MJ, Buck CA, Bechtol KB, et al: Soluble 80-kd
fragment of cell-CAM 120/80 disrupts cell-cell adhesion. J
Cell Biochem 34:187-202, 1987
185. Griffiths TR, Brotherick I, Bishop RI, et al: Cell adhesion
molecules in bladder cancer: soluble serum E-cadherin
correlates with predictors of recurrence. British Journal of
Cancer 74:579-84, 1996
186. Matsumoto K, Shariat SF, Casella R, et al: Preoperative
plasma soluble E-cadherin predicts metastases to lymph
nodes and prognosis in patients undergoing radical
cystectomy. The Journal of Urology 170:2248-52, 2003
187. Blaveri E, Simko JP, Korkola JE, et al: Bladder cancer
outcome and subtype classification by gene expression.
Clin Cancer Res 11:4044-55, 2005
188. Sanchez-Carbayo M, Socci ND, Lozano J, et al: Defining
molecular profiles of poor outcome in patients with invasive
bladder cancer using oligonucleotide microarrays. J Clin
Oncol 24:778-89, 2006
189. Lindgren D, Frigyesi A, Gudjonsson S, et al: Combined
gene expression and genomic profiling define two intrinsic
molecular subtypes of urothelial carcinoma and gene
signatures for molecular grading and outcome. Cancer
Res 70:3463-72
190. Takata R, Katagiri T, Kanehira M, et al: Predicting response
to methotrexate, vinblastine, doxorubicin, and cisplatin
neoadjuvant chemotherapy for bladder cancers through
genome-wide gene expression profiling. Clin Cancer Res
11:2625-36, 2005

229

230

Committee 4

Low Grade Ta Urothelial


Carcinoma of the Bladder
Chairs:
Badrinath Konety,
Willem Oosterlinck
Members:
Sam Chang,
Sigurdur Gudjonsson,
Raj Pruthi,
Mark Soloway,
Eduardo Solsona,
Paul Sved

231

Table of Contents
I. Urinary cytology, urinary
markers and the follow-up
of low grade non muscle
invasive bladder tumors

III. Expectant Management of


LG Ta Urothelial Carcinoma
IV. Intravesical Treatment for
LG Ta Tumors

II. Upper Tract Studies in


Patients with LG Ta Urothelial
Carcinoma

1. Intravesical Chemotherapy

V. Immunotherapy (BCG) in LG Ta
Patients

1. Upper Tract Studies at the Time


of Diagnosis of LG Ta Urothelial
Carcinoma of the Bladder
2. Upper Tract Surveillance of
Patients with Low Grade Ta
Urothelial Cancer

VI. Other Immunomodulating


Therapies

232

Low Grade Ta Urothelial Carcinoma


of the Bladder
Badrinath Konety, Willem Oosterlinck,
Sam Chang, Sigurdur Gudjonsson,
Raj Pruthi, Mark Soloway,
Eduardo Solsona, Paul Sved

In this section we have attempted to review all the


literature pertaining to non-muscle-invasive bladder
cancer and cull out the facts specifically pertaining
to low grade Ta (LG Ta) urothelial carcinoma. The
evidence available is often buried in broad studies of
all non-muscle-invasive bladder cancer and hence
the patients with LGTa urothelial carcinoma constitute
a smaller subgroup. Such subgroup analyses may
be inadequately powered, given the original primary
endpoints of many of the studies which would have
taken into account sample sizes for all non-muscleinvasive bladder cancer together. The evidence
available was analyzed using the Oxford method
of assigning the levels of evidence and summary
recommendations based on these levels of evidence
were graded as advised by the Oxford Centre
for Evidence-based Medicine, which are similar
to the GRADE working group recommendations
[1]. At the end of each section, a summary of the
recommendations and a ready reference table with
the recommendations and the levels of evidence
as well as the grade of the recommendation are
provided.

I. Urinary cytology, urinary


markers and the follow-up
of low grade non muscle
invasive bladder tumors
Extensive laboratory research has led to the
development of numerous urinary markers for
diagnosis of bladder cancer. The number of
publications identified through a PubMed search are
enormous when bladder cancer and biomarkers
or molecular markers are searched [2]. Many of
them exhibit sensitivity superior to urine cytology,
often with lower specificity. A few of them have been
applied clinically, but none has been accepted as a
standard diagnostic test in routine urology or included
in published guidelines to date. Insufficient and
inconsistent evidence for the use of such markers in
surveillance of non-muscle-invasive bladder cancer
is the reason for this [2,3].
The aim of surveillance in LG Ta tumors is mainly
for the timely detection of recurrences. LG Ta tumors
progress to more aggressive tumors in less than
1% of the cases. Recurrences are frequent, and the
main prognostic factors are multiplicity of the tumors
and rapid recurrence. Most guidelines advocate
cystoscopy at 3 months after initial resection in order
to detect rapid recurrence but subsequent cystoscopy
at yearly intervals thereafter if no tumor is detected
on first post-resection cystoscopy [4]. Besides that,
office fulguration of small, isolated recurrences is an
acceptable and widely used method for the treatment
of recurrent LG Ta tumors.

The initial diagnosis of LG Ta urothelial carcinoma


requires the performance of a transurethral resection
or biopsy. It is recommended that the presence
of muscle on the initial resection specimen be
documented to ensure accurate stage designation
as Ta. It is not required that biopsies for recurrent
LG Ta urothelial carcinoma have muscle present in
the specimen. This may be particularly the case with
office biopsies under local anesthesia where a deep
specimen will be difficult to obtain. Forcep biopsies
can only be pursued in patients with a documented
history of recurrent LG Ta urothelial carcinoma and
not for the initial presentation of bladder tumor.

What is the role of urinary cytology and markers in


this context?

233

Requirements for a good marker in this context


follow:

Ta tumors could be better identified by urinary


markers. Hence, urinary markers may have a role
in a low intensity surveillance program where the
cystoscopy interval is longer, such as 1 year, to
monitor for larger, recurrent LG Ta tumors during the
intervals. Large prospective randomized studies
of alternative surveillance regimens incorporating
urinary markers for low grade tumors are required
to definitely establish the utility of such markers in
the surveillance of LG Ta tumors.

1. It must detect all high grade tumors while they


are still amenable to curative treatment. Urinary
cytology is highly sensitive in this regard and
most commercially available urinary markers are
even slightly more sensitive. The problem with
cytology is interobserver variability [5,6] due to the
subjectivity of the test. Urinary markers are more
objective. The problem with the urinary markers
is the frequent occurrence of false positive tests
and thus the lack of specificity. As the likelihood
of LG Ta to high grade tumors is low, cytology and
markers are of limited value in this regard.

3. T
 he test should also be well accepted by the
patient, regarding the risk of overlooking a tumor
versus the discomfort and slight risks of the
cystoscopy.

2. In order to be able to replace cystoscopy, a urinary


marker should be able to detect recurrent tumors
before they are large and numerous. Cytology
performs poorly in this context. Several urinary
markers do better but still do not detect a large
proportion of the low grade tumors detectable by
cystoscopy. Among the commercially available
urinary markers, some of the best results for
detecting recurrent LG Ta tumors have been
obtained using immunocytology (uCyt). But
even this test has not demonstrated adequate
consistency and reliability to be able to replace
cystoscopy. One potential role for urinary markers
in the setting of LG Ta tumors is for reducing the
frequency of cystoscopy wherein these markers
can be utilized alternatively or interspersed with
cystoscopy for surveillance. Most protein-based
urinary markers such as NMP22 are affected
by the volume of the tumor, but recurrent low
grade tumors often are small. Larger, multiple LG

4. Preferentially tests should be point-of-care test,


with results available readily, easy to perform, with
a short learning curve, widely available, and not
too expensive. Table 1 [7] gives an overview of
how far the available urinary markers correspond
to these criteria.
Numerous reviews on urinary markers have
appeared in the last several years supporting the
statements made above [8-14]. They have all
come to the conclusion that urinary markers are
insufficiently tested in surveillance of non-muscleinvasive bladder cancer and that cystoscopy cannot
yet be omitted but at best can perhaps be delayed.
About the existing markers we can conclude the
following: The BTA test has a very limited role
because of its high false positive rate and low
sensitivity in low grade tumors [15,16]. NMP22
similarly suffers from high false positive rate but
may have higher sensitivity than cytology, and

Table 1. Summary of main urinary markers that could be used in Lg Ta UC

Markers

Overall

Overall

Interference by
BCG

Comments

UroVysion
Microsatellite
analysis
Gene microarray

sensitivity
(%)
30-72

specificity
(%)
63-95

and other bladder


conditions
No

Expensive and laborious

58
80-90

73
62-65

No
No

Immunocyt/ uCyt+
Nuclear matrix
protein 22

76-85

63-75

Yes

49-68

85.8-87.5

Yes

BTA stat

57-83

68-85.7

Yes

Low sensititivy

BTA TRAK

53.8-91

28.3-83.9

Yes

Low sensitivity

Cytokeratins

12.1-85

75-97.4

Yes

Survivin

53-90.4

88-100

No

Low sensitivity
Low sensitivity, expensive and
laborious

234

Expensive and laborious


Expensive and laborious
Good sensitivity in low-grade
tumors
Low sensitivity,

with selection of patients the specificity can be


improved [17-19]. Because of its high negative
predictive value, it can be used to prolong the
interval between cystoscopies. ImmunoCyt has the
highest sensitivity in detection of low grade tumors
and is less minimally affected by other urological
diseases [20], but with a 60% detection rate of low
grade tumors, the test remains largely insufficient to
replace cystoscopy. The combination of ImmunoCyt
with cytology appears to improve sensitivity and
negative predictive value, which can allow for more
accurate identification of those individuals who will
have a negative cystoscopy [21].

fluorescent cystoscopy has been shown to be able


to detect additional lesions compared to white light
cystoscopy. In at least one study, photodynamic
diagnosis using Hexvix was able to detect 25
additional LG Ta lesions in a cohort of 108 patients
with Ta bladder tumors [27]. Several randomized
studies have demonstrated the ability of fluorescent
cystoscopy-aided TUR to reduce recurrences as
well as improve recurrence-free survival. However,
these data do not specifically pertain to those with
LG Ta tumors but pertain to all non-muscle-invasive
bladder cancer [27]. At this time, the precise role of
photodynamic diagnosis in the diagnosis and followup of LG Ta tumors is evolving.

Urovysion/FISH adds little in the surveillance of low


grade tumors but can perhaps be used to adjudicate
inconclusive results of urinary cytology (e.g., atypical)
in order to prevent excessive work-up of such
patients [22]. FISH may have a larger role in high
grade and higher stage tumors as it can detect occult
disease and identify those patients who are likely to
recur and may be useful for predicting response to
intravesical therapy [23,24]. Microsatellite analysis
holds promise for predicting recurrences of low
grade tumors, as it appears to be able to identify
such recurrences in up to 80% of the cases but is
lacking in sensitivity [25,26].
More

recently,

photodynamic

diagnosis

Summary
Protein based urinary markers such as BTA and
NMP22 may have better sensitivity than cytology
for LG Ta tumors, but their specificity is too low to
allow them to replace cystoscopy. ImmunoCyt has
the best sensitivity for LG Ta tumors and can be
used to prolong intervals between cystoscopy but
cannot replace cystoscopy. FISH analysis does
not have a clear role in diagnosis of LG Ta tumors.
Photodynamic diagnosis may have benefit in
enhanced detection of LG Ta tumors but is limited by
low specificity and need for rigid cystoscopy.

or

Recommendation
Recommendation

Level of

Grade of
recommendation
B
A

BTA has little role in diagnosis of LG Ta tumors.


NMP22 has better sensitivity than cytology but cannot replace

evidence
2b
1b

cystoscopy.
ImmunoCyt has good sensitivity for LG Ta tumors and can be

2a

used to prolong intervals between cystoscopy.


UroVysion/FISH has no role in diagnosis of LG Ta tumors.

2b

II. Upper Tract Studies in


Patients with LG Ta Urothelial
Carcinoma

Information regarding the incidence of synchronous


upper tract urothelial carcinoma in patients with
noninvasive urothelial carcinoma is difficult to interpret
because most reports include patients with high risk
disease such as carcinoma in situ in addition to those
patients with low grade disease. Bajaj et al reported
that 4 of 120 (3.3%) patients with Ta bladder tumors
were diagnosed with a synchronous upper tract
tumor[28]. Palou et al, in a retrospective analysis of
1529 patients with noninvasive urothelial carcinoma,
found 6 synchronous upper tract tumors in 547
(1.1%) patients with LG Ta disease[29]. Goessl et al
reported no synchronous upper tract tumors in 207
patients with noninvasive bladder cancer [30].

1. Upper Tract Studies at the Time


of Diagnosis of LG Ta Urothelial
Carcinoma of the Bladder
Urothelial carcinomas of the urinary tract exhibit a
well-known tendency for multifocal growth. Despite
this, the use of imaging to detect synchronous renal
or ureteric urothelial tumors at the time of diagnosis
of LG Ta urothelial carcinoma of the bladder is
controversial. Some authors advocate routine
imaging of the upper tracts while others suggest
restricting such imaging to high risk patients,
including those with multiple bladder lesions[28-31].
In addition, the optimal imaging technique remains
undetermined.

Little information is available regarding the


clinicopathological features of low grade bladder
cancers most likely to be associated with a
synchronous upper tract urothelial carcinoma. It has

235

been suggested that multifocal non-invasive bladder


cancer is associated with a higher risk of concomitant
upper tract disease but there is no evidence that this
is the case if the bladder cancer is low grade [32].
Palou et al reported that only trigonal involvement
was associated with upper tract urothelial carcinoma
at diagnosis [29]. However, in this study low grade
tumors were not analyzed in isolation and it is
therefore possible that trigonal tumors were in fact
high grade.

additional staging information, including imaging of


pelvic lymph nodes and the detection of other intraabdominal abnormalities. The positive predictive
value of CT urography is over 80% for lesions greater
than 5 mm [33].

Summary
Synchronous upper tract urothelial carcinoma is rare
in patients with LG Ta urothelial carcinoma. Routine
IVU or CT urography is not recommended for these
patients at the time of diagnosis. However, thorough
upper tract evaluation is required for patients
presenting with gross hematuria and/or unexplained
positive urine cytology in the presence or absence of
a low grade noninvasive bladder tumor.

Intravenous urography (IVU) has traditionally been


the gold standard in the detection of upper tract
urothelial carcinoma. More recently, CT urography
has emerged as a technique to evaluate the upper
tract[33-35]. CT urography has a higher sensitivity
and specificity compared with IVU and provides

Recommendation
Recommendation

Routine upper tract studies are not recommended for patients


with LG Ta tumors at diagnosis of index tumor.
Upper tract evaluation in patients with LG Ta urothelial
carcinoma is recommended in the presence of:

Level of
evidence
2b

Grade of
recommendation
B

2b

o gross hematuria
o unexplained positive urine cytology

2. Upper Tract Surveillance of


Patients with Low Grade Ta
Urothelial Cancer

bladder cancer, Wright et al studied almost 100,000


patients in the Surveillance, Epidemiology and End
Results (SEER) registry [42]. Of 56,271 patients
with a Ta bladder tumor at diagnosis, 472 (0.3%)
were subsequently diagnosed with an upper tract
urothelial carcinoma. This report is especially
significant as it reflects the risk of recurrence in
patients in community practice rather than from
tertiary centers alone. The SEER analysis found that
patients with Ta bladder cancer were more likely to
be diagnosed with an upper tract recurrence than
those with muscle-invasive bladder cancer, although
this may reflect poorer overall survival in the latter
group. The mean time to upper tract recurrence in
the entire cohort was 33 months. Other reports place
the risk of upper tract urothelial carcinoma in patients
with Ta bladder cancer at 0.6- 2.5% [40,42-45], with
a wide range of upper tract recurrence intervals from
5 months to 6.2 years [38,39,46].

The presence of UC in the bladder is a wellrecognized risk factor for the subsequent diagnosis
of upper tract urothelial cancer [36]. Nevertheless,
there remains considerable uncertainty regarding
the value of routine upper tract imaging in the followup of patients with non-invasive urothelial carcinoma.
Some investigators have recommended annual or
biennial IVU, while others have concluded that there
is no need for regular monitoring of the upper tracts
[37-40]. This uncertainty is reflected in the wide
variance in practice of community urologists [41].

a) Incidence and Risk factors


In a retrospective review of 375 patients with Ta
urothelial carcinoma followed for a median of 6
years, Canales et al reported a subsequent upper
tract tumor in 13 (3.4%) patients [36]. The risk of
upper tract recurrence was the same regardless of
the grade or number of initial Ta bladder tumor(s).
Upper tract tumors were diagnosed at an average
of 22 months after initial bladder tumor diagnosis. In
the largest ever review investigating metachronous
upper tract urothelial carcinoma in patients with

Canales et al used multivariate analysis to determine


predictors of upper tract urothelial carcinoma in their
cohort of patients with Ta bladder cancer [36]. They
found that only patients with 2 or more recurrences
within 12 months of diagnosis were at higher risk of
upper tract tumor. Other studies group patients with
Ta and high grade T1 disease together, making the
data very difficult to interpret. These studies suggest

236

a weak link between tumor multiplicity, vesicoureteric


reflux, and upper tract recurrence in non-muscleinvasive bladder cancer [38,42,47,48].

surveillance for LG Ta urothelial carcinoma cannot


be recommended.

Summary

There is no evidence that the prognosis of patients


with upper tract urothelial carcinoma after Ta bladder
cancer is improved by early detection using routine
surveillance IVU, compared to patients presenting
with symptoms such as flank pain or gross hematuria
[36,48,49]. This is also the case for high grade noninvasive bladder cancer requiring BCG [49,50].
Hence routine upper tract surveillance does not
appear to have a clear benefit in patients with high
grade lesions, and therefore can be expected to have
even less relevance for those with low grade lesions.
Such routine surveillance would only add cost and
inconvenience to patients and not yield demonstrable
benefit. Hence the strategy of routine upper tract

The incidence of metachronous upper tract urothelial


carcinoma in patients treated for low grade,
noninvasive bladder cancer is low, particularly in
population-based analyses. There is no consistent
risk factor for upper tract recurrence, which may
occur years after initial diagnosis. There is no
evidence that upper tract lesions detected on routine
imaging are associated with a more favorable
prognosis. Therefore, routine upper tract surveillance
in patients with low grade Ta bladder cancer is not
recommended.

Recommendation
Recommendation

Routine upper tract studies are not recommended for


surveillance of patients with LG Ta tumors during follow-up.

Level of
evidence

Grade of
recommendation

2b

grade or stage progression [55-57]. These frequent


recurrences often result in repeat TURBTs, because
urologists are taught to remove tumors when they
are identified in accordance to past teachings and
guideline recommendations. In other words, past
and present practice has suggested that many
patients will undergo multiple TURBTs to manage
these small, recurrent, low-risk tumors.

III. Expectant Management of


LG Ta Urothelial Carcinoma
Complete transurethral resection is still considered
to be the gold standard in managing noninvasive
bladder tumors. Indeed, EAU guidelines note that the
goal of TURBT in Ta tumors is to make the correct
diagnosis and remove all visible lesions, and NCCN
guidelines explicitly state that TURBT is the standard
treatment for cTa, low grade tumors.[51,52] Current
AUA guidelines also suggest that TURBT is the
treatment of choice for most non-invasive bladder
tumors, but go on to acknowledge the potential role
of conservative management for certain patients with
recurrent low risk noninvasive disease.[53]

In deciding treatment options, other important aspects


are not only tumor recurrence rates but also grade
and stage progression, as this denotes a decidedly
poorer prognosis in bladder cancer. The natural
history of low risk bladder tumors is well-established,
and the rates of progression and mortality from
bladder cancer in patients with this type of tumor are
extremely low [55-58]. This is regardless of age at
presentation. The low likelihood of progression and
the indolent nature of certain types of bladder tumors
may therefore have a negligible effect on survival in
select patients who will die with their disease and not
from it. [55]

In this section, we will address the potential role of


observation of low risk bladder tumors by evaluating
the rationale for such an approach as well as the
clinical experience and reported evidence to date.
The terms observation, expectant management,
and active surveillance will be used interchangeably
for these discussions and recommendations.

Furthermore, though tumor resection is usually welltolerated, it is not without risks, especially in the elderly
or those with significant medical comorbidities. These
patients face greater risks with regional or general
anesthesia and often require postoperative hospital
stays, both of which increase cost and the burden to
our health care systems. [59] Complications directly
associated with TURBT must also be acknowledged,

Rationale
Approximately 50% of newly diagnosed cases of
bladder cancer are low risk papillary tumors (LG
Ta) [54]. Although the majority of these patients
will recur after initial TURBT, few will demonstrate

237

including bleeding and perforation in the short-term,


and bladder contracture and urethral stricture in the
long-term. [60]

stage progression as well. Both men had evidence of


moderate tumor growth and suspicious or malignant
cytology heralding such progression and resulting
in repeat TURBT. However, similar to the results of
Soloway et al, neither of these tumors progressed to
muscle-invasive disease. [64,66] Interestingly, these
2 men had prolonged histories of recurrent LG Ta
disease before progression, thus demonstrating the
ability of low risk tumors to progress even at distant
follow-up. Accordingly, the recommendations for
lifelong surveillance of all bladder tumor patients
seem justified. In what appears to be the largest
case series, Hernandez et al recently published their
experience with observation in 66 patients with low
risk bladder tumors. [67] In this series, the authors
selected patients with smaller tumors (less than 1
cm), but also included patients with T1a disease,
which are generally accepted as being higher risk
than LG Ta tumors. Despite this, at a median followup of over 10 months, the majority of patients had
neither stage (94%) nor grade (84%) progression,
and, like the prior studies, no patient progressed to
muscle-invasive disease.

Consequently, some have recently questioned as to


whether such an aggressive follow-up and treatment
regimen is warranted for these low risk patients.
Surveillance for bladder tumors, as compared
to prostate cancer or renal masses, is relatively
easy to perform, as the clinician has direct ways
to monitor the bladder urothelium through cytology
and cystoscopy. Indeed, it has been shown in a
study by Herr et al that low grade papillary tumors
were accurately identified at cystoscopy 93% of
the time, and when information from urine cytology
was added, the accuracy increased to 99%. [61]
Furthermore, the cystoscopic findings of bladder
cancer are accurate in predicting muscle-invasive
disease as well. [62] If staging and grading can be
accurately performed by cystoscopy, many patients
may be spared unnecessary trips to the operating
room.
Therefore, the rationale for observation of low risk
bladder tumors is to spare patients the morbidity and
inherent risks of repeat TURBT. Short- or long-term
surveillance of tumors has the potential to reduce the
number of resections that patients would undergo in
their lifetime. For the bladder cancer patient and the
treating urologist, the proper balance must be found
between the risk of progression and the patients
current state of health and medical comorbidities.
The decision to observe versus intervene is a balance
between the morbidity of treatment and follow-up
versus the risks of progression in each individual
patient. As has been suggested by Soloway, the
emphasis should be to do no harm and to avoid over
treatment. [63]

From these series, there does not appear to be any


clear patient selection criteria aside from the low
risk features that were observed in the majority of
patients. [64-67] Common findings in these studies
included patients with Ta tumors, low grade disease,
and low tumor burden (smaller tumors, fewer
lesions). The studies often recommend careful
and individualized patient selection. In fact, in the
studies of Soloway et al and Pruthi et al, only a
small percentage of their patient population were
included into an observation protocol (8% and 13%,
respectively). [44,66] Furthermore, most recommend
detailed discussion with patients as to the rationale
and strategy of observation. Indeed, in the study by
Hernandez, the authors note that once the treatment
alternatives were explained, a high percentage of
patients preferred to undergo close monitoring by
cystoscopy and to delay the surgery for as long as
possible. [67]

Published Data
Soloway et al were the first to expectantly observe
patients with low risk disease. They observed that in
32 patients with stage Ta or T1 tumors, only 3 (9%)
progressed in either stage or grade, and none had
progression to stage T2 disease.[64] Similarly, Gofrit
et al also demonstrated that expectant management
may be an acceptable course in patients with low
risk disease.[65] They followed 28 patients for 38
observation periods. Although 30 of these periods
were terminated because of recurrence or additional
growth, all resected tumors were found to be Ta
and grade 1 or 2 (i.e., none had grade or stage
progression). They also observed that initial tumor
diameter was predictive of tumor growth during the
observation period.

There also exists a growing body of evidence as


to the potential use of office fulguration for low risk
tumors, i.e. partial or total tumor ablation during
local cystoscopy, often at the time of surveillance.
Although the present discussion is focused on the
potential role of observation, office fulguration may
serve as an appropriate alternative to formal TURBT
in select patients. [68-70]
As urologists, we must weigh the risks and benefits
of aggressive versus conservative treatment in
dealing with non-invasive bladder tumors. The goal
of TURBT is to eliminate any visible tumor, and
this is the most assured way to prevent disease
progression. However, it does so with an increased
surgical and perioperative risk and cost. A balance
must be found between the risk of progression and

Pruthi et al showed that the majority of patients


with low-risk disease had no evidence of stage
or grade progression on a surveillance program.
[66] Two men (9%) were observed to have grade
progression of their disease, with 1 (4.5%) having

238

the patients current state of health and medical


comorbidities. Although expectant management
appears to be a safe treatment option, it is likely not
the best choice for everyone.

Future investigation should involve better quality


studies with higher levels of evidence to evaluate 1)
oncologic appropriateness of expectant management,
2) the impact of surveillance programs on the overall
health status and survival of the patients (including
potential benefits of avoiding repeated operating
room procedures), 3) effects on quality of life in
patients undergoing tumor observation versus repeat
TURBTs, and 4) potential impact of surveillance
programs on the economic costs and utilization of
our health care systems.

Of note, observation (or office fulguration), should


not replace formal TURBT with the patient under
anesthesia as primary treatment for the initial tumor
or for recurrence suspected on gross inspection to
represent a change in tumor stage or grade. TURBT
not only establishes the clinical stage of the tumor,
but it also serves as the basis for adjuvant treatment
decisions and surveillance intervals. Whenever there
is clinical doubt whether a patient is a candidate for
observation or fulguration, the patient should always
undergo a formal TURBT.

Summary
Observation can be considered for patients with
recurrent LGTa urothelial carcinoma (or PUNLMP) as
proven by prior TURBT. Optimal tumor characteristics
include those with low tumor burden. Optimal patient
characteristics include those with advanced age and
comorbidities, those with frequent recurrences, and
those without symptoms. Younger patients can also
be offered observation as long as they are aware
of the risks and benefits of immediate TUR vs.
observation. Observation includes cystoscopy and
urinary cytology at routine surveillance intervals.
Patients should be appropriately counseled as to the
treatment strategy and rationale, as well as to the
potential risks inherent with observation of a malignant
tumor. If the cytology is suspicious or positive, the
patient should be taken to the operating room for
a formal TURBT and other appropriate evaluation
under anesthesia. If there is increase in the size and
number of lesions, or other concerning changes on
gross appearance of the urothelium, then the patient
should be taken to the operating room for a formal
TURBT under anesthesia. For small lesions (less
than 1 cm) associated with negative cytology, office
fulguration is also an option. This can be followed by
immediate instillation of intravesical chemotherapy
to prevent recurrence (see below).

We
must
acknowledge
several
possible
shortcomings of the data available in regard to
the expectant management of LG Ta urothelial
carcinoma. Most of the current evidence is from
retrospective observational case series that do not
have the strength of a randomized control trial, the
gold standard of evaluating treatment efficacy and
superiority. Accordingly, there have been no clear and
uniform selection criteria for expectant management.
In addition, certain assessments in the reported case
series were of a subjective nature, including tumor
growth assessments, decisions to intervene, and
subsequent management. Nevertheless, qualitative
assessments of tumor change combined with more
quantitative evaluations of urine cytopathology seem
to best reflect the tools used in everyday clinical
practice. Such tools combined with the ability for
urologists to appropriately determine, cystoscopically,
when a recurrent tumor has progressed (without the
need for TURBT or biopsy) as demonstrated by
Herr et al, appear to be adequate in identifying early
disease progression before muscle-invasive disease
occurs. [61]

Recommendation
Recommendation




Expectant management should be pursued only in patients


with an established history of LG Ta urothelial carcinoma.
Optimal tumor characteristics include low tumor burden.
Observation is particularly applicable to those with advanced
age or comorbidity, but can be offered to younger patients
after careful and thorough discussion.
Surveillance includes periodic cystoscopy and cytology.
Patients who demonstrate increase in size, number or
appearance of tumor(s), or develop positive urine cytology
should cease expectant management and undergo formal
TUR and further evaluation.
Office fulguration is a good option for patients with small LG
Ta urothelial carcinoma. This can be followed by immediate
instillation of intravesical chemotherapy.

239

Level of
evidence
3

Grade of
recommendation
B

3
4

B
C

3
3

B
B

IV. Intravesical Treatment for


LG Ta Tumors

ultrasound bladder scanner), and a higher mitomycin


C concentration (40 mg in 20 mL of sterile water).
Other methods exist but have been examined in
higher risk individuals.

1. Intravesical Chemotherapy

c)
Intercalating
Agents
Epirubicin, and Valrubicin)

In most cases of non-muscle-invasive bladder


cancers, TURBT plays an essential diagnostic
and therapeutic role, and this is certainly the case
for LG Ta tumors. A careful cystoscopic evaluation
and eradication of tumor is an important initial
therapeutic intervention. While in some cases tumor
ablation may suffice as initial therapy, intravesical
chemotherapy plays an important role in decreasing
the recurrence rates of these tumors.[71-73]

(Doxorubicin,

The rationale for perioperative instillation includes the


destruction of residual microscopic tumor at the site
of TURBT and of floating cells, thereby preventing
reimplantation at the time of TURBT. Intravesical
therapy can also be employed in a maintenance
fashion as opposed to an induction course alone to
prolong the beneficial effects of the chemotherapy.

Because of its high molecular weight of 580 kD,


absorption and systemic toxicity of the anthracycline
derivative doxorubicin are extremely rare. Doses
vary widely, from 10 mg to 100 mg, in instillation
schedules that range from 3 times a week to
once a month. Epirubicin is related chemically to
doxorubicin and has been used in multiple studies
for initial post-TUR instillation as well as for multidose intravesical chemotherapy. It appears to have
a favorable toxicity profile and good efficacy. It offers
an attractive alternative to mitomycin C. Valrubicin, a
semi-synthetic analog of doxorubicin, was approved
by the US FDA in 1998 for the treatment of BCGrefractory carcinoma in situ of the bladder in patients
who are medically unfit or refuse a cystectomy.

Chemotherapeutic Agents

d) Gemcitabine

Chemotherapeutic agents include thiotepa, an


alkylating agent that cross-links with nucleic acids;
mitomycin, an antibiotic that inhibits DNA synthesis;
epirubicin, doxorubicin, and valrubicin, intercalating
agents that inhibit DNA synthesis; gemcitabine, a
deoxcytidine analog that inhibits DNA synthesis.

Gemcitabine has a broad spectrum of antitumor


activity and intravesical gemcitabine has been shown
to have activity in non-muscle-invasive bladder
cancer in intermediate risk and high risk patients.
Typical intravesical doses employed include 2 g in
50 to 100 mL of saline given weekly for 6 weeks with
2 hour dwell times. Gemcitabine is currently being
tested for post-TUR instillation in all tumors as part
of a study sponsored by the Southwest Oncology
Group in the US. In a randomized placebo controlled
study, Bohle et al [76] compared post-TUR instillation
of gemcitabine to bladder irrigation with saline and
found no difference in recurrence rates for tumors
of all stages/grades. Hence immediate post-TUR
instillation of gemcitabine does not appear to yield
any benefit over bladder irrigation in preventing
recurrences in patients with LG Ta urothelial
carcinoma.

a) Thiotepa
Introduced in 1961, thiotepa is the oldest and one
of the least expensive of the intravesical drugs.
Doses range from 30 mg in 30 mL of sterile water
or saline to 60 mg in 60 mL of water or saline. The
usual regimen consists of 6 to 8 weekly instillations
followed by monthly instillations for 1 year, but there
is no standard regimen. A low molecular weight of
189 KD increases possible systemic toxicity, and
myelosuppression is a risk. Importantly, the 2007
AUA Guidelines do not recommend thiotepa as an
effective chemotherapy agent. [74]

e) Therapeutic Schedules

b) Mitomycin C

The ideal intravesical chemotherapy regimen would


maximize effectiveness while decreasing the number
of therapies and maintaining a favorable side effect
profile. Data, although sometimes conflicting, is
available regarding the likelihood of treatment
efficacy for these agents. Recent meta-analyses
have attempted to clarify a sometimes murky picture.
Further clouding the situation is the fact that most
studies do not focus solely on LG Ta tumors but
instead on higher risk tumors.

This agent has a molecular weight of 329 KD,


and the side effect of myelosuppression is less
likely. Although the optimal method of mitomycin
C administration is uncertain, Au et al [75] have
demonstrated improved recurrence-free survival
and a prolonged median time to recurrence using
methods to enhance the effectiveness of this
medication. The strategy consisted of a period of
dehydration (no fluids for 8 hours prior to treatment),
urinary alkalinization (1.3 g NaHC03 by mouth, the
night prior, the morning of, and 30 minutes prior
to the intravesical therapy), confirmed complete
bladder drainage prior to intravesical instillation of
mitomycin C (post void residual less than 10 mL by

f) Perioperative Chemotherapy
A European Organisation for Research and Treatment
of Cancer (EORTC) study comparing epirubicin to
water immediately following TURBT demonstrated

240

that patients with primary and solitary recurrent


Ta tumors had a 50% decrease in recurrence with
epirubicin. [77]

further exploration.
The most common side effect of intravesical
chemotherapy is some form of lower urinary tract
symptoms (dysuria, frequency, and urgency) and
hematuria. Perioperative mitomycin C should
not be administered to patients with a known or
suspected bladder perforation following TURBT
as a small number of serious complications
related to mitomycin C extravasation have been
reported [82].

In a meta-analysis performed by the AUA Guidelines


Panel in 2007, a statistically significant 17%
decrease (95% CI 8-28%) in median recurrence
rate was found. No effect on progression was noted
[4]. Two trials were identified by the AUA Panel as
suitable for meta-analysis and included a combined
427 patients. In the larger trial by Tolley et al [78],
60% of the subjects treated with TURBT alone had
recurrences at 5 years compared to 45% of those
receiving a single, postoperative instillation of 40 mg
of mitomycin in 40 mL of water. Solsona et al [79]
reported a statistically significant decrease in early
recurrence of up to 2 years in patients who received
the single dose of mitomycin C. This study did focus
on lower risk patients; all patients had a 3 cm or less
single, papillary, primary, or recurrent tumor and
were disease-free for more than 1 year. Patients with
muscular invasion, G3 tumor, or bladder carcinoma
in situ on pathologic examination were excluded
from this study.

g) Induction Chemotherapy
At this point, it is unclear if any chemotherapeutic
agent holds superiority over another. A recent trial
advocated the superiority of gemcitabine over
mitomycin C, but the dosing schedule used was
different from those employed in prior studies [83].
Importantly, although maintenance regimens are
essential in higher risk disease such as CIS, there
is no evidence that multiple adjuvant instillations of
either BCG or chemotherapy have additional benefit
in patients at initial diagnosis of LG Ta urothelial
carcinoma [74]. Thus, for initial LG Ta tumors, it is
difficult to justify long-term and/or maintenance
doses of therapy. Therapy should be individualized
based on the prognostic factors of the patients
tumor.

Another large meta-analysis confirms the


effectiveness of single perioperative chemotherapy
versus TURBT alone for decreasing recurrence. This
meta-analysis was performed including 7 randomized
trials, each evaluating the clinical value of a single
immediate instillation of various chemotherapeutic
agents, and reported a 39% decrease in the odds
of recurrence compared to placebo, i.e., TUR alone
[80]. In recurrent and multiple tumors, the benefit of
the early instillation is insufficiently proven by a lack
of statistical power and therefore challenged [81].
To be effective, the instillation should be performed
on the same day as the TUR, as delayed instillation
does not appear to yield similar benefits.

Summary
A well-performed, complete transurethral resection
of the bladder tumor is a critical initial step in the
management of LG Ta tumors and prevention
of recurrences. Immediate post-TUR instillation
of intravesical mitomycin C or epirubicin is
recommended, as it has been shown to prevent
recurrences. Induction with or without maintenance
intravesical chemotherapy for LG Ta tumors may
have benefit, but this has not been established
unequivocally. Given that these patients have a very
low risk of progression and metastases, the risks
of repeated courses of intravesical chemotherapy
have to be weighed against the benefit. Hence we
recommend an extremely judicious and limited use
of repeated courses of intravesical chemotherapy in
this population.

In a recent randomized placebo-controlled trial,


the efficacy of a single postoperative instillation of
gemcitabine (2 g/100 mL of saline) was examined. In
this study of 355 patients with primary or recurrent Ta
or T1, grade1-3 tumors, gemcitabine was not superior
to saline bladder irrigation in terms of recurrencefree survival [76]. The recurrence rate was very low,
suggesting that the bladder irrigation plays a role in
prevention of recurrence, an observation that needs

Recommendation
Recommendation



The initial step in management is complete TUR.


Immediate post-TUR instillation of intravesical chemotherapy
in the form of mitomycin C or epirubicin is recommended as it
can prevent recurrences.
Induction chemotherapy with or without maintenance has
unclear but potential benefit.
The risks of repeated courses of intravesical chemotherapy
have to be weighed against the benefit.

241

Level of
evidence
2b
1a

Grade of
recommendation
A
A

2b

V. Immunotherapy (BCG) in LG Ta
Patients

chemotherapy compared to BCG therapy as rescue


therapy after initial intravesical chemotherapy failure
[95].

Low risk non-muscle-invasive bladder tumors are


characterized by a moderate recurrence rate from 2437% but with very low progression rates from 0-1.7%
[4]. Taking these figures into account the primary
therapeutic objective for low risk patients is to reduce
recurrence rate and the secondary objective is early
detection of progression. To decrease progression
rates in these patients would mean an overtreatment
including unnecessary side effects, as the vast
majority of patients will never develop progression,
so this is not a realistic objective for these patients.
To reach an appropriate balance between efficacy
of the different therapeutic approaches and their
potential toxicity is the main goal for patients with
LG Ta disease.

The role of intravesical BCG as rescue therapy after


chemotherapy failure in low risk groups has been
analyzed in a meta-analysis carried out by Shelley
et al [87]. They examined 1901 patients and found
no significant difference in the efficacy between
intravesical BCG and mitomycin C. However, BCG
was significantly superior to mitomycin C in a
subgroup analysis involving patients at high risk of
recurrence (p=0.0008). Corroborating these data,
in a randomized trial including 174 patients with
recurrent papillary Ta or T1 tumors and comparing
mitomycin C versus BCG, with a median follow-up
of 64 months, BCG achieved a 43% disease-free
survival superior to 34% with mitomycin C but the
difference was not statistically significant (p=0.22)
[96]. In another randomized trial [97] including 89
patients with frequent recurrent Ta or T1 tumors
without bladder Tis, at 20 years of follow-up, 80%
of patients who received mitomycin C developed
recurrence compared to 59.1% of those who
received BCG (p=0.005). These data also show the
efficacy of BCG as rescue therapy in patients with
recurrent tumors and BCG appears to be superior to
intravesical chemotherapy in the rescue setting.

In several comprehensive studies comparing the


efficacy of intravesical chemotherapy [84,85] or
BCG [86] versus observation, there was a significant
reduction of recurrence rate in the intervention
arm compared to the observation arm and BCG
was significantly superior to chemotherapy [87].
However, according to risk groups, BCG was clearly
superior to intravesical chemotherapy in terms of
decreasing recurrence and preventing or delaying
progression, but only when BCG maintenance
is applied [88-90]. However, in intermediate risk
groups there is a great controversy regarding the
efficacy between intravesical chemotherapy or BCG
[86,91-93], but the toxicity is significantly superior
for BCG. In summary, intravesical chemotherapy
or BCG represent alternative approaches for
these patients [4]. In low risk patients, immediate
instillation of a chemotherapy agent is considered
the standard of care as a meta-analysis clearly
demonstrates a significant reduction in recurrence
rate compared to observation [91]. However, in
cases of multiple tumors, this approach is insufficient
and a complete course of chemotherapy is required
[80,94]. Although there is no evidence that BCG is
superior to intravesical chemotherapy in this setting,
intravesical BCG constitutes a clear overtreatment
with unacceptable toxicity [4] following the results in
the intermediate risk group as mentioned above.

In a randomized trial comparing intravesical BCG


and mitomycin C with planned cross-over following
failure of initial therapy, of 39 patients who failed
mitomycin C and received rescue BCG, 19 (32%)
remained disease-free. On the other hand, 21
received mitomycin C after BCG failure and 4 (19%)
also remained free of disease [96]. In a metaanalysis including 9 randomized trials on 2261
patients comparing intravesical chemotherapy and
BCG, Huncharek [92] observed that in patients with
prior intravesical chemotherapy (1480 patients),
BCG significantly reduced recurrence rate compared
to intravesical chemotherapy with a reduction in
tumor recurrence rates of 46%, 49%, and 57% at 1,
2, and 3 years respectively [92]. In another recent
meta-analysis comparing intravesical mitomycin
C and BCG, analyzing patients who received prior
chemotherapy, BCG was superior to mitomycin C
(p=0.0264), reducing recurrence rate. These data
support the superiority of BCG as rescue therapy
in cases of recurrent tumor receiving intravesical
chemotherapy [89]. Whether BCG should be used
as rescue therapy after intravesical chemotherapy
failure or after a second line of intravesical
chemotherapy is unknown. Despite the fact that
many of these studies included patients of different
stages and grades and were not restricted to LG
Ta, it appears that BCG is effective in salvaging
a response in some low risk patients who do not
respond to intravesical chemotherapy.

Despite the significant decrease in recurrence


rate with an immediate post-TUR instillation of
chemotherapy and/or following a complete course
of chemotherapy in patients with multiple tumors,
around 30% of patients will have recurrence [4]. In
these cases, a different chemotherapeutic agent can
be effective in controlling recurrence. In a recent
randomized trial of 96 recurrent Ta or T1 tumors,
intravesical mitomycin C showed a 50% diseasefree survival with a mean follow-up of 65.7 months,
outlining the equivalent efficacy of intravesical

242

Summary
Based on a review of the available data, BCG is not
an appropriate initial therapy in low risk patients with
non-muscle-invasive bladder cancer but appears to
be effective in those patients who develop recurrence
after initial intravesical chemotherapy.

Recommendation
Recommendation

Level of
evidence

Grade of
recommendation

Intravesical BCG is not appropriate as initial intravesical


therapy for patients with LG Ta urothelial carcinoma.

Intravesical BCG could potentially be used in patients with


recurrent LG Ta urothelial carcinoma who have not responded
to courses of intravesical chemotherapy.

comparable efficacy to BCG for the treatment of


urothelial carcinoma. Of the several agents showing
promise, few have been evaluated in clinical trials.
Further, none of these agents have been specifically
tested in LG Ta tumors, and, while such tumors may
have been part of the cohorts tested, the value of
these agents in treating patients with recurrent LG
Ta tumors is unclear at this time.

VI. Other Immunomodulating


Therapies
As BCG is generally considered to be the most
effective intravesical agent for non-muscle-invasive
bladder cancer (91), there has been interest in
developing and testing other immunomodulating
agents against this disease entity. The aim of
researchers has been to search for an alternative
immunological agent with one of the following
properties:

Interferon-alpha and Keyhole limpet haemocyanin


(KLH) have been compared to BCG in randomized
clinical trials to test their utility in preventing recurrence
and progression in non-muscle-invasive bladder
cancer, but both were found to be significantly less
effective [98, 99].

- is more efficient than BCG in lowering recurrences


or progression rates.

Research
on
alternative
immunomodulating
agents has primarily focused on patients with high
risk of progression, that is those in whom BCG
currently is recommended as a first-line treatment,
which is covered in another chapter.
Future
immunotherapeutic efforts will undoubtedly benefit
from the gradually increasing understanding of the
immunobiology of bladder cancer. This will hopefully
lead to new immunologic treatment alternatives for
this group of patients.

- is as effective as BCG, but has lower rates of


adverse events.
- has acceptable efficacy and adverse events,
but can be used as a secondary agent in patients
intolerant or unresponsive to BCG.
Regarding patients with low grade Ta tumors, an
agent with less adverse events but with efficacy
similar to BCG would be of clinical value. Today,
BCG is not recommended as a first-line therapy
in this group of patients [4] given the higher risk of
adverse events in comparison to recommended firstline chemotherapeutic agents, such as mitomycin C
and epirubicin.

Summary
Secondary immunomodulating therapies do not
appear to have a role in management of LG Ta
urothelial carcinoma. Most of these agents have
been tested in high risk non-muscle-invasive bladder
cancer and hence their role and efficacy against LG
Ta tumors is unclear.

So far, despite considerable efforts, no alternative


immunologic treatment has been shown to have

Recommendation
Recommendation

Secondary immunomodulating therapies do not have a role in


management of LG Ta urothelial carcinoma at this time.

243

Level of
evidence
3

Grade of
recommendation
C

REFERENCES

20. Schmitz-Drager BJ, Beiche B, Tirsar LA, Schmitz-Drager


C, Bismarck E, Ebert T. Immunocytology in the assessment
of patients with asymptomatic microhaematuria. Eur Urol.
2007; 51:1582-1588.

1. Oxford Centre for Evidence-based Medicine Levels of


Evidence (May 2001).Produced by Bob Phillips, Chris
Ball, Dave Sackett, Doug Badenoch, Sharon Straus, Brian
Haynes, Martin Dawes since November 1998.

21. Lodde M, Mian C, Comploj E, Palermo S, Longhi E,


Marberger M, Pycha A. uCyt+ test: alternative to cystoscopy
for less-invasive follow-up of patients with low risk of
urothelial carcinoma. Urology. 2006 May;67(5):950-4
22. Schlomer BJ, Ho R, Sagalowsky A, Ashfaq R, Lotan Y.
Prospective Validation of the Clinical Usefulness of Reflex
Fluorescence In Situ Hybridization Assay in Patients With
Atypical Cytology for the Detection of Urothelial Carcinoma
of the Bladder. J Urol. 2010; 183:62-67.

2. Lotan Y, Shariat SF, Schmitz-Drager BJ, Sanchez-Carbayo


M, Jankevicius F, Racioppi M, et al. Considerations on
implementing diagnostic markers into clinical decision
making in bladder cancer. Urol Oncol. 2010;28:441-8.
3. Goebell PJ, Groshen SL, Schmitz-Drager BJ. Guidelines
for development of diagnostic markers in bladder cancer.
World J Urol. 2008;26:5-11.

23. Bergman J, Reznichek RC, Rajfer J. Surveillance of


patients with bladder carcinoma using fluorescent in-situ
hybridization on bladder washings. BJU Int. 2008;101:2629.

4. Babjuk M, Oosterlinck W, Sylvester R, Kaasinen E, Bohle A,


Palou-Redorta J. EAU guidelines on non-muscle-invasive
urothelial carcinoma of the bladder. Eur Urol. 2008; 54:303314.

24. van der Aa MNM, Zwarthoff EC, Steyerberg EW, Boogaard


MW, Nijsen Y, van der Keur KA, et al. Microsatellite Analysis
of Voided-Urine Samples for Surveillance of Low-Grade
Non-Muscle-Invasive Urothelial Carcinoma: Feasibility and
Clinical Utility in a Prospective Multicenter Study (CostEffectiveness of Follow-Up of Urinary Bladder Cancer Trial
[CEFUB]). Eur Urol. 2009; 55:659-668.

5. Raitanen MP, Aine R, Rintala E, Kallio J, Rajala P, Juusela


H, et al. Differences between local and review urinary
cytology in diagnosis of bladder cancer. An interobserver
multicenter analysis. Eur Urol. 2002; 41:284-289.
6. Glatz K, Willi N, Glatz D, Barascud A, Grilli B, Herzog M,
et al. An international telecytologic quiz on urinary cytology
reveals educational deficits and absence of a commonly
used classification system. Am J Clin Pathol. 2006;
126:294-301.

25. de Bekker-Grob EW, van der Aa MNM, Zwarthoff EC,


Eijkemans MJC, van Rhijn BW, van der Kwast TH, et al.
Non-muscle-invasive bladder cancer surveillance for which
cystoscopy is partly replaced by microsatellite analysis of
urine: a cost-effective alternative? BJU Int. 2009;104:4147.

7. Yutkin V, Nisman B, Pode D. Can urinary biomarkers replace


cystoscopic examination in bladder cancer surveillance?
Exp. Rev Anticanc. 2010; 10:787-790.

26. Roupret M, Hupertan V, Yates DR, Comperat E, Catto


JWF, Meuth M, et al. A comparison of the performance of
microsatellite and methylation urine analysis for predicting
the recurrence of urothelial cell carcinoma, and definition
of a set of markers by Bayesian network analysis. BJU Int.
2008;101:1448-1453.

8. Zwarthoff EC. Detection of tumours of the urinary tract in


voided urine. Scand J Urol Nephrol. 2008; 42:147-153.
9. Shirodkar SP, Lokeshwar VB. Bladder tumor markers: from
hematuria to molecular diagnostics - where do we stand?
Exp. Rev Anticanc. 2008; 8:1111-1123.

27. Kausch I, Sommerauer M, Montorsi F, Stenzl A, Jacqmin


D, Jichlinski P, Jocham D, Ziegler A, Vonthein R.
Photodynamic diagnosis in non-muscle-invasive bladder
cancer: a systematic review and cumulative analysis of
prospective studies. Eur Urol. 2010; 57:595-606.

10. Shariat SF, Karam JA, Lerner SP. Molecular markers in


bladder cancer. Curr Opin Urol. 2008; 18:1-8.
11. Kim WJ, Bae SC. Molecular biomarkers in urothelial bladder
cancer. Cancer Sci. 2008; 99:646-652.

28. Bajaj A, Sokhl H, and Rajesh A: Intravenous urography for


diagnosing synchronous upper-tract tumours in patients
with newly diagnosed bladder carcinoma can be restricted
to patients with high-risk superficial disease. Clin Radiol
2007; 62: 854-857.

12. Catto JWF. Old and new urinary markers: Which one is the
PSA for bladder cancer? Eur Urol Suppl. 2008; 7:422-425.
13. Agarwal PK, Black PC, Kamat AM. Considerations on the
use of diagnostic markers in management of patients with
bladder cancer. World J Urol. 2008; 26:39-44.

29. Palou J, Rodriguez-Rubio F, Huguet J et al: Multivariate


analysis of clinical parameters of synchronous primary
superficial bladder cancer and upper urinary tract tumour. J
Urol 2005; 174: 859-861.

14. van Rhijn BWG, van der Poel HG, van der Kwast TH.
Cytology and Urinary Markers for the Diagnosis of Bladder
Cancer. Eur Urol Suppl. 2009; 8:536-541.

30. Goessl C, Knispel HH, Miller K et al: Is routine excretory


urography necessary at first diagnosis of bladder cancer? J
Urol 1997; 157: 481-481.

15. Babjuk M, Soukup V, Pesl M, Kostirova M, Drncova E,


Smolova H, et al. Urinary cytology and quantitative BTA
and UBC tests in surveillance of patients with pTa pT1
bladder urothelial carcinoma. Urology. 2008; 71:718-722.

31. Herranz-Amo F, Diez-Cordero JM, Verdu-Tartajo F et al:


Need for intravenous urography in patients with primary
transitional carcinoma of the bladder. Eur Urol 1999; 36:
221-224.

16. Raitanen MP. The role of BTA stat Test in follow-up of


patients with bladder cancer: results from FinnBladder
studies. World J Urol. 2008; 26:45-50.

32. Millan-Rodriguez F, Chechile-Toniolo G et al: Upper


tract tumours after primary superficial bladder tumours:
Prognostic factors and risk groups. J Urol 2000; 164: 11831187.

17. Lotan Y, Shariat SF, Grp NS. Impact of risk factors on the
performance of the nuclear matrix protein 22 point-of-care
test for bladder cancer detection. BJU Int. 2008; 101:13621367.

33. Sadow CA, Wheeler SC, Kim J et al: Positive predictive


value of CT urography in the evaluation of upper tract
urothelial cancer. Am J Roentgonol 2010; 195: 337-343.

18. Shariat SF, Marberger MJ, Lotan Y, Sanchez-Carbayo M,


Zippe C, Ludecke G, et al. Variability in the performance
of nuclear matrix protein 22 for the detection of bladder
cancer. J Urol. 2006; 176:919-926.

34. Gray Sears CL, ward JF, Sears ST et al: Prospective


comparison of computerized tomography and excretory
urography in the initial evaluation of asymptomatic
microhaematuria. J Urol 2002; 168: 2457-2460.

19. Nguyen CT, Jones JS. Defining the role of NMP22 in


bladder cancer surveillance. World J Urol. 2008; 26:51-58.

244

35. Lang EK, Macchia RJ, Thomas R et al: Improved detection


of renal pathologic features on multiphasic helical CT
compared with IVU in patients presenting with microscopic
hematuria. Urology 2003; 61: 528-532.

54. Greenlee, R. T., Murray, T., Bolden, S. and Wings, P. A.:


Cancer statistics, 2000. CA Cancer J Clin, 50: 7, 2000.
55. Holmang S, Hedelin H, Anderstrom C, Holmberg E, Busch
C and Johansson SL: Recurrence and progression in low
grade papillary urothelial tumors. J Urol 1999; 162: 702.
56. Sylvester RJ, van der Meijden AP, Oosterlinck W, Witjes
JA, Bouffioux C, Denis L et al: Predicting recurrence and
progression in individual patients with stage Ta T1 bladder
cancer using EORTC risk tables: a combined analysis of
2596 patients from seven EORTC trials. Eur Urol 2006; 49:
466.
57. Prout, G. R., Jr., Barton, B. A., Griffin, P. P. and Friedell, G.
H. for the National Bladder Cancer Group: Treated history
of noninvasive grade 1 transitional cell carcinoma. J Urol,
1992; 148: 413.
58. Millan-Rodriguez F, Chechile-Toniolo G, Salvador-Bayarri J,
Palou J, Algaba F, Vicente-Rodriguez J. Primary superficial
bladder cancer risk groups according to progression,
mortality and recurrence. J Urol 2000; 164:680684.
59. Konety BR, Joyce GF and Wise M: Bladder and upper tract
urothelial cancer. In: Urological Diseases in America. Edited
by MS Litwin and CS Saigal. US Department of Health and
Human Services, Public Health Service, National Institutes
of Health, National Institute of Diabetes and Digestive
and Kidney Diseases. Washington, DC: US Government
Publishing Office 2007; NIH Publication No. 07-5512, pp
265.
60. Collado A, Chechile GE, Salvador J, et al. Early
complications of endoscopic treatment for superficial
bladder tumors. J Urol. 2000; 164:1529-1532.
61. Herr HW. Does cystoscopy correlate with the histology of
recurrent papillary tumours of the bladder? BJU Int. 2001;
88:683-685.
62. Satoh E, Miyao N, Tachiki H, et al. Prediction of muscle
invasion of bladder cancer by cystoscopy. Eur Urol. 2002;
41:178-181.
63. Soloway MS. Expectant treatment of small, recurrent,
low-grade, noninvasive tumors of the urinary bladder. Urol
Oncol. 2006; 24:58-61.
64. Soloway MS, Bruck DS, Kim SS. Expectant management
of small, recurrent, noninvasive papillary bladder tumors. J
Urol. 2003; 170:438-441.
65. Gofrit ON, Pode D, Lazar A, Katz R, Shapiro A. Watchful
waiting policy in recurrent Ta G1 bladder tumors. Eur Urol.
2006; 49:303-306.
66. Pruthi RS, Baldwin N, Bhalani V, Wallen EM. Conservative
management of low risk superficial bladder tumors. J Urol.
2008; 179:87-90.
67. Hernndez V, Alvarez M, de la Pea E, Amaruch N, Martn
MD, de la Morena JM, Gmez V, Llorente C. Safety of
active surveillance program for recurrent nonmuscleinvasive bladder carcinoma. Urology. 2009; 73:1306-1310.
68. Donat SM, North A, Dalbagni G and Herr HW: Efficacy of
office fulguration for recurrent low grade papillary bladder
tumors less than 0.5 cm. J Urol 2004; 171: 636.
69. German, K., Hasan, S. T. and Derry, C.: Cystodiathermy
under local anaesthesia using the flexible cystoscope. Br J
Urol, 1992; 69: 518.
70. Wedderburn AW, Ratan P, Birch BR. A prospective trial
of flexible cystodiathermy for recurrent transitional cell
carcinoma of the bladder. J Urol. 1999; 161:812-814.
71. Bouffioux C, Kurth KH, Bono A, Oosterlinck W, Kruger CB,
DePauw M et al: Intravesical adjuvant chemotherapy for
superficial transitional cell bladder carcinoma: results of
2 European Organization for Research and Treatment of
Cancer randomized trials with mitomycin C and doxorubicin
comparing early versus delayed instillations and short-term
versus long-term treatment. European Organization for
Research and Treatment of Cancer Genitourinary Group.
J Urol 1995; 153: 934.

36. Canales BK, Anderson JK, Premoli J et al: Risk factors for
upper tract recurrence in patients undergoing long-term
surveillance for stage Ta bladder cancer. J Urol 2006; 175:
74-77.
37. Oldbring J, Glifberg I, Mikulowski P et al: Carcinoma
of the renal pelvis and ureter following bladder
carcinoma:frequency, risk factors and clinicopathological
findings. J Urol 1989; 141: 1311-1313.
38. Palou J Farina LA, Villavicencio H et al: Upper tract
urothelial tumour after transurethral resection for bladder
tumour. Eur Urol 1992; 21: 110-114.
39. Solsona E, Ibarra I, Ricos JV et al: Upper urinary tract
involvement in patients with bladder carcinoma in situ (Tis):
its impact on management. Urology 1997; 49: 347-352.
40. Walzer Y and Soloway MS: Should followup of patients with
bladder cancer include routine excretory urography? J Urol
1983; 130: 672-673.
41. Enver MK, Miller PD and Chinegwundoh FI. Upper tract
surveillance in primary bladder cancer follow-up. BJU Int.
2004; 94: 790-792.
42. Wright JL, Hotaling J and Porter MP. Predictors of upper
tract urothelial cell carcinoma after primary bladder cancer:
A population based analysis. J Urol 2009; 181: 1035-1039.
43. Holmang S, Hedelin H, Anderstrom C et al: Long-term
follow-up of a bladder carcinoma cohort: routine follow-up
is not necessary. J Urol 1998; 160: 45-48.
44. Schwartz CB, Bekirov H and Melman A: Urothelial tumors of
upper tract folloing treatment of primary bladder transitional
cell carcinoma. Urology 1992; 40: 509-511.
45. Hurle R, Losa A, Manletti A et al: Upper urinary tract tumor
developing after treatment of superficial bladder cancer.
Urology 1999; 53: 1144-1148.
46. Yousem DM, Gatewood OM, Goldman SM et al:
Synchronous and metachronous transitional cell carcinoma
of the urinary tract: prevalence, incidence and radiographic
detection. Radiology 1988; 167: 613-618.
47. Millan-Rodriguez F, Chechile-Toniolo G, Salvador-Bayarri
J et al: Upper urinary tract tumors after primary superficial
bladder tumors: Prognostic factors and risk groups. J Urol
2008; 164: 1183-1187.
48. Amar A and Das S: Upper urinary tract transitional
cell carcinoma in patients with bladder carcinoma and
associated vesicouretric reflux. J Urol 1985; 133: 468-471.
49. Miller EB, Eure GR and Schellhammer PF: Uppper tract
transitional cell carcinoma following treatment of superficial
bladder cancer with BCG. Urology 42: 26-30, 1993.
50. Schwalb DA, Herr HW, Sogani PC et al: Upper tract
disease following intravesical BCG for superficial bladder
cancer: five years follow-up. J Urol 1992; 147(suppl): 273A.
Abstract 237.
51. Montie JE, Abrahams NA, Bahnson RR, Eisenberger
MA, El-Galley R, Herr HW et al: Bladder cancer. Clinical
guidelines in oncology. J Natl Compr Canc Netw 2006; 4:
984.
52. EAU guidelines on non-muscle-invasive urothelial
carcinoma of the bladder. Babjuk M, Oosterlinck W,
Sylvester R, Kaasinen E, Bhle A, Palou-Redorta J;
European Association of Urology (EAU). Eur Urol. 2008;
54:303-314.
53. Guideline for the management of nonmuscle invasive
bladder cancer (stages Ta, T1, and Tis): 2007 update. Hall
MC, Chang SS, Dalbagni G, Pruthi RS, Seigne JD, Skinner
EC, Wolf JS Jr, Schellhammer PF. J Urol. 2007; 178:231430.

245

72. Tolley DA, Parmar MK, Grigor KM, Lallemand G, Benyon


LL, Fellows J et al: The effect of intravesical mitomycin C on
recurrence of newly diagnosed superficial bladder cancer:
a further report with 7 years of follow up. J Urol 1996; 155:
1233.

85. Nilsson S. A systematic overview of chemotherapy effects


in urothelial bladder cancer. Acta Oncol 2001, 40: 371.
86. Shelley MD, Court J, Burgon K, et al: Meta-analysis
of intravesical therapy for superficial bladder cancer:
superiority of bacillus Calmette-Guerin may be confined to
high risk patients. Br J Cancer 2001; 85: 53.

73. Sylvester RJ, Oosterlinck W and van der Meijden AP: A


single immediate post operative instillation of chemotherapy
decreases the risk of recurrence in patients with stage Ta
T1bladder cancer: a meta-analysis of published results of
randomized clinical trials. J Urol 2004; 171: 2186.

87. Shelley M, Wilt T, Court J, Coles B, Kynaston H, Mason


M: Intravesical bacillus Calmette-Gurin in superior to
Mitomycin C in reducing tumour recurrence in high-risk
superficial bladder cancer: a meta-analysis of randomized
trials. BJU Int. 2004; 93: 485-490.

74. Hall MC, Chang SS, Dalbagni G, Pruthi RS, Seigne JD,
Skinner EC, Wolf JS Jr, Schellhammer PF. Guideline for
the management of non-muscle invasive bladder cancer
(stages Ta, T1, and Tis): 2007 update. J Urol. 2007;
178:2314-2330.

88. Bhle A, Jocham D and Bock PR: Intravesical bacillus


Calmette-Guerin versus Mitomycin C for superficial bladder
cancer: a formal meta-analysis of comparative studies on
recurrence and toxicity. J. Urol. 2003; 169: 90-95.

75. Au JL, Badalament RA, Wientjes MG, Young DC, Warner


JA, Venema PL, Pollifrone DL, Harbrecht JD, Chin JL,
Lerner SP, Miles BJ; International Mitomycin C Consortium.
Methods to improve efficacy of intravesical mitomycin C:
results of a randomized phase III trial. J Natl Cancer Inst.
2001; 93:597-604.

89. Malmstrm P-U, Sylvester R., Crawford D.E. at al. An


individual patient data meta-analysis of the long-term
outcome of randomised studies comparing intravesical
Mitomycin C versus bacillus Calmette-Gurin for nonmuscle-invasive bladder cancer. Eur. Urol. 2009; 56: 247256.

76. Bhle A, Leyh H, Frei C, Khn M, Tschada R, Pottek T,


Wagner W, Knispel HH, von Pokrzywnitzki W, Zorlu
F, Helsberg K, Lbben B, Soldatenkova V, Stoffregen
C, Bttner H; S274 Study Group. Single postoperative
instillation of gemcitabine in patients with non-muscleinvasive transitional cell carcinoma of the bladder: a
randomised, double-blind, placebo-controlled phase III
multicentre study. Eur. Urol. 2009 Sep;56(3):495-503.

90. Sylvester RJ, van der Meijden and DL Lamm; Intravesical


bacillus Calmette-Guerin reduces the risk of progression
in patients with superficial bladder cancer: a combined
analysis of the published results of randomized clinical
trials. J Urol. 2002; 168: 1964-1970.
91. Gontero P The role of bacillus Calmette-Guerin in the
treatment of non-muscle invasive bladder cancer. Eur Urol
2010;57:410.

77. Oosterlinck W, Kurth KH, Schroder F, Hammond B and


Sylvester R: A prospective European Organization for
Research and Treatment of Cancer Genitourinary Group
randomized trial comparing transurethral resection followed
by a single intravesical instillation of epirubicin or water in
single stage Ta, T1 papillary carcinoma of the bladder. J
Urol 1993; 149:749.

92. Huncharek M, Kupelnick B: Impact of intravesical


chemotherapy versus BCG immunotherapy on recurrence
of superficial transitional cell carcinoma of the bladder. Am
J Clin. Oncol. 2003; 26: 402-407.
93. Witjes JA and Hendricksen K: Intravesical pharmacotherapy
for non-muscle-invasive bladder cancer: A critical analysis
of currently available drugs, treatment schedules, and longterm results. Eur. Urol. 2008; 53: 45-52.

78. Tolley DA, Parmar MK, Grigor KM, Lallemand G, Benyon


LL, Fellows J et al: The effect of intravesical mitomycin
C on recurrence of newly diagnosed superficial bladder
cancer: a further report with 7 years of follow up. J Urol.
1996; 155: 1233.

94. Solsona E. Early single instillation is very beneficial and


should be the standard approach in non-muscle invasive
bladder cancer. Eur. Urol. 2009; suppl 8: 464.

79. Solsona E, Iborra I, Rics JV, Monrs JL, Casanova J,


Dumont R. Effectiveness of a single immediate mitomycin
C instillation in patients with low risk superficial bladder
cancer: short and long-term followup. J Urol. 1999;
161:1120-1123.

95. Reference needed Solsona


96. Mangiarotti B, Trinchieri A, Del Nero A and Montanari E:
A randomized prospective study of intravesical prophylaxis
in non-muscle invasive bladder cancer at intermediate
risk of recurrence: Mitomycin chemotherapy vs BCG
immunotherapy. Arch Ital Urol Androl 2008; 80: 167-171.

80. Sylvester RJ, Oosterlinck W and van der Meijden AP: A


single immediate postoperativeinstillation of chemotherapy
decreases the risk of recurrence in patients with stage Ta
T1bladder cancer: a meta-analysis of published results of
randomized clinical trials. J Urol 2004; 171: 2186.

97. Malmstrm P-U, Wijkstrm H, Lundholm C, Wester


K, Busch C, Norln BJ and Members of the SwedishNorwegian Bladder Cancer Study Group: 5-year followup
of a randomized prospective study comparing Mitomycin
C and bacillus Calmette-Gurin in patients with superficial
bladder carcinoma. J. Urol. 1999; 161: 1124-1127.

81. Gudjonsson S, Adell L, Merdasa F, et al. Should all patients


with non-muscle invasive bladder cancer receive early
intravesical chemotherapy after transurethral resection?
The results of a prospective randomised multicentre study.
Eur. Urol. 2009; 55:773780.

98. Jrvinen R, Kaasinen E, Sankila A and Rintala E: Longterm efficacy of maintenance bacillus Calmette-Gurin
versus maintenance Mitomycin C instillation therapy in
frequently recurrent TaT1 tumours without carcinoma in
situ: a subgroup analysis of the prospective, randomised
Finn bladder I study with a 20-year follow-up. Eur. Urol.
2009; 56: 260-265.

82. Nieuwenhuijzen JA, Bex A, Horenblas S. Unusual


complication after immediate postoperative intravesical
mitomycin C instillation. Eur Urol. 2003;43:711-712.
83. Faiola V, Montella L, Addeo R, et al. Intravesical gemicitabine
versus mitomycin for recurrent superficial bladder tumour
(stages pTA and pT1): a randomized prospective study.
Eur. J Cancer 2008; Suppl. 6:123126.

99. Jimenez-Cruz, J.F., et al., Intravesical immunoprophylaxis in


recurrent superficial bladder cancer (Stage T1): multicenter
trial comparing bacille Calmette-Guerin and interferonalpha. Urology 1997; 50:529-535.

84. Hunchareck M, Geschwind JF, Witherspoon B, McGarry


R, Adcock D: Intravesical chemotherapy prophylaxis in
primary superficial bladder cancer: a meta-analysis of 3703
patients from 11 randomized trials. J Clin Epidemiol 2000;
53: 676-680.

100. Kalble, T., et al., Intravesical prevention of recurrence of


superficial urinary bladder cancer with BCG and KLH. A
prospective randomized study. Urologe A 1991; 30:118121.

246

Committee 5

High grade Ta, CIS, and T1


Urothelial Carcinoma
Chairs:
Maximilian Burger,
Fred Witjes

Members:
Marko Babjuk,
Maurizio Brausi,
Chris Cheng,
Eva Comperat,
Colin Dinney,
Wolfgang Jager,
Wolfgang Otto,
Jay Shah,
Joachim Throff

247

Table of Contents
I. TaHigh Grade Ta Urothelial
Carcinoma

III. T1 Urothelial Carcinoma


1. Diagnosis and Risk Assessment

1. Introduction

2. Treatment

2. Diagnosis and Risk Assessment

Recommendations

3. Treatment

IV. Treatment Options for BCG


Failure

Recommendations

III. Carcinoma in Situ of the


Urinary Bladder

1. Introduction
2. Defining BCG Failure
3. Treatment Options

1. Introduction
2. Diagnosis

Recommendations

3. Pathology and Risk Assessment


4. Treatment
Recommendations

248

High grade Ta, CIS, and T1 Urothelial


Carcinoma
Maximilian Burger, Fred Witjes
Marko Babjuk, Maurizio Brausi, Chris Cheng,
Eva Comperat, Colin Dinney, Wolfgang Jager,
Wolfgang Otto, Jay Shah, Joachim Throff

3. Treatment

I. High Grade Ta Urothelial


Carcinoma

a) Turbt
Studies on the quality of transurethral resection of
bladder tumors (TURBT) showed that the residual
tumor rate of Ta urothelial carcinoma was 70%
4 weeks after first resection in a high-volume
department [10]. Fluorescence-guided resection
has been proven to improve the detection rates of
bladder tumors not only in CIS but even in papillary
carcinomas of the bladder, resulting in lower residual
tumor levels and a lower subsequent recurrence
rate [11,12,13,14]. Improving TURBT has become
a major focus of attention for two reasons. Firstly,
subsequent bacillus Calmette-Gurin (BCG) and
mitomycin C (MMC) instillation therapy is more
effective in patients after a complete TURBT
[15,16]. Secondly, an adequate TURBT, i.e., one
which provides sufficient tissue for histopathological
analysis, is crucial for proper staging. The presence
of detrusor muscle in the TURBT specimen is
regarded as a surrogate parameter of resection
quality [17]. A restaging TURBT provides more tissue
for pathological examination and thus allows better
staging. Accordingly, repeat TURBT is advised in
high risk non-muscle-invasive cases.

1. Introduction
Stage pTa urothelial carcinoma represents the
majority (60-75%) of bladder cancer cases [1]. In
general, these tumors show high recurrence but low
progression rates [2].

2. Diagnosis
Pathologists diagnose pTa urothelial carcinoma when
they find a mucosal layer of more than 8 cell lines in a
resected specimen with features of anaplasia [3]. In
contrast to stage T1 urothelial cancer, the basement
membrane is always intact. Only about 20% of pTa
cancer is high grade (grade 3) [3].
In contrast to low grade disease, the diagnosis of
high grade Ta urothelial carcinoma is facilitated
by urinary cytology with sensitivity and specificity
rates of over 90% [4,5]. Urinary tumor markers are
of limited importance for Ta tumors. Markers of the
hyaluronic acid family seem to be the most promising
urinary markers for this entity, exhibiting prognostic
significance for recurrence [6]. Nuclear matrix
protein 22 (NMP22) and other commercialized tests
have not been shown to be of any advantage when
compared to cystoscopy [7].

Equally, an adequate TURBT is crucially important in


determining the optimal treatment modality for any
given patient. The presence of detrusor muscle in
the TURBT specimen is of particular importance for
pT1 tumors. Several studies dealing with the quality
of TURBT or a second TURBT also indicated that
a better resection also lowers the recurrence and
progression rates and improves the reaction to BCG
as well as the prognosis of the patient [16-19].

Without taking into account associated factors,


patients with multifocal stage Ta high grade
urothelial carcinoma have the third highest risk
of both recurrence (around 40% after 1 year) and
progression (5% after 1 year) [2]. While external
evaluation of recurrence rates showed a good
correlation with the EORTC risk tables, prediction of
progression seems less concordant [8,9].

b) Instillation Therapy
While one immediate instillation is sufficient in

249

prophylaxis of recurrence and progression for


low grade Ta urothelial carcinoma, the prognosis
of patients with high grade Ta disease can only
be improved with maintenance BCG therapy [1].
Decreased recurrence rates subsequent to BCG
maintenance have also been reported in intermediate
risk tumors, although the more severe toxicity of BCG
when compared with intravesical chemotherapy has
to be taken into consideration [20]. The impact of
maintenance BCG on progression of high grade pTa
urothelial carcinoma remains unclear [21,22].

9. Seo KW, Kim BH, Park CH, Kim CI, Chang HS. The efficacy
of the EORTC scoring system and risk tables for the
prediction of recurrence and progression of non-muscleinvasive bladder cancer after intravesical bacillus calmetteguerin instillation. Korean J Urol. 2010 Mar;51(3):165-70.
Epub 2010 Mar 19.
10. Adiyat KT, Katkoori D, Soloway CT, De los Santos R,
Manoharan M, Soloway MS. Complete transurethral
resection of bladder tumour: are the guidelines being
followed? Urology. 2010 Feb;75(2):365-7.
11. Kausch I, Sommerauer M, Montorsi F, Stenzl A, Jacqmin
D, Jichlinski P, Jocham D, Ziegler A, Vonthein R.
Photodynamic diagnosis in non-muscle-invasive bladder
cancer: a systematic review and cumulative analysis of
prospective studies. Eur Urol. 2010 Apr;57(4):595-606.

High Grade Ta Urothelial Carcinoma:


Recommendations

12. Denzinger S, Wieland WF, Otto W, Filbeck T, Knuechel R,


Burger M. Does photodynamic transurethral resection of
bladder tumour improve the outcome of initial T1 high-grade
bladder cancer? A long-term follow-up of a randomized
study. BJU Int. 2008 Mar;101(5):566-9.

1. Suspicious urinary cytology and papillary tumor


on cystoscopy are highly suggestive of high grade
Ta urothelial carcinoma. (LOE 3, Grade B)

13. Otto W, Burger M, Fritsche HM, Blana A, Roessler W,


Knuechel R, Wieland WF, Denzinger S. Photodynamic
diagnosis for superficial bladder cancer: do all risk-groups
profit equally from oncological and economic long-term
results? Clin Med Oncol. 2009 Apr 30;3:53-8.

2. A complete resection has to be pursued. If there


is any question regarding the completeness of the
initial TURBT, repeat TURBT is advised. (LOE 2,
Grade A)

14. Mowatt G, Zhu S, Kilonzo M, Boachie C, Fraser C, Griffiths


TR, NDow J, Nabi G, Cook J, Vale L. Systematic review
of the clinical effectiveness and cost-effectiveness of
photodynamic diagnosis and urine biomarkers (FISH,
ImmunoCyt, NMP22) and cytology for the detection and
follow-up of bladder cancer. Health Technol Assess. 2010
Jan;14(4):1-331, iii-iv.
15. Divrik RT, ahin Af, Yildirim , Altok M, Zorlu F. Impact of
Routine Second Transurethral Resection on the Long-Term
Outcome of Patients with Newly Diagnosed pT1 Urothelial
Carcinoma with Respect to Recurrence, Progression Rate,
and Disease-Specific Survival: A Prospective Randomised
Clinical Trial. Eur. Urol. 2010 Apr; 58(2): 185-190.
16. Guevara A, Salomon L, Allory Y, Ploussard G, de la Taille
A, Paul A, Yiou R, Hoznek A, Dahan M, Abbou CC, Vordos
D. The role of tumour-free status in repeat resection before
intravesical bacillus Calmette-Guerin for high grade Ta, T1
and CIS bladder cancer. J Urol. 2010 Jun;183(6):2161-4.
17. Mariappan P, Zachou A, . Grigor KM, for the Edinburgh UroOncology Group . Detrusor Muscle in the First, Apparently
Complete Transurethral Resection of Bladder Tumour
Specimen Is a Surrogate Marker of Resection Quality,
Predicts Risk of Early Recurrence, and Is Dependent on
Operator Experience. Eur. Urol. 2010 May; 57(5): 843849.
18. Herr HW. Restaging transurethral resection of high risk
superficial bladder cancer improves the initial response
to bacillus Calmette-Guerin therapy. J Urol. 2005
Dec;174(6):2134-7.
19. Divrik RT, Sahin AF, Yildirim U, Altok M, Zorlu F. Impact of
routine second transurethral resection on the long-term
outcome of patients with newly diagnosed pT1 urothelial
carcinoma with respect to recurrence, progression rate,
and disease-specific survival: a prospective randomised
clinical trial. Eur Urol. 2010 Aug;58(2):185-90.
20. Sylvester RJ, Brausi MA, Kirkels WJ, Hoeltl W, Calais
Da Silva F, Powell PH, Prescott S, Kirkali Z, van de Beek
C, Gorlia T, de Reijke TM; EORTC Genito-Urinary Tract
Cancer Group. Long-term efficacy results of EORTC
genito-urinary group randomized phase 3 study 30911
comparing intravesical instillations of epirubicin, bacillus
Calmette-Gurin, and bacillus Calmette-Gurin plus
isoniazid in patients with intermediate- and high-risk stage
Ta T1 urothelial carcinoma of the bladder. Eur Urol. 2010
May;57(5):766-73.

3. BCG instillation therapy should be initiated in


patients with high grade pTa urothelial carcinoma.
(LOE 1a, Grade A)

References
1. Babjuk M, Oosterlinck W, Sylvester R, Kaasinen E, Bhle
A, Palou J, Rouprt M. Guidelines on TaT1 (non-muscle
invasive) Bladder Cancer. EAU, 2010.
2. Sylvester RJ, van der Meijden AP, Oosterlinck W, Witjes
JA, Bouffioux C, Denis L, Newling DW, Kurth K. Predicting
recurrence and progression in individual patients with stage
Ta T1 bladder cancer using EORTC risk tables: a combined
analysis of 2596 patients from seven EORTC trials. Eur
Urol. 2006 Mar;49(3):466-5; discussion 475-7.
3. Amling CL. Diagnosis and management of superficial
bladder cancer. Curr Probl Cancer. 2001 Jul-Aug;25(4):21978.
4. Brown FM. Urine cytology. It is still the gold standard for
screening? Urol Clin North Am. 2000 Feb;27(1):25-37.
5. Vom Dorp F, Pal P, Tschirdewahn S, Rossi R, Brgermann
C, Schenck M, Becker M, Szarvas T, Hakenberg OW,
Rbben H. Correlation of Pathological and CytologicalCytometric Grading of Transitional Cell Carcinoma of the
Urinary Tract. Urol Int. 2010 Dec 16. [Epub ahead of print
6. Kramer MW, Escudero DO, Lokeshwar SD, Golshani
R, Ekwenna OO, Acosta K, Merseburger AS, Soloway
M, Lokeshwar VB. Association of hyaluronic acid family
members (HAS1, HAS2, and HYAL-1) with bladder cancer
diagnosis and prognosis. Cancer. 2010 Oct 19. [Epub
ahead of print]
7. Smrkolj T, Miheli M, Sedlar A, Sterle I, Osredkar J,
Sedmak B. Performance of nuclear matrix protein 22 urine
marker and voided urine cytology in the detection of urinary
bladder tumours. Clin Chem Lab Med. 2010 Dec 1. [Epub
ahead of print]
8. Ather MH, Zaidi M. Predicting recurrence and progression
in non-muscle-invasive bladder cancer using European
organization of research and treatment of cancer risk
tables. Urol J. 2009 Summer;6(3):189-93.

250

prognosis, the initial reports also described its


diffuse appearance [4,5]. Indeed, CIS is more
typically related to multiple as opposed to single
tumor locations. Urothelial cancer within the urinary
bladder is accompanied by tumors of the prostatic
urethra generally 7 times more often and by tumors
of the upper urinary tract nearly 5 times more often if
CIS is also present [15,16].

21. Sylvester RJ, van der MEIJDEN AP, Lamm DL. Intravesical
bacillus Calmette-Guerin reduces the risk of progression in
patients with superficial bladder cancer: a meta-analysis of
the published results of randomized clinical trials. J Urol.
2002 Nov;168(5):1964-70.
22. Malmstrm PU, Sylvester RJ, Crawford DE, Friedrich
M, Krege S, Rintala E, Solsona E, Di Stasi SM, Witjes
JA. An individual patient data meta-analysis of the
long-term outcome of randomised studies comparing
intravesical mitomycin C versus bacillus Calmette-Gurin
for non-muscle-invasive bladder cancer. Eur Urol. 2009
Aug;56(2):247-56.

2. Diagnosis
For most stages of bladder cancer and indeed in
most cases, macroscopic hematuria is symptomatic
of CIS. CIS causes irritative LUTS more often than
papillary lesions. Three characteristics determine the
diagnostic workup: (a) CIS lacks profound cellular
cohesion within the lamina mucosa; (b) its growth
pattern is flat and lacks protrusion into the bladder;
and (c) it may be found throughout the urinary tract.

II. Carcinoma in Situ of the


Urinary Bladder
1. Introduction
Carcinoma in situ (CIS) is malignant neoplasia
of the urothelium, which is flat and not protruding
into the bladder and which does not invade the
lamina propria. That being said, it is nonetheless
considered to be an aggressive tumor entity due
to its level of dedifferentiation. The current WHO
classification defines CIS as a non-papillary, i.e.,
flat, lesion in which the surface epithelium contains
cells that are cytologically malignant. CIS shows
nuclear anaplasia identical to high grade urothelial
carcinoma [1]. This morphological appearance is
reflected in the molecular signature of CIS. While low
grade non-invasive lesions are thought to derive from
chromosome 9 abnormalities and mutations in the
fibroblast growth factor receptor gene, CIS is marked
by deletions of 19p13 resulting in TP53 mutations
with consecutive prevention of cell cycle arrest and
generalized genetic instability [2]. DNA aneuploidy is
found in more than 90% of CIS lesions [3].

The lack of profound cellular cohesion prompts the


use of urinary cytology to aid in the diagnosis of CIS
[17]. CIS can be reliably detected and followed up
by urinary cytology, and sensitivity and specificity
have been reported to exceed 90% [18]. However,
cytology may be hampered by inflammation (e.g.,
following instillation therapy) and distinguishing
malignant from reactive non-neoplastic cells may
become challenging [19]. Accordingly, novel tests
have been described. One of the more promising is
a multicolor fluorescence in situ hybridization (FISH)
assay assessing chromosomal abnormalities in DNA
in urine samples. While this FISH test has a low
sensitivity for low grade tumors lacking chromosomal
abnormalities, it detects high grade carcinomas
and CIS with reliable sensitivity [20]. In contrast, its
specificity does not seem to exceed that of urinary
cytology significantly, and FISH has not been broadly
used in the diagnosis of CIS, although a combination
with cytology has been suggested as advantageous
in the initial detection and follow-up of patients with
CIS [21,22]. To date, no further molecular tests have
been established in the diagnosis of CIS.

CIS was first reported as a distinct phenomenon


by Melicow in 1952. [4] Despite the non-invasive
character of CIS, it was suspected to possess
aggressive tumor biology and tendencies towards
early progression were described [5]. This notion
has been supported by the incidence of CIS. CIS is
found in approximately 3% of all patients with bladder
cancer, concomitantly with invasive bladder cancer
in 50%, and with muscle-invasive disease in 60%
[6,7,8]. As a consequence, CIS has been suggested
as a precursor lesion of invasive disease [9,10] and
concomitant CIS as a significant risk factor for poor
outcome. The tendency towards progression to
muscle-invasive stages is in part determined by the
presence of CIS; the EORTC risk score by Sylvester
et al attaches significant weight to its presence [11].
While CIS is also potentially related to an adverse
outcome in muscle-invasive stages, this has not
been implied in current nomograms [12,13,14].

The second trait, i.e., its flat growth pattern, may defy
cystoscopic detection. Classic white light cystoscopy
has been shown to miss up to 50% of cases.
Photodynamic diagnosis employs a fluorescent
substance, (hexyl-)aminolevulinic acid, which
accumulates fairly selectively in neoplastic cells.
Blue violet light causes the photosensitizing agent
to emit a reddish fluorescence [23]. Randomized
trials and a recent meta-analysis show a significantly
improved diagnostic accuracy of 39% for CIS
[24,25,26] and its use has been suggested to be
especially valuable in the detection of flat lesions
when malignant cells are present in urinary cytology
and there is an absence of suspicious lesions in white
light cystoscopy [23,27,17]. Definitive diagnosis is
based on the histopathological evaluation of tissue
samples by resection or cold-cup biopsy, although

In addition to the notion that CIS conveys a poor

251

the restrictions of inter- and intraobserver reliability


also have to be taken into account [28,29].

latter location shows a significantly higher rate of


lymph node metastasis and a lower 5-year survival
rate[42]. No difference was observed in recurrence
or disease-free survival for either uni- or multifocal
CIS or CIS location (urethral vs. vesical) [43].

The third trait, i.e., the presence of CIS throughout the


urinary tract, warrants a workup of the entire urinary
tract. Although little specific data exist, sampling the
prostatic urethra and each upper urinary tract for
urinary cytology needs to be considered should no
lesions be found on cystoscopy or IVU, CT, or renal
ultrasound[17].

3. Chronology of CIS

Because of the high likelihood that CIS will progress,


subsequent follow-up to specific therapy is generally
recommended, as with patients with non-muscleinvasive bladder cancer at high risk of progression:
cystoscopy and urinary cytology every 3 months
for a period of 2 years, every 4 months in the third
year, every 6 months in the fourth and fifth years,
and annually thereafter. In addition, ultrasound,
conventional IVU, or CT imaging of the upper urinary
tract should be performed periodically [17,30].

In the group of patients with primary CIS, progression


and death from disease seems to be rather unusual,
although more common in the group with concomitant
CIS [44]. Secondary CIS associated with pTa and/
or high grade infiltrating tumors historically has had
the worst prognosis. Recent data presented results
to the contrary, with a higher rate of progression in
primary CIS rather than in secondary CIS after BCG
therapies [45]. In this series, the authors observed a
better response to treatment in the group of primary
vs. secondary CIS and the 5-year cumulative
incidence of progression was also higher in the
primary vs. secondary group of CIS [45].

3. Pathology and risk assessment

4. Intravesical Therapy and CIS

As there is a frequent association with invasive


disease, CIS is clinically classified into different
types: (a) primary CISisolated with no previous
or concurrent papillary tumors; (b) secondary CIS
detected during the follow-up of patients with a
previous papillary tumor; and (c) concomitant CIS
in the presence of papillary tumors [31,32]. Genetic
and molecular approaches have underlined the
relationship between CIS and invasive carcinoma.
CIS has a variable histological appearance with
different morphologic subtypes, described by
McKenney in 2001 [33]. All these elements are part
of the prognosis and will be discussed below.

Numerous publications have shown that response to


intravesical treatment impacts prognosis. However,
few data are available concerning patients who have
CIS without any association with a pTa or pT1 tumor.
Several small studies could not confirm a correlation
between response to treatment, type of CIS, and
prognosis [34,46,47,48]. One reason for this might
be the frequent association of CIS with high grade
tumors, which could be responsible for the difficulties
in discerning an independent prognostic effect for
CIS [34]. A lack of response to BCG should therefore
be regarded as a poor prognosticator.

5. Molecular and Immunohistologic Parameters

a) Clinical prognostic factors

Molecular signatures can be used as diagnostic or


prognostic tools [49,50,51], such as p53 mutations,
amongst others [52].

1. Gender
Men are more likely to present with CIS (p=0.001),
although gender does not seem to affect prognosis
[34]. On the other hand, age does play a role, with
younger patients having a better outcome [35,36].

Aurora-A is a serine/threonine kinase that


activates the centrosome, which plays a role in cell
division. Aurora-A overexpression results in the
misaggregation of chromosomes and chromosomal
instability that eventually explains aneuploidy [53,54].
Aurora-A overexpression in CIS has been correlated
to the risk of relapse and progression [55].

2. Location
CIS of the bladder is also a known risk factor for
developing upper urinary tract tumors after a radical
cystectomy to treat urothelial carcinoma of the bladder
[37]. Volkmer et al reported in a series of 1420 patients
that the presence of CIS after a radical cystectomy
was a risk factor for recurrence in the upper urinary
tract (RiskRatio 2,3) [38]. Similarly, Youssef et al.
demonstrated that bladder CIS is an independent
predictor of disease recurrence and cancer-specific
mortality after a radical nephroureterectomy for upper
tract urothelial carcinoma (p=0.006 and p=0.045
respectively) [39]. Prostatic urethra involvement
of CIS has been observed in less than one third
of cystoprostatectomy specimens [40,41] and in
the prostatic urethra or prostatic ducts in 15%. The

In accordance with the histological appearance


of CIS and the discohesive pattern of cell growth,
adhesion molecule expressions were evaluated.
E-cadherin negative tumor cells were able to infiltrate
the normal urothelium as individual cells, whereas
tumor cells with a homogeneous expression of
E-cadherin exhibited sharp differences with normal
tissue [56,57]. Shariat showed that E-cadherin
expression was independently associated with
disease progression and cancer-specific survival
[57].
Cyclin D3 has a pivotal role in progression from G1

252

to S phase in the cell cycle. Cyclin D3 expression is


amplified in secondary concomitant CIS as opposed
to primary isolated CIS and is related to recurrence
and progression-free survival (p=0.002) [46]. In
clinical use, however, no marker is used routinely,
and the value of other single markers, genomic
profiling, and epigenetic markers for progression in
CIS remains to be determined.

urologist should be aware of their danger, rebiopsy,


especially if the biopsy lacks detrusor muscle, and
closely follow the patient.
CIS is an aggressive high grade disease. Pathologists
should be able to recognize the different features of
CIS. Bladder biopsies and cytology are a precondition
for accurate diagnosis and targeted treatment.
An association with urothelial carcinoma of the
upper urinary tract, prostatic urethra involvement,
secondary CIS associated with other urothelial
tumors, multifocality, age, and non-response to BCG
are all negative clinical prognostic factors. Loss of
E-cadherin, MUC-1, and overexpression of Aurora-A
and CyclinD3 are considered bad molecular markers
for patient outcome.

6. Pathology
In 2001, Mc Kenney et al described four different
morphological subtypes of CIS (large cell
pleomorphic, large cell nonpleomorphic, small cell,
and clinging/denuded). The most frequent pattern is
the large cell pleomorphic type, with cells showing
considerable nuclear variations and architecture
characterized by a loss of polarity. The most difficult
form to recognize is the clinging/denuded form, when
the urothelium may be extensively denuded (Figure
2). In the latter case, cytology is very important,
as detached malignant clusters or isolated cells
in voided urine permit differential diagnosis from
artifacts of biopsy or cystitis.

4. Treatment
a) Cystectomy
In patients with CIS and concomitant muscle-invasive
bladder tumors, cystectomy is the treatment of choice.
A partial cystectomy or radiotherapy is insufficient.
In patients with CIS with and without concomitant
non-muscle-invasive (Ta or T1) papillary tumors,
there is no consensus regarding the optimal course
of treatment: cystectomy or a more conservative
treatment strategy. There are no randomized trials
that have addressed this question.

Mc Kenney et al suggested that denudation reflects


rapid cell turnover and therefore potentially more
aggressive tumors. Nevertheless, they underlined
that individual patterns do not carry any known clinical
significance and that subclassification might confuse
urologists [33]. Molecular arguments and histological
observations strongly support the hypothesis that
a natural history in the development of CIS might
exist. Molecular alterations precede phenotypic
changes and end in morphologic changes seen on
routine staining. Early alterations during mitosis,
chromosomal instability and aneuploidy are reflected
by nuclear atypia and chromatin densification,
leading to the development of malignant urothelial
cells. In a further step, polarity and cell-to-cell
adherence are altered, resulting in denuded forms
of CIS with detached cell clusters in cytology and
few cells remaining on the surface. Eventually, via
mechanisms not yet understood, CIS develops into
invasive urothelial carcinoma.

Several recent papers have discussed the cystectomy


results in patients thought to have only CIS prior to
cystectomy [60,61,62]. A common theme in these
studies was the clinical understaging in a third or
more of the patients. Although the disease-specific
survival rates in these studies ranged from 85-90%,
early cystectomy represented overtreatment in
approximately 50% of patients.
When a conservative treatment strategy is chosen,
the transurethral resection (TUR) of all concomitant
papillary tumors is mandatory for correct staging and
grade determination. A TUR alone is insufficient,
as CIS can be missed at cystoscopy and, without
further treatment, 50% or more of patients with CIS
will progress to muscle-invasive disease [63,64].

Three rare subtypes of CIS must be considered


separately for their potential prognostic significance.
The first is non-invasive micropapillary CIS. It
was recently demonstrated that CIS associated
with micropapillary invasive cancer increased
recurrence upon univariate and multivariate
analysis. Furthermore, all superficial micropapillary
CIS displayed an invasive component [58]. Another
rare variant is the in situ squamous cell carcinoma
(SCC). SCC is frequently associated with an invasive
component. The last of these rare subtypes is in situ
adenocarcinoma, which is associated with classical
CIS and frequently with invasive carcinomas with
a poor prognosis such as micropapillary and small
cell CIS [59]. If these subtypes are reported, the

b) Intravesical Chemotherapy
As previously summarized, various intravesical
chemotherapeutic regimens have been compared in
patients with CIS [65,66,67]. The number of patients
included in these studies has often been small.
Consequently, definitive conclusions cannot be drawn
from the individual studies. Complete response rates
approaching 50% have been observed; nonetheless,
there may be important differences between these
studies with respect to patient characteristics
and assessment of response to treatment (level
1). Since 50% of the complete responders on
chemotherapy recurred during follow-up, complete
response alone is not a suitable endpoint because

253

of the risk of invasion and extravesical recurrence.


Indirect comparisons in a meta-analysis of 355
CIS patients treated with intravesical mitomycin C
(MMC), Adriamycin, epirubicin, or sequential MMC/
Adriamycin suggest that long-term disease-free
rates might be better with MMC [67]. However, direct
randomized comparisons do not exist (level 2).

chemotherapy (p=0.0002. After a median follow-up


of 3.6 years, 161 (46.7%) of 345 patients on BCG
exhibited no evidence of disease compared to 93
(26.2%) of 355 patients on chemotherapy (p<0.0001).
The long-term benefit of BCG was superior to MMC
in trials with maintenance BCG (p=0.04). A reduction
of 26% in the risk of progression for patients on BCG
when compared to those receiving chemotherapy
was observed (p=0.20), consistent with the results
of the meta-analysis (level 1).

c) Intravesical BCG
Intravesical bacillus Calmette-Gurin (BCG) has
been widely used in the treatment of patients
with CIS. The standard induction schedule of 6
weekly instillations yields a complete response
rate of approximately 70%, although complete
response rates over 80% have been reported in
individual studies (level 1) [65,67,68]. About 4060% of patients who do not respond to the initial
6 instillations will respond to a second series of 6
weekly instillations (level 2) [69,66]. Based on the
results of several studies and meta-analyses, 1 to 3
years of maintenance BCG therapy is recommended
in complete responders (level 1) [70, 67, 71, 72, 73].
Only half of all CIS patients will remain disease-free;
approximately one-third of all complete responders
will eventually recur. If recurrence occurs more than
1 year after BCG treatment, a re-induction with BCG
can be considered (level 3) [74].

More recently, in patients with high grade (grade 2


or 3) T1 tumors, BCG was found to be superior to
the combination of epirubicin and interferon alpha
2b for the endpoint of disease-free survival, but only
in the subgroup of 76 patients with concomitant CIS
[76]. In a recent meta-analysis comparing BCG to
MMC, only 12% of patients had CIS and a separate
analysis of treatment efficacy was not carried out in
this patient subgroup [77].

e) Comparison of Alternating or Sequential


MMC/BCG to BCG alone or to MMC alone
There is only a limited amount of evidence assessing
alternating MMC/BCG in patients with CIS. Alternating
MMC/BCG was compared to MMC alone in 133 CIS
patients in 2 studies, with higher complete response
and disease-free rates on MMC/BCG [78,79]. In the
largest trial comparing alternating MMC/BCG to BCG
alone in patients with CIS, there was a significantly
longer disease-free interval in the BCG monotherapy
arm: 80 (55%) of 145 patients were disease-free on
BCG alone compared to 72 (45%) of 159 patients on
the alternating arm [80]. Thus, it can be concluded
that an alternating schedule of MMC and BCG is not
superior to BCG alone (level 1).

These studies have all been carried out with full


dose BCG. However, a study in 155 patients with
high risk T1G3 and CIS disease concluded that a
one-third dose was as effective as a full dose BCG
course, and was associated with significantly less
toxicity (level 1) [75].

d) Comparison of Intravesical
Intravesical Chemotherapy

BCG

to

However, in a trial comparing sequential BCG and


electromotive MMC to BCG alone in 212 patients
with high grade (grade 2 or 3) T1 tumors, subgroup
analyses based on a median follow-up of 88 months
in the 57 patients with concomitant CIS showed
a statistically significant benefit of the sequential
treatment with regard to time to recurrence,
progression, and both disease-specific and overall
survival (level 1) [81]. It was suggested that BCGinduced inflammation might increase the permeability
of the bladder mucosa such that electromotive MMC
could reach the target tissue more easily and exert
greater anticancer effect.

Various randomized studies in high risk patients


have compared intravesical BCG to either no further
treatment, intravesical chemotherapy, or other
immunotherapies post TUR. Unfortunately, separate
results in CIS patients are often unavailable in many
of these studies that also included patients with
papillary tumors. A meta-analysis of 403 patients
with CIS showed that BCG reduced the risk of
progression by 35% compared to either intravesical
chemotherapy or a different immunotherapy (level
1). Twelve percent of patients with CIS on BCG
progressed as compared to 16% of patients receiving
other treatment strategies [70].

f) Treatment of Extravesical CIS

Previous reviews in 2005 identified 12 randomized


trials that compared intravesical BCG to intravesical
chemotherapy (MMC, Adriamycin, epirubicin,
Thiotepa, sequential MMC/Adriamycin) in 845
patients with CIS [66,69]. Data for 700 patients with
CIS included in 9 trials were available for inclusion in
a meta-analysis [67]. Among 298 patients receiving
BCG, 203 (68.1%) had a complete response as
compared to 158 (51.5%) of 307 patients receiving

No randomized trials have been carried out in patients


with extravesical CIS and, as such, no definitive
treatment recommendations can be provided.
Intravesical BCG does not reach either the upper
urinary tract or the prostatic urethra. Therefore, the
treatment of CIS in the upper urinary tract involves
rinsing the renal unit with BCG or chemotherapy.
CIS may be found in the prostatic urethra at various

254

levels: in the epithelial lining of the prostatic urethra


only, in the prostatic tissue following the prostatic
ducts, or in the prostatic tissue stroma (T4),
carrying the worst prognosis. In the latter case, a
cystoprostatectomy is advised. In the other two
cases, a TUR of the prostate can be followed by
intravesical BCG. Further details concerning results
obtained in a small series of non-randomized studies
have been summarized by van der Meijden (level 3)
[69].

hexaminolevulinic acid should be considered.


(LOE 1a, Grade A)
2. Diagnosis of CIS is based on histopathological
assessment of biopsy or resection specimen.
Pathology need not report subtypes but should
note the presence of micropapillary and small
cell CIS (LOE NA, Grade A).
3. Radical cystectomy at the time of CIS diagnosis
provides for excellent disease-free survival, but
is overtreatment in up to 50% of the patients.
(LOE 2, Grade A)

g) Treatment of BCG-refractory CIS


Up to 40-50% of patients with CIS may eventually
fail intravesical BCG. These patients have a very
poor prognosis, with a high risk of progression to
muscle-invasive disease and death as a result of
bladder cancer. The 3-year bladder cancer specific
survival is 67% in patients initially presenting with
muscle-invasive disease but falls to 37% in patients
who progress after intravesical treatment [82]. There
is no universally agreed definition of BCG failure
and there are no randomized trials comparing BCG
to cystectomy. However, treatment with BCG is
generally considered to have failed in the following
cases [83,84,85]: (1) detection of muscle-invasive
disease; (2) presence of high grade papillary tumors
and/or carcinoma in situ at both 3 months and 6
months; or (3) disease worsening (stage, grade, CIS)
under BCG treatment. Herr et al concluded that a
minimum treatment and follow-up period of 6 months
is required to identify high risk, non-muscle-invasive
bladder tumors as being truly BCG-refractory [85].

4. Intravesical BCG is recommended, as it provides


the highest complete response rate as well as
the highest long-term disease-free rate among
intravesical treatments. (LOE 1a, Grade A)
5. Concomitant papillary tumors should be
completely resected prior to the start of
intravesical treatment (LOE 1b, Grade A).
Following the resection, one immediate
instillation of chemotherapy may be considered
(LOE 3, Grade C).
6. Maintenance BCG treatment is required but the
optimal maintenance schedule is unknown. In
the absence of treatment failure, at least 1 year
of maintenance BCG is recommended. (LOE 2,
Grade A)
7. T
 he response to intravesical BCG should be
assessed 3 months after starting treatment. If
no response is seen, one might offer the patient
either cystectomy, another 6-week course of
BCG or 3 weekly boosters. The optimal time to
abandon conservative treatment and proceed to
cystectomy is unknown. (LOE 2, Grade B)

In BCG failures, cystectomy is the recommended


treatment option (level 3). For patients who are
unable or unwilling to have a cystectomy after BCG
failure, various alternative treatment possibilities exist
(level 3). Intravesical chemotherapy, device-assisted
chemotherapy, photodynamic therapy, and BCG
plus interferon alpha are the options most frequently
employed [86]. After a literature review in 2005 [69],
additional publications of intravesical treatment for
patients with BCG-refractory CIS became available
concerning BCG plus interferon alpha 2b [87],
hyperthermia and MMC [88], mycobacterial cell
wall-DNA complex [89], and gemcitabine [90,91]. In
general, the results for conservative treatment after
BCG failure are disappointing (level 3) and research
to discover new treatment options in these patients
is a top priority.

8. Cystoscopy and urinary cytology should be


performed every 3 months for a period of 2
years, every 4 months in the third year, every 6
months in the fourth and fifth year, and once per
year thereafter. If no recurrence, imaging of the
upper urinary tract by ultrasound, conventional
IVU, or CT should be performed periodically.
(LOE NA, Grade C)

References
1. Sauter G, Algaba F, Amin M, et al. Tumours of the urinary
system: non-invasive urothelial neoplasias. In: Eble
JN, Sauter G, Epstein Jl, Sesterhenn I, editors. WHO
Classification of Tumours of the Urinary System and Male
Genital Organs. Lyon: IARCC Press; 2004.

CIS: Recommendations

2. Hartmann A, Schlake G, Zaak D, et al. Occurrence of


chromosome 9 and p53 alterations in multifocal dysplasia
and carcinoma in situ of human urinary bladder. Cancer
Res 2002;62:809-18.

1. If CIS is clinically suspected, urinary cytology


should be obtained and white light cystoscopy
and imaging of the upper urinary tract should
be performed. (LOE 2, Grade A) In patients
with positive cytology but normal white light
cystoscopy, photodynamic diagnosis with

3. Hodges KB, Lopez-Beltran A, Davidson DD, Montironi R,


Cheng L. Urothelial dysplasia and other flat lesions of the
urinary bladder: clinicopathologic and molecular features.
Hum Pathol. 2010 Feb;41(2):155-62

255

4. Melicow MM. Histological study of vesical urothelium


intervening between gross neoplasms in total cystectomy.
J Urol 1952;68:261-79.

Mnsson W (2008): The Value of the UroVysion Assay for


Surveillance of Non-Muscle-Invasive Bladder Cancer. Eur
Urol 54(2):402-8

5. Melicow MM, Hollowell JW. Intra-urothelial cancer:


carcinoma in situ, Bowens disease of the urinary system:
discussion of thirty cases. J Urol 1952;68:763-72.

21. Fritsche HM, Burger M, Dietmaier W, Denzinger S, Bach E,


Otto W, Doblinger M, Schwarz S, Buchner H, Hartmann A.
Multicolor FISH (UroVysion) facilitates follow-up of patients
with high-grade urothelial carcinoma of the bladder. Am J
Clin Pathol. 2010 Oct;134(4):597-603.

6. Witjes JA: Bladder carcinoma in situ in 2003: state of the


art. Eur Urol 45: 142146, 2004.

22. Wild PJ, Fuchs T, Stoehr R, Zimmermann D, Frigerio S,


Padberg B, Steiner I, Zwarthoff EC, Burger M, Denzinger
S, Hofstaedter F, Kristiansen G, Hermanns T, Seifert HH,
Provenzano M, Sulser T, Roth V, Buhmann JM, Moch H,
Hartmann A. Detection of urothelial bladder cancer cells in
voided urine can be improved by a combination of cytology
and standardized microsatellite analysis. Cancer Epidemiol
Biomarkers Prev. 2009 Jun;18(6):1798-806. Epub 2009
May 19. Erratum in: Cancer Epidemiol Biomarkers Prev.
2010 Feb;19(2):629-30.

7. Denzinger S, Fritsche HM, Otto W, Blana A, Wieland WF,


Burger M. Early versus deferred cystectomy for initial highrisk pT1G3 urothelial carcinoma of the bladder: do risk
factors define feasibility of bladder-sparing approach? Eur
Urol. 2008 Jan;53(1):146-52.
8. Shariat SF, Palapattu GS, Karakiewicz PI, et al. Concomitant
carcinoma in situ is a feature of aggressive disease in
patients with organ-confined TCC at radical cystectomy.
Eur Urol 2007;51:152-60.
9. Dyrskjt L, Zieger K, Kruhffer M, Thykjaer T, Jensen JL,
Primdahl H, Aziz N, Marcussen N, Mller K, Orntoft TF.
A molecular signature in superficial bladder carcinoma
predicts clinical outcome. Clin Cancer Res. 2005 Jun
1;11(11):4029-36.

23. Witjes JA, Redorta JP, Jacqmin D, Sofras F, Malmstrm


PU, Riedl C, Jocham D, Conti G, Montorsi F, Arentsen HC,
Zaak D, Mostafid AH, Babjuk M. Hexaminolevulinate-guided
fluorescence cystoscopy in the diagnosis and follow-up of
patients with non-muscle-invasive bladder cancer: review
of the evidence and recommendations. Eur Urol. 2010
Apr;57(4):607-14.

10. Knowles MA. Molecular subtypes of bladder cancer: Jekyll


and Hyde or chalk and cheese? Carcinogenesis. 2006
Mar;27(3):361-73.

24. Kausch I, Sommerauer M, Montorsi F, Stenzl A, Jacqmin


D, Jichlinski P, Jocham D, Ziegler A, Vonthein R.
Photodynamic diagnosis in non-muscle-invasive bladder
cancer: a systematic review and cumulative analysis of
prospective studies. Eur Urol. 2010 Apr;57(4):595-60

11. Sylvester RJ, van der Meijden AP, Oosterlinck W, Witjes


JA, Bouffioux C, Denis L, Newling DW, Kurth K. Predicting
recurrence and progression in individual patients with stage
Ta T1 bladder cancer using EORTC risk tables: a combined
analysis of 2596 patients from seven EORTC trials. Eur
Urol. 2006 Mar;49(3):466-5;

25. Fradet Y, Grossman HB, Gomella L, et al. A comparison


of hexaminolevulinate fluorescence cystoscopy and white
light cystoscopy for the detection of carcinoma in situ in
patients with bladder cancer: a phase III, multicenter study.
J Urol 2007;178:68-73.

12. Shariat SF, Karakiewicz PI, Palapattu GS, Amiel GE, Lotan
Y, Rogers CG, Vazina A, Bastian PJ, Gupta A, Sagalowsky
AI, Schoenberg M, Lerner SP. Nomograms provide
improved accuracy for predicting survival after radical
cystectomy. Clin Cancer Res. 2006 Nov 15;12(22):666376.

26. Jocham D, Witjes F, Wagner S, Zeylemaker B, van


Moorselaar J, Grimm MO, Muschter R, Popken G, Knig
F, Knchel R, Kurth KH. Improved detection and treatment
of bladder cancer using hexaminolevulinate imaging:
a prospective, phase III multicenter study. J Urol. 2005
Sep;174(3):86

13. Zaak D, Burger M, Otto W, Bastian PJ, Denzinger S, Stief


CG, Buchner H, Hartmann A, Wieland WF, Shariat SF,
Fritsche HM. Predicting individual outcomes after radical
cystectomy: an external validation of current nomograms.
BJU Int. 2010 Aug;106(3):342-8

27. Karl A, Tritschler S, Stanislaus P, Gratzke C, Tilki D,


Strittmatter F, Knchel R, Stief CG, Zaak D. Positive
urine cytology but negative white-light cystoscopy: an
indication for fluorescence cystoscopy? BJU Int. 2009
Feb;103(4):484

14. van Rhijn BW, van der Poel HG, Boon ME, Debruyne FM,
Schalken JA, Witjes JA. Presence of carcinoma in situ and
high 2C-deviation index are the best predictors of invasive
transitional cell carcinoma of the bladder in patients with
high-risk Quanticyt. Urology. 2000 Mar;55(3):363-7.

28. Sharkey FE, and Sarosdy MF: The significance of central


pathology review in clinical studies of transitional cell
carcinoma in situ. J Urol 157: 6870, 1997.

15. Solsona E, Iborra I, Ricos JV, Dumont R, Casanova JL,


Calabuig C. Upper urinary tract involvement in patients with
bladder carcinoma in situ (Tis): its impact on management.
Urology 1997;49:347-52.

29. Witjes JA, Moonen PM, van der Heijden AG. Review
pathology in a diagnostic bladder cancer trial: effect of
patient risk category. Urology. 2006 Apr;67(4):751-5. Epub
2006 Mar 29.

16. Witjes JA. Bladder carcinoma in situ in 2003: state of the


art. Eur Urol 2004;45:142-6.

30. van Rhijn BW, Burger M, Lotan Y, et al. Recurrence and


progression of disease in non-muscle-invasive bladder
cancer: from epidemiology to treatment strategy. Eur Urol
2009;56:430-42.

17. Babjuk M, Oosterlinck W, Sylvester R, Kaasinen E, Bhle A,


Palou-Redorta J. EAU guidelines on non-muscle-invasive
urothelial carcinoma of the bladder. European Association
of Urology (EAU). Eur Urol. 2008 Aug;54(2):303-14.

31. Lamm DL. Superficial bladder cancer. Curr Treat Options


Oncol. 2002 Oct;3(5):403-11.

18. Sylvester RJ, van der Meijden A, Witjes JA, et al. Highgrade Ta urothelial carcinoma and carcinoma in situ of the
bladder. Urology 2005;66:90-107.

32. Lamm DL. Clinical evaluation of immunotherapy: are


there differences between papillary and flat in situ bladder
cancer? Cancer Surv 1998;31:99-108.

19. Kipp BR, Karnes RJ, Brankley SM, Harwood AR, Pankratz
VS, Sebo TJ, Blute MM, Lieber MM, Zincke H, Halling
KC (2005): Monitoring intravesical therapy for superficial
bladder cancer using fluorescence in situ hybridisation. J
Urol 173:401-404

33. McKenney JK, Gomez JA, Desai S, Lee MW, Amin MB


Morphologic expressions of urothelial carcinoma in situ: a
detailed evaluation of its histologic patterns with emphasis
on carcinoma in situ with microinvasion. Am J Surg Pathol.
2001 Mar;25(3):356-62.

20. Gudjnsson S, Isfoss BL, Hansson K, Domanski AM,


Warenholt J, Soller W, Lundberg LM, Liedberg F, Grabe M,

34. Boorjian SA, Zhu F, Herr HW.The effect of gender on

256

response to bacillus Calmette-Gurin therapy for patients


with non -muscle-invasive urothelial carcinoma of the
bladder. BJU Int. 2010 Aug;106(3):357-61.

49. Zieger K, Marcussen N, Borre M, rntoft TF, Dyrskjt L.


Consistent genomic alterations in carcinoma in situ of the
urinary bladder confirm the presence of two major pathways
in bladder cancer development. Int J Cancer. 2009 Nov
1;125(9):2095-103.

35. Takahashi T, Habuchi T, Kakehi Y, Mitsumori K, Akao T,


Terachi T, Yoshida O. Clonal and chronological genetic
analysis of multifocal cancers of the bladder and upper
urinary tract. Cancer Res. 1998 Dec 15;58(24):5835-41

50. Herbsleb M, Christensen OF, Thykjaer T, Wiuf C, Borre M,


Orntoft TF, Dyrskjt L. Bioinformatic identification of FGF,
p38-MAPK, and calcium signaling pathways associated
with carcinoma in situ in the urinary bladder. BMC 2008; 8:
37-48.

36. Griffiths TR, Charlton M, Neal DE, Powell PH. Treatment of


carcinoma in situ with intravesical bacillus Calmette-Guerin
without maintenance. J Urol. 2002 Jun;167(6):2408-12

51. Dyrskjt L, Zieger K, Real FX, Malats N, Carrato A, Hurst


C, Kotwal S, Knowles M, Malmstrm PU, de la Torre M,
Wester K, Allory Y, Vordos D, Caillault A, Radvanyi F, Hein
AM, Jensen JL, Jensen KM, Marcussen N, Orntoft TF. Gene
expression signatures predict outcome in non-muscleinvasive bladder carcinoma: a multicenter validation study.
Clin Cancer Res. 2007 15;13(12):3545-51

37. Hurle R, Losa A, Manzetti A, Lembo A. Intravesical bacille


Calmette-Gurin in Stage T1 grade 3 bladder cancer
therapy: a 7-year follow-up. Urology. 1999 Aug;54(2):25863.
38] Volkmer BG, Schnoeller T, Kuefer R, Gust K, Finter F,
Hautmann RE. Upper urinary tract recurrence after radical
cystectomy for bladder cancer--who is at risk?J Urol. 2009
Dec;182(6):2632-7.

52. Hopman AH, Kamps MA, Speel EJ, Schapers RF, Sauter G,
Ramaekers FC. Identification of chromosome 9 alterations
and p53 accumulation in isolated carcinoma in situ of the
urinary bladder versus carcinoma in situ associated with
carcinoma. Am J Pathol. 2002 Oct;161(4):1119-25.

39. Youssef RF, Shariat SF, Lotan Y, Wood CG, Sagalowsky


AI, Zigeuner R, Langner C, Montorsi F, Bolenz C, Margulis
V. Prognostic Effect of Urinary Bladder Carcinoma In Situ
on Clinical Outcome of Subsequent Upper Tract Urothelial
Carcinoma. Urology. 2010 Dec

53. Lentini L, Amato A, Schillaci T, Di Leonardo A. Simultaneous


Aurora-A/STK15
overexpression
and
centrosome
amplification induce chromosomal instability in tumour cells
with a MIN phenotype. BMC Cancer. 2007 13;7:212

40. Varinot J, Camparo P, Roupret M, Bitker MO, Capron F,


Cussenot O, Witjes JA, Comprat E. Full analysis of the
prostatic urethra at the time of radical cystoprostatectomy
for bladder cancer: impact on final disease stage. Virchows
Arch. 2009 Nov;455(5):449-53.

54. Comprat E, Camparo P, Haus R, Chartier-Kastler E,


Radenen B, Richard F, Capron F, Paradis V. Aurora-A/
STK-15 is a predictive factor for recurrent behaviour in noninvasive bladder carcinoma: a study of 128 cases of noninvasive neoplasms. Virchows Arch. 2007;450(4):419-24.

41. Liedberg F, Anderson H, Blckberg M, Chebil G, Davidsson


T, Gudjonsson S, Jahnson S, Olsson H, Mnsson W.
Prospective study of transitional cell carcinoma in the
prostatic urethra and prostate in the cystoprostatectomy
specimen. Incidence, characteristics and preoperative
detection. Scand J Urol Nephrol. 2007;41(4):290-6

55. Park HS, Park WS, Bondaruk J, Tanaka N, Katayama H,


Lee S, Spiess PE, Steinberg JR, Wang Z, Katz RL, Dinney
C, Elias KJ, Lotan Y, Naeem RC, Baggerly K, Sen S,
Grossman HB, Czerniak B. Quantitation of Aurora kinase A
gene copy number in urine sediments and bladder cancer
detection. J Natl Cancer Inst. 2008; 100(23):1740

42. Shen, SS, Lerner, SP, Muezzinoglu, B, et al. Prostatic


involvement by transitional cell carcinoma in patients with
bladder cancer and its prognostic significance. Hum Pathol
2006; 37:726.

56. Horikawa Y, Sugano K, Shigyo M, Yamamoto H, Nakazono


M, Fujimoto H, Kanai Y, Hirohashi S, Kakizoe T, Habuchi
T, Kato T. Hypermethylation of an E-cadherin (CDH1)
promoter region in high grade transitional cell carcinoma
of the bladder comprising carcinoma in situ. J Urol.
2003;169(4):1541-5

43. Hassan JM, Cookson MS, Smith JA Jr, Johnson DL, Chang
SS. Outcomes in patients with pathological carcinoma
in situ only disease at radical cystectomy. J Urol. 2004
Sep;172(3):882-4
44. Orozco RE, Martin AA, Murphy WM. Carcinoma in situ of
the urinary bladder. Clues to host involvement in human
carcinogenesis. Cancer. 1994 Jul 1;74(1):115-22.

57. Bindels EM, Vermey M, van den Beemd R, Dinjens WN,


Van Der Kwast TH. E-cadherin promotes intraepithelial
expansion of bladder carcinoma cells in an in vitro model of
carcinoma in situ. Cancer Res. 2000 1;60(1):177-83.

45. Chade DC, Shariat SF, Adamy A, Bochner BH, Donat


SM, Herr HW, Dalbagni G. Clinical outcome of primary
versus secondary bladder carcinoma in situ. J Urol. 2010
Aug;184(2):464-9.

58. Comprat E, Roupret M, Yaxley J, Reynolds J,


Varinot J, Ouzad I, Cussenot O, Samaratunga H.
Micropapillary urothelial carcinoma of the urinary bladder:
a clinicopathological analysis of 72 cases. Pathology.
2010;42(7):650-4

46. Lopez-Beltran A, Ordez JL, Otero AP, Blanca A, Sevillano


V, Sanchez-Carbayo M, Muoz E, Cheng L, Montironi R, de
Alava E. Cyclin D3 gene amplification in bladder carcinoma
in situ. Virchows Arch. 2010;457(5):555-61.

59. Tavora F, Epstein JI. Bladder cancer, pathological


classification and staging. BJU Int. 2008;102(9 Pt B):121620.

47. Andius P, Damm O, Holmng S. Prognostic factors in


patients with carcinoma in situ treated with intravesical
bacille Calmette-Gurin. Scand J Urol Nephrol.
2004;38(4):285-90.

60. Tilki D, Reich O, Svatek RS, Karakiewicz PI, Kassouf W,


Novara G, Ficarra V, Chade DC, Fritsche HM, Gerwens N,
Izawa JI, Lerner SP, Schoenberg M, Stief CG, Skinner E,
Lotan Y, Sagalowsky AI, Shariat SF. Characteristics and
outcomes of patients with clinical carcinoma in situ only
treated with radical cystectomy: an international study of
243 patients. J Urol. 2010 May;183(5):1757-63.

48. Fernandez-Gomez J, Solsona E, Unda M, Martinez-Pieiro


L, Gonzalez M, Hernandez R, Madero R, Ojea A, Pertusa
C, Rodriguez-Molina J, Camacho JE, Isorna S, Rabadan
M, Astobieta A, Montesinos M, Muntaola P, Gimeno A,
Blas M, Martinez-Pieiro JA; Club Urolgico Espaol de
Tratamiento Oncolgico (CUETO). Prognostic factors in
patients with non-muscle-invasive bladder cancer treated
with bacillus Calmette-Gurin: multivariate analysis
of data from four randomized CUETO trials. Eur Urol.
2008;53(5):992-1001.

61. Huang GJ, Kim PH, Skinner DG, Stein JP Outcomes of


patients with clinical CIS-only disease treated with radical
cystectomy. World J Urol. 2009 Feb;27(1):21-5.
62. Shariat SF, Palapattu GS, Amiel GE, Karakiewicz PI, Rogers
CG, Vazina A, Schoenberg MP, Lerner SP, Sagalowsky AI,
Lotan Y. Characteristics and outcomes of patients with

257

carcinoma in situ only at radical cystectomy. Urology. 2006


Sep;68(3):538-42.

subsequent response to bacille Calmette-Gurin plus


interferon intravesical therapy. Urology. 2008 Feb;71(2):297301.

63. Long-term results of intravesical therapy for superficial


bladder cancer. Lamm DL. Urol Clin North Am. 1992
Aug;19(3):573-80.

75. Martnez-Pieiro JA, Martnez-Pieiro L, Solsona E,


Rodrguez RH, Gmez JM, Martn MG, Molina JR, Collado
AG, Flores N, Isorna S, Pertusa C, Rabadn M, Astobieta
A, Camacho JE, Arribas S, Madero R; Club Urolgico
Espaol de Tratamiento Oncolgico (CUETO). Has a 3-fold
decreased dose of bacillus Calmette-Guerin the same
efficacy against recurrences and progression of T1G3 and
Tis bladder tumours than the standard dose? Results of
a prospective randomized trial. J Urol. 2005 Oct;174(4 Pt
1):1242-7.

64. Lamm DL, Blumenstein BA, Crissman JD, Montie JE,


Gottesman JE, Lowe BA, Sarosdy MF, Bohl RD, Grossman
HB, Beck TM, Leimert JT, Crawford ED. Maintenance
bacillus Calmette-Guerin immunotherapy for recurrent TA,
T1 and carcinoma in situ transitional cell carcinoma of the
bladder: a randomized Southwest Oncology Group Study.
J Urol. 2000 Apr;163(4):1124-9.
65. Oosterlinck W, Kirkali Z, Sylvester R, Silva FC, Busch
C, Algaba F, Collette S, Bono A. Sequential Intravesical
Chemoimmunotherapy with Mitomycin C and Bacillus
Calmette-Gurin and with Bacillus Calmette-Gurin Alone
in Patients with Carcinoma in Situ of the Urinary Bladder:
Results of an EORTC Genito-Urinary Group Randomized
Phase 2 Trial (30993). Eur Urol. 2010 Dec 7. [Epub ahead
of print]

76. Duchek M, Johansson R, Jahnson S, Mestad O, Hellstrm


P, Hellsten S, Malmstrm PU; Members of the Urothelial
Cancer Group of the Nordic Association of Urology. Bacillus
Calmette-Gurin is superior to a combination of epirubicin
and interferon-alpha2b in the intravesical treatment of
patients with stage T1 urinary bladder cancer. A prospective,
randomized, Nordic study. Eur Urol. 2010 Jan;57(1):25-31.
77. Malmstrm PU, Sylvester RJ, Crawford DE, Friedrich
M, Krege S, Rintala E, Solsona E, Di Stasi SM, Witjes
JA. An individual patient data meta-analysis of the
long-term outcome of randomised studies comparing
intravesical mitomycin C versus bacillus Calmette-Gurin
for non-muscle-invasive bladder cancer. Eur Urol. 2009
Aug;56(2):247-56.

66. Sylvester RJ, van der Meijden A, Witjes JA, Jakse G,


Nonomura N, Cheng C, Torres A, Watson R, Kurth KH.
High-grade Ta urothelial carcinoma and carcinoma in situ
of the bladder. Urology. 2005 Dec;66(6 Suppl 1):90-107.
67. Sylvester RJ, van der Meijden AP, Witjes JA, Kurth K.
Bacillus calmette-guerin versus chemotherapy for the
intravesical treatment of patients with carcinoma in situ
of the bladder: a meta-analysis of the published results of
randomized clinical trials. J Urol. 2005 Jul;174(1):86-91;
discussion 91-2.

78. Shelley MD, Court JB, Kynaston H, Wilt TJ, Coles B, Mason
M. Intravesical bacillus Calmette-Guerin versus mitomycin
C for Ta and T1 bladder cancer. Cochrane Database Syst
Rev. 2003;(3):CD003231.

68. Takenaka A, Yamada Y, Miyake H, Hara I, Fujisawa M.


Clinical outcomes of bacillus Calmette-Gurin instillation
therapy for carcinoma in situ of urinary bladder. Int J Urol.
2008 Apr;15(4):309-13.

79. Kaasinen E, Wijkstrm H, Malmstrm PU, Hellsten S,


Duchek M, Mestad O, Rintala E; Nordic Urothelial Cancer
Group. Alternating mitomycin C and BCG instillations versus
BCG alone in treatment of carcinoma in situ of the urinary
bladder: a nordic study. Eur Urol. 2003 Jun;43(6):637-45.

69. van der Meijden AP, Sylvester R, Oosterlinck W, Solsona E,


Boehle A, Lobel B, Rintala E; for the EAU Working Party on
Non Muscle Invasive Bladder Cancer EAU guidelines on
the diagnosis and treatment of urothelial carcinoma in situ.
Eur Urol. 2005 Sep;48(3):363-71.

80. Rintala E, Jauhiainen K, Rajala P, Ruutu M, Kaasinen E,


Alfthan Alternating mitomycin C and bacillus CalmetteGuerin instillation therapy for carcinoma in situ of the bladder.
The Finnbladder Group. J Urol. 1995 Dec;154(6):2050-3.

70. Witjes JA, v d Meijden AP, Collette L, Sylvester R,


Debruyne FM, van Aubel A, Witjes WP .Long-term followup of an EORTC randomized prospective trial comparing
intravesical bacille Calmette-Gurin-RIVM and mitomycin
C in superficial bladder cancer. EORTC GU Group and the
Dutch South East Cooperative Urological Group. European
Organisation for Research and Treatment of Cancer
Genito-Urinary Tract Cancer Collaborative Group. Urology.
1998 Sep;52(3):403-10.

81. Di Stasi SM, Giannantoni A, Giurioli A, Valenti M, Zampa G,


Storti L, Attisani F, De Carolis A, Capelli G, Vespasiani G,
Stephen RL. Sequential BCG and electromotive mitomycin
versus BCG alone for high-risk superficial bladder
cancer: a randomised controlled trial. Lancet Oncol. 2006
Jan;7(1):43-51.
82. Prognosis of muscle-invasive bladder cancer: difference
between primary and progressive tumours and implications
for therapy. Schrier BP, Hollander MP, van Rhijn BW,
Kiemeney LA, Witjes JA. Eur Urol. 2004 Mar;45(3):292-6.

71. Intravesical bacille Calmette-Gurin versus mitomycin


C in superficial bladder cancer: formal meta-analysis of
comparative studies on tumour progression. Bhle A, Bock
PR. Urology. 2004 Apr;63(4):682-6; discussion 686-7.

83. Martin FM, Kamat AM. Definition and management of


patients with bladder cancer who fail BCG therapy. Expert
Rev Anticancer Ther. 2009 Jun;9(6):815-20.

72. Jrvinen R, Kaasinen E, Sankila A, Rintala E; FinnBladder


Group. Long-term efficacy of maintenance bacillus CalmetteGurin versus maintenance mitomycin C instillation therapy
in frequently recurrent TaT1 tumours without carcinoma in
situ: a subgroup analysis of the prospective, randomised
FinnBladder I study with a 20-year follow-up. Eur Urol.
2009 Aug;56(2):260-5.

84. Zlotta AR, Fleshner NE, Jewett MA. The management of


BCG failure in non-muscle-invasive bladder cancer: an
update. Can Urol Assoc J. 2009 Dec;3(6 Suppl 4):S199205.
85. Huguet J, Crego M, Sabat S, Salvador J, Palou J,
Villavicencio H. Cystectomy in patients with high risk
superficial bladder tumours who fail intravesical BCG
therapy: pre-cystectomy prostate involvement as a
prognostic factor. Eur Urol. 2005 Jul;48(1):53-9; discussion
59.

73. Jrvinen R, Kaasinen E, Sankila A, Rintala E; FinnBladder


Group. Long-term efficacy of maintenance bacillus CalmetteGurin versus maintenance mitomycin C instillation therapy
in frequently recurrent TaT1 tumours without carcinoma in
situ: a subgroup analysis of the prospective, randomised
FinnBladder I study with a 20-year follow-up. Eur Urol.
2009 Aug;56(2):260-5.

86. Sylvester RJ.Bacillus Calmette-Gurin treatment of


non-muscle invasive bladder cancer. Int J Urol. 2011
Feb;18(2):113-20.

74. Gallagher BL, Joudi FN, Maym JL, ODonnell MA. Impact
of previous bacille Calmette-Gurin failure pattern on

87. Joudi FN, Smith BJ, ODonnell MA; National BCGInterferon Phase 2 Investigator Group. Final results from

258

in patients with stage T1 urothelial carcinoma was


unclear; of 111 patients with stage T1 disease who
underwent radical cystectomy, only 10 (9%) had
lymphovascular invasion.

a national multicenter phase II trial of combination bacillus


Calmette-Gurin plus interferon alpha-2B for reducing
recurrence of superficial bladder cancer. Urol Oncol. 2006
Jul-Aug;24(4):344-8.
88. Topic issue on new treatments in bladder cancer. Witjes JA.
World J Urol. 2009 Jun;27(3):285-7. Witjes JA, Hendricksen
K, Gofrit O, Risi O, Nativ O. Intravesical hyperthermia and
mitomycin-C for carcinoma in situ of the urinary bladder:
experience of the European Synergo working party. World
J Urol. 2009;27:319-24

A number of molecular prognostic markers have been


reported for T1 urothelial carcinoma. It has been
suggested that alterations of cell cycle regulatory
proteins involved in the progression from G1 to S
phase are among the most promising markers. The
p53 tumor suppressor gene is commonly altered
in human malignancies. Pretreatment p53 nuclear
overexpression in non-muscle-invasive bladder
tumors is associated with a high risk of disease
recurrence, progression, and cancer death after
BCG therapy [3]. Alterations of p53 were associated
with BCG failure. Others have not found that p53
expression before BCG treatment correlated with
BCG failure [2]. Controversial results have also
been reported with regard to p53 expression as an
independent predictor of progression [4]. While a
trend towards poorer clinical outcomes in high risk
patients with a p53 mutation has been observed, no
significant differences were seen in clinical outcome
parameters. A number of cell cycle regulators that
appear to be independent predictors of the survival
of patients with high grade T1 urothelial carcinoma
include p27kip1 and the cyclins D1 and D3 [5]. Many
other prognostic factors, including genetic alterations,
cell adhesion molecules, a family of proteases,
growth factors, and other molecular markers, have
been studied. To date, they do not have enough
specificity for clinical use in the management of T1
urothelial carcinoma. For patients with pT1 disease
at cystectomy, assessment of p53, p27, and Ki-67
in bladder specimens has been shown to improve
prediction of recurrence-free and disease-specific
survival. [6]

89. Morales A, Phadke K, Steinhoff G. Intravesical mycobacterial


cell wall-DNA complex in the treatment of carcinoma in situ
of the bladder after standard intravesical therapy has failed.
J Urol. 2009 Mar;181(3):1040-5.
90. Dalbagni G, Russo P, Bochner B, Ben-Porat L, Sheinfeld
J, Sogani P, Donat MS, Herr HW, Bajorin D. Phase II trial
of intravesical gemcitabine in bacille Calmette-Gurinrefractory transitional cell carcinoma of the bladder. J Clin
Oncol. 2006 Jun 20;24(18):2729-34.
91. Bassi P, De Marco V, Tavolini IM, Longo F, Pinto F, Zucchetti
M, Crucitta E, Marini L, Dal Moro F. Pharmacokinetic study
of intravesical gemcitabine in carcinoma in situ of the
bladder refractory to bacillus Calmette-Gurin therapy. Urol
Int. 2005;75(4):309-13.

III. T1 Urothelial Carcinoma


1. Diagnosis and Risk Assessment
The initial assessment of T1 urothelial carcinoma does
not differ decisively from Ta urothelial carcinoma. If
pT1 cancer has been confirmed pathologically, there
are additional considerations.
Clinical predictors, including histopathologic markers,
are most useful in forecasting progression in patients
with stage T1 urothelial carcinoma. Traditional
factors used to predict the clinical outcome of T1
urothelial bladder carcinoma after an initial TURBT
include time to recurrence, tumor grade, multiplicity,
tumor extent and size, concomitant CIS, urothelial
carcinoma involving the prostatic mucosa or ducts,
and depth of lamina propria invasion [1]. The
response to intravesical therapy is a reliable predictor
of progression within 3-6 months. Eighty percent of
patients found to have evidence of tumor recurrence
at the time of their 3-month cystoscopy after BCG
induction ultimately progressed to invasive disease
[2]. The relative importance of clinical and pathologic
factors varies according to adjuvant therapy. For
example, T1 substaging has not been found to be
useful in BCG-treated patients, and recurrence or
persistence of downstaged or downgraded disease
may be evidence of response in an individual patient
[2]. Nevertheless, these prognostic factors are not
accurate enough to predict the individual clinical
behavior of T1 urothelial tumors. Therefore, more
reliable indicators of biologic aggressiveness are
needed.

In summary, useful clinical prognostic factors for


T1 urothelial carcinoma include tumor grade, early
recurrence, multiplicity, tumor size, concomitant CIS,
urothelial carcinoma involving the prostatic mucosa
or ducts, and depth of lamina propria invasion. The
response to intravesical therapy is a very useful
clinical marker. A great number of molecular and
genetic prognostic markers, including alterations of
p53, have been studied for T1 urothelial carcinoma.
However, most of these markers have not been
validated and are not yet available for clinical use.

2. Treatment
a) Turbt
A restaging TURBT is warranted in patients with pT1
urothelial carcinoma. In addition to the previously
stated aspects, some special considerations also
apply. The rate of residual tumors is high after initial
TURBT for pT1 urothelial carcinoma. Schwaibold
et al [7] showed that 52% of patients had residual
tumors at restaging TURBT for initial T1 urothelial

Lotan et al examined the association of lymphovascular


invasion with prognosis at radical cystectomy.
Unfortunately, the role of lymphovascular invasion

259

carcinoma, of which 86% were located in the base or


margins of the previous resection and 14% in other
locations. The risk of residual tumor correlates with
the initial tumor grading. Divrik et al [8] reported a
rate of 6% for initial G1, 38% for G2, and 63% for G3
tumors. Even after fluorescence diagnosis at the first
TURBT for initial high grade T1 tumors, a restaging
TURBT revealed residual tumors in 12% of cases
[9].

Hendricksen [49] studied the additive effect of early


instillation or maintenance instillations of adjuvant
intravesical epirubicin compared to the standard
schedule of epirubicin in patients with intermediate
and high risk urothelial carcinoma of the bladder.
In a randomized study (n=753), 4 weekly and 5
monthly instillations of epirubicin 50 mg in 50 mL
saline for 1 hour (group 1) were compared to the
same schedule with an additional instillation 1) < 48
hours after TUR (group 2) and 2) following the same
schedule of group 1 with additional instillations at
9 and 12 mo (maintenance schedule). The results
showed that with this quasi intention-to-treat strategy
there was no difference in the 5-year recurrence-free
period between the treatment groups. T1 disease
was present in 156 (21%) of 731 patients. Grade 2
and grade 3 disease was present in 343 (47%) and
78 (11%), respectively. The immediate instillation
was given to 68% of the patients in group 2 within
24 hours. They concluded that one immediate
postoperative instillation of epirubicin preceding a
standard schedule also was not effective in reducing
recurrences in patients with intermediate and high
risk urothelial carcinoma of the bladder.

Dutta et al [10] reported a 64% risk of tumor


understaging in the absence of muscle in the first
TURBT specimen. As 30% of tumors were upstaged
when muscle was present in the first obtained
specimen, a restaging TURBT should be performed
for all high grade T1 tumors. Dalbagni et al [11] also
reported a rate of 20% upstaging after restaging
TURBT for initial high grade T1 tumors and thus
modified patient treatment strategies. A restaging
TURBT also provides prognostic markers for disease
outcome. In a retrospective study, Herr [12] found a
significantly higher risk of tumor progression over
5 years when residual pT1 disease was found at
restaging TURBT (82% vs. 19%). Thus, a restaging
TURBT helps to identify patients who should be
immediately referred for radical cystectomy.

In a multicenter study in Sweden [15], 305 patients


with primary or recurrent Ta or T1, G1 or G2 urothelial
carcinoma of the bladder (low or intermediate risk)
were randomized to receive epirubicin 80 mg in
50 mL saline within 24 hours of TURBT or TURBT
alone. The results showed that after a median
follow-up time of 3.9 years, the recurrence rate
in the epirubicin group was 62% vs. 72% in the
control group (p= 0.016). However, in a subgroup
analysis, the treatment with epirubicin only had
profound recurrence-reducing effect in patients with
primary, single tumors, while providing no benefit in
patients with recurrent or multiple Ta or T1, G1 or G2
intermediate risk tumors.

b) Immediate Instillation of Chemotherapy


The European Association of Urology (EAU)
guidelines recommend that all patients with nonmuscle-invasive bladder cancer should receive 1
immediate instillation of chemotherapy after a TUR
(Grade B recommendation) [13]
Patients with T1 urothelial carcinoma of the bladder
are at at least intermediate or high risk for progression,
depending on the grade. After TUR, these patients
should receive further adjuvant chemotherapy or
immunotherapy (BCG) once a week for 6 weeks
with or without maintenance. The role of one single
instillation of chemotherapy immediately after TUR
in these patients (T1, any grade) has been recently
questioned due to the publication of new data.

Similar findings were reported by Berrum-Svennung


et al [16]. They reported on 307 evaluable patients
randomized between receiving 50 mg epirubicin
within 6 hours after TURBT or TURBT alone.
The recurrence rate in the epirubicin group was
significantly lower (51% vs. 62.5%, p=0.04), but
recurrences were prevented only if initial tumors were
smaller than 5 mm. The authors also concluded that
an immediate single postoperative instillation was
only effective in small (i.e., very low risk) tumors.

In a randomized placebo-controlled phase III


multicenter study, Bohle [14] compared a single
postoperative instillation of gemcitabine to placebo
in patients with low or intermediate risk urothelial
carcinoma of the bladder. In this series, 25% and
29% of patients receiving gemcitabine and placebo,
respectively, had T1 disease; 10.5% and 11.3% were
high grade. In these patients, additional BCG with
maintenance was also used. The results showed that
after a follow-up period of 24 months, the recurrencefree survival was high in both groups (77.7% vs.
76.3%). There was no significant difference between
the 2 groups. Both groups had continuous bladder
irrigation for 20 hours. The authors hypothesized
that improved TUR (cystoscopy) techniques may
have contributed to the high recurrence-free survival
in both groups.

Finally, the value of a single instillation after a TURBT


in high risk patients has been recently challenged by
Brausi [17]. In Sylvesters meta-analysis, the number
of high risk patients who received 1 perioperative
instillation and further therapy was small. The
consistency of the conclusions regarding these
patients was unclear.
In the literature, there are few randomized studies
addressing this issue. The first was published by

260

Zincke in 1983 [18]. He compared 1 single instillation


of Thiotepa or doxorubicin after TURBT followed
by BCG therapy vs. TURBT without postoperative
instillation followed by BCG in patients with CIS.
The results indicated no benefit in those patients
receiving a single instillation. A second prospective
randomized study by Cai [19] compared 1 single
instillation of epirubicin after a TURBT followed by
BCG vs. TURBT without instillation and BCG in
high risk tumors. No difference in the recurrence
or progression rate was demonstrated between
the 2 groups. The authors concluded that 1 single
instillation of chemotherapy after a TURBT in high
grade, high risk tumors is not effective in reducing
recurrence or progression.

c) Initial Bladder-sparing
Cystectomy

Approach

reports of increased mortality in patients who develop


muscle-invasive disease while receiving conservative
treatment who then require cystectomy. Obviously,
this raises concern about delaying cystectomy.

d) Intravesical Therapy
Before the advent of BCG immunotherapy, the
incidence of progression in high grade, stage T1
urothelial carcinoma ranged from 27-65% with
follow-up ranging from 36-84 months. Important and
valid information about the prognosis of patients with
non-muscle-invasive bladder cancer can be obtained
from the scoring system developed by the EORTC)
According to these risk tables, the risk of progression
in high grade T1 tumors ranges between 17% and
45% at 5 years, depending on further variables [22].
These results, however, were before the era of a
second TURBT or fluorescence-guided resection,
maintenance BCG, and the use of an immediate
postoperative instillation of chemotherapy.

vs.

The goal for the treatment of bladder cancer is to


minimize the morbidity and mortality of the disease
while maximizing patient quality of life. This is
particularly true for cases of new or recurrent T1
disease. The substantial majority of stage T1 bladder
cancer is high grade and therefore at particular risk
of progression to incurable metastases. Treatment
with TURBT alone without adjuvant intravesical
therapy carries an unacceptable risk of recurrence
and progression. Even with optimal intravesical
chemotherapy or immunotherapy, however, a
substantial proportion of patients progress to
incurable disease stages [20].

With the advent of (maintenance) BCG, reported


results of intravesical therapy in T1 disease have
improved. With follow-up times ranging from 22-78
months, overall progression is in the range of 12%,
varying from 0-35%. It is quite remarkable that the
overall incidence of progression in patients with T1
disease is less than the incidence of occult muscleinvasive disease (up to 62%) in patients who undergo
an immediate cystectomy [23,24,25].
The role of BCG in the treatment of high grade nonmuscle-invasive bladder cancer is becoming clearer.
Intravesical BCG is more effective than intravesical
chemotherapy or TURBT alone in decreasing
recurrences. This was addressed by many
prospective randomized trials and meta-analyses.
Unfortunately, these trials enrolled patients both with
intermediate and high risk tumors. For this reason,
there is no separate analysis available based on
prospective randomized data that deals with the
efficacy of different treatment modalities in high
grade T1 tumors.

The 2 treatment options available today for patients


with high grade Ta or T1 bladder cancer are
immediate cystectomy or intravesical therapy with
a delayed cystectomy for persistent or recurrent
tumors. The selection of one of these options over
the other is among the most difficult in bladder
cancer management. Quality of life data in this area
is interesting. If faced with this choice, patients are
willing to accept disabilities incurred by cystectomy
to avoid compromising survival.
Selection of the treatment strategy for stage T1
urothelial carcinoma requires consideration of the
risks and benefits of cystectomy versus a repeat
TURBT with immediate instillation of chemotherapy
and BCG immunotherapy. Understaging is frequent.
When a cystectomy is performed for clinical stage
T1 disease, 30% of patients will be found to have
unsuspected disease that is stage T2 or greater. It
was recently shown that this risk increases to 49.7%
in patients with clinical high grade T1 disease [21].
Cystectomy is therefore an option to be considered.
However, cystectomy survival is far from optimal
and around one-third of patients with pT1 urothelial
carcinoma die of the disease [21]. Because of the
lack of evidence from randomized trials, there is no
recommendation for early cystectomy in patients
with high grade T1 disease. Nevertheless there are

It has been confirmed in 5 meta-analyses that


BCG therapy after TURBT is better than TURBT
only or TURBT plus chemotherapy in preventing
recurrences of non-muscle-invasive bladder tumors
[26,27,28,29,40]. A recently published metaanalysis, which was based on individual data from
2820 patients from 9 randomized trials, compared
the efficacy of BCG and MMC. In those trials where
BCG was used with maintenance, a 32% reduction
in the risk of recurrence was found (p<0.0001) [30].
Secondly, BCG lowers tumor progression in high
grade non-muscle-invasive bladder cancer but has
not been shown to improve survival. Patients with
high grade T1 tumors are at high risk for tumor
progression, which is more important and dangerous
than the risk of tumor recurrence. The question

261

as to whether BCG can reduce the risk of tumor


progression was addressed by a meta-analysis
carried out by the EORTC that evaluated data from
4863 patients enrolled in 24 randomized trials. The
results demonstrated a 27% reduction in the odds of
progression with BCG treatment (p=0.0001). In the
20 trials in which BCG maintenance was provided,
a reduction of 37% in progression was observed
(p=0.00004) [31] (Level 1). Unfortunately, 1) the
median follow-up was only 2.5 years; and 2) for
papillary tumors, only 7.6% had high grade disease.
This individual data meta-analysis, with a median
follow up of 4.4 years, compared BCG with MMC
and has not confirmed any significant difference in
progression and mortality rates [30].

BCG instillations are given according to the empirical


6-weekly induction schedule introduced by Morales
30 years ago. However, many different maintenance
schedules have been used, ranging from a total of
10 instillations given in 18 weeks, to 30 instillations
given over 3 years according to the Southwest
Oncology Group (SWOG) schedule [37]. The quality
of the initial TURBT in adequate staging and optimal
resection has a crucial role in providing optimal
results in a bladder sparing approach to high grade
T1 cases.
Conservative therapy is an attractive option for
many patients with high grade non-muscle-invasive
bladder cancer. There is a subgroup of patients with
the highest risk of progression who benefit from
early cystectomy. These may include patients with
adverse factors, for example, patients with tumor
multifocality, tumor size of over 3 cm, and associated
CIS. It has been shown that patients with at least
two of these factors could benefit from immediate
cystectomy [38].

Indeed, there are some data indicating that the longterm risk of tumor progression remains significant
in spite of BCG treatment. It has been reported that
after 15 years of follow-up, 53% of patients with initial
high-risk non-invasive bladder cancer progressed to
muscle-invasive disease, 36% eventually underwent
cystectomy, and 34% died of bladder cancer [32].
Only 27% were living with an intact bladder. Similar
results have been reported by Shahin et al [33] from
a study with 153 patients treated with either TURBT
and BCG or TURBT alone. With a median followup of 5.3 years, disease recurred in 70% and 75%
of patients treated with BCG and TURBT alone,
respectively. Progression was seen in 33% and 36%
of cases, respectively. A deferred cystectomy was
performed in 29% and 31% of cases, respectively.
Overall and disease-specific survival for those
receiving BCG compared with TURBT alone was
42% and 77% versus 48% and 79%, respectively.
The investigators stated that BCG therapy is unlikely
to substantially alter the final outcome for patients
with high grade T1 urothelial carcinoma.

The most important prognostic factor has been


shown to be the clinical response to intravesical
therapy. A study of patients treated with intravesical
doxorubicin found that 50% of those with stage T1 or
high grade recurrent tumors progressed. [39] The
pathologic tumor stage at recurrence was the only
factor shown to be predictive of recurrence in this
study. In 80 patients given intravesical therapy for T1
disease, Solsona et al [40] demonstrated that 78%
had a complete clinical response. However, as is
the case of many studies of high risk non-muscleinvasive bladder cancer, in this study, T1 tumors
were not separated from CIS in the analysis of risk
factors for progression. In conclusion, the absence of
a 3-month clinical response seems highly predictive
of progression [41]. It is not known exactly whether
a cystectomy should be performed immediately
at 3 months in non-responders or whether we can
repeat BCG instillations and wait for a response at
6 months. There is no doubt, however, that an early
cystectomy is strongly recommended in patients with
persistent high grade tumors despite 3 and 6 month
BCG treatments, since the prognosis for progression
increases significantly [42].

BCG instillation is associated with side effects. Up


to 90% of patients have irritative lower urinary tract
symptoms, and a small number of patients have
serious, debilitating complications, such as sepsis
and contracted bladders. Lamm [34] reported that
only 16% of patients were able to tolerate a full
maintenance course of BCG secondary to adverse
effects, albeit in an earlier era. With increased
experience in the use of BCG, the side effects now
appear to be less severe. Serious side effects are
encountered in less than 5% of patients and can be
effectively treated in virtually all cases. [35] (Level
3).

e) Early Cystectomy
Disease-specific survival after cystectomy for nonmuscle-invasive urothelial carcinoma is higher than
for patients with muscle-invasive disease. However,
the benefits must be balanced against the morbidity
and mortality rates of surgery. Amling et al [43]
reported on a large series of 531 patients undergoing
cystectomy and found a perioperative mortality rate
of 2.3% and a complication rate of 20.5% (level 3).
Similar complication rates have been reported in
other large cystectomy series by Stein et al [44] and
Ghoneim et al [45] (level 3).

For optimal efficacy, BCG must be given according


to a maintenance schedule [29,30,31,36]. It is
not possible, however, to determine which BCG
maintenance schedule is the most effective [31]. In
their meta-analyses, Bhle et al concluded that at
least 1 year of maintenance BCG was required to
show the superiority of BCG over MMC in preventing
recurrence or progression [29,30,31,35,36]. Today,

262

There is significant support for early cystectomy in


T1 disease in view of the high rate of late failure
after initially successful intravesical therapy and
the good quality of life possible post-cystectomy.
[46] Fritsche [21] reported on a multicenter series
of 1136 patients with high grade T1 disease
who underwent cystectomy without undertaking
neoadjuvant chemotherapy. Upstaging to muscleinvasive disease occurred in 49.7% and 35.5%
died of metastatic disease within 8 years of their
cystectomies (level 3).

were 85% and 82% at 5 and 10 years, respectively.


A total of 246 patients (24%) had lymph node tumor
involvement. The 5- and 10-year recurrence-free
survival rates for these patients were 35% and 34%,
respectively (Level 3).
Based on the current literature, patients with high
risk disease should be offered an early cystectomy.
Likewise, those with recurrent or persistent disease
after BCG immunotherapy should also be offered
cystectomy.

T1 Urothelial Carcinoma:
Recommendations

However, the proponents of an aggressive initial


approach acknowledge that a significant number of
patients will be rendered disease free with bladdersparing strategies. Current proposed indications
for immediate surgery include younger patients
with T1 tumors with at least one additional bad
prognostic factor including multifocality, associated
CIS, prostatic involvement, and tumor at a site that
is difficult to resect. Bianco et al [47] performed
a multivariate analysis to identify risk factors in
patients undergoing a cystectomy that influenced
cancer-specific survival. They found that patients
with concomitant CIS and those who had persistent
disease after an initial course of BCG therapy were
at significant risk (level 3). Herr and Sogani [48]
retrospectively evaluated 90 patients with highrisk non-invasive bladder cancer who ultimately
underwent cystectomy. These investigators reported
improved 15-year disease-specific survival for those
who underwent cystectomy within 2 years after an
initial BCG treatment (Level 3). Moreover, those who
underwent a cystectomy for recurrent non-invasive
disease had an improved outcome compared
to those who underwent surgery for progressive
disease. The investigators thus concluded that
deferring cystectomy until progression to muscleinvasive disease may decrease a patients overall
disease-specific survival. Nevertheless, 217 patients
from their original cohort of 307 with high-risk noninvasive disease never required cystectomy and
were thus spared the morbidity associated with
cystectomy.

1. T
 he assessment of T1 urothelial carcinoma
should be based on tumor grade, early
recurrence, multiplicity, tumor size, concomitant
CIS, urothelial carcinoma involving the prostatic
mucosa or ducts, and depth of lamina propria
invasion. (LOE1a, grade A).
2. T
 he use of 1 immediate postoperative instillation
of chemotherapy is not supported by consistent
data and therefore it should not be recommended
for standard practice. (LOE1a, grade A)
3. H
 igh risk patients and those with recurrent or
persisting disease after BCG should be offered
a cystectomy. (LOE2a, grade A)
4. If a bladder-sparing approach is desired, a
secondary TURBT should be performed and
followed with intravesical BCG therapy. (LOE3,
grade B)

References
1. Babjuk M, Oosterlinck W, Sylvester R, Kaasinen E, Bhle
A, Palou J, Rouprt M. Guidelines on TaT1 (non-muscle
invasive) Bladder Cancer. EAU, 2010.
2. van Rhijn BW, Burger M, Lotan Y, Solsona E, Stief CG,
Sylvester RJ, Witjes JA, Zlotta AR. Recurrence and
progression of disease in non-muscle-invasive bladder
cancer: from epidemiology to treatment strategy. Eur Urol.
2009 Sep;56(3):430-42. Epub 2009 Jun 26

The presence of metastasis has the most profound


impact on survival. A large series of 1054 patients
who had undergone radical cystectomy was reviewed
retrospectively by Stein et al [44]. The investigators
identified 401 patients who had pathologically nonmuscle-invasive tumors (pT0, pTa, pT1, or pTis).
These patients demonstrated improved 5-year
recurrence-free survival compared with those with
non-organ-confined tumors (pT3b, pT4, and those
with positive lymph nodes). No survival differences
were observed when comparing non-invasive (pTa,
pTis), lamina propriainvasive (pT1), and muscleinvasive tumors (pT2, pT3a), providing that there
was no evidence of metastatic disease to the lymph
nodes. The overall recurrence-free survival rates in
those with organ-confined, node-negative disease

3. Saint F, Le Frere Belda MA, Quintela R, Hoznek A, Patard


JJ, Bellot J, Popov Z, Zafrani ES, Abbou CC, Chopin DK,
de Medina SG. Pretreatment p53 nuclear overexpression
as a prognostic marker in superficial bladder cancer treated
with Bacillus Calmette-Gurin (BCG). Eur Urol. 2004
Apr;45(4):475-82.
4. Moonen PM, van Balken Ory B, Kiemeney LA, Schalken
JA, Witjes JA. Prognostic Value of p53 for High Risk
Superficial Bladder Cancer With Long-Term Followup. J
Urol 2007;177:80-3
5. Lopez-Beltran A, Luque RJ, Alvarez-Kindelan J, Quintero
A, Merlo F, Requena MJ, Montironi R. Prognostic factors in
survival of patients with stage Ta and T1 bladder urothelial
tumours: the role of G1-S modulators (p53, p21Waf1,
p27Kip1, cyclin D1, and cyclin D3), proliferation index,
and clinicopathologic parameters. Am J Clin Pathol. 2004
Sep;122(3):444-52.
6. Shariat SF, Bolenz C, Godoy G, Fradet Y, et al: Predictive

263

value of combined immunohistochemical markers in patients


with pT1 urothelial carcinoma at radical cystectomy. J Urol
2009, 182(1): 78-84 (Shariat et al#3)

with clinical T1 grade 3 urothelial carcinoma treated with


radical cystectomy: results from an international cohort. Eur
Urol 2010;57:300-9.

7. Schwaibold HE, Sivalingam S, May R, Hartung R: , The


value of a second transurethral resection for T1 bladder
cancer BJU Int. 2006, 97 (6): 1199-1201

22. Sylvester RJ, van der Meijden AP, Oosterlinck W, Witjes


A, Bouffioux C, Denis L, Newling DW, Kurth K. Predicting
recurrence and progression in individual patients with stage
TaT1 bladder cancer using EORTC risk tables: a combined
analysis of 2596 patients from seven EORTC trials. Eur
Urol 2006;49(3):466-77.

8. Divrik T, Yildirim , Erolu A, Zorlu F, zen H: Is a Second


Transurethral Resection Necessary for Newly Diagnosed
pT1 Bladder Cancer? J Urol 2006, 175 (4): 1258-1261

23. Hautmann RE, Gschwend JE, de Petriconi RC, Kron M,


Volkmer BG. Cystectomy for transitional cell carcinoma
of the bladder: results of a surgery only series in the
neobladder era. J Urol 2006;176:486-92.

9. Denzinger S, Wieland WF, Otto W, Filbeck T, Knuechel


R, Burger M:Does photodynamic transurethral resection
of bladder tumour improve the outcome of initial T1 highgrade bladder cancer? A long-term follow-up f a randomized
study. BJU Int 101:566-569, 2008.

24. Madersbacher S, Hochreiter W, Burkhard F, Thalmann GN,


Danuser H, Markwalder R, Studer UE. Radical cystectomy
for bladder cancer todaya homogeneous series without
neoadjuvant therapy. J Clin Oncol 2003;21:690-6.

10. Dutta SC, Smith Jr. JA, Shappell SB, Coffey DS, Chang
SS, Cookson MS: Clinical under staging of high risk
nonmuscle invasive urothelial carcinoma treated with
radical cystectomy. J Urol 2001, 166 (2): 490-493

25. Stein JP, Lieskovsky G, Cote R, Groshen S, Feng AC,


Boyd S, Skinner E, Bochner B, Thangathurai D, Mikhail
M, Raghavan D, Skinner DG. Radical cystectomy in the
treatment of invasive bladder cancer: long-term results in
1,054 patients. J Clin Oncol 2001;19:666-75.

11. Dalbagni G, Vora K, Kaag M, Cronin A , Bochner B, Donat


SM, Herr HW. Clinical outcome in a contemporary series of
restaged patients with clinical T1 bladder cancer: Eur Urol
2009; 56(6): 903-910

26. Shelley MD, Kynaston H, Court J, Wilt TJ, Coles B,


Burgon K, Mason MD. A systematic review of intravesical
bacillus Calmette-Gurin plus transurethral resection vs
transurethral resection alone in Ta and T1 bladder cancer.
BJU Int 2001;88(3):209-16.

12. Herr HW, Donat SM, Dalbagni G: Can Restaging


Transurethral Resection of T1 Bladder Cancer Select
Patients for Immediate Cystectomy? J Urol 2006, 177 (1):
75-79
13. Babjuk M, Oosterlinck W, Sylvester R, Kaasinen E, Bhle
A, Palou J, Rouprt M. Guidelines on TaT1 (non-muscle
invasive) Bladder Cancer. EAU, 2010.

27. Han RF, Pan JG. Can intravesical bacillus Calmette-Gurin


reduce recurrence in patients with superficial bladder
cancer? A meta-analysis of randomized trials. Urology
2006;67(6):1216-23.

14. Bohle A, Leyh H, Frei C et al. Single postoperative


instillation of Gemcitabine in Patients with Non-Muscleinvasive Transitional Cell Carcinoma of the bladder: A
Randomised, Double-blind, Placebo-controlled Phase III
Multicentre Study

28. Shelley MD, Wilt TJ, Court J, Coles B, Kynaston H, Mason


MD. Intravesical bacillus Calmette-Gurin is superior to
mitomycin C in reducing tumour recurrence in high-risk
superficial bladder cancer: a metaanalysis of randomized
trials. BJU Int 2004;93(4):485-90

15. Gudjonsson S, Adell L, Mersada F et al: Should all patients


with non muscle invasive bladder cancer receive early
intravesical chemotherapy after transurethral resection ?
The results of a prospective randomised multicentre study.
Eur Urol 2009: 55: 773

29. Bhle A, Jocham D, Bock PR. Intravesical bacillus CalmetteGuerin versus mitomycin C for superficial bladder cancer: a
formal meta-analysis of comparative studies on recurrence
and toxicity. J Urol 2003;169(1):90-5.

16. Berrum-Svennung I, Granfors T, Jahnson S, Boman


H, Holmng S. A single instillation of epirubicin after
transurethral resection of bladder tumours prevents
only small recurrences. J Urol. 2008 Jan;179(1):101-5;
discussion 105-6.

30. Malmstrm P-U, Sylvester R, Crawford DE et al. An


Individual Patient Data Meta-Analysis of the Long-Term
Outcome of Randomised Studies Comparing Intravesical
Mitomycin C versus Bacillus Calmette-Guerin for Non
Muscle-Invasive Bladder Cancer. Eur Urol 2009;56:24756.

17. Brausi M: Challenging the EAU guidelines on non muscle


invasive bladder cancer (NMIBC): single immediate
instillation of chemotherapy after transurethral resection
of NMIBC and chemotherapy versus BCG in treatment of
intermediate-risk tumours. Eur Urol. Suppl., 2010; 9:406

31. Sylvester RJ, van der Meijden AP, Lamm DL. Intravesical
bacillus Calmette-Guerin reduces the risk of progression in
patients with superficial bladder cancer: a meta-analysis of
the published results of randomized clinical trials. J Urol
2002;168(5):1964-70.

18. Zincke H, Utz DC, Taylor WF, Myers RP et al: Influence of


thiotepa and doxorubicin instillation at time of transurethral
surgical treatment of bladder cancer on tumour recurrence:
a prospective, randomised, double-blind, controlled trial . J
Urol; 129:505

32. Herr HW and Sogani PC. Does early cystectomy improve


the survival of patients with high risk superficial bladder
tumours? J Urol, 2001;166:1296-9.
33. Shahin O, Thalmann GN, Rentsch C, Mazzucchelli L,
Studer UE. A retrospective analysis of 153 patients treated
with or without intravesical bacillus Calmette-Guerin for
primary stage T1 grade 3 bladder cancer: recurrence,
progression and survival. J Urol, 2003;169:96-100.

19. Cai T, Nesi G, Tinacci G et al: Can early single dose


instillation of epirubicin improve bacillus Calmette-Guerin
efficacy in patients with non muscle invasive high risk
bladder cancer ? Results from a prospective, randomised,
double-blind controlled study. J Urol 2008; 180:110

34. Lamm DL, Blumenstein BA, Crissman JD, Montie JE,


Gottesman JE, Lowe BA, Sarosdy MF, Bohl RD, Grossman
HB, Beck TM, Leimert JT, Crawford ED. Maintenance
bacillus Calmette-Guerin immunotherapy for recurrent TA,
T1 and carcinoma in situ transitional cell carcinoma of the
bladder: a randomized Southwest Oncology Group Study.
J Urol, 2000;163:1124-9.

20. Babjuk M. New insights in intravesical treatment for


intermediate- and high-risk non-muscle-invasive urothelial
bladder carcinoma. Eur Urol. 2010 May;57(5):774-6;
21. Fritsche HM, Burger M, Svatek RS, Jeldres C, Karakiewicz
PI, Novara G, Skinner E, Denzinger S, Fradet Y, Isbarn
H, Bastian PJ, Volkmer BG, Montorsi F, Kassouf W, Tilki
D, Otto W, Capitanio U, Izawa JI, Ficarra V, Lerner S,
Sagalowsky AI, Schoenberg M, Kamat A, Dinney CP, Lotan
Y, Shariat SF. Characteristics and outcomes of patients

35. van der Meijden AP, Sylvester RJ, Oosterlinck W, Hoeltl


W, Bono AV; EORTC Genito-Urinary Tract Cancer Group.

264

Maintenance bacillus Calmette-Guerin for Ta, T1 bladder


tumours is not associated with increased toxicity: results
from a European Organisation for Research and Treatment
of Cancer Genito-Urinary Group Phase III Trial. Eur Urol
2003;44(4):429-34.

IV. Treatment Options for BCG


Failure

36. Bhle A, Bock PR. Intravesical bacillus CalmetteGuerin versus mitomycin C in superficial bladder cancer:
formal meta-analysis of comparative studies on tumour
progression. Urology 2004;63(4):682-7

1. Introduction
Although BCG is an effective therapy for urothelial
carcinoma, BCG failure remains a significant
problem. Greater patient age, previous intravesical
therapy, and failure to achieve a complete response
after induction BCG are associated with an increased
risk of progression or death for patients with nonmuscle-invasive bladder cancer [1]. Studies using
more intensive BCG treatment and extended BCG
maintenance schedules tend to show better results.
The median time to progression generally exceeds
12 months, with an estimated progression rate
of 5% by 6 months. For these patients, radical
cystectomy continues to be the gold standard.
However, patients are sometimes reluctant to
undergo major surgery for a condition that does not
pose an immediate threat to their lives. Furthermore,
radical cystectomy is not a suitable treatment option
for a subset of patients with severe comorbidities. A
number of alternatives have thus been developed.

37. Lamm DL: Efficacy and safety of bacille Calmette-Guerin


immunotherapy in superficial bladder cancer. Clin Infect
Dis.31 Suppl 3:S86-90, 2000.
38. Denzinger S, Fritsche H-M, Otto W, Blana A,Wieland W-F,
Burger M: Early Versus Deferred Cystectomy for Initial
High-Risk pT1G3 Urothelial Carcinoma of the Bladder:
Do Risk Factors Define Feasibility of Bladder-Sparing
Approach? Eur Urol 2008;53:146-52
39. Bouffioux C, Kurth KH, Bono A, et al. Intravesical adjuvant
chemotherapy for superficial transitional cell bladder
carcinoma: results of 2 European Organization for Research
and Treatment of Cancer randomized trials with mitomycin
C and doxorubicin comparing early versus delayed
instillations and short-term versus long-term treatment.
European Organization for Research and Treatment of
Cancer Genitourinary Group. J Urol. Mar 1995;153(3 Pt
2):934-941.
40. Solsona E, Iborra I, Dumont R, Rubio-Briones J, Casanova
J and Almenar S. The 3-month clinical response to
intravesical therapy as a predictive factor for progression
in patients with high risk superficial bladder cancer. J. Urol.
164 (2000), 685689.

2. Defining BCG Failure

41. Herr HW, Dalbagni G. Defining bacillus CalmetteGuerin refractory superficial bladder tumours. J Urol
2003;169(5):1706-8

In evaluating salvage therapies for use after BCG


failure, Herr and Dalbagni noted that comparisons
between therapies have been hampered by the lack
of standard definitions for BCG failure and BCGrefractory urothelial carcinoma [2]. Some series have
defined BCG failure after a single induction course
of BCG, others after 2 courses [3,4,5]. In addition,
result reporting methods have been inconsistent [6].
Most studies have included all patients who received
1 or more courses of BCG [7,8,9]. Investigators have
often combined patients with persistent disease
(nonresponders) and patients with recurrent disease
after an initial response. Some studies combined
patients who were BCG nonresponders and patients
who could not complete BCG therapy due to the
associated toxicity (BCG intolerant). Furthermore,
many studies have combined all patients with
papillary tumors both with and without CIS. Finally,
most studies did not indicate the disease-free interval
after the last BCG course. These inconsistencies
have made it more difficult to compare outcomes in
this very heterogeneous population.

42. Schrier BPH, Hollander MP, van Rhijn BWG, Ijzer


S, Kiemeney LALM, Witjes JA. Prognosis of muscle
invasive bladder cancer: difference between primary and
progressive tumours: possible implications for therapy. Eur
Urol 2004;45:292-6
43. Amling CL, Thrasher JB, Frazier HA, Dodge RK, Robertson
JE and Paulson DF: Radical cystectomy for stages Ta, Tis
and T1 transitional cell carcinoma of the bladder. J Urol,
1994;151:31-5.
44. Stein JP, Lieskovsky G, Cote R, Groshen S, Feng AC, Boyd
S, Skinner E, Bochner B, Thangathurai D, Mikhail M et al.:
Radical cystectomy in the treatment of invasive bladder
cancer: longterm results in 1,054 patients. J Clin Oncol,
2001;19: 666-75.
45. Ghoneim MA, el-Mekresh MM, el-Baz MA, el-Attar IA and
Ashamallah A: Radical cystectomy for carcinoma of the
bladder: critical evaluation of the results in 1,026 cases. J
Urol, 1997;158:393-9.
46. Malavaud B. T1G3 bladder tumours: the case for radical
cystectomy. Eur Urol 2004;45:406-10.
47. Bianco FJ Jr, Justa D, Grignon DJ, Sakr WA, Pontes
JE and Wood DP Jr. Management of clinical T1 bladder
transitional cell carcinoma by radical cystectomy. Urol
Oncol 2004;22:290-4.

In the most general sense, any recurrent disease


after initiation of BCG therapy can be referred to as
BCG failure. However, to provide more uniformity
in reporting, the following alternative descriptive
terms for specific types of BCG failure should be
used whenever possible:

48. Cookson MS, Herr HW, Zhang ZF, Soloway S, Sogani PC,
Fair WR. The treated natural history of high risk superficial
bladder cancer: 15-year outcome. J Urol, 1997;158:62-7.
49. Hendricksen K, Witjes W, Idema J, Kums J, van Vierssen
Trip O,de Bruin M, Vergunst H, Caris C, Janzing-Pastors
M, Witjes J. Comparison of Three Schedules of Intravesical
Epirubicin in Patients with NonMuscle-Invasive Bladder
Cancer . Eur Urol 2008;53:984-91

BCG-refractory disease is the term used when a


disease-free state is not achieved within 6 months
of the initial BCG therapy with either maintenance

265

or retreatment at 3 months due to either persistent


or rapidly recurrent disease. The definition also
includes any progression in stage, grade, or disease
extent within 3 months of the first BCG cycle (i.e.,
non-improving or worsening disease, despite BCG).
BCG-resistant disease is the term used when there
is recurrence or persistence of disease 3 months
after the induction cycle. If it is of lesser degree,
stage, or grade, and is no longer present at 6
months following BCG retreatment with or without
TURBT, the disease has improved and is resolved
with further BCG. BCG-relapsing disease is the
term used when there is a recurrence of disease
after disease-free status has been achieved within
6 months (i.e., disease resolved after BCG but
subsequently returns). Relapse is further defined
according to the time of recurrence as either early
(within 12 months), intermediate (12 to 24 months),
or late (> 24 months). There is some overlap with
BCG-refractory disease, as disease relapse while
on active maintenance, i.e., within 3 months, may
qualify as BCG-refractory disease. BCG-intolerant
disease is the term used when disease recurs after
a less-than-adequate course of therapy is applied
because of a serious adverse event or symptomatic
intolerance that mandates discontinuation of further
BCG therapy (i.e., recurrent disease in the setting
of inadequate BCG treatment because of drug
toxicity).

patients with any grade T1 disease; another 70%


(19 of 27) responded to further TURBT and BCG,
including 64% (7 of 11) with recurrent T1 disease.
Similar results were reported by Brake et al [12] who
found a 70% (89 of 128) enduring response after 1
BCG cycle and 51% (19 of 37) after a second BCG
cycle for the remaining 37 patients (13 of whom
had already progressed). Pansadoro and De Paula
[13] reported that the response rate of 47 patients
to TUR plus 1 cycle of 6 weekly BCG instillations
was 53% (25 of 47). Of the 22 failures treated with
a repeat TUR and a second cycle of BCG, 27% (6
of 22) responded, and, of the remaining 16 patients
who received a third cycle of BCG, only 6% (1 of
16) responded. These results are similar to those
reported in BCG studies in nonT1-restricted tumors
and illustrate that a second (but not greater) course
of BCG may be appropriate in select patients with
original stage T1 disease who have a recurrence of
non-BCG-refractory disease. Unfortunately, there
are insufficient data to assess the effectiveness of
repeated BCG treatments in refractory patients or
those with recurrent T1 disease [14].

c) Intravesical salvage chemotherapy


Of the various standard intravesical chemotherapeutic
agents (Thiotepa, doxorubicin, MMC) there is
only minimal reported experience with their use
in patients failing prior BCG therapy. Malmstrom
et al [15] reported a 19% 3-year disease-free rate
among intermediate and high risk patients treated
with MMC who had failed a prior first induction
cycle of BCG. The results for T1 recurrences are
unknown. Similarly, there are few data on the newer
anthracycline derivative valrubicin. Of 90 valrubicintreated patients with CIS with and without papillary
urothelial carcinoma who had failed 2 or more cycles
of prior intravesical therapy (most commonly BCG),
only 21% had a complete response at 6 months and
8% by 24 months [16]. Notably, all 5 patients with
stage T1 disease (previously resected) plus CIS
failed to achieve a complete response. Gemcitabine
has been suggested as a potential alternative for
patients intolerant to or otherwise unable to receive
BCG due to its favorable toxicity profile. However,
gemcitabine has been shown to have an almost
two-fold recurrence rate at 44 months (53% versus
28%) when compared to BCG in a randomized
prospective study [17]. Two other studies, however,
suggest that gemcitabine could be used in BCG
failures. Addeo et al treated 120 patients, of which
109 were assessable for response [18]. Ninetyone of these 109 patients failed previous BCG, but
details on these failures were not provided. Patients
were randomized between gemcitabine and MMC,
with a 3-year recurrence-free survival of 72% and
61%, respectively. Progression was noted in 6 of
54 patients and 10 of 55 patients, respectively. Di
Lorenzo et al [19] compared 1 year of gemcitabine
to BCG Connaught. The only difference in the

Prophylactic use of ofloxacin has been reported to


decrease incidence of moderate to severe adverse
events associated with BCG intravesical therapy, but
this strategy has not been widely adopted. [10]

3. Treatment Options
a) Cystectomy
Great variation exists in actual practice with regard
to the timing of cystectomy in patients who fail
induction BCG therapy. These vary from immediate
surgery after the first follow-up cystoscopy if residual
disease is present to delaying surgery for 6 months if
a complete response is not achieved. The evidence
for this interval is not well-established at the present
time. There is general consensus that the prognosis
for some patients with aggressive disease can be
adversely affected by delaying surgery, but the
prediction of this risk in individual cases is difficult.
As such, alternative bladder-sparing strategies are
frequently used (see below).

b) Repeat BCG Treatments


This treatment may be appropriate for both BCGresistant and BCG-relapsing disease. However,
the success of a second course of BCG for stage
T1 disease has not been extensively reported,
and only a few published studies have addressed
this issue. Cookson and Sarosdy [11] reported an
initial 69% (59 of 86) complete response to BCG in

266

treatment schedule was that the 6 week induction


course was biweekly for gemcitabine in comparison
to weekly for BCG. All patients had high risk nonmuscle-invasive bladder cancer, having failed
1 course of BCG. The 2-year recurrence-free
survival was 19% for gemcitabine versus 3% for
BCG (p<0.008). Progression rates were 33% and
37.5%, respectively. In all, gemcitabine seems to
have some effect in patients who have failed BCG;
nevertheless, many patients experience recurrence
and progression, especially in the high risk group.

for BCG failures has not yet been established [29].


Furthermore, for T1 disease alone, local recurrence
or progression occurs in approximately 50% of
cases [30]. Thermo-chemotherapy has recently
been reported on in patients with recurrent disease
following BCG [31]. One hundred and eleven patients
(29 pT1) were treated with 6 weekly induction
treatments, and thereafter 4 to 6 sessions with a 6
week interval. The disease-free survival estimates
were 85% and 56% after 1 and 2 years, respectively.
Only 3 patients experienced progression. There
was no difference between the different BCG failing
groups (definitions used as outlined above), nor,
for example, for Ta versus T1. Only maintenance
hyperthermia treatment was associated with
increased efficacy. In all, hyperthermia treatment
would appear to be a possibility for these patients,
although more studies are needed.

Similarly, prospective evaluation has found


epirubicin with interferon alpha to be inferior to BCG
for prophylaxis of recurrence and therefore has
a limited role in patients experiencing BCG failure
[20]. Given these poor results, it appears that current
intravesical salvage chemotherapy has little to offer
patients who have failed BCG, especially those with
stage T1 disease.

Recommendations for BCG Failure

d) Interferon alpha Immunotherapy

1. T
 he threat of progression remains real but
comfortably low enough within the first 6 months
of initiating BCG to consider alternatives to
cystectomy for those patients unfit or not willing
to undergo this standard management option
(LOE 2, Grade B).

In 2 studies, findings suggest that interferon alpha


is unlikely to be of benefit in stage T1 BCG failures.
The long-term (>2 years) success rate of interferon
alpha monotherapy of patients who fail BCG (CIS
and/or papillary urothelial carcinoma) is generally
15% [21]. Furthermore, in a study of interferon
alpha monotherapy for primary stage T1 disease,
interferon alfa was found to be no better than placebo
at 43 month follow-up [22].

2. In the case of any BCG non-response or failure,


cystectomy is recommended (LOE 2a, Grade
A).
3. T
 he current best option for alternative treatment
includes repeat resection and repeat BCG (LOE
1a, Grade A), possibly with interferon alpha as a
costimulant (LOE 2, grade C). Gemcitabine and
thermo-chemotherapy have shown efficacy,
but more studies are needed. There is no
reported evidence of significant efficacy using
current intravesical chemotherapy, interferon
alpha monotherapy, photodynamic therapy, or
radiation therapy.

e) Combination BCG plus Interferon alpha


Several
single
institutional
studies
have
demonstrated that the combination of low-dose BCG
plus interferon alpha may be useful as a salvage
regimen for BCG failures [23,24,25,26]. With followup times ranging from 12-30 months, disease-free
rates were 50-60%, even in patients with recurrent T1
disease. Furthermore, no patient with an expedient
cystectomy after BCG plus interferon alpha failure
had unresectable or metastatic disease. Interim
results of an even larger group of 231 patients with
BCG failure in a multi-institutional study showed a
disease-free rate of 42% at a median follow-up time
of 24 months [27]. The efficacy results for stage T1
patients have not yet been published, but preliminary
analysis reveals a similar degree of durable response
among an approximate 10% progression rate (M.
ODonnell, personal communication).

References
1. Lerner SP, Tangen CM, Sucharew H, Wood D, Crawford
ED: Failure to achieve a complete response to induction
BCG therapy is associated with increased risk of disease
worsening and death in patients with high risk non-muscle
invasive bladder cancer. Urol Oncol, 2009; 27(2): 155-9
2. Herr HW, Dalbagni G. Defining bacillus Calmette-Guerin
refractory superficial bladder tumors. J Urol 2003;
169:1706-8.

f) Other Alternatives

3. Klein EA, Rogatko A, Herr HW. Management of local


bacillus Calmette-Guerin failures in superficial bladder
cancer. J Urol 1992;147:601-5.

There are no reliable data published on the use of


photodynamic therapy for recurrent stage T1 cancer,
although it is generally believed to be more effective
with surface disease, such as CIS [28]. Likewise,
although 5-year disease-free survival rates of 5060% for radiation therapy for stage T1 disease
have been reported, its role as a salvage therapy

4. Glashan RW. A randomized controlled study of intravesical


alpha- 2b-interferon in carcinoma in situ of the bladder. J
Urol 1990; 144:658-61.
5. Sarosdy MF, Manyak MJ, Sagalowsky AI, et al. Oral
bropirimine immunotherapy of bladder carcinoma in situ
after prior intravesical bacille Calmette-Guerin. Urology
1998;51:226-31.

267

6. Martin FM and Kamat AM: Definition and management of


patients with bladder cancer who fail BCG therapy. Expert
Rev Anticancer Ther, 2009; 9(6): 815-20

2b in the intravesical treatment of patients with stage T1


urinary bladder cancer. A prospective randomized, Nordic
study. European Urol, 2010; 57: 25-31

7. Dalbagni G, Russo P, Sheinfeld J, et al. Phase I trial of


intravesical gemcitabine in bacillus Calmette-Guerinrefractory transitional- cell carcinoma of the bladder. J Clin
Oncology 2002; 20:3193-8.

21. ODonnell MA, Lilli K, Leopold C and the National Bacillus


Calmette-Guerin/Interferon phase 2 Investigator Group.
Interim results from a national multicenter phase II trial
of combination bacillus Calmette-Guerin plus interferon
alpha-2b for superficial bladder cancer. J Urol 172, 888 893, 2004.

8. ODonnell MA, Krohn J, DeWolf WC. Salvage intravesical


therapy with interferon-alpha 2b plus low dose bacillus
Calmette-Guerin is effective in patients with superficial
bladder cancer in whom bacillus Calmette-Guerin alone
previously failed. J Urol 2001; 166:1300-4.

22. Williams RD, Gleason DM, Smith AY, Zinner NR, Sagalowsky
AI, Montie JE, Brosman SA, Marks LS, Brito G, Boxer RJ,
Blank BH, Neri R and Rudeen J. Pilot study of intravesical
alpha-2b interferon for treatment of bladder carcinoma in
situ following BCG failure. J Urol 155 (supplement), 494A,
abstract 735, 1996.

9. Luciani LG, Neulander E, Murphy WM, et al. Risk of


continued intravesical therapy and delayed cystectomy in
BCG-refractory superficial bladder cancer: an investigational
approach. Urology 2001;58:376-9.

23. ODonnell MA, Krohn J, DeWolf WC. Salvage intravesical


therapy with interferon-alpha 2b plus low dose bacillus
Calmette-Guerin is effective in patients with superficial
bladder cancer in whom bacillus Calmette-Guerin alone
previously failed. J Urol 2001; 166:1300-4.

10. Colombel M, Saint F, Chopin D, Malavaud B, Nicolas L,


Rischmann P and the ITB01 Study Group: The effect of
ofloxacin on Bacillus Calmette-Guerin induced toxicity
in patients with superficial bladder cancer: results of a
randomized, prospective, double-blind, placebo controlled,
multicenter study. J Urol, 2006; 176: 935-9

24. Luciani LG, Neulander E, Murphy WM, et al. Risk of


continued intravesical therapy and delayed cystectomy in
BCG-refractory superficial bladder cancer: an investigational
approach. Urology 2001;58:376-9.

11. Cookson MS, Sorosdy MF. Management of stage T1


superficial bladder cancer with intravesical bacillus
Calmette-Guerin therapy. J Urol 1992; 148:797-801.

25. Punnen SP, Chin JL, Jewett MA. Management of bacillus


Calmette-Guerin (BCG) refractory superficial bladder
cancer: results with intravesical BCG and Interferon
combination therapy. Can J Urol 2003; 10:1790-5.

12. Brake M, Loertzer H, Horsch R, Keller H. Long-term results


of intravesical bacillus Calmette-Gurin therapy for stage T1
superficial bladder cancer. Urology. 2000 May;55(5):673-8

26. Lam JS, Benson MC, ODonnell MA, et al. Bacillus


Calmette-Guerin plus interferon-alpha2B intravesical
therapy maintains an extended treatment plan for
superficial bladder cancer with minimal toxicity. Urol Oncol
2003; 21:354-60.

13. Pansadoro V, Emiliozzi P, De Paula F, et al. Long-term


follow-up of G3T1 transitional cell carcinoma of the bladder
treated with intravesical bacille Calmette-Guerin: 18-year
experience. Urology 2002; 59:227-31.
14. Sylvester RJ, van der Meijden AP, Lamm DL: Intravesical
bacillus Calmette-Guerin reduces the risk of progression in
patients with superficial bladder cancer: a meta-analysis of
the published results of randomized clinical trials. J Urol.
168(5):1964-70, 2002.

27. ODonnell MA, Lilli K, Leopold C. Interim results from a


national multicenter phase II trial of combination bacillus
Calmette-Guerin plus interferon alfa-2B for superficial
bladder cancer. J Urol 2004; 176:888-92.

15. Malmstrm PU, Sylvester RJ, Crawford DE, Friedrich


M, Krege S, Rintala E, Solsona E, Di Stasi SM, Witjes
JA. An individual patient data meta-analysis of the
long-term outcome of randomised studies comparing
intravesical mitomycin C versus bacillus Calmette-Gurin
for non-muscle-invasive bladder cancer. Eur Urol. 2009
Aug;56(2):247-56.

28. Berger AP, Steiner H, Stenzl A, et al. Photodynamic therapy


with intravesical instillation of 5-aminolevulinic acid for
patients with recurrent superficial bladder cancer: a single
center study. Urology 2003; 61:338-41.

16. Steinberg G, Bahnson R, Brosman S, et al. Efficacy and


safety of valrubicin for the treatment of bacillus CalmetteGuerin refractory carcinoma in situ of the bladder. J Urol
2000; 163:761-7.

30. Rodel C, Dunst J, Grabenbauer GG, et al. Radiotherapy


is an effective treatment for high-risk T1-bladder cancer.
Strahlenther Onkol 2001; 177:82-8.

29. Dunst J, Sauer R, Schrott KM, et al. Organ-sparing treatment


of advanced bladder cancer: a 10-year experience. Int J
Radiat Oncol Biol Phys. 1994; 30:261-6.

31. Nativ O, Witjes JA, Hendricksen K, Cohen M, Kedar


D, Sidi A, Colombo R, Leibovitch I: Combined thermochemotherapy for recurrent bladder cancer after bacillus
Calmette-Guerin. J Urol. 2009 Oct;182(4):1313-7.

17. Porena M, Del Zingaro M, Lazzeri M, Mearini L, Giannantoni


A, Bini V, Costantini E: Bacillus Calmette-Guerin versus
gemcitabine for intravesical therapy in high-risk superficial
bladder cancer: a randomised prospective study. Urol Int,
2010; 84: 23-27
18. Addeo R, Caraglia M, Bellini S, Abbruzzese A, Vincenzi B,
Montella L, Miragliuolo A, Guarrasi R, Lanna M, Cennamo
G, Faiola V, Del Prete S: Randomized Phase III Trial on
Gemcitabine Versus Mitomycin in Recurrent Superficial
Bladder Cancer. Evaluation of Efficacy and Tolerance. JCO
Feb 1, 2010:543-548
19. Di Lorenzo G, Perdon S, Damiano R, Faiella A, Cantiello
F, Pignata S, Ascierto P, Simeone E, De Sio M, Autorino
R. Gemcitabine versus bacille Calmette-Gurin after initial
bacille Calmette-Gurin failure in non-muscle-invasive
bladder cancer: a multicenter prospective randomized trial.
Cancer 2010 Apr 15;116(8):1893-900
20. Duchek M, Johansson R, Jahnson S, Mestad O, Hellstrm
P, Hellsten S, and Malmstrm P: Bacillus Calmette-Guerin
is superior to a combination of epirubicin and interferon-

268

Committee 6

Muscle-invasive,
Presumably Regional, Tumor
Chairs:
Jason Efstathiou,
David I. Quinn,
Arnulf Stenzl
Members:
Joaquim Bellmunt,
Michael S. Cookson,
Georgios Gakis,
Khurshid A. Guru,
Axel Heidenreich,
Patrick Keegan,
Seth P. Lerner,
Edward M. Messing,
Arthur I. Sagalowsky,
Mark P. Schoenberg,
William U. Shipley,
Arlene O. Siefker-Radtke,
Cora Sternberg

269

Table of Contents
I. Overview

V. Perioperative Radiotherapy

II. Indications and Algorithm


of Treatment

VI. Bladder-sparing
Treatments for Localized
Disease

III. Radical Surgery

1. Transurethral Monotherapy
2. Partial Cystectomy

1. Removal of the Tumor-bearing


Bladder

3. Radiation-based Bladderpreserving Strategies

2. Surgical Outcome: Morbidity and


Mortality

VII. Treatment of Mixed


Histology Urothelial
Carcinoma

3.Oncological Outcome of Radical


Surgery According to TNM Staging

IV. Perioperative
Chemotherapy

VIII. Follow-up

1. Neoadjuvant Chemotherapy
1. Site of Recurrence

2. Adjuvant Chemotherapy

270

Muscle-invasive,
Presumably Regional, Tumor
Jason Efstathiou, David I. Quinn, Arnulf Stenzl
Joaquim Bellmunt, Michael S. Cookson, Georgios Gakis,
Khurshid A. Guru, Axel Heidenreich, Patrick Keegan,
Seth P. Lerner, Edward M. Messing, Arthur I. Sagalowsky,
Mark P. Schoenberg, William U. Shipley,
Arlene O. Siefker-Radtke, Cora Sternberg
Summary of International Consultation on
Urological Diseases (ICUD) Modified Oxford Center
for Evidence-Based

I. Overview
A panel of an international multidisciplinary group
consisting of urologists, oncologists and radio-oncologists provide data on evidence-based management
of patients with muscle-invasive, clinically presumably node-negative, bladder cancer. The guidelines
focus on treatment options and outcome, including
radical surgery, neoadjuvant and adjuvant treatment
modalities, bladder-sparing approaches, and treatment of mixed and secondary urothelial recurrences.
In addition, specific recommendations are given for
the surgical treatment and extent as well as for follow-up strategies after curative intent. The level of
evidence and grade of recommendation given are
as outlined by the Oxford Centre for Evidence-based
Medicine [1].

Medicine
grading
recommendations

system

for

guideline

Levels of evidence
Level 1 Meta-analysis of RCTs or a good-quality
RCT
Level 2 Low-quality RCT or meta-analysis of goodquality prospective cohort studies
Level 3 Good-quality retrospective case-control
studies or case series
Level 4 Expert opinion based on first principles
or bench research, not on evidence
Grades of recommendation

References

Grade A Usually consistent level 1 evidence


Grade B Consistent level 2 or 3 evidence or
majority evidence from RCTs

1. Modified from Oxford Centre for Evidence-based Medicine


Levels of Evidence (May 2001). Produced by Bob Phillips,
Chris Ball, Dave Sackett, Doug Badenoch, Sharon Straus,
Brian Haynes, Martin Dawes since November 1998.http://
www.cebm.net/index.aspx?o=1025 [access date February
2009].

Grade C Level 4 evidence, majority evidence


from level 2 or 3 studies, expert opinion
Grade D No recommendation possible because of
inadequate or conflicting evidence

II. Indications and Algorithms


of Treatment

RCT _ randomized controlled trial.


Adapted with permission from Incontinence.[1]

Radical
cystectomy
and
bilateral
pelvic
lymphadenectomy provide excellent loco-regional
control for patients with clinical stage T2-T4a, N0NX, M0 disease. Local pelvic recurrence rates are
as low as 4% in patients with node negative disease
and 15-20% for patients with node metastases.
[1,2] Cystectomy may also be appropriate initial

treatment for some patients with first occurrence of


high risk non-muscle-invasive cancer or recurrence.
Patients with high grade papillary disease (Ta, T1)
or carcinoma in situ that recurs following treatment
with intravesical BCG or patients with intermediate

271

risk papillary disease that cannot be controlled with


transurethral resection and intravesical therapy alone
should be offered cystectomy. Surgery should not be
withheld or delayed in these patients as the longterm survival probability is decreased when there is
evidence of progression to muscle-invasive cancer.[3]
Morris et al studied how physician behavior affected
outcomes in bladder cancer management and
determined that the persistent use of conservative
therapy delays necessary, aggressive treatment.[4]
In this setting, changing physician decision-making
and/or practice patterns may decrease mortality.
Salvage cystectomy is indicated for non-responders
to conservative therapy, recurrences after bladdersparing treatments, non-urothelial carcinomas (these
tumors may respond poorly to chemotherapy and
radiotherapy), and as a purely palliative intervention
in the setting of fistula formation, pain, or recurrent
gross hematuria.

been overcome. Several randomized clinical trials


are underway comparing open versus robotic but to
date there is a paucity of long-term data to provide
guidance regarding long-term oncologic outcomes.
Surgical modifications that spare a portion of the
prostate have been developed in order to improve
continence and potency following cystectomy. It is
not clear that this technique offers an incremental
improvement in quality of life outcomes over
meticulous nerve-sparing and careful dissection
of the apical prostate and membranous urethra.
The oncologic risk for prostatic involvement with
urothelial cancer is up to 40%.[13] The incidence
of prostate adenocarcinoma is similar, and at least
one-half of these cancers are considered clinically
significant.[14]
In sexually active women, vaginal preservation and/
or reconstruction must be discussed and planned
preoperatively. Recent reports challenge the dogma
of removal of the internal female organs as the rate
of involvement of the uterus, cervix, and ovaries
is low.[15] Preservation of the vagina and uterus
provides better support for a neobladder and the
pelvic floor when the extent of the cancer or the age
and general health status of the patient does not
warrant anterior pelvic exenteration.[16] One also
needs to balance the benefits of retaining the uterus,
Fallopian tubes, and ovaries versus the increased
surgical risks of potentially requiring hysterectomy
later after a neobladder. Involvement of the urethra
or bladder neck is an absolute contraindication to
urethra-sparing in women and a posterior-based
invasive cancer is a relative contraindication.[17,18]

Much has been written about the timing of cystectomy


and long-term survival, which is inconclusive due in
part to the retrospective nature of the research and
the biology and demographics of the unique patient
populations studied. A recent population based study
from the United States SEER Medicare database
looked at patients operated on between 1992 and
2001.[5] The authors findings suggest that 12 weeks
is the outside limit beyond which there is a negative
impact on outcome (level 3b).
The goals of therapy are complete eradication of the
loco-regional disease including the primary tumor
and regional lymph nodes. The primary landing
zone for node metastases includes the obturator
and internal and external iliac lymph nodes. The
incidence of node metastases increases with tumor
stage. The majority of patients with node metastases
have multiple positive nodes and at least one-third
have node metastases in secondary landing zones
including the pre-sacral and common iliac nodes.
[6-8] Neoadjuvant or adjuvant chemotherapy
targeting occult regional node and/or visceral
metastases should be integrated with loco-regional
therapy when the risk of this occurring is high
enough.[9-12] Patients should have volitional control
of urination via the retained urethra or by a continent
catheterizable stoma (in properly selected patients)
and treatment-related morbidity should be minimized
wherever possible.

The morbidity of cystectomy can be significant


though the mortality has been reduced to 2-3% in
most contemporary series.[19] A recent report from
Memorial Sloan Kettering Cancer Center described
1145 patients operated on between 1995 and
2005 using a modified Clavien system.[20,21] The
majority of patients (64%) experienced one or more
complications and the majority of these were grade
2-5. The hospital re-admission rate was 26%. The
authors suggest that this degree of peri-operative
morbidity may limit use of adjuvant chemotherapy in
as many as 30%[20].

Recommendations
Radical cystectomy is indicated for definitive locoregional control of non-muscle-invasive urothelial
cancer that is refractory to intravesical therapy or
cannot be controlled with transurethral resection
(LE 2b GR B).

Radical cystectomy provides excellent local


control of the primary tumor. In men, this should
include the bladder and surrounding perivesical
soft tissue, prostate, and seminal vesicles, and, in
women, the ovaries, uterus with cervix, and anterior
vagina. With the pervasive use of robotic-assisted
laparoscopic radical prostatectomy, many surgeons
are applying these skill sets to radical cystectomy
(discussed in Section 2.1.6). Proof of concept has
been demonstrated as the technical hurdles have

T
 he most common indication is primary treatment
for patients with initial occurrence of muscleinvasive cancer or progression to muscle-invasive
cancer that has failed intravesical therapy (LE 2a
GR B). The evidence supporting this is based on

272

an open surgical technique incorporating bilateral


pelvic lymphadenectomy.

11. Adjuvant chemotherapy in invasive bladder cancer: a


systematic review and meta-analysis of individual patient
data Advanced Bladder Cancer (ABC) Meta-analysis
Collaboration. Eur Urol. Aug 2005;48(2):189-199;
discussion 199-201.

M
 inimally-invasive techniques including roboticassisted laparoscopic radical cystectomy have
been reported in increasing numbers. Shortterm
outcomes, pathologic findings, and assessments
of morbidity compared to the open technique have
been reported. Continued research with these
techniques is required and specifically results of
several randomized trials underway should be
reported prior to accepting this as an equivalent
option to open radical cystectomy (LE 2a GR C).

12. Grossman HB, Natale RB, Tangen CM, et al. Neoadjuvant


chemotherapy plus cystectomy compared with cystectomy
alone for locally advanced bladder cancer. N Engl J Med.
Aug 28 2003;349(9):859-866.
13. Shen SS, Lerner SP, Muezzinoglu B, Truong LD, Amiel
G, Wheeler TM. Prostatic involvement by transitional cell
carcinoma in patients with bladder cancer and its prognostic
significance. Hum Pathol. Jun 2006;37(6):726-734.
14. Revelo MP, Cookson MS, Chang SS, Shook MF, Smith JA,
Jr., Shappell SB. Incidence and location of prostate and
urothelial carcinoma in prostates from cystoprostatectomies:
implications for possible apical sparing surgery. J Urol. May
2008;179(5 Suppl):S27-32.

C
 ystectomy is also indicated as palliation for
intractable pelvic pain and bleeding from locally
advanced or metastatic bladder cancer (LE 4 GR
C).

15. Chang SS, Cole E, Smith JA, Jr., Cookson MS. Pathological
findings of gynecologic organs obtained at female radical
cystectomy. J Urol. Jul 2002;168(1):147-149.

(Level of Evidence = LE,


Grade of Recommendation = GR)

16. Ali-El-Dein B, Gomha M, Ghoneim MA. Critical evaluation


of the problem of chronic urinary retention after orthotopic
bladder substitution in women. J Urol. Aug 2002;168(2):587592.

References

17. Schilling D, Horstmann M, Nagele U, Sievert KD, Stenzl


A. Cystectomy in women. BJU Int. Nov 2008;102(9 Pt
B):1289-1295.

1. Hautmann RE, Gschwend JE, de Petriconi RC, Kron M,


Volkmer BG. Cystectomy for transitional cell carcinoma
of the bladder: results of a surgery only series in the
neobladder era. J Urol. Aug 2006;176(2):486-492;
discussion 491-482.

18. Stein JP, Penson DF, Lee C, Cai J, Miranda G, Skinner DG.
Long-term oncological outcomes in women undergoing
radical cystectomy and orthotopic diversion for bladder
cancer. J Urol. May 2009;181(5):2052-2058; discussion
2058-2059.

2. Stein JP, Lieskovsky G, Cote R, et al. Radical cystectomy in


the treatment of invasive bladder cancer: long-term results
in 1,054 patients. J Clin Oncol. Feb 1 2001;19(3):666-675.

19. Lowrance WT, Rumohr JA, Chang SS, Clark PE, Smith
JA, Jr., Cookson MS. Contemporary open radical
cystectomy: analysis of perioperative outcomes. J Urol. Apr
2008;179(4):1313-1318; discussion 1318.

3. Dalbagni G, Vora K, Kaag M, et al. Clinical outcome in a


contemporary series of restaged patients with clinical T1
bladder cancer. Eur Urol. Dec 2009;56(6):903-910.

20. Donat SM, Shabsigh A, Savage C, et al. Potential impact of


postoperative early complications on the timing of adjuvant
chemotherapy in patients undergoing radical cystectomy:
a high-volume tertiary cancer center experience. Eur Urol.
Jan 2009;55(1):177-185.

4. Morris DS, Weizer AZ, Ye Z, Dunn RL, Montie JE,


Hollenbeck BK. Understanding bladder cancer death:
tumor biology versus physician practice. Cancer. Mar 1
2009;115(5):1011-1020.

21. Shabsigh A, Korets R, Vora KC, et al. Defining early


morbidity of radical cystectomy for patients with bladder
cancer using a standardized reporting methodology. Eur
Urol. Jan 2009;55(1):164-174.

5. Gore JL, Lai J, Setodji CM, Litwin MS, Saigal CS. Mortality
increases when radical cystectomy is delayed more than 12
weeks: results from a Surveillance, Epidemiology, and End
Results-Medicare analysis. Cancer. Mar 1 2009;115(5):988996.

III. Radical Surgery

6. Dangle PP, Gong MC, Bahnson RR, Pohar KS. How do


commonly performed lymphadenectomy templates influence
bladder cancer nodal stage? J Urol. Feb;183(2):499-503.

1. R
 emoval of the Tumor-bearing
Bladder

7. Roth B, Wissmeyer MP, Zehnder P, et al. A new multimodality


technique accurately maps the primary lymphatic landing
sites of the bladder. Eur Urol. Feb;57(2):205-211.

a) Background

8. Leissner J, Ghoneim MA, Abol-Enein H, et al. Extended


radical lymphadenectomy in patients with urothelial bladder
cancer: results of a prospective multicenter study. J Urol.
Jan 2004;171(1):139-144.

Radical cystectomy is the standard treatment for


muscle-invasive bladder cancer in most countries
worldwide [1]. In the pre-cystectomy era, 5-year
survival rates for patients with muscle-invasive
disease rarely exceeded 3% and performing radical
surgery was associated with a considerably increased
perioperative morbidity and mortality [2]. In the last
decades, advances in the surgical technique as well
as perioperative anesthesia care have significantly

9. Neoadjuvant cisplatin, methotrexate, and vinblastine


chemotherapy for muscle-invasive bladder cancer: a
randomised controlled trial. International collaboration of
trialists. Lancet. Aug 14 1999;354(9178):533-540.
10. Neoadjuvant chemotherapy in invasive bladder cancer:
update of a systematic review and meta-analysis of
individual patient data advanced bladder cancer (ABC)
meta-analysis collaboration. Eur Urol. Aug 2005;48(2):202205; discussion 205-206.

273

decreased the complication rates associated with


this procedure and have led to radical cystectomy
now being considered the mainstay of treatment for
this muscle-invasive disease.

morbidity of radical cystectomy in select elderly patients


with bladder cancer. J Urol 2002; 167(3): 1325-8
8. Lughezzani G, Sun M, Shariat SF, Budus L, Thuret R,
Seldres C, Liberman D, Montorsi F, Perrotte P, Karakiewicz
PI. A population based competing-risks analysis of the
survival of patients treated with radical cystectomy for
bladder cancer. Cancer 2011; 117(1):103-9

In the last decade, increased interest in quality of


life issues has increased the trend toward bladder
preservation with transurethral resection followed by
chemotherapy and radiation therapy [3, 4, 5]. Also,
performance status and age were formerly reported
to influence the choice of therapy, with cystectomy
being reserved for younger patients without
concomitant disease and better performance status
[6]. Nonetheless, recent studies have demonstrated
the feasibility and safety of radical surgery in
octogenarians [7]. The value of assessing overall
health before recommending and proceeding with
surgery was emphasized in a prior report, which
demonstrated an association between comorbid
illness and adverse pathological and survival
outcome following radical cystectomy [5]. Similar
to these results, a recent analysis from the SEER
registries evaluated the impact of comorbidities
on cancer-specific and other-cause mortality in a
population-based competing risk analysis of more
than 11,260 patients. Age was found to confer the
highest risk for other-cause mortality but not for
increased cancer-specific death, whereas locally
advanced tumor stage was the strongest predictor
for decreased cancer-specific survival [8]. Stratifying
elderly patients according to their individual riskbenefit profile within a multidisciplinary team might
help to select those who may benefit most from
radical surgery and optimize their outcomes.

b) Timing and Delay of Cystectomy


It has been suggested that delaying radical
cystectomy in organ-confined disease is associated
with decreased survival [1]. These concerns regard
primary surgery as monotherapy; the time spent for
neoadjuvant chemotherapy is another matter. Even
though it is well accepted that treatment should
be instituted once a diagnosis of cancer has been
made, several factors play a role in delaying such
treatment. Some of these factors are linked to
the health care system, while others are patientrelated. The question of whether there is a window
of opportunity for the treatment of invasive bladder
cancer remains unanswered.
Reviewing all studies on this issue, it becomes evident
that there is no linear relationship between delay of
radical cystectomy and prognosis. However, delays
are associated with worse outcome. The majority
of studies suggest that a treatment delay of more
then 12 weeks from diagnosis of invasive disease
to cystectomy is associated with a significantly
impaired prognosis.
Lee et al [2] analyzed the outcome of 214 patients who
underwent radical cystectomy for muscle-invasive
bladder cancer and they observed a significant
disease-specific (p=0.05) and overall survival
(p=0.02) advantage in patients who underwent radical
cystectomy within 3 months or less as compared to
those undergoing radical cystectomy greater than
3 months after diagnosis. Interestingly, the most
common factor contributing to cystectomy delay was
a scheduling delay. In a similar study, Fahmy et al [3]
characterized various periods of delay sustained by
bladder cancer patients before radical cystectomy
across Quebec. The authors analyzed the 1633
patients being treated with radical cystectomy
between 1990 and 2002 and they identified a delay
of more than 84 days to be associated with a 1.4
times increased risk of death from bladder cancer.
Gore et al [4] examined the survival impact of a delay
in radical cystectomy using nationally representative
SEER data. The authors identified 441 patients
with stage II bladder cancer who underwent radical
cystectomy between 1992 and 2001. Delay in
radical cystectomy beyond 12 weeks and 24 weeks
resulted in a 2-fold increased risk for both diseasespecific and overall mortality when compared to
immediate radical cystectomy performed within 4 to
8 weeks after diagnosis of muscle-invasive bladder
cancer. Kulkarni et al [5] identified 2535 patients who
underwent radical cystectomy between 1992 and

References
1. Stein JP, Lieskovsky G, Cote R, Groshen S, Feng AC,
Boyd S, Skinner E, Bochner B, Thangathurai D, Mikhail
M, Raghavan D, Skinner DG. Radical cystectomy in the
treatment of invasive bladder cancer: long-term results in
1,054 patients. J Clin Oncol 2001; 19(3):666-75.
2. Bostrm PJ, Mirtti T, Kssi J, Laato M, Nurmi M. Twenty
year experience of radical cystectomy in a medium volume
center. Scand J Urol Nephrol. 2009;43(5):357-64.
3. Herr HW, Bajorin DF, Scher HI. Neoadjuvant chemotherapy
and bladder- sparing surgery for invasive bladder cancer:
ten-year outcome. J Clin Oncol. 1998; 16(4):1298-301.
4. International collaboration of trialists. Neoadjuvant cisplatin,
methotrexate, and vinblastine chemotherapy for muscleinvasive bladder cancer: a randomised controlled trial.
Lancet. 1999; 354(9178):533-40
5. Zietman AL, Shipley WU, Kaufman DS. Organ-conserving
approaches to muscle-invasive bladder cancer: future
alternatives to radical cystectomy. Ann Med. 2000; 32(1):3442.
6. Miller DC, Taub DA, Dunn RL, Montie JE, Wei JT. The impact
of co-morbid disease on cancer control and survival following
radical cystectomy. J Urol. 2003Jan;169(1):105-9.
7. Soulie M, Straub M, Game X, Seguin P, De Petriconi
R, Plante P, Hautmann RE. A multicenter study on the

274

2004 in Ontario. The median waiting time between


transurethral resection and radical cystectomy was
50 days. Prolonged waiting times were significantly
associated with a lower overall survival rate. A cubic
splines regression analysis revealed that the risk
of death began to increase after a waiting time of
40 days. Similar data were reported by Jger et al
[6] who identified a waiting time of longer than 4
months to be associated with a significantly reduced
5-year cancer specific survival of 77% versus 86%
(p=0.04). Ayres et al [7] recently analyzed whether
a delay to radical cystectomy affects the survival
of patients with bladder cancer in England. The
analysis of organ-confined bladder cancer revealed
that a delay of greater than 90 days was associated
with a 1.4 times increased risk of death from bladder
cancer whereas there was significant correlation
between delay and prognosis for patients with locally
advanced disease.

5. Kulkarni GS, Urbach DR, Austin PC, Fleshner NE, Laupacis


A. Longer wait times increase overall mortality in patients
with bladder cancer. J Urol. 2009 Oct;182(4):1318-24.
Epub 2009 Aug 14.

There is one recent publication in the literature which


does not demonstrate any correlation between the
delay from the last TUR to radical cystectomy and
decreased survival times [8]. The authors analyzed
the outcomes of 592 patients who underwent radical
cystectomy within a median time from TUR to radical
cystectomy of 1.8 months. However, it has to be
taken into consideration that 35% of all patients who
underwent radical cystectomy demonstrated nonmuscle-invasive disease on final pathology.

In male patients, the literature over the last two


decades has set the standard of surgical limits for
curative radical cystectomy as complete removal
of the bladder with all macroscopically visible and
resectable bladder perforating tumor extensions,
removal of the adjacent distal ureters, and the lymph
nodes corresponding to the tumor-bearing bladder.

6. Jger W, Thomas C, Haag S, Hampel C, Salzer A, Throff


JW, Wiesner C. Early vs delayed radical cystectomy for
high-risk carcinoma not invading bladder muscle: delay of
cystectomy reduces cancer-specific survival. BJU Int. 2011
Jan 18. doi: 10.1111/j.1464-410X.2010.09980.x. [Epub
ahead of print]
7. Ayres BE, Gillatt D, McPhail S, Cottrell A, McGrath J,
Cottier B, Verne J, Persad RA. A DELAY IN RADICAL
CYSTECTOMY OF > 3 MONTHS IS NOT ASSOCIATED
WITH A WORSE CLINICAL OUTCOME. BJU Int 2008;
102: 1045
8. Nielsen ME, Palapattu GS, Karakiewicz PI, Lotan Y,
Bastian PJ, Lerner SP, Sagalowsky AI, Schoenberg MP,
Shariat SF. A delay in radical cystectomy of >3 months is
not associated with a worse clinical outcome. BJU Int. 2007
Nov;100(5):1015-20. Epub 2007 Sep 3.

c) Technique

Preservation of the anterior and membranous urethra


including the rhabdosphincter in order to enable an
orthotopic neobladder; parts of the prostate and
seminal vesicles for reasons of fertility, potency, and
continence; and intrapelvic autonomic and sensory
nerves to enhance potency and continence are all
technical variations to this standard that may improve
patients quality of life but must be attentively judged
against possible oncological risks [1] (Level of
evidence: 3, grade of recommendation: C).

Recommendations
D
 elay in cystectomy for muscle-invasive disease
is associated with significantly reduced overall
and cancer-specific survival (LE: 3, GR: B).
R
 adical cystectomy in patients with muscleinvasive bladder cancer should be performed
within 3 months after initial diagnosis of stage T2
T4 disease (LE: 3, GR: B)

In case of leaving parts of the prostatic gland


during the resection, the hazard of an unsuspected
adenocarcinoma may be as high as 23-54% of which
up to 29% may be clinically significant, lead to local
recurrence, or even metastasis [2-4]. Furthermore,
urothelial carcinoma may be present in the prostate
and in some series up to 27% of patients undergoing
cystoprostatectomy had prostate cancer [5]. A rather
new technical variation is deliberately leaving the
seminal vesicles and prostatic capsule in order to
better preserve the surrounding autonomic nerves.
The results regarding potency versus oncological risk
in small series of selected patients are encouraging
but need long-term confirmation in larger series [6, 7].
To date these technical modifications have not been
documented to improve continence and they remain
highly controversial regarding oncologic safety.

References
1. Fahmy NM, Mahmud S, Aprikian AG. Delay in the surgical
treatment of bladder cancer and survival: systematic review
of the literature. Eur Urol. 2006 Dec;50(6):1176-82
2. Lee CT, Madii R, Daignault S, Dunn RL, Zhang Y, Montie
JE, Wood DP Jr. Cystectomy delay more than 3 months
from initial bladder cancer diagnosis results in decreased
disease specific and overall survival. J Urol. 2006
Apr;175(4):1262-7
3. Fahmy N, Kassouf W, Jeyaganth S, Amin M, Mahmud
S, Steinberg J, Tanguay S, Aprikian A. An analysis of
preoperative delays prior to radical cystectomy for bladder
cancer in Quebec. Can Urol Assoc J. 2008 Apr;2(2):102-8.

In female patients, standard anterior pelvic


exenteration includes the entire urethra, adjacent
vagina, uterus, distal ureters, and respective lymph
nodes (level of evidence: 3, grade of recommendation:
C). Unless the primary tumor is located at the

4. Gore JL, Lai J, Setodji CM, Litwin MS, Saigal CS; Urologic
Diseases in America Project. Mortality increases when
radical cystectomy is delayed more than 12 weeks: results
from a Surveillance, Epidemiology, and End ResultsMedicare analysis. Cancer. 2009 Mar 1;115(5):988-96.

275

bladder neck or in the urethra, a major part of the


functioning female urethra and, provided a complete
tumor resection is possible, its autonomous nerves
can be preserved in case of a planned orthotopic
neobladder [8, 9, 10, 11] (level of evidence: 3, grade
of recommendation: C). New data also question the
necessity to remove the uterus or any portion of the
vagina, arguing for a better anatomical support of the
neobladder and better preservation of surrounding
autonomous nerves.

cystoprostatectomy: defining risk with prostate capsule


sparing cystectomy. Urol Oncol, 2007. 25(6): p. 460-4.
5. Gakis, G., D. Schilling, J. Bedke, K.D. Sievert and A. Stenzl,
Incidental prostate cancer at radical cystoprostatectomy:
implications for apex-sparing surgery. BJU Int,
2010. 105(4): p. 468-71.
6. Ong, C.H., M. Schmitt, G.N. Thalmann and U.E. Studer,
Individualized seminal vesicle sparing cystoprostatectomy
combined with ileal orthotopic bladder substitution achieves
good functional results. J Urol, 2010. 183(4): p. 1337-41.
7. Colombo, R. and R.E. Hautmann, Open to debate. The
motion: seminal-nerve sparing radical cystectomy is an
efficacious and safe treatment for selected bladder cancer
patients. Eur Urol, 2008. 53(1): p. 203-7.

In both sexes, the length of the distal ureteral


segment to be removed with the bladder has not
been specified and depends on oncological issues
such as tumor extension or presence of carcinoma
in situ and type of subsequent urinary diversion. In
one recent study, frozen section of the distal ureteral
margins had a sensitivity of 74% and a specificity
of 99.8%, resulting in an overall accuracy of 98.3%
[12]. With a serial sectioning strategy, most initially
positive ureteral margins can be converted into
negative final margins. Those patients were also at
decreased risk of developing recurrent disease in
the upper urinary tract [13, 14].

8. Sternberg, C.N., P.H. de Mulder, J.H. Schornagel, C.


Theodore, S.D. Fossa, A.T. van Oosterom, F. Witjes,
M. Spina, C.J. van Groeningen, C. de Balincourt, and
L. Collette, Randomized phase III trial of high-doseintensity methotrexate, vinblastine, doxorubicin, and
cisplatin (MVAC) chemotherapy and recombinant human
granulocyte colony-stimulating factor versus classic MVAC
in advanced urothelial tract tumors: European Organization
for Research and Treatment of Cancer Protocol no. 30924.
J Clin Oncol, 2001. 19(10): p. 2638-46.
9. Herr, H.W., D.F. Bajorin and H.I. Scher, Neoadjuvant
chemotherapy and bladder-sparing surgery for invasive
bladder cancer: ten-year outcome. J Clin Oncol, 1998.
16(4): p. 1298-301.

Recommendations

10. Hautmann, R.E., H. Abol-Enein, K. Hafez, I. Haro, W.


Mansson, R.D. Mills, J.D. Montie, A.I. Sagalowsky, J.P.
Stein, A. Stenzl, U.E. Studer, and B.G. Volkmer, Urinary
diversion. Urology, 2007. 69(1 Suppl): p. 17-49.

Preservation of the anterior and membranous


urethra, including parts of the prostate and seminal
vesicles for reasons of fertility, potency, and
continence, are technical variations to the nervesparing approach that may improve patients
quality of life but must be attentively judged against
possible oncological risks. (LE 3, GR C)

11. Stenzl A, Jarolim L, Coloby P, Golia S, Bartsch G, Babjuk


M, Kakizoe T, Robertson C. Urethra-sparing cystectomy
and orthotopic urinary diversion in women with malignant
pelvic tumors. Cancer. 2001 Oct 1;92(7):1864-71.
12. Gakis, G., D. Schilling, S. Perner, C. Schwentner, K.D.
Sievert and A. Stenzl, Sequential resection of malignant
ureteral margins at radical cystectomy: a critical assessment
of the value of frozen section analysis. World J Urol, 2010.

In female patients, standard anterior pelvic exenteration includes the entire urethra, adjacent vagina,
uterus, distal ureters, and respective lymph nodes
(LE: 3, GR: C). Sparing the urethra-supplying autonomous nerves can be performed in case of a
planned orthotopic neobladder (LE: 3, GR: C).

13. Tollefson, M.K., M.L. Blute, S.A. Farmer and I. Frank,


Significance of distal ureteral margin at radical cystectomy
for urothelial carcinoma. J Urol, 2010. 183(1): p. 81-6.
14. Schumacher, M.C., M. Scholz, E.S. Weise, A. Fleischmann,
G.N. Thalmann and U.E. Studer, Is there an indication for
frozen section examination of the ureteral margins during
cystectomy for transitional cell carcinoma of the bladder? J
Urol, 2006. 176(6 Pt 1): p. 2409-13; discussion 2413.

References

d) Lymphadenectomy
1. Pelvic Lymphadenectomy in Bladder Cancer

1. Stenzl, A., U. Nagele, M. Kuczyk, K.D. Sievert, A.


Anastasiadis, J. Seibold and S. Corvin, Cystectomy Technical Considerations in Male and Female Patients.
EAU Update Series, 2005. 3: p. 138-46.

Radical cystectomy with pelvic lymphadenectomy


represents the treatment of choice for muscleinvasive bladder cancer. Pelvic lymphadenectomy is
included as part of the surgical procedure to control
loco-regional disease and to potentially improve
cancer-specific survival. Survival after radical
cystectomy is usually predicted by the pathologic
tumor stage, status of surgical margins, and the
involvement of lymph nodes.

2. Abdelhady, M., A. Abusamra, S.E. Pautler, J.L. Chin and


J.I. Izawa, Clinically significant prostate cancer found
incidentally in radical cystoprostatectomy specimens. BJU
Int, 2007. 99(2): p. 326-9.
3. Pettus, J.A., H. Al-Ahmadie, D.A. Barocas, T.M. Koppie, H.
Herr, S.M. Donat, G. Dalbagni, V.E. Reuter, S. Olgac and
B.H. Bochner, Risk assessment of prostatic pathology in
patients undergoing radical cystoprostatectomy. Eur Urol,
2008. 53(2): p. 370-5.

Although
earlier
studies
have
already
demonstrated a prognostic benefit of extended
pelvic lymphadenectomy as compared to a limited
lymphadenectomy, the anatomically adequate extent

4. Weizer, A.Z., R.B. Shah, C.T. Lee, S.M. Gilbert, S. Daignault,


J.E. Montie and D.P. Wood, Jr., Evaluation of the prostate
peripheral zone/capsule in patients undergoing radical

276

of lymph node dissection to obtain reliable staging


results is still controversial.

most important prognostic factor in these patients,


predicting significantly decreased recurrence-free
survival and overall survival compared to patients
without node metastasis [23, 33, 34]. In patients
with negative nodes, the total number of lymph
nodes removed and the anatomic extent of the node
dissection are both useful measures in evaluating
the proper extent of surgery and predicting outcome.
In patients with nodal metastasis, the number of
nodes removed and the number and percent of
positive nodes may both be independent predictors
of recurrence and survival [35-38].

In a recent prospective randomized phase III trial on


the clinical efficacy of neoadjuvant chemotherapy
plus cystectomy [1], it was shown that surgical
factors, including the extent of lymph node dissection
and the individual surgeons experience, have a
major impact on the therapeutic outcome and overall
survival [2] (Level of Evidence: 1b). The data of
this trial also indicate that chemotherapy was more
likely to be of beneficial value if patients received
a high-quality surgery by an experienced surgeon.
It was concluded that it is extremely important to
develop universally accepted standards for radical
cystectomy and pelvic lymph node dissection in
patients with invasive bladder cancer in order to
improve outcome [3].

4. Extent of Pelvic Lymphadenectomy and


Therapeutic Outcome
Despite the aforementioned data which demonstrate
that long-term survival is possible in patients with
lymph node-positive bladder cancer, the anatomical
extent of pelvic lymphadenectomy and the minimum
of lymph nodes to be retrieved for an accurate
staging have not been defined. Some others take the
crossing of the ureters with the common iliac vessels
as the most cranial limit for lymph node dissection,
[9] whereas others extend lymphadenectomy up to
the aortic bifurcation. It is generally agreed that the
more lymph nodes are removed, the higher is the
number of patients with positive lymph nodes [10]
LE 3; furthermore, it has been demonstrated that
survival after radical cystectomy is predicted by the
pathologic stage of the primary bladder tumor and
pelvic nodes. Leissner et al [11] suggested that a
significant survival benefit was maintained if more
than 16 lymph nodes were removed. Stein et al [12]
reported that survival in patients with positive lymph
node disease was better if more than 15 pelvic
lymph nodes had been retrieved. On the other hand,
Abdel-Latif and coworkers [13] and Herr [14] could
not reproduce the relationship between survival
and number of dissected lymph nodes by using
multivariate statistical analysis. In this context, it has
to be underlined that the number of retrieved lymph
nodes can be influenced by many factors such as the
surgeon[47], the extent of lymphadenectomy [9-11],
presentation of the pathological specimen [16],
and pathologic work-up and techniques of analysis
[17]. The clinical reality, however, is significantly
different: an evaluation of the SEER database in
2003 demonstrated that the majority of patients in
a population-based analysis had 4 or fewer nodes
removed with cystectomy [40, 41].

2. Threshold Number of Lymph Nodes for


Accurate Staging
Recently, Capitanio et al [17] evaluated the likelihood
of finding one or more positive lymph nodes (LNs)
according to the number of LNs removed at radical
cystectomy (Level of Evidence: 2b). Results from
731 patients who underwent radical cystectomy and
bilateral pelvic lymphadenectomy at three different
institutions were analyzed. ROC curve coordinates
were used to determine the probability of identifying
1 or more positive LNs according to the total number
of removed LNs. Lymph node metastases were
present in 174 (23.8%). The mean (median, range)
number of LNs removed was 18.7 [17, 1-80]. The
ROC coordinate-based plots of the number of
removed LNs and the probability of finding 1 or
more LN metastases indicated that removing 45
LNs yielded a 90% probability. Conversely, removing
either 15 or 25 LNs indicated, respectively, 50%
and 75% probability of detecting 1 or more LNs
metastases. These data indicate that removing 25
LNs might represent the lowest threshold for the
extent of lymphadenectomy at radical cystectomy.
In a similar approach, Koppie et al reported a study
designed to determine if there was a minimum
number of lymph nodes analyzed above which there
was no improvement in survival [53] (LE: 2b). The
cohort included 1121 patients from MSKCC accrued
over a 14-year period. The investigators determined
that there was no plateau in the dose-response curve
with an increasing number of nodes up to 23 nodes,
as very few had 24 or more nodes removed. The
authors did not indicate the percent of patients who
underwent an extended node dissection, and, in fact,
13% had no nodes identified in the pathology report.
The median number of nodes removed was 9.

Despite these suboptimal clinical factors, there are a


few studies having demonstrated a significant impact
of the technique of pelvic lymph node dissection
with regard to therapeutic outcome. Poulsen et al
[8] were among the first to compare the prognostic
significance of limited versus extended pelvic
lymphadenectomy in a retrospective analysis of
194 patients undergoing radical cystectomy [LE
3]. Limited pelvic lymphadenectomy began at

3. Anatomic Extent of Pelvic Lymphadenectomy


Lymph node metastases are found in 20-25% of
patients who undergo radical cystectomy and pelvic
lymphadenectomy for bladder cancer, and it is the

277

the iliac bifurcation including the lymph nodes


along the external and internal iliac artery and the
obturator fossa. Extended pelvic lymphadenectomy
began at the aortic bifurcation and included the
common, external, and internal iliac artery and the
obturator fossa. The authors observed a substantial
improvement of 5-year recurrence-free survival in
patients with tumors confined to the bladder wall (85%
vs. 64%, p<0.02) and without lymph node involvement
(90% vs. 71%, p<0.02). Five-year probability for
loco-regional (7% vs. 2%) and systemic recurrences
(21% vs. 10%) was reduced substantially in patients
with bladder cancer confined to the bladder wall in
the extended pelvic lymphadenectomy group, but
did not reach statistical significance.

the nodal regions of standard dissection. Although


the median number of positive lymph nodes differed
significantly between both groups (22.5 vs. 8), there
was no difference with regard to the percentage
of positive nodes, which was 21% in both groups.
Interestingly, all patients with positive nodes above
the aortic bifurcation also had positive nodes
detected in the lower packages indicating that only
extensive locoregional metastatic disease might
involve the retroperitoneal areas associated with a
dismal prognosis. Including the lymph nodes along
the common iliac artery above the iliac bifurcation,
however, appears to be of prognostic value and of
clinical significance. In the study of Bochner et al
[21], 4 patients had unexpected micrometastatic
lymph node disease at the common iliac region only.
Reflecting the survival data of patients exhibiting
micrometastatic lymph node disease at time of radical
cystectomy, most of these patients are expected to
have a relatively favorable outcome. Morbidity of
pelvic lymphadenectomy is not increased by including
the common iliac region in routine pelvic lymph node
dissection, so this area should be removed as a
standard part of staging lymphadenectomy.

In another retrospective analysis of 484 patients


undergoing radical cystectomy and pelvic
lymphadenectomy, Leissner et al [17] demonstrated
that the total number of lymph nodes retrieved had
a significant impact on recurrence-free survival
(p<0.01) (LE 2b). The 5-year recurrence-free survival
rates were 25% and 53% in patients with 14 or fewer
and 15 or more lymph nodes removed, respectively.
Furthermore, the surgeon had a significant impact
on the prognosis as it was shown the number
of lymph nodes dissected ranged between 10.6
and 25.7 and differed significantly between the
11 different surgeons in the study. These data
are further corroborated by a recent paper on the
standardization of radical cystectomy and pelvic
lymphadenectomy [3]. However, the authors did
not demonstrate a significant overall and cancerspecific survival advantage for patients undergoing
extended pelvic lymphadenectomy as compared to
those undergoing limited dissection. The authors
further evaluated the concept of extended pelvic
lymphadenectomy in a prospective clinical trial of
290 patients [20]. The cranial limit of dissection was
the inferior mesenteric artery, the lateral border was
the genitofemoral nerve, and the caudal limit was the
pelvic floor. A mean number of 43.1 16.1 lymph
nodes were removed with 27.9% of the patients
having positive lymph nodes. Although the group
identified a preferred pattern of metastatic spread,
they were not able to identify a well-defined sentinel
lymph node or lymph node area.

Further evidence to include the common iliac region


derives from the prospective multi- institutional study
published recently by Leissner et al [20], (LE 2b).
Eighty-one (27.9%) patients demonstrated lymph
node involvement and 35% of all positive lymph
nodes derived from above the iliac bifurcation.
Furthermore, 20 (6.9%) patients were shown to
harbor positive lymph nodes above the bifurcation
of the common iliac artery only. Although no data
with regard to the prognostic significance in terms
of cancer-specific or progression-free survival are
available, these data strongly support the idea to
include the lymph nodes of the common iliac region
up to the aortic bifurcation in routine lymph node
dissection for muscle-invasive bladder cancer.
In another study, Abol-Enein et al [22] evaluated the
loco-regional distribution of positive pelvic lymph
nodes in 200 consecutive patients undergoing
radical cystectomy (LE 3). The authors also
attempted to identify the probability of lymph node
clearance with increasingly wide fields of node
dissection. In their investigation, extended pelvic
lymphadenectomy included the lymphatic tissue up
to the inferior mesenteric artery and the common,
external, and internal iliac region. A mean number of
50.6 lymph nodes were retrieved per patient with 48
(24%) patients exhibiting positive nodes. More than
one-third of these patients (39.6%) demonstrated
bilateral involvement. A single positive lymph node
was identified in 22 (45.8%) patients. The authors
demonstrated that close to 80% of all positive nodes
could be cleared completely if the field of pelvic
lymph node dissection included all lymphatic tissues
along the common, external, and internal iliac
region. Metastases outside the true pelvis were only

These data are in contrast to the recently published


prospective trial of Borchner et al [21] on the
evaluation of lymph node count and lymph node
mapping (LE 3). One hundred forty-four consecutive
patients were included in this monocentric evaluation
with 56 and 88 patients undergoing standard and
extended pelvic lymphadenectomy, respectively.
Standard dissection included the nodal regions of
the external iliac, hypogastric, and obturator fossa
with the iliac bifurcation representing the cranial
limit of lymph node dissection. Extended dissection
included the lymph nodes at the aortic bifurcation to
no more than 2 cm cranially to the bifurcation and

278

detected in multinodal disease and these metastatic


deposits were always associated with metastases at
the obturator fossa and/or the internal iliac region.
Therefore, the authors conclude that standard
lymphadenectomy in bladder cancer should always
include all lymphatic tissues in the true pelvis; lymph
node dissection might be extended up to the inferior
mesenteric artery if frozen section examination
exhibits positive lymph nodes in the sentinel region
of the true pelvis.

similar in patients with lymph node involvement


above the bifurcation of the common iliac vessels
(37%) compared to the entire population with lymph
node metastasis (41%) and to those with lymphatic
metastases in the true pelvis below the bifurcation
of the common iliac vessels (42%). The survival rate
was significantly higher in patients with 5 or fewer
involved lymph nodes (50% vs. 13%, p<0.002)
and in those with a lymph node density (number of
lymph nodes involved/total number of lymph nodes
removed) less than 20% (25% vs. 47%, p<0.05),
but it did not relate to the total number of retrieved
lymph nodes. These data underscore the contention
that extended dissection not only provides the most
accurate staging but also offers the patient the best
chance of survival. Following radical cystectomy,
patients can be stratified into risk groups according
to tumor stage, lymph node involvement, number
of metastatic nodes, and lymph node density.
The results of Steven et al [24] support the idea
that the benchmark for radical cystectomy should
include extensive pelvic lymph node dissection with
anatomic boundaries including the common iliac and
presacral nodes.

Dhar and coworkers [23] evaluated the impact


of limited and extended pelvic lymphadenectomy
in a cohort of 336 and 322 patients, respectively,
who were treated at 2 different institutions (LE 3).
The overall lymph node positive rate was 13% for
patients with limited and 26% for those who had
extended pelvic lymph node dissection. The authors
identified a significantly better recurrence-free
survival for patients who underwent extended pelvic
lymphadenectomy. These figures held true for both
organ-confined and locally advanced disease. The
5-year recurrence-free survival of patients with lymph
node positive disease was 7% for limited and 35% for
extended pelvic lymph node dissection. The 5-year
recurrence-free survival for pT2pN0 cases was
67% for limited and 77% for extended pelvic lymph
node dissection, and the respective percentages for
pT3pN0 cases were 23% and 57% (p<0.0001). The
5-year recurrence-free survival for pT2pN0-2 cases
was 63% for limited and 71% for extended pelvic
lymph node dissection, and for pT3pN0-2 cases the
respective figures were 19% and 49% (p<0.0001).
These data confirm that extended pelvic lymph
node dissection allows for more accurate staging
and improved survival of patients with non-organ
confined and lymph node-positive disease.

5. Extent of an Anatomically Accurate Pelvicstaging Lymphadenectomy in Bladder Cancer


As has been shown by the previous studies, the
anatomic extent of pelvic lymph node dissection in
patients undergoing radical cystectomy for muscleinvasive bladder cancer can be well-defined. Standard
lymphadenectomy should include all lymphatic
tissues around the common iliac, intercommon iliac,
internal iliac, d the obturator groups bilaterally since
up to one-third of all positive nodes are located
around the common iliac artery. This technical variant
will allow one to clear 80% of all positive nodes; if
frozen section does not demonstrate positive lymph
nodes in the true pelvis, lymph node dissection is not
needed to be carried out further cranially. If, however,
frozen section examination is not performed or if
it identifies positive nodes, the inferior mesenteric
artery represents the cranial border of lymph node
dissection.

In another single institution analysis, the clinical


importance of dissecting all lymphatic tissue up to
the aortic bifurcation became evident when analyzing
the outcome of 336 patients who underwent radical
cystectomy and extended pelvic lymphadenectomy
including the common and external iliac lymph
nodes and the periaortic, presacral, and obturator
fossa nodes [24], (LE 3). The lymphatic tissue
removed above and below the bifurcation of the
common iliac vessels was submitted separately for
histopathological analysis. Overall, 64 (19%) patients
had lymph node metastases of whom 22 (34.4%)
had lymph node involvement above the bifurcation of
the common iliac vessels outside the template of the
standard lymph node dissection. The median number
of retrieved lymph nodes was 27 (range 7 to 78) and
in those with lymph node metastases 27 (range 11 to
49) included 8 (range 0 to 17) above the bifurcation
and 18 (range 8 to 41) below the bifurcation of the
common iliac vessels in the true pelvis. Lymph node
involvement proved a significant adverse prognostic
factor with a 5-year probability of survival of 39%
versus 76%. The overall 5-year survival rates were

6. Critical Issues in Anatomic Pelvic


Lymphadenectomy for Bladder Cancer
Although the above-cited studies have apparently
demonstrated the clinical importance of extended
pelvic lymph node dissection with regard to the most
efficient retrieval of micrometastatic lymph nodes,
there are still some unresolved critical issues.
Pathologic examination of dissected lymph node
specimens in these studies has been done more
thoroughly and extensively than in other studies
concentrating on issues such as overall survival,
cancer-specific survival, and regional versus
distant failure. Therefore, some critical facts have
to be considered for the general community if

279

extended pelvic lymphadenectomy becomes


common practice in all patients undergoing radical
cystectomy. Lymphatic tissues dissected from
different areas should be sent separately instead
of en bloc submissions for pathologic evaluation
since it has been demonstrated recently that
the yield of lymph nodes increases significantly
thereby increasing the frequency of micrometastatic
deposits [16]. Intrainstitutional standardization of
pelvic lymphadenectomy appears to be of utmost
importance to generate reliable and reproducible
results since staging lymphadenectomy is extremely
surgeon-dependent, as has been demonstrated by
Leissner et al. [17].

There is some evidence from retrospective and


prospective analysis that an extended pelvic
lymphadenectomy might be associated with an
improvement in 5-year progression-free survival
(p<0.01) (LE 2b-3, GR: B)
Standard lymphadenectomy should include all
lymphatic tissues around the normal common iliac,
external iliac, internal iliac, and the obturator groups
bilaterally since up to one-third of all positive nodes
are located around the common iliac artery (LE: 3,
GR: B).

References

Last, but not least, although the various types of


extended pelvic lymphadenectomies have been
associated with an improved progression-free
survival, none of the trials has demonstrated an
advantage with regard to cancer-specific survival.
Benefit in terms of progression-free survival might
be just due to a stage migration associated with
extensive lymph node dissection. The majority of
patients with positive lymph nodes will die due to
distant metastatic spread in the long run and only
the few patients with a single or two metastatic
lymph nodes might benefit from the extended
variant of lymphadenectomy. In order to solve this
question, the Association of Urological Oncology
of the German Cancer Society has initiated a
prospective randomized clinical phase III trial to
evaluate the true clinical efficacy of extended pelvic
lymphadenectomy.

1. Grossmann HB, Natale RB, Tangen CM, et al. Neoadjuvant


chemotherapy plus cystectomy compared with cystectomy
alone for locally advanced bladder cancer. N Engl J Med
2003; 349: 859 866
2. Herr HW, Faulkner JR, Grossman HB, et al. Surgical factors
influence bladder cancer outcomes: a cooperative group
report. J Clin Oncol 2004; 22: 2781 2789
3. Herr H, Lee C, Chang S, Lerner S for the Bladder Cancer
Collaborative Group. Standardization of radical cystectomy
and pelvic lymph node dissection for bladder cancer: a
collaborative group report. J Urol 2004; 171: 1823 1828
4. Capitanio U, Suardi N, Shariat SF, Lotan Y, Palapattu GS,
Bastian PJ, Gupta A, Vazina A, Schoenberg M, Lerner SP,
Sagalowsky AI, Karakiewicz PI. Assessing
the minimum n
umber of lymph nodes needed at radical
cystectomy in p
atients with bladder cancer. BJU Int. 2009
May;103(10):1359-62
5. Koppie, T.M., A.J. Vickers, K. Vora, G. Dalbagni, and B.H.
Bochner, Standardization of pelvic lymphadenectomy
performed at radical cystectomy: can we establish a
minimum number of lymph nodes that should be removed?
Cancer, 2006. 107(10): p. 2368-74

In bladder cancer, the biological aggressiveness of a


particular cancer may have already been expressed
at time of diagnosis and the impact of less or more
extensive surgery in terms of recurrence and survival
might be completely overestimated. For the future,
it will be necessary to shed light on the important
clinical questions of the biologic role of lymph
node metastases for the likelihood of synchronous
or metachronous distant metastases. It will be
necessary to evaluate the expression of various
growth factors and mediators of systemic spread on
patients with lymph node metastases and to correlate
these findings with clinically important end points
such as cancer-specific survival, overall survival,
and locoregional versus systemic recurrence.

6. Stein, J.P., G. Lieskovsky, R. Cote, S. Groshen, A.C. Feng,


S. Boyd, et al., Radical cystectomy in the treatment of
invasive bladder cancer: long-term results in 1,054 patients.
J Clin Oncol, 2001. 19(3): p. 666-75.
7. Lerner, S.P., D.G. Skinner, G. Lieskovsky, S.D. Boyd, S.L.
Groshen, A. Ziogas, et al., The rationale for en bloc pelvic
lymph node dissection for bladder cancer patients with
nodal metastases: long-term results. J Urol, 1993. 149(4):
p. 758-64; discussion 764-5.
8. Poulsen AL, Horn T, Steven K. Radical cystectomy:
extending limits of pelvic lymph node dissection improves
survival for patients with bladder cancer confined to the
bladder wall. J Urol 1998; 160: 2015 2019
9. Mills RD, Turner WH, Fleischmann A, Markwalder R,
Thalmann GN, Studer UE. Pelvic lymph node metastases
from bladder cancer: outcome in 83 patients after radical
cystectomy and pelvic lymphadenectomy. J Urol 2001; 166:
19 23

Recommendations
The extent of lymph node dissection and the
individual surgeons experience has a major impact
on the therapeutic outcome and overall survival [2]
(LE: 1b, GR: A.)

10. Herr WH, Bochner BH, Dalbagni G, et al. Impact of the


number of lymph nodes retrieved on outcome in patients
with muscle invasive bladder cancer. J Urol 2003; 167:
1295 1297

The more lymph nodes removed, the higher is


the probability to detect at least 1 positive lymph
node. However, there is no defined threshold of the
numbers of lymph nodes that need to be removed
[4, 5] (LE: 2b, GR: B)

11. Leissner J, Hohenfellner R, Throff JW, Wolf HK.


Lymphadenectomy in patients with transitional cell
carcinoma of the urinary bladder: significance for staging
and prognosis. BJU Int 2000; 85: 817 821

280

2. Patient Selection

12. Stein JP, Cai J, Groshen S, Skinner DG. Risk factors for
patients with pelvic lymph node metastases following
radical cystectomy with en-bloc pelvic lymphadenectomy: the concept of lymph node density. J Urol 2003; 170:
35
13. Abdel-Latif M, Abol-Eneim H, El-Baz M, Ghoneim MA.
Nodal involvement in bladder cancer cases treated with
radical cystectomy: incidence and prognosis. J Urol 2004;
172: 85 89
14. Herr WH. Superiority of ratio based lymph node staging for
bladder cancer. J Urol 2003; 169: 943
16. Borchner BH, Herr HW, Reuter VE. Impact of separate
versus en bloc lymph node dissection in the number of
lymph nodes retrieved in cystectomy specimens. J Urol
2001; 166: 2295 17. Leissner J, Allhoff EP, Hohenfellner R, Wolf HK. Ranking
of pelvic lymphadenectomy in therapy and prognosis of
carcinoma of the bladder. Akt Urol 2003; 34: 392 397

Decreased blood loss and need for transfusion, less


opioid requirement, and shortened length of stay, with
comparable complications to open radical cystectomy
has led to radical cystectomy being performed in a
minimally invasive fashion. Meanwhile, immediate
oncologic results are acceptable, while mature data
on long-term oncologic results are still awaited.
Patients with clinical T3 and T4 disease should
be carefully selected for such approach following
consideration for neoadjuvant chemotherapy.[2]
Due to concerns for oncologic control and technical
difficulty, surgeons should selectively choose their
patients for this approach early in their experience.
As minimally invasive approach has evolved
since 1993, key progress has been achieved
with adoption of the robot-assisted approach.
Despite recent advances, approximately 50% of all
patients undergoing radical cystectomy experience
recurrence and subsequent mortality.[3]

18. Hollenbeck, B.K., Z. Ye, S.L. Wong, J.E. Montie, J.D.


Birkmeyer. Hospital lymph node counts and survival after
radical cystectomy. Cancer, 2008. 112(4): p. 806-12.
19. Konety, B.R., S.A. Joslyn, M.A. ODonnell. Extent of Pelvic
Lymphadenectomy and its Impact On Outcome in Patients
Diagnosed With Bladder Cancer: Analysis of Data From the
Surveillance, Epidemiology and End Results Program Data
Base. J Urol, 2003. 169(3): p. 946-950.

3. Surgical Margins
Despite advantages with magnified 3-dimensional
vision and precision achieved by robotic instruments,
in robotic interface the lack of tactile feedback has
created concerns about adequacy of wider excision
in advanced disease and avoiding soft tissue surgical
margins. Positive soft tissue surgical margins are
associated with high local recurrence, resulting
in poor overall survival. Open expert consensus
in 2005 recommended less than 10% of all cases
and less than 15% for bulky tumors as acceptable
positive soft tissue surgical margins rates in radical
cystectomy.[6]

20. Leissner J, Ghoneim MA, Abol-Eneim H, et al. Extended


radical lymphadenectomy in patients with urothelial bladder
cancer: results of a prospective multicenter study. J Urol
2004; 171: 139 144
21. Bochner BH, Cho D, Herr HW, et al. Prospectively packaged
lymph node dissections with radical cystectomy: evaluation
of node count variability and node mapping. J Urol 2004;
172: 1286 1290
22. Abol-Enein H, El-Baz M, Abd El-Hameed MA, Abdel-Latif
M, Ghoneim MA. Lymph node involvement in patients with
bladder cancer treated by radical cystectomy: a pathoanatomical study a single center experience. J Urol
2004; 172: 1818 1821
23. Dhar NB, Klein EA, Reuther AM, Thalmann GN,
Madersbacher S, Studer UE. Outcome
after
radical
cystectomy with limited or extended pelvic lymph node
dissection. J Urol. 2008 Mar;179(3):873-8

Data from the International Laparoscopic Cystectomy


Registry in 2008 revealed a soft tissue surgical margin
rate of 2%.[7] The International Robotic Cystectomy
Consortium, which is a consortium of over 20 robotic
surgeons from 15 institutions, demonstrated in 513
patients an overall positive soft tissue surgical margin
rate of 6.8%. Advanced stage was independently
associated with increased likelihood of a positive
margin while case number and institution volume
were not.[8] These rates are similar to larger open
series. These findings suggest that positive margins
are associated with infiltration of the soft tissue
boundaries of the bladder rather than learning curve
and surgical error. In patients with large volume
tumors and/or suspected extravesical disease, wide
dissection of the perivesical tissue is recommended
to reduce positive soft tissue surgical margin rates.
[9] Collection of data based on surgical pathology
rather than clinical stage and absence of centralized
pathology review might skew the results from the
International Robotic Cystectomy Consortium.

24. Steven K, Poulsen AL. Radical cystectomy and extended


pelvic lymphadenectomy: survival of patients with lymph
node metastasis above the bifurcation of the common iliac
vessels treated with surgery only. J Urol. 2007 Oct;178(4 Pt
1):1218-23

e) Laparoscopic
and
Robotic-assisted
Laparoscopic Cystectomy
1. Introduction
Literature on a minimally invasive approach to
radical cystectomy has evolved since 1993.[1] Over
the past decade, laparoscopic and robot-assisted
approaches have been seen as alternatives to open
radical cystectomy with the possible advantages of
decreased postoperative pain and quicker return
to normal day-to-day activity. The robotic approach
is similar to laparoscopic with the added benefits
of improved range of motion of instruments and
3-dimensional vision, affording a less steep learning
curve.

4. Lymph Node Yield


Bilateral extended pelvic lymphadenectomy is a

281

crucial part of radical cystectomy. The incidence


of positive nodes at the time of radical cystectomy
is in excess of 20%.[5] Based on the observed
therapeutic and prognostic advantages, experts
suggest removal of a minimum of 10-15 lymph
nodes and its adequacy has been criticized in
the minimally invasive approach. (see discussion
chapter 3.1.4) Guru et al reported the feasibility and
safety of performing adequate robot-assisted lymph
node dissection and demonstrated higher yield with
increasing case volume.[11] In direct comparison,
Wang and colleagues found no difference in the
number of lymph nodes retrieved via open or robotic
approach (20 vs. 17).[12] A recent prospective,
randomized, non-inferiority study by Nix et al.
demonstrated a mean lymph node yield of 19 in
the robotic-assisted group vs. 18 in the open group.
[13] Second look [open] lymphadenectomy by a
different experienced open surgeon showed minimal
additional lymph node yield (range 0 8) to the
previous median 43 lymph nodes removed by a robotassisted approach.[14] Collaborative evaluation by
the International Robotic Cystectomy Consortium,
has demonstrated a mean lymph node yield of 19
with high volume centers (>100 cases per year)
having the highest yields and high surgeon volume
(>50 cases) independently predicting extended
(rather than standard) lymphadenectomy.[15]

experienced significantly fewer total complications


and major complications at 30 and 90 days.[26]
Moreover, being operated with robot-assistance was
an independent predictor of fewer overall and major
complications.
Regardless, published series are limited in patient
number and long-term follow-up. Although the only
randomized controlled trial of radical cystectomy
was designed as a non-inferiority study regarding
lymphadenectomy and node retrieval, it also
reported no difference in complications between the
2 small groups (n=20),[13] although in this study this
endpoint was underpowered to clearly define this
relationship. Reporting of complications following
cystectomy is also limited by patient selection bias,
possible under-reporting, and non-standardized
documentation.[28] In an attempt to overcome these
biases, a group from MSKCC reported complications
on 156 consecutive (non-selected) patients who
underwent robot-assisted radical cystectomy using
standardized methodology and fulfilling the Martin
criteria for adequate reporting of adverse events
following surgery. The 90-day complication rate of
52% and overall complication rate of 65% at median
follow-up of 9 months are similar to open series.[27]
The 90-day complication-related mortality rate was
2.6% and the majority of complications were low
grade (63%). Careful reporting of complications
following the robot-assisted approach revealed rates
similar to high volume open cystectomy centers.

The number of lymph nodes retrieved depends


on node viability, method of submission (en bloc
or separately), and the processing technique. In
conclusion, a thorough anatomical dissection around
the pelvic vessels and complete clearance of all
nodal tissue within the anatomical boundaries using
a minimally invasive approach is advocated.

6. Quality of Life
Health-related quality of life (HRQoL) outcomes after
surgery remain critical for measuring the impact of
any surgical modality. Studies investigating HRQoL
in patients undergoing radical cystectomy and
urinary diversion are lacking proper credence and
power.[30] Many of the studies are retrospective or
cross-sectional by design and do not use validated
questionnaires, while lacking baseline or preoperative
assessment.

5. Complications
Despite preoperative optimization, advances
in surgical techniques, and postoperative care,
radical cystectomy remains a morbid operation
with complication rates of up to 2664% and
mortality rates of 1-7%.[16-19] Radical cystectomy
readmission rates are high with bowel-related
and urinary infectious complications being most
common.[20] One of the major attractions towards a
minimally invasive approach to radical cystectomy is
that the morbidity of this procedure could possibly be
reduced. Estimated blood loss reported in minimally
invasive series ranges from 1001100 mL, with most
series reporting a mean of less than 500 mL and
transfusion rates below 2%. Reported complication
rates following a robot-assisted approach are
similar to an open approach at 6-65%. Two small
prospective comparisons of open and robot-assisted
cases showed similar complication rates between
the two approaches. In a prospective comparison
of complications in 187 consecutive patients who
underwent radical cystectomy with either open or
robot-assisted approach, the robot-assisted group

Gilbert et al evaluated HRQoL outcomes for patients


with bladder cancer using Bladder Cancer Index (BCI)
in 315 patients. Domains of the BCI included urinary,
sexual, and bowel function and bother domain.[31]
Patients undergoing radical cystectomy had lower
sexual function scores than patients who kept their
native bladder. The difference in urinary and bowel
domains between cystectomy and non-cystectomy
patients differed by type of urinary diversion, ileal
conduit or neobladder.
As robot-assisted radical cystectomy is a relatively
new technique, there is a paucity of literature
evaluating HRLQoL in patients undergoing this
treatment modality. Yuh et al prospectively evaluated
short-term HRQoL outcomes after robot-assisted
radical cystectomy and ileal conduit diversion
using the FACT-BL questionnaire.[32] The FACT-

282

BL questionnaire includes 5 domains: well-being,


physical, social/family, emotional, and functional
as well as 12 additional bladder cancer specific
questions. Thirty-four patients were included in the
study with follow-up questionnaires at 1, 3, and
6-month postoperative time periods. As expected,
scores decreased significantly at the initial period
with improvement to baseline at the 6-month period.
Emotional domains improved almost immediately
after surgery and exceeded baseline scores at 6
months.

extravesical disease and 26% (195) had positive


lymph nodes.[38]

8. Intracorporeal Urinary Diversion


Haber et al compared outcomes of patients
undergoing complete laparoscopic intracorporeal
conduit and their modified open assisted
laparoscopic approach.[39] Specific technical
difficulties with the completely intracorporeal group
included a high anastomotic leak rate and long
operative times (mean 9.3 hours). This approach
was eventually abandoned for an open-assisted
laparoscopic approach.

It is important to keep in mind that much of the longterm morbidity of radical cystectomy is associated
with urinary diversion, not extirpation. There is
currently no evidence regarding quality of life after
intracorporeal urinary diversion.

Robot-assisted surgery has helped overcome


some of the difficulties in performing intracorporeal
urinary diversion. Two small series have described
robot-assisted radical cystectomy with the creation
of an intracorporeal ileal conduit. The construction
of an intracorporeal conduit was performed with
combination of both robotic and laparoscopic
techniques. The robot was undocked after completion
of the radical cystectomy and construction of the
ileal conduit was performed in a pure laparoscopic
fashion. The robot was repositioned and used to
perform the ureterointestinal anastomosis. When the
series are combined, ileus was the only postoperative
complication in 1 of 5 patients. Operative times
ranged from 10-13 hours and stay ranged from 5-10
days (mean 6.8).

7. Oncologic Outcomes
The long-term oncologic efficacy of robot-assisted
radical cystectomy has yet to be determined,
however, surrogate markers of oncologic outcomes
which include margin status and lymph node yield
have been described earlier in this text. Similar to
laparoscopic radical cystectomy, early studies of the
robot-assisted approach suffer from selection bias
including younger patients, lower stage, and minimal
comorbidities compared to most contemporary
open series.[33-35] That selection bias makes
interpretation of oncologic outcomes difficult.

Hayn et al recently presented a series of 96 patients


who underwent robot-assisted radical cystectomy with
creation of either an extracorporeal or intracorporeal
ileal conduit.[42] There were no differences in age,
body mass index, ASA score, estimated blood loss,
or pathologic stage between the two groups. Mean
diversion time was similar between the intracorporeal
and extracorporeal group (123 vs. 136 min) although
mean overall operative time was shorter in the
intracorporeal group (390 vs. 356 min). Thirty and
ninety-day complication rates were similar between
the two groups and there was no difference in the
grade of complications. There were no statistically
significant differences in bowel-related complications
between the extracorporeal and intracorporeal
groups (25% and 12%, respectively). Patients in the
intracorporeal group, however, had better short-term
body image scores compared to the extracorporeal
group.

Previous open series have reported that patients


who experience recurrence after cystectomy do so
at a median of 12 months.[3] Martin described shortterm survival of patients undergoing robot-assisted
radical cystectomy. Fifty-nine patients with at least
6-month follow-up were included in that study with
mean follow-up of 25 months. Clinical characteristics
of these patients appear to be similar to that of open
series with 27% of patients having extravesical
disease and 34% of patients having lymph node
positive disease.[29] Overall survival at 12, 24, and
36 months was 82, 72%, and 72%, respectively.
Recurrence-free survival was 82%, 72%, and 72%,
respectively. This is comparable to early results of
modern open series studies. The retrospective nature
of this study with potential selection bias hampers
direct comparisons at this time and further studies
will need to be performed to evaluate the oncologic
outcomes of minimally invasive approaches.

Even smaller series have reported the creation of


urinary diversions performed in an intracorporeal
fashion.
Intracorporeal
continent
cutaneous
diversions and pouches have been reported but
are rare. Gaboardi, in 2002, described the first
intracorporeal neobladder.[43]

Kaufmann et al described early oncologic outcomes


of 85 patients who underwent robot-assisted radical
cystectomy.[37] Two-year recurrence-free survival,
cancer-specific, and overall survival rates were
74%, 85%, and 79%, respectively. These outcomes
are similar to those described by Martin.[29] Recent
reports from the International Robotic Cystectomy
Consortium of 820 patients revealed overall survival
rates of 82% and 68% at 1 and 2 years, respectively.
Thirty-seven percent (288) of those patients had

Pruthi et al described a series of 8 ileal conduits and


3 orthotopic neobladders performed intracorporeally.
Estimated blood loss, recovery of bowel function,
and time to discharge were all comparable to their
series of extracorporeal urinary diversions [44].

283

A recent study of intracorporeal urinary diversions


from the International Robotic Cystectomy
Consortium included 73 patients who underwent ileal
conduit formation and 61 patients who underwent
neobladder creation.[45] ]The mean age of those
patients undergoing intracorporeal neobladder was
60. Mean diversion time was longer in the neobladder
group compared to the conduit group (224 min vs.
138 min). The 90-day complication rate was 42% for
both groups with the majority of the complications
being low grade. The neobladder group had a lower
risk of both overall complications and high grade
complications. Improvements in technology and
intracorporeal techniques may allow urologists to
become more facile with completely intracorporeal
urinary diversion, thus improving outcomes in the
future.

8. Hellenthal NJ, Hussain A, Andrews PE, et al. Surgical margin


status after robot assisted radical cystectomy: results from
the International Robotic Cystectomy Consortium. J Urol.
Jul 2010;184(1):87-91.
9. Yuh B, Padalino J, Butt ZM, et al. Impact of tumour volume
on surgical and pathological outcomes after robot-assisted
radical cystectomy. BJU Int. Sep 2008;102(7):840-843.
10. Stein JP, Cai J, Groshen S, Skinner DG. Risk factors for
patients with pelvic lymph node metastases following
radical cystectomy with en bloc pelvic lymphadenectomy:
concept of lymph node density. J Urol. Jul 2003;170(1):3541.
11. Guru KA, Sternberg K, Wilding GE, et al. The lymph node
yield during robot-assisted radical cystectomy. BJU Int. Jul
2008;102(2):231-234; discussion 234.
12. Wang GJ, Barocas DA, Raman JD, Scherr DS. Robotic
vs open radical cystectomy: prospective comparison of
perioperative outcomes and pathological measures of early
oncological efficacy. BJU Int. Sep 20 2007.
13. Nix J, Smith A, Kurpad R, Nielsen ME, Wallen EM, Pruthi
RS. Prospective randomized controlled trial of robotic versus
open radical cystectomy for bladder cancer: perioperative
and pathologic results. Eur Urol. Feb 2010;57(2):196-201.

Recommendations:

14. Davis JW, Gaston K, Anderson R, et al. Robot assisted


extended pelvic lymphadenectomy at radical cystectomy:
lymph node yield compared with second look open
dissection. J Urol. Jan 2011;185(1):79-83.

R
 obot-assisted radical cystectomy is a surgical
option for locally advanced bladder cancer.
Larger series and long-term oncologic data are
still awaited and will define its permanent role in
urologic oncology. (LE 2a, GR: B)

15. Marshall SJ, Hayn MH, Stegemann AP, Argawal PK,


Badani KK, Balbay MD. Lymph node yield and predictors
of extended lymphadenectomy at the time of robotassisted radical cystectomy: Results from the International
Cystectomy Consortium. American Urologic Association
Annual Meeting. Chicago, IL2011.

H
 igh volume centers with dedicated minimally
invasive surgical teams have shown better results
than smaller centers. (LE: 2c, GR: C)

16. Chang SS, Cookson MS, Baumgartner RG, Wells N,


Smith JA, Jr. Analysis of early complications after radical
cystectomy: results of a collaborative care pathway. J Urol.
May 2002;167(5):2012-2016.
17. Quek ML, Stein JP, Daneshmand S, et al. A critical analysis
of perioperative mortality from radical cystectomy. J Urol.
Mar 2006;175(3 Pt 1):886-889; discussion 889-890.

References

18. Shabsigh A, Korets R, Vora KC, et al. Defining early


morbidity of radical cystectomy for patients with bladder
cancer using a standardized reporting methodology. Eur
Urol. Jan 2009;55(1):164-174.

1. Sanchez de Badajoz E, Gallego Perales JL, Reche


Rosado A, Gutierrez de la Cruz JM, Jimenez Garrido A.
Laparoscopic cystectomy and ileal conduit: case report. J
Endourol. Feb 1995;9(1):59-62.

19. Novara G, De Marco V, Aragona M, et al. Complications and


mortality after radical cystectomy for bladder transitional
cell cancer. J Urol. Sep 2009;182(3):914-921.

2. Davis JW, Castle EP, Pruthi RS, Ornstein DK, Guru KA.
Robot-assisted radical cystectomy: an expert panel review
of the current status and future direction. Urol Oncol. SepOct 2010;28(5):480-486.

20. Stimson CJ, Chang SS, Barocas DA, et al. Early and late
perioperative outcomes following radical cystectomy: 90day readmissions, morbidity and mortality in a contemporary
series. J Urol. Oct 2010;184(4):1296-1300.

3. Stein JP, Lieskovsky G, Cote R, et al. Radical cystectomy in


the treatment of invasive bladder cancer: long-term results
in 1,054 patients. J Clin Oncol. Feb 1 2001;19(3):666-675.

21. Menon M, Hemal AK, Tewari A, et al. Nerve-sparing robotassisted radical cystoprostatectomy and urinary diversion.
BJU Int. Aug 2003;92(3):232-236.

4. Dotan ZA, Kavanagh K, Yossepowitch O, et al. Positive


surgical margins in soft tissue following radical cystectomy
for bladder cancer and cancer specific survival. J Urol. Dec
2007;178(6):2308-2312; discussion 2313.

22. Hemal AK, Abol-Enein H, Tewari A, et al. Robotic radical


cystectomy and urinary diversion in the management of
bladder cancer. Urol Clin North Am. Nov 2004;31(4):719729, viii.

5. Herr HW, Faulkner JR, Grossman HB, et al. Surgical


factors influence bladder cancer outcomes: a cooperative
group report. J Clin Oncol. Jul 15 2004;22(14):2781-2789.

23. Abraham JB, Young JL, Box GN, Lee HJ, Deane LA,
Ornstein DK. Comparative analysis of laparoscopic and
robot-assisted radical cystectomy with ileal conduit urinary
diversion. J Endourol. Dec 2007;21(12):1473-1480.

6. Herr H, Lee C, Chang S, Lerner S. Standardization of


radical cystectomy and pelvic lymph node dissection for
bladder cancer: a collaborative group report. J Urol. May
2004;171(5):1823-1828; discussion 1827-1828.

24. Murphy DG, Challacombe BJ, Elhage O, et al. Roboticassisted laparoscopic radical cystectomy with extracorporeal urinary diversion: initial experience. Eur Urol. Sep
2008;54(3):570-580.

7. Haber GP, Crouzet S, Gill IS. Laparoscopic and robotic


assisted radical cystectomy for bladder cancer: a critical
analysis. Eur Urol. Jul 2008;54(1):54-62.

284

25. Pruthi RS, Wallen EM. Is robotic radical cystectomy an


appropriate treatment for bladder cancer? Short-term
oncologic and clinical follow-up in 50 consecutive patients.
Urology. Sep 2008;72(3):617-620; discussion 620-612.

outcomes between intra-corporeal and extra-corporeal


ileal conduit after robot-assisted radical cystectomy. 2011
Annual AUA Meeting. Vol Abstract. Washington, DC2011.
43. Gaboardi F, Simonato A, Galli S, Lissiani A, Gregori A,
Bozzola A. Minimally invasive laparoscopic neobladder. J
Urol. Sep 2002;168(3):1080-1083.

26. Ng CK, Kauffman EC, Lee MM, et al. A comparison


of postoperative complications in open versus robotic
cystectomy. Eur Urol. Feb 2010;57(2):274-281.

44. Pruthi RS, Nix J, McRackan D, et al. Robotic-assisted


laparoscopic intracorporeal urinary diversion. Eur Urol. Jun
2010;57(6):1013-1021.

27. Hayn MH, Hellenthal NJ, Hussain A, Stegemann AP, Guru


K. Defining morbidity of robot-assisted radical cystectomy
using a standardized reporting methodology. Eur Urol.
2011;59(2):213-218.
28. Donat SM. Standards for surgical complication reporting
in urologic oncology: time for a change. Urology. Feb
2007;69(2):221-225.

45. Guru K, Hayn MH, Stegemann AP, et al. Total intra-corporeal


urinary diversion after robot-assisted radical cystectomy:
The International Robotic Cystectomy Consortium results
after 134 cases. 2011 Annual AUA meeting. Washington,
DC 2011.

29. Martin RC, 2nd, Brennan MF, Jaques DP. Quality of


complication reporting in the surgical literature. Ann Surg.
Jun 2002;235(6):803-813.

2. Surgical Outcome: Morbidity and


Mortality

30. Porter MP, Penson DF. Health related quality of life after
radical cystectomy and urinary diversion for bladder cancer:
a systematic review and critical analysis of the literature. J
Urol. Apr 2005;173(4):1318-1322.

a) Introduction
Complications related to radical cystectomy may be
directly related to pre-existing patient comorbidities,
the surgical procedure, the bowel anastomosis, or the
urinary diversion. Variables such as hospital volume,
case mix, and surgeon skill and experience influence
the rate, type, and severity of surgical complications.
Other factors, including the availability and breadth
of consultative, diagnostic, and ancillary services,
may explain the association between cystectomy
volume and short-term surgical outcomes.[1]

31. Gilbert SM, Wood DP, Dunn RL, et al. Measuring healthrelated quality of life outcomes in bladder cancer patients
using the Bladder Cancer Index (BCI). Cancer. May 1
2007;109(9):1756-1762.
32. Yuh B, Butt Z, Fazili A, et al. Short-term quality-of-life
assessed after robot-assisted radical cystectomy: a
prospective analysis. BJU Int. Mar 2009;103(6):800-804.
33. Pruthi RS, Nielsen ME, Nix J, Smith A, Schultz H, Wallen
EM. Robotic radical cystectomy for bladder cancer: surgical
and pathological outcomes in 100 consecutive cases. J
Urol. Feb 2010;183(2):510-514.

When reporting surgical complications during


cystectomy, regardless of the technique, a
standardized and reproducible classification of
surgical complications is applied. The modified
Clavien system is one such paradigm and has been
used to evaluate complications in more than 6300
surgical procedures.[2] Recently, complications of
both open [3] and laparoscopic radical cystectomy[4]
have been reported using the modified Clavien
system. When possible, overall mortality in the
perioperative period should be captured and
reported for both the 30- and 90-day postoperative
period.[5] Morbidity occurring within 90 days should
be considered an early complication, while those
complications after 90 days are considered late.
[6,7] Despite improvements in surgical technique,
anesthetic delivery, and peri-operative care, radical
cystectomy remains a challenging procedure.
Adverse events of any grade may occur in up to 58%
of patients and mortality rates up to 3.9% occur at 30
days after cystectomy. [8, 9, 10, 11]

34. Cha EK, Wiklund NP, Scherr DS. Recent advances in


robot-assisted radical cystectomy. Curr Opin Urol. Jan
2011;21(1):65-70.
35. Hautmann RE. The oncologic results of laparoscopic radical
cystectomy are not (yet) equivalent to open cystectomy.
Curr Opin Urol. Sep 2009;19(5):522-526.
36. Stein JP, Skinner DG. Results with radical cystectomy for
treating bladder cancer: a reference standard for highgrade, invasive bladder cancer. BJU Int. Jul 2003;92(1):1217.
37. Kauffman EC, Ng CK, Lee MM, Otto BJ, Wang GJ, Scherr
DS. Early oncological outcomes for bladder urothelial
carcinoma patients treated with robotic-assisted radical
cystectomy. BJU Int. Feb 2011;107(4):628-635.
38. Hayn MH, Stegemann AP, Agarwal PK, et al. Pathologic
and early oncologic outcomes after robot-assisted radical
cystectomy: Results from the International Robotic
Cystectomy Consortium. 2011 AUA Annual Meeting.
Washington, DC2011.
39. Haber GP, Campbell SC, Colombo JR, Jr., et al.
Perioperative outcomes with laparoscopic radical
cystectomy: pure laparoscopic and open-assisted
laparoscopic approaches. Urology. Nov 2007;70(5):910915.

The following will detail rates of morbidity and mortality


associated with cystectomy and urinary diversion,
as well as candidate quality of care indicators for
treatment of bladder cancer and recommendations
from the literature to minimize complications.

40. Yohannes P, Chiou RK, Pelinkovic D. Rapid communication:


pure robot-assisted laparoscopic ureteral reimplantation for
ureteral stricture disease: case report. J Endourol. Dec
2003;17(10):891-893.
41. Balaji KC, Yohannes P, McBride CL, Oleynikov D, Hemstreet
GP, 3rd. Feasibility of robot-assisted totally intracorporeal
laparoscopic ileal conduit urinary diversion: initial results of
a single institutional pilot study. Urology. Jan 2004;63(1):5155.

b) Morbidities
1. Comorbidities

42. Hayn MH, Stegemann AP, Consiglio J, Wilding G, Guru K.


Comparison of complications and validated quality of life

There is a significant body of literature evaluating

285

patient age as a prognostic indicator for radical


cystectomy. [12,13] In general, advanced age is a
risk factor for the development of complications due
to radical cystectomy. However, chronologic age is
less the determinant than physiologic age. Other
risk factors for morbidity include female gender, prior
abdominal surgery, extravesical disease, and prior
radiotherapy.[11] Furthermore, elevated body mass
index is associated with an increased rate of wound
dehiscence and hernia.[14] Unfortunately, most
series evaluating radical cystectomy do not include
indices of morbidity in the patient evaluation. Suffice it
to say, patients with pre-existing neurologic disease,
cardiopulmonary compromise, renal insufficiency,
autoimmune disease, and bowel disease experience
higher rates of complication. [11]

bowel anastomosis, or the diversion. Paralytic ileus


is quite common during the postoperative course
of many cystectomy patients, plaguing patients as
much as 22.7% of the time. [22] Fortunately, true
small bowel obstructions or anastomotic leak are less
common, but do occur in up to 8.7% of patients. [17]
Lymphocele rates vary on the degree of dissection
utilized during pelvic lymph node dissection. With
that said, an appropriate lymph node dissection may
carry a risk of symptomatic lymphocele in up to 5%
of patients. [18] Rates of wound infection, incisional
hernia, pelvic hematoma, and fascial dehiscence are
widely variable, but may occur in as many as 9% of
patients. [23]

3. Laparoscopic Radical Cystectomy


Data from laparoscopic and robotic cystectomy
series are relatively immature and with small patient
populations. Nonetheless, complication rates are
similar to open series and demonstrate similar shortterm oncologic control. Adverse events are typically
of the same variety as those that occur during open
surgery. The typical rate of any form of complication
is 30-35% in the literature. [4,24] Not surprisingly,
in some series, there does appear to be a slightly
greater rate of higher grade complications earlier in
a surgeons experience. [24]

2. Perioperative Complications
Intraoperative complications may include acute
blood loss requiring transfusion and injury to adjacent
structures. Acute bleeding during radical cystectomy
is common and is typically associated with ligation of
the bladder pedicles or dorsal vein of the prostate in
men and excision of the anterior vagina in women.
The development of bipolar cautery devices,
surgical staplers, and improved understanding of the
prostatic anatomy have helped to reduce blood loss
during cystectomy. Regardless, average blood loss
during the procedure ranges from 600 mL to 1700
mL in large series. [14, 15, 16] Transfusion rates
are as high as 66% in some series. [17] Injuries
to associated structures, such as the rectum, may
occur in as many as 1.7% of procedures.[18]

4. Urinary Diversion
Complications due to the urinary diversion vary
depending on the type of diversion and may occur
in either an early or late fashion. Early complications
may occur in the form of urine leak, pouch leak,
excessive mucus, urethro-enteric stricture, and
ureteroenteric stricture. Urine leak from either a
pouch or ureteroenteric anastomosis is noted in as
many as 7.7% of patients. [25] Typically, prevention
involves the placement of suction drains or ureteral
catheters until adequate time for healing. However,
there is debate regarding the necessity of routine
use and duration of ureteral catheters. Ureteral
stricture may be an early or late complication and has
occurred in up to 14% of patients in some series. [26]
The etiology of the stricture may be benign or, more
concerning, a malignant recurrence. Methods for
treatment may include percutaneous nephrostomy
with antegrade stenting, ureteroscopic balloon
dilation, or open revision. The type of anastomosis
(Bricker vs. Wallace), does not appear to effect the
incidence of ureteroenteric stricture. [27, 28]

Medical complications comprise a large proportion


of the morbidity experienced by patients after
radical cystectomy. These complications include
thromboembolic, cardiac, pulmonary, infectious,
and renal adverse events. The rates of deep venous
thrombosis (DVT) and pulmonary embolus (PE)
are up to 5%. [19] Prophylaxis in the form of low
molecular weight heparin has been shown to reduce
the rates of both DVT and PE. [20] Cardiac events
such as congestive heart failure, arrhythmia, and
myocardial infarction occur in as many as 7% of
patients. [16] Pulmonary compromise in the form of
acute respiratory distress, reintubation, or pneumonia
complicate the postoperative course of up to 7.8%
of cystectomy patients. [21] As many as 13% of
radical cystectomy patients develop an infectious
complication such as pyelonephritis, sepsis, wound
infection, or UTI.[20] It is quite common for patients
to have a colonized urine specimen. However,
symptomatic urinary tract infections and pouchitis
require treatment with appropriate antibiotics. Lastly,
renal insufficiency requiring dialysis may occur in as
many as 7% of patients in some series. [17]

The urinary diversion may be the etiology of a variety


of late complications in the radical cystectomy patient,
which may be benign due to scar from ischemia,
technical error, or recurrent tumor. Stomal stenosis
has been described in as many as 1.7% of patients
and is likely related to ischemia to the conduit.
Stomal hernia is a more common occurrence,
occurring in up to 5.2% of cases. [21] Both entities
may require open revision; the latter may require

Surgical complications may arise from the


cystectomy, the pelvic lymph node dissection, the

286

transfer to the contralateral side or reinforcement with


mesh. Worsening renal function, hydronephrosis,
and ureteral reflux may occur in as many as 50%
of patients at 15 years. [25] This suggests the
importance of long-term follow-up, serial imaging,
and laboratory assessments. Metabolic changes
are noted in up to 3% of patients with urinary
diversions and include Vitamin B12 deficiency,
metabolic acidosis, and electrolyte derangements.
Concomitant urinary stone disease may occur due
to these metabolic changes. Furthermore, chronic
bacterial colonization, mucus production, urinary
retention, and enteric hyperoxaluria may exacerbate
stone formation in patients with urinary diversions.
Rates of stone formation may approach 30% in
some series. [25]

reported in a uniform grading system. Currently


the best adapted graded system for cystectomy is
the Clavien grading system (LE: 2, GR: B).

Surgical complications associated with radical


cystectomy and urinary diversion should include
the length of follow-up for the patient cohort and
a minimum of 30-day, but preferably 90-day,
reported outcome (LE 3, GR C).

A
 SA score, age, previous treatment for bladder

cancer or other pelvic diseases, surgeon and


hospital volume of radical cystectomy, and type
of urinary diversion influence surgical outcome.
(LE: 3, GR: C)

Reduction in blood loss and blood transfusion is

afforded by meticulous technique, use of modern


surgical devices, and improved understanding of
pelvic anatomy. (LE: 3, GR C)

5. Mortality
Clearly, given the surgical complexity and high
rates of surgical morbidity, it is not surprising that
mortality from radical cystectomy ranges as high
as 3.9% in larger series. [15] Additionally, the rate
of mortality climbs for those patients with greater
morbidity or advanced age. Recently, Morgan and
colleagues performed a retrospective analysis of
220 patients over the age of 75 who underwent
radical cystectomy at a high volume center. They
noted a sobering 90-day mortality rate of 12.7%. In
the multivariate analysis, preoperative albumin was
as great a predictor of patient outcome as patient
age. [29]

Reduction of urinary extravasation and leak can


be achieved with careful closure of anastomosis or
pouch, stenting of the ureteroenteric anastomosis,
and maintenance of appropriate drainage. (LE: 3,
GR: C)

Reduction of symptomatic lymphocele formation

can be achieved with appropriate identification of


lymphatic channels, careful surgical technique,
and an open peritoneal window. Initial treatment
should begin with percutaneous drainage. (LE: 3,
GR: C)

6. Quality of Care Indicators

Reduction of anastomotic strictures requires

The treatment and management of bladder cancer is


complex and challenging. Metrics are being developed
to measure the quality of healthcare provided by
physicians and will likely play an increasing role in
healthcare delivery in the future. Cooperberg and
colleagues recently defined candidate measures
for quality of care in the treatment of bladder cancer
and their relationship to surgical outcomes.[30]
They note that time to cystectomy after diagnosis,
hospital volume, surgeon volume, nodal yield, and
utilization of orthotopic diversion are associated
with improved healthcare outcomes in patients
receiving radical cystectomy. However, despite
clear recommendations from multiple urologic
associations, compliance with treatment guidelines
range from a dismal 3% for postoperative mitomycin
C to 20% for surveillance cystoscopy and cytology.
[31] Likely, improved adherence to recommended
guidelines and recognition of important quality of
care measures can reduce the substantial morbidity
of radical cystectomy for bladder cancer.

meticulous surgical technique, minimal ureteral


dissection, well-perfused segment, copious
spatulation, and careful apical suture placement.
(LE: 3, GR: C)

Reduction of metabolic disorders after urinary

diversion requires preservation of distal ileum,


serial monitoring of electrolytes and vitamin
B12 levels, understanding of bowel segment
physiology, and appropriate emptying of urinary
diversion. (LE: 3, GR: C)

Reduction of DVT and PE can be achieved

with use of low molecular weight heparin, early


ambulation, and sequential compression devices.
(LE: 1B, GR: B)

References
1. Hollenbeck BK, Daignault S, et al. Getting under the hood
of the volume-outcome relationship for radical cystectomy.
2007 J Urol 177(6): 2095-2099

Recommendations

2. Dindo D, Demartines N, et al. Classification of surgical


complications: a new proposal with evaluation in a cohort
of 6336 patients and results of a survey. 2004 Ann Surg
240(2): 205-213.

Surgical complications associated with radical


cystectomy and urinary diversion should be

287

3. Hautmann RE, de Petriconi, et al. Lessons learned from


1,000 neobladders: the 90-day complication rate. 2010 J
Urol 184(3): 990-994.

23. Arumainayagam N, McGrath J, Jefferson KP, Gillatt DA.


Introduction of an enhanced recovery protocol for radical
cystectomy. BJU Int 2008;101:698701

4. Khan MS, Elhage O, et al. Analysis of early complications


of robotic-assisted radical cystectomy using a standardized
reporting system. . 2011 Urology 77(2): 357-362.

24. Hayn MH, Hellenthal NJ, Mansour AM, et al. BJU Int. Is
patient outcome compromised during the initial experience
with robot-assisted radical cystectomy? Results of 164
consecutive cases. BJUI 2010 Dec 16 [Epub ahead of
print]

5. Stimson CJ, Chang SS, et al. Early and late perioperative


outcomes
following
radical
cystectomy:
90-day
readmissions, morbidity and mortality in a contemporary
series. 2010 J Urol 184(4): 1296-1300.

25. Hautmann RE, de Petriconi R, Gottfried HW, Kleinschmidt


K, Mattes R, Paiss T. The ileal neobladder: complications
and functional results in 363 patients after 11 years of
followup. J Urol 1999;161:4227

6. Hollenbeck BK, Miller DC, et al. Aggressive treatment for


bladder cancer is associated with improved overall survival
among patients 80 years old or older. 2004 Urology 64(2):
292-297.

26. Madersbacher S, Schmidt J, Eberle JM, et al. Long-term


outcome of ileal conduit diversion. J Urol 2003;169:985
90

7. Donat SM Standards for surgical complication reporting in


urologic oncology: time for a change. 2007 Urology 69(2):
221-225

27. Kouba E, Sands M, Lentz A, Wallen E, Pruthi RS. A


comparison of the Bricker versus Wallace ureteroileal
anastomosis in patients undergoing urinary diversion for
bladder cancer. J Urol 2007;178: 9458.

8. Novara G, De Marco V, et al. Complications and mortality


after radical cystectomy for bladder transitional cell cancer.
2009 J Urol 182(3): 914-921

28. Pagano S, Ruggeri P, Rovellini P, Bottanelli A. The anterior


ileal conduit: results of 100 consecutive cases. J Urol
2005;174:95962

9. Hautmann RE, de Petriconi RC, et al. Lessons learned


from 1,000 neobladders: the 90-day complication rate.
2010 J Urol 184(3): 990-994

29. Morgan TM, Keegan KA, Barocas DA, et al. Predicting the
Probability of 90-day Survival in Elderly Bladder Cancer
Patients Treated with Radical Cystectomy. J Urol 2011.
Epub ahead of print

10. Svatek RS, Fisher MB, et al. Risk factor analysis in a


contemporary cystectomy cohort using standardized
reporting methodology and adverse event criteria. 2010 J
Urol 183(3): 929-934.

30. Cooperberg MR, Porter MP, Konety BR. Candidate quality


of care indicators for localized bladder cancer. Urol Oncol
2009; 27:435-442

11. Lawrentschuk N, Colombo R, et al. Prevention and


management of complications following radical cystectomy
for bladder cancer. 2010 Eur Uro 57(6):983-1001

31. Chamie K, Saigal C, Hanley J, et al. Quality of care for


patients with high-grade non-muscle invasive bladder
cancer. J Urol 2010 183:568-569

12. Game X, Soulie M, Seguin P, et al. Radical cystectomy in


patients older than 75 years: assessment of morbidity and
mortality. Eur Urol 2001;39:5259.
13. Clark PE, Stein JP, Groshen SG, et al. Radical cystectomy
in the elderly: comparison of clinical outcomes between
younger and older patients. Cancer 2005;104:3643.
14. Novara G, De Marco V, Aragona M, et al. Complications and
mortality after radical cystectomy for bladder transitional
cell cancer. J Urol 2009;182:91421.
15. Bostrom PJ, Kossi J, Laato M, Nurmi M. Risk factors for
mortality and morbidity related to radical cystectomy. BJU
Int 2009;103: 1916.
16. Lowrance WT, Rumohr JA, Chang SS, Clark PE,
Smith Jr JA, Cookson MS. Contemporary open radical
cystectomy: analysis of perioperative outcomes. J Urol
2008;179:13138
17. Shabsigh A, Korets R, Vora KC, et al. Defining early
morbidity of radical cystectomy for patients with bladder
cancer using a standardized reporting methodology. Eur
Urol 2009;55:16476.
18. Chahal R, Sundaram SK, Iddenden R, Forman DF, Weston
PMT, Harrison SCW. A study of the morbidity, mortality and
long-term survival following radical cystectomy and radical
radiotherapy in the treatment of invasive bladder cancer in
Yorkshire. Eur Urol 2003;43:24657.
19. Novotny V, Hakenberg OW, Wiessner D, et al. Perioperative
complications of radical cystectomy in a contemporary
series. Eur Urol 2007;51:397402
20. Golhaber, HZ Risk factors of venous thromboembolism J
Am Coll Cardiol. 2010 Jun 29;56(1):1-7.
21. Nieuwenhuijzen JA, de Vries RR, Bex A, et al. Urinary
diversions after cystectomy: the association of clinical
factors, complications and functional results of four different
diversions. Eur Urol 2008;53:83444
22. Cookson MS, Chang SS, Wells N, Parekh DJ, Smith Jr JA.
Complications of radical cystectomy for nonmuscle invasive
disease: comparison with muscle invasive disease. J Urol
2003;169:1014

3. Oncological Outcome of Radical


Surgery According to TNM Staging
a) Survival According to AJCC/TNM Staging
Long-term oncological outcome of radical cystectomy
has been investigated in a multitude of studies which
derive mainly from high volume centers across
Europe, North Africa, and North America [1-3]. The
oncological outcomes reported in these series are
based on the TNM staging system, which consists
of three important parameters for survival. These
parameters are local tumor invasiveness (T-tumor
stage), presence of positive lymph nodes (N-nodal
stage), and distant metastatic disease (M-metastasis).
Based on this concept, patients can be categorized
according to six prognostic groups (0a, 0is, I, II, III,
IV) as proposed by the American Joint Committee on
Cancer (AJCC) [4, 5]. The TNM staging system can
be used for determining the clinical and pathological
stage of patients with invasive bladder cancer.
Tables 1 and 2 list the current 2009 TNM staging
system with the corresponding AJCC prognostic
groups [6].
One of the largest retrospective single-center series
reported oncological long-term outcomes in a total
of 1054 patients treated with radical cystectomy and
pelvic lymphadenectomy[1]. In the entire cohort,
after a median follow-up of 10.2 years, recurrencefree and overall survival rates at 5 years were 68%

288

and 66%, respectively, and at 10 years 60% and


43%, respectively. In organ-confined disease (pTaT2b N0; AJCC stages 0a, 0is, I, and II) recurrencefree and overall survival rates at 5 years were 80%
and 74%, respectively, and 77% and 54% at 10
years, respectively. By contrast, in extravesical,
node-negative disease (pT3a-4a, pN0; AJCC stage
III), recurrence-free and overall survival at 5 years
dropped to 58% and 47%, respectively, and to 55%
and 27% after 10 years, respectively. Among 236
patients with histologically confirmed lymph-nodepositive bladder cancer (stage IV according to AJCC)
the 5-/10-year recurrence-free and overall survival
were only 35%/34% and 31%/23%, respectively
[1]. In the meantime, similar single- and multicenter
series have confirmed these first cutting-edge
results and led to conclude that radical cystectomy
is the mainstay of treatment in patients with muscleinvasive bladder cancer [2, 3, 7].

Table 1. 2009 TNM Staging (6)

T Category
TX: No primary tumor assessment
T0: No primary tumor detectable
Ta: Noninvasive papillary carcinoma
Tis: Carcinoma in situ (CIS)
T1: Invasion of the subepithelial connective tissue
T2: Invasion of the muscle layer
T2a: Invasion into the inner half of the muscle layer
T2b: Invasion into the outer half of the muscle layer
T3: Invasion of the perivesical fatty tissue
T3a: Microscopic perivesical invasion
T3b: Macroscopic perivesical invasion

b) C
 ontroversial Issues of the Current TNM
Staging System

T4: Invasion of organs or structures such as prostatic


stroma, the seminal vesicles, uterus, vagina, pelvic
wall, or abdominal wall

1. pT2 Substaging

T4a: Invasion of prostatic stroma, uterus, and/or


vagina

In 1997, the AJCC updated the TNM staging system


and introduced new substagings for the tumor stages
T2 and T3 [5]. The latest version was published in
2009 but without any changes to the former version
of 2002 [6]. Both substratifications were thought
to provide a better risk assessment for follow-up
strategies and improve counseling of patients for
adjuvant treatment options [4]. However, in patients
with node-negative, pT2a-T2b bladder cancer,
classified as AJCC stage II (4), recent retrospective
studies have challenged the prognostic importance
of substratifying pT2 tumors into those involving
the inner (T2a) or outer half of the detrusor muscle
(T2b) and suggested consolidating both substages
into one [8-10]. The largest of these series included
311 patients with pT2 bladder cancer but found that
pT2b classified patients still had a higher risk for
lymph node tumor involvement than pT2a classified
patients [10]. Furthermore, limitations of these
studies were that the extent of lymphadenectomy
and number of retrieved lymph nodes were not
exactly reported, which might have biased the final
survival analysis [10]. Additionally, patients with nonurothelial cell carcinoma or those who underwent
neoadjuvant chemotherapy were not excluded from
analysis [8, 9].

T4b: Invasion of the pelvic or abdominal wall


N Category
NX: Regional lymph nodes cannot be assessed due
to lack of information
N0: No regional lymph node spread
N1: Single lymph node metastasis in the true pelvic
region
N2: Multiple lymph nodes in the true pelvic region
N3: Single or multiple lymph nodes in the common
iliac region
M Category
MX: Distant metastases cannot be assessed
M0: No signs of distant metastasis
M1: Distant metastasis present
Table 2. AJCC Stages of Bladder Cancer (6)

Stage 0a:

Ta, N0, M0

Stage 0is:

Tis, N0, M0

Stage I:

T1, N0, M0

Stage II:

T2a- T2b, N0, M0

Stage III:

T3a-T4a, N0, M0

Therefore, a recent multicenter series on 565 patients


with pT2 urothelial carcinoma of the bladder has
attempted to overcome these limitations and reported
significant differences in survival between both
substages in node-negative pT2 disease [11]. These
findings were also confirmed in a mixed cohort of
1737 patients with pT2 bladder cancer of whom 54%
had squamous cell carcinoma [7]. In this study, the

Stage IV: T4b, N0, M0 or any T, N1 to N3,


M0 or any T, any N, M1

289

5-year disease-free survival was significantly higher


for patients with pT2aN0 (meanSE: 76.62.2%)
compared to those with pT2bN0 bladder cancer
(58.31.6%). In a recent series, this significant
difference in recurrence-free and cancer-specific
survival was also confirmed for patients with pT2
urothelial carcinoma of the bladder who were treated
with an extended pelvic lymphadenectomy approach
[12]. Moreover, another recent multicenter study
proposed a weighted prognostic model for patients
with node-negative pT2 bladder cancer. Among
different independent risk factors (presence of high
grade disease or lymphovascular invasion), pT2
substaging was the strongest one for recurrence-free
survival [13]. In conclusion, these data support the
prognostic importance of the current substratification
in node-negative pT2 bladder cancer.

5-year recurrence-free survival and cause-specific


survival for patients with pT3apN0 versus pT3bpN0
classified disease were not significantly different with
68% versus 72% and 54% versus 42%, respectively.
However, a larger multicenter study on 808 patients
with pT3N0 bladder cancer (median follow-up 43
months) treated with radical cystectomy without
any neoadjuvant modality reported significantly
improved 5-year recurrence-free survival (61% vs.
48%) and cause-specific survival (64% vs. 55%)
rates for patients with pT3a versus pT3b disease
[17]. This assumption is further supported by another
multicenter study in which a weighted prognostic
model was constructed for 578 patients with nodenegative pT3 bladder cancer. pT3 substaging was
found to independently contribute to a 2.05-fold
increase in the relative risk of decreased recurrencefree survival [20]. In conclusion, the present data
do support the current concept of substaging in
node-negative pT3 bladder cancer whereas in
node-positive pT3 disease a prognostic significance
cannot be attributed to this substratification [15, 18].

2. Definition of Organ-confined Bladder Cancer


The definition of organ-confined bladder cancer has
been controversially discussed in recent literature.
A prior series assumed that microscopic extension
of the tumor into the perivesical fat (pT3a) does
not significantly confer a higher risk for lymph node
involvement and decreased survival compared to
muscle-invasive bladder cancer [14]. This study
included 381 patients (pT2: 172 patients, pT3a: 88
patients, pT3b: 121 patients) and found no significant
difference for recurrence-free survival between the
stages pT2 versus pT3a, but for the stages pT2
versus pT3b. However, a trend towards significance
was reported for cancer-specific survival between
patients with pT2 versus pT3a disease (p=0.06)
[14]. By contrast, in another series with 134 pT2b
patients and 236 patients with pT3a or T3b bladder
cancer, recurrence-free and cancer-specific survival
was significantly improved for pT2b compared to
pT3a patients [15]. Moreover, a current study of the
SEER database analyzed the outcomes of 2238
patients with pT2b-T3b bladder cancer and found a
significantly higher rate of node-positive disease and
all-cause mortality in patients with node-negative
pT3a versus pT2b bladder cancer [16]. Regarding
these data, it seems anatomically intuitive to define
organ-confined muscle-invasive bladder cancer as
stages pT2bpN0cM0 or less.

4. P
 rognostication in Lymph Node Positive
Bladder Cancer
Oncological outcome in patients with node-positive
bladder cancer at radical cystectomy is reported
to be generally poor with a 5-year recurrence-free
survival ranging from 34-43% [1,3, 21]. Nonetheless,
long-term survival has been described especially in
patients with low-volume lymph node metastasis [2123]. The following pathological and clinical parameters
have been investigated as critical determinants
for survival in node-positive disease: number and
size of positive lymph nodes [22], extracapsular
extension of lymph node metastasis [24], number
of retrieved lymph nodes [25], aggregate lymph
node metastasis diameter [26], and the anatomic
extent of lymphadenectomy reflecting the surgical
meticulousness of the lymph node removal [27]. The
current pTNM staging system differentiates lymph
node tumor involvement according to the number
and size of positive lymph nodes (see table 1) but
does not reflect sufficiently the surgeons capability
of removing all the affected nodal tissue.
With growing evidence of the prognostic role of the
extent of lymphadenectomy for improving outcomes
in invasive bladder cancer [22, 27], the concept of
lymph node density (LND) has been proposed [28,
29]. It is defined as the number of positive lymph
nodes divided by the overall number of retrieved
lymph nodes. In most series, a cut-off value of
20% has been reported to statistically optimally
distinguish between different outcomes [25, 30, 31].
Various studies also found that LND outperformed
the pTNM based predictions (pN1-N3) in terms of
recurrence-free and disease-specific survival [28, 29,
31]. Even more, LND was a superior prognosticator
in node-positive patients after adjusting for the

3. pT3 Substaging
The prognostic significance of substaging patients
with pT3 bladder cancer into those with microscopic
(pT3a) and macroscopic (pT3b) perivesical fat
invasion as implemented by the 2002 TNM staging
system is also controversially discussed. Most larger
series found that the risk of lymph node metastases
was significantly higher for patients with pT3b
versus pT3a classified disease [15-17] while some
smaller series did not [18]. A recent single-center
analysis focussed on outcomes in 75 patients with
node-negative pT3 bladder cancer [19]. Actuarial

290

d) Stage pT4 Bladder Cancer

use of adjuvant chemotherapy [31, 32]. On the


contrary, Fleischmann et al. reported that LND lost
its independent prognostic value after accounting for
the presence of extracapsular extension of lymph
node metastasis [24]. Although the concept of LND
seems to be a promising alternative for improving the
prognostication of patients with lymph-node positive
bladder cancer, there are several unresolved issues
which hinder its unquestionable adoption into clinical
practice. Besides the retrospective design of the
studies in favor of the concept of LND, the proposed
cut-off values are based on statistical calculations
and are highly dependent on the surgical extent
and meticulousness of lymphadenectomy. Even
when considering a defined anatomical template of
extended pelvic lymphadenectomy, the number of
retrieved lymph nodes can vary between individuals
significantly [27] and this might have implications on
the proposed cut-off values. Furthermore, different
techniques in the pathological processing and
evaluation of lymph nodes have to be taken into
account [33]. Moreover, in the neoadjuvant setting,
LND was not found to be of prognostic value [34]. In
this respect, prospective trials are certainly needed to
address more specifically the role of LND in patients
with lymph-node positive bladder cancer.

Radical cystectomy in patients with T4 bladder


cancer has been shown to be a feasible option in
terms of therapy-related morbidity [38]. Nevertheless,
the oncological outcome of these patients treated
without neoadjuvant therapy is generally poor,
although some patients may achieve long-term
survival. Risk factors that determine independently
an adverse oncological outcome are the presence
of positive soft tissue surgical margins, lymph
node metastasis, lymphovascular invasion, tumors
infiltrating the abdominal or pelvic wall (staged as
pT4b), and female patients [39]. In this respect,
several series have demonstrated that a positive soft
tissue surgical margin per se is a strong independent
risk factor for decreased recurrence-free and cancerspecific survival [40, 41]. Especially female patients
are at increased risk for positive soft tissue surgical
margins at radical cystectomy [42]. However, since
complete tumor removal cannot be achieved by
surgery alone in case of abdominal or pelvic wall
infiltration, cystectomy should be preserved in this
stage for symptom relief, such as recurrent gross
hematuria, pain, fistula formation, and therapyrefractory urgency [43].

e) Impact of Lymphovascular Invasion on


Oncological Outcomes

c) Outcomes of Patients with Stage pT0 or


Carcinoma-in-situ-only Disease

Lymphovascular invasion in pathologically nodenegative bladder cancer was found to independently


predict worse cancer-specific and overall survival
after radical cystectomy, whereas in node-positive
disease its independent prognostic significance was
not confirmed [44, 45]. Therefore, in node-negative
disease, its presence might indicate micrometastasis
and, thus, help to improve risk assessment and
guide clinicians for counseling patients for adjuvant
treatment options or clinical trials [46]. However, when
assessing lymphovascular invasion on conventional
histological sections, its accurate detection can be
hampered due to retraction artifacts and difficulties
in identifying small lymphatic or blood vessels [47]
which, in turn, limits its prognostic value and stresses
the importance of additional immunohistochemical
studies.

The percentage of patients without evidence of


the primary tumor (classified as pT0) at radical
cystectomy has been shown to range from 5-7%
with a risk of concomitant lymph node metastasis
of approximately 3-7.5% [35, 36]. In one study,
the following clinical stages were reported at
preoperative transurethral resection in 56 patients
with final pT0 disease: Tis in 5 (9%), Ta in 2 (4%), T1
in 18 (32%), T2 in 29 (52%), and T3 in 2 (4%) (36).
The probability of remaining disease-free at 5 years
in patients with invasive bladder cancer (pT1-T2)
staged pT0 at radical cystectomy is approximately
89-90% and is significantly higher than in cases with
residual pT1-T2 bladder cancer [35, 36]. In a large
international study of 228 patients with pT0 bladder
cancer at final pathologic analysis, risk factors which
independently correlated with decreased survival
were the presence of lymph node metastasis and
female gender [35].

f) Molecular Markers and Outcome


The predictive value of molecular markers for risk
assessment in invasive bladder cancer has been
evaluated in a subset of series. These markers
play an important role in cell-cycle regulatory or
angiogenic mechanisms. The following markers
have been recently investigated: E-cadherin, pRB,
survivin, p53, p16, p21, p27, cyclin E, and Ki-67 [4850]. Expression levels of these markers are assessed
in cystectomy specimens by immunohistochemical
methods and the combination of different markers
have been shown to improve the predictive accuracy
of standard clinical and pathological risk factors [49].

For patients with carcinoma in situ only, who had


undergone radical cystectomy due to refractory
conservative treatment, the overall disease-free
and cancer-specific survivals were 74% and 85%,
respectively. However, 36% of these patients
experienced disease upstaging (pT1) at radical
cystectomy. Similar to patients with pT0 disease, risk
factors for decreased cancer-specific survival were
found to be the presence of lymph node metastases,
lymphovascular invasion, and female gender [37].

291

A recent analysis incorporated serum preoperative


C-reactive protein concentration as a serological
marker in a new predictive model, termed TNR-C
score, and showed that increased levels were not
only associated with decreased cancer-specific
survival but also with an increase in the predictive
ability of standard pathologic risk factors including
tumor stage, lymph node density, and resection
margin status [51]. Nonetheless, up to now, there
are no established molecular markers that can be
unequivocally recommended for risk assessment in
invasive bladder cancer on a routine basis.

nomogram, the standard predictors of the AJCC


based prediction model, the pT and pN stage, were
complemented by the following parameters: age,
gender, tumor grade at cystectomy, presence of
lymphovascular invasion, presence of carcinoma
in situ, neoadjuvant chemotherapy, and adjuvant
chemo- and radiotherapy. The addition of these
parameters increased significantly the predictive
accuracy of the BCRC nomogram by 3.2% as
compared to the AJCC-based predictions [56].
By contrast, the IBCNC nomogram relies on more
than 9000 cystectomy patients who were treated in
12 international centers worldwide. In this nomogram,
the following parameters have been added to the pT
and pN stage: age, gender, tumor grade, number
of days from diagnosis to cystectomy, and final
histology. This nomogram has shown to improve the
AJCC based predictions by approximately 7% [55].
The original data sets of both nomograms used 200bootstrap resamples for reducing overfit bias and
for interval validation. Currently, only a few smaller
series have externally validated these data but have
confirmed an approximately 4% improvement in the
predictive accuracy for both nomograms [58].

g) Impact of Clinical Parameters on Outcome:


Follow-up Duration, Surgical Expertise, and
Age
An ongoing discussion in invasive bladder cancer
is the validity of the reported survival analyses of
retrospective studies with short or intermediate
follow-up time interval. Considering the fact that
the median time to local or distant recurrence after
radical cystectomy in studies with a median followup of more than 10 years ranges from 7-18 months
[1], a recent study was set up to investigate the
validity of disease-free survival rates reported at 2
or 3 years after radical cystectomy. This study found
that disease-free survival rates reported at 2 or 3
years correlated well with the 5-year overall survival
regardless of the use of adjuvant chemotherapy
[52].

Nevertheless, some limitations must be taken


into account when addressing the patients
individual risk of recurrence and death by the use
of these nomograms. In the IBCNC nomogram, a
considerable number of patients with squamous cell
carcinomas were included but their primary clinical
and pathological parameters were not published,
which makes its general applicability difficult [55]. In
this respect, the BCRC nomogram provides detailed
patient data but the number of included patients is
relatively low. In addition, the independent prognostic
relevance of some of the included parameters
(i.e., adjuvant treatment modalities) currently is
controversial (see subsection 4.2). Nonetheless,
both nomograms can be regarded as an important
tool for estimation of outcomes in patients treated
with radical surgery.

Another important parameter that might influence


the outcomes of bladder cancer patients is surgical
expertise. A recent meta-analysis addressed the
ongoing debate on the relationship between highvolume centers and oncologic outcome. A positive
significant association on improved survival was not
found for hospital (HR 0.89; p=0.06) or for surgeon
volume (HR 0.83; p=0.26) [53].
Nonetheless, older patients might derive the highest
benefit when treated in a high-volume center [54]. In
this respect, the largest series on cystectomy to date
from the SEER database analyzed the outcomes of
13,796 bladder cancer patients and demonstrated
that patients above 80 years had an increased risk
for postoperative morbidity but not mortality. The
ability of patients older than 80 years to undergo
cystectomy had the highest impact on risk reduction
of cancer-related and non-cancer-related mortality
[43].

Recommendations
The AJCC substratifications in node-negative pT2
and pT3 bladder cancer are of prognostic value
(LE: 3, GR: C).
According to the TNM staging system, organconfined bladder cancer has to be defined as
pT2bN0M0 or less (LE: 3, GR: B).

h) Nomograms
In 2006, bladder cancer nomograms were published
from the Bladder Cancer Research Consortium
(BCRC) and the International Bladder Cancer
Nomogram Consortium (IBCNC), and are freely
available on the Internet [55-57]. The BCRC
nomogram is based on a total of 731 patients
treated in 3 North American institutions. In this

Nomograms
provide
improved
prognostic
information for oncological outcomes after
radical surgery as compared to the pTNM stage
predictions. However, its general applicability
has not been established sufficiently by external
validation (LE: 3, GR: B).

292

In patients older than 80 years, radical cystectomy


is associated with the highest risk reduction on
cancer-related and non-cancer-related mortality
(LE: 3, GR: C).

14. Bastian PJ, Hutterer GE, Shariat SF, Rogers CG, Palapattu
GS, Lotan Y, Vazina A, Amiel GE, Gupta A, Sagalowsky
AI, Lerner SP, Schoenberg MP, Karakiewicz PI; Bladder
Cancer Research Consortium ( 2008) Macroscopic, but not
microscopic, perivesical fat invasion at radical cystectomy
is an adverse predictor of recurrence and survival. BJU Int
101., 450-454

Based on the scarce data available, the routine


use of molecular markers for risk assessment after
radical cystectomy in invasive bladder cannot be
recommended (LE: 3, GR: B).

15. Quek ML, Stein SP, Clark PE, Daneshmand S, Miranda


G, Cai J, Groshen S, Cote RJ, Lieskovsky G, Quinn DI,
Skinner DG (2004) Microscopic and gross extravesical
extension in pathological staging of bladder cancer. J Urol
171, 640-645

References

16. Scosyrev E, Yoo J, Messing E (2010) Microscopic invasion


of perivesical fat by urothelial carcinoma: implications for
prognosis and pathology practice. Urology 76, 908-913

1. Stein JP, Lieshovsk G, Cote R, Groshen S, Feng AC,


Boyd S, Skinner E, Bochner B, Thangathurai D, Mikhail M,
Raghavan D, Skinner DG (2001) Radical cystectomy in the
treatment of invasive bladder cancer: long-term results in
1,054 patients. J Clin Oncol 19, 666-675

17. Tilki D, Svatek RS, Karakiewicz PI, Novara G, Seitz M,


Sonpavde G, Gupta A, Kassouf W, Fradet Y, Ficarra V,
Skinner E, Lotan Y, Sagalowsky AI, Stief CG, Reich O,
Shariat SF (2010) pT3 Substaging is a prognostic indicator
for lymph node negative urothelial carcinoma of the bladder.
J Urol 184, 470-474

2. Hautmann RE et al (2006) Urinary Diversion. Urology 69(1


Suppl. 1), 17-49
3. Madersbacher S, Hochreiter W, Burkhard F, Thalmann GN,
Danuser H, Markwalder R, Studer UE ( 2003) Radical cystectomy for bladder cancer today--a homogeneous series
without neoadjuvant therapy. J Clin Oncol 21, 690-696

18. Quek ML, Stein SP, Clark PE, Daneshmand S, Miranda


G, Cai J, Groshen S, Lieskovsky G, Quinn DI, Raghavan
D, Skinner DG (2003) Natural history of surgically treated
bladder carcinoma with extravesical tumor extension.
Cancer 98, 955-961
19. Boudreaux KJ Jr, Chang SS, Lowrance WT, Rumohr JA,
Barocas DA, Cookson MS, Smith JA Jr, Clark PE (2009)
Comparison of American Joint Committee on Cancer
pathologic stage T3a versus T3b urothelial carcinoma:
analysis of patient outcomes. Cancer 115, 770-775
20. Sonpavde G, Khan MM, Svatek RS, Lee R, Novara G,
Tilki D, Lerner SP, Amiel GE, Skinner E, Karakiewicz PI,
Bastian PJ, Kassouf W, Fritsche HM, Izawa JI, Scherr DS,
Ficarra V, Dinney CP, Lotan Y, Fradet Y, Shariat SF (2011)
Prognostic Risk Stratification of Pathological Stage T3N0
Bladder Cancer After Radical Cystectomy. J Urol Epub
ahead of print

4. Greene FL, Page DL, Fleming ID, Fritz A, Balch CM, Haller
DG, et al. (2002) AJCC Cancer Staging Manual. In pp. 335337, Lippincott-Raven, Philadelphia
5. Sobin LH, Wittekind C, (2002) TNM Classification of Malignant Tumours. In Vol. 199-202, Wiley-Liss, New York
6. Sobin LH, (2009) TNM Classification of Malignant Tumours.
In p. 263265, Wiley-Liss, New York
7. Ghoneim MA, Abdel-Latif M, E-Mekresh M, Abol-Enein
H, Mosbah A, Ashamallah A, el-Baz MA (2008) Radical
cystectomy for carcinoma of the bladder: 2,720 consecutive
cases 5 years later. J Urol 180, 121-127
8. Boudreaux KJ Jr, Clark PE, Lowrance WT, Rumohr JA,
Barocas DA, Cookson MS, Smith JA Jr, Chang SS (2009)
Comparison of american joint committee on cancer
pathological stage T2a versus T2b urothelial carcinoma:
analysis of patient outcomes in organ confined bladder
cancer. J Urol 181, 540-545

21. Bruins HM, Huang GJ, Cai J, Skinner DG, Stein JP, Penson
DF ( 2009) Clinical outcomes and recurrence predictors of
lymph node positive urothelial cancer after cystectomy. J
Urol 182, 2182-2187
22. Steven K, Poulsen AL, (2007) Radical cystectomy and
extended pelvic lymphadenectomy: survival of patients
with lymph node metastasis above the bifurcation of the
common iliac vessels treated with surgery only. J Urol 178,
1218-1223

9. Tokgoz H, Reich O, Turholmez K, Resorlu B, Kose K,


Tulunay O, Beduk Y (2007) Pathological staging of muscle
invasive bladder cancer. Is substaging of pT2 tumors really
necessary? Int Braz J Urol 33, 777-783
10. Yu RJ, Stein JP, Cai J, Miranda G, Groshen S, Skinner DG
(2006) Superficial (pT2a) and deep (pT2b) muscle invasion
in pathological staging of bladder cancer following radical
cystectomy. J Urol 176, 493-498

23. Dhar NB, Klein EA, Reuther AM, Thalmann GN,


Madersbacher S, Studer UE (2008 .) Outcome after radical
cystectomy with limited or extended pelvic lymph node
dissection. J Urol 179, 873-878

11. Tilki D, Reich O, Karakiewicz PI, Novara G, Kassouf W,


Ergn S, Fradet Y, Ficarra V, Sonpavde G, Stief CG,
Skinner E, Svatek RS, Lotan Y, Sagalowsky AI, Shariat
SF (2010) Validation of the AJCC TNM substaging of pT2
bladder cancer: deep muscle invasion is associated with
significantly worse outcome. Eur Urol 58, 112-117
12. Gakis G, Schilling D, Renninger M, Seibold J, Sievert KD,
Stenzl A (2010) Comparison of the new American Joint
Committee on Cancer substratification in node-negative
pT2 urothelial carcinoma of the bladder: analysis of patient
outcomes in a contemporary series. BJU Int doi:10.1111/
j.1464-410X.2010.09548.x
13. Sonpavde G, Khan MM, Svatek RS, Lee R, Novara G, Tilki
D, Lerner SP, Amiel GE, Skinner E, Karakiewicz PI, Bastian
PJ, Kassouf W, Fritsche HM, Izawa JI, Ficarra V, Dinney
CP, Lotan Y, Fradet Y, Shariat SF (2010) Prognostic risk
stratification of pathological stage T2N0 bladder cancer
after radical cystectomy. BJU Int doi: 10.1111/j.1464410X.2010.09902.x.]

24. Fleischmann A, Thalmann GH, Markwalder R, Studer


UE (2005) Extracapsular extension of pelvic lymph node
metastases from urothelial carcinoma of the bladder is an
independent prognostic factor. J Clin Oncol 23, 2358-2365
25. Fang AC, Ahmad AE, Whitson JM, Ferrell LD, Carroll PR,
Konety BR (2010) Effect of a minimum lymph node policy
in radical cystectomy and pelvic lymphadenectomy on
lymph node yields, lymph node positivity rates, lymph node
density, and survivorship in patients with bladder cancer.
Cancer 116, 1901-1908
26. Stephenson AJ, Gong MC, Campbell SC, Fergany AF,
Hansel DE (2010) Aggregate lymph node metastasis
diameter and survival after radical cystectomy for invasive
bladder cancer. Urology 75, 382-386
27. Dhar NB, Klein E, Reuther AM, Thalmann GN, Madersbacher
S, Studer UE (2008) Outcome after radical cystectomy with
limited or extended pelvic lymph node dissection. J Urol
179, 873-878

293

28. Stein JP, Cai J, Groshen S, Skinner DG (2003) Risk factors


for patients with pelvic lymph node metastases following
radical cystectomy with en bloc pelvic lymphadenectomy:
concept of lymph node density. J Urol 170, 35-41

40. Novara G, Svatek RS, Karakiewicz PI, Skinner E, Ficarra


V, Fradet Y, Lotan Y, Isbarn H, Capitanio U, Bastian PJ,
Kassouf W, Fritsche HM, Izawa JI, Tilki D, Dinney CP,
Lerner SP, Schoenberg M, Volkmer BG, Sagalowsky AI,
Shariat SF (2010) Soft tissue surgical margin status is a
powerful predictor of outcomes after radical cystectomy: a
multicenter study of more than 4,400 patients. J Urol 183.,
2165-2170

29. Herr HW, (2003) Superiority of ratio based lymph node


staging for bladder cancer. J Urol 169, 943-945
30. May M, Hermann D, Bolenz C, Tiemann A, Brookman-May
S, Fritsche HM, Burger M, Buchner A, Gratzke C, Wlfing
C, Trojan L, Ellinger J, Tilki D, Gilfrich C, Hfner T, Roigas
J, Zacharias M, Gunia S, Wieland WF, Hohenfellner M,
Michel MS, Haferkamp A, Mller SC, Stief CG, Bastian
PJ (2011) Lymph Node Density Affects Cancer-Specific
Survival in Patients with Lymph Node-Positive Urothelial
Bladder Cancer Following Radical Cystectomy. Eur Urol
doi:10.1016/j.eururo.2011.01.030

41. Hadjizacharia P, Stein JP, Cai J, Miranda G (2009) The


impact of positive soft tissue surgical margins following
radical cystectomy for high-grade, invasive bladder cancer.
World J Urol 27, 33-38
42. Dotan ZA, Kavanagh K, Yossepowitch O, Kaag M, Olgac S,
Donat M, Herr HW (2007) Positive surgical margins in soft
tissue following radical cystectomy for bladder cancer and
cancer specific survival. J Urol 178, 2308-2312

31. Kassouf W, Agar wal PU, Herr HW, Munsell MF, Spiess PE,
Brown GA, Pisters L, Grossman HB, Dinney CP, Kamat
AM (2008) Lymph node density is superior to TNM nodal
status in predicting disease-specific survival after radical
cystectomy for bladder cancer: analysis of pooled data
from MDACC and MSKCC. J Clin Oncol 26, 121-126

43. Hollenbeck BK, Miller DC, Taub D, Dunn RL, Underwood


W 3rd, Montie JE, Wei JT (2004) Aggressive treatment for
bladder cancer is associated with improved overall survival
among patients 80 years old or older. Urology 64., 292297
44. Lotan Y, Gupta A, Shariat SF, Palapattu GS, Vazina A,
Karakiewicz PI, Bastian PJ, Rogers CG, Amiel G, Perotte
P, Schoenberg MP, Lerner SP, Sagalowsky AI (2005)
Lymphovascular invasion is independently associated
with overall survival, cause-specific survival, and local and
distant recurrence in patients with negative lymph nodes at
radical cystectomy. J Clin Oncol 23, 6533-6539

32. Kassouf W, Leibovici D, Luongo T, Munsell MF, Vakar F,


Dinney CP, Grossman HB, Kamat AM. (2006) Relevance
of extracapsular extension of pelvic lymph node metastasis
in patients with bladder cancer treated in the contemporary
era. Cancer 107, 1491-1495
33. Stein JP, Penson DF, Cai J, Miranda G, Skinner EC, Dunn
MA, Groshen S, Lieskovsky G, Skinner DG (2007) Radical
cystectomy with extended lymphadenectomy: evaluating
separate package versus en bloc submission for node
positive bladder cancer. J Urol 177, 876-881

45. Shariat SF, Svatek RS, Tilki D, Skinner E, Karakiewicz PI,


Capitanio U, Bastian PJ, Volkmer BG, Kassouf W, Novara
G, Fritsche HM, Izawa JI, Ficarra V, Lerner SP, Sagalowsky
AI, Schoenberg MP, Kamat AM, Dinney CP, Lotan Y,
Marberger MJ, Fradet Y (2010) International validation of
the prognostic value of lymphovascular invasion in patients
treated with radical cystectomy. BJU Int 105, 1402-1412

34. Kassouf W, Agar wal PU, Grossman HB, Leibovici D,


Munsell MF, Siefker-Radtke A, Pisters LL, Swanson
DA, Dinney CP, Kamat AM (2009) Outcome of patients
with bladder cancer with pN+ disease after preoperative
chemotherapy and radical cystectomy. Urology 73, 147152

46. Bolenz C, Hermann E, Bastian PJ, Michel MS, Wlfing


C, Tiemann A, Buchner A, Stief CG, Fritsche HM, Burger
M, Wieland WF, Hfner T, Haferkamp A, Hohenfellner M,
Mller SC, Strbel P, Trojan L (2010) Lymphovascular
invasion is an independent predictor of oncological
outcomes in patients with lymph node-negative urothelial
bladder cancer treated by radical cystectomy: a multicentre
validation trial. BJU Int 106, 493-499

35. Tilki D, Svatek RS, Novara G, Seitz M, Godoy G, Karakiewicz PI, Kassouf W, Fradet Y, Fritsche HM, Sonpavde G,
Izawa JI, Ficarra V, Lerner SP, Schoenberg M, Stief CG,
Dinney CP, Skinner E, Lotan Y, Sagalowsky AI, Reich O,
Shariat SF (2010) Stage pT0 at radical cystectomy confers
improved survival: an international study of 4,430 patients.
J Urol 184, 888-894

47. ODonnell RK, Feldman M, Mick R, Muschel RJ (2008)


Immunohistochemical method identifies lymphovascular
invasion in a majority of oral squamous cell carcinomas
and discriminates between blood and lymphatic vessel
invasion. J. Histochem Cytochem 56, 803-810

36. Palapattu GS, Shariat SF, Karakiewicz PI, Bastian PJ, Rogers CG, Amiel G, Lotan Y, Vazina A, Gupta A, Sagalowsky
AI, Lerner SP, Schoenberg MP; Bladder Cancer Research
Consortium (2006) Cancer specific outcomes in patients
with pT0 disease following radical cystectomy. J Urol 175,
1645-1649

48. Youssef RF, Shariat SF, Kapur P, Kabbani W, Ghoneim T,


King E, Cockburn A, Mosbah A, Abol-Enein H, Ghoneim M,
Lotan Y (2011) Expression of cell cycle-related molecular
markers in patients treated with radical cystectomy for
squamous cell carcinoma of the bladder. Hum Pathol 42,
347-355

37. Tilki D, Reich O, Svatek RS, Karakiewicz PI, Kassouf W,


Novara G, Ficarra V, Chade DC, Fritsche HM, Gerwens N,
Izawa JI, Lerner SP, Schoenberg M, Stief CG, Skinner E,
Lotan Y, Sagalowsky AI, Shariat SF (2010) Characteristics
and outcomes of patients with clinical carcinoma in situ only
treated with radical cystectomy: an international study of
243 patients. J Urol 183., 1757-1763

49. Shariat SF, Bolenz C, Godoy G, Fradet Y, Ashfaq R, Karakiewicz PI, Isbarn H, Jeldres C, Rigaud J, Sagalowsky
AI, Lotan Y (2009) Predictive value of combined immunohistochemical markers in patients with pT1 urothelial carcinoma at radical cystectomy. J Urol 182, 78-84

38. Nagele U, Anastasiadis A, Merseburger AS, Corvin S,


Hennenlotter J, Adam M, Sievert KD, Stenzl A, Kuczyk
MA (2007) The rationale for radical cystectomy as primary
therapy for T4 bladder cancer. World J Urol 25, 401-405

50. Shariat SF, Tokvnaga T, Zhou J, Kim J, Ayala GE, Benedict


WF, Lerner SP ( 2004) p53, p21, pRB, and p16 expression
predict clinical outcome in cystectomy with bladder cancer.
J Clin Oncol 22, 1014-1024

39. Tilki D, S. R., Svatek RS, Karakiewicz PI, Isbarn H, Reich


O, Kassouf W, Fradet Y, Novara G, Fritsche HM, Bastian
PJ, Izawa JI, Stief CG, Ficarra V, Lerner SP, Schoenberg M,
Dinney CP, Skinner E, Lotan Y, Sagalowsky AI, Shariat SF
(2010) Characteristics and outcomes of patients with pT4
urothelial carcinoma at radical cystectomy: a retrospective
international study of 583 patients. J Urol 183, 87-93

51. Gakis G, Todennoefer T, Renninger M, Schilling D, Sievert


KD, Schwentner C, Stenzl A (2011) Development of a new
outcome prediction model in invasive bladder cancer based
on preoperative serum C-reactive protein and standard
pathologic risk factors: the TNR-C Score. BJU Int Epub
ahead of print

294

b) Rationale for Perioperative Therapy:


Advantages and Disadvantages of Neoadjuvant
Chemotherapy

52. Sonpavde G, Khan MM, Lerner SP, Svatek RS, Novara G,


Karakiewicz PI, Skinner E, Tilki D, Kassouf W, Fradet Y,
Dinney CP, Fritsche HM, Izawa JI, Bastian PJ, Ficarra V,
Schoenberg M, Sagalowsky AI, Lotan Y, Shariat SF (2011)
Disease-free survival at 2 or 3 years correlates with 5-year
overall survival of patients undergoing radical cystectomy
for muscle invasive bladder cancer. J Urol 185, 456-461
53. Goossens-Laan CA, Gooike GA, van Gijn W, Post PN,
Bosch JL, Kil PJ, Wouters MW (2011) A Systematic Review
and Meta-analysis of the Relationship Between Hospital/
Surgeon Volume and Outcome for Radical Cystectomy: An
Update for the Ongoing Debate. Eur Urol doi: 10.1016/j.
eururo.2011.01.037
54. Konety BR, Dhawan V, Allareddy V, Joslyn SA (2005)
Impact of hospital and surgeon volume on in-hospital
mortality from radical cystectomy: data from the health care
utilization project. J Urol 173, 1695-1700
55. International Bladder Cancer Nomogram Consortium,
Kattan MW, Bochnas B, Vora KC (2006) Postoperative
nomogram predicting risk of recurrence after radical
cystectomy for bladder cancer. J Clin Oncol 24., 39673972
56. Karakiewicz PI, Shariat SF, Palapattu GS, Gilad AE, Lotan
Y, Rogers CG, Vazina A, Gupta A, Bastian PJ, Perrotte
P, Sagalowsky AI, Schoenberg M, Lerner SP (2006)
Nomogram for predicting disease recurrence after radical
cystectomy for transitional cell carcinoma of the bladder. J
Urol 176, 1354-1361
57. Shariat SF, Karakiewicz PI, Palapattu GS, Amiel GE, Lotan
Y, Rogers CG, Vazina A, Bastian PJ, Gupta A, Sagalowsky
AI, Schoenberg M, Lerner SP (2006) Nomograms provide
improved accuracy for predicting survival after radical
cystectomy. Clin Cancer Res 12, 6663-6676
58. Zaak D, Burger M, Otto W, Bastian PJ, Denzinger S,
Stief CG, Buchner H, Hartmann A, Wieland WF, Shariat
SF, Fritsche HM (2010.) Predicting individual outcomes
after radical cystectomy: an external validation of current
nomograms. BJU Int 106, 342-348

Neoadjuvant or adjuvant chemotherapy has the


potential of eradicating micrometastases and
improving survival in patients with muscle-invasive
urothelial carcinoma. This seems to be particularly
true for patients with pathologic extravesical and
lymph node-positive disease. [7]
Administration of chemotherapy prior to surgery
versus after (adjuvant) offers several potential
advantages. Patients may be able to tolerate
treatment better and the response of the primary
tumor to chemotherapy can be assessed [8-12],
providing prognostic significance. In a study of
patients treated with neoadjuvant cisplatin-based
therapy followed by definitive surgery, 91% of
patients who responded to chemotherapy (defined
as pathologic stage T1) were alive at a median
follow-up of 25 months in contrast to 37% of nonresponders [10].
Downstaging of the tumor may provide an indication of
the activity of neoadjuvant chemotherapy, especially
in patients who have a pathologic complete response
and in patients who are pT1 stage after therapy.
Those patients with residual disease at radical
cystectomy should probably be offered clinical trials
evaluating non-cross-resistant alternative agents. A
complete response after neoadjuvant therapy may
also permit consideration of organ preservation
in selected cases. The standard of care is that the
majority of patients require and undergo cystectomy
or radical radiotherapy.

IV. Perioperative
Chemotherapy

An important potential disadvantage of neoadjuvant


chemotherapy is the discordance between clinical
and pathologic staging. In a study reported by
Scher et al, [13] while 57% of patients achieved a
clinical and cystoscopic complete response following
neoadjuvant M-VAC, only 30% had a pathologic
complete response at subsequent cystectomy.

1. Application of Chemotherapeutic
Agents in Bladder Cancer:
Neoadjuvant Chemotherapy
a) Introduction
Radical cystectomy is considered to be the gold
standard of treatment for patients with clinically
localized muscle-invasive bladder cancer. However,
despite potentially curative surgery, approximately
one-half of patients with deep, muscle-invasive
urothelial carcinoma (stages T2b-4,) develop
metastatic disease within 2 years.[1] At 5 years,
the survival rate after cystectomy is at best 65%,
ranging from 36-48% (level 3) [2-4] depending on
the presence of extravesical extension (pT3) and
lymph node metastases (N1-N3). Both factors are
associated with an increased risk for recurrence
following cystectomy. In contemporary series, 5-year
overall survival rates up to 57% have been reported
in patients with clinically unsuspected N1 disease,
as compared to 0-27% for those with larger volume
N2 or N3 disease [3, 5, 6].

A potential disadvantage of neoadjuvant chemotherapy is that patients achieving a complete clinical


response at the transurethral resection of bladder
tumor after chemotherapy may refuse cystectomy.
Another theoretical disadvantage of the neoadjuvant
approach is the possibility that some low-stage, lowrisk patients may unnecessarily receive neoadjuvant
chemotherapy. Conversely, delay of definitive local
treatment could potentially be associated with disease progression [14].
The primary disadvantage of the adjuvant
chemotherapy paradigm may be that it does not
appear feasible in a third of patients within 90
days after radical cystectomy due to postoperative
complications or slow recovery of functional status

295

[15, 16]. Also, approximately 40% of patients who


would be candidates for neoadjuvant chemotherapy
may not be candidates for postoperative cisplatin
because of a perioperative decline in renal function
[17].

provide benefit to patients, whether it is given before


cystectomy or before radiotherapy. In the United
States, radical cystectomy is preferred for patients
who have a good performance status. In most of
Europe, radical cystectomy is also the preferred
option, although some institutions consider local
radical radiotherapy as an alternative.

Besides all these pros and cons, both approaches


are targeting microscopic disease and the question
is to know what the best option for an individual
patient is. No published trials have directly
compared pure populations of neoadjuvant versus
adjuvant chemotherapy. In the absence of a
specific randomized trial that has compared optimal
neoadjuvant and adjuvant chemotherapy regimens
in association with cystectomy, it is not possible to
make a definitive recommendation about the utility of
adjuvant chemotherapy as compared to neoadjuvant
treatment. This will be further discussed in the
subsequent section of adjuvant chemotherapy.

Several randomized trials have explored whether


neoadjuvant chemotherapy improves survival in
bladder cancer. The results of these randomized
trials are presented in Table 3 [18, 19]. Some
studies suffered from small sample size, suboptimal
chemotherapy, premature closure, or inadequate
follow-up [20]. Among these trials, single agent
regimens failed to show a survival benefit from
neoadjuvant therapy [21]. However, well-designed
multi-agent chemotherapy trials utilizing effective
chemotherapeutic regimens have demonstrated an
improvement in survival. These trials have shifted
the treatment paradigm in muscle-invasive disease
favoring more the use of neoadjuvant chemotherapy
[18, 22].

Before reviewing the accumulating data from


controlled, prospective trials on neoadjuvant
chemotherapy for bladder cancer, it is beneficial
to acknowledge a growing consensus among
many investigators for at least selective use of
neoadjuvant chemotherapy in subgroups of patients
who are known to be at high risk for micrometastatic
disease including those with bulky primary tumors,
hydronephrosis due to the primary tumor, mixed
histology, and possibly lymphovascular invasion
within the primary tumor.

The SWOG Intergroup Trial 0080 randomized


patients with T2-T4a urothelial carcinoma of
the bladder to radical cystectomy alone (154
patients) versus 3 cycles of M-VAC followed by
radical cystectomy (153 patients) [22]. The use of
neoadjuvant chemotherapy was associated with a
higher rate of complete pathologic response (38%
versus 15%, p<0.001). At a median follow-up of 8.7
years, improvements in median survival (77 versus
46 months, p=0.06) and 5-year survival (57% versus
43%, p=0.06) favored the neoadjuvant M-VAC arm.
Because of its size, this trial had limited potential to
discern a clinically meaningful difference. This trend

c) Data Supporting a Survival Benefit for


Neoadjuvant Chemotherapy: Randomized
Trials of Neoadjuvant Chemotherapy
Neoadjuvant chemotherapy theoretically should

Table 3.

296

toward improved survival favoring M-VACtreated


patients with an estimated reduction in the risk of
death by 25% (hazard ratio [HR], 1.33) provides
some evidence of the benefit (level 1) [22]. There
were no treatment-related deaths and neoadjuvant
chemotherapy did not adversely impact the ability to
proceed with radical cystectomy or increase adverse
events related to surgery.

with follow-up of 7.4 years, a statistically significant


improvement in survival was observed for patients
who received neoadjuvant chemotherapy (p=0.048;
HR=0.85; 95% CI 0.72-1.0). This trial, well-powered
and with adequate follow-up, demonstrated both a
survival benefit and improved loco-regional control
with neoadjuvant CMV chemotherapy, however, the
predefined endpoint with an improvement in survival
of 10% was in fact not reached. Survival at 5 years
was 50% with CMV compared with 44% with radiotherapy; at 8 years, it was 43% with CMV and 37%
with radiotherapy. (level 1).

Several studies have been published based on


retrospective analysis of this trial database. As an
example, surgical factors were evaluated in 268
patients with muscle-invasive bladder cancer who
underwent radical cystectomy in this Intergroup
trial [23]. Cystectomies were performed by 106
surgeons at 109 institutions. Half of the patients
received neoadjuvant M-VAC. The 5-year postcystectomy survival and local recurrence rates in all
patients who underwent cystectomy were 54% and
15%, respectively. Surgical variables associated
with longer post-cystectomy survival were negative
margins (hazard ratio 0.37; p=0.0007) and removal
of 10 or more nodes (hazard ratio 0.51; p=0.0001).
These associations did not differ by treatment
arm (p=0.21 for all tests of interactions between
treatment and surgical variables). Predictors of local
recurrence were positive margins (odds ratio 11.2;
p=0.0001) and removal of fewer than 10 nodes
(odds ratio 5.1; P = 0.002). The quality of surgery
was an independent prognostic factor for outcome
after adjustments were made for pathologic factors
and neoadjuvant chemotherapy usage (LE 2).

A trial that was almost identical to the SWOG study


was performed by the Gruppo Uro-Oncologico Nord
Est (GUONE) cooperative group in Italy [25]. Over
a 6.5-year period, 206 patients were randomly
assigned to neoadjuvant M-VAC before cystectomy
or to cystectomy alone. No clear differences in
survival were demonstrated; as 3-year survival was
62% for the M-VACtreated patients and 68% for
patients in the cystectomy alone arm (LE 2).
The Nordic cystectomy I trial evaluated neoadjuvant
doxorubicin, cisplatin, and preoperative radiotherapy
before cystectomy versus preoperative radiotherapy
and cystectomy alone. A 15% survival difference in
favor of patients treated with chemoradiotherapy
was seen in only a subset analysis of patients with
T3 or T4 disease. Investigators were unable to
confirm this survival advantage in the subsequent
Nordic cystectomy II trial, in which 317 patients
were randomly assigned cystectomy or cystectomy
preceded by methotrexate and cisplatin (without
radiotherapy) [26]. However, combining the 2 trials
provided positive results in favor of neoadjuvant
chemotherapy (LE 2) [27].

Another recent analysis evaluated the impact of


histology when neoadjuvant M-VAC was given in this
trial. There was evidence of a survival benefit from
chemotherapy in patients with mixed tumors [24].
Presence of squamous or glandular differentiation in
locally advanced urothelial carcinoma of the bladder
does not seem to confer resistance to M-VAC and in
fact may be an indication for the use of neoadjuvant
chemotherapy before radical cystectomy.

d) Meta-Analysis
Because of the uncertainties of the definitive value
of neoadjuvant chemotherapy in terms of survival,
a meta-analysis of neoadjuvant chemotherapy trials
was performed [28]. Data from 2688 patients treated
in 10 randomized trials evaluating neoadjuvant
chemotherapy for invasive urothelial carcinoma
was reviewed. Of note, this analysis did not include
data from the SWOG Intergroup trial. Compared to
local treatment alone, neoadjuvant platinum-based
combination chemotherapy was associated with a
significant benefit in overall survival [HR=0.87 (95%
CI 0.78-0.98, p=0.016)], a 13% decrease in the risk of
death, and a 5% absolute survival benefit at 5 years
(overall survival increased from 45-50%). When
trials utilizing single-agent cisplatin were included,
the survival benefit did not achieve statistical
significance [HR=0.91 (95% CI 0.83-1.01, p=0.084)].
(LE 2). However, single-agent cisplatin did not show
an improvement in survival (P=0.26) compared
with no neoadjuvant therapy. As all platinum-based
combination trials were analyzed as a group, it is not

The MRC/EORTC performed a large trial in which


976 patients were enrolled and randomized to neoadjuvant CMV (cisplatin - methotrexate - vinblastin)
(491 patients) or no neoadjuvant chemotherapy
(485 patients) over a 5.5-year period from 106 institutions. This trial was performed more or less during the same time as the SWOG trial. The results
of this trial were updated at a median follow-up of
approximately 7 years [18]. Management of the primary tumor involved cystectomy, radiation therapy,
or both and was left to the choice of investigators.
An initial 8% improvement in time to progression
and a 5.5% difference in absolute 3-year survival
(HR=0.85; 95% CI 0.71-1.02) favoring the neoadjuvant chemotherapy arm was reported. When results
were published in 1999, a non-significant trend toward improvement in survival was observed in patients in the CMV arm. In a 2002 update from the
American Society of Clinical Oncology (ASCO),

297

possible to discern the best combination for use in


neoadjuvant therapy.

without paclitaxel in patients with metastatic disease


have led to their investigation in the neoadjuvant and
adjuvant setting. Although these newer regimens are
promising, there are no data from randomized trials
supporting their use in the neoadjuvant setting [35]
and limited data form phase II trials.

A subsequently reported meta-analysis that included


individual patient data from 3005 individuals enrolled
in 11 randomized trials, including the SWOG data
extrapolated from the published report [22], confirmed
the survival benefit for neoadjuvant cisplatin-based
compared to local therapy alone [28, 29]

In a phase II trial of 68 patients with adequate renal


function and clinical T3 or T2 with hydronephrosis, N0
M0 bladder cancer received 3 cycles of neoadjuvant
paclitaxel, carboplatin, gemcitabine (PCaG) with a
primary endpoint of pCR. Patients with T4 or nodepositive disease received 6 cycles of PCaG with
an endpoint of resectability [36]. The caveat is that
this regimen was fairly toxic in a population with
adequate baseline renal function and may often
warrant prophylactic granulocyte growth factors in
accordance with guidelines. ***

A very similar meta-analysis of neoadjuvant


randomized controlled trials was conducted in
Canada.[30] A total of 16 eligible trials that included
3315 patients were identified, and 2605 patients
provided data suitable for a meta-analysis of overall
survival. The pooled HR was 0.90 (95% CI 82%
to 99%; P=0.02). Eight trials used cisplatin-based
combination chemotherapy and the pooled HR was
0.87 (95% CI 78% to 96%; P=0.006), consistent
with an absolute overall survival benefit of 6.5%
from 50% to 56.5% (95% CI 2% to 11%). A major
pathologic response was associated with improved
overall survival in 4 trials. Neoadjuvant cisplatinbased chemotherapy improved overall survival in
muscle-invasive urothelial carcinoma, but the size of
the effect was modest (LE 2).

The SWOG conducted a phase II trial of 3 cycles of


neoadjuvant paclitaxel, gemcitabine, and carboplatin
followed by cystoscopic surveillance or immediate
radical cystectomy for patients with cT0 status after
chemotherapy [37]. Patients with cT0 status could
elect immediate radical cystectomy or cystoscopic
surveillance, and those who did not achieve cT0
status underwent immediate cystectomy. There was
an unacceptably high rate (60%) of persistent cancer
at radical cystectomy in patients presumed to have
pT0 status, which suggests that radical cystectomy
is a critical component of therapy.

The use of perioperative chemotherapy has been


limited until 2003-2005, when these meta-analyses
were published. Among 7161 analyzable patients
in the National Cancer Data Base with stage III
bladder cancer diagnosed between 1998 and 2003,
perioperative chemotherapy was administered to
11.6% of patients with 10.4% receiving adjuvant
chemotherapy and 1.2% receiving neoadjuvant
chemotherapy [31]. After 2003, there has been a
slight increase in its use. In a more recent report on
40,388 patients 18 to 99 years old diagnosed with
muscle invasive (stages II to IV) bladder cancer in
2003 to 2007 from the National Cancer Database,
the incidence of those who received chemotherapy
increased from 27.0% in 2003 to 34.5% in 2007 due
to an increase in neoadjuvant chemotherapy and
chemotherapy without surgery [32]. Clinical Practice
Guidelines can help to increase the implementation
of neoadjuvant chemotherapy. A Canadian study [33]
has shown that neoadjuvant referral and treatment
rates increased after publication of the Clinical
Practice Guidelines. However, overall referral and
treatment rates remained low.

While recently, some potential benefits have been


reported with the use of the triple regimen in the
adjuvant setting [38], only the M-VAC regimen
has been extensively evaluated. On that basis
M-VAC has the strongest evidence-based data for
neoadjuvant use.
While the gemcitabine/cisplatin (table 3) doublet has
not been validated in the perioperative setting, recent
retrospective data from the Memorial Sloan-Kettering
Cancer Center (MSKCC) shows that the GC regimen
produces a pCR rate of 35%, similar to M-VAC [39].
Multiple other reports on use of this doublet for
metastatic disease show very similar response rates
and survival to that obtained with M-VAC, and with
lower toxicity. In contrast to the above MSKCC study,
data from the Cleveland Clinic showed that only 7%
of patients achieved a pCR with mostly GC and other
non-MVAC-based regimens mainly administered in
community oncology practices [35].

Based on these observations and despite


level I evidence, neoadjuvant cisplatin-based
chemotherapies continue to be underutilized in the
management of bladder cancer, even at high-volume
tertiary centers [34].

e) Novel Combinations as
Therapy for Bladder Cancer

In the absence of definitive supportive data for GC


in the neoadjuvant setting, M-VAC remains the
preferred regimen.
Trials are building upon combinations of GC
or dose-dense M-VAC (DD-MVAC) with novel
biological agents administered in 3 to 4 cycles in the
neoadjuvant setting with pCR as a key intermediate
endpoint [7] (Table 2).

Neoadjuvant

The promising results from newer combinations


such as gemcitabine and cisplatin/carboplatin with or

298

f) Developing a Personalized Neoadjuvant


Therapy: A paradigm to define risk of relapse
and response to systemic therapy

to better define the subset of patients who should


have neoadjuvant chemotherapy. Such signatures
may provide a matrix into which chemosensitivity
genes can be intercalated to prediction not only of
stage variables for risk of relapse but also benefit
from specific chemotherapy agents [44,45]. These
hypotheses require testing and validation in
neoadjuvant cohorts and more definitively in trials.

The ability to predict response to a specific therapy


is still a major challenge in oncology. Advances in
molecular research have led to the identification
of genetic markers that impact upon response
to chemotherapy. Based on detailed molecular
information for each individual tumor, the clinician
will ultimately be able to more accurately select
the appropriate therapy for each patient according
to individual predicted response. This customized
treatment using chemosensitivity markers such
as intratumoral molecular pharmacology markers
(pharmacogenomics and genetics) should aid in
improving outcome [40].

There are several studies looking at candidate


markers in cohorts of patients treated with
neoadjuvant chemotherapy. These are summarized
below:

p53 aberrancy as inferred by IHC overexpression,

confers a poor prognosis in muscle-invasive bladder


carcinoma [46, 47]. In one report of 90 patients
undergoing neoadjuvant M-VAC chemotherapy,
patients with mutant p53 were 3 times more likely
to die from their disease than those with wild-type
p53 [47]. The impact of p53 overexpression on
survival was predominantly in T2 and T3a tumors.
Conversely, a retrospective analysis of data
suggested that adjuvant chemotherapy enhanced
survival in patients with p53 mutant tumors [48].

Several reports have outlined a variety of potential


predictive markers either in localized disease or
in advanced disease. Data remains limited, but it
is conceivable that in coming years a marker or a
panel of markers may be available that achieves this
predictive goal, allowing improved patient selection
for chemotherapy.

K
 i67,

p53, and angiogenesis (microvessel


staining with CD34) were assessed by
immunohistochemistry in 94 patients who accrued
to the SWOG 8710 phase III neoadjuvant M-VAC
trial [22, 49]. There was a trend toward shorter
disease-free and overall survival in patients with
higher Ki67 expression. Increased p53 nuclear
expression was associated with a shorter overall
survival but this was not statistically significant. No
association or trend between microvessel density
and outcome parameters was noted. The study
was very limited in power given the small number
of specimens available.

g) Muscle-invasive urothelial carcinoma


Neoadjuvant chemotherapy followed by cystectomy
improves survival compared to surgery alone.
However, 20 patients must be treated to cure one
additional person compared to surgery alone.
To identify those patients who will derive benefit,
limited studies have provided some hints on single
markers and which specific treatments may be of
benefit. At present, many of the published results of
gene expression profiling are preliminary, based on
small sample sizes, and it is not possible to make a
definitive statement on the role of gene expression
profiling in the molecular prognostication of invasive
urothelial carcinoma. Work in this area is directed
along 2 concurrent themes: risk stratification and
chemotherapy response prediction.

BRCA1 mRNA expression was analyzed by


quantitative PCR in tumor biopsies obtained by
transurethral resection from 51 patients with locally
advanced bladder cancer receiving neoadjuvant
chemotherapy. A close correlation was found
between BRCA1 mRNA levels and pathological
response. Low levels of BRCA1 were shown to
predict response to neoadjuvant cisplatin-based
chemotherapy, and correlated with longer diseasefree survival [50].
In vitro, XAF1 expression enhances the apoptotic
response of tumor cells to chemotherapeutic
agents. In vivo, in a paired sample study from 14
bladder cancer patients treated with a combination
of neoadjuvant gemcitabine and cisplatin, patients
with high levels of XAF1 in their tumor had increased
rates of response, progression-free survival (PFS),
and overall survival (OS)[51].

To this point, defining risk of relapse has been


dependent on clinical and imaging based staging,
which often understages the patient, with resultant
upstaging at surgery [41]. A recent study developed
a gene expression model (GEM) in primary bladder
cancer tissue to predict the presence of lymph node
involvement at cystectomy in the AUO-AB-05/95 trial
[42, 43]. A 20-gene signature has an Area under the
Receiver Operator Curve of 0.67 for prediction of
lymph node metastases and was an independent
predictor of their presence in multivariate analysis
with age, sex, pathological stage, and lymphovascular
space invasion. The gene signature divided patients
into high relative risk (1.74, 95% CI 1.03-2.93) and
low relative risk (0.70, 95% CI 0.51-0.96) of nodepositive disease. Further precision in determining
risk of lymph node involvement has the potential

Baseline tumor genomics appear promising


as predictors of pCR but limited small studies
have been reported. The Takata study [52], in
a retrospective study of patients with invasive

299

bladder cancer who received neoadjuvant M-VAC


chemotherapy, found that 14 predictive genes
separated the responder group (defined as no
residual muscle-invasive disease) from the nonresponder group. This system accurately predicted
the drug responses of 8 of 9 test cases. To further
validate the clinical significance of the system, the
investigators applied it to 22 additional cases of
patients with bladder cancer and found that the
scoring system correctly predicted clinical response
for 19 of the 22 test cases [53].

treatment for localized muscle-invasive bladder


cancer (grade B).

A
 discrepancy

between clinical/cystoscopic
and pathologic staging can be anticipated
after neoadjuvant chemotherapy and therefore
cystectomy is not obviated by response (grade
B).

T
 oxicity and mortality associated with neoadjuvant

chemotherapy are acceptable (grade B).


However, few data on quality of life are available.

In summary, 2 large randomized trials and a metaanalysis support the concept that neoadjuvant
chemotherapy for patients with muscle-invasive
bladder cancer provides a survival benefit greater
than with surgery alone. This approach should
be considered for patients who are candidates for
cisplatin-based combination chemotherapy and
radical cystectomy. For patients who have not
received neoadjuvant chemotherapy and who have
extravesical or node-positive disease following
cystectomy, enrollment in a clinical trial should be
encouraged. If patients are not protocol-eligible,
adjuvant cisplatin-based combination chemotherapy
is a reasonable consideration.

Meta-analysis of cisplatin-containing combination

neoadjuvant chemotherapy trials revealed


a modest difference in favor of neoadjuvant
chemotherapy (GR B).

We recommend using M-VAC [methotrexate,

vinblastine, doxorubicin, and cisplatin, (table


2)] for appropriately selected cases as the
neoadjuvant chemotherapy regimen (GR 1A).

A
 lthough other regimens, such as gemcitabineplus

cisplatin[GC], have similar activity in patients


with metastatic disease, there are no data from
randomized trials in the neoadjuvant setting to
support the use of regimens other than M-VAC.
(Grade 1A)

Summary
Cystectomy is considered to be the gold standard
of treatment for patients with localized muscleinvasive bladder cancer. Neoadjuvant chemotherapy
was intended for patients with operable clinical stage
T2 to T4a muscle-invasive disease. The rationale
for giving chemotherapy before cystectomy or fulldose radiotherapy is based on the intent to treat
micrometastatic disease present at diagnosis. A
discrepancy between clinical and pathologic staging
can be expected. Toxicity and mortality associated
with neoadjuvant chemotherapy are acceptable.
Available data suggest that for average-risk
patients with cT2 cancer, the benefit of adding
chemotherapy to local therapy is at best modest.
Likewise, available studies suggest a much more
substantial benefit for patients with high-risk disease,
such as cT3b cancers. Presence of squamous or
glandular differentiation in locally advanced urothelial
carcinoma of the bladder does not seem to confer
resistance to M-VAC. The quality of the surgery is a
confounding factor in these studies. Meta-analysis
of cisplatin-containing combination neoadjuvant
chemotherapy trials revealed a 5% difference in
favor of neoadjuvant chemotherapy. Unfortunately,
in this case, in which small differences in survival
can be seen, it is regrettable that the data on quality
of life are inadequate.

A
 vailable data suggest that for average-risk

cancer patients with cT2, the benefit of adding


chemotherapy to local therapy is at best modest.
Likewise, all available studies suggest a much
more substantial benefit for patients with high-risk
disease, such as cT3b cancers or those thought
to have lymph node involvement (GR B).

T
 he quality of the surgery is a confounding factor
in interpreting these studies (GR B).

Following cystectomy in patients who did not


receive neoadjuvant chemotherapy, we suggest
consideration of adjuvant chemotherapy (see
below) with a cisplatin-based regimen for patients
who have perivesical tumor extension (stage T3
or higher) or regional lymph node involvement
(GR 2C).

Presence of squamous or glandular differentiation

in locally advanced urothelial carcinoma of the


bladder does not seem to confer resistance to
M-VAC and in fact may be an indication for the
use of neoadjuvant chemotherapy before radical
cystectomy (GR 3C).

References

Recommendations
Neoadjuvant Chemotherapy

1. Raghavan D, Shipley WU, Garnick MB et al. Biology and


management of bladder cancer. N Engl J Med 1990; 322:
1129-1138.

Cystectomy is considered the gold standard of


300

2. Dalbagni G, Genega E, Hashibe M et al. Cystectomy for


bladder cancer: a contemporary series. J Urol 2001; 165:
1111-1116.

20. Sylvester R, Sternberg C. The role of adjuvant combination


chemotherapy after cystectomy in locally advanced bladder
cancer: what we do not know and why. Ann Oncol 2000; 11:
851-856.

3. Stein JP, Lieskovsky G, Cote R et al. Radical cystectomy in


the treatment of invasive bladder cancer: long-term results
in 1,054 patients. J Clin Oncol 2001; 19: 666-675.

21. Martinez-Pineiro JA, Gonzalez Martin M, Arocena F et


al. Neoadjuvant cisplatin chemotherapy before radical
cystectomy in invasive transitional cell carcinoma of the
bladder: a prospective randomized phase III study. J Urol
1995; 153: 964-973.

4. Ghoneim MA, el-Mekresh MM, el-Baz MA et al. Radical


cystectomy for carcinoma of the bladder: critical evaluation
of the results in 1,026 cases. J Urol 1997; 158: 393-399.

22. Grossman HB, Natale RB, Tangen CM et al. Neoadjuvant


chemotherapy plus cystectomy compared with cystectomy
alone for locally advanced bladder cancer. N Engl J Med
2003; 349: 859-866.

5. Herr HW, Donat SM. Outcome of patients with grossly node


positive bladder cancer after pelvic lymph node dissection
and radical cystectomy. J Urol 2001; 165: 62-64; discussion
64.

23. Herr HW, Faulkner JR, Grossman HB et al. Surgical factors


influence bladder cancer outcomes: a cooperative group
report. J Clin Oncol 2004; 22: 2781-2789.

6. Leissner J, Hohenfellner R, Thuroff JW, Wolf HK.


Lymphadenectomy in patients with transitional cell
carcinoma of the urinary bladder; significance for staging
and prognosis. BJU Int 2000; 85: 817-823.

24. Scosyrev E, Ely BW, Messing EM et al. Do mixed


histological features affect survival benefit from neoadjuvant
platinum-based combination chemotherapy in patients with
locally advanced bladder cancer? A secondary analysis
of Southwest Oncology Group-Directed Intergroup Study
(S8710). BJU Int.

7. Sonpavde G, Sternberg CN. Neoadjuvant systemic therapy


for urological malignancies. BJU Int 106: 6-22.
8. Scher HI, Splinter TA. Neoadjuvant chemotherapy for
invasive bladder cancer: future directions. Semin Oncol
1990; 17: 635-638.

25. Bassi P, Pagano F, Pappagallo G, al: e. Neo-adjuvant


M-VAC of invasive bladder cancer: G.U.O.N.E. multicenter
phase III trial. Eur Urol. 1998; 33:142. (abstract 567).

9. Splinter TA, Denis L, Scher HI et al. Neoadjuvant


chemotherapy of invasive bladder cancer. The prognostic
value of local tumor response. Prog Clin Biol Res 1989;
303: 541-547.

26. Sherif A, Rintala E, Mestad O et al. Neoadjuvant cisplatinmethotrexate chemotherapy for invasive bladder cancer -Nordic cystectomy trial 2. Scand J Urol Nephrol 2002; 36:
419-425.

10. Splinter TA, Scher HI, Denis L et al. The prognostic value
of the pathological response to combination chemotherapy
before cystectomy in patients with invasive bladder cancer.
European Organization for Research on Treatment of
Cancer--Genitourinary Group. J Urol 1992; 147: 606-608.

27. Shipley WU, Winter KA, Kaufman DS et al. Phase III trial
of neoadjuvant chemotherapy in patients with invasive
bladder cancer treated with selective bladder preservation
by combined radiation therapy and chemotherapy: initial
results of Radiation Therapy Oncology Group 89-03. J Clin
Oncol 1998; 16: 3576-3583.

11. Splinter TA, Scher HI, Denis L et al. The prognostic value
of the pT-category after combination chemotherapy for
patients with invasive bladder cancer who underwent
cystectomy. EORTC-GU group. Prog Clin Biol Res 1990;
353: 219-224.

28. Neoadjuvant chemotherapy in invasive bladder cancer: a


systematic review and meta-analysis. Lancet 2003; 361:
1927-1934.

12. Schultz PK, Herr HW, Zhang ZF et al. Neoadjuvant


chemotherapy for invasive bladder cancer: prognostic
factors for survival of patients treated with M-VAC with
5-year follow-up. J Clin Oncol 1994; 12: 1394-1401.

29. Neoadjuvant chemotherapy in invasive bladder cancer:


update of a systematic review and meta-analysis of
individual patient data advanced bladder cancer (ABC)
meta-analysis collaboration. Eur Urol 2005; 48: 202-205;
discussion 205-206.

13. Scher HI, Yagoda A, Herr HW et al. Neoadjuvant M-VAC


(methotrexate, vinblastine, doxorubicin and cisplatin) effect
on the primary bladder lesion. J Urol 1988; 139: 470-474.

30. Winquist E, Kirchner TS, Segal R et al. Neoadjuvant


chemotherapy for transitional cell carcinoma of the bladder:
a systematic review and meta-analysis. J Urol 2004; 171:
561-569.

14. Nielsen ME, Palapattu GS, Karakiewicz PI et al. A delay in


radical cystectomy of >3 months is not associated with a
worse clinical outcome. BJU Int 2007; 100: 1015-1020.
15. Donat SM, Shabsigh A, Savage C et al. Potential impact of
postoperative early complications on the timing of adjuvant
chemotherapy in patients undergoing radical cystectomy:
a high-volume tertiary cancer center experience. Eur Urol
2009; 55: 177-185.

31. David KA, Milowsky MI, Ritchey J et al. Low incidence of


perioperative chemotherapy for stage III bladder cancer
1998 to 2003: a report from the National Cancer Data Base.
J Urol 2007; 178: 451-454.
32. Fedeli U, Fedewa SA, Ward EM. Treatment of muscle
invasive bladder cancer: evidence from the National Cancer
Database, 2003 to 2007. J Urol 185: 72-78.

16. Shabsigh A, Korets R, Vora KC et al. Defining early


morbidity of radical cystectomy for patients with bladder
cancer using a standardized reporting methodology. Eur
Urol 2009; 55: 164-174.

33. Miles BJ, Fairey AS, Eliasziw M et al. Referral and


treatment rates of neoadjuvant chemotherapy in muscleinvasive bladder cancer before and after publication of a
clinical practice guideline. Can Urol Assoc J 4: 263-267.

17. Canter D, Viterbo R, Kutikov A et al. Baseline renal function


status limits patient eligibility to receive perioperative
chemotherapy for invasive bladder cancer and is minimally
affected by radical cystectomy. Urology 77: 160-165.

34. Porter MP, Kerrigan MC, Donato BM, Ramsey SD. Patterns
of use of systemic chemotherapy for Medicare beneficiaries
with urothelial bladder cancer. Urol Oncol 2009.

18. Neoadjuvant cisplatin, methotrexate, and vinblastine


chemotherapy for muscle-invasive bladder cancer: a
randomised controlled trial. International collaboration of
trialists. Lancet 1999; 354: 533-540.

35. Weight CJ, Garcia JA, Hansel DE et al. Lack of pathologic


down-staging with neoadjuvant chemotherapy for muscleinvasive urothelial carcinoma of the bladder: a contemporary
series. Cancer 2009; 115: 792-799.

19. Sternberg CN, Donat SM, Bellmunt J et al. Chemotherapy


for bladder cancer: treatment guidelines for neoadjuvant
chemotherapy, bladder preservation, adjuvant chemotherapy, and metastatic cancer. Urology 2007; 69: 62-79.

36. Smith DC, Mackler NJ, Dunn RL et al. Phase II trial of

301

paclitaxel, carboplatin and gemcitabine in patients with


locally advanced carcinoma of the bladder. J Urol 2008;
180: 2384-2388; discussion 2388.

genome-wide gene expression profiling. Clin Cancer Res


2005; 11: 2625-2636.
53. Takata R, Katagiri T, Kanehira M et al. Validation study
of the prediction system for clinical response of M-VAC
neoadjuvant chemotherapy. Cancer Sci 2007; 98: 113117.

37. deVere White RW, Lara PN, Jr., Goldman B et al. A


sequential treatment approach to myoinvasive urothelial
cancer: a phase II Southwest Oncology Group trial (S0219).
J Urol 2009; 181: 2476-2480; discussion 2480-2471.

2. Application of Chemotherapeutic
Agents in Bladder Cancer:
Adjuvant Chemotherapy

38. Paz-Ares LG, Solsona E, Esteban E et al. Randomized


phase III trial comparing adjuvant paclitaxel/gemcitabine/
cisplatin (PGC) to observation in patients with resected
invasive bladder cancer: Results of the Spanish Oncology
Genitourinary Group (SOGUG) 99/01 study. J Clin Oncol
2010; 28:18s.

a) Introduction
Despite the high rate of down-staging and response
in the neoadjuvant and metastatic setting, cisplatinbased chemotherapy is underutilized in the treatment
of urothelial cancer. As a consequence, more than
50% of patients with high grade bladder cancer
and muscle invasion ultimately die of disseminated
disease. High-risk patients with pT3-pT4 and node
negative disease have no more than a 5-year overall
survival of 47% after cystectomy; patients with lymph
node metastases have an overall 5-year survival rate
of up to 31% after radical cystectomy [1, 2]. Despite
a high risk of relapse, translating the high response
seen in advanced disease into long-term survival
in the locally advanced setting has proved difficult
[3, 4]. The chemotherapy agents used in urothelial
cancer have recently been reviewed and will not be
discussed in detail here [5].

39. Dash A, Pettus JAt, Herr HW et al. A role for neoadjuvant


gemcitabine plus cisplatin in muscle-invasive urothelial
carcinoma of the bladder: a retrospective experience.
Cancer 2008; 113: 2471-2477.
40. Bellmunt J, Albiol S, Suarez C, Albanell J. Optimizing
therapeutic strategies in advanced bladder cancer: update
on chemotherapy and the role of targeted agents. Crit Rev
Oncol Hematol 2009; 69: 211-222.
41. Canter D, Long C, Kutikov A et al. Clinicopathological
outcomes after radical cystectomy for clinical T2 urothelial
carcinoma: further evidence to support the use of
neoadjuvant chemotherapy. BJU Int 107: 58-62.
42. Lehmann J, Retz M, Wiemers C et al. Adjuvant cisplatin plus
methotrexate versus methotrexate, vinblastine, epirubicin,
and cisplatin in locally advanced bladder cancer: results of
a randomized, multicenter, phase III trial (AUO-AB 05/95).
J Clin Oncol 2005; 23: 4963-4974.
43. Smith SC, Baras AS, Dancik G et al. A 20-gene model for
molecular nodal staging of bladder cancer: development
and prospective assessment. Lancet Oncol 2011; 12: 137143.

Adjuvant chemotherapy for bladder cancer is controversial. This controversy is fueled by suboptimal
outcome for locally advanced patients treated with
radical cystectomy alone, a small potential benefit
of chemotherapy and a sequence of trials that have
been underpowered and/or closed early due to poor
accrual as well as the presence of more definitive
evidence for neoadjuvant chemotherapy. Neoadjuvant cisplatin-based combination chemotherapy is
the standard of care for medically fit patients with
high-grade stage T2 or greater bladder cancer based
on level 2B evidence for improved overall survival
and is discussed in detail above. An alternative is
the use of concurrent cisplatin and radiation therapy
for which there is also level 2B evidence, albeit in a
more highly selective cohort of patients. At this time,
there is no proven value to neoadjuvant or adjuvant
chemotherapy for patients undergoing definitive
chemo-radiation (see discussion in the section on
radiation therapy).

44. Bellmunt J, Paz-Ares L, Cuello M et al. Gene expression of


ERCC1 as a novel prognostic marker in advanced bladder
cancer patients receiving cisplatin-based chemotherapy.
Ann Oncol 2007; 18: 522-528.
45. Hoffmann AC, Wild P, Leicht C et al. MDR1 and ERCC1
expression predict outcome of patients with locally
advanced bladder cancer receiving adjuvant chemotherapy.
Neoplasia 2010; 12: 628-636.
46. Esrig D, Elmajian D, Groshen S et al. Accumulation of
nuclear p53 and tumor progression in bladder cancer. N
Engl J Med 1994; 331: 1259-1264.
47. Sarkis AS, Bajorin DF, Reuter VE et al. Prognostic value
of p53 nuclear overexpression in patients with invasive
bladder cancer treated with neoadjuvant MVAC. J Clin
Oncol 1995; 13: 1384-1390.
48. Cote RJ, Esrig D, Groshen S et al. p53 and treatment of
bladder cancer. Nature 1997; 385: 123-125.
49. Grossman HB, Tangen CM, Cordon-Cardo C et al. Evaluation
of Ki67, p53 and angiogenesis in patients enrolled in a
randomized study of neoadjuvant chemotherapy with or
without cystectomy: a Southwest Oncology Group Study.
Oncol Rep 2006; 16: 807-810.

Despite definitive evidence for its use, relatively


few patients are offered and receive neoadjuvant
chemotherapy before surgery [6, 7], although this
may be increasingly driven by clinical practice
guidelines [8-11]. In tandem with commonly
observed upstaging of bladder cancer patients
at surgery [12, 13], slow adoption of neoadjuvant
treatment has resulted in clinicians being confronted
with patients who have not had the potential benefit
of chemotherapy in combination with surgery but

50. Font A, Taron M, Gago JL et al. BRCA1 mRNA expression


and outcome to neoadjuvant cisplatin-based chemotherapy
in bladder cancer. Ann Oncol 22: 139-144.
51. Pinho MB, Costas F, Sellos J et al. XAF1 mRNA expression
improves progression-free and overall survival for patients
with advanced bladder cancer treated with neoadjuvant
chemotherapy. Urol Oncol 2009; 27: 382-390.
52. Takata R, Katagiri T, Kanehira M et al. Predicting response
to methotrexate, vinblastine, doxorubicin, and cisplatin
neoadjuvant chemotherapy for bladder cancers through

302

who have pathological staging that portends a risk


of relapse of up to 70%. This scenario begs the
question as to whether patients would be better
treated with immediate postoperative chemotherapy
or observed for possible relapse and treated at that
time. This question is currently the subject of some
studies where results are awaited.

[20]. Logothetis et al administered CISCA to a


group of 71 post-cystectomy patients with resected
nodal metastases, extravesicular extension,
lymphovascular invasion, or pelvic visceral invasion
[21]. These patients were compared in a nonrandomized fashion to 62 high-risk patients and
206 low-risk patients who did not receive adjuvant
chemotherapy. They concluded that adjuvant CISCA
conferred a 2-year disease-free survival advantage
to patients with unfavorable pathologic findings (70%
vs. 30%, p=0.00012). The earliest randomized control
trial of combination chemotherapy administered to
patients after radical cystectomy was conducted at
USC [22]. Ninety-one patients with p3, p4, or nodepositive urothelial carcinoma were randomized
to receive 4 cycles of CAP or to observation.
Chemotherapy resulted in a significant improvement
in the risk of disease recurrence at 3 years (30% vs.
54%, p=0.011, unstratified Wilcoxon test) but only a
trend to benefit in the overall risk of death (34% vs.
50% p=0.099, unstratified Wilcoxon). The median
survival of patients on chemotherapy was reported
to be 4.25 years versus 2.4 years for patients in the
observation group. This study has been criticized for
the fact that only 33 out of 44 patients assigned to the
chemotherapy arm received one or more cycles of
CAP, for the small sample size, and for deficiencies
in statistical analysis such as the use of the Wilcoxon
test emphasizing early differences. Nonetheless,
the study was provocative in revealing the potential
benefit of adjuvant chemotherapy and in highlighting
the difficulties involved in conducting such trials.

Unfortunately, there have been methodological


issues with many of the studies undertaken in the
postoperative chemotherapy setting and we are left
with meta-analyses for our best evidence. A pooled
analysis supported adjuvant chemotherapy, with a
more extensive meta-analysis demonstrating a benefit
to adjuvant cisplatin combination chemotherapy [14,
15]. Since these analyses were undertaken, 4 major
phase III studies were commenced and closed:
Spanish Oncology Genitourinary Group SOGUG
99/01 [16], CALGB, Italian Multicentric [17], and
EORTC 30994 [18]. These trials continue the pattern
of premature closure for poor accrual seen in earlier
studies, but may contribute in composite to the field.
Recently, a large cohort analysis assessing the effect
of adjuvant chemotherapy from several large centers
has been published [19] suggesting the greatest
impact of adjuvant chemotherapy is seen in patients
with extravesical extension or N+ disease.

b) A Short History of Early Clinical Trials of


Adjuvant Chemotherapy in Bladder Cancer
Multiple cisplatin-based combinations have been
evaluated in the adjuvant setting (see Table 4)

Table 4. Summary of key studies of adjuvant chemotherapy in bladder cancer


Center

Regimen

Outcome

University of
Mainz

MVEC/MVAC

Early
stopping Underpowered
due
to interim
analysis favoring
chemotherapy
Modest
benefit Methodological issues
for chemotherapy

University
Cisplatin-based
of Southern
California
Stanford University Cisplatin, methotrexate, vinblastine Early
stopping
due to
interim
analysis favoring
chemotherapy
1.
SOCUG Cisplatin, gemcitabine, paclitaxel Early termination
vs. observation
due
to
poor
99/01
accrual
CALGB-90104
Rapid
sequence
AG-ITP Early termination
chemotherapy with G-CSF vs. due
to
poor
cisplatin, gemcitabine
accrual
Italian Multicentric Cisplatin,
gemcitabine
vs. Early termination
study
observation
due
to
poor
accrual
and
futility
EORTC 30994
GC, M-VAC, or DD-M-VAC vs. Early termination
chemotherapy at relapse
due
to
poor
accrual

303

Comment

Ref.
(26)

(50) (1)

Underpowered,
delayed
time to progression (p=0.01)
effect on survival

(27)

Major
benefit
chemotherapy arm

(16)

to

(30)

Non-significant,
underpowered; trend to
better outcome in nonchemotherapy arm

(17)

(18)

A subsequent trial of adjuvant chemotherapy with 3


cycles of cisplatin alone did not result in any survival
benefit in a randomized study of 77 patients [23].
Potential explanations for the lack of significant
benefit include the usage of single agent cisplatin,
the small sample size, the inclusion of patients with
lower T stage and lymph node-negative disease,
and the administration of the planned 3 cycles
of chemotherapy to only 65% of patients in the
treatment arm.

patients who were treated in adjuvant chemotherapy


trials compared to observation studies that were
published before September 2004. The analysis
was limited in its ability to be definitive in most eyes,
with only 491 patients from 6 randomized controlled
trials included. The overall hazard ratio for survival
of 0.75 (95% CI 0.60-0.96, p=0.019) suggested
a 25% relative reduction in the risk of death for
chemotherapy compared to controls (see Figures
1 and 2). The authors commented on the small
number of patients and relative poor quality of data
going into the meta-analysis and highlighted the
need for accrual to ongoing phase III trials examining
adjuvant therapy.

Given the superiority of the M-VAC combination


over single agent cisplatin in the metastatic setting
[24], it became important to evaluate the M-VAC or
M-VEC (using epirubicin rather than Adriamycin or
doxorubicin) combinations in the adjuvant setting.
Stockle et al randomized patients with pT3, pT4,
and or pelvic lymph nodes to 3 cycles of M-VAC or
M-VEC versus observation[25, 26]. While planned to
accrue 100 patients, the study was closed after an
interim analysis of 49 randomized patients revealed
a significant advantage in relapse-free survival
with chemotherapy (p=0.0015). This trial has been
interpreted with caution given its early closure and
the fact that only 62% of patients randomized to
chemotherapy completed the 3 cycles of treatment.
Furthermore, patients in the observation arm were
not offered chemotherapy at relapse. Two to three
years later, the same authors reported their longer
experience with adjuvant M-VAC/M-VEC in 83
patients. Forty-nine of the patients had been enrolled
in the prospective trial before being closed while
the remaining 38 had received M-VAC/M-VEC as a
routinely recommended therapy based on the interim
results of the trial. Longer follow-up of the patients
(38 to 78 months) who were on the trial confirmed
significant improvement in progression-free survival
in the adjuvant chemotherapy group (p=0.0005).
The continued advantage in progression-free
survival with more mature data offered support to the
beneficial role of chemotherapy.

Contemporaneously, Dr. Ruggeri and colleagues


undertook a composite analysis from published data
from all phase III studies of adjuvant chemotherapy
published [14]. While less stringent than the Cochrane
review, the conclusions were similar with a benefit
to adjuvant chemotherapy for overall survival (RR
0.74; 95% CI 0.62-0.88 [P= 0.001]) and disease-free
survival (RR 0.65; CI 0.54-0.78, [P<0.001]).
Concurrently investigators have begun to compare
different adjuvant regimens. Investigators at MD
Anderson presented data comparing 2 cycles of
preoperative chemotherapy with M-VAC and 3
afterwards with the same chemotherapy given only
postoperatively [28] in a group of patients at high risk
of extravesical extension or nodal involvement at
surgery. This study demonstrated a high likelihood of
extravesical extension or nodal involvement in nearly
80% of patients treated with initial surgery, and no
difference in survival whether the chemotherapy was
given in the neoadjuvant or adjuvant setting. While
there was no difference between the two approaches
in terms of outcome, it did demonstrate the feasibility
of preoperative chemotherapy and, in particular, that
such therapy did not result in deterioration or toxicity
so that the patient had delayed surgery or missed it
altogether. The German Urologic Oncology groups
ran a phase III trial for patients with stage pT3a-4a
and/or pathologic node-positive urothelial carcinoma
of the bladder after radical cystectomy, randomizing
327 patients to either cisplatin and methotrexate
(CM) or methotrexate, vinblastine, epirubicin, and
cisplatin (M-VEC) (29). The 5-year progression-free,
tumor-specific, and overall survival rates were not
significantly different between the 2 arms although
patients given M-VEC had higher rates of grade 3
or 4 leukopenia (22%) than those given CM (7%,
p=0.0001).

The combination of cisplatin, vinblastine, and


methotrexate was utilized as adjuvant therapy in
a prospective randomized trial of 4 cycles of CMV
versus observation following cystectomy at Stanford
University [27]. Patients accrued to this trial had
pT3b and pT4 urothelial carcinoma with or without
lymph node involvement. Data was reported on 50
out of 55 enrolled patients. Twenty-two out of 25
patients randomized to adjuvant therapy received
the total number of 4 planned cycles. With a median
follow-up of 62 months, a significant difference in
freedom from progression was noted between the
chemotherapy and the observation group (median of
37 months vs. 12 months respectively, p=0.01). No
significant difference in overall survival was noted.

The meta-analysis and composite analysis represent


a watershed in perioperative chemotherapy for
bladder cancer in part because they suggested
benefit from chemotherapy but highlighted the relative
poor quality of the trials undertaken in the area.
The advent of phase III evidence for neoadjuvant
M-VAC at around this time also shaped thinking with

c) Meta-analysis and Composite Analysis


In 2006, an analysis was published of both composite
trials and a meta-analysis of individual trials of

304

Figure 1. Composite Kaplan-Meier curve of overall survival in bladder cancer patients treated with adjuvant
chemotherapy or observation that were included in the Cochrane individual patient data meta-analysis (15).

84

Figure 2. Hazard ratios for overall survival in bladder patients treated with adjuvant chemotherapy or
observation that were included in the Cochrane individual patient data meta-analysis (15).

305

neoadjuvant therapy becoming a standard of care.


Despite this, most patients were not being offered
chemotherapy before surgery. When presented with
the all too common upstaging and/or more definitive
evidence of risk of relapse in the surgical pathology
report, many clinicians and patients considered
postoperative treatment despite deficiencies in
evidence to support this approach.

42 days and no more than 3 months after surgery.


The accrual target was 800 patients, however, the
study was halted due to slow accrual less than 2 years
after opening with fewer than 100 patients enrolled.
Subsequent analysis of Memorial Sloan Kettering
Cancer Center data from studies used to develop
the rapid cycling regimen in more advanced disease,
which also incorporated ifosfamide, suggested
issues with toxicity in the accelerated therapy arm
that was also an issue in this trial [31]. In retrospect,
CALGB 90104 may have been overly ambitious in
attempting to move a dose-dense regimen forward
in the same step as the integration of newer agents.
Published results from this trial are awaited.

d) More Recent Phase III Trials


Subsequently, 4 major phase III trials have been
formulated and opened to accrual. They share
a common theme: adjuvant chemotherapy for
transitional cell-predominant urothelial cancer
coupled with the acrimony of early closure for slow
accrual.

The Spanish Urologic Oncology Group opened


the 99/01 trial comparing 4 cycles of paclitaxel,
cisplatin, and gemcitabine (PCG) to observation
[16]. This regimen was used based on the results of
a phase I/II trial in the advanced disease setting [32].
In a subsequent phase III trial in advanced urothelial
cancer, the addition of paclitaxel increased the
efficacy in terms of response rate when compared
to GC with a benefit in overall survival seen only in
patients having the bladder as primary (post-hoc
analysis). In the intention-to-treat population, only
a trend was observed (p=0.07) [33]. The adjuvant
99/01 trial accrued patients with pT3-4 and/or lymph
node positive bladder cancer with a creatinine
clearance greater than 50 mL/min and mandated
chemotherapy commencement within 8 weeks of
cystectomy, whereas prior studies had allowed up
to 12 weeks after surgery. The trial enrolled 142
patients between July 2000 and July 2007 when it
was closed prematurely due to poor accrual. Toxicity
in the triple drug chemotherapy arm was acceptable
with a single treatment-related death due to sepsis.
At a median follow up of 30 months, overall survival
was significantly prolonged in the PCG arm (median
not reached; 5-year overall survival 60%) compared
to observation (median 26 months; 5-year overall
survival 31%) (p<0.0009). Disease-free survival
(p<0.0001) and disease-specific survival (p<0.0002)
were also superior in the PGC arm. The results from
this trial raise several questions.

The Italian multicentric trial saw 194 patients


with pT2G3, pT3-4, N0-2 transitional cell bladder
carcinoma treated with radical cystectomy who
were then randomized to immediate chemotherapy
or a control arm of observation and chemotherapy
at relapse [17]. Patients were stratified by center
and lymph node metastases. Those patients given
adjuvant chemotherapy were randomized to two
slightly different schedules of gemcitabine and
cisplatin over a 4-week cycle given for 4 cycles.
The primary endpoint was overall survival. The
3-year overall survival was 67% ( 6% SE) for
chemotherapy at relapse patients and 48% ( 6%
SE) for those given immediate adjuvant therapy
with corresponding 3-year disease-free survival of
47% and 35%, respectively. These differences were
not statistically significant. The authors concluded
that there was no role for adjuvant chemotherapy
after cystectomy for locally advanced bladder. The
outcome data from this trial run counter to prior
experience where disease-free survival has been
increased on chemotherapy arms. Publication in a
peer-reviewed journal with more data on surgical
treatment given and chemotherapy delivery in each
arm is awaited.
The Cancer and Leukemia Group B 90104 trial
saw patients with urothelial carcinoma of the bladder
treated with cystectomy and with a creatinine
clearance greater than 60 mL/min randomized to
either a rapid cycling regimen of chemotherapy with
4 cycles of doxorubicin-gemcitabine given at 14 day
intervals with granulocyte colony stimulating factor
support followed by 4 cycles of paclitaxel-cisplatin
given at 21 day intervals compared to adjuvant
gemcitabine with cisplatin in a 4-week cycle [30].
Up to 4 cycles of 4 weeks each of each regimen
were administered. Accrued patients were stratified
based on pathological criteria according to primary
tumor status (<T4 vs. T4), number of positive lymph
nodes (0 or unknown vs. 1-5 vs. >5), and number of
dissected nodes (0-10 or unknown vs. more than 10).
Patients commenced chemotherapy no earlier than

Does the addition of paclitaxel increase the efficacy


of PCG in this setting compared to GC or MVAC,
when data for the addition of paclitaxel to GC in the
advanced setting are lacking [33]?
Did the stringency of time to commencement
of chemotherapy in this trial contribute to the
difference seen in overall survival?
Given data from several centers suggesting
diminished survival outcomes in patients with
later commencement of chemotherapy compared
to those starting earlier [28], should clinical trials
and clinical practice outside trials mandate
commencement within 8 weeks of surgery?

306

The EORTC-30994 trial accrued patients with pT3-4


and/or lymph node positive urothelial cancer and
with a creatinine clearance greater than 60 mL/min
to either 4 cycles of cisplatin-based combination
of chemotherapy of physician preference (GC,
MVAC, or Dose-Dense MVAC [34]) or chemotherapy
with the same regimen at first relapse [18]. Both
standard M-VAC and GC were given on a 4-week
cycle whereas DD-MVAC was given every 2 weeks.
Patients were stratified by pT stage and number of
lymph nodes dissected. Those patients randomized
to the chemotherapy arm were required to start
treatment within 90 days of surgery. The study was
unique in allowing a range of chemotherapy regimens
including a more intense deliver of cytotoxic drug
with DD-MVAC. The accrual target was 660 patients
but this trial also closed prematurely in June 2008
due to slow accrual after accruing 278 patients.

chemotherapy but may be key in planning studies


with an initial phase II accrual in the adjuvant setting
before expanding to a larger phase III cohort of
disease-free survival and toxicity endpoints are met.

f) Biomarkers and Other Indicators of Potential


Adjuvant Chemotherapy Benefit
The literature is beset with multiple analyses of
individual markers of outcome after cystectomy.
p53 aberrance, cycle cell gene dysregulation, and
presence of lymphovascular invasion identify patients
with low risk (<pT2) disease who are at heightened
risk of relapse (36-38). Recent attempts have been
made to link these markers with systemic therapy
interventions. In the p53 M-VAC trial, patients were
screened for p53 abrogation and randomized to
either 3 cycles of M-VAC or observation [39]. The
trial was closed due to futility contingent upon
slow accrual and low event rate. The final analysis
did not demonstrate an advantage for M-VAC
chemotherapy in patients whose tumors contained
abnormal p53, in fact those patients had a nonsignificant trend to a higher relapse and death rate
with chemotherapy. This result proved disappointing
and once again highlighted the difficulty of running
trials at the adjuvant interface in bladder cancer.
Current biomarker efforts led by the International
Bladder Cancer Consortium are directed at large
scale tissue microarray construction and analysis for
putative markers of chemotherapy response such
as ERCC1 (platinum drugs) [40-43], ribonucleotide
reductase (gemcitabine) [44, 45], topoisomerase II
(doxorubicin, epirubicin) [46, 47], and beta-tubulin
(taxanes) [45, 48, 49]. Hopefully, these studies
will delineate relationships between markers and
therapies as well as defining magnitude of effect to
help power those studies. Targeted monoclonal and
small molecule agents remain of interest in urothelial
cancer and a focus of studies that will attempt to
treat patients that have the target present in their
tumor and therefore are more likely to respond.

A further meta-analysis incorporating individual


patient data from the Italian multicentric, Spanish
Urologic Oncology Group and EORTC-30994 trials
was prospectively planned and will proceed once
data from each trial becomes mature.

e) Data from Larger Cohort Studies


A recent collaborative effort between 11 major
centers has yielded an international cohort analysis
of off trial adjuvant chemotherapy [19]. Patients were
grouped into quintiles based on risk characteristics
for relapse and death and chemotherapy impact
assessed across the cohort as a whole but also within
each risk segment. The cohort consisted of 3947
patients undergoing cystectomy and lymph node
dissection between 1979 and 2008, 932 (23.6%) of
whom received adjuvant chemotherapy; the largest
analysis of an adjuvant chemotherapy cohort to
date. Adjuvant chemotherapy was independently
associated with improved survival (hazard ratio,
0.83; 95% CI 0.72-0.97%, P=0.017). Significantly,
risk group predicated the survival impact of
chemotherapy on outcome. Increasing benefit from
adjuvant chemotherapy was seen across higherrisk subgroups (P<0.001), especially in those with
extravesical extension or nodal involvement. There
was a significant improvement in survival between
the treated and non-treated patients in the highestrisk quintile (hazard ratio, 0.75; 95% CI 0.62-0.90;
P=0.002). This group was characterized by an
estimated 32.8% 5-year probability of cancer-specific
survival, with 86.6% of patients having both stage
T3 or greater and nodal involvement. These data
may be useful in stratifying and selecting patients for
future studies.

Summary
The body of evidence supports the use of
perioperative chemotherapy, however, the best
evidence is for neoadjuvant rather than adjuvant
therapy. Several studies have suggested a 5-15%
absolute advantage for chemotherapy in the
postoperative setting. Given the small incremental
benefit to adjuvant chemotherapy, the demonstration
of a survival advantage may take a trial with several
thousand patients unless patients accrued can be
stratified for risk: benefit by clinical or pathological
parameters and/or biomarkers predictive of relapse
risk and/or chemotherapy benefit. The optimal
timing and intensity of chemotherapy in the adjuvant
setting remains to be determined. Accrual to trials
of adjuvant therapy in urothelial cancer represents a
major challenge.

Analysis from the same group of investigators


suggests that 2- and 3-year disease-free survival
after cystectomy is a strong surrogate for 5-year
overall survival [35]. This surrogacy needs to be
tested in prospective cohorts treated with adjuvant

307

Recommendations

9. Sternberg CN, Donat SM, Bellmunt J, et al. Chemotherapy


for bladder cancer: treatment guidelines for neoadjuvant
chemotherapy,
bladder
preservation,
adjuvant
chemotherapy, and metastatic cancer. Urology. 2007;69(1
Suppl):62-79.

Adjuvant
cisplatin-based
chemotherapy
is
supported by a recent large cohort analysis (LE:
2A), several relatively small randomized clinical
trials (LE: 1B), and the results of a meta-analysis
and composite analysis or randomized trials (LE:
1A). However, the consensus in the writing group
is that the trials used in meta-analyses were flawed
so as to make definitive conclusions difficult. On
that basis, the group provides a Grade B level
of recommendation for adjuvant cisplatin-based
chemotherapy in the patient with pT3/4 and/or
lymph node positive cancer at cystectomy, who
has not had neoadjuvant chemotherapy and is
medically fit.

10. Miles BJ, Fairey AS, Eliasziw M, et al. Referral and treatment
rates of neoadjuvant chemotherapy in muscle-invasive
bladder cancer before and after publication of a clinical
practice guideline. Can Urol Assoc J. 2010;4(4):263-7.
11. Bellmunt J, Orsola A, Maldonado X, Kataja V, Group EGW.
Bladder cancer: ESMO Practice Guidelines for diagnosis,
treatment and follow-up. Ann Oncol. 2010;21 Suppl
5:v134-6.
12. Canter D, Long C, Kutikov A, et al. Clinicopathological
outcomes after radical cystectomy for clinical T2 urothelial
carcinoma: further evidence to support the use of
neoadjuvant chemotherapy. BJU Int. 2010;107(1):58-62.
13. Svatek RS, Shariat SF, Novara G, et al. Discrepancy
between clinical and pathological stage: external validation
of the impact on prognosis in an international radical
cystectomy cohort. BJU Int. 2011.

Adjuvant regimens not containing cisplatin


(including those containing carboplatin) should
not be routinely used outside of clinical trials
because of a lack of evidence for their benefit in
that setting. We therefore offer GR against use
of non-cisplatin containing regimens. Patients
who cannot tolerate cisplatin-based combination
therapy should be observed unless a clinical trial
is available.

14. Ruggeri EM, Giannarelli D, Bria E, et al. Adjuvant


chemotherapy in muscle-invasive bladder carcinoma:
a pooled analysis from phase III studies. Cancer.
2006;106(4):783-8.
15. Advanced Bladder Cancer Meta-analysis C. Adjuvant
chemotherapy for invasive bladder cancer (individual patient
data). Cochrane Database Syst Rev. 2006(2):CD006018.

Until the current equipoise is resolved for adjuvant


chemotherapy in this setting, clinical trial remains
the best choice for patients with locally advanced
bladder cancer.

16. Paz-Ares LG, Solsona E, Esteban E, et al. Randomized


phase III trial comparing adjuvant paclitaxel/gemcitabine/
cisplatin (PGC) to observation in patients with resected
invasive bladder cancer: Results of the Spanish Oncology
Genitourinary Group (SOGUG) 99/01 study. J Clin Oncol,
ASCO Annual Meeting Proceedings. Chicago, IL; 2010:
Abstr LBA4518.
17. Cognetti F, Ruggeri EM, Felici A, et al. Adjuvant
chemotherapy (AC) with cisplatin + gemcitabine (CG)
versus chemotherapy (CT) at relapse (CR) in patients (pts)
with muscle-invasive bladder cancer (MIBC) submitted to
radical cystectomy (RC). An Italian multicenter randomised
phase III trial. J Clin Oncol, ASCO Annual Meeting
Proceedings. Chicago, IL; 2008: abstr 5023.

References
1. Stein JP, Lieskovsky G, Cote R, et al. Radical cystectomy in
the treatment of invasive bladder cancer: long-term results
in 1,054 patients. J Clin Oncol. 2001;19(3):666-75.
2. Frank I, Cheville JC, Blute ML, et al. Transitional cell
carcinoma of the urinary bladder with regional lymph node
involvement treated by cystectomy: clinicopathologic features
associated with outcome. Cancer. 2003;97(10):2425-31.

18. Sternberg CN, Sylvester R, Theodore C, et al. Comparison


of Immediate and Delayed Adjuvant Chemotherapy
in Treating Patients Who Have Undergone a Radical
Cystectomy for Stage III or Stage IV Transitional Cell
Carcinoma of the Bladder Urothelium: EORTC 30994.
2001.

3. Igawa M, Urakami S, Shiina H, Ishibe T, Kadena H, Usui


T. Long-term results with M-VAC for advanced urothelial
cancer: high relapse rate and low survival in patients with a
complete response. Br J Urol. 1995;76(3):321-4.

19. Svatek RS, Shariat SF, Lasky RE, et al. The effectiveness
of off-protocol adjuvant chemotherapy for patients with
urothelial carcinoma of the urinary bladder. Clin Cancer
Res. 2010;16(17):4461-7.

4. Juffs HG, Moore MJ, Tannock IF. The role of systemic


chemotherapy in the management of muscle-invasive
bladder cancer. Lancet Oncol. 2002;3(12):738-47.
5. Quinn DI, Creaven PJ, Raghavan D. Principles of
chemotherapy for genitourinary cancer. In: Richie JP,
DAmico AV, eds. Urological Oncology. Philadelphia, PA:
Elsevier Inc.; 2005:57-81.

20. Hussain SA, James ND. The systemic treatment of


advanced and metastatic bladder cancer. Lancet Oncol.
2003;4(8):489-97.
21. Logothetis CJ, Johnson DE, Chong C, et al. Adjuvant
cyclophosphamide,
doxorubicin,
and
cisplatin
chemotherapy for bladder cancer: an update. J Clin Oncol.
1988;6(10):1590-6.

6. David KA, Milowsky MI, Ritchey J, Carroll PR, Nanus DM.


Low incidence of perioperative chemotherapy for stage III
bladder cancer 1998 to 2003: a report from the National
Cancer Data Base. J Urol. 2007;178(2):451-4.

22. Skinner DG, Daniels JR, Lieskovsky G. Current status of


adjuvant chemotherapy after radical cystectomy for deeply
invasive bladder cancer. Urology. 1984;24(1):46-52.

7. Donat SM. Integrating perioperative chemotherapy into


the treatment of muscle-invasive bladder cancer: strategy
versus reality. J Natl Compr Canc Netw. 2009;7(1):40-7.

23. Studer UE, Bacchi M, Biedermann C, et al. Adjuvant


cisplatin chemotherapy following cystectomy for bladder
cancer: results of a prospective randomized trial. J Urol.
1994;152(1):81-4.

8. Fedeli U, Fedewa SA, Ward EM. Treatment of muscle


invasive bladder cancer: evidence from the National Cancer
Database, 2003 to 2007. J Urol. 2010;185(1):72-8.

308

24. Loehrer PJ, Sr., Einhorn LH, Elson PJ, et al. A randomized
comparison of cisplatin alone or in combination with
methotrexate, vinblastine, and doxorubicin in patients with
metastatic urothelial carcinoma: a cooperative group study.
J Clin Oncol. 1992;10(7):1066-73.

38. Lotan Y, Gupta A, Shariat SF, et al. Lymphovascular


invasion is independently associated with overall survival,
cause-specific survival, and local and distant recurrence in
patients with negative lymph nodes at radical cystectomy. J
Clin Oncol. 2005;23(27):6533-9.

25. Stockle M, Meyenburg W, Wellek S, et al. Advanced bladder


cancer (stages pT3b, pT4a, pN1 and pN2): improved
survival after radical cystectomy and 3 adjuvant cycles of
chemotherapy. Results of a controlled prospective study. J
Urol. 1992;148(2 Pt 1):302-6; discussion 306-7.

39. Stadler W, Lerner SP, Groshen S, et al. Randomized trial


of p53 targeted adjuvant therapy for patients with organconfined node-negative urothelial bladder cancer. J Clin
Oncol (Meeting Abstracts). 2009;27(15S):A5017.
40. Dabholkar M, Bostick-Bruton F, Weber C, Bohr VA,
Egwuagu C, Reed E. ERCC1 and ERCC2 expression
in malignant tissues from ovarian cancer patients. J Natl
Cancer Inst. 1992;84(19):1512-7.

26. Stockle M, Meyenburg W, Wellek S, et al. Adjuvant


polychemotherapy of nonorgan-confined bladder cancer
after radical cystectomy revisited: long-term results of a
controlled prospective study and further clinical experience.
J Urol. 1995;153(1):47-52.

41. Houtsmuller AB, Rademakers S, Nigg AL, Hoogstraten


D, Hoeijmakers JH, Vermeulen W. Action of DNA repair
endonuclease ERCC1/XPF in living cells. Science.
1999;284(5416):958-61.

27. Freiha F, Reese J, Torti FM. A randomized trial of radical


cystectomy versus radical cystectomy plus cisplatin,
vinblastine and methotrexate chemotherapy for muscle
invasive bladder cancer. J Urol. 1996;155(2):495-9;
discussion 499-500.

42. Olaussen KA, Dunant A, Fouret P, et al. DNA repair by


ERCC1 in non-small-cell lung cancer and cisplatin-based
adjuvant chemotherapy. N Engl J Med. 2006;355(10):98391.

28. Millikan R, Dinney C, Swanson D, et al. Integrated therapy


for locally advanced bladder cancer: final report of a
randomized trial of cystectomy plus adjuvant M-VAC versus
cystectomy with both preoperative and postoperative
M-VAC. J Clin Oncol. 2001;19(20):4005-13.

43. Bellmunt J, Paz-Ares L, Cuello M, et al. Gene expression of


ERCC1 as a novel prognostic marker in advanced bladder
cancer patients receiving cisplatin-based chemotherapy.
Ann Oncol. 2007;18(3):522-8.

29. Lehmann J, Retz M, Wiemers C, et al. Adjuvant cisplatin


plus methotrexate versus methotrexate, vinblastine,
epirubicin, and cisplatin in locally advanced bladder cancer:
results of a randomized, multicenter, phase III trial (AUOAB 05/95). J Clin Oncol. 2005;23(22):4963-74.

44. Gandhi V, Mineishi S, Huang P, et al. Cytotoxicity,


metabolism, and mechanisms of action of 2,2difluorodeoxyguanosine in Chinese hamster ovary cells.
Cancer Res. 1995;55(7):1517-24.
45. Rosell R, Scagliotti G, Danenberg KD, et al. Transcripts
in pretreatment biopsies from a three-arm randomized
trial in metastatic non-small-cell lung cancer. Oncogene.
2003;22(23):3548-53.

30. Bajorin DF. Phase III Study Comparing Sequential


Chemotherapy (AG-ITP) To Cisplatin And Gemcitabine As
Adjuvant Treatment After Cystectomy For Transitional Cell
Carcinoma Of The Bladder: CALGB-90104. 2001.

46. Tewey KM, Rowe TC, Yang L, Halligan BD, Liu LF.
Adriamycin-induced DNA damage mediated by mammalian
DNA topoisomerase II. Science. 1984;226(4673):466-8.

31. Milowsky MI, Nanus DM, Maluf FC, et al. Final results of
sequential doxorubicin plus gemcitabine and ifosfamide,
paclitaxel, and cisplatin chemotherapy in patients with
metastatic or locally advanced transitional cell carcinoma
of the urothelium. J Clin Oncol. 2009;27(25):4062-7.

47. Hazlehurst LA, Valkov N, Wisner L, et al. Reduction in druginduced DNA double-strand breaks associated with beta1
integrin-mediated adhesion correlates with drug resistance
in U937 cells. Blood. 2001;98(6):1897-903.

32. Bellmunt J, Guillem V, Paz-Ares L, et al. Phase I-II


study of paclitaxel, cisplatin, and gemcitabine in
advanced transitional-cell carcinoma of the urothelium.
Spanish Oncology Genitourinary Group. J Clin Oncol.
2000;18(18):3247-55.

48. Kavallaris M, Kuo DY, Burkhart CA, et al. Taxol-resistant


epithelial ovarian tumors are associated with altered
expression of specific beta-tubulin isotypes. J Clin Invest.
1997;100(5):1282-93.

33. Bellmunt J, von der Maase H, Mead G, et al. Randomized


Phase
III
Study
Comparing
Paclitaxel/Cisplatin/
Gemcitabine and Gemcitabine/Cisplatin in Patients with
Locally Advanced or Metastatic Urothelial Cancer without
Prior Systemic Therapy: EORTC/Intergroup Study 30987 J
Clin Oncol, ASCO Annual Meeting Proceedings. Chicago,
IL; 2007: LBA5030.

49. Monzo M, Rosell R, Sanchez JJ, et al. Paclitaxel resistance


in non-small-cell lung cancer associated with beta-tubulin
gene mutations. J Clin Oncol. 1999;17(6):1786-93.
50. Skinner DG, Daniels JR, Russell CA, et al. The role of
adjuvant chemotherapy following cystectomy for invasive
bladder cancer: a prospective comparative trial. J Urol.
1991;145(3):459-64; discussion 464-7.

34. Sternberg CN, de Mulder P, Schornagel JH, et al. Seven


year update of an EORTC phase III trial of high-dose
intensity M-VAC chemotherapy and G-CSF versus classic
M-VAC in advanced urothelial tract tumours. Eur J Cancer.
2006;42(1):50-4.

V. Perioperative Radiotherapy
Adjuvant and neoadjuvant radiotherapy for muscleinvasive and locally advanced bladder cancer were
performed in the 1970s and 1980s of the last century
without many convincing results [1]. However, use
of modern radiotherapy in combination with surgery
resulted in prospective randomized clinical trials
or case-control studies which improved the local
control rate, frequency of distal metastases, and
overall survival in patients with locally advanced
bladder cancer [1] (level of evidence: 2b, grade of
recommendation: B)

35. Sonpavde G, Khan MM, Lerner SP, et al. Disease-free


survival at 2 or 3 years correlates with 5-year overall
survival of patients undergoing radical cystectomy for
muscle invasive bladder cancer. J Urol. 2011;185(2):45661.
36. Esrig D, Elmajian D, Groshen S, et al. Accumulation of
nuclear p53 and tumor progression in bladder cancer. N
Engl J Med. 1994;331(19):1259-64.
37. Shariat SF, Chade DC, Karakiewicz PI, et al. Combination
of multiple molecular markers can improve prognostication
in patients with locally advanced and lymph node positive
bladder cancer. J Urol. 2009;183(1):68-75.

309

1. Background

to freedom from distant metastasis (67% vs. 54%),


disease freedom (59% vs. 47%), and overall survival
(52% vs. 40%) although these differences did not
reach statistical significance. It has to be considered
that 65% of the T3b patients in the cystectomy
only group received systemic chemotherapy so
that the results might be better than expected. In
a further analysis of the patients data, Pollack et
al [8] identified pretreatment hemoglobin, blood
urea nitrogen, and preoperative radiotherapy as
the only independent predictors of local control.
Similar results were reported by Pollack et al [8]
in an additional analysis of the patient cohort who
identified pathological response (p<0.0001), clinical
stage (p<0.01), hemoglobin level (p<0.02), pathologic
complete response (p<0.05) and BUN concentration
(p<0.05) to be significantly associated with the pelvic
control rate. Restricting the analysis to pretreatment
parameters only, clinical stage, hemoglobin, BUN
level, and sex were predictive factors of pelvic
control. Pathologic response and tumor size were
independent predictive factors associated with
overall survival and pretreatment hemoglobin and
tumor size were the only factors associated with a
long-term disease-free status.

Radical cystectomy with urinary diversion represents


the treatment of choice for muscle-invasive bladder
cancer according to the most recent guidelines.
However, especially for patients with locally
advanced disease, the survival rates are dismal at
19-59% and 9-49% for patients with pT3 and pT4
disease, respectively [2, 3]. Although the causes
of failure are mainly distant metastases, local
recurrences are described in 23-51% of the cases.
In one series, rates of local recurrences were 6%,
18%, and 51% in patients with stage pT1, pT2,
and pT3 disease, respectively [4]. In another series
11%, 23%, and 31% of the patients developed local
recurrences within 5 years after radical cystectomy
[5]. The considerably high rate of local failures in
the various stages of bladder cancer suggests the
presence of residual micrometastasis in the pelvis,
either in the remaining pelvic lymph nodes or at
the site of the tumor bed. The median time to local
recurrence was 7=12 months whereas the median
time to the development of distant metastases was
16-20 months suggesting that local recurrences
might precede distant metastases [4-7]. It has been
demonstrated in some studies that local failures
besides pT stage and nodal metastases represent
an independent predictor of distant metastasis.
Furthermore, it has been shown that the presence
of local recurrences is associated with a poor
outcome and a median survival of 5-8 months
[4-7]. Based on these data, it seems reasonable to
sterilize microscopic metastasis or residues either by
preoperative or postoperative radiation therapy.

In another retrospective study, Granfors et al [9]


compared the outcome of 90 and 97 patients with
muscle-invasive bladder cancer who received
or did not receive neoadjuvant radiation therapy,
respectively. Neoadjuvant radiotherapy resulted in a
significant downstaging effect so that 7% and 57% of
patients without and with radiotherapy demonstrated
pT0 disease in the radical cystectomy specimen.
Among cT3 tumors, 0% and 56% of the patients
without and with radiotherapy achieved pT0 disease.
The progression-free survival was significantly
longer for patients who underwent neoadjuvant
radiotherapy (p<0.001); the disease-specific survival
time of patients with cT3 disease was significantly
higher after neoadjuvant radiotherapy as compared
to cystectomy alone (p = 0.007).

2. Preoperative Radiotherapy
Preoperative radiation therapy was used in the
1970s. However, conflicting results have been
produced over the years so that the concept of
preoperative radiotherapy to prevent intraoperative
tumor cell seeding and to eradicate microscopic
metastasis in the perivesical fat was not introduced
in daily routine.

b) Prospective Studies

a) Retrospective Studies

So far only 6 prospective randomized clinical trials


have been published with only one trial demonstrating
a statistically significant difference in the 2-year
disease-free survival [10-14].

Cole et al [8] and Pollack et al [8] published the largest


retrospective series on preoperative radiotherapy
and cystectomy versus cystectomy alone in muscleinvasive bladder cancer comprising 560 patients of
whom 338 and 232 patients received the combination
treatment of surgery alone, respectively. A mean total
dose of 49.3 Gy at 2 Gy per fraction was delivered 4
to 6 weeks prior to cystectomy. The median followup was 91 months and 54 months in the radiation
and the cystectomy alone groups, respectively. The
5-year local control rate in patients with stage T3b
was significantly improved in the radiation group
with 91% as compared to the cystectomy group
with 72% (p=0.03). Patients with T3b disease also
demonstrated a better 5-year outcome with regard

In one of the first trials, Slack et al [11] analyzed


the outcome of 234 patients with muscle-invasive
bladder who underwent preoperative radiotherapy
with 45 Gy and who underwent radical cystectomy
30 days after radiotherapy. The 5-year survival after
preoperative radiotherapy was 55% as compared to
only 32% in the cystectomy-only group. In another
prospective trial, Awwad et al [12] compared the
outcome of 36 and 17 patients with cT3 carcinoma
in bilharzial bladders who underwent preoperative
radiotherapy with 40-41 Gy followed by radical

310

cystectomy as compared to radical cystectomy. The


authors identified a significant 2-year survival benefit
of 53% in the radiotherapy group versus 19% in the
cystectomy group. However, the trial has several
drawbacks: (1) only patients with T3 disease were
included, (2) the patient numbers were small with only
53 recruited patients, and (3) patient numbers per arm
were not equally distributed, and (4) squamous cell
carcinoma of the bladder might have another clinical
course than urothelial cancer. Anderstrom et al [13]
evaluated the prognosis of a group of 44 patients
with cT1-3a urothelial bladder cancer of whom 22
received preoperative radiotherapy with 32 to 54 Gy.
The 3- and 5-year survival rates were 81% and 61%
in patients without radiotherapy and they were 81%
and 75% in patients with radiotherapy. Patients with
a complete response after radiotherapy (pT0) had
the best prognosis. Ghoneim et al [14] prospectively
randomized 93 patients with cT1-T4 bilharzial
bladder cancer to receive preoperative radiotherapy
with 20 Gy followed by radical cystectomy versus
cystectomy alone. After a minimum follow-up of 5
years, there was no statistically significant difference
between both groups with survival rates of 39%
and 32%. In the most recent study, Smith et al [15]
randomized a group of 140 patients with invasive
bladder cancer or rapidly recurring superficial high
grade tumors to receive 20 Gy of pelvic radiation
followed by cystectomy within 1 week or cystectomy
alone. The 5-year survival rate was 53% in the
combination group versus 43% in the cystectomy
group (p=0.23).

on the frequency and type of complications following


preoperative radiotherapy. Tomic et al [18] analyzed
the outcome of 103 patients who underwent radical
cystectomy after preoperative radiotherapy and
they observed intestinal complications in 39% of the
patients with an overall mortality rate in intestinal
complications in 3.3% of the cases. Reisinger et
al [19] described a 37% frequency of intestinal
complications with 67% of the patients requiring
salvage surgery and 20% of the patients dying
due to these complications. Cole et al [8] describe
a 2-month mortality rate of 3.6% which was higher
than the 2.1% rate in the cystectomy-only group.
However, the authors do not provide any long-term
data on treatment-associated complications and
mortality.

d) Postoperative Radiotherapy

c) Morbidity of Preoperative (Neoadjuvant)


Radiation Therapy

Postoperative radiotherapy has the benefit that the


group of patients who might benefit from radiotherapy can be better identified as compared to preoperative radiotherapy. Postoperative radiotherapy,
however, bears the significant problem that 2 doselimiting tissues are present in the pelvis: the intestinal epithelium on one hand, including intact loops of
bowel and also those of an orthotopic neobladder,
and the late responding connective tissues around
the anastomosis of the ureter, the urethra, and the
pelvis. In some reports, the frequency of radiation-induced obstruction of the small bowel was as high as
37% with 25% and 10% of the patients who required
surgery or died, respectively [19]. There are no prospective randomized clinical trials of postoperative
radiation therapy in patients with muscle-invasive
urothelial cancer of the bladder available. The only
prospective randomized clinical trial has been performed in patients with muscle-invasive carcinomas
in bilharzial bladders [20]. In this trial, 236 patients
with locally advanced cT3a to cT4 bladder cancer
were included and they received either postoperative radiotherapy multiple daily fractionation (MDF),
using 3 daily fractions of 1.25 Gy each, with 3 hours
between fractions, up to a total dose of 37.5 Gy in
12 days (75 patients); or postoperative radiotherapy
conventional fractionation (CF), for a total dose of 50
Gy/5 weeks (78 patients). Postoperative radiotherapy demonstrated a significant improvement with regard to 5-year disease-free survival which was 49%
and 44% for hyperfractionated and conventional
postoperative radiotherapy as compared to 25% for
cystectomy alone. The therapeutic benefit of postoperative irradiation was consistent for all tumor types,
histological grades, and pathological stages for both
the disease-free survival and local control in patients
with negative lymph nodes. Patients with nodal metastases demonstrated lower recurrence rates in the
postoperative radiotherapy groups, but this was not
associated with improved disease-free survival.

There are only a few studies available that concentrate

The only retrospective study on adjuvant radiation

A recent meta-analysis pooled the data from 5


randomized trials to compare 3- and 5-year survival
rates between patients who received preoperative
radiation therapy followed by cystectomy versus
patients who were treated by radical cystectomy
alone [16]. The corrected odds ratio was 0.94 (95%
CI 0.57-1.55) indicating a non-statistically significant
result. The clinical trial data achieved with this
meta-analysis do not support a significant role of
preoperative radiotherapy in patients with muscleinvasive bladder cancer, which was confirmed in
another recent systematic review in the role of
radiotherapy in bladder cancer [17]. However, it
is questionable if the data achieved in these trials
can be transferred into modern management of
bladder cancer in the 2000s for several reasons.
All of the randomized trials have been performed in
1970s and the 1980s and none of the trials could
consider modern high-precision techniques such as
intensity modulated radiotherapy or image-guided
radiotherapy. Therefore, there might be space for new
prospective randomized clinical trials in patients with
muscle-invasive bladder cancer without evidence of
distant metastases.

311

therapy in patients with urothelial bladder cancer


was published by Cozzarini et al [22], who analyzed
the oncological outcome of 150 patients. All men
received a total dose of 50 Gy to the pelvis and
achieved a disease-free survival rate and local
control rate of 44.6% and 88.4%, respectively. There
are, however, 3 updated reports from nonrandomized
postoperative radiotherapy studies in patients with
squamous cell and adenocarcinoma of the urinary
bladder [23-25]. The latest update comprised 216
patients who were treated with radical cystectomy
and pelvic lymphadenectomy with (82 patients)
or without (134) postoperative radiotherapy.
Postoperative radiotherapy was given to the whole
pelvis in a dose of 50 Gy/25 fractions over 5 weeks,
and started 4-10 weeks after surgery. Postoperative
radiotherapy improved the disease-free survival
significantly from 33 +/- 6% for cystectomy alone
to 58 +/- 6% for patients receiving postoperative
radiotherapy (P=0.002). The independent prognostic
factors for disease-free survival were the pathological
stage, histological subtypes, nodal involvement, and
the addition of postoperative radiotherapy.

References
1. Zaghloul MS. Adjuvant and neoadjuvant radiotherapy for
bladder cancer: revisited. Future Oncol. 2010 Jul;6(7):117791.
2. Stein JP, Lieskovsky G, Cote R, Groshen S, Feng AC,
Boyd S, Skinner E, Bochner B, Thangathurai D, Mikhail
M, Raghavan D, Skinner DG. Radical cystectomy in the
treatment of invasive bladder cancer: long-term results in
1,054 patients. J Clin Oncol. 2001 Feb 1;19(3):666-75
3. Dhar NB, Campbell SC, Zippe CD, Derweesh IH, Reuther
AM, Fergany A, Klein EA. Outcomes in patients with
urothelial carcinoma of the bladder with limited pelvic lymph
node dissection. BJU Int. 2006 Dec;98(6):1172-5. Epub
2006 Sep 6
4. Greven KM, Spera JA, Solin LJ, Morgan T, Hanks GE. Local
recurrence after cystectomy alone for bladder carcinoma.
Cancer. 1992 Jun 1;69(11):2767-70.
5. Visser O, Nieuwenhuijzen JA, Horenblas S; Members of the
Urological Oncology Working Group of the Comprehensive
Cancer Centre Amsterdam. Local recurrence after
cystectomy and survival of patients with bladder cancer: a
population based study in greater amsterdam. J Urol. 2005
Jul;174(1):97-102
6. Dhar NB, Jones JS, Reuther AM, Dreicer R, Campbell
SC, Sanii K, Klein EA. Presentation, location and overall
survival of pelvic recurrence after radical cystectomy for
transitional cell carcinoma of the bladder.BJU Int. 2008
Apr;101(8):969-72

Recommendations for Preoperative


Radiotherapy
Preoperative radiation therapy prior to radical
cystectomy results in a significant downstaging
effect on muscle-invasive bladder cancer (LE: 3,
GR: C)

7. Pejcic T, Hadzi-Djokic J, Acimovic M, Markovic B,


Maksimovic H, Milkovic B, Kajmakovic B. Local recurrence
of bladder cancer after cystectomy with orthotopic
bladder substitution and ileal conduit. Acta Chir Iugosl.
2007;54(4):63-7.

Preoperative radiation therapy results in a


significantly increased rate of pT0 tumors, which
is associated with a significantly improved local
control rate (LE 2b; GR: C)

8. Cole CJ, Pollack A, Zagars GK, Dinney CP, Swanson DA,


von Eschenbach AC. Local control of muscle-invasive
bladder cancer: preoperative radiotherapy and cystectomy
versus cystectomy alone. Int J Radiat Oncol Biol Phys.
1995 May 15;32(2):331-40

Preoperative radiation therapy should be


delivered with total dose of 49.3 Gy at 2 Gy per
fraction 4 to 6 weeks prior to cystectomy (LE: 2b,
GR 2)

9. Pollack A, Zagars GK, Dinney CP, Swanson DA, von


Eschenbach AC. Preoperative radiotherapy for muscleinvasive bladder carcinoma. Long term follow-up and
prognostic factors for 338 patients. Cancer. 1994 Nov
15;74(10):2819-27.

Preoperative radiation therapy does not result in


a significant survival benefit as compared to radical
cystectomy alone (LE: B, GR: ?)

10. Granfors T, Tomic R, Ljungberg B. Downstaging and survival


benefits of neoadjuvant radiotherapy before cystectomy for
patients with invasive bladder carcinoma. Scand J Urol
Nephrol. 2009;43(4):293-9.

Preoperative radiation therapy might be


associated with an increased risk of postoperative
intestinal complications and patients should be
informed accordingly (LE: 3, GR:C).

11. Blackard CE, Byar DP. Results of a clinical trial of surgery


and radiation in stages II and 3 carcinoma of the bladder. J
Urol. 1972 Dec;108(6):875-8.
12. Slack NH, Bross ID, Prout GR Jr. Five-year follow-up
results of a collaborative study of therapies for carcinoma
of the bladder. J Surg Oncol. 1977;9(4):393-405.

Recommendations for Postoperative


Radiotherapy

13. Awwad HK, Baki HA, El Bolkainy et al. Preoperative


irradiation of T3 carcinoma in Bilharzial bladder. Int J Radiat
Oncol Biol Phys 1979; 5: 787 794

Postoperative radiotherapy cannot be


recommended in patients with urothelial cancer of
the bladder (LE: 3, GR: C)

14. Anderstrm C, Johansson S, Nilsson S, Unsgaard B,


Wahlqvist L. A prospective randomized study of preoperative
irradiation with cystectomy or cystectomy alone for invasive
bladder carcinoma. Eur Urol. 1983;9(3):142-7.

Postoperative radiotherapy with a total dose of


50Gy to the whole pelvis improves disease-free
survival and local control in patients with locally
advanced, lymph node negative adenocarcinoma
of the bladder (LE: 2a, GR: B)

15. Ghoneim MA, Ashamallah AK, Awaad HK, Whitmore WF Jr.


Randomized trial of cystectomy with or without preoperative
radiotherapy for carcinoma of the bilharzial bladder. J Urol.
1985 Aug;134(2):266-8.

312

b) Background

16. Smith JA Jr, Crawford ED, Paradelo JC, Blumenstein B,


Herschman BR, Grossman HB, Christie DW. Treatment
of advanced bladder cancer with combined preoperative
irradiation and radical cystectomy versus radical
cystectomy alone: a phase III intergroup study. J Urol. 1997
Mar;157(3):805-7; discussion 807-8.

TUR monotherapy of invasive bladder cancers has


been discussed in the urologic literature since the
1940s[1]. Multiple groups evaluated the utility of
this approach either in isolation or in combination
with systemically administered poly chemotherapy.
[2-4] Comparing TUR to other standards of the day
including radical cystectomy, radiation monotherapy,
and combination therapy, Henry et al observed that
the 5-year survival rate for patients with T2 tumors
treated by TUR only was 63%.[5] In 1998, Solsona
and coworkers reported the results of a comparative,
non-randomized study of 133 patients with T2-3
bladder cancer treated by TUR. The clinical course
of these patients was compared to a concurrent
group of patients treated by radical cystectomy in
the same center.[6] These investigators reported
that for patients with negative TUR bed biopsies
following initial TUR, disease-specific survival was
equivalent to that observed in the cystectomy group.
Furthermore, patients in the TUR group with negative
muscle biopsies but with carcinoma in situ were found
to respond favorably to intravesical therapy. Herr et
al. reported on a similar experience in 155 patients
treated at a single institution with 10-year follow-up.
[7] Investigators form MD Anderson Cancer Center
reported that while TUR as monotherapy was
effective in appropriately selected individuals with
invasive bladder tumors, this approach was only
applicable to approximately 11% of the patients at
their center presenting with invasive bladder cancers.
[8] Recently, the Valencia group has updated its
experience with 15-year follow-up data. This most
recent report supports the authors contention that
in their hands at least, TUR monotherapy produces
high rates of disease-specific survival across all age
groups treated.[9]

17. Shelley MD, Wilt TJ, Barber J, Mason MD. A metaanalysis of randomised trials suggests a survival benefit for
combined radiotherapy and radical cystectomy compared
with radical radiotherapy for invasive bladder cancer: are
these data relevant to modern practice? Clin Oncol (R Coll
Radiol). 2004 May;16(3):166-71.
18. Widmark A, Flodgren P, Damber JE, Hellsten S, CavallinSthl E. A systematic overview of radiation therapy effects
in urinary bladder cancer. Acta Oncol. 2003;42(5-6):56781.
19. Tomic R, Granfors T, Modig H. Morbidity after preoperative
radiotherapy and cystectomy in patients with bladder
cancer. Scand J Urol Nephrol. 1997 Apr;31(2):149-53.
20. Reisinger SA, Mohiuddin M, Mulholland SG. Combined
pre- and postoperative adjuvant radiation therapy for
bladder cancer--a ten year experience. Int J Radiat Oncol
Biol Phys. 1992;24(3):463-8.
21. Zaghloul MS, Awwad HK, Akoush HH, Omar S, Soliman
O, el Attar I. Postoperative radiotherapy of carcinoma in
bilharzial bladder: improved disease free survival through
improving local control. Int J Radiat Oncol Biol Phys.
1992;23(3):511-7.
22. Cozzarini C, Pelegrini D, Fallini M, et al. Reappraisal of the
role of adjuvant radiotherapy in muscle-invasive transitional
cell carcinoma of the bladder. Int J Radiol Oncol Biol Phys
1999; 45: 221
23. Zaghloul MS, Mohran TZ, Saber RA, Agha N. Postoperative
radiotherapy in bladder cancer. J Egypt Nat Cancer Inst
2002; 14: 161 - 168
24. Zaghloul MS, Nouh A, Nazmy M, Ramzy S, Zaghloul
AS, Sedira MA, Khalil E. Long-term results of primary
adenocarcinoma of the urinary bladder: a report on 192
patients. Urol Oncol. 2006 Jan-Feb;24(1):13-20.
25. Zaghloul MS, El Baradie M, Nouh, Abdel-Fatah S, Taher A,
Shalaan M. Prognostic index for primary adenocarcinoma
of the urinary bladder. Gulf J Oncolog. 2007 Jul;(2):47-54.

VI. Bladder-sparing
Treatments for Localized
Disease

c) Indications
TUR monotherapy is appropriate, based on the
literature cited above, for the treatment of patients
with T2-T3 N0Mx bladder cancer in whom local
endoscopic resection is likely to produce complete
removal of the tumor exclusive of concomitant
non-invasive disease (i.e., CIS). Patients with
locoregional extension of disease as evidenced
by hydronephrosis or lymphadenopathy on axial
imaging are not considered good candidates for this
approach.

1. Transurethral Monotherapy for


the Treatment of Muscle-invasive
Bladder Cancer
a) Level of Evidence Reviewed
To date, no randomized studies have been performed
comparing transurethral resection (TUR) of invasive
(TNM stages T2-T4, No MX) bladder cancer as
monotherapy to other standard of care modalities
such as radical cystectomy or combined modality
therapy. The literature on this subject consists of
a few carefully performed clinical trials that are
observational in nature; these studies are generally
comparative, non-randomized, and uncontrolled
clinical experiences and are consistent at best with
level of evidence 2b, grade B.

c) Surgical Technique
Successful TUR of invasive bladder cancer is
confirmed by negative biopsies of the base and
periphery of the resection bed. In work cited above,
negative confirmatory biopsies correlated with longterm survival in those treated with TUR only.

313

d) Oncologic Outcome

8. Leibovici D, Kassouf W, Pisters LL, et al. Organ preservation


for muscle-invasive bladder cancer by transurethral
resection. Urology 2007;70:473-6.

TUR monotherapy should produce oncologic


outcomes equivalent to radical cystectomy in
appropriately selected patients according to longterm, single institution experiences reported by
Solsona and Herr (see bibliography). Bladder
preservation rates are also high in individuals with
small tumors completely resected with negative
follow-up biopsies.

9. Solsona E, Iborra I, Collado A, Rubio-Briones J, Casanova


J, Calatrava A. Feasibility of radical transurethral resection
as monotherapy for selected patients with muscle invasive
bladder cancer. J Urol 2010;184:475-80.

2. Partial Cystectomy
a) The Role of Partial Cystectomy
Partial cystectomy is an alternative form of therapy
for muscle-invasive disease that may be used in a
highly selected cohort of patients [1]. There is only
level 3 or 4 evidence regarding the appropriate use
of partial cystectomy for urothelial disease and it is
the treatment of choice for urachal adenocarcinoma,
which is not the subject of these guidelines. There are
no prospective phase II studies of partial cystectomy
nor are there prospective comparisons or randomized
trials assessing the relative efficacy of partial versus
radical cystectomy or radical transurethral resection
for invasive urothelial cancer.

Recommendations
1. TUR monotherapy is an alternative to radical
cystectomy in appropriately selected (see 2. below)
and counseled patients with T2-T3a N0Mx bladder
cancer (LE 2b, GR: B)
2. Patients most appropriate for this approach have
tumors that:
a. Are small
b. Are completely resectable
c. Have negative tumor bed and periphery
biopsies

A very small subset of patients (<5%) presenting with


muscle-invasive urothelial cancer will be eligible for
partial cystectomy when applying strict criteria. The
principal limiting factor is tumor location which should
be high on the dome or anterior wall away from the
bladder neck. Tumors on the posterior or lateral
walls may also be treated with partial cystectomy
if anatomically accessible [2, 3]. Tumors should be
primary rather than recurrent, and there should be
no CIS on bladder biopsies. A 2 cm circumferential
margin of normal mucosa is recommended and
tumors should be 3 cm or less in diameter.

d. Are not associated with upper tract compromise


(i.e., hydronephrosis)
e. Are not associated with radiographic evidence
of locoregional extension of disease (T3b, N+) at
the time of first treatment
3.
Should be discussed as part of the informed
consent process to patients contemplating
management options for invasive bladder tumors
(TNM Stages T2-3a, N0Mx; LE: 3, GR: C)

Partial cystectomy may also used in patients with a


tumor in a diverticulum in sites other than the dome
and may require ureteral re-implantation in select patients [4]. A recent series studied 39 patients treated
in a variety of ways, including partial cystectomy, for
tumor in a diverticulum [5]. Thirteen patients demonstrated T2 or greater disease and had a 45% 5-year
survival rate. Those patients with Ta and T1 disease
had better long-term survival (83% and 72%, respectively). A few small case series also report the
technical aspects and safety of laparoscopic/robotic
partial cystectomy for urothelial cancer and the use
of cystoscopy for tumor localization and initial identification of resection margins (LE 4C, GR C) [6, 7].

References
1. Flocks RH. The treatment of infiltrating carcinoma of the
bladder by transurethral resection. J Urol 1948;60:244.
2. Flocks RH. Treatment of patients with carcinoma of the
bladder. J Am Med Assoc 1951;145:295-301.
3. Barnes RW, Dick AL, Hadley HL, Johnston OL. Survival
following transurethral resection of bladder carcinoma.
Cancer Res 1977;37:2895-7.
4. Hall RR, Newling DW, Ramsden PD, Richards B, Robinson
MR, Smith PH. Treatment of invasive bladder cancer by
local resection and high dose methotrexate. Br J Urol
1984;56:668-72.

Several recent studies suggest that partial


cystectomy is over-utilized, particularly in nonacademic settings. In a population-based study
in Quebec, 30% of patients with invasive bladder
cancer underwent partial cystectomy over a 22-year
period [8]. Equally concerning is that only 23% of
patients had a pelvic lymphadenectomy and 24%
of patients required a salvage radical cystectomy.
Hollenbeck et al queried the SEER and National
Inpatient Sample databases from 1988-2000 and

5. Henry K, Miller J, Mori M, Loening S, Fallon B. Comparison


of transurethral resection to radical therapies for stage B
bladder tumors. J Urol 1988;140:964-7.
6. Solsona E, Iborra I, Ricos JV, Monros JL, Casanova J,
Calabuig C. Feasibility of transurethral resection for muscle
infiltrating carcinoma of the bladder: long-term followup of a
prospective study. J Urol 1998;159:95-8; discussion 8-9.
7. Herr HW. Transurethral resection of muscle-invasive
bladder cancer: 10-year outcome. J Clin Oncol 2001;19:8993.

314

found that in 2000, partial cystectomy was still


performed in 13-17% of patients and more commonly
in rural, non-teaching, low volume hospitals [9].
More recently, Fedeli and colleagues reviewed the
US National Cancer Database from 2003-2007 and
found a lower utilization that decreased over time
from 10% to 7% [10]. Capitanio et al reviewed the
SEER-9 database from 1988-2004 which included
7243 patients with stages pT1-4 N1-2 M0 disease
treated with partial (22%) or radical cystectomy [11].
They performed a matched analysis utilizing pT and
pN stage, grade, race, age, and year of surgery
and suggested that the use of partial cystectomy
did not undermine long-term cancer control. These
data should be interpreted with caution as within this
same database, 24% of patients did not appear to
have any node dissection and an additional 18% had
only 1-5 nodes removed, suggesting that surgical
quality was less that optimal regardless of the use of
partial or radical cystectomy.

many series in an effort to minimize the risk of wound


implantation, but there is no evidence to support their
routine use (LE 4C, GR C). Despite strict criteria of
cT1 or cT2 disease, 36% were upstaged to pT3 and
12% had node metastases, though only two-thirds of
patients underwent a pelvic lymphadenectomy. Fiveyear recurrence-free and disease-specific survival
probabilities were 62% and 84%, respectively, and
tumor size was the only variable associated with
recurrence. These data support a highly selective
use of partial cystectomy in patients with muscleinvasive bladder cancer (LE 3, GR B)

b) Partial Cystectomy after Neoadjuvant


Chemotherapy
Partial cystectomy has been incorporated into
bladder-sparing protocols after initial neoadjuvant
multi-agent chemotherapy in highly selected patients
with localized tumors. Sternberg et al reported on 13
patients among 104 treated with a bladder sparing
approach in mind who had partial cystectomy [12].
The 5-year survival for this select cohort was 69%.
In a separate report from MSKCC, 115 achieved
cT0 status after neoadjuvant chemotherapy and
28 underwent TUR alone and 15 underwent partial
cystectomy as definitive surgical management with
similar long-term survival of 74% for the combined
group with a median follow-up of 10 years [13]. There
may be a role for partial cystectomy as an alternative
to radical cystectomy following chemoradiation
therapy [14].

Two recent papers describe the contemporary


experiences from Memorial Sloan Kettering Cancer
Center and MD Anderson Cancer Center [2, 4]. The
former series described 58 patients with primary
non-urachal bladder cancer treated from 1995-2001
[4]. This represented 6.2% of patients presenting for
surgical therapy. All but 5 patients had a unilateral or
bilateral pelvic lymphadenectomy and 9 patients had
node metastases. A total of 7 patients had tumor in a
diverticulum. Of 5 patients with a final positive margin,
3 had negative intra-operative frozen sections of
the margins presumably due to sampling error.
On univariate analysis, CIS and multifocality were
associated with non-muscle-invasive recurrence
and lymph node metastases, and positive surgical
margins were associated with advanced recurrence.
CIS and node metastases were independent
predictors of advanced recurrence. Median followup was 31 months and 69% were alive and diseasefree while 22% died of disease.

Recommendations
Highly selected patients with focal invasive
cancers and cT0 or minimal residual disease after
neoadjuvant chemotherapy may be candidates
for bladder sparing with either TUR or partial
cystectomy (LE: 3, GR: C).

References

In the MD Anderson series, 37 patients underwent


partial cystectomy for curative intent between 1982
and 2003 [2]. All patients had pT2 or pT3 disease
and 14% had node metastases. Long-term cancer
control was achieved in 65% with an intact bladder
with a median follow-up of 53 months. Non-muscleinvasive recurrences occurred in 9 (24%) patients
and all were treated successfully. On multivariate
analysis, only pathologic tumor stage was associated
with recurrence-free survival.

1. Malkowicz SB, van Poppel H, Mickisch G, et al. Muscleinvasive urothelial carcinoma of the bladder. Urology. Jan
2007;69(1 Suppl):3-16.
2. Kassouf W, Swanson D, Kamat AM, et al. Partial cystectomy
for muscle invasive urothelial carcinoma of the bladder: a
contemporary review of the M. D. Anderson Cancer Center
experience. J Urol. Jun 2006;175(6):2058-2062.
3. Smaldone MC, Jacobs BL, Smaldone AM, Hrebinko
RL, Jr. Long-term results of selective partial cystectomy
for invasive urothelial bladder carcinoma. Urology. Sep
2008;72(3):613-616.

Smaldone et al reported a single surgeon series of 25


patients operated on over a 10-year period [3]. Their
protocol included 25 Gy of preoperative radiotherapy
delivered to the abdominal wall in 5 fractionated
doses, intraoperative intravesical Thiotepa, and
postoperative 6 weeks of intravesical BCG.
Preoperative radiotherapy and/or intra-operative
intravesical chemotherapy have been reported in

4. Holzbeierlein JM, Lopez-Corona E, Bochner BH, et al. Partial


cystectomy: a contemporary review of the Memorial SloanKettering Cancer Center experience and recommendations
for patient selection. J Urol. Sep 2004;172(3):878-881.
5. Golijanin D, Yossepowitch O, Beck SD, Sogani P, Dalbagni
G. Carcinoma in a bladder diverticulum: presentation and
treatment outcome. J Urol. Nov 2003;170(5):1761-1764.
6. Nerli RB, Reddy M, Koura AC, Prabha V, Ravish IR,

315

achieving local tumor control and patients are often


cured by contemporary cystectomy with major series
reporting 5-year pelvic control rates of 80-90%
and 5-year overall survival rates from 40% to 60%
[1-5]. It is this standard that treatment with bladderpreserving strategies must meet.

Amarkhed S. Cystoscopy-assisted laparoscopic partial


cystectomy. J Endourol. Jan 2008;22(1):83-86.
7. Mariano MB, Tefilli MV. Laparoscopic partial cystectomy in
bladder cancer -- initial experience. Int Braz J Urol. MayJun 2004;30(3):192-198.
8. Fahmy N, Aprikian A, Tanguay S, et al. Practice patterns
and recurrence after partial cystectomy for bladder cancer.
World J Urol. Aug;28(4):419-423.

Contemporary radiation-based bladder-sparing


therapy algorithms consist of: (1) maximal
transurethral resection of the tumor (TURBT),
(2) induction external beam radiation therapy
with concurrent chemotherapy, (3) cystoscopic
assessment of treatment response with prompt
cystectomy for non-responders, and (4) active
cystoscopic surveillance with a cystectomy at the
first sign of invasive recurrence. These algorithms
were developed as a result of the lack of adequate
local control of muscle-invasive urothelial carcinoma
treated by TURBT, by chemotherapy, or by
radiotherapy when used alone.

9. Hollenbeck BK, Taub DA, Dunn RL, Wei JT. Quality of


care: partial cystectomy for bladder cancer--a case of
inappropriate use? J Urol. Sep 2005;174(3):1050-1054;
discussion 1054.
10. Fedeli U, Fedewa SA, Ward EM. Treatment of muscle
invasive bladder cancer: evidence from the National Cancer
Database, 2003 to 2007. J Urol. Jan;185(1):72-78.
11. Capitanio U, Isbarn H, Shariat SF, et al. Partial cystectomy
does not undermine cancer control in appropriately
selected patients with urothelial carcinoma of the bladder:
a population-based matched analysis. Urology. Oct
2009;74(4):858-864.
12. Sternberg CN, Pansadoro V, Calabro F, et al. Can patient
selection for bladder preservation be based on response to
chemotherapy? Cancer. Apr 1 2003;97(7):1644-1652.

b) External Beam Radiation Alone with Salvage


Cystectomy Reserved for Tumor Recurrence

13. Herr HW, Bajorin DF, Scher HI. Neoadjuvant chemotherapy


and bladder-sparing surgery for invasive bladder cancer:
ten-year outcome. J Clin Oncol. Apr 1998;16(4):12981301.

From the 1960s through the 1980s, the most common


type of bladder-sparing treatment was external
beam radiation therapy alone. Since the 1980s in the
United States radiation treatment has generally been
reserved for patients judged too unfit for cystectomy
on the basis of comorbid conditions or due to
disease extent. These negative selection criteria
may have contributed to the relatively poor results
achieved with radiation therapy alone compared
to cystectomy (see Table 5). Approximately 1015% of patients are excluded from treatment by
radical cystectomy at the time of operation because
previously unrecognized unresectable tumor spread
is found. Thus in cystectomy series, but not radiation
series, some of the patients with advanced local
spread tumor are excluded so this may be another
selection bias favoring cystectomy.

14. Koga F, Kihara K, Fujii Y, et al. Favourable outcomes of


patients with clinical stage T3N0M0 bladder cancer treated
with induction low-dose chemo-radiotherapy plus partial
or radical cystectomy vs immediate radical cystectomy: a
single-institutional retrospective comparative study. BJU
Int. Jul 2009;104(2):189-194.

3. Radiation-based Bladderpreserving Strategies: Radiation


Alone or Combined with Other
Modalities
a) Introduction
The treatment options for muscularis propria-invasive
bladder tumors can broadly be divided into those
that involve removal of the bladder and those that
spare it. There is a big difference among countries
in the path of care that they most often use to treat
these patients. For instance, in the United States
only a minority of the patients are offered radiation
therapy. However, in the United Kingdom, 60% of
eligible patients receive radical radiation therapy,
with surgery reserved for failures. The treatment of
many other cancers in North America, Europe, and
around the world, include organ-preserving therapy
as the, or one of the, current standards of care
for malignancies of the breast, larynx, anus, head
and neck, soft tissues (sarcoma), and prostate. In
each case, radical surgical extirpation can often
be avoided without compromising patient survival.
Improved radiotherapy techniques combined
with an enhanced understanding of the optimal
chemotherapeutic regimens have promoted multimodality therapy in selected patients with muscleinvasive bladder cancer as a viable alternative to
radical surgery alone. Cystectomy is effective in

In the 1960s, the Cooperative Surgical Adjuvant


Bladder Study group randomized 475 patients
with Stage T2-T4a bladder cancer to preoperative
radiation therapy (45 Gy) followed by open surgery
or surgery alone [13]. The study was large but
compromised by incomplete data collection and
follow-up. Of the 138 patients who completed
preoperative radiation therapy and surgery on
protocol, 34% had a complete pathologic response
of their bladder tumor (stage pT0). Furthermore,
those with a complete pathologic response at radical
cystectomy had a survival advantage of 55% versus
32% compared to those who had residual tumor at
the time of cystectomy. This led in the 1970s to 4
randomized trials comparing external beam radiation
therapy alone (60 Gy) with radical cystectomy
reserved for local recurrence to the standard group
receiving preoperative radiation therapy (40-50
Gy) with immediate radical cystectomy [6,10,14]

316

(Table 6). Three of these trials showed equivalent


overall survival with either approach. These studies
give Level 1b evidence that a bladder-preserving
approach with radiation therapy alone and salvage
radical cystectomy for local recurrence was not
significantly different in overall survival in this preneobladder era.

salvage cystectomy for the non-responders were


studied. The randomized trial from the United
Kingdom studied accelerated (twice-daily) radiation
monotherapy in 229 patients with muscle-invasive
urothelial carcinoma treated with radiation therapy
alone [15]. There was no advantage with the twicedaily schedule over the conventional once-daily
radiation schedule, and acute bowel and bladder
toxicities were higher. The biological rationale
underlying accelerated fractionation radiation therapy
in bladder cancer therefore remains unproven. Thus,
off protocol, the oncedaily radiation treatments
remain entirely reasonable.

Alternative radiation therapy fractionation schemes,


including the use of twice-daily radiation with the
potential advantage of improved biological effect to
better control of rapidly proliferating tumors, and with
the practical advantage of a more rapid completion
of radiation therapy and thus a shorter interval to

Table 5. (Section 5.3) Results of radical radiation therapy alone (Monotherapy): Muscle-Invading
Bladder Cancer
Study
M.D. Anderson Cancer Center [6]
National Bladder Cacer Group a
Eudinburgh [7]
London Hospital [8]
Princess Margaret Hospital [9]
Danish National Study [10]
Norway [11]
UK Cooperative Group [12]

Number
32
35
889
182
121
95
308
91

T2

40%
46%
59%
38%

5-year survival rates


T3(T4a)
All Stages

26%
35%
39%
23%
14%
28%

40%
36%
40%
45%
24%

SD Cutler, National Cancer Institute, Unpublished observations, 1983.


Modified from Zietman et al. Organ-Conserving approaches to muscle-invasive bladder cancer: future alternatives
to radical cystectomy. Ann Med 2000;32: 42-47
a

Table 6. (Section 5.3) Five year survival data from four randomized trials comparing preoperative radiation therapy (40-50 Gy) with immediate cystectomy to radiation therapy alone (60 Gy) with salvage
cystectomy recurence.
Study

No. of patients

5-year
Survival with
Pre-op RT and
cystectomy
(%)

Satistical
Significance

39

5-year
Survival with
RT and
Salvage
cystectomy
(%)
28

Urologie
Cooperative\
Group, UK
[14]
Danish
National
Cancer Group
[10]
National
Bladder
Cancer Group
(a)
MD Anderson
Cancer Center
[6]

189

183

29

23

none

72

27

40

none

67

45

22

significant

RT = radiation therapy
(a) SD Cutler, National Cancer Institute, unpublished observations, 1983

317

Notes

none

Large
T3
tumors
included

c) External Beam Radiation Therapy Combined


with Other Modalities and with Salvage
Cystectomy for Recurrence

of chemotherapy and radiation therapy combined


with TURBT. A study of 104 patients treated with
TURBT and M-VAC showed a T0 tumor response
rate of 49% and in these 52 patients the 5-year
survival rate was 67% [26]. However, 66% of the
104 patients required radical cystectomy, suggesting
a relatively low local bladder tumor control with
chemotherapy and TURBT alone. As described
below, modern trimodality therapy combining
radiation and chemotherapy with TURBT has led to
substantially higher T0 tumor response rates (6487%) and lesser need for salvage cystectomy. This
has tempered further interest in the treatment with
only chemotherapy and TURBT.

By the late 1980s, single institutions reported the


combination of a visibly complete TURBT followed
by radiation therapy lead to improved local control
[16, 17]. By 1994, the group from the University of
Erlangen referred to the visibly complete TURBT,
when followed by a tumor bed negative biopsy, as
an R0 resection [17]. The importance of the visibly
complete TURBT was also seen in many subsequent
cooperative group trials using radiation concurrent
with chemotherapy such as in RTOG 89-03 [18].
Combining external beam radiation beam therapy
with brachytherapy (for the delivery of precision
partial bladder radiation therapy) has been used by
some groups in Holland, Belgium, and in France.
Brachytherapy is done by an open cystotomy with
an implant using iridium-192 as low dose rate
irradiation to doses of 40 Gy combined with external
beam radiation doses of 30 Gy. The majority of the
reported series include patients who first underwent
maximal tumor resection by either partial cystectomy
or an open TURBT. The approach is reserved for
patients with solitary bladder tumors less then 5 cm
in diameter. The 5-year survival rates reported are
from 62-84% with a disease-specific survival rate
approaching 80% [19-21].

A second randomized trial comparing radiation


alone to radiation plus concurrent chemotherapy
with 5-fluorouracil (5-FU) and mitomycin C (MMC)
has been reported by a multi-center group led
from Birmingham, England, involving 360 patients.
The results showed no measurable differences in
toxicities. There was a significant increase in pelvic
disease-free rates at 2 years (67% free of recurrence
compared to 54% with radiation alone, P=0.02) and
at 5 years (62% and 51% respectively) [28]. The fiveyear overall survival rate was increased from 35% to
51%. Seventy-five percent of the patients treated
with 5-FU and MMC concurrent with radiation
therapy reported no late side effects. Of the 25%
who did report side effects, fewer then 5% reported
them as serious. Patients on this protocol underwent
urologic evaluation of their bladder capacity before
and after treatment. There was a median reduction
in bladder capacity of 10%, which was the same in
both groups.

1. E
 xternal Beam Radiation Combined with
Concurrent Radiosensitizing Chemotherapy
Encouraging results were obtained when cisplatin
was first made available by the NCI to the National
Bladder Cancer Group in the early 1980s for
patients with muscle-invasive bladder cancer who
were unsuitable for radical cystectomy. In a protocol
combining the use of concurrent cisplatin with
conventional doses of radiation was carried out from
1981 to 1986 in 68 patients with muscle-invasive
bladder cancer, 64% of patients with clinical stage
T2 tumors and 22% of those with clinical stage
T3-T4a disease had long-term survival rates [22].
These promising findings of combining cisplatin
with radiation led to a randomized trial of radiation
therapy with or without concurrent cisplatin [23] in
99 patients with clinical stage T3 muscle-invasive
bladder cancer conducted by the National Cancer
Institute of Canada. This trial showed a significant
improvement in pelvic tumor local control at 5
years in patients treated with cisplatin and radiation
therapy (68%) versus radiation therapy alone (47%).
Similarly, early prospective studies at the MSKCC
that combined chemotherapy and TURBT resulted
in T0 tumor response rates nearly twice that of
chemotherapy alone, but these rates were less than
50% [24,25]. The results of combining TURBT with
chemotherapy but without radiation have not been as
successful in organ-preservation as the combination

Cooperative group and single institution trials


with concurrent chemotherapy and radiation
combined with TURBT. Over the last 25 years,
single institutions in North America and Europe and
multi-institutional cooperative groups including the
Radiation Therapy Oncology Group (RTOG) and the
Southwestern Oncology Group (SWOG) have enrolled
over 1000 patients with muscle-invasive bladder
cancer in bladder-preserving protocols. Several
variables have been tested including evaluating
more than cisplatin alone as the radiation-sensitizing
chemotherapy and evaluating alternative radiation
schemes. Beginning in the late 1980s some single
institutions and the RTOG were studying neoadjuvant
chemotherapy in addition to trimodality therapy for
the operable patients with muscle-invasive bladder
cancer (Table 7). Encouraging results led to the
opening of RTOG 8903, a Phase III trial comparing
neoadjuvant MCV chemotherapy, or not, prior to
concurrent cisplatin and radiation. This study was
closed prematurely after accrual of 123 of the planned
174 patients because there was an unexpectedly
high rate of leukopenia in the MCV arm [18]. With
a median follow-up of 5 years, the overall survival
rate was 48%, and was 49% in patients who were

318

Table 7. (Section 5.3) Results of tri modality treatment for muscle-invasive bladder cancer for selective
bladder preservation
5-year
survival
with intact
bladder
(%)

Study
loention

Multimodality therapy
used

Number of
patients

5-year
overall
survival
(%)

TURBT, MCV, external


beam
radiation + cisplatin

91

62 (4 yr)

44 (4 yr)

RTOG 8802

Tester et al.
[29]

TURBT, 5-FU, external


beam
radiation + cisplatin

120

63

NA

University of
Puris

Housset et.
[30]

TURBT, +/- MCV,


external beam
radiation + cisplatin

123

49

38

RTOG 8903

Shipley et al.
[18]

TURBT, external beam


radiation + cisplatin,
carboplatin, or 5-FU
and cisplatin

415

50

42

University of
Erlangen

Rodel et al.
[31]

TURBT, external beam


radiation + TAX and
cisplatin, adj. CEM
and cisplatin

80

56

47

RTOG 9906

Kaufman et al.
[32]

TURBT, MCV, external


beam
radiation + cisplatin

348

52

44

MGH

Efstathiou et
al. [33]

References

TURBT, transurethral resection of bladder tumor; tumor, 5-FU, 5-fluorouracil; MCV, methotrexate, cisplatin, vinblastine, TAX pactiaxel; GEM gemcitabine; RTOG, Radiation Therapy Oncology Group, MGH, Massachuscits General
Hospital; NA, not available.

randomized to the neoadjuvant MCV arm. Likewise,


there was no statistically significant difference in the
T0 tumor response rate, distant metastasis, or the
5-year survival with an intact bladder. Two European
phase III trials have studied the role of neoadjuvant
chemotherapy before radiation alone. The Danish
Group [36] treated patients with muscle-invasive
bladder cancer with neoadjuvant chemotherapy
before radiation and showed an insignificant 5%
decrease in 5-year survival compared to patients
treated with radiation alone (19% v. 24%). The
other phase III trial led in the United Kingdom by
the MRC Group studied neoadjuvant MCV before
radical radiation or radical cystectomy (34-35). In
the radical radiation sub-group of 415 patients, there
was an insignificant trend to better survival in those
treated with neoadjuvant MCV than those treated
with radiation alone. A meta-analysis of those 2
trials showed no significant difference in survival
(30.4% vs. 28.1%, P=0.334) with the addition of the
neoadjuvant chemotherapy [37].

therapy was reserved for those patients who had


a complete clinical response at the mid-point of
concurrent chemoradiation after radiation dosage of
40 (Gy). These patients then received consolidation
with additional chemotherapy and radiation for the
total dose of 64-65 Gy. Incomplete responders were
advised to undergo a radical cystectomy as were
patients whose invasive tumors recurred after the
full 64-65 Gy treatment. All patients were treated with
an aggressive TURBT, which was visibly complete
in 66% of patients. All patients were treated with
cisplatin concurrently with radiation therapy. With
the median follow-up for all surviving patients of 7.7
years, the 5-year, 10-year, and 15-year actuarial
overall survival rates were 52%, 35%, and 22%,
respectively (stage T2, 62%, 43%, and 28%; stage
T3-T4a, 41%, 27%, and 16%). The 5-year, 10-year,
and 15-year disease-specific survivals was 64%,
59%, and 57% (stage T2, 74%, 67%, and 63%; stage
T3-T4a, 53%, 49%, and 49%). The disease-specific
survival rates stratified by clinical stage are shown
in Figure 4, which demonstrates that there were
very few late recurrences at least up to 15 years.
These results are similar to those in contemporary
cystectomy series. In this series, the 5-year, 10year, and 15-year disease-specific survival for the

The Massachusetts General Hospital experience with


348 patients with muscle-invasive bladder cancer
who were entered on the successive prospective
trimodality protocols from 1986-2002 have recently
been updated [33]. Bladder-sparing trimodality

319

TURBT

Induction radiation and concurrent


chemotherapy

Repeat cystocopy with


transurethral biopsy
Complete Response

Incomplete Response

Consolidation chemotherapy and


radiation # adjuvant
chemotherapy AND long-term
cystoscopic surveillance

Radical cystectomy # adjuvant


chemotehrapy

Recurrent tumor
Figure 3. Current Schema for Trimodality Treatment of Muscle-invasive Bladder Cancer With Selective
Bladder Preservation.

Figure 4. Long-term Disease-specific Survival with Selective Bladder Preservation from the Massachusetts
General Hospital Experience (33).

320

102 patients undergoing cystectomy were 55%,


44%, and 44%, respectively. This indicates the
very important contribution of prompt cystectomy
for disease control in patients whose tumors recur.
A recent evaluation of patients undergoing salvage
cystectomy at the Massachusetts General Hospital
over these years indicates quite acceptable surgical
morbidity or mortality compared to major primary
cystectomy series. (Eswara & Heney, submitted
2010).

patients with tumors pathologic stage pT2-pT4, the


5-year overall survival rate was 36%[2]. The actuarial survival rate of all 269 patients with pathologic
stages ranging from pT0 to pT4 in this series was
45%. These results are similar to the Massachusetts
General Hospital Series [33] as well as those from
University of Erlangen (17,31) and RTOG [18]. The
similarity in survival is likely in part to the prompt use
of cystectomy when necessary for recurrence in the
bladder-preservation series.

d) Comparison of Survival Outcomes Following


Curative Therapy in Contemporary Series
by Cystectomy or by Bladder-preserving
Trimodality
Therapy
with
Cystectomy
Reserved for Recurrence

e) Quality of Life After Radiation-based


Bladder-preserving Therapy for Muscleinvasive Bladder Cancer
The instruments to assess quality of life have
been well established for prostate cancers and
gynecologic cancers, but not for bladder cancer.
The instruments that are currently used for bladder
cancer patients are adaptations and thus their
validity is somewhat uncertain. These studies are
also limited by incomplete sampling of all potential
participants which leaves unclear whether or not the
non-participants are those who have had a worse
outcome or who are the most satisfied. Despite
these limitations, some general principles can be
derived from this literature.

Comparing results of bladder-preserving therapy


to those of contemporary radical cystectomy series
is confounded by the discordance between clinical
staging (TURBT) and pathologic (cystectomy) staging. Clinical staging is more likely to underestimate
the extent of disease with regard to penetration into
the muscularis propria or beyond than is pathologic
(cystectomy) staging [38]. Thus if any favorable outcome bias exists among these selected, it is in favor
of the pathologically-reported cystectomy series. For
patients with muscle-invasive bladder cancer, the
overall survival outcomes following either contemporary radical cystectomy at major single institutions or
by trimodality therapy are shown in Table 8. The University of Southern California reported on 633 patients undergoing radical cystectomy with pathologic
stages T2-T4a with an overall survival rate at 5 years
of 48%, and at 10 years of 32% [1]. The Memorial
Sloan Kettering Cancer Center (MSKCC) contemporary radical cystectomy series showed that in 184

Minimal late pelvic toxicity is certainly required for


successful implementation of a selective bladderpreserving protocol. Long-term bowel and bladder
toxicity after chemoradiotherapy was investigated
in patients enrolled in prospective sequential RTOG
trials (8903, 9506, 9706, and 9906). One study
reported on 157 patients who underwent combined
modality therapy and who survived at least 2 years
from the start of treatment with their bladders intact.
The median follow-up was 5.4 years [39]. Seven

Table 8. (Section 5.3) Muscle-Invasive Bladder Cancer: Survival Outcomes Following Curative Therapy in
Contemporary Series

Series
Cystectomy:
U.S.C.
2001 [1]
M.S.K.C.C.
2001 [2]
SWOF/ECOG/CALGB*#
2002 [3]
Selective Bladder Preservation:
U. Erlangen*
2002 [17,31]
M.G.H.*
2009 [33]
R.T.O.G.
1998 [18]

Stages

Number

pT2-pT4a

633

pT2-pT4a

181

cT2-cT4a

317

cT2-cT4a

326

cT2-cT4a

348

cT2-cT4a

123

* These series include all patients by their intention-t-treat.


# 50% of patients were randomly assigned to receive 3cycles of neoadjuvant M-VAC.

321

Overall Survival
5 year
10 year
48%

32%

36%

27%

49%

34%

45%

29%

52%

35%

49%

--

percent of the patients experienced late grade 3 or


4 pelvic toxicity (5.7% GU and 1.9% GI). In only 1 of
9 patients, did a grade 3-4 GU toxicity persist. This
indicates that rates of late pelvic toxicity for patients
who undergo selective bladder preservation and
retain their native bladder are low.

dissatisfaction with their sexual lives. In the Swedish


and Italian series 38% and 25% of the men retained
functional erections as compared to 15% and 8% of
cystectomy controls. Use of sildenafil by patients in
the MGH series may have been the major reason for
better retained erectile function.

Zeitman and colleagues reported a study on patients


receiving TURBT, chemotherapy, and radiation
in the treatment of the bladder cancers at the
Massachusetts General Hospital (MGH) [40]. Of
221 patients with clinical stage T2-T4a cancer of
the bladder treated at the MGH from 1986-2000, 71
were alive with their native bladders and diseasefree in 2001. These patients were asked to undergo
a urodynamic study (UDS) and to complete a quality
of life questionnaire. Sixty-nine percent participated
in some component of the study with a median time
from the trimodality therapy of 6.3 years. This long
follow-up is sufficient to capture the majority of the late
radiation side effects. Seventy-five percent of patients
have normal functioning bladders by UDS. Reduced
bladder compliance, a recognized complication of
radiation, was seen in 22% of patients. However,
distressing bladder symptoms were seen only in
one-third of these patients. Two women showed
bladder hypersensitivity, involuntary contractions,
and incontinence. The questionnaires showed that
bladder symptoms were uncommon overall in both
sexes. However, 19% of women reported problems
with bladder control, and 11% of them wore pads.
The distress from urinary symptoms was only half as
common as symptom prevalence. Bowel symptoms
occurred in 22% of patients and caused distress in
14%. The majority of men retained sexual function
with or without use of sildenafil. Global healthrelated quality of life was high. The majority of the
patients treated by trimodality therapy therefore
retained good bladder function. It was concluded
that there is a small but detectable level of lasting
bowel dysfunction and distress and that this might
be judged as the additional price that these patients
have to pay to retain their bladders.

f) Translational Research: Prognostic or Predictive Molecular Tumor Markers as Predictors of Response to Radiation Treatment
Tumor suppressor genes such as p53 and pRB
have been studied in detail in bladder cancer, but
both markers have led to contradictory data in the
assessment of risk for disease--progression and
survival [43]. Cell-cycle regulatory proteins p27
and Ki-67 might predict recurrence and disease
progression but are not ready for routine clinical use
[44-46].
The bladder tumors pre-treatment apoptotic index or
altered expression of the RB1 or the BCL2 genes
might alter tumor response to radiation therapy [4749]. RTOG investigated the outcome of 73 patients
treated in 4 RTOG bladder-preserving protocols and
noted that among patients treated with transurethral
surgery and chemotherapy concurrent with radiationaltered expression of p53, CDKN2A, and pRB had no
prognostic significance but overexpression of Her-2
(ERBB2) correlated significantly with a reduced
complete response rate (50% vs. 81%, p=0.026).
The aim of targeted therapies is to interfere with
molecular events related to tumor proliferation [50].
Examples of these therapies are cetuximab, an antiEGFR monoclonal antibody; gefitinib and erlotinib,
EGFR-specific inhibitors; trastuzumab, an antihuman EGFR Type 2 (Her-2)-related monoclonal
antibody; and bevacizumab, a VEGF monoclonal
antibody [51]. Both EGFR and HER-2 are targets
identified on cancer cells, whereas VEGF is a target
that acts on the tumor microenvironment. Studies
have shown reduced complete response rates when
HER-2 is over expressed [51-53]. These results
have led to an ongoing RTOG protocol (RTOG0524) for patients with muscle-invasive bladder
cancer who are not fit for a cystectomy. This phase
I / II trial investigates paclitaxel and daily radiation
therapy with trastuzumab given to patients whose
tumors over express HER-2. This study is one of
the first examples of a molecular targeted therapy
being added to treatment for patients with localized
muscle-invasive bladder cancer. To date 55 patients
have been enrolled with the results still pending.
Recently the group in Leeds and in Oxford in the
United Kingdom, in collaboration with the Ontario
Cancer Institute in Toronto, have evaluated MRE
11 expression in patients with muscle-invasive
bladder cancer treated with radical radiation or with
cystectomy [55]. MRE 11 is one of a panel of DNA

Two cross-sectional questionnaire studies, one


from Sweden and one from Italy, have compared
the outcome following radiation with the outcome
following cystectomy [41-42]. The questionnaire
results for urinary function following radiation are
very similar to those recorded in the MGH study. Over
74% of the patients recorded good urinary function.
Both studies compared bowel function in irradiated
patients with that seen in patients undergoing
cystectomy. In both, the bowel symptoms were
greater for those receiving radiation than for those
receiving cystectomy (10% vs. 3% and 32% vs.
24% respectively), but in neither was this statistically
significant. In contrast to men who had been irradiated
for prostate cancer, the majority of the male bladdersparing patients reported adequate erectile function
(full or sufficient for intercourse) and only 8% reported

322

damage signaling proteins active in the process of


DNA double strand break repair. DNA double stand
breaks are the most lethal form of injury produced
by ionizing radiation and by some chemotherapy
agents. MRE 11 had been singled out as one
possible predictor of radiation treatment outcome.
The cohorts of patients treated with radical radiation
and a separate cohort of patients also treated in
Leeds by radical cystectomy have documented
that high MRE 11 protein expression by the tumor
predicts improved outcome with radiation therapy
but not in the cystectomy cohorts. Comparing the
high MRE 11 patients by treatment showed that the
patients treated with radiation have a significantly
higher disease-specific survival than those treated
with immediate cystectomy (P=0.02) while those
with low MRE 11 protein expression do insignificantly
better with surgery than with radiation. These results
tentatively identify MRE 11 overexpression as a
predictive molecular marker of improved causespecific survival following radiation therapy for
muscle-invasive bladder cancer.

and survival in selected patients comparable


to radical cystectomy, and affords a 70%
chance of maintaining a well-functioning native
bladder (see Table 8). Quality of life studies
have demonstrated that the retained native
bladder functions well and long-term toxicity of
chemoradiation to pelvic organs is low. These
reports support the acceptance of modern
bladder-sparing trimodality therapy for selected
patients as a proven alternative to cystectomy
(LE 3, GR C).
See Table 9 for
recommendations.

summary

of

these

Conclusion
In selected patients with muscle-invasive bladder
cancer, bladder-preserving therapy with cystectomy
reserved for tumor recurrence represents a safe
and effective alternative to an immediate radical
cystectomy. Cumulative published data of more
than 1000 patients in single institution and multiinstitution cooperative group trials demonstrate that
trimodality therapy results in excellent local control in
70% of patients with muscle-invasive bladder cancer
while preserving a native functional bladder without
compromising long-term survival. The 10-year
overall survival and disease-specific survival rates in
the bladder-sparing protocols are comparable to the
overall results reported with contemporary radical
cystectomy. Moreover, the 15-year results indicate
a plateau in disease-specific survival, suggesting
no evidence of increased rates of recurrence with
longer follow-up times. Life-long bladder surveillance
is essential. Prompt cystectomy for tumor recurrence
is necessary to prevent tumor dissemination. Thus,
bladder-preserving therapy is a bona fide option and
valid alternative to radical cystectomy in selected
patients. This approach should be discussed along
with all of the other treatment options during overall
initial treatment planning. This approach contributes
significantly to the quality of life of the patients so
treated and represents a unique opportunity for
urologic surgeons, radiation oncologists, and medical
oncologists to work hand-in-hand in a joint effort to
provide patients with the best treatment option for
this disease.

g) Levels of Evidence (LE) and Grades of


Recommendation (GR) for Radiation-based
Bladder-preserving Strategies for Muscleinvasive Bladder Cancer
1. Radiation therapy followed by salvage cystectomy
for tumor recurrence has comparable survival to
preoperative radiation therapy and cystectomy.
LE 1b, GR A).
2. Radiation therapy and chemotherapy result in
a higher rate of pT0 status than does radiation
therapy alone (LE 1b, GR A) .
3. Combined radiation and chemotherapy allow
good preservation of bladder function in the
great majority of patients. (LE 2a, GR B).
4.There is no clinical trial basis to indicate that
neoadjuvant chemotherapy prior to radiation
therapy improves survival. (LE 1a, GR A).
5. Complete TURBT, when possible, is associated
with higher rates of local tumor control and higher
cure rates than does incomplete initial tumor
resection for selected patients in trimodality
radiation/chemotherapy trials. (LE 2a, GR B).
6. Limited data suggests that high expression of
the molecular marker MRE 11 may be a putative
predictor for cause-specific survival following
radical radiation therapy for muscle-invasive
bladder cancer. (LE 3, GR B).
7. T
 rimodality therapy consisting of TURBT plus
concurrent radio sensitizing chemotherapy and
radiation is judged safely possible, and, when
combined with early salvage cystectomy for
recurrence, this bladder-preserving treatment
approach offers a chance for long-term cure

323

Table 9. Summary of Recommendations


Treatment/
Comparison

Evidence

Level of
Evidence

Grade of
Recommendation

RT alone vs
40Gy+Cystectomy

3 of 4 RCTs report
similar survival

1b

ChemoRT vs
RT alone

1b

Neoadjuvant CT
with RT or ChemoRT

2 RCTs report significant


improvement in bladder
tumor eradication
3 RCTs and 1 meta-analysis
report no benefit

1a

ChemoRT preserves
good bladder function

3 QOL studies and RTOG protocols


report good tolerance

2a

Complete TURB
with ChemoRT

3 reports (1 phase III, 2 phase II)


show benefit

2a

Predictive Biomarkers
of outcome after RT

MRE 11 expression predicts


improved CSS (1 study)

Trimodality therapy vs
immediate cystectomy

Comparison of 3 contemporary
series of each report similar
5- and 10-yr survival

9. Gospodarowicz MK, Hawkins NV, Rawlings GA, et al.


Radical radiotherapy for muscle invasive transitional cell
carcinoma of the bladder: failure analysis. J Urol. Dec
1989;142(6):1448-1453; discussion 1453-1444.

REFERENCES

10. Sell A, Jakobsen A, Nerstrom B, Sorensen BL, Steven


K, Barlebo H. Treatment of advanced bladder cancer
category T2 T3 and T4a. A randomized multicenter study
of preoperative irradiation and cystectomy versus radical
irradiation and early salvage cystectomy for residual tumor.
DAVECA protocol 8201. Danish Vesical Cancer Group.
Scand J Urol Nephrol Suppl. 1991;138:193-201.

1. Stein JP, Lieskovsky G, Cote R, et al. Radical cystectomy in


the treatment of invasive bladder cancer: long-term results
in 1,054 patients. J Clin Oncol. Feb 1 2001;19(3):666-675.
2. Dalbagni G, Genega E, Hashibe M, et al. Cystectomy
for bladder cancer: a contemporary series. J Urol. Apr
2001;165(4):1111-1116

11. Fossa SD, Waehre H, Aass N, Jacobsen AB, Olsen DR,


Ous S. Bladder cancer definitive radiation therapy of
muscle-invasive bladder cancer. A retrospective analysis of
317 patients. Cancer. Nov 15 1993;72(10):3036-3043.

3. Grossman HB, Natale RB, Tangen CM, et al. Neoadjuvant


chemotherapy plus cystectomy compared with cystectomy
alone for locally advanced bladder cancer. N Engl J Med.
Aug 28 2003;349(9):859-866

12. Horwich A, Pendleburg S, Dearnaley D. Organ Conservation


in bladder cancer. Eur J Cancer. 1995;31(Suppl. 6): 208.

4. Madersbacher S, Hochreiter W, Burkhard F, et al. Radical


cystectomy for bladder cancer today--a homogeneous
series without neoadjuvant therapy. J Clin Oncol. Feb 15
2003;21(4):690-696.

13. Slack NH, Bross ID, Prout GR Jr. 5-year follow up results
of a collaborative study of therapies for carcinoma of the
bladder. J. Surg. Oncol. 9: 393-405, 1977.

5. Hautmann RE, Gschwend JE, de Petriconi RC, Kron M,


Volkmer BG. Cystectomy for transitional cell carcinoma
of the bladder: results of a surgery only series in the
neobladder era. J Urol. Aug 2006;176(2):486-492;
discussion 491-482.

14. Bloom HJ, Hendry, WF, Wallace, BM, et. al. Treatment of T3
Bladder Cancer: controlled trial of preoperative radiotherapy
and radical cystectomy vs. radical radiotherapy. Br. J. Urol.
54: 136-151, 1982
15. Horwich A, Dearnaley D, Huddart R, et al. A randomised
trial of accelerated radiotherapy for localised invasive
bladder cancer. Radiother Oncol. Apr 2005;75(1):34-43.

6. Miller LS. Bladder cancer: superiority of preoperative


irradiation and cystectomy in clinical stages B2 and C.
Cancer. Feb 1977;39(2 Suppl):973-980.

16. Shipley WU, Prout GR, Jr., Kaufman SD, Perrone TL.
Invasive bladder carcinoma. The importance of initial
transurethral surgery and other significant prognostic
factors for improved survival with full-dose irradiation.
Cancer. Aug 1 1987;60(3 Suppl):514-520

7. Duncan W, Quilty PM. The results of a series of 963


patients with transitional cell carcinoma of the urinary
bladder primarily treated by radical megavoltage X-ray
therapy. Radiother Oncol. Dec 1986;7(4):299-310
8. Jenkins BJ, Caulfield MJ, Fowler CG, et al. Reappraisal of
the role of radical radiotherapy and salvage cystectomy in
the treatment of invasive (T2/T3) bladder cancer. Br J Urol.
Oct 1988;62(4):343-346.

17. Dunst J, Sauer R, Schrott KM, Kuhn R, Wittekind C,


Altendorf-Hofmann A. Organ-sparing treatment of advanced
bladder cancer: a 10-year experience. Int J Radiat Oncol
Biol Phys. Sep 30 1994;30(2):261-266.

324

18. Shipley WU, Winter KA, Kaufman DS, et al. Phase III trial
of neoadjuvant chemotherapy in patients with invasive
bladder cancer treated with selective bladder preservation
by combined radiation therapy and chemotherapy results
of Radiation Therapy Oncology Group 89-03. J Clin Oncol.
Nov 1998;16(11):3576-3583.

therapy for invasive bladder cancer: the long-term MGH


experience. ASTRO Abstract #21, Nov. 2009.
34. International collaboration of trialists. Lancet. 354:533-540,
1999.
35. Hall R. Updated results of a randomised controlled trial of
neoadjuvant cisplatin (C), methotrexate (M) and vinblastine
(V) chemotherapy for muscle-invasive bladder cancer. Proc
Am Soc Clin Oncol. 2002;21:178A abstract 710.

19. Gosparodarowicz M, Blandy J, eds. Radiation therapy alone


for organ conservation for invasive bladder carcinoma.
. 3rd ed. Philadelphia, PA: Lippincott Williams & Wilkins;
2000. Vogelzang N, Scardino P, Shipley W, al. e, eds.
Comprehensive textbook of genitourinary oncology.

36. Sengelov L, von der Maase H, Lundbeck F, et al.


Neoadjuvant chemotherapy with cisplatin and methotrexate
in patients with muscle-invasive bladder tumours. Acta
Oncol. 2002;41(5):447-456.

20. Nieuwenhuijzen JA, Pos F, Moonen LM, Hart AA, Horenblas


S. Survival after bladder-preservation with brachytherapy
versus radical cystectomy; a single institution experience.
Eur Urol. Aug 2005;48(2):239-245

37. Advance
Bladder
Cancer
(ABC)
Meta-analysis
Collaboration. Neoadjuvant chemotherapy in invasive
bladder cancer; a systematic review and meta-analysis.
Eur. Urol.48: 202-205,2005.

21. Pos F, Horenblas S, Dom P, Moonen L, Bartelink H. Organ


preservation in invasive bladder cancer: brachytherapy, an
alternative to cystectomy and combined modality treatment?
Int J Radiat Oncol Biol Phys. Mar 1 2005;61(3):678-686.

38. Wijkstrom H, Norming U, Lagerkvist M, Nilsson B,


Naslund I, Wiklund P. Evaluation of clinical staging before
cystectomy in transitional cell bladder carcinoma: a longterm follow-up of 276 consecutive patients. Br J Urol. May
1998;81(5):686-691.

22. Shipley WU, Prout GR, Jr., Einstein AB, et al. Treatment of
invasive bladder cancer by cisplatin and radiation in patients
unsuited for surgery. JAMA. Aug 21 1987;258(7):931-935.

39. Efstathiou JA, Bae K, Shipley WU, et al. Late pelvic toxicity
after bladder-sparing therapy in patients with invasive
bladder cancer: RTOG 89-03, 95-06, 97-06, 99-06. J Clin
Oncol. 27:4055-4061, 2009.

23. Coppin CM, Gospodarowicz MK, James K, et al. Improved


local control of invasive bladder cancer by concurrent
cisplatin and preoperative or definitive radiation. The
National Cancer Institute of Canada Clinical Trials Group. J
Clin Oncol.14:2901-2907, 1996.

40. Zietman AZ, Sacco D, Skowronski U, et al. Organ


conservation in invasive bladder cancer by transurethral
resection, chemotherapy and radiation: results of a
urodynamic and quality of life study on long-term survivors.
J Urol. 170: 1772-1778,2003.

24. Hall RR, Newling DW, Ramsden PD et. Al. Treatment of


invasive bladder cancer by local resection and methotrexate.
Br. J. Urol. 56: 668-672, 1984.
25. Herr HW, Bajorin DF, Scher HI. Neoadjuvant and bladdersparing surgery for invasive bladder cancer: 10-year
outcome. J. Clin. Oncol. 16: 1298-1301,1998.

41. Caffo O, Fellin G, Graffer U, Luciani L. Assessment of


quality of life after cystectomy or conservative therapy
for patients with infiltrating bladder carcinoma. A survey
by a self-administered questionnaire. Cancer. Sep 1
1996;78(5):1089-1097.

26. Sternberg CN, Pansodoro V, Calandro F et al. Patient


selection for bladder preserving based on response to
chemotherapy? Cancer 97: 1644-1652, 2003

42. Henningsohn L, Wijkstrom H, Dickman PW, Bergmark


K, Steineck G. Distressful symptoms after radical
radiotherapy for urinary bladder cancer. Radiother Oncol.
Feb 2002;62(2):215-225.

27. Prout GR, Jr., Shipley WU, Kaufman DS, et al. Preliminary
results in invasive bladder cancer with transurethral
resection, neoadjuvant chemotherapy and combined pelvic
irradiation plus cisplatin chemotherapy. J Urol. 144:11281134; discussion 1134-1126, 1990.

43. Habuchi T, Marberger M, Droller MJ, et al. Prognostic


markers for bladder cancer: international consensus panel
on bladder tumor markers. Urology 2005; 66 (suppl 6a):
6474

28. James N Huddurt RA, Hussain, S et al. BC2001 a Phase


III randomized trial of synchronous chemo-radiotherapy
compared to radiation therapy alone in muscle invasive
bladder cancer . Plenary presentation : ASTRO Annual
Meeting, November 1, 2011.

44. Korkolopoulou P, Christodoulou P, Konstantinidou AE, et


al. Cell cycle regulators in bladder cancer: a multivariate
survival study with emphasis on p27Kip1. Human Pathol
2000; 31: 751 60.

29. Tester W, Caplan R, Heaney J, et al. Neoadjuvant combined


modality program with selective organ preservation for
invasive bladder cancer: results of Radiation Therapy
Oncology Group phase II trial 8802. J Clin Oncol. 14119126., 1996.

45. Kamai T, Takagi K, Asami H, et al. Decreasing of p27(Kip1)


and cyclin E protein levels is associated with progression
from superficial into invasive bladder cancer. Br J Cancer
2001; 84: 124251.

30. Housset M, Dufour B, Maulard C. Concomitant 5-fluorouracil


-cisplatin and bifractionated split course radiation therapy
for invasive bladder cancer. Proc Am Soc Clin Oncol
1997;16:319a.

46. Liukkonen T, Rajala P, Raitanen M, et al. and Finnbladder


Group. Prognostic value of MIB-1 score, p53, EGFr, mitotic
index and papillary status in primary superficial (stage pTa/
T1) bladder cancer; a prospective study. Eur Urol 1999;
136: 393400.

31. Rodel C, Grabenbauer GG, Kuhn R, et al. Combinedmodality treatment and selective organ preservation
in invasive bladder cancer: long-term results. J Clin
Oncol.20:3061-3071, 2002.

47. Pollack A, Wu CS, Czerniak B, Zagars GK, Benedict WF,


McDonnell TJ. Abnormal bcl-2 and pRb expression are
independent correlates of radiation response in muscleinvasive bladder cancer. Clin Cancer Res 1997; 3: 1823
29.

32. Kaufman DS, Winter KA, Shipley WU, et al. Phase I-II
RTOG study (99-06) of patients with muscle-invasive
bladder cancer undergoing transurethral surgery, paclitaxel,
cisplatin, and twice-daily radiotherapy followed by selective
bladder preservation or radical cystectomy and adjuvant
chemotherapy. Urology. 73:833-837, 2009.

48. Chyle V, Pollack A, Czerniak B, et al. Apoptosis and


downstaging after preoperative radiotherapy for muscleinvasive bladder cancer. Int J Radiat Oncol Biol Phys
1996; 35: 28187.

33. Efstathiou J, Coen J, Spiegel D, et al. 15-year outcomes


of selective bladder preservation by combined-modality

49. Rodel C, Grabenbauer GG, Rodel F, et al. Apoptosis,

325

urologists and pathologists will see and need to be


familiar with: nested variant urothelial cancer, micro
papillary urothelial cancer, carcinogenesis (sarcoma
urothelial cancer), and urothelial cancer with small
cell, squamous, and carcinomas components. While
not all aspects of management of each variant will be
discussed, we will highlight recent information which
in most has significantly affected management.

p53, bcl-2 and Ki-67 in invasive bladder carcinoma:


possible predictors for response to radiochemotherapy and
successful bladder preservation. Int J Radiat Oncol Biol
Phys 2000; 46: 1213
50. Chakravarti A, Winter K, Wu CL, et al. Expression of the
epidermal growth factor receptor and Her-2 are predictors
of favorable outcome and reduced complete response
rates, respectively, in patients with muscle-invading bladder
cancers treated by concurrent radiation and cisplatinbased chemotherapy: a report from the Radiation Therapy
Oncology Group. Int J Radiat Oncol Biol Phys 2005; 62:
30917.

1. Nested variant urothelial cancer


It was originally identified as an aggressive entity
by Talbert and Young[5] who characterized the
tumor as one with a deceptively benign histologic
appearance. Small nests and tubules of urothelial
cells infiltrate through the lamina propria and usually
the muscularis. Its appearance can be mistaken for
a proliferative variant of von Brunns nests, a finding
present in many individuals. This entity can appear
as an isolated tumor or along with more classic
urothelial canceralmost invariably high grade,
in other portions of the bladder or adjacent to the
nested variant portion. It has been estimated to
represent 0.3% of all urothelial cancers.[6]. Little is
known about its response to non-surgical treatments.
This malignant variant has even a higher male
predominance than standard urothelial cancer (3:1
males to females in the United StatesSEER[7]).
Perhaps its most important clinical feature is mistaking
it for either a benign condition or a bland non-muscleinvasive urothelial cancer. Because these cancers
cannot be diagnosed without at least lamina propria
invasion, the standard treatment should be radical
cystectomy in the absence of distant metastases
and because the vast majority are more invasive
than this in the absence of distant metastases.
Whether neoadjuvant or adjuvant chemotherapy for
urothelial cancer is beneficial is unknown. Although
adjuvant radiation or chemotherapy have not been
shown to be beneficial, it is likely that the rarity of
this disease and the advanced nature of the cases
treated in such ways would leave this issue up for
further study.[8]

51. Press MF, Lenz HJ. EGFR, HER2 and VEGF pathways:
validated targets for cancer treatment. Drugs 2007; 67:
204575
52. Krger S, Witsch G, Bttner H, et al. HER2 overexpression
in muscle-invasive urothelial carcinoma of the bladder:
prognostic implications. Int J Cancer 2002; 102: 51418.
53. Jimenez RE, Hussain M, Bianco FJ, et al. Her-2/neu
overexpression in muscle-invasive urothelial carcinoma
of the bladder: prognostic significance and comparative
analysis in primary and metastatic tumors. Clin Cancer Res
2001; 7: 244047.
54. Hussain MH, MacVicar GR, Petrylak DP, et al. Trastuzumab,
paclitaxel, carboplatin, and gemcitabine in advanced human
epidermal growth factor receptor-2/neu-positive urothelial
carcinoma: results of a multicenter phase II National Cancer
Institute trial. J Clin Oncol 2007: 25: 221824.
55. Choudhury, A, Nelson; Mark TW, et al: MRE 11 expression
is predictive of cause-specific survival following radical
radiotherapy for muscle-invasive bladder cancer. Cancer
research 70: 7017-7026, 2010.

VII. Treatment of Mixed


Histology Urothelial
Carcinoma
Urothelial carcinoma of the bladder represents by
far the most common histology of bladder cancer.
However, cancers classified as urothelial contain
non- urothelial components in 5-40% of cases.[1,2]
In general, the higher the grade of the urothelial
component and more invasive the cancer, the
more likely that non-urothelial components will be
found. Whether this is because such cancers are
less differentiated and more easily develop aberrant
histologic features or because the mixed urothelial
cancers are generally very aggressive and are
often not diagnosed until they are far advanced is
unknown.

2. Ur
 othelial carcinoma with
micropapillary differentiation
It is another variant often considered to have bland
histology, was first described as part of the urothelial
carcinoma complex by Amin et al in 1994 [9]. It is
now considered to represent between 1 and 2%
of all urothelial cancers and has a strong male
predominance.[1,10,11] The histologic appearance
resembles micropapillary tumors arising in the ovary,
lung, breast, and other anatomic locations.[12] The
immunohistologic phenotype of these tumors has
been elegantly outlined.[12] All 13 of the cases
tested in that report[12] were positive for MUC1,
MUC2, CK7, PTEN, p53, and ki67. CK20 and
3BVE12 were negative in half the cases and Her/neu
was negative in 9 of the 13. In general, except for
p53 staining, at least 50% of cells were positive for

In general, however, molecular and genetic markers


of urothelial cancer are present in non-urothelial
as well as urothelial components, implying a clonal
origin.[3,4] Most of these cancers are fairly rare
and most information about them comes from case
series. However, for the two most common mixed
histologies, urothelial carcinoma with squamous
and/or carcinomas elements, some data from
randomized prospective studies existalthough
collected retrospectively. These will be reviewed. This
section is not meant to be encyclopedic, but instead
will focus on 6 urothelial carcinoma variants that all

326

the 6 markers present in every specimen. All of the


patients in this series and several of the others[10]
had at least muscle invading cancer and most had
disease beyond the bladder. In a recent review of
cases from MD Anderson Cancer Center, Kamat et
al[11] found that these cancers may be detectable
prior to muscle invasion and that cystectomy still is
recommended. The response to BCG was extremely
poor, with 18 of 27 patients experiencing progression
to stage T2 or greater at a median 8 months followup, one-third of whom had progressed to metastatic
disease. At 30 months median follow-up, only 5 of 29
patients were free of disease with an intact bladder.
[11] Neoadjuvant chemotherapy did not appear
superior to immediate cystectomy. The authors
recommended against such an approach unless
there was surgically unresectable (cN1 or cT4b)
disease. Following immediate cystectomy, 71%
of 32 cT4a or lower stage patients were alive at 5
years. In contrast, only 32% of 23 similar patients
treated with neoadjuvant chemotherapy followed by
cystectomy were alive at 5 years.

role of chemotherapy or radiation remains at best


anecdotal for this extremely aggressive malignancy;
even with disease confined to the bladder, 5- and 10year survival rates in the 25-35% range have been
reported.[18] However, since nearly 75% of patients
have more advanced disease at diagnosis,[18]
surgery alone is often not curative. Anecdotally good
responses of advanced cancers with combined
surgical treatment and chemotherapy have been
reported,19 but they remain the rarity. It should
be remembered that over 40% of these cancers
have multiple sites within the bladder[18] so that
careful bladder mapping in the rare circumstances
where partial cystectomy would seem feasible is
mandatory.
While the role of neoadjuvant chemotherapy is
unclear, Black et al13 reported downstaging to pT0
in 5 of 11 (45%) patients with clinical stage T2-T3
tumors using urothelial cancer-like regimens.

4. Small cell carcinoma


It can exist in a pure form, but, in the urinary tract over
50% have an epithelial component. Occasionally,
the epithelial component can be of non-urothelial
type such as squamous cell or adenocarcinoma, but
the vast majority of mixed tumors are with urothelial
cancers.[1]

3. S
 arcomatoid carcinoma and
carcinosarcoma of the urinary
bladder
They consist of high grade cancers with malignant
epithelial and mesenchymal elements. While pathologists debate if there is a distinction between the
two, classically sarcomatoid carcinomas consist of
urothelial cancer with a spindle cell (mesenchymal)
component which retains expression of epithelial
molecular markers, while carcinosarcoma contains
recognizably more differentiated sarcomatous elements such as osteosarcoma.[13,14,15] As opposed
to nested variant or micropapillary urothelial cancers, the aggressive nature of this tumor is not overlooked because of bland histology. Extremely rarely,
this condition can be confused with other benign
conditions such as pseudosarcomatous, myofibroblastic proliferationsmost commonly postoperative
spindle cell nodules and pseudotumors.[1] The presence of outright urothelial carcinoma, however, usually makes the distinction quite obvious. Molecular
analyses including those for p53 indicate a common
clonal origin for both carcinomatous and sarcomatoid components.[16,17]

As opposed to other urothelial cancers with mixed


histologies which are generally treated as urothelial
cancers, it has been recognized by the World Health
Organization that any amount of small cell histology
should be regarded as a small cell carcinoma.[20]
This is presumably because this small cell component
determines the patients prognosis.
As with small cell carcinomas elsewhere, the majority
are positive for chromogranin A and synaptophysin,
and over 90% present with muscle invasion.[1] As
with carcinosarcoma, molecular analyses have
indicated similar genetic changes in the small cell and
coexistent urothelial carcinomatous components.
[1,21]
In general, mixed histology small cell carcinoma is
recognized as a systemic disease[1,3] and probably
should receive systemic chemotherapy aimed at
small cell carcinoma (e.g., cisplatin and etoposide
based therapy) along with extensive local treatment.
[1] In an MD Anderson series, the pT0 rate in patients
with cT2-4a mixed urothelial and small cell cancers
treated with neoadjuvant neuroendocrine regimens
was 10/12 (83%), compared with 3/9 (33%) with
a urothelial regimen.[13] Moreover, the diseasespecific survival at 5 years for these 21 patients
treated with neoadjuvant chemotherapy was 78%,
compared to 36% in 25 patients who underwent
immediate cystectomy.[22]

A review of data from the SEER cancer registry[18]


reported that between 1973 and 2004, this entity
represented 0.6% of all primary bladder cancers.
Nearly 75% of cases had regional or distant disease
at the time of diagnosis. Two-thirds of the patients
were male, actually a slightly lower percentage
than pure urothelial cancers and far lower than
micropapillary or nested variant cancers. In general
the treatment recommended is cystectomy even for
non-muscle-invasive carcinosarcomas, almost all
of which invade at least the lamina propria.[18] The

327

Bladder sparing approaches have also been reported.


Bex et al have published results for chemoradiation
that approach those for chemotherapy and
cystectomy,[23] although primarily only patients with
limited disease were reported.

Nested variant and micropapillary urothelial cancers


have deceivingly bland histologies and at times may
be confused for benign lesions. Even when detected
in non-muscle-invasive stages, these should be
treated with aggressive local extirpative therapy.
Endoscopic and intravesical management should be
discouraged.

5. The most common of the


mixed urothelial cancers are
those with squamous and/or
adenocarcinomatous elements

Carcinosarcomas and urothelial cancers with small


cell elements usually are diagnosed correctly, but
again are extremely aggressive malignancies.
Carcinosarcomas should be treated with local extirpative surgery. The role of additional systemic
chemotherapy is being explored and has received
some favorable reports.[19] Urothelial cancer with
small cell components should be treated as small
cell carcinoma with neoadjuvant cisplatin/etoposide
followed by aggressive local treatmentbe it extirpative surgery or chemoradiation. Muscle-invasive
urothelial cancers with squamous or adenocarcinomatous elements have responded particularly
well to multi-drug cisplatin-based urothelial cancer
regimens, especially M-VAC, and it would appear
that neoadjuvant chemotherapy should be used for
them before definitive surgery. Awareness of these
features of these common mixed urothelial cancers
by urologists and pathologists should assist in their
management.

They account for as many as 30% of patients


undergoing cystectomy for what has been termed
muscle-invasive urothelial cancer. In general,
these patients fare less well than those with pure
urothelial carcinoma even when stratified by stage.
[2] While no randomized prospective studies have
specifically addressed these cancers or any of the
other mixed variant urothelial cancers, because of
their high proportion, they have been included in
randomized prospective phase III trials of treatments
for invasive urothelial cancer and, at least in one
of the mature long-term studies of neoadjuvant
chemotherapy with cystectomy, results have been
analyzed.[24] In SWOG 8710, 59 (19.2%) of 307
patients with cT2-4a cN0 cM0 urothelial cancer had
squamous (N=37), adenocarcinomatous (N=20),
or squamous and adenocarcinomatous (N=2)
elements. Patients were randomized to receive 3
cycles of M-VAC chemotherapy before cystectomy
or undergo immediate cystectomy. There was
evidence of a survival benefit for patients with mixed
histology cancers who received M-VAC compared
to those who went directly to cystectomy (HR =
0.46, 96% CI 0.25-0.87) (p=0.02). This was so for
both clinical T2 and cT3-4a tumors (5-year survival
0.73 [M-VAC + cystectomy] vs. 0.54 [cystectomy
alone] for cT2; or 0.58% [M-VAC + cystectomy]
vs. 0.34 [cystectomy only] for cT3, 4a) and was far
better than the benefit of M-VAC for pure urothelial
cancers (HR=0.90, p=0.48). Not surprisingly M-VAC
also improved downstaging to pT0 compared to
cystectomy for mixed histology tumors (34% vs.
4%, p=0.004). The strengths of this study were its
randomized prospective design and the presence
of central histology review. The weaknesses were
the relatively small number of mixed histology cases
(N=59) and lack of quantification of the non-urothelial
cancer component. Despite these shortcomings,
until more information becomes available, patients
with cT2-T4a urothelial cancers with squamous or
adenocarcinoma elements should receive cisplatinbased combination chemotherapy before cystectomy
to optimize their chances of survival.

Recommendations
Nested variant and micropapillary urothelial
cancers, regardless of whether they are detected
in non-muscle-invasive or muscle-invasive
stages, should be treated with aggressive local
extirpative therapy (LE: 3, GR: B)
Carcinosarcomas should be treated if possible
with local extirpative surgery (LE: 3, GR: B)
Urothelial cancer with small cell components
should be treated with neoadjuvant chemotherapy
including neoadjuvant cisplatin and etoposide
followed by aggressive local treatment (LE: 3,
GR: B)
Muscle-invasive
urothelial
cancers
with
squamous or adenocarcinomatous elements
should be treated with neoadjuvant multi-drug
cisplatin-based urothelial cancer regimens before
definitive surgery (LE: 3, GR: C).

References
1. Amin MB. Histological variants of urothelial carcinoma:
diagnostic, therapeutic and prognostic implications.
Modern Pathology 22:S96-118, 2009.

In summary

2. Wasco MJ, Daignault S, Zhang Y, et al. Urothelial carcinoma


with divergent histologic differentiation (mixed histologic
features) predicts the presence of locally advanced bladder
cancer when detected at transurethral resection. Urology
70:69-74, 2007.

Mixed urothelial cancers with variant histologies


are generally found with only high grade urothelial
cancer and behave at least as aggressively as high
grade urothelial cancer and often even more so.

328

3. Chalasani V, Chin JL, Izawa JI. Histologic variants of


urothelial cancer and nontheatrical histology in bladder
cancer. Can Urol Assoc J 3:5193-8, 2009.

22. Siekfer-Radtke AO, Gee J, SHen Y, et al. Multi-modality


management of urachal carcinoma: The MD Anderson
Cancer Center Experience. J Urol 169:1295-8, 2003

4. Kipp BJ, Tyner H, Campion MB, et al. Chromosomal


alterations detected by fluorescence in situ hybridization in
urothelial carcinoma and rare histologic variants of bladder
cancer. Am J Clin Path 130:552-559, 2008.

23. Bex A, DeVries R, Pos F, et al. Long-term survival after


sequential chemoradiation for limited disease small cell
carcinoma of the bladder. World J Urol 27:101-106, 2009
24. Scosyrev E, Ely BW, Messing EM, et al. Do mixed
histological features affect survival benefit from neoadjuvant platinum-based combination chemotherapy
in patients with locally advanced bladder cancer? A
secondary analysis Southwest Oncology Group-Directed
Intergroup Study (8710). BJU 2010

5. Talbert ML, Young RH. Carcinoma of the urinary bladder


with deceptively benign-appearing foci: a report of three
cases. Am J Surg Path 13:374-381, 1989.
6. Holmang S, Johansson SL.
The nested variant of
transitional cell carcinoma. A rare neoplasm with poor
prognosis. Scan J Urol 35:102-105, 2001
7. Scosyrev E, Noyes K, Feng CY, Messing EMM. Sex
and racial differences in bladder cancer presentation and
mortality in the United States. Cancer, 1;115(1):68-74,
2009.

VIII. Follow-up

8. Dhall D, Al-Ahmadie, Olgac S. Nested variant of urothelial


carcinoma.
Archives Path Lab Med. 131:1725-1727,
2007

1. Distant Recurrences
Recently the long-term oncological outcome of
radical cystectomy was analyzed in a contemporary
series of 2287 patients who underwent radical
cystectomy between 1998 and 2008 [1], LE 3. The
mean and median follow-up was 35 and 29 months,
respectively. The 5-year overall, recurrence-free, and
cancer-specific survival rates were 57%, 48%, and
67%, respectively, with distant and local recurrence
rates of 37% and 6%, respectively. In another series,
Canter et al [2] reported on a 5-year recurrencefree survival and cancer-specific survival of 56.5%
and 59.5%, respectively, among 212 patients with a
mean follow-up time of 28 months, LE 3. Again, the
majority of patients developed distant metastases.
Contemporary cystectomy series have demonstrated
a 5-15% probability of pelvic recurrence. Most
recurrences manifest during the first 24 months,
often within 6-18 months after surgery. However,
late recurrences have occurred up to 5 years after
cystectomy.

9. Amin MB, JY Ro, el-Sharkawy T, et al. Micropapillary


variant of transitional cell carcinoma of the urinary bladder.
Histologic pattern resembling ovarian papillary serous
carcinoma. Am J Surg Path 18:1224-1232, 1994
10. Heudel P, El Karak F, Ismaili N, et al. Micropapillary
bladder cancer: A review of Leon Berard Cancer Center
experience. BMC Urology 9:5, 2009.
11. Kamat AM, Dinney C, Gee JR, et al. Micropapillary Bladder
Cancer: A review of the University of Texas MD Anderson
Cancer Center Experience with 100 Consecutive Patients.
Cancer 110:62-70, 2007.
12. Lopez-Beltran A, Montironi R, Blanca A, et al. Invasive
micropapillary urothelial cancer of the bladder. Human
Pathology 41:1159-1164, 2010.
13. Black PC, Brown GA, Dinney C. The impact of variant
histology on the outcome of bladder cancer treated with
curative intent. Urologic Onc: Seminars and Original
Invest. 27:3-7, 2007
14. Baschinsky DY, Chen JH, Vadmal MS, et al. Carcinoma
of the urinary bladder an aggressive tumor with diverse
histogenesis. A clinicopathologic study of four cases and
review of the literature. Arch Pathol Lab Med 124:1172-8,
2000.

The goal of all follow-up strategies is to detect


systemic metastases as early as possible to allow
adequate systemic treatment with or without
removal of metastatic lesions with a benefit in terms
of progression-free and cancer-specific survival.
To achieve the goal of an individualized follow-up
scheme, natural timing of recurrence, probability of
recurrences, functional deteriorations at particular
sites, and possible treatment options of a recurrence
should be considered [3].

15. Lopez- Beltran A, Pacelli A, Rothenberg HJ, et al.


Carcinosarcoma and sarcomatoid carcinoma of the
bladder: Clinicopathological study of 41 cases. J Urol
159:1497-503, 1998.
16. Sung MT, Wang M, MacLennan GT, et al. Histogenesis
of sarcomatoid urothelial carcinoma of the urinary bladder:
Evidence for a common clonal origin with divergent
differentiation.. J Path. 211:420-430, 2007
17. Armstrong AB, Wang M, Eble JN, et al. TP53 mutational
analysis supports monoclonal origin of biphasic
sarcomatoid urothelial carcinoma (carcinosarcoma) of the
urinary bladder. Mod Path 22:1113-118, 2009.
18. Wang J, Wang FW, LaGrange CA, et al. Clinical features
of sarcomatoid carcinoma (carcinosarcoma) of the urinary
bladder: Analysis of 221 cases. Sarcoma. Article ID
454792, 2010.
19. Damiano R, DAmiento M, Amorosi A, et al. Gemcitabine
and cisplatin following surgical treatment of urinary bladder
carcinosarcoma. Tumori 90:458-60, 2004
20. Eble JN, Sauter G, Epstein JI, et al eds. Path and Genetics
of Tumors, IARC Press: Lyon;359:2004
21. Abbosh PH, Wang M, Eble JN, et al. Hypermethylation of
tumor-suppressor gene CpG islands in small cell carcinoma
of the urinary bladder. Mod Path 21:355-362, 2008.

Recently, a nomogram based on 728 patients who


underwent cystectomy was presented. Standard
predictors were pathological stage of the primary
tumor (pTN) and nodal status (pN), LE 2b. The
prediction of recurrent disease increased by 3.2%
when the nomogram included: age, lymphovascular
invasion, CIS, neoadjuvant chemotherapy, adjuvant
chemotherapy, and adjuvant radiotherapy [4]. This
nomogram can be used to predict the individual risk
of systemic relapse and to develop a risk-adapted
follow-up protocol.

329

a) Distant Recurrences

87 were symptomatic and 87 were asymptomatic.


The 5-year cancer-specific and overall survival rate
was significantly higher in the group of patients
with asymptomatic metastasis so that the authors
recommend routine cross-sectional imaging of
the chest and the abdomen. On the other hand,
Volkmer et al [7] identified 444 out of 1270 patients
with recurrences after radical cystectomy, of which
154 and 290 were asymptomatic and symptomatic,
respectively. The overall survival rates at 1, 2, and
5 years were not significantly different between
both groups. The authors draw the conclusion that
symptom-guided follow-up might provide similar
results to a strict follow-up protocol at lower costs.

Distant recurrences are seen in up to 50% of patients


treated with cystectomy. Most recurrences are seen
in the first 24 months, although progression has
been observed after more than 10 years [5-8]. The
most likely sites for distant recurrences are the lungs,
liver, and bones [9]. Treatment of metastatic disease
with cisplatin-based combination chemotherapy with
either M-VAC or cisplatin and gemcitabine results in
a mean survival time of around 14 months.
Helical CT represents the standard imaging technique
of choice to detect lung metastases [9, 10]. Use of
a 5-mm contiguous reconstruction allows detection
of a minimum-sized lesion of 10 mm (LE 2b). For
the detection of liver metastases, CT should be
performed with native images as well as after using
nonionic, iodine-containing water-soluble contrast
agents; multidetector helical CT is todays standard
[11, 12]. For helical CT, 5 mm reconstructions should
be used, allowing a minimum-sized lesion of 10 mm
to be detected.

Recommendations
Postoperative follow-up after radical cystectomy
should be performed in a risk-adapted approach
using currently available nomograms (LE: 3, GR:
C)
Symptom-oriented follow-up might result in the
same long-term outcome as standardized followup protocols but at a lower cost (LE: 3; GR: C)

The majority of lymph node metastases originating


from urogenital cancer are located in the
retroperitoneum so CT scans of the abdomen and
the pelvis are the imaging procedures of choice
[13, 14]. On these cross-sectional modalities,
nodal metastases are usually suspected according
to location and size criteria, i.e. a maximum short
axis diameter 1cm is considered malignant [13].
However, CT scans of the abdomen and pelvis might
give false-negative results in up to 30% of cases due
to difficulties in the interpretation of lymph nodes
based on morphology and size alone [14]. Magnetic
resonance tomography (MRT) scans of the abdomen
and pelvis do not provide additional information and
should be restricted to patients with contraindications
to CT [10] (LE 2b). To date, PET has not been shown
to improve sensitivity in patients with metastatic
urogenital cancer overstaging by doing CT scanning
alone [15, 16], (LE 3). Bone scintigraphy (BS),
conventional radiographic techniques, computed
tomography, (18) FFDG PET/CT and whole body
MRI represent potential imaging studies to diagnose
and to monitor skeletal metastases [17-20].

Helical CT represents the imaging modality of


choice to identify lung, lymph node, and liver
metastasis (LE 2b; GR: B)

References
1. Yafi FA, Aprikian AG, Chin JL, Fradet Y, Izawa J, Estey
E, Fairey A, Rendon R, Cagiannos I, Lacombe L, Lattouf
JB, Bell D, Drachenberg D, Kassouf W. Contemporary
outcomes of 2287 patients with bladder cancer who were
treated with radical cystectomy: a Canadian multicentre
experience. BJU Int. 2010 Dec 16. doi: 10.1111/j.1464410X.2010.09912.x. [Epub ahead of print]
2. Canter D, Long C, Kutikov A, Plimack E, Saad I, Oblaczynski
M, Zhu F, Viterbo R, Chen DY, Uzzo RG, Greenberg RE,
Boorjian SA. Clinicopathological outcomes after radical
cystectomy for clinical T2 urothelial carcinoma: further
evidence to support the use of neoadjuvant chemotherapy.
BJU Int. 2011 Jan;107(1):58-62. doi: 10.1111/j.1464410X.2010.09442.x.
3. Malkowicz SB, van Poppel H, Mickisch G, Pansadoro
V, Throff J, Soloway MS, Chang S, Benson M, Fukui
I. Muscle-invasive urothelial carcinoma of the bladder.
Urology 2007 Jan;69(1 Suppl):3-16

However, based on 2 recent retrospective clinical


studies it appears questionable if early detection
of recurrences is associated with a survival benefit
when compared to recurrences that are detected
due to symptoms [6, 7], (LE 2b). In the first study,
479 patients who underwent radical cystectomy with
adjuvant therapy and who were followed according
to a standardized follow-up protocol were evaluated
to determine whether diagnosis of asymptomatic
recurrence after radical cystectomy by routine followup investigations conferred a survival benefit versus
symptomatic recurrence. After a median follow-up of
4.3 years, 174 recurrences were detected, of which

4. Karakiewicz PI, Shariat SF, Palapattu GS, Gilad AE, Lotan


Y, Rogers CG, Vazina A, Gupta A, Bastian PJ, Perrotte
P, Sagalowsky AI, Schoenberg M, Lerner SP. Nomogram
for predicting disease recurrence after radical cystectomy
for transitional cell carcinoma of the bladder. J Urol 2006
Oct;176(4 Pt 1):1354-61; discussion 1361-2.
5. Mathers MJ, Zumbe J, Wyler S, Roth S, Gerken M, Hofstdter
F, Klotz T. Is there evidence for a multidisciplinary followup after urological cancer? An evaluation of subsequent
cancers. World J Urol 2008 Jun;26(3):251-6.
6. Giannarini G, Kessler TM, Thoeny HC, Nguyen DP,
Meissner C, Studer UE. Do patients benefit from routine

330

2. Follow-up and Treatment of


Secondary Ureteral and Urethral
Tumors

follow-up to detect recurrences after radical cystectomy


and ileal orthotopic bladder substitution? Eur Urol. 2010
Oct;58(4):486-94.
7. Volkmer BG, Kuefer R, Bartsch GC Jr, Gust K, Hautmann
RE. Oncological followup after radical cystectomy
for bladder cancer-is there any benefit? J Urol. 2009
Apr;181(4):1587-93; discussion 1593.

a) Acquisition of Evidence
The recommendations given on follow-up and
treatment of secondary ureteral and urethral tumors
after radical cystectomy are based on an extensive
Medline search.
Regarding surgical treatment
of invasive bladder cancer, there is only limited
prospective, randomized data. Evidence obtained
from publications on this topic is mainly based
on retrospective multicenter series. Articles were
selected according to the following criteria: quality of
study evaluation, topic relevance, and presence of
intermediate and long-term outcome. Older studies
were only included if data were scarce in recent
publications. Generally, follow-up regimens for
ureteral and urethral recurrences should be adapted
according to the natural timing and probability of
recurrence, possible functional deterioration at
recurrence-specific sites, as well as the availability
of an efficient treatment option. The level of evidence
and grade of recommendation given are based as
outlined by the Oxford Centre for Evidence-based
Medicine [1].

8. Bochner BH, Montie JE, Lee CT. Follow-up strategies and


management of recurrence in urologic oncology bladder
cancer: invasive bladder cancer. Urol Clin North Am 2003
Nov;30(4):777-89
9. Heidenreich A, Albers P, Claen J, Graefen M, Gschwend J,
Kotzerke J, Krege S, Lehmann J, Rohde D, Schmidberger
K, Uder M, Zeeb H. Imaging studies in metastatic
urogenital cancer patients undergoing systemic therapy:
recommendations of a multidisciplinary consensus meeting
of the Association of Urological Oncology (AUO) of the
German Cancer Society. Urol Int 2010; 85: 1 - 10
10. Winer-Muram HT. The solitary pulmonary nodule. Radiology
2006; 239: 34 49 Wormanns D, Ludwig K, Beyer F,
Heindel W, Diederich S. Detection of pulmonary nodules
at multirow-detector CT: effectiveness of double reading to
improve sensitivity at standard-dose and low-dose chest
CT. Eur Radiol. 2005;15:14-22
11. Kanematsu M, Kondo H, Goshima S, Kato H, Tsuge U,
Hirose Y, Kim MJ, Moriyama N. Imaging liver metastases:
review and update. Eur J Radiol. 2006; 58:217-28
12. Khan SA. Imaging of liver cancer. World J Gastroenterol
2009; 15: 1289 1300
13. Barrett T, Choyke PL, Kobayashi H. Imaging of the lymphatic
system: new horizons. Contrast Med Mol Imaging 2006; 1:
230 - 245

b) Secondary Ureteral Tumors

14. Morisawa N, Koyama T, Togashi K. Metastatic lymph nodes


in urogenital cancers: contribution of imaging findings.
Abdom Imaging 2006; 31: 620 629

1. Incidence and Time to Ureteral Recurrence

15. Powles T, Murray I, Brock C, Oliver T, Avril N. Molecular


positron emission tomography and PET/CT imaging in
urological malignancies. Eur Urol 2007; 51: 1511 - 1521

According to recent series, secondary urothelial


tumors account for approximately up to 20% of all
recurrences after radical cystectomy for invasive
bladder cancer [2]. As opposed to local and distant
recurrences, metachronous upper urinary tract
recurrences after invasive bladder cancer are
considered late oncological events. Most ureteral
recurrences have been reported to occur after a
median time of 24-41 months after radical cystectomy
[3-7]. As a result of improved survival rates in
invasive bladder cancer, an increasing number of
patients will achieve long-term survival and will
therefore remain at risk for tumor recurrence in the
upper urinary tract. In numerous studies, the 5-year
rate of ureteral recurrences after radical cystectomy
is reported to range from 2-9% [2, 4, 6, 8-15]. In a
longitudinal study using landmark time analysis of a
total of 1329 patients treated with radical cystectomy,
it was shown that the risk of upper urinary tract
recurrence does not decrease over time with 3- and
5-year cumulative incidences of 4% (95% CI 3-6%)
and 7% (95% CI 5-8%) [15]. Furthermore, ureteral
recurrences have been reported to occur even 9
years after curative surgery, [6] which highlights the
importance of long-term surveillance of the upper
tracts after radical cystectomy.

16. Kibel AS, Dehdashti F, Katz MD, Klim AP, Grubb RL,
Humphrey PA, Siegel C, Cao D, Gao F, Siegel BA.
Prospective study of [18F]fluorodeoxyglucose positron
emission tomography/computed tomography for staging
of muscle-invasive bladder carcinoma. J Clin Oncol. 2009
Sep 10;27(26):4314-20. Epub 2009 Aug 3.
17. Pollen JJ, Gerber K, Ashburn WL, Schmidt JD. The value of
nuclear bone imaging in advanced prostate cancer. J Urol
1981; 125: 222 - 233
18. Ghanem N, Uhl M, Brink I, Schfer O, Kelly T, Moser E,
Langer M. Diagnostic value of MRI in comparison to
scintigraphy, PET, MS-CT and PET/CT for the detection of
metastases of bone. Eur J Radiol. 2005; 55: 41-55
19. Nakanishi K, Kobayashi M, Nakaguchi K, Kyakuno M,
Hashimoto N, Onishi H, Maeda N, Nakata S, Kuwabara M,
Murakami T, Nakamura H. Whole-body MRI for detecting
metastatic bone tumor: diagnostic value of diffusionweighted images. Magn Reson Med Sci. 2007; 6:147-55
20. Schlemmer HP, Schfer J, Pfannenberg C, Radny P,
Korchidi S, Mller-Horvat C, Ngele T, Tomaschko K,
Fenchel M, Claussen CD. Fast whole-body assessment
of metastatic disease using a novel magnetic resonance
imaging system: initial experiences. Invest Radiol. 2005;
40: 64-71

331

2. Rationale of a Risk-adapted Strategy for


Follow-up

10, 11, 17]. In two recent studies, the sensitivity and


specificity of frozen section analysis for the detection
of malignant ureteral margins was reported to
range between 74-75% and 98-99%, respectively,
with a positive predictive value of 94%, resulting in
an overall accuracy of 98% [4, 10]. On the other
hand, heterogeneous data exist on whether a
sequential resection of malignant ureteral margins
can unequivocally be advocated to reduce the risk
of a malignant anastomotic margin at cystectomy [4,
5, 10]. In two recent studies, the rates of converting
initially positive ureteral margins into negatives with
a sequential resection strategy were as high as 3941%, [4, 10] whereas in another study a conversion
rate of up to 82% was reported [5]. In this study, those
with initially positive but finally negative margins still
had a 4.4-fold increased risk as compared to those
patients with initially negative margins [5). However,
their risk decreased after conversion to a finally
negative ureteral margin since those patients with
positive anastomotic margins had an even higher
7.4-fold risk of recurrence [5].

Several clinical and pathological parameters have


been identified as risk factors for metachronous
upper urinary tract recurrence and should be taken
into consideration when addressing the patients
individual risk of recurrence. The rationale of a
risk-adapted strategy is supported by a recently
performed large retrospective analysis evaluating
the long-term risk of upper tract recurrence in
1420 patients who had undergone cystectomy[16].
In this study, a number of risk factors were shown
to potentiate the risk of upper tract recurrence.
Multivariate analysis revealed that patients with no
risk factors (history of carcinoma in situ (CIS) and
recurrent bladder cancer, high-grade pTa-T1 bladder
cancer, and distal ureteral malignancy at radical
cystectomy) had only a 0.8% risk of developing
ureteral recurrence 15 years after surgery whereas it
increased to 13.5% in patients with 3-4 existing risk
factors [16]. Furthermore, a recent large retrospective
single center series on 174 patients with recurrence
after radical cystectomy showed that the rate of
concomitant distant recurrences in patients with
secondary urothelial carcinoma was only 11%[2].

Since the incidence of malignant ureters at


cystectomy is highest in the distal ureters [18], some
authors have suggested resecting the ureters more
proximally at the crossing with the iliac vessels [11]. A
retrospective series among 755 patients undergoing
cystectomy reported CIS in the most proximally
resected ureters in only 1.2% of the patients [11]. In
this study, a considerable number of 17% of patients
with CIS on frozen section analysis had upper tract
recurrence and 80% of the patients with CIS at the
ureteral margin had also CIS of the bladder. The
authors suggested performing frozen section analysis
only in patients with known CIS of the bladder. This
may minimize the risk of ureteral strictures due to
ischemia of the distal part of the ureter and also be
beneficial for patients after prior pelvic irradiation
therapy, but makes the use of an afferent tubular
ileal segment for ureteral reconstruction necessary
[11].

Approximately 75% of upper tract recurrences are


detected when patients present with symptoms, such
as gross hematuria or flank pain. These symptoms
are often associated with locally advanced disease
and, consequently, are associated with poor
outcomes after radical nephroureterectomy [6].
Thus, new strategies for earlier detection of upper
tract recurrences while they are still localized
are necessary for effective treatment by radical
nephroureterectomy [6] given the opportunity of local
curative treatment [6]. Assessment of risk factors for
upper tract recurrence might help to identify high-risk
patients and tailor surveillance regimes according
to a risk-adapted strategy, thereby reducing the
need of unnecessary follow-up examinations and
surveillance costs in low-risk patients.

3. Risk Factors for Upper Urinary Tract


Recurrence

Carcinoma in situ of the Bladder


In patients with non-muscle-invasive bladder cancer
(pTa, pTis, pT1), the presence of CIS of the bladder
was found to be a risk factor for metachronous
upper tract carcinoma [10]. Conversely, in patients
with muscle-invasive disease, concomitant CIS
was not found to be independently associated with
upper tract recurrence [4, 6, 8, 14, 19], whereas
in patients with CIS-only disease at cystectomy, a
significantly higher rate of upper tract recurrences
was reported [20]. This finding may be artificially
skewed towards the patient cohort with low-stage
disease at cystectomy since a confounding effect of
survival differences between different tumor stages
has to be assumed.

U
 reteral Margin and Frozen Section Analysis
Ureteral tumor involvement at final pathological
analysis of radical cystectomy specimens was found
in 4.8-13.0% of the patients and in 3.6-8.3% of the
examined ureters [4, 10, 11, 15, 17]. It has been
shown in various studies that tumor involvement of
the distal ureter at cystectomy is an independent
risk factor for upper tract recurrence [15, 16] with
an approximately 2.6-fold increase in the relative
risk [16]. While there is substantial evidence that
suggests that intraoperative frozen section analysis
is a reliable tool to detect malignant ureteral margins
at radical cystectomy, its role is controversial [4, 5,

332

T
 umor Stage and Multifocality

patients with anastomotic tumor recurrence tended


to develop early progression to distant metastatic
disease, whereas patients with more proximal upper
tract recurrences showed improved recurrence-free
survival after radical cystectomy [25].

In several studies, tumor stage was investigated


as a possible risk factor of upper tract recurrence.
Approximately 65-100% of upper tract recurrences
occur in patients with organ-confined bladder cancer
(pT2bpN0M0 or lower stage disease) [8, 9, 14, 19,
21]. In one study, patients with pTa-T1 bladder cancer
were reported to have a 1.8-3.8 fold higher risk of
upper tract recurrence as compared to patients with
muscle-invasive disease [3]. Similar to the outcome of
patients with CIS-only disease at cystectomy, tumor
stage was not an independent risk factor for upper
tract recurrence, but was a rather strong predictor
of prolonged survival after radical cystectomy. These
data may be biased by the fact that patients with
high-stage disease were at considerably higher risk
for early local or systemic recurrence as compared to
patients with non-muscle-invasive disease [2, 22].

5. Diagnosis of Upper Tract Recurrence


R
 adiological Imaging
According to current studies, patients diagnosed
with asymptomatic upper tract recurrence during
routine follow-up have a significantly higher survival
advantage than those with symptomatic recurrences
[6]. This stresses the importance of an early detection
of metachronous ureteral malignancy for a timely
initiation of a curative treatment.
In a recent analysis, the accuracy of a total of
1064 intravenous pyelography (IVP) studies of 322
patients, who underwent routine follow-up of the
upper tract at 1, 2, 3, 5, 7, and 10 years after radical
cystectomy, was investigated. Of these patients,
upper tract recurrence was detected in only 15
(4.7%), but only 8 of them had suspicious findings
on IVP. Patients with positive final ureteral margins
were at highest risk for recurrence. Therefore, the
authors concluded that routine excretory urography
should be limited to patients at high risk for upper
tract recurrence. Likewise, Slaton et al developed a
stage-specific surveillance protocol in patients after
radical cystectomy and suggested the use of upper
tract imaging every 1-2 years after radical cystectomy
for all tumor stages [26], but specific risk factors were
not taken into consideration in this study.

In a recent study, multifocality of the initial bladder


tumor was found to independently contribute to
a higher risk of ureteral involvement at radical
cystectomy and subsequent upper tract recurrence,
[4] which supports the results of a prior study [23].
Limited data exist as to whether or not urethral tumor
involvement predicts a higher likelihood for upper
tract recurrence [6, 14]. Urethral tumor involvement
in female patients has been significantly associated
with a higher risk for upper tract recurrence. Likewise,
in male patients, urothelial carcinoma of the prostatic
ducts invading the lamina propria was associated with
a higher risk for secondary ureteral tumors but not
for patients with a continuous expansion of urothelial
carcinoma into the prostate (classified as pT4a stage)
[6, 14]. Most presumably, this observation may also
be biased due to impaired survival rates in patients
with pT4a urothelial carcinoma of the bladder.

Although former studies reported similar sensitivity


rates of IVP and multidetector computed tomography
(MD-CT) urography for the detection of upper tract
recurrence ranging from 0-55%, [8, 9, 14, 19, 21, 26]
some current studies have adopted MD-CT urography
as the preferred diagnostic option compared to
conventional IVP for the detection of upper tract
tumors [27]. Contemporary series reporting on the
value of MD-CT derive mainly from series in which
patients had either primary upper tract tumors
or underwent upper tract imaging for a history of
microscopic or gross hematuria and are also limited
by a relatively small number of included patients. To
provide some recent aspects and developments in
this diagnostic field the results of the most current
studies will be outlined in the following.

4. Survival After Upper Urinary Tract


Recurrence
The 3-year survival rate for patients with upper
tract recurrence lies between 0 and 25%, but longterm survival of more than 9 years has also been
reported in some series [6, 19, 24]. Survival after
secondary ureteral recurrence is mainly predicted
by tumor stage and lymph node tumor involvement.
In a single-center series it was demonstrated that in
patients with pT3 or greater secondary upper tract
carcinoma, the median survival was only 1.3 years,
compared to 3.4 years in patients with pTa-T1 cancers
[6]. Similar results were seen in another series of 85
female patients who had an upper tract recurrence
rate of 2.4%. All recurrences were muscle-invasive
at nephroureterectomy [21].

In patients with primary upper tract tumors, MD-CT


urography shows a significantly higher sensitivity,
specificity, and test accuracy of 96%, 100%, and
99% as compared to IVP at 75%, 86%, and 85%,
respectively [27]. For MD-CT urography, the location
of the primary tumor seems to have a direct impact
on the detection accuracy. The sensitivity for the
detection of tumors in the renal pelvis was found
to range from 78-94% but decreased to 19-54% for

Some data suggest that tumor location also affects


survival in patients with upper tract recurrence
[25]. In some series, recurrence at the ureteroileal
anastomosis was found in up to 40-60% of the patients
with upper tract recurrence [4, 25]. Importantly,

333

tumor lesions in the ureter [28]. In a retrospective


series of 188 patients with a history of urothelial
carcinoma (using diagnostic ureterorenoscopy as the
diagnostic reference standard), MD-CT urography
showed a positive predictive value (PPV) of 63-67%
for the identification of a tumor lesion in the upper
tract [29, 30]. In both of these studies, characteristic
findings on MD-CT urography suspicious of an upper
tract tumor were filling defects and urothelial wall
thickening. Interestingly, when stratified by location,
urothelial wall thickening was more predictive of
tumors in the pelvicalyceal system (PPV 88%) than
in the ureter (PPV 33%). Conversely, filling defects
were more predictive in the ureter (PPV 88%) than
in the pelvicalyceal system (PPV 50%) [29]. MD-CT
was able to correctly predict the pTNM staging in
58-88% of the patients with upper tract malignancies
[28, 31]. However, even in contemporary series,
the main drawback of MD-CT urography is the low
positive predictive value for small tumors or urothelial
wall thickening (0-46%) [32].

cancer from intestinal epithelial cells [40]. Most


importantly, the additional use of urinary cytology
in established follow-up regimens (with radiological
and clinical examinations) does not significantly lead
to an earlier detection of upper tract recurrences
[40]. Selective ureteral urinary cytology results in an
improved positive predictive value of greater than
85% for high-grade muscle-invasive upper urinary
tract urothelial carcinomas, [41] but, for screening
purposes, invasiveness and additional costs have
to be considered.
Alternatively, urine-based markers have recently
become a popular tool to overcome the limitations of
conventional urinary cytology. Evidence in literature
derives mainly from patients with primary upper tract
tumors with relatively small numbers of included
patients. In a comparative analysis of 30 patients with
urothelial carcinoma of the upper tract, fluorescence in
situ hybridization (FISH) showed significantly higher
sensitivity rates as compared to urinary cytology
(77% vs. 36%) but not for specificity (95% vs. 100%),
and this has been confirmed in other studies [39, 42].
Moreover, FISH analysis, which uses gene-labeled
probes for the chromosomes 3, 7, 9, and 17, is less
interference-prone than urinary cytology [38]. For
immunocytology, a prior small study in 16 patients
with upper tract carcinoma reported a sensitivity rate
of 100% for the detection of low grade (G1-2) tumors
[43]. Altogether, due to their non-interference probe
analysis, FISH is a promising tool to contribute to a
more non-invasive diagnosis of high-grade upper
tract carcinomas. Nevertheless, taken together, there
is not enough evidence to support the routine use of
urinary cytology and urine-based markers alone for
the follow-up of the upper tracts [44].

The use of a split-bolus technique among 200


patients presented initially with hematuria was
prospectively assessed in a recent series [33].
Split-bolus techniques provide improved imaging
by simultaneous nephrographic and excretory CTphase acquisition. The corresponding sensitivity,
specificity, and PPV rates were 100%, 99%, and
99%, respectively [33]. Alternatively, in one study,
the use of MRI urography images was investigated
in 17 patients with a total of 23 upper tract lesions. A
sensitivity of 74% for the detection of small urothelial
carcinomas (2 cm or less) was reported [34].
Possible further technical improvements include
the acquisition of diffusion-weighted MRI images in
patients with upper tract obstruction due to upper
tract tumor lesions [35], but, still, these techniques
have failed to detect flat ureteral lesions [36].

U
 reterorenoscopy
Ureterorenoscopy with biopsy using semirigid or
flexible instruments is the method of choice for the
histological diagnosis of an upper tract carcinoma [45].
Access to the upper tract after urinary diversion can
be particularly challenging and may require the use
of combined retrograde and antegrade techniques
[46]. For evaluating the local T stage, upper tract
biopsies are less reliable because of the limited
capability of biopsy instruments to retrieve adequate
tissue specimens sufficient for the evaluation of the
complete pelvicalyceal or ureteral wall [6]. This fact
is supported by a prior study that demonstrated that
even when lamina propria was present within the
specimens, almost 50% of the tumors were initially
misdiagnosed as superficial and were found to be
invasive in the final nephroureterectomy specimens
[47].

For local staging of the primary tumor lesion and


regional lymph nodes, magnetic resonance imaging
(MRI) has not proven diagnostic superiority to MDCT but is indicated in patients with contraindications
to MD-CT (iodine allergy, renal insufficiency) [34].
MRI remains contraindicated in patients with
pacemakers. Gadolinium infusion must be omitted
in patients with severe renal insufficiency (creatinine
clearance less than 30 mL/min) to avoid the risk of
subsequent nephrogenic systemic fibrosis. [37].
U
 rinary Cytology and Molecular-based Urinary
Markers
Some authors have used urinary cytology at least
annually in patients after radical cystectomy [2, 13].
However, despite its high specificity (80-100%) the
sensitivity of voided urine cytology for the detection
of upper tract tumors is low (36-60%) (38, 39). After
urinary diversion, the detection of malignant cells
in voided urine samples is significantly hampered
because of the difficulty in distinguishing urothelial

6. Treatment of Upper Urinary Tract Recurrence


R
 adical Nephroureterectomy
Radical nephroureterectomy is the standard treat-

334

ment for patients with invasive upper tract recurrence [6, 13, 21]. Eligibility of patients (in terms of
performance status and tumor extent) to undergo
open radical nephroureterectomy is a predictive factor for improved survival. A prior study demonstrated
that while the overall median survival of patients with
upper tract recurrence was only 10 months, those
who were eligible for radical nephroureterectomy
had a median overall survival of 26 months [8].

survival. Interestingly, in this study, patients who


did not receive any lymphadenectomy at the time of
surgery had comparable survival rates to those with
pathologically confirmed node-positive disease at
radical nephroureterectomy. By contrast to patients
with upper tract recurrence, the rate of lymph-node
positive disease in primary upper tract carcinoma
was only 25% [52, 53]. Moreover, the ideal anatomic
template of regional lymphadenectomy according
to the primary site of recurrence has not been
established, which is mainly due to the inconsistent
lymphatic drainage in the retroperitoneum in
urologic malignancies [54]. Altogether, given the
lack of evidence for a therapeutic benefit of regional
lymphadenectomy, its role in patients with upper tract
recurrence remains unclear and therefore currently
is considered as a diagnostic procedure.

Increasing experience with laparoscopic radical


nephroureterectomy suggests that it is oncologically
equivalent to open surgery for tumors of equal
grade, stage, and lymph node status. These
data derive mainly from patients with primary
upper tract carcinoma. A recent study comparing
retrospectively the outcomes of patients treated
with open (N=704 patients) and laparoscopic radical
nephroureterectomy (N=70 patients) for primary
upper tract urothelial carcinoma found no significant
differences in 5-year cancer-specific survival
between both the open and laparoscopic groups [48].
One limitation of this analysis is that patients treated
with laparoscopic radical nephroureterectomy had
significantly less likelihood for lymph node tumor
involvement at the time of surgery as compared to
patients treated with open surgery, which might have
biased the final survival analysis. Nonetheless, given
the lack of data and the challenging postoperative
anatomy in the retroperitoneum and pelvis after radical
cystectomy and urinary diversion, with the need to
enter and reconstitute the ureteroileal anastomosis,
laparoscopic
radical
nephroureterectomy
in
secondary upper tract recurrence has still to be
regarded as an investigational treatment option.

A
 djuvant Chemotherapy
The role of adjuvant chemotherapy in node-positive
upper tract recurrence has only been addressed in
a few studies [13]. In locally advanced upper tract
carcinoma, the results of adjuvant chemotherapy
remain poor as only 5% or fewer of the patients will
achieve long-term survival [13]. The low incidence
of upper tract recurrences makes it unlikely that
randomized prospective studies on the role of
adjuvant therapy can be performed. Controlled
observational and registration type multi-institutional
studies with pooling of data are encouraged.
C
 onservative Treatment for Upper Tract Recurrence
Conservative treatment options for patients with
upper tract recurrence after invasive bladder
cancer include endoscopic and percutaneous
resections. Indications for performing renal-sparing
treatment in patients with upper tract recurrence
are bilateral tumors, solitary kidney, or severe renal
insufficiency, and tumor features of small size,
low grade, and low stage. Studies reporting on
outcomes after endoscopic treatment of upper tract
carcinomas derive almost exclusively from patients
with primary carcinomas who have a significantly
improved prognosis as compared to patients with
metachronous tumors. Approximately 50-75% of the
patients with upper tract recurrence present initially
with high-stage, high-grade disease compared to
only about 30% of the patients with de novo urothelial
carcinoma of the upper tract [6, 8, 9, 19, 21, 24, 55,
56]. Moreover, in patient series performing radical
nephroureterectomy and regional lymphadenectomy,
lymph node tumor involvement has been reported
in 32-83% [6, 8, 14]. These data indicate that upper
tract recurrences are more aggressive malignancies
than primary urothelial cancers. Optimal indications
for conservative nephron-sparing treatment of
secondary upper tract tumor recurrences remain to
be defined.

L
 ymphadenectomy
The role and extent of regional lymphadenectomy at
radical nephroureterectomy in patients with upper tract
recurrence has not been well-defined and remains
controversial. Open radical nephroureterectomy was
accompanied by regional lymphadenectomy only in
a few retrospective series [6, 8, 14]. In these series,
the rate of locally advanced disease (pT3-T4) ranged
from 54-66% with lymph-node positive disease found
in 32-83% of the patients.
Similar to the increasing evidence for a therapeutic
benefit of extended lymphadenectomy for invasive
bladder cancer, [49, 50] there is recent evidence that
also suggests that extended lymphadenectomy in
invasive upper tract carcinoma is a strong predictor
for improved disease-free and cancer-specific
survival [51]. A retrospective series on 132 patients
with primary muscle-invasive upper tract urothelial
carcinoma showed that pathological tumor stage and
lymph node tumor involvement were independent
risk factors for disease-free survival [52]. Moreover,
the number of removed lymph nodes and the
extent of lymphadenectomy was an independent
risk factor for recurrence-free and cancer-specific

335

Recommendations for Follow-up and


Treatment of Secondary Ureteral
Tumors

63]. Patients with symptomatic urethral recurrences


have no significantly different median time to
diagnosis as compared to patients diagnosed by
cytological abnormalities [58]. Urethral recurrences
in superficial tumor stages (pTa, pTis) are diagnosed
in many cases by abnormal cytology (59-100%) as
opposed to invasive tumors (pT1-T4), which are
often detected after patients have developed local
symptoms (~79%) such as gross hematuria, bloody
urethral discharge, palpable mass, change in voiding
habits, or local pain [58, 64].

Surveillance regimens should be based on a riskadapted strategy (LE: 3, GR: B).


T
 he number of risk factors potentiates the risk
of recurrence. Risk factors include: (a) positive
ureteral margin, (b) carcinoma in situ of the
bladder and ureter, (c) tumor multifocality, (d)
urethral tumor involvement, and (e) male gender
(LE: 3, GR: B).

2. Risk Factors in Male Patients


In a large, retrospective single-center series of 729
male patients treated with radical cystectomy with
a median follow-up of 38 months, independent risk
factors for urethral recurrence were found to be a
history of non-muscle-invasive bladder cancer,
prostatic urethral involvement, and NMIBC (pTa,
pTis, pT1) at radical cystectomy. In this study,
superficial prostatic involvement (pTa, pTis) was a
stronger predictor for urethral recurrence compared
to invasive prostatic involvement [58)]. Similar
findings were found in another large, retrospective,
single-center series of 768 male patients with a
longer median follow-up of 13 years in which prostatic
urethral involvement, but not tumor multifocality
or bladder CIS, was an independent risk factor for
urethral recurrence. By contrast to the prior report
[58], patients with invasive prostatic involvement
had a higher risk for urethral recurrence compared
to patients with superficial prostatic urethral tumors
[57].

Patients with final positive ureteral margins in


the definitive pathology report at surgery are at
increased risk of ureteral recurrence. Frozen
section analysis for this finding has a high
sensitivity and specificity for the detection of
malignant ureteral margins (LE: 3, GR B).
Ureterorenoscopy with biopsy is the diagnostic
procedure of choice for the diagnosis of a
metachronous upper tract recurrence after
radical cystectomy (LE: 3, GR: C).
Cytology and urine-based markers such
as fluorescence in situ hybridization and
immunocytology are indicative of upper tract
recurrence, but their sole use cannot be
recommended for exclusive follow-up of the
upper tracts (LE: 3, GR: B).
From an oncologic standpoint, routine upper
tract imaging is only indicated in patients with
high risk or clinical symptoms suspicious of a
metachronous upper tract recurrence and the
optimal interval for follow-up imaging is unclear
(LE: 3, GR: C).

3. Urinary Diversion and Urethral Recurrence


Some series suggest that patients with orthotopic
diversion have a decreased risk for urethral
recurrence as compared to patients with
supranational diversions [57, 61, 65]. However,
other comparable series did not find orthotopic
diversion to be a significant independent risk factor
compared to cutaneous diversions [58]. Prior studies
have hypothesized that the continued exposure
of the remnant urothelium to urine in orthotopic
diversions may have a protective effect against the
development of a urethral recurrence [66]. These
conflicting results may also be due to a patient
selection bias since patients with lower tumor stages
might rather be treated with orthotopic diversion than
patients with locally advanced disease [58].

Nephroureterectomy is the treatment of choice


for invasive upper tract recurrence, providing
prolonged survival (LE 3, GR: B).
Based on scarce data, regional lymphadenectomy
in patients undergoing radical nephroureterectomy
is of limited diagnostic value, and its therapeutic
role is undetermined. (LE: 3, GR: C).

c) Secondary urethral tumors


1. Incidence of Urethral Recurrences after
Radical Cystectomy
According to contemporary series, the risk of urethral
recurrence is low, occurring after a median time of 1424 months in 3.7-8.1% of male patients undergoing
radical cystectomy for bladder cancer, [57-62] with a
resulting 5-year urethral recurrence-free probability
after radical cystectomy of approximately 95% [57,
60]. Most urethral recurrences are detected by
evaluation of symptoms (57-61%), whereas 31-39%
are detected by abnormal urethral cytology [58,

4. Preoperative Prostatic Biopsy and Frozen


Section Analysis
In one study, preoperative prostatic urethral biopsies
showed a high negative predictive value of 99%
compared to final urethral margin status whereas a
positive finding correlated in only 68% of the cases
with a positive final urethral margin status [67].
Confirmatory results of the relatively low positive

336

predictive value of transurethral loop biopsies at


the level of the verumontanum as to the final results
of cystoprostatectomy specimens were reported in
246 patients. The sensitivity and specificity was only
53% and 77%, resulting in a positive and negative
predictive value of 45% and 77%, respectively [68].
By contract, in this study, sensitivity and specificity
of frozen section analysis was 100% [67]. A recent
retrospective study of 294 patients confirmed that
a positive urethral margin status is the strongest
independent risk factor for urethral recurrence as
compared to prostatic urethral and stromal invasion
[60)]. This strengthens the importance of achieving a
tumor-free urethral margin at radical cystectomy.

of the urethral margin for frozen section analysis


[73]. Altogether, the present data underscore the
value of the intraoperative frozen section analysis
for the detection of urethral malignancy at radical
cystectomy. Furthermore, they add to the growing
body of evidence that patients with invasive bladder
cancer at the bladder neck should not be excluded
from an orthotopic approach in advance unless
the intraoperative frozen section analysis reveals
evidence of malignancy at the distal urethral margin
[73].

6. Survival after Urethral Recurrence


The 5-year disease-specific and overall survival after
urethral recurrence is reported to be only 35% and
52%, with a median overall survival after diagnosis of
28-54 months [58, 60, 63]. There are heterogeneous
results in the current literature whether bladder or
urethral recurrence pathology is the most important
factor for overall survival after urethral recurrence.
While Huguenot et al found no differences in survival
[58], Lin et al reported in 24 patients with urethral
recurrence that bladder pathology determined
predominantly their survival [77]. Conversely,
Clark et al found among 47 patients with urethral
recurrence that stage of urethral recurrence but
not bladder cancer pathology was more predictive
for overall survival [63]. Possible explanations for
these contradictory observations may be different
patient selection criteria and the different use of
prophylactic urethrectomies at radical cystectomy
which may have biased the survival analyses of
both studies. Second, the number of patients with
pT0 disease after urethrectomy differs considerably
in the 2 studies. Lin et al included 9 patients (37%)
with pT0 disease [77], whereas Clark et al excluded
all pT0 patients from final survival analysis [63].
In their analysis, Huguet et al found none of the
patients to have pT0 stage at urethrectomy [58],
and, most importantly, 26% of their patients with
urethral recurrence had upper tract recurrence, [58]
which tends to be high stage and high-grade disease
at initial diagnosis [6]. Third, survival in patients
with urethral recurrence and concomitant distant
metastases is mainly determined by the presence of
metastatic disease [2]. In conclusion, it remains still a
matter of debate whether urethral or bladder cancer
pathology is more predictive for overall survival after
urethral recurrence.

5. Risk Factors in Female Patients


In female patients undergoing radical cystectomy,
the rate of urethral recurrence was reported to range
between 0.8-4.3% [21, 69-71]. By contrast to the low
rate of concomitant urethral malignancy in females
with primary bladder cancer (~2%), [72] a higher
rate of urethral malignancy at radical cystectomy of
12% has been reported [73]. A history of multifocal
or recurrent bladder cancer was found to contribute
to an increased risk for urethral recurrence [21].
Furthermore, the risk of tumor involvement of the
distal urethra at radical cystectomy is significantly
higher in patients with bladder neck or vaginal
involvement and preoperatively enlarged inguinal
lymph nodes [74-76]. A pathological reevaluation
of 67 female cystectomy specimens showed that
concomitant urethral tumors were exclusively
located in the proximal or mid portion of the urethra.
Moreover, women with urethral tumors often showed
localization of the primary tumor at the bladder neck.
These tumors were of higher grade and stage and
harbored a higher risk of node-positive disease
[75]. Stein et al prospectively evaluated 71 female
cystectomy specimens and reported that tumor
localization at the bladder neck and urethra was
present in 19% and 7%, respectively. The presence
of a bladder neck tumor significantly correlated with
the highest risk of concomitant urethral malignancy.
Conversely, approximately 60% of the patients with
malignancy at the bladder neck had no evidence of
tumor involvement of the proximal urethra. In this
study, frozen section analysis of the distal urethral
margin showed a 100% sensitivity and specificity for
detection of a positive urethral margin as compared
with the final pathological analysis [74]. Another study
on 85 female patients confirmed the usefulness of
frozen section analysis with a sensitivity of 100%
for the detection of a malignant urethral margin
[21]. Nonetheless, one must remember that
urethral recurrence may occur even in patients
with negative final urethral margins possibly due
to a submucosal extension of the primary bladder
tumor into the urethra with unaffected urothelial
layer at frozen section analysis [76]. This highlights
the importance of an adequate full-thickness biopsy

7. Diagnosis and Staging of Urethral Recurrence


Given the low rate of urethral recurrence, there is
little data or agreement on the optimal follow-up of
the asymptomatic patient with a retained urethra
after radical cystectomy. All patients with urethral
bleeding, pain, or mass should be evaluated
promptly. The follow-up regimens described in
current literature derive mainly from different highvolume centers and are based on urinary cytology,

337

urethral washings, urethroscopy, and different


imaging modalities [57, 59, 61, 62, 77, 78]. In terms
of frequency of follow-up, some authors suggest the
routine use of urinary cytology, urethral washings,
and urethroscopy on a quarterly basis for the first
2 years continued semiannually afterwards, [61]
whereas others only conduct clinical follow-up [58].
It seems therefore reasonable to tailor surveillance
regimens according to the patients individual risk
profile. Urinary cytology is certainly a useful
and noninvasive diagnostic tool to detect urethral
recurrence but its sensitivity is considerably reduced
in patients after urinary diversion [40]. Controversial
data exist to the question whether the routine use of
urethral washings in patients with a retained urethra
after cystoprostatectomy is of prognostic advantage.
Among 24 patients with a median follow-up of 28
months after urethrectomy, a significant impact on
disease-free survival was not observed between
those who routinely received urethral washings and
those who did not [77]. Urethroscopy with biopsy is
certainly the method of choice for the histological
diagnosis of urethral recurrence but the use of
urethroscopy on a regular basis has not appear to
provide any survival benefit [9, 57, 59, 60, 62]. For
local staging after detection of a urethral malignancy,
there is increasing evidence that magnetic resonance
imaging is superior to MD-CT in patients with urethral
malignancy in terms of staging accuracy[78].

urethrectomy at the time of urethral recurrence [58,


61]. Similar results were obtained from a recent
analysis of the SEER database investigating the
timing of urethrectomy as a possible predictor for
improved survival among 2401 male patients who
underwent radical cystoprostatectomy for bladder
cancer. A total of 195 men (8.1%) developed urethral
recurrence and were treated with either concurrent
or salvage urethrectomy. The use of concurrent
urethrectomy at radical cystoprostatectomy versus
salvage urethrectomy was not found to confer any
significant independent survival benefit (HR=0.775,
95% CI 0.592-1.014, p=0.0632). Interestingly, by
contrast to patients treated at urban non-teaching
hospitals, patients treated at teaching hospitals were
more likely to undergo salvage urethrectomy for
recurrence versus immediate urethrectomy. In this
study, tumor stage at radical cystectomy was the only
independent variable that predicted performance
of concurrent urethrectomy [62]. Furthermore,
complication rates and intraoperative blood loss
are not significantly different in patients treated
with delayed or immediate urethrectomy [79]. In
female patients, given the overall low rate of urethral
recurrence, the use of prophylactic urethrectomies
without evidence of invasive malignancy at the
urethral margin does not have a significant impact
on the oncological long-term outcome after radical
cystectomy [72, 80, 81]. As for upper tract recurrence,
adjuvant treatment modalities for urethral recurrence
have yet to be addressed sufficiently in the literature
and should be performed only within prospective
randomized trials.

8. Management of Urethral Recurrence


Non-invasive Urothelial Tumors (pTa-pTis)
In patients with a non-invasive urethral recurrence, a
urethra-preserving strategy can be attempted using
transurethral resection of the tumor lesion [64].
However, in a series with high-grade pTa urothelial
carcinoma of the urethra after radical cystectomy
treated with TUR, only 3 of 4 patients with recurrence
progressed to invasive disease. After recurrence,
all were managed with salvage urethrectomy and
remained disease-free after a median time of 2444 months [64]. The authors concluded that an
additional adjuvant local treatment, i.e., BCG, in high
grade tumors may have reduced the risk of further
recurrence [64]. In a prospective analysis of patients
with carcinoma in situ in the urethra after radical
cystectomy, the use of intraurethral BCG perfusion
weekly for 6 weeks showed a complete remission rate
of 80% but did not show any therapeutic response in
patients with papillary or invasive disease [59].

Recommendations for Follow-up and


Treatment of Secondary Urethral
Tumors
Risk factors for urethral recurrence in male
patients are prostatic tumor involvement (either
superficial or invasive) and bladder neck
involvement in female patients (LE: 3, GR: B).
Intraoperative frozen section analysis has a
high sensitivity and specificity for the detection
of a malignant urethral margin for both male and
female patients (LE: 2b, GR: B).
Patients with positive urethral margins in the final
pathology report are at increased risk of urethral
recurrence (LE: 3, GR: B).
T
 he use of routine surveillance urinary cytology,
urethral washings, and diagnostic urethroscopy in
asymptomatic patients has not shown a survival
benefit (LE: 3, GR: B).

Invasive Urothelial Tumors (pT1-pT4)


Urethrectomy is the treatment option of choice in
male patients with invasive urethral recurrence
providing long-term survival [62]. In 2 recent studies,
the use of prophylactic urethrectomies in male
patients with invasive prostatic tumor involvement
at radical cystectomy was not found to confer any
survival benefit compared to patients treated with

Urethrectomy is the preferred treatment in patients


with invasive carcinoma at the urethral margin at
radical cystectomy or at a later recurrence (LE
3, GR B)

338

Local conservative treatment is an option in


patients with non-invasive tumor(s) or carcinoma
in situ of the urethra (LE: 3, GR: C).

14. Sved PD, Gomez P, Nieder AM, Manoharan M, Kim SS,


Soloway MS (2004) Upper tract tumour after radical
cystectomy for transitional cell carcinoma of the bladder:
incidence and risk factors. BJU Int 94, 785-789

In patients with urethral recurrence and


concomitant distant recurrence outside the
urinary tract, systemic chemotherapy is indicated
(LE: 3, GR: C).

15. Tran W, Serio AM, Raj GV, Dalbagni G, Vickers AJ,


Bochner BH, Herr H, Donat SM (2008) Longitudinal risk
of upper tract recurrence following radical cystectomy for
urothelial cancer and the potential implications for longterm surveillance. J Urol 179, 96-100

References

16. Volkmer BG, Schnoeller T, Kuefer R, Gust K, Finter F,


Hautmann RE (2009) Upper urinary tract recurrence after
radical cystectomy for bladder cancer--who is at risk? J
Urol 182, 2632-2637
17. Osman Y, El-Tabey N, Abdel-Latif M, Mosbah A, Moustafa
N, Shaaban A (2007) The value of frozen-section analysis
of ureteric margins on surgical decision-making in patients
undergoing radical cystectomy for bladder cancer. BJU Int
99, 81-84

1. Modified from Oxford Centre for Evidence-based


Medicine Levels of Evidence (May 2001). Produced by
Bob Phillips, Chris Ball, Dave Sackett, Doug Badenoch,
Sharon Straus, Brian Haynes, Martin Dawes since
November 1998. http://www.cebm.net/index.aspx?o=1025
[access date February 2009].

18. Herr HW, Whitmore WF Jr (1987) Ureteral carcinoma in


situ after successful intravesical therapy for superficial
bladder tumors: incidence, possible pathogenesis and
management. J Urol 138, 292-294

2. Giannarini G, Kessler TM, Thoeny HC, Nguyen DP,


Meissner C, Studer UE (2010) Do patients benefit from
routine follow-up to detect recurrences after radical
cystectomy and ileal orthotopic bladder substitution? Eur
Urol 58, 486-494

19. Kenworthy P, Tanguay S, Dinney CP (1996) The risk of


upper tract recurrence following cystectomy in patients with
transitional cell carcinoma involving the distal ureter. J Urol
155, 501-503

3. Meissner C, Giannarini G, Schumacher MC, Thoeny H,


Studer UE, Burkhard FC (2007) The efficiency of excretory
urography to detect upper urinary tract tumors after
cystectomy for urothelial cancer. J Urol 178, 2287-2290

20. Solsona E, Iborra I, Rics JV, Dumont R, Casanova JL,


Calabuig C (1997) Upper urinary tract involvement in
patients with bladder carcinoma in situ (Tis): its impact on
management. Urology 49, 347-352

4. Gakis G, Schilling D, Perner S, Schwentner C, Sievert KD,


Stenzl A (2010) Sequential resection of malignant ureteral
margins at radical cystectomy: a critical assessment of the
value of frozen section analysis. World J Urol Epub ahead
of print
5. Tollefson MK, Blute ML, Farmer SA, Frank I (2010)
Significance of distal ureteral margin at radical cystectomy
for urothelial carcinoma. J Urol 183, 81-86

21. Akkad T, Gozzi C, Deibl M, Mller T, Pelzer AE, Pinggera


GM, Bartsch G, Steiner H (2006) Tumor recurrence in
the remnant urothelium of females undergoing radical
cystectomy for transitional cell carcinoma of the bladder:
long-term results from a single center. J Urol 175, 12681271

6. Sanderson KM, Cai J, Miranda G, Skinner DG, Stein JP


(2007) Upper tract urothelial recurrence following radical
cystectomy for transitional cell carcinoma of the bladder:
an analysis of 1,069 patients with 10-year followup. J Urol
177, 2088-2094

22. Stein JP, Lieskovsky G, Cote R, Groshen S, Feng AC,


Boyd S, Skinner E, Bochner B, Thangathurai D, Mikhail M,
Raghavan D, Skinner DG (2001) Radical cystectomy in the
treatment of invasive bladder cancer: long-term results in
1,054 patients. J Clin Oncol 19, 666-675

7. Schoenberg MP, Carter HB, Epstein JI (1996) Ureteral


frozen section analysis during cystectomy: a reassessment.
J Urol 155, 1218-1220

23. Sharma TC, Melamed MR, Whitmore WF Jr (1970)


Carcinoma in-situ of the ureter in patients with bladder
carcinoma treated by cystectomy. Cancer 26, 583-587

8. Balaji KC, McGuire M, Grotas J, Grimaldi G, Russo P (1999)


Upper tract recurrences following radical cystectomy: an
analysis of prognostic factors, recurrence pattern and stage
at presentation. J Urol 162, 1603-1606

24. Malkowicz SB, Skinner DG (1990) Development of upper


tract carcinoma after cystectomy for bladder carcinoma.
Urology 36, 20-22

9. Huguet-Prez J, Palou J, Milln-Rodrguez F, SalvadorBayarri J, Villavicencio-Mavrich H, Vicente-Rodrguez J


(2001) Upper tract transitional cell carcinoma following
cystectomy for bladder cancer. Eur Urol 40, 318-323

25. Yossepowitch O, Dalbagni G, Golijanin D, Donat SM,


Bochner BH, Herr HW, Fair WR, Russo P ( 2003) Orthotopic
urinary diversion after cystectomy for bladder cancer:
implications for cancer control and patterns of disease
recurrence. J Urol 169, 177-181

10. Raj GV, Tal R, Vickers A, Bochner BH, Serio A, Donat


SM, Herr H, Olgac S, Dalbagni G (2006) Significance of
intraoperative ureteral evaluation at radical cystectomy for
urothelial cancer. Cancer 107, 2167-2172

26. Slaton JW, Swanson DA, Grossman HB, Dinney CP (1999)


A stage specific approach to tumor surveillance after radical
cystectomy for transitional cell carcinoma of the bladder. J
Urol 162, 710-714

11. Schumacher MC, Scholz M, Weise ES, Fleischmann A,


Thalmann GN, Studer UE (2006) Is there an indication for
frozen section examination of the ureteral margins during
cystectomy for transitional cell carcinoma of the bladder? J
Urol 176, 2409-2413

27. Wang LJ, Wong YC, Huang CC, Wu CH, Hung SC, Chen
HW (2010) Multidetector computerized tomography
urography is more accurate than excretory urography for
diagnosing transitional cell carcinoma of the upper urinary
tract in adults with hematuria. J Urol 183, 48-55.

12. Silver DA, Stroumbakis N, Russo P, Fair WR, Herr HW


(1997) Ureteral carcinoma in situ at radical cystectomy:
does the margin matter? J Urol 158, 768-771
13. Solsona E, Iborra I, Rubio J, Casanova J, Dumont R,
Monrs JL (2003) Late oncological occurrences following
radical cystectomy in patients with bladder cancer. Eur Urol
43, 489-494

28. Scolieri MJ, Paik ML, Brown SL, Resnick MI (2000)


Limitations of computed tomography in the preoperative
staging of upper tract urothelial carcinoma. Urology 56.,
930-934
29. Xu AD, Ng CS, Kamat A, Grossman HB, Dinney C, Sandler
CM (2010) Significance of upper urinary tract urothelial

339

thickening and filling defect seen on MDCT urography in


patients with a history of urothelial neoplasms. AJR Am J
Roentgenol 1955, 959-956

Voided urine fluorescence in situ hybridization testing for


upper tract urothelial carcinoma surveillance. J Urol 184,
879-882

30. Caoili EM, Cohan RH, Inampudi P, Ellis JH, Shah RB,
Faerber GJ, Montie JE (2005) MDCT urography of upper
tract urothelial neoplasms. AJR Am J Roentgenol 184,
1873-1881

45. Brien JC, Shariat SF, Herman MP, Ng CK, Scherr DS,
Scoll B, Uzzo RG, Wille M, Eggener SE, Terrell JD, Lucas
SM, Lotan Y, Boorjian SA, Raman JD (2010) Preoperative
hydronephrosis, ureteroscopic biopsy grade and urinary
cytology can improve prediction of advanced upper tract
urothelial carcinoma. J Urol 184., 69-73

31. Fritz GA, Schoellnast H, Deutschmann HA, Quehenberger


F, Tillich M ( 2006) Multiphasic multidetector-row CT (MDCT)
in detection and staging of transitional cell carcinomas of
the upper urinary tract. Eur Radiol 16, 1244-1252

46. Sanderson KM, Rouprt M (2007) Upper urinary tract


tumour after radical cystectomy for transitional cell
carcinoma of the bladder: an update on the risk factors,
surveillance regimens and treatments. BJU Int. 100, 11-16

32. Sadow CA, Wheeler SC, Kim J, Ohno-Machado L,


Silverman SG (2010) Positive predictive value of CT
urography in the evaluation of upper tract urothelial cancer.
AJR Am J Roentgenol 195, W337-343

47. Lee BR, Jabbour ME, Marshall FF, Smith AD, Jarrett TW
(1999) 13-year survival comparison of percutaneous and
open nephroureterectomy approaches for management
of transitional cell carcinoma of renal collecting system:
equivalent outcomes. J Endourol 13., 289-294

33. Maheshwari E, OMalley ME, Ghai S, Staunton M, Massey


C (2010) Split-bolus MDCT urography: Upper tract
opacification and performance for upper tract tumors in
patients with hematuria. AJR Am J Roentgenol 194, 453458

48. Walton TJ, Novara G, Matsumoto K, Kassouf W, Fritsche


HM, Artibani W, Bastian PJ, Martnez-Salamanca JI,
Seitz C, Thomas SA, Ficarra V, Burger M, Tritschler S,
Karakiewicz PI, Shariat SF (2010) Oncological outcomes
after laparoscopic and open radical nephroureterectomy:
results from an international cohort. BJU Int doi: 10.1111/
j.1464-410X.2010.09826.x.

34. Takahashi N, Kawashima A, Glockner JF, Hartman RP,


Leibovich BC, Brau AC, Beatty PJ, King BF. (2008) Small
(<2-cm) upper-tract urothelial carcinoma: evaluation with
gadolinium-enhanced three-dimensional spoiled gradientrecalled echo MR urography. Radiology 247, 451-457

49. Dhar NB, Klein EA, Reuther AM, Thalmann GN,


Madersbacher S, Studer UE (2008) Outcome after radical
cystectomy with limited or extended pelvic lymph node
dissection. J Urol 179, 873-878

35. Takeuchi M, Matsuzaki K, Kubo H, Nishitani H (2008)


Diffusion-weighted magnetic resonance imaging of urinary
epithelial cancer with upper urinary tract obstruction:
preliminary results. Acta Radiol 49, 1195-1199.

50. Dhar NB, Campbell SC, Zippe CD, Derweesh IH, Reuther
AM, Fergany A, Klein EA (2006) Outcomes in patients with
urothelial carcinoma of the bladder with limited pelvic lymph
node dissection. BJU Int 98, 1172-1175

36. Nishizawa S, Imai S, Okaneya T, Nakayama T, Kamigaito


T, Minagawa T ( 2010) Diffusion weighted imaging in the
detection of upper urinary tract urothelial tumors. Int Braz J
Urol 36, 18-28

51. Kondo T, Hashimoto Y, Kobayashi H, Iizuka J, Nakazawa H,


Ito F, Tanabe K (2010) Template-based lymphadenectomy
in urothelial carcinoma of the upper urinary tract: impact on
patient survival. Int J Urol 17, 848-854

37. Kim KH, Fonda JR, Lawler EV, Gagnon D, Kaufman JS


(2010) Change in use of gadolinium-enhanced magnetic
resonance studies in kidney disease patients after US Food
and Drug Administration warnings: a cross-sectional study
of Veterans Affairs Health Care System data from 20052008. Am J Kidney Dis 56, 458-467

52. Roscigno M, Shariat SF, Freschi M, Margulis V, Karakiewizc


P, Suardi N, Remzi M, Zigeuner R, Bolenz C, Kikuchi E,
Weizer A, Bensalah K, Sagalowsky A, Koppie TM, Raman
J, Fernndez M, Strbel P, Kabbani W, Langner C, Wheat
J, Guo CC, Kassouf W, Haitel A, Wood CG, Montorsi F.
(2009) Assessment of the minimum number of lymph
nodes needed to detect lymph node invasion at radical
nephroureterectomy in patients with upper tract urothelial
cancer. Urology 74, 1070-1074

38. Akkad T, Brunner A, Pallwein L, Gozzi C, Bartsch G, Mikuz


G, Steiner H, Verdorfer I (2007) Fluorescence in situ
hybridization for detecting upper urinary tract tumors--a
preliminary report. Urology 70, 753-757
39. Marn-Aguilera M, Mengual L, Ribal MJ, Musquera M,
Ars E, Villavicencio H, Algaba F, Alcaraz A (2007) Utility
of fluorescence in situ hybridization as a non-invasive
technique in the diagnosis of upper urinary tract urothelial
carcinoma. Eur Urol 51, 409-415

53. Roscigno M, Shariat SF, Margulis V, Karakiewicz P, Remzi


M, Kikuchi E, Zigeuner R, Weizer A, Sagalowsky A,
Bensalah K, Raman JD, Bolenz C, Kassou W, Koppie TM,
Wood CG, Wheat J, Langner C, Ng CK, Capitanio U, Bertini
R, Fernndez MI, Mikami S, Isida M, Strbel P, Montorsi F.
Eur Urol 56, 512-518

40. Yoshimine S, Kikuchi E, Matsumoto K, Ide H, Miyajima


A, Nakagawa K, Oya M (2010) The clinical significance
of urine cytology after a radical cystectomy for urothelial
cancer. Int J Urol 17, 527-532

54. Pao B, Sebasti C, Buesch L, Mestres J, Salvador R,


Macas NG, Nicolau C (2011) Pathways of lymphatic spread
in male urogenital pelvic malignancies. Radiographics 31,
135-160

41. Messer J, Shariat SF, Brien JC, Herman MP, Ng CK, Scherr
DS, Scoll B, Uzzo RG, Wille M, Eggener SE, Steinberg G,
Terrell JD, Lucas SM, Lotan Y, Boorjian SA, Raman JD (2011)
Urinary cytology has a poor performance for predicting
invasive or high-grade upper-tract urothelial carcinoma.
BJU Int doi: 10.1111/j.1464-410X.2010.09899.x.

55. Rouprt M, Hupertan V, Traxer O, Loison G, Chartier-Kastler


E, Conort P, Bitker MO, Gattegno B, Richard F, Cussenot
O (2006) Comparison of open nephroureterectomy and
ureteroscopic and percutaneous management of upper
urinary tract transitional cell carcinoma. Urology 67, 11811187

42. Shan Z, Wu P, Zheng S, Tan W, Zhou H, Zuo Y, Qi H, Zhang


P, Peng H, Wang Y (2010) Evaluation of upper urinary
tract tumors by FISH in Chinese patients. Cancer Genet
Cytogenet 203, 238-246
43. Lodde M, Mian C, Wiener H, Haitel A, Pycha A, Marberger
M (2001) Detection of upper urinary tract transitional cell
carcinoma with ImmunoCyt: a preliminary report. Urology
58, 362-366.

56. Guarnizo E, Pavlovich CP, Seiba M, Carlson DL, Vaughan


ED Jr, Sosa RE (2000) Ureteroscopic biopsy of upper
tract urothelial carcinoma: improved diagnostic accuracy
and histopathological considerations using a multi-biopsy
approach. J Urol 163, 52-55

44. Johannes JR, Nelson E, Bibbo M, Bagley DH (2010)

57. Stein JP, Clark P, Miranda G, Cai J, Groshen S, Skinner DG

340

(2005) Urethral tumor recurrence following cystectomy and


urinary diversion: clinical and pathological characteristics in
768 male patients. J Urol 173, 1163-1168

74. Stein JP, Esrig D, Freeman JA, Grossfeld GD, Ginsberg


DA, Cote RJ, Groshen S, Boyd SD, Lieskovsky G, Skinner
DG (1998) Prospective pathologic analysis of female
cystectomy specimens: risk factors for orthotopic diversion
in women. Urology 51, 951-955

58. Huguet J, Monllau V, Sabat S, Rodriguez-Faba O, Algaba


F, Palou J, Villavicencio H. (2008) Diagnosis, risk factors,
and outcome of urethral recurrences following radical
cystectomy for bladder cancer in 729 male patients. Eur
Urol 53, 785-792

75. Stein JP, Cote RJ, Freeman JA, (1995) Indications for lower
urinary tract reconstruction in women after cystectomy for
bladder cancer: a pathological review of female cystectomy
specimens. J Urol 154, 1329-1333

59. Varol C, Thalmann GN, Burkhard FC, Studer UE (2004)


Treatment of urethral recurrence following radical
cystectomy and ileal bladder substitution. J Urol 172, 937942

76. Maralani S, Wood DP Jr, Grignon D, Banerjee M, Sakr


W, Pontes JE (1997.) Incidence of urethral involvement
in female bladder cancer: an anatomic pathologic study.
Urology 50, 537-541

60. Cho KS, Seo JW, Park SJ, Lee YH, Choi YD, Cho NH, Yang
SC, Hong SJ (2009) The risk factor for urethral recurrence
after radical cystectomy in patients with transitional cell
carcinoma of the bladder. Urol Int 82, 306-311

77. Lin DW, Herr HW, Dalbagni G (2003) Value of urethral


wash cytology in the retained male urethra after radical
cystoprostatectomy. J Urol 169, 961-963

61. Nieder AM, Sved PD, Gomez P, Kim SS, Manoharan


M, Soloway MS (2004) Urethral recurrence after
cystoprostatectomy: implications for urinary diversion and
monitoring. Urology 64., 950-954

78. Stewart SB, Leder RA, Inman BA (2010 Aug;) Imaging


tumors of the penis and urethra. Urol Clin North Am 37,
353-367

62. Nelles JL, Konety BR, Saigal C, Pace J, Lai J; Urologic


Diseases in America Project (2008) Urethrectomy following
cystectomy for bladder cancer in men: practice patterns
and impact on survival. J Urol 180, 1933-1936

79. Spiess PE, Kassouf W, Brown G, Highshaw R, Wang X,


Do KA, Kamat AM, Czerniak B, Dinney CP, Grossman HB.
(2006) Immediate versus staged urethrectomy in patients
at high risk of urethral recurrence: is there a benefit to either
approach? Urology 67, 466-471

63. Clark PE, Stein JP, Groshen SG, Miranda G, Cai J,


Lieskovsky G, Skinner DG (2004) The management of
urethral transitional cell carcinoma after radical cystectomy
for invasive bladder cancer. J Urol 172, 1342-1347

80. Stenzl A, Colleselli K, Poisel S, Feichtinger H, Pontasch


H, Bartsch G (1995) Rationale and technique of nerve
sparing radical cystectomy before an orthotopic neobladder
procedure in women. J Urol 154, 2044-2049

64. Yoshida K, Nishiyama H, Kinoshita H, Matsuda T, Ogawa O


(2006) Surgical treatment for urethral recurrence after ileal
neobladder reconstruction in patients with bladder cancer.
BJU Int 98, 1008-1011

81. Stenzl A, Jarolim L, Coloby P, Golia S, Bartsch G, Babjuk


M, Kakizoe T, Robertson C ( 2001) Urethra-sparing
cystectomy and orthotopic urinary diversion in women with
malignant pelvic tumors. Cancer 92, 1864-1871

65. Hassan JM, Cookson MS, Smith JA Jr, Chang SS (2004)


Urethral recurrence in patients following orthotopic urinary
diversion. J Urol 172, 1338-1341
66. Freeman JA, Tarter TA, Esrig D, Stein JP, Elmajian DA, Chen
SC, Groshen S, Lieskovsky G, Skinner DG (1996) Urethral
recurrence in patients with orthotopic ileal neobladders. J
Urol 156, 1615-1619
67. Kassouf W, Spiess PE, Brown GA, Liu P, Grossman HB,
Dinney CP, Kamat AM (2008) Prostatic urethral biopsy
has limited usefulness in counseling patients regarding
final urethral margin status during orthotopic neobladder
reconstruction. J Urol 180., 164-167
68. Donat SM, Wei DC, McGuire MS, Herr HW. (2001) The
efficacy of transurethral biopsy for predicting the long-term
clinical impact of prostatic invasive bladder cancer. J Urol
165, 1580-1584
69. Stein JP, Penson DF, Lee C, Cai J, Miranda G, Skinner
DG (2009) Long-term oncological outcomes in women
undergoing radical cystectomy and orthotopic diversion for
bladder cancer. J Urol 181, 2052-2058
70. Ali-El-Dein B. (2009) Oncological outcome after radical
cystectomy and orthotopic bladder substitution in women.
Eur J Surg Oncol 35, 320-325
71. Ali-el-Dein B, Abdel-Latif M, Ashamallah A, Abdel-Rahim M,
Ghoneim M (2004) Local urethral recurrence after radical
cystectomy and orthotopic bladder substitution in women:
a prospective study. J Urol 171., 275-278
72. Stenzl A, Draxl H, Posch B, Colleselli K, Falk M, Bartsch G
(1995) The risk of urethral tumors in female bladder cancer:
can the urethra be used for orthotopic reconstruction of the
lower urinary tract? J Urol 153, 950-955
73. Stein JP, Penson DF, Wu SD, Skinner DG. (2007)
Pathological guidelines for orthotopic urinary diversion in
women with bladder cancer: a review of the literature. J
Urol 178, 756-760

341

342

Committee 7

Urinary Diversion
Chairs:
Richard E. Hautmann,
Bjoern G. Volkmer

Members:
Hassan Abol-Enein, Thomas Davidsson, Sigurdur Gudjonsson,
Stefan H. Hautmann, Henriette V. Holm, Cheryl T. Lee,
Fredrik Liedberg, Stephan Madersbacher, Murugesan Manoharan,
Wiking Mansson, Robert D. Mills, David F. Penson,
Eila C. Skinner, Raimund Stein, Urs E. Studer,
Joachim W. Thueroff, William H. Turner

Corresponding Author:
Richard E. Hautmann,
Professor of Urology
University of Ulm
Boschstrasse 4 a
89231 Neu-Ulm
Tel.: +0049 731 153 8750
FAX +0049 731 153 8752
e-mail: richard.hautmann@uni-ulm.de

343

Table of Contents
III. TYPES OF URINARY DIVERSION

I. INTRODUCTION

1. ORTHOTOPIC BLADDER SUBSTITUTION


IN MEN

II. GENERAL ASPECTS OF URINARY


DIVERSION

2. ORTHOTOPIC DIVERSION IN FEMALES


3. C
 ONTINENT CUTANEOUS URINARY
DIVERSION

1. URINARY DIVERSION AND RENAL


FUNCTION

4. CONDUIT URINARY DIVERSION

2. SECONDARY TUMORS AFTER URINARY


DIVERSION

5. URETEROSIGMOIDOSTOMY

3. COMPLICATIONS

6. PALLIATIVE URINARY DIVERSION

4. URINARY DIVERSION AFTER PELVIC


IRRADIATION

7. U
 RINARY DIVERSION IN THE PEDIATRIC
AGE GROUPSPECIAL CONSIDERATIONS

IV. RECOMMENDATIONS

5. PREGNANCY AND URINARY DIVERSION

344

Urinary Diversion
Richard E. Hautmann, Bjoern G. Volkmer
Hassan Abol-Enein, Thomas Davidsson, Sigurdur Gudjonsson ,
Stefan H. Hautmann, Henriette V. Holm, Cheryl T. Lee,
Fredrik Liedberg, Stephan Madersbacher, Murugesan Manoharan,
Wiking Mansson, Robert D. Mills, David F. Penson,
Eila C. Skinner, Raimund Stein, Urs E. Studer,
Joachim W. Thueroff, William H. Turner
recommendations given are again of Grade C
only [1]. Grade C recommendation is given when
expert opinion is delivered without a formal analytical
process. In order to check the power of the expert
opinion and the validity of the consensus reached
by the diversion group, each committee member
disclosed the experience, surgical volume, and types
of diversions used at his or her home institution. The
results are presented in Table 1.

I. Introduction
Radical cystectomy and urinary diversion have
been assessed the highest relative value in terms
of difficulty of the surgery for any procedure in urology. They are also the most difficult laparoscopic or
robotic procedures and more so if the diversion is
performed totally intracorporeally. The risk of cystectomy and urinary diversion is based not only on
the technical challenges of the procedure but also
on the nature of the patients need. The incidence
of bladder cancer increases continually with advancing age; thus, the responsibility of providing optimal
surgical treatment for elderly and possibly frail patients is common among urologists. Improvement in
patient rehabilitation is noteworthy through continent
cutaneous diversions and neobladders and better
enterostomal therapy support. In this context, there
must remain continued emphasis on refining the surgical technique of radical cystectomy and urinary diversion to provide utmost safety for the patient.

Our expert opinion is based on almost 16,000


diversions and radical cystectomies performed in 3
continents (Africa, USA, Europe). This committee
represents a well-balanced combination of pioneering
institutions of any type of diversion, high volume
centers and surgeons, as well as data from low
volume institutions, plus a leading pediatric urology
institution and the Swedish registry for bladder
cancer, which reports any case from Sweden,
including treatment, that has been observed in the
respective period.
Some conclusions from Table 1 are:

Only 3/11 institutions have experience with any

The International Consultation on Urological


Diseases (ICUD) has looked at published evidence
and produced recommendations at various levels.
For proper assignment to levels of evidence,
one has to consider study design (prospective,
retrospective), number of patients enrolled, if the
study cohort consists of all patients available or
not, type of assessment tool and its psychometric
properties
(validity/reliability),
and
response
rate. Unfortunately, not a single randomized
controlled study within the field of urinary
diversion exists. Consequently, almost all studies
used in this report are of Level 3 evidence good
quality retrospective studies or case series or
Level 4 evidence including expert opinion based on
first principles research. Therefore, the grades of

type of diversion.

A
 nal diversions play no role in the US, but are of
value in pediatric patients and in the third world.

Continent cutaneous diversions play a secondary

role; even former pioneering institutions (LA) use it


with decreasing frequency.

Conduit (42.2%) and neobladder (38%) are the


standard diversions at large centers.

T
 ruly population-based data from the USA [2] and

from the Swedish Bladder Cancer Registry (S.


Jahnson, Linkping, Sweden) show a neobladder
rate in the range of 15%; with increasing hospital
volume the neobladder/continent diversion rate

345

approaches 75%, addressing the impact of hospital


volume on the use of continent reconstruction and
the question of continent diversion being a new
parameter for measurement of quality of care. Our
group considers these data a further argument for
the centralization of radical cystectomy and urinary
diversion.

hospital case load. We need realistic criteria! Our


committee considers a minimum annual case load
of 25 radical cystectomies done by not more than
two surgeons, to be the basis of a high volume
center.

A
 nother major difficulty the committee had to master

was the lack of standardized reporting in the


urological literature. Recently Donat emphasized
the need for standardized reporting of outcomes
and complications, especially in urological
oncology, because of the mostly elderly population
and the associated comorbidities [3]. Radical
cystectomy and urinary diversion are 2 steps of
one operation. However, the literature uniformly
reports the complications of radical cystectomy
while ignoring that the majority complications are
diversion-related. While this may seem semantic,
it is not [4].

A
 detailed analysis of Table 1 brings us to the key

question for criteria used to define the limits for


low, intermediate, and high annual hospital and
surgical case load. In most publications, one single
radical cystectomy per year was defined as low
annual surgical case load, 2 radical cystectomies
per year are intermediate, and 3 or more radical
cystectomies per year as high. Even though 2
cases per year represent 100% more than one
radical cystectomy per year, from a practical point
of view, these surgeons have a very low case load.
On the other end, a surgeon who performs 3 or 4
radical cystectomies is already defined as a high
volume surgeon. So, the limits appear to be chosen
totally arbitrarily. The same applies to the annual

Until recently, significant disparity in the quality of


surgical complication reporting as well as the lack of
universally accepted reporting guidelines, definitions,

Table 1. Numbers and Types of Urinary Diversions (%) performed by the Authors

Period

Neobladder

Cont.Cut.

643

95-04

45.1

1.4

53.5

962

00-09

40.0

2.0

58.0

0.9

611

85-99

51.5

1.5

42.5

1.6

2.5

0.4

Ann Arbor
Bern

Conduit

UC
Anal
TUUC

No
diversion/ others
unknown

No. RCX

708

00-10

51.0

8.0

39.0

1.5

0.5

0.1

Kassel

765

94-10

30.2

6.8

60.5

0.7

2.0

0.1

Los
Angeles

1359

71-01

51.6

25.8

22.3

0.3

1012

00-10

74.2

5.3

20.2

0.3

119

00-04

28.6

31.1

40.3

134

04-09

6.0

30.6

63.4

335

68-80

55.0

45

593

81-90

2.0

33

41.0

15

982

91-00

6.0

39

41.0

12

1023

01-10

15.0

24

53.0

Mansoura

3157

80-04

39.1

3.5

34.4

23.1

246

02-09

10.6

89.4

158

1997

19.0

19

55.0

221

2003

17.0

12

70.0

208

2006

9.0

80.0

Lund

Mainz

Norwich
Swedish
Registry

229

2008

15.0

81.0

1613

86-09

66.0

0.4

22.0

10.0

1.3

0.2

Vanderbilt

789

00-07

35.5

0.4

63.5

0.1

0.5

15867

38.0

10.4

42.2

1.2

7.5

0.1

0.8

Ulm

Total

346

and grading systems have made it impossible to


compare the surgical morbidity and outcomes in
patients who have undergone cystectomy. There
is a clear case for reporting of complications in a
standardized way. The Clavien system provides
such a straightforward and validated instrument, and
it has already been successfully adopted by several
urological centers [4, 5].

stenosis; stone formation; and reflux of infected


urine.
The pathological changes seen with varying
etiologies are the same, and thus the cause at the
time of detection may not be obvious. As well a
natural age-related decline, estimated to be 1 mL/
min each year over the age of 50 [8], may occur, and
there may be many non-urologic causes for declining
renal function including hypertension, diabetes, and
drugs.

This article discusses the reconstructive options


after radical cystectomy due to bladder cancer or
other conditions, the criteria for selection of the
most appropriate procedure, and the outcomes and
complications associated with the available diversion
options.

Many studies that report on renal function after


urinary diversion are small with limited follow-up
and report retrospectively on serum creatinine with
or without ultrasound or intravenous urography
(IVU) appearance. This is unfortunate as the serum
creatinine remains within normal limits until there has
been approximately a 50% reduction in GFR and the
presence of upper tract dilation does not necessarily
equate to a reduction in renal function.

II. General Aspects of


Urinary Diversion
1. Urinary Diversion and Renal
Function

Ideally, for an accurate assessment of function


we should rely on serial isotopic GFR. 51 Cr-EDTA
is used frequently as it is freely filtered through
glomeruli and minimally reabsorbed from intestinal
segments. Measurement of pre- and postoperative
renal function using radioisotope studies would
provide the best information on the safety of lower
urinary tract reconstruction with regard to the upper
tracts.

a) Preoperatively
Renal function has an important bearing on outcome
following urinary reconstruction as there is an
increase in acid load as a result of absorption of
urinary constituents through the bowel mucosa. The
larger the surface area of bowel and the greater the
contact time with urine, the greater the acid load will
be. Thus, continent reconstruction poses a greater
potential problem than conduit urinary diversion. In
the presence of normal renal function, most patients
with a continent reservoir are able to compensate,
provided the reservoir is emptied to completion at
regular intervals. In the presence of renal impairment,
the ability to cope with the increase in acid load is
reduced, which may result in a metabolic acidosis.

Kristiansson et al reported 10-year follow-up using


51
Cr-EDTA isotopic GFR on patients who were
randomized to an ileal (n=18) or colonic (n=20)
conduit as well as a refluxing or non-refluxing
ureteroileal anastomosis [9]. Thirty-four percent
of patients had a drop in GFR of more than 25%,
40% in the colonic group and 28% in the ileal group.
There was no difference in GFR reduction between
refluxing and non-refluxing ureteroileal anastomoses,
although renal scarring and bacteriuria were more
common with a refluxing anastomosis. Ureteroileal
strictures developed up to 14 years postoperatively.
In these patients, ureteric reimplantation improved
renal function in 6 of 7 renal units, and balloon
dilation improved renal function in 1 of 5 renal units.

There is no exact renal function cutoff below which


continent urinary reconstruction should not be
performed, but, as a general guide, in the presence
of a serum creatinine >150 mol/l or glomerular
filtration rate (GFR) < 50 mL/min, one should avoid a
continent diversion [6].
One proviso is in the presence of obstructive renal
impairment, drainage may be performed to allow
renal function to recover before a final decision is
made. Continent urinary diversion has also been
performed as a staged procedure in preparation for
renal transplantation in patients with end-stage renal
disease [7].

b) Postoperatively

More recently, Samuel and colleagues using serial


isotope GFR have reported follow up of more than 4
years in 178 patients who had an ileal conduit with
freely refluxing ureteroileal anastomoses [10]. Fiftytwo (29%) demonstrated a worsening GFR of more
than 5% (mean 31%) at a mean of 8.2 years. In 33
(18%), there was no identifiable obstruction as a
cause for deteriorating GFR.

There are many potential reasons for deterioration in


renal function following urinary diversion including:
transmission of high pressure to the upper urinary
tract secondary to obstruction, physical or functional,
at any sitein particular ureteroileal and stomal

In a single center non-randomized retrospective


comparison of 275 patients with minimum follow-up
of 12 months (mean 52) who underwent ileal conduits
or orthotopic bladder substitution, Song et al reported
hydronephrosis in 1.4%, 8.3%, and 6.3% following

347

ileal conduit (n=78), orthotopic bladder substitution


with refluxing anastomoses (n=86), and orthotopic
bladder substitution with antirefluxing anastomoses
(n=11), respectively [11]. Chronic renal failure
(defined as creatinine above 3.0 mg/dL) occurred in
7.7% of those in the conduit group compared with
3.5% of those with a refluxing anastomosis and 2.7%
of those without in the orthotopic bladder substitution
groups. The mean preoperative serum creatinine
was significantly higher in the conduit group and
all of those in the conduit group who progressed
to chronic renal failure were thought to have renal
deterioration unrelated to the urinary diversion.
Pyelonephritis was significantly more common in the
orthotopic bladder substitution groups.

They concluded that ureteroileal obstruction was


the leading cause of renal deterioration following
orthotopic bladder substitution and that it is less
prevalent with refluxing anastomoses to a low
pressure afferent limb detubularized ileal reservoir.

e) Ureterocolonic Diversion
As a result of significant complications, the popularity
of ureterosigmoidostomy declined once conduit
urinary diversion became established. Deterioration
in renal function as a result of high pressure reflux
of infected urine was one of the main concerns.
Modifications in technique have made colonic
diversion a much safer procedure; however, its
popularity remains modest.

c) Continent Cutaneous Diversion


Following continent cutaneous diversion, Kristjansson
and colleagues, using 51 Cr-EDTA isotopic GFR,
reported a decrease in renal function of more than
25% total GFR in 28% of 18 patients 11 years after a
cecal reservoir [9].

Conclusions
Urinary diversion into bowel segments is not
inherently damaging to the kidneys. In general, renal
function after diversion into continent detubularized
reservoirs compares favorably with ileal conduit
diversion. However, the literature is insufficient to
recommend one form of diversion over another.
There remains a long-term risk of renal deterioration,
which is often asymptomatic, and thus close followup is necessary for all patients who have undergone
urinary diversion in order to identify correctable
causes early.

In another paper, Jonsson reported follow-up on


126 patients up to 25 years (median 9.5) post Kock
pouch [12]. They recorded a reduction in 51 Cr-EDTA
GFR from 6 years onwards, but this was equivalent
to that expected in relation to aging. It was concluded
that renal function was not affected by the diversion
as long as obstruction was recognized early and
managed appropriately.

Those with renal pathology prior to surgery seem to


be at greatest risk of postoperative renal deterioration.
Serum creatinine is an imprecise measure of renal
function. Isotopic GFR measurement will detect
renal function deterioration most accurately and at
an early stage. The latter, however, is not available
to all patients. In these situations, follow-up with
serum creatinine and ultrasound should be followed
by diuresis renography if upper tract dilation is
seen. Early intervention for physical obstruction
often results in a sustained improvement in renal
function.

Fontaine et al reported over 10 year follow-up of a


heterogeneous group of 53 patients with bladder
exstrophy 37 continent cutaneous diversions and
16 with the urethra as the continence mechanism
[13]. Using 51 Cr-EDTA clearance, they found no
deterioration in renal function in 43 of the 53. In
the remaining 10 (19%), which consisted of 8 with
a Mitrofanoff combined with either ileal or sigmoid
augmentation or pouch, 1 with a Kock pouch, and 1
with an ileocecal patch and reconstruction onto the
native urethra, GFR fell by 20% or more.

d) Orthotopic Bladder Substitution

2. Secondary Tumors after Urinary


Diversion

As with other types of reconstruction, there are


few long-term reports of GFR following orthotopic
reconstruction. Thoeny and colleagues reported a
prospective analysis in 76 patients with a median
follow-up of 84 months (range 60-155) after a low
pressure ileal afferent limb bladder substitute [14].

a) U
 rinary Diversion
Segments

in

Isolated

Bowel

Our review included the patient data presented by


Austen and Klble in their paper published in the
Journal of Urology [15] and in their reply to a letter to
the editor of that journal [16]. Those data had been
obtained through a MEDLINE search up to October
2004. We used the same headings as Austen and
Klble (i.e., urinary diversion and carcinoma), and,
similar to those investigators, we analyzed the data
on latency period, histological findings, tumor location
(ureterointestinal anastomosis vs. distance to the
anastomosis), initial diagnosis leading to urinary
diversion (malignant vs. benign), and treatment of

Preoperative mean serum creatinine +/- 1 SD was 98


+/- 19 mol/L and at 10 years thereafter was 83 +/- 27
mol/L. Twelve (16%) of 76 patients had increased
serum creatinine, 5 of whom had preoperative
dilation or obstruction. In this prospective study,
renal deterioration was seen only in the presence
of pre-existing renal pathology or postoperative
obstruction; however, more precise assessment of
GFR is not available.

348

the secondary tumors. We also included 4 patients


from Lund and Oslo.

20 years ago [19]. All 15 adenocarcinomas appeared


in patients who had undergone surgery due to bladder
exstrophy. Median latency was 37 years (range 1356 years). In 3 patients, conversion to other forms
of diversion had been performed after 8 months,
14 years, and 18 years, respectively, but it was not
reported whether the ureterocolic anastomoses were
removed in those cases.

Table 2 shows the total number of tumors found


in isolated bowel up to August 2010, including the
cases reported by Austen and Klble [15, 16] and 2
cases from Lund. During the period 2004 to August
2010, 15 carcinomas after ureterosigmoidostomy
were described.

It is likely that the etiology of tumor development


after ureterointestinal diversion is multifactorial.
There is evidence of the involvement of increased
levels of N-nitrosamine and/or free oxygen radicals
in response to infection and chronic inflammation.

Until recently, all risk calculations were severely


hampered by the absence of easily accessible hard
data on the total number of procedures performed. As
a courtesy, Klble et al. provided some information
of the first multicenter study that allows us to relate
the number of secondary tumors with the number of
various types of urinary diversions done [18].

Conclusions
Patients who have undergone conduit diversion,
continent cutaneous diversion, or orthotopic bladder
substitution do not seem to be at increased risk of
secondary malignancy. By comparison, the risk is
slightly higher after cystoplasty, albeit not increased
enough to support endoscopic surveillance.
However, the present knowledge regarding gastric
cystoplasty is insufficient, and hence patients should
be followed after such surgery. Furthermore, yearly
colonoscopy is recommended in cases involving
ureterosigmoidostomy, beginning 10 years after the
procedure.

They analyzed the operative records of 44 German


hospitals for urinary diversions performed from
1970-2007 and registered all reported tumors till
2009. In 17,758 urinary diversions, 32 secondary
tumors occurred (Table 3). The tumor risk in
ureterosigmoidostomy (2.58%) and cystoplasty
(1.58%) was significantly higher than in other
continent forms or urinary diversion (p<0.0001).
The risk in orthotopic (ileo)colonic neobladders
(0.97%) was significantly higher (p=0.0001) than in
ileal neobladders (0.05%). The difference between
ileocecal pouches (0.18%) and ileal neobladders
was not significant (p=0.46), and the tumor risk with
ileal conduits was minimal (0.02%), (Table 4).

3. Complications
Any critical analysis and comparison of urinary
diversion technique centers around the questions:
Is there a perfect solution? or: Which diversion
technique is the best? or: Which one has the lowest
complication rate? At this point our committee must
disappoint the readers of this report and rather
explain why it is absolutely inappropriate to ask
these questions.

b) Ureterosigmoidostomy
Since 2004, there have been reports of 15
adenocarcinomas
and
8
adenomas
after
ureterosigmoidostomy (a reference list can be
obtained from the present authors), and descriptions
of more than 200 cases had already been published

Table 2. Secondary tumors after diversion using isolated gut segments


Ileum

Colon

Ileocecal

Stomach

Total no.

15+1*

22

Continent cutaneous

14**

14

Neobladder

Rectal bladder

Cystoplasty

41

15

10

68

Total

60

42

10

114

Conduit

From publications by Austen and Klble (15,16), since 2004, and 4 Lund cases.
* This case was transferred from continent cutaneous diversion using colonic segment.
**Includes 2 Lund cases (described in this paper) but excludes a case (17) from Austen and
Klble, that is now transferred to the ileal conduit group.

349

Table 3. Numbers of Secondary Tumors in 17,758 Urinary Diversions


no.
secondary
tumors

no. diversions
%

median
latency period
(years)

Neobladder Ileal

4190

(0.05 %)

3.0

6.3

Neobladder Ileocoecal

239

(1.26 %)

4.0

13.6

Neobladder Colonic

70

(1.43 %)

6.0

9.9

Pouch Ileocoecal

2181

(0.18 %)

12.0

8.8

Ileocystoplasty

233

(1.71 %)

21.5

10.5

Colocystoplasty

20

Conduit Ileal

8637

(0.02 %)

11.0

8.9

Conduit Colon

430

(0.23 %)

40.0

24.4

Ureterocutaneostomy

1138

Ureterosigmoidostomy

16

620

(2.58 %)

32

17,758

Total

mean time
operation to
analysis
(years)

4.9

10.3

26.0

18.9
Klble 2011

Table 4. Tumor Risk in Continent and Incontinent Urinary Diversion


Ileum

(ileo-)colon

total

Conduit

0.02%

0.23%

0.03%

p=0.84

p=0.0091

Pouch/Neobladder

0.05%

0.28%

0.13%

Total

0.03%

p=0.00001

0.08%

p<0.0001
Cystoplasty

0.27%

1.71%

0.00%

1.58%

2.58%

p=0.46
Ureterosigmoidostomy

Klble 2011

350

Muscle-invasive bladder cancer occurs in a

data on long-term complications that consider


their time-dependency will be comparable.
Otherwise, complications that occur early are
overrepresented, while those occurring late might
be underrepresented, since the number of patients
will decrease over time [20].

comorbid population with a mean age of 65-69


years in contemporary radical cystectomy series.
The incidence of early complications, defined as
occurring either during the hospital stay or within 90
days of surgery, has been reported in the range of
20-57% [5]. Therefore, accounting for the impact of
surgical morbidity on patient outcome is essential
for treatment planning, for clinical trial design, and
for assessing new surgical techniques.

An ideal methodology for reporting adverse events


related to surgical therapy should include 10
established basic reporting criteria [5]: (1) a clear
description of the method of data acquisition, (2)
an indication of the duration of follow-up, (3) an
indication of whether or not outpatient complication
data are included, (4) definitions/inclusion criteria
of at least one complication, (5) an indication of the
mortality rate and cause of death, (6) an indication
of the morbidity rate (number of patients and the
total number of complications recorded), (7) an
indication of procedure-specific complications, (8)
the utilization of a grading system to clarify severity
of complications, (9) an indication of the median
or mean length of stay, and (10) an indication of
the methodology utilized to assess patient risk
stratification (e.g., Charlson-Romano index, ASA
scoring).

A recent evaluation of the urologic oncology


literature revealed that the majority of series that
reported on radical cystectomy morbidity had not
employed a formal complication reporting system,
had not utilized grading systems other than
categorizing them into major versus minor, had
not accounted for comorbidities, or had not defined
complications. This makes it impossible to compare
data and most certainly leads to an underestimation
of morbidity for the procedure. In addition, the
incidence of perioperative complications have
often been utilized as surrogate measures of
surgical competency, institutional quality of care,
success of new surgical techniques, and have
even been suggested as benchmarks for financial
reimbursement [5].

Recently, efforts have been made to standardize


the reporting of early complications. The Clavien
system in the Memorial Sloan-Kettering Cancer
Center modification has proven to be a useful tool to
classify these complications according to domains
(e.g., genitourinary, gastrointestinal, infectious)
and grades [5]. Unfortunately, this system is less
suitable for reporting long-term complications [20].

T
 he difficulty in comparing complications across

series is manifold. The lack of standardization is


hampering the progress of reducing morbidity and
complications associated with radical cystectomy.
Radical cystectomy and urinary diversion are two
steps of one operation. However, the literature
notoriously reports on complications of radical
cystectomy, ignoring that the vast majority of
complications are diversion-related. This may
seem semantic, but it is not [4]. This is of particular
importance for this consensus report.

The recent literature on urinary diversion shows a


trend to the increasing use of standardized reporting
of complications in urologic oncology, in particular of
complications of urinary diversion (4, 5, 20]. Table
5 presents a representative example of adequate
reporting: It is the summary of neobladder-specific
early ( 90 days) complications in 1013 patients
with neobladders. Table 6 presents an attempt
to compare diversion-related complications in a
realistic way [20, 21, 22].

Until recently, any publication on cystectomy and

urinary diversion complication rates has reported


the rates by simply dividing the absolute number
of events by the total number of patients treated,
irrespective of the time to the complication onset.
Early complications occur in a short and well-defined
interval of 90 days postoperatively. Stratification of
complication rates to the first, second, or third month
postoperatively is not meaningful. In contrast, longterm complications may occur decades later. Since
most complications develop time-dependently,
they should be reported using the Kaplan-Meier
method for time-dependent events in analogy to
reporting tumor recurrence. This can easily be
explained by the example incisional hernias.
We observed 41 hernias in 923 patients (4.4%).
In contrast, the Kaplan-Meier method revealed a
rate of 6.4% at 10 yearsa 45% higher rate than
that published usually. Standardization of reporting
long-term complications does not only impact on
definitions but also an statistical analysis. Only

Conclusions
Surgical morbidity following urinary diversion is
significant and, when strict reporting guidelines
are incorporated, higher than previously published.
Accurate reporting of postoperative complications
after radical cystectomy is essential for counseling
patients, combined modality treatment planning,
clinical trial design, and assessment of surgical
success. Specific diversion related complications,
i.e., neobladder-vaginal fistula or hypercontinence
following orthotopic bladder substitution in women
are reported in the respective sections.

351

Table 5. Neobladder specific complications within 90 days of surgery by category in 1013 patients with
neobladders
no
complication

category

Grade
1

Grade
2

Grade
3

Grade
4

Grade
5

complications

38

146

35

12

236 (23.3%)

Total

infection

777 (76.7%)

abscess

988

20

25

UTI

837

38

138

176

sepsis

978

15

12

35

genitourinary complications

843 (83.2%)

100

64

170 (16.8%)

renal failure

995

11

18

ureteral obstruction

965

16

31

48

urinary leak

948

44

21

65

urinary fistula

1002

11

urinary retention

985

23

28

miscellaneous

918 (90.6%)

95 (9.4%)

metabolic acidosis despite treatment


surgical

1007

985 (97.2%)

28 (2.8%)

1010

incisional hernia

Grade 1: oral medications


Grade 2: intravenous medications
Grade 3: interventional radiology or operation
Grade 4: lasting disability or organ resection
Grade 5: death

Hau J Urol, 2010

Table 6. Diversion-specific (not cystectomy) long term complications * using non-standardized reporting

f.u. median
patients

Conduit
Bern

Neobladder
Ulm

Mayo

(years):

8.1

6.3

6.0

(n=)

131

1,057

923

complications

(n=)

192

1453

---

patients with

(n=):

87/131

643/1057

376/923

patients with

(%):

66.0

61.0

40.8

within years:

5:

45

50

10:

50

68

15:

54

70

Reoperation rate (%):

20:

94
40

80
6

Complications (pts/%):

bowel

32

(24%)

215

(20.3%)

31

(3.4%)
(5.0%)

UTI

30

(23%)

174

(16.5%)

46

stoma

32

(24%)

163

(15.2%)

anastomosis

18

(14%)

122

(11.5%)

102

(11.5%)
(0.2%)

urolithiasis

12

(9%)

162

(15.3%)

renal function

35

(27%)

213

(20.2%)

* Bern :
Mayo:
Ulm:

> complications 3 months ; > 5 years follow-up

Madersbacher J Urol, 2003

> 30 days of surgery

Shimko

J Urol, 2011

> 3 months of surgery

Hautmann

J Urol, 2011

352

4. Urinary Diversion after Pelvic


Irradiation

another method to increase or boost the dose to the


prostate. While such dose escalation may improve
local control of the prostate cancer, any salvage
surgery may become that much more difficult.
Earlier detection of treatment failure using PSA may
increase the pool of patients potentially suitable for
salvage surgery, but it is uncertain if earlier and more
aggressive surgery will translate into better results
with acceptable morbidity. Bladder preservation
strategies will also provide a greater challenge for
salvage surgery. Fortunately, the dose of external
radiation therapy to the pelvis for bladder cancer
is less than that delivered for prostate cancer ( 60
Gy).

Tumor recurrence of a urologic, gynecologic or


gastrointestinal tumor or a defunctionalized bladder in
patients who had received definitive radiation therapy
may be followed by salvage radical cystectomy.
Historically, these patients have been considered to
have a risk of significant postoperative morbidity and
unsatisfactory functional results. Complications have
been attributed to radiation damage to the ureter and
bowel, resulting in increased rates of anastomotic
problems, upper urinary tract obstruction, and
infection. Therefore, most centers use supravesical
urinary diversion with a transverse colonic segment
[23-29) or cutaneous ureterostomy.

During pelvic radiotherapy for cancer, the cecal pole


as well as parts of the ascending colon, appendix,
and ileum are exposed to considerable doses of
radiation. Since these segments of bowel are used
for reservoir and afferent segment construction, it
is plausible that the high complication rate that was
observed was secondary to radiation damage of the
intestinal segments.

In a series of patients with urinary diversion after


pelvic irradiation [30], 58.5% of these patients
experienced one or more complications occurring
within a 90 day period from the date of surgery.
Seventy-seven percent
of the
patients with
neobladders versus 52.2% of the patients without
neobladders had complications. The rate of grade
3-5 (major) complications, according to the modified
Clavien system of the Memorial Sloan-Kettering
Cancer Center, was 28% in patients with neobladders
and 31.9% in patients without neobladders. The
majority of these complications were classified
as genitourinary or infection-related. In-hospital
mortality was 0% for patients with neobladders and
11.6% for patients without neobladders . In 10 of 25
patients with orthotopic reconstruction, serious and
highly unusual late complications were observed,
e.g., spontaneous ileal perforation, ileal stenosis,
stenosis of the descending colon, spontaneous
neobladder perforation, and neobladder-vaginal
fistula.

Limited data are available on the outcome of


continent urinary diversion in patients with previous
pelvic irradiation [23, 24, 26, 31].
Bochner et al [32] reported the outcome of
orthotopic ileal urinary diversion during salvage
cystoprostatectomy after failed radiation for
bladder and prostate cancers. They noted a total
of 6 complications (33%) in 18 patients, including
prolonged urinary leakage in 2, and afferent nipple
stenosis, ureteral stenosis, enteropouch fistula,
and urinary retention in 1 each. The reoperation
rate during the mean 28 month follow-up was 17%.
Of the patients, 33% were incontinent during the
day and 44% were incontinent at night. Gheiler et
al [33] reported a similar experience. Brand [34]
described cecal rupture after Indiana ileocecal
continent diversion as a postoperative complication,
underlining the vulnerability of the irradiated cecal
region. Other than this report, all published studies
of the results of continent cutaneous and orthotopic
urinary diversion in irradiated patients have an
average follow-up of only 1-2 years. Nevertheless,
alarmingly high open reoperation rates of up to 25%
have been reported during this limited follow-up.
Comparing results in early publications [35, 36] with
subsequent data published many years later by the
same groups clearly demonstrates that conclusions
based on small series with limited follow-up may be
misleading.

This series confirms the high morbidity of radical


cystectomy in a comorbid population with
experienced surgeons in a high volume institution.
Salvage treatment after failure of the initial therapy
for localized cancer is always more difficult. These
patients are an adversely selected group because
the first attempt of curative treatment had failed.
Judgment is required to decide who should have a
surgical resection after previous radiation therapy,
and then technical expertise is needed to do the
operation with acceptable success. This series adds
to the evidence that salvage surgery for bladder
and prostate cancer, as well as gastrointestinal and
gynecologic cancers, can be done safely. Overall
complication rates are higher than those noted in
patients without previous radiation therapy [30].
In the future, the problems with salvage are likely
to increase. Dose escalation for prostate cancer
radiation therapy is a potential reason. Currently,
the dose to the prostate with conformal radiation
therapy is 74-78 Gy. Implant radiation, either by
itself or combined with external radiation therapy, is

Since radiation damage is historically known to


increase with time, these aspects are especially
important when evaluating complications of surgical
procedures in irradiated patients. Proper patient
selection for salvage surgery has also contributed
to the improvements in long-term outcome [37].
Selection of appropriate surgical candidates for

353

salvage therapy depends on several factors:


recurrent prostate or bladder cancer versus
gastrointestinal or gynecologic cancer, extent of
recurrent disease, and existence of fistula formation.
It is believed that patients with more advanced local
disease, refractory voiding symptoms related to a
fibrosed non-functional bladder, or severe symptoms
related to other complications associated with the
prior irradiation will be better served with cystectomy
and lower urinary tract reconstruction. Based on the
published long-term experience, salvage surgery
(cystoprostatectomy, anterior exenteration) with
orthotopic lower urinary tract reconstruction is a
safe, effective procedure that can provide a potential
curative intervention and a functional lower urinary
tract for properly selected patients in whom previous
definitive radiation therapy has failed, with results
only marginally worse as compared to non-irradiated
patients, at least in high volume centers [38].

diversion [39].
There are no contraindications against a Cesarean
section.
Cesarean section
cautiously in:

should

be

performed

very

patients with intraperitoneal hydrocephalus shunts


patients with pouches, enterocystoplasties, or
neobladders

c) Special Aspects
1. W
 hich form of urinary diversion should be
recommended in young patients who plan to have
children?

Our review of the literature shows a successful


course of pregnancy is possible with all types of
urinary diversion. We therefore strongly recommend
that the choice of the type of urinary diversion should
consider first the individual situation and needs of
the patient, but with secondary respect to a potential
pregnancy.

5. Pregnancy and Urinary Diversion


Fertility-sparing cystectomy is restricted to children or
premenopausal women with benign diseases. Once
the problems associated with urinary incontinence
have been solved and the patients have reached
puberty, sexuality and fertility become more
significant. At this time the medical responsibility
shifts from the pediatrician and pediatric urologist to
the gynecologist and general practitioner. However,
these colleagues rarely have major experience with
the psychological aspects of sexuality, fertility, and
pregnancy in these patients. Meanwhile, all functions
of female sexuality can be preserved, if a fertilitysparing cystectomy and urinary diversionincluding
a neobladderis performed. Consequently, the
patients desire for a normal sexual life has to be
respected.

2. Is simultaneous sterilization an option?


In the twentieth century, the majority of patients with
urinary diversion had simultaneous sterilization. The
first patient who became pregnant following urinary
diversion did so despite previous sterilization.
Nowadays, sterilization of patients undergoing
urinary diversion is not performed routinely, since
pregnancy is not considered contraindicated in
patients with urinary diversion.

3. Are there medical indications for an


interruption?
In 1972, Olesen recommended a therapeutic
abortion if the renal function was abnormal when
pregnancy was diagnosed [40]. Since the difficulties
of saving the renal function during pregnancy
have been ameliorated, we think that this general
recommendation should be obsolete in 2011.

a) Childbirth
In all patients with urinary diversion, the form of
delivery must be planned considering many relevant
aspects: potential damages to the urinary diversion
by the Cesarean section, potential damages to the
pelvic floor by vaginal delivery, special anatomic
problems of the mother, and, last but not least, the
health of the child.

4. W
 hat examinations are recommended before,
during and following pregnancy?
Richmond et al suggest that whenever possible,
counseling of patients with myelomeningocele
should be done before conception, when radiological
pelvimetry and renal function assessment could be
carried out without risk to the fetus in utero [41].

Vaginal delivery should be performed with caution


in:

patients with ureterosigmoidostomy


patients with malpresentation

All pregnant women with any type of urinary


diversion should be assessed (ideally antepartum)
by an multidisciplinary team of experts, involving
the obstetrician (perinatologist), urologist, obstetric
anesthesiologist, general surgeon, neonatologist,
and nursing staff who work together to develop
(tailor) the optimal peripartum plan on a case-bycase basis [42].

patients with cervical prolapse


b) Cesarean section:
Hensle et al consider the elective scheduling of a
Cesarean delivery before the onset of labor the
preferred method in patients with urinary diversion,
in particular in those with orthotopic continent urinary

354

patients with ureteral reflux,

Kennedy et al recommend an initial screening


maternal ultrasound examination at 20 weeks of
gestation to look for maternal hydronephrosis.
Follow-up maternal renal ultrasound examinations
should be performed if hydronephrosis is noted
initially or for recurrent episodes of pyelonephritis
[43].

patients with impaired renal function,


patients requiring catheterization, and patients with
ureterosigmoidostomy.

7. Pregnancy following urinary diversion and


renal transplantation

In Germany, the patient is seen by the obstetrician


every 4 weeks during the first 4 months of gestation,
every 3 weeks during the next 3 months, every 2
weeks during the next 2 months, and then in weekly
intervals until delivery. We strongly recommend
that the patient be seen by a urologist at the same
intervals until delivery. The urologist should perform
urinalysis, microbiologic controls of the urine, renal
ultrasound (including estimation of the resistive
index, when hydronephrosis is present), and controls
of the storage function and emptying of the urinary
diversion.

The first successful pregnancy after renal


transplantation ended in 1958 with the delivery
of a healthy child [45]. Since that time, a growing
number of reports on successful pregnancies have
been published. Problems like immunosuppression,
impaired renal function, and obstructive uropathy
may lead to a much less favorable prognosis as for
pregnancy in patients with urinary diversion alone.
Until today, only 3 pregnancies in women with a kidney
transplanted to an ileal conduit were reported. In 2
cases, pregnancy had to be terminated in the tenth
and twenty-second weeks, respectively, because of
progressing renal failure from compression of the
ileal loop [46]. The only successful pregnancy was
published by Hein et al [47]. Their patient had a
neurogenic bladder caused by an occult spina bifida.
At the age of 14, an ileal conduit was created. Nine
years later, the patient suffered from progressive
renal failure caused by chronic pyelonephritis. She
underwent bilateral nephrectomy. At the age of 26
years, a living donor kidney transplantation was
performed. Four months later, the patient became
pregnant. In the thirty-fifth week, an intermittent
obstruction of the left ureter was noted, leading to the
decision of terminating the pregnancy. Delivery of a
healthy male was accomplished by a low segment
Cesarean section because of a narrow pelvic cavity.
The postoperative course was uneventful.

5. What information on reconstructive urology


does an obstetrician need?
The obstetrician needs complete information on
the anatomic situation of the patient with urinary
diversion. The knowledge of all reports on surgical
procedures performed in the particular patient in
the past is compulsory. Precise information on the
fixation points of the urinary diversion and of the
course of the ureters is necessary. Urological input
at the time of surgery helps identification of the lower
urinary tract anatomy, prompt recognition of any
inadvertent urinary tract injury, and assistance in the
repair of any damage.

6. Is long-term antibiotic prophylaxis required


during pregnancy with a urinary diversion?

8. Pregnancy tests in patients with urinary

Symptomatic urinary tract infection is a problem


encountered in a large proportion of patients with
urinary diversion. Contributing factors to urinary
tract infections are urinary stasis, difficulty with
clean intermittent catheterization, and ureteric
compression.

diversions

In a recent study, Nethercliffe and coworkers


analyzed urine of patients with urinary diversion
using the Clearview pregnancy test (Unipath
Ltd. Bedford, England) [48]. Although all of these
patients were not pregnant, positive pregnancy tests
were found in 8 women and 5 men (false-positive
rate: 57%). The authors suggest that the mucus
produced in enterocystoplasties may interfere with
the pregnancy test. Patients and doctors have to be
aware of this [49] .

Twenty-one percent of the patients reviewed had


premature labor, presumably related to the high
incidence of lower urinary tract infections and acute
episodes of pyelonephritis. The authors advocated
aggressive antibiotic therapy for all urinary tract
infections in this group of patients [44]. Therefore,
Hensle et al considered it advisable to keep all these
women on antibiotic prophylaxis through pregnancy
[39].

III. Types of Urinary Diversion

We think that a low-dose antibiotic prophylaxis is at


least advisable in the following patients:

1. Orthotopic Bladder Substitution


in Men

patients with a history of febrile UTI during

a) Indications and Tumor Stage

patients with hydronephrosis,

The indication for cystectomy is almost always


bladder cancer. The extent of pelvic disease has little

pregnancy,

355

bearing on the appropriateness of orthotopic bladder


substitution. If pelvic recurrence does develop, it
does not usually have a significant impact on the
function of an orthotopic bladder substitute, and
patients who have positive pelvic nodes can achieve
good functional results with orthotopic bladder
substitution [50, 51].

chance of maintaining erectile function. In men with


an orthotopic bladder substitute, who had attempted
nerve-sparing, nighttime continence was better
than in those who did not [54]. The rate at which
nighttime continence was achieved and the final
nighttime continence rate were both improved. This
results from preservation of the afferent intrapelvic
anatomic sensory nerve supply to the pelvic floor
and external urethral sphincter [55] with preserved
urethral sensation. An impaired urethral sensation
is associated with poorer continence [56]. Nerve
sparing also preserves the efferent autonomic
innervation to the sphincteric smooth muscle, which
may mediate resting pressure generation (tonus).

The risk of urethral recurrence after orthotopic


bladder substitution is generally 5-10% and it usually
occurs in the first 3 years when it does happen. A
high risk is a contraindication to orthotopic bladder
substitution, but prediction of risk is not simple.
Reported risk factors for urethral recurrence include
multifocal disease, carcinoma in situ, prior intravesical
chemotherapy, ureteric disease, and urothelial
cancer in the distal prostatic urethra. Furthermore,
prostatic urothelial carcinoma is common. If whole
mount pathologic assessment of the prostate is done
after cystoprostatectomy, detection rates are nearly
50% [52, 53]. Most patients with prostatic disease,
however, do not develop urethral recurrence. So,
unless there is obvious disease in the prostatic
urethra, we perform distal prostatic urethral biopsies
before surgery and then construct an orthotopic
bladder substitute if these are negative. Paraffinembedded biopsies are more reliable than an
intraoperative frozen section. Nevertheless, an
intraoperative frozen section of the resection margin
is considered sufficient by many centers. Biopsies
done before cystectomy also enable discussion of
the result with the patient, who then has greater
certainty that orthotopic bladder substitution will be
possible at the time of cystectomy.

If tumor characteristics permit, then nerve sparing


should be attempted. This can be bilateral if disease
is not muscle-invasive, or unilateral if there is
lateralized muscle-invasive disease. In men, the
nerves are at particular risk dorsolateral to the
seminal vesicles, in the vesicoprostatic angle, and in
the region of the prostatic apex.

e) Reservoir Configuration
An orthotopic bladder substitute must be a lowpressure reservoir of adequate capacity (allowing a
socially acceptable voiding interval without urinary
leakage) and must empty to completion. If this is
so, the upper urinary tracts will be preserved, and
metabolic disturbance will be minimal. Many surgical
techniques have been reported but some key factors
are alike. Detubularization and a spherical shape
ensure that an orthotopic bladder substitute has low
pressure and maximum volume for the length of bowel
used. A spherical reservoir has 4 times the capacity
and a quarter of the pressure compared to a cylinder
made from the same length of bowel. The use of
less bowel reduces the reservoirs surface area and
therefore the chance of absorptive complications,
but larger reservoirs have lower end-filling pressures
[57] and better continence, particularly in the early
postoperative period. This advantage is short-term,
though, and the functional capacity of an orthotopic
bladder substitution that is made from a shorter
segment increases rapidly from 150 mL to 400-500
mL [58]. The use of excessive length of bowel may
produce low-pressure floppy reservoirs that empty
less effectively. Larger reservoirs have a greater
chance of rupture, because when they are allowed
to fill to capacity, the wall tension at a given pressure
is greater. The use of 40-45 cm of ileum is sufficient
for the reservoir itself. Popular techniques include
an ileal afferent limb orthotopic bladder substitute
using 55 cm of distal ileum, preserving the 25 cm of
terminal ileum, and the W-shaped bladder substitute
[59, 60].

b) Age and Motivation


Although there is no age cut-off for orthotopic bladder
substitution, in practice many patients over the age
of 70 years will opt for a simpler conduit urinary
diversion as the postoperative course is less arduous
and urinary incontinence is less likely. The motivation
of the patient is probably the most important factor
when considering their suitability for an orthotopic
bladder substitute, although it is difficult to assess
this objectively. Patients must be prepared to commit
to the long-term follow-up program necessary. There
is no point in the surgeon being enthusiastic if the
patient is not; this will result in poor function and the
risk of complications.

c) Sphincter Function
Urinary continence after orthotopic bladder
substitution depends, amongst other factors, on
adequate urethral sphincter function. Caution should
be exercised before offering orthotopic bladder
substitution in patients with significant urethral
strictures.

f) Minimally-invasive Surgery

d) Surgical Technique

There is increasing interest in laparoscopic and


robotic cystectomy, with either intracorporeal or

Nerve-sparing cystectomy does not just increase the

356

extracorporeal formation of ileal conduit or the


orthotopic bladder substitute [61-64]. Whether
reports with intermediate follow-up suggesting
equivalent pathologic and oncologic outcome will be
confirmed remains to be determined [65]. But there
are advantages in terms of blood loss, transfusion
rate, postoperative pain, and return of bowel function
[64]. In most reported series, cases have been highly
selected; it remains to be seen what the limitations
of minimally-invasive surgery might be in terms of
tumor size and bladder mobility. With regard to the
reconstruction, although there are a number of reports
of intracorporeal surgery, most surgeons currently
prefer to perform an extracorporeal technique as this
is faster. However, to do this, the ureters must be
kept long and there is concern about an increased
risk of ureteroileal stenosis [61]. Once a reservoir
has been constructed the urethral anastomosis
may be carried out by an open technique or by reestablishing a pneumoperitoneum. The technology
will probably need to improve before intracorporeal
reconstruction becomes standard practice.

delaying voiding when the patient feels that the urge


to void is irresistible and that leakage may occur. A
rapid increase in reservoir capacity following surgery
allows daytime continence to be achieved. Nighttime
continence is established less quickly. During sleep,
a detrusor-sphincter reflex normally increases outlet
pressure as the bladder wall stretches during filling;
this reflex is lost after cystectomy. As the reservoir
fills at night, therefore, additional outlet contraction
is not recruited, and when the rise in reservoir
pressure exceeds outlet pressure, leakage occurs.
This tendency to leak is normally compounded
because vasopressin is secreted at night, and urine
is therefore concentrated. However, the intestinal
wall of the bladder substitute will secrete water
into the reservoir, to try to render its contents isoosmolar, and so overnight urine output is greater
after orthotopic bladder substitution than before
cystectomy [68].
Men achieve self-assessed continence by day and
by night in 92% and 76% of cases, respectively.
Attempted
nerve-sparing
improves
daytime
continence, and increasing age worsens it [54]. Men
with type II diabetes gained daytime continence
more slowly than controls and were less likely to
achieve nighttime continence. Urinary tract infection
also worsens continence [69], and an unexpected
deterioration in continence should prompt exclusion
of infection and residual urine. Urine infection should
be uncommon.

g) Functional Aspects
1. Voiding
The mechanism of voiding for patients with an
orthotopic bladder substitute is well-described [57,66]
in case of an ileal reservoir (not cecal!). A passive
pressure rise occurs in the reservoir during filling,
and this stretches the reservoir wall, activating the
intestinal stretch receptors, resulting in a sensation
of filling. The patient learns this new sensation
during postoperative voiding rehabilitation. For an
ileal reservoir, voiding occurs by reducing outlet
pressure and pelvic floor and sphincter relaxation,
perhaps combined with slight straining. If straining
occurs, the upper urinary tracts and the reservoir
are equally subjected to the resulting abdominal
pressure rise, and so no pressure gradient is created
from the reservoir towards the upper tracts. In the
early postoperative period, men should sit to void, to
help them to learn how to empty the reservoir. Late
voiding difficulty occurs in 20% of men [67] within the
first 10 postoperative years for a variety of causes,
but almost all of these men void normally after
endoscopic assessment and relief of the underlying
bladder outlet obstruction.

3. Upper Urinary Tract Preservation


Voiding with an orthotopic bladder substitute
cannot produce reflux; this has been confirmed
scintigraphically [70]. Long-term upper tract
outcomes are excellent: as few as 2.7% of patients
should develop ureteroileal stricture, if a direct endto-side ureteroileal anastomosis is used [71, 72].
The use of stents in ureteroileal anastomosis after
orthotopic bladder substitution has been shown
to improve outcomes [73]. Up to half of patients
who develop a short ureteroileal stricture can be
successfully managed endourologically [72].
These data are consistent with the outcome of a
randomized surgical trail done to determine the effect
of different afferent mechanisms with an orthotopic
bladder substitute. This showed clearly that the use
of an antireflux nipple valve was associated with a
worse outcome than a dynamic isoperistaltic afferent
tubular ileal segment [74]. This was confirmed in
another randomized study recently [75]. In the longterm, the upper tracts are well preserved, and upper
tract calculi are rare [76].

2. Continence
Continence is achieved when outlet pressure
exceeds reservoir pressure. This requires careful
preservation of the urethral sphincter and the
formation of a low pressure reservoir, by doubly
cross-folding detubularized ileum (see above) to
achieve the desired reservoir volume of 400-500 mL.
The reservoir is constructed with an initial volume
of around 150 mL, and in the first postoperative
weeks, the patient has to work actively to stretch
the reservoir. Stretching the reservoir is done by

Table 7 presents specific long-term complications


when standardized reporting is used [20].

4. Postoperative Management
Of paramount importance is the active postoperative

357

management and regular long-term follow-up of


patients with an orthotopic bladder substitute. The
key issues [77] are achieving a capacity of 400-500
mL, residual free voiding of sterile urine, and the
treatment of any outlet obstruction.

are at higher risk of incontinence but some of them


will have excellent neobladder function (unpublished
data). Medical comorbidities, cardiac and pulmonary
function, and cognitive function are all important
factors that should be considered, along with the
patients social support and patient preference.

2. Orthotopic Diversion in Females


a) Patient Selection: Patient-related Factors

b) Patient Selection: Oncologic Factors

There are both patient-related and cancer-related


factors influencing the appropriateness of continent
orthotopic diversion in women (Table 8). The
patient-related factors are generally based on the
recommendations of surgeons performing these
diversions rather than on evidence-based studies.
A number of the contraindications to continent
diversion are identical for men and women. These
include the basic requirements of adequate renal
function, available healthy bowel, and a functional
urethral sphincter. Preexisting incontinence is a
relative contraindication for women considering a
neobladder. Some bladder cancer patients experience
urge incontinence due to the tumor itself, which
may be expected to improve after the cystectomy.
A woman with stress urinary incontinence may be
willing to continue to wear pads rather than deal with
a stoma, or may be considered for a sling or Burch
procedure at the time of diversion with planned selfcatheterization.

Prior to the wide adoption of orthotopic diversion for


females, it was necessary to show that it was safe
to preserve the urethra during the cystectomy. De
Paepe et al initially reviewed 22 female cystectomy
specimens and found urethral involvement in
36% [80]. Coloby and colleagues evaluated 47
consecutive cystectomy specimens with step
sectioning through the urethra, finding only 7% with
urethral involvement, all of whom also had bladder
neck involvement [81]. Stenzl had similar findings in
a review of 356 specimens with localized invasive
cancer, as did Marlani and Chen [82-84]. Finally,
Stein and colleagues evaluated 67 specimens with
13% urethral involvement. Again only bladder neck
involvement and anterior vaginal wall invasion
predicted urethral involvement, though 50% of those
with bladder neck tumor had no tumor in the urethra
[85]. In most of these studies, tumor involvement of
the trigone, the presence of carcinoma in situ (CIS),
or multifocal primary tumors did not predict urethral
recurrence.

Age alone is not a criterion for offering continent


diversion [78, 79]. The impact of age on outcomes
with orthotopic diversion has not been fully examined
in females. In our experience, women over age 75

It is now standard to require a negative frozen


section of the urethral margin prior to proceeding
with neobladder construction in women [78, 86, 87,
88, 89, 115]).

Table 7. Diversion specific long-term complications (> 3 months post-op) at 10 years in 923
patients with neobladder operated between
1986-2008 using the Kaplan-Meier method

Table 8. Contraindications For Orthotopic


Diversion In Women

Absolute contraindications:

R enal insufficiency (estimated creatinine


clearance of < 35-40)

ureteroileal stenosis13.80%

refluxing anastomosis (Wallace) 6.40%


non-refluxing (Le Duc)17.00%
urinary retention (female) 42.20%
subneovesical obstruction in men 8.20%
(correctable)
functional obstruction in men 5.10%
(clean intermittent catheterization required)
incisional hernia 6.40%
(41 in 923 patients = 4.4%)
Hautmann, J Urol 2011

Cancer invading the anterior vagina

Positive frozen section at the urethral margin

Inadequate available bowel

Relative contraindications:

Invasive cancer at the bladder neckon TURBT


S evere co-morbidities requiring very short
operative time

Significant stress incontinence

L ocally extensive disease at surgery (stage


T4)

Prior pelvic radiation


358

preservation [95]. However, others have routinely


dissected the presacral tissue as part of the node
dissection without a clear reduction of continence
[97, 101], To date, no randomized comparison of
these 2 techniques has been performed.

In summary, it is reasonable to advise against


neobladder reconstruction for a woman with invasive
bladder neck involvement or suspected invasion of
the vaginal wall or cervix. However, such patients
may be considered for neobladder diversion if
intraoperative frozen section of the urethral margin
is negative. It appears that overall 60-70% of
women undergoing cystectomy might be reasonable
candidates for continent diversion (88).

d) Complications
Most of the early and late complications of women
undergoing radical cystectomy and neobladder are
identical to those of men and are managed in a
similar fashion [20, 102, 103]. Two complications are
different in female patients:

c) Anatomic Basis of the Preservation of


Continence in Women
Orthotopic neobladder was introduced as an option
for women in the 1990s. Prior to that time, it was
generally believed that the primary continence
mechanism in women was located in the bladder
neck itself [90].

1. Pouch-vaginal Fistula
This complications occurs in 1-5% of patients even
in experienced hands [86, 101, 104]. Awareness of
this potential complication is important at the time
of surgery. When possible, the anterior vaginal wall
should be left intact, taking a strip of vagina only
in cases with close approximation of the tumor.
Any vaginal incision should be carefully closed in
a watertight manner with absorbable suture. An
omental flap may be transposed to the pelvis and
tacked to the pelvic floor on either side of the urethral
stump [79, 97].

However, it was ultimately recognized that the urethra


alone could provide continence if the sphincter
mechanism was carefully preserved [91].
Colleselli and colleagues did elegant cadaver
studies showing that the primary rhabdosphincter in
adult women is an omega shaped structure under
the pubic symphysis deep to the endopelvic fascia
surrounding the distal third of the urethra. The main
nerves supplying this sphincter are the somatic
pudendal nerves which run under the endopelvic
fascia. The autonomic nerves coursing through the
pelvic plexus and along the lateral vagina supply the
smooth muscle of the bladder neck and urethra. The
smooth muscle extends throughout the length of the
urethra [92]. Similar findings had been described
by others [93]. The distal urethra was found to
correspond to the area providing continence on
videourodynamic studies performed in women
who had undergone orthotopic reconstruction
after cystectomy [94]. There is consensus that
preservation of the rhabdosphincter function is critical
to maintaining continence in women undergoing
neobladder reconstruction [79, 86, 87, 95-97].
Avoiding dissection into or below the endopelvic
fascia surrounding the urethra helps ensure that this
muscle and its nerve supply are not disturbed [96].
Stenzl, Studer, and others suggest that preservation
of this plexus in a nerve-sparing approach is crucial
to maintaining urethral tone and may decrease the
risk of late urinary retention due to spasticity of the
denervated smooth muscle of the urethra [93, 98100]. Stenzl compared 66 patients from 3 centers
who had bilateral nerve preservation to 28 who had 1
side preserved and 7 who had no nerve preservation
(due to tumor infiltration or scarring). The report
does not indicate how these data were collected or if
there was a difference in follow-up or other variables
between the groups. Self-catheterization was
required in only 9% of the first group (66 patients)
compared to 0% of the group with 1 nerve spared
(28 patients) and 72% of the 7 patients without nerve

These fistulae rarely heal spontaneously except


in the first few weeks after surgery. Therefore, a
prolonged trial of catheter drainage or more proximal
diversion is probably not warranted. Vaginal
estrogen supplementation may help healing prior
to such a repair. Repair can be attempted vaginally,
with a Martius labial fat pad flap or muscle flap from
the leg interposed between the layers in difficult or
irradiated cases. When attempts to repair a vaginal
fistula have failed, the patient may be better served
by conversion to a cutaneous diversion.

2. Urinary Retention
Urinary retention is clearly more common in women
than men undergoing orthotopic diversion (Table 9).
Such retention may occur early, but often appears
after a year or more of good neobladder function
and emptying. In the Ulm series of 116 women,
the rate of retention increased steadily over time to
approximately 50% by 5 years [78]. The etiology has
been debated, but most authors believe it is due to
a mechanical kink in the uretha-pouch anastomosis
as the full pouch falls posteriorly during Valsalva
maneuver [79, 97, 103, 110-112]. This can often
be documented on a lateral straining cystogram.
However, not all patients with retention have this
finding. Other suggested etiologies include autonomic
denervation of the urethral stump or disordered reinnervation resulting in inability to relax the sphincter
[78, 94, 113, 114].
Since the first description of this potentially undesirable
late complication of orthotopic bladder substitution
in women [109] in 1996, a number of authors have

359

suggested modifications in surgical technique to try


to prevent this problem and have presented data to
suggest improved outcomes [79, 95, 97, 110, 112].
However, all are consecutive series and because
the complication may appear late, such reports may
be biased by shorter follow-up in the new group.
Nevertheless, some attempt to fill the posterior
pelvis and re-establish anterior and superior fixation
of the new bladder seems to be warranted. At the
University of Southern California, a sacrocolpopexy
with mesh and omental transposition laid between
the bladder and vagina have been routinely
performed [79]. Ali-el-Dein described anchoring the
vaginal apex to the preserved round ligaments and
also used an omental flap [97]. However, others
believe these maneuvers are unnecessary [118],and
recent results suggest that they have not prevented
retention (see below). Studer has suggested that
the location of the urethral opening in the pouch is
an important variable [115]. A recent study from the
group in Ulm suggested that patients in whom the
bladder neck itself is preserved (for example those
with non-urothelial tumors) have a higher risk of
retention than those in whom the urethra is divided
just below the bladder neck [108]. Preservation of the
uterus and its supporting ligaments when possible
may be an effective way to help prevent retention,
though again this question has not been subjected
to a prospective or randomized trail [116].

It is clear that every woman undergoing neobladder


reconstruction should be advised that intermittent
catheterization may be required for adequate
emptying and must be willing and able to learn
how to perform this. Many women who are dry but
require self-catheterization seem quite happy with
the diversion in spite of this [117].

e) Continence (Table 9)
Continence results in the literature are difficult
to compare between series because of a lack of
consensus on definitions, varied follow-up periods,
and varied mechanisms of data collection. Some
of the larger series reporting continence results
with neobladders have not separated out patients
by gender [86, 110, 115]. Few studies have used
the gold standard of a validated, anonymous
questionnaire to evaluate continence. In addition,
urinary retention may develop as a late event, so
follow-up time is an important variable in these
reports. In one of the largest series, Ali-el-Dein and
associates reported on a total of 136 women who
underwent a radical cystectomy and orthotopic
substitution, with 100 patients evaluable at a mean
follow-up of 36 months. Overall, 95% of the women
were continent in the day, 86% were continent at
night, 2 were completely incontinent, and 16% were
in chronic retention [102].
Stein and colleagues completed a mailed questionnaire
study using the validated Bladder Cancer Index with
56 women returning the questionnaire (64% of the 87
surviving women). Significant daytime incontinence
was reported in 23% of the women and nighttime
incontinence in 34%. Somewhat surprisingly, 61%

Treatment of retention is intermittent catheterization.


Alpha blockers are not effective [97]. Transurethral
resection of a urethral fold and open reduction of the
pouch size with anterior fixation to the abdominal
wall have also been described [97, 116].

Table 9. Continence In Reported Series Limited to Females


Follow-up
Author

Hautmann

# pts

(mos -mean)

Daytime

Nighttime

Continence % Continence%

Self Catheterize
%

(1996)
(2000)

116

60

83

83

50

(2006)
Granberg

(2008)

59

29

90

57

35

Nesrallah

(2005)

29

37

97

86

10

Ali-el-Dein

(2008)

192

51

92

72

16

Stenzl

(2001)

83

26

82

72

11

Stein

(2002)

81

78

78

N/A

33

Stein

(2009)#

56

103

87

66

61

# Anonymous validated questionnaire

360

reported that they catheterized at least once per day


and 39% always voided by self-catheterization. Only
18% of the women who catheterized reported that it
was a moderate bother and 56% reported it was no
bother at all [117].

because of the risk of a false negative report from


the pathologist. Urothelial cell carcinoma located in
the urethra or involving prostatic ducts or stroma is
the main indication for urethrectomy.

In female patients, biopsies should also be obtained

f) Sexual Function and Quality of Life

from the bladder neck, and, if positive, urethrectomy


should be performed. Furthermore, it is necessary
to exercise caution if there is a tumor close to
the bladder neck, as well as in cases involving
widespread CIS. Optimal knowledge regarding the
urothelial pathology of the lower urinary tract is of
importance when informing and discussing with the
patient preoperatively.

Only a few studies have examined the postoperative


sexual function of women undergoing cystectomy
and urinary diversion [118-122]. Results suggest that
sexual dysfunction is common and may be potentially
improved by leaving the uterus intact when possible
and preserving the autonomic nerves lateral to the
vagina [119, 120]. Comparisons of most aspects of
quality of life in both men and women between types
of urinary diversion have been limited but do not
show any convincing differences [122,123, 124].

Some patients may prefer continent diversion to


orthotopic reconstruction because of the risk of
urine leakage after the latter procedure.

Conclusion

c) Prerequisites

Orthotopic neobladder reconstruction is an attractive


option for selected women undergoing radical
cystectomy for bladder cancer. Oncologic outcomes
appear to be excellent with appropriate selection
criteria. Careful attention to patient selection, surgical
technique, and follow-up are all important to optimize
functional results.

Although antirefluxing ureteric anastomosis is not


necessary in orthotopic bladder substitution, it is
required in continent cutaneous diversion because
the efficient outlet mechanism can allow high
intrareservoir pressure. Both ileum and the right
colon can be used for construction of the pouch.
Different types of inlet, pouch, and outlet have
been combined. The stoma is usually in the right
lower quadrant or in the umbilicus. In this review,
the techniques that are most widely used today are
described. The majority of these were developed in
the 1980s and 1990s, although a few new methods
were also reported in the last decade.

Additional studies are necessary to allow surgeons


to minimize incontinence and urinary retention in
these patients.

3. Continent Cutaneous Urinary


Diversion
a) Introduction

d) The Outlet

The modern era of continent urinary diversion began 30


years ago with the introduction of continent cutaneous
diversion, which at that time was represented by
the Kock pouch. Since then, numerous techniques
for this type of diversion have been described, but
some of these appeared only once in the literature,
indicating that they were associated with technical
problems, high complication rates, and suboptimal
functional results. Today, only a handful of methods
are in use, and, in general, they are the second
choice after orthotopic bladder substitution for
patients undergoing radical cystectomy.

T
 he intussuscepted ileal nipple valve
T
 he Kock pouch: The Kock pouch was the

first method to be accepted by the urologic


community after it was described by Kock and
refined by Skinners group in the 1980s [125,
126]. However, the initial enthusiasm about this
procedure was soon dampened by increasing
reports of complications, including the following:
the afferent ileal nipple valve was prone to stenosis
with subsequent upper tract dilatation and risk of
renal impairment; the efferent counterpart was
subject to erosion by mesh, pinhole fistula, sliding,
and prolapse with risk of urine leakage and/or
difficulties in catheterization. The risk of afferent
nipple stenosis is estimated to be 30% after 5
years [127]. Because of the technical complexity
of the Kock pouch and the high complication rate,
this procedure is no longer used in routine clinical
practice. A modified technique for constructing the
efferent nipple [128] has been described, but no
further information is available in the literature from
the past 5 years.

b) Indications

In patients with bladder cancer undergoing radical

cystectomy, the main indication for continent


cutaneous diversion is when urethral removal is
deemed necessary due to a high risk of recurrence
of urothelial carcinoma. This risk can be estimated
based on the pathology report from the preoperative
transurethral resection biopsies of the prostate.
Such biopsies should be taken from the bladder
neck to the verumontanum on both sides before
cystectomy. Relying on frozen sections of the
urethra obtained during surgery may be dangerous

361

T
 he Mainz pouch I: The ileal nipple valve was also

nocturnal continence in 74%, and 22% indicated


having satisfactory continence [135].

used in the first Mainz pouch that was described


[129] but modifications were subsequently
introduced due to the above-mentioned problems.
Today, the appendix is the first choice, but if it
missing or unsuitable, a nipple valve is configured
and fixed to the ileocecal valve and the reservoir
wall with staples [130]. In terms of complexity, the
Mainz pouch I with an intussuscepted nipple valve
cannot be considered inferior to the Kock pouch.

T
 he Lundiana pouch: In the Lundiana pouch, the
ileum is also stapled, and the ileocecal valve is then
incorporated into the outlet in such a way that it
adds to the efferent length and closure mechanism
[136]. In Lund, 200 patients received this type
of diversion in the years 1992-2007, 160 at time
of cystectomy and 40 to treat benign disease.
Early (90-day) mortality was 2.5%. Evaluation of
191 of the patients who started intermittent selfcatheterization revealed that continence was
achieved in 177 (93%). Revision of the outlet due
to incontinence and/or difficulties in catheterization
was done in 13 patients, and revision of the stoma
was performed in 14 (unpublished). In one patient,
conversion to an ileal conduit was necessary after
unsuccessful attempts to revise the outlet. At last
follow-up, 14 patients reported that they had had
one or more episodes of difficult catheterization.

T
 he appendix: Advantages of using the appendix

are that it is ready made and a smaller portion


of bowel is used. Inasmuch as the appendix is not
always available or suitable, it is necessary to be
familiar with other methods for continent cutaneous
diversion. The main problem associated with using
the appendix is the tendency towards stomal
stenosis. In the Mainz study [131], 63 (32%) of 196
patients underwent a total of 107 interventions.
Using a similar technique, other investigators [132]
observed 28% stenosis within a mean follow-up of
only 33 months. More recently, it was found that 6
of 52 patients needed early operative intervention
to revise the stoma, and late stomal complications
occurred in 12 of 52 [133]. This clearly offsets the
excellent continence (100%) reported in several
publications.

T
 he serous-lined extramural valve/T-pouch:

The serous-lined extramural valve as a continent


urinary outlet was first described by the Mansoura
team, and soon thereafter a similar technique
was reported by the Los Angeles group (double
T-pouch). The principle of both these methods is
the creation of a serous-lined trough in which a
tapered ileal segment is placed. Two articles have
described the use of this technique in Europe [137,
138]. As correctly stated by the authors of this
report, construction of the pouch is sophisticated.
Indeed, this is the main drawback of the method
and represents a serious obstacle to general
acceptance.

T
 he

tapered/stapled ileal outlet: Several


variations of the tapered/ stapled ileal outlet have
been described, all of which comprise a narrowed
ileal segment that creates passive resistance. A
true intraluminal closure mechanism is absent in
most, which may have an impact on continence.
The advantage of this approach is its simplicity.

e) Stone Formation

T
 he Indiana pouch: In this procedure, the ascending

Stone formation is a common phenomenon after


urinary tract reconstruction, and its etiology is
multifactorial, including residual urine, chronic
bacteriuria, mucus, and foreign material such as
staples [139]. Although an incidence as high as
44% has been reported [140], most reports indicate
rates of 5-20%. In the series of 191 Lundiana outlets,
pouch stones occurred in 22 (12%) patients, and
in 18 of those patients, the stones were removed
endoscopically (unpublished data). the importance
of regular pouch irrigation has been recognized and
should be encouraged.

colon patched with an ileal segment constituted the


pouch, the outlet being a 10-cm plicated or stapled
ileal segment. Some modifications in fashioning
the outlet were also introduced, such as stapling
the ileum with a gastrointestinal anastomosis
(GIA) stapler and metallic staples, and using only
the right colonic segment. In addition, Lambert
sutures were placed at the angulated portion
of the ileocecal valve so as to roll cecal tissue
around the ileocecal valve and narrow it; this was
done to increase outflow resistance. However, in
a series of 125 patients in which the technique
had been modified to some extent with regard to
the ileocecal valve [134], the vast majority had
complications, 73 of which were related to the
efferent limb, such as incontinence, parastomal
hernia, difficult catheterization, or stomal stenosis.
These problems resulted in 31 reoperations. The
final outcome with regard to continence was not
stated. The last published study concerning the
Indiana pouch indicated that complete daytime
continence was achieved in 96% of patients and

f) Perforation/Rupture
The risk of this complication is higher after continent
cutaneous diversion than after orthotopic bladder
substitution, because the former lacks a pop-off
mechanism [141]. In the series of 191 Lundiana
pouches (unpublished data), this complication was
seen in 8 patients; 6 of these individuals underwent
open surgery, whereas 2 were cured by antibiotics
and drainage.

362

g) Other Important Aspects

complication rates [145]. It is likely that ileum is


used more commonly because it is the technically
simplest conduit to perform. Acknowledging this,
there are still settings where it is preferable that
colon, as opposed to ileum, be used. Specifically,
in the setting of patients with short bowel syndrome
or in patients who have had prior irradiation of the
ileum or distal ureters, a colonic conduit should be
considered. Another setting where a sigmoid colon
conduit should be considered is in the patient who is
undergoing en bloc resection of the colon or rectum,
as this may eliminate the need for an additional
bowel anastomosis.

There is virtually 100% bacterial colonization after


intestinal urinary diversion, but, provided the flow
of urine is not obstructed, clinical symptoms are
exceedingly rare. This is discussed in another review
published by Wullt and co-workers [142].

Conclusion
Continent cutaneous diversion has a place as an
option for reconstruction of the urinary tract in patients
who undergo cystectomy. The main indications
seem to be in patients in whom urethrectomy has to
be performed and in those in whom the prospect of
possible urine leakage after orthotopic neobladder is
repugnant. Multiple techniques have been described.
However, many of them are too complicated to gain
widespread acceptance. Simplicity characterizes
the appendiceal outlet and the outlet of the different
modifications of the Indiana pouch, and excellent
functional results can be obtained. However,
complications from the pouch and the outlet are not
infrequent and lifelong surveillance of these patients
is necessary.

Ileum or colon are the preferred intestinal segments


for use in urinary conduit diversion. Jejunum is
technically feasible, but should be used as a
last resort, given the higher rate of metabolic
abnormalities. A recent Cochrane review of the
available evidence comparing ileal conduit to colonic
conduit failed to show significant differences in upper
urinary tract infection or ureterointestinal stenosis
[147]. This review did not specifically look at rates
of stomal stenosis or metabolic abnormalities,
but others have noted similar rates of these longterm complications when comparing ileal to colonic
conduit [145, 146]. To this end, the choice of ileum
versus colon in conduit diversion should be based
upon surgeon preference and patient characteristics,
as discussed earlier.

4. Conduit Urinary Diversion


The conduit is still the most commonly performed
urinary diversion today [2, 143]. This is also confirmed
by our group (Table 1). Nevertheless, there is very
little new information in the literature and a lack of
hard data, including standardized reporting. Recent
efforts to overcome this dilemma at least in part are
the 2003 and 2011 series published by Madersbacher
et al and Shimko et al.

a) C
 omplications Following Conduit Diversion
The discussion of complications again is hampered
by insufficient standardized reporting.

The introduction of continent urinary diversion


minimized the impact of urinary diversion on body
image and allowed the patient to live a more normal
life. Despite these potential benefits of continent
urinary diversion, ileal loop conduit has remained the
more commonly used form of urinary diversion, likely
due to the fact that it is technically simpler to perform.
Rarely conduits are formed from large bowel. Whether
or not this last point regarding complications is true
remains a point of debate. Large bowel surgery has a
higher infectious complication rate than small bowel
surgery, the mesocolon is shorter, and the vascular
supply more critical, depending on the segment
used. Advantages are, for example, a bigger stoma!
The metabolic abnormalities associated with jejunal
conduits provide a compelling reason to limit the
use of jejunal conduits to patients who, for whatever
reason (i.e., prior irradiated bowel or distal ureters
and/or inflammatory bowel disease) cannot undergo
either an ileal or colonic urinary conduit.

Short-term Complications
A recent review of complications in 1057 patients
undergoing ileal loop diversion at the Mayo clinic
documented a 3.6% incidence of abscess at
the bowel anastomosis and a 2.7% incidence of
enteric fistula formation, although some of these
complications occurred more than 3-6 months after
surgery [22]. Given that bowel leak, abscess, and/or
fistula formation are potentially lethal complications,
meticulous surgical technique should be employed
when performing the ileoileal bowel anastomosis
[147].
The other short-term complication that is highly
relevant to urinary conduit diversion is urine leak
at the ureteroileal anastomosis. Estimated to occur
in 2-5.5% of patients [145, 148], this complication
was associated with an astonishingly high rate of
mortality in early ileal conduit series [149, 150].
While more recent reports do not demonstrate such
high mortality rates with this complication [146], this
is still a complication that should be avoided with the
use of proper surgical technique and placement of
ureteral stents. A recent randomized clinical trial of

Urinary conduit using ileum is the most commonly


performed conduit procedure [144]. Studies
comparing ileal or colonic urinary conduit diversion
have documented fairly similar long- and short-term

363

54 patients undergoing either ileal loop conduit or


ileal neobladder documented more frequent urine
leak on postoperative day 1 in the unstented group,
although no differences were noted on postoperative
days 3 and 7. Importantly, stenting reduced the risk
of early upper tract dilatation and was associated
with improved bowel function. Stenting was not
associated with an increased risk of ureteroileal
stricture or upper tract infection [147]. To this end, we
recommend routine ureteral stenting in all patients
undergoing ileal loop conduit.

rate with a median time to stricture of 1.1 years [22].


It has been suggested that the individual end-toside ureteroileal anastomosis technique described
by Bricker may have a higher rate of stricture when
compared to using a conjoined (Wallace) technique
[149]. However, 2 recent retrospective series failed
to document any difference in the rate of ureteral
stenosis between the 2 different types of reimplant
[152, 153]. The choice of ureteral reimplant, therefore,
should be based on individual patient characteristics
and surgeon preference.

Long-term Complications

Long-term complications associated with the urinary


stoma are not infrequent occurrences following ileal
loop conduit. Stomal stenosis has been reported
to occur in 2-4% of patients undergoing ileal loop
diversion, with a median time to diagnosis of 9.2
years in one study [22, 148]. Chechile et al noted
a stomal revision rate of 5.5% in their patients
undergoing ileal loop and, importantly, noted no
difference in the re-operative rates between patients
undergoing end versus loop type stomas [154). The
incidence of parastomal hernia following ileal loop
varies from 3.7-13.9% [22, 148] and was noted to
occur at a median of 2.4 years after surgery.

Ileal conduits have been used for many years.


However, standardized and detailed information
is only rarely available regarding the incidence of
complications. Madersbacher et al for example,
reported on a cohort of 131 patients undergoing
ileal conduit diversion who survived at least 5 years,
where an overall complication rate of 66% was
observed [21]. In fact, many patients experienced
more than one complication during the first 5 years,
and complications can continue to develop even
beyond 15 years; 24% of patients developed stoma
problems, 24% had bowel complications, 23% had
symptomatic urinary tract infection (UTI), and 27%
had deterioration of renal function.

Table 6 presents a comparison of conduit versus


neobladder The data must be interpreted with
caution. The neobladder long-term complication rate
seems to be much lower (40% vs. 60%!) .

In the Mayo series, 19% of patients undergoing ileal


loop diversion developed new onset chronic renal
insufficiency (defined as a serum creatinine greater
than 2.0 mg/dL) at a median of 2.3 years after
surgery [22].

The conduit complication rate seems to be endless!


As expected, the UTI rate favors the neobladder
(5% vs. 25%) and the ureteroileal anastomosis
complication rates are identical.

There are a number of reasons why these patients


experience deterioration in renal function. One
common etiology is renal scarring secondary to
infection. Reported long-term pyelonephritis rates
following ileal loop conduit vary from 6-23% [151,
21]. In the Mayo series, 12% of patients reported a
pyelonephritis episode at a median of 2.3 years after
surgery. The overall infectious complication rate in
this series was 16.5% [22]. There has been some
debate concerning the value of nonrefluxing ureteral
anastomoses in conduit surgery, but the available
evidence does not support the use of nonrefluxing
reimplants in the setting of urinary conduit. Given
that refluxing ureters allow easier radiologic followup of the upper tracts, as loopograms should allow
free reflux of contrast facilitating assessment of the
remaining urothelium, refluxing ureteral reimplants
should be routinely performed at the time of ileal
conduit.

Conclusion
Ileal conduit diversion remains the most commonly
used method for reconstructing the urinary tract in
conjunction with radical cystectomy. It is probably
technically easier than continent reconstruction.
However, the complications, early as well as late, are
legion. Several studies confirm a high incidence of
upper tract complications, probably increasing with
length of follow-up.
It is difficult to draw definitive comparisons with other
diversion techniques as surgical techniques have
improved markedly over the last 25 years, and few
series report comparable long-term outcomes ( 20
years) in patients with neobladders.

5. Ureterosigmoidostomy
Regarding this type of diversion, there is really
no new information. This corresponds to conduit
diversion.

Another potential cause of long-term renal


insufficiency is ureteroileal anastomotic stricture.
This is thought to occur primarily due to distal
ureteral ischemia and is associated with a high reoperation rate. A recent systematic review of the
literature noted an 8.2% incidence of anastomotic
stricture following ileal conduit [148]. The group at
the Mayo clinic reported a 10% anastomotic stricture

a) Preoperative Considerations
The success of ureterosigmoidostomy depends
on patient selection. Factors are anal sphincteric
function, renal function, liver function, degree of

364

- Urine Leakage

ureteral dilation, history of prior radiation therapy,


primary colonic disease, and patients compliance with
long-term medications [155]. Paramount to success
is an adequate anal sphincter mechanism. This
should be assured both by history and preoperative
testing using large volume enemas with the patient
assuming normal activities. Inability of the patient
to retain 400 to 500 mL in the upright position for 1
hour is a contraindication to ureterosigmoidostomy
[155]. Patients with neurogenic bladder are not
suitable candidates for ureterosigmoidostomy, as
there may be associated anal sphincter dysfunction
[156]. Preoperative evaluation of the patient
selected for ureterosigmoidostomy should include
studies for bowel disease that might contraindicate
the operation. The presence of diverticulitis or colon
polyps may be investigated by barium enema or
colonoscopy [155]. Compromised renal function
results in accentuation of hypercritical acidosis that
results from ureterosigmoidostomy. Consequently,
patients should have normal renal function to avoid
this complication. Inability to handle the excess
ammonia absorption has been reported to lead
to hyperglycemia encephalopathy. Therefore,
ureterosigmoidostomy should be avoided in patients
with liver disease [157]. In the presence of ureteral
dilation, creation of a nonrefluxing anastomosis may
be difficult to achieve. Compliance with long-term
medication, such as antibiotics and bicarbonate, is
mandatory in patients with ureterosigmoidostomy
[155].

Leakage of urine occurs as an early complication


of ureterosigmoidostomy with a reported incidence
of 3.5% after combined technique of ureterocolic
anastomosis [161]. Small leaks usually seal
spontaneously. Extensive leaks, especially those
associated with prolonged ileus or signs of peritonitis,
are indications for immediate reoperation. Bilateral
percutaneous nephrostomy tubes and defunctioning
colostomy are viable conservative therapeutic
options [162].
- Pelvic Abscess
This diagnosis should be considered if a
patient develop otherwise unexplained fever.
Ultrasonography and computed tomography with
or without aspiration establish the diagnosis in most
cases. Once diagnosed, open or percutaneous
drainage should be carried out [162].

e) Late complications
- Reflux
Reflux is rare except after the Nesbit technique
of direct mucosa-to-mucosa anastomosis [163].
The incidence is high with dilated ureters, so
if one or both ureters are significantly dilated,
ureterosigmoidostomy should not be considered
[164]. Allen interposed an ileal nipple valve between
the dilated ureters and the sigmoid colon to solve
this problem [165].

b) Preoperative preparation
An adequate bowel preparation is mandatory
prior to ureterosigmoidostomy to avoid infectious
complications. The regimen includes a low residue
diet, oral laxatives, cleansing enemas, and oral
intestinal antiseptics for 3 days preoperatively [155].

- Pyelonephritis
Pyelonephritis is one of the most common and most
dangerous complications of ureterosigmoidostomy,
even with the combined or transcolonic techniques.
Wear and Barquin reported an incidence of 81% with
the older refluxing techniques of Coffey [166] and
Nesbit [167], as compared with an incidence of 5.7%
after the Leadbetters [168] combined technique
[164]. Williams et al reported an incidence of 45%
with the Goodwin transcolonic technique [169, 170].
Zincke and Segura reported an incidence of 20% with
the combined technique of ureterosigmoidostomy
[171]. The combination of high bacterial counts
inside rectal contents and the high rectosigmoid
pressure during voiding might be responsible for the
high incidence of pyelonephritis. Thus, long-term
antibacterial suppressive therapy is recommended
for patients with ureterosigmoidostomy [172].

c) Advantages
In 1973, Wear and Barquin enumerated the following
advantages of ureterosigmoidostomy over Brickers
ileal loop: voluntary sphincteric control of urination
without an external stoma and its complications, no
indwelling tubes or external collecting device, shorter
operating time and easier technique, ability to stage
the operation, ability to perform via intraperitoneal
or extraperitoneal approach, requirement of 2 rather
than 5 suture lines, and better acceptance by the
patient and his relatives [158, 159].

d) Early Complications
- Postoperative Anuria
The most serious immediate complication from
ureterosigmoidostomy is anuria if the anastomosis
is not stented. Anuria after removal of the ureteral
stents indicates bilateral obstruction caused by tissue
edema. In both cases, percutaneous nephrostomy
tubes should be placed, especially in case of fever
or flank pain [160].

- Ureterocolic anastomotic stricture


Ureteral obstruction at the ureterocolic anastomotic
site may occur as a late complication in 32% of patients
when the Leadbetter technique [160] is used and in
49% with other types of ureterocolic anastomosis
[164]. The incidence of upper tract dilatation after

365

ureterosigmoidostomy with a submucosal tunnel


was 12.845% [161, 173, 174, 175, 176].

cystectomy in geriatric patients. Several reports


have documented feasibility and oncological efficacy
of cystectomy in the elderly.

- Incontinence

Although perioperative mortality was substantially


higher in the advanced age group (5.3% vs. 1%),
postoperative morbidity was similar (34% in those <70
years, 39% in those 70 years [186]. The Memorial
Sloan-Kettering Cancer Center retrospectively
analyzed 44 octogenarians who underwent radical
cystectomy [187]. Postoperative mortality was 4.5%
and 29 patients (66%) had to be hospitalized acutely
after surgery, emphasizing the considerable mortality
and morbidity of this procedure in octogenarians
[187]. The major advantage of radical cystectomy is
the avoidance of local problems caused by locally
advanced disease; following radical cystectomy, the
rate of local recurrences is small [188]. One has to be
aware, however, that in contrast to younger patients,
the oncological superiority of radical cystectomy
compared to bladder-sparing approaches in elderly
patients is not convincingly demonstrated [189].

Incontinence after ureterosigmoidostomy is a


drastic social problem. It can be attributed to high
rectal pressure. Patients should be advised to
empty on a regular basis, including once or twice
during the night [166]. Imipramine hydrochloride
at bedtime is indicated for those with persistent
nocturnal enuresis [177]. Daytime continence rates
for ureterosigmoidostomy range from 8397% [161,
173, 174, 175, 176] and nighttime continence rates
from 58-92%.

Conclusion
Although the mortality and initial morbidity following
ureterosigmoidostomy have been significantly
reduced, some inherent chronic complications
remain problematic.

f) The sigma rectum pouch (Mainz Pouch II)

b) Types of Palliative Urinary Diversion

The sigma rectum pouch is another low pressure


modification of ureterosigmoidostomy. The procedure
was introduced by the Mainz group in 1993 [178].
The procedure entails single folding and antimesenteric splitting of the rectosigmoid colon to improve
the urodynamic characteristics of the reservoir. The
Goodwin technique is utilized for ureteral reimplantation [179]. In 1996, the same group reported on
their experience with the sigma rectum pouch in 73
patients [180]. Stenosis at the ureteral implantation
site was encountered in 5 patients (6.8%). Daytime
continence was achieved in 95% of patients. Nocturnal enuresis was encountered in only 2% of patients.
All patients were on prophylactic alkalinization. Similar satisfactory short-term results were reported [181
- 183]. However, no long-term results were reported
following this technique.

Cutaneous Ureterostomy
This is the most popular form of alternative nonbowel form of diversion in elderly patients or in a
palliative setting [190, 191]. In a recent series of
72 patients 80 years of age or older undergoing
radical cystectomy and urinary diversion, cutaneous
ureterostomy was performed in 87.5% [191].
Operative time is short, and renal function is not a
selection factor. Construction of a single stoma in the
lateral or midline position is generally feasible and
ensures easy care with minimal patient discomfort.
Stenting is usually necessary to avoid stomal
stenosis. Saika et al assessed health related quality
of life in elderly patients (75 years) who underwent
radical cystectomy and received either an ileal
conduit, cutaneous ureterostomy, or orthotopic
bladder substitution [190]. After a follow-up of over 3
years, no relevant differences in any area of quality
of life was seen in the 3 groups, underlining the role
of cutaneous ureterostomy in frail, elderly patients
[190].

6. Palliative Urinary Diversion


The issue of palliation and urinary diversion centers
around 2 issues: (1) management of elderly/geriatric
patients with muscle-invasive bladder cancer in whom
radical cystectomy/urinary diversion is associated
with a considerable morbidity and mortality, and
(2 patients with tumor-induced upper urinary tract
dilatation and renal insufficiency in a palliative setting.
A recent analysis of the SEER-database revealed
a peak age incidence of bladder cancer around 85
years [184]. Due to demographic changes with a
3-fold rise in the number of octogenarians expected
in the next 25 years, a substantial increase of elderly
patients with bladder cancer can be expected [185].

Defunctionalization of the Contralateral Renal


Unit

If only 1 kidney is diverted and urine continues


to flow downstream, it may be necessary to
defunctionalize the latter [192]. These patients
are usually poor candidates for nephrectomy. In a
previously obstructed kidney, ligation of the ureter
usually causes significant pain and spontaneous
ureteral recanalization [192]. In these cases, renal
arterial embolization should be considered.

a) Bladder sparing or radical cystectomy with


urinary diversion in elderly patients?

Palliative Treatment of Ureteral Obstruction

There exists no randomized study to compare


a bladder-sparing strategy versus palliative

Common causes of malignant compression of the

366

7. Urinary Diversion in the Pediatric


Age GroupSpecial Considerations

ureterovesical junction and the prevesical ureter


are locally advanced bladder, prostate, and cervical
cancers. Malignant obstruction of the mid- and upper
ureter is usually caused by metastatic lymphatic
spread. Indication for treatment should be considered
very carefully [192]. Septic episodes are rare in
malignant ureteral obstruction if there has been no
retrograde endoscopic manipulation. The insertion
of a regular double J stent or percutaneous catheter
will result in frequent consultations to change the
catheters, even under anesthesia. Any manipulation
carries the risk of infection and dislocation, and
quality of life should be the main treatment goal.

In the pediatric age group, continent and incontinent


types of urinary diversion can be applied, each
with specific advantages and disadvantages [197].
Bladder augmentation and heterotopic continent
cutaneous diversion are the options for continent
urinary diversion in patients with and without
neurogenic deficits. Orthotopic bladder substitution
and continent anal diversion [198] should be offered
only to patients with competent urethral or anal
sphincters. For incontinent urinary diversion, a
colonic conduit is the preferred solution in children
[199].

Percutaneous Nephrostomy
Drainage of an obstructed upper urinary tract caused
by a locally advanced or metastatic urothelial cancer
leads to an ethical dilemmais such drainage
going to facilitate treatment with chemotherapy or
radiotherapy, or is it perpetuating and allowing other
problems to develop? Aravantinos et al concluded,
based on an analysis of 270 cases, that only patients
with specific cancers (e.g., prostate cancer) that
progress slowly by nature may substantially benefit
from the procedure [193]. Bilateral nephrostomies
are generally poorly tolerated and usually only the
renal unit with the better function should be diverted
by nephrostomy.

Indications for urinary diversion in childhood include


neurogenic bladder, functional or anatomical
loss of the lower urinary tract, radical surgery for
malignancies, and failure of reconstruction of the
lower urinary tract (e.g., failure of primary bladder
closure in patients with bladder exstrophy).

a) Postoperative Considerations
The level of the neurological defect is a crucial aspect
in patients with myelomeningocele. This determines
whether a patient is ambulatory or wheelchair-bound,
and whether she or he will be able to adequately
perform clean intermittent catheterization (CIC). An
important aspect is the transition from ambulatory to
a wheelchair-bound state with aging and increasing
obesity [200].

Pyelovesical Bypass
Desgrandchamps et al introduced a technique of
pyelovesical bypass with a composite prosthesis
[194]. In the initial series, 21 patients were treated
with a success rate of over 90% [194]. The authors
concluded that subcutaneous pyelovesical diversion
ensures a better quality of life than classical
percutaneous nephrostomy in cancer patients at the
palliative stage [195]. Similar encouraging data were
reported by others [196].

Urine storage at low pressures in an adequate


capacity reservoir for preservation of the upper
urinary tract and achieving urinary continence are the
main goals of bladder augmentation or substitution
in childhood.

b) Surgical Techniques

Ureteral Stenting

1. Bladder Augmentation and Substitution

These data suggest that in patients with a limited life


expectancy permanent stents might be an option.

Bladder augmentation is indicated in patients who


are incontinent due to an overactive detrusor and/
or a low-compliance bladder or small contracted
bladder after radiation or multiple surgeries. Patients
should be able to perform CIC via the urethra. If this
is impossible for anatomic or orthopedic reasons,
a Mitrofanoff cutaneous stoma is preferable for
catheterization [201].

Conclusion
The decision regarding bladder sparing or radical
cystectomy in the elderly/geriatric patient with
invasive bladder cancer should be based on tumor
stage and comorbidity best quantified by a validated
score, such as the Charlson score. Chronological
age is of limited relevance. Cutaneous ureterostomy
is the most popular non-bowel urinary diversion in
this setting, providing adequate quality of life. The
issue of decompression of an obstructed urinary
tract in a patient with advancing pelvic malignancy
(particularly bladder cancer) remains a difficult
clinical situation. The indication for drainage should
only be made when the views and wishes of the
patient and caregivers are taken into account. The
prognosis remains very poor.

Gastric segments, small and large bowel, and the


ureter can be used for bladder augmentation. The
bladder is opened widely (clamtechnique)
to prevent an hourglass phenomenon of the
bladder, with the augmented segment becoming a
diverticulum at the bladder dome.
In patients, who are unable to perform CIC via the
urethra, the presence of the appendix is a distinct
advantage. The appendix can be used as a continent

367

cutaneous stoma of an augmented bladder. The


appendix is embedded in the taenea libera in a
similar manner as in heterotopic continent cutaneous
pouches [202]. The ileocecal valve can be used
as an antireflux mechanism for ureters, which are
implanted into the terminal ileum, which works even
for severely dilated or short ureters [203, 204]. Nondilated ureters can be implanted into the large bowel
by the submucosal tunnel technique [205].

popular form of urinary diversion until the latter half


of the last century [242-245]. In the long run, high
postoperative complication rates and deterioration of
the upper urinary tract became evident. Electrolyte
imbalances, upper urinary tract infections, chronic
renal failure, and uremia were common in the
past [243, 244, 246-250]. However, refinements in
surgical techniques, improved antibiotic treatment,
and better suture materials allowed a revival of
this surgical strategy as an attractive alternative
for continent urinary diversion [245, 251, 252]. This
is especially true for developing countries where
cultural and economic difficulties are encountered
in accepting and obtaining catheters for intermittent
self-catheterization of a continent cutaneous urinary
diversion and external appliances for incontinent
urinary diversion [253]. Both a previous dilated
upper urinary tract and an insufficient anal sphincter
are contraindications [254].

Complications from the use of gastric segments


are hyponatremic, hypochloremic alkalosis and
hematuria-dysuria syndrome in up to 37% of patients
[206, 207]. A further late complication is development
of secondary malignancies starting from the tenth
postoperative year [208-211]. Therefore, the use of
the stomach for bladder augmentation has become
obsolete.
Ureterocystoplasty avoids most of the adverse
effects of intestinal cystoplasties. However, the
association of a small contracted bladder with a
severely dilated ureter of a non-functioning kidney
is rare. In 1973, Eckstein described the technique of
ureterocystoplasty [212]. Its success depends very
much on patient selection. A reaugmentation rate of
up to 73% is reported [213-216].

The rectosigmoid pouch (Mainz Pouch II) is based


on the classic ureterosigmoidostomy but applies
the principles of detubularization and spherical
reconfiguration of bowel to eliminate high-pressure
peristaltic contractions and create a low pressure,
high capacity rectosigmoid reservoir [198]. In a series
of 35 children with a mean follow-up of 112 months,
all were continent during the day and 3 (9%) had
nighttime incontinence requiring pads. Ten (15%) of
69 ureters required reimplantation (7 for obstruction
and 3 for reflux) [255]. Most of the submucosally
imbedded ureters that required reimplantation were
already preoperatively dilated. Implanting dilated
ureters in a serosa-lined extramural tunnel provides
better results [255-257]. The augmented and valved
rectum, and the double-folded rectosigmoid bladder
are further modifications of the ureterosigmoidostomy
strategy with excellent continence rates [258260]. However, as the surgery of the augmented
and valved rectum technique is quite complex and
requires temporary colostomy, the technique did not
find wide acceptance.

In patients with bladder exstrophy, bladder


augmentation with or without closure of the bladder
neck with a Mitrofanoff stoma is an option after
failure of primary reconstruction [217-220]. In the
long-term, these patients become continent with
their Mitrofanoff stoma. However, the complication
rate is high at 40% [220-222].

2. Continent Cutaneous Diversion


When there are irreparable sphincter defects, small
bladder capacity, obstruction of the upper urinary
tract, patients being unable to perform transurethral
CIC, and impossible reconstruction of the bladder
neck, continent cutaneous diversion is a reasonable
alternative to bladder substitution. Ileum [223-227],
the ileocecal segment [228-235], and colon can be
used for the construction of a continent pouch [236238].

4. Incontinent Diversion
For patients who are either physically or mentally
unable to perform CIC and for patients with chronic
renal failure, incontinent urinary diversion remains
the diversion of choice. Ileum [261-263], the ileocecal
segment [264, 265], transverse colon [266, 267], and
sigmoid colon can be used for incontinent diversion
[268, 269].

High continence rates are the advantage of


continent cutaneous diversion as compared to
bladder augmentation only. Using the embedded
appendix or an ileal intussusception valve as the
continence mechanism, more than 90% of patients
were continent [239]. If the appendix is not available,
the Yang-Monti technique is a viable option for the
creation of a continent catheterizable tube [240,
241].

In 1937, Seifert described incontinent urinary diversion


by a jejunum conduit and Bricker popularized the use
of the ileal conduit in 1950 [261, 263]. Early results
were promising and hence the procedure was also
commonly used for urinary diversion in children.
However, the number of reported complications
increased with the length of follow-up [270].
Deterioration of the upper urinary tracts and renal
function, calculus formation, and stoma stenosis

3. Continent Anal Diversion in Children (see also


III. 5)
Ureterosigmoidostomy was the first form of continent
urinary diversion to be performed and became a

368

resulted in late complication rates of 19-86% [250,


271-278]. In the 1970s, Hendren and Middleton
concluded that this operation should be abandoned
in children or any patient with potential longevity
[273, 279]. The alternative form of incontinent urinary
diversion is the colonic conduit introduced by belhr
in 1952 and popularized by Mogg in the early 1960s
[268, 269]. Elder and coworkers suggested that it
had the same long-term complications as the ileal
conduit [280]. However, they used an isoperistaltic
and refluxing conduit. Using an isoperistaltic colonic
conduit with a antirefluxive ureter implantation, good
results after a mean follow-up period of more than 16
years were reported (range 5-26 years) [199].

failure of primary reconstruction, dilatation of the


upper urinary tracts is common when lower urinary
tract reconstruction is indicated. Implantation of these
ureters into the bowel has a higher complication rate
as compared to normal ureters [197, 205]. Using a
submucosal tunnel implantation into large bowel,
16% of the renal units required reimplantation. Using
the serosa-lined extramural tunnel technique [293]
in dilated ureters, the reimplantation rate was much
lower (3%) [239].
Recently, the necessity of antirefluxing ureteral
implantation in children has been discussed
controversially [299-302]. Clearly, there is a need
for a prospective randomized study assessing the
impact of reflux from a potentially contaminated
diversion. Until evidence is obtained that even lowpressure reflux will not harm the kidney in the long
run, in young patients with a long life expectancy
the kidneys should be protected by an appropriate
antirefluxing technique.

c) Surgical Complications
The use of bowel segments for urinary tract
reconstruction in children is challenging and surgical
complications are common [197, 281-284]. After
bladder augmentation, Herschorn et al reported a reoperation rate of 36% over 6 years [285]. Using the
ileocecal segment for bladder augmentation and a
cutaneous stoma for for CIC (Hemi-Indiana pouch),
Husman and Cain reported a re-operation rate of
48% in 62 patients [186]. Using the Kock pouch in
20 children and adolescents, Abd-el-Gawad et al
reported pouch-related complications after a followup of 6.5 years in 10 (88%) of 13 children and in 3
(43%) of 7 adolescents [287, 288]. In a series of 91
patients with bladder exstrophy, Surer and colleagues
observed stomal stenosis or stomal prolapse in 24,
calculus formation in 24, bowel obstruction in 3,
fistulae in another 3, and bladder perforation in 2
patients [283].

2. Stomal Stenosis
In continent urinary diversion, stomal stenosis is the
most frequent single complication, with an incidence
of up to 32% [148, 239, 283, 284, 303].
It is treated by simple excision or incision of the
scar at skin level. In children with an ileal conduit,
stenosis of the stoma occurred in up to 42% [273278]. In patients with a colonic conduit, the incidence
was 16% after a mean of 119 years (range 520)
postoperatively [304].

3. Bone Density
Among other mechanisms, chronic acidosis may play
a major role in a decrease of bone mineral density
after bladder augmentation or urinary diversion.

1. Ureterointestinal Stenosis
Ureteral implantation is a crucial point in urinary
diversion. For implantation of the ureter into a bowel
segment, refluxive and antirefluxive techniques
are available [227, 289-293]. The incidence of
ureterointestinal stenosis for refluxive techniques
(e.g., Nesbit or Wallace technique) is between 2%
and 7%, and for antirefluxive techniques (Le Duc,
submucosal tunnel, serosa lined tunnel, afferent
nipple valve) is between 2% and 8% [205, 294298].

4. Growth Retardation
It is very unlikely that the changes in position on
the growth curve after enterocystoplasty are a
consequence of enterocystoplasty. It seems to be a
non-specific phenomenon that must be considered
in any clinical population of similar size and age
distribution after a similar time period [305].

5. Vitamin B12

In a recent review article, Somani and co-workers


demonstrated that the average incidence of
ureterointestinal stenosis in patients with an
ileal conduit was 11% in prospective and 8% in
retrospective studies. In patients with a continent
cutaneous diversion, the stenosis rate was 7% in
prospective and 8% in retrospective studies, and in
bladder substitution it was 7% in prospective and 5%
in retrospective studies [148]. Most of these studies
were performed in patients with bladder cancer and
normal upper urinary tracts.

About 1% of orally administered vitamin B12 is


absorbed by an additional, yet unknown pathway
[306, 307]. Studies, unlike older textbooks suggest,
demonstrate that the terminal ileum can not be
considered the only site of vitamin B12 absorption
[306-311].
After urinary diversion with ileal segments, substitution
therapy is performed in up to 35% of patients [312,
313, 314, 315. 316]. However, in the literature there
are only reports of 5 patients with clinical symptoms
(1 with neurological symptoms) seemingly related to
vitamin B 12 deficiency [314, 316, 317].

In patients with neurogenic bladder or in those after

369

A randomized study demonstrated that oral therapy


(2 mg per day) is as effective as parenteral application
(1 mg intramuscularly) of vitamin B12 in patients with
cobalamin deficiency [307]. Hence, oral treatment of
low vitamin B12 levels may be considered especially
in children.

Based on these, our committee strongly


recommends that this type of surgery only be
performed at high volume hospitals. Our committee
also considers a minimum annual case load of 25
surgeries, done by not more than 2 surgeons, to be
the definition of a high volume center.

6. Bowel Dysfunction

V. References

There are relatively few reports in the literature


concerning bowel dysfunction after urinary diversion.
An increased stool frequency was reported after ileal
or colonic conduit diversion in 4-33% of patients, after
bladder augmentation, in 7-59% after substitution, and
after continent cutaneous diversion in 3-23% [296,
318-324]. On the other hand, stool incontinence is
in epidemiological studies surprisingly common with
an estimated prevalence of up to 20% depending on
age, gender, and population [325, 326].

1. Hautmann RE, Abol-Enein H, Hafez K, et al.: Urinary


Diversion. Urology 2007; 69(Suppl 1A):17-49,
2. Gore JL, Litwin MS: Quality of care in bladder cancer:
trends in urinary diversion following radical cystectomy.
WorldJUrol 2009;27:45-50.
3. Donat SM: Standards for Surgical Complication Reporting
in Urologic Oncology: Time for a Change. Urology
2007;69:221-225.
4. Hautmann RE, Petriconi RC, Volkmer BG: Lessons Learned
From 1,000 Neobladders: The 90-Day Complication Rate.
JUrol 2010; 184:990-994.
5. Shabsigh A, Korets R, Vora KC, et al: Defining Early
Morbidity of Radical Cystectomy for Patients with Bladder
Cancer Using a Standardized Reporting Methodology.
EurUrol. 2009;55:164-176.

IV. Recommendations
At high volume hospitals, orthotopic reconstruction
has become the procedure of choice for urinary
diversion in both men and women undergoing radical
cystectomy. In these patients, the construction of a
neobladder allows the elimination of a stoma and
preservation of body image without compromising
the cancer control. However, the patient must be
committed to the labor-intensive rehabilitation
process. He or she must also have adequate
manual dexterity to perform self-catheterization
should it become necessary. When involvement of
the lower urinary tract by tumor excludes the use of
a neobladder, a continent cutaneous reservoir may
still offer some advantages over an ileal conduit.
For patients who are not candidates for either type
of continent diversion, the ileal loop remains a timehonored option.

6. Studer UE, Hautmann RE, Hohenfellner M, et al: Indications


for continent diversion after cystectomy and factors affecting
long-term results.UrolOncol.1998;4(4):172-182.
7. Kocot A, Spahn M, Loeser A, et al: Long-Term Results
of a Staged Approach: Continent Urinary Diversion in
Preparation for Renal Transplantation:JUrol.2010;184:203
8-2042.
8. Granerus G, Aurell M: Reference values for 51 CrEDTA
clearance as a measure of glomerular filtration rate.
ScandJClinLabInvest 1981;41:611.
9. Kristjansson A, Wallin L, Mansson W: Renal function
up to 16 years after conduit (refluxing or anti-refluxing
anastomosis) or continent urinary diversion. 1. Glomerular
filtration rate and patency of uretero-intestinal anastomosis.
BJUrol 1995;7:539-545.
10. Samuel JC, Bhatt RI, Montagnue RJ, et al: The natural
history of post-operative renal function in patients
undergoing ileal conduit diversion for cancer measured
using serial isotopic glomerular filtration rate and 99m
Technetiummercaptoacetyltriglycine
renography.JUrol
2006;176:2518-2522.

Regrettably, almost all the many recommendations


given by our committee are of grade C.
Nevertheless, grade C recommendations can
be very helpful, as long as the experts have the
necessary experience. There are 3 important
facts:

11. Song C, Kang T, Hong J, et al: Changes in upper urinary


tract after radical cystectomy and urinary diversion: a
comparison od antirefluxing and refluxing orthotopic bladder
substitutes and the ileal conduit. JUrol 2006; 175:185-189.
12. Jonsson O, Olofsson G, Lindholm E, et al: Long-time
experience with the Kock Ileal reservoir for continent
urinary diversion. EurUrol 2001;40:632-640.

Radical cystectomy and urinary diversion have

been assessed the highest relative value of


difficulty of the surgery or any procedure in urology
(open, laparoscopic, and robotic), resulting in a low
acceptance of orthotopic reconstruction as seen
from population-based data.

13. Fontaine E, Leaver R, Woodhouse CRJ: The effect of


intestinal urinary reservoirs on renal function: a 10-year
follow-up.BJU Int 2000; 86:195-198.
14. Thoeny HC, Sonnenschein MJ, Madersbacher S, et al: Is
ileal orthotopic

The morbidity of the procedure is up to 75%

The perioperative and long-term complication rate

15. Austen M, Klble T: Secondary malignancies in different


forms of urinary diversion using isolated gut. JUrol
2004;172:831-838.

diversion-related.

is significant, even in the most experienced hands.

370

bladder substitution with an afferent tubular segement


detrimental to the upper urinary tract in the long term. JUrol
2002;168:2030-2034.

16. Austen M, Klble T. Reply to comment by Lavelle J.Re:


Secondary malignancies in different forms of urinary
diversion using isolated gut. JUrol 2005; 173:1831-1832.

37. Penalver MA, Angioly R, Mirhashemi R et al: Management


of early and late complications of ileocolonic continent
urinary reservoir (Miami pouch). GynecolOncol 1989;
69:185

17. Ahlstrand C, Herder A: Primary adenocarcinoma of distal


ileum used as outlet from a right colonic urinary reservoir.
ScandJUrolNephrol 1998;32:70-72.

38. Eisenberg MS, Dorin RP, Bartsch G, et al.: Early


complications of cystectomy after high dose pelvic radiation.
JUrol 2010; 184:2264-2269.

18. T. Klble, I. Hofmann, H. Riedmiller, and D. Vergho:


Secondar Malignancies in Urinary Diversion: A Multicenter
Analysis. Eur Urol 2011; 60: in press

39. Hensle TW, Bingham JB, Reiley EA, et al.: The urologic
care and outcome of pregnancy after urinary tract
reconstruction. BJUint 2004; 93:588-590.

19. Klble T, Tricker AR, Friedl P, et al.: Ureterosigmoidostomy:


long-term results, risk of carcinoma and etiological factors
for carcinogenesis. JUrol 1990; 144:1110-1114

40. Oelsen S: Pregnancy complicated by ureteroileocutaneostomy. DanMedBull 1972; 19:108-109.

20. Hautmann RE, Petriconi RC, Volkmer BG: Twenty-five


years of experience with 1000 neobladders: The long term
complication rate. JUrol 2011, 186 in press.

41. Richmond D, Zabarievski I, Bond A: Management of


pregnancy in mothers with spina bifida. EurJ ObstetReprod
Biol 1987; 25:341-345.

21. Madersbacher S, Schmidt J, Eberle JM, et al: Long-term


outcome of ileal conduit diversion. JUrol 2003;169:985990.

42. Kuczkowski KM, Hay B: Peripartum care of the parturient


with Indiana continent urinary: a need for a multidisciplinary
approach. AnnFrAnesthReanim 2004; 23:927-928

22. Shimko MS, Tollefson MK, Umbreit EC, et al: Longterm Complications of Conduit Urinary Diversion. JUrol
2011;185:562-567.

43. Kennedy WA, Hensle TW, Reiley EA, et al.: Pregnancy after
orthotopic continent urinary diversion. SurgGynecolObstet
1993; 177:405-409.

23. Ahlering TA, Kanellos AW, Boyd SD, et al: A comparative


study of perioperative complications with Kock pouch
urinary diversion in highly irradiated versus nonirradiated
patients. JUrol 139:1202-1204.

44. Barret RJ, Peters WA: Pregnancy following urinary


diversion. ObstetGynecol 1983; 62:582-586.
45. Murray JE, Reid DE, Harrison JH, et al.: Successful
pregnancies
after
human
renal
transplantation.
NewEnglJMed 1963; 269:341.

24. Mannel RS, Braly PS, Buller RE. Indiana pouch continent
urinary reservoir in patients with previous pelvic irradiation.
ObstetGynecol 1990; 75:891-893.

46. Sciarra JJ, Toledo-Pereyra LH, Bendel RP, et al.: Pregnancy


following renal transplantation. AmJObstetGynec 1975;
123:411-425.

25. Leissner J, Black P, Fisch M et al: Colon pouch (Mainz


Pouch III) for continent urinary diversion after pelvic
irradiation.Urology 2000;56:798-802.

47. Hein PR, Jansen JLJ, Debruyne FMJ: Successful


pregnancy after renal transplantation in a patient with
a urinary diversion. EurJObstetGynecReprodBiol 1978;
8:129-132.
48. Nethercliffe J, Trewick A, Samuell C, et al.: False-positive
pregnancy tests in patients with enterocystoplasties. BJUInt
2001; 87:780-782.

26. Beckley S, Wajsman JE, Pontes JE. Transverse colon


conduit: a method of urinary diversion after pelvic irradiation.
JUrol 184;128:464-468.
27. Gschwend JE, May F, Paiss T, Gottfried HW, et al:Highdose pelvic irradiation followed by ileal neobladder urinary
diversion: complications and long-term results. BrJUrol
1996;77:680-683.

49. Hautmann RE, Volkmer BG: Pregnancy and urinary


diversion. UrolClinNAm 2007; 34:71-88.

28. Wammack R, Wricke C, Hohenfellner R: Long-term


results of ileocaecal continent urinary diversion in patients
treated with and without previous pelvic irradiation.JUrol
2001;167:2058-2062.
29. Hautmann RE, Petriconi R, Volkmer BG: Neobladder
formation after pelvic irradiation. WJUrol 2009; 27:57-62.
30. Hautmann RE, Volkmer BG, Schumacher MC, et al: Longterm results of standard procedures in urology: the ileal
neobladder. WJUrol 2006;24:305-314.
31. Orovan WL, Davis JR: Urinary diversion using the Kock
ileal reservoir in patients with irradiated bowel. CanJSurg
1987;31:1136.
32. Bochner BH, Figueroa AJ, Skinner EC, et al: Salvage
radical cystoprostatectomy and orthotopic urinary diversion
following radiation failure. JUrol 1998;160:29-33.
33. Gheiler EL, Wood FP jr, Montie JE, Pontes JE: Orthotopic
urinary diversion is a viable option in patients undergoing
salvage cystoprostatectomy for recurrent prostate cancer
after definitive radiation therapy. Urology 1997; 50:580584.
34. Brand E: Caecal rupture after continent ileocaecal urinary
diversion during total pelvic exenteration. ObstetGynecol
1991; 78:570.
35. Penalver MA, Bejany DE, Averette HE et al.: Continent
urinary diversion in gynecologic oncology. GynecolOncol
1989; 34:274.
36. Mannel RS, Manetta A, Buller RE et al.: Use of ileocaecal
continent urinary reservoir in patients with previous pelvic
irradiation. GynecolOncol 1995; 59:376.

50. Hautmann RE, Abol-Enein H, Hafez K, et al.: Urinary


diversion. Urology 2007; 69 (1Suppl):17-49.
51. Hautmann RE, Simon J: Ileal neobladder and local
recurrence of bladder cancer: patterns of failure and impact
on function in men. JUrol 1999; 162(6):1963-1966.
52. Revelo MP, Cookson MS, Chang SS, et al: Incidence and
location of prostate and urothelial carcinoma in prostates
from cystoprostatectomies: implications for possible apical
sparing surgery. JUrol 2004; 171 (2Pt 1):646-651.
53. Wood DP, Jr., Montie JE,Pontes JE, et al.: Transitional cell
carcinoma of the prostate in cystoprostatectomy specimens
removed for bladder cancer. JUrol 1989; 141(2):346-349.
54. Kessler TM, Burkhard FC, Perimenis P, et al.: Attempted
nerve sparing surgery and age have a significant effect
on urinary continence and erectile function after radical
cystoprostatectomy and ileal orthotopic bladder substitution.
JUrol 2004; 172: 1323-1327.
55. Costello AJ, Brooks M, Cole OJ: Anatomical studies of the
neurovascular bundle and cavernosal nerves. BJU Int.
2004; 94:1071-1076.
56. Hugonnet CL, Danuser H, Springer JP, et al.: Urethral
sensitivity and the impact on urinary continence in patients
with an ileal bladder substitute after cystectomy. JUrol
2001; 165:1502-1505.
57. Turner WH, Studer UE: Orthotopic Bladder Substitution. In:
Vogelzang NJ Scardino PT, Shipley WU, Coffey DS. eds:
Comprehensive Textbook of Genitourinary Oncology.2nd
ed: Lippincott Williams and Wilkins 2000:459-472.

371

58. Studer UE, Zingg EJ: Ileal orthotopic bladder substitutes.


What we have learned from 12 yearsexperience with 200
patients. UrolClinNorthAm 1997; 24:781-793.

78. Hautmann RE, Volkmer BG, Schumacher MC, et al.: Longterm results of standard procedures in urology: the ileal
neobladder. WorldJUrol 2006; 24:305-14.

59. Studer UE, Danuser H, Hochreiter W, et al.: Summary of


10 yearsexperience with an ileal low-pressure bladder
substitute combined with an afferent tubular isoperistaltic
segment. WorldJUrol 1996; 14:29-39.

79. Stein JP, Ginsberg DA, Skinner DG: Indications and


technique of the orthotopic neobladder in women.
UrolClinNorthAm 2002; 29:725-34.
80. De Paepe ME, Andre R, Mahadevia P: Urethral involvement
in female patients with bladder cancer. A study of 22
cystectomy specimens. Cancer 1990; 65:1237.

60. Hautmann RE: The ileal neobladder to the female urethra.


UrolClinNorthAm 1997; 24:827-835.
61. Haber GP, Crouzet S, Gill IS: Laparoscopic and robotic
assisted radical cystectomy for bladder cancer: a critical
analysis. EurUrol 2008; 54:54-62.

81. Coloby PJ, Kakizoe T, Tobisu K, et al.: Urethral involvement


in female bladder cancer patients: Mapping of 47
consecutive cysto-urethrectomy specimens. JUrol 1994;
152:1438-42.

62. Hayn MH, Hussain A, Mansour AM, et al.: The learning


curve of robot-assisted radical cystectomy: results from
the International Robotic Cystectomy Consortium. EurUrol
2010; 58:197-202.

82. Stenzl A, Draxl H, Posch B, et al.: The risk of urethral


tumors in female bladder cancer: Can the urethra be used
for orthotopic reconstruction of the lower urinary tract?
JUrol 1995: 153:950-5.

63. Ng CK, Kauffman C, Lee MM, et al.: A Comparison of


Postoperative Complications in Open versus Robotic
Cystectomy. EurUrol 2010; 5:274-282.
64. Pruthi RS, Nielsen ME, Nix J, et al.: Robotic Radical
Cystectomy for Bladder Cancer: Surgical and Pathological
Outcomes in 100 Consecutive Cases. JUrol Dec 12, 2009.

83. Chen ME, Pisters LL, Malpica A, et al.: Risk or urethral,


vaginal and cervical involvement in patients undergoing
radical cystectomy for bladder cancer: Results of a
contemporary cystectomy series from M.D. Anderson
Cancer Center. JUrol 1997; 157:2120-23.

65. Hautmann RE: The oncologic results of laparoscopic radical


cystectomy are not (yet) equivalent to open cystectomy.
CurrOpinUrol 2009; 19:522-526.

84. Maralani S, Wood DP Jr, Grignon D, et al.: Incidence of


urethral involvement in female bladder cancer: An anatomic
pathologic study. Urology 1997; 50:537-41.

66. Burkhard FC, Kessler TM, Springer J, et al.: Early and


late urodynamic assessment of ileal orthotopic bladder
substitutes combined with an afferent tubular segment.
JUrol 206; 175: 2155-2160; discussion 2160-2151.

85. Stein JP, Cote RJ, Freeman JA, et al: Indications for lower
urinary tract reconstruction in women after cystectomy for
bladder cancer: A pathological review of female cystectomy
specimens. JUrol 1995; 154:1329-33.

67. Thurairaja R, Studer UE: How to avoid clean intermittent


catheterization in men with ileal bladder substitution. JUrol
2008; 2504-2509.

86. Abol-Enein H, Ghoneim MA: Functional results of orthotopic


ileal neobladder with serous-lined extramural ureteral
reimplantation: Experience with 450 patients. JUrol 2001;
165:1427-32.

68. Mills RD, Studer UE: Metabolic consequences of continent


urinary diversion. JUrol 1999; 161:1057-1066.
69. Zehnder P, Dhar N, Thurairaja R, Ochsner K, et al.: Efect
of urinary tract infection o reservoir function in patients with
ileal bladder substitute. JUrol 2009; 181:2545-2549.

87. Hautmann RE, Abol-Enein H, Hafez K, et al.: World Health


Organization (WHO) Consensus Conference on Bladder
Cancer: Urinary diversion. Urology 2007; 69 (1 Suppl): 1749.

70. Waidelich R, Rink F, Kriegmair, et al.: A study of reflux in


patients with an ileal orthotopic bladder. BrJUrol 1998;
81:241-246.

88. Stein JP, Esrig D, Freeman JA, et al.: Prospective pathologic


analysis of female cystectomy specimens: Risk factors for
orthotopic diversion in women. Urology 1998; 51:951-5.

71. Nesbit RM: Ureterosigmoid anastomosis by direct elliptical


connection: a preliminary report. JUrol 1949; 61:728-734.

89. Stenzl A, Colleselli K, Poisel S, et al.: Anterior exenteration


with subsequent ureteroileal urethrostomy in females.
Anatomy, risk of urethral recurrence, surgical technique,
and results. EurUrol 1998; 33:18-20.

72. Studer UE, Burkhard FC, Schumacher M, et al.: Twenty


years experience with an ileal orthotopic low pressure
bladder substitute lessons to be learned. JUrol 2006;
176:161-166.

90. Tanagho EA, Miller ER, Meyers FH, et al.: Observations on


the dynamics of he bladder neck. BrJUrol 1966; 38:72.
91. Hubner WA, Trigo-Rocha F, Plas EG, et al.: Functional
bladder replacement after radical cystectomy in the female:
experimental investigation of a new concept. EurUrol 1993;
23:400-4.

73. Mattei A, Birkhaeuser FD, Baermann C, et al.: To stent or


not to stent perioperatively the ureteroileal anastomosis
of ileal orthotopic bladder substitutes and ileal conduits?
Results of a prospective randomized trial. JUrol 2008;
179:582-586.

92. Colleselli K, Stenzl A, Eder R, Strasser H, et al.: The female


urethral sphincter: A morphological and topographical
study. JUrol 1998; 160:49-54.

74. Studer UE, Danuser H, Thalmann GN, et al.: Antireflux


nipples or afferent tubular segments in 70 patients with
ileal low pressure bladder substitutes: longterm results of a
prospective randomized trial. JUrol 1996; 156:1913-1917.

93. Bhatta DN, Kessler TM, Mills RD, et al: Nerve-sparing


radical cystectomy and orthotopic bladder replacement in
female patients. EurUrol 2007; 52:1006-14.

75. Shaaban AA, Abdel-Latif M, Mosbah A, et al.: A randomized


study comparing an antireflux system with a direct ureteric
anastomosis in patients with orthotopic ileal neobladders.
BJU Int 2006; 97:1057-1062.

94. Borirakchanyavat S., Aboseif SR, Carroll PR, et al.:


Continence mechanism of the isolated female urethra: an
anatomical study of the intrapelvic somatic nerves. JUrol
1997; 158:822-26.

76. Dhar NB, Hernandez AV, Reinhardt K, et al.: Prevalence


of nephrolithiasis in patients with ileal bladder substitutes.
Urology 2008; 71:128-130.

95. Stenzl A., Colleselli K, Poisel S, et al.: Rationale and


technique of nere sparing radical cystectomy before an
orthotopic neobladder procedure in women. JUrol 1995;
154:2044-49.

77. Thurairaja R, Burkhard FC, Studer UE: The orthotopic


neobladder. BJU Int 2008; 102:1307-1313.

372

96. Stein JP, Stenzl A, Esrig D, et al: Lower urinary tract


reconstruction following cystectomy in women using the
Kock ileal reservoir with bilateral ureteroileal urethrostomy:
Initial clinical experience. JUrol 1994; 152:1404-8.

116. Dhar NB, Kessler TM, Mills RD, et al.: Nerve-sparing


radical cystectomy and orthotopic bladder replacement in
female patients. EurUrol 2007; 52:1006-14.
117. Stein JP, Penson DF, Wu SD, et al.: Pathological guidelines
for orthotopic urinary diversion in women with bladder
cancer: a review of the literature. JUrol 2007; 178:756-60.

97. Ali-El-Dein B, Gomha M, Ghoneim MA: Critical evaluation


of the problem of chronic urinary retention after orthotopic
bladder substitution in women. JUrol 2002; 168:587-92.

118. Volkmer BG, Gschwend JE, Herkommer K: Cystectomy


and orthotopic ileal neobladder: the impact on female
sexuality. JUrol 2004; 172:2353-7.

98. Grossfeld GD, Stein JP, Bennett CJ, et al.: Lower urinary
tract reconstruction in the female using the Kock ileal
reservoir with bilateral uretoileal urethrostomy: update of
continence results and fluorourodynamic findings. Urology
1996; 48:383-88.

119. Zippe C, Nandipati KC, Agarwal A, et al.: Female sexual


dysfunction after pelvic surgery: the impact of surgical
modifications. BJU Int 2005; 96:959-63.

99. Park JM, Montie JE: Mechanisms of incontinence and


retention after orthotopic neobladder diversion. Urology
1998; 51:601-9.

120. Bhatt A, Nandipati K, Dhar N: Neurovascular preservation


in orthotopic cystectomy: impact on female sexual function.
Urology 2006; 67:742-5.

100. Turner WH, Danuser H, Moehrle K, et al.: The effect of


nerve sparing cystectomy technique on postoperative
continence after orthotopic bladder substitution. JUrol
1997; 158:2118-22.

121. Ali-El Dein B, Shaaban A: Assessment of female sexual


dysfunction following radical cystectomy and urinary
diversion. JUrol 2008; 179(4 supp):299A.
122. Hart S, Skinner Ec, Meyerowitz BE, Boyd SD, et al.:
Quality of life after radical cystectomy for bladder cancer
in patients with an ileal conduit or cutaneous or urethral
Kock pouch. JUrol 1999; 162:77-81.

101. Stein JP, Grossfeld GD, Freeman JA, et al.: Orthotopic


lower urinary tract reconstruction in women using the Kock
ileal neobladder: Update experience in 34 patients. JUrol
1997; 158:400-5.

123. Gerharz EW, Mansson A, Hunt S, et al.: Quality of life


after cystectomy and urinary diversion: an evidence based
analysis. JUrol 2005; 174:1729-36.

102. Ali-el-Dein B, Shaaban AA, Abu-Eideh RH, et al.: Surgical


complications following radical cystectomy and orthotopic
neobladder in women. JUrol 2008; 180:206-10.

124. Gerharz EW: Is there any evidence that one continent


diversion is any better than any other or than ileal conduit?
CurrOpinion in Urology 2007; 17:402-7.

103. Hautmann RE, Petriconi RC, Volkmer BG: Lessons Learned


From 1,000 Neobladders: The 90-Day Complication Rate.
JUrol 2010; 184:990-4.

125. Kock NG, Nilson AE, Nilsson LO, et al.: Urinary diversion
via a continent ileal reservoir: clinical results in 12 patients.
JUrol 182; 128:469-475.

104. Rapp DE, OConnor RC, Katz EE, et al.: Neobladdervaginal fistula after cystectomy and orthotopic neobladder
construction. BJU Int 2004; 94:1092.

126. Skinner DG, Lieskovsky G, Boyd SD: Continuing


experience with the continent ileal reservoir (Kock pouch)
a san alternative to cutaneous urinary diversion: An update
after 250 cases. JUrol 1987; 137:1140-1145.

105. Granberg CF, Boorjian SA, Crispen PL, et al.: Functional


and oncological outcomes after orthotopic neobladder
reconstruction in women. BJU Int 2008; 102:1551-5.
106. Nesrallah LJ, Almeid FG, Dall`oglio MF, et al.: Experience
with the orthotopic ileal neobladder in women: mid-term
follow-up. BJU Int 2005; 95:1045-7.

127. Jonsson O, Dahlstrand C, Edlund C, et al.: Afferent


intussuscepted antireflux nipple valve complications in the
Kock pouch for continent urinary diversion. Early results
with a modified technique. ScandJUrol Nephrol 2006;
40:472-478.

107. Stenzl A, Jarolim L, Coloby L, et al.: Uretra-sparing


cystectomy and orthotopic urinary diversion in women with
malignant pelvic tumors. Cancer 2001; 92:1864-71.

128. SoulieM, Seguin P, Martel P,


et al.: A modified
intussuscepted nipple valve in the Kock pouch urinary
diversion: assessment of perioperative complications and
functional results.

108. Hautmann RE, Petriconi R, Kleinschmidt K, et al.:


Orthotopic ileal neobladder in females: impact of the
urethral resection line on functional results. Int Urogynecol
J PelvicFloor Dysfunc 2000; 11:224-29.

109. Hautmann RE, Paiss TH, Petriconi R: The ileal neobladder


in women: 9 years of experience with 18 patients. JUrol
1996; 155:76-81.

BJU Int 2002; 90:397-402.

129. Throff, JW, Alken P, Riedmiller H, et al.: The Mainzpouch (mixed augmentation ileum and cecum) for bladder
augmentation and continent diversion. JUrol 1986; 136:1726.

110. Stein JP, Dunn MD, Quek ML, et al.: The orthotopic T
pouch ileal neobladder: Experience with 209 patients.
JUrol 2004; 172:584-7.
111. Lee KS, Montie JE, Dunn RL, et al.: Hautmann and Studer
orthotopic neobladders: a contemporary experience. JUro
2003; 169:2188-91.
112. Nagele U, Kuczyk M Anastasiadis G, et al.: Radical
cystectomy and orthotopic bladder replacement in
females. EurUrol 2006; 50:249-57.
113. Hugonnet CL, Danuser H, Springer JP, et al.: Decreased
sensitivity in the membranous urethra after orthotopic ileal
bladder substitution. JUrol 1999; 161:418-21.
114. Kessler TM, Studer UE, Burkhard FC: Increased proximal
urethral sensory threshold after radical pelvic surgery in
women. Neurourol & Urodynamics 2007; 26:208-12.
115. Studer UE, Burkhard FC, Schumacher M, et al.: Twenty
years experience with an ileal orthotopic low pressure
bladder substitute lessons to be learned. JUrol2006;
176:161-66.

130. Throff JW, Alken P, Riedmiller H, et al.: 100 cases of


Mainz pouch: continuing experience and evolution. JUrol
1988; 140:283-288.
131. Wiesner C, Stein R, Pahernik S, et al.: Long-term
followup of the intussuscepted ileal nipple valve and the
in situ, submucosally embedded appendix as continence
mechanisms of continent urinary diversion with the
cutaneous ileocecal pouch (Mainz pouch I). JUrol 2006;
176:155-160.
132. Stein P, Daneshmand S, Dunn M, et al.: Continent right
colon reservoir using a cutaneous appendicostomy.
Urology 2004; 63:577-580.
133. Bochner BH, Karanikolas N, Barakat RR, et al.:
Ureteroileocecal appendicostomy based urinary reservoir
in irradiated and nonirradiated patients. JUrol 2009;
182:2376-2380.

373

134. Holmes DG, Trasher JB, Park GY, et al.: Long-term


complications related to the modified Indiana pouch.
Urology 2002; 60:603-606.

154 Chechile G, Klein EA, Bauer L, et al.: Functional


equivalence of end and loop ileal conduit stomas. JUrol
1992; 147:582-6.

135. Nieuwenhuijzen JA, de Vries RR, Bex A, et al.: Urinary


diversion after cystectomy: the association of clinical
factors, complications and functional results of four
different diversions. EurUrol 2008; 53:834.

155. Spirnak JP, Caldamone AA: Ureterosimoidostomy. Urol


Clin NorthAm 1986; 13:285-294.
156. Richie JP, Skinner DG: Ureterointestinal diversion in
Walsh PC: Cambells Urology 5th ed. Philadelphia. W.B.
Saunders 1986; 2601-2619.

136. Mansson W, Davidsson T, Knyves T, et al.: Continent


urinary tract reconstruction Lund experience. BJU Int
2003; 92:271-276.

157. McDougal WA: Metabolic complications of urinary


intestinal diversion. JUrol 1992; 147:1199-1208.

137. Marino G, Laudi M: Ileal T-pouch as a urinary continent


cutaneous diversion:: clinical and urodynamic evaluation.
BJU Int 2002; 90:47-50.

158. Bricker EM: Bladder substitution after pelvic evisceration.


SurgGynObstet 1950; 30:1511-1521.
159. Wear JB Jr, Barquin OP: Ureterosigmoidostomy. Longterm-results. Urology 1973; 1:192-200.

138. Seifert H-G, Obaje A, Mller-Mattheis V, Mller M, et al.:


Clinical and functional results after continent cutaneous
urinary diversion with the double-T-pouch. Urol Int 2008:
80:8-12.

160. Truss MC: Supravesical diversion. First Edition ed. vol


Chap. 49. Philadelphia, Lippincott Comp. 1994.
161. Koo HP, Avolio L, Duckett JW Jr: Long-term results of
ureterosigmoidostomy in children with bladder exstrophy.
JUrol 1996; 156:2037-2040.

139. Beiko DT, Razvi H: Stones in urinary diversions update on


medical and surgical issues. Curr Opin Urol 2002; 12:297303.

162. Hirokawa M. Iwasaki A, Yamazaki A, et al.: Improved


technique of tubeless cutaneous ureterostomy and results
of permanent urinary diversion. EurUrol 1989; 16:125132.

140. Okada Y, Shichiri Y, Terai A, et al.: Management of late


complications of continent urinary diversion using the
Kock pouch and the Indiana pouch procedures. Int J Urol
1996; 3:334-339.

163. Nesbit R: Ureterosigmoid anastomosis by direct elliptical


correction. JUrol 1949; 61:728-734.

141. Mansson W, Bakke A, Bergmann B, et al.: Perforation of


continent urinary reservoirs.

164. Yoshimura K, Maekawa S, Ichioka K, et al.: Tubeless


cutaneous ureterostomy: the Toyoda method revisited.
JUrol 2001; 165:785-788.

Scandinavian experience. Scand J Urol Nephrol 1997;


31:529-53.

142. Wullt B, Agace W, Mansson W: Bladder, bowel and bugs


bacteriuria in patients with intestinal urinary diversion.
World J Urol 2004; 22:186-195.

165. Allen TD: Salvaging the obstructed ureterosigmoidostomy


using an ileal interposition technique. JUrol 1993;
150:1195-1198.

143. Jahnson S, Damm O, Hellsten S, et al.: Urinary diversion


after cystectomy for bladder cancer: a population-based
study in Sweden. Scand J Urol Nephrol.; 44:69-75.

166. Coffey RC: Physiologic implantation of the severed


ureter or common bile duct into the intestine. JAMA 1911;
56:397.

144. Gore JL, Saigal CS, Hanley JM, et al.: Variations in


reconstruction after radical cystectomy. Cancer 2006;
107:729-37.
145. Kanofsky JA,Godoy G, Taneja SS: Complications of
Urologic Surgery. In Taneja SS ed, 4 edn, Philadelphia
Saunders Elsevier 2010; 533-46.
146. Dahl DM, McDougal WS: Use of Intestinal Segments in
Urinary Diversion. Campbell-Walsh Urology, Philadelphia
Saunders Elsevier 2007; 2534-78.
147. Yong SM, Dublin N, Pickard R, et al.: Urinary diversion
and
bladder
reconstruction/
replacement
using
intestinal segments for intractable incontinence or
following cystectomy. Cochrane Database SystRev.
2009:CDOO3306.
148. Somani BK, Nabi G, Wong S, et al.: How close are we to
knowing whether orthotopic bladder replacement surgery
is the new gold standard? evidencre from a systematic
review update. Urology 2009; 74:1331-9.
149. Wallace DM: Ureteric diversion using a conduit: a simplified
technique. Br J Urol 1966; 38:522-7.
150. Nichols WK, Krause AH, Donegan Wl. Urinary fistulas
after ureteral diversion. A J Surg. 1972; 124:311-6.
151. Koehler PR, Bowles WT, McAlister WH: Roentgenographic
evaluation of late results of ileal loop urinary diversion in
infants and children. Am J Roentgenol Radium Ther Nucl
Med. 1967; 100:177-85.
152. Evangelidis A, Lee EK, Karellas ME: Evaluation of
ureterointestinal anastomosis: Wallace vs Bricker. J Urol
2006; 175:155-8; discus.8.
153. Kouba E, Sands M, Lentz A, et al.: A comparison of
the Bricker versus Wallace ureteroileal anastomosis in
patients undergoing urinary diversion for bladder cancer.
JUrol 2007;178:945-8; discus.8-9.

167. Hendren WH: Reconstruction of previously diverted


urinary tracts in children. JPediatrSurg 1973; 8:135-150.
168. Leadbetter WF: Considerations of problems incident to
performance of ureteroenterostomy: Report of a technique.
TransAmAssocGenitourinSurg 1950; 42:39.
169. Goodwin WE, Harris AP, Kaufman JJ: Open transcolonic
ureterointestinal anastomosis. SurgGynecolObstet 1953;
97:295-300.
170. Williams
DF,
Burkholder
GV,
Goodwin
WE:
Ureterosigmoidostomy: a 15-year experience. JUrol 1969;
101:168-170.
171. Zincke H, Segura JW: Ureterosigmoidostomy: critical
review of 173 cases. JUrol 1975; 113:324-327.
172. Duckett
JW,
Gazak
JM:
Complications
of
ureterosigmoidostomy. UrolClinNorthAm 1983; 10:473481.
173. Connor IP, Hensle TW, Lattimer JK, et al.: Long-term followup of 207 patients with bladder exstrophy: an evolution in
treatment. JUrol 1989; 142(3):793-5; discussion 795-6.
174. Stckle M, Becht E, Voges G, et al.: Ureterosigmoidostomy:
an outdated approach to bladder exstrophy? JUrol 1990;
143(4):770-4; discussion 774-5.
175. Zabbo A, Kay R: Ureterosigmoidostomy and bladder
exstrophy: a long-term follog-up. JUrol 1986; 136:396398.
176. Wynant HP, Morelle V, Fonteyne E, et al.:
Ureterosigmoidostomy: reevaluation of a forgotten
continent urinary diversion. ActaUrolBelg 1991;
59(4):103-7.
177. Ghoneim MA, Shehab-El-Din AB, Ashamallah AK:
Evolution of the rectal bladder as a method for urinary
diversion. JUrol 1981; 126:737-740.

374

178. Fisch M, Wammack R, Muller SC: The Mainz pouch II


(sigma rectum pouch). JUrol 1993: 149:258-263.

198. Fisch M, Wammack R, Muller SC, et al.: The Mainz pouch


II (sigma rectum pouch). JUrol 1993; 149(2):258-63.

179. Mathisen W: A new method of ureterointestinal


anastomosis: A preliminary report. SurgGynecolObstet
1953; 96:255-258.

199. Stein R, Fisch M, Stockle M, et al.: Colonic conduit in


children: protection of the upper urinary tract 16 years
later? JUrol 1996; 156(3):1146-50.

180. Fisch M, Wammack R, Hohenfellner R: The sigma rectum


pouch (Mainz pouch II). WorldJUrol 1996; 14:68-72.

200. Hunt GM, Poulton A: Open spina bifida: A complete cohort


reviewed 25 years after closure. DevMedChildNeurol
1995; 37:19-29.

181. Stein R, Fisch, Beetz R: Urinary diversion in children and


young adults using the Mainz Pouch I technique. BrJUrol
1997; 79:354-361.

201. Mitrofanoff P: Cystostomie continente trans-appendiculaire


dans le traitement des vessies neurologiques. ChirPediat
1980; 21:297-305.

182. Gerharz EW, Kohl UN, Weingartner K: Experience with


the Mainz modification of ureterosigmoidostomy. BrJSurg
1998; 85:1512-1516.

202. Riedmiller H, Burger R, Muller S, et al.: Continent appendix


stoma: a modification of the Mainz pouch technique. JUrol
1990; 143(6):1115-7.

183. Obek C, Kural AR, Ataus S: Complications of the Mainz


pouch II (sigma rectum pouch). EurUrol 2001; 39:204211.

203. Roth S, Weining C, Hertle L: Simplified uretero-intestinal


implantation in continent urinary diversion using
ileovalvular segments as afferent loop and appendix as
continent outlet. JUrol 1996; 155:1200-4.

184. Schultzel M, Saltzstein SL, Downs TM, et al.: Late age (85
years or older) incidence of bladder cancer. JUrol 2008;
179:1302-1306.

204. Throff JW, Franzaring L, Gillitzer R, et al.: Simplified


orthotopic ileocaecal pouch (Mainz pouch) for bladder
substitution. BJU Int. 2005; 96(3):443-65.

185. Statistik
Austria:
http://www.statistik.at/web-de/
statistiken/bevoelkerung/demographische prognosen/
bevoelkerungsprognosen/027331.html.

205. Wiesner C, Pahemik S, Stein R, et al.: Long-term follow-up


of submucosal tunnel and serosa-lined extramural tunnel
ureter implantation in ileocaecal continent cutaneous
urinary diversion (Mainz pouch I). BJU Int. 2007;
100(3):633-7.

186. Wood DP, Montie JE, Maatman TJ, et al.: Radical


cystectomy for carcinoma of the bladder in the elderly
patient. JUrol 1987; 138:46-48.
187. Stroumbakis N, Herr HW, Cookson MS, et al.: Radical
cystectomy in the octogenarian. JUrol 1997; 158:21132117.

206. Bogaert GA, Mevorach RA, Kogan BA: Urodynamics and


clinical follow-up of 28 children after gastrocystoplasty.
BritishJUrol 1994; 74:469.

188. Madersbacher S, Hochreiter W, Burkhard F, et al.: Radical


cystectomy for bladder cancer today a homogeneous
series without neoadjuvant therapy. JClinOncol 2003;
21:690-696.

207. Nguyen DH, Bain MA, Salmonson KL, et al.: The syndrome
of dysuria and hematuria in pediatric urinary reconstruction
with stomach. JUrol 1993; 150:707-9.
208. Vemulakonda VM, Lendvay TS, Shnorhavorian M, et al.:
Metastatic adenocarinoma after augmentation gastrocystoplasty. JUrol 2008; 179(3):1094-6; discussion 7.

189. Chamie K, Hu B, de Vere White R, et al.: Cystectomy in the


elderly: does survival benefit in younger patients translate
to the octogenarians? BJUI 2008; 102:284-290.

209. Castellan M, Gosalbez R, Perez-Brayfield M, et al.:


Tumor in Bladder Reservoir After Gastrocystoplasty. JUrol
2007 Aug 16.

190. Saika T, Arata R, Tsushima t; ET AL:.Health-related quality


of life after radical cystectomy for bladder cancer in elderly
patients with an ileal conduit, ureterocutaneostomy, or
orthotopic urinary reservoir: a comparative questionnaire
survey. ActaMedOkayama 2007; 61:199-203.
191. Yamanaka K, Miyake H, Hara I, et al.: Significance of
radical cystectomy for bladder cancer in patients over 80
years old. Int UrolNephro. 2007; 39:209-14.
192. Ubrig B, Roth S: Palliative urinary diversion. Glenns
Urologic Surgery, Lippincott Williams & Wilkins 2009; 94:
618.
193. Aravantinos E, Anagnostou T, Karatzas AD, et al.:
Percutaneous nephrostomy in patients with tumors of
advanced stage: treatment dilemmas and impact on
clinical course and quality of life. JEdourol 2007; 21:12971302.
194. Desgrandchamps F, Cussenot O, Meria P, et al.:
Subcutaneous urinary diversions for palliative treatment of
pelvic malignancies. JUrol 1995; 154:367-370.
195. Desgrandchamps F, Leroux S, Ravery V, et al.:
Subcutaneous pyelovesical bypass as replacement for
standard percutaneous nephrostomy for palliative urinary
diverion: prospective evaluation of patientss quality of
life. JEndourol 2007; 21:173-176.
196. Lloyd SN, Tirukonda P, Biyani CS, et al.: The detour
extraanatomic stent a permanent solution for benign and
malignant ureteric obstruction? EurUrol 2007; 52:193198.
197. Stein R, Fisch M, Ermert A, et al.: Urinary diversion and
orthotopic bladder substitution in children and young adults
with neurogenic bladder: a safe option for treatment? JUrol
2000; 163(2):568-73.

210. Baydar DE, Allan RW, Castellan M, et al.: Anaplastic signet


ring cell carcinoma arising in gastrocystoplasty. Urology.
2005; 65(6):1226.
211. Husmann DA, Rathbun SR: Long-term follow up of
enteric bladder augmentations: the risk for malignancy.
JPediatrUrol 2008; 4(5):381-5; discussion 6.
212. Eckstein HB, Chir M, Martin R: Uretero-Cystoplastik.
AktUrol 1973; 4:255-7.
213. Youssif M, Badawy H, Saad A, et al.: Augmentation
ureterocystoplasty in boys with valve bladder syndrome.
JPediatrUrol 2007; 3(6):433-7.
214. Husmann DA, Snodgrass WT, Koyle MA, et al.:
Ureterocystoplasty: indications for a successful
augmentation. JUrol 2004; 171(1):376-80.
215. Churchill BM, Jayanthi VR, Landau EH, et al.:
Ureterocystoplasty: importance of the proximal blood
supply. JUrol 1995; 154(1):197-8.
216. Bellinger MF: Ureterocystoplasty: a unique method for vesical augmentation in children. JUrol 1993; 149(4):811-3.
217. Kibar Y, Roth CC, Asley RA, et al.: Evaluation of need for
salvage continence procedures after failed modern staged
repair. Urology 2010; 76(1):39-42.
218. Cervellione RM, Bianchi A, Fishwick J, et al.: Salvage
procedure sto achieve continence after failed bladder
exstrophy repair. JUrol 2008; 179(1):304-6.
219. Woodhouse CR, North AC, Gearhart JP: Standing the
test of time: long-term outcome of reconstruction of the
exstrophy bladder. WorldJUrol 2006 Mar 4.

375

220. Gearhart JP, Peppas DS, Jeffs RD: The application of


continent urinary stomas to bladder augmentation or
replacement in the failed exstrophy reconstruction. BrJUrol
1995; 75(1):87-90.

for directing the orifices of the ureter into the rectum;


temporary success; subsequent death; autopsy. Lancet
1852; 2:568-70.
243. Bennett AH: Exstrophy of bladder treated by ureterosigmoidostomies: long-term evaluation. Urol 1973; 2:165-8.

221. Mouriquand PD, Bubanj T, Feyaerts A, et al.: Long-term


results of bladder neck reconstruction for incontinence in
children with classical bladder exstrophy or incontinent
epispadias. BJU Int 2003;92(9):997-1001; discussion 2.

244. Segura JW, Kelalis PP: Long-term results of


ureterosigmoidostomy in children with bladder exstrophy.
JUrol 1975; 114(1):138-40.

222. Woodhouse CRJ, Redgrave NG: Late failure of the


reconstructed exstrophy bladder. BJU 1996; 77:590-2.
223. Kock NG: Continent ileostomy. ProgSurg 1973; 12:180201.

245. Gobet R, Weber D, Renzulli P, et al.: Long-term follow up


(37-69 years) of patients with bladder exstrophy treated
with ureterosigmoidostomy: uro-nephrological outcome.
JPediatrUrol 2009; 5(3):190-6.

224. Kock NG, Nilson AE, Norlen LJ, et al.: Urinary diversion
via a continent ileum reservoir: Clinical experience.
JUrolNephrol Suppl. 1978; 49:2j3-31.

246. Silverman SH, Woodhouse CR, Strachan JR, et al.:


Long-term management of patients who have had urinary
diversions into colon. BJU 1986; 58(6):634-9.

225. Leisinger HJ, Suberli H, Schauwecker H, et al.: Continent


ileal bladder: First clinical experience. EurUrol 1976; 2:812.

247. Creevy CD: The Disadvantage of Ureterosigmoidostomy.


CanadMAJ 1954; 70:172-5.
248. Clarke BG, Leadbetter WF: Ureterosigmoidostomy:
Collective review of results in 2897 reported cases. JUrol
1955; 73:999-1008.

226. Madigan MR: The continent ileostomy and the isolated


ileal bladder. AnnRoyCollSurgEng 1976; 58:62-9.
227. Skinner DG, Boyd SD, Lieskovsky G: Clinical experience
with the Kock continent ileal reservoir for urinary diversion.
JUrol 1984; 132(6):1101-7.

249. Zabbo A, Kay R: Ureterosigmoidostomy and bladder


exstrophy: a long-term followup. JUrol 1986; 136(2)396-8.
250. Mesrobian HG, Kelalis PP, Kramer SA: Long-term followup
of 103 patients with bladder exstrophy. JUrol 1988;
139:719-22.

228. Mansson W: The continent caecal reservoir for urine.


ScandJUrolNephrol Suppl 1984; 85:1-137.
229. Rowland RG, Mitchell ME, Bihrle R: The cecoileal continent
urinary reservoir. WorldJUrol 1985; 3:185-90.

251. Koo HP, Avolio L, Duckett JW, Jr: Long-term results of


ureterosigmoidostomy in children with bladder exstrophy.
JUrol 1996; 156(6):2037-40.

230. Thuroff JW, Alken P, Engelmann U, et al.: the Mainz


pouch (mixed augmentation ileum n zecum) for bladder
augmentation and continent urinary diversion. EurUrol
1985; 11(3):152-60.

252. Stckle M, Becht E, Voges G, et al.: Ureterosigmoidostomy:


an outdated approach to bladder exstrophy? JUrol 1990;
143(4):770-4; discussion 4-5.

231. Ashken MH: An appliance-free ilecaecal urinary diversion:


Preliminary communication. BJU 1974; 46:631.

253. Mteta KA, Mbwambo JS, Eshleman JL, et al.: Urinary


diversion in children with mainly exsdtrophy and epispadias:
alternative to primary bladder closure. CentAfrJMed 2000;
46(12):318-20.

232. Benchekroun A: Continent caecal bladder. EurUrol 1977;


3:248-50.

254. Stein R, Fisch M, Stckle M, et al.: Urinary Diversion in


Bladder Exstrophy and Incontinent Epispadias: 25 years
of experience. JUrol 1995; 154:1177-81.

233. Gallo AG: A Propsito de operacin de triana: nueva


tcnica de exclusion de la vejiga utilizando el ciego
aislando como receptacullo urinario. Bol Trab Acad
Argentina Cirujanos 1946; 30:604-10.

255. Pahernik S, Beetz R, Schede J, et al.: Rectosigmoid pouch


(Mainz Pouch II) in children. JUrol 2006; 175(1):284-7.

234. Makkas M: Zur Behandlung der Blasenektopie.


Umwandlung des ausgeschalteten Coecum zur Blase und
der Appendix zur Urethra. Zentralblatt fr Chirurgie 1910;
37:1073-6.

256. Abol-Enein H, Ghoneim MA: A novel uretero-ileal


reimplantation technique: the serous lined extramural
tunnel. A preliminary report. JUrol 1994; 151(5):1193-7.

235. Verhoogen J: Nostomie urtro-caecale. Formation


dune nouvelle poch vsicale et dun nouvel urtre.
AssFranc dUrol 1908; 12:362-5.

257. Fisch M, Abol Enein H, Hohenfellner R: Ureterimplantation


mittels sersem extramuralem Tunnel in Mainz-Pouch
I und Sigma-Rektum-Pouch (Mainz-Pouch II). AktUrol
1995; 26:I-X.

236. Leissner J, Black P, Fisch M, et al.: Colon pouch (Mainz


pouch III) for continent urinary diversion after pelvic
irradiation. Urol 2000; 56(5):798-802.

258. Kock NG, Ghoneim MA, Lycke KG, et al.: Urinary diversion
to the augmented and valved rectum: preliminary results
with a novel surgical procedure. JUrol 1988; 140:1375-9.

237. Bihrle R, Foster RS, Steidle CP, et al.: Creation of a colongastric composite reservoir: a new technique. JUrol 1989;
141:1217-20.

259. el Mekresh MM, Hafez AT, Abol-Enein H, et al.: Double


folded rectosigmoid bladder with a new ureterocolic
antireflux technique. JUrol 1997; 157(6):2085-9.

238. Fisch M, Wammack R, Hohenfellner R: Kolon-transversumascendens/descendens Pouch. AktUrol 1996; 27:I-VIII.

260. Hafez AT, Elsherbiny MT, Dawaba MS, et al.: Long-term


outcome analysis of low pressure rectal reservoirs in 33
children with bladder exstrophy. JUrol 2001; 165/6 Pt
2):2414-7.

239. Stein R, Wiesner C, Beetz R, et al.: Urinary diversion in


children and adolescents with neurogenic bladder: the
Mainz experience. PediatrNephrol 2005; 20(7):926-31.
240. Monti PR, Lara RC, Dutra MA, et al.: New technique for
construction of efferent conduits based on the Mitrofanoff
principle. Urol 1997; 49:112-5.

261. Bricker EM: Bladder substitution after pelvic evisceration.


SurgClinNAmer 1950; 30:1511-20.
262. Schoemaker J: Discussie op voordracht van JM. van Dam
over intra-abdominale plastieken. NedTGeneesk 1911;
55:836-7.

241. Yang WH: Yang needle tunnelling technique in creating


antireflux and continent mechanisms. JUrol 1993;
150(3):830-4.

263. Seifert L: Indikation und Operationstechnis der DarmSiphon-Blase, ihrer Leistungs-fhigkeit in Rntgenbild
und Film. Zeitschr f Urologie 1937; 1:23-33.

242. Simon J: Ectopia vescia (Absence of the anterior walls


of the bladder and pubic adominal parities); operation

376

264. Bricker EM, Eiseman B: Bladder reconstruction from


cecum and ascending colon following resection of pelvic
viscera. AnnSurg 1950; 132:77-84.

287. Abd-el-Gawad G, Abrahamsson K, Hanson E, et al.: Kock


urinary reservoir maturation in children and adolescents:
consequences for kidney and upper urinary tract. EurUrol
1999; 36(5):443-9.

265. Zinman L, Libertino JA: Ileocecal conduit for temporary and


permanent urinary diversion. JUrol 1975; 113:317-23.

288. Abd-el-Gawad G, Abrahamsson K, Hanson E, et al.:


Evaluation of Kock urinary reservoir function in children and
adolescents at 3-10 yearsfollow-up. ScandJUrolNephr
1999; 33(3):149-55.

266. Nelson JH: Atlas of radical pelvic surgery. New York:


Appleton centura-Crofts 1969.
267. Schmidt JD, Hawtrey CE, Buchsbaum HJ: Transverse
colon conduit: a preferred method of urinary diversion
for radiation-treated pelvic malignancies. JUrol 1975;
113(3):308-13.

289. Le Duc A, Camey M: Un procd d`implantation ureteroileal antireflux dans lenterocystoplastie. JUrol 1979;
85:449.
290. Fallace DM: Ureteric diversion using a conduit: a simplified
technique. BJU 1966; 38(5):522-7.

268. belhr R: Die Darmblase. Langenbecks Ar u DtschZChir


1952; 1952:271.

291. Nesbit RM, Bohne AW: Ureterosigmoid anastomosis


by direct elliptic connection. NY State JMed 1949;
49/24):2933-7, illust.

269. Mogg RA: Treatment of the neurogenic urinary incontinence


using the colonic conduit. BJU 1965; 37:681.
270. Madersbacher S, Schmidt J, Eberle JM, et al.: Long-term
outcome of ileal conduit diversion. JUrol 2003; 169(3):98590.

292. Goodwin WE, Harris AP, Kaufmann JJ, et al.: Open


transcolic ureterointestinal anastomosis: a new approach.
Surg Gynecol Obstet1953; 97:295.

271. Althausen AF, Hagen-Cook K, Hendren WH 3rd: Nonrefluxing colon conduit: experience with 70 cases. JUrol
1978; 120(1):35-9.

293. Abol EH, Ghoneim MA: A novel uretero-ileal reimplantation


technique: the serous lined extramural tunnel. A preliminary
report. JUrol 1994; 151:1193-7.

272. Graham AG: Long-term results of ileal conduit diversion in


children a brighter picture? BJU 1982; 54(6):632-4.

294. Shaaban AA, Abdel-Latif M, Mosbah A, et al.: A randomized


study comparing an antireflux system with a direct ureteric
anastomosis in patients with orthotopic ileal neobladders.
BJU Int. 2006; 97:1057-62.

273. Middleton AW, Hendren WH: Ileal conduit in children at the


Massachusetts General Hospital from 1955 to 1970. JUrol
1976; 115:591.

295. Perimenis P, Studer UE: Orthotopic continent urinary


diversion an ileal low pressure neobladder with an afferent
tubular segment: how I do it. EurJSurgOncol 2004;
30:454-9.

274. Shapiro SR, Lebowitz R, Colodny AH: Fate of 90 children


with ileal conduit urinary diversion a decade later:
analysis of complications, pyelography, renal function and
bacteriology. JUrol 1975; 114(2)289-95.

296. Webster C, Bukkapatnam R, Seigne JD, et al.: Continent


colonic urinary reservoir (Florida pouch): long-term
surgical complications (greater than 11 years). JUrol 2003;
169:174-6.

275. Steiner MS, Morton RA: Nutritional and gastrointestinal


complications of the use of bowel segments in the lower
urinary tract. UrolClinNorthAm 1991; 18/4):743-54.
276. Fontaine E, Barthelemy Y, Houlgatte A, et al.: Twenty-year
experience with jejunal conduits. Urol 197; 50(2) 207-13.

297. Holmes DG, Thrasher JB, Park GY, et al.: Long-term


complications related to the modified Indiana pouch. Urol
2002; 60:603-6.

277. Gilbert SM, Hensle TW: Metabolic consequences and


long-term complications of enterocystoplasty in children:
a review. JUrol 2005; 173(4):1080-6.

298. Mansson W, Davidsson T, Konyves J, et al.: Continent


urinary tract reconstruction the Lund experience. BJU
Int. 2003; 92:271-6.

278. Schwarz GR, Jeffs RD: Ileal conduit urinary diversion in


children: computer analysis of followup from 2 to 16 years.
JUrol 1975; 114:285.

299. Abol-Enein H, Ghoneim MA: functional results of orthotopic


ileal neobladder with serous-lined extramural ureteral
reimplantation: experience with 450 patients. JUrol 2001;
165:1427-32.

279. Hendren WH: Nonrefluxing colon conduit for temporara or


permanent urinary diversion in children. JPedSurg 1975;
10:381.

300. Helal M, Pow-Sang J, Sanford E, et al.: Direct (nontunneled)


ureterocolonic reimplantation in association with continent
reservoirs. JUrol 1993; 150:835-7.

280. Elder DD, Moisey CU, Rees RW: A long-term follow-up


of the colonic conduit operation in children. BJU 1979;
51(6):462-5.

301. Pantuck AJ, Han Kr, Perrotti M, et al.: Ureteroenteric


anastomosis in continent urinary diversion: long-term
results and complications of direct versus nonrefluxing
techniques. JUrol 2000; 163:450-5.
302. Hohenfellner R, Black P, Leissner J, et al.: Refluxing ureterointestinal anastomosis for continent cutaneous urinary
diversion. JUrol 2002; 168:1013-6; discussion 6-7.
303. Wiesner C, Stein R, Pahernik S: Long-term folloup of the
intussuscepted ileal nipple and the in situ, submucosally
embedded appendix as continence mechanisms of
continent urinary diversion with the cutaneous ileocecal
pouch (Mainz pouch I). JUrol 2006; 176:155-9; discussion
9-60.
304. Stein R, Wiesner C, Beetz R, et al.: Urinary diversion
in children and adolescents with neurogenic bladder:
the Mainz experience. Part III: Colonic conduit. Pediatr
Nephrol 2005; 20(7)932-6.
305. Gerharz EW, Preece M, Duffy PG, et al.: Enterocystoplasty
in childhood: a second look at the effect on growth. B J U
Int. 2003; 9(1):79-83.

281. Khoury JM, Timmons SL, Corbel L, et al.: Complications of


enterocystoplasty. Urol 1992; 40(1):9-14.
282. Stein R, Fisch M, Black P, et al.: Strategies for reconstruction
after unsuccessful or unsatisfactory primary treatment of
patients with bladder exstrophy or incontinent epispadias
(see comments). JUrol 1999; 161(6):1934-41.
283. Surer I, Ferrer FA, Baker LA, et al.: Continent urinary
diversion and the exstrophy-epispadias complex. JUrol
2003; 169(3):1102-5.
284. Frimberger D, Lakshmanan Y, Gearhart JP: Continent
urinary diversions in the exstrophy complex: why do they
fail? JUrol 2003; 170/4 Pt 1): 1338-42.
285. Herschorn S, Hewitt RJ: Patients perspective of longterm outcome of augmentation cystoplasty for neurogenic
bladders. Urol 1998; 52:672-8.
286. Husmann OA, Cain MP: Fecal and urinary continence
after ileal cecal cystoplasty for the neurogenic bladder.
JUrol 2001; 165(3):922-5.

377

306. Elia M: Oral or parenteral therapy for B 12 deficiency.


Lancet 1998;352:1721-2.
307. Kuzminski AM, Del Giacco EJ, Allen RH, et al.: Effective
treatment of cobalamin deficiency with oral cobalamin.
Blood 1998; 92(4):1191-8.
308. Hagedom CH, Alpers DH: Distribution of intrinsic factorvitamin B12 receptors in human intestine. Gastroenterology
1977; 73(5):1019-22.

309. Baker SJ, Mackinnon NL, Vasudevia P: The site of


absorption of orally administered vitamin B12 in dogs.
IndianJMedRes 1958; 46:812-7.
310. Booth CC, Mollin DL: The site of absorption of vitamin B12
in man. Lancet 1959; 1(7062):18-21.
311. Fleming WH, King ER: The site of absorption of orally
administered Co60-labelled vitamin B12 in dogs: the effect
of dose. Gastroenterology 1962; 42:164-8.
312. Pfitzenmaier J, Lotz J, Faldum A, et al.: Metabolic
evaluation of 94 patients 5 to 16 years after ileocecal
pouch (Mainz pouch 1) continent urinary diversion. JUrol
2003; 170:1884-7.
313. Olofsson G, Kilander A, Lindgren A, Ung KA, et al.: Vitamin
B 12 metabolism after urinary diversion with a Kock ileal
reservoir. ScandJUrolNephrol 2001; 35:382-7.
314. Matsui U, Topoll B, Miller K, et al.: Metabolic long-term
follow-up of the ileal neobladder. EurUrol 1993; 24:197200.
315. Akerlund S, Delin K, Kock NG, et al.: Renal function and
upper urinary tract configuration following urinary diversion
to a continent ileal reservoir (Kock pouch): a prospective
5 to 11-year followup after reservoir construction. JUrol
1980; 142:964-8.
316. Steiner MS, Morton RA, Marshall FF: Vitamin B12
deficiency in patients with ileocolic neobladders. JUrol
1993; 149:255-7.
317. Ooi BC, Barnes GL, Tauro GP: Normalization of
vitamin B12 absorption after ileal resection in children.
JPaediatrChildHealth 1992; 28:168-71.
318. Singh G, Thomas DG: Bowel problems
enterocystroplasty. BJU 1997; 79:328-32.

after

319. N`Dow J, Leung HY, Marshall C, et al.: Bowel dysfunction


after bladder reconstruction. JUrol 1998;159:14704;discussion 4-5.
320. Roth S, Semjonow A, Waldner M, et al.: Risk of bowel
dysfunction with diarrhea after continent urinary
diversion with ileal and ileocecal segments. JUrol 1995;
154:1696-9.
321. Henningsohn L, Wijkstrom H, Dickman PW, et al.:
Distressful symptoms after radical cystectomy with urinary
diversion for urinary bladder cancer: a Swedish populationbased study. EurUrol 2001; 40:151-62.
322. Awad SA, Al-Zahrani HM, Gajewski JB, et al.:Long-term
results and complications of augmentation ileocystoplasty
for idiopathic urge incontinence in women (see comments).
BJU 1998; 81:569-73.
323. Stein R, Schenk A, Wiesner C, et al.: Bowel movements
after urinary diversion matched pairs of illeocecal pouch
and ileal conduit. JUrol 2005; 173:486.
324. Somani BK, Kumar V, Wong S, et al.: Bowel dysfunction
after transposition of intestinal segments into the urinary
tract: 8-year prospective cohort study. JUrol 2007.
177(5):1793-8.
325. Giebel GD, Lefering R, Troidl H, et al.: Prevalence of fecal
incontinence: what can be expected? IntJColorectal Dis.
1998; 13(2)73-7.
326. Madoff RK, Parker SC, Varma MG, et al.: Faecal incontinene
in adults. Lancet 2004 Aug. 14-20; 364(9434):621-32.

378

Committee 8

Urothelial Carcinoma of the


Prostate
Chairs:
Joan Palou,
David Wood

Members:
H. Al-Ahmadie,
B. Bochner,
Henk van der Poel,
Ofer Yossepowitch

379

Table of Contents
I. INCIDENCE

III. Urothelial Carcinoma


Prostatic Stromal Invasion

1. Carcinoma in Situ

1. Incidence

2. Papillary Lesions

2. Mechanisms of Stromal Invasion

3. Stromal Invasion

3. Diagnosis

4. Cystectomy Specimens

4. P
 redicting Prostatic Stromal
Invasion

II. Diagnosis of Superficial


Urothelial Carcinoma or CIS
in the Prostatic Urethra

5. Treatment and Outcome Analyses

IV. RECOMMENDATIONS

1. Diagnosis
2. Treatment
3. Follow-up of the Prostatic
Urethra

380

Urothelial Carcinoma of the Prostate


Joan Palou, David Wood
H. Al-Ahmadie, B. Bochner,
Henk van der Poel, Ofer Yossepowitch

a tumor in the prostatic urethra is associated with


urothelial cancer elsewhere in the urinary tract
[14]. Mungan et al reported multiplicity as the only
risk factor of PU involvement in patients with nonmuscle-invasive bladder cancer [18]. Although in
cystoprostatectomy specimens the incidence of
urothelial carcinoma of the prostatic urethra has
been reported as high as 43% (27), most of these
series consisted of patients with larger, muscleinvasive bladder cancer [19].

I. INCIDENCE
1. Carcinoma in Situ
In 1952 Melicow and Hollow described Bowens
disease of the prostatic urethra [1]. Ortega et al were
the first to describe carcinoma in situ (CIS) of the
prostatic urethra in 1953 (Level 3)[2]. In approximately
90% of the cases, CIS is associated with a papillary
or solid bladder tumor [3]. However, in about 10%
of cases, it may present as an isolated lesion. It is
often diagnosed in the context of multifocal disease
of the bladder (Level 3)[2]. Prostatic urethra (PU)
involvement by CIS is relatively rare (Level 3)[4].
Its prevalence and significance has been clarified
by the use of routine random biopsies including the
prostatic urethra in patients with superficial bladder
tumors or positive urine cytology. In 1529 patients
with primary bladder tumors who had random
biopsies, 19% had carcinoma in situ, and 2.7%
had CIS in the prostatic urethra (Level 3)[5]. The
presence of PU CIS increased with the duration of
bladder cancer: secondary tumor involvement of
the prostatic urethra and ducts in bladder cancer
may be detected in 10-15% of patients with highrisk superficial bladder disease within 5 years and
in 20-40% within 15 years (Level 3)[6-8] (Fig.1).The
majority of patients also relapse in the bladder (Level
3) [8].

Figure 1. Urothelial carcinoma in situ involving prostatic


urethra. Markedly atypical and pleomorphic tumor cells
undermining urethral lining (highlighted in inset).

3. Stromal Invasion
Stromal invasion is present in 37-64% of men with
prostate involvement at cystoprostatectomy for
invasive bladder cancer [19;20] (Fig. 2). It was
generally shown to be a sign of poor prognosis [2022]. Survival in men with PU involvement without
stromal invasion was more than 60% 5 years after
the cystoprostatectomy, whereas less than 30% of
men survived when stromal invasion into the prostate
was present [20]. Stromal invasion into the prostate
was shown to be associated with an increased risk
of nodal metastases [21].

The most frequently reported risk factors for PU CIS


are multifocality of bladder cancer and CIS in the
bladder [9-12]. Others found bladder neck location
of the bladder cancer predictive of PU CIS [13;14].

2. Papillary Lesions
Primary papillary lesions in the prostatic urethra do
occur in only 1-4% of cases [15-17]. In most cases

381

Figure 2. Left: Urothelial carcinoma in situ involving prostatic ducts (UC) without stromal invasion. Note
adjacent normal prostatic ducts that are free of tumor involvement (NP). Right: Urothelial carcinoma with
invasion of prostatic stroma (arrows). Note adjacent ducts involved by in situ disease.

Besides stromal invasion, a distinction can be made


between contiguous and non-contiguous tumor
growth into the prostate. The former describes direct
invasion of the bladder tumor into the prostate,
whereas the latter refers to synchronous presence
of PU tumors. In a series of 320 cystoprostatectomy
specimens, Ayyathurai et al describe contiguous
tumor growth into the prostate in 9% of cases,
and non-contiguous in 15% [20]. Several studies
showed worse prognosis for contiguous versus noncontiguous growing tumors [20;23;24].

involvement in 143 of 489 cystectomy specimens


(Level 3)[23]. Nineteen cases showed direct
penetration from bladder to prostate and had a bad
prognosis. The remaining 124 patients had different
stages of indirect prostatic involvement. Thirty-seven
had low grade tumor or CIS and 29 had prostatic
disease involving the ducts only; these patients had
a relatively good prognosis. Wood et al reported
a 43% incidence of urothelial carcinoma of the
prostate in cystectomy specimens (Level 3)[14]. In
this series, 94% of those with prostatic involvement
had disease present in the prostatic urethra,
including 67% with CIS of the prostatic ducts or
acini. Risk factors included CIS of the bladder neck
or trigone, prior intravesical therapy, and ureteral
involvement by urothelial carcinoma. In a prospective
pathologic assessment, Revelo et al reported
results of 121 consecutive cystoprostatectomy
specimens analyzed by whole mount (Level 3)[28].
Of 121 prostates, 58 (48%) had urothelial carcinoma
involving the prostatic urethra, of which 19 (33%)
had apical involvement. All patients with prostatic
apical involvement by urothelial carcinoma uniformly
had involvement of more proximal (toward the base)
portions of the prostate. These results validate the
concept of cystoprostatectomy. Nixon et al reviewed
192 consecutive radical cystectomy specimens and
noted prostatic urethral involvement in 30 (15.6%)
specimens (Level 3)[11]. Of patients with CIS in
the bladder specimen, 31.3% (25/80) had prostatic
urethral involvement. Similarly, 34.7% (25/72) with
multifocal urothelial carcinoma of the bladder had
prostatic urethral involvement. Multivariate analysis
revealed the risk of prostatic urethral involvement to

4. Cystectomy Specimens
Most reports of prostatic urethral involvement at the
time of radical cystectomy are not only retrospective
but lack careful pathologic assessment of the prostate
and thus are likely to underreport the true incidence
of involvement with urothelial carcinoma. However,
the incidence of prostatic urethral involvement
approaches 50% in series where detailed pathologic
assessment of the prostate is performed [25;26].
In a prospective histopathologic study by Sakamoto
et al, 31 of 134 cystoprostatectomy specimens in
patients with primary bladder cancer revealed prostatic involvement (Level 3)[27]. They differentiated
between superficial (periurethral) and deep prostatic glands and demonstrated that there was prostatic urothelial carcinoma involvement of superficial
glands around the verumontanum in 93% of the cases, and at 5 and 7 oclock in 84%; only 1 patient had
deeper gland involvement without superficial gland
involvement.
Esrig et al analyzed their findings of prostatic

382

be 12- to 15-fold higher if CIS or multifocal urothelial


carcinoma of the bladder was present. More recent
data showed a prostatic tumor location in 38% of 248
cystoprostatectomy specimens [25]. In the majority
of men with prostatic involvement of urothelial
cancer, the cancer originated from the urothelial
mucosa, whereas in 4% the tumor extended through
the bladder wall into the prostate [29].

lesions in the prostate (Level 3)[35]. Donat et al


reported a 39% rate of prostate relapse in patients
with superficial bladder cancer (Level 3)[33]. They
recommended a systematic transurethral loop
biopsy extending through the bladder neck, trigone,
and prostatic urethra, and ultrasound-guided biopsy
of the prostate to evaluate patients with high grade
superficial disease; they do not justify the proposed
timing or frequency.

Predictors of PU involvement in cystoprostatectomy


specimens are:

a. multiplicity [11;18;25].

b. bladder neck tumor location [19;26].

c. bladder carcinoma in situ [11;19].

The staging of bladder carcinoma should include


independent evaluation and staging of any prostatic
lesion (Tables 1-4) (Level 3)[22]. Involvement of
the prostate by CIS is almost always associated
with bladder CIS (Level 3)[3]. Superficial disease
in the prostate is sometimes evident as a papillary
tumor. More commonly, it is silent and insidious with
no evidence of macroscopic disease. Once stromal
invasion is detected, the management changes
completely (see Section III. Prostatic Stromal
Invasion). Cystoscopy is a valuable measure of
prostatic urethral involvement in macroscopic
lesions, with a sensitivity of 83.3% and specificity of
95.1% (Level 3)[11]. But it is not always a valuable
tool; in patients with recurrent superficial urothelial
disease after BCG treatment, when considering
suspicious or visible lesions, Orihuela et al had a
48% negative biopsy rate (Level 3)[36]. CIS in the
prostatic urethra is common in high-risk superficial
bladder cancer (9-25%), and can evolve to prostatic
stromal disease (Level 3)[37]. Rikken et al stress
the need to perform a biopsy of the prostatic urethra
in every patient with high grade superficial bladder
cancer, with the intention of identifying patients with
tumors in that location to allow better therapeutic
planning (Level 3)[31]. Eight of 9 patients with
CIS of the prostatic urethra have coexistent CIS
in the bladder. Resectoscope loop biopsies of the
prostatic urethra are taken from the lateral lobes
and floor beginning distal to the bladder neck and
extending just proximal to the verumontanum. In
1 study, 76.5% of 17 prostatic urethral biopsies
contained identifiable prostatic ducts, and ductal
urothelial carcinoma was identified in 7 [33]. Hillyard
et al recommended taking a biopsy of the prostatic
urethra with the resectoscope loop from the lateral
lobes distal to the bladder neck and performing a
complete resection of the prostatic urethra (Level 3)
[38]. Bretton et al also recommended performing a
complete resection with 1 to 3 passes and separate
identification to determine the extent of prostatic
involvement (Level 3)[39].

II. Diagnosis of Superficial


Urothelial Carcinoma or CIS
in the Prostatic Urethra
1. Diagnosis
Urothelial carcinoma of the prostate is often
understaged; correct diagnosis has to be achieved in
order to avoid progression due to undetected stromal
invasion (Level 3)[8;23]. Interestingly, prostatic
involvement of urothelial cancer was the only
predictor of understaging in recurrent high risk nonmuscle-invasive bladder cancer: 53% understaging
in men with prostate involvement versus 20% in
men without. This suggests that prostatic sampling
plays a role in management of non-muscle-invasive
bladder cancer [30]. Cystoscopy was described as
a valuable tool to diagnose PU involvement [11].
However, Rikken et al showed that cystoscopically
hard to detect lesions, such as carcinoma in situ, can
be expected in around 10% of men with (a history
of) non-muscle-invasive bladder cancer at prostate
biopsy diagnosis [31].Transurethral resection of
the PU had highest accuracy for detecting stromal
invasion when compared to FNA or core-biopsy
prostate sampling [32]. Still detection of TUR biopsies
in the PU for stromal invasion was reported to be
only 56% [33]. Even at cystoprostatectomy specimen
pathology, an accurate protocol is required to prevent
undersampling of the prostatic urethra [34].
The only way to approach 100% accuracy would
be to perform an extensive transurethral resection
of the prostate, an impractical approach for routine
staging of patients with bladder cancer (Level 3)
[35]. Multiple random biopsies of the bladder and
prostatic urethra are not recommended as a normal
routine in superficial bladder cancer. There is a very
low incidence of CIS in that location, but this means
that CIS of the prostate will be undetected in patients
with primary bladder tumors (Level 3)[35]. Solsona
et al recommended a prostatic biopsy in patients
with positive cytology or in those with macroscopic

Survival is optimized if radical treatment precedes


stromal invasion (Level 3)[8]. The detection
of prostatic relapse, particularly early stromal
invasion, is difficult. That is why frequent and
lifelong biopsies of the bladder neck and prostate
are recommended in patients with recurrent high
grade papillary and in situ tumors of the bladder
(Level 3)[8].

383

Table1. Classification of Prostatic Urothelial Carcinoma According to Chibber et al [49]

Group 1 CIS of prostatic urethra and ducts, with or without similar disease in the bladder
Group 2 Superficial papillary bladder tumor of prostatic urothelium and major ducts with similar bladder
tumor
Group 3 Invasive tumor of bladder base and neck with prostatic extension
Table 2. Classification of Prostatic Involvement According to Hardeman and Soloway [70]

Stage 1 Tumor confined to prostatic urothelium


Stage 2 Invasion of ducts and acini, but confined to the basal membrane
Stage 3 Stromal invasion
Table 3. TNM (2001) Classification of Prostatic Urothelial (Transitional Cell) Carcinoma (71)

Tis pu CIS, affecting prostatic urethra


Tis pd CIS, affecting prostatic ducts
T1 Tumor invading subepithelial connective tissue
T2 Tumor invading prostatic stroma, spongiosum body, or periurethral muscle
T3 Tumor invading cavernous body prostatic capsule or bladder neck (extraprostatic extension)
T4 Tumor that invades surrounding organs
Table 4. TNM Classification (2009) of T4 Urothelial (Transitional Cell) Carcinoma of the Bladder (72)

T4 Tumor invades any of the following: prostate stroma, seminal vesicles, uterus, vagina, pelvic wall, abdominal
wall
T4a Tumor invades prostate stroma, seminal vesicles, uterus, or vagina
T4b Tumor invades pelvic wall or abdominal wall

2. Treatment

of urothelial neoplasms that cannot be managed


with intravesical therapy. Since Morales introduced
bacillus Calmette-Gurin (BCG) as intravesical
therapy of superficial urothelial carcinoma of the
bladder, several authors have evidenced its efficacy
both on bladder CIS and on CIS of the prostatic
urethra (Level 3)[22;42-45]. Superficial tumors of
the prostate can be treated conservatively if they are
completely resected and treated with intravesical
instillations. TUR with intravesical BCG provides a
significantly better prophylaxis of tumor recurrence
in Ta and T1 bladder cancer than TUR alone (Level
2)[46]. However, topical antineoplastic agents
may neither gain sufficient access to the prostatic
urothelium nor are incubated upon it for sufficient
duration of time (Level 3)[47]. In a few patients, there
is contact of the chemotherapeutic agent with the
prostatic urethra, but in the majority it is only during
micturition. In a non-randomized study comparing
BCG and epirubicin, there seemed to be a better
prostate response to BCG (Level 3)[47]. Hillyard et al
advocated radical surgery in patients with ductal and
stromal involvement; however, this is not justified with
numbers (Level 3)[38]. Solsona et al also advocated
radical surgery in patients with ductal involvement

a) General Comments
Local resection and staging of the often present
bladder tumor are essential for management of PU
urothelial cancers. Often the bladder cancer dictates
treatment, in particular in the cases with a muscleinvasive bladder tumor. Management of superficial
(non-stroma-invading) and non-superficial (stromal
or contiguous) PU tumors will be discussed.

b) Management of High Grade or CIS PU


Cancers
The prognostic implications of detecting CIS of
the prostatic urethra are defined by the risk of
progression (Level 3)[37]. Some authors feel that
this finding mandates the need to offer radical
cystoprostatectomy as primary therapy (Level 3)
[40;41]. Esrig et al analyzed their findings of prostatic
involvement at radical cystectomy; 37 of their patients
with low grade tumors or CIS and 29 with ductal only
prostatic disease had a relatively good prognosis
(Level 3)[23]. Both the prostatic urethra and distal
ureter have been considered extravesical locations

384

(Level 3)[48]. Although it seems that these patients


do worse with conservative treatment, this is likely
due to selection bias.

to BCG. In fact, treatment of CIS of the prostatic


urethra associated with superficial bladder tumor
has resulted in complete response rates of 70-100%
in the prostatic urethra and 47-72% when both the
bladder and prostate are considered (Table 6). In
the series of Bretton et al, therapeutic failure has
always been evidenced as disease progression in
the bladder; they believe that transurethral resection
of the prostate contributed significantly to the
successful control of tumor in the prostatic urethra
(Level 3)[39]. Hillyard et al described 2 cases in
which radical cystoprostatectomy was performed
to treat tumor progression in the prostatic urethra
or due to progression in the bladder (Level 3)[38].
Schellhammer et al reported an update of the
series of Hillyard et al and described persistence
or recurrence of disease in 9 of 17 patients (Level
3, [33]). Seven were treated with radical surgery.
Four presented with recurrence or progression in
the bladder, 1 in the prostate and 2 in both bladder
and prostate. They obtained good results in 8 of
10 patients with mucosal carcinoma and in 4 of 7
patients with ductal involvement. Therefore, a more
aggressive approach is warranted in cases with
ductal or acinar involvement.

c) BCG Penetration in the Prostatic Urethra


There is not complete agreement on the degree of
BCG penetration of the prostatic urethra in bladder
instillations. Some authors propose a primary
resection of the bladder neck to obtain direct
contact between the drug or immunotherapy and
the prostatic urothelium [43;49]. Several studies
on the prostatic tissue of patients who received
BCG without prior resection of the prostate have
demonstrated the presence of granulomas,
indicating that BCG penetrates into the prostate
(Level 3)[50-52]. Oates et al performed biopsy in
13 of 32 patients who received BCG because a
prostatic nodule or enhanced consistency of the
gland was palpated, and granulomas were found in
all of them (Level 3)[52]. Mukamel et al diagnosed
granulomas in 40% of the patients and distinguished
an early period (1.5-3 months after BCG) during
which these granulomas formed caseum (Level 3)
[51]. At a later stage (4-14.5 months after BCG), they
did not. Hillyard et al performed several cystograms
in patients without resection of the bladder neck or
prostate and only occasionally observed contrast
medium in the prostatic urethra (Level 3)[38].
However, patients with previous TUR of the prostate
show good opening of the bladder neck on the one
hand, and reflux to the prostatic ducts on the other.
Both changes could facilitate the penetration of BCG
into the prostatic gland.

In selected cases, a disease-specific survival of


89% was reported at 7.5 years follow-up for 28
patients with CIS of T1 or lower grade PU cancers
[55]. However, 28% required a cystectomy for local
recurrence or progression. In a series of 33 patients,
general progression of disease was evidenced in 8
cases, 3 of them at the prostatic urethra (Level 3). The
therapeutic response of this form of disease reaches
70%, while the global vesicoprostatic response is
only 40%. Palou et al showed an 82% response rate
of BCG for intraprostatic duct CIS even without TURP
[54]. Therefore, the results support the view that
BCG is a valid option to treat carcinoma in situ of the
prostatic urethra even without previous transurethral
resection and that it achieves an improved response.
However, with longer follow-up, disease progression
occurs in 24% of the cases. These patients have a
high disease-specific mortality.

BCG instillations were shown to result in lower local


recurrences in patients after TURP compared to men
without TURP [18;39;46;53-55].
Leibovici et al found a clinically significant elevation
of prostate specific antigen (PSA) in 41.6% of the
patients receiving BCG therapy, which reverted
to normal in 3 months (Level 3)[56]. The variation
of PSA levels following instillation of BCG was
evaluated (Level 3)[57]. The variation was higher in
patients who had previously undergone TUR of the
prostate (Table 5), which may be related to better
penetration of BCG into the prostate. Given these
results, resection of the bladder neck or prostate
is advisable, with separate superficial and deep
sampling at 5 and 7 oclock on both sides of the
verumontanum. This allows better initial staging
and gives the best chance to detect prostatic duct
involvement. This maneuver will also facilitate
penetration of BCG in the prostate.

Table 5. Total PSA Values During BCG Instillations


in Patients With and Without TURP [57]

TURP No TURP

Before BCG

1.2 2.1

1st instillation

7.7 3.4

2nd instillation

10.9 3.5

3rd instillation

8.8 3.8

d) BCG Treatment

4th instillation

7.1 4.5

The response rate of bladder CIS to BCG


immunotherapy is approximately 70%. When only
the mucosa of the prostatic urethra is involved, this
form of prostatic involvement should also respond

5th instillation

6.2 4.7

One month

5.3 4.7

Three months

3.5 3.2

385

Table 6. Conservative treatment of CIS in the prostatic urethra with BCG with or without TUR

Reference

Number of
patients (%)

Pathology

TURP

Bretton et al
(39)
Ovesen et al
(73)
Gofrit et al
(53)
Palou et al
(46)
Taylor et al
(55)
Mungan et al
(18)

23 (8.7)

CIS

10 (23.8)*
20
33 (20.7)*
28
14

Follow-up
(months)

Yes

Global/
prostate
response
56/100

CIS

Yes

-/ 80

26

CIS and
exophytic
CIS

Yes

46.5/95.3

58.5

No

39.4/69.6

64

CIS and
exophytic
CIS and
exophytic

No

35.7/50

90

No

NA/64

66

42

* Incidence in patients with bladder CIS

A good staging should be initially performed,


comprising a transurethral resection of the prostate,
together with strict follow-up and aggressive
management of recurrence or persistence of
ductal disease. Very few cases of recurrence in the
prostatic urethra after BCG failure have been treated
conservatively. Orihuela et al treated 5 patients,
and only 2 patients with papillary lesions sustained
complete response (Level 3)[36]. The other 3
patients with CIS or high grade tumors recurred
and progressed. Initial management of superficial
prostatic urethral disease with intravesical treatment
with BCG has reasonably good results with a
response rate of approximately 70%.

noninvasive and in 38% there was stromal invasion.


Solsona et al strongly recommend frequent random
biopsies of the prostatic urethra during initial and
repeated cystoscopic examinations (Level 3)[35].
Details of frequency and technique were not provided.
For patients with positive cytology in follow-up in the
absence of macroscopic bladder carcinoma, it is
mandatory to evaluate the bladder and the prostatic
urethra with multiple biopsies (Level 3)[35]. With a
positive cytology during the first 6 months of followup after conservative management of a superficial
bladder carcinoma, bladder recurrence is most likely
to be the cause (Level 3)[58]. The prostatic urethra
should be considered if there has been associated
carcinoma in situ, tumor near the bladder neck,
or multifocal disease (Level 3)[11;39]. If positive
cytology appears in longer term follow-up, the upper
urinary tract should be evaluated (Level 3)[59].

Patients who progress and those with urothelial


carcinoma of the bladder undergoing radical
cystectomy should be considered for prostatectomy
as well, given the high degree of prostatic involvement
(Level 3)[11;14;28].

Summary

3. Follow-up of the Prostatic


Urethra

Superficial urothelial cancers of the prostate are rare


and most often associated with bladder cancers. In
radical cystectomy specimens, the incidence of prostatic stromal invasion by urothelial carcinoma of the
prostate may be related to the extent of pathologic
evaluation and is associated with poor prognosis.
Most reports of prostatic urethral involvement at the
time of radical cystectomy are not only retrospective,
but lack careful pathologic assessment of the prostate. Underreporting of the true incidence of prostatic
involvement is common. Series of patients who underwent cystectomy for urothelial carcinoma of the
bladder report an incidence of urothelial carcinoma
of the prostate from 12-48% with stromal invasion in
7.6-16.6% [8;11;14;22-24;28;60-62]. Transurethral
resection of the prostate may improve response to
instillation therapy such as BCG. Careful follow-up
is required since recurrence in the prostate predicts
prognosis in non-muscle-invasive bladder cancer.

With increasing numbers of patients receiving initially


longer courses of intravesical therapy for high grade
superficial bladder cancer or CIS rather than radical
surgery, there will be an increased number of patients
at risk of developing urothelial carcinoma of the
prostate. Urothelial carcinoma of the prostate may
occur in 8-48% of patients (Level 3)[11;14;36;39].
Prostatic urethral involvement tends to be more
common in patients with CIS of the bladder, as well
as those with multifocal tumors and involvement of
the bladder neck (Level 3)[11;14;36;39].
In the series of Herr and Donat, with a minimum
follow-up of 15 years, 39% of patients with superficial
bladder cancer (72% with associated CIS treated with
BCG) relapsed in the prostate at a median follow-up
of 28 months (Level 3)[39]. In 62%, the tumors were

386

specimens, Ayyathurai et al describe contiguous


tumor growth into the prostate in 9% of cases
and non-contiguous in 15% [20]. Several studies
showed worse prognosis for contiguous versus noncontiguous growing tumors [20,23,24](Table 7).

III. Urothelial Carcinoma


Prostatic Stromal Invasion
1. Incidence

Table 7. Classification of Involvement of the Prostate


According to Pagano et al [24]

Adequate specimen processing by whole


mount sectioning of the prostate is essential for
determining the true incidence and pattern of
stromal invasion. Using this technique, several
contemporary cystoprostatectomy series provide
evidence suggesting prostatic stromal invasion is
more common than generally appreciated. Stromal
invasion is present in 37-64% of men with prostate
involvement at cystoprostatectomy for invasive
bladder cancer [19,20]. Shen et al studied 214
consecutive cystoprostatectomy specimens and
found invasive prostatic urothelial carcinoma in 39
patients (18%) [63] A similar incidence of stromal
invasion was observed by Richards et al (19%)[26]
and by Pettus et al (20%)[19]. It should be noted
that the prostatic stroma might be involved not only
in those with invasive bladder cancer but also in
men with non-invasive bladder cancer previously
managed by transurethral resection and intravesical
therapy. Herr and Donat evaluated 186 consecutive
men with superficial bladder cancer and found a
total of 14% to subsequently relapse with prostatic
stromal involvement [8].

Stage

Patients (%) Survival (%)

Contiguous

44 (61)

7%

Noncontiguous

28 (39)

46%

Urethral mucosa

100%

Ductal/acinar

14

50%

Stromal

40%

Extracapsular invasion 0

A third pattern has been described as tumors


located at the bladder neck trigone extending into
the prostatic stroma through thin lamina propria of
the bladder neck without histological evidence of
extravesical or intraurethral spread [33], but the
prognostic significance remains unknown.

3. Diagnosis
In men diagnosed with invasive urothelial carcinoma,
accurate determination of prostatic involvement
prior to cystoprostatectomy remains hampered by
inadequate diagnostic tools. Wood et al were first to
assess the diagnostic yield of transurethral prostate
biopsies, needle biopsies of the prostate, and fine
needle aspiration of the prostate to identify prostatic
involvement in 25 consecutive patients [14]

Stromal invasion is generally considered a sign of


poor prognosis [20-22]. Survival in men with PU
involvement without stromal invasion was more than
60% 5 years after the cystoprostatectomy, whereas
less than 30% of men survived when stromal invasion
was present [20]. Stromal invasion into the prostate
was shown to be associated with an increased risk
of nodal metastases [21].

While superficial prostate involvement was


accurately identified in the majority of patients, the
overall diagnostic accuracy for stromal invasion was
poor with only 2 of the 5 patients being identified by
TUR biopsies. In a more recent study from the same
group, 6 of 35 cystectomy specimens (17%) harbored
urothelial carcinoma involving the prostatic stroma
(either tumor extension from the bladder involving
periprostatic tissue or direct extension via the urethra)
[64]. In the 5 patients who had a TUR biopsy prior to
surgery, the diagnostic yield of prostatic involvement,
including superficial urethral and stromal invasion,
was 100%. Donat et al evaluated 246 male patients
undergoing cystectomy for bladder cancer and
analyzed the ability of transurethral loop biopsies
to adequately determine the presence and extent of
prostatic involvement [65]. Biopsies were obtained at
the 4 and 8 oclock positions from the bladder neck
to the verumontanum. With a sensitivity of 53%,
specificity of 77%, and positive predictive value of
45%, the authors concluded that transurethral biopsy,
while remaining the primary diagnostic modality, is
an imperfect tool for detecting stromal invasion by
urothelial carcinoma.

2. Mechanisms of Stromal Invasion


Prostatic stromal invasion is defined by the
presence of irregular invasive tumor nests or single
cells within the dense fibromuscular stroma of the
prostate or admixed within benign prostate glands.
Accompanying ductal/acinar carcinoma in situ and a
desmoplastic inflammatory response often coexists.
Besides stromal invasion, a distinction can be made
between contiguous and non-contiguous tumor
growth into the prostate. The former describes
direct invasion of bladder tumor into the prostate
as an extravesical tumor extension (stage pT3b or
greater) penetrating into the stroma directly through
the prostatic capsule whereas the latter refers to
synchronous presence of PU tumors as a pagetoid
spread of urethral tumor into the stroma via the ducts
and acini. In a series of 320 cystoprostatectomy

387

4. Predicting Prostatic Stromal


Invasion

While lacking support from level l evidence, the


importance of extended lymph node dissection in the
context of stromal invasion cannot be overstated. It
has been well acknowledged that the more extensive
the dissection and the more nodes removed,
the higher the staging accuracy and therapeutic
benefit to patients.[65,67,68] Mapping studies have
provided important data regarding the incidence
and location of node metastases in patients with
bladder cancer, specifically pT4 disease. Dangle et
al have recently demonstrated that extending the
lymph node dissection to include the presacral and
common iliac nodes changed the pN stage from pN0
to pN1 in 69% of cases and from pN1 to pN2 in 33%
of patients [65]. These findings reiterate results from
an earlier mapping study by Leissner and colleagues
who demonstrated that 7% of patients with Iymph
node spread had node metastases in the common
iliac or presacral regions only [69].

Given the lack of accurate diagnostic tools,


clinical factors that might help clinicians predict
stromal invasion would be particularly useful. Most
studies, however, assessing predictors of prostatic
involvement by urothelial carcinoma have combined
superficial urethral involvement and stromal invasion
into a single endpoint in their multivariate analyses.
Wood et al identified carcinoma in situ (CIS) involving
the trigone, bladder neck, periurethral structures,
and ureter, and the use of intravesical instillations
as predictors of prostatic urethral involvement [14].
Similarly, Pettus and colleagues found prostatic
urothelial carcinoma in 77 of 235 cystoprostatectomy
specimens (stromal invasion present in about half) and
identified CIS at the trigone as a significant predictor
of prostatic involvement (OR 6.3) [19]. In fact, only
6% of the patients with a bladder tumor proximal to
the trigone and no associated CIS had pathological
evidence of urothelial cancer involving the prostate.
In a more recent series of 96 cystoprostatectomy
specimens, Richards et al found prostatic stromal
involvement in 18 patients (19%) [26]. After adjusting
for clinical stage, Iymphovascular invasion, and
tumor multifocality, the presence of carcinoma in
situ (OR 3.2) and location of tumor at or below the
trigone (OR 3.3) were the only independent clinical
predictors of stromal invasion. Taken together, even
in the absence of positive findings on transurethral
loop biopsies, the presence of cancer, in particular
CIS, at the bladder neck trigone should be considered
a strong risk factor for prostatic stromal invasion.
Recently, in another study by Arce et al, they found
in multivariate analysis of a series of 717 cystectomy
specimens the presence of solitary T2-T3 bladder
tumor at the trigone or bladder neck was predictive
of invasive prostatic involvement [66].

In addition to lymph node status, it has been


suggested that the outcome of men with stromal
invasion can be stratified by the pattern and depth
of stromal invasion as aforementioned. Spread of
urothelial carcinoma into the ducts via the urethra
may not be as ominous as extravesical tumor
extension into the prostate or silent invasion into
the stroma emanating directly from a bladder neck
tumor. The recent study published by Barocas and
colleagues demonstrated the overall survival of men
with carcinoma in situ involving the prostatic urethra
or involvement of prostatic ducts to be similar to the
overall survival of men with no prostatic involvement
[21]. The outcome of those with stromal invasion was
uniformly poor irrespective of the mode of spread.

IV. RECOMMENDATIONS
Non-invasive Urothelial Carcinoma of the
Prostate

5. Treatment and Outcome Analyses


The
application
of
neoadjuvant
systemic
chemotherapy followed by radical cystoprostatectomy
with extended pelvic lymph node dissection is the
standard of care for men with prostatic stromal
invasion by urothelial cancer. Ample evidence is now
available to attest that prostatic stromal invasion is
associated with a higher likelihood of lymph node
metastases and a profound negative impact on
survival independent of lymph node status. Shen
and colleagues, for example, detected lymph node
metastases in 74% of the patients with stromal
invasion; the 5-year survival in this group was 32%
[63]. Similar findings were reported by Barocas et
al who found lymph nodes metastases in 74% of
men with bladder tumor extending into the stroma
and estimated the 3-year overall survival of men
with stromal invasion with and without lymph node
involvement to be 7% and 17%, respectively [21].

1. Macroscopic evidence of superficial bladder


cancer in the prostatic urethra is highly specific.
TUR of any macroscopic or suspicious lesion
is required for determination of the stage and
grade of tumors in the prostatic urethra. Once a
non-muscle-invasive high-grade tumor or CIS of
the bladder has been diagnosed, careful followup of the prostatic urethra is mandatory (grade
C).
2. High- and low-grade non-invasive urothelial
carcinoma and CIS of the prostate should be
treated with intravesical BCG (grade C).
3. Evidence (level 3) shows that TUR may improve
contact of BCG with the prostatic urethra and it
looks that response rates to BCG are increased
(grade C).
4. CIS or tumor in the prostatic ducts warrants

388

further study because very few patients have


been treated. It is advisable for the clinician to
perform radical surgery if there is any doubt, to
avoid understaging in patients with prostatic duct
involvement (grade C).

References
1. Melicow MM, Hollowell JW. Intra-urothelial cancer:
carcinoma in situ, Bowens disease of the urinary system:
discussion of thirty cases. J Urol 1952;68:763-72.

5. Patients with non-invasive urothelial carcinoma


in whom conservative therapy fails should be
considered for cystoprostatectomy (grade C).

2. Ortega LG, Whitmore WF, Jr., Murphy AI. In situ carcinoma


of the prostate with intraepithelial extension into the urethra
and bladder. Cancer 1953;6:898-923.
3. Seemayer TA, Knaack J, Thelmo WL, Wang NS, Ahmed
MN. Further observations on carcinoma in situ of the urinary
bladder: silent but extensive intraprostatic involvement.
Cancer 1975;36:514-20.

6. Patients with intermediate- to high-risk


superficial urothelial carcinoma of the bladder or
CIS, especially with involvement of the bladder
neck and multifocality, require monitoring of the
prostatic urethra when prostatic duct involvement
is observed (grade B or C).

4. Matzkin H, Soloway MS, Hardeman S. Transitional cell


carcinoma of the prostate. J Urol 1991;146:1207-12.
5. Millan-Rodriguez F, Chechile-Toniolo G, Salvador-Bayarri
J, Palou J, Vicente-Rodriguez J. Multivariate analysis of the
prognostic factors of primary superficial bladder cancer. J
Urol 2000;163:73-8.

Prostatic Stromal Invasion


1. T
 he incidence of prostatic urothelial carcinoma
in men with superficial or invasive bladder
cancer ranges from 12- 48%, and 7.6-16.6%
have stromal invasion. The prostate is a site of
relapse for patients with non-muscle-invasive
bladder cancer after intravesical therapy (grade
C).

6. Bassi P. BCG (Bacillus of Calmette Guerin) therapy of highrisk superficial bladder cancer. Surg Oncol 2002;11:77-83.
7. Davis JW, Sheth SI, Doviak MJ, Schellhammer PF.
Superficial bladder carcinoma treated with bacillus CalmetteGuerin: progression-free and disease specific survival with
minimum 10-year followup. J Urol 2002;167:494-500.
8. Herr HW, Donat SM. Prostatic tumor relapse in patients
with superficial bladder tumors: 15-year outcome. J Urol
1999;161:1854-7.

2. T
 ransurethral biopsy of the prostatic urethra is
effective in identifying prostatic involvement
but does not accurately reveal the extent of
involvement, particularly with stromal invasion
(grade C).

9. Grabstald H. Prostatic biopsy in selected patients with


carcinoma in situ of the bladder: preliminary report. J Urol
1984;132:1117-8.
10. Hardeman SW, Soloway MS. Urethral recurrence following
radical cystectomy. J Urol 1990;144:666-9.

3. Retrospective and prospective studies are


needed to determine prognostic factors for
prostatic stromal invasion (grade D).

11. Nixon RG, Chang SS, Lafleur BJ, Smith JA JA, Cookson
MS. Carcinoma in situ and tumor multifocality predict the
risk of prostatic urethral involvement at radical cystectomy
in men with transitional cell carcinoma of the bladder. J Urol
2002;167:502-5.

4. Radical cystoprostatectomy is the treatment of


choice for locoregional control in patients with
prostatic stromal invasion (grade B).

12. Tongaonkar HB, Dalal AV, Kulkarni JN, Kamat MR.


Urethral recurrences following radical cystectomy for
invasive transitional cell carcinoma of the bladder. Br J Urol
1993;72:910-4.

5. In patients with pT4 disease, the incidence of


positive nodes ranges from 40-50%, and node
mapping studies indicate that multiple sites are
involved (grade C).

13. Kefer JC, Voelzke BB, Flanigan RC, et al Risk assessment


for occult malignancy in the prostate before radical
cystectomy. Urology 2005;66:1251-5.
14. Wood DP, Jr., Montie JE, Pontes JE, VanderBrug
Medendorp S, Levin HS. Transitional cell carcinoma of the
prostate in cystoprostatectomy specimens removed for
bladder cancer. J Urol 1989;141:346-9.

6. Data show that the extent of lymphadenectomy


may have an impact on survival (grade C).
7. T
 he number of patients with prostatic stromal
invasion treated in radiation therapy trials is too
small to allow definitive conclusions regarding
survival outcome (grade C).

15. Bates HR, Jr. Transitional cell carcinoma of the prostate. J


Urol 1969;101:206-7.
16. Rubenstein AB, Rubnitz ME. Transitional cell carcinoma of
the prostate. Cancer 1969;24:543-6.
17. Wolfe JH, Lloyd-Davies RW. The management of transitional
cell carcinoma in the prostate. Br J Urol 1981;53:253-7.

8. A
 lthough the data are limited and the number of
patients with prostate stromal invasion cannot be
determined from the literature, the use of radical
cystectomy as a salvage procedure does not
appear to diminish disease-specific or overall
survival probability (grade C).

18. Mungan MU, Canda AE, Tuzel E, Yorukoglu K, Kirkali Z.


Risk factors for mucosal prostatic urethral involvement in
superficial transitional cell carcinoma of the bladder. Eur
Urol 2005;48:760-3.
19. Pettus JA, Al Ahmadie H, Barocas DA, et al Risk
assessment of prostatic pathology in patients undergoing
radical cystoprostatectomy. Eur Urol 2008;53:370-5.
20. Ayyathurai R, Gomez P, Luongo T, Soloway MS, Manoharan
M. Prostatic involvement by urothelial carcinoma of the
bladder: clinicopathological features and outcome after
radical cystectomy. BJU Int 2007;100:1021-5.

389

21. Barocas DA, Patel SG, Chang SS, et al Outcomes of


patients undergoing radical cystoprostatectomy for bladder
cancer with prostatic involvement on final pathology. BJU
Int 2009;104:1091-7.

mucosal involvement of the prostatic urethra: results of


treatment with intravesical bacillus Calmette-Guerin. J Urol
1988;139:290-3.

22. Schellhammer PF, Bean MA, WHITMORE WF, Jr. Prostatic


involvement by transitional cell carcinoma: pathogenesis,
patterns and prognosis. J Urol 1977;118:399-403.

39. Bretton PR, Herr HW, WHITMORE WF, Jr., et al Intravesical


bacillus Calmette-Guerin therapy for in situ transitional
cell carcinoma involving the prostatic urethra. J Urol
1989;141:853-6.

23. Esrig D, Freeman JA, Elmajian DA, et al Transitional cell


carcinoma involving the prostate with a proposed staging
classification for stromal invasion. J Urol 1996;156:1071-6.

40. Lamm DL, Stogdill VD, Stogdill BJ, Crispen RG.


Complications of bacillus Calmette-Guerin immunotherapy in
1,278 patients with bladder cancer. J Urol 1986;135:272-4.

24. Pagano F, Bassi P, Ferrante GL, et al Is stage pT4a (D1)


reliable in assessing transitional cell carcinoma involvement
of the prostate in patients with a concurrent bladder cancer?
A necessary distinction for contiguous or noncontiguous
involvement. J Urol 1996;155:244-7.

41. Schellhammer PF, Ladaga LE, Moriarty RP. Intravesical


bacillus Calmette-Guerin for the treatment of superficial
transitional cell carcinoma of the prostatic urethra
in association with carcinoma of the bladder. J Urol
1995;153:53-6.

25. Mazzucchelli R, Barbisan F, Santinelli A, et al Prediction


of prostatic involvement by urothelial carcinoma in
radical cystoprostatectomy for bladder cancer. Urology
2009;74:385-90.

42. Herr HW. Carcinoma in situ of the bladder. Semin Urol


1983;1:15-22.

26. Richards KA, Parks GE, Badlani GH, et al Developing


selection criteria for prostate-sparing cystectomy: a review
of cystoprostatectomy specimens. Urology 2010;75:111620.

44. deKernion JB, Huang MY, Lindner A, Smith RB, Kaufman


JJ. The management of superficial bladder tumors and
carcinoma in situ with intravesical bacillus Calmette-Guerin.
J Urol 1985;133:598-601.

43. Lamm DL. Bacillus Calmette-Guerin immunotherapy for


bladder cancer. J Urol 1985;134:40-7.

27. Sakamoto N, Tsuneyoshi M, Naito S, Kumazawa J. An


adequate sampling of the prostate to identify prostatic
involvement by urothelial carcinoma in bladder cancer
patients. J Urol 1993;149:318-21.

45. Morales A, Eidinger D, Bruce AW. Intracavitary Bacillus


Calmette-Guerin in the treatment of superficial bladder
tumors. J Urol 1976;116:180-3.
46. Palou J, Xavier B, Laguna P, Montlleo M, Vicente J. In
situ transitional cell carcinoma involvement of prostatic
urethra: bacillus Calmette-Guerin therapy without
previous transurethral resection of the prostate. Urology
1996;47:482-4.

28. Revelo MP, Cookson MS, Chang SS, et al Incidence and


location of prostate and urothelial carcinoma in prostates
from cystoprostatectomies: implications for possible apical
sparing surgery. J Urol 2004;171:646-51.
29. Barbisan F, Mazzucchelli R, Scarpelli M, et al Urothelial
and incidental prostate carcinoma in prostates from
cystoprostatectomies for bladder cancer: is there a
relationship between urothelial and prostate cancer? BJU
Int 2009;103:1058-63.

47. Canda AE, Tuzel E, Mungan MU, Yorukoglu K, Kirkali Z.


Conservative management of mucosal prostatic urethral
involvement in patients with superficial transitional cell
carcinoma of the bladder. Eur Urol 2004;45:465-9.
48. Solsona E, Iborra I, Ricos JV, et al The prostate involvement
as prognostic factor in patients with superficial bladder
tumors. J Urol 1995;154:1710-3.

30. Huguet J, Crego M, Sabate S, et al Cystectomy in patients


with high risk superficial bladder tumors who fail intravesical
BCG therapy: pre-cystectomy prostate involvement as a
prognostic factor. Eur Urol 2005;48:53-9.

49. Chibber PJ, McIntyre MA, Hindmarsh JR, et al Transitional


cell carcinoma involving the prostate. Br J Urol
1981;53:605-9.

31. Rikken CH, van Helsdingen PJ, Kazzaz BA. Are biopsies
from the prostatic urethra useful in patients with superficial
bladder carcinoma? Br J Urol 1987;59:145-7.

50. LaFontaine PD, Middleman BR, Graham SD, Jr., Sanders


WH. Incidence of granulomatous prostatitis and acidfast bacilli after intravesical BCG therapy. Urology
1997;49:363-6.

32. Wood DP, Jr., Montie JE, Pontes JE, Levin HS.
Identification of transitional cell carcinoma of the prostate
in bladder cancer patients: a prospective study. J Urol
1989;142:83-5.

51. Mukamel E, Konichezky M, Engelstein D, et al Clinical and


pathological findings in prostates following intravesical
bacillus Calmette-Guerin instillations. J Urol 1990;144:1399400.

33. Donat SM, Wei DC, McGuire MS, Herr HW. The efficacy
of transurethral biopsy for predicting the long-term
clinical impact of prostatic invasive bladder cancer. J Urol
2001;165:1580-4.

52. Oates RD, Stilmant MM, Freedlund MC, Siroky MB.


Granulomatous prostatitis following bacillus CalmetteGuerin immunotherapy of bladder cancer. J Urol
1988;140:751-4.

34. Varinot J, Camparo P, Roupret M, et al Full analysis of the


prostatic urethra at the time of radical cystoprostatectomy
for bladder cancer: impact on final disease stage. Virchows
Arch 2009;455:449-53.
35. Solsona E, Iborra I, Ricos JV, et al Recurrence of superficial
bladder tumors in prostatic urethra. Eur Urol 1991;19:8992.

53. Gofrit ON, Pode D, Pizov G, et al Prostatic urothelial


carcinoma: is transurethral prostatectomy necessary
before bacillus Calmette-Guerin immunotherapy? BJU Int
2009;103:905-8.

36. Orihuela E, Herr HW, WHITMORE WF, Jr. Conservative


treatment of superficial transitional cell carcinoma
of prostatic urethra with intravesical BCG. Urology
1989;34:231-7.

54. Palou Redorta J, Schatteman P, Huguet Perez J, et al


Intravesical instillations with bacillus calmette-guerin for the
treatment of carcinoma in situ involving prostatic ducts. Eur
Urol 2006;49:834-8.

37. Thelmo WL, Seemayer TA, Madarnas P, Mount BM,


Mackinnon KJ. Carcinoma in situ of the bladder with
associated prostatic involvement. J Urol 1974;111:491-4.

55. Taylor JH, Davis J, Schellhammer P. Long-term followup of intravesical bacillus Calmette-Guerin treatment
for superficial transitional-cell carcinoma of the bladder
involving the prostatic urethra. Clin Genitourin Cancer
2007;5:386-9.

38. Hillyard RW, Jr., Ladaga L, Schellhammer PF. Superficial


transitional cell carcinoma of the bladder associated with

390

56. Leibovici D, Zisman A, Chen-Levyi Z, et al Elevated prostate


specific antigen serum levels after intravesical instillation of
bacillus Calmette-Guerin. J Urol 2000;164:1546-9.
57. Palou J, Lopez H, Millan F, et al Effect of intravesical
instillations of BCG on total PSA. Preliminary results. BJU
Int 2000;86:254.
58. Schwalb DM, Herr HW, Fair WR. The management of
clinically unconfirmed positive urinary cytology. J Urol
1993;150:1751-6.
59. Schwalb MD, Herr HW, Sogani PC, et al Positive urinary
cytology following a complete response to intravesical
bacillus Calmette-Guerin therapy: pattern of recurrence. J
Urol 1994;152:382-7.
60. Coutts AG, Grigor KM, Fowler JW. Urethral dysplasia
and bladder cancer in cystectomy specimens. Br J Urol
1985;57:535-41.
61. Njinou Ngninkeu B, Lorge F, Moulin P, Jamart J, Van Cangh
PJ. Transitional cell carcinoma involving the prostate: a
clinicopathological retrospective study of 76 cases. J Urol
2003;169:149-52.
62. Reese JH, Freiha FS, Gelb AB, Lum BL, Torti FM. Transitional
cell carcinoma of the prostate in patients undergoing radical
cystoprostatectomy. J Urol 1992;147:92-5.
63. Shen, S. S., Lerner, S. P., Muezzinoglu, B. et al: Prostatic
involvement by transitional cell carcinoma in patients with
bladder cancer and its prognostic significance. Hum Pathol.
2006;37: 726.
64. Weizer, A. Z., Shah, R. B., Lee, C. T. et al: Evaluation of the
prostate peripheral zone/capsule in patients undergoing
radical cystoprostatectomy: defining risk with prostate
capsule sparing cystectomy. Urol Oncol. 2007; 25: 460.
65. Dangle, P. P., Gong, M. c., Bahnson, R. R. et al: How
do commonly performed lymphadenectomy templates
influence bladder cancer nodal stage? J Urol. 183: 499.
66. Arce, J., Gaya, JM., Huguet, J., Rodrguez, O., Palou, J.,
Villavicencio, H. Can we identify those patients who will
benefit from prostate-sparing surgery? Predictive factors
for invasive prostatic involvement by transitional cell
carcinoma. Can J Urol.2011 18(1);5345-5352
67. Hellenthal, N. J, Ramirez, M. L., Evans, C. P. et al: Trends in
pelvic lymphadenectomy at the time of radical cystectomy:
1988 to 2004. J Urol. 2009; 181: 2490.
68. Dhar, N. B., Klein, E. A. Reuther, A. M. et al: Outcome after
radical cystectomy with limited or extended pelvic lymph
node dissection. J Urol. 2008; 179: 873.
69. Leissner, J., Ghoneim, A., Abol-Enein, H. et al Extended
radical lymphadenectomy in patients with urothelial bladder
cancer: results of a prospective multicenter study. J Urol.
2004; 171: 139.
70. Hardeman SW, Soloway MS. Transitional cell carcinoma of
the prostate: diagnosis, staging, and management. World J
Urol 1988;6:170-4.
71. Wittekind Ch, Henson DE, Hutter RVP, Sobin LH. TNM
supplement. 2nd edition. A commentary on uniform use.
John Wiley & Sons, Inc.; 2001.
72. International Union Against Cancer: TNM classification of
malignant tumours, 6th ed. Edited by M.H. Harmer. J Urol
2009.
73. Ovesen H., Poulsen AL., Steven K. Intravesical
Bacillus Calmette-Gurin with Danish strain for
treatment of carcinoma in situ of the bladder. Br J Urol
1993;75(5Pt2):744-8

391

392

Committee 9

Chemotherapy for Metastatic


Urothelial Carcinoma
Chair:
CN. Sternberg
Members:
D. Bajorin,
R. Dreicer,
M. Galsky,
D. George,
M. Milowsky,
D. Quinn,
G. Sonpavde,
WM. Stadler,
D. Theodorescu,
D. Vaughn
Correspondence:
Cora N. Sternberg, MD, FACP,
San Camillo and Forlanini Hospitals
Padiglione Flajani
Circonvallazione Gianicolense 87
00152 Rome, Italy
Phone: +39 06 6641 8008
Fax: +39 06 6630771
E-mail:cstern@mclink.it

393

Table of Contents
I. Background and
Introduction

VIII. Combination Therapy


IX. Level of Evidence and
Grades of Recommendation
for Second-line and Salvage
Therapy

II. First-line Treatments


III. Modifications of M-VAC

X. Novel Agents for Advanced


Urothelial Carcinoma

IV. Doublet Chemotherapy


V. Triplet Drug Combination
with Platinum Analogues

XI. New Trial Designs and the


Development of Tailored
Therapy

VI. Cisplatin-ineligible Patients

XII. Predictive Models

VII. Second-line Treatments/


Salvage Chemotherapy
Monotherapy

XIII. Abstract

394

Chemotherapy for Metastatic


Urothelial Carcinoma
CN. Sternberg,
D. Bajorin, R. Dreicer, M. Galsky,
D. George, M. Milowsky, D. Quinn,
G. Sonpavde, WM. Stadler, D. Theodorescu, D. Vaughn,

I. Background and
Introduction

III. Modifications of M-VAC


Several phase II trials have evaluated administration
of dose-dense M-VAC (DD-M-VAC) with either GMCSF (granulocyte macrophage colony stimulating
factor) or G-CSF (granulocyte colony stimulating
factor). These studies have demonstrated the
possibility of increasing the doses as much as 60%
[7]. Unfortunately, in some of these trials, toxicity
was also significantly increased [8-10].

Bladder cancer is a major health problem as defined


by incidence, mortality, and cost of care. While
a major part of the expense is directed towards
treatment of non-muscle-invasive and organconfined bladder cancer, the major morbidity and
mortality is a result of recurrent and/or metastatic
disease. Although urothelial cancer is typically a
chemo-sensitive disease, only a small sub-set of
patients with advanced disease achieve long-term
disease-free status.

The European Organisation for Research and


Treatment of Cancer (EORTC) conducted a
randomized trial comparing DD-M-VAC given on
an every 2 week schedule with classic M-VAC.
With DD-M-VAC, it was possible to deliver higher
doses of doxorubicin and cisplatin as compared to
classic M-VAC. Although there was not a significant
difference found in median survival (approximately
15 months in both arms), 21.8% were alive on the
DD-M-VAC arm versus 13.5% on the M-VAC arm
at 5 years (HR=0.76, P=0.042). There was also
a significant difference in favor of DD- M-VAC in
response rate (72% vs. 58%, P=0.016), complete
response rate (25% vs. 11%, P=0.006) and median
progression-free survival (9.5 vs. 8.1 months,
P=0.017) [7;11].

II. First-line Treatments


Cisplatin combinations such as M-VAC (Methotrexate,
Vinblastine, Adriamycin, Cisplatin) were extensively
studied in the 1980s and yielded partial and complete
remissions in patients with metastatic disease, with
long-term survival reported in complete responders
[1;2]. Patients with nodal disease only and a good
performance status have a 35% complete response
rate, with the majority enjoying long-term survival
and apparent cure [3].
On the basis of trials comparing M-VAC to cisplatin
chemotherapy [4;5] and to CISCA (cisplatin,
cyclophosphamide,
Adriamycin)
combination
therapy, M-VAC became the standard of care for the
treatment of metastatic bladder cancer in the 1990s
with reproducible response rates in the range of 4070% and a median survival of approximately 1 year
in several series [2;4;6].

A phase III randomized trial comparing M-VAC plus


prophylactic G-CSF with docetaxel-cisplatin (DC) plus
G-CSF was reported by the Hellenic Cooperative
Oncology Group [12]. Two hundred twenty patients
were randomly assigned to M-VAC (109 patients) or
DC (111 patients). The overall response rate (54.2%
vs. 37.4%; P=0.017), median time to progression
(TTP; 9.4 vs. 6.1 months; P=0.003) and median
survival (14.2 vs. 9.3 months; P =0.026) favored
the M-VAC regimen. After adjusting for prognostic
factors, the difference in time to progression

Nevertheless, the toxicity of the M-VAC regimen,


without grouth factor support, consisting mainly of
myelosuppression and mucositis, with a low but
consistent therapy related death rate, has challenged
routine use.

395

remained significant, whereas the survival difference


was non-significant at the 5% level. M-VAC caused
more frequent grade 3 or 4 neutropenia (35.4% vs.
19.2%; P=0.006), thrombocytopenia (5.7% vs. 0.9%;
P=.046), and neutropenic sepsis (11.6% vs. 3.8%;
P=0.001), but toxicity of M-VAC was considerably
lower than that previously reported for M-VAC
administered without G-CSF. It can be concluded
from this study that not only was M-VAC supported
by G-CSF more effective than DC but that it was
also better tolerated than classic M-VAC as first-line
treatment in advanced urothelial carcinoma (level of
evidence 1b).

survival was 7.7 months for GC and 8.3 months


for M-VAC, with a hazard ratio of 1.09. The 5-year
progression-free survival rates were 9.8% and
11.3%, respectively (P=0.63). On the basis of a more
favorable risk-benefit ratio, many consider the GC
regimen as a new standard treatment in metastatic
urothelial cancer (level of evidence 1b).
Several phase II trials explored the combination
of paclitaxel and cisplatin [19;20] with an overall
response rate of approximately 50%. At least three
trials evaluated the combination of docetaxel and
cisplatin with complete responses ranging from
19- 26% and partial responses from 34-58%. [2123]. The toxicity of these regimens was considered
moderate with neutropenia and neurotoxicity being
most common. As noted above, a phase III trial
demonstrated that this regimen was inferior to
M-VAC plus G-CSF [12].

These trials confirm that classic M-VAC with


appropriate hematopoietic growth factor support
is an effective treatment. DD-M-VAC is another
excellent option that may potentially lead to higher
long-term cure rates.

Several phase II trials have studied the combination


of paclitaxel with varying doses (150-225 mg/
m) and carboplatin (AUC 5-6) reporting overall
response rates of 14-65%, with complete responses
in 0-40% [24-28]. This regimen is well-tolerated with
predominantly mild hematologic and neurologic
toxicities.

IV. Doublet Chemotherapy


Several phase II trials in both pretreated and
untreated patients evaluated the combination of
cisplatin and gemcitabine (GC). Different schedules
and dosages in locally advanced and/or metastatic
bladder cancer have been conducted [13] with
overall response rates ranging from 41- 57% and
complete response from 18- 22% [14-16]. Toxicity
was generally acceptable and median survival rates
were 12.5, 14.3, and 13.5 months, respectively.

A phase III trial conducted by the ECOG compared


M-VAC with paclitaxel-carboplatin [29]. Patients with
previously untreated metastatic urothelial carcinoma
were randomized to either standard M-VAC or
paclitaxel 225 mg/m plus carboplatin AUC 6 every
3 weeks. After 2.5 years, the study was closed due
to slow accrual. Of the planned 330 patients, only 85
were enrolled. Compared with carboplatin-paclitaxel,
patients treated with M-VAC had more severe
myelosuppression, mucositis, and renal toxicity.
Interestingly, a quality of life evaluation revealed
non-significant differences between the two arms.
At a median follow-up of 32.5 months, there was
no significant difference in response rate (35.9%
M-VAC vs. 28.2% PC, P=0.34) or median survival
(15.4 months M-VAC vs. 13.8 months PC, P=0.41)
between the two arms. Definitive conclusions
are not possible, given that the trial was severely
underpowered.

These results led to a large industry-sponsored


randomized trial of GC versus M-VAC. When first
published, at a median follow-up of 19 months,
overall survival in the M-VAC group was 14.8 months
as compared to 13.8 months in the GC group [17].
This trial was not designed as an equivalency trial,
and many more patients would have been needed
to prove that the two regimens were equivalent in
efficacy.
The toxicity profile was different in the two arms,
with lower rates of grade 3-4 neutropenia (72% vs.
84%), neutropenic fever (2% vs. 14%), and grade
3-4 mucositis (1% vs. 22%) in the GC arm, and a
lower rate of thrombocytopenia and anemia in the
M-VAC arm. It is possible that some of the toxicities
in the M-VAC group could have been avoided with
the prophylactic use of G-CSF, which was not
routinely prescribed. The quality of life profile was
similar for both arms with the exception of a nonstatistically significant improvement in fatigue in the
GC group.

The combination of gemcitabine and pemetrexed has


been tested in 64 chemo-naive patients with locally
advanced and/or metastatic urothelial carcinoma. The
overall response rate among 47 patients evaluable
for response was 28% and overall response rate for
the intention-to-treat population was 20%. Median
response duration was 11.2 months and median
overall survival 10.3 months. There was one toxic
death due to neutropenic sepsis. The efficacy was
not superior to that of single-agent gemcitabine
[30].

In an update of this trial [18], overall survival was


similar in both arms with median survival of 14.0
months for GC and 15.2 months for M-VAC. The
5-year overall survival rates were 13.0% and 15.3%,
respectively (P=0.53). The median progression-free

The combination of gemcitabine and paclitaxel has


been evaluated in both pretreated and untreated

396

patients with different doses and schedules [31-35]


with responses in the range of 35-55%.

investigated non-cisplatin-based chemotherapy


regimens in cisplatin-ineligible patients, however,
these trials have generally included heterogeneous
eligibility criteria due to a lack of consensus
regarding what actually constitutes cisplatinineligibility. Recently, a working group surveyed a
large group of medical oncologists in an attempt to
develop a uniform definition of patients unfit for
cisplatin that could be used to standardize eligibility
criteria for clinical trials in this patient population
moving forward [42;43]. Based on the survey
results and the available literature regarding the
safety of cisplatin in patients with urothelial cancer
and other solid tumors, a consensus definition was
generated. According to this definition, patients are
deemed cisplatin-ineligible if they meet at least
one of the following: ECOG performance status
of 2, creatinine-clearance less than 60 mL/min,
common terminology for adverse events (CTCAE)
grade 2 or greater hearing loss, CTCAE v4 grade 2
or greater peripheral neuropathy, or New York Heart
Association Class III heart failure. These criteria are
rooted in common clinical practice, are designed
to optimize the general population of patients with
urothelial cancer who are eligible for clinical trials,
and may lead to a greater likelihood of developing a
viable strategy for regulatory approval of therapeutic
regimens for this important patient subset.

Given the activity observed in several two-drug


combinations, partially non-overlapping toxicities,
and different mechanisms of action, combination
therapy in three-drug regimens was the next logical
step.

V. Triplet Drug Combination


with Platinum Analogues
Investigators have attempted to enhance treatment
efficacy by adding a third drug to the doublet
combinations. Data emerging from the phase II
studies have suggested that triplet regimens might
produce higher response rates and improve median
survival [36-38].
To elucidate the role of paclitaxel when added to
GC, a large international Intergroup Study EORTC
30987 compared the triplet paclitaxel-cisplatingemcitabine (PCG) with the conventional GC doublet
in 627 chemo-naive patients with locally advanced
or metastatic urothelial carcinoma. The study was
designed to identify a 4-month difference in median
survival between the two treatment groups. Overall
response rate was 57.1% for PCG (CR 15%)
and 46.4% for GC (CR 10%, P=0.02). Median
progression-free survival was 8.4 months and 7.7
months for PCG and GC respectively (P=0.10).
Median survival was 15.7 months for PCG and 12.8
months with GC, with no significant difference in
overall survival (P=0.10). Both treatments were welltolerated, with more thrombocytopenia and bleeding
on GC (12% vs. 7%) and more febrile neutropenia
on PCG (13% vs. 4%). In a retrospective subgroup
analysis, a survival benefit was seen in patients with
the bladder as their primary tumor site and in those
with no risk factors or one factor (risk factors were
the presence of visceral metastases and a WHO
performance status of 1) [39].

The most commonly used non-cisplatin-based


combination chemotherapy regimen in patients
with advanced urothelial cancer is gemcitabine and
carboplatin. This combination has been evaluated as
first-line treatment in phase II trials [44;45] .
In one of the gemcitabine-carboplatin trials, 60
unselected patients with advanced urothelial
carcinoma were treated with carboplatin AUC 5
day 1 and gemcitabine 1000 mg/m2 days 1 and 8
every 21 days in a study of the Hellenic Cooperative
Oncology Group [46] . The objective response rate
was 38.4% with 11.7% complete responses and
26.7% partial responses. The median time to disease
progression was 7.6 months and the median overall
survival was 16.3 months. The median survival was
comparable to that reported for the combination of
M-VAC according to the Memorial Sloan-Kettering
Cancer Center prognostic model (3) based on similar
baseline prognostic features.

In light of the results, this triplet regimen cannot be


recommended as first-line therapy (level of evidence
1b).

VI. Cisplatin-ineligible Patients

The authors concluded that the combination of


gemcitabine and carboplatin appears to have
considerable activity as the first-line treatment
of unselected patients with advanced urothelial
carcinoma with manageable toxicity, and that it
deserved further evaluation in this setting.

As the majority of patients with advanced urothelial


cancer are older (median age 68 years) and renal
function declines with age, alternative non-cisplatinbased therapy is needed [40]. In fact, approximately
40% of patients with advanced urothelial cancer are
cisplatin-ineligible due to renal dysfunction alone
[41], and the proportion of cisplatin-ineligible patients
is likely much higher when including treatmentlimiting factors such as poor performance status
and comorbidities. Several phase II studies have

The University of Michigan led a phase II trial


of carboplatin, gemcitabine, and paclitaxel for
advanced urothelial cancer in patients with advanced
urothelial malignancy of any histology, no previous
chemotherapy for metastatic disease, Southwest

397

Oncology Group performance status of 2 or less,


and serum creatinine levels of 2 mg/dL or less [38].
Of 49 patients accrued, 15 (32%) experienced a
complete response while 17 (36%) patients had
a partial response. Febrile neutropenia occurred
in only 1.4% of patients while grade 3 neuropathy
occurred in 4 patients. The median survival was
14.7 months. The trial was notable for significant
responses and good median survival with a noncisplatin regimen, including responses in patients
with squamous histology, who historically tend to
be resistant to cytotoxic therapy. On that basis, this
regimen has been widely used for subjects who are
not cisplatin-ideal even though it has never been
formally compared to cisplatin-containing regimens.

study 30986, a phase III trial comparing M-CAVI


versus GCa in patients with advanced urothelial
cancer unfit for cisplatin-based chemotherapy,
were presented [50;51]. Eligibility criteria for this
trial included a WHO performance status 2 and/
or impaired renal function (GFR greater than 30
but less than 60 mL/min). Overall survival was not
different between the two arms [8.1 months versus
9.3 months; HR 0.94 (95% CI, 0.72-1.22); P=0.64].
However, severe acute toxicity (defined as grade
3-4 mucositis, grade 4 thrombocytopenia associated
with bleeding, grade 3-4 neutropenic fever, grade
3-4 renal toxicity, death) was substantially greater
for M-CAVI than GCa (21.2 % vs. 9.3%). This
trial provides level 1 evidence for the use of GCa
chemotherapy in patients deemed ineligible for
cisplatin-based chemotherapy; however, adoption
of a standard definition of cisplatin-ineligible and
additional randomized studies of chemotherapy
and novel agents in this patient population are
desperately needed [42].

A randomized trial in patients with surgically


incurable advanced bladder cancer evaluated
whether the carboplatin-based regimen M-CAVI
(methotrexate, carboplatin, and vinblastine) offered
an advantage over the cisplatin-based regimen
M-VAC (methrotrexate, vinblastine, doxorubicin,
and cisplatin) [47]. In this 47 patient trial, the overall
response rates were 39% for M-CAVI and 52% for
M-VAC (P=0.3) with 3 complete responses among
the patients receiving M-VAC and none in the
M-CAVI group. The median disease-related survival
was 16 months with M-VAC versus 9 months with
M-CAVI (P=0.03). The authors concluded that
M-CAVI was less toxic but less active than M-VAC
and the M-CAVI regimen should be reserved for noncisplatin candidates.

VII. Second-line Treatments/


Salvage Chemotherapy
Monotherapy
Vinflunine, a microtubule inhibiting vinca alkaloid
is the most thoroughly investigated agent in the
second-line setting. One phase II trial recruited 51
patients, most of whom had progressed within 12
months, and demonstrated an 18% response rate
and a median duration of response of 9.1 months
[52] . Another phase II trial accrued 175 patients with
disease progressing within 12 months of platinumbased chemotherapy and demonstrated a response
rate of 15% and median duration of response of 6
months [53]. Subsequently, a randomized phase III
trial accrued 370 patients and compared vinflunine
plus best supportive care (BSC) to BSC alone as
second-line therapy [54]. This trial accrued patients
progressing after frontline platinum-containing
chemotherapy for metastatic disease, and those
who had received prior peri-operative chemotherapy
only were excluded. More than 80% had progressed
within 6 months after prior chemotherapy and greater
than 70% of patients had visceral disease. An
extension of survival, the primary endpoint, was not
demonstrated by an intention-to-treat (ITT) analysis
(6.9 vs. 4.6 months, P=0.287), but there was a
statistical improvement in response rate (8.6 vs. 0%)
and median progression-free survival (3.0 vs. 1.5
months). Approximately 30% of patients in both arms
received subsequent systemic therapy, which may
have confounded survival. Multivariate Cox analysis
adjusting for prognostic factors demonstrated a
significant extension of survival with vinflunine
(P=0.036), reducing the risk of death by 23%. In
another analysis of only the eligible population

A second randomized phase II study evaluated the


cisplatin-containing regimen M-VEC (methotrexate, vinblastine, epirubicin, and cisplatin) and the
carboplatin-containing regimen M-VECa (methotrexate, vinblastine, epirubicin, and carboplatin) in 57
patients with recurrent or metastatic bladder cancer
[48]. The overall clinical response rate was 71% (with
25% complete responses) in the M-VEC group and
41% (with 11% complete responses) in the M-VECa
group (P=0.04).
In a third randomized phase II trial comparing the
toxicity and efficacy of gemcitabine plus cisplatin
(GP) and gemcitabine plus carboplatin (GCa) in
patients with advanced urothelial cancer, the median
survival times were 12.8 months and 9.8 months,
respectively [49]. Although the trial was not designed
with sufficient power to detect significant differences
in efficacy between the arms, there is the suggestion
of a better outcome in the cisplatin arm. The
authors concluded that the GCa combination might
be considered as an alternative to GP in selected
patients such as those with renal dysfunction. The
trial only included patients with a creatinine clearance
equal to or greater than 60 mL/min and a Zubrod
performance status of 0-2.
More recently, the updated results of the EORTC

398

(n=357), the median survival was significantly


longer for vinflunine with BSC compared to BSC
(6.9 vs. 4.3 months, P=0.04). Based on this study,
vinflunine was approved by the European Medicine
Agency (EMA) and is the only agent approved by a
regulatory agency for this indication. The key grade
3 or 4 toxicities for vinflunine were neutropenia
(50%), febrile neutropenia (6%), anemia (19%),
fatigue (19%), and constipation (16%). A recent
retrospective analysis of patients who received
second-line vinflunine identified ECOG performance
status greater than 0, hemoglobin less than 10 g/dL,
and liver metastasis as poor prognostic factors [55].

gemcitabine-paclitaxel with continuation beyond 6


cycles until progression [72]. None of the patients
had received previous paclitaxel, and approximately
half had received previous gemcitabine. The median
disease-specific survival was 7-8 months and the
median progression-free survival was approximately
3.5 months in both groups. The strategy of a
fixed number of cycles versus continuation until
progression could not be evaluated since a mean of
only 4 cycles were delivered in both groups.
Table 1. Reported Phase II Trials of Second-line
Therapy for Metastatic Urothelial Carcinoma

Numerous other chemotherapeutic and targeted


agents have been evaluated as second-line therapy
in non-randomized phase II trials and modest
or marginal activity has been demonstrated for
some agents (Table 1). Eligibility criteria for these
reported phase II trials have been highly variable
and heterogeneous, enrolling patients treated with
peri-operative chemotherapy followed by frontline
therapy for metastatic disease or enrolling those who
had received peri-operative chemotherapy alone
and with no requirement for a defined treatmentfree interval after frontline therapy. Prior treatment
may not be clearly defined. This renders comparison
of outcomes across trials difficult; the corollary is
that going forward, it might be important to enroll
homogeneous patients in predefined chemosensitive
or chemoresistant strata. Generally, these trials
report response rates of 5-25%, median progressionfree survival of 2 to 3 months and median survivals
of 6 to 9 months (Table 1) [56-70].
Taxanes (paclitaxel, docetaxel, nanoparticle-albuminbound paclitaxel) have been evaluated following firstline GC, while gemcitabine and the taxanes, alone
or in combination have been employed following
M-VAC. Second-line pemetrexed, a multitargeted
antifolate and active as first-line therapy, illustrates
the degree of heterogeneous outcomes in previouslytreated patients. One study demonstrated modest
activity of 8% and did not attain the level of activity
required to continue accrual to the second stage in
another trial (60;64). Conversely, another phase II
trial in presumably comparable patients employed
pemetrexed in the second-line setting of metastatic
urothelial carcinoma and demonstrated a response
rate of 28% [71].

VIII. Combination Therapy


Combination regimens have also been evaluated
as second-line therapy in phase II trials (Table 1),
although a potential incremental benefit compared
to monotherapy remains undefined. A German
randomized phase II trial of 102 patients compared
the strategy of administering 6 cycles of second-line

Drug

N
56

RR
%
20

PFS
(mo)
2.4

OS
(mo)
5.5

Ifosfamide (62)
Gemcitabine (66)

30

11

4.9

8.7

Gemcitabine (67)

35

22.5 -

5.0

Weekly Paclitaxel (68)

31

10

2.2

7.2

Docetaxel (65)

30

13

9.0

Nab-paclitaxel (70)

35

44

Paclitaxel-Gemcitabine
41
(31)
Ifosfamide-Gemcitabine
34
(105)
Carboplatin-Paclitaxel (73) 44

60

14.4

21

4.0

9.0

16

4.0

6.0

Pemetrexed (60)

47

27.7 2.9

9.6

Pemetrexed (64)

12

Ixabepilone (58)

42

11.9 2.7

8.0

Oxaliplatin (61)

18

1.5

7.0

Vinflunine (53)

175 15

2.8

8.2

Vinflunine (52)

51

18

3.0

6.6

Irinotecan (57)

40

2.1

5.4

Topotecan (69)

44

9.1

1.5

6.3

Gefitinib (106)

31

3.0

Lapatinib (63)

59

4.5

Sorafenib (80)

27

6.8

Sunitinib (77)

45

2.4

6.9

Pazopanib (79)

18

22

Index: (-)=Not available or stipulated in publication; N=number of patients; RR=response rate;


PFS=progression-free survival; OS=overall survival

399

Another trial evaluated carboplatin-paclitaxel


following prior cisplatin-based chemotherapy not
including paclitaxel, and reported a response rate of
16%, median progression-free survival of 4 months,
and median survival of 6 months [73]. A phase I/II trial
evaluated a combination of salvage weekly cisplatin,
gemcitabine, and ifosfamide in a heterogeneous
group of patients who had received previous
platinum-based chemotherapy [74]. The response
rate was 40.8% but hematologic toxicities appeared
prohibitive. Similarly, the combination of pemetrexed
and gemcitabine has demonstrated moderate activity
coupled with substantial myelosuppression [30;75].
Scant data support re-administration of secondline M-VAC following prior first-line M-VAC in those
with an excellent previous quality of response and
relatively prolonged time to progression [76]. M-VAC
appears to have activity after GC, although this has
not been formally reported.

Akt/mTOR, Ras), immune modulation (CTLA-4


antagonists), and strategies to target the putative
tumor initiating stem cells. Given the rapid proliferation
that characterizes metastatic urothelial carcinoma,
chemotherapeutic agents targeting DNA replication
and/or cell division may continue to play a role even
in the era of biologic agents (Table 2). While early
disease appears to segregate into two molecular
pathways (papillary tumors are characterized by
gain-of-function mutations affecting oncogenes such
as RAS and FGFR3, and carcinoma in situ (Tis)
and invasive tumors are characterized by loss-offunction mutations affecting tumor suppressor genes
p53, RB, and PTEN). A critically important driver of
urothelial carcinogenesis in the setting of advanced
disease remains elusive.

1. Chemotherapy
Eribulin is a synthetic derivative of the marine
sponge product halichondrin-B that inhibits
tubulin polymerization, and is being evaluated as
monotherapy in the first- or second-line setting,
and in combination with GC. Novel anti-mitotic
agents that inhibit the kinesin spindle protein (AZD4877), the polo-like kinase (BI-6727) and a potent
methotrexate analogue, pralatraxate are all being
evaluated in the second-line setting (Table 2). Nabpaclitaxel is being evaluated as a single agent in
previously treated patients and in combination with
carboplatin and gemcitabine as neoadjuvant therapy
preceding radical cystectomy.

IX. Level of Evidence and


Grades of Recommendation
for Second-line and Salvage
Therapy
Since a survival benefit was not demonstrated with
vinflunine by intention-to-treat analysis in the only
phase III trial conducted in this setting and modest
efficacy (that appears similar to activity of other
available agents), in our opinion, these data represent
level 2a evidence and a grade B recommendation.
Additionally, other agents that have been evaluated
in smaller numbers of patients in phase II trials
also represent level 2a evidence and a grade B
recommendation. Given the poor outcomes with
currently available agents and regimens and the
palliative nature of such therapy, clinical trials should
remain the preferred option for the second-line and
salvage therapy of metastatic urothelial carcinoma
(Tables 2 and 3). Given the observation that these
patients progress rapidly and poor performance might
frequently preclude systemic therapy, the strategy of
early second-line therapy in patients with stable or
responding disease after first-line chemotherapy is
being investigated (Table 3).

2. Angiogenesis-targeting Agents
Modest activity has been demonstrated in phase II
trials of the multi-targeted VEGFR tyrosine kinase
inhibitor, sunitinib as frontline or salvage therapy
of metastatic urothelial carcinoma [77;78]. In the
salvage setting of a heavily pretreated population
that received sunitinib 50 mg daily for 4 of 6 weeks
or 37.5 mg daily, a partial response was seen in 3
of 45 patients, and in 1 of 32 patients respectively.
Clinical regression or stable disease was achieved
in 33 of 77 (43%) patients. The median progressionfree survival (2.4 months) and overall survival (6-7
months) were disappointing. There were substantial
toxicities with one treatment-related death and 47
patients experiencing grade 3 or 4 toxicities.
While sunitinib has demonstrated modest activity in
the salvage setting (described earlier), a frontline trial
enrolled cisplatin-ineligible patients with a creatinine
clearance of 30-60 mL/min and ECOG performance
status 1 or greater to receive sunitinib 50 mg daily
for 4 weeks of every 6 weeks [77]. Of 14 evaluable
patients, 2 partial responses (14.3%) were obtained,
and 9 patients (64.3%) had SD lasting greater than
3 months and the median progression-free survival
was 6 months. This agent is under further study in
combination with GC as neoadjuvant treatment and
as maintenance therapy in patients experiencing

X. Novel Agents for Advanced


Urothelial Carcinoma
A better understanding of the molecular biology of
urothelial carcinoma is yielding promising, albeit
unvalidated results. Molecular targets including
angiogenic pathways (VEGF, angiopoietin/Tie2,
FGFR3), growth factor tyrosine kinase receptors
(EGFR, Her2, Met, IGF-1R), cytoplasmic signaling
pathways of survival and proliferation (PI3K/

400

complete response, partial response, or stable


disease after first-line therapy.

dose reduction [82]. Based on these promising results, the CALGB has launched a critically important
first-line phase III trial that compares GC with GCbevacizumab. Another phase II trial is evaluating the
combination of carboplatin, gemcitabine, and bevacizumab for cisplatin-ineligible patients [83]. Separate phase II trials are evaluating neoadjuvant GC
or DD-M-VAC plus bevacizumab followed by radical
cystectomy with a goal of improving pathologic complete remissions and long-term outcomes.

Another single institution Italian experience with


pazopanib (a multi-targeted VEGFR tyrosine kinase
inhibitor) pre-treated patients was reported at the
ESMO meeting in October 2010. Of 18 heavily
pre-treated patients with a median of 5 (312)
prior regimens, 4 (22%) patients obtained a partial
response and 11 (61%) had stable disease [79].
Sorafenib did not demonstrate significant activity
as first- or second-line therapy by itself, but its
evaluation in combination with frontline gemcitabine
plus carboplatin is ongoing (Table 2) [80;81].

Aflibercept (VEGF-trap), a VEGF receptor fusion


protein that has greater affinity for VEGF than
bevacizumab and also targets placental growth factor
(PlGF), demonstrated poor activity as second-line
monotherapy [84]. Based on preliminary evidence
for activity in a phase I trial, a randomized phase
II trial in the second-line setting will evaluate the
value of adding AMG-386 (angiopoietin neutralizing
peptibody) to carboplatin-paclitaxel [85]. Another
randomized phase II trial is planning the secondline evaluation of a combination of docetaxel with
monoclonal antibodies against either VEGFR1 (IMC18F1) or VEGFR-2 (IMC-1121B) (Table 3).

A randomized phase II trial is evaluating salvage


docetaxel alone or with vandetanib (a VEGFR and
EGFR targeting agent) in patients who have received
up to 3 prior regimens (Table 3).
Dovitinib, a multitargeted inhibitor of VEGF and
FGF receptors is being evaluated in the secondline setting after evaluation of FGFR-3 expression
(Table 2 and 3). A phase II trial evaluating first-line
GC plus bevacizumab for metastatic urothelial carcinoma demonstrated promising activity (overall response rate 58%, median progression-free survival
8.2 months, median overall survival 19.1 months),
accompanied by a significant rate of venous thromboembolism that was ameliorated with gemcitabine

3. EGFR and Her2-targeting Agents


Disappointingly, gefitinib alone or in combination with
GC demonstrated poor or no activity [86;87]. Erlotinib
has demonstrated biologic activity in the neoadjuvant
setting and further evaluation may be warranted in

TABLE 2. Key Ongoing or Planned Non-randomized Trials Employing Novel Agents for Urothelial
Carcinoma*
Line of Therapy

Drug/Regimen

Molecular Target

First or second
First
Neoadjuvant
First (cisplatin-ineligible)
Neoadjuvant
First
Neoadjuvant, first-line
Neoadjuvant
First-line
Neoadjuvant
Second
Second
Second
Second
Second
Second
Second
First-line cisplatin-ineligible,
second-line
First-line

Eribulin
GC-Eribulin
GCa-Nab-paclitaxel
GCa-Bevacizumab
DD-MVAC-bevacizumab
GCa-Sorafenib
GC-Sunitinib
GC-Sorafenib
GC-lenalidomide
Dasatinib
AZD-4877
Pralatrexate
Volasertib
Pazopanib
Dovitinib
Tamoxifen
Everolimus
Everolimus-Paclitaxel

Tubulin
Tubulin
Tubulin
VEGF
VEGF
VEGFR, PDGFR, Raf
VEGFR, PDGFR
VEGFR, PDGFR
Immune-modulation, angiogenesis
Src
Kinesin spindle protein
Folate
Polo-like kinase
VEGFR, PDGFR
VEGFR, PDGFR, FGFR
Estrogen receptors
mTOR
mTOR and tubulins

GC-Temsirolimus

mTOR

Index: *- From www.ClinicalTrials.Gov accessed 20 October 2010 (107), GC=gemcitabine plus cisplatin,
GCa=gemcitabine plus carboplatin, Nab-paclitaxel=nanoparticle albumin bound paclitaxel, DD-MVAC= dose dense
M-VAC

401

selected patients. An ongoing randomized phase II


trial is evaluating the combination of cetuximab with
front-line GC.

treated with first-line trastuzumab in combination


with paclitaxel, carboplatin, and gemcitabine in a
phase II trial [88]. Thirty-one (70%) of 44 patients
responded (5 complete responses and 26 partial
responses), although it is unclear if these patients
unequivocally had Her2-driven tumors since gene
amplification was infrequent. Median time to
progression and survival were 9.3 and 14.1 months,
respectively. Disappointingly, a randomized phase II
trial that attempted to compare GC with or without
trastuzumab as front-line therapy and another phase
II trial that attempted to evaluate trastuzumab as
second-line monotherapy for Her2 over-expressing
metastatic urothelial carcinoma closed early
due to a lower than expected frequency of Her2
overexpression. Also, a phase II trial is evaluating
trastuzumab in combination with paclitaxel and
radiotherapy as a bladder-conserving regimen in
patients with localized/locally advanced urothelial
carcinoma of the bladder.

In a phase II trial of patients with EGFR and/or Her2


expression, lapatinib demonstrated partial responses
in 3% and response plus stability 16 weeks or longer
in 12% of patients with a trend towards apparent antitumor effect in those with EGFR or Her2 expression
2+/3+ by immunohistochemistry (IHC) [63]. A
randomized British trial is evaluating maintenance
lapatinib or placebo in patients with EGFR and/or
Her2-expressing tumors with stable or responding
disease after frontline chemotherapy (Table 3).
In the EORTC Genitourinary Group, escalating
doses of lapatinib in combination with GC are under
evaluation in a dose-finding study.
In one study, patients with metastatic urothelial or
squamous carcinoma whose tumors expressed
Her2/neu (by IHC, serology, or fluorescence in situ
hybridization) in the primary or metastatic site were

Table 3. Key Ongoing or Planned Randomized Trials of Systemic Therapy*


Target

Line of
therapy

Phase

Eligibility

Group 1

Group 2

First-line

III

Chemo-naive or 1 year after


peri-operative chemotherapy

GC

GC-Bevacizumab

First-line

II

Chemo-naive or 1 year after


peri-operative chemotherapy

GC

GC
Cetuximab

Cisplatin-ineligible

GCa

GCa-Vandetanib

Docetaxel

Docetaxel Cetuximab

EGFR

Placebo

Sunitinib

VEGFR, PDGFR

Placebo

Lapatinib

Her2, EGFR

Cetuximab Paclitaxel
Docetaxel Vandetanib

EGFR
VEGFR, EGFR

Sunitinib

VEGFR, PDGFR

Docetaxel-IMC18F1

VEGFR2,
VEGFR1

First-line
Second-line

II
II

Second-line

II

Second-line

II

Second-line

II

Salvage

II

Consolidation following
response/stability after
frontline therapy
Consolidation following
response/stability after
frontline therapy
Consolidation following
response/stability after
frontline therapy
Progression after frontline
regimen
Progression after 3 prior
regimens
Progressive disease,
intolerant or ineligible for
cisplatin

First or secondline

II

Second-line

II

Progressive disease

Second-line

II

Progressive disease

Cetuximab
Docetaxel
Placebo
DocetaxelIMC1121B
Carboplatinpaclitaxel

Carboplatinpaclitaxel- AMG386

VEGF

EGFR
VEGFR, EGFR

Angiopoietin/Tie2

Index: * From www.Clinicaltrials.gov accessed October 20, 2010, GC=gemcitabine plus cisplatin, GCa=gemcitabine,
carboplatin, IMC1121B=monoclonal antibody against VEGFR2, IMC18F1=monoclonal antibody against VEGFR1

402

4. Other Molecular Targets

would potentially be extremely beneficial and advance the field. Several studies have made progress
towards this end. In the neoadjuvant setting, Takata
et al have employed a cDNA microarray platform to
profile gene expression from biopsies of 27 patients
with invasive bladder cancer who were subsequently treated with M-VAC neoadjuvant chemotherapy
[100]. Based on supervised analyses designed to
discover genes differentially expressed between
14 M-VAC responders (defined as downstaging to
T1 or less) and 13 non-responders (defined as T2
or greater stage), a set of 50 genes was identified,
25 overexpressed and 25 underexpressed, that differed significantly between these groups. Splitting
their data into an 18 sample learning and 9 sample
test set, they developed a 14-gene M-VAC response
classifier based on the learning samples that correctly classified 8 of 9 test set samples. Quantitative
RT-PCR to assess expression of these genes for
classification purposes was used to validate these
results [100]. However, such signatures are of much
greater potential value if they are independently and
prospectively validated on novel test sets. To that
end, the same group performed a validation study of
their predictive platform on 22 additional cases, profiled as before [97]. The signature correctly predicted
response in 19 of 22 additional cases, providing a
sensitivity of ~100% with a specificity of ~73%. Most
importantly, in terms of potential application of this
classification scheme for patients with invasive bladder cancer, in this independent test set, the positive
predictive value for the assay was approximated at
~79% (11 of 14 cases), while the negative predictive value was approximated at ~100% (8 of 8 nonresponders identified).

Expression of the IGF-1R appears to increase with


stage and correlate with poor outcomes [89]. A phase
II trial is planning the evaluation of combination GC
and cixutumumab, a fully human monoclonal IgG1
antibody against IGF-1R, as first-line therapy for
metastatic urothelial carcinoma.
Based on ER- expression in urothelial carcinoma
and the inhibitory effect of selective estrogen
receptor modulators in preclinical models, salvage
therapy with oral tamoxifen is being evaluated (Table
3) [90;9].
The mTOR inhibitors, everolimus and temsirolimus
are being evaluated as monotherapy or in combination
with chemotherapy (Table 2). Immune-modulation
with ipilimumab, a monoclonal antibody against
CTLA-4 has demonstrated immune-modulating
activity in the neoadjuvant setting, and a trial is being
planned in combination with frontline chemotherapy
(Table 2) [92]. Lenalidomide is another agent with
immune-modulating and anti-angiogenic activity
and is being evaluated in combination with frontline
combination chemotherapy (Table 2). Src inhibitors
have demonstrated preclinical activity, and a trial is
evaluating dasatinib as neoadjuvant therapy with
correlative endpoints [93;94].

XI. New Trial Designs and the


Development of Tailored
Therapy
Innovative clinical trial paradigms may enable the
rapid evaluation and approval of novel agents. The
consolidation second-line strategy following frontline
therapy in stable or responding patients employing
biologic agents is being pursued with sunitinib and
lapatinib. Neoadjuvant therapy before surgery in
localized disease permits rapid in vivo assessment
of pathologic response and biologic activity.

A second study examined cisplatin-based


combination chemotherapy (M-VAC and GC) in 30
patients with locally advanced or metastatic invasive
bladder cancer (T4b, N2-3, or metastatic M1). An
oligonucleotide microarray platform was used to
profile the gene expression of tumors from these
patients and related it to survival [98]. A set of 55
genes was differentially expressed between these
groups; interestingly, all 55 were underexpressed
in patients with long-term survival. No difference in
the genes correlated with survival between patients
treated with M-VAC versus GC was noted, suggesting
that gene expression differences may underlie
the similar efficacy of these regimens. To answer
the question whether any of these genes might
be adaptable to traditional immunohistochemical
validation, they selected two, emmprin and survivin,
for immunohistochemistry, and found that they were
highly significantly associated with overall survival
singularly or in combination (all cases P 0.001).
Lastly, addressing the issue of specificity of these
genes to response of these drugs as opposed to
some other survival parameter, they found that
emmprin positive and negative tumors exhibited

The goal of personalized therapy may be


attainable. Factors predictive of response may
facilitate personalized therapy and enable judicious
patient selection even in the early stages of drug
development. Low levels of DNA-repair genes
(ERCC1 and BRCA1) and tumor gene expression
profiling or genomics based on in vitro drug
sensitivities in the NCI-60 cancer cell line panel
appear to be associated with better long-term
outcomes [95-99]. A vigorous and multidisciplinary
international effort to validate these observations will
enable more rapid advances.

XII. Predictive Models


Selection of patients with a high likelihood of benefit

403

differential response rates of 39% versus 74%,


while survivin positive and negative tumors exhibited
response rates of 47% and 70%, respectively. Double
negative tumors exhibited a high response rate of
82%, whereas double positive tumors exhibited a
response rate of only 27%. This resulted in an odds
ratio for response in double negative tumors of
11.9 (95% CI, 3.2-42.3). Based on these findings,
the authors proposed that if emmprin and survivin
immunohistochemical detection were independently
validated in large patient cohorts and staining were
inter-institutionally standardized, these markers
might be used to stratify patients into high and low
likelihood of response groups, with likely responders
treated by traditional cisplatin-based methods and
likely non-responders assigned to investigational
therapies [98].

gene expression-based or immunohistochemicaldetected biomarkers could become part of the


diagnostic workup and management in the future
(level of evidence 2b, Grade of recommendation: Not
recommended outside a clinical trial or biomarker
study). Unfortunately, the majority of these promising
biomarkers have not yet undergone appropriate
assay validation in prospectively collected samples
so that full biomarker qualification studies can be
performed. Implementation of molecular markers
on which to base treatment decisions in bladder
cancer lagged behind other diseases but it is hoped
that the strategies discussed above can be validated
in prospective studies in order to move the field
forward.

In the adjuvant setting, a study [101] assessed the


gene expression of the chemotherapy response
modifiers multidrug resistance gene 1 (MDR1) and
excision repair cross-complementing 1 (ERCC1) to
determine if they are useful in identifying a group
of patients benefiting from cisplatin-based adjuvant
chemotherapy. Formalin-fixed paraffin-embedded
tumor samples from 108 patients with locally
advanced bladder cancer, who had been enrolled
in AUO-AB05/95, a phase III trial randomizing
a maximum of 3 courses of adjuvant cisplatin
and methotrexate (CM) versus methotrexate,
vinblastine, epirubicin, and cisplatin (M-VEC), were
included in the study. Tumor cells were retrieved by
laser-captured micro-dissection and analyzed for
MDR1 and ERCC1 expression using a quantitative
real-time reverse transcription-polymerase chain
reaction assay. Importantly, expressions of MDR1
and ERCC1 were independently associated with
overall progression-free survival. The correlation of
high MDR1 expression with inferior outcome was
stronger in patients receiving M-VEC, whereas
ERCC1 analysis performed equally in the CM and
M-VEC groups.

Combination cisplatin regimens are a standard of

Recommendations
care for patients with a Grade A recommendation
for first-line advanced urothelial cancer with
adequate renal function.

In patients with renal impairment, advanced


age or poor performance status, carboplatin
and gemcitabine is a grade A recommendation
and for patients with renal impairment only use
of gemcitabine/carboplatin/paclitaxel can be
considered (Grade C).

U
 se of triplet regimens adding paclitaxel
to
gemcitabine
recommended.

plus

cisplatin

is

not

F
 or second-line therapy after a platinum-based

treatment, vinflunine has been adopted as a


standard in Europe but not in North America.

U
 se of vinflunine is a grade B recommendation.

N
 ovel agents and approaches are urgently needed
in the treatment of urothelial cancer and clinical
trial accrual needs to be actively encouraged by
urologists, medical oncologists, and radiation
oncologists. In this regard, the standard of care
for a majority of patients with advanced bladder
cancer is the best clinical trial available.

Gene expression-based predictive models may


not only be predictive of single drug therapeutic
response, but could also be combined to successfully
predict combination responses [99;102-104].
This approach uses a novel algorithm term CoeXpression ExtrapolatioN (COXEN), which uses
expression microarray data as a Rosetta Stone for
translating between drug activities in a cancer cell
line panel to drug activities in a set of clinical tumors.
This technique has been used in bladder cancer and
has also successfully predicted clinical outcome in
breast and ovarian cancer. More than 500 patients
have been evaluated, of which 233 patients were
enrolled in clinical studies.

XIII. Abstract
Advanced metastatic urothelial carcinoma of the
bladder and other bladder cancers are responsive to
chemotherapy. Cisplatin-containing chemotherapy
regimens such as M-VAC, dose dense M-VAC
(DD-M-VAC) and cisplatin with gemcitabine can
improve quality of life and overall survival. Cisplatinbased regimens can produce complete responses
in approximately 15% of patients, which can be
durable. Major unmet needs exist for patients who

In summary, while not yet extensively validated


in a sufficient number of samples obtained from
prospective clinical trials, there is the potential that

404

fail first line standard cisplatin-based regimens, are


precluded from cisplatin treatment because of renal
impairment, or are not suitable for cisplatin due to
comorbid conditions or advanced age. For cisplatinineligible patients, the combination of carboplatin
with gemcitabine has emerged as a standard of care.
For patients with platinum-refractory cancer, several
chemotherapeutic agents have demonstrated activity.
These include: vinflunine, taxanes, ifosfamide, folate
modulators, and anthracyclines; of these, only
vinflunine has been compared to best supportive
care in a phase III trial. Novel targeted agents under
investigation may be of interest in the second line or
maintenance setting. To date, no second line agent
has yet demonstrated definitive clinical benefit in
a randomized controlled trial. Therefore, a clinical
trial remains an excellent option for all patients with
advanced urothelial cancer.

or without recombinant human granulocyte-macrophage


colony-stimulating factor for the initial treatment of advanced
malignant urothelial tumors: results of a randomized trial. J
Clin Oncol 1995;13(9):2272-7.
11. Sternberg CN, de Mulder P, Schornagel JH et al. Seven
year update of an EORTC phase III trial of high dose
intensity M-VAC chemotherapy and G-CSF versus classic
M-VAC in advanced urothelial tract tumors (EORTC
protocol 30924). Eur J Cancer 2006;42(1):50-4.
12. Bamias A, Aravantinos G, Deliveliotis C. Docetaxel
and cisplatin with granulocyte colony-stimulating factor
(G-CSF) versus M-VAC with G-CSF in advanced urothelial
carcinoma: a multicenter, randomized, phase III study from
the Hellenic Cooperative Oncology Group. J Clin Oncol
2004;22:220-8.
13. von der Maase H. Gemcitabine and Cisplatin in Locally
Advanced and/or Metastatic Blaccer Cancer. Eur J Cancer
2000;36 Suppl 2:13-6.
14. Kaufman D, Raghavan D, Carducci M et al. Phase II trial
of gemcitabine plus cisplatin in patients with metastatic
urothelial cancer. J Clin Oncol 2000;18 (9):1921-7.
15. Lorusso V, Manzione L, De Vita F, Antimi M, Selvaggi
FP, De Lena M. Gemcitabine plus Cisplatin for Advanced
Transitional Cell Carcinoma of the Urinary Tract: A Phase II
Multicenter Trial. J Urol 2000;164(1):53-6.

References
1. Logothetis CJ, Dexeus F, Finn L et al. A prospective
randomized trial comparing CISCA to MVAC chemotherapy
in advanced metastastic urothelial tumors. J.Clin.Oncol.
1990;8:1050-5.

16. Moore MJ, Winquist EW, Murray N et al. Gemcitabine


Plus Cisplatin, an Active Regimen in Advanced Urothelial
Cancer: A Phase II Trial of the National Cancer Institute of
Canada Trials Group. J Clin Oncol 1999;17(9):2876-81.

2. Sternberg CN, Yagoda A, Scher HI et al. M-VAC for


advanced transitional cell carcinoma of the urothelium:
Efficacy, and patterns of response and relapse. Cancer
1989;64:2448-58.

17. von der Maase H, Hansen SW, Roberts JT et al. Gemcitabine


and cisplatin versus methotrexate, vinblastine, doxorubicin,
and cisplatin in advanced or metastatic bladder cancer:
results of a large, randomized, multinational, multicenter,
phase III study. J Clin Oncol 2000;18(17):3068-77.

3. Bajorin DF, Dodd PM, Mazumdar M et al. Long-term


survival in metastatic transitional-cell carcinoma and
prognostic factors predicting outcome of therapy. J Clin
Oncol 1999;17(10):3173-81.

18. von der Maase H, Sengelov L, Roberts JT et al. Long-term


survival results of a randomized trial comparing gemcitabine
plus cisplatin, with methotrexate, vinblastine, doxorubicin,
plus cisplatin in patients with bladder cancer. J Clin Oncol
2005;23(21):4602-8.

4. Loehrer P.J.Sr, Einhorn LH, Elson PJ et al. A randomized


comparison of cisplatin alone or in combination with
methotrexate, vinblastine, and doxorubicin in patients with
metastatic urothelial carcinoma: a cooperative group study.
J Clin Oncol 1992;10(7):1066-73.

19. Dreicer R, Manola J, Roth BJ, Cohen MB, Hatfield AK,


Wilding G. Phase II study of cisplatin and paclitaxel
in advanced carcinoma of the urothelium: an Eastern
Cooperative Oncology Group Study. J Clin Oncol
2000;18(5):1058-61.

5. Saxman SB, Propert KJ, Einhorn LH et al. Long-term


follow-up of a phase III intergroup study of cisplatin
alone or in combination with methotrexate, vinblastine,
and doxorubicin in patients with metastatic urothelial
carcinoma: a cooperative group study. J Clin Oncol
1997;15(7):2564-9.

20. Burch PA, Richardson RL, Cha SS et al. Phase II Trial of


Combination Paclitaxel and Cisplatin in Advanced Urothelial
Carcinoma (UC). J Urol 2000;164:1538-42.

6. Tannock I, Gospodarowicz M, Connolly J, Jewett M. M-VAC


(methotrexate, vinblastine, doxorubicin and cisplatin)
chemotherapy for transitional cell carcinoma: the Princess
Margaret Hospital experience. J.Urol. 1989;142:289-92.

21. Garcia del Muro X, Marcuello E, Guma J et al. Phase II


multicentre study of docetaxel plus cisplatin in patients with
advanced urothelial cancer. Br J Cancer 2002;86(3):32630.

7. Sternberg C, de Mulder P, van Oosterom AT, et al.


Escalated M-VAC chemotherapy and recombinant human
granulocyte macrophage colony stimulating factor (GMCSF) in patients with advanced urothelial tract tumors. Ann
Oncol 1993;4(5):403-7.

22. Sengelov L, Kamby C, Lund B, Engelholm SA. Docetaxel


and cisplatin in metastatic urothelial cancer: a phase II
study. J Clin Oncol 1998;16(10):3392-7.
23. Dimopoulos MA, Bakoyannis C, Georgoulias V et al.
Docetaxel and cisplatin combination chemotherapy in
advanced carcinoma of the urothelium: a multicenter phase
II study of the Hellenic Cooperative. Ann Oncol 1999;10
(11):1385-8.

8. Loehrer PJS, Elson P, Dreicer R et al. Escalated dosages


of methotrexate, vinblastine, doxorubicin, and cisplatin plus
recombinant human granulocyte colony-stimulating factor
in advanced urothelial carcinoma: an Eastern Cooperative
Oncology Group trial. J Clin Oncol 1994;12(3):483-8.

24. Small EJ, Lew D, Redman BG et al. Southwest Oncology


Group Study of Paclitaxel and Carboplatin for Advanced
Transitional Cell Carcinoma: The Importance of Survival as
a Clinical Trial End Point. J Clin Oncol 2000;18 (13):253744.
25. Redman BG, Smith DC, Flaherty L, Du W, Hussain M.
Phase II trial of paclitaxel and carboplatin in the treatment
of advanced urothelial carcinoma. J Clin Oncol 1998;16
(5):1844-8.

9. Seidman AD, Scher HI, Gabrilove JL et al. Doseintensification of methotrexate, vinblastine, doxorubicin,
and cisplatin with recombinant granulocyte-colony
stimulating factor as initial therapy in advanced urothelial
cancer. J.Clin.Oncol. 1992;11:414-20.
10. Logothetis CJ, Finn LD, Smith T et al. Escalated MVAC with

405

26. Vaughn DJ, Malkowicz SB, Zoltick B et al. Paclitaxel plus


carboplatin in advanced carcinoma of the urothelium: an
active and tolerable outpatient regiment. J Clin.Oncol.
1998;16 (1):255-60.

42. Galsky MD, Hahn NM, Rosenberg J et al. A Consensus


Definition of Patients with Metastatic Urothelial Carcinoma
Unfit for Cisplatin-based Chemotherapy. Lancet Oncol .
2010. Ref Type: In Press

27. Pycha A, Grbovic M, Posch B et al. Paclitaxel and


carboplatin in patients with metastatic transitional cell
cancer of the urinary tract. Urol 1999;53(3):510-5.

43. Galsky MD, Hahn NM, Rosenberg JE et al. Defining


cisplatin ineligible patients with metastatic bladder cancer.
Genitourinary Cancer Symposium , 105. 2011. Ref Type:
Abstract

28. Vaughn DJ, Manola J, Dreicer R, See W, Levitt R, Wilding


G. Phase II study of paclitaxel plus carboplatin in patients
with advanced carcinoma of the urothelium and renal
dysfunction (E2896): a trial of the Eastern Cooperative
Oncology Group. Cancer 2002;95(5):1022-7.

44. Xu N, Zhang XC, Xiong JP et al. A phase II trial of


gemcitabine plus carboplatin in advanced transitional cell
carcinoma of the urothelium. BMC Cancer 2007;7:98.
45. Linardou H, Aravantinos G, Efstathiou E et al. Gemcitabine
and carboplatin combination as first-line treatment
in elderly patients and those unfit for cisplatin-based
chemotherapy with advanced bladder carcinoma: Phase II
study of the Hellenic Co-operative Oncology Group. Urol
2004;64(3):479-84.

29. Dreicer R, Manola J, Roth BJ et al. Phase III trial of methotrexate, vinblastine, doxorubicin, and cisplatin versus carboplatin and paclitaxel in patients with advanced carcinoma
of the urothelium. Cancer 2004;100(8):1639-45.
30. von der Maase H, Lehmann J, Gravis G et al. A phase II trial
of pemetrexed plus gemcitabine in locally advanced and/
or metastatic transitional cell carcinoma of the urothelium.
Ann Oncol 2006;17(10):1533-8.

46. Bamias A, Moulopoulos LA, Koutras A et al. The combination


of gemcitabine and carboplatin as first-line treatment in
patients with advanced urothelial carcinoma. A Phase II
study of the Hellenic Cooperative Oncology Group. Cancer
2006;106(2):297-303.

31. Sternberg CN, Calabr F, Pizzocaro G, Marini L, Schnetzer


S, Sella A. Chemotherapy with Every-2-Week Gemcitabine
and Paclitaxel in Patients with Transitional Cell Carcinoma
who have received prior Cisplatin-based therapy. Cancer
2001;92(12):2993-8.

47. Bellmunt J, Ribas A, Eres N et al. Carboplatin-Based


versus Cisplatin-Based chemotherapy in the treatment of
surgically incurable advanced bladder carcinoma. Cancer
1997;80:1966-72.

32. Meluch AA, Greco FA, Burris HA3 et al. Paclitaxel and
gemcitabine chemotherapy for adanced transitional-cell
carcinoma of the urothelial tract: a Phase II trial of the
Minnie Pearl Cancer Research Network. J Clin Oncol
2001;19(12):3018-24.

48. Petrioli R, Frediani B, Manganelli A et al. Comparison


between a cisplatin-containing regimen and a carboplatincontaining regimen for recurrent or metastatic cancer
patients: A randomized phase II study. Cancer 1996;77:34451.

33. Kaufman DS, Carducci MA, Kuzel TM et al. A multiinstitutional phase II trial of gemcitabine plus paclitaxel
in patients with locally advanced or metastatic urothelial
cancer. Urol Oncol 2004;22(5):393-7.

49. Dogliotti L, Carten G, Siena S et al. Gemcitabine plus


cisplatin versus gemcitabine plus carboplatin as first-line
chemotherapy in advanced transitional cell carcinoma of
the urothelium: results of a randomized phase 2 trial. Eur
Urol 2007;52(1):134-41.

34. Calabr F, Lorusso V, Rosati G et al. Gemcitabine and


paclitaxel every 2 weeks in patients with previously untreated
urothelial carcinoma. Cancer 2009;115(12):2652-9.

50. De Santis M, Bellmunt J, Mead G et al. Randomized


phase II/III trial assessing gemcitabine/ carboplatin and
methotrexate/carboplatin/vinblastine in patients with
advanced urothelial cancer unfit for cisplatin-based
chemotherapy: phase II--results of EORTC study 30986. J
Clin Oncol 2009;27(33):5634-9.

35. Srinivas S, Guardino AE. A nonplatinum combination in


metastatic transitional cell carcinoma. Am J.Clin Oncol
2005;28(2):114-8.
36. Bellmunt J, Guillem V, Paz-Ares L et al. Phase I-II study
of paclitaxel, cisplatin, and gemcitabine in advanced
transitional-cell carcinoma of the urothelium. J Clin Oncol
2000;18(18):3247-55.

51. De Santis M, Bellmunt J, Mead G et al. Randomized


phase II/III trial comparing gemcitabine/carboplatin (GC)
and methotrexate/carboplatin/vinblastine (M-CAVI) in
patients (pts) with advanced urothelial cancer (UC) unfit for
cisplatin-based chemotherapy (CHT): Phase III results of
EORTC study 30986. Proc Annu Meet Am Soc Clin Oncol
28(18s). 2010. Ref Type: Abstract

37. Bajorin DF, McCaffrey JA, Dodd PM et al. Ifosfamide,


Paclitaxel, and Cisplatin for Patients with Advanced
Transitional Cell Carcinoma of the Urothelial Tract: Final
Report of a Phase II Trial evaluating two Dosing Schedules.
Cancer 2000;88(7):1671-8.

52. Culine S, Theodore C, De Santis M et al. A phase II


study of vinflunine in bladder cancer patients progressing
after first-line platinum-containing regimen. Br J Cancer
2006;94(10):1395-401.

38. Hussain M, Vaishampayan U, Du W, Redman B, Smith DC.


Combination Paclitaxel, Carboplatin, and Gemcitabine is
an active treatment for advanced urothelial cancer. J Clin
Oncol 2001;19(9):2527-33.

53. Vaughn D, Srinivas S, Stadler WM et al. Vinflunine in


Platinum-Pretreated Patients with Locally Advanced
or Metastatic Transitional Cell Carcinoma of the
Urothelium: Results of a Large Phase 2 Study. Cancer
2009;115(18):4110-7.

39. Bellmunt J, von der Maase H, Mead GM et al. Randomized


phase III study comparing paclitaxel/cisplatin/gemcitabine
and gemcitabine/cisplatin in patients with locally advanced
or metastatic urothelial cancer without prior systemic
therapy: EORTC 30987 / Intergroup Study. Proc Annu
Meet Am Soc Clin Oncol 25(18S), 242s. 2007. Ref Type:
Abstract

54. Bellmunt Molins J, von der Maase H, Theodore C et al.


Phase III trial of vinflunine plus best supportive care
compared with best supportive care alone after a platinumcontaining regimen in patients with advanced transitional
cell carcinoma of the urothelial tract. J Clin Oncol
2009;27(27):4454-61.

40. Fehrman-Ekholm I, Skeppholm L. Renal function in the


elderly (>70 years old) measured by means of iohexol
clearance, serum creatinine, serum urea and estimated
clearance. Scand J Urol Nephrol 2004;38(1):73-7.
41. Dash A, Galsky MD, Vickers AJ et al. Impact of renal
impairment on eligibility for adjuvant cisplatin-based
chemotherapy in patients with urothelial carcinoma of the
bladder. Cancer 2006;107(3):506-13.

55. Bellmunt J, Chow WH, Fougeray R et al. Prognostic factors


in patients with advanced transitional cell carcinoma of the
urothelial tract experiencing treatment failure with platinumcontaining regimens. J Clin Oncol 2010;28(11):1850-5.

406

56. Dreicer R, Gustin DM, See WA, Williams RD. Paclitaxel


in advanced urothelial carcinoma: its role in patients
with renal insufficiency and as salvage therapy. J Urol
1996;156(5):1606-8.

prolonged treatment [German Association of Urological


Oncology (AUO) trial AB 20/99]. Ann Oncol 2010;2 Aug
2010 (Epub ahead of print).
73. Vaishampayan UN, Faulkner JR, Small EJ et al. Phase
II trial of carboplatin and paclitaxel in cisplatin-pretreated
advanced transitional cell carcinoma: a Southwest
Oncology Group study. Cancer 2005;104(8):1627-32.

57. Beer TM, Goldman B, Nichols CR et al. Southwest


Oncology Group phase II study of irinotecan in patients
with advanced transitional cell carcinoma of the urothelium
that progressed after platinum-based chemotherapy. Clin
Genitourin Cancer 2008;6(1):36-9.

74. Pagliaro LC, Millikan RE, Tu SM et al. Cisplatin, gemcitabine,


and ifosfamide as weekly therapy: a feasibility and phase
II study of salvage treatment for advanced transitional-cell
carcinoma. J Clin Oncol 2002;20(13):2965-70.

58. Dreicer R, Li S., Manola J, et al. Phase II trial of epothilone B


analog BMS-247550 (ixabepilone) in advanced carcinoma
of the urothelium (E3800): a trial of the Eastern Cooperative
Oncology Group. Cancer 2007;110:759-63.

75. Dreicer R, Li H, Cooney MM, Wilding G, Roth BJ, Eastern


Cooperative Oncology Group. Phase 2 trial of pemetrexed
disodium and gemcitabine in advanced urothelial cancer
(E4802): a trial of the Eastern Cooperative Oncology
Group. Cancer 2008;112(12):2671-5.

59. Rosenberg JE, Halabi S, Sanford B.L. et al. Phase II trial


of bortezomib in patients with previously treated advanced
urothelial tract transitional cell carcinoma:CALGB 90207.
Ann Oncol 2008;19(5):946-50.

76. Kattan J, Culine S, Theodore C, Droz JP. Second-line


M-VAC therapy in patients previously treated with the
M-VAC regimen for metastatic urothelial cancer. Ann Oncol
1993;4(9):793-4.

60. Sweeney CJ, Roth BJ, Kabbinavar FF et al. Phase II


study of pemetrexed for second-line treatment of transitional cell cancer of the urothelium. J Clin Oncol
2006;24(21):3451-7.

77. Bellmunt J, Maroto P, Mellado B et al. Phase II study


of sunitinib as first line treatment in patients with
advanced urothelial cancer ineligible for cisplatin-based
chemotherapy. Genitourinary Cancer Symposium . 2008.
Ref Type: Abstract

61. Winquist E, Vokes E, Moore MJ, et al. A phase II


study of oxaliplatin in urothelial cancer. Urol Oncol
2005;23(3):150-4.
62. Witte RS, Elson P, Bono B et al. Eastern Cooperative
Group phase II trial of ifosfamide in the treatment of
previously treated advanced urothelial carcinoma. J Clin
Oncol 1997;15 (2):589-93.

78. Gallagher DJ, MIlowsky MI, Gerst SR et al. Phase II study


of sunitinib in patients with metastatic urothelial cancer. J
Clin Oncol 2010;28(8):1373-9.

63. Wuelfing C, Machiels J, Richel D et al. A single arm,


multicenter, open label, phase II study of lapatinib as secondline treatment of pts with locally advanced or metastatic
transitional cell carcinoma. Cancer 2009;115(13):2881-90.

79. Necchi A, Nicolai N, Guglielmi A et al. Phase II Study


of Pazopanib Monotherapy for Patients with relapsed/
refractory Urothelial Cancer (INT70/09, NCT01031875).
Ann Oncol 21(8). 2010. Ref Type: Abstract

64. Galsky MD, Mironov S, Iasonos A, Scattergood J, Boyle


MG, Bajorin DF. Phase II trial of pemetrexed as second-line
therapy in patients with metastatic urothelial carcinoma.
Invest New Drugs 2007;25(3):265-70.

80. Dreicer R, Li H, Stein M et al. Phase II trial of sorafenib


in patients with advanced urothelium cancer: A trial
of the Eastern Cooperative Oncology Group. Cancer
2009;115(18):4090-5.

65. McCaffrey JA, Hilton S, Mazumdar M. Phase II trial of


docetaxel in patients with advanced or metastatic transitional
cell carcinoma. J Clin Oncol 1997;15 (5):1853-7.

81. Sridhar SS, Winquist E, Eisen A et al. A phase II study of


first-line sorafenib (Bay 43-9006) in advanced or metastatic
urothelial cancer. A trial of the PMH Phase II Consortium.
Genitourinary Cancer Symposium . 2008. Ref Type:
Abstract

66. Albers P, Siener R, Perabo FG et al. Gemcitabine


monotherapy as 2nd-line treatment in cisplatin refractory
transitional cell carcinoma - prognostic factors for response
and improvement of quality of life. Onkologie 2002;25:4752.

82. Hahn NM, Stadler WM, Zon RT et al. A multicenter phase


II study of cisplatin (C), gemcitabine (G), and bevacizumab
(B) as first-line therapy for metastatic urothelial carcinoma
(UC): Hoosier Oncology Group GU04-75. Proc Annu Meet
Am Soc Clin Oncol 27(15S,I), 239s. 2009. Ref Type:
Abstract

67. Lorusso V, Pollera CF, Antimi M et al. A phase II study of


gemcitabine in patients with transitional cell carcinoma of
the urinary tract previously treated with platinum. Eur J
Cancer 1998;34 (8):1208-12.

83. www.clinicaltrials.gov. 6-2-2010. Ref Type: Internet Communication

68. Vaughn DJ, Broome CM, Hussain M, Gutheil JC, Markowitz


AB. Phase II trial of weekly paclitaxel in patients with
previously treated advanced urothelial cancer. J Clin Oncol
2002;20(4):937-40.

84. Twardowski P, Stadler WM, Frankel P et al. Phase II study


of Aflibercept (VEGF-Trap) in patients with recurrent or
metastatic urothelial cancer, a California Cancer Consortium
Trial. Urology 2010;76(4):923-6.

69. Witte RS, Manola J, Burch PA, Kuzel T, Weinshel EL,


Loehrer PJS. Topotecan in previously treated advanced
urothelial carcinoma: an ECOG phase II trial. Invest New
Drugs 1998;16(2):191-5.

85. Mita AC, Takimoto CH, Mita M et al. Phase 1 study of AMG
386, a selective angiopoietin 1/2-neutralizing peptibody, in
combination with chemotherapy in adults with advanced
solid tumors. Clin Cancer Res 2010;16(11):3044-5.

70. Sridhar SS, Canil CM, Mukherjee SD et al. A phase II study


of single-agent nab-paclitaxel as second-line therapy in
patients with metastatic urothelial carcinoma. Genitourinary
Cancer Symposium . 2010. Ref Type: Abstract

86. Philips GK, Halabi S, Sanford B.L., Bajorin D, Small EJ,


Cancer and Leukemia Group B. A phase II trial of cisplatin,
fixed dose-rate gemcitabine and gefitinib for advanced
urothelial tract carcinoma: results of the Cancer and
Leukaemia Group B 90102. BJU Int. 2008;101(1):20-5.

71. Paz-Ares L, Ciruelos E, Garcia-Carbonero R, Castellano D,


Lopez-Martin A, Cortes-Funes H. Pemetrexed in bladder,
head and neck, and cervical cancers. Sem Oncol 2002;29(6
Suppl 18):69-75.

87. Philips G, Halabi S, Sanford B.L., Bajorin D, Small EJ,


Cancer and Leukemia Group B. A Phase II trial of cisplatin
(C), gemcitabine (G) and gefitinib for advanced urothelial
carcinoma: Results of Cancer and Luekemia Group B
(CALGB) 90102. Ann Oncol 2009;10(6):1074-9.

72. Albers P, Park SI, Niegisch G et al. Randomized phase III


trial of 2nd line gemcitabine and paclitaxel chemotherapy in
patients with advanced bladder cancer: short-term versus

407

88. Hussain MH, MacVicar GR, Petrylak DP et al. Trastuzumab,


paclitaxel, carboplatin, and gemcitabine in advanced human
epidermal growth factor receptor-2/neu-positive urothelial
carcinoma: results of a multicenter phase II National
Cancer Institute trial. J Clin Oncol 2007;25(16):2118-24.

105. Pectasides D,Aravantinos G, Kalofonos H et al. Combination


chemotherapy with gemcitabine and ifosfamide as secondline treatment in metastatic urothelial cancer. A phase
II trial conducted by the Hellenic Cooperative Oncology
Group. Ann Oncol 2001;12(10):1417-22.

89. Rochester MA, Patel.N., Turney BW et al. The type 1


insulin-like growth factor receptor is over-expressed in
bladder cancer. BJU Int. 2007;100(6):1396-401.

106. Petrylak DP, Tangen CM, Van Veldhuizen PJJr et al. Results
of the Southwest Oncology Group phase II evaluation
(study S0031) of ZD1839 for advanced transitional cell
carcinoma of the urothelium. BJU Int. 2010;105(3):31721.

90. Shen SS, Smith CL, Hsieh JT, et al. Expression of


estrogen receptors-alpha and -beta in bladder cancer
cell lines and human bladder tumor tissue. Cancer
2006;106(12):2610-6.

107. Olive KP, Jacobetz MA, Davidson CJ et al. Inhibition of


Hedgehog signaling enhances delivery of chemotherapy
in a mouse model of pancreatic cancer. Science
2009;324(5933):1457-61.

91. Sonpavde G, Okuno N, Weiss H, et al. Efficacy of selective


estrogen receptor modulators in nude mice bearing human
transitional cell carcinoma. Urol 2007;69(6):1221-6.
92. Liakou CI, Kamat A, Tang DN et al. CTLA-4 blockade
increases IFNgamma-producing CD4+ICOShi cells to shift
the ratio of effector to regulatory T cells in cancer patients.
Proc Natl Acad Sci USA 2008;105(39):14987-92.
93. Chiang GJ, Billmeyer BR, Canes D et al. The src-family
kinase inhibitor PP2 suppresses the in vitro invasive
phenotype of bladder carcinoma cells via modulation of Akt.
BJU Int. 2005;96(3):416-22.
94. Levitt JM, Yamashita H, Jian W, Lerner SP, Sonpavde G.
Dasatinib is preclinically active against Src-overexpressing
human transitional cell carcinoma of the urothelium with
activated Src signaling. Mol Cancer Ther 2010;9(5):1128-35.
95. Bellmunt J, Paz-Ares L, Cuello M et al. Gene expression of
ERCC1 as a novel prognostic marker in advanced bladder
cancer patients receiving cisplatin-based chemotherapy.
Ann Oncol 2007;18(3):522-8.
96. Font A, Taron M, Gago JL et al. BRCA1 mRNA expression
and outcome to neoadjuvant cisplatin-based chemotherapy
in bladder cancer. Ann Oncol 2010;Jul 5 [Epub ahead of
print].
97. Takata R, Katagiri T, Kanehira M et al. Validation study
of the prediction system for clinical response of M-VAC
neoadjuvant chemotherapy. Cancer Sci 2007;98(1):113-7.
98. Als AB, Dyrskjot L, von der Maase H et al. Emmprin and
survivin predict response and survival following cisplatincontaining chemotherapy in patients with advanced bladder
cancer. Clin Cancer Res 2007;13(15 Pt1):4407-14.
99. Williams PD, Cheon S, Havaleshko DM et al. Concordant
gene expression signatures predict clinical outcomes of
cancer patients undergoing systemic therapy. Cancer Res.
2009;69(21):8302-9.
100. Takata R, Katagiri T, Kanehira M et al. Predicting response
to methotrexate, vinblastine, doxorubicin, and cisplatin
neoadjuvant chemotherapy for bladder cancers through
genome-wide gene expression profiling. Clin Cancer Res
2005;11:2625-36.
101. Hoffmann AC, Wild P, Leicht C, Bertz S, Danenberg PV,
et al. MDR1 and ERCC1 expression predict outcome of
patients with locally advanced bladder cancer receiving
adjuvant chemotherapy. Neoplasia 2010;12:628-36.
102. Lee JK, Havaleshko DM, Cho H et al. A strategy for
predicting the chemosensitivity of human cancers and its
application to drug discovery. Proc Natl Acad Sci USA
2007;104:13086-91.
103. Smith SC, Baras AS, Lee JK, Theodorescu D. The COXEN
principle: translating signatures of in vitro chemosensitivity
into tools for clinical outcome prediction and drug discovery
in cancer. Cancer Res. 2010;70:1753-8.
104. Smith SC, Theodorescu D. Learning therapeutic lessons
from metastasis suppressor proteins. Nat Rev Cancer
2009;9:253-64.

408

Committee 10

Non-urothelial Cancer of the


Urinary Bladder
Chair:
Bruce R Kava,
Members:
Hassan Abol-Enein,
Jack Baniel,
Adrienne Carmack,
Alexandra Coquhoun,
William Turner

Address all Correspondence


Bruce R Kava,
Associate Professor of Urology
Interim Chair, University of Miami Miller School of Medicine
Chief of Urology, Department of Veterans Affairs Medical
Center, Miami
PO Box M016960 (M814)
Miami, Florida, USA

409

Table of Contents
I. Non-urothelial Bladder
Cancer

IV. Adenocarcinoma
1. Epidemiology
2. Primary Adenocarcinoma

II. SCC Not Associated with


Schistosomiasis

3. Urachal Adenocarcinoma
4. Adenocarcinoma in Bladder
Exstrophy

1. Epidemiology

5. Secondary (Metastatic) Bladder


Adenocarcinoma

2. A
 ssociation with Spinal Cord
Injury (SCI)
3. Smoking

V. Small Cell Carcinoma

4. Etiology

1. EPIDEMIOLOGY

5. Clinical and Pathologic Features

2. Clinical Features

6. Treatment

3. Pathologic Features
4. Treatment

III. Squamous Cell Carcinoma


in Association with
Schistosomiasis (Bilharziasis)

VI. Other Bladder Tumors


1. Bladder Sarcoma

1. Epidemiology

2. C
 arcinosacroma and arcinomatoid
Tumors

2. Etiology

3. Paraganglioma and Pheochromocytoma

3. Clinical and Pathologic Features


4. Treatment

4. Bladder Pseudotumors
5. Melanoma
6. Lymphoma

410

Non-urothelial Cancer of the Urinary


Bladder
Bruce R Kava,
Hassan Abol-Enein, Jack Baniel, Adrienne Carmack,
Alexandra Coquhoun, William Turner

and treatment recommendations are different. Each


will be discussed separately in the sections below.

I. Non-urothelial Bladder
Cancer

II. SCC Not Associated with


Schistosomiasis

Non-urothelial bladder cancer refers to a


heterogeneous group of malignancies, which are
considerably less prevalent than urothelial carcinoma.
Data from several contemporary epidemiologic
studies highlighting the relative percentage of the
various histological subtypes of bladder tumors by
geographic region is depicted in Figure 1 [1,2].

1. Epidemiology
Often referred to as non-Bilharzial SCC, this
represents the most common non-urothelial bladder
malignancy, accounting for 2-5% of cases in most
contemporary cystectomy series [4-8]. These tumors
are most often diagnosed during the seventh decade
of life.

The relative scarcity of these tumors has precluded


the development of large-scale prospective clinical
trials. In a recent review of the National Cancer
Institutes Surveillance Epidemiology and End
Results (SEER) database, [3] slightly more than
3700 non-urothelial tumors were identified over a
34-year period from over 125,000 cases of bladder
cancer. Of these non-urothelial tumors, only small
cell carcinoma appears to have demonstrated a
slight increase in incidence during that time period
(Figure 2).

The incidence of SCC demonstrates less of a male


predominance than does urothelial carcinoma.
Compiling data from 915 patients in 10 series
of patients with SCC, Johansson and Cohen [9]
reported that the ratio of males to females was 1.4:1.
Similar to urothelial carcinoma, however, women
are more likely than men to present with advanced
disease [10]. Data from the Netherlands Cancer
Registry [11] corroborates this higher incidence of
T3 and T4 tumors noted in women relative to men
(21.7% vs. 14.5% in T3, and 14.5% vs. 8.4% in T4).

As a result of these data, the level of evidence


provided regarding screening recommendations,
staging, and treatment strategies of non-urothelial
malignancies is limited. Nevertheless, the clinician
must be aware of these tumors, which often
warrant a different clinical approach than urothelial
carcinoma. Each of the tumors will be addressed
individually, and evidence-based recommendations
will be provided when applicable.

Data from Surveillance, Epidemiology, and End


Results (SEER) program conducted between 1973
and 1997 [12] shows that there is a large racial
disparity in the incidence of SCC. With an annual
incidence of 1.2 per 100,000 person-years, African
Americans were twice as likely to develop these
tumors as Caucasians, who had an annual incidence
of 0.6 per 100,000 person-years.

Squamous Cell Carcinoma (SCC)


SCC may occur de novo, or in individuals who
have been infected with the parasite Schistosoma
hematobium. It is important to recognize the
distinction between these two populations of
patients, because the epidemiology, natural history,

2. A
 ssociation with Spinal Cord
Injury (SCI)
In the United States, patients with SCI represent
the largest cohort of patients affected by SCC. This

411

Figure 1. Percentage and histologic subtype of non-urothelial bladder cancers seen in various tumor
registries worldwide (1,2).

Figure 2. Age-adjusted incidence rates of nonurothelial bladder cancer by histological subtypes in the USA,
nine SEER registries, 1973-2007 (3).

412

condition is believed to arise as a result of chronic


urinary tract inflammation, a factor which may
contribute to the higher likelihood of SCC in patients
with chronic cystitis, persistent calculi, and bladder
diverticula [13].

tobacco abuse. Any history of hematuria should be


evaluated.

3. Smoking
The relationship between SCC and cigarette
smoking is not clear; Johansson and Cohen [9]
found a higher incidence of SCC in smokers. SEER
data support a direct correlation between quantity
of cigarettes smoked and relative risk of developing
SCC [23]. Indirect evidence from the Swedish
Cancer Registry [24], however, does not support an
association between SCC and smoking. A review
of this data, which plotted trends in bladder cancer
in Sweden between 1960 and 1993, revealed that,
despite a rising incidence of urothelial carcinoma in
Swedish women (which correlated with an increased
prevalence of smoking), the incidence of SCC
remained relatively constant.

Historically, the incidence of bladder cancer in


patients with SCI was believed to be as high as 2.310%, with most cases representing SCC [14-16].
SCI patients with indwelling catheters were found to
have the highest risk for the development of SCC;
up to 10% of these individuals developed SCC at
10 years [14]. Interestingly, studies determined
that patients performing clean intermittent selfcatheterization were significantly less likely to
develop SCC than those SCI patients with indwelling
catheters [17, 18].
Recently, these concepts have been scrutinized by
data emerging from several contemporary series. A
United States Department of Veterans Affairs review
of admission data from 33,560 SCI patients identified
only 130 patients with bladder cancer, for an overall
incidence of 0.39% [19]. Some 42 patient records
were available for review, including 23 (55%) with
urothelial carcinoma, 14 (33%) with SCC, and 4
(10%) with adenocarcinoma. It is noteworthy that in
26 patients with indwelling catheters, the incidences
of SCC and urothelial carcinoma were equal,
implicating chronic inflammation and tobacco abuse
as competing risk factors.

4. Etiology
Because urinary tract infections are more common
in women, a relationship between squamous
metaplasia, leukoplakia, and the development of
SCC was proposed by Connery [25] and by Holly
and Mellinger [26]. In a series of 20 patients with
long-standing leukoplakia, OFlynn and Mullaney
[26] observed the development of 5 cases of SCC.
Although squamous metaplasia is not considered
a premalignant lesion, it is often found in areas
adjacent to SCC [13]. Many studies, however,
have confirmed that there is a high prevalence of
squamous metaplasia in the general population, and
that it is not necessarily implicated in the development
of SCC [28].

In another study of 2900 SCI patients from several


centers in the southwest United States, bladder
cancer was detected in only 8 (0.32%) patients, none
of whom had an indwelling catheter [20]. Another
study evaluated 15 SCI patients with SCC, who were
followed over a 24-year period from a VA Center in
Northern California [21]. Interestingly, the majority of
these patients had never had an indwelling catheter.
Finally, in the largest study to date, Pannek et al. [22]
evaluated 43,561 SCI patients from multiple urologic
centers in Eastern Europe. In all, 48 patients with
bladder cancer were identified, for an overall
incidence of 0.11%. Only 7% of patients in this series
had indwelling catheters, and the prevalence of SCC
was only 0.02%.

A few case reports have documented associations


between cyclophosphamide, bacille CalmetteGurin (BCG), human papillomavirus, and SCC
of the urinary bladder [29, 30]. Other studies have
revealed possible genetic and chromosomal changes
that may be associated with SCC. Abnormalities of
chromosomes 3, 8, 10, 13, and 17 have been detected
in SCC [31]. Studies on uroplakin II gene expression
found a significant difference in expression between
urothelial carcinoma and SCC, with expression
being greater in SCC. Uroplakins are the major
differentiation products of the urothelium that control
the various pathways of urothelial differentiation [8].

Based upon these epidemiologic data, guidelines


regarding the necessity, frequency, and the diagnostic
testing needed for bladder cancer screening in the
SCI patient population cannot be determined. Initial
reports, in which a high percentage of patients
with spinal cord injury developed SCC are flawed,
primarily related to the retrospective manner in which
data were obtained, small patient cohorts, and less
rigorous statistical evaluation than would be expected
for a more contemporary evaluation. Although the
incidence of SCC in a contemporary SCI population
appears to be less than 1%, it is recommended that
these patients should be monitored, particularly if
they have indwelling catheters or have a history of

5. Clinical and Pathologic Features


The presenting symptoms in patients with nonschistosoma-related SCC are not distinguishable
from urothelial carcinoma. Hematuria is the main
clinical feature in 63-100% of patients. Irritative
bladder symptoms are seen in two-thirds of patients.
Weight loss, back or pelvic pain, and frank obstructive
symptoms are less common and suggest advanced
disease [4-6,32]. A urinary tract infection is present
in 30-93% of patients at the time of diagnosis
[4-6,33,34], and symptoms have often been present

413

b) Radical Surgery

for a prolonged period of time before the diagnosis


is made [4,4,8,33].

While most of the surgical series are subject to


selection bias, radical cystectomy seems to offer
some advantages in patients with SCC. Richie et
al [37] reported a 5-year survival rate of 48% in 25
patients treated with radical cystectomy. While the
investigators stated that this compared favorably with
results in patients with urothelial carcinoma, they did
not include in the analysis 3 patients (9%) who died
during the perioperative period and an additional 5
patients in whom insufficient pathologic or follow-up
data were obtained. Given the small cohort of study
patients, and adjusting for these cases, the 5-year
survival rate is significantly lower than was reported.
Tumor stage was identified as the most important
predictor of outcome.

Relatively few series of patients from Europe and the


Americas have addressed pure SCC. Most of these
are retrospective observational series that are more
than 10 years old, and use older staging and grading
systems. Despite this, we know that the majority
of patients present with a bulky, solitary tumor that
extensively involves the bladder wall. These tumors
are sessile lesions, often with ulceration and areas of
squamous metaplasia adjacent to the primary tumor.
A predilection for the trigone has been noted, but
SCC can arise anywhere within the bladder. Tumors
may occupy a bladder diverticulum and have been
described in association with bladder calculi [35].
SCC is often locally advanced at the time of
diagnosis. Debbagh et al [36] reported that 10 of 14
(71%) patients in their series had a palpable tumor on
rectal examination, and 11 (79%) had upper urinary
tract obstruction. Pretreatment imaging studies may
demonstrate hydronephrosis in 33-59% of cases
[5,6,8]. Of 114 patients with SCC of the bladder in
one series, 92% had T2 to T4 disease at the time
of diagnosis, and most tumors were high grade [5].
As with urothelial carcinoma, clinical understaging is
seen in as many as 73% of patients [6,37].

Similarly Serretta et al [8] reported on 19 patients


with pure SCC of the bladder undergoing radical
cystectomy. With a mean follow-up of 52 months,
63% had died of local recurrence, with only one
patient developing distant metastases.
Emerging data on the surgical management and
outcome of treated SCC continues to be derived from
relatively small single center series. In a recent series
of 45 patients [38], in which 67% presented with T3
tumors, 37% of patients were alive without disease
and 29% had died of disease at a median followup of 15 months following cystectomy. Similarly,
Kassouf et al [39] reported a 50% recurrence in 10
of 20 patients who had undergone cystectomy at
only 5 months. The majority of patients who died of
SCC had isolated pelvic recurrence in the absence
of disseminated metastases, emphasizing the
importance of local surgical control.

6. Treatment
Pure SCC of the bladder has a poor prognosis,
with most patients succumbing within 1 to 3 years
of diagnosis.
Failure to provide loco-regional
control is the hallmark of the difficulty in managing
these patients. In a series of 120 patients from the
Royal Marsden Hospital, the overall 5-year survival
rate was 16%, with only 8% of patients developing
metastatic disease [33].

c) P
 rognostic Comparison with Urothelial
Carcinoma

a) Radiation
Irrespective of whether radiation is used as
neoadjuvant or as primary treatment, results have
been uniformly poor [6,32]. In a report by Quilty
and Duncan [34], 51 patients were treated with
radical radiotherapy, delivered with a 3-field beamdirected technique, covering the entire bladder, to a
prescribed dose of 55 Gy over 4 weeks. Patients
were treated prone, immediately after emptying
their bladders. Only 4 patients in this series had T2
cancer, with a median survival of 14.3 months, and
a 3-year survival of 26.8%. Of 48 patients treated
with radiation therapy by Rundle et al [6], the 5-year
survival for patients with T2 and T3 disease was
16.7% and 4.8%, respectively. No patient with T4
disease was alive beyond 11 months. Similarly, in
a series of 17 patients with T2 and T3 tumors who
were treated with radiation therapy alone, Johnson
et al [5] reported a 20% 5-year survival rate, which
was not statistically different than the 34.6% 5-year
survival rate seen in 7 patients in which preoperative
radiation was followed by radical cystectomy.

Several large contemporary cystectomy series in the


literature have compared outcomes in patients with
urothelial carcinoma with those in patients with SCC.
In a large series from Japan [40], evaluating 1042
patients treated with urothelial carcinoma and 89
patients with SCC, there was no significant difference
observed in 5-year post-cystectomy survival (68.0%
urothelial carcinoma vs. 60.8% SCC). In a review of
SEER data between 1988 and 2003 [41], patients
with SCC had worse outcomes than those with
urothelial carcinoma except for those patients with
localized cancers who were treated by cystectomy.

d) Neoadjuvant Radiation
Several studies have retrospectively evaluated the
efficacy of pre-cystectomy irradiation. Data is limited
because of the small number of patients studied, and
the lack of randomized clinical trials. Johnson et al
[5] integrated preoperative radiation therapy followed

414

Recommendations

by cystectomy and reported a 5-year survival rate of


34%. Swanson et al [42] reported similar results with
the same approach. Patients with T2 disease showed
the highest survival figures. Given the relative rarity
of these tumors, the advanced stage of presentation,
and the overwhelmingly poor overall survival with
these tumors, it would be difficult to devise clinical
trials of sufficient power to demonstrate a difference
in the outcome between the two groups.

In summary, non-schistosoma-related SCC


is an uncommon form of bladder cancer that
usually presents at an advanced stage, generally
recurs loco-regionally, and has an extremely
poor prognosis. Death is most often related to
loco-regional failure, and not to disseminated
metastasis.

e) Impact of Urinary Diversion

The current literature supports cystectomy as the


treatment of choice, when possible (grade B) [53].

The impact of the type of urinary diversion in patients


with SCC is the subject of some discussion in the
literature. In a series of 19 patients undergoing radical
cystectomy and urinary diversion, all 3 patients with
an orthotopic ileal neobladder developed recurrence
at the anastomosis between the neobladder and the
urethra [43]. In another series reported by Stenzl et
al [44], intraoperative frozen section biopsies were
obtained from the bladder neck before orthotopic
reconstruction. No local recurrences occurred in the
5 female patients in this series.

III. Squamous Cell Carcinoma


in Association with
Schistosomiasis (Bilharziasis)
1. Epidemiology
This type of cancer is prevalent where urinary
Schistosoma hematobium is endemic. It is often
referred to as schistosoma-related SCC or bilharzial
SCC. The highest incidence of SCC of the bilharzial
bladder occurs in Egypt [1]. In a report by Ghoneim et
al [54], SCC accounted for 59% of 1026 cystectomy
specimens. A high incidence of SCC is also found
in Iraq, the Jizan region in southern Saudi Arabia,
Yemen, and Sudan. In Africa, the disease has been
reported in the Gold Coast region and in South
Africa. However, the incidence in these countries
is lower because Schistosoma hematobium is
less endemic and less severe [55]. With a mean
age at presentation of 46 years, the mean age of
schistosoma-related SCC patients is 10 to 20 years
younger than that seen with non-schistosomarelated SCC [56]. In areas in which schistosomiasis
is endemic, some 80% of cancer specimens have
shown histologic evidence of bilharzial infestation
[55]. A lag period of approximately 30 years has been
reported between infection with the parasite and
subsequent development of the disease. The maleto-female ratio is 5:1 [57]. This male predominance
is attributed to sustained contact with infected water
supplies that laborers often endure while working in
various outdoor environments (Figure 3).

f) Chemotherapy
The role of neoadjuvant or adjuvant chemotherapy in
the treatment of patients with pure SCC of the bladder
is uncertain. SCC is considerably less responsive to
standard chemotherapy regimens used for urothelial
carcinoma [45, 46]. Neoadjuvant M-VAC has
been tried with no objective response [47]. Newer
combination regimens consisting of gemcitabine,
paclitaxel, and docetaxel, when combined with a
platinum compound, may yield sustained disease
remission in up to 50% of cases; these hold promise
for the future [48].

g) Prevention and Early Detection


Several screening protocols have been advocated
in an attempt to diagnose these tumors earlier,
thereby improving outcomes. Broecker et al [49]
recommended annual cystoscopy and urine cytology
in patients with SCI. Others have suggested routine
random bladder biopsies every 1 to 2 years. Navon
et al [50] advocated urine cytology or random
bladder biopsies in patients with spinal cord injuries
for more than 10 years or in those with recurrent or
chronic urinary tract infection. Celis et al [51] showed
that psoriasin, a calcium-binding protein expressed
by squamous epithelia, is a potential marker for
SCC. Other biomarkers, such as SCC antigen and
bcl-2 and p53 oncoproteins, may have a possible
role in early diagnosis [52]. The role of these new
markers in the early detection and follow-up of
bladder SCC requires further study before they can
be recommended.

Contemporary epidemiological studies have


demonstrated a decline in the incidence of bilharzial
SCC, as public health efforts have been successful
in eradicating schistosomiasis in many areas of the
Nile delta and in rural Egypt [58-62]. Interestingly,
despite the decline in SCC, the incidence of bladder
cancer remains high. This is believed to be a result
of the high prevalence of tobacco use that has now
created a significant rise in the incidence of urothelial
cancer. Figure 4 depicts some of these trends [63].

415

2. Etiology

metaplasia of the bladder epithelium and the


predominance of SCC in patients with bilharziasis.

Evidence from animal models suggests that the


biogenesis of bladder cancer is a multistage process.
It involves initiation by carcinogens, followed by
promotion of tumor growth [64]. Schistosomarelated bladder cancer may be initiated by exposure
to an environmentally or locally-produced chemical
carcinogen that is excreted in urine. This carcinogen
reacts with the mucosal surface of the bladder to
produce irreversible and potentially carcinogenic
changes in the DNA of some urothelial cells. Chronic
bacterial infection, which commonly complicates
urinary schistosomiasis, has been implicated in the
production of nitrosamines, which are well-known
potent carcinogens derived from precursors in the
urine. Additionally, bacterial infections may induce
the secretion of beta-glucuronidase, which has been
shown to split conjugated carcinogens to yield free
carcinogenic products [65,66].

Recent laboratory studies have shown that


overexpression of cyclooxygenase-2 is associated
with increasing pathological stage and grade and
a worse outcome after radical cystectomy [67,68].
Abdulamir et al found that bilharzial and nonbilharzial SCC have different molecular profiles of
tumor development of progression [69].

3. Clinical and Pathologic Features


Patients usually present with symptoms of cystitis,
including painful micturition, frequency, and
hematuria. A diffuse, irregular filling defect is usually
detected on cystogram. Computed tomography (CT)
scanning or magnetic resonance imaging (MRI) is
helpful for diagnosis and staging, which relies on
cystoscopy, biopsy, and bimanual examination under
anesthesia [70].

The possibility that carcinogenic products may be


of parasitic origin is not supported by several recent
investigations [55]. However, local mechanical
irritation by Schistosoma eggs appears to be an
important promoting factor [66]. Vitamin A deficiency
may explain the high frequency of squamous

Urine cytology may provide valuable diagnostic


information in patients with schistosoma-related SCC
[71]. In fact, cytokeratin shedding in urine has been
used as a biologic marker for the early detection of
SCC [72].

Figure 3. Gender disparities in stage of bladder SCC at diagnosis. Data derived from the Netherlands Cancer
Registry (11).

From Felix, et al : Cancer Causes Control (2008):19,421-9


Figure 4. Changing Patterns of Egyptian Bladder CancerData abstracted from NCI- Cairo (4).

416

Most patients present for treatment at an advanced


stage, and 25% of cases are inoperable when
first seen [56]. This is because of the nonspecific
symptoms of schistosoma-related SCC, which often
mimic simple bilharzial cystitis.

circumvent the physiologic and social inconvenience


of urinary diversion and the possible loss of sexual
potency. Once again, local resection is feasible only
in select cases. Solitary tumors must not involve
the trigone, and their size must allow resection with
adequate oncologic margins; the rest of the bladder
mucosa must be free of any associated precancerous
lesions. These strict criteria are met in only a minority
of cases. El-Hammady et al. [79] found partial
cystectomy feasible in only 19 of 190 (10%) patients.
Additionally, augmentation cystoplasty was required
in 5 of these patients, in order to increase residual
bladder capacity. The 5-year survival rate was 26.5%,
and patients with low-grade tumor had approximately
twice the survival as those with high-grade disease.
Less favorable results with partial cystectomy were
reported by Omar et al [80] in a series of 22 cases.
All patients except 1 developed tumor recurrence
within 2 years. This different outcome may be related
to the wide variability of selection criteria.

Similar to urothelial bladder cancer, clinical


understaging occurs in a high percentage of cases
[70] and schistosoma-related SCC is often locally
advanced at the time of diagnosis. In a study of 608
patients with schistosoma-related SCC, pT1 disease
was found in 2.6%, pT2 in 10.5%, pT3 in 80.0%,
and pT4 in 6.9% [54]. Lymph node metastases were
only present in 18.7% of cystectomy specimens.
Interestingly, the vast majority of tumors were low
and medium grade (49.7% and 33.2%, respectively),
a factor that may account for the low incidence of
lymph node positivity [73].
Grossly, tumors are generally of the nodular,
fungating type and are located in the dome or
posterior or lateral walls of the bladder. Five gross
types have been identified: nodular (60%), ulcerative
(23%), verrucous (7%), papillary (7%), and diffuse
(3%) [55]. A variety of atypical changes in the
bladder mucosa, including metaplasia, dysplasia,
and, rarely, carcinoma in situ, may be associated
with the disease [74].

d) Radical Cystectomy
Radical cystectomy and urinary diversion provides
a logical treatment approach for patients with
resectable tumors [81, 82]. In an early series of
138 cases, Ghoneim et al [82] reported a high
perioperative mortality rate of 13.7%. This was
primarily due to peritonitis, intestinal obstruction, and
liver failure. Cardiopulmonary complications were
uncommon among this relatively young group of
patients. In this older series, overall 5-year survival
was 32.6%. Patients with pT1 and pT2 disease
had a 43% overall 5-year survival, which compared
favorably with 30% in patients with pT3 and pT4
tumors. Low-grade tumors were associated with a
46% overall 5-year survival, and high-grade disease
was associated with a 5-year survival of 21%.
Lymph node metastases were associated with a
poor outcome, and reduced the 5-year survival rate
to 20% [81].

4. Treatment
a) Radiation
The growth characteristics of carcinoma of the
bilharzial bladder have been studied with the goal
of evaluating its potential response to radiation
therapy [74]. Two growth features were noted: [1]
high cell mitotic rate with a potential doubling time of
6 days, and [2] an extensive cell loss factor. Tumors
with such growth characteristics were expected to
exhibit a better response to radiation therapy [75].
Nevertheless, early experiences with external beam
radiation therapy for definitive control of these tumors
were disappointing [76]. Factors that interfered with
the efficiency of radiation treatment in these cases
included coexisting Schistosoma-related urologic
lesions, which interfere with local tissue tolerance,
and considerable tumor bulk, which reduces local
tumor control. Furthermore, the presence of radioresistant hypoxic tumor cells is suspected in light of
the capillary vascular pattern of this cancer [77].

A more contemporary report evaluating 1026


cystectomy patients from an area in which
schistosomiasis is endemic [54] found that 59% of
tumors were SCC. Bilharzial ova were identifiable
in 88% of specimens. Interestingly, extravesical
extension was not significantly different among
patients with SCC or urothelial carcinoma (13.5%
and 14.9%, respectively). Overall the 5-year survival
rate with SCC was 50.3%. Only tumor stage, grade,
and lymph node involvement had independent
significant effects on survival.

b) Endoscopic Resection
In view of the usual high tumor volume and its
advanced stage, transurethral resection is not
definitive treatment in the management of this
disease. Endoscopic resection should be used only
for obtaining a biopsy specimen for histopathologic
diagnosis.

e) Neoadjuvant Radiation
As discussed above, schistosoma-related SCC
is often understaged. Unlike urothelial cancer,
recurrences in patients undergoing radical surgery
for schistosoma-related SCC often occur locally.
The aim of preoperative radiation is to eradicate
microextension of tumor into the perivesical tissues
and lymphatics.

c) Segmental Resection (Partial Cystectomy)


Segmental resection is an attractive alternative to

417

Awaad et al [83] compared the results of cystectomy


after preoperative administration of 40 Gy radiation
therapy with outcomes in a control group treated
with cystectomy only. Reported 2-year survival
rates were significantly improved in the irradiated
group. Ghoneim et al [84] compared the results
of cystectomy after preoperative irradiation using
20 Gy with those of cystectomy alone. Ninety-two
patients were divided into 2 groups, and followed
for 60 months. Patients who received preoperative
radiation showed a trend towards longer survival,
although this was not statistically significant. In lowstage tumors, regardless of grade, survival was not
influenced by preoperative irradiation, as it was in
high-stage tumors.

Secondary prevention includes early detection


with urine cytology and selective screening of the
population at risk. The yield of a single screening
study done in a rural area in Egypt was 2 per 1000
individuals [71]. Such a detection rate would not
justify regular screening.

Recommendations
In summary, bilharzial SCC is the most common
form of bladder cancer in endemic areas. It most
often presents at an advanced stage but with low
grade cells. Cystectomy is the standard treatment,
but long-term survival remains disappointing
(grade B). Limited evidence (Grade B) supports
a potential role of neoadjuvant chemotherapy
and radiation therapy, but it is not yet sufficient to
facilitate a recommendation.

The presence of a large proportion of hypoxic


cells within bulky tumors could explain the modest
improvement that was observed after this regimen.
To enhance the therapeutic value of irradiation,
misonidazole, a hypoxic cell sensitizer, was
given before the radiation regimen was delivered
[84,85]. A prospective trial was divided into 3 arms:
cystectomy only, 20 Gy of preoperative radiation
followed by cystectomy, and preoperative radiation
followed by cystectomy with misonidazole added
as a radio-sensitizer. The addition of misonidazole
did not provide any additional survival advantage to
patients who were given preoperative radiotherapy
(level 1 evidence).

IV. Adenocarcinoma
1. Epidemiology
Adenocarcinoma of the bladder is the third most
common histologic type of bladder carcinoma. In
regions in which schistosomiasis is not endemic, it
comprises 0.5- 2.0% of all bladder tumors [89-91].
The incidence in areas where schistosomiasis is
endemic is higher, at 5.2-11.4% [54, 92,93].

f) Chemotherapy

Annual incidence rates for adenocarcinoma are


correspondingly very low. One study from the United
States using SEER data identified just 32 patients
(0.7%) with adenocarcinoma from 4045 patients with
newly diagnosed bladder cancer over a 1- year period
(1977-1978) [94]. Similar to urothelial carcinoma,
adenocarcinoma shows a male predominance. In a
total of 11 series comprising 247 patients, the gender
ratio between males and females was 2.7:1 [9].

Several agents have been evaluated by Gad-elMawla et al [86] in order to treat patients who are
initially not believed to be surgically resectable.
All trials were phase 2 studies in which a single
agent was used, and the optimal results were
obtained with epirubicin. Neoadjuvant and adjuvant
epirubicin chemotherapy were used in a prospective,
randomized study involving 71 patients with invasive
cancer in bilharzial bladders. Two-thirds of treated
patients had SCC. Disease-free survival rates were
73.5% and 37.9%, favoring the chemotherapy group
[87]. Additional long-term follow-up results have not
been published.

The relative rarity of this tumor makes determination


of risk factors difficult [94]. Nevertheless, there appears to be a higher likelihood of developing adenocarcinoma in patients with a history of schistosomiasis, endometriosis, bladder augmentation, and other
irritative conditions of the urinary bladder [95,96].
Adenocarcinoma also carries the unique distinction
of being the most common tumor arising in the bladder of patients with exstrophy. Such patients carry a
4% lifetime risk of developing adenocarcinoma [97].

Bilharzial SCC does not appear to be responsive


to chemotherapy that has traditionally been used
for urothelial carcinoma. In a recent multicenter
study that consisted of 120 patients treated with
neoadjuvant cisplatin and gemcitabine [88], patients
with SCC had no survival benefit over those who
underwent cystectomy alone.

Adenocarcinoma of the urinary bladder is classified as either: primary adenocarcinoma, urachal


adenocarcinoma, or secondary (metastatic) adenocarcinoma. Secondary tumors are indicative of local
extension of a primary colonic, prostatic, or ovarian
cancer [98]. Primary, urachal, exstrophy-associated,
and metastatic adenocarcinoma will be discussed
separately.

g) Prevention and Early Detection


Bilharzial SCC is a preventable malignant disease.
Primary prevention entails control of bilharziasis
through snail control (the intermediate host of the
parasite) and mass treatment of the rural population
with oral antibilharzial drugs such as praziquantel.

418

2. Primary Adenocarcinoma

poor prognosis regardless of the treatment modality


utilized. Five-year survival varies from 11-57%; the
small number of patients in each series precludes
individual comparison of treatment undertaken.
Consistently reported independent prognostic
factors for survival include age, stage, grade, and
lymph node involvement [92,93,96,99,100,105].

a) Epidemiology
El-Bolkainy and associates reported an incidence of
8.1% in a series of 229 cases of bladder cancer [56].
Between 1970 and 1995, 1870 cystectomies were
carried out in the Urology and Nephrology Center of
Mansoura. Of these, 185 cases (9.9%) proved to be
primary non-urachal adenocarcinoma of the bladder
on histopathologic examination [55]. Similarly,
between 1994 and 2003, 3659 cystectomies were
performed at the National Cancer Institute, Cairo and
Minia Oncology Centre, Egypt. Of these, 192 (5.2%)
were for primary non-urachal adenocarcinoma of the
bladder [92]. In addition to these 2 large schistosomaassociated series of patients, two recent series from
the United States have been evaluated. Using the
SEER database, Wright et al identified 1374 cases
of primary adenocarcinoma of the bladder recorded
between 1973-2002 [99]. Lughezzani et al [100] also
evaluated SEER data and found a total of 306 (2.5%)
primary adenocarcinoma patients among 12,003
bladder cancer patients undergoing cystectomy
between 1988 and 2006.

c) Pathological features
In order for a diagnosis of primary adenocarcinoma
of the bladder to be made, it must be clearly
distinguished from urothelial carcinoma with areas of
glandular metaplasia. The pathogenesis of primary
non-urachal adenocarcinoma is based on the ability
of the urothelium to undergo metaplastic changes
[108]. Mostofi [109] proposed that the metaplastic
potential of the urothelium has 2 distinct patterns.
Progressive invagination of hyperplastic epithelial
buds into the lamina propria (von Brunns nests) leads
to the formation of cystitis cystica. Subsequently,
metaplasia of the urothelial lining of these cysts,
which become columnar mucin-producing cells,
results in the production of cystitis glandularis, which
is a premalignant lesion. As a result, careful follow
up of patients with cystitis glandularis is necessary
[109]. Chronic bladder irritation and infection are the
predisposing factors for these changes [107,110].
This explains, at least partly, the higher incidence of
these tumors among patients with cystitis resulting
from schistosomiasis.

b) Clinical features
Primary adenocarcinoma of the bladder presents
with hematuria in most patients, which may be
associated with irritative voiding symptoms and,
occasionally, mucus passage in the urine [89, 93,101104]. Cystoscopically, the tumor is usually sessile,
but can be papillary [103]. It can arise anywhere
along the lateral walls, trigone, dome, and anterior
wall of the bladder [93,101,103,104]. Multiple tumors
are present approximately 50% of the time [54,101].
Adenocarcinoma is virtually always invasive with
only 1 series documenting 2 non-invasive tumors
[105]. Interestingly, both patients were alive at 51
and 61 months following TUR alone.

Histologically, adenocarcinoma may be non-mucinproducing or mucin-producing. Most of these tumors


are mucin-producing, but the gross passage of mucus
during micturition remains uncommon [105,110].
This is because most of the mucus is secreted into
the extracellular (interstitial) matrix of the urinary
bladder, rather than into the bladder lumen [93].
Occasionally, mucin is secreted within the lumen
of the acini. In other cases, mucin accumulates
within the intracellular space displacing the nucleus
to a peripheral crescent, giving the cells a signet
ring appearance. It is generally accepted that this
histological subtype denotes a poor prognosis [111114].

Similar to urothelial carcinoma, the in situ form of


adenocarcinoma appears to portend a poor prognosis.
Chan et al [106] reviewed all the case files for a series
of 19 patients who had bladder biopsies performed
at Johns Hopkins Hospital between 1984 and 2000
confirming the presence of adenocarcinoma in situ.
Nearly 80% of these patients also exhibited some
form of concomitant urothelial carcinoma in situ.
Although none of the patients progressed to develop
primary adenocarcinoma of the bladder, 74%
ultimately received treatment for invasive carcinoma
(often for an aggressive histological subtype: 5 cases
of small cell and 4 cases of micropapillary urothelial
carcinoma). Clearly, it is difficult to determine whether
the urothelial or adenocarcinoma component of the
in situ disease determined prognosis, but certainly
the co-existence of both entities mitigated a poor
outcome.

No uniformly accepted grading system for


adenocarcinoma of the bladder exists. On the
basis of histopathologic findings, Anderstrom
et al classified vesical adenocarcinoma into 5
patterns: glandular with columnar, sometimes
entericappearing cells; colloid carcinoma; papillary
adenocarcinoma; signet ring cell carcinoma; and
clear cell carcinoma [96]. Several other histologic
subtypes have been described, including mucinous,
enteric (colonic) type, adenocarcinoma not otherwise
specified (NOS), clear cell, hepatoid, and mixed type
[104]. Unfortunately, no clear data exists regarding
the impact of these different histologic patterns on
survival or prognosis. The exception is signet ring
cell carcinoma, which usually follows a very rapid

Primary adenocarcinoma of the bladder has a very

419

course, resulting in death in the majority of patients


within 6 months [111].

Others have found no correlation between DNA


ploidy and outcome [93,12] whereas Song et al [124]
showed that DNA ploidy significantly correlated with
tumor progression and subsequent poor outcome.
In short, there are no molecular markers predictive
of outcome for primary adenocarcinoma that have
been validated in an independent patient cohort.

d) Treatment
Several treatment strategies have been used in the
management of primary adenocarcinoma of the
bladder. In general, the available series are too
small to permit meaningful comparisons between
differing treatment approaches.

2. Radiation Therapy
Adenocarcinoma is not a radioresponsive disease,
with 5-year survival rates of less than 20% in
patients treated by external beam irradiation alone
[91,106,114]. The addition of preoperative irradiation
to partial cystectomy did not improve the survival in
2 early studies [91,96]. One of the 2 large Egyptian
cystectomy series for adenocarcinoma did assess
the benefit of adjuvant radiation treatment [93].
Patients receiving adjuvant radiation experienced
improved outcome (61% vs. 37% 5-year survival),
but the study was not a randomized trial and patient
selection may have accounted for the survival
differences.

1. Surgical Resection
Surgical strategies include transurethral resection
(TUR), partial cystectomy, and radical cystectomy.
The therapeutic yield after TUR, with or without
adjuvant radiotherapy, is poor (19-31% 5-year
survival) [102,115]. Partial cystectomy for localized
disease in a mobile part of the bladder has been
advocated by several authors, including a series of
15 patients who had a 5-year survival of 54% [96].
Other series however, have shown poorer 5-year
survival rates of 16-37% following partial cystectomy
[91,116].
The reported 5-year disease-free survival in several
cystectomy series range from 0-80% in several
historical series [91,96,115-120]. Contemporary
series quote more consistent 5-year survival rates,
ranging from 35-55% [92,93,99,100]. Such survival
rates are not dissimilar to those of patients with either
urothelial or squamous cell carcinoma [54,121].
Indeed, Lughezzani et al [100] showed that while
patients with primary bladder adenocarcinoma are
significantly more likely to present with high stage
disease, stage and grade, the adjusted survival
was equivalent to that of patients with urothelial
carcinoma.

3. Chemotherapy

A number of factors have been shown to have


independent prognostic significance in patients with
adenocarcinoma of the bladder. Those consistently
reported include age, stage, grade and lymph
node involvement. Potential molecular prognostic
factors have also been assessed. El Sobky et
al [122] assessed the prognostic significance
of markers of angiogenesis in bilharzial primary
bladder adenocarcinoma. Fifty-five cases of
primary bladder adenocarcinoma were assessed
for immunohistochemical markers of angiogenesis
(anti-CD31), microvessel density (MVD) and DNA
ploidy in relation to survival. Only lymph node
involvement and MVD were independent prognostic
factors for survival. However, previously defined
clinical prognostic factors (stage and grade) did not
predict survival in this series suggesting possible
selection bias in the series (only 55 cases were
assessed from a previously reported larger cohort
of 185 patients). The prognostic significance of
DNA ploidy has been assessed by other groups with
conflicting results. Shaaban et al [123] showed that
DNA ploidy did not correlate with tumor stage or
grade but was predictive of lymph node metastases.

Urachal carcinoma is extremely rare and represents


less than 1% of all bladder tumors [121,131-133].
Previous reports consistently suggested that 33%
of bladder adenocarcinomas were urachal in
origin [105,118,120,131]. More recently, literature
suggests that just 10% of such tumors are urachal
[99]. Tumors present at a significantly younger age
(median age at presentation 47-56 years) and the
male:female distribution is less pronounced than
for other forms of bladder adenocarcinoma at 1.3:1
[99,103,132-134].

Reports of the use of chemotherapy in the treatment


of primary bladder adenocarcinoma are limited. The
literature is confined to case reports or very small case
series. The most successful chemotherapy regimens
utilize 5-fluorouracil (5-FU), which is extrapolated
from its use for gastrointestinal malignancies. Most
of the published series involve small numbers and
the response is universally unsatisfactory [126-130].

3. Urachal Adenocarcinoma
a) Epidemiology

b) Clinical and pathological features


Due to the frequent extravesical location of the
tumor, presentation is often late. The most common
presenting symptoms are hematuria, dysuria, and
frequency. Less common symptoms and signs
include mucosuria, abdominal pain, abdominal
mass, and, rarely, umbilical discharge [133,134].
Cross sectional imaging can be a useful diagnostic
tool (Figure 5). CT appearances consistent with
urachal adenocarcinoma include a mixed solid/
cystic midline mass, an unencapsulated caudal
tumor component involving the bladder wall with

420

a cystic encapsulated supravesical portion, mass


calcification, and presence of low attenuation areas
(consistent with mucin content) [135,136].

of glandular or polypoid proliferation, presence


of urachal remnants within the tumor, extension
of tumor into the bladder wall with involvement of
the space of Retzius, anterior abdominal wall, or
umbilicus, and no evidence of a primary neoplasm
elsewhere. These criteria are felt to be too strict
by many, who have adopted the more pragmatic
approach of the group at the MD Anderson Cancer
Center of considering all adenocarcinomas located
at the bladder dome or midline as adenocarcinoma
until proven otherwise [131].

There are no known risk factors for development of


urachal adenocarcinoma. The presence of a urachal
anomaly per se does not increase the risk of urachal
adenocarcinoma [137] However, patients with a
urachal mass who experience hematuria or are over
55 years of age, have a 17- and 3-fold increased
likelihood of urachal carcinoma, respectively [138].
The histological appearance of urachal carcinoma is
indistinguishable from enteric type adenocarcinoma
[131,139]. Recognized histological forms include:
enteric, mucinous, signet cell, and adenocarcinoma
not otherwise specified [140]. Two theories propound
as to the origin of these unique tumors. According
to Begg et al [141], this type of tumor arises from
enteric rests left behind following cloacal closure
during embryological development. The alternative
theory suggests development along a metaplastic
pathway [110,131].

Sheldon et al initially described the staging system


that is still widely used for urachal carcinoma [131]:
Stage I - tumor confined to the urachal mucosa
Stage II - local invasion confined to the urachus
Stage III - local extension into A) bladder, B)
abdominal wall, C) peritoneum, or D) local viscera
Stage IV - metastases to A) regional lymph nodes or
B) distant sites

Several diagnostic classification systems exist for


urachal adenocarcinoma. Sheldon and Mostofi
[110,131] listed several strict criteria for diagnosis
including tumor located at the dome of the bladder,
absence of cystitis cystica or cystitis glandularis,
predominant invasion of the muscularis or deeper
tissues with a sharp demarcation between the
tumor and surface bladder urothelium that is free

The Mayo clinic group proposed a simpler staging


system (Stage 1, confined to urachal mucosa,
stage 2, extending beyond the urachus/bladder,
stage 3, regional lymph node metastasis, stage 4,
distant metastasis) [138]. Given the similar survival
between patients with stage 3 and 4 disease, the MD
Anderson group suggested consolidation of these

Figure 5. 59 year old woman with urachal adenocarcinoma. The


anterior bladder wall is thickened (red arrows), with mild perivesical
fat stranding, and prominence of the urachus (blue arrows). A large
discrete mass is not noted here.

421

2 stages to create a simplified three tier staging


system [133]. Notwithstanding these modifications,
the Sheldon staging system is still the most widely
used system in current practice.

total urachectomy should be undertaken as part of


surgical treatment of all urachal tumors.

c) Treatment

A number of chemotherapeutic regimens have been


assessed in patients with either metastatic disease
at presentation, or high risk of disease recurrence
following surgical resection (positive surgical margins
or presence of lymph node metastases). Most reports
involve single patients or very small case series. In
common with primary bladder adenocarcinoma, the
most promising regimens appear to be 5-FU based
[127,133]. Of 20 patients receiving chemotherapy in
the MD Anderson Cancer Center series for metastatic
disease, there were only 4 patients with significant
objective responses, of whom 3 had received a 5-FU
based regimen. Subsequently the group is presently
performing a multi-center phase II clinical trial, using
5-FU, gemcitabine, cisplatin, and leucovorin [146].

2. Chemotherapy

1. Surgical Resection
Correct identification that a bladder tumor is urachal
in origin is imperative as this dictates modification
of standard bladder cancer surgical techniques,
to include complete excision of the urachus and
umbilicus. Surgical treatment options include
either 1) partial cystectomy with en bloc excision
of the urachus, posterior rectus fascia, peritoneal
reflections, and umbilicus, or 2) radical cystectomy
with en bloc excision of the urachus/umbilicus.
There are no randomized trials comparing these
surgical techniques, but durable results appear
achievable with either approach. Reported 5-year
survival rates following partial cystectomy with en
bloc urachal/umbilical excision range from 31-83%
[91,96,99,105,118,120,132-134,138,139,142]. The
more contemporary series estimate a 40-50% 5year survival with either partial or radical cystectomy
[132,138]. Some authors emphasize the need to
utilize frozen section at the time of surgery to ensure
negative surgical margins are achieved when
performing a partial cystectomy [133,138].

There are no diagnostic tumor markers for urachal


adenocarcinoma. However, elevated Ca19-9 and
CEA levels have been reported by a number of
authors [133,147,148]. These tumor markers have
been shown to correlate well with response to
chemotherapy [133].

4. Adenocarcinoma in Bladder
Exstrophy
a) Epidemiology

There are currently no data to support the therapeutic


benefit of bilateral pelvic lymphadenectomy as part
of the surgical management of urachal tumors.
Lymphadenectomy undoubtedly allows more
accurate disease staging and thereby provides
prognostic information. However, in the absence of
effective adjuvant therapies, its routine use cannot
be endorsed.

Adenocarcinoma is the most common bladder


cancer in patients with bladder exstrophy. Patients
with exstrophy have a reported 4% lifetime risk of
developing this type of malignancy [97]. In a review
of 40 patients with untreated bladder exstrophy,
McIntosh and Worley [149] found that 33 cases were
adenocarcinoma, the mean age at diagnosis was
44 years, and two-thirds of the patients were male.
Even with the advent of urinary diversion surgery, the
risk of developing carcinoma in the bladder remnant
is significant. Smeulders et al reported on a series of
102 patients born with bladder exstrophy, who were
followed for a minimum of 35 years [97]. There were
4 cases of bladder cancer, which was calculated to
be 694 times greater than the risk in the normal adult
population. Three of the 4 patients had undergone
a simple cystectomy before the age of 5, and all
were males. This implies that there was retention
of a portion of the bladder in these individuals, which
is speculated to become malignant in the presence
of sperm or secretions from the male genitourinary
tract.

There is very little available information describing


the feasibility or oncologic outcome associated
with minimally-invasive approaches (laparoscopic
or robotic) for urachal tumors [143-145]. None of
the follow-up data from these case reports are
mature enough to compare outcomes with those of
conventional open surgery.
Survival rates for patients with urachal adenocarcinoma appear to be superior to those of patients
with non-urachal adenocarcinoma [99,105]. Factors
predictive of disease recurrence following surgical resection include positive surgical margins and
high disease stage [133,138]. The Mayo clinic series
showed low stage, low pathologic grade, the undertaking of total urachectomy, and negative surgical
margin status to be associated with improved outcome on univariate analysis. Interestingly, stage and
undertaking total urachectomy no longer retained
statistical significance on multivariate analysis [138].
The authors (probably correctly) attribute this to the
low number of patients studied and emphasize that

b) Pathology and clinical features


The histopathology of the bladder in exstrophy
patients has been well-described [150]. The
development of adenocarcinoma of these bladder
remnants is thought to arise from either metaplasia
of the urothelium, which is exposed to inflammatory

422

stimuli, or as a result of the displacement of ectopic


colonic or rectal epithelium that occurs during
division of the cloaca. The fact that the majority of
the bladder malignancies occur during the fifth and
sixth decades of life and that adenocarcinoma has
been found to arise in other conditions resulting in
defunctionalized bladders and after augmentation
enteroplasty favors the metaplasia theory [151153]. The prognosis of these tumors is poor as a
result of late presentation. As a result, all patients
with exstrophy who have retained their bladders
should be followed closely, although no specific
follow-up regimen can be recommended based upon
the existing literature.

urinary bladder being one of the most common


organs involved [161,162]. Talamonti et al reported
on a series of 70 patients undergoing en bloc
bladder resection for colorectal carcinoma [163].
There were 58 primary and 12 recurrent tumors
in the series. Median survival was 34 months in
patients with negative surgical margins and 11
months for patients with positive margins. There
were no 5-year survivors. A more recent series
by Carne et al [164], retrospectively evaluated 53
patients who had secondary involvement of the
bladder with a colorectal cancer over a 15-year
period. The most common site of the primary tumor
was the sigmoid colon [46/53], with the remainder
invading from the rectum [2/53], the ascending colon
[4/53], and the transverse colon [1/53]. All 4 patients
undergoing blunt dissection of the tumor from the
bladder, without resection of the involved bladder,
developed local recurrences and subsequently died
of their disease. There were no local recurrences
in 4 patients undergoing total cystectomy and 3 of
the 4 patients with follow-up were alive, 2 without
and 1 with disease. Finally, in the 45 patients who
underwent partial cystectomy, there were a total of 8
(19%) local recurrences. The authors emphasized
the importance of obtaining frozen sections of the
resection margins at the time of partial cystectomy.
However, whether frozen sections were taken in
the patients who had local recurrences was not
reported.

5. Secondary (Metastatic) Bladder


Adenocarcinoma
Secondary bladder adenocarcinomas are tumors
that involve the bladder by either direct extension
from adjacent organs, reimplantation from adenocarcinoma located higher up within the urinary tract, or
lympho-hematogeneous spread. A retrospective review of surgical specimens collected over a 90-year
time period [154] showed that secondary bladder
tumors comprised 2.3% of 5533 surgical specimens
assessed. The most common sites of secondary tumor origin were, in order of their occurrence, colonic,
prostatic, rectal and cervical. Over 50% of tumors
assessed were adenocarcinomas. Indeed, secondary adenocarcinoma of the bladder is more common than primary adenocarcinoma [155]. Some of
these secondary adenocarcinomas are histologically
distinct, allowing rapid identification of tumor origin,
such as prostatic tumors. However, many present
the clinician with a diagnostic challenge when trying to determine tumor origin, which is clearly of importance in planning subsequent disease management. In the first instance, the clinical history, endoscopic exams, and radiological studies are valuable
in identifying tumor origin. When such studies are
not informative, immunohistochemical staining of
tumor biopsy specimens may be undertaken. Two
groups have reported that lack of CDX2 and villin
staining indicates that an adenocarcinoma sampled
from the bladder is unlikely to be colorectal in origin
[156,157]. Positive PSA staining is pathognomonic
of prostatic origin and vimentin staining indicates
endometrial/ovarian origin [158]. Many other immunostains have been assessed, including CK7, CK20,
CEA and Ca19-9, however, most lack the specificity
to confidently discriminate tumor origin [158]. Similarly, assessment of mucin produced by the tumor
has failed to add clarity when assessing tumor origin
[159]. Urine cytology can be useful at identifying tumors of secondary origin, for example, prostatic or
colonic, although it lacks sensitivity in the diagnosis
of primary bladder adenocarcinoma[160].

Treatment Recommendations

The treatment of adenocarcinoma depends upon


its subclassification. Primary adenocarcinoma is
poorly responsive to radiation and chemotherapy
and should be treated with radical cystectomy
(Grade B).

Urachal adenocarcinoma should be treated with

en bloc resection of the urachus and umbilicus with


partial or radical cystectomy (Grade B).

The incidence of adenocarcinoma is much higher

in exstrophy patients. Any patient with bladder


exstrophy who has retained his or her bladder
should be closely followed, though an exact
regimen cannot be defined based on currentlyavailable evidence (Grade C).

Patients with metastatic primary or metastatic


urachal adenocarcinoma of the bladder should be
considered for enrollment into a clinical trial using
5-FU based chemotherapy (Grade C).

In patients with locally advanced colonic or


rectal tumors that secondarily involve the bladder,
patients may undergo complete resection of the
involved portion of the bladder, either with partial
cystectomy and verified negative margins or radical
cystectomy (Grade B).

Approximately 10% of colorectal adenocarcinomas


will be attached to adjacent structures, with the

423

nuclei. This gives a characteristic appearance of


round blue cells under the microscope with routine
hematoxylin and eosin staining. Frequent mitotic
figures and extensive necrosis are common in
small cell carcinoma, and are indicative of its often
aggressive behavior [176]

V. Small Cell Carcinoma


1. EPIDEMIOLOGY
Primary small cell carcinoma is a neuroendocrine
tumor that is rare, with just under 300 cases reported
in the English-language literature. Despite this, this
histologic subtype accounts for 0.48-0.7% of all
cases of primary bladder tumors [165-169].

The diagnosis of small cell carcinoma may be confirmed by immunohistochemical stains for neuronspecific enolase, chromogranin A, and synaptophysin [168, 177]. In more complex cases, additional
immunohistochemical stains for pan-cytokeratin and
p63 may be needed to distinguish it from urothelial
carcinoma [177].

The incidence of small cell carcinoma appears to


be increasing in the United States. A recent review
of SEER data identified 642 cases of small cell
carcinoma of the urinary bladder between 1991
and 2005 [169]. During that time period, there was
a 300% increase in the incidence of this tumor,
from 0.05 to 0.14 cases per 100,000 population.
Additionally, relative to all other bladder tumors,
small cell carcinoma increased from 0.3% to 0.6% of
all histologic types of bladder cancer,

A number of studies have shown that small cell


carcinoma exists as either pure small cell carcinoma
or in combination with urothelial carcinoma [170172]. Rarely, small cell carcinoma may coexist with
squamous cell carcinoma or adenocarcinoma [171,
178]. The implications of this are that in pure small
cell carcinoma, the chemotherapeutic regimen should
be specifically directed to small cell carcinoma.
Extremely poor responses have been noted in those
patients treated with regimens that are specific for
urothelial carcinoma (MVAC or taxol, methotrexate,
and cisplatin) [179].

2. Clinical Features
Most of the data regarding the clinical features of
small cell carcinoma of the bladder is derived from
only a few series and case reports. The vast majority
of patients diagnosed with small cell carcinoma are
male, with a propensity for it to affect patients in
their late 60s and early 70s [169-171]. Caucasians
account for 90% of patients with small cell carcinoma
in the SEER-cancer registry [169].

4. Treatment
Local treatment with either radiation therapy [170]
or surgery alone [170, 171,179] yields very poor
results in patients with clinically localized small cell
carcinoma of the bladder. Based upon the treatment
paradigm developed for small cell lung cancer,
primary chemotherapy followed by either radiation
therapy or cystectomy seems to provide the optimal
means of managing patients with this tumor [180184]. Based upon the chemotherapeutic regimen
used for small cell lung cancer, the agents used are
etoposide and cisplatin, possibly alternating with
ifosfamide and doxorubicin.

Similar to other bladder malignancies, hematuria is


the presenting complaint in close to 90% of patients
[171,172]. On occasion, paraneoplastic syndromes
may herald the diagnosis [173,174].
Small cell carcinoma usually cannot be visually distinguished from urothelial carcinoma on cystoscopy.
Urine cytology may be helpful in providing a suspicion of small cell carcinoma. A retrospective analysis of 29 urine specimens from patients diagnosed
with small cell carcinoma showed that 56% could be
diagnosed by urinary cytopathology [175]. The other
cases were interpreted as high grade urothelial carcinoma. Five patients had both small cell carcinoma
and urothelial carcinoma.

a) Chemotherapy and cystectomy


In the small series reviewed by Sved et al [170],
13 of 18 patients who underwent cystectomy plus
chemotherapy were alive at a mean of 27 months.
Similarly, Walther [165] reported favorable response
rates in seven patients treated with systemic etoposide and cisplatin in neoadjuvant and adjuvant protocols. In a retrospective analysis from MD Anderson Cancer Center, median and 5-year disease-free
survival was significantly improved in patients who
received preoperative chemotherapy [179]. In this
study, 5-year survival was 36% for 25 patients who
underwent cystectomy alone compared with 78% in
those who received preoperative chemotherapy. Interestingly, an evaluation of the surgical specimens
indicated that 10/12 patients receiving a regimen
directed towards small cell carcinoma (either etoposide and cisplatin or ifosfamide and doxorubicin)
had no residual tumor in the bladder versus 3 of 9

At the time of presentation, 90% of patients with


small cell carcinoma have muscle-invasive disease,
approximately 50% have T3 or T4 disease [169],
and 24-67% have metastases [169,170]. Small cell
carcinoma metastasizes most commonly to lymph
nodes, liver, bone, lung, and brain [170].

3. Pathologic Features
Microscopically, small cell carcinoma of the bladder
resembles small cell carcinoma of the lung. The tumor
is comprised of a population of relatively uniform
cells with scant cytoplasm and hyperchromatic

424

patients receiving chemotherapy directed towards


urothelial cancer (MVAC or taxol, methotrexate, and
cisplatin).

hematuria, and the diagnosis is made early. The


majority of the tumors are high grade and may attain
very large size before recognition (see Figure 6).
Tumor grade is established on the basis of mitotic
rate and proliferation indices rather than nuclear
atypia [190]. The preferred treatment for localized
disease is radical cystectomy with negative margin
resection. In the largest series to date, encompassing
10 patients, the 5-year survival rate was 80%.
Metastatic sarcomas are treated with multimodality
protocols. Doxorubicin and ifosfamide are the most
active single agents available [191].

b) Chemotherapy and radiation therapy


Radiation therapy has been used as an alternative to
cystectomy in many centers, and seems to provide
comparable results when used with neoadjuvant
chemotherapy. In the largest series to date, 17
patients with localized small cell bladder cancer
were treated over a 14-year period using platinumbased chemotherapy followed by radiation therapy
to 56-70 Gy [184]. The median overall survival was
32.5 months, with 88% of patients showing an initial
complete response. In another retrospective study
evaluating 10 patients, the 2- and 5-year survival
rates in patients treated with chemotherapy and
radiation were 70% and 44%, respectively [185].
Of note, local recurrences warranting surgery have
occurred with this approach [184]

2. C
 arcinosacroma and
carcinomatoid Tumors
Carcinosarcoma is a rare type of primary tumor
composed of an admixture of malignant epithelial
(carcinoma) and malignant soft tissue elements
(sarcoma). The term sarcomatoid has been used to
describe a malignant spindle cell tumor with epithelial
differentiation [192]. In a series of 41 patients,
Lopez-Beltran reported that both carcinosarcoma
and sarcomatoid tumors had similar presentations
[192]. The most common epithelial component was
urothelial in both types. Most patients had locally
advanced tumors at the time of diagnosis. Despite
aggressive surgical management, the outcome is
poor. Treatment failure occurs within 1 to 2 years
following treatment [192]. Metastatic disease showed
a favorable response to cisplatin and gemcitabine
[193].

c) Chemotherapy alone
This is primarily used in patients with metastatic
disease. Similar to small cell lung cancer, there is
an initial response seen in most patients. In a phase
II trial evaluating 30 patients with metastatic small
cell bladder cancer [185], 72% of patients had a
complete response, and 28% of patients had a
partial response. The durability of the responses is
short-lived, however, as the median overall survival
with this disease is between 5 and 13.3 months
[184,186].

3. Paraganglioma and Pheochromocytoma

Second line therapy has been evaluated in isolated


case reports using vinorelbine [187]. Prophylactic
brain irradiation has also been evaluated, but limited
data is available. [185,188]

These are extra-adrenal neoplasms derived from


neural crest cells. Bladder pheochromocytoma
accounts for 0.05% of bladder tumors. It may be
derived from embryonic rests of chromaffin cells in
the detrusor sympathetic plexus. It accounts for 10%
of extra-adrenal pheochromocytoma. Malignancy
was demonstrated in 10% and characterized by
local invasion, regional lymph node metastases, or
distant spread.

Conclusions
Small cell carcinoma of the bladder is an aggressive
disease that often presents in advanced stages.
Because of the rarity of this lesion, evidence is
limited. Aggressive multimodal therapy is warranted,
yet altogether this subtype of bladder cancer carries
a dismal prognosis.

Bladder pheochromocytoma may be hormonally


active and presents with attacks of paroxysmal
hypertension, headaches, palpitations, blurred vision,
and sweating associated with the act of micturition
[194]. If the disease is suspected, cystoscopy should
be performed under adrenergic blockade in the
operating room. The gross appearance is often a
solitary, submucosal, or intramural nodule. Biopsy
should be avoided. The diagnosis depends on CT
scan or MRI for anatomical location and the extent
of the lesion. Isotopic scanning using 131iodine
metaiodinebenzylguinidine (MIBG) is the study of
choice for localizing small pheochromocytomas with
more than 90% specificity [195]. Positron emission
tomography was recently used with high sensitivity
as well [194].

VI. Other Bladder Tumors


1. Bladder Sarcoma
Malignant soft tissue tumors represent the most
common histologic type of the non-epithelial
bladder tumors. Half of bladder sarcomas are
leiomyosarcoma, 20% are rhabdomyosarcoma, and
the remainder are angio-, osteo-, and carcinosarcoma
[189]. The histological pattern of leiomyosarcoma
is characterized by interwoven bundles of spindleshaped cells. The incidence is higher in patients
with previous local radiation treatment or systemic
chemotherapy [190]. Most patients present with

425

Figure 6. 62 year old male with leiomyosarcoma of the


urinary bladder. The mass fills the entire urinary bladder
and pelvis, leaving only a small area in which urine can
accumulate (arrows). Despite its massive size, the patient
underwent cystoprostatectomy with cancer- free surgical
margins and remains disease free at 5 years.

The standard treatment is local excision via partial


cystectomy combined with pelvic lymph node
dissection. The surgery is employed under the same
precautions as in adrenal pheochromocytoma with
controlled adrenergic blockade.91

urethra more than the bladder. Secondary melanoma


of the bladder was found in patients with widespread
metastatic melanoma of the skin [190, 198]. The
patients history and careful examination of the skin
is essential to confirm the primary nature of the
tumor. The histologic picture of bladder melanoma
is similar to other melanomas. It is composed of
large malignant cells arranged in nests with variable
amounts of pigments. The cell origin of bladder
melanoma is undefined. Treatment of primary
bladder melanoma is radical surgery. The prognosis
is poor [199, 200].

It may be difficult to distinguish benign and malignant


lesions if the disease is localized. Lifelong follow-up is
important as the malignant pheochromocytoma may
show local recurrence with or without metachronous
metastases.

4. Bladder Pseudotumors

6. Lymphoma

Bladder pseudotumors are rare and may resemble


malignancy. The etiology and histogenesis
remain unclear. Some of these lesions present
as postoperative spindle-cell tumors. It may be
difficult to distinguish them from leiomyosarcoma
on a histopathologic basis [196]. However, absence
of significant nuclear atypia, less than three mitotic
figures per high power field, and presence of spindle
cells with myxoid degeneration and eosinophilic
cytoplasm favor pseudotumor. Some tumors may
show a fascicular growth pattern with deposition of
interstitial collagen [197]. Local recurrence or distant
metastases are rare following tumor excision. If the
diagnosis is clear, transurethral resection or partial
cystectomy is sufficient. Radical cystectomy may be
required if the diagnosis is difficult to distinguish from
bladder sarcoma.

Bladder lymphoma is usually a part of metastatic


spread of systemic disease. Primary lymphoma
is very rare. Microscopic analysis shows diffuse
infiltration of lymphoid cells into the normal structures
of the bladder [201]. Primary lymphoma is more
common in women [202]. It is mostly localized
and of low grade with good prognosis [201]. Local
irradiation is the recommended treatment with a high
recurrence-free survival [201, 202].

VIII. summary of the recommendations


Squamous Cell Carcinoma

5. Melanoma

Patients with SCI should be aware of the risk of


development of SCC. A surveillance schedule,
however, cannot be determined from the currently
available evidence. (Grade D)

Primary bladder melanoma is very rare. It affects the

Cystectomy is the best primary therapy for SCC,

426

Paraganglioma and Pheochromocytoma

whether it is related to schistosomiasis or is not


associated with schistosomiasis (Grade B).

The standard treatment for patients with


paraganglioma and pheochromocytoma is partial
cystectomy with pelvic lymphadenectomy, with
the same precautions taken as for any other
pheochromocytoma with preoperative adrenergic
blockade (Grade C).

Neoadjuvant chemotherapy or radiation therapy


prior to cystectomy for SCC is not advisable, given
the information available to date (Grade D).
In patients with metastatic disease, epirubicinbased chemotherapy regimens offer a higher
likelihood of an initial complete response than other
agents. The long-term survival benefits remain to
be determined (Grade A).

References

Adenocarcinoma

1. Ibrahim AS and Khaled HM: Urinary Bladder Cancer. from:


Freedman LS, Edwards BK, Ries LAG, Young JL (eds):
Cancer incidence in four member countries (Cyprus, Egypt,
Israel, and Jordan) of the Middle East Cancer Consortium
(MECC) compared with US SEER. Chapter 11;, National
Cancer Institute. NIH Pub No 06-5873. Bethesda, MD. P
105.

Patients with bladder exstrophy who have retained


their bladders should be closely followed, but no
particular regimen can be recommended on the
basis of currently available evidence (Grade D).
Patients with primary adenocarcinoma of the
urinary bladder should be treated with radical
cystectomy, when possible (Grade B). There is
no role for neoadjuvant radiation or chemotherapy,
based upon the available literature (Grade D).

2. Ploeg M, Aben KK, Hulsbergen-van de Kaa CA, Schoenberg


MP, Witjes JA, Kiemeney LA: Clinical epidemiology of
nonurothelial bladder cancer: analysis of the Netherlands
Cancer Registry. J Urol 2010, 183, 915-20.
3. Zhang Y, Zhu C, Curado MP, Zheng T, Boyle P: Changing
patterns of bladder cancer in the USA: evidence of
heterogeneous disease.
2011, BJU International, in
press

Patients with urachal adenocarcinoma should be


treated with en bloc excision of the urachus and the
umbilicus with either partial or radical cystectomy
(Grade B).

4. Miller A, Mitchell JP, and Brown NN. The Bristol Bladder


Tumor Registry. Br J Urol 1969;41(suppl):1 64.
5. Johnson DE, Schoenwald MB, Ayala AG, et al. Squamous
cell carcinoma of the bladder. J Urol 1976;115:542544.

Patients with metastatic adenocarcinoma of the


bladder should undergo complete resection of
the involved portion of the bladder, with partial
cystectomy with verified negative margins or with
the use of radical cystectomy (Grade B).

6. Rundle JSH, Hart AJL, McGeorge A, et al. Squamous cell


carcinoma of the bladder: a review of 114 patients. Br J Urol
1982; 54:522526.
7. Lopez JI, Angulo Cuesta J, Flores Corral N, et al. Squamous
cell carcinoma of the urinary bladder: clinico-pathologic
study of 7 cases [in Spanish]. Arch Esp Urol 1994;47:756
760.

Small Cell Carcinoma


When small cell carcinoma is identified on a
TURBT specimen, the patient should undergo a full
metastatic work-up including a CT of the abdomen
and pelvis, a bone scan, a chest x-ray, and a
neurologic examination. (Grade C)

8. Serretta V, Pomara G, Piazza F, et al. Pure squamous cell


carcinoma of the bladder in Western countries. Eur Urol
2000;37: 85 89.
9. Johansson SL, and Cohen SM. Epidemiology and etiology
of bladder cancer. Semin Surg Oncol 1997;13:291298.
10. Fleshner N, Herr HW, Stewart AK, et al. The National
Cancer Data Base report on bladder carcinoma. Cancer
1996;78: 15051513.

Patients with small cell carcinoma of the urinary


bladder require aggressive combination therapy to
achieve cures, such as combined chemotherapy
and radical cystectomy or chemotherapy and
radiation therapy. (Grade B)

11. Mungan NA, Kiemeney LA, van Dijck JA, et al. Gender
differences in stage distribution of bladder cancer. Urology
2000; 55:368 371.
12. Porter MP, Voigt LF, Penson DF, et al. Racial variation in the
incidence of squamous cell carcinoma of the bladder in the
United States. J Urol 2002;168:1960 1963.

Bladder Sarcoma
Patients with bladder sarcoma should be treated
with radical cystectomy with negative surgical
margins (Grade C).

13. Cohen SM, Shirai R, and Steineck G. Epidemiology and


etiology of premalignant and malignant urothelial changes.
Scand J Nephrol Suppl 2000;205:105115.

Those with metastatic sarcoma should be treated


with a multimodality protocol (Grade C).

14. Locke JR, Hill DE, and Walzer Y. Incidence of squamous


cell carcinoma in patients with long term catheter drainage.
J Urol 1985;133:1034 1035.

Carcinosarcoma and Sarcomatoid Tumors

15. Bejany DE, Lockhart JL, and Rhamy RK. Malignant vesical
tumors following spinal cord injury. J Urol 1987;138:1390
1392.

Carcinosarcoma and sarcomatoid tumors have


a poor prognosis and surgical management is
inadequate. Multimodality therapy is recommended
(Grade C).

16. Kaufman JM, Fam B, Jacobs SC, et al. Bladder cancer and
squamous metaplasia in spinal cord injury patients. J Urol
1977;118: 967971.

427

17. Sene AP, Massey JA, McMahon RTF, et al. Squamous


cell carcinoma in a patient on clean intermittent self
catheterization. Br J Urol 1990;65:213214.

39. Kassouf W, Spiess PE, Siefker-Radtke A, Swanson D,


Grossman HB, Kamat AM, et al. Outcome and patterns of
recurrence of nonbilharzial pure squamous cell carcinoma
of the bladder: a contemporary review of The University of
Texas M D Anderson Cancer Center experience. Cancer.
2007 Aug 15;110(4):764-9.

18. Zaidi SZ, Theaker JM, and Smart CJ. Squamous


cell carcinoma in a patient on clean intermittent self
catheterization [case report]. Br J Urol 1997;80:352353.

40. Nishiyama H, Habuchi T, Watanabe J, et al. Clinical


outcome of a large-scale multi-institutional retrospective
study for locally advanced bladder cancer: a survey
including 1131 patients treated during 1990-2000 in Japan.
Eur Urol 2004;45:176181.

19. West DA, Cummings JM, Longo WE, et al. Role of chronic
catheterization in the development of bladder cancer in
patients with spinal cord injury. Urology 1999;53:292297.
20. Bickel A, Culkin DJ, Wheeler JS Jr. Bladder cancer in spinal
cord injury patients. J Urol 1991;146:12401242.

41. Scosyrev E, Yao J, Messing E. Urothelial carcinoma versus


squamous cell carcinoma of bladder: is survival different
with stage adjustment? Urology. 2009 Apr;73(4):822-7.

21. Kalisvaart JF, Katsumi HK, Ronningen LD, Hovey RM.


Bladder cancer in spinal cord injury patients. Spinal Cord
2010 Mar;48(3):257-61.

42. Swanson DA, Liles A, and Zagars GK. Pre-operative


irradiation and radical cystectomy for stages T2 and
T3 squamous cell carcinoma of the bladder. J Urol
1990;143:3740.

22. Pannek J. Transitional cell carcinoma in patients with spinal


cord injury: a high risk malignancy? Urology 2002;59:240
244.
23. Kantor AF, Hartge P, Hoover RN, et al. Epidemiological
characteristics of squamous cell carcinoma and adenocarcinoma of the bladder. Cancer Res 1988;48:38533855.

43. Wishnow KI, and Dmochowski R. Pelvic recurrence after


radical cystectomy without preoperative radiation. J Urol
1988;140: 42 43.

24. Thorn M, Bergstrom R, Johansson AM, et al. Trends in


urinary bladder cancer incidence in Sweden 1960-93 with
special reference to histopathology, time period birth cohort,
and smoking. Cancer Causes Control 1997;8:560 567.

44. Stenzl A, Jarolim L, Coloby P, et al. Urethra-sparing


cystectomy and orthotopic urinary diversion in women with
malignant pelvic tumors. Cancer 2001;92:18641871.
45. Khaled HM, Hamza MR, Mansour O, et al. A phase II study
of gemcitabine plus cisplatin chemotherapy in advanced
bilharzial bladder carcinoma. Eur J Cancer 2000;36(suppl
2):3437.

25. Connery DB. Leukoplakia in the urinary bladder and its


association with carcinoma. J Urol 1953;69:121127.
26. Holley PS, and Mellinger GT. Leukoplakia of the bladder
and carcinoma. J Urol 1961;86:235241.

46. Loehrer PJ Sr, Einhorn LH, Elson PJ, et al. A randomized


comparison of cisplatin alone or in combination with
methotrexate, vinblastine, and doxorubicin in patients with
metastatic urothelial carcinoma: a cooperative group study.
J Clin Oncol 1992; 10:1066 1073.

27. OFlynn JD, and Mullaney J. Vesical leukoplakia progressing


to carcinoma. Br J Urol 1974;46:3137.
28. Wiener DP, Koss LG, Sablay B, et al. The prevalence and
significance of Brunns nests, cystitis cystica, and squamous
metaplasia in normal bladders. J Urol 1979;122:317321.

47. Sternberg CN, Yagoda A, Scher HI, et al. Preliminary results


of M-VAC (methotrexate, vinblastine, doxorubicin and
cisplatin) for transitional cell carcinoma of the urothelium.
J Urol 1985;133:403 407.

29. Westenend PJ, Stoop JA, and Hendriks JGH. Human


papilloma viruses 6/11, 16/18 and 31/33/51 are not
associated with squamous cell carcinoma of the urinary
bladder. BJU Int 2001;88:198201.

48. Raghavan D. Progress in the chemotherapy of metastatic


cancer of the urinary tract. Cancer 2003;97(suppl 8):2050
2055.

30. Tatsura H, Ishiguro Y, Okamuro T, et al. Bladder squamous


cell carcinoma with human papilloma virus type 6 (HPV 6).
Int J Urol 1995;2:347349.

49. Broecker BH, Klein FA, and Hackler RH. Cancer of the
bladder in spinal cord injury patients. J Urol 1981;125:196
197.

31. Fadl-Elmula I, Gorunova L, Lundgren R, et al. Chromosomal


abnormalities in 2 bladder carcinomas with secondary
squamous cell differentiation. Cancer Genet Cytogenet
1998;102:125130.

50. Navon JD, Soliman H, Khonsari F, et al. Screening


cystoscopy and survival of spinal cord injured patients with
squamous cell cancer of the bladder. J Urol 1997;157:2109
2111.

32. Bessett Pl, Abell MR, and Herwig KR. A clinicopathologic


study of squamous cell carcinoma of the bladder. J Urol
1974;112: 6667.

51. Celis JE, Rasmussen HH, Vorum H, et al. Bladder squamous


cell carcinomas express psoriasin and externalize it to the
urine. J Urol 1996;155:21052112.

33. Jones MA, Bloom HJG, Williams G, et al. The management


of squamous cell carcinoma of the bladder. J Urol
1980;52:511514.

52. Tsukamoto T, Kumamoto Y, Ohmura K, et al. Squamous cell


carcinoma associated antigen in uro-epithelial carcinoma.
Urology 1992;40:477 483.
53. Abrams P, Grant A, and Khoury S. Evidence-based
medicine: overview of the main steps for developing and
grading guideline recommendations. In Abrams P, Cardozo
L, Khoury S, et al (Eds), Incontinence, Volume 1. Plymouth,
United Kingdom, Health Publications Ltd., pp 10 11.

34. Quilty PM, and Duncan W. Radiotherapy for squamous cell


carcinoma of the urinary bladder. Int J Radiat Oncol Biol
Phys 1986;12:861 865.
35. Costello AJ, Tiptaft RC, England HR, et al. Squamous cell
carcinoma of bladder. Urology 1984;23:234236.
36. Debbagh A, Bennani S, Hafiani M, et al. Epidermoid
carcinoma of the bladder: a report of 14 cases. Ann Urol
(Paris) 1997;31: 199 203.

54. Ghoneim MA, El-Mekresh MH, El-Baz MA, et al. Radical


cystectomy for carcinoma of the bladder: critical evaluation
of the results in 1026 cases. J Urol 1997;158:393399.

37. Richie JP, Waisman J, Skinner DG, et al. Squamous


carcinoma of the bladder: treatment by radical cystectomy.
J Urol 1976;115: 670 671.

55. El-Bolkainy MN (Ed): Topographic Pathology of Cancer.


Cairo, Egypt, Cairo University, The National Cancer
Institute of Cairo University, 1998, pp 59 63.

38. Lagwinski N, Thomas A, Stephenson AJ, Campbell S,


Hoschar AP, El-Gabry E, et al. Squamous cell carcinoma of
the bladder: a clinicopathologic analysis of 45 cases. Am J
Surg Pathol. 2007 Dec;31(12):1777-87.

56. El-Bolkainy MN, Ghoneim MA, and Mansour MA. Carcinoma


of the bilharzial bladder in Egypt: clinical and pathological
features. Br J Urol 1972;44:561570.

428

57. El-Sebai l, Sherif M, El-Bolkainy MN, et al. Verrucose


squamous carcinoma of bladder. Urology 1974;4:407
410.

75. Awaad HK, Hegazy M, Ezzat S, et al. Cell proliferation of


carcinoma in bilharzial bladder: an autoradiology study.
Cell Tissue Kinet 1979;12:513520.

58. Salem S, Mitchell RE, El-Dorey AE, Smith JA, Barocas DA:
Successful control of schistosomiasis and the changing
epidemiology of bladder cancer in Egypt. BJU International.
2010, 107, 206-11.

76. Denekamp J. The relationship between the cell loss factor


and the immediate response to radiation in animal tumors.
Eur J Cancer 1972;8:335340.
77. Awaad HK. Radiation therapy in bladder cancer. Alexandria
Journal of Medicine 1958;4:118 122.

59. Fedewa SA, Soliman AS, Ismail K, Hablas A, Seifeldin IA,


Ramadan M, Omar HG, Nriagu J, Wilson ML: Incidence
analyses of bladder cancer in the Nile delta region of Egypt.
Cancer Epidem 2009, 33, 176-81.

78. Omar AH, Shalaby MA, and Ibrahim AH. On the capillary
vascular bed in carcinoma of the urinary bladder. East Afr
Med J 1975;51:3441.

60. Zaghloul MS, Nouh A, Moneer M, El-Baradie M, Nazmy M,


Younis A. Time-trend in epidemiological and pathological
features of schistosoma-associated bladder cancer. J
Egypt Natl Canc Inst. 2008 Jun;20(2):168-74.

79. El-Hammady SM, Ghoneim MA, Hussein ES, et al.


Segmental resection for carcinoma of the bladder.
Mansoura Medical Bulletin 1975;3:191197.
80. Omar SM. Segmental resection for carcinoma of the
bladder. J Egypt Med Assoc 1969;52:975978.

61. Zarzour AH, Selim M, Abd-Elsayed AA, Hameed DA,


Abdelaziz MA. Muscle invasive bladder cancer in Upper
Egypt: the shift in risk factors and tumor characteristics.
BMC Cancer. 2008;8:250.

81. Ghoneim MA, and Awaad HK. Results of treatment in


carcinoma of the bilharzial bladder. J Urol 1980;123:850
852.

62. Gouda I, Mokhtar N, Bilal D, El-Bolkainy T, El-Bolkainy NM.


Bilharziasis and bladder cancer: a time trend analysis of
9843 patients. J Egypt Natl Canc Inst. 2007 Jun;19(2):15862.

82. Ghoneim MA, Ashamallah AG, Hammady S, et al.


Cystectomy for carcinoma of the bilharzial bladder: 138
cases 5 years later. Br J Urol 1979;51:541544.
83. Awaad HK, Abdel Baki H, El-Bolkainy N, et al. Pre-operative
irradiation of T3 carcinoma in bilharzial bladder. Int J Radiat
Oncol Biol Phys 1979;5:787794.

63. Felix AS, Soliman AS, Khaled H, Zaghloul MS, Banerjee


M, El-Baradie M, El-Kalawy M, Abd-Elsayed AA, Ismail
K, Hablas A, Sifeldin IA, Ramadan M, Wilson ML: The
changing patterns of bladder cancer in Egypt over the past
26 years. Cancer Causes Control 2008, 19,421-9.

84. Ghoneim MA, Ashamallah AK, Awaad HK, et al. Randomized


trial of cystectomy with or without pre-operative radiotherapy
for carcinoma of the bilharzial bladder. J Urol 1985;134:266
268.

64. Hicks RM. Multistage carcinogenesis in the urinary bladder.


Br Med Bull 1980;36:3946.

85. Denekamp J, McNally NJ, Fowler JF, et al. Misonidazole in


fractionated radiotherapy: are many small fractions best?
Br J Radiol 1980;53:981990.

65. Hicks RM, Walters CL, El-Sebai I, et al. Demonstration


of nitrosamines in human urine: preliminary observations
on a possible aetiology for bladder cancer in association
with chronic urinary tract infections. Proc R Soc Med
1977;70:413417.

86. Gad-el-Mawla N, Hamza MR, Zikri ZK, et al. Chemotherapy


in invasive carcinoma of the bladder: a review of phase II
trials in Egypt. Acta Oncol 1989;28:7376.

66. El-Merzabani MM, and El-Aaser AA. Etiological factors of


bilharzial bladder cancer in Egypt: nitrosamines and their
precursors in urine. Eur J Cancer 1979;15:287291.

87. Gad-el-Mawla N, Mansour MA, Eissa S, et al. A randomized


pilot study of high-dose epirubicin as neoadjuvant
chemotherapy in the treatment of cancer of bilharzial
bladder. Ann Oncol 1991;2: 137140.

67. Youssef R, Kapur P, Kabbani W, Shariat SF, Mosbah A,


Abol-Enein H, et al. Bilharzial vs non-bilharzial related
bladder cancer: pathological characteristics and value of
cyclooxygenase-2 expression. BJU International. 2010
Nov 24.

88. Khalid H, Zaghloul M, Ghoneim M, et al. Gemcitabine


and cisplatin as a neoadjuvant chemotherapy for invasive
bladder cancer: effect on bladder preservation and survival.
Proc Am Soc Clin Oncol 2003;22:411, Abstract 1652.

68. Youssef RF, Shariat SF, Kapur P, Kabbani W, Mosbah A,


Abol-Enein H, et al. Prognostic value of cyclooxygenase-2
expression in squamous cell carcinoma of the bladder. The
Journal of Urology. 2011 Mar;185(3):1112-7.

89. Jacobo E, Loening S, Schmidt JD, Culp DA. Primary


adenocarcinoma of the bladder. J Urol 1977;17:54-56.
90. Bennett JK, Wheatley JK, Walton KN. 10-year experience
with adenocarcinoma of the bladder. J Urol 1984;131:262263.

69. Abdulamir AS, Hafidh RR, Kadhim HS, Abubakar F. Tumor


markers of bladder cancer: the schistosomal bladder
tumors versus non-schistosomal bladder tumors. J Exper
Clin Cancer Res. 2009; 28:27.

91. Thomas DC, Ward AM, Williams JL. A study of 52 cases of


adenocarcinoma of the bladder. Br J Urol 1971;43:4-15.

70. Ghoneim MA, Mansour MA, and El-Bolkainy MN. Staging


of carcinoma of bilharzial bladder. Urology 1974;3:4042.

92. Zaghloul MS, Nouh A, Nazmy M, Ramzy S, Zaghloul AS,


Sedira AM and Khalil E. Long term results of primary
adenocarcinoma of the urinary bladder: A report on 192
patients. Urol Oncol 2006:24:13-20.

71. El-Bolkainy MN, Ghoneim MA, El-Morsey BA, et al.


Carcinoma of bilharzial bladder: diagnostic value of urine
cytology. Urology 1974;3:319 323.

93. El-Mekresh M, El-Baz M, Abol-Enein H, Ghoneim M.


Primary adenocarcinoma of the urinary bladder: a report of
185 cases. Br J Urol 1998;82:206-212.

72. Basta MT, Attallah AM, Seddek MN, et al. Cytokeratin


shedding in urine: a biological marker for bladder cancer?
Br J Urol 1988; 61:116 121.
73. Ghoneim MA. Nontransitional cell bladder cancer. In
Krane RJ, Siroky MB, Fitzpatrick JM (Eds), Clinical Urology.
Philadelphia, Lippincott, 1994,679-87.

94. Kantor AF, Hartge P, Hoover RN, Fraumeni JF Jr.


Epidemiological characteristics of squamous cell
carcinoma and adenocarcinoma of the bladder. Cancer
Res 1988;48:3853-3855.

74. Kafagy MM, El-Bolkainy MN, and Mansour MA. Carcinoma


of the bilharzial urinary bladder: a study of the associated
lesions in 86 cases. Cancer 1972;30:150 159.

95. Makar N. Some observations on pseudoglandular proliferation in the bilharzial bladder. Acta Union Internationalis
Contra Cancrum 1962;18:599.

429

96. Anderstrom C, Johansson SL, von Schultz L. Primary


adenocarcinoma of the urinary bladder. A clinicopathologic
and prognostic study. Cancer 1983;52:1273-1280.

117. Abenoza P, Monivel C, Fraley E. Primary adenocarcinoma


of the urinary bladder. A clinico-pathologic study of 16
cases. Urology 1987;31:9-14.

97. Smeulders N and Woodhouse CR. Neoplasia in adult


exstrophy patients. BJU Int 2001;87:623-628.

118. Burnett AL, Epstein JI, Marshall FF. Adenocarcinoma of


urinary bladder: classification and management. Urology
1991;67:315-321.

98. Wheeler JD and Hill WT. Adenocarcinoma involving the


urinary bladder. Cancer 1954;7:119-135.

119. Kamat MR, Kulkarni JN and Tongaonkar HB.


Adenocarcinoma of the bladder: study of 14 cases and
review of the literature. Br J Urol 1991;86:254-257.

99. Wright JL, Porter MP, Li CL, Lang PH and Lin DW.
Differences in survival among patients with urachal and
non-urachal adenocarcinomas of the bladder. Cancer
2006;107:721-728.

120. Wilson TG, Pritchatt RR, Lieskovsky G, Warner NE and


Skinner DG. Primary adenocarcinoma of bladder. Urology
1991;38:223-226.

100. Lughezzani G, Sun M, Jeldre C, Alasker A, Budaus


L, Shariat S, Latour M, Widmer H, Duclos A, JolivetTremblay M, Montorsi F, Perrotte P and Karakiewicz
P. Adenocarcinoma versus urothelial carcinoma of the
urinary bladder: comparison between pathologic stage at
radical cystectomy and cancer-specific mortality. Urology
2010;75:376-381.

121. Skinner DG and Leiskovsky G. Contemporary cystectomy


with pelvic node dissection compared to preoperative
radiation therapy plus cystectomy in management of
invasive bladder cancer. J Urol 1984;131:1069-1073.
122. El Sobky E, Gomha M, El-Baz M, Abol-Enein H and
Shaaban AA. Prognostic significance of tumour
angiogenesis in schistosoma-associated adenocarcinoma
of the urinary bladder. BJU Int 2002;89:126-132.

101. Khalid H, Zaghloul M, Ghoneim M, Sabr R, Manie M, AbolEnein H, Mageed H, Mansour O, Sherbiny M, Mahran T.
Gemcitabine and cisplatin as a neoadjuvant chemotherapy
for invasive bladder cancer: Effect on bladder preservation
and survival. Proc Am Soc Clin Oncol 22: page 411, 2003
(abstr 1652).

123. Shaaban AA, Elbaz MA and Tribukait B. Primary nonurachal


adenocarcinoma in the bilharzial urinary bladder:
deoxyribonucleic acid flow cytometric and morphological
characterization in 93 cases. Urology 1998;51:469-476.

102. Kramer S, Bredael J, Croker B, Paulson D, Glenn J.


Primary non-urachal adenocarcinoma of the bladder. J
Urol 1979;121:278-281.

124. Grignon DJ, El-Naggar A, Ro JY, Johnson DE and Ayala


AG. Deoxyribonucleic acid flow cytometry on primary
adenocarcinoma of the bladder: an analysis of 36 cases. J
Urol 1989;142:1206-1210.

103. Dandekar NP, Dalal AV, Tongaonkar HB, Kamat MR. Adenocarcinoma of bladder. Eur J Surg Oncol 1997;23:157160.

125. Song J, Farrow GM and Lieber MM. Primary


adenocarcinoma of the bladder: favourable prognostic
significance of deoxyribonucleic acid diploidy measured
by flow cytometry. J Urol 1990;144:1115-1118.

104. Gill HS, Dhillon HK, Woodhouse CRJ. Adenocarcinoma of


the urinary bladder. Br J Urol 1989;64:138-142.

126. Nevin JE, Melnick I, Baggerly. JT, Easley CA, Landes R.


Advanced carcinoma of the bladder: Treatment using
hypogastric artery infusion with 5-fluorouracil, either as a
single agent or in combination with bleomycin or adriamycin
and supervoltage radiation. J Urol 1974;112:752-758.

105. Grignon DJ, Ro JY, Ayala AG, Johnson DE, Ordonez NG.
Primary adenocarcinoma of the urinary bladder: a clinicopathologic analysis of 72 cases. Cancer 1991;67(8):21652172.
106. Chan TY and Epstein. In situ adenocarcinoma of the
bladder. Am J Surg Pathol 2001;25:892-899.

127. Logothetis CJ, Samuels ML, Ogden S. Chemotherapy


for adenocarcinoma of bladder and urachal origin.
5-Fluorouracil, doxorubicin and mitomycin-C. Urology
1985;26:252-255.

107. Allen TD and Henderson BW. Adenocarcinoma of the


bladder. J Urol 1965;93:50-56
108. Mostafi FK. Potentialities of bladder epithelium. J Urol
1954;71:705-714.

128. Hatch RR and Fuchs EF. Intra-arterial infusion of


5-fluorouracil for recurrent adenocarcinoma of bladder.
Urology 1989;33:311-312.

109. Kittredge WE, Coliett AJ, Morgan C. Adenocarcinoma of


the bladder associated with cystitis glandularis. A case
report. J Urol 1964;91:145-150.

129. Awakura Y, Yamamoto M, Fukuzawa S, Nonomura M and


Fukuyama S. Intra-arterial infusion of 5-FU, leucovorin
and cisplatin for primary adenocarcinoma of the urinary
bladder. Int J Urol 1999;6:581-584.

110. Mostofi FK, Thompson RV, Dean AL.


Mucous
adenocarcinoma of the bladder. Cancer 1955;18:599.

130. Kasahara K, Inoue K and Shuin T.


Advanced
adenocarcinoma of the urinary bladder successfully
treated by the combination of cisplatinum, mitomycin-C,
etoposide and tegafur-uracil chemotherapy. Int J Urol
2001;8:133-136.

111. Choi H, Lamb S, Pintar K, Jacobs SC. Primary signet


ring cell carcinoma of the urinary bladder. Cancer
1984;53:1985-1990.
112. Blute ML, Engen DE, Treavis WD, Kvols LK. Primary
signet ring cell adenocarcinoma of the bladder. J Urol
1989;141:17-21.

131. Sheldon CA, Clayman RV, Gonzalez R, Williams RD,


Faley EE. Malignant urachal lesions. J Urol 1984;131:1.

113. Fiter L, Gimeno F, Martin L, Gimeno F, Martin L, Comez


Te Jeda L. Signet-ring adenocarcinoma of the bladder.
Urology 1993;41:30-33.

132. Henly R, Farrow GM and Zincke H. Urachal cancer: role of


conservative surgery. Urology 1993;42:635-639.
133. Siefker-Radktke AO, Gee J, Shen Y, Wen S, Daliani D,
Millikan RE and Pisters LL. Multimodality management
of urachal carcinoma: The M.D. Anderson Cancer Center
experience. J Urol 2003;169:1295-1298.

114. Akamatsu S, Takahashi A, Ito M and Ogura K. Primary


signet-ring cell carcinoma of the urinary bladder. Urology
2010;75:615-618.
115. Malek R, Rosen J and ODea M. Adenocarcinoma of the
urinary bladder. Urology 1983;21:26-29.

134. Gopalan A, Sharp DS, Fine SW, Tickoo SK, Herr


HW, Reuter VE and Olgac S. Urachal carcinoma: a
clinicopathological analysis of 24 cases with outcome
correlation. Am J Surg Pathol 2009;33:659-668.

116. Nocks B, Heney N, Daly J. Primary adenocarcinoma of


urinary bladder. Urology 1983;21:26-29.

430

135. Yu JS, Kim KW, Lee HJ, Lee YJ, Yoon CS and Kim MJ.
Urachal remnant diseases: spectrum of CT and US
findings. Radiographics 2001;21:451-461.

157. Raspollini MR, Nesi G, Baroni G, Girardi LR and Taddei GL.


Immunohistochemistry in the differential diagnosis between
primary and secondary interstitinal adenocarcinoma of
the urinary bladder. Appl Immunohistochem Mol Morphol
2005;13:358-362

136. Lee SH, Kitchens HH and Kim BS. Adenocarcinoma


of the urachus: CT features. J Comput Assist Tomogr
1990;13:232-235.

158. Torenbeek R, Lagendijk JH, Van Diest PJ, Bril H, Van de


Molengraft FJJM and Meijer CJLM. Value of a panel of
antibodies to idetify the primary origin of adenocarcinoma
presenting as bladder carcinoma. Histopathology
1998;32:20-27.

137. Schubert GE, Pavkovic MB and Bethke-Bedurfit BA.


Tubular urachal remnants in adult bladders. J Urol
1982;12:40-42.
138. Ashley RA, Inman BA, Sebo TJ, Leibovich BC, Blute
ML, Kwon ED and Zincke H.
Urachal carcinoma:
clinicopathological features and long-term outcomes of an
aggressive malignancy. Cancer 2006;107:712-720.

159. Nakanishi K, Tominaga S, Kawai T, Torikata C, Aurues T and


Ikeda T. Mucin histochemistry in primary adenocarcinoma
of the urinary bladder (of urachal or vesicular origin) and
metastatic adenocarcinoma originating in the colorectum.
Pathology International 2000;50:297-303.

139. Johnson DE, Hodge GB, Abdul-Karim FW, Ayala AG.


Urachal carcinoma. Urology 1985;26:218.

160. Bardales RH, Pitman MB, Stanley MW, Korourian S and


Suhrland MJ. Urine cytology of primary and secondary
urinary bladder adenocarcinoma. Cancer (cancer
Cytopathol) 1998;84:335-343.

140. Young RH and Eble JN. Unusual forms of carcinoma of


the urinary bladder. Hum Pathol 1991;22:948-965.
141. Begg RC. The urachus, its anatomy, histology and
development J Anat 1930;64:170.

161. Gall FP, Tonak J, Altendorf A. Multivisceral resections in


colorectal cancer. Dis Colon Rectum 1987;30:337-341.

142. Herr HW. Urachal carcinoma: the case for extended


partial cystectomy. J Urol 1994 Feb;151(2):365-366.

162. Lopez MJ and Monafo WM. Role of extended resection


in the initial treatment of locally advanced colorectal
carcinoma. Surgery 1993;113:365-372.

143. Hong S-H, Kim JC and Hwang T-K. Laparoscopic partial


cystectomy with en bloc resection of the urachus for
urachal adenocarcinoma. Int J Urol 2007;14:963-965.

163. Talamonti MS, Shumate CR, Carlson GW, Curley SA.


Locally advanced carcinoma of the colon and rectum
involving the urinary bladder. Surg Gynecol Obstet
993;177:481-487.

144. Wadhwa P, Kolla SB and Hemal AK. Laparoscopic en bloc


partial cystectomy with bilateral pelvic lymphadenectomy
for urachal adenocarcinoma. Urology 2006;67;837-843.
145. Madeb R, Knopf JK, Nicholson C. The use of robotically
assisted surgery for treating urachal anomalies. BJU Int
2006;98:838-842.

164. Carne PWG, Frye JNR, Kennedy-Smith A, Keating J,


Merrie A, Dennett E, Frizelle FA. Local invasion of the
bladder with colorectal cancer: surgical management
and patterns of local recurrence. Dis Colon Rectum
2004;47:44-47.

146. Siefker-Radktke AO. Urachal carcinoma: surgical and


chemotherapeutic options. Expert Rev Anticancer Ther
2006;6:1715-1721.

165. Walther PJ. Adjuvant/neoadjuvant etoposide/cisplatin


and cystectomy for management of invasive small cell
carcinoma. J Urol 2002; 167(4): 285 [Abstract 1124].

147. Kikuno N, Urakami S, Shigeno K, Shina H and Igawa


M. Urachal carcinoma associated with increased
carbohydrate antigen 19-9 and carcinoembryonic antigen.
J Urol 2001;166:604.

166. Blomjour CE, Vos W, De Voogt HJ, Van der Valk P,


Meijer CJ. Small cell carcinoma of the urinary bladder.
A clinicopathologic, morphologic, immunohistochemical,
and ultrastructural study of 18 cases. Cancer 1989; 64(6):
1347-57.

148. Guarnaccia S. Adenocarcinoma of the urachus associated


with elevated levels of Ca125. J Urol 1991;145:140-141.
149. McIntosh JF and Worley G Jr. Adenocarcinoma arising in
exstrophy of the bladder: report of two cases and review
of the literature. J Urol 1955;73:820-829.

167. Holmang S, Borghede G, Johansson SL. Primary small


cell carcinoma of the bladder: a report of 25 cases. J Urol
1995; 153: 1820-1822.

150. Williams DI. Epispadias and exstrophy. In Eckstein HB,


Hohenfellner R, Williams DI eds, Surgical Paediatric
Urology. Stuttgart: Georg Thieme, 1997:298-312.
151. Yap RL, Weiser A, Ozer O, Pazona J, Schaeffer A.
Adenocarcinoma arising from a defunctionalized bladder.
J Urol 2002;167:1782-1783.

168. Trias I, Algaba F, Condom E, Espanol I, Segui J, Orsola


I, et al. Small cell carcinoma of the urinary bladder.
Presentation of 23 cases and review of 134 published
cases. Eur Urol 2001; 39(1):85-90.

152. Witters S and Baert-Van Damme L. Bladder exstrophy


complicated by adenocarcinoma. Eur Urol 1987;13:415416.

169. Koay, EJ, The BS, Paulino AC, Butler EB: A surveillance,
epidemiology, and end results analysis of small cell
carcinoma of the bladder. Cancer 2011, in press.

153. Barrington JW, Fulford S, Griffiths D, Stephenson


TP. Tumors in bladder remnant after augmentation
enterocystoplasy. J Urol 1997;157:482-486.

170. Sved P, Gomez P, Manoharan M, Civantos F, Soloway


MS. Small cell carcinoma of the bladder. BJU Int 2004;
94(1): 12-17. (Level 3)

154. Bates AW and Baithun SI. Secondary neoplasms of the


bladder are histological mimics of nontransitional cell
primary tumours: clinicopathological and histological
features of 282 cases. Histopathology 2000;36:32-40.

171. Cheng L, Pan CX, Yang XJ, Lopez-Beltran A, MacLennan


GT, Lin H, Kuzel TM, Papavero V, Tretiakova M, Nigro K,
Koch MO, Eble JN. Small cell carcinoma of the urinary
bladder: a clinicopathologic analysis of 64 patients.
Cancer. 2004 Sep 1; 101(5): 957-62.

155. Melicow MM. Tumors of the urinary bladder: A


clinicopathological analysis of over 2500 specimens and
biopsies. J Urol 1995;74:498-521.

172. Choong NW, Quevado JF, Kaur JS: Small cell carcinoma of
the urinary bladder. The Mayo Clinic experience. Cancer
2005, 103, 1172-8.

156. Suh N, Yang XJ, Tretiakova MS, Humphrey PA and Wang


HL. Value of CDX2, villin, and -methylacylcoenzyme A
racemase immunostains in the distinction between primary
adenocarcinoma of the bladder and secondary colorectal
adenocarcinoma. Modern Pathology 2005;18:1217-1222

173. Abbas F, Civantos F, Benedetto P, Soloway MS. Small cell


carcinoma of the bladder and prostate. Urology 1995; 46:
617-630.

431

174. Kanat O, Evrensel T, Filiz G, Usta M, Baskan E, Dilek


K, Manavoglu O. Systemic AA amyloidosis and nephritic
syndrome associated with small cell carcinoma of the
bladder. Nephrol Dial Transplant. 2003 Nov; 18(11): 24532454.

190. Helpap B. Nonepithelial neoplasms of the urinary bladder.


Virchows Arch 2001; 439(4): 497-503.

175. Helpap B: Morphology and therapeutic strategies for


neuroendocrine tumors of the genitourinary tract. Cancer
2002,95,7,1415-20.

192. Lopez-Beltran A, Pacelli A, Rothenberg HJ, Wollan


PC, Zincke H, Blute ML, et al. Carcinosarcoma and
sarcomatoid carcinoma of the bladder: clinicopathological
study of 41 cases. J Urol 1998; 159(5): 1497-503.

191. Russo P, Brady MS, Conlon K, Hajdu SI, Fair WR, Herr
HW, et al. Adult urological sarcoma. J Urol 1992; 147(4):
1032-6 [Discussion 1036-7].

176. Grignon DJ. Neoplasms of the urinary bladder. In: Bostwick


DG, Eble JN, editors. Urologic surgical pathology. St. Louis
(MO): Mosby-Year Book, Inc; 1997. p. 216-305.

193. Froehner M, Gaertner HJ, Maseck A, Wirth MP. Durable


complete remission of metastatic sarcomatoid carcinoma
of the bladder with cisplatin and gemcitabine in an 80year-old man. Urology 2001; 58(5): 799.

177. Thompson S, Cioffi-Lavina M, Chapman-Fedricks J,


Gomez-Fernandez C, Fernandez-Castro G, Jorda M:
Distinction of high grade neuroendocrine carcinoma/small
cell carcinoma from conventional urothelial carcinoma of
urinary bladder: an immunohistochemical approach. Appl
Immunohistochem Mol Morphol. 2011, in press.

194. Hwang JJ, Uchio EM, Patel SV, Linehan WM, Walther
MM, Pacak K. Diagnostic localization of malignant bladder
pheochromocytoma using 6-18F fluorodopamine positron
emission tomography. J Urol 2003; 169(1): 274-5.

178. Abrahams NA, Moran C, Reyes AO, Siefker-Radtke


A, Ayala AG: Small cell carcinoma of the bladder: a
contemporary clinicopathological study of 51 cases.
Histopathology 2005, 46,57-63.

195. Klingler HC, Klingler PJ, Martin Jr JK, Smallridge RC,


Smith SL, Hinder RA. Pheochromocytoma. Urology 2001;
57(6): 1025-32.
196. Iczkowski KA, Shanks JH, Gadaleanu V, Cheng L, Jones
EC, Neumann R, et al. Inflammatory pseudotumor and
sarcoma of urinary bladder: differential diagnosis and
outcome in thirty-eight spindle cell neoplasms. Mod Pathol
2001; 14(10): 1043-51.

179. Siefker-Radtke AO, Dinney CP, Abrahams NA: Evidence


supporting preoperative chemotherapy for small cell
carcinoma of the bladder: a retrospective review of the MD
Anderson cancer experience. J Urol 2004, 172,481-4.
180. Grignon DJ, Ro JY, Ayala AG, et al. Small cell carcinoma
of the urinary bladder. A clinicopathologic analysis of 22
cases. Cancer 1992: 69: 527-536.

197. Jones EC, Clement PB, Young RH. Inflammatory


pseudotumor of the urinary bladder. A clinicopathological,
immunohistochemical, ultrastructural, and flow cytometric
study of 13 cases. Am J Surg Pathol 1993; 17(3): 264-74.

181. Oesterling JE, Brendler CB, Burgers JK, Marshall FF,


Epstein JL. Advanced small cell carcinoma of the
bladder. Successful treatment with combined radical
cystoprostatectomy
and
adjuvant
methotrexate,
vinblastine, doxorubicin and cisplatin chemotherapy.
Cancer 1990: 65: 1928-1936.

198. Tainio HM, Kylmala TM, Haapasalo HK. Primary


malignant melanoma of the urinary bladder associated
with widespread metastases. Scand J Urol Nephrol 1999;
33(6): 406-7.
199. De Torres I, Fortuno MA, Raventos A, Tarragona J, Banus
JM, Vidal MT. Primary malignant melanoma of the bladder:
immunohistochemical study of a new case and review of
the literature. J Urol 1995; 154(2 Pt 1): 525-7.

182. Nejat RJ, Purohit R, Goluboff ET, Petrylak D, Rubin MA,


Benson MC. Cure of undifferentiated small cell carcinoma
of the urinary bladder with M-VAC chemotherapy. Urol
Oncol 2001; 6: 53-55.

200. Ito K, Matsuo Y, Takahashi O, Yajima H, Yamanaka H, Ito


H, Synchronous malignant melanoma of the male bulbar
urethra and transitional cell carcinoma of the bladder. BJU
Int 1999; 84(7): 877-8.

183. Ismaili N, Arifi S, Flechon A, et al: Small cell cancer of the


bladder : pathology, diagnosis, treatment, and prognosis.
Bull Cancer 2009, 96,6, E30-44.

201. Al-Maghrabi J, Kamel-reid S, Jewett M, Gospodarowicz M,


Wells W, Banerjee D. Primary low-grade B-cell lymphoma
of mucosa associated lymphoid tissue type arising in the
urinary bladder: report of 4 cases with molecular genetic
analysis. Arch Pathol Lab Med 2001; 125(3): 332-6.

184. Bex A, Niewuwenhuijzen JA, Kerst M, et al: Small cell


carcinoma of the bladder: a single center prospective
study of 25 cases treated in analogy to small cell lung
cancer. Urology 2005,65,2, 295-9.
185. Lohrisch C, Murray N, Pickles T, Sullivan L: Small cell
carcinoma of the bladder: long term outcome with
integrated chemoradiation. Cancer 1999, 86, 2346-52.

202. Kempton CL, Kurtin PJ, Inwards DJ, Wollan P, Bostwick


DG. Malignant lymphoma of the bladder: evidence from
36 cases that low-grade lymphoma of the MALT- type is
the most common primary bladder lymphoma. Am J Surg
Pathol 1997; 21(11): 1324-33.

186. Siefker- Radtke AO, Kamat AM, Grossman HB, et al:


Phase II clinical trail of neoadjuvant alteranating doublet
chemotherapy with fosfamide/doxorubicin and etoposide/
cisplatin in small-cell urothelial cancer. J Clin Oncol 2009,
27, 16, 2592-7.
187. June RR, Dougherty DW, Reese CT, Harpster LE,
Hoffman SL, Drabick J: Significant activity of single agent
vinorelbine against small cell cancer of the bladder as
second line chemotherapy: a case series and review of
the literature. Urol Oncol 2011, in press.
188. Mackey JR, Au HJ, Hugh J, Venner P: Genitourinary small
cell carcinoma. Determination of clinical and therapeutic
factors associated with survival. J Urol 1998, 159, 5,
1624-9.
189. Parekh DJ, Jung C, OConner J, Dutta S, Smith
Jr ER. Leiomyosarcoma in urinary bladder after
cyclophosphamide therapy for retinoblastoma and review
of bladder sarcomas. Urology 2002; 60(1): 164.

432

ICUD Guidelines on Bladder Cancer:


Diagnosis and Evaluation
Ashish M. Kamat
(MD Anderson Cancer Center, Houston, TX)
Paul K. Hegarty
(Guys Hospital , London, UK)

Guidelines
A group of 17 experts from 5 countries
thoroughly assessed 125 references related to diagnosis
and screening of bladder cancer
> 250 page document

No new level 1 or 2 studies since last ICUD


guidelines published
Most reports level 3 or 4, few level 2b

Most recommendations are Grade C


some B and D

Screening Recommendations
Bladder cancer screening is not helpful in
improving survival (Grade C)
If undertaken, screening should be confined to
high risk patients (Grade C).
Screening in high-risk can consist of annual
cytology & dipstick (Grade C).
Further studies of screening are warranted

Hematuria
No correlation between degree of hematuria &
diagnosis of bladder cancer (Grade B).
Microscopic hematuria is variable & intermittent in
bladder cancer

Lack of hematuria on a single urinalysis does not


exclude bladder cancer (Grade C).

Evaluation
Imaging of the upper tracts necessary in
investigation of hematuria (Grade B).
For asymptomatic microscopic hematuria without
risk factors for urothelial carcinoma, urine cytology
or cystoscopy can be used (Grade B).
In patients with irritative voiding symptoms and
hematuria, carcinoma in situ (CIS) must be ruled out
(Grade C).

Cystoscopy and Cytology


Cystoscopy alone is the most cost effective
method to detect recurrence of bladder cancer
(Grade C).
Among urinary markers, urine cytology is the
gold standard for surveillance (Grade B).
Cytopathologists should use uniform
nomenclature (Grade C).

Cystoscopy and Cytology


Bladder wash cytology provides better
diagnostic yield than voided urine cytology
(Grade B).
At cystoscopy minimal manipulation should be
performed prior to bladder wash. Residual urine
mixed with bladder lavage specimen should be
sent for cytology (Grade D).

Cytology
For follow-up urine cytology is most useful for
diagnosis of high-grade tumor recurrence (Grade B).
Cytology valuable for differentiating high-grade
from low- grade urothelial carcinomas (Grade B).
No consensus on other urinary markers in the
diagnosis of bladder cancer (Grade D).
Markers eg FISH may be most useful in the
setting of a negative cystoscopy and atypical
cytology. (Grade C).

Endoscopic Techniques
White light cystoscopy (WLC) is the gold standard
for evaluation of the lower urinary tract (Grade B).
Fluorescence cystoscopy
improves the rate of detection of CIS (Grade B).
decreases risk of residual tumor after TURBT (Grade B).
useful for positive cytology but negative findings on WLC
(Grade C).

Imaging
Imaging of the upper tracts necessary in
investigation of haematuria (Grade B).
CT urography most useful
intravenous urography, regular CT, ultrasound & MRI
are options (Grade C).

Imaging for staging should be obtained prior to


TURBT
or >2 weeks after TURBT to avoid artifacts (Grade C).

Imaging
CT & MRI not accurate for staging of primary
bladder tumors but may be demonstrate metastatic
disease (Grade B).
With invasive bladder tumors, metastatic work-up
includes chest radiography, liver function tests, and
alkaline phosphatase measurement (Grade C).
Bone scan for patients with bone pain or elevated
alkaline phosphatase concentration (Grade B)

Technique
Insufficient information to support specific
energy modality for TURBT (Grade D).
Patients undergoing TURBT should be given
appropriate prophylactic antibiotics (Grade B).

Technique of TURBT
Document shape, size & location of the tumor
(Grade C).
Document appearance base of the tumor (Grade C).
these details provide important prognostic information

Obtain separate tumor base & margin biopsies for


larger tumors (Grade C).
Attempt complete tumor resection except in
patients with diffuse CIS (Grade C).

Technique
If ureteral orifice is resected,
cutting current should be used.
Functional study should be performed 3-6 weeks
later (Grade C).

Diverticulum
Aggressive resection risks perforation.
Low-grade, noninvasive tumors may be treated TUR
or fulguration with or without intravesical therapy
(Grade C).

Technique
A second TURBT should be performed in all
patients with a high-grade T1 lesion or select HG
Ta lesion (Grade B).
The optimal timing of repeat TURBT is 4-6 weeks
after the first resection (Grade C).

Technique: Random Biopsies


Not routinely recommended (Grade C).
Indicated in
Patients with positive findings on urine cytology &
normal cystoscopy (Grade B).
Candidates for partial cystectomy (Grade C).

Prostatic Urethral Biopsy


Indicated in cases of multifocal urothelial
carcinoma of the bladder, CIS & visible
abnormalities of the prostatic urothelium (Grade
B).
Prostate biopsy not useful in counselling patients
for neobladder, but useful in indentifying cT4
disease (Grade C).

Contributors
AM Kamat (USA)
PK Hegarty (UK)
JR Gee (USA)
S Rawal (India)
Clark (USA)
JD Davies (USA)
SS Chang (USA)
A Patel (UK)
M Manoharan (USA)

TS OBrien (UK)
J Fan (China),
RS Svatek (USA)
M Arianayagam (Australia)
P Black (Canada)
NJ Hegarty (Ireland)
R Sanchez-Salas (France)

INTERNATIONAL CONSORTIUM FOR UROLOGIC


DISEASES EUROPEAN ASSOCIATION OF UROLOGY
INTERNATIONAL CONSULTATION ON BLADDER
CANCER 2011:
CONSENSUS GUIDELINES BY THE PATHOLOGY OF
BLADDER CANCER WORK GROUP
Mahul B. Amin, MD, Victor E. Reuter, MD
And
Members of the Pathology of Bladder Cancer Work Group
Vienna Austria, March 18, 2011

35 Urologic Pathologists from 13 Countries

Participants
Hikmat A. Al-Ahmadie, MD (New York, NY, USA)
Ferran Algaba, MD (Barcelona, Spain)
Isabel Alvardo-Caberrera, MD (Mexico City, Mexico)
Liang Cheng, MD (Indiana, IN, USA)
John C. Cheville, MD (Rochester, MN, USA)
Glenn Kristiansen, MD (Zurich, Switzerland)
Richard J. Cote, MD (Miami, FL, USA)
Brett Delahunt (Wellington, New Zealand)
John N. Eble, MD (Indianapolis, IN, USA)
Mahmoud Elbaz, MD (Cairo, Egypt)
Elizabeth Genega, MD (Boston, MA, USA)
Christian Gulmann, MD (Dublin, Ireland)
Arndt Hartmann, MD (Erlangen, Germany)

Participants
Cord Langer, MD (Graz, Austria)
Antonio Lopez-Beltran, MD (Cordoba, Spain)
Cristina Magi-Galluzzi, MD, PhD (Cleveland, OH, USA)
Jorda Merce, MD, PhD (Miami, FL, USA
George J. Netto (Baltimore, MD, USA)
Esther Oliva, MD (Boston, MA, USA)
Priya Rao, MD (Houston, TX, USA)
Jae Y. Ro, MD, PhD (Seoul, S. Korea)
John R. Srigley, MD (Toronto, ON, Canada)
Satish K. Tickoo, MD (New York, NY, USA)
Toyonori Tsuzuk, MD, PhD (Nagoya, Japan)
Saleem A. Umar, MD (Miami, FL, USA)
Theo Van der Kwast, MD (Toronto, ON, Canada)
Robert H. Young, MD (Boston, MA, USA)

Whats New in this Consensus Offering?:


An Update on:

The knowledge of the histoanatomy of the bladder


as it pertains to the diagnosis and staging of
bladder cancer
The prognostic significance of histologic grading
by the WHO (2004)/ISUP system, its contributions,
shortfalls and opportunities for refinement
The information needed from the clinicians by the
pathologist in order to provide optimal pathologic
reporting of bladder cancer
On nomenclature for inverted lesions of the
bladder

Whats New in this Consensus Offering?:


An Update on:

The histologic types of bladder cancer and the


variants of urothelial cancer, some relatively
recently described, with a focus on definition for
diagnosis and their implied clinico-pathologic
significance
Role of immunohistochemistry and molecular
studies in contemporary routine practice diagnosis,
prognostication or predicition of bladder cancer
The reporting of bladder cancer in transurethral
resection of bladder tumors (TURBT) and
cystectomy specimens

Urothelial denudation

Seen with
instrumentation with
CIS
Extensive or complete
denudation in a bladder
biopsy should be
reported as correlation
with concurrent
cytology results may
yield a positive
diagnosis of
malignancy in the
patient

Urothelial hyperplasia

Papillary or flat
Thickened urothelium
without cytologic atypia
Flat hyperplasia may be
seen adjacent to low
grade papillary tumors
Papillary hyperplasia is
suggested to be a
precursor lesion for a
subset of papillary
urothelial neoplasms

Squamous Metaplasia

Should be reported and


specified whether
keratinizing or not and
whether focal or
multifocal
Current evidence does
not support any
precursor potential with
focal changes, although
multifocal and/or
extensive lesions
precede or may be
associated concurrent
with neoplasia

Glandular
Metaplasia

Cystitis cystica &


cystitis cystica
glandularis are
proliferative lesions

They may be associated


with intestinal metaplasia

Glandular metaplasia
is not a risk factor for
adenocarcinoma or
UCa

Multifocal disease
Metaplasia with dysplastic
changes
- Associated with risk for
adenocarcinoma

pT1 invasion into


the lamina propria
pT1
pT2

pT3

67-75% 5-year survival

pT2 invasion into


the m. propria
60-63% 5-year survival

pT3 invasion into


the perivesical fat
31-50% 5-year survival

pT4 invasion of
adjacent organs
10-25% 5 year survival

Muscularis Mucosae Muscle

Hyperplastic Muscularis Mucosae:

Patterns frequently observed but exceed thickness of


previously described MM

To avoid confusion in diagnosis, documentation


of MM-only invasion by carcinoma is not
recommended in the main pathologic diagnosis,
and should be reported as urothelial carcinoma
with lamina propria invasion (at least pT1)
Involvement of MM may be included in a comment
to provide information on depth/extent of invasion

M. mucosae muscle patterns

Typical

Variant

Wispy or thin muscle bundles

Hypertrophic MM muscle

Muscularis Propria

Several terms are currently being used to describe


both MP (deep muscle, muscle proper or detrusor
muscle) and MM (superficial muscle). These terms
are not recommended and use of standardized
nomenclature of MP muscle bundles is
recommended
It is important to document the presence or
absence of MP in a biopsy irrespective of
involvement
Some institutions recommend not mentioning the
presence or absence of MP in low grade tumors
and require documentation in high grade tumors
only

Muscularis Propria
In cystectomy specimens, a
more objective and
reproducible anatomical or
extent of disease criterion
is needed between pT2a
and pT2b subcategories
Substaging of pT2 disease
and recognition of pT3
disease is not tenable in
TURBT specimens

Documenting Amount of
Muscularis Propria in Specimen

Formal studies are lacking

Amount of MP included in the specimen (isolated and rare versus


established groups)
MP in the area of the tumor or not in the vicinity

Information may be useful in the management of


invasive cancer
Inclusion of the information is optional but may be
communicated in a multidisciplinary setting

Muscle Involved by
UCa, Indeterminate
Type

Muscle bundles
indeterminate between
MM and MP should be
reported with terminology
such as "invasive
urothelial carcinoma with
invasion of muscle,
indeterminate type" to
prompt the urologist for a
restaging biopsy
procedure

Adipose Tissue

Adipose tissue is frequently present in the LP and


MP of the urinary bladder, usually scant in the
former and abundant in the latter
Involvement of adipose tissue by tumor in biopsy or
TURBT specimens should not be automatically
interpreted as pT3 disease
There is limited reliability of pT3a vs pT3b
subcategorization as gross involvement of
perivesical fat (extravesical fat) may not always be
readily recognizable and may be mimicked by
reactive changes

Assessment of pT2 vs. pT3 cystectomy alone

Lamina Propria Invasion

There are problems with interobserver


reproducibility in the diagnosis of early
LP invasion
When early invasive urothelial carcinoma
is suspected, diagnosis through
examinational of additional levels, or in a
consensus manner either with other
pathology colleagues or thru quality
assurance meeting is encouraged

Microinvasive Urothelial Carcinoma

To diagnose early invasion, stringent criteria such


as: only focal invasion, less than 1 high power
field, or 0.5 mm from the nearest basement
membrane, should ideally be employed
Studies are needed to establish a clinically
significant definition of microinvasive urothelial
carcinoma. Until there is understanding of the
definition, it is recommended that only stringent
criteria be used or the term microinvasive
carcinoma not be used

Substratification or Substaging of
Lamina Propria Invasion (pT1)

Depth of invasion may be established either:

Up to MM: pT1a or beyond the MM or vascular plexus: pT1b;


Up to MM: pT1a, in to MM: pT1b, or beyond the MM or vascular
plexus: pT1c

Currently substaging is not recommended due to the lack


of widely accepted and reproducible criteria although there
is much need for such studies
It is recommended for the pathologist to provide an
estimate of the LP invasion in pT1 tumors with respect to
depth and/or quantity of invasion (e.g. focal, multifocal,
extensive, etc.)

Minimal (micro) invasion

Extensive invasion

Grading of Urothelial Carcinoma

The non-invasive urothelial lesions and neoplasms


can be flat, papillary (exophytic) and inverted
(endophytic). All 3 growth patterns may be seen in
a single tumor
The World Health Organization (WHO) (2004) /
International Society of Urologic Pathologists
(ISUP) classification system is the recommended
system. It has also been endorsed by the WHO
2004 Blue Book, the 4th Series Armed Forces
Institutes of Pathology Fascicle on Bladder and the
7th edition AJCC Cancer Staging Manual

WHO (2004)/ISUP Grading System


Flat Lesions:
Normal
Hyperplasia
Reactive
Dysplasia
CIS
Atypia of
unknown
significance

Papillary Lesions:
Non invasive

Papilloma
PUNLMP
Low grade
High grade

Invasive

Inverted Lesions:
Non invasive

High grade
Low grade (rare)

Papilloma
PUNLMP
Low grade
High grade

Invasive

High grade

WHO (2004)/ISUP Classification


The diagnosis of
papillary urothelial
neoplasia is made
on the basis of the
presence of a
fibrovascular core
Grading is
performed
Based primarily on
cytology
& some architectural
features

N
O
R
M
A
L

D
Y
S
P
L
A
S
I
A

H
Y
P
E
R
P
L
A
S
I
A

C
A

I
N
S
I
T
U

P
A
P
I
L
O
M
A

L
O
W
G
R
A
D
E

L
M
P

H
I
G
H

G
R
A
D
E

Inverted Neoplasms

Papillary tumors may be associated with a variable


degree of inverted growth patterns; although focal
areas of inverted growth are not uncommon,
prominent or exclusive inverted growth is much
rarer and when encountered may pose problems
related to grading or assessment of invasion
From the clinical standpoint during cystoscopy,
inverted tumors made be polypoid, dome-shaped
or frequently raise the suspicion of invasive
urothelial carcinoma

Courtesy R. Montironi, Italy

Exophytic tumor

Inverted tumor

I
N
V
P
A
P
I
L
O
M
A

I
N
V
L
M
P

I
N
V

I
N
V

L
G

H
G

Inverted High Grade without


invasion

Inverted High Grade with


invasion

Urothelial Dysplasia

Overall features are those of a neoplastic


atypia but which fall short of the criteria for
CIS; dysplasia is not further graded
There is some evidence, largely genetic, that
dysplasia shares some abnormalities with CIS
and therefore likely represents a precursor
lesion
Few studies, most dated, indicate a 5-19% risk
of developing cancer

Urothelial Dysplasia

The diagnosis of de novo dysplasia (i.e. in a


patient without history of bladder neoplasia)
should not be made or should be made with
great caution as the vast majority of patients,
in the limited studies, do not progress to
cancer
While dysplasia likely represents a marker of
urothelial genetic instability, the diagnosis
should not by itself invoke any therapy;
continued surveillance is recommended

UROTHELIAL DYSPLASIA

Urothelial CIS

Biologically, therapeutically and prognostically


significant flat lesion
There is a spectrum of nuclear and architectural
atypia. CIS defined by WHO (2004) / ISUP
includes cases diagnosed previously as severe
dysplasia and even some cases previously
diagnosed as moderate dysplasia
Development of invasion is seen in 20 to 30% of
the cases

WHO (2004) /ISUP : Prognostic Significance


Papillom PUNLMP
a

LG pap
ca

HG
pap ca

34-78
(50)
7-12

34-74

Recurrence

9-18

Grade
progression

2%

17-62
(34)
11

Stage
progression

0-4

3-18

27-61

100

93-100

82-96

65-90

Survival

Does not predict risk at individual patient level


Need for incorporation with other clinical parameters in
nomograms to provide personalize risk stratification

Contributions of WHO (2004) /ISUP

Establishment of uniform terminology and common


definitions for papillary neoplasms
Establishment of detailed criteria of various preneoplastic
conditions and various grades of tumor
Correlation with urine cytology terminology, facilitating
cyto-histologic correlation and making it easier for
urologists to manage patients
Creation of a category of tumor that identifies a tumor with a
negligible risk of progression (PUNLMP), whereby patients
avoid the label of having cancer which has psychosocial
and financial implications. Neither is a benign lesion
(papilloma) diagnosed in these patients, so they may still be
followed up closely

Contributions of WHO (2004) /ISUP

Identification of a distinct group of patients


(high-grade papillary UCa) who would be
ideal candidates for intravesical therapy
Removal of ambiguity in diagnostic
categories in WHO 1973 system (e.g., TCC
grade 1-2, TCC grade 2-3)
Stratification of bladder tumors into
prognostically significant categories
Emergence of molecular correlates for high
grade tumors and tumors at the low-grade
end of the spectrum which may help provide
ancillary grading tools, and possibly guide in
future refinements of the current grading
system

Grading Papillary Urothelial Neoplasm


with Histologic Heterogeneity

Grade heterogeneity is not uncommonly


encountered in papillary urothelial neoplasia.
There are studies showing that pure HG
papillary U Ca has a higher disease
progression rate than tumors with mixed highgrade and low-grade areas.
The WHO (2004)/ISUP system recommends
grading of heterogeneous tumors to be based
on the highest grade present in a tumor

Grading Papillary Urothelial Neoplasm


with Histologic Heterogeneity

There is no current widely acceptable


definition or criteria to provide quantitative
estimate of size of smallest focus required
to upgrade a lesion.
Studies are needed to establish
quantitative/semi-quantitative criteria that
need to be present to alter assignment of
grade in tumors with grade heterogeneity

Grading Papillary Urothelial Neoplasm


with Histologic Heterogeneity

The distinction between PUMLMP and low grade papillary U


Ca is not that critical in an individual patient
Distinction of low grade from high grade carcinoma is more
important
When assigning the tumor grade in tumors with borderline
grade histology, other tumor parameters such as
multifocality, previous grade of the tumor, size of lesions,
frequency of recurrence, presence of concurrent CIS,
positive cytology may be factored in the grading
Currently, no reliable or acceptable IHC or molecular
markers to assist in grading

Papillary Hyperplasia with Cytologic


Atypia

Terminology used when criteria of papillary


neoplasia are not fulfilled but there is a
background of cytologic atypia

Dysplasia with early papillary formations


CIS with early papillary formations

Usually occurs in the backdrop of treatment


setting
These terms are only descriptive diagnoses and
outcome studies are not available

Epstein, Reuter, Amin Bladder Biopsy


Interpretation,
2nd edition

CIS with early


papillary features

Dysplasia with early


papillary features

Grading Invasive Cancer

Practice in the United States

Almost all cases, once diagnosed as

invasive, are called high grade


Exceptionally rare cases are called
low grade including variants such as
nested variants
Most studies show that once invasive,
depth of invasion is more important

Grading Invasive Cancer

Practice in the Europe

Further grading of high grade (WHO 2004/ISUP


system)
G2 and G3 categories (of WHO 1973 system)
Incorporated into algorithms such as the EORTC
system
Criteria for distinction of G2 versus G3 not well
defined but studies show prognostic significance
in at least pT1 tumors

Conventional Urothelial Ca

Histologic Variants of Bladder Cancer


Prognosis:
Aggressive

Micropapillary
Small cell
Sarcomatoid
With rhabdoid
features
Signet ring
adenocarcinoma
Giant cell carcinoma

Favorable

Pure LELC
Verrucous Ca
Carcinoid tumor

Therapy:
Sarcomatoid Ca
Small cell
Large cell
neuroendocrine
Lymphoepitheliomalike
Micropapillary

Diagnosis
Nested variant
UCa with small
tubules
Plasmacytoid UCa
UCa with clear cell
features

Histologic Variants of Bladder Cancer

Bladder carcinoma can demonstrate many different


morphologies; major categories include urothelial
carcinoma, squamous cell carcinoma,
adenocarcinoma and small cell carcinoma
All forms of bladder carcinoma originate from
changes in the overlying urothelium
Urothelial carcinoma is the most common category
of bladder carcinoma, comprising approximately
95% of all bladder carcinomas
Urothelial carcinoma may be subdivided into
conventional and variant subtypes

Histologic Variants of Bladder Cancer

Variant forms include the 15 subtypes; it remains unclear


what percentage of variant form impacts survival in most
cases
Most variants often admixed with conventional UCa,
squamous or glandular differentiation
Level I evidence is limited for most bladder carcinoma
variants, given their relative rarity. Prospective or
consecutive cases studies in large series evaluating the
impact of various histologic variants on outcome or therapy
have not yet been performed

Histologic Variants of Bladder Cancer


The percentage of variant component required to
make a diagnosis of a particular variant is not defined
When mixed with conventional or other variants of U
Ca, the relative percentages of the different histologies
must be specified
e.g. U Ca with micropapillary histology (40%),
conventional U Ca histology (50%) and with
squamous differentiation (10%)

Papillary tumor

Micropapillary U Ca

Micropapillary UCa

Sarcomatoid UCa

Small cell Ca

Nested UCa

Microcystic UCa

Plasmacytoid UCa

Lipoid-rich UCa

UCa with Squamous and Glandular


Differentiation
Controversial and limited data
Some studies have shown a worse prognosis when
compared with pure urothelial cancer
Others show no differences between stage matched
cohorts
The presence of divergent differentiation in a TURBT
may predict locally advanced cancer at cystectomy

Squamous Cell Carcinoma

Invasive squamous cell carcinoma


Schistosomal
Non-schistosomal

Verrucous carcinoma

Should be pure in morphology


If associated with destructive invasion,
should be classified as squamous cell
carcinoma invasive with verrucous
features

Molecular Subtyping of UCa

There are promising emerging tests which have the


potential to impact prognostication and prediction
to therapy in U Ca
Prospective studies determining their efficacy and
best practices for performing the tests are
underway (e.g. the UROMOL study in Europe
investigating and validating gene expression
profile)
Further studies need to be defined for professional
and technical quality assurance and optimal
patient care

IHC Markers
in Staging UCa
Smoothelin
- Contractile protein
- Fully differentiated muscle
cells
- Hyperplastic muscularis
mucosae (-) to weak
- Diffuse positivity in muscularis
propria

Caution: Susceptible to variation in antibody titer


Must be interpreted in the context of morphology

Reporting of Bladder Cancer

Guidelines : CAP and European Society of


Uropathology etc.
Reporting in synoptic pattern is preferred
Clinical information important for pathology
diagnosis

Cystoscopic impression
Previous history of bladder cancer
Previous history of therapy (intravesical or systemic)
History of stones, indwelling, catheter, etc.

A deeper bite for assessment of MP involvement in


TURBTs should be preferably submitted separately,
for larger tumors

Reporting at Cystectomy

No residual cancer in cystectomy:


? pT0 or pT2 (if TURBT showed muscle

invasive disease)
By AJCC 6th edition: pT2
Studies in literature have shown pT0 (no
tumor at cystectomy) have better prognosis
than pT2 tumor at cystectomy
Need conference multidisciplinary input to
resolve reporting tumor (-) cystectomies

Reporting of Bladder Cancer

Minimum number of lymph nodes in cystectomy


not established
Size of largest metastatic focus and extranodal
extension, if present should be documented
Frozen section is not an optimal method for a
primary diagnosis of invasive urothelial carcinoma
or to perform pathologic staging prior to a
cystectomy

Chapter 3: Biological Markers


Shahrokh F. Shariat
Bernd J. Schmitz-Drger
Michael Droller
Vinata B. Lokeshwar
Yair Lotan
MLiss A. Hudson
Bas van Rhijn
Michael Marberger
Yves Fradet
George P. Hemstreet
Per-Uno Malmstrm
Osamu Ogawa
Pierre Karakiewicz

Weill Cornell Medical College, New York, NY, USA


EuromedClinic, Frth, Germany
Mount Sinai School of Medicine, New York, NY, USA
University of Miami School of Medicine, Miami, FL, USA
UT Southwestern Medical Center, Dallas, TX, USA
Washington University, St. Louis, Missouri, USA
Netherlands Cancer Institute, Amsterdam, Netherlands
Medical University of Vienna, Vienna, Austria
Laval University, Laval, Quebec, Canada
University of Nebraska, Omaha, Nebraska, USA
University Hospital, Uppsala, Sweden
Kyoto University, Kyoto, Japan
University of Montreal, Montreal, Quebec, Canada

Disclosures
Patents
Method to determine prognosis after therapy for prostate
cancer
Granted 2002-09-06
Methods to determine prognosis after therapy for bladder
cancer
Granted 2003-06-19
Prognostic methods for patients with prostatic disease
Granted 2004-08-05
Soluble Fas urinary marker for the detection of bladder
transitional cell carcinoma
Granted 2010-07-20

Definition of Biological Marker


NIH definition: a characteristic that is objectively measured and
evaluated as an indicator of normal biological processes, pathogenic
processes, or pharmacologic responses to a therapeutic intervention
For our purposes: a biomarker is the end result of a bioassay,
a laboratory technique for processing biological material from humans,
expressed quantitatively or categorically
Imaging results could also be included as a biomarker if we interpret
bioassay broadly enough
Measurement outcome of a biomarker:
binary (e.g., positive or negative)
categorical (e.g., high, medium, low)
quantitative (e.g., the level of some urine/serum protein), or
more complicated (e.g., a genomic signature or metagene)

Clinical Uses of Biomarkers


Clinical practice
Determine the risk of developing disease
Early detection/screening
Establish diagnosis

Diagnostic
Markers

Determine prognosis
Predict response to therapy
Therapeutic target

Monitor disease/treatment response


Clinical Trials
Stratifying study populations
Conducting interim analysis of efficacy/safety
Applied toward regulatory approval

Prognostic
Markers

Biomarker Conundrum for Bladder Cancer

Number of published articles

PubMed Search on bladder cancer AND (biomarker OR biomarkers OR


molecular marker OR molecular markers) yields 4125 hits (3/17/2011)

Number of biomarkers routinely used by urologists: cytology


Why havent biomarkers lived up to their promise?

A variety of issues and


barriers can affect the
movement of clinical tests

from research to clinical


practice.

Shariat et al., Eur Urol 2007

Analytical and regulatory barriers to the

application of biomarkers
Status of intellectual property protection
Availability of standard reference materials for the assay
Complexity of assay format
Implementation of quality control to assure reproducibility and accuracy
Sufficient market testing size to assess methods of commercialization
Lack of clear guidelines for good manufacturing/laboratory practice and
quality control requirements for all phases of biomarker development

Cost and effort required to accumulate clinical data under appropriately


designed, Institutional Review Board-approved prospective trials
The interval required for resolution of patent issues, assay
standardization, validation, testing, and regulatory approval
Shariat et al., Eur Urol 2007

Why havent biomarker studies led to translation?

Biomarker research done in context of usual clinical care, not


clinical trials
Biomarker assays not standardized
Most biomarkers findings not reproducible
Biomarkers that appear biomedically and statistically
significant at one center - NOT significant in other centers
Lack of minimal set of CDEs, use cases, metadata
Pre-analytical variation not controlled or annotated

Usually small, single institution studies

Bringing the ideal marker to clinical practice

Easier: performed easily and promptly in a


clinical environment

Better: does the biomarker add to established


predictors?
Faster: available in an efficient and timely
manner
Cheaper: cost-effectiveness

Shariat et al., Eur Urol 2008

Modification of the structured phase-approach to the


systematic discovery, evaluation, and validation of biomarkers
PHASE

GOALS/AIMS

EXPERIMENTATION

SAMPLE DETAILS

Preclinical

Exploratory; nominate

Preclinical study

Possible bias: small size and

Testing

and rank candidate

for hypothesis generation

convenience sampling

biomarker profiles
0

Develop an assay with

Reproducibility and robustness

clinically reproducible results

of assay; No assessment of
benefit

II

Test on small sample to

Marker optimization, establish

Sample population assay

determine benefit

prediction rules, determine cut-

developed from candidate

offs

biomarker profile

Retrospective design

Sample population should be

Determine operating
characteristics & internal

the target population

validation
III

External validation

Retrospective or prospective,
Generaliziability, Impact on

clinical decision-making
IV

Assess whether biomarker


reduces the burden of disease

Post-approval reporting and


testing for other disease
processes or disease stages

Multi-institutional, large study

Reporting

recommendations
for tumor marker
prognostic studies

(REMARK)
NCI-EORTC

McShane et al.,2005

Levels of evidence for grading clinical utility of biomarkers: Tumor

Marker Utility Grading System (TMUGS)

Hayes et al. JNCI 1996

Predictive modeling - Personalized medicine


The only useful function of a statistician is to make predictions,
and thus to provide a basis for action.

W.E. Deming
Modeling strategies
Logistic regression
Cox proportional hazards regression (caveat: competing risks
lead to overestimation of the probability of an event)
Recursive partitioning and regression trees
Artificial neural networks
Shariat et al., Cancer 2008

To determine the value of a new biomarker

It is NOT sufficient to show that it is:


1. significantly related to the outcome
2. statistically significant in a multivariate model including the
standard clinical and pathologic factors
3. more significant than the standard clinical and pathological
factors in a multivariate model
A variable that is statistically significant in a multivariate model
might not improve the models predictive accuracy
- Odds ratio and hazard ratio do not meaningfully describe a
biomarkers ability to classify patients
- Neither the p-value

Gain in predictive accuracy

It is necessary to show that adding a biomarker to an


existing model based on the most important clinical and
pathologic factors improves the predictive accuracy of
the model.
Does the biomarker add predictive/prognostic information to the
current risk group classification?

Does the classifier improve the predictive inaccuracy and proportion


of explained variation?
Does the classifier improve the Area Under the ROC Curve (AUC)
or C-index (Concordance Index)?

Clinical validation of biomarker-based model

A biomarker is optimal for the patients in the training set


used to develop it (overly optimistic results)
Therefore biomarker-based models need to be validated
on data not used to develop it:
- Internal validation on original dataset
- External validation on independent dataset (best!!!)

Statistics in marker research


Attention to sound statistical practice
- should be delivered as early as possible
- will help maximize the promise of biomarkers for patient care
Studies should include a measure of predictive accuracy
External validation using data from a completely different study
provides the highest irrefutable evidence that a model is valid
Other issues such as over-reliance on statistical significance
tests, over-fitting of models, small sample sizes, and lack of prespecification of the validation process continue to be problematic
Statistical methodology is an area of research

Shariat et al., Urol Oncology, 2010

Clinical States Model of Bladder Cancer


Death From Other
Causes

Initial Bladder
Evaluation: No
Cancer Diagnosis

Non-invasive
UCB

Invasive
UCB

Metastatic
UCB

Death From UCB

Shariat et al., Eur Urol 2003

Urinary Markers for Screening and Monitoring


Death From Other
Causes

Initial Bladder
Evaluation: No
Cancer Diagnosis

Non-invasive
UCB

Death From UCB

Potential indications for marker use


Screening (Voiding symptoms, hematuria, risk populations e.g.,
occupational exposure/life style)
Reflex testing
Follow-up of patients with bladder cancer

Shariat et al., Eur Urol 2003

Why are new markers attractive?


Cystoscopy gold standard for diagnosis and
surveillance
Invasive, expensive, time consuming

Up to 10% of significant lesions still missed by


cystoscopy

NCCN UCB Surveillance Protocol


Cystoscopy
0-2 years Every 3 months
2-4 years Every 6 months
5 years + Annually
Cytology
Obtained at time of cystocopy
Butonly 40% complete recommended surveillance

protocol

Cytology
Cells readily available in urine

Relies on abnormal cellular morphology


Reasonable specificity (87-95%)
Poor sensitivity (12-48%)
Higher sensitivity for high grade disease
Useless: infection during intravesical Rx

Need for highly trained cytopathologist


Interpreter dependent

Accuracy affected by:

Low inter- and intra-observer

Collection technique
Pelvic radiation, intravesical
therapy, instrumentation
Cytopathologist

reproducibility

Variability of urinary cytology is very high


Positive in 38% to 65% of recurrent UCB
Sensitivity for grade 3:

33% to 95%

Sensitivity for stage T2:

37% to 100%

Accuracy from poor (63%) to excellent (89%)

even after adjusting for age and gender


regardless of tumor stage and grade

Gold Standard?
Sensitivity of Cytology
(Review of Literature)
Grade 1
37%

Grade 2
75%

Grade 3
94%

After 1990

11%

31%

60%

All studies

21%

53%

78%

Before 1990

Halling et al JUrol 2002

Ideal Tumor Marker


Potential to replace, delay, or complement cystoscopy and/or
cytology
Characteristics
- Non-invasive - Rapid - Objective - Reproducible - Cost-effective

- Easy to perform and interpret - High sensitivity and specificity


For monitoring, two different paradigms:
- Low risk tumors: reduction of diagnostic cystoscopies

- High risk tumors: recognition of tumor recurrence as early as


possible and prevention of progression

Commercially available bladder tumor markers


Test/
marker
Cytology

Marker detected/
Marker type
Tumor cells

Hematuria
detection
BTA-Stat

A: Hemoglobin
B: Red blood cells
Complement factor Hrelated protein (and also
Complement factor H)
Complement factor Hrelated protein (and also
Complement factor H)
Nuclear mitotic apparatus
protein
Nuclear mitotic apparatus
protein
Nuclear matrix protein

BTA-TRAK

NMP-22

NMP-22
BLCA-4
Survivin

Specimen

Assay type

FDA approval

Voided urine
Barbotage specimen
A: Voided urine
B: Voided urine
Voided urine

Microscopy

n.a.

Voided urine

Sandwich ELISA

Diagnosis, follow- Bard/Bion


up
Diagnostics

Voided urine

Sandwich ELISA

follow-up

Matritech, Inc.

Voided urine

Point-of-care device

Matritech, Inc.

Voided urine

ELISA (using a rabbit


polyclonal antibody)
Bio-dot test

Diagnosis high
risk, follow-up
-

A member of inhibitors of
apoptosis gene family
UBC
Cytokeratin 8 and 18
(cytoskeletal proteins)
CYFRA 21-1 Cytokeratin 19 (a
cytoskeletal protein)
DD23
185-kDa tumor
associated Antigen

Voided urine

uCyt+

UroVysion

Carcinoembryonic
antigen, 2 bladder tumor
cell-associated mucins
Alt in chromosomes 3, 7,
17 and 9p21

Voided urine

A: Dipstick, B: Interference- microscopy /RBC analyzer


Dipstick immunoassay
Diagnosis, followup

Manufacturer

A: Bayer
Corp. B: Bard/Bion
Diagnostics

Eichrom
Technologies
Fujirebio
Diagnostics
IDL Biotech.

Sandwich ELISA or a point


of care test
Immunoradiometric assay
or ELISA
Immunocytochemistry

Bio Int; Roche


Diag
UroCor Labs

Voided urine,
Exfoliated cells

Immunocytochemistry

Follow-up

Scimedx, Inc.

Voided urine,
Exfoliated cells

Multi-colored, multi-probe
FISH

Diagnosis, follow- Abbott, Vysis


up

Voided urine
Exfoliated cells

Urine Biomarkers for Detection and Surveillance


Marker

# Studies

# Pts

Sensitivity

Range

Specificity

Range

BTA
BTAstat
BTAtrak
NMP22
FDP

5
17
4
15
4

436
1377
360
838
168

48
58
71
71
54

32-58
29-74
60-83
47-100
47-68

92
73
66
73
61

91-92
56-86
60-79
55-98
25-80

ImmunoCyt
Cytometry
Quanticyt
Hb-dipstick
Lewis X

6
5
5
2
3

276
364
129
117
95

67
60
58
40
75

52-100
45-85
45-65
37-41
68-79

75
82
76
87
85

62-82
50-92
68-87
87
67-86

FISH
Telomerase
Microsatellite
CYFRA21-1

4
3
6
3

165
146
108
156

79
39
82
85

70-86
29-66
75-92
75-88

70
Na
89
82

66-93
Na
79-100
73-95

Cytology

26

2213

35

13-75

94

85-100

Shariat et al., Minerva Urol, 2009

Marker Sensitivity and Specificity: Meta-analyses


Marker

Cytology
Cytology
Cytology
Cytology

Median Sensitivity Median Specificity


(range)
(range)
55 (48-62)*
94 (90-96)*
34 (20-53)
99 (83-99)
35 (13-75)
94 (85-100)
44 (38-51)*
96 (94-98)*

Total Number
of Patients
3,444
2,767
5,545
14,260

BTA stat
BTA stat
BTA stat

70 (66-74)*
71 (57-82)
58 (29-74)

75 (64-84)*
73 (61-82)
73 (56-86)

1,160
2,534
3,461

NMP22
NMP22
NMP22
NMP22
NMP22 BladderChek

67 (60-73)*
73 (47-87)
71 (47-100)
68 (62-74)*
65 (50-85)

78 (72-83)*
80 (58-91)
73 (55-98)
79 (74-84)*
81 (40-87)

2,290
2,413
2,041
10,119
2,426

Immunocyt
Immunocyt
This assessment

67 (52-100)
84 (77-91)*
81 (42-100)

75 (62-82)
75 (68-83)*
75 (62-95)

959
3,041
4,899

FISH (Urovysion)
FISH (Urovysion)
This assessment

72 (69-75)*
76 (65-84)*
72 (23-100)

83 (82-85)*
85 (78-92)*
80 (40-100)

2,477
3,101
2,852

Sensitivity of Select Markers Based on Tumor Grade and Stage


Marker

# studies

Grade 1

Grade 2

Cytology
Cytology

8
9

0.12 (.04-.31)
0.17

0.26 (.17-.37) 0.64 (.38-.84)


0.34
0.58

BTA stat
BTA stat

8
7

0.47 (.38-.56)
0.45

0.73 (.59-.83) 0.94 (.55-.99)


0.6
0.75

NMP 22

0.41

NMP 22

0.61(.35-.81)

Immunocyt

0.78

0.9

FISH (Urovysion)

0.56

0.78

0.95

0.53

Grade 3

0.8

0.71 (.41-.90) 0.79 (.63-.89)

Marker

# studies

Ta

T1

>T2

TIS

Cytology

.15 (.09-.25)

.46 (.34-.59)

.55 (.35-.73)

.63 (.29-.87)

BTA stat

.57 (.47-.67)

.82 (.63-.92)

.91 (.74-.97)

.66 (.42-.83)

NMP 22

.60 (.42-.76)

.85 (.27-.97)

.89 (.50-.98)

.73 (.54-.86)

NMP22 improves prediction of UCB


in high risk patients

May improve referrals from PCP


May improve timing of cysto in systems with limited resources
Lotan and Shariat, BJU Int. 2009

AUC for detection: 74% (95% CI: 72-76%)


Range across institutions: 68 to 89%

Question: (How) can molecular markers support screening of


patients at risk of having or developing bladder cancer?
Statement:

Feasibility of bladder cancer screening has been demonstrated in several


prospective trials. The results from one study using dip-stick testing for
hematuria suggests a survival benefit of individuals undergoing

hematuria screening. Despite these encouraging reports validation of the


results and improved definition of risk populations suited for screening is
required.
Recommendation:
Bladder cancer screening using urine for testing is promising but
cannot be recommended at present.
LoE: 1b

Grade: B

Agreement: 100%

Question: (How) can molecular markers support follow-up of


patients with non-muscle invasive low risk bladder cancer?
Statement:
Marker-guided follow-up of patients with superficial low risk tumors

appears feasible. However, studies proving the efficacy of this concept


and demonstrating an added value for the patients or the health system
are lacking.
Recommendation:
Marker-guided follow-up of patients with non-muscle invasive low
grade bladder cancer appears attractive, however, based upon the
current level of evidence this procedure cannot be recommended
at present
LoE: 1b

Grade: B

Agreement: 100%

Question: (How) can molecular markers support follow-up of


patients with non-muscle invasive high risk bladder cancer?
Statement:

Molecular markers detect high grade bladder cancer with high


sensitivity. At this stage it remains unclear how molecular markers can
support surveillance of patients with high grade bladder cancer.
Recommendation:
A use of molecular markers in surveillance of patients with high
grade bladder cancer cannot be recommended.
LoE: 2b

Grade: B

Agreement: 100%

Question: (How) can molecular markers be used in reflex testing for


bladder cancer?
Statement:
At present experience with reflex testing is very limited (mostly restricted to
FISH technique) and therefore does not permit a definite statement. reflex

testing should be exploited in more detail within prospective controlled


studies.
Recommendation:
Reflex testing is considered experimental at present and should be
investigated further.

LoE: 2b

Grade: B

Agreement: 100%

Summary I

All test sensitivities > cytology (low grade!)

All test specificities < cytology

CIS sensitivities surprisingly low

No single marker has demonstrated superior clinical utility


over cytology and cystoscopy

There is no ideal marker

We will have to select appropriate markers according clinical needs


Screening setting: high specificity
Follow-up setting: sensitivity
Diagnostic strategy affects selection of a marker

Summary II
Despite numerous publications, the utility of markers in
clinical decision-making remains limited
meanwhile the first prospective trials on a marker-guided
follow-up in bladder cancer patients are underway.
In addition, several screening studies could demonstrate
the feasibility of molecular markers in this setting.
Probably no single assay will be sufficient because of the
molecular heterogeneity of bladder cancer

Clinical states of advanced UCB

Determine prognosis
Predict response to therapy
Therapeutic target
Monitor disease/treatment response

Clinically
Advanced Disease

Clinically
Non-muscle
Invasive Disease

Death from
Disease

Local and/or
Distant
Recurrence

Clinical
Metastases

Death from
Other Causes

Clinical Needs

Accuracies of prognostic models (i.e., nomograms) not perfect


there may be markers that help improve outcome prediction
Molecular markers established as independent predictors of disease
recurrence and survival in pts treated with bladder cancer
Novel markers may identify patients with biologically and
clinically aggressive BCa and thereby help improve current
staging and prognostic models

Tumors are heterogeneous

Potential Role for Biomarkers


Identify high risk patients (staging and prognosis)
Advanced disease at cystectomy
Recurrence after cystectomy
Predict response to chemotherapy
Identify pathways for targeted therapy

Shariat et al, 2009

PROMISING TISSUE-BASED PROGNOSTIC MOLECULAR MARKERS


Cell cycle related markers

p53, pRB, p21, p27, cyclin E1, cyclin D1, Ki67

Apoptosis

Fas (CD95), Caspase-3, Survivin, Bcl-2

Angiogenesis

Microvessel density, Thrombospondin-1, VEGFR, bFGF

Signaling proteins

ERBB family receptors, RAS, PI3K pathway genes, FGFR3

Hormone receptors

Her2, AR, ER, PR

PROMISING BLOOD-BASED PROGNOSTIC MOLECULAR MARKERS


Plasma-insulin growth factor binding protein-1 and -3
Transforming growth factor-1

Interleukin-6 and its receptor


Soluble E-cadherin
Urokinase plasminogen activation axis (uPA, PAI-1, PAI-2, uPAR)
Matrix metalloproteinase

HIGH-THROUGHPUT TUMOR PROFILING (DNA microarray, metabolomics, etc)

Investigated pathways
Cell Cycle regulation
p53

inhibits G1/S progression

pRb

sequesters E2F, inhibits cell cycle progression

p21/p16/p27

cyclin-dependent kinase inhibitors

Cyclin D1/cyclin E cell cycle control at G1/S transition


KI-67

cell proliferation

Apoptotic pathway
p53

induces apoptosis or DNA repair

Bcl-2

inhibits caspase activation

Bax

activates cytochrome C release and apoptosis

Caspase 3

induces apoptosis

Survivin

protects from apoptosis and regulates mitosis

Angiogenesis pathway
TSP1

inhibits angiogenesis

VEGF

promotes angiogenesis through NOS

uPA

degrades extracelullar matrix

bFGF

growth factor stimulating angiogenesis

Is one marker optimal?

Combined cell-cycle biomarkers


Shariat et al., J Clin Oncol 2003

Combined apoptosis biomarkers


Karam et al., Lancet Oncol 2007

Nomograms
Recurrence

BCa-specific Death

Shariat et al., Cancer 2008

Competing-risk nomograms for prediction of clinical


outcomes in patients with pT1-2 N0 UCB

after radical cystectomy

Recurrence

Points
Age at surgery
P. Grade

10

20

30

40

50

60

Cancer-specific Survival

70

80

90

100

Age at surgery

30 40 50 60 70 80 90
2

Lymphova

3
1

Lymphova

Total Points
1(Y)REC-Free.Surv.
2(Y)REC-Free.Surv.
3(Y)REC-Free.Surv.
5(Y)REC-Free.Surv.
7(Y)REC-Free.Surv.

0
0

2
20
0 .9 9 5

40
0 .9 9

0 .9 9

60

20

30

40

50

60

70

30

40

50

60

70

80
1

80

4
100

120

0 .9 8 0 .9 7 0 .9 50 .9 3 0 .9 0 .8 6 0 .8

0 .9 8 0 .9 7 0 .9 50 .9 3 0 .9 0 .8 6 0 .8

140

160

0 .7 0 .6 0 .5 0 .4

180

Total Points
1(Y)BCa-Spec.Surv.

20

0.998

2(Y)BCa-Spec.Surv.

40

0.995

0.995

0.99

60

0.99

80

0.98 0.97

0.98 0.97

4
100

120

0.95 0.93 0.9 0.86 0.8

0.95 0.93 0.9 0.86 0.8

140

160

0.7 0.6

0.7 0.6 0.5 0.4 0.3

0.995

0.99

0.98 0.97

0.95 0.93 0.9 0.86 0.8

0.7 0.6 0.5 0.4 0.3 0.2

0 .7 0 .6 0 .5 0 .4 0 .3 0 .2 0 .1
0.99

0 .7 0 .6 0 .5 0 .4 0 .3 0 .2 0 .1

7(Y)BCa-Spec.Surv.
0 .9 8 0 .9 7 0 .9 50 .9 3 0 .9 0 .8 6 0 .8

100

5(Y)BCa-Spec.Surv.
0 .9 8 0 .9 7 0 .9 50 .9 3 0 .9 0 .8 6 0 .8

90

0 .7 0 .6 0 .5 0 .4 0 .3 0 .2

3(Y)BCa-Spec.Surv.
0 .9 8 0 .9 7 0 .9 50 .9 3 0 .9 0 .8 6 0 .8

80

90

No. of Markers

0
1

No. of Markers

10

Points

0 .7 0 .6 0 .5 0 .4 0 .3 0 .2 0 .1 0 .0 3

0.99

0.98 0.97

0.98 0.97

0.95 0.93 0.9 0.86 0.8

0.95 0.93 0.9 0.86 0.8

0.7 0.6 0.5 0.4 0.3 0.2

0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.03

Shariat et al., J Urol, 2011

Ongoing clinical trials at UT


Southwestern Dallas, Texas
1. Prospective staining of
all TUR and cystectomy
specimens for markers
2. Prospective, randomized
single center study:
adjuvant chemotherapy after
radical cystectomy in patients
with pT1-3 N0 who have 3
markers altered

Shariat et al., Cancer 2008

UCB Molecular Pathways- Opportunities for Targeted


Therapies
Non Invasive LG UCB pathway: HRAS or FGFR3 mutant proteins
Inhibition of RTK-as signaling pathway
e.g. Receptor tyrosine kinase inhibitors: gefitinib, erlotinib
ErbB2 blocker: trastuzumab
EGFR blocker: cetuximab
Invasive High Grade UCB pathway:
Restoration p53/RB function: gene therapy/small molecules
Inhibition VEGF/VEGFR: Sorafenib, sunitinib, pazopanib,
bevacizumab

Inhibition of anti-apoptotic molecules (survivin)

Summary
Currently NO prognostic marker is ready for integration
into clinical decision-making
Bladder cancer develops along multiple molecular pathways
multiple molecular markers to capture biological
potential of tumor
Molecular medicine holds the promise that clinical outcomes
will be improved by directing therapy toward tumor
mechanisms and targets
The advent of high-throughput technologies is allowing
comprehensive identification of molecular targets and
molecular markers

The Future: integration of molecular panels


Improve prognosis of heterogeneous disease (UCB)
e.g. early cystectomy in pT1HG
Improve selection and thereby increase use of current

chemotherapy
Prediction of response to current chemoRx
Therapeutic response indicator
e.g. select patients who achieve pT0 after neoadjuvant
chemotherapy for bladder preservation

Target for therapy (combination therapies)

Questions?

57

ICUD-EAU 2011
Recommendations of committee
on LG Ta disease
Badrinath Konety (USA), Willem Oosterlinck
(Belgium), Paul Sved (Aus), Raj Pruthi (USA),
Sam Chang (USA), Eduardo Solsona (Spain),
Sigurdur Gudjonsson (Sweden), Mark
Soloway (USA)

Levels of Evidence used


(modified from Sackett et al. OCEBM 2001)

Level

Type of evidence

1a

Evidence obtained from meta-analysis of randomized trials

1b

Evidence obtained from at least one randomized trial

2a

Evidence obtained from one well designed, controlled study


without randomization

2b

Evidence obtained from at least one other type of well designed,


quasi-experimental study

Evidence obtained from well designed non-experimental studies


such as comparison studies, correlation studies and case reports

Evidence obtained from expert committee reports or opinions or


clinical experience of respected authorities

Grade of recommendation
Grade

Nature of recommendation

Based on clinical studies of good quality and consistency addressing the


specific recommendations and including at least one randomized trial

Based on well conducted clinical studies but without randomized clinical


trials

Made despite the absence of directly applicable clinical studies of good


quality

Grade classification (WHO-ISUP)


Papilloma
PUNLMP
Low grade

High grade

Incidence of LG Ta SEER-Medicare
2000-2005
Year of Diagnosis

Stage I Disease

2000
2001
2002
2003
2004
2005

807 (19.28)
814 (19.28)
797 (18.33)
835 (17.37)
930 (17.96)
848 (18.06)

LG Ta
Progression rates are low <20%
Cancer specific survival is extremely high
>95%
PUNLMP more likely to recur but
progression extremely rare
Lifelong risk of recurrence with late
recurrence following years of no tumor

Overall survival LG Ta
Kaplan-Meier Estimate for Overall Survival
1.00

Survival Distribution Function

0.75

0.50

0.25

0.00
0

10

20

30

40

50

60

70

survival_months
STRATA:

gradecat=High Grade
gradecat=Undifferentiated Grade

Censored gradecat=High Grade


Censored gradecat=Undifferentiated Grade

gradecat=Low Grade
gradecat=Unknown Grade

Censored gradecat=Low Grade


Censored gradecat=Unknown Grade

80

Cancer specific survival LG Ta


1.00

Survival Distribution Function

0.75

0.50

0.25

0.00
0

10

20

30

40

50

60

survival_months
STRATA:

gradecat=High Grade
gradecat=Low Grade
gradecat=Undifferentiated Grade
gradecat=Unknown Grade

Censored gradecat=High Grade


Censored gradecat=Low Grade
Censored gradecat=Undifferentiated Grade
Censored gradecat=Unknown Grade

70

80

Prior recommendations
Oosterlinck et al. 2005
Lesions >0.5cm, multiple (>5),
papillonodular and positive cytology need
further evaluation (B)
Routine upper tract studies not
recommended during follow up (B)
To be considered in symptomatic patients and
those with positive cytology (A)

Random biopsy for positive cytology (B)


Oosterlinck et al. Urology 2005

Previous recommendations
Immediate post TUR intravesical chemo recommended
(A)

MMC (40mg or Epirubicin (50mg) (A)


Avoid chemo if perforation present (D)
Instillation same day as TUR (B)
Adjuvant intravesical chemo for multiple tumors (A)

Intravesical chemo first line (B)


BCG second line (A)
Cysto at 3 months if negative next at 9 months (C)
Office fulguration acceptable (C)
Urine cytology or markers not needed at diagnosis (A)
or follow up (B)

Literature review
PubMed search conducted
Search terms used
Bladder cancer; bladder neoplasm, low grade, low
stage, diagnosis, intravesical therapy; cystoscopy,
smoking, prevention, observation, expectant
management, CT urography, intravenous
pyelogram/IVP, cytology, urinary markers, fluorescent
cystoscopy, transurethral resection, surveillance, BCG,
Mitomycin.
Human, 25 years
Other association guidelines (AUA, EAU, BAUS, Int.
Consultation on bladder tumors/SIU)

Urinary tumor markers


Recommendation
BTA has little role in diagnosis of LG Ta
tumors
NMP22 has better sensitivity than
cytology but cannot replace cystoscopy
ImmunoCyt has good sensitivity for LG Ta
tumors and can be used to prolong
intervals between cystoscopy
UroVysion/FISH has no role in diagnosis
of LG Ta tumors

Level of
evidence
2b

Grade of
recommendation
B

1b

2a

2b

Performance characteristics of
markers
Marker
UroVysion
Microsatellite analysis
Gene microarray
Immunocyt/ uCyt+
Nuclear matrix protein 22
BTA stat
BTA TRAK
Cytokeratins
Survivin

Sensitivity

Specificity

30-72
58
80-90
76-85
49-68
57-83
53.8-91
12.1-85
53-90.4

63-95
73
62-65
63-75
85.8-87.5
68-85.7
28.3-83.9
75-97.4
88-100

Upper tract evaluation and surveillance


Recommendation

Level of evidence

Grade of
recommendation

Routine upper tract studies are not


recommended for patients with LG Ta
tumors at diagnosis of index tumor
Upper tract evaluation in patients with
LG Ta UC is recommended in the
presence of:
o gross hematuria
o unexplained positive urine cytology

2b

2b

Routine upper tract studies surveillance is


not recommended for patients with LG
Ta tumors during follow up

2b

Expectant management
Recommendation

Level of
Grade of
evidence recommendation

Expectant management should be pursued only in


patients with established history of LG Ta UC
Optimal tumor characteristics include low tumor burden

Particularly applicable to those with advanced age or


co-morbidity but can be offered to younger patients
after careful and thorough discussion
Surveillance includes periodic cystoscopy and cytology

Patients who demonstrate increase in size, number or


appearance of tumor(s) or develop positive urine
cytology should cease expectant management and
undergo formal TUR and further evaluation
Office fulguration is a good option for patients with small
LG Ta UC. This can be followed by post fulguration
instillation of intravesical chemotherapy

Intravesical chemotherapy
Recommendation
Initial step in management is complete
TUR
Immediate post-TUR instillation of
intravesical chemotherapy in the form of
Mitomycin C or epirubicin is
recommended to prevent recurrences
Induction chemotherapy with or without
maintenance has unclear but potential
benefit
The risks of repeated courses of
intravesical chemotherapy have to be
weighed against the benefit

Level of
evidence
2b

Grade of
recommendation
A

1a

2b

Intravesical immunotherapies
Recommendation

Intravesical BCG is not appropriate as initial


intravesical therapy for patients with LG Ta UC
Intravesical BCG could potentially be used in
patients with recurrent LG Ta UC who have not
responded to courses of intravesical
chemotherapy
Secondary immunomodulating therapies do not
have a role in management of LG Ta UC at this
time

Level of
evidence

Grade of
recommendation

High grade Ta, T1 and CIS

ICUD update session


dr. Max Burger, Regensburg
prof. Fred Witjes, Nijmegen

March 18, 2011

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

Members committee 4

Marko Babjuk, Prague


Maurizio Brausi, Modena
Chris Cheng, Singapore
Eva Comperat, Paris
Jay Shah, Colin Dinney, Houston
Joachim Throff, Wolfgang Jger, Mainz
Wolfgang Otto, Max Burger (co chair), Regensburg
Fred Witjes (chair), Nijmegen

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

Dr Max Burger

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

High grade Ta

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

Ta HG recommendation 1
Ta HG should be suspected if
papillary tumour is seen in cystoscopy
urinary cytology is positive

The assessment Ta HG should be based on

multiplicity
tumour size
early recurrence
concomitant CIS

LOE 3, recommendation grade B


Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

Ta HG recommendation 2
Ta HG should be completely resected
photodynamic diagnostic may be considered
uncertain radicality should prompt re- TURB

Ta HG should undergo instillation therapy


effect of early instillation uncertain
BCG should be given
duration of BCG should exceed one year

LOE 2, recommendation grade A


Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

CIS

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

CIS recommendation 1
If CIS is suspected clinically
cytology, cystoscopy, upper tract imaging

TURBT in CIS:
PDD considered in pos. cytology/ neg. cystoscopy

LOE 1a, recommendation grade A


Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

CIS recommendation 2
Cystectomy in CIS:
provides excellent survival
however, overtreatment in 50%

LOE 2, recommendation grade A

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

CIS recommendation 3
Intravesical therapy
immediate instillation may be considered
should follow complete resection of pap. tumours
should be based on BCG

LOE 1a, recommendation grade A


Maintenance therapy
required, best schedule unkown, >1 year

LOE 2, recommendation grade A


Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

CIS recommendation 4
Treatment response

should be assessed at 3 months


if no response: cystectomy or further BCG
BCG: another 6 wk. course or 3 wkly boosters
Unknown when best to abandon conservation

LOE 2, recommendation grade B

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

CIS recommendation 5
Follow up
white light cystoscopy and cytology

every 3 months for 2 years


every 4 months in 3rd year
every 6 months in 4th/ 5th year
once per year thereafter

imaging of upper tract yearly

LOE NA, recommendation grade C

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

Q&A

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

Dr Witjes

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

T1 disease

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

T1 recommendation 1
The assessment T1 BC should be based on

tumour grade and size


early recurrence
Multiplicity
concomitant CIS
UC involving the prostatic mucosa or ducts
depth of lamina propria invasion

LOE1a, recommendation grade A


Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

Prediction of prognosis
impossible on an individual level

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

Other potential prognostic factors


Response to intravesical therapy
Tumor at 3 months after BCG: up to 80% progression

Insufficiently validated in clinical trials are:


LVI??
Markers
P53, p27, Ki67

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

T1 recommendation 2
The use of one immediate postoperative
instillation of chemotherapy is not supported
by consistent data and therefore it should not
be recommended as standard practice
LOE1a, recommendation grade A

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

New literature on immediate instillation


Hendricksen (2008):
course of epirubicin with or without one instillation

Gudjonsson (2008), Berrum-Svennung (2008):


epi versus nothing
only effect in very low risk tumors

Bhle (2009):
Gemcitabin versus placebo
20 hrs contineous bladder irrigation

Brausi (2010):
data in high risk very limited
Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

EAU guideline 2011

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

EAU guideline 2011

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

T1 recommendation 3
Patients with high risk and with recurrent or
persistent disease after BCG immunotherapy
should be offered cystectomy
LOE2a, recommendation grade A

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

Who already initial cystectomy


Especially consider immediate cystectomy
in patients with 2 of the following factors:
Tumor multifocality
Size > 3 cm
Associated CIS

Obviously non responders to BCG


3 or 6 months
Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

Why cystectomy
No RCTs, risk of overtreatment, but.

Frequent understaging
Up to 30% in T1
Up to 50% in high grade T1

Better survival after surgery


QOL: pts seem to accept surgery and the
consequences for better survival
Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

Why cystectomy

Prognosis in case of progression is bad

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

T1 recommendation 4
If a bladder sparing approach is desired, a
secondary TURB should be performed and
followed by intravesical BCG therapy
LOE3, recommendation grade B

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

BCG
Some kind of maintenance schedule should
be used
Impact on progression unclear

BCG has more toxicity

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

BCG failure

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

BCG failure recommendation 1


In any BCG nonresponse or failure
cystecomy is recommended
LOE 2a, recommendation grade A.

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

BCG failure recommendation 2


The threat of progression remains real but
comfortably low enough within the first 6
months of beginning BCG to consider
alternatives to cystectomy for those patients
unfit or refusing this standard management
option
LOE 2, recommendation grade B

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

Progression
comfortably low enoughmeans
<5% within 12 months
Usually exceeds 12 months

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

Know the different BCG failures


Refractory:
disease free not achieved within 6 mths (persistence,
recurrence, progression)

Resistant:
recurrence of persistence at 3 months, free at 6 months

Relapsing:
disease after 6 months nd after CR

Intolerant:
disease after less than adequate therapy
Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

BCG failure recommendation 3a


Alternative treatment options
repeat resection and repeat BCG (LOE 1a, Gr.A)
More than 2 cycles BCG seems ineffective

possibly combine with IFN- (LOE 2, grC)


Gemcitabin has shown effectivity, but more
studies are needed (2 recent trials)
Thermo-chemotherapy has shown effectivity, but
more studies are

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

BCG failure recommendation 3b


No alternative treatment options

There is no reported evidence of significant


efficacy using

current intravesical chemotherapy


IFN- monotherapy
photodynamic therapy
radiation therapy.
Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

Q&A

Department of Urology
Radboud University Nijmegen Medical Centre
The Netherlands

ICUD-EAU International Consultation on


Bladder Cancer

Committee #6
Muscle-invasive bladder cancer

26th EAU Congress, Vienna, 2011


Eberhard-Karls University Tuebingen

Department of Urology

Eberhard-Karls University Tuebingen

Department of Urology

Committee Members
Joaquim Bellmunt
Michael S. Cookson
Georgios Gakis
Khurshid A. Guru
Axel Heidenreich
Seth P. Lerner
Edward M. Messing
David I. Quinn
Mark P. Schoenberg
William U. Shipley
Arlene O. Siefker-Radtke
Arnulf Stenzl (Chair)
Cora Sternberg
Eberhard-Karls University Tuebingen

Department of Urology

Contents
1. OVERVIEW
2. INDICATIONS AND ALGORITHM OF TREATMENT
3. RADICAL CYSTECTOMY
1. Removal of the tumor bearing bladder
1. Timing and delay of cystectomy
2. Technique
3. Lymphadenectomy
4. Laparoscopic and robotic-assisted lap. cystectomy
2. Surgical outcome: morbidity and mortality
3. Oncological outcome of radical surgery according to TNM
staging
4. PERIOPERATIVE CHEMOTHERAPY
1. Neoadjuvant chemotherapy
2. Adjuvant chemotherapy
Eberhard-Karls University Tuebingen

Department of Urology

Contents contd.
5. PERIOPERATIVE RADIOTHERAPY
6. BLADDER-SPARING TREATMENT FOR LOCALISED DISEASE

1. Transurethral monotherapy
2. Partial Cystectomy
3. External beam radiotherapy

4. Multimodality treatment
7. TREATMENT OF MIXED HISTOLOGY UROTHELIAL CARCINOMA
8. FOLLOW-UP

1. Site of recurrence
1. Distant recurrences
2. Follow-up and treatment of secondary ureteral and urethral

tumors
Eberhard-Karls University Tuebingen

Department of Urology

Timing and delay of cystectomy


Minimally invasive techniques including robotic assisted
laparoscopic radical cystectomy have been reported in
increasing numbers. Shortterm outcomes, pathologic findings
and assessments of morbidity compared to the open technique
have been reported. Continued research with these techniques
is required and specifically results of several randomized trials
underway should be reported prior to accepting this as an
equivalent option to open radical cystectomy (Level 2a Grade
C).
Cystectomy is also indicated as palliation for intractable pelvic
pain and bleeding from locally advanced or metastatic bladder
cancer (Level 4 Grade C)
Eberhard-Karls University Tuebingen

Department of Urology

Timing and delay of cystectomy


Delay in cystectomy for muscle-invasive disease is
associated with significantly reduced overall and cancer
specific survival.
(level of evidence: 3, grade of recommendation: B)

Radical cystectomy in patients with muscle-invasive


bladder cancer should be performed within 3 months
after initial diagnosis of stage localized T2 T4 disease.
(level of evidence: 3 grade of recommendation: C)

Eberhard-Karls University Tuebingen

Department of Urology

Treatment of mixed histology urothelial


carcinomas
Nested variant and micropapilary urothelial cancers regardles
of detected in non-muscle invading or muscle-invading stages
should be treated with aggressive local extirpative therapy
(Level of Evidence 3, Grade of recommendation B)
Carcinosarcomas should be treated if possible with local
extirpative surgery
(Level of Evidence 3, Grad of recommendation B)
Urothelial cancer with small cell components should be treated
with neoadjuvant chemotherapy including cisplatin and
etoposide neoadjuvantly followed by aggressive local treatment
(Level of Evidence 3, Grade of recommendation B)
Eberhard-Karls University Tuebingen

Department of Urology

Radical Cystoprostatectomy
apex-sparing
Urethra

Bladder
P
D

Rhabdosphincter

SV

enucleation

Neurovascular bundle

seminal vesicle sparing


LK
Eberhard-Karls University Tuebingen

Department of Urology

Technique

Preservation of the anterior and membranous urethra


including parts of the prostate and seminal vesicles for
reasons of fertility, potency and continence are technical
variations to the nerve-sparing approach which may improve
patients quality of life but must be attentively judged
against possible oncological risks.
(level of evidence 3, Grade C)

Eberhard-Karls University Tuebingen

Department of Urology

Cystectomy in FemalesIndications and extent

Urethra

T2-T4a, N0-NX, M0 B
high-risk pTa-T1
BCG Non-responder
Transurethrally uncontrollable
pTa-disease
Prostate (GR B)

Sphincter

Uterus

Bladder

Vagina

Pelvic Lymphadenectomy
Eberhard-Karls University Tuebingen

Department of Urology

Technique
In female patients standard anterior pelvic exenteration
includes the entire urethra, adjacent vagina, uterus, distal
ureters and respective lymph nodes
(level of evidence: 3, Grade: C)

The urethra + supplying autonomous nerves can be


preserved in case of a planned orthotopic neobladder
(level of evidence: 3, Grade: C)

Eberhard-Karls University Tuebingen

Department of Urology

Oncology

Function

Eberhard-Karls University Tuebingen

Department of Urology

Laparoscopic and robotic-assisted


laparoscopic cystectomy (RALC)
Minimally invasive approach to radical cystectomy has a significant
role in management of locally advanced bladder cancer (level 3,
grade C)

Immediate oncologic results are acceptable and comparable to


open radical cystectomy. Meanwhile long-term oncological data is
still awaited and will define its permanent role in urologic oncology.
(level 2b, grade B)

Eberhard-Karls University Tuebingen

Department of Urology

Laparoscopic and robotic-assisted


laparoscopic cystectomy (RALC) - contd
Centers across the world with expertise in minimally invasive
surgery especially surgical expertise should presently be offering
these options for bladder cancer. (level 3, Grade C)

High-volume centers with dedicated minimally invasive surgical


teams have shown better and safe results. (level 3, Grade B)

Eberhard-Karls University Tuebingen

Department of Urology

Cystectomy:
Short/long costs of BCa/UUT Ca
Hospital/Surgeon Volume vs. costs

Konety et al, J Urol, 2004

Cystectomy:
Hospital/Surgeon Volume vs. Mortality
Hospital volume

Surgeon volume

% Mortality

5
4
3
2
1
0
<3

3-10

11-25

26-50

>50

Cystectomies per year


Barbieri et al, J Urol, 178:1418-22, 2007

Mc Cabe et al, Postgrad Med J, 83:556-60, 2007

Eberhard-Karls University Tuebingen

Department of Urology

Surgical outcome-morbidity and mortality


Surgical complications associated with radical cystectomy and urinary
diversion should be reported in a uniform grading system. The currently
best adapted graded system for cystectomy is the Clavien grading
system (level of evidence: 2, grade of recommendation B).
Surgical complications associated with radical cystectomy and urinary
diversion should include the length of follow-up for the patient cohort,
and a minimum of 30-day but preferable for 90-day reported outcome
(level of evidence 3, grade of recommendation C).
ASA score, age, previous treatment for bladder cancer or other pelvic
diseases, surgeon and hospital volume of radical cystectomy, and type
of urinary diversion influence surgical outcome (level of evidence: 3,
grade of recommendation: C)
Reduction in blood loss and blood transfusion is afforded by meticulous
technique, use of modern surgical devices, and improved understanding
of pelvic anatomy. (level of evidence: 3, grade of recommendation C)
Eberhard-Karls University Tuebingen

Department of Urology

Follow-up and treatment of secondary


ureteral tumours
Ureterorenoscopy with biopsy is the diagnostic procedure of
choice for the diagnosis of a metachronous upper tract
recurrence after RC (Level 3, Grade C).
The use of cytology and urine-based markers, i.e. fluorescence
in-situ hybridization and immunocytology, cannot be
recommended for routine follow-up of the upper tracts so far
since their use has only been evaluated in patients with primary
upper tract cancers (Level 3, Grade B).
Routine upper tract imaging is only indicated in patients with
clinical symptoms suspicious of a metachronous upper tract
recurrence (Level 3, Grade B).

Eberhard-Karls University Tuebingen

Department of Urology

Follow-up and treatment of secondary


ureteral and urethral tumours

Nephroureterectomy is the treatment of choice for


invasive upper tract recurrence providing prolonged
survival.
(Level 3, Grade B)

A regional lymphadenectomy in patients undergoing


radical nephroureterectomy is only of diagnostic value.
(Level 3, Grade C).

Eberhard-Karls University Tuebingen

Department of Urology

Eberhard-Karls University Tuebingen

Department of Urology

Eberhard-Karls University Tuebingen

Department of Urology

Spare for discussion only

Eberhard-Karls University Tuebingen

Department of Urology

Oncological outcome according to TNM


staging
The AJCC substratifications in node-negative pT2 and pT3
bladder cancer are of prognostic value (Level 3, Grade B)
Nomograms provide improved prognostic information
oncological outcomes after radical surgery as compared to
pTNM staged predictions. However, its general applicability
not been established sufficiently by external validation so
(Level 3, Grade B)

for
the
has
far

In patients older than 80 years, radical cystectomy is associated


highest risk reduction on cancer-related and non-cancer related
mortality (Level 3, Grade B).

Eberhard-Karls University Tuebingen

Department of Urology

Oncological outcome
Radical cystectomy is the mainstay of treatment in muscleinvasive bladder cancer providing long-term survival
(Level 2b, Grade B).
The AJCC substratifications in node-negative pT2 and pT3
bladder cancer are of prognostic value (Level 2b, Grade B)
In patients older than 80 years, radical cystectomy is associated
highest risk reduction on cancer-related and non-cancer related
mortality (Level 3, Grade B).
Based on the scarce data available the routine use of molecular
markers for primary risk assessment in invasive bladder cannot
be recommended so far (Level 3, Grade B).
Eberhard-Karls University Tuebingen

Department of Urology

Follow-up and treatment of secondary


ureteral tumours
Surveillance regimens should be based on a risk-adapted strategy
(Level 3, Grade C).
The number of risk factors potentiates the risk of recurrence: (a)
positive ureteral margin (b) carcinoma in situ of the bladder and
ureter, (c) tumor multifocality, (d) urethral tumor involvement
and (e) male gender (Level 3, Grade B).
Patients with finally positive ureteral margins at surgery are at
increased risk of ureteral recurrence. Frozen section analysis has
a high sensitivity and specificity for the detection of malignant
ureteral margins (Level 2b, Grade B).
Eberhard-Karls University Tuebingen

Department of Urology

Follow-up and treatment of secondary


urethral tumours
Risk factors for urethral recurrence in male patients are
prostatic tumor involvement (either superficial or invasive) and
bladder neck involvement in female patients (Level 3, Grade B).
Intraoperative frozen section analysis has a high sensitivity and
specificity for the detection of a malignant urethral margin for
both male and female patients (Level 2b, Grade B).
Patient with positive final urethral margins are at increased risk
of urethral recurrence (Level 3, Grade B).
Prophylactic urethrectomy is not justifiable in patients with
carcinoma in situ at the urethral margin (Level 3, Grade C).

Eberhard-Karls University Tuebingen

Department of Urology

Follow-up and treatment of secondary


urethral tumours
The use of urinary cytology, urethral washings and diagnostic
urethroscopy for follow-up has not proven to provide a survival
benefit (Level 3, Grade B).
Urethrectomy should be considered in patients with invasive
carcinoma at the urethral margin at radical cystectomy or in
case of invasive urethral recurrence (Level 3, Grade B).
Local conservative treatment is an option in patients with noninvasive tumor(s) or carcinoma in situ of the urethra (Level 3,
Grade C).
In patients with urethral recurrence and concomitant distant
recurrence outside the urinary tract systemic chemotherapy is
indicated (Level 3, Grade C).
Eberhard-Karls University Tuebingen

Department of Urology

ICUDEAU GUIDELINES ON BLADDER CANCER

Committee 7: Urinary diversion


Richard E. Hautmann, Hassan Abol-Enein,
Thomas Davidsson,
Sigurdur Gudjonsson
Stefan H. Hautmann
Henriette V. Holm
Cheryl T. Lee
Fredrik Liedberg
Stephan Madersbacher Wiking Mansson
Murugesan Manoharan Robert D. Mills
David F. Penson
Eila C. Skinner
Raimund Stein
Urs E. Studer
Joachim W. Thueroff
William H. Turner
Bjoern G. Volkmer

ICUD-EAU GUIDELINES BLADDER CANCER

RCX AND URINARY DIVERSION


Highest relative value
in terms of difficulty of the surgery for
any procedure in urology
They are also the most difficult laparoscopic or robotic
procedures and more so if the diversion is performed
totally intracorporeally.
Risk of procedure is based
not only on the technical challenges
but also on the nature of the patients needing.

ICUDEAU GUIDELINES ON BLADDER CANCER

RCX AND URINARY DIVERSION (UD)


ARE 2 STEPS OF ONE OPERATION
Literature uniformly reports the
complications of RCX
While ignoring that the majority of
complications are diversion related (75%)!
This may seem sematic, but it is not
Hautmann, J Urol 2010

ICUD-EAU GUIDELINES ON BLADDER CANCER

COMMON MEASURES
FOR SURGICAL OUTCOMES OF RCX AND UD

Estimated blood loss


Operative time
Analgesic use
LOS
Time to return to work
Perioperative death
Complication rates
Hospital costs

ICUD-EAU GUIDELINES ON BLADDER CANCER

FEW STUDIES ON RCX AND UD


REPORT ON

Readmission rates
Reoperation rates
Intensive care stays
Additional interventional radiologic
procedures
DONAT, UROLOGY, 2007

ICUDEAU GUIDELINESS ON BLADDER CANCER

THE MAJORITY OF RCX SERIES ON MORBIDITY HAD:

NOT defined complications


NOT utilized a grading system other
than categorizing into
major versus minor
NOT accounted for comorbidities
NOT employed a formal complication
reporting system
Shabsigh EUR Urol 2009

ICUD-EAU GUIDELINES ON BLADDER CANCER

CONCLUSIONS
The lack of consistency in the method used to
accrue, define, and report complication data
makes it impossible to compare morbidity rates
among different surgical techniques, surgeons
or institutions
Improved surgical outcome reporting
(f.i. Clavien system) will allow more meaningful
comparisons when randomized trials are sparse
and f.i. for comparison between open and
minimaly invasive techniques
DONAT, UROLOGY, 2007

ICUD-EAU GUIDELINES OF BLADDER CANCER

EVIDENCE AND URINARY DIVERSION

Not a single randomized study *


Almost all studies used in this report are of
Level 3 or Level 4 evidence including
expert opinion based on first principles
research
The grades of recommendations given are
mostly of Grade C
* except studies on ureteroileostomy

Table 1: Numbers and Types of Urinary Diversions (%) performed by the Authors
No
diversion/
Anal unknown others

No. RCX

Period

Neobladder

Cont.Cut.

Conduit

UC
TUUC

643

95-04

45.1

1.4

53.5

962

00-09

40.0

2.0

58.0

0.9

611

85-99

51.5

1.5

42.5

1.6

2.5

0.4

708

00-10

51.0

8.0

39.0

1.5

0.5

0.1

765

94-10

30.2

6.8

60.5

0.7

2.0

0.1

1359

71-01

51.6

25.8

22.3

0.3

1012

00-10

74.2

5.3

20.2

0.3

119

00-04

28.6

31.1

40.3

134

04-09

6.0

30.6

63.4

335

68-80

55.0

45

593

81-90

2.0

33

41.0

15

982

91-00

6.0

39

41.0

12

1023

01-10

15.0

24

53.0

Mansoura

3157

80-04

39.1

3.5

34.4

23.1

Norwich

246

02-09

10.6

89.4

158

1997

19.0

19

55.0

221

2003

17.0

12

70.0

208

2006

9.0

80.0

229

2008

15.0

81.0

Ulm

1613

86-09

66.0

0.4

22.0

10.0

1.3

0.2

Vanderbilt

789

00-07

35.5

0.4

63.5

0.1

0.5

Total

15867

38.0

10.4

42.2

1.2

7.5

0.1

0.8

Ann Arbor
Bern
Kassel
Los Angeles
Lund

Mainz

Swedish
Registry

Table 1: Numbers and Types of Urinary Diversions (%) performed by the Authors

No. RCX

Period

643

95-04

45.1

1.4

53.5

962

00-09

40.0

2.0

58.0

0.9

611

85-99

51.5

1.5

42.5

1.6

2.5

0.4

708

00-10

51.0

8.0

39.0

1.5

0.5

0.1

765

94-10

30.2

6.8

60.5

0.7

2.0

0.1

1359

71-01

51.6

25.8

22.3

0.3

1012

00-10

74.2

5.3

20.2

0.3

119

00-04

28.6

31.1

40.3

134

04-09

6.0

30.6

63.4

335

68-80

55.0

45

593

81-90

2.0

33

41.0

15

982

91-00

6.0

39

41.0

12

1023

01-10

15.0

24

53.0

Mansoura

3157

80-04

39.1

3.5

34.4

23.1

Norwich

246

02-09

10.6

89.4

158

1997

19.0

19

55.0

221

2003

17.0

12

70.0

208

2006

9.0

80.0

229

2008

15.0

81.0

Ulm

1613

86-09

66.0

0.4

22.0

10.0

1.3

0.2

Vanderbilt

789

00-07

35.5

0.4

63.5

0.1

0.5

Total

15867

38.0

10.4

42.2

1.2

7.5

0.1

0.8

Ann Arbor
Bern
Kassel
Los Angeles
Lund

Mainz

Swedish
Registry

Cont.Cut.

Conduit

UC
TUUC

No
diversion/
Anal unknown others

Neobladder

Table 1: Numbers and Types of Urinary Diversions (%) performed by the Authors

others

Period

643

95-04

45.1

1.4

53.5

962

00-09

40.0

2.0

58.0

0.9

611

85-99

51.5

1.5

42.5

1.6

2.5

0.4

708

00-10

51.0

8.0

39.0

1.5

0.5

0.1

765

94-10

30.2

6.8

60.5

0.7

2.0

0.1

1359

71-01

51.6

25.8

22.3

0.3

1012

00-10

74.2

5.3

20.2

0.3

119

00-04

28.6

31.1

40.3

134

04-09

6.0

30.6

63.4

335

68-80

55.0

45

593

81-90

2.0

33

41.0

15

982

91-00

6.0

39

41.0

12

1023

01-10

15.0

24

53.0

Mansoura

3157

80-04

39.1

3.5

34.4

23.1

Norwich

246

02-09

10.6

89.4

158

1997

19.0

19

55.0

221

2003

17.0

12

70.0

208

2006

9.0

80.0

229

2008

15.0

81.0

Ulm

1613

86-09

66.0

0.4

22.0

10.0

1.3

0.2

Vanderbilt

789

00-07

35.5

0.4

63.5

0.1

0.5

Total

15867

38.0

10.4

42.2

1.2

7.5

0.1

0.8

Ann Arbor

Bern
Kassel
Los Angeles

Lund

Mainz

Swedish
Registry

Conduit

Anal

No. RCX

Cont.Cut.

UC
TUUC

No
diversion/
unknown

Neobladder

Table 1: Numbers and Types of Urinary Diversions (%) performed by the Authors

No. RCX

Period

643

95-04

45.1

1.4

53.5

962

00-09

40.0

2.0

58.0

0.9

611

85-99

51.5

1.5

42.5

1.6

2.5

0.4

708

00-10

51.0

8.0

39.0

1.5

0.5

0.1

765

94-10

30.2

6.8

60.5

0.7

2.0

0.1

1359

71-01

51.6

25.8

22.3

0.3

1012

00-10

74.2

5.3

20.2

0.3

119

00-04

28.6

31.1

40.3

134

04-09

6.0

30.6

63.4

335

68-80

55.0

45

593

81-90

2.0

33

41.0

15

982

91-00

6.0

39

41.0

12

1023

01-10

15.0

24

53.0

Mansoura

3157

80-04

39.1

3.5

34.4

23.1

Norwich

246

02-09

10.6

89.4

158

1997

19.0

19

55.0

221

2003

17.0

12

70.0

208

2006

9.0

80.0

229

2008

15.0

81.0

Ulm

1613

86-09

66.0

0.4

22.0

10.0

1.3

0.2

Vanderbilt

789

00-07

35.5

0.4

63.5

0.1

0.5

Total

15867

38.0

10.4

42.2

1.2

7.5

0.1

0.8

Ann Arbor
Bern
Kassel
Los Angeles
Lund

Mainz

Swedish
Registry

Cont.Cut.

Conduit

UC
TUUC

No
diversion/
Anal unknown others

Neobladder

Table 1: Numbers and Types of Urinary Diversions (%) performed by the Authors

Ann Arbor

No. RCX Period


643
95-04
962
00-09

Neobladder
45.1
40.0

Cont.Cut. Conduit
1.4
53.5
2.0
58.0

UC
TUUC
-

No
diversion
/
Anal unknown others
0.9
-

611

85-99

51.5

1.5

42.5

1.6

2.5

0.4

708

00-10

51.0

8.0

39.0

1.5

0.5

0.1

765

94-10

30.2

6.8

60.5

0.7

2.0

0.1

1359

71-01

51.6

25.8

22.3

0.3

1012

00-10

74.2

5.3

20.2

0.3

119

00-04

28.6

31.1

40.3

134

04-09

6.0

30.6

63.4

335

68-80

55.0

45

593

81-90

2.0

33

41.0

15

982

91-00

6.0

39

41.0

12

1023

01-10

15.0

24

53.0

Mansoura

3157

80-04

39.1

3.5

34.4

23.1

Norwich

246

02-09

10.6

89.4

158

1997

19.0

19

55.0

221

2003

17.0

12

70.0

208

2006

9.0

80.0

229

2008

15.0

81.0

Ulm

1613

86-09

66.0

0.4

22.0

10.0

1.3

0.2

Vanderbilt

789

00-07

35.5

0.4

63.5

0.1

0.5

Total

15867

38.0

10.4

42.2

1.2

7.5

0.1

0.8

Bern
Kassel

Los Angeles
Lund

Mainz

Swedish
Registry

Table 1: Numbers and Types of Urinary Diversions (%) performed by the Authors

others

Period

643

95-04

45.1

1.4

53.5

962

00-09

40.0

2.0

58.0

0.9

611

85-99

51.5

1.5

42.5

1.6

2.5

0.4

708

00-10

51.0

8.0

39.0

1.5

0.5

0.1

765

94-10

30.2

6.8

60.5

0.7

2.0

0.1

1359

71-01

51.6

25.8

22.3

0.3

1012

00-10

74.2

5.3

20.2

0.3

119

00-04

28.6

31.1

40.3

134

04-09

6.0

30.6

63.4

335

68-80

55.0

45

593

81-90

2.0

33

41.0

15

982

91-00

6.0

39

41.0

12

1023

01-10

15.0

24

53.0

Mansoura

3157

80-04

39.1

3.5

34.4

23.1

Norwich

246

02-09

10.6

89.4

158

1997

19.0

19

55.0

221

2003

17.0

12

70.0

208

2006

9.0

80.0

229

2008

15.0

81.0

Ulm

1613

86-09

66.0

0.4

22.0

10.0

1.3

0.2

Vanderbilt

789

00-07

35.5

0.4

63.5

0.1

0.5

Total

15867

Ann Arbor
Bern
Kassel
Los Angeles

Lund

Mainz

Swedish
Registry

Conduit

Anal

No. RCX

Cont.Cut.

UC
TUUC

No
diversion/
unknown

Neobladder

38.0 10.4 42.2

1.2 7.5 0.1

0.8

ICUDEAU GUIDELINES ON BLADDER CANCER

UD AND RENAL FUNCTION

Age related GFR decline: 1 ml /min/year age 50


Serum creatinine remains normal until GFR < 50 %
Ideal: Isotopic GFR (51Cr-EDTA)
Cut off for continent diversion:
Serum creatinine
> 150 mol/l
GFR
< 50 %
Upper tract dilatation:
Does not equate to
reduction of renal function
Preexisting renal pathology: Greatest risk of postop
renal deterioration
Urinary diversion into bowel segments
is not inherently damaging to the kidneys

Table 2: Secondary tumours after diversion using isolated gut


segments
Ileum

Colon

Ileocecal

Stomach

Total no.

15+1*

22

Continent Cutaneous

14**

14

Neobladder

Rectal bladder

Cystoplasty

41

15

10

68

Total

60

42

10

114

Conduit

From publications by Austen and Klble (15,16), since 2004, and four Lund cases.

This case was transfered from continent cutaneous diversion using colonic segment.

** Includes 2 Lund cases (described in this paper) but excludes a case (17) from
Austen and Klble, that is now transferred to the ileal conduit group.

Mansson 2011

Table 3: Secondary Tumors in 17,758 Urinary Diversions


no.
secondary
tumors

no.
diversions

median
latency period
(years)

mean time:
operation to analysis
(years)

Neobladder Ileal

4190

(0,05 %)

3,0

6,3

Neobladder Ileocecal

239

(1,26 %)

4,0

13,6

Neobladder Colonic

70

(1,43 %)

6,0

9,9

Pouch Ileocecal

2181

(0,18 %)

12,0

8,8

Ileocystoplasty

233

(1,71 %)

21,5

10,5

Colocystoplasty

20

Conduit Ileal

8637

(0,02 %)

11,0

8,9

Conduit Colon

430

(0,23%)

40,0

24,4

Ureterocutaneostomy

1138

Ureterosigmoidostomy

16

620

Total

4,9

10,3
(2,58 %)

32 17,758

26,0

18,9

Klble 2011

Table 4: Tumor Risk in Continent and Incontinent Diversion

Ileum
Conduit

(ileo-)colon

0,02%

0,23%

p = 0.84

Pouch/Neobladder
Total

0,03%
p = 0.0091

0,05%
0,03%

total

p = 0.00001

0,28%

0,13%

0,27%

0,08%

p < 0.0001

Cystoplasty

1,71%

0,00%

1,58%

p = 0.46

Ureterosigmoidostomy

2,58%
Klble 2011

ICUDEAU GUIDELINES ON BLADDER CANCER

SECONDARY TUMORS AFTER DU


CONCLUSIONS
Ureterosigmoidostomy, cystoplasty and
probably (ileo)colonic OBS bear a
significantly increased tumor risk and
require regular endoscopic evaluation
beginning at the 5th postop year
After OBS and conduits regular endoscopy is
not mandatory
In ileocecal pouches, it is recommended in
the presence of symptoms such as
hydronephrosis, chronic urinary infection,
and hematuria

IDUC-EAU GUIDELINES ON BLADDER CANCER

ORTHOTOPIC BLADDER SUBSTITUTION


IN MEN (OBS)

INDICATION AND TUMOR STAGE


Extent of pelvic disease
little bearing on appropriateness
Pelvic recurrence
no impact on OBS function
LN + patients
achieve good functional results
HAUTMANN, J UROL, 1999

ICUD-EAU GUIDELINES ON BLADDER CANCER

OBS INDICATION AND TUMOR STAGE

NO CONSENSUS
PREOPERATIVE DISTAL PROSTATIC
BIOPSIES
Paraffin-embedded biopsies more reliable
Discuss result with patient before surgery

FROZEN SECTION OF THE RESECTION


MARGIN AT THE TIME OF SURGERY

ICUD-EAU GUIDELINES ON BLADDER CANCER

FUNCTION PRESERVING CANCER SURGERY

OBS WITH NERVE SPARING RCX

If tumor characteristics permit


Chance of maintaining erectile function
Preserved urethral sensation
Improved resting pressure
Better night time continence

KESSLER, J UROL, 2004


HUGONNET, J UROL, 2001

ICUD-EAU GUIDELINES ON BLADDER CANCER

OBS - AGE AND MOTIVATION


No age cut off for OBS
In practice: Many patients > 70 years will opt
for a simpler conduit as the postoperative
course is less arduous and incontince is less
likely
Motivation of the patient is the most
important factor when considering suitability
for an OBS, although it is difficult to assess
this objectively
Patients must be prepared to commit to the
long term follow up program necessary

ICUD-EAU GUIDELINES ON BLADDER CANCER

OBS - UPPER TRACT PRESERVATION


Voding with OBS cannot produce reflux
Antireflux nipple worse
than isoperistaltic tubular segment
2 RANDOMIZED SURGICAL TRIALS !
Low anastomotic stricture rate (<2,7 %)
Stented uretero-ileal anastomosis improves outcomes
WAIDELICH
STUDER
STUDER
SHAABAN

BJU
J.UROL
J.UROL
BJU

1998
2006
1996
2006

ICUD-EAU GUIDELINES O BLADDER CANCER

OBS POSTOPERATIVE MANAGEMENT


OF PARAMOUNT IMPORTANCE:

Active postoperative management


and regular long term follow-up
Key issue is achieving a capacity of 400-500 ml
Residual free voiding of sterile urine
Treatment of any outlet obstruction

THURAIRAJA BJU 2008

ICUD-EAU GUIDELINES ON BLADDER CANCER

OBS IN FEMALES
PATIENT SELECTION: ONCOLOGIC FACTORS
It is reasonable to advise against OBS for invasive
bladder neck involvement or suspected invasion of
the vaginal wall or cervix
Such patients may be considered for OBS if
intraoperative frozen section of the urethral margin
is negative
Overall 60-70% of women are resonable candidates
for OBS.

STEIN, UROLOGY 1998

ICUD-EAU GUIDELINES ON BLADDER CANCER

OBS IN FEMALES:
ANATOMIC BASIS OF THE PRESERVATION OF CONTINENCE

Prior to the 1990s the primary continence mechanism


was located in the bladder neck itself
The urethra alone could provide continence if the
sphincter mechanism was carefully preserved
The primary rhabdosphincter is an OMEGA shaped
structure surrounding the distal third of the urethra
The main nerves supplying this sphincter are the
somatic pudendal nerves which run under the
endopelvic fascia. The autonomic nerves coursing
through the pelvic plexus and along the lateral vagina
supply the smooth muscle of the bladder neck and
urethra
THANAGHO
HBNER

BJU
1966
BHATTA
EUR.UROL 2007
EUR.UROL 1993 COLLESELLI J UROL
1998

ICUD-EAU GUIDELINES ON BLADDER CANCER

OBS IN FEMALES SURGICAL BASIS OF NERVE SPARING


NO CONSENSUS

NERVE SPARERS:
Avoiding dissection below the endopelvic fascia surrounding the
urethra helps insure that this muscle and its nerve supply are not
disturbed
Nerve sparing aproach is crucial to maintaining urethral tone
and may decrease the risk of retention due to spasticity of the
denervated urethra
MECHANICS:
Routine dissection of presacral tissue as part of the node
dissection without clear reduction of continence
To date no randomized comparison of these two techniques has
been performed

STENZL
TURNER
ALI-EL-DEIN
STEIN

J.UROL 1995
J.UROL 1997
J.UROL 2002
J.UROL 1997

TABLE 9: Continence In Reported Series Limited to Females


(78, 79, 102, 105 - 109 )
# pts
Author
Hautmann

Follow-up

Daytime

Nighttime

(mos -mean)

Continence
%

Continence
%

Self
Catheterize
%

(1996)
(2000)
(2006)

116

60

83

83

50%

Granberg

(2008)

59

29

90

57

35

Nesrallah

(2005)

29

37

97

86

10

Ali-el-Dein

(2008)

192

51

92

72

16

Stenzl

(2001)

83

26

82

72

11

Stein

(2002)

81

78

78

N/A

33

Stein

(2009)#

56

103

87

66

61

# Anonymous validated questionnaire

ICUD-EAU GUIDELINES ON BLADDER CANCER

CONTINENT CUTANEOUS DIVERSION (CCD)

Second choice after OBS


Indications in 2011
Simultaneous urethrecomy
TCC lower urinary tract
TCC prostatic ducts (?)
Patient preference
Risk of leakage after OBS
Patient refuses labor intensive rehabilitation
Pediatric age group

ICUD-EAU GUIDELINES ON BLADDER CANCER

CCD PRINCIPAL DISADVANTAGES


No leak point (pop off)
Need for antireflux mechanism

NO CONSENSUS
Pouch pioneers
+
Pediatric urologists +
Risk of complications
CCD
Rupture
10 %
Stone formation
10-20 %
Bacterial colonization
100 %

>

OBS
<1%
<1%
15%

ICUD-EAU GUIDELINES ON BLADDER CANCER

CCD - OUTLET
Nipples abandoned
Appendix: Ready made
Availability, suitability
Stomal stenosis (1020%)
Tapered/plicated/stapled ileum: Simplicity
Serous lined extramural valve (T-pouch)
Sophisticated
No general acceptance

ICUD-EAU GUIDELINES ON BLADDER CANCER

CONDUIT
AND URETEROSIGMOIDOSTOMY

No / very little new information


Lack of hard data
No standarized reporting
Recent efforts to overcome this dilemma

Table 6: Diversion specific (not RCX) long term complications *


using non-/ standardized reporting
Conduit
Bern
f.u. median

(years):

patients

Neobladder
Mayo

Ulm

8.1

6.3

6.0

(n=)

131.0

1,057.0

923.0

complications

(n=)

192.0

1453.0

---

patients with

(n=):

87/131.0

643/1,057.0

patients with

(%):

66.0

61.0

within years:

5:

45

50

10:

50

68

15:

54

70

20:

94

80

Reoperation rate
(%):

40%

376/923.0

40.8

6%

Complications
(pts/%):
bowel

32

(24%)

UTI

30

(23%)

stoma

32

(24%)

anastomosis

18

(14%)

12

(9%)

35

(27%)

urolithiasis
renal function
*Bern :
Mayo:
Ulm:

> complications 3 months ; > 5 years follow up


> 30 days of surgery
> 3 months of surgery

215

(20.3%)

31

(3.4%)

(16.5%)

46

(5.0%)

163

(15.2%)

122

(11.5%)

102

(11.5%)

(15.3%)

(0.2%)

(20.2%)

174

162
213

Madersbacher J.Urol. 2003


Shimko
J.Urol. 2011
Hautmann
J.Urol. 2011

Table 6: Diversion specific (not RCX) long term complications *


using non-/ standardized reporting
Conduit
Bern
f.u. median

(years):

patients

Neobladder
Mayo

Ulm

8.1

6.3

6.0

(n=)

131.0

1,057.0

923.0

complications

(n=)

192.0

1453.0

---

patients with

(n=):

87/131.0

643/1,057.0

patients with

(%):

66.0

61.0

within years:

5:

45

50

10:

50

68

15:

54

70

20:

94

80

Reoperation rate
(%):

40%

376/923.0
40.8

6%

Complications
(pts/%):
bowel

32

(24%)

215

(20.3%)

31

(3.4%)

UTI

30

(23%)

174

(16.5%)

46

(5.0%)

stoma

32

(24%)

163

(15.2%)

18

(14%)

(11.5%)

102

(11.5%)

urolithiasis

12

(9%)

162

(15.3%)

(0.2%)

renal function

35

(27%)

213

(20.2%)

anastomosis

*Bern :
Mayo:
Ulm:

> complications 3 months ; > 5 years follow up


> 30 days of surgery
> 3 months of surgery

122

Madersbacher J.Urol. 2003


Shimko
J.Urol. 2011
Hautmann
J.Urol. 2011

ICUD-EAU GUIDELINES ON BLADDER CANCER

CONDUIT - CONCLUSION
Remains the most commonly used diversion in
conjunction with RCX
Technically easier than continent reconstruction
Complications, early as well as late, are legion
Incidence of upper tract complications increasing
with length of follow up
Comparisons with other techniques are difficult as
surgical techniques have improved markedly over
the last 25 years
Few series report comparable long-term outcomes
(>20 years) in patients with neobladders

ICUD-EAU GUIDELINES ON BLADDER CANCER

PEDIATRIC DIVERSION: INDICATIONS


Neurogenic bladder
Functional or anatomical loss of the lower
urinary tract
Failure of primary reconstruction of the
lower urinary tract (bladder exstrophy)
Radical surgery for malignancies

ICUD-EAU GUIDELINES ON BLADDER CANCER

PEDIATRIC DIVERSION GOALS

Urine storage at low pressures in


an adequate capacity reservoir for
preservation of the upper tract
Achieving urinary continence

ICUD-EAU GUIDELINES ON BLADDER CANCER

URINARY DIVERSION IN THE PEDIATRIC AGE GROUP


SPECIAL CONSIDERATIONS

Options for continent diversion in patients


with and without neurogenic deficits:
Augmentation and CCD
Anal diversion should be offered only to
patients with competent anal sphincters
Preferred solution for incontinent
diversion: colonic conduit

Table 7: Diversion specific long term complications (> 3 month


post-op) at 10 years in 923 neobladder pts operated
between 1986-2008 using the Kaplan-Meier method
Uretero-ileal stenosis
refluxive anastomosis (Wallace)
non refluxive
(Le Duc)

13,80 %
6,40 %
17,00 %

Urinary retention (female)

42,20 %

Subneovesical obstruction in men


(correctable)

8,20 %

Functional obstruction in men (CIC)

5,10 %

Incisional hernia
(41 in 923 pts = 4,4 %)

6,40 %

Hautmann, JUrol 2011

Table 5: Neobladder specific complications within 90 d of surgery


by category in 1013 patients (modified CLAVIEN)
category

no complication

infection

777 (76,7%)

abscess

988

UTI

837

Sepsis

978

GU complications

843 (83,2%)

renal failure

Grade 1

Grade 2

Grade 3

Grade 4

Grade 5

38

146

35

12

20

38

complications

Total
236 (23,3%)
25
176

138
6

15

12

35

100

64

170 (16,8%)

995

11

18

ureteral obstruction

965

16

31

48

urinary leak

948

44

21

65

urinary fistula

1002

11

urinary retention

985

23

28

miscellaneous

918 (90,6%)

metabol. Acidosis
despite TX

1007

incisional hernia

1010

Grade 1:oral medications


Grade 2: intravenous medications
Grade 3: interventional radiology or operation
Grade 4: lasting disability or organ resection
Grade: 5: death

95 (9,4%)

6
3

Hautmann JUrol.2010

ICUD-EAU GUIDELINES ON BLADDER CANCER

COMPLICATIONS OF UD
Conclusions:
Surgical morbidity following UD is significant and,
when strict reporting guidelines are incorporated,
higher than previously published
Accurate reporting of postoperative
complications after RCX is essential for:
counseling patients
combined modality treatment planning
clinical trial design
assessment of surgical success

ICUD-EAU GUIDELINES ON BLADDER CANCER

COMMITTEE 7
FINAL RECOMMENDATION FOR RCX AND UD

Centralisation / regionalization
High volume surgeons / hospitals
Annual case load of 25 procedures
Not more than 2 surgeons

PROSTATE INVOLVEMENT BY UROTHELIAL CELL


CARCINOMA

Joan Palou (Chair)


David Wood (co-chair)
Bernie Bochner
Henk van der Poel
H. Al-Ahmadie
Ofer yossepowitch

UROTHELIAL CELL CARCINOMA AND CARCINOMA


IN SITU OF THE PROSTATIC URETHRA
INCIDENCE
Primary bladder tumours: from 0.5 to 28 %

Depending on the population screened.


Palou et al, Urology 1995

The incidence of prostatic UCC in patients with superficial or invasive


bladder cancer oscillates between 12 and 48%, and between 7.6 and
16.6% of stromal invasion.

UROTHELIAL
CELL
CARCINOMA
IN SITU OF THE PROSTATIC URETHRA

AND

CARCINOMA

INCIDENCE
CIS
First described by Melicow in 1952

Rarely present as an isolated lesion


In a series of 1529 pats. with NMIBC it was 19 % bladder CIS and
2.7% CIS in PU (Millan et al, J Urol 2000)
It may be detected from 15 to 40 % in the followup of High grade
NMIBC (Herr HW et al J Urol 1999; Davis JW et al, J Urol 2002)

PAPILLARY UROTHELIAL CELL CARCINOMA


OF THE PROSTATIC URETHRA
INCIDENCE
Papillary lesions

Primary papillary lesions occur from 1-4 %


Bates HR,Jr, J Urol 1969
Multiplicity as a risk factor
Mungan MU et al, Eur Urol 2005

INCIDENCE OF PROSTATIC UC IN CYSTECTOMY SPECIMENS

Publication

Year

Patients

TCC in prostate (%)

J Urol

1977

300

12

Coutts et al

Brit J Urol

1985

43

36

Wood et al

J Urol

1989

84

23

Reese et al

J Urol

1992

115

29

Pagano et al

J Urol

1996

570

13

Esrig et al

J Urol

1996

489

29,2

Herr et al

J Urol

1999

186

39

Ngninkeu et al

J Urol

2003

283

27

Revelo et al

J Urol

2004

121

48

Nixon et al

J Urol

2002

192

15,6

Author
Schellhammer et al

STROMAL INVASION BY UROTHELIAL CELL CARCINOMA

INCIDENCE
Stromal invasion

Is present in 37-64 % of men with prostate involvement at


cystoprostatectomy
The differences of incidence between the different studies are

related to a careful pathologic assessment (whole mount section)


There is an increase risk of nodal metastases

Pagano et al, J Urol 1996

HISTOPATHOLOGY AND SURVIVAL


Stromal invasion
Stage

Patients (%)

Survival (%)

Contiguous

44 (61)

Not contiguous

28 (39)

46

Urethral mucosa

100

Acinal/ductal

14

50

Stromal

40

Extracapsular invasion

Pagano et al, J Urol 1996

Donat et al, et al, J Urol 2000

STAGING OF UCC OF THE PU

Stage 1

Tumour confined to the


prostatic urothelium

Stage 2

Duct and acini invasion,


but confined to the basal
membrane

Stage 3

Stromal invasion

Hardeman and Soloway, Worl J Urol 1988

CLASSIFICATION OF THE UCC OF THE PU

TNM BLADDER: Pt4a (Stromal invasion)


TNM PROSTATIC URETHRA:

Tis pu CIS, affecting the prostatic urethra


Tis pd CIS, affecting the prostatic ducts
T1

Tumour invading the connective subepithelial tissue

T2

Tumour invading the prostatic stroma, the corpus spongiosum

or

the periurethral muscle

T3

Tumour invading the prostatic capsule of the corpus

cavernosum or the bladder neck (extraprostatic extension)


T4

Tumour invading the neighbouring organs

ANATOMICAL AND HISTOPATHOLOGICAL HISTORY


RECOMMENDATION

TNM : pT4a only in cases of stromal invasion.


For the rest, the stage of the bladder tumour and of the PU
should be indicated using suffix Tis pu or Tis pd, meaning extension
towards prostatic urethra or towards prostatic ducts.

Extravesical invasion of the seminal vesicles is a sign of locally


advanced disease.

Volkmer et al, J Urol 2003

Daneshmand et al, J Urol 2004

DIAGNOSIS OF UCC OF THE PU

Macroscopic evidence of superficial bladder cancer of the prostatic


urethra is highly specific.
Nixon et al, J Urol 2002

CONSIDERATIONS:
Correct sampling of the PU: extensive TUR not feasible in all the

patients
Biopsy with the resectoscope loop from the lateral lobes distal to the
bladder neck until the verum montanum

In any suspicious lesion and recurrent high grade or bladder CIS

DIAGNOSIS OF UCC OF THE PU

Prostatic sampling plays an important role in the management of the


urothelial cell cancer

Prostatic involvement may become a predictor of understaging in BCG


failure in NMIBC.
Huguet et al, Eur Urol
2005

Eur Urol 2006

Huguet et al, Eur Urol 2006

Huguet et al, Eur Urol 2006

TREATMENT OF CIS OF THE PU (TIS PU)

BCG
Palou J, et al: Urology, 47: 482, 1996

Hillyard RW, et al: J Urol, 139: 290, 1988

BCG PENETRATION IN THE PU

Granulomas in the prostate up to 40 % after intravesical BCG


Mukamel et al, J Urol 1990

Increase of PSA after BCGin 41.6 % of the patients


Leibobici et al, J Urol 2000

Higher increase of PSA in BCG treated patients after TUR of the


prostate.
Palou et al, BJU Int 2000

EFFECT OF BCG ON PSA SERUM LEVELS

mo

TURp:

Yes

No

mos

Palou et al, BJU 2000

TREATMENT

MESSAGE:
High- and low-grade non-invasive UCC or CIS of the prostate should be
treated with endovesical BCG.
The TUR seems to improve BCG contact with the PU.

It seems that TUR improves efficacy of BCG endovesical treatment.

SUPERFICIAL UCC OF THE PROSTATE


(mucosa involvement)

LOW GRADE

HIGH GRADE or CIS

BCG

BCG

No recurrence

Recurrence

LOW GRADE

2nd COURSE
OF BCG

No recurrence

HIGH GRADE
or
CIS
Progression

Recurrence

CYSTOPROSTATECTOMY
Consider

TREATMENT OF CIS OF THE PROSTATIC URETHRA,


WITH PROSTATIC DUCT INVOLVEMENT (TIS PD)

BCG ?

TREATMENT OF CIS OF THE PU


WITH PROSTATIC DUCT INVOLVEMENT (TIS PD)

Soloway MS, et al. : J Urol, 167: 1573, 2002


Hardeman SW, et al. : World J Urol, 6: 170, 1988

TREATMENT PROSTATIC DUCTS

Palou et al, Urology 2007

Palou et al, Eur Urol 2006

RESULTS

BCG is a feasible therapeutic option for this kind of patients with TCC
affecting the prostatic ducts.
In this trial:

73% of the patients had bladder preservation

Only one death by tumour was observed

It is obvious that these patients need to be strictly followed with cystoscopy


and cytology to detect any recurrences and/or progression.
Palou et al, Eur Urol 2006

SUPERFICIAL UCC OF THE PROSTATE


Duct involvement

TUR + BCG

Re-TUR at 3 mos

NO TUMOUR

FOLLOW-UP

CYSTOPROSTATECTOMY

TUMOUR

PROSTATIC URETHRA FOLLOW-UP

Patients

Prostatic urethra
tumour

Follow-up

Herr and Donat,


1999

186

39 %
62 % not invasive
38 % stromal inv.

15 years

Hardeman et al.
1988

63

16 %

43 months

Solsona et al.
1991

276

13.3 %

34.3 months

Tumour incidence in the prostatic urethra when following-up superficial bladder


carcinoma.

TREATMENT: TIS IN PU AFTER BCG TREATMENT

Mean age: 72 years (63-82)


Mean follow up: 47 months (12-84)

PROSTATIC URETHRA FOLLOW-UP


Secondary tumours: In high-risk superficial bladder cancer it occurs in
some 10-15% at 5 years and some 20-40 % at 15 ys.
Herr et al, J Urol 1999

RECOMMENDATION
In patients with high-risk superficial bladder UCC and/or CIS, particularly

with bladder neck involvement and multifocality, the prostatic urethra


should be monitored.
Be careful with understaging

RECOMMEDATIONS
1. Macroscopic evidence of superficial bladder cancer in the prostatic urethra is
highly specific.
TUR of any macroscopic or suspicious lesion is required
Once a
non-muscle invasive high-grade tumor or CIS, careful follow-up of the prostatic
urethra is mandatory
(grade C).
2. High- and low-grade non-invasive urothelial carcinoma and CIS of the prostate
should be treated with intravesical BCG (grade C).
3. Evidence (level 3) shows that TUR may improve contact of BCG with the
prostatic urethra and it looks that response rates to BCG are increased (grade C).

RECOMMEDATIONS

4. CIS or tumor in the prostatic ducts warrants further study. It is advisable for the
clinician to perform radical surgery if there is any doubt, to avoid understaging
(grade C).
5. Patients with non-invasive urothelial carcinoma in whom conservative therapy
fails should be considered for cystoprostatectomy (grade C).
6. Patients with intermediate- to high-risk superficial urothelial carcinoma of the
bladder or CIS, especially with involvement of the bladder neck and multifocality,
require monitoring of the prostatic urethra (grade B or C).

PROSTATIC STROMAL INVASION


DIAGNOSIS

Transurethral biopsies of the prostatic urethra have shown to be efficacious for


identifying prostatic involvement, but they do not reliably determine the extension of

such involvement, particularly regarding stromal invasion.


53 % sensitivity and 77 % specificity in the diagnosis of stromal invasion.
New methods are necessary to detect prostatic stromal invasion.
Wood et al, J Urol 1989
Donat et al, J Urol 2001

Pagano et al, J Urol 1996

PROSTATIC STROMAL INVASION


PREDICTIVE FACTORS

CIS at the trigone


Tumor multifocaclity

Location of the tumour at or below the trigone (or at the bladder neck)
Lymphovascular invasion

Pettus et al, Eur urol 2008

Richards et al, Urology 2010


Arce et al, Canad J urol 2011

TREATMENT

ROLE OF CYSTECTOMY
RECOMMENDATION
Grade C
Radical cystoprostatectomy with lymphadenectomy, with or without
urethrectomy, is the choice treatment for loco-regional control in patients
with prostatic invasion.

TREATMENT

ROLE OF LYMPHADENECTOMY
FACTS
There is a 40-50 % incidence of positive nodes in a pT4a tumour.
Data exist that show lymphadenectomys potential impact on survival.

See: Invasive bladder cancer

RECOMMEDATIONS: STROMAL INVASION

1. The incidence of prostatic urothelial carcinoma in men with superficial or


invasive bladder cancer ranges from 12% to 48%, and 7.6% to 16.6% have
stromal invasion. The prostate is a site of relapse for patients with non-muscle
invasive bladder cancer (grade C).
2. Transurethral biopsy of the prostatic urethra is effective in identifying prostatic
involvement but does not accurately reveal the extent of stromal invasion
(grade C).
3. Retrospective and prospective studies are needed to determine better
prognostic factors for prostatic stromal invasion (grade C).

RECOMMEDATIONS: STROMAL INVASION

4. Radical cystoprostatectomy is the treatment of choice for locoregional control in


patients with prostatic stromal invasion (grade B).
5. In patients with pT4 disease, the incidence of positive nodes ranges from 40%
to 50%, and node mapping studies indicate that multiple sites are involved (grade
C).

6. Data show that the extent of lymphadenectomy may have an impact on survival
(grade C).

RECOMMEDATIONS: STROMAL INVASION

7. The number of patients with prostatic stromal invasion treated in radiation


therapy trials is too small to allow definitive conclusions regarding survival
outcome (grade C).
8. Although the data are limited and the number of patients with prostate stromal
invasion cannot be determined from the literature, the use of radical cystectomy
as a salvage procedure does not appear to diminish disease-specific or overall
survival probability (grade C).

Chemotherapy for metastatic transitional cell


carcinoma of the urothelium
Cora N. Sternberg, San Camillo Forlanini Hospital, Rome, Italy
Guru Sonpavde, US Oncology Baylor College, Webster
Walter M. Stadler, University of Chicago, Chicago
Dean Bajorin, Memorial Sloan Kettering Cancer Center, New York
Robert Dreicer, Cleveland Clinic, Cleveland
Daniel George, Duke University School of Medicine, Durham
Matthew Milowsky, Memorial Sloan Kettering Cancer Center New York
Dan Theodorescu, University of Colorado Cancer Center, Denver, Colorado
David Vaughn, University of Pennsylvania, Philadelphia
Matthew D. Galsky, Mount Sinai, New York
David Quinn, University of Southern California, Los Angeles
ICUD March, 2011

First line randomized trials in


advanced transitional cell carcinoma
Author

Treatment

RR (%)

MDS (mo)

Best arm

Loehrer

M-VAC
CDDP

126
120

39
12

12.5
8.2

M-VAC >CDPP

Logothetis

M-VAC
CISCA

65
55

65
46

12.6
10.0

M-VAC >CISCA

Von der Maase

M-VAC
GC

202
203

46
49

14.8
13.8

M-VAC ~ GC

HD-MVAC
M-VAC

134
129

62
50

14.5
14.1

HD-M-VAC >
M-VAC

Bamias

M-VAC
DC

109
111

54
37

14.2
9.3

M-VAC > DC

Dreicer

M-VAC
PC

44
41

36
28

15.4
13.8

M-VAC > PC

PCG
GC

312
315

57.1
46.4

15.7
12.8

PCG ~ GC

Sternberg

Bellmunt

MDS: median duration of suvival. M-VAC: cisplatin, methotrexate, adriamycin, vinblastine. CDDP: cisplatin.
CISCA: cisplatin, adriamicin, cyclophosphamide. GC: gemcitabine, cisplatin. HD-MVAC: high dose M-VAC. DC:
docetaxel, cisplatin. PC: Paclitaxel, carboplatin. PCG: paclitaxel, cisplatin, gemcitabine.

Better patient selection, earlier diagnosis and screening, better


supportive care (growth factors)

High Dose M-VAC vs M-VAC


Study design (n=263)
Classic M-VAC

Stratify
Who PS

R
A
N
D
O
M
I
Z
E

MTX: 30mg/m2 DAYS 1, 15, 22


VLB: 3mg/m2 DAYS 2, 15, 22
ADM: 30mg/m2 DAY 2
CDDP: 70 mg/m2 DAY 2
Q 28 d

HD- M-VAC + G-CSF


MTX: 30mg/m2 DAY 1
VLB: 3mg/m2 DAY 2
ADM: 30mg/m2 DAY 2
CDDP: 70mg/m2 DAY 2
Q 14 d
Sternberg CN, J Clin Oncol 2001;19(10):2638-1646

High Dose M-VAC vs M-VAC


7 Year update of Overall Survival

100

90

Unadjusted p=0.0417
HR=0.76 (0.58 0.99)

80
70

35% survival

60
50

40

HD-M-VAC: median 15.1 months

30
20
10

M-VAC: median 14.9 months

25% survival

0
0
O N
112129
101134

Number of patients at risk :


32
15
11
45
29
23

10

(years)
12
Treatment

4
8

2
0

M-VAC
HD M-VAC

Sternberg CN, Eur J Cancer 2006 Jan;42(1):50-4

M-VAC with G-CSF support


Hellenic trial of M-VAC vs docetaxel and cisplatin:
classic M-VAC with appropriate hematopoietic
growth factor support is an effective treatment.
(level of evidence 1b)
EORTC trial of HD-MVAC vs M-VAC is another
excellent option that may potentially lead to higher
long term cure rates.
(level of evidence 1b)

Doublet chemotherapy
GC vs M-VAC - 5 year update
Overall Survival
GC

14.0 months (12.3-15.5 )

M-VAC

15.2 months (13.2-17.3 )


HR: 1.09 (0.88-1.34)

Von der Maase H, J Clin Oncol 2005 Jul 20;23(21):4602-8

EORTC30987/Intergroup study design

T4bN0M0 or TxN2-3
or M1 TCC of
urothelium
No prior
chemotherapy

R
a
n
d
o
m
i
s
e

Gemcitabine: 1000 mg/m2 days 1, 8 and 15


Cisplatin :
70 mg/m2 day 1 or 2
Arm 1 is given as a 4 week cycle (28 days)
Every 21 days if d15 is withheld or missed
Paclitaxel
Cisplatin
Gemcitabine

80 mg/m2 on days 1 and 8


70 mg/m2 on day 1
1000 mg/m2 on days 1 and 8

Arm 2 is given as a 3 week cycle (21 days)

Opened in May 2001. Closed in June 2004


Centres activated: 107. Patients entered/required: 627/610
Treatment: until disease progression; a max. of 6 cycles

Bellmunt J, et al. J Clin Oncol 2007;25 (Suppl 18): 242s. Abst LBA5030

Median FU = 3.2 years

Overall Duration of Survival

Maximum FU = 5.5 years

100
90

Gem/Cis

247/315

12.8 mo

Pac/Cis/Gem

239/312

15.7 mo

0.86 (0.72-1.03)

80

Overall Logrank test: p=0.10


70
60

14% r. reduction risk of death (n.s.)

50
40
30
20
10
(years)

O
247

N
315

Number of patients at risk :


159

76

34

239

312

185

86

35

13

6
Treatment

0
2

Gem+Cis
Gem+Cis+Pac

Survival for all patients grouped according to


number of risk factors present at baseline
n=199
Risk factors:
0= KPS > 80, no visceral mets
1= KPS < 80 or visceral mets
2= KPS < 80 and visceral mets

33.0 m.
13.4/13.6 m.

9.3 m.

Bajorin, D. F. et al. J Clin Oncol; 17:3173-3181 1999


Copyright American Society of Clinical Oncology

First line chemotherapy for platinum


eligible patients
Recommended as first line therapy
M-VAC, HD-MVAC and GC
(level of evidence 1b)
Triplet PCG not recommended as first line
therapy (level of evidence 1b)

Cisplatin ineligible patients


The majority of patients with advanced urothelial
cancer are older (median age, 68 years)
Renal function declines with age

Alternative non-cisplatin based therapy is needed


Approximately 40% of patients with advanced
urothelial cancer are cisplatin-ineligible due to renal
dysfunction alone
The proportion of cisplatin-ineligible patients is much
higher when including treatment-limiting factors such
as poor performance status and comorbidities

Defining cisplatin-ineligible
patients with metastatic bladder
cancer
M. D. Galsky, N. M. Hahn, J. E. Rosenberg, G.
Sonpavde, W. K. Oh, R. Dreicer, N. J. Vogelzang, C.
N. Sternberg, D. F. Bajorin, J. Bellmunt, CisplatinIneligibility in Bladder Cancer Working Group

28% of all patients had CrCl < 60 ml/min


40% of patients > 70 had CrCl < 60 ml/min

40

60

68

36

20

16
7

Proportion ineligible

80

100

A large proportion of patients are unfit for cisplatinbased therapy

Age < 60

Age 60-70

Age 70-80

Age 80+

Dash et al, Cancer, 2006

EORTC 3098
Randomize

At least one of the following:


WHO PS 2
CrCL > 30 mL/min and <
60 mL/min

VINCENT
Randomize

At least one of the following:


NYHA Class III-IV CHF
CrCL of 60 mL/min

Gemcitabine
Carboplatin

Methotrexate
Carboplatin
Vinblastine

Vinflunine
Gemcitabine

Placebo
Gemcitabine

Hypothesis

Tremendous variability exists in the definition


of unfit patients
A uniform definition of unfit patients will lead
to:
more uniform clinical trials
a greater likelihood of developing a viable
strategy for regulatory approval

Types of Definitions
Pick and Choose

At least one of the following:


WHO PS 2
Creatinine clearance of <
60 mL/min

Traditional
inclusion/exclusion

All of the following:


Karnofsky performance
status 60%
Creatinine clearance 30
mL/min
Cardiac ejection fraction
40%

Survey of GU Medical Oncologists


Survey sent on three occasions over 3 weeks

65/120 (54%) completed survey


The majority (62%) cited prior involvement with
clinical trials in unfit patients

What renal function threshold should be used?

N=65

How should renal function be measured?

N=65

How should renal function be measured?

N=65

What age threshold should be used?

N=65

What PS threshold should be included?

N=65

What comorbidities should be included?


(neuropathy)

N=65

Proposed eligibility criteria for clinical trials


enrolling unfit patients
At least one of the following:
1. ECOG PS of 2 (KPS of 60-70)

2. Creatinine clearance < 60 ml/min


3. CTCAE v4 Grade 2 audiometric hearing loss
4. CTCAE v4 Grade 2 peripheral neuropathy
5. NYHA Class III heart failure

Conclusions
Substantial heterogeneity exists among criteria
defining unfit patients

A simple and concise definition of unfit is


proposed for use as eligibility criteria for clinical
trials moving forward.

Randomized phase II/III study assessing


gemcitabine/carboplatin (GC) and
methotrexate/carboplatin/vinblastine (M-CAVI)
in previously untreated patients (pts)
with advanced urothelial cancer unfit for
cisplatin based chemotherapy:
phase III results of EORTC study 30986
M. De Santis 1, J. Bellmunt 2, G. Mead 3, J.M. Kerst 4, M. Leahy 5,
G. Daugaard 6, T. Gil 7, P. Maroto 8, S. Marreaud 9, R. Sylvester 9
1 Kaiser

Franz Josef - Spital and ACR-ITR VIEnna, Vienna; 2 Hospital Vall d'Hebrn, Barcelona;
3 Royal South Hants Hospital, Southampton; 4 Netherlands Cancer Institute, Amsterdam;
5 St James Hospital, Leeds; 6 Rigshospitalet, Copenhagen; 7 Institut Jules Bordet, Brussels;
8 Hospital Santa Creu, Barcelona; 9 EORTC Data Center, Brussels

Phase II/III EORTC randomized trial in


unfit metastatic bladder patients
GCa; gemcitabine & carboplatin
M-CAVI; methotrexate, carboplatin & vinblastine
Performance status WHO 2 and/or
Impaired renal function
(30 ml/ min GFR < 60 ml/min)

De Santis M et al, J Clin Oncol 2010 (suppl: abstr LBA 4519)


De Santis M et al, J Clin Oncol 2009, 27, no. 33, 5634-5639

Phase III results of EORTC study 30986

First valid PFS data in this patient population


100
90
80
70
60
%
50
40
30
20
10
0
0
O N
113 119
115 119

HR=1.04 (95%CI: 0.80, 1.35)


p=0.78

4.2 months (95%CI: 3.7, 5.9)


5.8 months (95%CI: 4.8, 6.9)

(years)
1
2
3
4
5
Number of patients at risk :
20
5
5
3
1
15
3
3
2
2

6
1
1

8
Treatment
1
M-CAVI
1
GC

De Santis M et al, J Clin Oncol 2010 (suppl: abstr LBA 4519)

Phase III results of EORTC study 30986

First valid OS data in this patient population


100
90
80
70
60
%
50
40
30
20
10
0
0
O N
108 119
110 119

HR=0.94 (95%CI: 0.72, 1.22)


p=0.64

8.1 months (95%CI: 6.1, 10.3)


9.3 months (95%CI: 7.6, 11.3)

(years)
1
2
3
4
5
Number of patients at risk :
37
13
7
3
1
44
15
5
2
2

1
1

8
Treatment
1
M-CAVI
1
GC

De Santis M et al, J Clin Oncol 2010 (suppl: abstr LBA 4519)

GCa vs M-CAVI in poor PS or CrCl< 60 ml/min


The first randomized phase III - trial comparing
GCa and M-CAVI in cisplatin ineligible patients
Patients ineligible for cisplatin, benefit from
carboplatin based combination chemotherapy
although toxicities were important
Both combinations were active in this well defined
group of unfit patients, although toxicity was
higher on the M-CAVI arm
Patients with CrCl 40-60 and those with poor PS
could be evaluated separately
Sternberg CN, ASCO discussion 2010

GCa vs M-CAVI in poor PS or CrCl< 60 ml/min

This trial provides level 1 evidence for the use


of GCa chemotherapy in patients deemed
ineligible for cisplatin-based chemotherapy,
Adoption of a standard definition of cisplatinineligible and additional randomized studies
of chemotherapy and novel agents in this
patient population are desperately needed

2nd line chemotherapy: single agent phase II trials


Author and year
of publication

Molecule

No. of
patients
evaluated

ORR
(%)

Median
survival time
in months

Blumenreich, 1982

Vinblastine

37

18

NR

Pronzato, 1997

Ifosfamide

20

0.8

Witte, 1997

Ifosfamide

56

20

5.1

Mc Caffrey, 1997

Docetaxel

30

13

9.0

Papamichael, 1997

Paclitaxel

14

NR

Vaughn, 2002

Paclitaxel

31

10

7.2

Joly, 2009

Paclitaxel

6.8

Lorusso, 1998

Gemcitabine

31

23

5.0

Gebbia, 1999

Gemcitabine

24

29

13.0+

Albers, 2002

Gemcitabine

28

11

8.7

Witte, 1998

Topotecan

44

6.3

Dodd, 2000

Pyrazoloacridine

14

9.0

Roth, 2002

Piritrexim

27

7.0

Wulfing, 2009

Lapatinib

59 (34)

4.0

Moore, 2003

Oxaliplatin

18

NR

Sridhar, 2005

Bortezomib

11

NR

Gomez-Albuin, 2007

Bortezomib

19

3.8

Rosenberg, 2008

Bortezomib

25

5.7

Dreicer, 2007

Epothilone B

45

12

Sweeney, 2006

Pemetrexed

47

28

9.6

Galsky, 2007

Pemetrexed

12

NR

Gallagher, 2007

Sunitinib

45

NR

Dreicer, 2009

Sorafenib

27

6.8

45 (37)

Phase III trial of Vinflunine plus Best Supportive


Care vs BSC alone (n=370)
2nd line
T4bN0M0 or
TxN2-3 or M1
Progression after 1st
line Platinum
treatment

Stratify
Center
Refractory vs. nonrefractory

R
A
N
D
O
M
I
Z
E

VFL+BSC (PS 0: 320mg/m, q3w; PS


0 with previous pelvic irradiation
and PS1: 280mg/m (n=253)

BSC with treatment upon


progression permitted (n=117)

2:1

Primary endpoint: Overall survival


Secondary Endpoints: ORR, PFS, Disease control, QOL and Clinical Benefit
Bellmunt J, J Clin Oncol. 2009 Sep 20;27(27):4454-61

Phase III trial of Vinflunine plus Best Supportive


Care vs BSC alone: OS

N=357

N=370

ITT population

Eligible patients
Bellmunt J, J Clin Oncol. 2009 Sep 20;27(27):4454-61

Overall Survival: Intent to Treat Population (N= 370)

Survival Distribution Function

7 month median survival is a benchmark in the second line setting!


VFL + BSC
N= 253

BSC
N= 117

No. of events

204

103

No. censored (%)

49 (19.4)

14 (12.0)

Median OS
(months)
(95% Cl)

6.9
(5.7, 8.0)

4.6
(4.1, 7.0)

VFL + BSC
BSC

Hazard ratio (95%


CI)

0.88 (0.69, 1.12)

p value *

0.2868

Overall survival (months)


* Stratified log rank test
Bellmunt J, J Clin Oncol. 2009 Sep 20;27(27):4454-61

Vinflunine - second line therapy


Based on this study, vinflunine was approved by the
European Medicine Agency (EMA)
The only agent approved by a regulatory agency for this
indication.
A retrospective analysis of patients that received
second-line vinflunine identified ECOG performance
status >0, hemoglobin <10g/dL and liver metastasis as
poor prognostic factors

Bellmunt J, J Clin Oncol 2010;28(11)1850-5

Combination therapy as second-line therapy


Paclitaxel and gemcitabine:
German randomized phase II trial of 102 patients
compared 6 cycles of second-line gemcitabinepaclitaxel with continuation beyond 6 cycles until
progression
Italian trial of every 2 week gemcitabine-paclitaxel
Carboplatin-paclitaxel following prior cisplatinbased chemotherapy
Albers P, Ann Oncol. 2011 Feb;22(2):288-94.
Sternberg CN, Cancer; 2001 Dec 15;92(12):2993-8.
Vaishampayan UN, Cancer 2005, 104(8):1627-32

Second-line chemotherapy
Vinflunine: Level 2a and Grade B
recommendation
Clinical trials should remain the preferred option

Novel agents: ongoing or planned phase II trials of


salvage therapy for metastatic TCC
Drug/Regimen

Molecular target

Institution

Nab-paclitaxel

Tubulin

Canadian

AZD-4877

Kinesin spindle protein

Multicenter

Pralatrexate

Folate

Multicenter

Pazopanib

VEGFR2, PDGFR

Italy, NCI (US)

Tamoxifen

Estrogen receptors

Baylor

Everolimus

mTOR

Belgian

Everolimus-Paclitaxel

mTOR and tubulins

SWOG

Aflibercept

VEGF, PlGF

NCI (US)

TKI-258 (Dovitinib)

VEGFR, PDGFR, FGFR

Multicenter

Vorinostat

HDAC

California

Docetaxel-ASA404

Tumor and endothelial tubulin

Mt. Sinai/HOG

BI-6727

Polo-like kinase

Multicenter (US)

Phase II trials with HER 1/2 inhibition


Author

Regimen

Results

Comments

Petrylak 2003
(SWOG)

2nd line Gefitinib N=29, 3%RR 4%


PFS at 6 mos

Minimal effect
with single agent
Gefitinib

Philips 2008
(CALGB)

Fixed Dose-Rate
Gemcitabine,
DDP + Gefitinib

No better than
GC with
excessive
toxicity

Hussain 2007

HER 2 positive
N=44/57 70% RR, No Phase III trial
Trastuzumab,
OS 14.1 mos
Paclitaxel, Carbo
+ Gemcitabine

Wlfing, 2009

2nd line
Lapatanib

N=25, 36% RR
OS 11.1 mos

N=59, 1.7% OR,


SD 31% med
TTP 8.6 wks, OS
17.9 wks

Clinical benefit
correlated with
EGFR and HER-2
overexpression

Modified from Sternberg C, ASCO Education 2008

Trials with HER 1/2 Inhibition


Advanced Disease

Maintenance or
Consolidation

Lapatinib + GC
phase I (EORTC)

Lapatinib (Powles)

GC +/- Cetuximab
(Hussain)

Docetaxel +/- Gefitinib


(Siefker-Radtke,
MDACC)

Neo adjuvant
Erlotinib (SiefkerRadtke, MDACC)

Cetuximab +/Paclitaxel (Fox


Chase)

Modified from Sternberg C, ASCO Education 2008

Stage IV bladder cancer


4-8 cycles of
1st line chemotherapy

HER1/HER2
status

Negative

(European Consortium)

Maintenance
HER1/HER2 positive
Responders
Randomized
Selected

Positive

Exclusion criteria

Response to
chemo

PD

Increase PFS from the


time of completion of
chemo. from 6 to 9 mos

CR/PR/SD

Randomisation

Lapatinib

n=204
Placebo

Progression free survival (primary end point)


Follow up for survival
Coordinator, T Powles

Targeting angiogenesis in urothelial cancer


VEGF, bFGF and IL-8 levels correlate with stage and
outcome
MVD: predictor of progression, vascular inv, N+, recurrence,
and S in invasive TCC

VEGF levels in the metastatic setting appears to correlate


with poor DFS

VEGF mRNA levels and MVD independent prognostic factors


for recurrence and M1 in neoadjuvant M-VAC

Elfiky AA and Rosenberg JE, Curr Oncol Rep. 2009 May;11(3):244-9

Single Agent Angiogenesis Inhibition in TCC


Presented

Advanced

Ongoing

Advanced

Bellmunt
ASCO GU 2008

1st line Unfit


Sunitinb
14.3% PR, 78.6%
benefit and 6 mos
TTP

Necci ESMO
2010)
Pili ( ASCO GU)

2nd and 3 line


Pazopanib (TKI)

Gallagher ASCO
2008/ 2009*

2nd line Sunitinb


7% PR, 24% Stab
but 17% PFS > 3
mos. Cont dose:
less activity*

California
Cancer
Consortium

2nd line
VEGF Trap (antiVEGF antibody
with broad
activity)

Dreicer (ECOG)
ASCO 2008

2nd line Sorafenib


no significant
activity

Hussain
Consortium
ASCO 2009

Maintenance
Randomized
Phase II Sunitinib
Modified from Sternberg C, ASCO Education 2008

Advanced Combination Chemotherapy and


Anti-angiogenesis Trials
Presented

Phase II

Ongoing

Phase II/III

Hahn (HOG)
ASCO 2009

Cisplatin +
Gemcitabine +
Bevacizumab

Choueiri
(DFCI)

2nd line
Docetaxel +/ZD6474
(Zactima)

Bajorin
(MSKCC)
ASCO 2009
(tox)

Carbo + Gem +
Bevacizumab

Kelly (Jefferson)

Carboplatin +
Gem+Sorafenib

Rosenberg
(CALGB)
Phase III

Cisplatin +
Gemcitabine +/Bevacizumab

Modified from Sternberg C, ASCO Education 2008

Personalized therapy using emerging


technologies
Pharmacogenomics (blood, tumor)
Genomic profiling (molecular profiling)

Genomic profiling/drug sensitivity: i.e. CO-XEN


Mutational analysis (high-throughput)

Proteomics still in its infancy


Micro RNA
Circulating tumor cells

Pharmacogenomics in bladder cancer


Differences in DNA repair capacity modify
individual susceptibility to carcinogenesis, but
also affect individual response to therapy
Nucleotide excision repair (NER) is one of the
major DNA repair pathways ( ERCC1, BRCA1)
The product of ERCC-1 gene functions in the
NER pathway and is a molecular marker of
platinum resitance

Predictive factors of response:ERCC1


Nucleotide excision repair
Excision repair cross
complementing 1 (ERCC1)
predicts response and
resistance to cisplatin

57 patients with advanced TCC,


Rx with GC or the triplet
containing paclitaxel and GC

High mRNA

Low mRNA

Significant correlation
between relative gene
expression levels > 7 and
decrease in survival (p=
0.035)
Bellmunt J, Ann Oncol. 2007 Mar;18(3):522-8

Co-expression extrapolation (COXEN)


algorithm
Predicting clinical outcomes by using
molecular signatures (i.e. breast cancer)
Incorporates gene profiling with drug sensitivity

Validates the clinical setting with patient


outcomes
Forecasts drug responses
Lee JK, Proc Natl Acad Sci U S A. 2007 Aug 7;104(32):13086-91

Resistant

Evaluation of COXEN treatment response


biomarkers in human bladder cancer
REFERENCE Sets

2
3

2
3

1
2

2
3
1

1
2

575A

1
2

BLA-40 is 40 Expression
Human Bladder Profiling
Cancer Cell Lines
mRNA profiled

1
3
2

1
2

1
2

1
3
2

IC50 for Test


Compounds

1
2

1
2

1
2

2
1

2
1

1
2
1
2

1
2

Sensitive

1
x
1
2

1
2

1
2

1
2

1
2

1
2
1
2

1
2

Resistant

1
2

Bladder Tumor
samples
from patients

GI50
Actual
Resistant:
log(GI50)
resistent
resistant
Prediction
COXEN
Resistant:
predicted
resistent

resistant

1.0
0.8

Predicted non-responders

(P-value
= 0.021)
(p-value
= 0.021)

0.0

COXEN Score

Predicted responders

0.6

Prediction
COXEN
Sensitive:
predicted
sensitive

0.4

(p-value = 0.016)
GI50
Actual
Sensitive:
log(GI50)
sensitive

0.2

umuc9
X253jp
slt4p3
X253jbv
umuc14
crl7833
fl3p10
ku7
umuc3
umuc3e
rt4
crl7193
crl2169
ht1197
X575a
cubIII
mghu3
jon
kk47
bc16.1

Bladder
Tumors
Fraction Disease-Free

BLA-40 Cell line Cisplatin normalized log(GI50) & MiPP prediction scores

-1.5

-1.0

-0.5

0.0

0.5

1.0

1.5

2.0

Predict: in vitro
Drug Sensitivity

50

100

150

Survival
time
Time
(months)

Standardized log(GI50)
Standardized COXEN Score

VALIDATION Sets

Resistant

1
2
1
2

1
2

Drug Response Prediction Model


(Gene Expression)
Evaluation of Model on
Cells or Tumors
(RT-PCR)

2
1

Predict: Therapy
Response

Focus on the subset of probes that consititutes a signature of drug sensitivity!

1
2

SW1710
x

COXEN

Bladder
Cell Lines

1
2

KK47

Resistant

CRL2742
x
1
2

1
2

1
2

2
1

VMCUB1

2
1

Resistant

x
2
1

2
1

1
2
1
2

UMUC1

Sensitive

1
2

2
1

Sensitive

x
1
2

UMUC3

HTB9

Sensitive

x1
3
2

2
1

x
2
1

NCI-60 Panel
3
2
1

2
1

J82

Sensitive

2
1

2
1
RT4
1 Sensitive
2
3
x 1 1

1
2

2
1
3

1
2
3

1
2

1
2
3

2
3

Resistant

Sensitive

x
1

Sensitive

1
x
2

Sensitive

x
1
3
2

2
3
1

1
2

1
2

1
2
1
2

Validation of new drug effectiveness in human


bladder cancer cells: 45,545 compounds tested

1
2
3

RT4
1 Sensitive
2 1 1
x3

2
3
1

1
x
2

1
2
2
1

1
2
1
2

575A

1
2

1
3
2

1
2

1
3
2

1
2

1
2

2
1

2
1

2
1

2
1

1
x

1
2

1
2

Resistant

1
2

1
2

1
2

2
1

1
2

2
1
1
2

1
2

Resistant

1
2

Resistant

1
2

SW1710

1
2
1
2

1
2

2
1

2
1

CRL7193
Resistant

x
1
2

1
2

1
2

CRL7833
2
1

2
1
1
2

1
2
1
2

253J-P
Resistant

1
2

1
2

1
2

KK47

1
2

Resistant

x
1
2

1
2

Resistant

CRL2742

Resistant

1
2
2
1

1
2

1
2

HT1376

1
2

1
2

1
2
1
2

1
2

Resistant

1
2

Resistant

x
1
2

253J-BV

Resistant

1
2

1
2

2
1

T24

1
2
x

UMUC13D

UMUC1

2
1

1
2

1
2

1
2
1
2

2
1

1
2
1
2

1
2

KU7

1
2
2
1

1
2

1
253JLaval 2
Resistant
1
x
2
1
1
2 2

Resistant

x
2
1

VMCUB1

1
2

1
2

Sensitive

Sensitive

1
2

1
2

1
2

UMUC9
Resistant

UMUC3

1
2

2
1

Sensitive

x
1
2

2
1

CUBIII

1
2

2
1

1
2

Sensitive

HTB9

Sensitive

x1
3
2

2
1

1
2

Sensitive

MGH-U4

1
2

J82

x2
1

1
2

x
1

Sensitive

2
3

60 100

2
3
1

Sensitive

1
2
3
2
1

1
2

HU456

2
1

1
2

1
2

Sensitive

1
2

1
2

1TCCSUP 1
2
Sensitive
2
x
1 1

2
1

MGH-U3

1
2

1
2

1
2

Resistant

1
2

2
1
2

UMUC2

1
2

x
1

1
2

1
2
3

Sensitive

x
1
3
2

20

with C-1311

Sensitive

SLT4

2
3
1

BC16.1
x
1
2

1
2

1
2

Sensitive

1
2
1
2

1
2

1
2
3

2
3

Bladder Cancer Clinical Trial

1
2

1
2
3

UMUC14

T24T

1
2
3

2
1
3

1
2

Sensitive

x
2
1
3

1
2
3

1
2
3

3
2

2
3
1

1
2

40 human bladder
cancer cell lines
(BLA-40)

2
1
3

2
3

NSC 637993

Percent of cell counts

2
3

2
1

JON

2
1

Sensitive

1
2

1
2
3

Sensitive

x
2
3
1

x
1

2
3
1

UMUC6

Sensitive

1
3
2

2
3
VMCUB3 1

1
2
3

1
2
3

1
2
3

1
3
2

1
2
3

3
2
1

2
1

2
HT1197 1
Sensitive
1
2
1
2 2
1

1
2

1
2

Sensitive

VMCUB2

2
3
1

2
1
3

FL3

Sensitive

2
3

3
2
1

1
2

1
2

-8 -7 -6 -5 -4

Log base 10 of molar concentration of NSC 598354


637993
Courtesy of D. Theodorescu

51

Gene expression
model

Genome-wide gene expression profiling


Predicting response to neoadjuvant M-VAC

Takata R, Aktuelle Urol. 2010 Jan;41 Suppl 1:S41-5


Takata R Clin Cancer Res 2005;11, 2005

Genetic signatures
associated with
.
aggressive behavior

High
throughput
DNA
microarrays
identify
biomarkers for
bladder cancer

Sanchez-Carbayo M, J Clin Oncol. 2006 Feb 10;24(5):778-89

Gene expression based or immunohistochemical


detected biomarkers
Promising biomarkers have not yet been
extensively validated in sufficient numbers in
prospective clinical trials
Could become part of the diagnostic workup
and management in the future
Grade of recommendation: Not
recommended outside a clinical trial or
biomarker study (level of evidence 2b)

Conclusions for first line therapy


First-line advanced urothelial cancer with adequate renal
function: combination cisplatin regimens are the standard
of care (Grade A recommendation).
Use of triplet regimens adding paclitaxel to gemcitabine
plus cisplatin is not recommended.
Unfit patients: renal impairment, advanced age or poor
performance status: carboplatin and gemcitabine
(Grade A recommendation).
Renal impairment only gemcitabine/carboplatin/paclitaxel
can be considered (Grade C).

Conclusions for second line therapy


For second line therapy after a platinum based
treatment, vinflunine has been adopted as a
standard in Europe but not in North America.
(Grade B recommendation).

Novel agents and approaches are urgently needed


and clinical trial accrual needs to be actively
encouraged.
The standard of care for a majority of patients with
advanced bladder cancer is the best clinical trial
available.

Non-urothelial cancer of the Bladder


Bruce R. Kava (USA) sub committee chair
Jack Baniel (Israel)- presentation

Hassan Abol-Enein (Egypt)


Alexandra J Colquhoun (USA)

William H Turner (UK)


Adrienne J. K. Carmack (USA)

Non-urothelial Bladder Cancer


TCC is the most prevalent bladder tumor (superficial
and invasive)
The other end of bladder tumors:
Squamous Cell Carcinoma
Adenocarcinoma
Small cell carcinoma
Others
(Level 3 evidence)

Worldwide Incidence
Urothelial and Nonurothelial Bladder Cancer

Squamous Cell Carcinoma (SCC)


Two main categories:
SCC in western countries (non-bilharzial)
SCC in the bilharzial bladder
(Rundle,1982) (Level 3 evidence)

Non-bilharzial SCC of the Bladder


Epidemiology
2nd bladder malignancy
age 70 years
2-5% in contemporary cystectomy series
Black to white 2:1
Male to female ratio 1.4:1
Presents at advanced stage
(Serratta,2000) (Level 3 evidence)

Non Bilharzial SCC of The Bladder


Gender Differences

In USA:
incidence of urothelial ca in men >3x women
incidence of non-urothelial ca in men =1.4x women

(Johansson SL and Cohen SM 1997: compiled data from 915 patients in 10 series)

Level 3

Women present with more advanced stage tumors:

Mungan AL, et al (2000):


Data derived from the
Netherlands Cancer Registry:
Level 3

Non Bilharzial SCC of The Bladder


Etiology
Relation to cigarette smoking is not clear.
*Chronic inflammatory disorders.
*Calculi, diverticula, indwelling catheters.
Spinal cord injury patients.
Squamous metaplasia is common.
(Thorn,1997) ( Level 3 evidence)
(- *Level 4 evidence)

Non Bilharzial SCC of The Bladder


Association with Spinal Cord Injury
Incidence of SCC in Spinal Cord Injured patients:
2.3- 10% in early, small retrospective reports
Kaufman et al (1977), Locke et al (1985), Bejany, et al (1987)

Level 3

West et al (Urology 1999):US Dept of Veterans Affairs Admissions


Data:
33,560 patients with SCI identified
130 (0.39%) patients with Bladder cancer identified
of: 42 records available for review:
May reflect: increase in cigarette
23 (55%) Urothelial ca
smoking and reduced indwelling
13 (33%) SCC
catheters
Level 3
4 (10%) Adenoca

Non Bilharzial SCC of The Bladder


Association with Spinal Cord Injury
Bickel et al (1991): report on 3 Louisiana medical centers
8/2900 (0.32%) SCI patients developed bladder cancer
2/8 SCC, 6/8 Urothelial ca
Level 3
Pannek J (2002): by questionnaire sent to Urology Depts in
Eastern Europe
43,561 SCI patients identified
Only 7% of patients had indwelling catheters
48 (0.11%) patients developed bladder cancer
19% of the patients with bladder cancer were SCC
Level 3

SCC of the Bladder in


Spinal Cord Injury
Recommendations:

(Grade B)

The incidence is less than 1%

Monitoring of patients with indwelling catheters is


important
Hematuria should be evaluated
Frequency of surveillance and diagnostic work up
can not be determined

Non-bilharzial SCC of the Bladder


Pure SCC (Clinico-pathologic features):
Usually present with hematuria

30-90% have associated infection irritative


symptoms

Latent period from symptoms to diagnosis


Superficial SCC is rare, majority are muscle- invasive
(Jones, 1980; Rundle, 1982; Quilty, 1986)
(Level 3 evidence)

Non-bilharzial SCC of the Bladder


Pure SCC (Clinico-pathological features):
Solitary large tumor
Sessile lesion with ulceration surrounded with
metaplasia
Trigone, but can occur in other areas
May develop in diverticula, associated with stones
Imaging and work-up the same as for TCC
(Costello, 1984; Sereta, 2000)
(Level 3, 4 evidence)

Non-bilharzial SCC of the Bladder:


Treatment
Prognosis is generally poor (most die 1-3 years after dg.)
Radical surgery is the most effective treatment
5-year survival rates range from (16-48%)
Treatment failure is usually due to local pelvic
recurrence
Incidence of distant mets 8-10%
(Jones, 1980)

(Level 3 evidence)

Non-bilharzial SCC of the Bladder: Treatment

Radiotherapy Alone:
5-year survival 16-26%
The response was more in T2 disease.
T3 disease (4.8%) 5 years.
(Quilty, 1986; Johnson, 1976; Rundle, 1982)

(Level 3 evidence)

Non-bilharzial SCC of the Bladder:


Treatment
Preoperative radiotherapy?
Some improvement 5 years 48% (many patients
had been excluded).
No reliable randomized studies to draw a
conclusion.
(Richie,1976)

(Level 3 evidence)

Non-bilharzial SCC of the Bladder: Treatment


Chemotherapy ?
Role is uncertain
Effective chemotherapy protocol has not been found
Neoadjuvant M-VAC no response
(Khaled, 2000)
(Level 3 evidence)

The results of new combination regimens (paclitaxelgemcitabine-cisplatin) are awaited.


(Level 4 evidence)

Bilharzial SCC of The Bladder


Pathogenesis
Latent period up to 30 years.
Locally produced chemical carcinogen

DNA changes
Bacterial infections (Nitrosamines)
Metaplasia (Mechanical irritation)
Level 4

Changing Patterns of Egyptian Bladder Cancer


Data abstracted from NCI- Cairo

From Felix, et al : Cancer Causes Control (2008):19,421-9

Bilharzial SCC of The Bladder


Clinical presentation
Latency period of approx 30
years between infestation
and development of cancer
Similar symptoms between
bilharzial cystitis and
malignant cystitis. Painful
and frequent voiding, with
hematuria
many cases have advanced
disease when first seen

Diagnostic work-up is the


same as TCC (CK in the urine)

Level 3 and 4

Life cycle of
Schistosomiasis

Retrospective Egyptian Cystectomy Series


Pathologic Stage and Prognosis
Tumor grade:
GI: 50%
GII: 32%
GIII: 18%

Understaged in 33%

Ghoneim MA, et al (J Urol 1997): Level 3

Bilharzial SCC of The Bladder


Clinico-pathological features:
Most of the tumors are located in posterior or lateral walls
5 types based upon gross appearance
Nodular fungating: 60%
Ulcerative 23%
Verrucous 7%
Papillary 7%
Diffuse 3%
Atypical changes - Metaplasia, dysplasia, leukoplakia are
common but CIS is rarely seen.

Very rare to invade the ureters or the urethra

Bilharzial SCC of The Bladder


EUA/Endoscopic Resection
In many cases, limited to biopsy
Not a known tool for treatment because of bulky
disease.
In case of complete endoscopic excision, follow up
cystoscopy and deep biopsy is required

Level 4

Bilharzial SCC of The Bladder


Partial cystectomy
Indications limited:

Low grade, solitary lesion

Small tumor allows good safety margin

Away from the trigone.


These criteria were present in 19/190 pat. (10%).
Augmentation cystoplasty needed in 25%.
5 year survival in 26.5%.
Local recurrence was high.
El-Hammady, (1975): Level 3

Bilharzial SCC of The Bladder


Radical Cystectomy
608 patients with SCC.
5 year survival = 50.3%.
Tumor stage, grade and LN involvement were
independent prognostic factors.
Cystectomy as single modality is inadequate

Ghoneim (1997): Level 3

Bilharzial SCC of The Bladder


Neoaduvant Radiation Therapy
20 cGy prior to cystectomy vs cystectomy alone
Prospective, randomized trial evaluating 90
patients.
60 months follow-up
Trend in favor of irradiation but not statistically
significant.

Ghoneim, et al (1985): Level 1

Bilharzial SCC of The Bladder


Radiation Therapy alone:
Few studies
Poor response

Tumor bulk with hypoxic cells


Bilharzial cystitis (intolerance)
Level 4

cystectomy vs. 20 rad + cystectomy vs. 20 rad + misondzl +


cystectomy - Addition of misonidazole as a radiosensitizer
did not alter survival advantage.
Ghoneim 1985
Level 1

Bilharzial SCC of The Bladder


Chemotherapy
Phase II studies showed some activity of epirubicin.
Level 4

Neoadjuvant epirubicin showed an initial promising


results in one study but was not reproducible.
Level 2

In recent multicenter trial using neoadjuvant


cisplatin + gemcitabine, patients with SCC did not
show survival advantage. Level 2
New chemotherapy protocol has to be identified.

Squamous Cell Carcinoma


Grade B:Recommendations
Non-Bilharzial SCC:
Predominantly single center, retrospective series
Radical Cystectomy appears to offer the best survival in localized
disease
Bilharzial SCC:
Predominantly single center, retrospective series
Radical Cystectomy appears to offer the best survival in
localized disease

Adenocarcinoma of the bladder

Adenocarcinoma of The Bladder

Primary
Urachal
Metastatic

Adenocarcinoma of the Bladder


Incidence
0.5-2%

Western series
(Bennet, 1984) (Level 3 evidence)

8-11%

In Endemic Bilharzial areas


(El-Mekresh, 1998) (Level 3 evidence)

Exstrophy 4% life time risk


Male to female: 2.7:1
(Level 3 evidence)

Adenocarcinoma of The Bladder


Urachal Carcinoma:
Rare tumor (1 in 1400)

Younger age (49 years vs. 58 nonurachal)


Male to female 4:3
(Dandeker, 1997)
(Level 3 evidence)

Adenocarcinoma of The Bladder


Urachal Carcinoma:
Location (bladder dome)
Extravesical extension
Sharp demarcation
Not associated with cystitis glandularis
Indistinguishable from enteric adenocarcinoma
(Johnson, 1985)
(Level 4 evidence)

Adenocarcinoma of The Bladder


Urachal Carcinoma:
Staging (Sheldon, 1984)
I: Urachal mucosa
II: Invasion to the urachus
III: Invasion to the bladder, abdominal
wall, local viscera
IV: regional nodes or distant metastases

Adenocarcinoma of The Bladder


Urachal Carcinoma:
Treatment:
Wide excision of the umbilicus,
urachus, and partial cystectomy
Survival:
40% at 5 years (MD Anderson)
Prognosis:
Lymph node involvement
Positive surgical margin
(Level 3 evidence)

Adenocarcinoma of The Bladder


Urachal Carcinoma:
systemic treatment:
Chemotherapy

5-fluorouracil and cisplatin


No benefit

(Level 3 evidence)

Radiotherapy
Not enough trials seems poor
(Level 4 evidence)

Adenocarcinoma of The Bladder


Adenocarcinoma in bladder exstrophy:
The most common malignancy in the retained bladder
More in males
(McIntoch and Woly 1955) (Level 3 evidence)
4/102 patients followed for more than 35 years
700 times risk > normal population
(Smeulder, 2001)

(Level 3 evidence)

metaplasia of the urothelium d/t defunctionalised


bladder and bowel augmentation
(Level 4 evidence)

Adenocarcinoma of The Bladder


Primary adenocarcinoma
Chronic vesical irritation, metaplastic changes
Virtually always invasive

(Dandekar,1997)

Mucin or non-mucin producing


Stage, Grade, LNs are the prognostic factors
(El-Mekresh,1998)

(Level 3 evidence)

Adenocarcinoma of The Bladder


Primary adenocarcinoma
Treatment
TUR
(19% -5 years)
(Kramer,1979)
Partial cystectomy
(dismal ~25%)
(Abenoza,1987)
Radical cystectomy
(0-80%)
(El-Mekresh,1998)
Radiotherapy
(<20%)
(Thomas,1971)
Preoperative irradiation (No benefit) (Anderstrom,1983)
Chemotherapy
(Dismal)
(Hatch,1989)
(Level 3 evidence)

Adenocarcinoma of The Bladder


Secondary adenocarcinoma
Direct extension
Lymphohematogeneous spread
Colon 21%
Prostate
19%
Rectum
12%
Cervix 11%
(Bates and Baithun,2000) (Level 3 evidence)
Deleted in Colon Cancer protein (DCC) immunostain +ve may
differentiate primary from 2 (Yossepowitch, urology 2004) level 3

Adenocarcinoma of the Bladder


Secondary adenocarcinoma
Treatment:
Wide excision with a wide safety margin is required
(20% recurrence)
Bladder-sparing surgery (100% recurrence)
The treatment failure depends on the stage of the
primary disease
(Carne,2004) (Level 3 evidence)

Small Cell Carcinoma


Rare tumor (neuroendocrine tumor in the lungs)
+/- 300 cases reported
0.4 0.7% of all bladder tumors
Age above 60, 80% are males
Pathological diagnosis is difficult
Usual appearance at stage T2
Metastatic disease in 67% (LN, liver, bone, lung and brain).
(Sved et al.,2004)

(Level 3 evidence)

Small Cell Carcinoma


Treatment: multimodal therapy
Lung small cell chemotx. cis+etoposide/ifos+doxo,
median survival 10-14 months
MDA experience (JUrol,2004) 46 pts. 5-year survival:
cystectomy + post-op chemo 36%
neoadjuvant chemo + cystectomy -78% (chemo
directed to small cell and not MVAC/gemzar+cis)
2009, 2 studies (UK+North Africa) survival 7-33 months
(Siefker-Radtke et al.,2004)

(Level 3 evidence)

Small Cell Carcinoma


GRADE B RECOMMENDATION:
Cure is most likely with a combined multimodality
approach including neoadjuvant chemotherapy
and local therapy

Bladder Sarcoma
Leiomyosarcoma (50%)
Hematuria early diagnosis
Radical cystectomy in early cases (80% survival)
(Rosso,1992) (Level 3 evidence)

Multimodal treatment in metastatic disease


(Level 4 evidence)

Rare Bladder Malignancies


Carcinosarcoma
Aggressive disease (41 cases)
Poor outcome following cystectomy alone
(Lopenz-Beltran,1998) (Level 3 evidence)

Multimodal treatment

(Froehner,2001) (Level 4 evidence)

Paraganglioma, pheochromocytoma
Paroxysmal symptoms with micturition
Cystoscopy under precaution (Biopsy !!!)
Local excision with lymphadenectomy
May be difficult to distinguish (benign or malig)
(Klingler,2001) (Level 3 evidence)

Rare Bladder Malignancies

Melanoma:
Primary is rare
Secondary from the skin is more prevalent

Clinical picture and biopsy are diagnostic urethra and not so


much in the bladder
Treatment is radical surgery
Prognosis is poor
(De Torres, 1995) (Level 3 evidence)

Rare Bladder Malignancies

Lymphoma:
Part of metastatic spread
More in women
Localized disease has good prognosis
Local irradiation is satisfactory
(Kempton,1997) (Level 3 evidence)

Bladder Pseudotumors
Rare

Resemble malignancy
Postoperative spindle cell tumors
Can mimic leiomyosarcoma
Absence of mitotic figures
TUR or partial resection is enough

Cystectomy if the diagnosis is difficult


(Jones, 1993) (Level 3 evidence)

Você também pode gostar