Você está na página 1de 991

PUMPS

Sec 0 n d Ed it ion
Dedicated to the memory of Igor J. Karassik, an engineer of note
whose career contributed handsomely to the general comprehen-
sion of pumps, hence their benefit to society at large.
PUMPS
Second Edition
Igor J. Karassik
(Deceased)
formerly Senior Consulting Engineer
Ingeool-Dressor Pump Comp;!flY
Liberty Comer, New Jersey

Terry cGuire
Consulting Engineer and Director, AlhllOCes
IngersoIl-Dressor Pump ComP<!flY
Liberty Comer, New .Jmey

CHAPMAN & HAll

I JlONAI. THOMSON PUBUSHING


Thomson SctenCf:

New YorI> • AlbGny • 80m • Boston • ClIlOmlltI • DetrOit


laldon • M!Idod • Mdbolme • Meloco CIty . Poclflc GrO'<'e
PIYl5 • Sen FrllnClSCO • SI09IlPOI'I! • Tokyo • Torooto • Wasn'l19too
Join Us on the Internet

WWW: http://www.thomson.com
EMAIL: findit®kiosk.thomson.com

thomson.com is the on-line portal for the products, services and resources available from International
Thomson Publishing (ITP).

This Internet kiosk gives users immediate access to more than 34 ITP publishers and over 20,000 products.
Through thomson.com Internet users can search catalogs, examine subject-specific resource centers and
subscribe to electronic discussion lists. You can purchase ITP products from your local bookseller, or
directly through thomson.com.

Visit Chapman & Hall's Internet Resource Center for information on our new publications,
links to useful sites on the World Wide Web and an opportunity to join our e-mail mailing list.
Point your browser to: http://www.chaphaIl.com or http://www.thomson.comlchaphalllmecheng.htmi
for Mechanical Engineering
A service of I®p®

Copyright © 1998 by Chapman & Hall


Softcover reprint of the hardcover 2nd edition 1998

Chapman & Hall Chapman & Hall


115 Fifth Avenue 2-6 Boundary Row
New York, NY 10003 London SEl 8HN
England
Thomas Nelson Australia Chapman & Hall GmbH
102 Dodds Street Postfach 100 263
South Melbourne, 3205 D-69442 Weinheim
Victoria, Australia Germany
International Thomson Editores International Thomson Publishing-Japan
Campos Eliseos 385, Piso 7 Hirakawacho-cho Kyowa Building, 3F
Col. Polanco 1-2-1 Hirakawacho-cho
11560 Mexico D.F Chiyoda-ku, 102 Tokyo
Mexico Japan
International Thomson Publishing Asia
221 Henderson Road #05-10
Henderson Building
Singapore 0315

All rights reserved. No part of this book covered by the copyright hereon may be reproduced or used in
any form or by any means-graphic, electronic, or mechanical, including photocopying, recording, taping,
or information storage and retrieval systems-without the written permission of the publisher.

1 2 3 4 5 6 7 8 9 10 XXX 01 00 99 98

Library of Congress Cataloging-in-Publication Data


Karassik, Igor T., 1911-
Centrifugal pumps: selection, operation, and maintenance I by
Igor J. Karassik, J.T. MtGuire. -- 2nd ed.
p. em.
Includes index.
ISBN 978-1-4615-6606-9 ISBN 978-1-4615-6604-5 (eBook)
DOI 10.1007/978-1-4615-6604-5
1. Centrifugal pumps. I. MCGuire, J. T., 1947- II. Title.
TJ919.K3 19%
621.6'7--dc20 96-1819
CIP

British Library Cataloguing in Publication Data available

"Centrifugal Pumps" is intended to present technically accurate and authoritative information from highly
regarded sources. The publisher, editors, authors, advisors, and contributors have made every reasonable
effort to ensure the accuracy of the information, but cannot assume responsibility for the accuracy of all
information, or for the consequences of its use.

To order this or any other Chapman & Hall book, please contact International Thomson Publishing,
7625 Empire Drive, Florence, KY 41042. Phone: (606) 525-6600 or 1-800-842-3636.
Fax: (606) 525-7778. e-mail: order@chaphall.com.
For a complete listing of Chapman & Hall titles, send your request to Chapman & Hall, Dept. BC,
115 Fifth Avenue, New York, NY 10003.
Contents

Preface to the Second Edition vii

Foreword (First Edition) IX

PUMP TYPES AND CONSTRUCTION


1. Classification and Nomenclature 3
2. Casings and Diffusers 18
3. Multistage Pump Casings 44
4. Impellers and Wearing Rings 62
5. Axial Thrust in Single- and Multistage Pumps 91
6. Hydraulic Balancing Devices 104
7. Shafts and Shaft Sleeves 116
8. Stuffing Boxes 136
9. Mechanical Seals 161
10. Breakdown Seals 195
11. Bearings 211
12. Couplings 272
13. Baseplates and Other Pump Supports 298
14. Special Designs: Vertical Pumps 311
15. Special Designs: Self-Priming Pumps 347
16. Special Effect Pumps 352
17. Materials of Construction 367

II PUMP PERFORMANCE 399


18. Heads, Conditions of Service, Performance Characteristics, and Specific Speed 401
19. Suction Conditions and Limitations on Suction Performance 473
20. System-Head Curves 506
21. Pumps and Energy Conservation 545

v
vi Contents

22. Pump Operation at Off-Design Conditions 558

III CONTROLS, DRIVERS, AND PRIMING 593


23. Controls 595
24. Drivers 621
25. Priming 686

IV SERVICES AND SELECTION OF PUMPS 715


26. Services 717
27. Procuring Centrifugal Pumps 800

V INSTALLATION, OPERATION, MAINTENANCE, AND DIAGNOSTICS 831


28. Installation 833
29. Operation 882
30. Monitoring and Performance Testing 893
31. Maintenance 916
32. Diagnostics of Field Problems 927

VI DEVELOPMENT 947
33. The Centrifugal Pump of Tomorrow 949

Data Section 957

Index 975
Preface to the
Second Edition

New Yorkers woke up to face a dismal morning, at least the New Yorkers who did not depend on
electric alarm clocks. It was cold, with the temperature hovering around 20 degrees. The electricity was
off and so was the heat. No hot water, no water of any kind issued from the faucets. No radio, no news.
Even battery-operated radios were no help, because the stations were not broadcasting. No gas to cook
breakfast. Badly shaven or not shaven at all, hungry, grumpy, and bewildered, men issued into the streets
to find no subways running, no traffic except for an occasional horse-drawn peddler's wagon. Those
who tried to start their cars had no success. A few men took off for their offices on foot. Most congregated
on street corners and asked each other questions. There were no newspapers because there was no way
of delivering them. Before long, all sorts of rumors were flying. A little later, rioting and looting broke
out allover New York. The police were handicapped by the fact that all normal communication had
failed. By noon, emergency telephone communication operated by batteries had been reestablished
between critical points in the city. But by noon, the situation itself had become critical.
Elsewhere in the world the picture was essentially the same in the cities and small towns, although
the smaller the towns, the less panic. In the country, matters were simpler. When farmers found that
there was no water, they went to the rivers, lakes, and ponds with buckets and brought some water back.
Just like their ancestors had done hundreds of years before.
All this happened because on that morning all the pumps in the world stopped running. But of course
this could not have happened. For had all the pumps stopped, I would not be writing this, nor would
you be reading it, because the human heart is also a pump and it too would have stopped.
Every industrial process that underlies our modern civilization involves the transfer of liquids from
one level of pressure or static energy to another level, and as a result, pumps have become an essential
part of all industrial processes. Carried farther, this means that pumps are an integral part of all modern
economic and social development. At the same time, as I have said in one of my papers, the role of the
pump industry should be much more than the development of new lines of pumps, the manufacturing
of these pumps, and their selling. It should also include making efforts to build more efficient and longer
lived equipment, to do so with less expenditure of natural resources, and, especially, to educate pump
users in practices that consume less energy, provide trouble-free operation for longer periods, and reduce
the incidence of premature failure.

vii
viii Preface to the Second Edition

This last, the education of pump users, is precisely what this book was intended to do. To what extent
we must have achieved our purpose, our readers must decide.
My good friend and associate, J. T. (Terry) McGuire, and I have been working very closely together
for a long time. Our view of engineering problems and of their solutions coincide to an astonishing
degree. When I was asked to prepare a second edition of my book Centrifugal Pumps, it was logical
that I turned to Terry and suggested that he be my coauthor on this project. He agreed to do so, and his
cooperation has been most valuable, both in improving the resultant work and in easing my burden.
It would be presumptuous on my part to pretend that nothing has changed in the technology of
centrifugal pumps during the 30 years since I prepared the manuscript for the first edition of this book.
Let me, then, speak of some of these changes.
In the area of pump hydraulics, the most important addition to our knowledge has been the understand-
ing of the phenomena that take place in a pump impeller as the capacity of a pump is reduced below
that at which best efficiency is achieved. These phenomena can lead to pntssure pulsations, vibrations,
and damage to the impeller and to certain adjacent pump areas. The understanding of the causes of these
phenomena has, in turn, led to a better evaluation of what should be the minimum operating flows for
centrifugal pumps. This subject is treated in a completely new chapter (Chapter 22).
Greater understanding exists today with respect to conditions at the pump suction and of the effect
of liquid properties on the behavior of pumps under cavitating conditions. The part of the book dealing
with this subject has been much expanded and is now covered separately in Chapter 19. The portion of
the chapter discussing guidelines for pump suction conditions recommended by the Hydraulic Institute
has been revised in a much simplified form.
Some improvements have been made in the range of attainable efficiencies, and means are now available
to correct these attainable efficiencies when pump design practices depart from certain preselected standard
constraints. These more refined guidelines are incorporated in Chapter 18.
In pump construction, the last 30 years have seen the application of finite element analysis to the
design of pump components, resulting in better hydraulic designs, refinement of the technology of rotor
dynamics, more effective pressure containing parts, and a significant increase in the service life of
mechanical seals.
The emerging technological improvements in the area of variable frequency motors will have a very
marked influence on the selection of drivers for centrifugal pumps and in the availability of pump speeds
other than the synchronous speeds at 60 and 50 cycles.
Pump applications have shown a trend to higher capacities, pressures, and, consequently, power levels.
There has also been a greater tendency to use centrifugal pumps to handle a variety of slurries and other
solids-laden liquids.
Finally, two new chapters are devoted to pumps and energy conservation and to the centrifugal pump
of tomorrow.
As in the case of the first edition of this book, we have tried to avoid using this as a vehicle to discuss
subjects that are possibly of great interest to pump designers but can contribute nothing but confusion
to pump users. We have guarded ourselves against the temptation to discuss the theoretical intricacies
of velocity diagrams, vane angles, or other similar design details. A doctorate in fluid dynamics should
not, we firmly believe, be a prerequisite to understanding what the user must do to achieve a successful
pump installation.
This second edition is dedicated to Henry R. Worthington. We've done this to mark the 150th
anniversary, albeit now 6 years ago, of the development of the direct acting steam pump by Henry R.
Worthington-an invention that laid the foundation for the entire pump industry.

Igor J. Karassik
Foreword
(First Edition)

The subject of centrifugal pumps has received much attention in technical literature both here and abroad.
However, the authors felt that most of this literature placed greatest emphasis on centrifugal pump theory,
with insufficient stress on the more practical side of the problem. This practical side is more important
to most engineers and users, as these people put centrifugal pumps to use while only a small minority
actually design the equipment.
One aim of this book is to guide the centrifugal pump user in system design and equipment selection
for the most satisfactory combination of the two. It is also intended to provide useful information about
equipment already installed as a guide to maximum service with minimum maintenance and unscheduled
outage. The structural details and component parts of centrifugal pumps are described and methods are
recommended for restoring each component to its initial condition after deterioration in service. In
addition, special chapters are devoted to vertical pumps, self-priming units, and the so-called regenerative
pumps. These are followed by a discussion of construction materials. A detailed presentation is given
on the concept of "heads," conditions of service, and performance characteristics of various types
of centrifugal pumps. System-head curves and their effect on pump output and selection are also
fully discussed.
An important factor in centrifugal pump application and operation that has often been neglected is
the controls. This subject has been given special attention. Because successful pump application also
depends on a harmonious combination of pump and driver, a chapter is included on pump drivers.
Another important subject, priming, is discussed in great detail.
Nearly all centrifugal pumping services have their individual problems and requirements. These
services range from general water supply, sewage, drainage, and irrigation to power plant, process work,
and other specialized applications. Growth and change in processes and industries have contributed to
the development of new designs for the ever-increasing number of pumps. Many special designs are
therefore available today that may be severely limited in application flexibility. Centrifugal pump users
should have a general knowledge of specialized designs to help assure proper application. The chapter
on services covers these special types and presents related operational information. It is supplemented
by a chapter on the preparation of inquiries and ordering procedures.
One important section of the book is devoted to the installation, operation, and maintenance of

ix
x Foreword

centrifugal pumps. Finally, to make this book as useful as possible, a general Data Section contains
valuable data required for engineering pumping installations and analyzing the performance of exist-
ing units.
The authors have attempted to avoid complex technical explanations and involved theoretical discus-
sions having little practical value to centrifugal pump users. Theoretical design data would only suggest
that the user is expected to judge the excellence of the designer. This aim is not part of, nor compatible
with, the objectives of the book, which are to provide practical and useful knowledge of centrifugal
pump construction, application, control, installation, operation, maintenance, and trouble-shooting.
The data in this book apply to all makes and types of centrifugal pumps. Wherever possible, therefore,
illustrations have been selected from a wide group of manufacturers. For obvious reasons, however, we
had greatest access to the extensive files of the Worthington Corporation. For many subjects-for
example, individual pump parts-the illustrations would be similar, regardless of source, and therefore
most of these were selected from the Worthington files. Wherever photographs of complete pumps or
sectional drawings that are not from Worthington are reproduced, the captions give credit to the pump
manufacturer responsible for the design. The authors wish to extend their thanks to the Worthington
Corporation, the Allis-Chalmers Mfg. Co., Byron-Jackson Co., the DeLaval Steam Turbine Co., Ingersoll-
Rand Co., Pacific Pump Co., and many others who very graciously granted the right to reproduce
equipment photographs and drawings.
The authors also wish to thank numerous magazines including Air Conditioning, Heating and Ventilat-
ing, Power, Power Engineering, Southern Power and Industry, Water and Sewage Works, and many
others for their kind permission to utilize material from articles by the authors that had originally appeared
in their pages.
We are indebted to the Hydraulic Institute for the permission to reproduce a number of charts and
data from its Standards.
Finally, the authors wish to express their thanks to Messrs. A. H. Borchardt, G. F. Habach, L. H.
Garnar, W. C. Krutzsch, C. J. Tullo, and many other associates at Worthington Corporation for providing
valuable advice and constructive criticism.
Roy Carter and I decided to undertake this book a number of years ago. Unfortunately, Mr. Carter
did not live to see it completed. He passed away unexpectedly in September 1958. I decided that the
book should be completed and therefore continued the task alone. I hope it measures up to our mutual
expectations. Because its writing reflects our many discussions, and its chapters contain much that we
had conceived together as articles for technical magazines, this book carries both our names as coauthors.

Igor J. Karassik
PUMPS

Sec 0 n d Ed i tiD n
I
PUMP TYPES
and
CONSTRUCTION
1
Classification and Nomenclature

Pumping can be defined as the addition of energy to a fluid to move it from one point to another. It is
not, as frequently thought, the addition of pressure. Because energy is capacity to do work, adding it to
a fluid causes the fluid to do work, normally flowing through a pipe or rising to a higher level.
A centrifugal pump is a kinetic device, meaning that it adds energy to the pumped liquid by increasing
its velocity. Because the addition of energy depends on liquid velocity, the amount of energy added
varies with the rate of flow through the pump. These are the fundamental physics behind the usual head
versus capacity characteristic of the centrifugal pump (Fig. 1.1). It is important to note the distinct
difference between the head (energy added) versus flow characteristic of a centrifugal pump and that of
a displacement pump (Fig. 1.1). Centrifugal pumps have low "flow regulation"; their flow varies widely
with variations in system resistance, a characteristic that lends itself to easy flow control. Displacement
pumps exhibit high flow regulation; their flow is largely independent of variations in system resistance,
making them ideal for services where a constant flow is necessary over varying system conditions. The
limitations of displacement pumps are machine size versus capacity and mechanical complexity.
A centrifugal pump is a simple machine consisting of a set of rotating vanes enclosed within a housing
or casing. Torque applied by the pump's driver is converted to total head by the action of the vanes on
the pumped liquid, and these vanes are the only component that adds energy to the liquid. This action
follows Euler's equation faithfully, provided it is recognized that the effective liquid velocities, magnitude,
and direction cannot be determined directly from the geometry of the vanes. That determination is part
of the centrifugal pump designer's "art." Stripped of all refinements then, a centrifugal pump has just
two main parts: (1) a rotor, made up of the vaned component, known as an impeller, and a shaft, and
(2) a stator, made up of the impeller enclosure, known as a casing, some form of seal where the shaft
passes thru the casing, and bearings to support the rotor. The structural details of these parts and all
refinements applied in modern pump construction are covered in Chapters 2 through 17.
Most people find it difficult to visualize the path of the liquid passing through a centrifugal pump.
Figure 1.2 shows this path for a single-stage end-suction pump operating at its design capacity, that at
which best efficiency is obtained. The liquid, forced by either atmospheric or other pressure, enters the
impeller vanes at a relatively low velocity, is accelerated to a high velocity as it passes through the

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
4 Classification and Nomenclature

Displacement I
"'i

Energy
Added @ Constant Speed

Flow

Fig. 1.1 Head-capacity characteristic of centrifugal (kinetic) versus displacement pumps.

impeller. then is discharged into the casing, where the high velocity is converted into pressure, a process
known as "diffusion."
One of the most important factors contributing to the increasing use of centrifugal pumps has been
the universal development of electric power. This century has seen electricity replacing small steam
plants as the main industrial power source. Although reciprocating pumps were ideal for steam drive,
the development of the electric motor permitted use of the much lighter and cheaper direct-connected
centrifugal pump. Even though early centrifugal pumps would be considered inefficient by modem
performance standards, their lower first cost more than compensated for this shortcoming. The centrifugal
pump also immediately demonstrated other important advantages over the reciprocating pump. For
example, the centrifugal pump gives steady flow at uniform pressures without pressure surges. It provides
the greatest possible flexibility, developing a specific maximum discharge pressure under any operating
condition with delivery controlled by either speed variation or throttling.
Naturally, manufacturers working to widen the field of centrifugal pump applications through experi-
ence and research have greatly improved the operating range of pressures, the efficiency, and the
mechanical and hydraulic design of their product. Concurrently, electric motor builders improved their
designs, permitting pump manufacturers to use higher rotative speeds and develop pumps suitable for
higher heads. So, over the last 80 years, the application of centrifugal pumps has been greatly extended
in both pressure and capacity. Centrifugal pumps have been built in sizes ranging from tiny swimming
pool pumps handling a few gallons per minute to the very large reversible pump turbines at Bath County,
which deliver 116 m3/s (1.8 million gpm) against 387 m (1,270 ft) total head and absorb 410 MW
(549,000 hp) when running in pump mode. The range of differential pressures is just as wide, starting
with a single-stage cellar drainer pump which develops around 0.3 bar (5 psi), and rising to the multistage
pumps used for oil field water injection, which develop as much as 360 bar (5,250 psi) in a single casing.
And centrifugal pumps have been designed to run at speeds as high as 25,000 rpm for industrial
applications. Compounding all that, the centrifugal pumps used in the space shuttle for fuel supply are
the highest energy density machines yet made: 3,400 m3Jhr (15,000 gpm), 56,700 m (186,000 ft), 52,000
FLOW LINE

()

SECTION THROUGH IMPELLER ANO VOLUTE


"LONG "'E"~ FLOW SURFAt.~

1 rPICAt PUMP SECTION

80
OISCHARGE I I I I
~ VANE TIP
I I I I
8w 70 ..:>,
I I
'" VOLU TE THROAT
i!i0. 60 --- I- 1-1-
200 GPM ./ f"-.
166 FT TOTAL HEAD ... V
3,500 RPM 50 --f-
2t IN. SUCTION DIAM
...t:l
2 IN. DISCHARGE OIAM z SUCT ION V
- 40 r- VANE TIP
6i IN. IMPELLER DIAM
f IN. IMPELLER WIDTH
,: /
I-
g 30 V \.
..J /
~ 1\
20
...'":> -1-1-'
~ 10
CD

" o
DE VELOP E 0 PATH
SUC TION OISCHARGE
FL ANGE FLANG E

'"
Fig. 1.2 Flow path and velocity variation through an end-suction volute pump.
6 Classification and Nomenclature

Capacity, m3/h
10 100 1000 10,000
100,000 ""T"""~,r-.,......L.. _ _ _.J..-_ _--L_ _ _..L..-_ _---L----,

"
"" Reciprocating
1000
10,000 -
-----~'" Centrifugal
....
III
.0
100 ~
1000 - ::J
CJ)
CJ)

~
Q..

I 10
100 I
I
I I
I i
10;----r_---,----r--~r_--~ ~ i
10 100 1000 10,000 100,000
Capacity U.S. gpm

Fig. 1.3 Approximate upper limit of pressure and capacity by pump class.
(Reprinted with permission/rom "Pump Handbook," Me-Graw Hill, NY, 1976.)

kW (70,000 hp) at 37,000 rpm in a machine weighing 780 kg (1720 lb) including the integral gas
turbine driver.
Although the coverage of centrifugal pumps is very broad, it is not unlimited. For high-differential
pressures at low capacities, reciprocating pumps are necessary. Similarly for some services within the
normal coverage of centrifugal pumps, a displacement pump, either rotary or reciprocating and in that
order, may be the better choice for the service conditions. Figure 1.3 shows the approximate upper limits
of energy added (pressure rise) and flow for each of the three basic pump classes.

MODERN PRACTICE IN SMALL- AND MEDIUM-CAPACITY RANGE

In small- and medium-size centrifugal pumps, about 60 percent of all pumps in use are of centrifugal
design and over 75 percent of these are in a head capacity range that can be met by standardized end-
suction pumps.
A typical example of this standardization is a line of pumps consisting of a number of liquid ends,
all suitable for mounting (1) on a motor in a close-coupled arrangement, (2) on a bearing frame for
separately coupled or belt drive, or (3) on a turbine in a close-coupled arrangement, although the last
is rare today. Many standard alternatives are then incorporated into the basic plan, both in materials
selected and mechanical construction, thereby eliminating "specials." The use of such an integrated line
can result in better delivery service, a wider selection of standardized units, and dollar savings through
the maximum use of interchangeable parts. In one typical case, some 100 sizes of pumps, using inter-
changeable parts, can produce over 60,000 different, standard combinations to suit almost any user's needs.
Classification and Nomenclature 7

CENTRIFUGAL PUMP CLASSIFICATION

Centrifugal pumps are produced in a wide range of design variations. To bring some order to these
variations, several classification systems are in use. The original system is based on the form of the
collector that surrounds the impeller and then the geometry of the impeller itself. Pumps employing a
volute collector (Fig. 1.2) are generally called volute pumps, and those having a multiple vane diffuser
(Fig. 1.4) are called diffuser pumps. Diffuser pumps were once quite commonly called turbine pumps,
but this term has become more selectively applied to vertical deep-well centrifugal diffuser pumps, now
called vertical turbine pumps. The impellers, in tum, are classified as radial flow, mixed flow, and axial
flow, now usually called propeller type. The impeller classifications are further subdivided by suction
arrangement, single or double, and vane closure (see Chap. 4).
Associated with classification based on the hydraulic components, there are terms related to the
arrangement of those components. If the pump is one in which the total head is developed by a single
impeller, it is called a single-stage pump. Often the total head to be developed requires the use of two
or more impellers in series, each taking its suction from the discharge of the preceding impeller. For
this purpose, two or more single-stage pumps may be connected in series, or all the impellers may be
incorporated in a single casing. The machine is then called a multistage pump.
In the early development of centrifugal pumps, birotor pumps (Fig. 1.5) and even trirotor pumps were
fairly common. In effect, these were two one-half capacity or three one-third capacity pumps built into

Casing

Fig. 1.4 Typical diffuser pump.


8 Classification and Nomenclature

Fig. 1.5 Birotor pump.

the one casing and operated in parallel. Modem versions of such designs, typically used for pipeline
service, are generally a series-parallel arrangement, for example, a three-impeller two-stage pump (two
single-suction first-stage impellers in series with a double-suction second-stage impeller; Fig. 1.6) or a
four-impeller two-stage pump (two single-suction first-stage impellers in series with two single-suction
second-stage impellers; Fig. 1.7.
The mechanical design of the casing provides the added classification of axially split or radially split,
and the orientation of the axis of rotation determines whether the pump is horizontal, vertical, or
(occasionally) inclined. Horizontal pumps are classified still further according to suction and discharge
nozzle locations, the more common arrangements being

Fig. 1.6 Two-stage, three-impeller pump.


(Courtesy Sulzer.)
Classification and Nomenclature 9

Fig. 1.7 Two-stage, four-impeller pump.

1. End suction, top discharge


2. Top suction, top discharge
3. Side suction, side discharge
4. Bottom suction, side discharge.

Some pumps operate with the total liquid flow conducted to and from the unit by piping. Other
pumps, most often vertical types, are submerged in their suction supply. Vertical pumps are therefore
either dry-pit or wet-pit types. If the wet-pit pumps are axial flow, mixed flow, twin volute, or vertical
turbine types, the liquid is discharged up through the supporting drop or column pipe to a point either
above or below the supporting floor. These pumps are consequently designated as above-ground or
below-ground discharge units.
An alternative to the classification system described here has been developed by the Hydraulic Institute
[1.1]. Figure 1.8 shows a subtle variation of this system, which first differentiates between the various
classes of kinetic pumps, of which the centrifugal pump is one. Centrifugal pumps are then classified
by their basic rotor construction, their drive arrangement, and finally their casing configuration. The
virtue of this system is that it focuses on pump configuration versus use, thereby leaving the choice of
hydraulic detail to the designer. Because the basic rotor construction has a significant effect on the
balance of a pump's design, having it as the first distinguishing feature means several fundamental design
issues are settled once the class of rotor is selected.
....
<:)
Suppooted
---, - BlIICket
E ~ MouI*d
Overhung-{-~_~~~
Sepandely ~_ PIt Vol
Centrifugal -----i

SIngle SIage C A*I CIIIIing


SpItCasIng
_ SpIt
Between CasIng
BearIng. SpIll CIIIIing
L _0$'- c:: _
_ SplIt
Vertical C VerIIeaI Tu_
Diffuser ",-,Ie,
r-SingIeSlllge
Regenerative I Overhung - - - - - - - i L - MuItI-"-
Turbine L Between SIIIge
,Kineuc Bearings
c:: Single
Mulll-Slllge

Special E Reversible Centrifugal


Effect Rowllng Casing (Pilot)
Barske
Pumps----i t Viseous Shear (Tesla)
Induced Vortex

Displacement

Fig. 1.8 Classification of kinetic pumps (after Hydraulic Institute).


Classification and Nomenclature 11

CENTRIFUGAL PUMP COMPONENT NOMENCLATURE

The basic elements of a centrifugal pump are its impeller, casing, shaft, and bearings, but there are other
necessary parts. Various names have been given to these parts by different manufacturers, often leading
to confusion. Figures 1.9, 1.10, 1.11, 1.12 and 2.9 show typical constructions of a horizontal overhung
pump, a horizontal double-suction volute pump, a horizontal multistage pump (radially split), the bowl
section of a single-stage axial-flow propeller pump, and a vertical overhung pump. Names recommended
by the Hydraulic Institute for the various parts are given in Table 1.1.

EVOLUTION

The reader may be interested in comparing centrifugal pumps of the 1900s with their modem counterparts.
These are illustrated in Figs. 1.13 through 1.19, which clearly show many of the changes in mechanical
construction that were necessary for improved service life and maintenance. Other changes simply reflect
refinements in design, foundry, or machine shop practice.

Table 1.1 Recommended Names of Centrifugal Pump Parts


These parts are called out in Figs. 1.9, 1.10, 1.11, 1.12 and 2.9

Item no. Name of part Item no. Name of part

I Casing 35 Cover, bearing, inboard


IA Casing (lower half) 37 Cover, bearing, outboard
IB Casing (upper half) 39 Bushing, bearing
2 Impeller 40 Deflector
5 Diffuser 40A Deflector, inboard
5A Diffuser, first stage 40B Deflector, outboard
5B Diffuser, discharge 50 Locknut, coupling
6 Shaft 55 Bell, suction
7 Ring, casing 56 Drum, balancing
8 Ring, impeller 59 Cover, handhole
8A Ring, impeller, eye 60 Ring, oil
8B Ring, impeller, hub 63 Bushing, seal housing
9 Cover, suction 65 Seal, mechanical, stationary element
11 Cover, casing 72 Collar, thrust
13 Packing 73 Gasket
14 Sleeve, shaft 73A Gasket, casing
15 Discharge bowl 73B Gasket, bearing cover, inboard
16 Bearing, inboard 73C Gasket, bearing cover, outboard
17 Gland 73D Gasket, compensator
18 Bearing, outboard 78 Bearing spacer
19 Frame 80 Seal, mechanical, rotating element
20 Nut, shaft sleeve 85 Tube, shaft enclosing
22 Locknut, bearing 86 Ring, thrust, split
24 Nut, impeller 89 Seal
25 Ring, suction cover 99 Housing, bearing
27 Ring, casing cover 101 Pipe, column
29 Seal cage 103 Bearing, connector
31 Housing, bearing, inboard 115 Bushing, balancing
32 Key, impeller 123 Cover, bearing, end
12 Classification and Nomenclature

22

46

7 2 27 63 80 65 40A 16 60 37.

Fig. 1.9 Horizontal sil tgle-stage overhung refinery pump.


(Numbers refer 0 parts listed in Table 1.1.)

Fig. 1.10 Horizontal single-stage double-suction volute pump.


(Numbers n(er to parts listed in Table 1.1.)
Classification and Nomenclature 13

25

Fig. 1.11 Horizontal multistage double casing (barrel) diffuser pump.


(Numbers refer to parts listed in Table 1.1.)

Fig. 1.12 Vertical wet-pit diffuser pump bowl.


(Numbers refer to parts listed in Table 1.1.)
14 Classification and Nomenclature

Fig. 1.13 Old double-suction pump with separately cast suction elbows.

Fig. 1.14 Double-suction pump evolved from that in Fig. 1.13.


(Features integrally cast suction and discharge passages, separate casing heads, and radially split casing.)
Classification and Nomenclature 15

Fig. 1.15 Modern double-suction single stage pum p with an axially split casing.

Fig. 1.16 Old style multistage pump with radially split casing.
16 Classification and Nomenclature

Fig. 1.17 Modem high-speed high-pressure multistage pump (for over 175 bar gauge [2,500 psig))
with a radially split casing.

Fig. 1.18 Early motor-mounted pump (around 1905).


Classification and Nomenclature 17

Fig. 1.19 Modem close-coupled pump.

BIBLIOGRAPHY

[1.1] ANSI/HI 1.1-1.5, 1994, Centrifugal Pumps, Hydraulic Institute, Parsiopany, NJ, USA.
2
Casings and Diffusers

The term "casing" is generally used to describe the component that "contains" the pump. In practice a
pump's casing has five functions:

1. Provide pressure containment.


2. Incorporate the collector, either as an integral part of the casing or as a separate piece.
3. Permit rotor installation and removal.
4. Support the pump or provide a structural connection to the pump's support.
5. Maintain alignment of the pump and its rotor under the action of pressure and reasonable piping loads.

Traditionally pump casings have been classified by the type of collector used, so that aspect is dealt
with first.
The purpose of the collector is to gather and diffuse the high velocity liquid discharged by the impeller.
This process is necessary to (1) slow the liquid to a usable velocity and (2) convert the kinetic energy
into pressure energy, thus recovering more of the pump's energy input. Two basic forms of collector
are in use: the volute and the diffuser.
The volute-casing pump (see Fig. 1.2) derives its name from the spiral-shaped casing surrounding
the impeller. This casing section collects the liquid discharged by the impeller and converts velocity
energy into potential energy. A centrifugal pump volute increases in area from its initial point until it
encompasses the full 360 deg around the impeller and then flares out to the final discharge opening.
The wall dividing the initial section and the discharge nozzle portion of the casing is called the tongue
of the volute, or the "cut-water." The diffusion vanes and concentric casing of a diffuser pump fill the
same function as the volute casing in energy conversion.
In propeller and other pumps in which axial-flow impellers are used, it is not practical to use a volute
casing; rather, the impeller is enclosed in a pipe-like casing. Normally, diffusion vanes are used behind
the impeller proper, but in certain extremely low head units, these vanes may be omitted.
A diffuser-type centrifugal pump was illustrated in Fig. 1.4. The development of the diffuser appreciably
improved the efficiency of the rather crude volute forms characteristic of the early days of centrifugal

18
I. J. Karassik et al., Centrifugal Pumps
© Chapman & Hall 1998
Casings and Diffusers 19

pump construction. Later improvements in the hydraulic desigJ I of impellers and volute casings made
the diffuser of little, if any, value in increasing pump efficiency. It is therefore seldom applied to a
single-stage volute pump, although it possesses structural as '",ell as hydraulic advantages that may
sometimes be useful.
The principal advantages of a diffuser-type collector are that °or a given stage performance, it allows
a smaller stage volume than an equivalent volute-type collector, and, with appropriate design and
construction, produces negligible radial reaction on the rotor. As a consequence, diffuser-type collectors
are used for all propeller and vertical turbine pumps (Fig. 2. L), most high-pressure, double-casing,
multistage pumps (Fig. 1.11), and some special design of meditm head pumps (Figs. 1.6 and 2.13).
Against the advantage of a smaller pump, a poorly designed diffuser can impair the hydraulic
characteristics of the pump. The fast-moving liquid from the inlpeller can meet the fixed vanes of the
diffuser without shock only when the pump is operating at rated capacity, for only then does the angle
of the vanes correspond to the angle at which the liquid leaves 1he impeller. At all other rates of flow,
the multiple vanes cause shock and turbulence, so that the pumr may operate in an unstable condition.
As a matter of fact, when flow is restricted to as low as 5-10 peJcent of normal capacity, the shock and
turbulence may become sufficiently severe to reduce the total llead generated. As a result, the head-

Fig. 2.1 Vertical mixed-flow pump with diflusion vanes.


20 Casings and Diffusers

capacity curve of diffuser-type pumps could easily acquire a "droop" in the shut-off capacity area,
making the pump unsatisfactory for parallel operation. Do not interpret this to mean that a diffuser pump
always produces a drooping characteristic. However, such a curve can result from this design unless
extreme care is taken in layout of the impeller and diffuser combination. Pump manufacturers have long
tried to stabilize diffuser pump head-capacity curves, and various solutions are available, based on proper
selection of impeller vane angles, curvature of the impeller blades, and careful design of diffuser pas-
sageways.
Another problem arising from the use of diffusers is potential pump flexibility. Obviously, pump
manufacturers try to obtain as much coverage from a single pump pattern as possible, to keep the number
of patterns comprising a complete line of pumps at a minimum and to reduce the necessary number of
parts in stock. With a volute pump, the impeller diameter may be decreased as much as 20 percent from
its maximum value without appreciably reducing the pump efficiency caused by increased hydraulic
losses. On the other hand, similar reduction in diameter of a diffuser-type pump impeller would produce
unacceptable performance. The increased gap between the impeller periphery and the diffuser inlet vanes
would result in excessive hydraulic losses. For this reason, a maximum-diameter impeller can be cut
only from 5 to 10 percent. Further reduction requires a different diffuser pattern with a smaller inlet
vane diameter.
Impeller cutdown restrictions necessitate an increased parts inventory. Also, the flexibility of con-
structed units is limited because a change in conditions of service, otherwise taken care of by an impeller
cutdown, may also require a new diffuser for satisfactory performance.

RADIAL THRUST

In a single-volute pump casing design (Fig. 2.2), uniform or near-uniform pressures act on the impeller
when the pump is operated at design capacity (which coincides with the best efficiency). At other
capacities (Fig. 2.3), the pressures around the impeller are not uniform, and there is a resultant radial

Fig. 2.2 Zero radial reaction in single-volute casing. Fig. 2.3 Radial reaction in a single-volute casing.
Pressure distribution is uniform at design capacity. Pressure distribution is not uniform at
off-design capacity.
Casings and Diffusers 21

100

80
Radial
Thrust
% Thrust 60
@O%
Flow
40

20

20 40 60 80 100 120 140 160 180


Flow-% BEP

Fig. 2.4 Characteristic of radial reaction in a single-volute casing.


F decreases from shutoff to design capacity and then increases with overcapacity. At overcapacity,
the reaction is roughly in the opposite direction from that at partial capacity.

reaction (F). A graphical representation of the typical change in this force with pump capacity is shown
in Fig. 2.4; note that the force is greatest at shutoff 1
For any pt>r(,~!l!::gc of ~apacIty, radial reaction is a function of the total head and of the width and
diameter of the impeller. The magnitude of the reaction, Fr in kN (16), can be estimated using the
following relationship:

kD 2W 2H(SG) . . . F kD2WzH(SG). US .
Fr = 10.21(104) III metnc umts or r = 2.31 III umts

where: k = radial thrust factor


D2 = impeller diameter, mm (in)
W2 = impeller width, induding the shrouds, mm (in)
H =pump total head, m (ft)

The radial thrust factor k is determined by experiment, and varies with both percentage of design capacity
and pump specific speed (see Fig. 2.5). Zero radial reaction is not often realized. The more usual
characteristic is that shown in Fig. 2.4, in which the reaction is greatest at shutoff, least in the region
of best efficiency point (BEP). To complicate matters a little further, the direction of radial reaction
varies with capacity, typically being 90 deg away from the tongue at shutoff, moving to approximately
270 deg away at flows on the order 140 percent of BEP (see Fig. 2.6).
In a centrifugal pump design, shaft diameter and bearing size can be affected by allowable deflection
as determined by shaft span, impeller weight, radial reaction forces, and the torque to be transmitted.
Formerly, standard designs with maximum diameter impeller were only suitable for operation down to

lIn Fig. 2.2, 2.3, 2.4, and 2.7, no attempt has been made to show correct quantitative force values for a specific example nor
to locate the exact resultant force. The magnitude and direction of forces vary with the type of pump, casing design, and many
other factors.
22 Casings and Diffusers

0.4

0.3 ..
~
u: c5 0.2
a...

::.:::
0.1
".".".".~
• ". .... --- 100%
o 20 40 60 80 100 120 140 160
N5 - Specific Speed

Fig. 2.5 Radial thrust factor k.

90°

0% 0°

180° -------------------4~-----------------

140%

Ns = 1,700

Fig. 2.6 Direction of radial thrust.

50 percent of design capacity. For sustained operations at lower capacities, the pump manufacturer, if
properly advised, would supply a heavier shaft, usually at a much higher cost. More recently, sustained
operation at extremely low flows without informing the manufacturer at the time of purchase became a
common practice. The result-broken shafts, especially on high-head units.
Because of the increasing operation of pumps at reduced capacities, it has become desirable to design
Casings and Diffusers 23

standard units to accommodate such conditions. One solution is to use heavier shafts and bearings.
Except for low-head pumps in which only a small additional load is involved this solution is not
economical. The only practical answer is a casing design that develops a much smaller radial reaction
force at partial capacities. One of these is the double-, olute casing design, also called twin-volute or
dual-volute.
The application of the double-volute design principle to neutralize reaction forces at reduced capacity
is illustrated in Fig. 2.7. Basically, this design consists of two 1S0-deg volutes; a passage external to
the second joins the two into a common discharge. Although a pressure unbalance exists at partial
capacity through each ISO-deg arc, forces F, and F2 are approximately equal and opposite, thereby
producing little, if any, radial force on the shaft and bearings. Although the double-volute casing design
principle has been known for a long time, broad use of it had to await the development of improved
foundry techniques. The problems were two. In axially split casings, the second volute or "splitter"
spanned the split joint, and so casting accuracy had to be improved to avoid mismatch between the
casing halves. (Fig. 2.S) In all casings, there was a lower limit to the size of the cored passage that had
to be cast to produce the second volute. By the 1960's, these difficulties were overcome to the extent
that double volute construction became feasible in a commercial line of intermediate and large-size
double suction pumps. At the current state of the art, double volute construction is generally available
in pumps of 100 mm (4 in) discharge and larger.
The double-volute design has many "hidden" advantages. For example, in large-capacity medium-
and high-head single-stage vertical pump applications, the rib forming the second volute and separating
it from the discharge waterway of the first volute strengttens the casing (Fig. 2.9).
When the principle of the double volute is applied t< I individual stages of a multistage pump, it
becomes a twin-volute. The question has been broached wbether this design should be called twin-volute
or a two-vane diffuser, but the first has become the accepted form. A typical twin-volute is illustrated
in Fig. 2.10. The kinetic energy of the water discharged from the impeller must be transformed into
pressure energy, then turned back ISO deg to enter the impeller of the next stage. The twin-volute,
therefore, also acts as a return channel. The back view in Fig. 2.10 shows this, as well as the guide
vanes used to straighten the flow into the next stage.
A double volute is not always practicable. In small low· specific-speed designs the second or hidden

DOUBLE-VOLUTE WALL
Fig. 2.7 Radial reactions in double-volute pump. Fig . 2.8 Transverse view of double-volute
casing pump.
24 Casings and Diffusers

Fig. 2.9 Sectional view of vertical-shaft end-suction pump with a double-volute casing.
Numbers refer to parts listed in Table 1.1.

passage can become too small for reliable casting and cleaning. There may be a need to line the casing
waterway for corrosion or erosion resistance. The nature of the liquid pumped may preclude small
passages or fine tongues. In these cases, an alternative is a modified single volute. The modification
entails making part or all of the volute concentric. Figure 2.11 compares the shape and radial thrust
characteristic of these two modified forms with that of a conventional single volute.
The third approach to reducing radial thrust is to use a diffuser instead of a volute. Because a diffuser
is analogous to a series of small volutes spread equally around the impeller, any hydraulic reactions on
the impeller tend to be balanced. There are, however, limitations. Diffusers can develop quite high values
of radial reaction (Fig. 2.12) if there are irregularities in the vanes, if the impeller is not concentric with
the vanes or, in some designs, when the flow is reduced to the point where the diffuser develops rotating
Casings and Diffusers 25

Fig. 2.10 Twin-volute of a mul :istage pump.


Front view (left) and back view (right).

Spiral Volute
(conventional) Semi-concentric Concentric

t t t
Head Head
1-110 .
...\i....... .
........, Head .................,.,.,
HlO.
I I t " ."
...f.".#. I "'"
100% 100% ~
'.'.
100% m""
.'./,1 I,,.,,
'"

0-100% 0-100% 0-100%


Note: Radial force plotted as a percentage of the force at thrust off for spiral volute.

Fig. 2.11 Characteristic of radial reaction in modified volutes (semi-concentric and concentric) versus
normal single volute.
26 Casings and Diffusers

R
. egular@10-15%N
(
r Irregular

I. . .>---'-----t·..· ..

,,---,
, ,
~
I ,

, , Rotating Force
, ~ / (Diffuser Stall)
Radial
Force ,,
,,
~
~

,
,,
~

,, Stationary Force
(Rotor Eccentricity)

I
100
Flow-%BEP

Fig. 2.12 Possible radial reactions in a diffuser.

stall (Fig. 2.12). The first two reactions are generally stationary, the third rotates at a low frequency,
typically 10-15 percent of running speed.
Efficiency can also influence the choice of collector design. Taking a well designed single volute as
a basis, the following general observations can be made:

1. Double volute-lower peak efficiency in small sizes, comparable in larger; characteristic similar.
2. Modified volute-lower peak efficiency in all sizes; characteristic broader thus offering comparable overall
power consumption in pumps running over a wide flow range.
3. Diffuser-higher peak efficiency in all sizes; efficiency characteristic narrower, thus tending to offset the
advantage of higher peak efficiency in pumps running over a wide flow range.

Some designs have a collector arrangement incorporating a diffuser discharging into a single volute
(Fig. 2.13). Resorting to this arrangement offers two advantages. First, in small- and middle-size pumps
the higher peak efficiency of the diffuser is realized. Second, the point of peak efficiency can be shifted
while maintaining high peak efficiency by changing both the impeller and the diffuser. Because the
Fig. 2.13 Single stage pump with a separate diffuser discharging into a single volute casing.
(Courtesy Sulzer.)

~
28 Casings and Diffusers

diffuser is discharging into a volute, this collector arrangement still exhibits radial thrust but of lower
magnitude than a volute alone.
Volutes can be either an integral part of the casing (Figs. 2.8 and 2.9) or separate and bolted into the
casing (Fig. 2.10). Except in very large pumps, diffusers are separate (Fig. 2.13), a consequence of the
care needed to produce their small passages.

SOLID AND SPLIT CASINGS

Solid casing is a design in which the discharge waterways that lead to the discharge nozzle are all in
one casting or fabricated piece. It must have one side open so that the impeller can be introduced into
the casing; however, it cannot be completely solid, and designs normally called solid casing are really
radially split (Figs. 2.14, 2.16, 2.17, and 2.18).
Split casing is a casing made of two or more parts. The term "horizontally split" had regularly been
used to describe horizontal double-suction pumps, indicating that the casing was divided by a horizontal
plane through the shaft centerline or axis (Fig. 2.15). That designation was an unfortunate choice because
applications of the same pump design for vertical use or with the nozzle position rotated caused confusion.
The term "axially split" is now preferred. Since both the suction and discharge nozzles of axially split
pumps are usually in the same half of the casing, the other half may be removed (upper half in the case
of horizontal pumps) for inspection without disturbing the bearings or piping.
Like its counterpart, "horizontally split," the term "vertically split" is unfortunate. It refers to a casing
split in a plane perpendicular to the axis of rotation. The term "radially split" is now preferred.

Fig. 2.14 Radially split, foot mounted, overhung pump with stuffing box head or cover for
back-pull-out capability.
Casings and Diffusers 29

Fig. 2.15 Axially split casing, horizontal, double suction volute pump.

Fig. 2.16 Vertical overhung pump with an elbow t) pe suction nozzle.


30 Casings and Diffusers

Fig. 2.17 Horizontal overhung process pump with a fiat elbow type suction nozzle.

A now obsolete casing configuration for horizontal pumps had an axial split but with the joint surface
inclined to the horizontal. The virtue claimed for this construction was the utility of top discharge
combined with the convenience of an axial split.
Whether to split a casing radially or axially depends on the impeller configuration, whether the pump
is single or multistage, the pressure to be contained, and to some degree the shaft orientation. Pressure
containment is influenced by liquid specific gravity (SG) and temperature, low SG or high temperature
or both tending to increase the split joint design pressure necessary to ensure a tight seat. Shaft orientation
reflects the influence of gravity on assembling the impeller or element into the pump's casing. For the
various combinations of these factors there is a casing split that represents the lesser manufacturing cost.
Usual casing splits for the more common pump arrangements and how the various factors influence the
choice is shown in Table 2.1. When factors beyond those previously listed have a bearing on the design,
the choice of casing split can be quite different.

SUCTION NOZZLE

Centrifugal pumps are sensitive to the flow distribution at the impeller inlet, and the sensitivity increases
with specific speed. Depending on its design, the suction nozzle can have a significant influence on the
flow distribution at the impeller inlet, thus it is an important part of a casing. In many instances, the
suction nozzle arrangement that provides the easiest installation also penalizes pump performance, raising
Casings and Diffusers 31

Table 2.1 Typical Casing Split for Various Pump Configurations and Pressure Ratings

Impeller Single Suction Double Suction

Number of Stages Single


lMulti Single

Pressure All Low & Medium High

Shaft Orientation Either


I~
Horiz. Vertical Either

Casing Split Radial Axial Radial Radial

the net positive suction head required (NPSHR) and lowering efficiency. When this occurs, it is necessary
to strike a compromise between installation cost and energy wnsumption.
With only subtle variations, all pumps use one of three basic suction nozzle arrangements: end, elbow,
or flat elbow. End suction (Fig. 2.14) has the liquid entering the impeller eye without any turning.
Because the liquid approach is so direct, this arrangement offt~rs the best potential for low NPSHR and
high efficiency, and is therefore widely used in pumps with single-suction impellers.
An elbow suction is used for vertical single-suction pumps when the piping arrangement make it
desirable to tum the flow through 90 degrees before entering the impeller (see Chapter 14). With proper
design (large radius and tapered), these elbows do not seriously affect the pump's NPSHR and efficiency.
Refinement of the elbow suction to a flat elbow (Figs. 2.16 and 2.17) significantly shortens the axial
space required for the nozzle with only a minor effect on pump NPSHR and efficiency. Flat elbow
suction nozzles are particularly important for reducing bearing span when the impeller is located between
bearings (Figs. 2.13 and 2.15).
32 Casings and Diffusers

PREROTATION AND STOP PIECES

On occasion, the liquid in the suction pipe may swirl for some distance ahead of the actual impeller
entrance. This phenomenon is called "prerotation." When this occurs there are two possible causes: One
is that the pump is operating below its suction recirculation capacity and the prerotation is being induced
by same (see Chap. 22). The second is that the incoming flow is being distorted by the suction entrance,
piping, or nozzle. In the first case, the effect on pump performance is evident from test results, and can
only be corrected by changing the pump's impeller design. In the second case, the pump's performance
is often impaired (high NPSHR, low head and efficiency) because the liquid is not entering the impeller
as the design intended.
To confer some tolerance of flow distortions introduced by suction piping, most pump suction nozzles
include a stop vane or vanes to straighten the flow as it approaches the impeller. An exception to this
is pumps intended for services such as sewage and paper stock, in which solids in the liquid would
likely buildup on the stop vane and block the pump.

RADIALLY SPLIT CASINGS

At least one side of a radially split casing must be removable so the impeller or element can be installed
in the casing. Single-stage end-suction pumps are the most common example of radially split casings
and are built with one of three cover arrangements. If the cover is on the suction side, it becomes the
casing sidewall and contains the suction opening or nozzle (Fig. 2.18). This is called the suction cover
or casing suction head. Other designs are made with a stuffing box cover (Fig. 2.14), whereas still others
have both a suction cover and a stuffing box cover (Fig. 2.19).
For general service, the end-suction single-stage pump design is used extensively for both close
coupled (motor mounted) and separately coupled pumps up to 8-in. discharge size. In all of these, the
small size makes it feasible to cast the volute and one side integrally. Whether the stuffing box or suction
side is made integrally with the casing is usually determined by the service for which the pump is
intended. Designs for general, chemical, and process service have a so-called stuffing box or casing

Fig. 2.18 Section of frame-mounted end-suction pump with radially split casing.
Note the suction cover.
Casings and Diffusers 33

Fig. 2.19 Section of radially split, foot mounted end-suction pump with two covers (suction and stuffing box).

Fig. 2.20 Section of frame mounted, radially split, end-suction slurry pump with stuffing box cover and
replaceable casing liners.
34 Casings and Diffusers

cover (Fig. 2.14). For separately coupled pumps, provided the appropriate spacer coupling is furnished,
the virtue of this arrangement is the ability to remove the entire bearing frame, cover, and impeller
assembly without disturbing the pump's driver or suction and discharge piping connections. Such an
arrangement is often referred to as "back pull-out." When the casing is prone to high wear, as in slurry
service, or there is an advantage to being able to vary the discharge nozzle orientation (see nozzle
locations), the casing is frame mounted with a suction cover (Figure 2.18) or a stuffing box cover (Fig.
2.20). Dismantling pumps of this arrangement without disturbing the driver or the pump's bearing frame
requires a spool in the suction piping.
Radially split casings for other than end suction are available (see under nozzle locations). For pumps
of two or more stages, radially split casings are necessarily more complicated than for a single stage.
A detailed treatment of such casings is given in Chapter 3.

AXIALLY SPLIT CASINGS

Most axially split casings are designed with their mounting feet integral with the lower half (Fig. 2.15)
or with the fixed half in the case of pumps mounted with the split vertical. By being built in this way,

Fig. 2.21 Section of single stage volute pump with axially split casing and separate stuffing boxes.
Casings and Diffusers 3S

the pump can be dismantled without disturbing the driver (provided the coupling halves will separate
normal to the shaft axis) or the pump's suction and discharge connections.
Single-stage designs for general service usually have the stuffing boxes cast integrally with the casing
(Fig. 2.15), since that is an economical arrangement. For more severe service, where stuffing box cooling
or a special mechanical shaft seal is necessary, it is often better to employ a separate stuffing box (Fig.
2.21), to avoid the difficulty of having to cast a complex shape as part of the casing.
Bearing support is either by integral brackets (Fig. 2.15) or by brackets built into the bearing housings
and bolted to the casing sidewall (Fig. 2.21).

CASING CONSTRUCTION FOR OPEN IMPELLER PUMPS

In the simplest open impeller pump, the close axial clearances needed to seal the impeller are formed
by the casing and cover or head directly (Fig. 2.22). Such construction is employed in low-cost pumps
because it is simple; and in many chemical pumps, because it has the least number of potential crevice
corrosion sites. When the service is abrasive, the casing and in severe services the head are equipped
with a replaceable liner or wear plate (Figs. 2.20 and 2.23). Adopting this construction allows the wear
plate to be of a more abrasion-resistant material when necessary and enables ready replacement of the
wearing surface.
Except in very rare instances, the use of wear plates is limited to single-suction impellers. Early
designs extended this construction to double-suction impellers, but the difficulty of setting up and
maintaining the close axial clearances had rendered such arrangements obsolete.

Fig. 2.22 End suction chemical pump with semi-open impeller running directly against casing.
36 Casings and Diffusers

I
-1---- -j- -
[ -- - - - - -.- - - - - - - ]
--- , -
1
: l Discharge
I
I:

Fig. 2.23 End suction chemical and stock pump with semi-open impeller running against a wear-plate.

NOZZLE LOCATIONS

End suction single stage horizontal pumps are usually arranged for top-vertical discharge (see Figs. 2.14,
2.18, 2.20, and 2.22). Other positions, however, may be obtained when necessary. The most common
variations are top-horizontal (overshot), bottom-horizontal (undershot), and bottom-vertical. Figure 2.24

A B
o F

Fig. 2.24 Possible positions of discharge nozzles for horizontal end-suction solid-casing frame mounted pump.
Rotation illustrated is counterclockwise from suction end.
Casings and Diffusers 37

shows these positions plus intermediate positions (inclined) that may be available if the casing attachment
allows. Generally, the alternate nozzle positions are employed to simplify piping arrangements, save
space, or reduce pipeline erosion by the elimination of an elbow. Unless the casing is frame mounted
(Figs. 2.18 & 2.20), the provision of alternative nozzle locations requires a change in casting or a casing
with multiple mounting pads. With some older frame-mounted designs, interference between the discharge
flange and the bearing frame or base precludes the two bottom nozzle locations. In other instances,
access to auxiliary connections, such as stuffing box sealing, limits casing rotation.
Single-stage radially split horizontal pumps with an elbow or flat elbow suction are usually arranged
with both suction and discharge top-vertical (Fig. 2.17). Such an arrangement is generally known as
"top-top," referring to the suction first, then the discharge. Top-top nozzles are separated axially in
single-suction pumps (Fig. 2.17) and are in the same axial plane in double-suction pumps (Fig. 2.25).
When the piping arrangement dictates it, pumps of the form in Fig. 2.25 can be furnished with "side-
side" instead of top-top nozzles.
Axially split casings always have the nozzles in the lower half, or what is termed the fixed half when
the pump axis is vertical, so the pump can be dismantled without breaking the suction and discharge
piping connections. Horizontal pumps, whether single or double suction, almost invariably have a side
discharge nozzle and either side or a bottom suction nozzle. If the suction nozzle is placed on the side
of the pump casing with its axial centerline at right angles to the vertical centerline (see Fig. 2.15), the
pump is classified as a side-suction pump. If its suction nozzle points vertically downward (Fig. 2.26),
the pump is called a bottom-suction pump. Single-stage bottom-suction pumps are rarely made in sizes
below lO-in. discharge nozzle diameter.

Fig. 2.25 Centerline supported, radially split single stage pump with top suction and discharge nozzles.
38 Casings and Diffusers

Fig. 2.26 Bottom-suction single stage axially split casing, single stage pump.

Special nozzle positions can sometimes be provided for double-suction axially split casing pumps to
meet special piping arrangements, for example, a vertically split casing with bottom suction and top
discharge in one half of the casing. As these special designs are usually costly, they should be avoided.

CENTRIFUGAL PUMP ROTATION

The direction of rotation of a centrifugal pump is a fundamental element of its specification; there are
only two directions, and running a pump in the wrong direction has dire consequences on its performance.
In many designs, the direction of rotation also affects the location of the suction and discharge nozzles,
so it's important in this discussion of casings to address the means used to define the direction of rotation.
According to Hydraulic Institute Standards, rotation is defined as clockwise or counterclockwise by
looking at the driven end of a horizontal pump or looking down on a vertical unit. Some manufacturers
still designate rotation of a horizontal pump from its outboard end. Therefore, to avoid misunderstanding,
clockwise or counterclockwise rotation should always be clarified by including the direction from which
one looks at the pump.
Casings and Diffusers 39

The tenns "inboard end" and "outboard end" are used only with horizontal pumps. Inboard end is
the one closest to the driver, whereas the outboard end is the one farthest away. The tenns lose their
significance with dual-driven pumps and are not then used. Many centrifugal pump casings produced
from one pattern can be built into a pump of either clockwise or counterclockwise rotation. One such
design is the axially split, single-stage double suction pump (Fig. 2.15). Other designs such as end-
suction and some radially split double suction pumps have an integral head on one side, and therefore
require separate directional casing patterns.

CASING HAND HOLES

Casing hand holes are furnished primarily on pumps handling sewage or stringy materials that may
become lodged on the impeller suction vane edges or on the tongue of the volute. They pennit removal
of this material without dismantling the complete pump. End-suction pumps used for handling such
liquids are provided with hand holes for access to the suction side of the impellers. These are located
on the suction head or in the suction elbow. Hand holes are also included in drainage, irrigation,
circulating, and supply pumps if foreign matter may become lodged in the waterways. On very large
pumps, manholes provide access to the interior for both cleaning and inspection.

MECHANICAL FEATURES OF CASINGS

Most single-stage centrifugal pumps are intended for service with moderate pressures and temperatures.
As a result, pump manufacturers usually design a special line or lines of pumps for high operating
pressures and temperatures rather than make their standard line unduly expensive by having it cover too
wide a range of operating conditions.
Casings for higher pressure or temperature or both tend to be radially split, despite the difficulties
this arrangement entails in double-suction pumps and large horizontal pumps of either single- or double-
suction configuration. Axially split casings have inherent limitations, which generally restrict their
economical use to low and intennediate pressures. The two principal limitations are both products of
casing deflection under the action of pressure. First, when the bearing brackets are attached to the casing
sidewall, as they are in double-suction pumps (Fig. 2.15), "ballooning" or bulging of the casing sidewalls
rotates the brackets slightly, which raises the pump's rotor. In extreme cases, the rotor can be raised to
the extent it contacts the internal running clearances, thus rendering the pump inoperable. For the usual
design pressures, internal or external ribbing is used to avoid this deficiency.
Second, deflection of the casing, the bolting flange, and the bolting itself reduces the residual gasket
load, with the greatest reduction at the innennost diameter of the joint. If the gasket load is reduced too
much, the gasket is washed out, allowing internal leakage and eventually casing erosion. Chapter 3
discusses this problem and its solution in detail.
Most pumps are supported directly by feet on their casing or indirectly by feet on a frame or pedestal.
The virtue of this arrangement is a simple connection to the foundation, whether through a baseplate or
directly to the foundation. For higher pumping temperatures, the change in rotor position caused by
thennal expansion of the casing or frame poses a major coupling alignment problem. To avoid this
difficulty, pumps for higher temperatures are usually supported by feet located at or very close to the
casing centerline (Figs. 2.17 and 2.25). The temperature at which centerline support is necessary varies
with casing height and coupling misalignment capacity; a commrnon industry standard is 175°C (350°F).
With centerline support, the base or foundation must include pedestals to reach the pump's feet.
40 Casings and Diffusers

Contrary to past opinion, practicality and more sophistication in piping design mean that pump casings
now must also withstand moderate loads from the connected piping. Two effects should be considered
in assessing a particular design's piping load capability: distortion of the pump casing proper, leading
to contact at the internal clearances, and distortion of the pump as a whole, leading to misalignment at
the coupling. Close-coupled pumps are not, of course, influenced by the second effect, because driver-
to-pump alignment does not depend on a connection through earth. This feature affords close-coupled
pumps a notable advantage.
Casing distortion from piping loads is a function of pressure rating, material, and the path of load
transmission to the foundation. In general, casings designed for higher pressures will have relatively
higher piping load capability. The importance of material is stiffness, because for the same stress, bronze
and cast iron deflect about twice as much as steel. Thus, changing to a stiffer material offers lower
distortion if the stress levels are similar. Ideally, piping loads should pass directly from the nozzle to a
mounting foot and thence to the foundation. By doing this, the extent of the casing subjected to higher
loads is minimized, hence casing distortion is minimized. A casing with feet on its nozzles (Fig. 2.15
& 2.21) has the path of piping load transmission approaching the ideal. Centerline-supported pumps can
also approach the ideal provided the design seeks to realize a direct structural connection between the
nozzles and mounting feet rather than through the casing.
To maintain coupling alignment, the prime distinction is the number of points of support, whether
two, three, or four points. Most centerline-supported overhung pumps (Fig. 2.17) and some high-
temperature foot-supported pumps are two-point supported. Because even relatively small deflections of
the feet and pedestal are usually magnified substantially as displacement of the shaft at the coupling,
two-point-supported pumps generally have the lowest piping load capability. Depending on the stiffness
of the pump and its bearing frame, three-point support (Fig. 2.14) can significantly increase piping load
capacity, a result of the third support close to the shaft end. The difficulty with this design is that it
turns the bearing frame and the bracket connecting it to the casing into structural elements subject to
part of the piping loads. Provided the bearing frame is sufficiently stiff to accommodate the additional
loading while maintaining bearing, seal, and running clearance alignment, this arrangement is quite
viable. For high-temperature pumps, three-point support with a foot-mounted casing requires care lest
differences in thermal expansion of the supports cause bearing frame distortion. To a lesser degree, but
still of some consequence, the same concern exists for centerline-supported casings. Four-point support
(Fig. 2.25) offers the greatest piping load capability, and although more expensive, should be given
serious consideration for services where high piping loads are likely. Two factors account for this. First,
the greater piping load capability may allow a less expensive piping layout. Second, the pump will likely
require less maintenance than an equivalent two- or three-point-supported overhung pump.

SERIES UNITS

For large-capacity, medium-high-head service conditions that require such an arrangement, two single-
stage double-suction pumps can be connected in series on one baseplate with a single driver. Such an
arrangement is very common in waterworks applications for heads of 75 to 120 m (250 to 400 ft). One
series arrangement uses a double-extended shaft motor in the middle, driving two pumps connected in
series by piping (Fig. 2.27). In a second type, a standard motor is used with one pump having a double-
extended shaft (Fig. 2.28). This latter arrangement may be limited, because the shaft of the pump next
to the motor must be strong enough to transmit the total pumping horsepower.
If the total pressure generated by such a series unit is relatively high, the casing of the second pump
may require ribbing.
Casings and Diffusers 41

Fig. 2.27 Series unit (motor in middle).

Fig. 2.28 Series unit (motor at end).


42 Casings and Diffusers

CASING MAINTENANCE

Pumps that handle noncorrosive water or liquids are not usually subject to extensive casing wear.
However, the casing waterways should always be thoroughly cleaned and repainted during a complete
overhaul. A suitable paint should be used that finnly adheres to the metal so that the water velocity will
not wash or jet it off.
An enamel-like finish is the most efficient. A program of casing cleaning and repainting should be
established on the basis of local conditions. This will prevent the protective coat from ever fully eroding
before replacement, thereby preventing corrosion.
Pumps handling gritty or sandy water naturally are more subject to casing troubles. Erosion can be
reduced by selecting pumps with low-liquid velocities-that is, low head per stage-or employing more
erosion resistant materials, or a combination of both. Caution is needed, however, because conventional
water pump designs and materials can tolerate only relatively low concentrations of sand or silt. As a
general guide, special materials should be investigated if the concentration exceeds 1 percent, and a
pump designed for slurry should be used if the concentration exceeds 2-3 percent.
Progress has recently been made in processes of rubber-coating pump waterways, and this technique
may be desirable in some applications. If the casings of pumps handling sandy or gritty water are to be
protected primarily by periodic painting, a suitable type pump should be carefully selected for local
water conditions and a special maintenance schedule established.
In these difficult pumping applications, the casing should be regularly examined for corrosion, which
will be indicated by cast-iron graphitization. This occurs when the ferrous particles are washed out by
electrolytic action and deposited on bronze pump parts. If severe graphitization takes place, the manufac-
turer should be consulted on the possibility of substituting materials more impervious to the pumped liquid.
If the casing is pitted or eroded in places, it can be restored by welding, brazing, silver soldering, or
metal spraying, depending on the material and the facilities available. The authors know of several large
centrifugal pumps in waterworks service in which corroded areas, located where water velocities are
low, are actually filled with properly anchored concrete.
Special care must be taken to examine and recondition meta1-to-metal fits where stationary parts such
as casing rings, diffusers, or stage-pieces seat in the casing. If the casing is steel, and these fits show
signs of erosion, it might be advantageous to face them with 18-8 stainless steel and refinish.
Frequently, the cut-water, or volute tongue as it is also called, becomes eroded, for example, when
a pump handles water with some sand in suspension or when the periphery of the impeller is located
too close to the tongue. Another fairly common cause of erosion in this area is galvanic action between
a cast-iron casing and bronze fittings. The cast iron graphitizes and wears away most perceptibly in
areas of high velocity, like those near the volute tongue.
The best way to correct this condition is to cut back the tongue so that it is straight across .and then
file it to a smooth rounded edge (Fig. 2.29). This cut back does not affect pump capacity unfavorably;
on the contrary, it is often used to squeeze out an extra small percentage of capacity without putting in
an impeller of larger diameter. The added capacity comes from the increase in the casing throat area,
which causes an increase for a given casing velocity. If the pump is double volute, it is prudent to cut
back both the tongues so they are 180 deg apart after correction. Doing this helps to ensure there is
minimal radial thrust after correction. In cases where this is not feasible, consult the manufacturer to
detennine whether asymmetric tongues can be tolerated.
Care should be taken not to distort or warp the casing during overhaul. After repairs are completed,
the horizontal flanges of an axially split casing should be finished to a flat surface with hand tools. Of
course, if the repairs are very serious, the pump is serviced at the manufacturer's shop. The casing
flanges may have to be refinished at that time and the casing rebored.
Except for some special designs, every pump has gaskets that are subject to damage when the pump
Casings and Diffusers 43

Fig. 2.29 Method of filing worn volute tongue.

is opened. If the old gasket adheres to the lower half of the casing and is in good condition, it is not
necessary to replace it. However, it should be replaced if it is damaged in any way, and for this reason
a new gasket should always be available.
The new gasket should be of the same thickness as the original and, if possible, of the same type of
material so that it will have the same compression characteristics. Too thick a gasket usually leads to
leakage. If the gasket is thinner than the original, tightening of the two casing halves may exert undue
force on casing wearing rings and distort them.
In installing a new gasket, the inner edge must be accurately trimmed along the edge of the stuffing
box bore. At all points where the gasket abuts on the outer diameter and the sides of stationary parts,
the edges must be trimmed squarely and neatly, allowing sufficient gasket overlap. Tightening the upper
half of the casing will effectively press the gasket edges against the stator parts, insuring proper sealing.
This trimming operation is best accomplished by first cementing the gasket to the upper casing half with
shellac (this makes for easier gasket removal at the next overhaul) and then cutting all edges square
with a razor blade. Of course, all foreign matter must be removed from the casing flanges before the
gasket is applied to the lower casing half.
In reassembling the pump, it is recommended that powdered graphite be rubbed into the gasket before
the top casing half is replaced. This action will prevent the gasket from sticking to the lower half when
the casing is next dismantled.
3
Multistage Pump Casings

Although most single-stage pumps have volute casings, both volute and diffuser casings are used in
multistage pumps. Advantages of the volute casing are described in Chapter 2. However, the diffuser
casing or collector is a very strong competitor of the volute design in high-pressure applications. The
principal reason for this is that for the same pump capacity, a diffuser design is smaller (occupies less
volume) than a volute, thus making a diffuser pump less expensive to manufacture. When compared to
integrally cast volutes, the volume advantage of a diffuser design is compounded by symmetry, an
absence of complicated castings, suitability for radially split construction, and more uniform expansion
in high-temperature service. Arranging twin volutes in the same manner as diffusers (Fig. 2.10) overcomes
all the advantages except that of lower volume for the same capacity. At that point, the choice between
twin volutes and diffusers must be made based on pump cost versus operating cost, the broader efficiency
characteristic of the volute being an advantage when the pump has to run over a wide flow range. There
is evidence, too, that at conditions far from BEP, twin volutes produce lower rotor forces than diffusers.
In smaller, lower pressure multistage pumps, the cost advantage of diffusers is not so significant. At
the same time, ease of manufacturing favours volute pumps, at least down to around 2-in. discharge, so
most such pumps have volute casings. In all but the smallest sizes, twin volutes are used. The smallest
sizes use single volutes to avoid the casting problems inherent in small double volutes, and balance
radial thrust by staggering successive volutes 180 deg. apart. Figure 3.1 shows the principle for a two-
stage pump. Older designs successfully used staggered single volutes in larger sizes than is generally
done today. The advantage these designs had was a relatively robust shaft, able to withstand the moment
produce between impeller pairs. Modem designs tend to use more slender shafts in the interests of higher
efficiency and lower NPSHR, and have therefore had to use a twin volute at each stage.

AXIALLY AND RADIALLY SPLIT CASINGS

Both axially and radially split casings are used for multistage centrifugal pumps. Variety in pump design
is based on these two successful solutions for casing construction problems, which were arrived at in

44

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
Multistage Pump Casings 45

Radial
Thrust

Stage 1

Fig. 3.1 Multistage volute pump arrangement for radial thrust balance.

the early days of the centrifugal pump. Each of these designs, incorporated in pump models over many
years, has followed its own separate evolution.
Although the radially split casing design might logically seem to be a recent development motivated
by higher pressure requirements, it is really a reversion to the original type. The first centrifugal pumps
had a vertically split casing, with radial joints in a plane perpendicular to the shaft axis. Radially split
casings have been in continuous use both here and abroad ever since, especially in high-pressure multistage
centrifugal pumps.
The idea of splitting the casing axially, or parallel to the axis of rotation, was introduced to overcome
radially split casing limitations on accessibility for inspection and repairs. The casing suction and
discharge connections were located in the lower half. The upper half could be lifted and the rotor laid
bare for inspection or removal without disconnecting the pump proper either from its foundations or
from the suction and discharge piping. This was the key influence in the development of the axially
split casing. The change was made solely for reasons of cost. Less machining was deemed likely to
reduce the manufacturing cost, and the greater convenience of dismantling and reassembly would certainly
reduce maintenance costs.
As long as pump working pressures remained within modest limits, axially split casings proved
satisfactory. Higher working pressures, however, exposed a basic weakness. As pressures increased, it
became more and more difficult to maintain a tight axial split flange. The consequences of this were
intemalleakage between the stages with subsequent casing erosion and occasional axial thrust problems
at the least; leakage of the pumped liquid to the atmosphere at the worst. Distortion of the joint flange
46 Multistage Pump Casings

was recognized as the problem. The first remedies were to increase the flange thickness, stiffen the
casing with ribs, and increase the number and size of casing bolts. Although these remedies increased
the allowable working pressures to some extent, there was still an upper limit beyond which reliable
axial split operation could not be ensured.
Axially split against radially split pump casings was one of the most eontroversial design topics until
late in the nineteen-thirties. Then, the 1938 to 1948 period produced a definite trend, with 85 bar gauge
(1,250 psig) becoming the average working pressure at which a user changed from axially to radially
split casings. For some time after that, designer opinion on the proper pressure limits for the two casing
designs varied considerably, with some favoring a reduction in the changeover point to as low as 70
bar gauge (1,000 psig) and others convinced that the axially split casing limit could be safely increased
to 110 bar gauge (1,600 psig).
The limiting factor in the history of axially split casings was the difficulty in making even rough
estimates of the stress and strain (distortion) in such a complex shape (see the discussion later). An
entirely satisfactory design could be realized by "cut and try" methods, but usually at greater expense
than was commercially tolerable. To realize consistently well behaved axial split joints at reasonable
cost, it was necessary for the technique known as "finite element analysis" (FEA) to come into general
use. With this technique, the casing is modeled as many connected blocks, and its stress and strain
contours are calculated for a given set of loads, including piping loads and constraints. The beauty of
FEA is that it takes account of the interdependence between stress and strain and allows the casing to
be "tested" while it is still just a design. The drawback is that the solutions can only be run on a computer;
the extent of calculation is well beyond manual techniques. With FEA the working pressure limit for
axially split casings has risen to 275 bar gauge (4,000 psig). Such casings, however, tend to be more
expensive than an equivalent radially split design, so are used only when other circumstances, such as
floor space or ease of maintenance, dictate them. The general economical limit of axially split casings
in services such as boiler feed is 170 bar gauge (2,500 psig). For applications handling petroleum based
liquids, the limit can be lower.
API-61O [3.1], taking account of the consequence of casing leakage in oil refineries, the greater risk
of leakage when handling petroleum products at elevated temperatures, and the difficulty of remaking
axial split joints in the field, limits axially split casings to rated pressures of 100 bar gauge (1,450 psig),
pumping temperatures of 230°C (450°F) and liquids of SG 0.7 and higher at the pumping temperature.
These are conservative limits and there are many axially split casings operating reliably beyond them.
In contemplating axially split casings beyond the limits of API-61O, the pump user needs to consider
the ratio of the casing's maximum allowable working pressure to the rated pressure (the higher the
better), actual operating experience with the same casing in similar conditions, and finally the availability
of the skills necessary to successfully remake the casing joint.

CASING UPPER PRESSURE LIMITS

To analyze a bolted and gasketed flange joint, we must visualize the phenomena occurring during the
initial tightening process and when internal pressure is applied. As the flanges are pulled tight by bolt
action, the bolts are subjected to a tensile stress. The gasket itself is subjected to a compression stress
determined by the bolt stress and the relative bolt and gasket areas. As internal pressure is applied, the
bolt stress increases and the bolts stretch. Assuming that the deformation of the gasket under initial
tightening is not permanent, gasket thickness will increase by the same amount that the bolts are increased
in length. This requires that the gasket stress not exceed the yield strength of the gasket material and
Multistage Pump Casings 47

that the material possess sufficient resiliency. Although the gasket compression stress is thus reduced
under internal pressure, it remains sufficiently high to prevent leakage between the flanges.
This analysis assumes that the flange surfaces are sufficiently smooth to give relatively uniform gasket
stress under internal pressure, the stress never falling below the minimum required for effective sealing.
If flange surfaces are rough, large local stresses may develop in the gasket at flange high spots during
the initial tightening. Conversely, the gasket may have low-stress areas at flange depressions even when
high stress is applied during the initial tightening. When internal pressure is applied, highly stressed
gasket areas may not match the increase in gasket thickness found in other areas, and low flange areas
not sufficiently stressed initially may fail to hold the internal pressure, causing leakage.
For these reasons, planed or milled flange surfaces are only suitable for low or medium pressures. If
a satisfactory flange joint is desired in a high-pressure range, the flange surfaces must be very carefully
ground and low spots eliminated.
Analyses of the tightness of bolted flange joints vary in complexity. If the flange is symmetrical (for
example, a pipe flange), the bolt load distributed uniformly over the entire gasket contact area, and the
gasket material perfectly homogeneous, it is relatively easy to calculate initial bolt and gasket stress,
elongation of bolts caused by internal pressure, and the resulting increase in gasket thickness. If the
restoration characteristics of the gasket material are known, it can then be determined whether the joint
is sufficiently tight for its intended service.
Axially split casing flanges in multistage pumps do not, however, lend themselves to simple analysis.
Figure 3.2 shows the horizontal flange of a six-stage pump. The bolting distribution is usually very
complex, making it necessary to use finite element analysis to study overall joint quality under varying
pressure at the individual stages. By definition any "model" of a part is an approximation, so prudence
requires that FEA be occasionally verified or calibrated by experiment. Stress is checked using strain
gages, and gross strain or deflection by checking the actual movement, or joint separation, at critical

Fig. 3.2 Horizontal flange of an axially split six-stage pump casing.


48 Multistage Pump Casings

areas within the casing under working pressure. The critical areas, of course, are the projecting casing
tongues that separate the various pump stages and that are farthest removed from the bolting. These
data must be supplemented by stress-deflection tests on the gasket material to permit comparison between
the casing deflection under internal pressure and restoration of gasket thickness under reduced stress.
The comparison will indicate whether the flange joint will remain tight under pressure or leak. Special
gages (Fig. 3.3) permit experimental measurement of casing stretch. One or more such gages are clamped
to the internal ring bores on the lower casing half. They are so arranged that the spring pushes the soft
copper wedges into the split at the adjacent tongue. The gage plunger is retracted before installation and
secured with the holding screw. To determine gasket compression induced by the initial bolting, small
pieces of gasket are removed at the tongues. The measurement is obtained by using feeler gages.
After the casing stretch gages are installed, the upper half of the casing is put in place and the bolts
tightened. As the shaft and impellers are not in place for this test, it is possible to reach into the stuffing

Holding Screw

___"""""--__ Slot

Clamp

Fig. 3.3 Casing flange deflection gage.


Used in experimental analysis of axially split casing pump designs.
Multistage Pump Casings 49

box opening and release the gage-holding screws. The soft copper wedges are thereby pressed against
the closed split under spring pressure, free to advance as casing deflection occurs. It is also possible at
this stage to reach in with a feeler gage and measure gasket compression under initial tightening.
Endplates are fastened over the stuffing box opening, and the casing subjected to the desired hydrostatic
test pressure. The copper wedges penetrate into the split as far as it is opened up. After the casing has
been held at test pressure for the required length of time, the pressure is released. The two casing halves
resume their original position, and an indentation is made on the soft copper wedges at an easily
measurable point. When the casing is opened and the gages removed, this indentation permits exact
determination of casing deflection. Comparison of this measurement with the restoration curve of the
gasket material under varying stresses will indicate whether the selection of the size, number, and location
of the bolting material, the stiffening ribbing on the casing, and the gasket material will assure a pump
casing that will remain tight under working pressure conditions.
Interstage pump leakage along the horizontal flange with accompanying annoyance, maintenance
expense, and pump outage were common characteristics of multistage pumps for many years. They were
previously accepted as an unavoidable consequence of man's attempt to generate hydraulic pressures in
excess of those provided by the free state of nature. Fortunately, designs have today improved to the
point that an axially split casing pump need not develop this leakage.

Fig. 3.4 Two-stage axially split casing volute pump.


For small capacities and pressures up to 17.5 bar gauge (250 psig).
SO Multistage Pump Casings

Fig. 3.5 Two-stage axially split casing volute pump.


For pressures up to 27.5 bar gauge (400 psig). Note the outline of integrally cast interstage passages.

Fig. 3.6 Five stage axially split casing volute pump.


For pressures up to 110 bar gauge (1,600 psig).
Multistage Pump Casings 51

Fig. 3.7 Ten stage axially split casing volute pump.


For pressures up to 175 bar gauge (2,500 psig).

AXIALLY SPLIT CASING DESIGN

The design and arrangement of stages within a casing is discussed in more detail later. Whatever the
arrangement, it is necessary to connect the successive stages of a multistage pump. In the low and
medium pressure and capacity range, these interstage passages are cast integrally with the casing (Figs.
3.4 and 3.5). Older designs for higher capacities and pressures used external crossovers (Fig. 3.6) to
keep casing dimensions small and avoid sudden changes in flow velocity or direction, hence higher
hydraulic losses. Improved understanding of casing flow, particularly in the crossover region between
stages, has yielded modern designs that realize equal or better efficiency without the added cost of
external crossovers (Fig. 3.7).

Interstage Construction
A multistage pump inherently has adjacent chambers at different pressures. These chambers must be
isolated from one another so that leakage from high to low pressure will occur only at the clearance
joints between stationary and rotating pump parts and will thus remain minimal. The isolating wall used
to separate two adjacent chambers of a multistage pump is called a stage-piece, a diaphragm, or an
interstage diaphragm. The stage-piece can be a single piece, or it may be fitted with a renewable stage-
piece bushing at the rotor section immediately inside the stage-piece.
The stage-pieces, which are usually solid, are assembled onto the rotor along with impellers, sleeves,
bearings, and similar components to make the pump's element. It is important to note the distinction
between rotor and element; the former is the assembly of all the rotating components, the latter is the
rotor plus the removable stationary parts. A typical element for an axially split pump is shown in Fig. 3.8.
52 Multistage Pump Casings

Fig. 3.8 Inner element of six-stage opposed-impeller axially split casing pump.
Stage-pieces are assembled on the rotor between the impellers.

Upper Half
Locking / OfCasing
Surface ....
"

Lower Half
Of Casing

Fig. 3.9 Locked tongue-and-groove joint in lower half of axially split casing.
Multistage Pump Casings 53

To prevent the stage-pieces from rotating, a locked tongue-and-groove joint is provided in the lower
half of the casing (Fig. 3.9). (This tongue-and-groove is clearly visible in Fig. 3.8.) Clamping the upper
casing half to the lower half securely holds the stage-piece and prevents rotation.
Proper seating of a solid stage-piece against an axially split casing has given designers considerable
trouble not only because it presents alignment problems of a three-way joint but also because the joint
must be tight and leakproof under a pressure differential without bolting the stage-piece directly to
the casing.
A small-diameter casing is an important pump design factor in combatting stage-piece leakage. It
helps ensure a seal fitting of the two casing halves adjacent to the stage-piece when the casing bolting
is pulled tight. The small diameter also helps eliminate stage-piece cocking, which leaves leakage
clearance in the upper casing half when it is pulled down tight. No matter how rigidly the stage-piece
may be seated in the lower casing half, there must be a sliding fit between the seat faces of the stage-
piece and the upper casing half so that the upper half may be pulled down. Also, each stage-piece must
be arranged so that pumping pressure differential will tend to seat it tightly against the casing (Fig. 3.10)
rather than open up the joint. Some designs incorporated an elastic seal ring in the stage-piece and
wearing ring (Fig. 3.11) to try to ensure pressure tightness at the face despite minor machining variations,
indifferent assembly, or thermal distortion. These worked as intended, but with better machining and
materials, the simpler arrangement shown in Fig. 3.10 has proved more cost effective. The use of
elastomer seals at axially split pump interstage joints is now limited to stage pieces with a radial seal
only or those with a very high pressure drop, where even minor leakage would quickly cause wire-
drawing of the casing.
Although simplification of interstage joints is highly desirable, it should not extend to the point where

Low Pressure
High Pressure

Seating --t-::n"''--t---f---t:--,tI''''V
Surface Stage - Piece

Fig. 3.10 Arrangement of stage-piece between pump pressure stages.


S4 Multistage Pump Casings

Neoprene
Seal Ring

Fig. 3.11 Arrangement of stage-piece elastic seal-ring.

joint perfonnance is impaired. To do so risks interstage leakage with a consequent reduction in pump
perfonnance and element service life. The heads per stage being developed in modem designs are a
product of the need for higher pressure and practical limitations on the number of stages. That these
higher heads can be developed reliably is in part the result of careful design and development of the
interstage joints.

General Considerations
Axially split casing pumps are used routinely for pressures up to 170 bar gauge (2,500 psig). Such
pressures introduce the extremely important subject of proper choice of material. High-pressure piping
systems, of which pumps fonn a part, are invariably made of steel because this material is stronger and
stiffer than iron, and is ductile, that is, it yields significantly before fracturing. Considerable piping strain
occurs in these systems, some of which is transmitted to the pump casing. The latter is essentially an
axially split barrel flanged at the split and fitted with two necks to serve as inlet and discharge openings.
Under the action of piping strain, these necks are the highest stressed regions of the casing and if not
made of a strong, ductile material they may fracture. Steel (or chrome steel; see materials in Chap. 17)
is therefore the safest material for pump casings whenever working pressures exceed 70 bar gauge (1,000
psig). This discussion points up a very important feature in suction and discharge flange design. Whereas
raised face flanges are perfectly satisfactory for steel casing pumps, their use is extremely dangerous
with cast-iron pumps. A lack of ductility in cast iron leads to flange breakage when the bolts are tightened
because the bending moment fulcrum is located inside the bolt circle. Therefore, be sure to avoid raised
face flanges on cast-iron casings, as well as the use of a raised-face flange pipe directly. against a flat-
face cast-iron flange.
Multistage Pump Casings 55

If a cast-iron pump casing is used, it remains possible to use a steel adapter piece in the shape of
either a straight pipe extension or an elbow. This adapter piece should have a flat flange face at the
pump discharge flange connection.
Most specifications are very conservative in specifying the type of discharge flange, but they frequently
overlook the most elementary safety requirements in describing the suction flange. Obviously, the latter
should be suitable for the pump casing hydrostatic test pressure. Therefore, if the hydrostatic test pressure
is over 35 bar gauge (500 psig), it would be improper to use an ANSI 150 flange.
Location of the pump casing support is not critical in smaller units operating under 17.5 bar gauge
(250 psig) and at moderate temperatures (see Fig. 3.4). Very little distortion is likely in this small unit,
whatever the support foot location. However, in larger units operating at higher pressures and perhaps
higher temperatures, it is most important to support the casing as close as possible to the horizontal
centerline and immediately below the bearings (Figs. 3.5-3.7).
Certain pump requirements are basic and normally accepted by all reputable pump manufacturers.
However, it is good practice to include them when preparing pump specifications.

1. The analysis and physical characteristics of the recommended casing material should be outlined.
2. The manufacturer should specify the casing hydrostatic test pressure. Usually this pressure is 1.5 times the
casing design pressure, which equals or exceeds the maximum allowable working pressure at 38°C (lOO°F).
Because the pressure increment between standard flanges is large in the high pressure range, the maximum
allowable working pressure of multistage pump casings is often less than that of its suction and discharge
flanges.
3. If the casing joint is axially split, it should be provided with dowels for accurate reassembly after opening.
4. The casing should be provided with a suitable vent at the suction of the first stage impeller.
5. Suitable valved openings should be provided for venting the high points of the casing and for draining.
6. Necessary pressure instrument taps should be provided at suction and discharge flanges.
7. If shaft seal leakage can accumulate in the brackets connecting the bearings to the casing, the bracket should
have a suitable drain connection and overflow holes should be provided to prevent flooding of a bearing if
a drain becomes blocked.

RADIALLY SPLIT DOUBLE-CASING DESIGN

The oldest form of radially split casing multistage pump is that commonly called the "ring-casing" or
the "doughnut" type. When it was originally found necessary to use more than one stage to generate
higher pressures, two or more single-stage units of the prevalent radially split casing type were bolted
together. Figure 3.12 shows this type of two-stage pump, about 1907 vintage. This is a far cry from a
modem radially split casing pump, but it clearly illustrates the origin of today's design principles.
In later radially split casing pump designs (Fig. 3.13), the individual stage sections and separate
suction and discharge heads were held together with large through-bolts. These pumps, still basically
an assembly of bolted-up sections, had serious dismantling and reassembly problems because suction
and discharge connections had to be broken each time the pump was opened. The double-casing pump
retained the advantages of the radially split casing design and solved the dismantling problem.
The development of the double-casing high-pressure pump has been most interesting. Like the evolution
of several related but different animal families, it has evolved into several types, similar in principle but
different in conception.
The basic principle consists in enclosing the working parts of a multistage centrifugal pump in an
inner casing and in building a second casing around this inner casing. The space between the two casings
is usually maintained at the discharge pressure of the last stage. Some newer designs, however, effectively
56 Multistage Pump Casings

Discharge

Suction

Fig.3.12 Two-stage radially split "ring section casing" pump (around 1907).

revert to single casing construction by sealing each stage directly to the outer casing. The objective of
this arrangement is to reduce the size of the outer casing. That it does, but at the expense of casing
material and manufacturing complexity since the casing must now have a sealing fit for each stage.
The inner casing design follows one of two basic principles: (1) axially split (Fig. 3.14) or (2) radially
split (Fig. 3.15). The first of the two is a simple evolutionary step from the usual axially split casing
for low-pressure pumps. This type of single casing will distort and breathe, with consequent leakage at
the axial split and between stages, whenever final discharge pressure exceeds certain design limits because
practical flange thickness and bolting strength maximums are not adequate. To help prevent the leakage
at the axial split, the casing is enclosed in a solid barrel of cast or forged steel. As the outside of the
axially split casing is subjected to a pressure greater than the average internal pressure, the inner casing
is under compression and the axial flanges will remain tight.
The inner casing is usually of volute design. The main shortcomings of this particular design are as
follows: (1) it lacks the symmetry of a radially split inner casing and (2) it eliminates possible leakage
at the axial joint without ensuring against interstage leakage at the stage pieces, or leakage at the three-
cornered joint where the axial split meets the main sealing face in the outer casing. The principal
advantage of this design is that the rotors of high-speed pumps, those running at more than 4,000 rpm,
can be dynamically balanced, then installed without subsequent dismantling and reassembly.
The double-casing pump with radially split inner casing is an evolution of the ring-casing pump, with
added provisions for ease of dismantling. After assembly, it is inserted inside a cast or forged cylindrical
casing and supported by one of two basic means (see later discussion) to remain aligned within the
pump yet be free to expand under temperature changes.
Multistage Pump Casings 57

Fig. 3.13 Multistage radially split "ring section casing" pump of the 1930s.

Radial joints between the stages offer the advantage of greater accuracy in the manufacture of the
hydraulic passages, and because all the joints are ring-type, lower risk of leakage between the stages or
across the element-to-casing sealing face. The one disadvantage is that after the rotor is balanced it must
be dismantled, then reassembled with the inner casing pieces to make up the complete inner assembly
or element. With the appropriate construction (see Ch. 7), the rotor's balance can be maintained as the
element is assembled.
The difference in construction and assembly sequence of axially and radially split inner casings can
lead to confused terminology. A rotor is defined as the assembly of all the rotating parts of a pump, and
an element as the assembly of the rotor plus the stationary internal parts. In a single casing axially split
pump, the element is the rotor plus the wearing rings, stage pieces, and interstage bushings (Fig. 3.8).
In a double casing pump with an axially split inner casing, the same assembly exists, but it is now only
a sub-assembly of the complete inner element which necessarily includes the inner casing.
58 Multistage Pump Casings

Byron Jackson
Double Case Pump

Fig. 3.14 Double-casing pump with axially split inner casing.


(Courtesy BW/IP International Inc.)

.... .. .
. -':i..-- .-
t
I I

',~-

Fig. 3.15 Double-casing pump with radially split inner casing.


Multistage Pump Casings 59

The assembled inner element is inserted into a cylindrical casing or "barrel", which is usually a
weldment of forgings. Support of the element within the casing and provision for its thermal expansion
are generally achieved by either: 1) a radial fit at the suction end of the casing and stay bolts through
the element to casing sealing face (Fig. 3.15), or 2) a radial fit at each end of the casing and a spring
or "compensator gasket" between the element and casing head to maintain contact at the casing sealing
face (Figs. 1.11 and 3.14). Of the two methods, the latter is less expensive, has proven reliable, and is
therefore used in most modem designs.
In Fig. 3.16 the inner element (radially split) of a double casing pump is being installed in the outer
casing. Fig. 3.17 shows the external appearance of this type of unit (see also Fig. 11.56). The suction
and discharge nozzles are an integral part of the outer casing, and the inner element can therefore be
withdrawn without disturbing piping connections. The most important advantages of this design are
symmetry of the outer casing and ring-type outer casing joints, both of which make for more reliable
sealing at high pressures and temperatures. Taking advantage of this, double casing pumps have been
built for pressures to 965 bar gauge (14,000 psig) and temperatures to 425°C (800°F).

CASING MAINTENANCE

Axially split single casings are maintained following the principles set out in Ch. 2.
In double casing pumps, the maintenance of axially split inner casings broadly follows the principles
for single casings, but does require special techniques to reduce the risk of leakage between the stages
and across the element-to-casing seal. Most axially split inner casings have opposed impeller rotors,
which means there is a second important seal, with a differential pressure equal to about half the pump's
pressure rise, between the inner casing and the casing head.
The maintenance of radially split inner casings typically involves checking the ring joint surfaces for
damage, checking that the locating fits between the pieces are correct, and inspecting waterway surfaces
for damage (erosion, cracks). Joint face damage and loss of fit between the pieces is repaired by welding,
stress relieving where necessary, and remachining to new dimensions. Any damage to the waterway
surfaces should first be investigated to determine and correct the cause. Once that is done, the damage
can usually be repaired by excavating to sound material, then welding, stress relieving where necessary,
and refinishing to the original surface contour.
Although the outer casings of double casing pumps usually do not require maintenance, this being
one of the objectives of the design, the following should be checked each time the pump is opened:

• Inner surfaces for corrosion (or casing wall thickness).


• Waterway overlays for substrate corrosion.
• Element-to-casing sealing face for mechanical damage and erosion or "wire drawing".
• Casing and head gasket faces for damage.
• Fillet adjacent to the element-to-casing sealing face for cracks.

Local corrosion can be repaired by welding and refinishing the surface to the original contour. Major
corrosion of the casing inner surfaces (beyond the design corrosion allowance) or substrate corrosion
under overlays is most efficiently done by machining out to sound material, building up with automatic
welding, stress relieving, then remachining to new dimensions. The build-up can include a corrosion
resistant overlay if required by the original design or deemed necessary from the actual corrosion rate
in service. Damage to sealing surfaces is repaired by welding, usually with austenitic stainless steel,
then remachining. Cracking in the casing is serious. Whether it can be repaired depends on the cause
60 Multistage Pump Casings

Fig. 3.16 Inner element of radially split double-casing pump being inserted into its outer casing.

Fig. 3.17 Multistage radially split double-casing pump.


Multistage Pump Casings 61

of the cracking and the extent of it. Extensive cracking caused by low cycle thennal fatigue should not
be repaired because there is a risk that all the material through the section has been damaged. For other
than minor repairs such as local corrosion or damage to sealing surfaces, the casing must be disconnected
from its piping and moved to a repair shop.

BIBLIOGRAPHY

[3.1] API-61O, 8th Edition, Centrifugal Pumps for Heavy-Duty Chemical, Gas Processing, and Refinery Service,
1995, American Petroleum Institute, Washington, D.C.
4
Impellers and Wearing Rings
--- - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - -

IMPELLER TYPES

The function of the impeller is to convert torque applied to the pump shaft to pressure and kinetic energy
in the pumped liquid. It does this by the action of its vanes on the liquid and is the only component of
a centrifugal pump that does so. The other components, even those that convert kinetic energy to pressure,
absorb some energy from the liquid. Given its unique function, the impeller can be considered the
fundamental component of a centrifugal pump.
Impellers are classified by the following four design distinctions:

1. Shaft mounting
2. Inlet or suction arrangement
3. Vane shape and form
4. Vane closure.

Before discussing classification in detail, it is beneficial to address impeller nomenclature. Referring to


Fig. 4.1, liquid enters the impeller through the "suction eye," the minimum area just ahead of the vanes,
passes through the waterways formed by the "vanes," "hub," and "shroud," and leaves at the "discharge
tip." Torque is transmitted from shaft to vanes through the "hub." Leakage of high-pressure liquid from
the impeller discharge back to suction is limited by a close-running clearance at the "outer hub" or
"wearing ring hub."
Shaft mounting, although nominally a mechanical feature, has a bearing on impeller hydraulics and
so is dealt with first. When the impeller is attached to one end of the shaft, such that the shaft does not
extend into the impeller eye, the impeller is termed overhung. For hydraulics, this is the ideal arrangement,
because the shaft does not block the impeller eye. The alternative is to have the shaft passing through
the impeller eye, an arrangement usually referred to as shaft through eye. Because the shaft now blocks
the impeller eye, the eye diameter must be larger than that of an overhung impeller with the same eye
area. The larger eye diameter increases the NPSHR for an otherwise equal design. A further possible

62

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
Impellers and Wearing Rings 63

Outer Hub or Center dividing wall of


Wearing Ring Hub Double Suction Impeller

Suction Eye

Suction Vane
edge or tip

Hub

Shroud

Fig. 4.1 Nomenclature of double suction impeller.

effect is lower pump efficiency as the ratio of eye diameter to impeller diameter increases, a consequence
of overloading the vanes (trying to add too much energy over too short a vane).
There are two inlet or suction arrangements: single suction and double suction (Fig. 4.1). As the
terms imply, a single-suction impeller has one eye; a double suction, two. Pumps with double-suction
impellers normally have the two casing suction passages joined to a common suction nozzle and flange.
Whether to use single- or double-suction impellers depends on a number of factors.
Small, radially split pumps, either horizontal or vertical shaft, almost invariably have single-suction
impellers, because it is an economical arrangement and the NPSHR is tolerable. Very large vertical-
shaft pumps, too, invariably have single-suction impellers, the choice in this case being a product of
structural considerations (vertical) and realizing the highest possible efficiency (overhung impeller).
In medium and large pumps, the use of double-suction impellers has two advantages. First, it allows
lower NPSHR than a single-suction impeller of equal capacity. Second, double-suction impellers are
theoretically in axial hydraulic balance, thus reducing the size of the necessary thrust bearing (but see
Chap. 22 for cautionary remarks). When the service conditions allow an axially split casing (see Chap.
2), the added casing complexity is fairly easily accommodated. For those services requiring a radially
split casing, the construction is more complex but often employed for both lower NPSHR and the
advantage of a between bearings rotor.
Multistage pumps usually have single-suction impellers to avoid the casing complexity associated
with getting liquid to and from a series of double-suction impellers. The only exception is the first stage
of sizes 3-in. discharge and larger, which is sometimes double suction to avoid having to provide
additional NPSH.
64 Impellers and Wearing Rings

Pumps designed to handle liquids laden with solids or abrasives incorporate special leakage restriction
arrangements to avoid binding and allow for adjustment (see wearing rings). These arrangements are
not suited to double-suction impellers, so such pumps also normally have single-suction impellers.
Vane shape and form are generally divided into four groups:

1. Plain vane (Figs. 4.2 and 4.9-4.11)


2. Francis vane (Figs. 4.3-4.5)
3. Mixed flow (Fig. 4.6)
4. Propeller or axial flow (Fig. 4.7).

In a plain vane impeller, the vanes are of single curvature, with all vane surfaces straight lines parallel
to the axis of rotation.
The vane surfaces in a Francis-vane impeller have a double curvature. This impeller is also often
called the Francis screw-vane or screw-vane impeller.
An impeller design that has both a radial and axial flow component is called a mixed-flow impeller.
It is generally restricted to single-suction designs with a specific speed above 4,200. Types with lower
specific speeds are called Francis-vane impellers (see Chap. 18). Mixed-flow impellers with a very small
radial flow component are usually referred to as "propellers." In a true propeller or axial-flow impeller,
the flow strictly parallels the axis of rotation. In other words, it moves only axially.
The relation of impeller profiles to specific speed for single-suction impellers is shown in Fig. 4.8.
Impeller classification according to vane shape is arbitrary, because impeller types overlap in various
pump types. For example, impellers in single- and double-suction pumps of low specific speed have
vanes extending across the suction eye. This arrangement provides a mixed flow pattern at the impeller
entrance for low pickup losses at high rotative speeds, but allows the discharge portion to follow the
plain vane principle. In pumps of higher specific speed operating against low heads, impellers have
double-curvature vanes extending over the full vane surface. They are therefore full Francis-type impellers.
The mixed-flow impeller, usually a single-suction type, is essentially one-half of a double-suction high-
specific-speed Francis-vane impeller.
Vane closure refers to the means used to "close" the waterways of the impeller. Some means of
closure is necessary lest the pressure difference across the impeller vanes (between the pressure and
suction sides, an inherent aspect of impeller action) cause leakage and a consequent drop in pump
performance. Three closure arrangements are generally recognized: (1) open, (2) semiopen, and (3) closed.
An open impeller (Figs. 4.6 and 4.9) essentially consists of vanes connected to a hub. Leakage across
the vane ends is limited by a close axial clearance at each side, formed by the casing and cover wall.
The design has two mechanical limitations. First, the vanes are weak in bending, and if at all long must
be stiffened with a rib or partial shroud. Second, either the casing or cover surface must be adjustable
or readily replaceable to allow economical restoration of running clearances.
With the appropriate vane design, open impellers are able to pump liquid containing significant solids
without suffering blockage of the waterways or binding between the impeller and casing. This capability
is a consequence of having a "cutting edge" at each end of the vane and not having a shroud, which
can accumulate material between itself and the adjacent casing walls. Open impellers are better suited
to stringy or bulky solids than abrasive materials, the latter being handled with longer service lives by
semi-open or closed impellers. Combined with the mechanical limitations, this restricts the normal use
of open impellers to small, general service solids-handling pumps.
In a semi-open impeller (Fig. 4.10), one side of the impeller waterway is closed by a shroud, usually
on the back side of the impeller, occasionally on the front side. With this refinement, the mechanical
limitations of the open impeller are largely overcome; vane bending strength is adequate and the impeller
Fig. 4.2 Plain-vane single-suction closed impeller. Fig. 4.3 Francis-vane double-suction closed impeller.

Fig. 4.4 High-specific-speed Francis-vane double- Fig. 4.5 Low-specific-speed Francis-vane single-
suction closed impeller. suction closed impeller.

Fig. 4.6 Open mixed-flow impeller. Fig. 4.7 Axial-flow impeller.

65
66 Impellers and Wearing Rings

I
0 0 0 0 0 0 o o o o o o o o 0 0 o
...
0 0
CD ,...
0 0
IJI
0
'"
0
~-
o
"'.
o
q
o
~
o
o
o
o.
o
q
o
~
o
0_
0
q
0
q
o
..,
0.
,- - ..... '"
¢-
'" '" CD '" 0

Radi al ' ~an, li,ld Fr anc:is - ler.. f ie ld MI.,d-fIO" f i' ld Al i a l - f1aoo l i,'d

Fig. 4.8 Variation in impeller profiles showing approximate range of specific speeds (US units).

Fig. 4.9 Open impellers.


Impellers at left and right strengthened by a partial shroud.

Fig. 4.10 Semi-open impeller.


Impellers and Wearing Rings 67

Fig. 4.11 Semi-open impeller.


Front view (left); back view (right) shows pump-out vanes.

to casing clearance can be maintained by an adjustable rotor. The shroud side normally has pump-out
vanes (Fig. 4.11) to (1) lower the pressure at the shaft seal and (2) keep the region between the shroud
and the casings free of any solids. When the shroud has balance holes in the region of the vane inlet
edge, the pump-out vanes also serve to minimize the leakage flow down the back shroud.
The compromise made by using a semi-open impeller is lower solids-handling capability than an
open impeller. For all but the most difficult services, the compromise is of little consequence, thus semi-
open impellers are widely used in pumps for chemical, paper, slurry, and severe industrial services. An
argument made frequently for their use in such services, aside from resistance to blockage, is that
clearance wear, hence pump performance, can be restored to a worthwhile degree without having to
open the pump.
A closed impeller (Figs. 4.2-4.5) has both sides of its waterways closed with a shroud, an arrangement
that eliminates leakage across the ends of the vanes. Of course high-pressure liquid still tends to leak
back to lower pressure, so some form of cylindrical or axial restriction is incorporated between the
casing and the shroud at a convenient location to minimize the leakage (see wearing rings).
For a given pump, except in very small sizes, a closed-impeller design is more economical to make
for high efficiency than either an open- or a semi-open-impeller design. The reasons for this are the
sensitivity of the latter two designs to clearance at the ends of their impeller vanes and the need for
clearance adjustment. Sensitivity to end clearance requires additional care in machining to achieve
consistently high performance. The need for clearance adjustment complicates pump construction. Closed
impellers are also sensitive to leakage, but not to the same degree as open and semi-open impellers (the
leakage is around the impeller rather than back into the succeeding channel), and the close clearances
used to limit leakage are easier to manufacture accurately. Given this advantage, closed impellers are
used almost exclusively for clean liquid services, the exception being small, low-cost, nonadjustable
pumps designed to be replaced when worn out.
68 Impellers and Wearing Rings

SPECIAL DESIGNS

Given the myriad services to which centrifugal pumps are applied, there are many impellers designed
for specific applications. Details of some of the more prevalent are given here.
Liquids containing rags, stringy material, and solids like sewage will quickly clog the inlet region of
conventional impellers. The preferred solution is a closed impeller with large waterways and thick,
rounded vane inlet edges (Fig. 4.12), a variation known as a nonclog impeller. For pumps up to 12- to
16-in. discharge, such impellers usually have only two vanes, larger sizes have three and four vanes.
With hydraulic design compromised to allow the passage of large solids, nonclog impellers are not quite
as efficient as conventional designs and are prone to noisy operation at capacities below design (see
Chap. 22).
Solids-laden liquids that cannot be handled by nonclog impellers are often pumped with what is
known as afree-flow or induced-vortex impeller (Fig. 4.13). In this arrangement, the impeller is recessed
into the back of the casing, out of the pumped flow to give a free-flow path. The impeller is a simple,
radial vaned, semiopen design. Head is produced in the pumped flow by momentum exchange. Low
momentum liquid enters the impeller near the hub, is accelerated through the impeller, then issues as
high momentum liquid into the pumped flow where it gives up energy to the flow. Pump efficiency is
lower th.an conventional designs, but the performance characteristics are similar.
Closed impellers for liquids containing abrasive solids follow the usual slurry pump practices of
simple vane shape, thick vanes and shrouds for tolerance of wear, and pump-out vanes on the back
shroud. To both restrict leakage down the front shroud and keep the region free of solids, the front
shroud also has pump-out vanes (Fig. 4.14).
For high-energy impellers (head above 200 m [600 ft] and power more than 225 kW [300 hpJ) or
when low-pressure pulsations and low noise are important, it becomes necessary to modify conventional
impeller designs. The essential objective is to reduce the pressure pulsation produced as the impeller
vanes pass the stationary volute tongue or diffuser vanes.
Although it is really a casing design consideration, a good start is to have a certain minimum clearance
between the impeller and collector vanes (Fig. 4.15). How much is still a matter of some controversy

Fig. 4.12 Phantom view of radial-vane Fig. 4.13 Induced vortex impeller.
non-clogging impeller.
Impellers and Wearing Rings 69

Fig. 4.14 Closed impeller for slurry pump (rough casting).

HIGH ENERGY PUMPS


DEFINITION
STAGE PERF GREATER COLLECTOR CLEARANCE
THAN: 200 M (650 FT) TYPE
& " R2
225 KW (300HP) DIFFUSER :3
VOLUTE 6

LEARANCE
% R2,,(R J - R,l 100
~
I R2
ql
1,1t
III,
II'I

:('1:,
I
I \
t , •

Fig. 4.15 Radial clearance between impeller and stationary vanes (gap "B").
70 Impellers and Wearing Rings

Fig. 4.16 Double suction impeller with split and staggered vanes.

because it is influenced by the head being developed, the impeller design, and the form of the stationary
vanes. There is, however, general agreement that values of diametral clearance equal to 1-2 percent of
impeller diameter are too small, and that increasing the clearance to 3-5 percent benefits both pump
efficiency and behavior (see Chap. 22 for further discussion).
Within the impeller itself, the objective is to minimize variations in the velocity of the liquid being
discharged. The most beneficial approach is to sacrifice efficiency and use a conservative design; one
with a relatively low discharge recirculation capacity (see Chap. 22). Further refinement involves such
measures as split and staggered vanes (Fig. 4.16), which is commonly used for double-suction impeller on
fan pump and similar service, and has also been used for single-suction impellers when circumstances
warranted it. The virtue claimed for split and staggered vanes is that the velocity variation (wake) associated
with each vane is halved, hence the pressure pulse caused by its passing is also halved. In very high energy
impellers (typically 750 m [2,500 feet] and 13,500 kW [18,000 hp)), split and staggered vanes have proven
necessary to ensure the bending strength of the vanes between the shrouds was sufficient.
An inducer (Fig. 4.17) is a special version of the propeller pump. Its function is to operate in series
with a conventional impeller and produce enough head to raise the net positive suction head available
(NPSHA) at the impeller to at least that required by the impeller. Being designed for only a very low
head, the inducer has inherently lower NPSHR and, to compound the benefit, is able to run with quite
a degree of cavitation without suffering damage or materially affecting the pump head. Correctly designed
and applied, an inducer enables pumping to higher heads with the same NPSHA or to the same heads
with lower NPSHA. In either case, the pumping installation is less expensive. There is need. however.
for emphasis on correct design and application. Inducers have a limited flow over which they operate
Impellers and Wearing Rings 71

Fig. 4.17 Inducer.

well. At higher flows, the head produced is not enough to provide the NPSH required by the impeller.
Recirculation (see Chap. 22) sets in at flows below the lower limit. Operation outside the "stable" flow
range therefore results in cavitation or noise and vibration or all three.

MANUFACTURE

In the past most impellers were cast. Usually the casting was one piece unless small size and low specific
speed dictated casting in two pieces and permanently assembling after machining. Modem practice has
fabrication and molding making notable inroads into impeller manufacture. Small, closed impellers for
mass-produced standard pumps are now routinely spot welded together from stainless steel pressings.
The justification for this approach is better performance (accuracy of shape and finish) and lower cost
provided the production runs are high enough. Using a combination of cast and wrought (pressed) pieces,
very large impellers for pumps and pump/turbines are also being fabricated. Adopting this approach has
72 Impellers and Wearing Rings

improved the accuracy of shape and lowered cost, while maintaining mechanical integrity equal to or
better than that of a one-piece casting. Casting is, of course, still an entirely valid means of making
impellers, but it is a skill that is becoming more concentrated, hence more expensive, and so is fostering
the development of alternatives.
Molding is the means used to produce plastic impellers. For small water pumps and severe service
chemical pumps, plastic is being used with increasing frequency. The need to separate the mold limits
one-piece molding to semi-open plain-vane impellers. Closed impellers can be made in two pieces and
bonded together, but the vane shape is still limited to plain.

IMPELLER TYPE SELECTION

From the preceding text it is evident there are many possible configurations of centrifugal pump impeller.
Figure 4.18 is intended as a broad guide to selecting or identifying the most appropriate basic impeller
type. It does not address, to any extent, special designs, and it obviously cannot be more than a broad
guide given the variations in service conditions and available product.

WEARING RINGS

Wearing rings provide an easy and economically renewable running clearance between the impeller and
casing. A running clearance without renewable parts is illustrated in Fig. 4.19. To restore original
clearances after wear, the operator must either (1) build up the worn surfaces by welding, metal spraying,
or other means, and then true up the part; or (2) buy new parts.
The new parts are not very costly in small pumps, especially if the stationary casing element is a
simple suction cover. As a matter of fact, the cost of a renewable stationary ring would sometimes differ
very little from that of a totally new suction cover for these units. This would not be true for larger
pumps of course, nor if the stationary element is part of a complicated casting. If the first cost of a
pump is of prime importance, the designer can benefit the operator by providing means for both stationary
parts and the impeller to be remachined. Renewable casing and impeller rings can then be installed
(Figs. 4.20 and 4.21). A preferable approach, now common practice for large pumps, is to furnish the
new pump with single-ring construction (Fig. 4.20). This simplifies pump construction, provides a readily
renewable wearing surface in the component of lesser material (casing typically iron or steel), and
recognizes the durability of the impeller wearing surface (typically 13 chrome or 13/4 chrome nickel steel).
Nomenclature for the casing or stationary part that forms the leakage joint surface varies as follows:
(1) "casing ring" (if mounted in the casing); (2) "suction cover ring" or "suction head ring" (if mounted
in a suction cover or head); and (3) "casing cover ring" or "head ring" (if mounted in the casing cover
or head). Some engineers like to identify the part further by prefixing the word "wearing" to the word
"ring," for example, "casing wearing ring." A renewable part for the impeller wearing surface is called the
"impeller ring." Pumps with both stationary and rotating rings are said to have "double-ring" construction.

WEARING RING TYPES

There are various types of wearing ring designs, and selection of the most desirable type depends on
the liquid being handled, the pressure differential across the running clearance, the rubbing speed, and
the particular pump design. In general, centrifugal pump designers use the ring construction they have
found most suitable for each particular pump service. The most common ring constructions are the flat
Impellers and Wearing Rings 73

CLEAN LIQUIDS

PUMP SIZE

SMALL
I
MEDIUM LARGE

I
'STAGES
I
'STAGES
I
SHAFT AXIS
I I I
I I I I
SINGLE MULTI SINGLE MULTI HORIZONTAL VERTICAL
I I I I I I
CD ® 0 ® 0 CD
0 ®

SOLIDS LADEN UQUIDS

ABRASIVENESS

NONE TO LOW MODERATE HIGH

I
PUMP SIZE
I
I
SMALL MEDIUM
I I
@ ® @ ®
® o ® o
o
Fig. 4.18 Impeller type selection chart.
KEY: 1. Overhung, single-suction, closed. 2. Shaft through eye, single suction, closed. 3. As 2 but with double-
suction first stage. 4. Shaft through eye, double-suction, closed. 5. Overhung, single-suction, open. 6. Overhung,
single-suction, semiopen. 7. Overhung, single-suction, special design for service.

type (Figs. 4.20 and 4.21) and the L type. The leakage path in the fonner is a straight annular clearance.
In the L-type ring illustrated in Fig. 4.22 the axial clearance between the impeller and casing ring is
large so that the velocity of the liquid flowing into the stream entering the suction eye of the impeller
is low. The L-type casing rings shown in Fig. 4.22 and 4.23 have the additional function of guiding the
liquid into the impeller eye; they are called "nozzle rings." Impeller rings of the L type shown in Fig.
4.23 also furnish protection for the face of the impeller wearing ring hub.
74 Impellers and Wearing Rings

IMPELLER

SUCTION
RING

SUCTION HEA

SUCTION HEAD

Fig. 4.19 Plain flat running clearance (no rings). Fig. 4.20 Single flat casing ring construction.

SUCTION
RING

SUCTION HEAD

Fig. 4.21 Double flat ring construction.


Impellers and Wearing Rings 7S

CASING ---....\" CASING CASING RING

CASING RING-__>/

IMPELLER

~~~LLER_+--_-++,

Fig.4.22 Casing ring with "L-nozzle". Fig.4.23 Double ring with "L-nozzle".

CASING--ooooOoooo\l/
CASING----i"
RELIEF
CHAMBER

CASING RING

IMPELLER RING -+----+-+'


r-t-"-"''"---.J-IMPELLER

Fig. 4.24 Single labyrinth of intenneshing type.


Double ring construction with nozzle-type casing ring. Fig. 4.25 Double labyrinth ring construction.

In the past, some designers favored labyrinth type rings (Figs. 4.24 and 4.25). The objective of these
designs was lower leakage for a given pressure drop. This was to be realized by two features: first, the
series of close clearances followed by expansions would increase the overall friction coefficient. Second,
the reentrant feature of the design increased the total length of close clearance. The objective was often
realized, but many pumps also suffered vibration problems, the cause of which was finally traced to
labyrinth type rings. By virtue of the liquid flow and pressure distribution within it, a leakage restriction
with both inside and outside close clearances is self-disturbing; that is, radial displacement of the rotor
develops a force causing even more radial displacement. As a result of this finding, labyrinth type rings
have been modified to have either inside or outside close clearances but not both. With this modification,
the single labyrinth (Fig. 4.24) offers little advantage. Double, or even more, labyrinth rings (Fig. 4.25)
are employed when space permits and the complexity is justified.
Wearing ring designs employing only intermediate expansions to raise the overall friction coefficient
are also used. One example is the stepped ring, which can be single step (Fig. 4.26) or multiple steps.
Another example is the serrated ring (Figure 4.27), in which a series of grooves serve as the intermediate
expansions. In addition to reducing leakage losses, serrating one of the surfaces of a radial running
clearance serves to significantly increase the clearance's resistance to galling in the event of internal
contact. For many applications, particularly those where materials with low galling resistance must be
76 Impellers and Wearing Rings

RELIEF
CHAMBER

Fig. 4.26 Single step-type running clearance


(double rings). Fig. 4.27 Serrated casing wearing ring.

used, the anti-galling characteristic of a serrated ring is more important than any improvement in pump
performance, the logic being that a pump not running has zero performance.
Wearing ring grooves or serrations can be either normal to the axis or helical. There are arguments
advanced that helical grooving which acts to pump against the leakage flow, improves pump efficiency
by reducing the leakage. Considering that such a pumping arrangement is relatively inefficient, any
reduction in leakage would be offset by an increase in power. This leads to the suggestion that it is
better to allow the leakage to be repumped by the more efficient impeller. In those cases where reverse
pumping wearing rings have yielded higher efficiency, the change in leakage rate must have been enough
to affect impeller performance. Helical grooving does seem to offer better resistance to damage from
incidental contact between the surfaces. This is attributed to the helical groove "wiping" across the
adjacent surface rather than "scoring" into it.
For raw water pumps in water works service and large pumps in sewage service in which the liquid
contains sand and grit, water-flushed rings have become popular (Fig. 4.28). Clear water under a pressure
greater than that on the discharge side of the rings is piped to the inlet and distributed by the cored

~ ___________ IMPELLER

,.--_ _ _ _ IMPELLER RING

,..-------- SUCTION HEAD RING

z----SUCTION HEAD

~---CI_E~,R WATER INLET

Fig.4.28 Water-flushed wearing ring.


Impellers and Wearing Rings 77

passage, the holes through the stationary ring, and the groove to the clearance. Ideally, the clear water
should fill the clearance with some flow both to the suction and discharge sides to prevent any sand or
grit from getting into the clearance space. To realize this, the pressure drop across the downstream
portion of the ring considering the flush flow alone, must be greater than the difference between the
static pressure at the pump suction and that just upstream of the running clearance. In all but very low
head pumps, this requires a flush flow higher than is usually practicable, which is why flushed wearing
rings have not always performed as expected. Similar difficulties have been encountered in the petroleum
industry, and are often remedied by adding pump-out vanes to the impeller shrouds to lower the pressure
adjacent to the joint. That solution has the disadvantage of noticeably increasing the pump's power, a
consequence of the power absorbed by the pump-out vanes. The trend today is to avoid the complexity
of flushing and employ wearing materials or coatings able to yield acceptable service lives running in
the pumped liquid.
In large pumps (roughly 36-in. discharge, or larger), particularly vertical end-suction single-stage
volute pumps, mere size alone permits some refinements not found in smaller pumps. One example is
the inclusion of inspection ports for measuring ring clearance (Fig. 4.29). These ports can be used to
check the impeller centering after the original installation as well as to observe ring wear without
dismantling the pump.
The lower rings of large vertical pumps handling liquids containing sand and grit on intermittent
service are highly susceptible to wear. During shutdown periods, the grit and sand settle out and naturally
accumulate in the region in which these rings are installed, as it is the lowest point on the discharge
side of the pump. When the pump is started again, this foreign matter is washed into the clearance and
causes wear. To prevent this action in medium and large pumps, a dam-type ring is often used (Fig.
4.30), for it permits the pocket on the discharge side of the dam to be periodically flushed.
One trouble with the simple water-flushed ring previously described (Fig. 4.28) is its failure to provide
uniform pressure in the stationary ring groove. If the pump size and design permit, two sets of regular
flat wearing rings in tandem and separated by a large water space (Fig. 4.31) provide the best solution.
The large water space allows uniform distribution of the flushing water to the full 360 deg of each
running clearance.
For pumps handling gritty or sandy water, the ring construction should provide an apron on which
the stream leaving the clearance can impinge, as sand or grit will erode any surface it hits. Thus, a form
of L-type casing ring similar to that shown in Fig. 4.28 should be used.

STUFFING BOX HEAD

STUFFING BOX IMPELLER


HEAD RING RING

IMPELLER

Fig.4.29 Wearing ring design with inspection hole.


78 Impellers and Wearing Rings

IMPELLER----------~

I't-.------SUCTION HEAD

Fig. 4.30 Dam-type ring construction.

IMPELLER
IMPELLER RING NO.2
SUCTION HEAD RING NO.2

FULL CIRCLE
CORED PASSAGE

CLEAR WATER
INLET

HEAD
SUCTION HEAD RING NO. I

Fig. 4.31 Two sets of rings with space for flushing water.

WEARING RING LOCATION

In some designs used by one or two sewage pump manufacturers, leakage is controlled by an axial
clearance (Fig. 4.32). Usually this design requires a means of adjustment of the shaft position for proper
clearance. Then, if uniform wear occurs over the two surfaces, the original clearance can be restored
Impellers and Wearing Rings 79

IMPELLER ----~

Fig. 4.32 Axial running clearance. Fig. 4.33 Adjustable axial-clearance


ring construction.

simply by adjusting the impeller position. This method does have its limitations, however, for the impeller
must be nearly central in the casing waterways.
Axial running clearances are not overly popular for double-suction pumps because a very close
tolerance is required in machining the ring fit in reference to the centerline of the volute waterways.
Radial clearances, however, allow some shifting of the impeller for centering. The only adverse effect
is a slight inequality in the lengths of the leakage paths on the two impeller sides. An ingenious ring
construction that overcomes this objection is illustrated in Fig. 4.33. However, this design is more
expensive than the conventional radial-clearance ring. Also, if it must be adjustable after wear, the
threaded parts must be corrosion proof for the liquid being handled.
So far, this discussion has treated only those running clearances located adjacent to the impeller eye
or at the smallest outside shroud diameter. Sometimes, however, the running clearance is located at the
impeller periphery (Fig. 4.34). In a vertical pump, this design is advantageous because the space between
the joint and the suction waterways is open and sand or grit cannot accumulate. But the design is
impractical in regular pump lines because of increased area and rubbing speed and because the impeller
diameters used in the same casing vary over a wide range.
There are, however, advantages to having some form of leakage restriction at or near the impeller
outside diameter (OD) in addition to the usual hub running clearance (Fig. 4.35). As used today, the
restriction is typically two-three times the clearance of the wearing ring and is not made up of renewable
surfaces. In solids-handling services such as sewage pumping, the restriction serves to minimize accumula-
tion of solids in the region between the shroud and casing or cover wall. A second advantage is partial
isolation of the shroud region from pressure pulsations occurring at the impeller discharge during off
design operation (see Chap. 22). In high-head pumps this can be an important factor in realizing reliable
80 Impellers and Wearing Rings

IMPELLER

CASING

Fig.4.34 Running clearance at periphery of impeller.

operation over wide flow ranges. A disadvantage of the arrangement is a reduction in efficiency caused
by not being able to recover part of the energy put into disk friction as useful head.

STATIONARY WEARING RING MOUNTING

In small single-suction pumps with suction heads, a stationary wearing ring is usually pressed into a
headbore and can be locked securely by several set screws located half in the head and half in the ring
(see Fig. 4.21). Larger pumps often use an L-type ring with the flange held against a face on the head
(see Fig. 4.36). Single-suction radially split pumps designed for rings on the stuffing box side usually
have either a flat or L-type ring located in a shouldered bore in the cover (Fig. 4.29). Ring retention
can be by a shrink fit, an interference fit with pins, or a slide fit with machine screws, from either the
liquid or atmospheric side (Fig. 4.36). In the interests of reliability, it is desirable to minimize the use
of fasteners for wearing rings. Internal pins and machine screws can and do work loose. External screws
are easier to secure, but are a potential source of leakage unless sealed carefully. For service temperatures
up to 200°C (400°F), a simple shrink fit has proven reliable. Higher service temperatures or the need
for easier removal have been accommodated by using several short welds to retain interference fit rings.
In axially split casing pumps, the cylindrical casing bore (in which the casing ring will be mounted)
should be slightly larger than the outside diameter of the ring. Unless some clearance is provided, ring
distortion may occur when the two casing halves are assembled. However, the joint between the ring
and the casing must be tight enough to prevent leakage. This is usually accomplished with a radial
metal-to-metal joint (like the one marked "J" in Fig. 4.23) so arranged that the discharge pressure will
press the ring against the casing surface.
Impellers and Wearing Rings 81

r "A" 2-3 TIMES "CLA"

CLR

+
t

Fig. 4.35 Shroud space isolation (gap "A").

Fig. 4.36 Double ring construction with machine screw fasteners.


Screws fasten wearing rings to casing and impeller.
82 Impellers and Wearing Rings

If the pressure differential is very high, leakage may occur between the casing and the ring, eroding
the casing. A sealed casing ring (Fig. 4.37) may be required, especially on multistage pumps with high
stage pressures.
As it is not desirable for the casing ring of an axially split design to be pinched by the casing, the
ring will not be held tightly enough to prevent its rotation (due to the impeller torque transmitted through
the liquid in the clearance space) unless special provisions are made to keep it in place. One common
way of preventing rotation is to place a pin in the casing that extends into a clearance hole in the ring
(Fig. 4.38). The same basic arrangement can be reversed, placing the pin in the ring, with equal
effectiveness. Pin-type anti-rotation devices are suitable for low-cost low-energy stages. They are not
really adequate for high-energy stages, often suffering pin or hole failure, and should not be used in
vertical-shaft axially split pumps, the single shoulder making for difficulty assembly. An alternative,
and very reliable, antirotation arrangement is to have a tongue on the casing ring, extending around 180
deg and engaging a corresponding groove in the one half of the casing (Fig. 4.39). This method can be
used with casing rings having a central flange (see Fig. 4.33) by making the flange of larger diameter
for 180 deg and cutting a deeper groove in that half of the casing.
Many methods are used to retain impeller rings. A shrink fit is the simplest, and has proven reliable
for temperatures to 150°C (300°F). At higher temperatures, there is a risk the ring may loosen and turn
should the pump be started from cold with hot liquid. In these circumstances, a legitimate question is
whether an impeller ring should be used at all. Many believe they should. Usual practice is to mount
the ring with an interference fit, then secure it with threaded axial pins (Fig. 4.21), machine screws (Fig.
4.36), or a number of short welds. The first two methods involve internal fasteners. Great care must be
taken to ensure the fasteners are locked, l~st they back out and cause premature failure (see comment
on casing rings also). Threaded axial pins can only be used when the impeller and ring materials are of
similar alloy and hardness. If the ring is appreciably harder than the impeller, a common requirement,
the tap drill will drift off center into the softer impeller. Short welds must be made with care to

._--IMPELLER

CASING RING

-----SEAL RING

ASING

Fig. 4.37 Sealed casing ring.


Impellers and Wearing Rings 83

UPPER HALF
CASING

LOWER HALF
~.-J.-~ CASING
CASING
Fig. 4.38 "Pin-in-casing hole-in-ring" locking device. Fig. 4.39 Tongue-and-groove casing ring design.

avoid cracking in the weld or adjacent material. Provided this is done, it is a simple and reliable ring
retention method.
Impeller rings of brittle materials such as cast iron, Ni-resist and fully hardened 13 chrome steel,
should not be mounted in tension. One solution in older designs was to thread the hub and ring and
screw the parts together but is rarely used today because it is deemed too complicated. When a hard
surface is required on the impeller, it is really better to avoid a ring altogether and apply a hard coating
directly to the hub. Hard coating offers simpler construction and better use of resources by applying
exotic materials only where necessary. This technology has been proven in other forms of turbomachinery,
and although its adoption into centrifugal pump practice has been slow, its use will increase.
In impeller ring design, ring stretch due to centrifugal force must be considered, particularly if the
pump is a high-speed unit. For example, some boiler feed pumps operate at speeds that would loosen
the rings if only a press fit were used. For such pumps shrink fits should be used or, preferably, the
impeller rings eliminated.

IMPELLER MAINTENANCE

An impeller removed from a pump casing should be carefully examined on all surfaces for unusual
wear, such as from abrasion, corrosion, or cavitation. Most pumps for general service use bronze impellers,
which have a reasonably long life. Occasionally, these pumps operate on high suction lifts or at part
capacities, both of which affect impeller life. Manufacturers can suggest a more suitable impeller design
for such conditions.
Pumps handling water containing low concentrations of sand or silt may use bronze, cast-iron, nickel-
cast iron, or even chrome-steel impellers, depending upon the amount of sand, its abrasiveness, and the
character of the water.
Generally, impeller materials that form a protective coating or film, which adheres firmly to the
underlying metals and is not washed off by the water stream, should always be used. However, abrasive
material naturally erodes this protective film on many metals, making their use undesirable.
84 Impellers and Wearing Rings

Abrasion wear can be best tested by a sedimentation test. Some of the pumped liquid is allowed to
stand in a glass container for a few hours, and the settled particles are examined for grit. A chemical
laboratory analysis of the pumped liquid is usually necessary to determine whether corrosion is responsible
for undue wear. Of course, if corrosion wear is detected, the substitution of better materials becomes
necessary.
Cavitation is often accompanied by pitting in the impeller suction areas and can be detected by a
crackling noise during operation. If impellers rapidly become pitted or eroded, check the NPSHA then
the range of flows over which the pump is being operated (see Chap. 22). Unless the original impeller
was cast iron, changing to a better material usually is not sufficient to correct rapid cavitation erosion.
In small pumps, impeller wear is best corrected by replacement of the impeller, because the pump
size does not permit its being rebuilt. Whereas rebuilding by brazing, soldering, welding, and the like,
is feasible, the cost is high, and so replacement is usually the better solution.
Most large impellers will provide many years of service, regardless of abrasion, if eroded areas are
treated by "building up" the metal. Although unlikely, wear may sometimes occur in the impeller hub
over the shaft mounting or at the keyway. The first may be caused by a porosity in the impeller casting,
permitting water to seep from the higher pressure region to the fit between the shaft and impeller.
Sometimes, the shaft material is the one more readily attacked. Wear at the keyway may occur if the
impeller fits loosely on the shaft or the key is not properly fitted.
Finally impeller cracks may develop, the usual causes being vibration caused by pressure pulsations
associated with internal recirculation or vane passing (see Chap. 22) or subsurface defects not detected
during manufacture. Whether to repair or replace cracked impellers depends on their size, the material,
and the location and extent of the cracking. If repair is elected, the method must respect that an impeller
is a dynamic part, subject to cyclic stresses, and prone to catastrophic failure if a critical crack develops.
Impeller balance should be rechecked whenever the impeller is removed from the pump rotor during
overhaul. Whether to balance statically or dynamically depends on the impeller's width-to-diameter ratio
and the pump's rotative speed. One widely used guide is to dynamically balance whenever the diameter
to width ratio is less than 6 or the rotative speed above 3,600 rpm. If there is any doubt, consult the
manufacturer. When dynamic balancing is required, note that an impeller in dynamic balance is necessarily
in static balance, but the reverse is not the case. In fact, an impeller in static balance can easily be so
badly out of dynamic balance that correction is impossible. The significance of this is that impellers
requiring dynamic balancing should not be balanced statically first. Metal removal for balancing must
be done in a manner that will not affect the pump's hydraulic performance or mech;prical reliability.
This means not removing metal from impeller waterways unless there is an obvious variation in thickness
or profile, and not drilling or milling. Both techniques produce sharp corners, which can lead to erosion
or initiate cracking.
For balancing a shrouded impeller, the best practice is to mount the impeller off-center in a lathe and
take a cut (which will be deepest at the periphery) from the shroud (Fig. 4.40). The cut can be taken
from both shrouds, depending on their actual thicknesses and the amount of metal to be removed.
In semiopen impeller pumps, the removed metal can be taken from the shroud if the design permits
or from underneath the vanes if those on the heavy side are thicker than the others. The latter method
is the one used for balancing open impellers.

WEARING RING MAINTENANCE

Installation
Most rings are now pressed on the impeller. As distortion may occur during the mounting process,
it is advisable to check the shaft and impeller assembly on centers to see if the new ring surfaces are
Impellers and Wearing Rings 8S

Fig. 4.40 Metal removal for balancing impeller.

true and, if not, to true them up. If the proper facilities are available, it would be just as easy to get
slightly oversize rings and turn their wearing surface to the proper diameter after mounting.

Clearance
One manufacturer's clearance and tolerance standards for nongalling wearing joint metals in general
service pumps are shown in Fig. 4.41. They apply to the following combinations: (1) bronze with a
dissimilar bronze, (2) cast iron with bronze, (3) steel with bronze, (4) monel metal with bronze, and (5)
cast iron with cast iron. If the metals gall easily (like the chrome steels), the values given should be
increased by about 0.050 mm (0.002 in.). In mUltistage pumps, the basic diameter clearance should be
increased by 0.075 mm (0.003 in.) for larger rings. The tolerance indicated is "plus" (+) for the casing
ring and "minus" (-) for the impeller hub or impeller ring.
In a single-stage pump with a running clearance of nongalling components, for example, the correct
machining dimension for a casing ring diameter of 225.000 mm (9.000 in.) would be 225.000 plus 0.075
and minus 0.000 (9.000 plus 0.003 and minus 0.000 in.) and for the impeller hub or ring, 225.000 minus
0.450, or 224.550 plus 0.000 and minus 0.750 mm (9.000 minus 0.018, or 8.982 plus 0.000 and minus
0.003 in.). Actual diametral clearances would be between 0.450 and 0.600 mm (0.018 and 0.024 in.).
Naturally, the manufacturer's recommendation for ring clearance and tolerance should be followed.
A widely used industry standard for we~ng ring clearances is given in American Petroleum Institute
(API) Standard 610. These clearances are intended for pumping temperatures up to 260°C (500°F), so
are somewhat larger than those in Fig. 4.38 to allow for thermal distortion. The standard provides for
pumping temperatures above 260°C (500°F) and the use of materials with low galling resistance by
requiring an increment over the standard clearance. There is also provision for closer than standard
clearances in multistage pump balancing devices and similar critical components.

Allowable Wear
It is difficult to generalize on the amount of wear allowable before a pump should be dismantled and
the running clearances renewed, because too many factors are involved. Internal leakage through the
rings naturally means an efficiency loss. Ring renewal should be such that the overhaul cost will be
offset by the power savings. Thus, with constant use and high power costs, more frequent renewal can
be justified. The power lost to leakage decreases with increasing specific speed. At the same time it is
higher in multistage pumps of a given specific speed because of the need to balance axial thrust (see
86 Impellers and Wearing Rings

DIAMETER - MM
50 70 100 150 200 300 400 SOO 700 1,000 1,500 2,000
I J L L I _J
0.040 I 1.00

,
~
~z ::i!
et -~ 0.030 ui
0.75 (,)
UJUJ I
UU ! Z
ZZ
etet
a:: a::
•• <C
a:
w
et UJ CLEARANCE
g
..J
~
d
UJ..J


....
~ 0.020 O.SO
, O!I
..J(!) ...,j
W
et Z (,)
a::z
J
Z
1-- <C
~l5 0.010
a:
0.25 <C

.+-
-~
etet - . w
o:t -f- -- TOLERANCE s,_
~--
f-'-- _. ..J
(,)

! I
,-
-t'
3 4 5 6 8 10 15 20 30 40 50 60 80
RUNNING CLEARANCE DIAMETER, IN.
CASING RING

DIAMETRAL /
CLEARANCE?
0--/ I \
RADIAL
CLEARANCE'
I

Fig.4.41 Wearing-ring clearances for single-stage pumps using nongalling materials.

Chap. 5). A general rule of thumb is to renew the running clearances in single stage pumps when they
have increased 100 percent (leakage about 3 times higher), and in multistage pumps when they have
increased 50 percent (leakage about 2 times higher).
Even though the clearance is not excessive and the pump can be reassembled without renewing the
components, always check the impeller hub diameter and the inside diameter of the stationary wearing
ring for eccentricity of wear.

Measurement of Clearances
Wearing ring clearances may sometimes be measured by inserting a feeler gage between the stationary
and rotating parts. If the wearing ring is L type and the lip of the L prevents inserting the gage, the
clearance may be approximately checked without dismantling the element in the following manner:

1. Mount a dial indicator on the impeller (Fig. 4.42), and with the stationary ring resting on the impeller
wearing-ring hub, set the dial reading to zero.
Impellers and Wearing Rings 87

I"'QICA'TOR SUPPORT MO\.lNTED ON IMPEI.I.ER

Fig. 4.42 Measuring wearing ring clearances in a multistage pump with a dial indicator.

2. Without moving the impeller or dial indicator, push up on the stationary ring from below and record the
maximum dial reading. This corresponds to the diametral clearance.
3. Repeat this operation for every clearance joint and make a record of all readings.

This operation is best carried out, however, with the element removed from the pump casing. It is
best suited to multistage pumps because once the element is out of the casing of single-stage pumps,
the stationary rings may be freely removed and the clearance determined by measuring the two diameters
and calculating the difference.
One note of warning: This short-cut method gives no clue to the condition of adjacent clearance
surfaces. In other words, burrs, grooves, or indentations caused by foreign matter passing through the
clearances will go undetected, as will the resultant damage to the surfaces.
If the pump has been dismantled, normal procedure is to measure independently the inside diameter
(ID) of the wearing ring fit and the OD of the impeller wearing ring hub. Use inside and outside
micrometers, respectively (Figs. 4.43 and 4.44) Several measurements will determine whether or not the
wearing ring or impeller has become worn in an egg-shaped manner. The clearance is considered to be
the maximum difference between the maximum ID and the minimum OD readings.
Clearances may also be measured directly by placing the impeller within the wearing ring (Fig. 4.45)
and moving it laterally against a dial indicator to determine total diametral clearance. To determine
inequality in wear around the circumference, the impeller should be rotated and the dial indicator attached
to several points of the stationary part. If the pump has been dismantled, however, the "difference"
method is more reliable.
The impeller and wearing rings should be the same temperature before measurements are made. Some
high-pressure and high-temperature pumps use shrunk-on impellers that must be heated before removal
from the shaft to at least 200°C (400°F) and possibly to as much as 260 to 320°C (500 to 600°F). These
should be allowed to cool down to about 50°C (l20°F) so that measurements can be made comfortably.
But if the wearing ring is at 28°C (80°F), say, there will be a 22°C (40°F) difference between the two
parts and this difference can be quite significant. If the coefficient of thermal expansion is taken as 11.7
X 10-6 mm per mm per degree Celsius (6.5 x 10-6 in per in per degree Fahrenheit), and if the wearing
ring fit diameter is 200 mm (8 in.), the apparent clearance will be about 0.05 mm (0.002 in.) less than
88 Impellers and Wearing Rings

Fig. 4.43 Measuring wearing ring bore with an inside micrometer.

Fig. 4.44 Measuring OD of impeller hub with outside micrometer.


Impellers and Wearing Rings 89

Fig. 4.45 Measuring wearing ring clearance with dial indicator.

the true clearance. This error will, of course, be magnified if the impeller diameter is measured when
its temperature is even higher than the 50°C (120 0 P) we have assumed.
This possibility of error is frequently overlooked, as many people assume that such a small difference
in metal temperatures is not of consequence.

Restoring Clearances When No Rings are Used


To restore the clearance between impeller and casing when no ring is provided, the operator must
(1) buy new parts, (2) build up the worn surfaces by welding, metal spraying, or other means, or (3)
install a wearing ring or rings if sufficient metal is available in the casing part or on the impeller hub.

Restoring Clearances of Pumps with Single Rings


There are three ways to restore the clearance of a pump with single flat or L-type wearing ring con-
struction:

1. Obtain a new casing ring bored undersize from the manufacturer. Then, true up the impeller wearing ring
hub by turning down in a lathe.
2. Build up the worn surface of the wearing ring by welding or metal spraying so that it can be bored undersize.
Then, true up the impeller wearing ring hub.
3. True up the wearing ring by boring oversize, build up the impeller wearing ring hub, and machine to give
correct clearance with the rebored ring.

Of these, the first is the best approach. With impeller hubs sized for say 3 undersizes, the time between
having to restore the impeller hubs to original size can be 20 years or more. Once the impeller hub has
90 Impellers and Wearing Rings

been worn below its minimum undersize, it is built up by welding and remachined to its standard
maximum size. When the impeller is a material that cannot be welded, such as cast iron, double-ring
construction is preferred for all but small pumps.

Restoring Clearances of Pumps with Double Rings


If the pump has double flat or L-type wearing rings, clearances may be renewed, by one of the
following methods:

1. Obtain a new oversize impeller ring and use the old casing ring bored out larger.
2. Obtain a new casing ring bored undersize and use the old impeller ring turned down.
3. Renew both rings if necessary.
4. Build up either the casing or impeller ring by welding or metal spraying and machine the other part. By
altering the ring buildup, the original leakage joint diameter can be closely maintained.

The relative merits of various methods for renewing clearances in single-ring pumps mentioned earlier
in this chapter also apply to double-ring construction.
For rings other than the flat type, the manufacturer's recommendations for maintenance should be
followed. In axial running clearances (see Fig. 4.32) the stationary part, the impeller, or both may be
altered to accommodate rings. The complicated labyrinth rings (see Figs. 4.24 and 4.25) do not generally
permit either the building up of worn surfaces or the remachining of parts. Replacement of both labyrinth
rings is thus advisable.
5
Axial Thrust in Single-
and Multistage Pumps

AXIAL THRUST IN SINGLE-STAGE PUMPS

The pressures generated by a centrifugal pump exert forces on both its stationary and rotating parts.
The design of these parts balances some of these forces, but separate means may be required to
counterbalance others.
Axial hydraulic thrust is the summation of unbalanced impeller forces acting in the axial direction.
As reliable large-capacity thrust bearings are now readily available, axial thrust in single-stage pumps
remains a problem only in larger units.
Theoretically, a double-suction impeller is in hydraulic axial balance with the pressures on one side
equal to and counterbalancing the pressures on the other (Fig. 5.1). In practice, this balance may not be
achieved for the following reasons:

1. The suction passages to the two suction eyes may not provide equal or uniform flows to the two sides.
2. External conditions, such as an elbow being too close to the pump suction nozzle, may cause unequal flows
to the suction eyes.
3. The two sides of the discharge casing may not be symmetrical, or the impeller may be located off-center.
These conditions will alter the flow characteristics between the impeller shrouds and casing, causing unequal
pressures on the shrouds.
4. Unequal leakage through the two leakage joints will tend to upset the balance.
5. The pump is operating with discharge recirculation, which causes random, unequal fluctuations in the effective
pressure acting on each impeller shroud (see chapter 22).

Combined, these factors create definite axial unbalance. To compensate for this, all centrifugal pumps,
even those with double-suction impellers, incorporate thrust bearings.
The ordinary single-suction radial-flow impeller with the shaft passing through the impeller eye (Fig.
5'!), is subject to axial thrust because a portion of the front wall is exposed to suction pressure, thus
exposing relatively more backwall surface to discharge pressure. If the net pressure generated by the
impeller were equal to the stage pressure rise, and the pressure at the impeller discharge were uniform

91

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
92 Axial Thrust in Single- and Multistage Pumps

1II -------, r - - - - - - III


"0:----_
IUW 11l nr.,---,~
"ex It
::J -
_-_
--_-_
- ___I
-I / 1 t - - - - - - ct
4:(/')
1:(/')--------<001 1~-------I / J t - - - - - - :J
At------ ~
UIll _ _ _ ____1
(/')0: l6-a:1L1 - - - - - - - I /11------1L1
li 0. -------I 00..-------4 J I - - - - - ct
a.
~II;:~~----------- ~
A""-----((
t------<
t------0
.1......_----1/)
", ..._ _ _ _ _ 0

-~----. .-~~-
DOUBLE. SUCTION SINGLE SUCTION
IMPELLER IMPELLER

Fig. 5.1 Origin of pressure acting on impeller shrouds to produce axial thrust.

over all the surfaces exposed to impeller discharge pressure, the axial force acting toward the suction
would be equal to the product of the net pressure generated by the impeller and the unbalanced annular area.
In reality this is not the case, for two reasons. First, the net pressure generated by the impeller is
lower than the stage pressure rise by the pressure recovered in diffusion (see Chap. 2). Typically this
pressure recovery is equivalent to 20-25 percent of stage head at BEP, and about 10 percent at shutoff.
Second, the pressure on the single-suction impeller shrouds is not uniform. The liquid between the
impeller shrouds and casing walls is in rotation, and the pressure at the impeller periphery is normally
appreciably higher than at the impeller hub. Although we need not be concerned with the theoretical
calculations for this pressure variation, Fig. 5.2 describes it qualitatively. Generally speaking, axial thrust
toward the impeller suction is about 20 to 30 percent less than the product of the net pressure and the
unbalanced area. Beyond pressure distribution over the impeller shrouds, two further factors can affect
the axial thrust in centrifugal pumps. The first is momentum. In any design where the flow changes
direction as it passes through the impeller, that is all designs other than axial flow, the change in direction
produces a momentum force. This force acts away from the suction (see Fig. 5.3), and its magnitude,
Fm , is estimated from

where m=mass flow rate through impeller


em! = velocity at impeller inlet
cp = angle through which flow is turned.

Momentum is often neglected, with little ill effect, in small low-flow pumps. In larger pumps, particularly
those developing relatively low net pressure, momentum can be the major component of axial thrust
and therefore cannot be neglected.
The second factor is the axial component of the vane force exerted in accelerating the liquid into the
impeller vane cascade. This force acts toward the suction, and exists only in those impellers whose inlet
vanes reach into the impeller eye. Acceleration thrust is well understood in mixed- and axial-flow
impellers (discussed later), but it can also be of consequence in medium specific speed radial flow
impellers. Its magnitude is a function of specific speed (impeller shape; Fig. 4.8), then the detail design
of the particular impeller. At the current state of the "art", the magnitude of acceleration thrust can only
Axial Thrust in Single- and Multistage Pumps 93

Pr'SWry Dcf,,., on
,1ft/WII,r shroud.
/
!
/

~------ / _ _~
------

- --
'---- -

Fig. 5.2 Actual pressure distribution on front and back shrouds of single-suction impeller with shaft through
impeller eye.

Fm

Fig. 5.3 Axial thrust produced by change in momentum of flow through impeller.
94 Axial Thrust in Single- and Multistage Pumps

Dischar~ prrssvre
/
Fronr wrorlng
ring .........
"
Balancing ......
hoI.
"
........

Sud/on pressure I I
Svcfion prrssv~

~fr-- t --f---
Fig. 5.4 Balancing axial thrust of single-suction impeller with wearing ring on the back and balancing holes.

be detennined with any accuracy from test data. Experience suggests attention be paid to it with radial
flow, single suction impellers of specific speed 2,000 and higher, and absorbing more than 300 kW
(400 HP).
To eliminate the axial thrust of a single-suction impeller, a pump can be provided with both front
and back wearing rings. To equalize thrust areas, the inner diameter of both rings is made the same
(Fig. 5.4). Pressure approximately equal to the suction pressure is maintained in a chamber located on
the impeller side of the back wearing ring by drilling so-called balancing holes through the impeller.
Leakage past the back wearing ring is returned into the suction area through these holes. However, with
large-single-stage single-suction pumps, balancing holes are considered undesirable because leakage
back to the impeller suction opposes the main flow, creating disturbances. In such pumps, a piped
connection to the pump suction replaces the balancing holes. Another way to eliminate or reduce axial
thrust in single-suction impellers is by use of pump-out vanes on the back shroud. The effect of these
vanes is to reduce the pressure acting on the back shroud of the impeller (Fig. 5.5).
Because axial thrust is reduced without additional leakage, such as occurs with back wearing rings,
pump out vanes are sometimes thought to offer higher efficiency. To realize the pressure reduction the
vanes absorb extra power, the net result being no notable difference in efficiency. For this reason, plus
the need to maintain a relatively close axial clearance, pump out vanes are generally used only in pumps
handling solids, where the design helps keep the space between the impeller shroud and casing free of
foreign material.
So far, our discussion of the axial thrust has been limited to single-suction impellers with a shaft
passing through the impeller eye and located in pumps with two shaft seals, one on either side of the
impeller. In these pumps, suction pressure magnitude does not affect the resulting axial thrust. On the
other hand, axial forces acting on an overhung impeller with one shaft seal (Fig. 5.6) are definitely
affected by suction pressure. In addition to the unbalanced force found in a single-suction, two-box
design (see Fig. 5.2), there is an axial force equivalent to the product of the equivalent shaft area through
Axial Thrust in Single- and Multistage Pumps 95

,,--Pump-ouf van~s
/'

- /
/
R~ducflon In pr~ssur~ cau~d by pump - ouf van~s

Fig. 5.5 Reducing axial thrust of single-suction impeller with pump-out vanes.

the seal (shaft area for packed box seals; area of balance diameter for mechanical seals) and the difference
between suction and atmospheric pressure. This force acts toward the impeller suction when the suction
pressure is less than the atmospheric or in the opposite direction when it is higher than the atmospheric.
When an overhung impeller pump handles a suction lift, the additional axial force is very low. For
example, if the shaft diameter through the seal is 50 mm ([2 in.] area 19.6 cm2 [3.14 in.2])-and if the
suction lift is 6.1m (20 ft) of water (pressure 0.42 bar (abs) [6.06 psia]), the axial force caused by the

..
Adddional thrust ..,dh
suction I,f't ac fs In
Addtf,onal fhrust ..,dh
pOJlftv~ suctIon acts
thiS dlr~c tlon In th,s dlr~ctlon

_=J
t=~:j
--~~W~~~r2~~~~--~ ~=~j
/
/
./
,,-- ./
Afmosph~rlc pr~ssur~

Fig. 5.6 Axial thrust problem with single-suction overhung impeller and single stuffing box.
96 Axial Thrust in Single- and Multistage Pumps

overhung impeller and acting toward the suction will be only 116 N (27 lb). On the other hand, if the
suction pressure is 6.9 bar gauge (100 psig), the force will be 1352 N (314 lb) and acts in the opposite
direction. Therefore, as the same pump may be applied for many conditions of service over a wide range
of suction pressures, the thrust bearing of pumps with single-suction overhung impellers must be arranged
to take thrust in either direction. They must also be selected with sufficient thrust capacity to counteract
forces set up under the maximum suction pressure established as a limit for that particular pump.

Special Applications

Special applications may induce extremely high axial thrusts, in boiler circulating pumps, for example.
To reduce shaft seal maintenance, these pumps are usually the single-suction, single-stage overhung impeller
type. Figure 5.7 shows that the force acting on the shaft toward the right is the product of 82.8 bar gauge
(1,200 psig) (the difference between suction and atmospheric pressures) and 78.5 cm2 (12.56 sq in.) the
unbalanced area of 10 cm (4-in. shaft), or 65 kN (15,072lb). Note that with such a high suction pressure
there is no condition which produces a thrust toward the left of the impeller. It is true that providing the
impeller with a back wearing ring would reduce stuffing box pressure by an amount equal to the net pump
pressure. But as this pressure is only 4.14 bar (60 psi), the back wearing ring would reduce the sealed pressure
by only 5 percent, which is negligible on a nominal 80 bar gauge (1,200-psig) boiler circulating pump. On
the other hand, the partial balance of the axial thrust obtained by eliminating the back wearing ring is very
useful. If the unbalanced area ofthe single-suction impeller is 258 cm2 (40 in.2), the net pressure, 4.14 bar
(60 psi), and the average pressure acting on the back shroud 80 percent of the net pressure rise, then the
unbalanced thrust acting to the left is 8.5 kN (1,920 lb) and the net force acting to the right is reduced to
56.5 kN (13,152Ib), or almost 13 percent less. When the pump is standing idle on standby duty, however,
this partial reduction disappears, and the end thrust returns to its former value of 65 kN (15,072lb). The
pump thrust bearing, therefore, must accommodate this higher thrust.
Except for very large units and certain special applications, the maximum thrust developed by mixed-
flow and axial-flow impellers is not of consequence because the operating heads are relatively low.
With axial-flow impellers, axial thrust is caused by the pressure on the vanes as they act on the liquid.
In addition, there is a difference in pressure acting on the two shaft ends, the end in the pump subject
to suction pressure and the other to atmospheric pressure. Occasionally, provision is made in an axial-

Line b~or,ng Tltrusf bearlflg


/

-
VI
Q. - .... ,;a;;;..;,.,.w- -,- -- _ . - -- - - - - - "
_ _ ____ -t~I/lrust

• IJ.l521b
()

~
..... .
\

' 4 m sit art af sfurrmg box

Fig. 5.7 Axial thrust in single-suction boiler circulating pump.


Axial Thrust in Single- and Multistage Pumps 97

flow pump for a running clearance at or near the discharge hub periphery, with balancing holes through
the hub. This construction is used mainly in vertical wet-pit pumps with covered shaft designs so that
the seal at the lower end of the cover pipe will be subject to suction rather than discharge pressure.
With mixed-flow impellers, axial thrust is a combination of forces caused by action of the vanes on
the liquid and those arising from the difference in the pressures acting on the various surfaces. Wearing
rings are often provided on the back of mixed-flow impellers, with either balancing holes through the
impeller hub or an external balancing pipe leading back to the suction.
In the past, some large mixed-flow impeller designs and some high-head vertical pumps with single-
suction radial-type impellers had a running clearance on the back side of the impeller that was larger
in diameter than the leakage joint on the suction side. This disparity caused the axial thrust to act upward,
balancing the dead weight of the rotor. This practice was discarded once more reliable thrust bearings
became available.
The use of wearing rings on the back of large-capacity sewage pumps with mixed-flow impellers has
not met with general approval. Therefore, larger capacity thrust bearings must be used.

AXIAL THRUST IN MULTISTAGE PUMPS

It might seem that the advantages of balanced axial thrust and greater available suction area in a double-
suction impeller would warrant applying such impellers to multistage pumps. But, there are definite
shortcomings to this practice. The average multistage pump has relatively low capacity when compared
to the entire range covered by modem centrifugal pumps. It is seldom necessary, therefore, to use double-
suction impellers just to reduce the net positive suction head required for a given capacity. Even if a
double-suction impeller is desirable for the first stage of a large-capacity multistage pump, it is hardly
necessary for the remaining stages. As to the advantage of the axial balance it provides, it must be
considered that a certain amount of axial thrust is actually present in all centrifugal pumps and the
necessity of a thrust bearing is therefore not eliminated.
Most important, the use of double-suction impellers in a multistage pump adds needless length to
the pump shaft span. Additional space is required for the extra passage leading to the second inlet of
each successive stage. In a pump with four or more stages (Fig. 5.8), this increase becomes quite
appreciable and causes additional casting difficulties. If shaft diameter is increased to compensate for
the longer span so as to maintain reasonable shaft deflection, the impeller inlet areas are correspondingly
reduced. The result is that the advantage of superior suction conditions usually offered by double-suction
impellers is considerably reduced. Finally, as it is impractical to arrange the various double-suction
impellers in any but the ascending order of the stages, the impeller at one end of the casing becomes
the last stage impeller and the pressure acting on the adjacent stuffing box becomes the discharge pressure
of the next-to-Iast stage. To reduce this pressure, a pressure-reducing bushing must be interposed between
the last-stage impeller and the stuffing box, and this bushing further increases the overall length. The
result of all these considerations is that most multistage pumps are built with single-suction impellers.
Two obvious single-suction impeller arrangements for a multistage pump are as follows:
1. Several single-suction impellers may be mounted on one shaft, each having its suction inlet facing
in the same direction and its stages following one another in ascending order of pressure (Fig. 5.9), an
arrangement known as "in-line" or "tandem" impellers. The axial thrust is then balanced by a hydraulic
balancing device (see Chap. 6).
2. An even number of single-suction impellers can be mounted on one shaft, one half of these facing
in an opposite direction to the second half. With this arrangement, axial thrust on the one half is
compensated by the thrust in the opposite direction on the other half (Fig. 5.10). This mounting of single-
suction impellers back-to-back is frequently called "opposed impellers."
98 Axial Thrust in Single- and Multistage Pumps

.....
...
<!)
....,'"
<;>
..,
.,....
<;>
...
<;>
<I
0- .... ....<I
III III
'" III

Sl UF F ING BOX UNDER Slur riNG BOl( UNDER


s ue T 0, 111 PRE SS u RE STAGE ;, PRE SSuRE

Fig. 5.8 Four-stage pump with double-suction impellers.

Fig. 5.9 Multistage pump with single-suction impellers facing in one direction (tandem) and
hydraulic balancing device.

An uneven number of single-suction impellers can be used with this arrangement, provided the correct
shaft and interstage bushing diameters are used to give the effect of a hydraulic balancing device that
will compensate for the hydraulic thrust on one of the stages.
Intuition suggests that opposed impeller rotors inherently have zero axial thrust. In current practical
designs this is not achieved; there is normally some residual thrust when the pump is new, and it either
increases or is developed as the pump's running clearances increase with wear. To explain why this is
Axial Thrust in Single- and Multistage Pumps 99

OP POS ED IMPELLERS

Fig. 5.10 Four-stage pump with opposed impellers.

the case, it's useful to first state the conditions needed to achieve zero axial thrust with an opposed
impeller rotor. These are:

1. The pump must have two shaft seals of equal size.


2. The opposing stages must be of identical configuration.
3. The major pressure breakdowns within the pump must be the same diameter.
4. There must be no leakage between the stages.

Except for some special pumps that have an internal product lubricated bearing at one end, and therefore
only one shaft seal, most multistage pumps fulfill the first condition. The second requirement is also
met in most designs, unless the pump has a special first stage impeller or an odd number of stages. And
when it does, the third requirement, which is normally met too, is varied to compensate as noted earlier
in this chapter. The practical difficulty arises with the fourth condition because there is leakage between
the stages. This has its greatest effect between the back-to-back impeller pairs, because the direction of
the leakage over the impeller shrouds affects the average pressure acting on the impellers' back shroud.
When the flow is radially inward, as it is for the higher pressure impeller, the average pressure is lower
than normal (Fig. 5.11). Similarly, when the flow is radially outward, as it is for the lower pressure
impeller, the average pressure is higher than normal (Fig. 5.11). The net result is that each pair of back-
to-back impellers has some residual thrust, whose magnitude increases with the leakage flow between
the stages. This residual thrust can be balanced for one condition with different diameter pressure
breakdowns, but for simplicity it is usually carried by the pump's thrust bearing. Good design practice
is to size the bearing to accommodate the thrust produced with internal running clearances worn to 2-3
times their new values.
100 Axial Thrust in Single- and Multistage Pumps

Residual Thrust
..

'--"-- Theoretical

L Effect of Ql
Outward
Effect of QL
Inward

Fig. 5.11 Residual thrust in back-to-back opposed impeller pair.

The effect on axial thrust of wear at the pressure breakdown between back-to-back opposed impellers
can be reduced in 2 stage pumps by using a face-to-face arrangement (Fig. 5.12). Because the residual
thrust in this arrangement is produced by a small difference in head per stage, the result of leakage
across the interstage bushing raising the gross capacity of the first stage impeller, the increase in axial
thrust with wear at the clearances is much lower. In the pump shown in Figure 5.12, the pressure at the
shaft seals is equal to impeller discharge pressure less the pressure reduction down the back shroud. For
many applications this is quite acceptable. When it is not, various combinations of impeller axial balancing
(back wearing ring and balancing holes) and breakdown bushings, with appropriately sized bleed-offs,
can be used to lower the pressure at the shaft seals.
Axial Thrust in Single- and Multistage Pumps 101

Fig. 5.12 Two stage pump with face-to-face opposed impellers.

ARRANGEMENT OF STAGES IN OPPOSED·IMPELLER


MULTISTAGE PUMPS

Once a multistage pump is balanced by an opposed-impeller design, the best sequence in which the
individual stages are to be arranged within the pump casing must be determined. This problem is not
simple, as illustrated by analyzing the best arrangement for a six-stage pump. The total number of
possible arrangements is the permutation of 6, that is, 6 x 5 x 4 x 3 x 2 xl, or 720. Of course, half
this number are duplicates, because they describe the same arrangement as viewed in a mirror. The
problem is actually simplified by the fact that most potential sequences are rather complicated and can
be eliminated entirely. Figure 5.13 illustrates four of the most logical possibilities for a six-stage pump.
Three factors enter into the analysis of a satisfactory stage arrangement, and the final solution must
be a compromise between the "best" individual solutions, each satisfying the following requirements:

1. The arrangement of stages provides the minimum possible leakage at the running clearances and maintains
this minimum over a long period of time.
2. The various stage impellers are arranged so that the shaft seals are subject to the lowest pressure in the pump.
3. The sequence of stages precludes excessive complications in the forming of interstage passages.

Running Clearance Leakage


To minimize internal leakage, the pump designer must determine whether it is preferable to keep a
single breakdown under relatively high pressure (as arrangement I in Fig. 5.13, in which breakdown A
102 Axial Thrust in Single- and Multistage Pumps

n
.
, - ---- - - - -- - - _.. - -
:- - --,
~ -- - - --
r- -- \
- - ---
;-- -)
-- - ..... ,

~ :JI-o-~~ i\J lkj


I

c
6 5 4 3
j1J 2

m ;-- - - ~:..-=-:-..=- :::.-_-___..:- =- ::. ~-_=_.:-~ =-:::. ~-~ _~ .:::.:::- =- -=---~~ --':

~ ~_~-~:~! --:-rJil~~) ~ J E

4 6 5 2

III ( ---~_. -_-~~ ~ ~ ..: =.=.;- ==::. = ~_-_-~_--.... ---',


\ I

I :

5 4
~j 2
- - - - - _.. - -_.---- -- - --

Fig. 5.13 Stage arrangements for six-stage axially balanced pump.


(I.) Breakdown A subject to three-stage pressure differential; one shaft seal under high pressure at B. (1/.)
Arrangement with two high-pressure breakdowns, including four-stage pressure differential at C and two-stage
differential at D. (Ill.) Breakdowns E and F under two-stage pressure differential. (IV.) All running
breakdowns subject to only one-stage pressure differential.

is subject to a three-stage pressure differential) or to have many breakdowns under a moderate pressure
differential (such as the five interstage breakdowns in arrangement IV (Figs. 5.13 and 5_14). Experimental
data show that the latter alternative allows the lowest internal leakage, hence the best hydraulic preference.
Before the development of more erosion resistant materials, operating experience showed multiple
breakdowns (arrangement IV in Fig. 5.13) was the arrangement most resistant to wear at the internal
clearances, thereby allowing longer periods between overhauls to restore hydraulic performance or correct
high axial thrust (see the discussion earlier in this chapter). As materials have improved, it has been
established that the simpler stage arrangement I, a single breakdown, relatively longer and with a closer
clearance to account for the higher differential pressure, yields comparable hydraulic performance and
achieves adequate periods between overhauls in most applications. The one exception is applications in
which the pumped liquid contains abrasive solids. For these, past experience suggests that the classic
stage arrangement IV, with essentially the same pressure differential across each running clearance,
would still achieve a longer period between overhauls. The difficulty with this solution is that the design
is very expensive to produce; see later discussion in this chapter.
Axial Thrust in Single- and Multistage Pumps 103

Fig. 5.14 Section of six-stage opposed-impeller pump.


Suction pressure equals zero; pressure generated by each impeller is indicated by P.

Shaft Seal Pressures


In all but one (arrangement I) of the stage arrangements shown in Fig. 5.13, the pressure at the shaft
seals is satisfactory. By placing the two lowest pressure stage impellers at the two ends of the pump,
the seals are subjected to only the lowest pressures in the pump, namely suction and first-stage discharge
pressures. For stage arrangement I, a second breakdown, B, is added to lower the pressure at the adjacent
shaft seal. The leakage from breakdown B is usually returned to pump suction, so the pressure at the
seals is nominally equal. To account for friction losses in the bleed-off piping, it is usual to say the
pressure at the seal adjacent to the breakdown is 1 bar (15 psi) above suction pressure.

Casing Simplicity
Comparing the four possible stage arrangements of a six-stage pump shown in Fig. 5.13, it can be
seen that arrangement I has two high pressure breakdowns, A and B, each with a pressure differential
of three stages, and requires only one casing crossover. Arrangement II has two high-pressure breakdowns:
C with a pressure differential of four stages, D with two, and requires two casing crossovers. Similarly,
arrangement III has two high pressure breakdowns, E and F, both with a pressure differential of two
stages, but requires five casing crossovers to achieve it. Arrangement IV achieves the ideal of having
only the differential pressure of one stage across each of its running clearances, but also requires five
casing crossovers to do so.
Casings with multiple external crossovers (Figs. 3.6 and 5.13) are far more expensive to produce
than those with a single integral crossover (Fig. 3.7). It is for this reason that pump manufacturers
pursued the development of running clearance materials to make stage arrangement I in Figure 5.13 viable,
and as a result established it as the dominant arrangement in modem opposed impeller multistage pumps.
6
Hydraulic Balancing Devices

A single-suction impeller is subject to axial hydraulic thrust caused by the pressure differential between
its two faces. If all the single-suction impellers of a multistage pump face in the same direction, the
total theoretical hydraulic axial thrust acting toward the suction end of the pump will be the sum of the
individual impeller thrusts. The thrust magnitude (in pounds) will be approximately equal to the product
of the net pump pressure (in pounds per square inch) and the annular unbalanced area (in square inches).
Actually the axial thrust turns out to be about 70 to 80 percent of this theoretical value (see Chap. 5).
Some form of hydraulic balancing device must be used to balance the axial thrust and to reduce the
pressure on the shaft seal adjacent to the last-stage impeller. This hydraulic balancing device may be a
balancing drum, a balancing disk, or a combination of the two.
In estimating axial thrust and assessing hydraulic balancing devices, there is a fundamental point to
be noted: axial thrust is only produced by pressures acting over the normal areas between rotating and
stationary surfaces. Figure 6.1 shows a simple axial thrust model. The thrust developed is equal to P
multiplied by the area between D3 and D2, and acts to the left. A much larger force, equal to P multiplied

p
Retention
Nut

_______
t
D1

Fig. 6.1 Elementary axial thrust model.

104

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
Hydraulic Balancing Devices 105

by the area between D2 and Db acts on the retention nut. This larger force is of great consequence to
rotor design but not to axial thrust.

BALANCING DRUMS

The simplest balancing device is a single diameter balancing drum (Fig. 6.2). This design produces a
balancing force by lowering the pressure acting on part of the back of the last stage impeller to close
to suction pressure. To do this the balancing drum, which is keyed to the shaft and therefore rotates
with it, runs with a close radial clearance in the stationary "balancing drum head" or "balancing bushing."
The balancing drum head is installed in a casing diaphragm or partition that separates the region behind
the last stage impeller from the balancing leak-off chamber.
The balancing leak-off chamber is connected either to the pump suction or to the vessel from which
the pump takes its suction. Thus, the back pressure in the balancing chamber is only slightly higher than
the suction pressure, the difference between the two being equal to the friction losses between this
chamber and the point of return. The leakage between the drum head is, of course, a function of the
differential pressure across the drum and of the clearance area.
To assess the effect of a balancing drum, the last-stage impeller and balancing drum can be considered
as a single piece (the pressure between the impeller and drum affects the separating force between them
but not the rotor axial thrust). Looking at Fig. 6.2, the thrusts acting on the balancing drum and last
stage impeller are

1. Toward the discharge end-last stage suction pressure multiplied by the net thrust area at the last stage
impeller inlet (area "A").
2. Toward the suction end--discharge pressure multiplied by the last stage unbalanced area (area "B"), plus
the back pressure in the balancing chamber multiplied by the back balancing area (area "C").

The first force is greater than the second, thereby counterbalancing the axial thrust exerted upon the
single-suction impellers. Actually, the drum diameter can be selected to balance axial thrust completely
or within 90 to 95 percent, depending on the desirability of carrying any thrust bearing loads.

[" --- --- --- --- --[-

Area 'A'
Area'S'
rhL
Area 'C'

Fig. 6.2 Balancing drum.


106 Hydraulic Balancing Devices

It has been assumed in the preceding simplified description that the pressure acting on the impeller
walls is constant over their entire surface and that the axial thrust is equal to the product of the total
net pressure generated and the unbalanced area. The actual pressure distributions are more complex.
They are not constant over the shrouds and generally not equal at two corresponding points on opposite
impeller shrouds (D and E, Fig. 6.2). Broadly, three factors account for this:

1. Rotation of the liquid adjacent to the impeller shrouds produces a pressure reduction in the radial direction
(see Fig. 5.2).
2. Leakage flow direction and velocity affect the pressure distribution over the impeller shroud. Flow outward
increases the average pressure, whereas flow inward decreases it. For all but the last stage impeller of tandem
impeller pumps (Fig. 6.2), the leakage flow is outward over the back shroud, inward over the front, giving
a higher average pressure on the back shroud, hence higher axial thrust. As clearances increase with wear,
the higher leakage velocities act to further increase axial thrust. Figure 6.3 illustrates the point.
3. The clearance between impeller shroud and casing wall influences the average angular velocity of the liquid
in that region, hence the pressure reduction due to centrifugal action. Reducing the clearance tends to increase
the pressure reduction, thus lowering the average pressure over the shroud. Depending on the position of
the impeller, this effect may reinforce or oppose that of leakage.

The effects just described are for a given flow. As flow through the pump varies, there are two addi-
tional effects:

1. The static pressure at the impeller discharge varies, ranging from 75-80 percent of stage pressure rise at
best efficiency point to about 90 percent at shutoff. As the static pressure at the impeller discharge varies
so does the average pressure over the shrouds.

- - - - - Centrifugal Force Only


- - - - With Influence of Leakage, QL, Direction
- - - - - Effect of Wear at Clearances

Pi Pi

,~ p.
.\
\~ sf-

\
. \\
\ \
. I

Fig. 6.3 Effect of leakage direction on shroud pressure distribution.


Hydraulic Balancing Devices 107

2. Axial thrust produced by the momentum of the liquid entering the impeller varies. Specifically, reducing
the flow lowers the momentum, thus increasing thrust toward the suction.

By drawing correction factors from actual thrust measurements and allowing for tolerable manufacturing
variations, axial thrust can be estimated with acceptable accuracy over the range of normal operating
flows. Because the thrust developed by the balancing drum is nearly constant, varying only with pump
differential pressure, the design task becomes one of trying to insure the thrust is always inboard over
the pump's operating flow range and its magnitude within the thrust bearing's capacity. The latter
requirement dictates care, as the thrust magnitude is the small difference between two large forces.
Figure 6.4 shows a typical axial thrust characteristic.
Depending On the actual pump design, the onset of secondary flows at the impeller discharge as flow
is reduced can significantly alter the shroud pressure distributions. The greater effect is usually over the
front shroud (inward leakage flow) and can produce thrust reversal (see Fig. 6.4). Being a product of
an unstable flow regime, these conditions also tend to fluctuate, which poses the possibility of fatigue
failure of the rotor or thrust bearing. When this problem is anticipated, designing the impeller and
adjacent casing wall to isolate the shroud space (Fig. 4.35) is an effective solution.
Most balancing drum designs incorporate some form of grooving on the drum. This is done to (1)
decrease the leakage flow by raising the friction coefficient and (2) reduce the risk of seizure at the
clearance should there be metal-to-metal COntact between the parts. The form of the grooves varies
markedly, a product of the degree of hydraulic rotor support sought by the design.
A balancing drum, correctly sized, satisfactorily balances most of the axial thrust of single suction
impellers and reduces the pressure at the discharge side shaft seal. It does this at some expense to
efficiency because the leakage is usually relatively high, but with the advantage of simplicity since there
is no close axial clearance to be set during assembly. What it lacks, however, is the virtue of automatic

Bypass
Bearing
Capacity

+ (positive)
Toward
Suction

Axial Q
Thrust Osap

- (negative)
Toward
Discharge CD Ideal: Flat characteristic, no thrust reversal
® Acceptable: Steep characteristic, thrust reversal at part load
® Undesireable: Sudden thrust reversal at part load

Fig. 6.4 Typical axial thrust characteristics with balancing drum.


108 Hydraulic Balancing Devices

compensation for any changes in axial thrust caused by variations in pressure distribution over the
impeller shrouds (see earlier discussion). The net result is a fairly wide variation in residual thrust over
the pump's operating flow range (Fig. 6.4). To achieve automatic compensation for changes in impeller
thrust, it is necessary to have a balancing device whose balancing force varies with the rotor's axial
position. The simplest of these devices is the balancing disk.

BALANCING DISKS

The operation of the simple balancing disk is illustrated in Fig. 6.5. The disk is fixed to and rotates with
the shaft. It is separated from the balancing disk head installed as a casing part, by a small axial clearance.
The leakage through this clearance flows into the balancing chamber and from there either to the pump
suction or to the vessel from which the pump takes its suction. The back of the balancing disk is subject
to the balancing chamber back pressure whereas the disk face experiences a range of pressures. These
vary from discharge pressure at its smallest diameter to back pressure at its periphery. The inner and
outer disk diameters are chosen so that the difference between the total force acting on the disk face
and that acting on its back will balance the impeller axial thrust.
If the axial thrust of the impellers should exceed the thrust acting on the disk during operation, the
latter is moved toward the disk head, reducing the axial clearance between the disk and the disk head.
The amount of leakage through the clearance is reduced so that the friction losses in the leakage return
line are also reduced, lowering the back pressure in the balancing chamber. This automatically increases
the pressure difference acting on the disk and moves it away from the disk head, increasing the clearance.
Now, the pressure builds up in the balancing chamber, and the disk is again moved toward the disk
head until an equilibrium is reached.
To ensure proper balancing disk operation, the change in back pressure in the balancing chamber

4 --
Restricting Orifice

To Suction---... Balancing Chamber

Balancin Disk Head

Fig. 6.5 Simple balancing disk.


Hydraulic Balancing Devices 109

must be of an appreciable magnitude. Thus, with the balancing disk wide open with respect to the disk
head, the back pressure must be substantially higher than the suction pressure to give a resultant force
that restores the normal disk position. This can be accomplished by introducing a restricting orifice in
the leakage return line that increases back pressure when leakage past the disk increases beyond normal.
The disadvantage of this arrangement is that the pressure on the shaft seal is variable-a condition that
will shorten the life of the seal and is therefore to be avoided.
Many older pump designs used a balancing disc without a supplementary thrust bearing to limit the
rotor's axial movement. Under stable operating conditions, this simple design demonstrated adequate
reliability. The same was not so during transient conditions (flow swings, cavitation, or similar), which
usually resulted in serious damage to the balancing disc and often the rest of the pump. It is for this
reason that the use of a balancing disc without a thrust bearing is now generally limited to small pumps.

COMBINATION DISK AND DRUM

For the reasons just described, the simple balancing disk is seldom used. The combination balancing
disk and balancing drum (Fig. 6.6) was developed to obviate the shortcomings of the disk while retaining
the advantage of automatic compensation for axial thrust changes.
The rotating component of this balancing device is a long cylinder with a wide-faced flange. Acting
with the stationary component, known as a "disk head," the cylinder forms a balancing drum, the flange
a balancing disk. In this design, radial clearance remains constant regardless of disk position, whereas

1- Axial
Clearance

B Radial
Clearance 1
E
-+ +
---1----+
A
-+ L..::~O-=:::I

Intermediate
Relief Chamber

Fig. 6.6 Combination balancing disk and drum.


110 Hydraulic Balancing Devices

the axial clearance varies with the pump rotor position. The following thrusts are developed by this
device acting in conjunction with the last stage impeller:

1. Toward the discharge end-last stage suction pressure multiplied by area A, plus the intermediate pressure
multiplied by area C, plus the average pressure across the axial clearance multiplied by area D.
2. Toward the suction end-last stage discharge pressure multiplied by area B, plus the back pressure multiplied
by area E.

Whereas the "position-restoring" feature of the simple balancing disk required an undesirably wide
variation of the back pressure, it is now possible to depend on a variation of the intermediate pressure
to achieve the same effect. Here is how it works. When the pump rotor moves toward the suction end
(to the left in Fig. 6.6) because of increased axial thrust, the axial clearance is reduced, and pressure
builds up in the intermediate relief chamber, increasing both the value of the intermediate pressure acting
on area C and the average pressure acting on area D. In other words, with reduced leakage, the pressure
drop across the radial clearance decreases, increasing the pressure drop across the axial clearance. The
increase in intermediate pressure forces the balancing disk towards the discharge end until eqUilibrium
is reached. Movement of the pump rotor toward the discharge end would have the opposite effect of
increasing the axial clearance and the leakage and decreasing the intermediate pressure acting on area
C and the average pressure on area D.
Figure 6.7 illustrates the pressure distribution in a combination balancing disk and drum. No attempt
is made to describe the exact manner in which the pressure decreases between any two points, although
this curve is not necessarily a straight line. Also, this illustration is not quantitatively correct. It only
serves to show that changes in the balancing device position vary the internal pressure distribution
without altering the back pressure. The only possible variation may be caused by pressure changes at
the point where the balancing device leakage is returned to the system. An orifice may still be located
in the return line. Its function now, however, is not that of changing back pressure but rather of gaging
the volume of leakage flow. This flow should not be throttled outside the balancing device; the orifice
pressure drop is negligible, ranging from about 0.15 to 1.5 bar (2 to 20 psi).

Balancing device modifications


Most pumps employing a combination disc drum to balance axial thrust also have a thrust bearing
to limit rotor movement (see Balancing Disks). The bearing is typically sized for the residual thrust
produced with the disc faces at some minimum clearance determined by the pump manufacturer. In high
temperature applications, such as hot charge in refining, there have been instances where differential
expansion during thermal transients increased the minimum disc clearance to the point where the thrust
bearing became overloaded and failed.
Various modifications of hydraulic balancing devices are now in use. All try to retain the self-
compensating feature of the disk or combined disk and drum, while avoiding the risk of face contact
and scoring inherent in a close running axial clearance (typically 0.050 to 0.125mm [0.002 to 0.005
in.]). One means of realizing this is to soften the restoring action by lowering the proportion of the total
pressure drop taken across the axial clearance. This is done by narrowing the disk face of a combined
disk and drum, and either leaving the disk at the end of the drum (Fig. 6.6) or locating it between two
drum regions (Fig. 6.8). The latter design, often referred to as a stepped drum, has the added virtues of
a substantial back pressure at the disk, hence no flashing across the disk, and the ability to "fine-tune"
net thrust by changing the clearance at one drum. In some pumps using stepped balancing drums, the
design axial clearance is fairly large, typically 0.4Omm (0.015 in.), to further reduce the risk of contact
at the disk faces. Set this way, the disk's self-restoring action is normally inactive, serving only as
backup protection to control rotor position should the thrust bearing cease doing so.
Hydraulic Balancing Devices 111

Pel

w
a::
::;)
(/) Pe2
(/)
w
a::
11.

Pc

RADIAL
CLEARANCE
ALANCING
CHAMBER

Fig. 6.7 Pressure distribution in combination balancing disk and drum.


Key:
PA = discharge pressure;
PB = pressure at intermediate relief chamber;
Pc = back pressure;
I = normal pressure distribution;
II = pressure distribution after disk moves away from disk head;
PA - PB = pressure drop through drum portion;
PB - Pc = pressure drop through disk portion.

Another approach to the problem of a close-running axial clearance is to change its detail configuration
so the restoring action is more stable. Tapering one of the faces so the axial clearance converges in
the flow direction (Fig. 6.9) achieves this. It does so by increasing the rate of pressure drop as the
clearance decreases.
Materials selection for balancing devices will be treated in a subsequent section. However, it is
imperative to remember that both the material and the design are extremely important. If the balancing
device wears appreciably, problems can develop, the nature of which depends on the type of balancing
device. With a single diameter balancing drum (Fig. 6.2), pump performance will deteriorate (from
increased leakage), and eventually axial thrust will increase as the back pressure in the balancing leak-
off chamber rises. A balancing disk (Fig. 6.5) or combination disk and drum (Fig. 6.6) without a thrust
bearing poses the risk of contact between the impellers and some internal pump part as the rotor moves
axially to compensate for wear. When a thrust bearing is used, rotor axial movement to compensate for
wear is limited, and so wear leads to a drop in performance (from higher leakage) and higher thrust
bearing loads as the disk portion becomes less effective.

INDIVIDUAL AXIAL THRUST BALANCING

A design sometimes used to balance the axial thrust of single-suction impellers without the use of a
hydraulic balancing device is illustrated in Fig. 6.10. It provides the individual impellers with wearing
112 Hydraulic Balancing Devices

Fig. 6.8 Combination balancing disk and drum with disk located in center portion of drum.

Disc
- 11- Clearance

Convergent
Parallel
Faces
Faces

Press "-
Drop ~

Clearance

Fig. 6.9 Balancing disk modification for improved stability.


Hydraulic Balancing Devices 113

2nd Stage 3rd Stage


Discharge Discharge

Fig. 6.10 Balancing axial thrust of single suction impellers with back wearing rings.

rings both in front and back, the inner diameter of both rings being the same to equalize the thrust areas.
Balancing holes are drilled through the impellers to equalize the pressures from front to back. This
prevents the leakage water that flows across the back sealing surface from collecting in the annular space
back of the impeller and building up the pressure at that point. For convenience, the back wearing ring
may form an integral part with the diffuser or stage-piece bushing.
This design theoretically provides axial balance. Although the pressures on the two impeller sides
may not be exactly equal in practice (because of unequal wearing ring leakage), the amount of unbalance
is rather small and can usually be accommodated by the thrust bearing. Unfortunately, the use of a back
wearing ring becomes less justifiable when one considers the effects of this construction on multistage
pump internal leakage and mechanical design.
Normally, pumps with impellers arranged in ascending stage order enjoy almost negligible pressure
difference and leakage across the stage-piece separating two consecutive stages. However, with a wearing
ring at the back of the impeller, this difference becomes equivalent to the pressure generated by one
stage. Thus, two additional clearance joints subject to a full one-stage pressure difference are now used,
namely, the back wearing ring and the stage-piece joints.
A back wearing ring design is lacking in one other respect, compared to a hydraulic balancing device.
That is, it does not reduce the pressure on the discharge-end shaft seal which is now subjected to the
suction pressure of the last stage. Unless this pressure can be reliably sealed (a pronounced improbability
in other than low pressure pumps), it is necessary to provide some form of pressure-reducing mechanism
ahead ofthe box. This requirement, in addition to the space needed for the back wearing rings, substantially
increases the total shaft span of a multistage pump and makes the design even less desirable.

COMPARISON OF BALANCING DEVICES AND OPPOSED IMPELLERS

On the surface, it would appear that the choice between using a balancing device and arranging impellers
in opposed sequence to balance axial thrust reflects a basic difference in design philosophy. Consequently,
this choice has always been controversial among designers and users. The argument supporting each
114 Hydraulic Balancing Devices

111
111 A
~ 'If?
--3'------1.4- ----'-------'6---7'------'8'- - - - - - - -
2

Fig. 6.11 Balancing axial thrust with balancing device in eight-stage pump.
Breakdown A is subject to a differential pressure of eight stages.

method centers on the presence or absence of a balancing device subject to a differential pressure equal
to the total pressure generated by the pump, and the effect that it has on potential wear rate, rotor
dynamics, and sensitivity to wear.
The argument over potential wear rate is strictly semantic. To balance axial thrust and keep the
pressure at the seals low, the total differential pressure has to be broken down in either case. Whether
this is accomplished across a single breakdown (Fig. 6.11) or the balancing device is split into two
separate portions distributed throughout the pump (Fig. 6.12) and given a different name is immaterial
so long as the running joints in each case are of proper length. Wear is essentially a function of the
pressure drop per unit length of break down and if the lengths of the breakdowns are chosen to maintain
the same pressure drop unit length the wear will not be affected by the number of breakdowns nor by
the pressure differential across them. This assumes the risk of rubbing contact at the major breakdowns
is similar, since contact within clearances is also a factor in wear rate.
There is a significant difference between the two designs in their effect on rotor dynamics. Because
the major pressure breakdowns also act as hydrostatic bearings (see the Lomakin effect in Chapter 7),
their position along the rotor becomes important. The center bushing in an opposed impeller arrangement
(Fig. 6.12) is located close to rotor midspan, effectively halving the bearing span when the pump is
running with liquid in it, and so changing the rotor's dynamics considerably. It is this effect that allows
the satisfactory operation of slender shaft rotors. The balancing device of tandem impeller rotors also
acts as a hydrostatic bearing, but because it is located relatively close to one end of the pump, its effect
on the rotor's dynamics is much less. See Chapter 7 for further discussion on rotor dynamics.
Sensitivity to wear at the running clearances has two aspects: residual thrust and rotor dynamics. In
pumps with tandem impellers, the impeller thrust increases with wear. Balancing discs and combined
disc drums compensate for this up to the point where the disc clearance is controlled by the thrust
bearing. Balancing drums cannot compensate for the additional impeller thrust, and so the thrust balance
changes. The residual thrust mayor may not increase depending on whether the original balance produced
reverse thrust at some capacity (Fig. 6.4). Opposed impellers develop additional residual thrust at the
opposed impeller pair (see Chapter 5) and from any differences in flow within the pump caused by

2 3 4 8 7 6 5

Fig. 6.12 Balancing axial thrust with opposed impellers in eight-stage pump.
Breakdowns A & B are each subject to a differential pressure of four stages, for a total of eight stages.
Hydraulic Balancing Devices 115

uneven wear at the clearances. For most designs within the limits of nonnal wear at the clearances, the
increase in residual thrust is quite small.
Unless the running clearances have been intentionally designed to maximize the Lomakin effect, the
rotor dynamics of pumps with balancing devices are not very sensitive to wear at the clearances. Opposed
impeller pumps, on the other hand, are very sensitive, particularly those with slender rotors. As the
major breakdowns in these designs wear, the stiffness of the hydrostatic bearing at rotor midspan
decreases, thereby significantly changing the rotor's dynamic characteristics (see Chapter 7).
With some sensitivity to wear in both designs, it is useful to be able to measure the leakage through
the major breakdowns. These data give a direct indication of the condition of the major breakdowns,
and an indirect indication of the condition of the other clearances which nonnally wear at about the
same rate. Pumps with a balancing device have only one major breakdown at one end of the pump (Fig.
6.11), where the leakage from it can be easily measured. The same can be done with the end or balance
bushing in the usual opposed impeller arrangement (Fig. 6.12), but it is not practical to measure the
leakage through the center bushing, the other major breakdown in this design.
The various points of this comparison between balancing devices and opposed impellers as means
of balancing hydraulic thrust are summarized in Table 6.1.
Despite all that can be stated in comparing the two designs, there are, in the history of multistage
pump applications, many instances in which a pump with a balancing device perfonned better than one
with opposed impellers and vice versa. Many factors besides those already discussed can account for
this, hydraulic design, rotor stiffness, materials, and casing construction to name four, but from the pump
user's point of view one design perfonned better than the other, and therefore personal preference often
makes the choice between the two designs.

Table 6.1 Comparison of Balancing Devices and Opposed Impellers

Balancing Device
Factor Drum Disc Drum Opposed Impellers

Suitable for high temperature Yes Nd 1) Yes


Wear rate All approximately equal(2)
Effect on rotor dynamics Minor Major
Residual thrust Small(3) Negligible Small
Effect of wear on:
-residual thrust Moderate(4) Negligible(5) Varies(6)
-rotor dynamics Minor Major
Measure leakage Yes No(7)

Notes:
1. Differential thermal expansion during thermal transients can open disc clearance leading to thrust
bearing failure from overload.
2. Provided pressure drop per unit length similar and no rubbing contact.
3. Residual thrust low at design capacity; generally increases as flow reduced.
4. Depends on thrust characteristic and original thrust balance.
5. Within limits of rotor movement allowed by thrust bearing.
6. Depends on number of stages, number of opposed impellers pairs, and equal wear at high
pressure breakdowns.
7. Leakage through the center bushing (Fig. 6.12) cannot be measured
7
Shafts and Shaft Sleeves

The primary function of a centrifugal pump shaft is to transmit torque to the impeller. At the same time,
the shaft and its support arrangements must maintain alignment of the rotor within its running clearances
and through the shaft seal.
These functions lead to the following design requirements: First, the shaft must be strong enough to
withstand sudden starting, such as occurs when a motor is started across the line. It should also tolerate
the higher power associated with expected abnormal operation such as starting a hot service pump on
cold liquid, variations in the process (resulting in high liquid sa or viscosity), reduced system head
(allowing the pump to run to high capacity), or, in axial flow pumps, increased system head (causing
the pump to run back at lower capacity, hence higher power). Second, the shaft must be designed for
tolerable deflection under the weight of the rotor and any hydraulic loads, either radial or axial. Third,
the shaft must not be liable to any significant response to a resonant condition (excitation of a critical
speed), lest it develop destructive vibration.

CRITICAL SPEEDS

As critical speed is a key factor in the selection of shaft diameters, the centrifugal pump user ought to
have a general knowledge of this subject.
Developments in pump design have compounded the need. Smaller pumps running at higher speeds
have brought about a significant improvement in the understanding of pump rotor dynamics. Applied
correctly, this improved understanding results in better pumps. Applied generally, it can result in unduly
restrictive, hence expensive, specifications for quite simple pumps.
Any object made of an elastic material has an infinite number of natural frequencies, each of a
different mode shape. When a pump rotor or shafting system rotates at any speed corresponding to a
natural frequency, the vibration produced by minor unbalance will be magnified. These speeds are called
the critical speeds.
In conventional pump designs, the rotating assembly is theoretically uniform around the shaft axis,
and the center of mass should coincide with the axis of rotation. This theory will not hold for two

116

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
Shafts and Shaft Sleeves 117

reasons. First, there are always minor machining or casting irregularities, and second, there will be
variations in metal density of each part. Thus, no matter how carefully a rotor is made, there will always
be some eccentricity of the center of mass to produce centrifugal force and therefore deflection as the
assembly rotates. A critical condition is reached when the rotative speed is such that the centrifugal
force equals the shaft's elastic restoring force. With these opposing forces equal, the only force restraining
the shaft, neglecting effects at the running clearances, is that produced by viscous friction within the
shaft material. If the rotor is allowed to run at this condition without external restraint, there is a high
risk that vibration will increase until the shaft fails. On occasion, the same phenomenon can occur at a
speed above or below the pump's rotative speed. For this to happen, the rotor must be excited by a
disturbing force rotating at or close to a critical speed. To rotate at a speed other than the pump speed,
the disturbing force must be of hydraulic origin.

Rigid and Flexible Shaft Designs


The lowest critical speed is called the first critical speed; the next higher is called the second, and
so forth. In centrifugal pump nomenclature, a "stiff shaft" means one with an operating speed lower
than its first critical speed, whereas a flexible shaft is one with an operating speed higher than its first
critical speed.

ROTOR MECHANICAL STIFFNESS

In the simplest analysis, the rotor of a centrifugal pump (Fig. 7.1) is considered running in air and
supported by infinitely stiff bearings (Fig. 7.2[a]). By taking the equation for the natural frequency of
a single mass rotor:

......:.... ............

_~~-~::~----- -- ---~~~ ':~~~.~~ ---- --- ---':'~_:~:'[~_ J

Fig. 7.1 Rotor assembly of a single-stage double-suction pump.


118 Shafts and Shaft Sleeves

(8)

(b)

Fig.7.2 Rotor support models.


a) Rotor supported by infinitely stiff bearings; critical speed determined by rotor mass and stiffness alone.
b) Rotor supported by bearings offinite stiffness and having damping, with additional support and damping at inter-
nal clearances; critical speeds determined by motor mass, rotor stiffness, support stiffness, and support damping.

O)n = (k/m)D.5

where k is stiffness and m is mass, both in consistent units, and rearranging it, the first bending critical
speed, NCh can be estimated from

NCI = 946/(Y,,)D.5 in metric units or NCI = 188/(y,,)D.5 in US units


where NCI = critical speed in rpm
y" = rotor static deflection in mm (in).

The rotor static deflection Yst is that resulting from rotor weight alone, and is therefore a measure of the
rotor's mechanical stiffness. Under the action of radial hydraulic forces, the running rotor deflection
Shafts and Shaft Sleeves 119

may well be higher. That is of consequence to the shaft seal and running clearances, but has no bearing
on critical speeds unless seal or clearance considerations dictate a change in shaft geometry. For all but
the simplest shafts, calculating a rotor's static deflection requires either a reliable approximation for
manual work or use of a computer routine if greater accuracy is necessary.
The significance of mechanical stiffness can be gleaned by considering two rotor designs, one to run
at 1,800 rpm, the other at 3,600 rpm, both to be "stiff." A usual margin for rigid motors is to have the
critical speed at least 20 percent above of the running speed, giving critical speeds of 2,160 and 4,320
rpm for the two rotors being considered. The corresponding static deflections are 0.193 and 0.048mm
(0.0076 and 0.0019 in.) A rotor whose static deflection is 0.193mm (0.0076 in.) is relatively easy to
design, and, in fact, would likely be somewhat stiffer to avoid contact at the running clearances. Except
for pumps of only one or two stages, limiting the static deflection to 0.048mm (0.0019 in.) makes the
rotor design a great deal more difficult unless unusual proportions are resorted to. Most 3,600-rpm
multistage pumps, therefore, have rotors that are "flexible" when analyzed on the basis of mechanical
stiffness alone.

HYDRAULIC STIFFNESS (LOMAKIN EFFECT)

While analyzing critical speeds considering only rotor mechanical stiffness is simple, it is not sufficiently
accurate for many designs, because other factors significantly affect the actual dynamic behavior. The
dynamic behavior of all pump rotors is influenced to some degree by the following factors:

1. Finite bearing stiffness, with sleeve bearings having the lowest stiffness of the arrangements used
2. Additional rotor support and damping from radial type shaft seals such as packed box and breakdown bushings
3. Additional rotor support and damping from hydraulic effects at internal running clearances
4. Increased effective mass from motion of the liquid adjacent to the vibrating rotor.

These factors change the rotor support arrangement from the simple two infinitely stiff bearings arrange-
ment in Fig. 7.2(a) to the much more complicated series of springs and dashpots in Fig. 7.2(b). The net
effect of these factors varies with pump detail design. In most cases where the net effect is of consequence,
typically multistage and high-speed pumps, the dominant component is that produced by a hydraulic
effect at the internal running clearances. This effect is known as the Lomakin effect after its discoverer
and can be termed hydraulic stiffness for comparison with a rotor's mechanical stiffness. Under the
action of significant Lomakin effect, the rotor's running critical speed is raised to a higher value (additional
support or stiffness) and the response through the critical speed reduced (damping; see Fig. 7.3). Note
that in very large low-speed pumps, the dominant factor tends to be increased effective mass, leading
to a reduction in the running critical speed.
The Lomakin effect is a product of hydrostatic (pressure) and hydrodynamic (velocity and viscosity)
action as liquid passes through a pump's internal running clearances under the action of a differential
pressure. As such, the magnitude of the Lomakin effect depends on the pressure drop across the running
clearances, the surface speed at them, their geometry (length, clearance, and surface form), and the liquid
viscosity. Given these dependencies, there is very little Lomakin effect until the pump is running at say
75 to 80 percent of rated speed. For the same fundamental reasons, clearances with a large length-to-
clearance ratio and smooth surfaces develop more Lomakin effect than those with a low length-to-
clearance ratio and one surface serrated. From the same reasoning, it is evident running clearance wear
will reduce the Lomakin effect. Tests show the reduction is greater in arrangements designed to produce
high Lomakin effect (Fig. 7.4).
120 Shafts and Shaft Sleeves

0.8
a.
I

J!1
'E
ai 6 / Without Lomakin Effect
"0
::I
:t:::
a..
E
<C 4
c:
0
~...
.0
:> 2 r With Lomakin
I Effect

1000 2000 3000 4000 5000


Pump Speed, rpm

Fig. 7.3 Influence of Lomakin effect on rotor dynamic response.


(I. J. Karrasik et ai, Pump Handbook, 1st Edition, 1976. Reproduced with permission of the
McGraw-Hill Companies)

Computer routines are used to estimate the running dynamic behavior of rotors, but the accuracy of
their estimates depends on how well the running clearances can be modeled, a process that does not yet
yield entirely consistent results. To check the accuracy of a prediction it is necessary to determine the
rotor's natural frequencies while it is running. Measurements during rundown give an approximation,
but they are on the low side since the Lomakin effect is decreasing rapidly as the pump's speed drops.
Measurements at running speed require some form of external excitation. Variable-speed shaking or
calibrated impact are two available methods; the second, with the appropriate instrumentation, is the easier.
To differentiate between the two approaches to analyzing rotor dynamics, the terms "dry critical"
and "wet critical" have been introduced. The former refers to an analysis assuming the rotor is supported
by infinitely stiff bearings alone, the latter refers to an analysis recognizing the influence of the Lomakin
effect, liquid around the impeller, shaft seal support, and the finite stiffness of bearings. Both values
tend to be used in pump evaluations because the difference gives a measure of the pump's sensitivity
to wear at its running clearances.
As a general rule, it is prudent to require rotors with low Lomakin effect to have at least a 20 percent
margin separating running and critical speeds. When the Lomakin effect is high, the rotor's relative
response to excitation at the critical speed becomes the governing criterion. This is necessary because
the high damping spreads the region of response (Fig. 7.3), making a clear separation hard to define.
The response of such rotors should always be calculated for both new and worn clearances, to reduce
the risk of destructive rotor vibration as the clearances increase with wear.

TORSIONAL CRITICAL SPEEDS

Two or more masses connected by a torsionally elastic shaft (Fig. 7.5), have n - I natural frequencies,
where n is the number of masses connected by the shaft. In a manner similar to lateral or bending critical
Shafts and Shaft Sleeves 121

7000

6500 \.

\\
\.
6000
~
a.
~
al 5500

"
~
(/J
oo~
"iii
o
+=' 5000
8
~
4500

4000
c:
Ol
·iii
........ r--......

Q)
0
c:
- ~O\led
~- ~ "
~
co
Q)
0 Q) D,. P across wear ring
(3
I = 300 PSI
3500
0 .01 .02 .03 .04 .05
Diametral Clearance (Inches)

Fig. 7.4 Effect of running clearance wear on wet critical speed.

12 (Driver)
(P ump) 11
r--""-'-

- - - -- -- t---- - - l-

I
- kt
Shaftin 9

Fig. 7.5 Torsional critical speed model.

speeds, sustained operation of a rotor under conditions where one of its torsional natural frequencies is
close to an exciting frequency, can lead to large amplitude torsional vibration with consequent shaft
failure. In centrifugal pump practice, torsional critical speeds are usually of consequence only when the
rotor has low torsional stiffness (long and slender) and the drive has a pulsating characteristic, for
example, an engine or variable-frequency-controlled motor. Since torsional natural frequencies are a
product of connected masses, they are a property of the pump plus its drive train, including couplings,
and can only be determined as such. A potential torsional resonance problem is best corrected by changing
122 Shafts and Shaft Sleeves

the stiffness of the connection, usually by modifying the coupling. In the rare instances where resonance
cannot be easily avoided, torsional stress analysis is carried out to determine whether the resonance can
be tolerated.

SHAFT SIZING

Sizing a shaft is an iterative process. In general, a minimum diameter is determined for strength, steps
to enable assembly are added, deflection is checked, then the dynamic characteristics are predicted. The
process is then repeated until an acceptable solution is realized. Pump designers, of course, have various
methods to enable them to quickly develop an approximate solution.
Sufficient shaft strength requires that the shaft not fail by fatigue, taking account of the loads imposed
during starting, normal operation, and any specified abnormal operation. This means that the endurance
strength of the shaft material, in the pumped liquid if that is the case, must be above the combined stress
at the critical point in the shaft. How far above (how big a design factor) depends on the certainty of
loading, liquid condition, shaft detail at the critical point, and material condition. Loading should assume
lO percent torsional fluctuation and make provision where applicable for some coupling misalignment.
Nominal torsional stress at the smallest diameter is a common evaluation approach. Conservative, but
not absolutely safe, values for nominal torsional stress are

Shaft Material Typical Torsional Stress-MPa (psi)


Carbon steel 48 (7,000)
Alloy steel, chrome steel 55 (8,000)
316 stainless steel 35 (5,000)

A discussion of deflection and dynamics is easiest if the two basic rotor configurations, overhung and
between bearings, are considered separately.
For overhung rotors accepted design practice now requires the following:

1. Deflection, static plus that caused by imposed forces under the worst expected operating conditions, should
be less than the available internal clearance.
2. Deflection at the shaft seal should be less than 0.050 mm (0.002 in).
3. First bending critical speed (dry basis) should be at least 20 percent above the running speed (except for
high-speed pumps).

These requirements have become accepted practice, because they have been proven reliable (running
clearance wear, seal, and bearing life), are simple to design for, usually do not exact any performance
penalty, and do not cause an undue increase in pump cost.
Determining all the data necessary to verify a rotor meets the preceding design requirements is quite
an extensive task. A simpler, and generally sufficiently accurate, alternative approach is to employ a
factor that is proportional to static deflection but can be calculated from shaft geometry alone. By
comparing factors or using an upper limit based on experience, a series of offers or a troublesome
pump can be quickly evaluated. For cantilever rotors of the form shown in Fig. 7.6, the deflection is
proportional to
Shafts and Shaft Sleeves 123

IMP BRG BRG IMP BRG BRG


<t <t <t <t <t <t
1-.-~1--·1- L2-1 ~ Li -1- ~2---1

Fig. 7.6 Cantilever shafts.

where Ll is the impeller overhang, Lz is the bearing span, Dl is the shaft diameter under the shaft sleeve,
and D2 is the shaft diameter between the bearings.
Common practice for shafts with the larger diameter between the bearings (Fig. 7.6(a)) is to evaluate
them using the first term of the equation just given, specifically

since the .error introduced by this simplification is quite small. When Dl is greater than D2 (Fig. 7.6(b)),
the error can be significant thus the result quite misleading.
Acceptable values of the simplified shaft flexibility factor, L I3/D14, decrease with increasing pump
size. Figure 7.7 shows the characteristic of shaft flexibility factor over the normal range of process
pumps for designs that meet API-610 [3.1] and for designs intended to have shafts 2 times as stiff as
required by API-61O. The latter is an informal standard that has evolved in the United States over the
past decade from the oil industry's emphasis on shaft stiffness (see Chap. 27).
Between bearings, rotors for one- and two-stage pumps are usually relatively short and therefore
easily made to meet the simple requirements for cantilever rotors. Between bearings rotors for pumps,
having three or more stages, cannot be designed to withstand any appreciable radial thrust, so twin
volutes, diffusers, or staggered single volutes (see Chaps. 2 and 3) are employed to reduce radial thrust
to negligible values. The design of the rotors themselves is controversial. One philosophy, resting on
lower cost and higher efficiency, advocates low mechanical stiffness (slender shaft) and high hydraulic
stiffness (Lomakin effect). The opposing philosophy, claiming that reliability is more important than
first cost and efficiency (the efficiency of a failed pump is zero) calls for high mechanical stiffness (large
shaft) and low hydraulic stiffness. In a more general sense, it is worth noting that this controversy is
peculiar to liquid handling turbomachines; those handling gases must rely on mechanical stiffness because
the hydraulic stiffness is very low.
Slender shaft rotors are sized for torque, then stepped up to permit assembly and perhaps reflect some
empirical stiffness requirement. Their static deflection is greater than the available internal clearance
and the dry first bending critical speed is consequently well below the running speed. As the pump is
accelerated, the Lomakin effect rapidly increases the hydraulic stiffness such that at 100 percent speed
the rotor is classically "stiff' (Fig. 7.8). Most successful slender shaft pumps have opposed impeller
rotors (see Chap. 5), an arrangement whose center bushing is ideally suited to developing the high
Lomakin effect necessary for success. Large shaft rotors, on the other hand, are designed for static
deflection less than the available internal clearance. The dry first bending critical speed is therefore
124 Shafts and Shaft Sleeves

QH
= USGPM(FT)
N RPM
5 10 50 100 500 1,000
100

50
1,000

500

10 250

~&~. -
L3/D4

l~ ,
5.0 100
(in· 1 )
..........<.\"-4,0 ~.s'
~ 'ISfiftJ
~ ......
50

1.0 25

0.5
10

0.1
0.1 0.5 1.0 5.0 10 50 100

QH = M3/HR {M}
N RPM

Fig. 7.7 Typical values of cantilever shaft flexibility factor UID" versus pump size factor.

significantly higher. There is also an increase in critical speed as the pump is brought up to speed, but
by design, it is small and the rotor typically remains classically "flexible" (Fig. 7.8).
Relying on high Lomakin effect to produce acceptable dynamics in slender shaft rotors is not an
unmixed blessing. Given the Lomakin effect's dependence on pressure differential and running clearance
condition, there are two cautionary comments. First, rubbing contact at the running clearances whenever
the pump speed is significantly below design (at startup and shutdown) dictates care in the choice of
materials for these clearances if damage is to be minimized, and even that may not prevent rapid wear
under frequent start/stop operation. Second, the dynamic behavior of such rotors may be quite sensitive
to running clearance conditions. In the worst case, the reduction in critical speed as clearances wear can
be sufficient to allow coincidence between the critical and running speeds. If this happens when the
clearances are worn to the point where damping is also significantly reduced, there is a risk the rotor
will develop severe vibration and further damage the pump.
As with overhung rotors, a factor proportional to static deflection is frequently used to quicldy evaluate
between bearings rotors. Various factors in use, with the most common being
Shafts and Shaft Sleeves 125

LOW MECHANICAL STIFFNESS,


HIGH HYDRAULIC

EFFECT
Cl~
UJ{!) OF WEAR
UJ~
a..Cl
C/)Z 100% RUNNING
...,JUJ ------------
«al SPEED
Q 1--
I-C/)
-a:
a:_
O!:S

HIGH MECHANICAL
STIFFNESS,
LOW HYDRAULIC

100
PERCENTAGE OF RUNNING SPEED

Fig. 7.8 Multistage pump rotor characteristics: variation of critical speed with running speed and running
clearance wear.

where L is the bearing span and D is the nominal shaft diameter. This factor is based on the stiffness
of the shaft alone and neglects the weight of components mounted on it. Such an approach is justified
on the grounds that the shaft provides most of a rotor's mechanical stiffness. The only published guideline
for L4/D 2 is a demarcation for rotor runout and balancing in the 8th edition of APlO-61O: rotors with
L4/D2 above 1.9x109 mm2 (3.0x106 in2) have higher allowable runout and are identified as difficult to
balance for operating speeds above 3,800 rpm. Beyond this, the choice of a limiting value of L4/D2 must
rely on the evaluator's experience.
There is opinion that neglecting the weight of the mounted components distorts the evaluation,
particularly for slender shaft rotors whose mounted components usually weigh more than the shaft, and
therefore have a significant influence on static deflection and dry critical speeds. A useful alternative
factor, one that is proportional to the rotor's first dry bending critical speed, is

where W is rotor weight and D and L are before. A notable advantage of this factor is that some guidelines
for limiting values have been published (Fig. 7.9). These allow a rotor to be classified as "too slender,"
"slender," suitable for "wet operation only," or capable of "running dry." Although a correlation such
as this can never be the last word on a particular design, it is consistent with the authors' experience.
Whether to use a slender or heavy shaft rotor is not a clear choice but depends on many factors. To
provide a starting point, the more significant factors are

• Neither design can tolerate high radial hydraulic forces, so the hydraulic design and selection must be good
or pump life will be short.
126 Shafts and Shaft Sleeves

1000
\ + I" L(mm)~

U'l
c:i
;;t
900
SLENDER SHAFTS
\
,
\
Om (m,f::::::: r.:::: :~ -

Possible Difficulties \
--
Cl
'"...J
800 - In Achieving Rub - -
Free Initial Build :-'\
W(N) -

~ SLENDER SHAFTS
II
:::.::: 700 \ Unduly Affected By Out Of -
Balance and Lack Of Initial
a: Straightness. Possibility Of

\ i\.
w Premature Wear Of Internal
I-
W Clearances
::1! 600 -
«
a:
~

~~
Recommended Design Line -----
For Wet Running Pumps
~
500 -

"
CJ)
CJ)
w
Z
LL
LL
i= 400 I":~
~
r::::::~
CJ)
l-
LL
«
I 300 '" " ...............
CJ)
Recommended Upper Limit For - - - - ~
Pumps With Dry - Running Capability
~~
200

100
r---
-
o 2 3 4 5 6 7 8 9
N = (RPM/100)
MAX OPERATING SPEED

Fig. 7.9 Guidance chart for stiffness of between-bearing rotors [7.1].

• At a given rotative speed, slender shaft pumps require less NPSH and are usually more efficient than large
shaft pumps. In modem designs, the latter advantage derives almost solely from slender shaft pumps having
more stages and consequently a higher specific speed. Hydraulic design refinements have overcome the loss
of efficiency traditionally associated with increasing the shaft size in a given impeller. The higher specific
speed of slender shaft pumps also makes them less expensive as cost varies with the 2nd or 3rd power of
diameter but only linearly with length, making a long, slender pump less expensive than one that is shorter
and larger in diameter.
• The service life (time between need to open the casing and renew internal clearances) varies markedly with
the application. Table 7.1 shows typical service lives for the two broad classes of multistage pump rotors in
various applications.
Shafts and Shaft Sleeves 127

Table 7.1 Typical Multistage Pump Service Lives in Various Applications

Typical Service Life-Hours


- - - _.- - - - - --- - - -
- - - - -- - -- - - - -

Application Slender Shaft Large Shaft

Continuous operation at essentially constant flow 50,000-75,000 150,000-200,000


pumping benign liquid
Frequent start/stop operation with wide flow swings, 7,500-10,000 30,000-40,000
pumping liquid of low lubricity or containing low
concentration of abrasive solids
Frequent start/stop operation with wide flow swings, Normally not used 7,500-10,000
pumping liquid with high concentration of
abrasive solids

ROTOR CONSTRUCTION
A centrifugal pump's rotor comprises the shaft plus impeller or impellers, balancing device if applicable,
sleeves, and some retention devices, usually nuts. The design and mounting of these parts is of consequence
to the rotor's integrity and sometimes even influences shaft sizing.
Mounting and retention of the impeller or impellers is determined by power, speed, and impeller
material. The most common arrangement is to center the impeller with a cylindrical slide fit, drive it
with a key, and locate it axially with a shoulder and nut or snap rings (Fig. 7.1). Small low-cost pumps,
some chemical pumps, and most slurry pumps have the impeller threaded onto the shaft and locked
against a shoulder (Fig. 7.10). Although the centering is not as accurate as a cylinderical slide fit, this

Fig. 7.10 Thread mounted impeller.


128 Shafts and Shaft Sleeves

arrangement does offer simplicity; low cost; and, for slurry pumps, a practicable means of mounting
hard metal, rubber-lined, or ceramic impellers. At high rotative speeds, dynamic balance requirements
dictate more accurate centering than can easily be achieved with slide fits. Taper mounting (single stage
only) or shrink fits are used to achieve this. Good practice requires that shrink fit multistage rotors have
the fits stepped to ease assembly. Most taper- and shrink-fit-mounted impellers have a key for torque
transmission. When the interference in the fit is high, the key serves as a backup device only, to prevent
rotation should the fit loosen during a thermal transient or similar event. At high powers or in severe
services, it is important that the impeller is not able to move on the shaft or its mounting fit could suffer
fretting corrosion. Medium-size pumps typically employ a tapered fit in these circumstances, large pumps
use a bolted flange (Fig. 7.11), in which the impeller is centered on a rabbet fit and the coupling is held
tight by fitted bolts.
Most rotors have renewable shaft sleeves to protect the shaft, locate parts on it, or both (see discussion
later in this chapter). There are, however, circumstances where sleeves are not desirable. For reasons of
manufacture and installation, shaft sleeves have a certain minimum thickness, typically 2.5mm (0.10
in.) for shafts to about lOOmm (4 in.) diameter. In small pumps, adding a minimum thickness sleeve to
a shaft sized for adequate stiffness usually results in an abnormally large diameter at the shaft seal. The
alternative is to make the shaft of a material able to offer reasonable wear resistance or hard-coat the
shaft with chrome or ceramic in the region of the seal. Figure 7.12 shows such a shaft, often referred
to as "solid" shaft construction. Solid shafts cost less to make than a shaft plus a sleeve. Taking account
of this and the simpler maintenance procedure, solid shafts are often the most cost-effective solution for
small pumps.
In multistage pumps, particularly those of large shaft design, the minimum thickness of a sleeve
would aggravate the effect of shaft size on impeller proportions. The prevailing practice, therefore, is
to make the shaft of a material able to resist erosion and corrosion of the exposed regions between
the impellers.

Fig. 7.11 Section of large vertical pump-turbine.


Note flange mounted impeller. tilting pad guide bearing. rigid line shaft coupling. (Courtesy: Voith)
Shafts and Shaft Sleeves 129

Fig. 7.12 Section of small centrifugal pump without shaft sleeves.

How the impeller and other rotor parts are mounted affects the rotor's mechanical stiffness. Parts
mounted with a shrink fit increase the effective diameter, hence the stiffness, in the region of the fit. In
most designs, however, the total length of shrink fits is a fairly small fraction of the rotor's bearing
span, so the overall increase in stiffness is not great. Slide fit parts held in compression with nuts, such
as the rotor in Fig. 7.1, often increase the stiffness significantly, although with notable dependence on
the tightness of the nuts. In multistage rotors, where such construction has many faces butting together,
the benefit in stiffness is usually offset by extreme difficulty in building and maintaining a straight rotor.
Most good specifications now require individually located impellers for multistage rotors to avoid this
vexing problem. Slide fit parts retained at one end, such as the hook-type shaft sleeve prevalent in
chemical and overhung process pumps, do not increase stiffness.

SHAFT MAINTENANCE

Except for small pumps with solid shafts, it is unusual to have to replace a centrifugal pump shaft.
Typically the shaft will last the life of the pump, unless damaged as a consequence of some other failure
or problem within the pump. Given this, any shaft failure should be investigated carefully to find the
cause, so it can be corrected and further failures prevented.
Although the shaft is nominally a lifetime component, it should be carefully cleaned and inspected
at each pump overhaul, and where necessary and feasible, repaired to ensure it is suitable for further
service. The important points in this process are

1. Check for straightness, supporting the shaft on rollers at its bearing journals (next to the journals when they
are chrome plated). Avoid checking for straightness with the shaft between centers; the centers are often
130 Shafts and Shaft Sleeves

damaged during assembly and dismantling. The allowable shaft runout depends on the pump's design and
service and should be specified in the manufacturer's manual. As a general rule, bent shafts should be
replaced rather than attempting to straighten them. The usual straightening techniques (e.g., pressing, peening,
hot spotting) rely on local residual stresses. During operation, vibration tends to relieve the residual stresses,
thus allowing the shaft to assume its original bent shape. Compounding this, some of the techniques leave
sufficient residual stress to become the site of subsequent fatigue failure. Shafting for vertical turbine pumps
is the one exception to the general "no straightening" rule. This long, slender shafting is routinely and
successfully straightened by repeated pressing and subsequent stress relief. The procedure appears simple,
but is best carried out by a shop experienced in doing it.
2. Inspect the shaft, paying attention to at least the following aspects:
• Coupling and antifriction bearing fits for fretting and scoring
• Impeller and sleeve fits for corrosion, erosion (from leakage), and scoring
• Wetted regions for corrosion and erosion (from liquid impingement)
• Keyways for wear, distortion, and corrosion
• Threads for corrosion and damage (e.g., torn or cracked threads)
• Fillets for damage (tool marks, scratches) and cracking.
Inspection for cracks can be visual, using some magnification, or by the various non-destructive evaluation
(NOE) methods (liquid penetrant, magnetic particle, ultrasonic).
3. Take measurements to determine the exact extent of any damage found during the inspection. Whether to
replace or repair a shaft depends on the nature and extent of the damage, the availability of a new shaft,
and the available repair processes. Refinishing to remove scoring, scratching, and minor corrosion is acceptable
provided it does not leave the shaft undersize at a critical location. Machining down and building up by
plating (chrome or nickel) or spraying (flame or plasma) are viable processes to repair local corrosion,
scoring, and wear, provided they are carried out correctly. The major risks are weakening of the shaft,
leading to fatigue failure and spalling of the plating or coating in areas of rubbing contact or heavy press
fits. Welding is generally not a suitable repair technique. This is not to say shafts are not welded, those for
large pumps sometimes are, it just recognizes that the materials used for most centrifugal pump shafts are
not readily welded or require postweld heat treatment, a process that would distort the shaft. Because welding
is not usually viable, cracks in a shaft are cause for its replacement.
4. If the shaft has been repaired, it should be checked again for straightness and the repaired areas carefully
inspected. It is well to note the location and extent of repair and the final inspection results in the pump's
maintenance file so there is some history should a problem develop later.

SHAFT SLEEVES

Pump shafts are usually protected from erosion, corrosion, and wear at stuffing boxes, running clearances,
internal bearings, and in the waterways by renewable sleeves. The most common shaft sleeve function
is that of protecting the shaft from wear at a stuffing box. Thus shaft sleeves serving other functions
are given specific names to indicate their purpose. For example, a shaft sleeve used between two
multistage pump impellers in conjunction with the interstage bushing to form an interstage running
clearance is called an interstage or distance sleeve.
In medium-size centrifugal pumps with two external bearings on opposite sides of the casing (the
common double-suction and multistage varieties), the favored shaft sleeve construction uses an external
shaft nut to hold the sleeve in axial position against the impeller hub. Sleeve rotation is prevented by a
key, usually an extension of the impeller key (Fig. 7.13). If the axial thrust exceeds the frictional grip
of the impeller on the shaft, it is transmitted through the sleeve to the external shaft nut.
In larger high-head pumps, a high axial load on the sleeve is possible, and a design like that in Fig.
7.14 may be favored. This design has the commercial advantages of simplicity and low replacement
Shafts and Shaft Sleeves 131

IMPELLER
KEY -=::;"-_j~

IMPELLER SLEEvE SHAF NUT

Fig. 7.13 Sleeve with external lock nut and impeller key.

IMPELLER SHAFT GLAND

SHAFT NUT

IMPELLER NUT SLEEVE SET SCREW

Fig. 7.14 Sleeve with internal impeller nut, external shaft sleeve nut, and separate key.
132 Shafts and Shaft Sleeves

IMPELLER SHAFT GLAND

SLEEVE

Fig. 7.15 Sleeve threaded onto shaft with no external lock nut.

cost. Some manufacturers favor the sleeve shown in Fig. 7.15, in which the impeller end of the sleeve
is threaded and screwed to a matching thread on the shaft. A key cannot be used with this type of sleeve
and right- and left-hand threads are substituted so that the frictional grip of the packing on the sleeve
will tighten it against the impeller hub. In the sleeve designs shown in Figs. 7.13 and 7.14, right-hand
threads are usually used for all shaft nuts because keys prevent the sleeve from rotating. As a safety
precaution, the external shaft nuts (Figs. 7.13 and 7.14) and the sleeve itself (Fig. 7.15) use set screws
for a locking device.
In pumps with overhung impellers, various types of sleeves are used. Often, stuffing boxes are placed
close to the impeller, and the sleeve actually protects the impeller hub from wear (Fig. 7.16). As a
portion of the sleeve in this design fits directly on the shaft, the impeller key can be used to drive the
shaft sleeve. Part of the sleeve is clamped between the impeller and a shaft shoulder to maintain its
axial position.

Shaft sleeve seals


Older designs relied on a metal-to-metal joint between the sleeve and impeller hub (see Fig. 7.13)
to prevent leakage under the sleeve. For pumps handling water at moderate pressures with a packed box
shaft seal, in which a little leakage did not matter, this simple arrangement performed well. Most modem
designs are intended to be applied over a wide range of services, with a variety of shaft seals, thus have
to employ a means of ensuring there is no leakage under the sleeve. A common arrangement for sleeves
on between bearings rotors is shown in Fig. 7.17. Hook-type shaft sleeves (Fig. 7.18), are retained at
one end and sealed with a gasket or "0" ring. Retention at one end reduces the risk of bending the shaft
and allows for differential thermal expansion. In services where there is a risk of scaling or carbonizing
under the sleeve if the pumped liquid gets into the fit, making sleeve removal difficult, it is better to
seal the inner end of the sleeve. This is an inherent feature with hook-type sleeves on overhung rotors.
When sleeves are used to protect a shaft against corrosion, it is crucial to ensure the shaft does not
Shafts and Shaft Sleeves 133

SHAFT STUFFING BOX

GL 0 SLEEVE IMPELLER

Fig. 7.16 Sleeve for pumps with overhung-impeller hubs extending into stuffing box.

RUBBER "0" RING

SLEEVE SHAFT NUT

Fig. 7.17 Shaft sleeve seal to prevent leakage along shaft.

get wet, or it could suffer crevice or concentration cell corrosion. Many users feel the added maintenance
procedures necessary to ensure this are time consuming and risky, so they insist on shafts that are able
to survive without protection.

Material for shaft seal sleeves


Most pumps are now equipped with mechanical seals. Many of these seals are assembled and mounted
as a cartridge, of which the shaft sleeve is an integral part. See Chap. 9 for details of the shaft sleeves
used with mechanical seals. The rest of this section, including Maintenance, deals with shaft sleeves for
packed box shaft seals, and has some relevance to breakdown-type seals (Chap. 10).
134 Shafts and Shaft Sleeves

Fig. 7.18 Hook-type sleeve for pump with overhung impeller.

Stuffing box shaft sleeves are surrounded in the stuffing box by packing; the sleeve must be smooth
so that it can tum without generating too much friction and heat. Thus the sleeve materials must be
capable of taking a very fine finish, preferably a polish. Cast iron is therefore not suitable. A hard bronze
is generally used for pumps handling clear water, but chrome or other stainless steels are sometimes
preferred. For services subject to grit, hardened chrome or other stainless steels give good results. For
more severe conditions, Stellite coated sleeves are often used and occasionally sleeves that are chromium
plated at the packing area. Sleeves made entirely of a hardened chrome steel are usually the most
economical and satisfactory.

SHAFf SLEEVE MAINTENANCE

Shaft sleeves are usually the fastest wearing pump part and the one most frequently requiring replacement.
Once sleeves are worn appreciably, the packing cannot be adjusted to prevent excessive leakage. As a
matter of fact, excessively worn sleeves frequently tear and score any new packing as soon as it is
inserted. Thus, sleeves frequently require repair or replacement when no other pump overhaul is necessary.
Sleeves of single-stage and low-head multi-stage pumps can be removed quite easily. As the long
sleeves sometimes used in high-pressure multistage pumps may be harder to remove, they are often
fabricated with external grooves so that a sleeve puller can be used (Fig. 7.19). In a design that uses an
impeller nut between the sleeve and the impeller (see Fig. 7.14), a tight sleeve can often be loosened
by backing off the impeller nut.
Shaft sleeves are occasionally reconditioned by welding or metal spraying and final grinding. This proce-
dure is not recommended for a pump on severe service, or if existing facilities for the final grinding are
inadequate. It is necessary to assure both concentricity of grinding and the perpendicularity of the sleeve
radial faces to the sleeve bores. Concentricity should be double checked after reassembly on the rotor.
Although it may be easier to pack a pump with brand new shaft sleeves, the sleeves do not have to
Shafts and Shaft Sleeves 135

Fig. 7.19 Shaft sleeve puller.

be replaced each time new packing is installed. The degree of permissible sleeve wear grooving depends
on the type of grooving. Usually, the sleeve surface is highly polished under the packing action, and
the grooving is undulated rather than composed of sharp separate grooves under each individual packing
ring. Sometimes, slight grinding of these worn sleeves is permissible to permit reuse if the pump service
is not too severe. The controlling factors are the availability of the necessary tools, shop facilities, and
trained shop personnel. Restored sleeves must have a good, smooth surface, and the refinished parts
should neither be run-out nor distorted. If the facilities are available, it may be advisable to try regrinding
and reusing one set of worn sleeves to establish the practicability of this procedure.
The shaft sleeve OD should not be reduced to a point at which excessive clearance at the bottom of
the stuffing box permits any packing to be squeezed inside the pump when the glands are tightened. As
a rule, sleeves for packed boxes should not be ground down more than 0.65 to 0.75mm (25 to 30
thousandths) on the diameter and should be given a 0.40 micron (l6-microinch) finish.
Worn sleeves, however, are ordinarily replaced rather than reconditioned. Hammering to expand or
crack the material will facilitate their removal, but extra care should be taken to prevent shaft damage.

BIBLIOGRAPHY

[7.1] A. B. Duncan and 1. F. Hood, The Application of Recent Pump Developments to the Needs of the Offshore
ai/Industry, Proc. of the Conference on Pumps and Compressors for Offshore Oil and Gas, London, UK,
June 29-July 1, 1976, pp. 7-24.
8
Stuffing Boxes

INTRODUCTION

Strictly, the term "stuffing box" is related to the packed box shaft seal, long the principal device used
to seal a pump's casing where the shaft passes through it. Although the number and variety of devices
used for this purpose has increased significantly as processes and the pumps used have evolved, the
term stuffing box is still widely understood to mean the shaft seal, regardless of the type. For this reason,
retaining the existing title of this chapter uses a familiar term to introduce a very important subject.
The sealing of a pump is a crucial aspect of its design. Although it is not always obvious, even a
relatively minor deterioration in sealing performance can render a pump inoperable. Sensitivity to sealing
performance has increased with environmental concerns, with many services now being the subject of
legislation limiting allowable leakage. Even where specific leakage limits do not exist, there are often
implicit limits in general environmental legislation.
All centrifugal pumps are sealed, either to keep the pumped liquid in the pump or, if the internal
pressure adjacent to the seal is below atmospheric, to keep the atmosphere out of the pump. Most seals
are some form of dynamic liquid or gas device where the shaft passes through the casing. All these
seals leak to some degree, although in many designs, the rate is so low the leakage is not obvious.
Generally, the leakage passes to atmosphere. In those services where leakage to atmosphere cannot be
tolerated or a dynamic liquid seal is difficult to realize, a so-called "sealless" pump is used. Such pumps
are not actually sealless; they are either arranged so seal leakage is returned to the suction vessel or
they are hermetically sealed.
The types of shaft sealing devices and sealing arrangements commonly used in modern centrifugal
pumps are listed next. For quick reference, the chapter in this book dealing with the particular type or
arrangement is also listed.

Seal Type or Arrangement Chapter


Packed box (soft packing) 8
Fixed packing 8
Hydrodynamic 8

136

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
Stuffing Boxes 137

Axial face (mechanical) 9


Breakdown bushing 10
Sealless; leakage to suction 14
Sealless; hermetically sealed 24

Sealless pumps involve special construction of either the pump or its drive arrangement, and are therefore
covered in the chapters dealing with the appropriate construction.

SEAL SELECTION

The various sealing devices and arrangements used in centrifugal pumps have different capabilities.
Selecting the most appropriate type of arrangement therefore depends on an accurate specification of
the conditions prevailing at the seal. All too frequently, this fundamental step is overlooked, only to
become the reason for a badly behaved or inoperable seal once a pump is put into service.
A comprehensive specification of sealing conditions, often referred to as seal environment, requires
the following:

1. Pressure at the seal, taking account of the pump's internal arrangement, and including any range in pressure
likely from operating conditions or wear at the pump's running clearances
2. Temperature at the seal, noting whether cooling or heating is being applied to the seal, and if so the estimated
temperature drop or rise
3. Liquid being sealed, its properties (pH, SG, viscosity, specific heat) if not common knowledge, its condition
(presence of solids or contaminants), and any unusual characteristics (e.g., tendency to crystallize, precipi-
tate solids)
4. Surface speed at the sealing interface
5. Rotor displacements, radial and axial, likely under expected operating conditions
6. Operating mode (Le., intermittent, continuous, frequent stop/start)
7. Allowable leakage rate.

Note that the only reference to size is the surface speed at the seal interface. This is intentional. Given
the importance of sealing, it is better if the appropriate device or arrangement determines the space
required rather than having the space available limit the seal design. To a degree, some of the more
modern pump specifications now recognize this by specifying minimum space provisions for certain
types of seals.
For the more difficult applications, selecting the sealing device or arrangement is an iterative process,
involving repeated checks of sealing feasibility against the prevailing sealing conditions or environment.
Figure 8.1 illustrates the process. Unless the pump selection or basic construction are changed, the only
characteristics of the sealing conditions that can be changed in this iterative process are pressure at the
seal, temperature at the seal, and liquid condition. Details of how these can be changed vary, to some
extent, with the seal type, and thus are addressed in the specific seal sections.
Selecting the appropriate sealing device or arrangement is a complex task. In a general sense, it is
possible to narrow the choices by first considering allowable leakage, then the liquid and service conditions
(see Figs. 8.2 and 8.3). Once the choice of seal types is narrowed, refinement of the selection depends
on the specific conditions, available seal capabilities and personal preference. Sealing technology is, of
course, continually evolving, so any selection guide is only approximate.
138 Stuffing Boxes

ACCURATELY DETERMINE
SEAL ENVIRONMENT

IS SEALING
FEASIBLE?
NO

CHANGE SEAL
ENVIRONMENT

FINAL SEAL
SELECTION

Fig. 8.1 Seal selection-fundamental.

Moderate-Soft Packing (packed box)

Allowable Fixed Packing


Leakage Minimum - [
Axial Face (mechanical)

None
i Hydrodynamic* (1)

Breakdown (to pump suction)

Axial Face (double) (2)

Hermetically Sealed (isolated from atmosphere)

(1) *Zero Leakage when running

(2) Barrier Fluid Leakage to atmosphere

Fig. 8.2 Seal type versus allowable leakage.


Stuffing Boxes 139

Water at Low Temp.


Intermittent Operation - - Soft Packing

Water, Continuous Axial Face


-{
Operation Soft Packing

Hydrocarbon (1) - { Axial Face


Continuous Operation Hermetically Sealed

Liquid
and Liquid Difficult to Seal
Service - { Breakdown
Suction at
Atmospheric Pressure Hydrodynamic
Dry Running Possible

Liquid Difficult to Seal


Toxic, Isolation from - - Hermetically Sealed
Atmosphere Required
Hydrodynamic
{ Soft Packing
High Solids Content
Fixed Packing
Axial Face

(1) Seal Selection depends on allowable leakage

Fig. 8.3 Seal type for liquid and service conditions.

PACKED BOX SEAL

In its elementary fonn, a packed box shaft seal consists of a number of rings of soft, compressible
packing in a cylindrical recess, generally known as a "stuffing box," around the pump shaft sleeve (Figs.
8.4 and 8.5). The packing is compressed axially by an adjustable gland to produce a close radial clearance
with the shaft, thereby minimizing leakage through the seal. To prevent extrusion of the packing into
the pump, it is necessary to have a close clearance at the bottom of the stuffing box. This clearance can
be an integral part of the casing (Fig. 7.16), a separate bottom ring (Fig. 8.5), or a separate throat bushing
(Fig. 8.4). The last arrangement is used as a manufacturing aid (it allows a larger boring bar) or when
it is considered necessary to be able to restore the close clearance (a legacy from pumps whose shafts
were too slender). Separate throat bushings are generally one piece and when the casing is axially split
are located with a tongue and groove fit.
To function correctly, a packed box seal must leak. The purpose of the leakage is twofold: First, it
lowers the friction between the shaft or sleeve and the packing, thus minimizing the heat generated.
Second, it helps remove the heat that is generated. See "Stuffing box maintenance" for instructions and
cautionary remarks on packing adjustment.
140 Stuffing Boxes

SLEEVE STUFFING 80X SEAL CAGE GLAND


THROAT BUSHING

Fig. 8.4 Conventional stuffing box with throat bushing.

SEALING LIQUID

BOTTOMING RING PACKING

Fig. 8.5 Conventional stuffing box with bottoming ring.


Stuffing Boxes 141

m/sec
o 10 20 30
100
1000

10 iu
III
100 i;
CD
en
II)

en
a.. ...u
II)
0

«
~
:::l
II)
10 II)
~
a..

0.1

01234567
Relative Velocity @ Seal-1000 FtiMin

Fig. 8.6 Packed box seal capabilities (after Durametallic).

Figure 8.6 shows the general upper limit of packed box seal capability in terms of pressure versus
surface speed at the seal interface. It is possible to exceed these limits, but unless the design, manufacture,
installation, and operation of the seal are all carried out very carefully, the seal is likely to be a continual
source of operating difficulty.

SEAL CAGES

Most pumps are designed to have the region adjacent to their shaft seal at or very close to suction
pressure. When a pump so designed operates with negative suction head, the inner end of the stuffing
box is under vacuum, and air tends to leak into the pump. For this type of service, packing is usually
separated into two sections by a lantern ring or seal cage (Fig. 8.4). Water or some other sealing fluid
is introduced under pressure into the space, causing flow of sealing fluid in both axial directions. This
construction is useful for pumps handling flammable or chemically active and dangerous liquids since
it prevents outflow of the pumped liquid. To ensure adequate lubrication of the packing, it is also
advisable to use a seal cage when the pressure at the seal is 0.3 bar gauge (5 psig) or lower. Seal cages
are usually axially split for ease of installation and removal (Fig. 8.7).
Some installations involve variable suction conditions, the pump operating part time with head on
suction and part time with suction lift. When the operating pressure inside the pump exceeds the
atmospheric pressure, the liquid seal cage becomes inoperative (except for lubrication). However, it is
maintained in service so that when the pump is primed at starting, all air can be excluded.
142 Stuffing Boxes

Fig. 8.7 Lantern ring or seal cage.

Sealing liquid arrangements


When a pump handles clean, cool water, sealing liquid is usually taken from the pump discharge, or,
in multistage pumps, from an intermediate stage. An independent supply of sealing water should be
provided if any of the following conditions exist:

1. A suction lift in excess of 4.5 m (15 ft)


2. A discharge pressure under 0.7 bar or 7.0 m (10 psig or 23-ft)
3. Hot water (over 120°C [250°F]) being handled without adequate cooling (except for boiler feed pumps, in
which seal cages are not used)
4. Muddy, sandy, or gritty water being handled
5. For all hotwell pumps
6. The liquid being handled is other than water-such as acid, juice, molasses, or sticky liquids-without
special provision in the stuffing box design for the nature of the liquid.

When sealing water is taken from the pump discharge, an external connection may be made through
small diameter piping (Fig. 8.8) or internal passages. In some pumps, these connections are arranged so
that a sealing liquid can be introduced into the packing space through an internal drilled passage either
from the pump casing or an external source (Fig. 8.9). When the pumped liquid is used for sealing, the
external connection is plugged. If sealing liquid from an external source is required, it is connected at
the external pipe tap and the internal connection is plugged.
Because the leakage flow from a correctly adjusted packed box is relatively low, seals with sealing
liquid supplied from the pump discharge (Figs. 8.8 and 8.9) are normally subject to a pressure close to
pump discharge. For pressures up to 1.4-1.7 bar (20-25 psig) this is quite acceptable. At higher pressures,
it is advisable to use only external piping (Fig. 8.8) and to carry out most the adjustment of seal leakage
using the needle valve provided. Not taking this precaution generally results in operating difficulty with
the seal, the difficulty increasing with pump discharge pressure.
Stuffing Boxes 143

Fig. 8.8 Piping connections from the pump discharge.

Fig. 8.9 Arrangements for sealing liquid supply in an end-suction pump.


A. Internal seal. B. External seal.
144 Stuffing Boxes

Radial P1
Clearance

Throat
Bushing

Fig. 8.10 Packed box arrangment with throat bushing and seal cage for pump handling gritty or dirty liquid.

Abrasive solids in the pumped liquid will cause accelerated seal wear (packing rings and sleeve) if
the liquid passes through the seal. To prevent this, pumps handling abrasive laden liquids and sealed
with a packed box often have a throat bushing and seal cage at the bottom of the stuffing box (Fig.
S.10). Clean liquid, compatible with the pumped liquid, is brought from an external source and injected
into the seal. The injection flow required is that to produce a liquid velocity through the throat bushing
clearance of 3-4.6 m/s (10-15 ftls). Experience shows velocities on this order are necessary to ensure
there is minimal mixing of pumped liquid and injection flow within the seal. The drawback of an injected
throat bushing is dilution of the pumped liquid, often a concern in mineral processing and petroleum
refining. When dilution is a concern, the choice is a seal cage located farther into the packing set (Fig.
S.4) and a higher rate of seal wear, or a different type of shaft seal (consult Fig. S.3).
In some services, pumps handle liquids containing low concentrations of fine solids, not sufficient to
erode the casing or impeller but enough to shorten sleeve and packing life if used as sealing liquid. Fine
filters can be used to remove the solids. The arrangement is complex, however, because filters tend to
plug, and therefore have to be monitored for pressure drop and arranged in duplex if on line cleaning
is required. This difficulty can be overcome in many instances by using cyclone separator to remove
the fine solids. Figure S.11 shows the flow diagram for such an arrangement. See Chapter 9 for a detailed
description of cyclone separators and the precautions necessary in their installation and use.
If clean, cool water is not available (as with drainage and irrigation pumps) or cannot be connected
directly to the pump (as with sewage pumps), grease or oil seals are often used. Most pumps for sewage
service have a single stuffing box subject to discharge pressure and are located below the liquid in the
suction reservoir. It is therefore not necessary to seal these pumps against air leakage, but forcing grease
into the sealing space and packing helps to exclude grit.
An automatic oil sealer that exerts discharge pressure in a cylinder on one side of a plunger, with
oil or light grease on the other side, is available for sewage service. The oil or grease line is connected
to the seal connection, which is near discharge pressure. As the inner end of the stuffing box would be
at about SO percent discharge pressure, there is a slow flow of grease or oil into the pump when the
unit is in operation. No flow takes place when the pump is out of service.
Some pumps handle water in which there are small, even microscopic, solids. Using water of this
kind as a sealing liquid introduces the solids into the leakage path, shortening the life of the packing
and sleeves. It is sometimes possible to remove these solids by installing small pressure filters in the
sealing water piping from the volute to the stuffing box.
Stuffing Boxes 145

Clean liquid
to seal cage
connection

Dirty liquid
to pump
suction

Fig. 8.11 Cyclone separator for packed box seal.

WATER-COOLED STUFFING BOXES

High temperatures or pressures or both complicate the operation of packed box seals by raising the
temperature of the leakage. Should the temperature rise to the point that the leakage flashes within the
packing set as the pressure drops (Fig. 8.12; from Ref. [8.1]), damage to the packing and the sleeve will
quickly render the seal inoperable. Direct injection with cool liquid is the most efficient means of
controlling seal temperature, but is not always acceptable for reasons of pumped liquid contamination
(e.g., boiler feed) or the availability of liquid at the correct condition. Therefore usual practice is to
provide pumps for the more difficult services with jacketed, water-cooled stuffing boxes. The cooling
water removes heat from the liquid leaking through the stuffing box and heat generated by friction in
the box, thus improving packing service conditions. In some special cases, oil or gasoline may be used
in the cooling jackets instead of water. Two water-cooled stuffing box designs are available. The first
(Fig. 8.13) provides cored passages in the casing casting. These passages, which surround the stuffing
box, are arranged with in-and-out connections. The second type uses a separate cooling chamber combined
with the stuffing box proper, the whole assembly being inserted into and bolted to the pump casing (Fig.
8.14). The choice between the two is based on manufacturing preferences.

Pressure and temperature conditions


With a more thorough understanding of the interrelation of stuffing box pressures, rubbing speeds,
and leakage temperatures, improved water-cooled stuffing boxes have been built for temperatures up to
200°C (400°F), and stuffing box pressures up to 35 bar (500 psig), without pressure-reducing breakdowns
or labyrinths. This type of stuffing box is illustrated in Fig. 8.15. For greatest cooling efficiency, the
temperature difference between the cooling liquid and the leakage through the box must be kept to a
maximum at all points. In this design (Fig. 8.15), the cooling water is introduced nearer to the outside
of the stuffing box. Before moving axially toward the interior of the pump, the cooling water is circulated
completely around that portion of the stuffing box which surrounds the packing. A cored passage is
146 Stuffing Boxes

Outer Ih ell cooliq

- ---- Frictional bell Dot removed - - Frictional beal removed

Precoolin. sec:1ion Sb.ftleal

Fig. 8.12 Pressure and temperature profile through packed box.


(Courtesy Sulzer)

provided from this annular chamber toward the interior of the pump. The cooling water then circulates
in a secondary annular chamber extending inside the pump beyond the packing. This allows precooling
of stuffing box leakage before it reaches the packing. The cooling water then escapes through a second
cored passage to the cooling chamber exit. In this design, the coldest cooling water is adjacent to the
coldest leakage. Having picked up some heat, the cooling water flows into the pump at a higher temperature
and cools a higher temperature leakage. One such unit has operated continuously for over a year without
renewing the packing, under stuffing box operating conditions of 22 bar (325 psig) and 200°C (400 0 P).
The shaft diameter at the stuffing box was 100 mm (4 in), and the operating speed 3,600 rpm.
Stuffing box pressure and temperature limitations vary with the pump type, because it is generally
not economical to use expensive stuffing box construction for infrequent high-temperature or high-
Stuffing Boxes 147

QUENCHING UOIJID

STUFFING BOX BUSHING COOLING WATER INLET

Fig. 8.13 Jacketed stuffing box with cored cooling passage cast in casing.

,
SECTION A-A
LE"MOF' TO LOWE~ ~1t[SSUIt[

COO"' ... W"TlEIt OUTLlT t


!

SlUFFING

C~I'" W"TlEIt CMAMIER

Fig. 8.14 Separate jacketed stuffing box assembly with pressure-reducing stuffing box bushing.
148 Stuffing Boxes

COOLING WATER
OUTLET

CORED PASSAGE
INTO PRECOOLING
ANNULAR SPACE

PRECOOLING
OF LEAKAGE

"f-E~ - --- ---~I- ~


INTERIOR

ANNULAR
COOLING
AREA
AROUND
PACKING

Fig. 8.1S Special water-cooled stuffing box for high pressures and high temperatures.

pressure applications. Therefore, whenever the manufacturer's stuffing box limitations for a given pump
are exceeded, the only solution is the application of pressure-reducing devices ahead of the stuffing box.

PRESSURE·REDUCING DEVICES

Essentially, pressure-reducing devices consist of a bushing or meshing labyrinth, ending in a relief


chamber located between the pump interior and the stuffing box. The relief chamber is connected to
some suitable low-pressure point in the installation, and the leakage past the pressure-reducing device
is returned to this point. The only drawback to application of these devices is the necessity of bleeding
a part of the effective pump capacity back to a lower pressure level, and the resultant reduction in
installation efficiency. If the pumped liquid must be salvaged, as with treated feed-water, it is returned
back into the pumping cycle. If the liquid can be wasted, the relief chamber can be connected to a drain.
There are many different pressure reducing device designs. Figure 8.14 illustrates a design for limited
pressures. A short, serrated stuffing box bushing is inserted at the bottom of the stuffing box, followed
by a relief chamber. The leakage past the serrated bushing is bled off to a low-pressure point.
With relatively high-pressure units, intermeshing labyrinths may be located following the balancing
device and ahead of the stuffing box (Fig. 8.16). Piping from the chamber following pressure-reducing
Stuffing Boxes 149

BALANCING DEVICE
LEAKOFF TO SUCTION

LEAKOFF TO
LAST LOWER PRESSURE
IMPEL

Fig. 8.16 Labyrinth-type pressure reducing bushing on the discharge side of a pump equipped
with a balancing device.

devices should be amply sized so that as wear increases leakage, piping friction will not increase stuffing
box pressure.
Mining and mineral processing occasionally require slurry pumps arranged in series to develop heads
beyond the capability of a single pump. When the pumps are next to each other, one of the major
problems is sealing the latter pumps against leakage of high pressure slurry. One successful approach
is to provide a clear liquid breakdown bushing (Fig. 8.17) ahead of a conventional packed box. High-
pressure clear liquid, generally water, is provided by a small reciprocating pump.
150 Stuffing Boxes

SEAL WATER
INLET

I
Pi +
SEAL WATER ILAND
I OUTLET
Pi P21.LOW PRESSURE)

RADIAL
CLEARANCE

PRESSURE BREAKDOWN
IUSHING

Fig. 8.17 Breakdown bushing injected with clear liquid for pumps handling gritty or dirty liquid.

STUFFING BOX PACKING

Basically, stuffing box packing is a pressure breakdown device. The packing must be somewhat plastic
so that it can be adjusted for proper operation. It must also absorb energy without failing or damaging
the rotating shaft or shaft sleeve. In a breakdown of this nature, friction energy is liberated. This generates
heat that must be dissipated in the fluid leaking past the breakdown or by means of cooling water
jacketing or both.
There are numerous stuffing box packing materials, each adapted to some particular class of service.
Until recently asbestos, lubricated with graphite or inert oil, was one of the principal materials. Now
that asbestos is a proven carcinogen, it is no longer used to any significant extent as a packing material.
Although there is no one alternative material possessed of all the properties of asbestos, research has
produced materials able to outperform asbestos on specific services. These developments have led to
the simple packed box being again considered a viable shaft seal for modern pumps. Some of the
principal packing materials now in use are described in the following summary and their typical properties
set out in Table 8.1. Note that the pressure and PV limitations in Table 8.1 are below those shown in
Figure 8.6. The lower values reflect a conservative approach by the packing manufacturers, the likely
intention being to avoid those applications where success is so dependent upon the pump operator's skill.

1. Plastic-a mixture of short strand synthetic fiber and lubricant, the latter usually graphite or mica and oil
or grease. Occasionally metal particles are added to act as a lubricant reservoir. Used alone, plastic packing
produces a low-leakage seal for moderate pressures and high temperatures. At higher pressures, the allowable
temperature decreases and back up rings of braided or metallic packing are recommended to avoid extrusion.
Chemical resistance is limited.
2. Synthetic fibers-a wide variety of materials has been developed in the pursuit of non-asbestos packings.
Those currently in general use are Teflon®, aramid, and graphite. Teflon (PTFE) is an attractive material
because it is inert in most pumped liquids. Developing a suitable packing, however, has been hampered by
the material's low thermal conductivity and high coefficient of thermal expansion. PTFE yam treated with
colloidal PTFE as a lubricant and a silicon compound for improved heat dissipation is a viable packing for
high pressures, moderate to high temperatures and moderate surface speeds. Flitney [8.2] reports expanded
PTFE with 50 percent encapsulated graphite further improves thermal conductivity and allows surface speeds
on the order of 23 mls (4,500 ft/min). Aramid fiber is serviceable to moderate temperatures but has less
chemical resistance than PTFE. The virtue of aramid fiber is high strength and abrasion resistance, making
it a packing suitable for pumps handling slurries or solids laden liquids. The packing itself tends to be
Stuffing Boxes 151

Table 8.1 Properties of Common Packing Materials

Pressure PV Temperature
Material bar (psig) (Note 1) °C (OF) pH Notes

Plastic 7.0(100) 70(190xI0 3) 315(600) 4.8


17.5(250) 165(470) 65(150) 4.8
PTFE yarn 17.5(250) 165(470) 260(500) 0-14 2
Aramid fiber 10.5(150) 116(380) 65(150) 3-10 3
Graphite filament 17.5(250) 165(470) 400(750) 0-14
Graphite foil 17.5(250) 165(470) 400(750) 0-14
Lead foil 17.5(250) 165(470) 230(450) 2-10
Aluminum or copper foil 17.5(250) 165(470) 400(750) 3-10
Notes
I. PV factor based on (m/s) x (bar) left/min) x (psi)].
2. Maximum surface speed 9.1 m/s (1,800 ft/min).
3. Maximum surface speed 6.6 m/s (1,900 ft/min); performance lowest of published values.
4. Values of pressure, PV, and temperature are maximum.

abrasive, and so is limited to moderate surface speeds and must run on a very hard shaft sleeve (Brinell
hardness 650 minimum, where BHN is the Brinell hardness number) shaft sleeve. Graphite or carbon filament
packings have wide chemical resistance, high temperature capability, and high pressure/speed capacity. The
material is expensive, however, and is therefore normally used only for severe chemical or high-temperature
service.
All the synthetic fiber packings are interbraid or interlace construction. This affords high flexibility and
enables the packing to remain intact even when individual strands wear through at the working surface.
3. Graphite foil-known generically as "exfoliated graphite" or more commonly as Graphol ®, graphite foil
realizes high thermal conductivity, which allows for a seal with very low leakage rates, nominally 50 percent
of that for asbestos packing. Chemical resistance and service capabilities are the same as for graphite filament,
the only notable difference being better thermal conductivity, therefore the capability of running with slightly
less leakage.
4. Metallic-composed of flexible metallic strands or foil with graphite or oil lubricant impregnation and with
either a synthetic or plastic core. The impregnation makes this packing self-lubricating for its start-up period.
The foils are made of babbitt or lead, copper, and aluminum. Babbitt or lead is used on water and oil service
for low and medium temperatures (up to 230°C or 450°F) and medium to high pressures. Copper is used
for medium to high temperatures and pressures with water and low-sulfur-content oils. Aluminum is used
mainly on oil service for medium to high temperatures and pressures. Babbitt or lead packing is not suitable
for operation on brass or bronze shaft sleeves. Cooper and aluminum packings require a shaft sleeve whose
Brinell hardness is at least 500.

Packing is supplied either in continuous coils of square cross section or in prefonned die-molded
rings. Plastic packing is sometimes supplied in bulk or in cartridges for injection by gun. Graphite foil
packing is produced as ribbon, but for pump packings, it is usually supplied as split die-fonned rings.
When coil-type packing (Figs. 8.18 and 8.19) is used, it is cut in lengths that make up individual rings.
The ends are cut with a diagonal, or scarf joint, and with a slight clearance to provide for expansion
and avoid buckling. The rings have a tendency to swell from the liquid action and the rise in temperature.
The scarf joint allows the ends to slide and laterally absorb expansion.
It is preferable, where possible, to use die-molded packing rings (Fig. 8.20), which are available to
152 Stuffing Boxes

Fig. 8.18 Synthetic fiber packing in continuous coil form.


(Courtesy John Crane Co.)

Fig. 8.19 Metallic packing in spiral form.


(Courtesy John Crane Co.)

exact size and in sets. A molded ring insures an exact fit to the shaft or shaft sleeve and to the stuffing
box bore and also establishes equal packing density throughout the stuffing box.
Hard packing such as metallic is not as resilient as fiber or plastic and is, therefore, more sensitive
to changes in operating conditions. One common means of improving seal performance under these
conditions is to use a combination of hard and soft packing, for example metallic and fiber or plastic
(Fig. 8.21). Combination packing sets always have the end rings of hard material to minimize extrusion.
Such sets are available in standard die-formed ring combinations from most packing manufacturers.
The pressure drop across a packed box seal is not linear. Figure 8.12 shows a five-ring arrangement
in which more than 50 percent of the pressure drop occurs across the outer ring, while the inner ring
Stuffing Boxes 153

Fig. 8.20 Metallic packing in ring fonn.


(Courtesy John Crane Co.)

Fig. 8.21 Combination of hard and soft packing.


(Courtesy John Crane Co.)
154 Stuffing Boxes

has essentially zero pressure drop. From the meager data available it seems this characteristic increases
with the number of rings, so there is little to be gained from using more than four or five rings of packing
for even the most severe services. In some instances, low pressure, for example, extra rings can cause
difficulties by limiting leakage through the packing.
Packing size typically ranges from 10 to 30 percent of the shaft or sleeve diameter. The minimum
practicable size is generally considered 6 mm (0.25 in.) square. Some industry standards, however,
dictate a larger minimum, API-610 [3.1] for instance, requiring 10 mm (0.38 in.) square minimum in
refinery pumps.
Packed box seal performance is influenced significantly by the shaft or sleeve surface, both hardness
and finish, and any cyclic radial motion of the surface during pump operation. Considering sealing
surface hardness first, good practice based on past experience suggests the following: For sealed pressures
below 2.5 bar (35 psig) conventional packing running on a soft sleeve (bronze or equal hardness) will
give good service. Raising the pressure to 5.2 bar (75 psig) requires a harder sleeve, the usual material
being 13 chrome steel (Brinell hardness 300-500 depending on grade and heat treatment). Conventional
packing will give reasonable service but the best results will be realized with metal foil and plastic or
graphite. At pressures above 5.2 bar (75 psig), the packing should be either metal foil and plastic or
graphite, and the sleeve fully hardened 13 chrome (Brinell hardness 450 minimum) or hard coated. Usual
hard coating materials are ColmonoyTM 6 or ceramic, either chrome oxide or tungsten carbide. Ceramic
is harder than ColmonoyTM but has lower resistance to thermal shock. In services involving abrasives,
both the nature of the liquid and the type of packing (aramid fiber or similar) dictate ceramic coated
sleeves. In all cases the sealing surface should be finished to Ra 0.8~m (32 ~in) or better.
Shaft or sleeve runout and vibration produce cyclic radial motion of the sealing surface. The packing
is not able to follow this motion so it runs with a larger clearance, hence more leakage. Flitney [8.2]
reports a linear increase in leakage with sealing surface runout. Good practice requires that sealing
surface runout not exceed 0.075 mm (0.003 in.) total indicated runout (TIR). Alignment of the shaft or
sleeve within the stuffing box is not as critical as runout and vibration because the packing is flexible
enough to accommodate small static eccentricities on the order of 0.25 mm (0.010 in). Care is necessary,
however, in high-pressure service where the risk of extrusion is higher.

STUFFING BOX GLANDS

Stuffing box glands may assume several forms, but basically they can be classified into two groups:

1. Solid glands (Fig. 8.22).


2. Split glands (Fig. 8.23).

Split glands are made in halves so that they may be removed from the shaft without dismantling the
pump, thus providing more working space when the stuffing boxes are being repacked. Split glands are
desirable for pumps that have to be repacked frequently, especially if the space between the box and
the bearing is restricted. The two halves are generally held together by bolts (Fig. 8.23), although other
methods are also used. Split glands are generally a construction refinement rather than a necessity, and
they are rarely used in smaller pumps. They are commonly furnished for large single-stage pumps, for
some multistage pumps, and for refinery pumps. Another common refinement is the use of swing bolts
in stuffing box glands. Such bolts may be swung to the side, out of the way, when the stuffing box is
being repacked.
Stuffing box leakage into the atmosphere might, in some services, seriously inconvenience or even
endanger the operating personnel-for example, when such liquids as hydrocarbons are being pumped
Stuffing Boxes ISS

Fig. 8.22 Solid stuffing box gland .

Fig. 8.23 Split stuffing box gland.



at vaporizing temperatures or temperatures above their flash point. As this leakage cannot always be
cooled sufficiently by a water-cooled stuffing box, smothering glands are used (see Fig. 8.13). Provision
is made in the gland itself to introduce a liquid-either water or another hydrocarbon at low temperature-
that mixes intimately with the leakage, lowering its temperature, or, if the liquid is volatile, absorbing it.
Stuffing box glands are usually made of bronze, although cast iron or steel may be used for all iron-
fitted pumps. Iron or steel glands are generally bushed with a nonsparking material like bronze in refinery
156 Stuffing Boxes

service to prevent the ignition of flammable vapors by the glands sparking against a ferrous metal shaft
or sleeve.

STUFFING BOX MAINTENANCE

Stuffing box maintenance primarily consists of packing replacement. Although this sounds simple, it
must be done correctly, or pump operation will not be satisfactory. The following procedure should be
followed in repacking a stuffing box:

1. Never try to add one or two rings to the old packing. This is false economy. Remove the old packing
completely, using a packing puller, if available, and thoroughly clean the box. Inspect the sleeve to make
sure it is in acceptable condition. Putting new packing in a box against a rough or badly worn sleeve will
not give satisfactory service. .
2. Be sure that the new packing is a proper type for the liquid, operating pressure, and temperature. Unless
the packing comes preformed in sets, make sure that each ring is cut square on a mandrel of correct size.
3. Insert each ring of packing separately, pushing it squarely into the box and firmly seating it, using split
pusher rings of proper length, fitting the box nicely. Successive rings of packing should be rotated so the
joints are 120 or 180 deg apart.
4. When a seal cage is used, make sure to install it between the proper two packing rings so it will correctly
handle the sealing liquid supply when the box is fully packed and adjusted.
5. After all the required packing rings have been inserted, install the gland and firmly tighten the gland nuts.
Make sure that the gland enters the stuffing box squarely and without cocking, so the full periphery of the
packing is under uniform pressure.
6. After the first tightening of the gland, back off the nuts until they are only finger tight. Start the pump with
the gland loose, so that there will be excessive initial leakage. Tighten up slightly and evenly on the gland
nuts, at 15 or 20 min intervals, so the leakage is reduced to normal after several hours.

Normal seal leakage is not easily defined. In principle, the leakage necessary for a particular seal
depends on the amount of heat generated and how that can be dissipated. The heat generated is, in tum,
a function of sealed pressure, surface speed at the seal, packing type, and sleeve condition. Data on the
relative influence of these factors are meager and often conflicting. As a practical guide, Table 8.2 shows
the range of usual leakage rates for various sealed pressures. These data are for conventional packing
and typical conditions. With carbon or graphite packing (self-lubricating) and well-finished sleeves
running with minimum vibration, lower leakage rates are attainable.
Take care when adjusting the packing. Unless enough liquid leaks through the seal to remove the
heat generated, the packing will be burned and the sleeve scored. (On pumps with quenching glands,
stop the supply of quenching water at intervals, and observe actual leakage through the box, otherwise,
visual inspection cannot distinguish between leakage past the packing and the quenching liquid supply.)
Repacking and adjusting stuffing boxes should only be done by experienced personnel. Others assigned

Table 8.2 Packing Leakage Rates

Sealed pressure bar (psig) Leakage rate liter/hr (gal/hr)

4.0 (60) 0.2-6.8 (0.06-1.8)


4.0 to 7.0 (60 to 100) 0.8-11 (0.20-3.0)
7.0 to 17 (100 to 250) 1.9-28 (0.50-7.4)
Stuffing Boxes 157

this work should be cautioned against putting too much pressure on the gland. It should be made clear
that excessive leakage is not as hannful as too little.
Packing removed from a stuffing box being repacked, should be examined in order to obtain as much
information as possible on the cause of packing wear. Often, correctable operating conditions or inadequate
packing procedures are revealed by this examination. Some of the more frequently encountered symptoms
are the following:

1. Excessive wear on rings nearest to the gland, while the bottom rings remain in good condition, is caused
by overtightening of the packing in one adjustment or by not inserting rings one at a time, and pushing each
home before inserting the following ring.
2. Charring or glazing of the inner circumference of the rings is caused by excessive heating, insufficient
lubrication, or inadequate packing material for the pressure and temperature conditions.
3. Wear on the outer circumference of the rings occurs when they rotate within the stuffing box.
4. Heavy packing ring wear on one selective portion of the inner circumference may be caused by excessively
worn bearings or eccentric rotor operation.
5. If some rings are cut too short or shrink excessively, the adjacent rings will bulge and be extruded into the
open space.

FIXED PACKING

Conventional packed box seals using soft packing suffer from two distinct disadvantages. First, the
packing requires periodic adjustment to maintain acceptable leakage rates. Second, in pumps handling
abrasive laden liquids, the flush water necessary to realize reasonable seal life is a source of pumped
liquid dilution. Quite often, pumped liquid dilution is the greater of the two disadvantages, particularly
as energy costs increase.
Responding to both these problems, several manufacturers have developed fixed packing type seals;
arrangements designed for zero adjustment and minimum pumped liquid dilution. Three recent develop-
ments in the mining industry, reported by Pearse [8.4], are typical of the arrangements used:

1. Hydrostatic gland-two feather-edged rubber seals are arranged with a seal cage between them (Fig. 8.24).
Water at a pressure above that inside the pump at the seal is injected into the region between the lip seals.
For typical slurry pumps, the injection pressure is approximately 75 percent of pump discharge pressure.
Leakage into the pump is reportedly on the order of 0.2 m3/hr (1 gpm).

Water @ 75% Pump Discharge

~
Rubber
Seal = 0.2 m3 /hr (1 gpm) into pump
Ring
J...../
(Wilkinson Rubber Linatex)

Fig. 8.24 Hydrostatic gland (diagrammatic).


158 Stuffing Boxes

Pressure
Actual Reducing Sand
Seal Element Barrier

Fig. 8.25 IHe Liquidyne seal.


(Courtesy [HC Holland)

2. "Simmering" seal-using the same principle as the hydrostatic gland, the Simmering seal employs two
opposed lip seals and has oil injected between them to effect a seal.
3. "Liquidyne" seal:-three sealing elements are retained in a four-piece bolted housing (Fig. 8.25). Flushing
water at a pressure above that inside the pump is applied between the inner and intermediate sealing elements
(sand barrier and pressure reducing elements). A small part of this flush flow passes into the pump, thus
excluding the pumped liquid from the seal. The balance of the flush liquid flows through the pressure-
reducing element, then to an atmospheric drain. The outer seal serves only to contain leakage passing to
the drain.

All three of these designs rely on resilient radial sealing elements and a flush or injection liquid source
at a pressure above that inside the pump. The first two differ from the third in that leakage from the
pressure reducing element passes directly to atmosphere instead of being collected and passed to a drain.

HYDRODYNAMIC SEALS

Although the construction and operation of hydrodynamic seals is quite different from packing, their
frequent use in place of packing for difficult services and their reliance on some form of packing when
the pump is shut down are deemed sufficient reasons to justify their inclusion in this chapter.
Stuffing Boxes 159

A hydrodynamic seal is a device that produces, by hydrodynamic action, a pressure below atmosphere
at the shaft opening. Since the pressure inside the pump is below atmosphere, the seal has to be arranged
to maintain a stable liquid/air interface while the pump is operating.
Because the pump configuration suits it, hydrodynamic seals are generally applied to single suction,
overhung impellers. Figure 8.26 shows an arrangement employing a "sealing impeller," or "expeller"
as it more commonly known. Operation of the seal can be determined from the pressure profiles through
the pump (see Fig. 8.27). At the impeller periphery there exists a static pressure rise (less than the total
head-see Chap. 4) across the impeller. Pump-out vanes on the back shroud of the impeller (see Chap.
4) produce a pressure drop down the back shroud (PPOy) to yield a lower pressure Ph at the impeller hub.
The plain side of the expeller causes a small increase in pressure P L resulting in a slightly higher pressure
P, at the expeller periphery. Provided the pump is operating within the seal's capability, the vaned side
of the expeller lowers the static pressure to atmosphere part way down the vanes at rio Unless unusual
proportions are resorted to (e.g., expeller or pump-out vanes at larger diameter than the impeller vanes),
expeller-sealed pumps are limited to suction pressures on the order of 15 percent of the pump's differen-
tial pressure.
Some early hydrodynamic seals were developed for high-speed pumps, there being at the time no
other suitable sealing device. More recently, most applications are for pumps handling abrasive laden

Fig. 8.26 Expeller seal.


160 Stuffing Boxes

Ppov-
I-
constant
XT.D.H.

1-
..... ,

r
Pg

!
't'

f
'i
--1-_
SHAFT EXPELLER IMPELLER
CENTERLINE

Fig. 8.27 Pressure profiles in pump with expeller seal.

or other difficult to seal liquids. The expeller has few parts and no close running clearances, and is
therefore well suited to such applications.
Beyond their tolerance of solids laden and other difficult liquids, hydrodynamic seals are often justified
on the basis of not having to provide a flush water system (difficult and expensive for a remote, arid
site), and the elimination of pumped liquid dilution.
Hydrodynamic seals have zero leakage while the pump is running. But when the pump is shut down
some form of auxiliary seal is necessary to prevent leakage if the suction pressure is above atmosphere.
Most designs use some form of packed seal, either soft packing grease lubricated or fixed packing such
as a lip seal. More sophisticated arrangements include pneumatic bushings (inflated once the pump is
shutdown) and centrifugally activated closures.
Being a pumping device, hydrodynamic seals absorb power, therefore their inclusion in a pump is
not a free means of effecting a seal. In most cases, however, the cost of additional power for the seal
probably equals that of a flush water system, so it is not significant in the total cost of the plant, and
may even be a saving in some cases. The one problem the seal power can cause is vaporization of
volatile liquids, In such cases, it is necessary to ensure a small circulation of liquid from the seal back
to the pump proper, so the heat generated is dissipated.

BIBLIOGRAPHY

[8.1] Schoffler, W. & Florjancic, D., Packed Stuffing Boxes in Pumps Circulating Hot Water, Sulzer Technical
Review, 3/1981, pp 99-103.
[8.2] Flitney, R. K., Soft Packings, Tribology International, vol. 19, #4, August '86, pp 181-183.
[8.3] Pearse, G., Pumps for the Mineral Industry, Mining Magazine, V152, #4, April '85, pp 299, 301, 303, 305,
307-309,311,313.
9
Mechanical Seals
- ~---~~------

Mechanical seals are widely considered superior to the other forms of shaft seals employed in centrifugal
pumps, particularly packed stuffing boxes and breakdown-type seals. This is generally correct, but there
are applications for which mechanical seals represent "too much technology," and services for which
they are not suitable (see the introduction to Chap. 8).
The general superiority of mechanical seals over packed boxes and breakdown seals derives from
the orientation of their sealing elements. Before proceeding with a detailed treatment of mechanical seals
themselves, it is worthwhile to develop the details of this fundamental difference.
Packed boxes and breakdown seals are classified as radial seals, those that act on the circumference
of the shaft (Fig. 9.1[aD. As such, these seals must be designed to accommodate variations in shaft
dimensions and radial position. Such variations result from manufacturing tolerances or movements
within the pump produced by operating loads or thermal expansion or both. Whether this accommodation
is provided by a resilient material, packing, or a specific clearance between shaft and seal, the seal
inherently has a certain amount of leakage. Compounding that, designs employing packing are limited
in surface speed, sealed pressure, and the nature of the sealed liquid.
Mechanical seals, sometimes referred to as face seals, are arranged so the seal acts against an axial
face (Fig. 9.1 [b D. With the sealing interface now normal to the shaft axis, it no longer has to accommodate
variations in radial dimensions and shaft position. Freed of this need, the sealing interface can now run
with a very close clearance and its design (proportions and materials) has only to consider wear resistance.
The advantages offered by mechanical seals are:
1. Greater sealing capability (defined as pressure containment with tolerable leakage and service life). With
appropriate face geometry and materials, acceptable leakage and seal life can be realized at surface speeds
on the order of SOmis (10,000 ft/min) and sealed pressures on the order of 140 bar (2,000 psi).
2. Lower leakage. Because the sealing interface can run with a very close clearance, the leakage from a
mechanical seal is significantly less than from an equivalent packed box seal. In most cases, the leakage is
so little it evaporates on coming into contact with the atmosphere and is therefore not obvious to an observer.
It is important, however, to recognize that there is some leakage from all mechanical seals, a fact of
consequence to plant safety and the more stringent pollution limitations coming out of environmental consider-
ations.

161

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
162 Mechanical Seals

Pa
---'----+-

a) Radial seal acts on b) Axial seal acts on plane surface to


cylindrical surface to produce pressure drop in radial direction.
produce pressure drop
in axial direction.

Fig. 9.1 Fundamental seal arrangements for rotating shafts.

3. Tolerance of liquids. Many of the liquids handled by pumps in modern refining, petrochemical, and chemical
processing have little or no lubricity and act as solvents to the lubricants used in most packings. To seal
these liquids with a packed box requires the introduction of a compatible seal oil into the packing, a
requirement that complicates the installation and causes dilution or contamination of the pumped liquid.
4. No adjustment. Mechanical seals are self-compensating for wear, therefore they do not have to be adjusted
over the course of their service life. In a large, complex plant this confers a substantial advantage to overall
pump maintenance costs.

The advantages of mechanical seals have led to their playing a fundamental role in the evolution and
development of processes employed by mankind to raise the standard of living. It is true to say that
many of the processes now taken for granted would not have been possible without mechanical seals
for the pumps used in them. Despite this contribution to modem technology, the mechanical seal is not
a new device. Mayer [9.1] reports the use of mechanical seals "about 1900" for difficult services, and
Fig. 9.2 shows a design patented in 1915. As is so often the case with mechanical devices, the 1915
seal contains features that are now considered in the forefront of design.
Although they were obviously known early in pump history, mechanical seals remained a relative
novelty until general service conditions started to really tax the capability of packed box seals. To quite
a degree, this was entirely rational; the packed box is a simple seal, and there is no real advantage to
using more technology than is necessary for a particular duty. Where a mechanical seal was necessary,
however, the legacy of the packed box unduly influenced seal design for a long time, a circumstance
that impaired mechanical seal performance and has only recently been overcome.

PRINCIPLES OF MECHANICAL SEALS

Although there are a number of mechanical seal arrangements and a great many detailed design variations,
all incorporate the same fundamental elements. The sliding seal interface is effected between the fiat,
Mechanical Seals 163

Fig.9.2 Mechanical seal patented in 1915.

polished mating faces of two rings, one connected and sealed to the pump's rotor, the other to its casing.
To accommodate manufacturing tolerances, minor axial movement of the pump's rotor, and wear of the
seal faces, one of the rings is flexibly mounted and provided with a means of ensuring the sealing faces
remain in contact.
In much the same manner as sleeve bearings, mechanical seals depend on some form of lubrication
between the seal faces to lower the coefficient of friction and help remove the heat generated. Without
this lubrication, the faces will come into intimate contact and wear rapidly. The presence and quality
of face lubrication depends on a number of interrelated factors, namely: sealed pressure, nominal surface
speed, mechanical distortion of the seal faces, seal face materials, heat generated at the faces, heat
dissipation, proximity of sealed liquid to its boiling point, and thermal distortion of the seal faces.
Resolving all these issues to produce a reliable seal is the seal designer's problem. There is merit,
however, in briefly reviewing the principles behind these factors.
All mechanical seals derive their face lubrication from leakage of the sealed fluid, be it a liquid or
a gas between the seal faces. Seal leakage, and therefore the quality of face lubrication, is influenced
significantly by the sealed pressure and the characteristic of that pressure drop across the seal faces. The
pressure drop characteristic depends on face orientation in the running seal. Figure 9.3 shows the three
basic pressure distributions and the face orientation associated with each. In general, seal capability and
leakage rate both increase as the pressure drop characteristic moves from concave to convex (or as seal
face orientation changes from divergent to convergent in the direction of leakage flow).
For a mechanical seal to function, the forces tending to close its faces must exceed those tending to
open the faces. This net closing force gives rise to what is termed face loading. The significance of face
loading is that it has an upper limit, largely dependent on face materials and the sealed liquid conditions,
beyond which the lubricating film between the faces breaks down. From the simple seal shown in Fig.
9.4, the net force tending to close the faces is the resultant of the hydrostatic closing force (FH), the
hydrostatic opening force (F0), the flexible element compression force (F d, and the dynamic gasket
friction force (F/). Neglecting flexible element seal friction for the time being (see Design), the least
certain of these forces is the hydrostatic opening force. The usual assumption is that the pressure drop
164 Mechanical Seals

70

60
System ~ Flat Seal Faces
Pressure, 50

~)
Percent
40

30

20 H
10 L Low Pressure Side

00 10 20
Face Width, Percent

Fig. 9.3 Seal face orientation (exaggerated) versus pressure drop profile. (Courtesy Durametallic Corp.)

-Face load (+ ve to left)

~n~
~------~I-.--· Ff Fe

Face load = FH - Fa + Fe - Ff

FH = PSAH
Fa = PA F
P= Mean
effective

rL pressure
between
seal faces
Fe = Load from
compression
device
Ff = Friction
force@
dynamic
gasket
(if used)

Fig. 9.4 Force balance at mechanical seal faces.


Mechanical Seals 165

across the seal faces is linear, giving a mean effective pressure pi equal to half the pressure drop. If,
however, the pressure drop is concave, a result of the faces being divergent (Fig. 9.3), the opening force
will be lower and the net face loading higher, possibly to the extent that the face load capacity is
exceeded. At the other extreme, a convex pressure drop, the opening force will be higher and the net
face loading lower. To a certain extent this is desirable because it increases seal capacity, although at
the expense of higher leakage; but if the face loading is too low, the seal can become unstable with the
risk that its faces may "blow apart" and allow gross leakage.
As already noted, the face loading capacity of a seal is affected by the condition of the sealed liquid.
In its passage through the seal surface, the leakage flow of sealed liquid undergoes a pressure drop from
sealed pressure to atmospheric and a temperature rise, the latter from seal friction and the pressure drop.
When the sealed liquid is already close to its vapor pressure, the seal leakage may vaporize as it passes
through the seal interface. If vaporization occurs only a short distance across the seal face (Fig. 9.S[a]),
the liquid film may not be sufficient to support the face loading, a condition that will cause the seal to
wear rapidly. Three approaches are used to try to avoid this difficulty.
The classical solution is to avoid any tendency to vaporize by ensuring the temperature of the leakage
is below its atmospheric boiling point (Fig. 9.5 [c]). When this approach is feasible and its expense can
be accommodated, it represents an excellent solution. In this connection, it is of particular importance
for sealing hot water, where by reducing the temperature of the water at the seal to 70°C (160°F), good
sealing performance can be achieved with quite straightforward seal designs.
It is not always feasible to cool the liquid below its atmospheric boiling point, and in other cases it
is not deemed viable. In these circumstances, the approach taken is to raise the pressure at the seal or
cool the liquid or both to ensure vaporization occurs closer to the face ID (Fig. 9.S[b]), thus providing
nearly full face load capacity. Two criteria are used to determine the conditions necessary to ensure this
behavior. The pressure at the seal is required to be above the liquid vapor pressure by some margin
(e.g. 3.S bar or SO psi minimum per API-682 [9.2]). Alternatively, the temperature of the liquid at the
seal must be below its boiling point, at the sealed pressure, by some margin whose value is dependent
on the sealing conditions and the seal face materials (Fig. 9.6).
Careful examination of Fig. 9.S reveals an apparent paradox: vaporization between the seal faces
raises the mean effective pressure acting between them, thus raising the opening force, Fo, and lowering
the net face loading. This has been confirmed by testing but the difficulty is that while the net face
loading is decreasing so is the extent of the liquid film and thereby the face load capacity. Seal operation
in this region is therefore dependent on a delicate balance between net face loading and load capacity
and is not a desirable state, particularly for a single component liquid such as water.
When the sealed liquid cannot be cooled sufficiently, the pump cannot be rearranged to raise the
pressure at the seal, or the sealing conditions exceed the capability of these two approaches, the third
approach is to modify the characteristic of the pressure drop across the seal faces. Two means are
employed to do this: shaped grooves in the rotating face to promote hydrodynamic or convergent face
orientation for a convex pressure drop (Fig. 9.3). These are dealt with in more detail in the following
discussion of face lubrication technology.
Because the performance of mechanical seals has not always been as good as expected, particularly
in severe services, there has been a great deal of research into the precise nature of seal face lubrication.
The present state of the art identifies three classes of face lubrication: boundary, mixed film, and full
film. At the same time, there now exist computer routines able to predict the class of face lubrication
likely to prevail in a given seal whose operating conditions are known. Figure 9.7 shows the general
limits of seal capability for these three classes of face lubrication.
Most mechanical seals run with boundary lubrication, and as Fig. 9.7 shows, this class of face
lubrication provides broad capabilities. Boundary lubrication is largely a product of a favorable pressure
166 Mechanical Seals

Ps
-I

E7fvaJ~t~n--F
I. Pv
-
Face load
capacity
LuqUid vapor
pressure @ TL

a) Vaporization close to seal face 0.0.

Ps,Ts
Ps
I- -I

P
Ts

-t~{'
Path
.1TL I. Pv_ Vaporization

b) Vaporization close to seal face 1.0. Sealed pressure, Ps, typically equivalent to
margin of 14°C (2S0F) over Ts; varies with liquid, face materials, & sealed pressure.

I--_l~l' V
Ps,Ts Ps

Path AT~I I-Pv


D
T L < atmospheric boiling point

c) No vaporization across seal face.

Fig. 9.5 Effect of vaporization across seal faces.


Mechanical Seals 167

Product Water

30 \Silicon
\ carbide
\
25

20

\
Alumina \ Vapour pressure
-~ ,,
ceramic \

\,
curve (water)

,,
10
,, \
,
\
5 \

o 50 100 150 200 250


Temperatu ref" C

Fig. 9.6 Variation in required" (temperature margin below boiling point) with sealing conditions
"~T
and seal face materials.
Operating limits for alumina ceramic, tungsten carbide, and silicon carbide inserted balanced seals
(stationary face-carbon): size 80 mm, speed 4,000 rpm. (Courtesy Flexibox)

drop (from sealed pressure to atmospheric), which forces liquid between the faces. Once the liquid is
between the faces, minor surface irregularities and face porosity serve to develop boundary lubrica-
tion conditions.
For conditions beyond the capability of boundary-lubricated seals, mixed-film lubrication offers a
worthwhile improvement in capability (Fig. 9.7). All seals designed for gases (sometimes used in pumps
as backup seals or for sealing cyrogenic liquids in the vapor phase) rely on mixed-film lubrication. The
distinguishing feature of mixed-film seals is the incorporation of specially shaped grooves in one of the
faces (Fig. 9.8), whose function is to develop mixed-film lubrication by hydrodynamic action. Most
designs have the face grooves at the pressure side of the face (usually the 00), and therefore act on the
sealed fluid to develop face lubrication. The difficulties with all mixed-film seals are performance
prediction (much of the design methodology is still largely empirical) and the accumulation of debris
in the grooves, which can result in unexpected wear.
168 Mechanical Seals

0, 6, c- PACKED BOX APPLICATIONS


mls
10 20 30 40 so 60 70
soo

200
•MIXED FILM LUBRICAnON 100

~ ~ooo 0_0_.,\. \ BREAKDOWN BUSHING


so

~ 1---"" . OR 20
: \ \ . MECHANICAL SEAL-
10 a:
i 100 I::-----T--lr-----·~" ~
'\ .
'"
~ s

~a *BOUNDARY~.\ .
.\
2

f LUBRICATION 1.0
10~-----------I--------------~\-.--~~~----~
O.S

\ \
10~-L--~~~-L~~--~~-~~--~~~~~~
\. 0\ 0.2

0.1
3 4 5 6 7 8 9 10 11 12 13 14 15
RELATIVE VELOCITY @ SEAL-1,OOO FT/MIN

Fig. 9.7 Seal capabilities (After Durametallic).

Full-film lubrication offers the highest sealing capability, and because the faces are separated by a
full-fluid film, such seals also offer long service lives. Two means are employed to develop full-film
lubrication. The older approach is to inject lubricating fluid directly between the faces (Fig. 9.9). This
arrangement is generally known as a hydrostatic seal. With the advent of computer analysis (finite
element methods), it is now possible to design a seal so that at operating conditions it runs with its/aces
convergent in the direction of leakage. From the earlier discussion of seal face orientation, this arrangement
allows the pressure gradient to force liquid between the faces thus ensuring full face load capacity. At
the same time, the convex face pressure distribution (Fig. 9.3) lowers the face loading. Leakage from
both designs of full-tilm-Iubricated seals is higher than from boundary- and mixed-film designs, but is
still lower than from alternative seal types. Hydrostatic seals have the disadvantage of relying on a
separate lubricating fluid system. Convergent face seals avoid this, but their design must take account
of the actual range of operating conditions, and their manufacture and installation must accurately
reproduce the design configuration.
Mechanical Seals 169

1. Lubrication Flats
2. Lubrication Slots
3. Spiral Grooves, Unidirectional
4. Lubrication Slots
5. Lubrication Grooves

Fig. 9.8 Various seal face groove fonns used to develop mixed-film face lubrication.
(Courtesy Durametallic Corp.)

HYDRAULIC BALANCE

In the discussion of face loading, the hydrostatic force tending to close the seal faces was identified as
one of the major components in that force balance. The magnitude of this force under a given pressure
depends on the seal's hydraulic balance. Two terms are in general use to classify this aspect of seal
design: unbalanced or balanced.
The seal in Fig. 9.1O(a) has the sealed pressure PSI acting over an annulus AH , bounded by the OD
of the face contact region and the ID of the flexible element gasket. As such, the area producing the
hydrostatic closing force is equal to or greater than AF , the area producing the opening force. A seal
with this ratio of effective hydrostatic force areas is termed unbalanced. Recognizing that the opening
hydrostatic force is produced by the average pressure acting over the face contact area AF , it is evident
the face loading in unbalanced seals increases significantly as the sealed pressure rises. This characteristic
limits the pressures and liquids that can be successfully sealed with unbalanced seals. Many users seek
to avoid the complexity of assessing whether an unbalanced seal will work in a particular service, and
instead specify a conservative upper limit on sealed pressure, or specify that all pusher-type seals are
to be balanced, as does API-682.
170 Mechanical Seals

Fig. 9.9 Axial face seal with hydrostatic injection.


(Courtesy Voith)

(8)

Fe
Fe-E::V
~

(b)

Fe

L BALANCE
DIAMETER
lL!::~~~ AH
AF < 1.0

Fig.9.10 Mechanical seal hydraulic balance-a) unbalanced, b) balanced.


Mechanical Seals 171

PUMPED
LIQUID ATMOSPHERIC
SIDE SIDE

Fig. 9.11 Internal mechanical seal.

As the sealed pressure rises beyond the capability of unbalanced seals, it is necessary to lower the
hydrostatic closing force to keep the seal face loading within the capacity of the seal faces. This is done
by making the hydrostatic closing force area less than that for the opening force (Fig. 9.1O(b». Such a
seal is termed balanced. The actual degree of balance is defined by the ratio of hydrostatic force areas,
AH/AF' and typically ranges from 0.6 to 0.9 depending primarily on service conditions and face materials.
Balance ratios below 0.6 offer lower face loading, hence longer seal life, but at the expense of higher
leakage and lower seal stability (the risk of a minor fluctuation in operating conditions causing the seal
to "blow open").

SEAL ARRANGEMENTS

There are two aspects to seal arrangements. The first deals with fundamental arrangements of the basic
parts and the second with various arrangements of complete seals.
Fundamental arrangements of the basic parts considers the location of the rotating element. Internal
seals (Fig. 9.11) have their rotating element inside the seal housing and the sealed liquid at the OD of
the seal interface, therefore leakage is normally inward toward the shaft. Centrifugal action in the liquid
tends to oppose leakage across the seal interface. External seals (Fig. 9.12) are the opposite; the rotating
element is outside the seal housing and the sealed liquid is at the ID of the seal interface. Leakage is
therefore normally outward away from the shaft, with centrifugal action in the liquid tending to promote
leakage, and to promote the centrifuging of any solids in the liquid into the sealing interface.

ROTATING ELEMENT

ATMOSPHERIC
SIDE

t SHAFT

Fig. 9.12 External mechanical seal.


172 Mechanical Seals

PUMPED ATMOSPHERIC
LIQUID SIDE
SIDE

\::i SHAFT

Fig. 9.13 Double mechanical seal.

Most centrifugal pumps equipped with mechanical seals employ a single seal, either intemal or
external. With this arrangement there is direct leakage of the sealed liquid to the atmosphere. This is
not always tolerable, for reasons of seal face life, liquid toxicity, liquid cost, or environmental pollution.
One solution is to use multiple seals, either double or tandem depending on the particular circumstance.
When the service is such that the pumped liquid cannot be allowed to leak into the atmosphere (toxic
or polluting) or must be kept out of the seal interface (abrasive solids in the liquid), a double seal offers
one solution (Fig. 9.13). The seal is made up of two single seals mounted either back to back (Fig. 9.14)
or face to face. The region between the seals is maintained at a pressure greater than the sealed liquid
pressure with a barrier fluid. By doing this, the sealed liquid is excluded from the seal, since the leakage
flow at the inner seal is from the barrier fluid to the sealed liquid. Until recently, the barrier fluid was
always a liquid. With a liquid, the heat generated at the seal faces is often high enough to require that
the liquid be cooled by circulation through some form of heat exchanger. For a long time, a double seal
with a liquid barrier was the only available solution for toxic, polluting, or abrasive pumped liquids.
This arrangement is complex, however, and somewhat unreliable as a consequence, a circumstance that
has lead to the development of better solutions for most services. Drawing from centrifugal compressor
technology, double seals using an inert gas as the barrier fluid (Fig. 9.14) have simplified the total sealing
system considerably. Such designs are now used quite extensively in pumps handling toxic and polluting
liquids, limited only by compatibility of the barrier gas with the pumped liquid. Similarly, research into
the behavior of seals in abrasive liquids has precipitated the development of specially designed single
seals for these applications; see Seal environment later in this chapter.
The objective of a tandem seal is to provide a backup so failure of the main or primary seal does
not result in gross leakage and the need to immediately shut down the pump. In services such as
hydrocarbon or crude oil, safety and pollution requirements alone can warrant tandem seals to prevent
even short-term gross leakage. A tandem seal has two or more single seals arranged in series, with each
successive seal serving to backup the preceding seal. The most common arrangement has two seals (Fig.
9.15), the inner acting as the primary, the outer as the secondary. Under normal conditions, the primary
seal contains the pumped liquid, whereas the secondary seal runs at a pressure below the sealed pressure,
usually atmospheric, in an inert buffer fluid. Leakage beyond a certain rate causes a pressure rise in the
secondary seal lubrication system, a condition that is used to close off the lubrication system vent, and
hence contain the primary seal leakage and sound an alarm.
For services where a high pressure must be sealed with the utmost reliability, staged mechanical seals
Mechanical Seals 173

Fig. 9.14 Section, non-contacting double seal with inert gas barrier.
(Courtesy John Crane, Inc.)

are sometimes used. The seals are arranged in tandem, but unlike in the tandem seal, the total pressure
drop is shared approximately equally between the seals. This is achieved by bleeding a small flow
through a throttling orifice in each seal housing (Fig. 9.16), and returning the flow to a point of lower
pressure in the system. The primary use for staged mechanical seals has been in nuclear reactor primary
coolant circulating pumps (see Chap. 26).

Buffer
Seal Liquid
Flush (circulated)
I I
Pumped Atmospheric
Liquid Side
Side

Shaft Rotating Elements


ct- - - - - - - - - - -

Fig. 9.15 Tandem mechanical seals.


174 Mechanical Seals

If Req'd:
Buffer
Bleed-off to
Liquid P1 > P2 > P3
at P > P1 point at P = P3

~ Os

Pumped Atmospheric
Liquid Side
Side

Fig. 9.16 Staged mechanical seals.

SEAL DESIGN

Most who review this section on seal design will either be seeking to apply a seal or determine why a
particular seal is misbehaving. A mechanical seal cannot be looked at in isolation; a pump depends on
a seal to prevent leakage, but the design and ultimate performance of the seal is influenced significantly
by the design of the pump itself.
It is better if the pump design is looked at first, because doing so can save a great deal of wasted
effort and expense. The aspects that need to be investigated and finally quantified are

1. What is the pressure at the seal? Does it change with operating conditions or pump wear or both? The pump
design needs to be studied carefully, and perhaps the manufacturer consulted, to determine the answers to
these questions. Sometimes the pressure at the seal is normally a vacuum. As already noted, mechanical
seals designed for liquids will suffer rapid face wear if run under conditions that do not produce a lubricating
liquid film. Therefore if the pressure is normally a vacuum, it is necessary to either change the pump design
to raise the pressure or resort to a more complicated seal, for example, a double seal.
2. Does rotor deflection under some operating conditions produce significant angular misalignment through
the seal? Mechanical seals, particularly those with rotating flexible elements (see detailed seal design), are
susceptible to premature failure if run with high angular misalignment. Several industry specifications address
this problem by limiting rotor deflection at the seal to 0.050 mm (0.002 in.) under the worst expected
operating conditions.
3. Is there significant axial rotor movement during pump operation? Pumps with hydrodynamic thrust bearings
(Kingsbury type; see Chap. 11) typically have 0.30 to 0.40 mm (0.012- to 0.Q15-in.) rotor endplay. Similarly,
but as a result of manufacturing tolerances rather than bearing clearance, pumps with antifriction type thrust
bearings can have as much as 0.50 mm (0.020-in.) rotor endplay. Most mechanical seals can accommodate
axial movements of this magnitude, provided the movement is slow and relatively infrequent. Usually this
is the case, but under adverse operating conditions (see Chap. 22) pump rotors can shuttle rapidly, a condition
that usually causes excessive seal leakage and premature failure. When this condition is the product of
operating flows alone, meaning that it is not influenced by an extreme hydraulic design, the better solution
is to avoid operating the pump down to such low flows.
4. What is the housing arrangement available for the seal? It is now well recognized that the space for
conventional packing has little radial room for optimum seal geometry and heat dissipation, resulting in
seals that fail frequently and prematurely. Two industry standards, ANSI B73.1[9.3] and API 610, now
Mechanical Seals 175

mandate minimum radial dimensions for mechanical seal housings as one facet of an effort to improve seal
life in chemical and refinery pumps. An extension of this concept is the arrangement generally known as
an open seal housing (Fig. 9.17). The rationale behind this design is to provide as much room as possible
for free circulation of the pumped liquid around the seal, thus allowing optimum heat removal and avoiding
any tendency for the housing geometry to force solids into the improved heat dissipation. Experience in the
chemical- and mineral-processing industries has shown that pumps employing this housing arrangement
have longer seal lives than those with more conventional arrangements.

There are myriad variations in detailed mechanical seal design, far too many to cover in this text.
Compounding that, the design of mechanical seals is still changing rapidly, so the best course for this
text is to address the function of the basic parts and the fundamental differences in design, paying
attention to the points that are thought to have a significant effect on seal performance and life. When
a need arises for more detailed information, the reader should consult the literature (e.g. Mayer [9.1],
Lebeck [9.4]), or the seal manufacturers.
Beyond a seal face design able to maintain an adequate lubricant film under the expected operating
conditions, the design of a mechanical seal must also address the following basic functions:

1. Sealing the flexibly mounted face; keeping it in contact with its mating face and driving it (resisting the
torque developed by friction between the faces).

Fig. 9.17 Open seal housing.


176 Mechanical Seals

2. Whether to have the face loading or compression device rotating or stationary; wet or dry.
3. How to mount the seal.
4. How to introduce liquid into the seal housing for cooling.

Sealing the flexibly mounted face (secondary sealing), keeping it in contact with its mating face, and
driving it are the functions of what is generally termed the compression unit or flexible element. Sealing
is difficult because the design must effect a positive seal while allowing the sealed face to move freely.
This is accomplished by one of two basic methods: a dynamic gasket, frequently an O-ring (Fig. 9.18),
between the flexibly mounted face and its sleeve or housing, or a bellows, either elastomer (Fig. 9.19)
or metal (Fig. 9.20). Seals employing a dynamic gasket are often referred to as pusher seals, a term
derived from the gasket being pushed along the sleeve or housing to compensate for face wear. Although
a bellows provides the flexible seal, it still must be sealed to the sleeve or housing. Elastomer bellows
do this with a metal-banded, tight-fitting sleeve or nose at the end of the bellows (Fig. 9.19). A metal
bellows relies on some form of static gasket (Fig. 9.20), the form and material dependent on the service
conditions. Each of the means of secondary sealing has advantages for particular service conditions. As
an aid to selection, Table 9.1 summarizes the salient features of the two basic designs. The means of
secondary sealing is a principal distinction between mechanical seal design, and can therefore be used
as a basis for classifying seal designs (Fig. 9.21).
Springs are the usual means of energizing or compressing the seal to keep its faces in contact. Most
designs use coil springs, either a large single spring or an arrangement of multiple small springs. A large
single spring (Fig. 9.19) is simpler, has a low spring constant (making it tolerant of setting variations

Compression Springs

Coil

Fig. 9.18 "Pusher" type mechanical seal with dynamic gasket.


(Courtesy Durametallic Corp.)
Mechanical Seals 177

Fig. 9.19 Elastomer bellows type mechanical seal.


(Courtesy John Crane, Inc.)

Fig. 9.20 Stationary metal bellows type mechanical seal.


(Courtesy EG & G Sealo/)
178 Mechanical Seals

Table 9.1 Mechanical seal capability versus secondary seal type

Bellows
Secondary seal Dynamic gasket Elastomer polymer Metal

Compression Single or multiple springs Bellows


Size range Broad; 12-500 mm Broad Narrower; 20-125 mm
(0.5 to 20 in.) (0.75 to 5.0 in.)
Pressure limit High; to 200 bar Low; to 2 bar Intermediate; to 35 bar
(3,000 psig) (30 psig) (500 psig)
Temperature range Intermediate; -57 to 150°C Low; -18 to 50°C High; -270 to 425°C
(-70 to 300°F) (0 to 120°F) (-450 to 800°F)
Balance Available Available Inherent
Hysteresis Moderate Negligible
Risk of fretting High! None
Special designs Available Not available Limited
Special materials Available Not available Limited
Number of parts More More Fewer
Cost Base Usually lower Usually higher
! Reduced with stationary compression unit.

and face wear), and is less likely to become clogged with foreign material that may be in the sealed
liquid. Size, rotative speed, and available materials determine the limits of single springs. A multiple
spring arrangement (Fig. 9.18) allows a more compact seal, is said to apply more even face loading,
can tolerate higher rotative speeds, and the smaller springs are easily made in the more exotic alloys.
The smaller springs have a high spring constant, making the seal more sensitive to setting variations
and face wear, and are more likely to clog if there is foreign material in the sealed liquid (there are
design variations to avoid this; see dry compression units).
Elastomer bellows seals are generally spring energized (Fig. 9.19), the bellows typically being too
resilient for this function. Metal bellows seals, however, have as one of their virtues a bellows that
serves as both the secondary seal and the means of energizing the seal (Fig. 9.20). The spring constant
of metal bellows is high, thus they are sensitive to setting variations and face wear in the same manner
as multiple spring seals.
Coil springs are not the only device employed to energize mechanical seals. Wave washers are used

DynamiC
gasket
'O'Ring
'i,v'Rin
Wedge
g
'U'Ring
Balance
Unbal'd
Comp'n
{ ."'"',,•.} ,m {
't
Rotating
Stationary
Wt
D' },~"""'
1
FI 'bl
eXI e
Single spring
Mit' .
U I spring
Wave washer
E"""",,,_ wosh"
Seal { 'C'Ring ry Elastomer washer

LElastomer
Bellows L Polymer
Metal -+-Inherently balanced-+ comPt 'n -{ SRt °t~ating }- Flexible -l Metal Bellows
Un! a lonary Element

Fig. 9.21 Mechanical seal design classification.


Based on form of secondary seal.
Mechanical Seals 179

in some designs instead of multiple coil springs; the characteristics are similar. Some seals designed for
abrasive-laden liquids have used a plastic encapsulated Belleville washer or a rubber block in shear as
both the compression device and the secondary seal (Fig. 9.22). These have very high spring constants,
requiring that the seal be adjusted periodically to compensate for wear. Very large seals have used sealed
pressure, or some fraction of it, applied to a piston built into the moving face (Fig. 9.23) to energize it
by hydrostatic force. A novel design for small seals uses magnetic attraction (Fig. 9.24). The magnetic
materials limit the services to which this elegantly simple design can be applied.
Torque is transmitted through the compression unit by one of three basic designs. The most common
is an arrangement of drive pins (Fig. 9.18) or tabs and grooves (Fig. 9.19). Although a little complicated,
these arrangements have the advantage of being bidirectional. An alternative design for pusher seals is
to use a single coil spring to both energize the seal and transmit the face torque (Fig. 9.25). This
arrangement has the virtue of simplicity but the seal is now unidirectional, requiring care in specification
and installation to ensure the applied torque tends to extend (unwind) the spring. In metal bellows seals
(Fig. 9.20), the torque is generally transmitted through the bellows.
At the time of this writing, the more common arrangement is to have the compression unit rotating.
By and large that is a legacy of seals designed to fit in the space provided for packing and a belief that
rotating the compression unit improved heat dissipation from the seal faces. Once the requirement to
have mechanical seals and packing interchangeable is dispensed with, the better design is to have the
compression unit stationary. There are two distinct advantages. First, a stationary compression unit
accommodates angular misalignment between the rotor axis and seal housing face by taking a "set,"
whereas a rotating compression unit is subject to continual cycling in this circumstance (Fig. 9.26).
Second, a stationary compression unit does not have its action influenced by inertia effects caused by
rotation. For seals operating at nominal surface speeds above 250 mls (4500 ft/min), a stationary
compression unit is mandatory to avoid operating problems due to inertia effects. In the case of metal
bellows seals sealing solids-laden liquids, the point is moot. One argument is that rotating the bellows
ensures that it does not become clogged with solids. Another argument is that a stationary bellows is
less likely to be eroded by the solids. On balance, the better solution is probably to have the bellows
stationary and ensure adequate circulation within the seal housing (see pump design).
For clean, innocuous liquids, the simpler arrangement is to have the compression unit wet (Figs. 9.18,
9.19,9.20, and 9.26). When the liquid contains solids likely to clog the seal parts, the compression unit
needs to be a special design (Fig. 22) or arranged so it is dry (Fig. 9.27). Metal bellows seals, by way
of their construction, always have the compression unit wet.
Seals are mounted in one of two ways: direct or cartridge. In direct mounting (Figs. 9.18, 9.19, and
9.20), the major components of the seal (compression unit, faces, sleeve and flange) are assembled into
the pump separately or as subassemblies. The setting of the seal is therefore dependent on locating
surfaces within the pump or the mechanic's skill in positioning the parts correctly. On top of this, the
parts are all susceptible to damage and contamination until the seal is finally closed up. By preassembling
the seal, sleeve, and flange into a unit, an arrangement known as cartridge mounting, these problems
can be eliminated as possible causes of premature seal failure. Cartridge-mounted seals (Figs. 9.25 and
9.27) are set with locking tabs between the flange and sleeve. These keep the cartridge assembled as it
is installed into the pump. Once the cartridge is installed and the sleeve and flange secured, the lock
tabs are removed. A further advantage of cartridge mounting, one not related to reliability, is greatly
reduced tum around time for seal changing (provided, of course, that a spare cartridge is available). Not
by any means a new idea, cartridge mounting awaited a critical need to improve seal reliability before
its inherently greater expense became acceptable.
The usual means used to introduce liquid into the seal housing is a tapped connection in the housing
or flange. When the sole purpose of doing this is to cool the seal, the flow necessary is quite low and
provided the flow is controlled, a simple flange tapping (Fig. 9.20) is adequate. For cooling, it is important
ISO Mechanical Seals

(a) 3 Fixed ring


LO a Ring insert
4 Rotating ring
L3 b Ring insert
3 5 Cone spring
membrane
6 Drive rings
7 Sleeve
2 Shaft

(b)

Clamping
plate for
Virtually quick seal
Indestructible readjustment
Rubber-In-Shear
element with
Rotating unit smooth contours
clamped between
impeller and sleeve
for easy installation

Fig. 9.22 Mechanical seals for slurry applications using two alternative compression devices.
a) Encapsulated conical spring. (Courtesy Cefilac)
b) Rubber in shear. (Type RIS; BWIIP International Inc.)
Mechanical Seals 181

1. Rotating Face (Composite)


2. Stationary Face
3. Inner Housing
4. Outer Housing
Shaft 5. Retaining Ring
Flange
6. Wear Indicator
7. Head Cover
8. Static Seal (Air Operated)

Shaft

Fig. 9.23 Axial face seal with hydrostatic loading.


(Courtesy Sterling)

that the flush liquid be directed into the region of the seal faces. If the flush or injection flow is intended
to both cool the seal and raise the pressure in the seal housing by pressure drop across a throat bushing,
the flow needed can be quite high. In this case, the velocity of liquid issuing from a simple flange tapping
may easily be high enough to disturb the bahavior of the seal by impact or eventually erode the seal
parts. This problem is avoided by either diffusing the flow through the flange connection (Fig. 9.28) or
adding a second connection in the housing proper and introducing most of the flow through that. Of the
two approaches, the former is the preferred because it is simpler. It is not always possible, however.

(1) stationary seat


(2) carbide-ceramic sealing head
(3) magnetic ring
(4) a-ring static seal,
also transmits the torque

Fig. 9.24 Magnetic axial face seal showing magnetic circuit.


182 Mechanical Seals

Fig. 9.25 Single spring, spring driven, and cartridge mounted seal.
(Courtesy Flexibox International)

~I''''~

9
¢ -Static Angularity
Between Rotor &
Stationary Face

(a) Rotating Compression Unit-Cycles Every


Revolution to Compensate

Rotating Face
Square To Shaft Axis

¢-Static Angularity
Between Rotor &
Stationary Unit
Mounting Face
(b) Stationary Compression Unit-Adopts
Set Position To Compensate

Fig. 9.26 Effect of angular misalignment on compression unit: a) rotating, b) stationary.


Mechanical Seals 183

Fig. 9.27 Seal for general purpose or slurry applications with stationary, external spring.
(Courtesy Flexibox International)

Materials
Face materials are chosen based on wear and corrosion resistance in the sealed liquid, heat dissipation,
and cost, there being a distinct connection between good performance and cost. The usual combination
is hard against soft, and the materials range from reinforced phenolic against cast iron to carbon against
silicon carbide (possibly zirconium carbide by the time this is published). As noted in the discussion of
vaporization at the faces, the choice of the hard face material can significantly affect the conditions
required for satisfactory operation of the seal (Fig. 9.6). General practice is to rotate the face with the
higher thermal conductivity. In some difficult hydrocarbon services (e.g., light ends), there are instances
where having the better conductor stationary seems to materially aid the development of a stable liquid
film between the faces. Seals for slurries employ hard against hard faces, typically silicon carbide against
tungsten carbide. This choice is dictated by the need for abrasion resistance; it is tolerable because the
surface speeds are relatively low.
Corrosion and endurance strength govern the selection of materials for the metal parts in mechanical
seals. Corrosion needs particular care because the relatively small parts usually cannot tolerate the degree
of metal loss that would be quite acceptable in, say, a pump casing. Metal parts are at least type 316
stainless steel. Welded metal bellows are typically Hastelby C® or similar materials suitable for high-
temperature service. Shaft sleeves are frequently hard coated in the region under the dynamic gasket in
pusher seals to reduce the risk of fretting corrosion. Seal flanges are at least chrome steel for adequate
corrosion resistance at the critical sealing and locating surfaces.
Dynamic gasket materials range from neoprene to Teflon.™ Resilience is an important characteristic
184 Mechanical Seals

Multi-Point
Flush Injection

Fig. 9.28 "Multipoint" injection to diffuse injection flow.


(Courtesy Flexibox International)

or the seal may either leak (too loose) or have high hysteresis (too tight). For this reason when the
service conditions require Teflon,TM it is only used in a form (C ring, encapsulated a-ring) where its
lack of resilience is compensated for. Static gaskets are chosen for chemical and temperatures resistance.
Materials range from neoprene through GraphojlTM to soft iron or stainless steel.

Auxiliary Seals
Many mechanical seals are equipped with some sort of auxiliary seal. These can function to allow
the introduction of a quench fluid (often low-pressure steam) to remove deposits formed at the atmospheric
Mechanical Seals 185

side of the seal, or to contain leakage for a short time should the seal fail (similar to a tandem seal but
without the sophistication and durability). Designs range from a simple close clearance bushing, lip seal
or packed box for quench fluid, to dry running (gas type) mechanical seals (Fig. 9.29) and abeyant
mechanical seals (faces not normally in contact; Fig. 9.30) to contain leakage should the main seal fail.

Seal Environment

The importance of providing the correct seal environment cannot be overstated. Attention to this
aspect of seal application ensures the best seal choice in the selection phase and avoids the extremely
high costs that can be associated with a chronically unreliable mechanical seal.
From the discussion of vaporization at the seal faces, it is clear that dissipation of the frictional heat
generated at the faces is important. If the liquid being sealed is at a temperature well below its boiling
point at the sealed pressure and the housing design allows good natural circulation, (Fig. 9.17), the
prevailing seal environment is satisfactory. Should the housing design not allow good circulation, piping
a small flow from the pump discharge (API Plan 11, Fig. 9.31) to the seal flange will correct the
deficiency to an acceptable degree.
When the liquid is being sealed at conditions close to its boiling point, the requirements for satisfactory
operation become more complex. Depending on the seal face design selected (see face lubrication), it
will be necessary to cool the sealed liquid below the pumping temperature or raise the pressure in the
seal housing. Lowering the temperature can be achieved by injecting cool liquid from an external source
(API plan 32; Fig. 9.32), by cooling a small flow from the pump discharge (API plan 21; Fig. 9.33), or
by limiting the heat flow from the pump into the seal housing and circulating liquid from the housing
through a heat exchanger and returning it across the seal faces (API plan 23; Fig. 9.34). Of these three
approaches, the third, plan 23, has the highest thermal efficiency; that is, it dissipates the least amount
of energy to provide the environment needed for the seal. Plan 23 is expensive, although not significantly
more than plan 21, and requires a certain minimum peripheral speed to effect adequate circulation of
the sealed liquid. Despite these limitations, plan 23 is used extensively because it is reliable and does
not rely on exotic seal face materials.
Raising the pressure in the seal housing is achieved by either eliminating the balance holes in a single-
suction impeller (see Chap. 4) and circulating liquid from the seal housing back to the pump suction
(API plan 13; Fig. 9.35), or taking a large flow from the pump discharge (API plan 11) or an external
source (API plan 32) and developing the required pressure drop across a throat bushing. The former is
the better design since the seal housing pressure is not influenced by pump wear. It is not always feasible,
however, because the resultant axial thrust can easily exceed bearing capacity. The flows required for
plan 11 or 32 are often larger than expected, particularly when the liquid has low SG (see seal design
for cautionary comments on introducing large flows into mechanical seal housings).
Liquids pumped at high temperatures but still below their atmospheric boiling point, e.g. heat transfer
oil and refinery bottoms), can be sealed at or close to the pumping temperature using metal bellows
seals. The heat developed by seal friction is dissipated with small flow from the pump discharge (API
plan 11). Seal face life will be longer, however, if the temperature at the seal is lowered. Some users
have achieved a moderate reduction in temperature by adding a liquid to air heat exchanger (making it
plan 21; Fig. 9.33) located in the motor cooling air draft. A still more conservative approach, one that
is yielding very long seal face lives, is to keep the seal both cool and clean by injecting cool light cycle
oil from an external source (API plan 32) during normal operation.
Modern pump designs has all but eliminated the jacketed seal housing as a means of cooling seals.
When new their effectiveness was relatively low; once in service for a while, particularly at high
temperature, scaling on the water side reduced their effectiveness still further. Given this, the simplification
realized by eliminating this cooling more than offset the small advantage. There is perhaps some advantage
186 Mechanical Seals

INJECTION THRU
CAGE RING TAP

CAUTION:
BOll HOLES MUST
ClEAR OIA "W"

A
C MIN
t - - - - - - - - J- - - - - -- + -- -- - - - G-2

Fig. 9.29 Dry running (gas type) auxiliary seal.


(Courtesy BWIIP Inc.)

Fig. 9.30 Abeyant auxiliary seal.


(Courtesy Flexibox International)
Mechanical Seals 187

Orifice; Min. dia. 0.12 in.

Fig. 9.31 Seal flush from pump discharge; API plan 11.

From external
source

Fig. 9.32 Seal flush from external source; API plan 32.

in some form of cooling jacket to act as an additional thermal barrier in pumps employing self-circulating
seals (API plan 23).
If by way of pump construction and suction pressure, the pressure at the seal would normally be less
than 0.35 bar (5 psig), the seal housing pressure must be increased (see face lubrication). Two approaches
are used: The first is to use one of the methods previously described for raising the sealed pressure when
pumping liquids close to their boiling point. Note that unless plan 32 is used, seals pressurized by these
methods will be at suction conditions in pumps on standby, and may therefore suffer considerable "air-
in" leakage if the suction pressure is a high vacuum. The alternate is a double seal, a complex solution
most users try to avoid.
With modem mechanical seals, it is generally not necessary to lower the pressure at the seal. Should
188 Mechanical Seals

Heat
exchanger

Fig. 9.33 Seal flush from discharge through heat exchanger; API plan 21.

Heat Exchanger

Seal Water In

Fig. 9.34 Circulation of sealed liquid through heat exchanger; API plan 23.
Mechanical Seals 189

Fig. 9.35 Seal flush from within pump back to suction; API plan 13.

doing so be necessary for some reason, a breakdown bushing with bleed-off to a point of low pressure
(see Chap. 8) is the method used.
Solids-laden liquid cannot be sealed effectively with conventional mechanical seals. Slurries are being
sealed with varying degrees of success, using seals designed specifically for several abrasive service
(Figs. 9.22 and 9.27). Unless legislation dictates otherwise, slurry seals have all but replaced double
seals for these duties. Installing a throat bushing in the bottom of the seal housing and injecting clean
liquid from an external source, thus excluding the pumped liquid from the seal housing, will allow a
conventional seal to work. The drawback is the loss of plant efficiency caused by product dilution or
returning refined product to the cycle.
In services where the liquid is merely contaminated with solids, a simple means of cleaning up the
flush liquid is to use a cyclone separator (API plan 31; Fig. 9.36). These devices (Fig. 9.37) function
by accelerating a liquid stream to high velocity through an orifice, then feeding it into a cone. The high-
velocity swirling in the cone centrifuges those solids heavier than the liquid to the outside of the liquid
body. Clean liquid is drawn from the center of the cone (outlet C) and piped to the seal flush connection.
The concentrated dirty liquid is piped from the bottom of the cyclone (outlet B) back to the pump
suction. For cyclone separators to function correctly, the pressures at outlets B and C need to be within
1.4 bar (20 psi) of each other. If the sealed pressure is more than 1.4 bar (20 psi) above suction pressure,
an orifice (not a valve) should be added in the line back to the pump suction to raise the pressure at the
cyclone outlet. By their very action, cyclone separators will not separate solids whose sa is lower than
that of the liquid. Their action is also impaired by high viscosity; the drag on the solids reduces the
degree of separation, particularly of the finer solids. To avoid having the cyclone plug at the inlet orifice
and stop all flow to the seal, it is prudent to install a Y-type strainer upstream (Fig. 9.37). The strainer
can be arranged for manual backflushing if on-line cleaning is deemed necessary.
Double mechanical seals, by definition, have the region between them filled with a barrier fluid that
is maintained at a pressure above that of the sealed liquid. When the barrier is a liquid, it is generally
necessary to circulate and cool it to dissipate the heat generated at the seal faces. In most cases, this
function is realized using an external, pressurized reservoir (API plan 53; Fig. 9.38). Circulation is by
a pumping device on the seal if that is feasible, otherwise by a separate pump. Multiple installations of
double seals are sometimes served by a single, central circulation system (API Plan 54; Fig. 9.39).
190 Mechanical Seals

'Y' type strainer

Cyclone
separator

Fig.9.36 Seal flush from discharge through cyclone (abrasives) separator; API plan 31.
Note addition of "Y" type strainer.

Clean Flow Outlet

.......... :>"'.......1.......... . .
.' . .......:.;-- Inlet

.'

Dirty Flow Outlet

Fig. 9.37 Abrasives separator; a) connections, b) principle of operation.


(Courtesy John Crane, Inc.)
Mechanical Seals 191

IJ:NII PRESSURE lOW LEVl:L

9 o
ALARM AlARM
FU

IlELIEF IALVI: I

I
I
I
I
I

IJ:NII PRESSUflE
SWITCH

PUMP SHAFT

DOUBLE SEAL

Fig. 9.38 API plan 53 system for double (pressurized dual) seals.
(Courtesy Flexibox International)

Barrier liquid is circulated by the system; the pressure at the seal is maintained by an orifice or regulator
in the common return line.
The primary seal in a tandem mechanical seal arrangement has its environment maintained as if it
were a single seal (see earlier discussion). The secondary seal is lubricated and cooled by a buffer liquid
at atmospheric pressure. This function is carried out by an external, nonpressurized reservoir (API plan
52, Fig. 9.40). Practice for circulating and cooling the barrier liquid follows that for plan-53 systems.
192 Mechanical Seals

1_ _ _ .__ .._._.

Fig. 9.39 Central circulator for double seals with liquid barrier fluid; API plan 54:
a) circulator unit, b) flow diagram.
(Courtesy Durametallic Corp.)

In addition to lubricating and cooling the secondary seal, plan-52 systems must also detect primary seal
leakage and isolate it. This function is achieved with an orifice and motorized valve in the reservoir's
vent line. A high flow through the vent line will cause a measurable pressure rise in the reservoir. A
pressure switch in the reservoir detects this, sound an alarm, and closes the motorized vent line valve
to contain the primary seal leakage. From this point on, the pump is being sealed by the secondary seal
until it is shut down and the seal assembly replaced.
So far the discussion of mechanical seal environment has centered on the sealed liquid side of the
interface. This is by far the more important side, but for many services the environment at the atmospheric
side is also of major importance. The services in which this is the case are those that may produce
freezing, crystallization, carbonization, or precipitation on the atmospheric side as a result of the small
leakage that is inherent in mechanical seals. These physical phenomena must be avoided lest seal
operation can be impaired or the seal damaged by the build up of solids in the region of the seal interface.
Freezing is prevented using a dead-ended blanket (API plan 51). An external fluid quench of water,
steam, or a gas (API plan 62), is used to prevent crystallization, carbonization, or precipitation. In many
instances, it is necessary to incorporate a special sleeve or bushing in the seal flange to ensure the quench
fluid circulates into the region under the sealing interface.

Operation
An all encompassing rule often quoted by seal manufacturers is: Never start up seals dry. With the
variations in modem mechanical seal designs, this means: Do not start the pump before ensuring its seal
has the means needed for face lubrication.
Mechanical Seals 193

o
HGH I'RESSUIIE lOW LEVEL

9
AlAAM AlARM

RElIEf -AlVE

I
r
I
I PRESSURE
I INOICATOR

HGH PRfSSURE
SWITCH
lEVEL INDICATOR

PUMP SHAFT

TANDEM SEAL

Fig. 9.40 API plan 52 system for tandem (unpressurized dual) seals.
(Courtesy Flexibox International)

In the simpler cases, just following normal pump operating practice is sufficient to ensure the seal
functions correctly. As the seal becomes more complicated, however, it becomes necessary to include
some additional procedures. What is necessary can be determined fairly easily by considering the
provisions made to ensure the correct seal environment. When the temperature at the seal is maintained
at other than the pumping temperature, there should be means to check the actual temperature and the
temperature needs to be verified periodically. The same facility and periodic check is necessary when
194 Mechanical Seals

the pressure at the seal is maintained at other than that inside the pump. By doing this, a change in
temperature or pressure can be used as an indicator of some deterioration in the pump or seal, thus
leading to investigation before seal failure indicates there was a problem. The even more complicated
auxiliary systems used with double and tandem seals should be checked for cleanliness and function
before being put into service, and once in service, their operation requires periodic verification. In critical
services, monitoring equipment is being used to warn of a significant change in seal environment, for
example, return liquid temperature in a plan-23 system or high seal leakage. The latter rely on accumulation
of leakage in a vessel or a rise in pressure in the vent space of the seal flange. When contemplating this
degree of sophistication, it is worthwhile to weigh what has to be monitored automatically against the
added complexity of doing so. Often it is better to adopt a slightly simpler approach. And in that simpler
approach it is always worthwhile to pay heed to visual observation; instrumentation cannot tell all.

Maintenance
As a general rule it is better not to open a seal for maintenance unless a change in operation (pressure,
temperature, or leakage) or seal wear (some seals have wear indicators) dictates doing so. Once the seal
is opened, it is important to carefully inspect the parts to learn whether the deterioration of the seal was
due to normal wear and tear or an abnormality (operating condition or component) that could be
avoided the next time. As an aid to this activity, the major seal manufacturers have available excellent
troubleshooting guides, complete with illustrations of the various types of damage that can be encountered.
If examination of the seal parts identifies a correctable problem (e.g., seal environment or component
materials), correcting the deficiency should be the next step in returning the seal to service.
Restoring the seal to new condition can be handled one of two ways. The plant can have a stock of
replacement parts, preferably obtained from the seal manufacturer to ensure continuity of quality, and
draw from these to rebuild the seal. Alternatively, the plant can just stock replacement units, sending
the entire worn seal to a specialist shop for restoration. Unless the plant maintenance shop is very
sophisticated, this approach is necessary for seals such as metal bellows, which must be leak tested with
gas to verify bellows integrity. The concept of exchanging complete seal units is entirely consistent with
that of cartridge-mounted seals, and explains why the practice is on the increase. When the maintenance
practice is to exchange the seal unit, examining the replaced seal for abnormal damage becomes the
responsibility of the specialist shop. The plant, however, still needs to maintain a certain awareness of
what is normal so any tendency toward premature failure is quickly detected, its cause identified, generally
in conjunction with the repair shop, and the problem corrected.
Following the lines of the discussion of what pump design does for seals, it is important to pay heed
to what pump condition can do to seals. During maintenance, this refers to checking that the pump's
rotor is sufficiently straight, that its axis is sufficiently concentric with and square to the surfaces that
locate the seal flange, and that the adjacent bearing clearances, radial and axial, are within those for
which the seal was designed. Detailed diagrams of the checks that need to be made are invariably
included in the seal manufacturers' trouble shooting guides. Acceptance criteria are given on other the
seal drawing or in the installation instructions.

BIBLIOGRAPHY

[9.1] E. Mayer, Mechanical Seals, 5th edition, 1977 Newnes-Butterworths, London, UK.
[9.2] API-682, Shaft Sealing Systems for Centrifugal and Rotary Pumps, American Petroleum Institute, Washington,
D.C., 1995.
[9.3] ANSI B73.1, 1991.
[9.4] A. O. Lebeck, Principles and Design of Mechanical Face Seals, John Wiley, New York, 1991
10
Breakdown Seals

In the history of centrifugal pump development, breakdown-type seals were employed when the sealing
conditions of surface speed, pressure drop, or both exceeded the capability of packed boxes. Their
greatest use came with the advent of the high-speed boiler feed pump. At the time of this development,
the early 1950s, mechanical seals were not able to demonstrate the dependability necessary for such a
critical application. Today the circumstances are different; mechanical seals are an equal contender for
such services.
Breakdown seals act on the circumference of the shaft or rotor, and are therefore termed radial seals,
as are packed boxes. They are distinguished from packed boxes by having a specific clearance between
the shaft or rotor and the stationary breakdown device (Fig. 10.1). This arrangement eliminates the heat
generated by rubbing in packed boxes, but does, of course, allow greater leakage through the seal. The
magnitude of the clearance is a compromise between leakage rate and the need to accommodate rotor
runout and minor variations in the rotor's radial position.
All conventional breakdown seal installations require an auxilliary system to gather and dispose of
the seal leakage. Whether the installation also requires an injection system to introduce liquid from an
external source into the seal depends on the nature of the pumped liquid.

CONDENSATE INJECTION SEALING

Most of the breakdown seals in use today are in boiler feed pumps. The pumped liquid in these cases
is at a temperature above its atmospheric boiling point, and so would flash into vapor if allowed to leak
through the seal. To avoid this, the seals are injected with cold condensate to cool the leakage. This
arrangement is commonly known as "condensate injection sealing." The seals, given their development
from packed boxes, are often termed "packless" stuffing boxes.
The construction of a pump with condensate injection sealing is illustrated in Fig. 10.1. A labyrinth
breakdown bushing is substituted for the conventional packing, and the pump shaft sleeve runs within
this bushing with a reasonably small radial clearance. Cold condensate, available at a pressure in excess
of the boiler feed pump suction pressure, is introduced centrally in this breakdown bushing. A small
portion of the injection water flows inwardly into the pump proper; the remainder flows out into a

195
I. J. Karassik et al., Centrifugal Pumps
© Chapman & Hall 1998
196 Breakdown Seals

Fig. 10.1 Breakdown seal construction for high-pressure boiler feed pump.

collecting chamber that is vented to the atmosphere. From this chamber, the leakage is piped back to
the condenser.

SOURCE OF SUPPLY

Cold condensate (at temperatures from 27-38°C (80 to 100°F)) is available at the condensate pump
discharge. In closed cycles (Fig. 10.2) and those open cycles where the feed pump takes its suction
directly from the deaerator (Fig. 10.3), the pressure at the condensate pump discharge is higher than
feed pump suction pressure. When the feed pump in an open cycle is preceded by a suction booster
pump (Fig. 10.4), feed pump suction pressure is generally higher than condensate pump discharge
pressure. In these cases, it is necessary to either lower the pressure at the feed pump seals (by bleeding
off to the deaerator) or use booster pumps to raise the pressure of the cold condensate. Some installations
employ a closed heater between the suction booster and boiler feed pump. Along with a certain penalty
in total pumping power, this arrangement dictates the need for condensate injection booster pumps, since
the pressure at the seals cannot be lowered appreciably.
The water for the injection in all cases should be taken immediately from the condensate pump
discharge before it has gone through any closed heaters. It is preferable to use injection water at
Breakdown Seals 197

SUCTION PRESSURE:
Z50 PSI AT FULL LOAD
FRICTION LOSSES: 337 PSI AT 1/4 LOAD
40 PSI AT FULL LOAD
3 PSI AT 114 LOAD

TEMPERATURES:
32O"F AT FULL LOAD
Z50-F AT 1/4 LOAD

'------11
STATIC CALIBRATED ORIFICE
HEAD
IOPSI

CONDENSER HOTWELL
I CLOSED
HEATERS

CONDENSATE INJ£CTION LINE


FRICTION LOSS:
5 PSI AT FULL LOAD

DISCHARGE PRESSURES'
300 PSI AT FULL LOAD
350 PSI AT 1/4 LOAD

Fig. 10.2 Application of injection breakdown seals in a closed feedwater cycle.


Pressure distribution indicated for full and one-quarter load.
DEAERATING
HEATER
316·F, 70 PSIG
AT FULL LOAD
23Z"F, 7 PSIG
AT 114 LOAD

STATIC HEAD
60FT
SUCTION PHESSURE:
906 PSIG AT FULL L0401
31 4 PSIG AT 1/4 LOAD

INJECTION PRESSURE·
I~~ PSIG AT FULL LOAD
200 PSIG AT 114 LOAD
CLOSED
HEATERS

DISCHARGE PRESSURE·
CONDENSATE 170 PSIG AT FULL LOAD
PUMP ZI5 PSIG AT 1/4 LOAD

Fig. 10.3 Application of injection breakdown seals in an open feedwater cycle.


198 Breakdown Seals

Deaerating Heater
316°F, 70 PSIG at full load
232°F, 7 PSIG at V4 load

Antiflash Orifice LiP = 20 PSI


1/
Suction End

1ft
20
Bleed-off Balancing Leak-off
\.------1~___f/ LiP = 5 PSI
Calibrated Orifice

j - - - LiP=5PSI
'----./-"'1

-, Suction Booster
Boiler Feed Pump
Suction Pressure:
15 ft Discharge Pressure 188 PSIG at full load

I
Closed
194 PSIG at Full Load
Heaters 151 PSIG at V4 load
157 PSIG at V4 Load

Condenser
Hot-well
-.1-ft- f-_P..."ip_in..."g=-Li_P_=_5_P_S_I_--{

I Allema.v~o \ Bleed-ofts back 10 deaerator

Condensate Pump
Discharge Pressure: tI 0 ,I

170 PSIG at full load Injection Booster Pump


215 PSIG at V4 load LiP = 59 PSI

Fig. 10.4 Application of injection breakdown seals in an open feedwater cycle with a suction booster pump.
Pressure distribution indicated for full and one-quarter load.

temperatures below 49°C (l20°F) so as to avoid the slight steaming at the seal covers (steaming takes
place if injection temperatures in excess of that figure are used). This steaming is undesirable partly
because of the concern it may arouse in the operators and partly because of the possibility of its
condensation near the pump bearings.
Pumps arranged for condensate injection sealing, moreover, are usually not provided with any other
cooling means in that area. If the shaft sleeves rotating within the condensate seal breakdown bushing
are not adequately cooled, the heat from inside the pump will travel through the shaft to the pump
bearings and may be injurious to the bearing life.
In the closed feedwater cycle, all de aeration takes place in the condenser, and therefore the injection
water is fully deaerated. Although this is not quite true in the open cycle, an appreciable amount of
deaeration takes place in the condenser even though a deaerating heater is provided in the feed cycle.
Thus the injection condensate in a modem steam power plant will contain almost no oxygen (0.01 cc
per liter or less). As the saturation level at a temperature of 38°C (lOO°F) and atmospheric pressure is
Breakdown Seals 199

about 4.7 cc per liter, the oxygen in the injection supply itself has no significance. Moreover, the amount
of injection water that enters the pump proper and is not returned to the condenser is very small. Thus
no appreciable contamination of the feedwater will take place through the condensate injection sealing.
The one exception to this general rule can occur in an open cycle with a suction booster ahead of the
feed pump (Fig. 10.4). If both the feed pump and its suction booster have breakdown-type seals, and
condensate injection is achieved with a booster pump, the injection pressure available at the booster
pump is appreciably higher than its suction pressure. Unless the seal injection flow into the booster
pump seals is carefully controlled, the flow of partially deaerated condensate into the feedwater can be
high enough to materially raise the oxygen content.
The amount of the injection water will depend on (1) the diameter of the seal, (2) the clearance
between the shaft sleeve and the pressure breakdown bushing, and (3) the injection pressure. To give
some general idea of the values in question, if the sleeve diameter is 125 mm (5 in.) and the diametral
clearance 0.225 mm (0.009 in.), the amounts measured in a 3,600-rpm pump will be approximately
as follows:

1. Total injection per seal-1.8 to 2.3 m3/hr (8 to 10 gpm)


2. Leakage into the pump interior, per seal-0.4-0.9 m 3/hr (2 to 4 gpm)
3. Return to condenser, per seal-1.4-1.8 m 3/hr (6 to 8 gpm).

It is essential that the injection supply be absolutely clear and free of foreign matter. It is, therefore,
necessary to install filters or strainers in the injection line to avoid the entrance of fine mill scale or
oxide particles into the close clearances between the stationary bushings and the sleeves. Pressure gages
should be installed upstream and downstream of these filters to permit the operator to follow the rate
at which foreign matter clogs up the filters and to clean these when the pressure drop across them
becomes excessive.
For overall reliability, the source of injection condensate must be adequate for all likely operating
conditions. To ensure this, the following influences need to be taken into account:

1. Variations in the pressure at the pump's seals over the entire range of operating flows (see open cycle and
closed cycle)
2. An increase in the pressure at the seals as the pump's running clearances wear. This refers specifically to
those clearances located adjacent to the shaft seals (e.g., the balancing device in Fig. 10.5). As these wear,
the higher leakage raises the back pressure in the region of the seal.
3. Higher seal injection flows to maintain the same injection pressure (or seal drain temperature) as the seals wear.

DRAINS FROM CONDENSATE INJECTION SEALING

Two different systems are used to dispose of the drains coming from the collecting chambers. The first
utilizes traps that drain directly to the condenser. The second collects the drains in a condensate storage
tank into which various other drains are also returned. As this tank is under atmospheric pressure, it
must be set at a reasonable elevation below the pump centerline so that the static elevation difference
will overcome friction losses in the drain piping. A pump then transfers the condensate drains from the
storage tank into the condenser.
To our knowledge, no specific difficulties have ever taken place in installations in which the injection
sealing condensate is evacuated through traps, except for an isolated case of trap malfunctioning. Proper
maintenance of this equipment should hold such occurrences to an absolute minimum. A minor problem
may arise if boiler feed pumps are operated during the start-up before condenser vacuum is established.
200 Breakdown Seals

FILTERED
CONDENSATE

t
INJECTION
STATION STATION
AIR SUPPLY AIR SUPPLY

I CONTROL
DIAPHRAGM
AIR
X-;>-_~_·~lTanf ER
VALVE SI

B~!!.S~~~~ I
t
SENSING LINE __ I
r---~
I
I
I
II I'
BLEEDOFF TO DEAERATORA
OR BF PUMP SUCnON
( IF NO BLEEDOFF IS REQUIRED )
DRAIN

BALANCING
DEVICE LEAKOFF

Fig. 10.5 Pressure control of condensate injection.

This operation causes a rise in the back pressure on the seal drains and, unless provision is made to
relieve this back pressure, some overflow of injection condensate may take place at the collecting
chamber covers.
Neither system of evacuation has major advantages over the other, and the choice between the two
is dictated primarily by personal preferences. However, if the boiler feed pumps are located at the lowest
plant elevation (in some outdoor plants, for instance), it becomes necessary to use traps because there
is insufficient elevation difference to drain from the collecting chambers into an open tank.
The clearances between the sleeves and the breakdown bushings will increase about 50% in a time
approximately equal to the life of the internal wearing parts. With 50% larger clearances, the leakage
will double. This factor should be considered when sizing the return drain piping back to the condenser
or to the collecting tank if friction losses are to be kept to a minimum in this piping. The collecting
chamber at the pump stuffing box is vented to the atmosphere; the only head available to evacuate it is
the static head between the pump and the point of return. This head must always be well in excess of
the frictional losses (even after the leakage doubles); otherwise the drains will back up and run off at
the collecting chamber.
The water thrown off the shaft into the ventilated collection chamber will probably reach 75 percent
saturation or more before it reaches the drain pipes. Assuming that there is a considerable length of
partially full piping between the collecting chamber and the trap in the drain line to the condenser, the
oxygen content of the returned condensate can well be assumed to have 100 percent saturation. Although
this figure may seem very high, the deaerating capacity of a modem condenser is greatly in excess of
average requirements and should be amply capable of handling the oxygen in a saturated return, for it
Breakdown Seals 201

makes up 2 percent or less of the normal flow. Its presence, nevertheless, makes it necessary to provide
vents in the drain lines to avoid the accumulation of air pockets produced by separation of the entrained
air. Without vents there is a risk of erratic drain operation, and some "cyclic" spillover at the seal covers
each time a slug of air forms in the drain piping. In installations where this has occurred, a smaller
"geyser" would rise from the collecting chamber vents; increasing static head in the collecting chambers
would then force the slug of air out through the piping, and the geyser would subside. The installation
of vents or breathers eliminated the difficulty entirely.

CONTROL OF THE INJECTION

To complement the advantages gained from breakdown seals in maintenance and availability, it is
desirable to reduce the consumption of condensate for injection to as little as possible. Throttling valves
in the seal injection lines are used to achieve this. Generally, a separate throttling valve is installed at
each injection point (i.e., two valves per pump). In some early applications of condensate injection
sealing, the desire to simplify the installation led to the use of two valves for each group of pumps
serving a common turbo-generator unit. Such an arrangement is unable to compensate for differences
in pump and seal condition (extent of wear), and is therefore no longer used. Two valves are used
because the pressure on the suction and discharge sides will not necessarily be the same. An examination
of Fig. 10.1 shows that the inward flow on the suction side has to overcome a pressure exactly equal
to the pump suction pressure. On the discharge side, the flow will proceed into the balancing-device
relief chamber. The pressure there will exceed the suction pressure by the amount of loss through the
calibrated orifice in the balancing relief line, which is used to measure the leakage past the balancing
device. This loss may be quite appreciable after the pump has become worn.
Control of the injection throttling valves is by one of three systems: manual, pressure, and temperature.

Manual
Each throttling valve is adjusted manually to give the desired drain conditions. As operating conditions
change, it may be necessary to readjust the valves, particularly for pumps operating in closed cycles.
The valves also have to be readjusted periodically to compensate for the effects of pump and seal wear.

Pressure
An automatic pressure controller adjusts the throttling valve to maintain a constant pressure differential
between the injection pressure and the pump's internal pressure, thus ensuring a flow of condensate into
the pump under all conditions (Fig. 10.5). Pressure control compensates for changes in operating pressures,
but injects a constant amount of condensate regardless of changes in condensate or pumped liquid
temperature. Further, pressure control is not able to distinguish between a running and a stationary pump.
At the suction end, this makes no difference, but at the discharge end, the balancing disk does quite a
deal of pumping when the pump is running. This results in the pressure at the seal being below the
balancing leakoff chamber pressure, which leads to higher condensate consumption since the control
has to be set for the higher pressure, which prevails when the pump is stationary.

Temperature
Instead of ensuring a certain condensate flow into the pump, temperature control automates manual
control by using a temperature controller to adjust the throttling valve to maintain a "safe" seal drain
202 Breakdown Seals

FILTERED
CONDENSATE
STATION INJECTION STATION

+
AIR SUPPLY AIR SUPPLY

DIAPHRAGM I 4
!==
TROI.L£ A!!!_~X I\-D- AIR
_~_,.....JI"II.,..,-.:tnI
SIGNAL VALVE VALVE SIGNAL

I I
I I
ITEMP • I TEMP.
I'EN SING A
jSENSING
LINE
fiNE
I
I I
I
I I
I BALANCING
DEVICE I
IL _____ ~~
_ __- i. .~L~E~A~KO~F__~~__~ _ ______ -1I

DRAIN

Fig. 10.6 Temperature control of condensate injection.

temperature (Fig. 10.6). With this control, the consumption of condensate is kept to the minimum
necessary for all conditions. Accompanying the minimal consumption of condensate, there is a need for
additional care in the design of the seal and its injection control system. This is best explained by first
examining the nature of flow within a breakdown seal with temperature-controlled injection. Given a
normal drain temperature of 60 to 65°C (140 to 150°F), and having injection condensate typically at 27
to 38°C (80 to 100°F), it is reasonable to hypothesize that a small flow of "hot" feedwater from the
pump mixes with "cool" injected condensate to give "warm" seal leakage (Fig. 10.7). Field tests have

TH>TD>Tc
QD = QL +Qj

Fig. 10.7 Flow within temperature-controlled seal.


Breakdown Seals 203

___ Approximately equal


to condensate
temperature

Injection Pressure, Pi

Fig. 10.8 Drain temperature versus injection pressure.

confinned this is the case. The existence of a small flow of hot water from the pump has three
ramifications:

1. The operating temperature of the shaft sleeve at the inner end of the seal will be higher than at the outer
end. Since the labyrinth bushing has cool condensate circulating around it, it will remain cooler than the
sleeve, particularly at the inner end. To compensate for this difference in operating temperatures, and hence
thermal expansion, it is necessary that the inner end of the seal have a larger "cold" clearance than the outer end.
2. Once the injection pressure is reduced to the point where there is a flow of hot water out of the pump, the
drain temperature changes rapidly with only small changes in injection pressure (Fig. 10.8). To avoid flashing
in the seal (see later), it is crucial to measure the temperature of the seal leakage proper, not some reservoir
that would take too long to register the change in drain temperature, and to have a control system able to
respond quickly to small changes in that temperature.
3. Because there is little or no flow of cool condensate into the pump (the control valves have stops to prevent
their closing entirely so there will be flow into the pump under some conditions), temperature control aids
the warmup process and is critical for turbine-driven pumps, which must operate on turning gear. In the
latter case, pressure control would admit large flows of cool condensate into the pump, leading to temperature
stratification and consequent casing distortion, with pump seizure the likely result.

Although the objective of condensate injection control is to reduce the consumption of condensate, it is
of prime importance that the system be designed to avoid the possibility of inadequate injection and
consequent flashing in the seals. Should flashing occur, it is likely the seal will be destroyed by rubbing
between the shaft sleeve and labyrinth bushing. Compounding that, there is the possibility of shaft failure
from overheating in the region of the seal rub. In this connection the control system design needs to
pay due attention to the following: .

1. Accurate measurement of the control parameter (Le., pump internal pressure of seal drain temperature).
2. Design pressure drop across the control valve high enough to ensure the valve has control of the injection
flow. A pressure drop of 0.35 bar (5 psi) at maximum seal flow (typically two times the flow with "new"
clearances) has proven operable.
3. Control loop with minimum hysteresis.
204 Breakdown Seals

We must now examine the relationship between the pressure available at the injection of the condensate
and the internal pressure over the complete range of operating station loads. For reasons that will become
obvious, this relationship is vastly different for open and for closed feedwater cycles. The application
of breakdown seals to the two cycles should thus be studied separately.

OPEN CYCLE

The relationship between the various pressures under consideration is illustrated in Figures 10.3 and
10.4, both of which represent typical installations with condensate injection seals in an open cycle. In
the simpler arrangement (Fig. 10.3), the feed pump takes its suction directly from the deaerator. The
condensate pump discharges into a deaerating heater through a series of closed heaters. The discharge
pressure at the condensate pump rises from 11.1 bar (170 psig) at full load to 14.8 bar (215 psig) at
one-quarter load. The static head between the condensate pump and the deaerating heater is 24.4 m (80
ft), or approximately 2.4 bar (35 psi). The friction losses in the piping and through the closed heaters
are 3.1 bar (45 psi) at full load and only 0.20 bar (3 psi) at one-quarter load. Thus the pressure immediately
ahead of the heater will be 6.2 bar (90 psig) at full load and 12.2 bar (177 psig) at one-quarter load.
The condensate control valve located at the entrance to the deaerating heater will vary the admission of
condensate in accordance with load requirements, throttling off approximately 1.4 bar (20-psi) pressure
at full load and as much as 11.1 bar (170-psi) pressure at one-quarter load.
The boiler feed pump centerline is located 18.3 m (60 ft) below the waterlevel in the deaerating
heater. The friction losses in the suction piping are 0.20 bar (3 psi) at full load and 0.01 bar (0.2 psi)
at one-quarter load. Thus, the suction pressure at the boiler feed pump varies from 6.25 bar (90.6 psig)
at full load down to 2.17 bar (31.4 psig) at one-quarter load. The internal pressure on the suction side
of the pump varies the same.
On the discharge side, the internal pressure will be somewhat higher. In an open feedwater cycle
arranged with the feed pump taking its suction directly from the de aerator, the balancing device leakoff
is either returned to the deaerator (Fig. 10.3) or directly to the pump suction. Returning the leakoff to
the de aerator has a greater effect on the pump internal pressure, and so will be the arrangement considered
here. The calibrated orifice in this return line can be assumed to have a loss of 0.35 bar (5 psi) and the
return piping itself another 0.35 bar (5-psi) friction loss. Thus the internal pressure on the discharge side
will exceed that on the suction side by approximately 0.70 bar (10 psi) and will therefore range from
6.94 bar (100.6 psig) at full load to 2.86 bar (41.4 psig) at one-quarter load.
As the condensate pump is located 6.1 m (20 ft) below the boiler feed pump, the injection pressure
will range from 10.7 bar (155 psig) at full load to 13.8 bar (200 psig) at one-quarter load if we assume
friction losses of approximately 0.40 bar (6 psi) in the injection line. Thus the injection pressure will
exceed the internal pressures at all loads. If it is desirable to minimize both the amount of inward flow
and of the condensate being returned to the condenser and being repumped, the injection lines may well
be provided with control throttling valves. If it is desirable to maintain, let us say, a 0.35 bar (5-psi)
differential between the injection pressure and the internal pressure, the amounts of pressure to be
throttled will be the following:

1. At the suction side-4.44 (64.4 psi) at full load and 11.63 (168.6 psi) at one-quarter load
2. At the discharge side-3.75 (54.4 psi) at full load and 10.94 (158.6 psi) at one-quarter load.

There should be no difficulty in selecting control valves that will maintain the desired pressure differential,
and the rather high value of the pressure to be throttled will permit selection of a reasonably small valve
for the purpose.
Breakdown Seals 205

When the feed pump takes its suction from a booster pump (Fig. 10.4), the relationship between seal
and injection pressures becomes more complicated. The available injection pressure is that at the conden-
sate pump discharge: 11.7 bar (170 psig) at full load, and rising to 14.8 bar (215 psig) at one-quarter
load. With the deaerator located 10.7 m (35 ft) above the booster pump centerline, deaerator pressure
ranging from 4.8 bar (70 psig) at full load to 0.5 bar (7 psig) at one-quarter load, and the suction piping
losses 0.20 bar (3 psi) at full load and 0.01 bar (0.2 psi) at one-quarter load, the suction pressure at the
booster pump varies from 5.6 bar (81 psig) at full load down to 1.5 bar (21 psig) at one-quarter load.
For a booster pump with breakdown type seals, the injection pressure exceeds that at the seals by 5.1
bar (74 psi), at full load and 12.0 bar (174 psi) at one-quarter load after allowing 1.0 bar (15 psi) for
head loss in the injection and filter piping. As discussed under source of injection, pressure differentials
on this order dictate injection controls to avoid the risk of oxygen contamination of the feedwater.
For the conditions described, the suction pressure at the boiler feed pump ranges from 13.0 bar (188
psig) at full load down to 10.4 bar (151 psig) at one-quarter load. Since the suction pressure at full load
exceeds the available injection pressure, injection is not available, so some modification is necessary.
Two approaches are used: lower the pressure at the seals or increase the available injection pressure.
Lowering the pressure at the seals is accomplished by adding a pressure breakdown bushing at the
suction end of the pump, then returning the suction end bleed-off and the balancing leakoff to the
deaerator (Fig. 10.4). The return line has to include an antiflash orifice sized to ensure the back pressure
at the balancing device is at least 0.35 bar (5 psi) above the leakage vapor pressure under all operating
conditions. For the installation being used as an example, the pressure drop necessary is 1.4 bar (20
psi), the limiting condition being pump start-up with the deaerator at full load temperature. With this
arrangement, the pressures at the seals and the consequent control valve pressure drops are the following

Seal Load Pressure at Seal bar (psig) Valve Pressure Drop bar (psi)
Discharge Full 7.5 (108) (2.9) 42
Quarter 3.1 (45) (10.3) 150
Suction Full 7.1 (103) (3.2) 47
Quarter 6.8 (98) (10.7) 155

As the balancing device and suction end bleed-off bushing clearances increase with wear, the pressure
at the seals will increase by way of the higher pressure drop through the return piping and orifices. As
an illustration, consider taking the suction end bleed-off flow as 25 percent of the balancing leakoff
flow, an increase of approximately 30 percent in leakage (equivalent to only a 20 percent increase in
running clearance) will raise the pressure at the discharge end seal to 10.3 bar (150 psig). With the
pump in this condition, available injection pressure will be approximately equal to the pressure at the
seal during full load operation. Given that it is generally desirable to not have to restore the balancing
device clearance until it is worn to 150 percent of its new value, the system being used as an example
is marginal at best.
Increasing the injection pressure can be used either as an alternative to lowering the pressure at the
seals, or as a means of correcting a marginal design such as the example above. For the first case, the
balancing leakoff would be returned to the pump suction, typically with a total pressure drop of 0.35
bar (5 psi), all of it across the calibrated orifice. With the balancing device clearance worn to 150 percent
of its new value, the leakoff pressure drop will rise to 1.1 bar (16 psi). At full load, the resultant pressure
at the discharge end seal is 14.0 bar (204 psig). Allowing for injection piping and filter losses (1 bar or
15 psi total), 0.35 bar (5 psi) pressure drop across the injection valve, and a 0.35 bar (5 psi) margin of
injection over seal pressure, the minimum booster pump pressure rise is 4.1 bar (59 psi), at the flow
corresponding to worn seals.
206 Breakdown Seals

Evaluating the marginal "lower seal pressure" design for the same worn conditions yields a pressure
at the discharge end seal of 15.2 bar (221 psig). This is higher than the pressure with the balancing
leakoff returned to suction, and serves to illustrate an important point. When it is necessary to include
relatively high pressure drops to avoid flashing in a leakoff system, the effect of higher leakage flows
as the pump wears may well negate the advantage sought. For the example used, it would be necessary
to reduce the effect of increased leakage (by keeping most the back pressure constant with a regulator
or elevated flash tank) if injection booster pumps are to be avoided. If that is not feasible, the simpler
approach would be to return the balancing leakoff to suction and resort to booster pumps to increase
the injection pressure.

CLOSED CYCLE

When the breakdown-type seal is applied to a closed feedwater cycle, the conditions prevailing at the
discharge (balancing-device) end of the pump are actually more severe than at the suction end (Fig.
10.9). At full load and at some reduced load conditions, the injection pressure, PB, is greater than the

RETURN TO SUCTION INJECTION FROM CONDENSATE PUMP

Fig. 10.9 Section through injection breakdown seal at the discharge end.
KEY:
(A) Suction pressure plus pressure drop through calibrated orifice in balancing device leak-off line;
(B) discharge pressure of condensate pump, less friction loss through supply piping;
(C) essentially same as (B);
(D) essentially atmospheric pressure.
Breakdown Seals 207

pressure in the balancing-device relief chamber, P A, because of the friction losses between the discharge
of the condensate pump and the suction of the boiler feed pump. As the pressure, Po is essentially the
same as the pressure at the injection point, it also exceeds the pressure in the balancing-device relief
chamber, and flow takes place inwardly from point C to point A as well as outwardly from point C into
the collecting chamber, D.
As the load is reduced, the friction losses between the condensate pump and the main feed pump
decrease approximately with the square of the capacity. Thus at some extremely low pump loads, the
boiler feed pump suction pressure may be only 0.10-0.15 bar (lor 2 psi) lower than the condensate
pump discharge pressure (neglecting the static elevation difference, which is the same for both suction
and injection piping).
A typical example of what happens is shown in Fig. 10.2. At full load, there is ample excess pressure
in the injection line to produce flow into the pump. When the load is reduced to one-quarter flow, the
following situation prevails:

1. At the suction end, the suction pressure becomes 23.2 bar (337 psig) whereas the discharge pressure of the
condensate pump (less the static head) is 23.4 bar (340 psig). If the injection flow remains essentially
unchanged and the friction losses in the injection piping are still assumed to be 0.35 bar (5 psi), the available
injection pressure is reduced to 23.1 bar (335 psig), or 0.14 bar (2 psi) less than the suction pressure.
2. At the discharge end, the pressure in the balancing-device relief chamber is equal to the suction pressure
plus the loss through that calibrated orifice in the balancing relief line which is used to measure the leakage
past the balancing device. When the pump is new, this loss is approximately 0.35 bar (5 psi), and the pressure
in this chamber at full load is 17.6 bar (255 psig). The injection-line pressure is 19.7 bar (285 psig), and a
2.1 bar (3D-psi) differential is available to cause inward flow of injection water. At one-quarter load, the
pressure in the relief chamber is 23.6 bar (342 psig), whereas the injection-line pressure is only 23.1 bar
(335 psig), or 0.5 bar (7 psi) less than the relief chamber pressure.

If the pump is worn and the pressure drop through the calibrated orifice is permitted to go up to 2.4
bar (35 psi), the relief chamber pressure will become 19.7 bar (285 psig), balancing exactly the injection
line pressure. At one-quarter load, the relief chamber pressure becomes 25.7 bar (372 psig) whereas the
injection pressure is only 23.1 bar (335 psig). Thus no excess pressure exists in a worn pump between
the injection pressure and the relief chamber pressure under any load conditions.
What actually does take place, then, in the packless stuffing box under these conditions? As the
difference between pressures at C and A in Fig. 10.9 diminishes, less and less flow takes place inwardly
from point C to point A. Finally, a condition prevails in which the pressure at A slightly exceeds the
pressure at C. At that time, a small amount of feedwater flows from point A to point C and mixes with
the injection water. The mixture proceeds as before toward the collecting chamber, from which it is
returned to the condenser.
Because of the breakdown between points A and C and because the pressure differential between
them is only a fraction of the pressure differential between points C and D, the amount of this "reverse"
flow is relatively small and should not raise the temperature of the mixture appreciably. Thus, even
though the reduction of the pressure drop between the condensate pump discharge and the boiler feed
pump suction results in a change of flow direction between points C and A, the operation of the breakdown
seal remains acceptable.
This description of the flow process eliminates the effects of the balancing device on the pressure at
point A for the sake of simplicity. Actually, this device develops a pumping action that leaves the pressure
at the shaft sleeve near point A some 1.4 to 2.1 bar (20 to 30 psi) below the pressure at the periphery
of the relief chamber in pumps running at 3,570 rpm, on the order of 3.5 to 5.2 bar (50 to 75 psi) in
208 Breakdown Seals

INJECTION PRESSURE-BALANCING RELIEF


CHAMBER PRESSURE, PSI
-80 -70 -60 -50 -40 -30 -20 -10 0

PUMP AT

~
3~70 RPM
or .....-
~ rJ.....}- 1--'+
-
~
V ~ r-- PUMP STATIONARY
IL. ",... lL ~
0 ,..,.-
-
C)
a: ~

--
~ ~
~ -2
IL.

i-4
i""""'"
--- ~
~
-

!
""-6

~ -B
""
o 10 20 30 40 50 60 70 80
INJECTION PRESSURE-BALANCING RELIEF
CHAMBER PRESSURE, PSI

Figo 10.10 Effect of pump rotation on injection flow.

high-speed pumps. Figure lD.lD which shows the inward leakage (from points C to A) under varying
pressure differences between injection line and the relief chamber, graphically illustrates this condition.
Paradoxically, some flow still takes place inwardly, even though the pressure difference is negative, as
long as the pump is running. When the pump is idle, of course, the condition disappears.
Of course, if a continuous flow of injection water from point C to point A is desired regardless of
load, a small booster pump should be installed in the injection line, taking its suction from the condensate
pump discharge and raising that pressure by some 3.5 bar (50 psi). It should be noted that no flow takes
place through the balancing device when the pump is idle and that the inward pressures at the suction
and discharge sides are essentially equal. Thus there should be no problem in maintaining inward sealing
flow to a pump kept idle on standby service.
The fundamental principle of the condensate injection seals applied so successfully to boiler feed
pumps, a pressure breakdown bushing to control leakage, is not restricted to boiler feed pumps. In
applications where the leakage can be readily disposed of, a breakdown-type seal often represents a low-
maintenance solution to sealing the shaft. This typically occurs when the pumped liquid is below its
atmospheric boiling point and free of large, abrasive solids, thus removing the need for any injection
into the seal, which simplifies the installation. Pumps applied to water injection, hydraulic descaling,
and hydraulic presses are common examples of such a choice of shaft seal.
A further use of breakdown seals is in a form of "sealless" pump, one in which the leakage is returned
directly to the pump's suction reservoir (see Chap. 14).
Breakdown Seals 209

MECHANICAL MODIFICATIONS

The compromise that has to be struck in the design of a breakdown seal is between running clearance
and leakage. Make the clearance too small and incidental contact between the sleeve and bushing will
result in rapid wear and a commensurate increase in leakage. With a solid bushing, the clearance is that
necessary to accommodate minor eccentricity and runout from manufacturing variations, and small

Fig. 10.11 Floating seal ring design.


210 Breakdown Seals

changes in clearance and alignment during pump operation. One solution advanced to allow smaller
effective clearances without the risk of premature wear is the so-called .floating ring seal (Fig. 10.11).
The essential feature of this design is to make the seal radially "flexible." This is achieved, in theory,
by building up the seal with a series of alternate rings: seal rings having a close clearance over the
sleeve but free to "float" radially within the seal housing; spacer rings accurately located in the seal
housing and with a large clearance over the sleeve. All the rings are prevented from rotating by some
form of pin-and-slot arrangement. In practice, floating ring seals have not functioned as well as the design
promised. Although there were, and still are, examples of well-behaved seals, the general experience was
poor; the seals typically wore at a higher rate than comparable solid bushings. The principal difficulty
appeared to be limited radial flexibility, a result of not achieving sufficient axial force balance on each
seal ring to allow ready radial movement. But for isolated installations, floating ring seals are no longer
used in boiler feed pumps. With axial face (mechanical) seals demonstrating adequate capability for
most services, there is little incentive to refine the design until it works as expected.
11
Bearings
------------~-~-----~--~- ... - ~~~- -.-~~--- - - - - - -

Centrifugal pumps require bearings to allow the rotor to tum while maintaining correct alignment between
the rotor and the pump's stationary parts under the action of radial and axial forces. Their design,
therefore, is as critical as any of the other major pump parts; a pump with unreliable bearings is a
continual cause of unnecessary outages and maintenance expense.
The provision of adequate bearings is not an easy task. Doing so is the pump designer's responsibility,
but some understanding of the process can aid equipment evaluation and trouble shooting, should that
be necessary. To start the process, the designer first determines accurately the bearing loads over the
pump's expected operating flow range. These data are supplemented with any unusual environmental
conditions that could influence the design, for example, high pumping temperature or pumps that are
exposed to wind-driven dust or frequent hosing down. Using these data, the designer then adopts a
sequence similar to the following to develop the design:

1. Select bearing type and size.


2. Select means of lubrication.
3. Check heat load and dissipation; provide supplementary means of dissipation if necessary.
4. Verify stiffness of bearing housing to avoid both bearing misalignment and resonant vibration.
5. Check the effect of bearing and bearing housing stiffness on the dynamic behavior of the pump's rotor.
6. Select means of sealing the bearing housing.

Each of these steps involves the consideration of several factors, all of which are dealt with in the
relevant sections of this chapter.

BEARING ARRANGEMENTS

Most centrifugal pumps are equipped with external oil- or grease-lubricated bearings in a classical two-
bearing arrangement. The bearing providing radial location is called a line bearing; that providing axial

211

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
212 Bearings

location, a thrust bearing. To realize the two-bearing arrangement, many designs have the thrust bearing
providing both radial and axial location. Others incorporate a radial and thrust bearing in one housing,
which is usually designated as the thrust bearing.
Overhung pumps, those with cantilever rotors (Fig. 11.1 [aD, have both the line and thrust bearings
on one side of the impeller. By convention the bearing nearer the impeller is designated the inboard
bearing and that farther away the outboard bearing. Generally, the outboard bearing is the thrust bearing
to better balance the total bearing loads. In pumps with the impeller or impellers between the bearings,
a simply supported rotor (Fig. 11.1 [bD, convention has the bearing adjacent to the coupling designated
the inboard bearing, that at the opposite end of the pump the outboard bearing. Again, usual practice is
to have the outboard bearing the thrust bearing to allow an optimum bearing selection (a thrust bearing
at the inboard location would be influenced by the shaft size necessary to transmit the pump's torque).

-·-- L1 - ----L2-
Fr 1
Inboard Outboard

- Coupling

- - - RT

(a) Overhung (cantilever) rotor

- L2 --··l. . >-~-r-- L1 -

Outboard Inboard
I I

Coupling

(b) Between bearings (simply supported) rotor

Fig. 11.1 Basic rotor configurations: (a) overhung (cantilever), (b) between bearings (simply supported).
Bearings 213

In modem pump designs, the bearings are almost invariably mounted in housings attached directly to
the pump casing. Occasionally the bearings for very large vertical and horizontal pumps are mounted
in housings attached to the pump's baseplate or foundation. In the case of horizontal pumps, this
arrangement is known as "pedestal bearings."
In many centrifugal pump configurations, the classical two-bearing arrangement cannot provide
sufficient rotor support to maintain internal alignment. This is primarily a consequence of low rotor
stiffness, and necessitates a multiple-bearing arrangement, with some of the bearings usually located
within the pump proper. These internal bearings pose a special set of problems (see internal bearings).
For a given set of imposed loads, radial and axial at the impeller, bearing loads vary significantly
with rotor configuration and geometry. To illustrate the point, consider the rotors shown in Fig. 11.1
are subject to a maximum radial load, F" of 2,200 N (500 lb) and an axial thrust, Fa, of 4,400 N (1,000
lb). Applying these loads to two versions of a cantilever rotor and one of a simply supported rotor
produces the bearing reactions, RL and Rr , set out in the table.

RT-N
Rotor LI-mm L2 RL Radial Axial Total

Cantilever 254 254 4,440 2,220 4,440 4,970


Cantilever 191 125 5,550 3,330 4,440 5,550
Simply supported 254 254 1,110 1,110 4,440 4,580
RT-lb
Rotor LI-in L2 RL Radial Axial Total

Cantilever 10 10 1,000 500 1,000 1,118


Cantilever 7.5 5 1,250 750 1,000 1,250
Simply supported 10 10 250 250 1,000 1,031

Two significant points need to be drawn from these results. First, although the second cantilever rotor
is, for the same diameters, notably stiffer than the first (65 percent less deflection; see shafts in Chap.
7), both the bearing reactions are higher, 25 percent at the line bearing, 12 percent at the thrust bearing.
For these two-bearing frames to have the same nominal bearing life, the bearings on the stiffer shaft
would have to have correspondingly higher basic capacities (see bearing life in this chapter). If rotor
geometry is not taken into account when evaluating cantilever bearing frames, the evaluation can be
quite misleading.
Second, changing to a simply supported rotor results in lower bearing reactions than even the best
case cantilever rotor, 75 percent lower at the line bearing and 8 percent lower at the thrust bearing. (The
lesser reduction at the thrust bearing is a product of the high axial thrust chosen for this example).
Evidently it is easier to achieve high bearing integrity (low bearing loads relative to basic capacity) with
a simply supported rotor than with a cantilever rotor. In small, low duty pumps (single stage, 10 to
12-in. discharge), the cost advantage of a simply supported rotor typically does not offset the higher
costs of other aspects of a between-bearings pump (heavier, more complex casing, additional shaft seal,
and bearing housing). For heavier duty single- and almost all multistage pumps, between bearings is the
normal configuration. In some services, however, such as sewage and slurry, the needs of the liquid end
design dictate overhung construction regardless of size. Very large pumps, those used for water transfer
and pumped storage, are overhung when vertical, the usual arrangement of modem designs, and between
bearings when horizontal.
214 Bearings

Fig. 11.2 Internal sleeve bearing.

BEARING TYPES

Centrifugal pumps employ a wide variety of bearing fonns, ranging from simple internal sleeve bearings
(Fig. 11.2) in low-cost water pumps to electronically controlled magnetic bearings in sophisticated high-
energy pumps. All of the fonns, however, can be classified as one of three types, based on their method
of operation:

1. Antifriction or rolling element


2. Plain; either boundary lubricated, hydrodynamic, or hydrostatic
3. Magnetic

ANTIFRICTION BEARINGS
Relative motion within the bearing is through the rolling action of balls or rollers between an inner and
outer race (Fig. 11.3). The tenn "antifriction" derives from the very low coefficient of friction associated
with rolling motion as distinct from sliding motion. To function correctly, the rolling elements in
antifriction bearings must be kept equally distributed around the bearing. This is achieved with a separator
or cage (Fig. 11.3).
Although the fundamental motion in antifriction bearings is rolling there is nevertheless some sliding
present. First, the rolling elements are subject to dynamic action (acceleration and deceleration) as they
pass into and out of the loaded zone of the bearing, or as the rotor's speed is varied during the course
of operation. Second, there is defonnation at the points of contact between rolling elements and races,
the consequent "imperfect geometry" resulting in local sliding motion. Third, the cage or separator comes
into contact with the rolling elements in the course of keeping them in position. To minimize the heat
generated at these points of sliding friction, antifriction bearings require lubrication.
Bearings 215

, . . . - - - - - - - Outer Ring Raceway

Cage ---_~~ r - - - - - - - Outer Ring Land


Inner Ring ---1tf);.~=--.)
----jL-_ _ _ _ _ Bore Surface
,...,J...J::::;;;T-=;----- Inner Ring Raceway
+---- Inner Ring Land

Outer Ring ---:".-c-+- ' - I-


-+- - - - 00 Surface
Side Faces

Figure 11.3 Rolling-element bearing.


(Courtesy SKF Industries, Inc.)

TYPES AND APPLICATIONS

There are many types of antifriction bearings. To make assessment a little easier, it is common to
differentiate between "ball" and "roller" bearings. Of these two, ball bearings are the more common in
centrifugal pump practice, since they offer an excellent compromise between capability and cost. Roller
bearings have greater capability, but are more expensive and so are only used when the higher cost
is justified.
Most centrifugal pumps are equipped with single-row deep-groove ball, or Conrad, bearings (Fig.
11.4). This configuration can accommodate both radial and axial loads, the latter in either direction, has
reasonable speed capability, and can tolerate misalignment up to 0.25 deg. An alternate design with a
filling slot in the inner race allows more rolling elements, which increases the radial load capacity by
20 to 40 percent, but reduces thrust capacity and tolerance of misalignment by approximately 70 percent.
When used at all, filling slot bearings should be used in an arrangement which allows only radial loads.
Some industry specifications preclude their use regardless of loading.
For loads beyond the capacity of single row ball bearings, the usual next choice is a double-row
arrangement (Fig. 11.5). Although the radial load capacity is approximately 50 percent higher, the
maximum speed is lower, a consequence of the greater number of rolling elements within the bearing.
Thrust capacity and tolerance of misalignment depend on the detail design of the races. Designs with
the ball load paths converging outward (usual) generally have filling slots, hence low thrust capacity
and can tolerate only 0.05 deg misalignment. Those bearings with the ball load paths converging inward
(less usual) have higher thrust capacity and can tolerate more misalignment, but in no case more than
a single-row bearing. These limitations in speed, thrust capacity and misalignment often dictate the
choice of another bearing type.
If the designer foresees significant misalignment at the bearing (up to 2.5 deg, which in these days
when the advantages of relatively stiff shafts are well understood should be the exception), a self-aligning
ball bearing is available (Fig. 11.6). The bearing is double row to achieve a radial load capacity of
216 Bearings

Figure 11.4 Single-row deep-groove ball bearing.


(Courtesy SKF Industries, Inc.)
Figure 11.5 Double-row deep-groove ball bearing.
(Courtesy SKF Industries, Inc.)

Figure 11.6 Self-aligning double-row ball bearing.


(Courtesy SKF Industries, Inc.)

Figure 11.7 Single-row angular-contact ball bearing.


(Courtesy New Departure.)

approximately 70 percent of the comparable single-row deep-groove bearing. With a spherical bore in
the outer race, the thrust load capacity of self-aligning ball bearings is negligible, and they should be
arranged for radial loading only.
As the ratio of axial to radial load increases, it becomes necessary to introduce some matching
angularity into the ball load paths. Such bearings are tenned angular contact and are available with
contact angles ranging from 15 to 40 deg, although for high-speed angles above 30 deg are not recom-
mended. Single-row angular-contact bearings (Fig. 11.7) have approximately the same radial load capacity
as comparable deep-groove bearings while offering up to 2.3 times the thrust capacity but in one direction
only. Misalignment tolerance is limited to 0.03 deg. Given their unidirectional thrust capacity, single-
row angular-contact bearings are used in centrifugal pumps only when it is certain the applied thrust
will always be in one direction, such as it generally is in vertical turbine pumps.
Bearings 217

Figure 11.8 Two single-row (duplex) angular-contact Figure 11.9 Two single row (duplex) angular-contact
bearings mounted back-to-back. bearings mounted face-to-face.
(Courtesy New Departure.) (Courtesy SKF Industries, Inc.)

Most centrifugal pumps exhibit axial thrust in both directions, the direction generally changing with
operating flow or pump condition or both. When the load in either direction requires an angular contact
bearing, the usual practice is to use a duplex arrangement of "ground face" or "matched" single-row
bearings to provide bidirectional thrust capacity. There are two possible arrangements, usually known
as mounts: "back-to-back" (Fig. 11.8) and "face-to-face" (Fig. 11.9). Several aspects of the application
of duplex angular-contact bearings are currently quite controversial, and so warrant discussion in
detail.

1. Mounting-From strict geometric considerations, duplex angular-contact bearings as applied to most centrifu-
gal pumps should be mounted face-to-face. With the ball load paths converging inward, the bearing assembly
offers the lower resistance to moments and so approaches the point support condition assumed in its loading
calculations. There is a body of opinion opposing this on the grounds that the higher moment capability of
back-to-back mounting contributes to rotor stiffness (which it undoubtedly does in the case of low stiffness
shafts but at considerable expense to bearing loading). Ease of assembly and handling favors back-to-back
mounting since it is not possible to accidentally dislodge the outer races. Both mountings tend to "pump"
lubricant, back to back from the center out, face to face in toward the center. With back-to-back mounting
it is therefore ideal if the lubricant is introduced between the bearings, which requires accurately matched
spacers for the inner and outer races. In the absence of this complication, it is often necessary to use an oil
ring at each side of the assembly to ensure enough lubricant finds its way into the bearings. Face-to-face
mounting requires care to ensure an excess of lubricant is not retained within the bearing leading to churning
and subsequent overheating.
218 Bearings

2. Preload-In applications involving high unidirectional thrust and pumps with duplex 40 degree angular-
contact bearings, there have been instances where skidding between the balls and races in the unloaded
bearing has caused premature failure. One remedy advanced to cure this problem has been to mount the
bearings with a "light" preload. The difficulty with this requirement is that component tolerances involved
in the bearing assembly produce a far wider range of installed bearing preload than was originally intended.
If the preload is too high, the additional loading from thermal expansion as the bearing warms up produces
loads high enough to cause immediate failure. Making the inner race a slide fit avoids much of this difficulty,
but poses the risk of fretting corrosion of the shaft under the bearing. Unless the pump is known to be
subject to high thrust in one direction, there is now general agreement that duplex angular-contact bearings
should not be preloaded.
3. Contact angle-Past practice has been, and some industry standards (e.g., API 610) have mandated it, to
have both bearings of a duplex arrangement equal, the logic being that the assembly cannot be made back
to front. In an attempt to overcome the problem that pre loading sought to correct, some bearing manufacturers
are offering duplex arrangements in which the nominally unloaded bearing is a 15-deg contact angle. In
this way, the bearing assembly offers bidirectional thrust capacity, although of unequal values, with the
lower capacity bearing better suited to running unloaded. Field experience suggests the arrangement realizes
its claims, the only qualification being that the assembly must be installed in the correct direction.

For thrust loads beyond the capacity of a single angular-contact bearing, there has been a practice of
using duplex angular-contact bearings mounted in tandem (Fig. 11.10) to provide higher unidirectional
capacity. This rather complex arrangement, usually involving three bearings in total for bidirectional
thrust capability, is particularly difficult to lubricate and is now rarely used. Modem practice is to employ
a single bearing of inherently higher capacity, for example, a tapered roller (see later).
Double-row angular-contact bearings (Fig. 11.11) are nominally the equivalent of a duplex pair. Their
usual arrangement is with ball load paths converging outward or back-to-back. Basic bearing capacity

Figure 11.10 Two single-row (duplex) angular Figure 11.11 Double-row angular-contact
contact bearings mounted in tandem. ball bearing.
(Courtesy SKF Industries, Inc.) (Courtesy New Departure)
Bearings 219

Figure 11.12 Cylindrical roller bearing (without end rings).


(Courtesy SKF Industries, Inc.)

Figure 11.13 Double-row spherical roller bearing.


(Courtesy SKF Industries, Inc.)

is approximately 50 percent higher than that of the same size single-row bearing. Thrust capacity,
however, is reduced in one direction by the presence of a filling slot, needed to allow bearing assembly.
Given this limitation, double-row angular-contact bearings are used only when the thrust loading is low
and incidental. Some industry specifications preclude their use regardless of loading.
Roller bearings generally have higher basic capacities than ball bearings of equal size. Except for
one type, cylindrical or barrel roller, roller bearings have lower speed limits than equal size ball bearings,
and in all cases they are more expensive.
Cylindrical roller bearings (Fig. 11.12) are intended for high radial loads and high speeds. The
standard bearing does not provide any location between the inner and outer race, so it is necessary to
positively locate the outer race in the bearing housing. The inner race typically is mounted with an
interference fit and located against a shoulder on the shaft. Inner-races with locating rings are available,
but in centrifugal pump practice these additional points of sliding friction are usually avoided. Although
termed "cylinders," the rolling elements are actually barreled slightly (relieved at each end) to avoid
stress concentration at the ends of the rollers. Even with this refinement, allowable misalignment is only
0.08 deg, meaning that shaft stiffness has to be particularly high to ensure the bearing functions as
intended. Care is needed in bearing and related component design to avoid premature failure caused by
roller skidding in lightly loaded, high-speed bearings. European pump designs employ cylindrical roller
bearings quite extensively. As of this writing, the opposite is the case in the United States.
For high loads at low to moderate speeds, spherical roller bearings are frequently the only feasible
antifriction bearing selection. In centrifugal pump practice, the double-row convex configuration (Fig.
11.13) has been used quite extensively as the line and thrust bearing when the expected axial thrust was
220 Bearings

low, and as the line bearing only when high axial thrust was expected. This usage derives from the
bearing's geometry; with the axis of the rollers inclined at a small angle from the shaft axis, thrust
capacity is low, typically on the order of 25 to 30 percent of radial load capacity. With a spherical bore
in the outer race, tolerance of misalignment is high, 1.5 deg being the limit normally quoted. Recent
advances in bearing geometry have increased speed capability. Applications at the upper end of published
speed limits need care with lubrication (see lubrication in this chapter) to ensure overheating is not
a problem.
Thrust loads beyond the capacity of angular contact bearings can be accommodated by a spherical
roller thrust bearing (Figure 11.14) provided the speed is low. Two roller designs are used: symmetric
and asymmetric, the latter offering 50 percent higher bearing capacity. Misalignment tolerance is quoted
at 3.0 deg. Radial load capacity is limited, again a function of the bearing's geometry; therefore if the
loading includes a high radial component, a separate radial bearing is used. Typical applications of
spherical roller thrust bearings have been large vertical wet pit pumps, and large sealed (as distinct from
hermetically sealed or "sealless") boiler circulating pumps.
Tapered roller bearings accommodate both radial and axial loads. Single-row bearings (Fig. 11.15)
typically have thrust capacity on the order of 60 percent of radial load capacity, but this does vary with
the design of the bearing (inclination of the roller axes), and is significantly higher in bearings designed
for high thrust. Except in the case of thrust bearings subjected to a high unidirectional load, tapered
roller bearings must be used in pairs, either separate (Fig. 2.19) or duplex (Fig. 11.16). This is necessary
to maintain the bearings in their correct position and to resist the axial reaction that is produced when
radial load is applied to one of the bearings. (Note that when the axial reaction is opposed by a higher

~~'~
!f\ ,\--
------. ..... t~

\. ~
~".. .-
~
:1_ -
I

~-~~
J.~
----..;;;;;;:

Figure 11.14 Spherical roller thrust bearing. Figure 11.15 Single row tapered roller bearing.
(Courtesy SKF Industries, Inc.) (Courtesy Timken, Inc.)
Bearings 221

(a) (b)

Figure 11.16 Duplex tapered roller bearings:


(a) mounted back-to-back (b) mounted face-to-face.
(Courtesy Timken, Inc.)

external thrust, only one bearing of the pair has a resultant axial load.) Tapered roller bearings have
limited tolerance of misalignment, therefore designs using separate bearings require high-stiffness shafts
and accurately aligned housing bores. Such designs also require provision for adjustment to set the
bearings at the correct "cold" or "bench" and play to allow for thermal expansion as the assembly warms
to operating temperature.
In centrifugal pumps, duplex tapered roller bearings are frequently resorted to when the loading and
speed exceed the capabilities of duplex angular contact bearings. Of the two arrangements face-to-face
is preferable because it more closely approached point support and is therefore less susceptible to
extraneous loads from minor misalignment. As with separate bearings, duplex assemblies must have the
correct "bench" end play when cold. To reduce the risk of incorrect adjustment, duplex assemblies are
generally furnished preset with a precision spacer between the cups or cones depending on the mounting
arrangement. As tapered roller bearings are loaded, the taper of the rolling elements results in a small
force tending to push the rollers out from between the races. This force is resisted by the cone rib. As
the speed of the bearing increased, lubrication at this point of sliding friction becomes critical (see
lubrication in this chapter).
222 Bearings

ANTIFRICTION BEARING LIFE

When correctly installed, operated, and maintained, antifriction bearings ultimately fail by fatigue of
their rolling elements and receways. Out of a great deal of testing, the bearing manufacturers collectively
have developed standard equations to determine the basic capacity, designated C, for the various bearing
types. Because the mode of failure is fatigue, there is inherently some spread in the test results, so the
basic capacity has a probability assigned to it. For regular machine applications such as centrifugal
pumps, the basic capacity of a bearing is that load at which 90 percent of production will meet or exceed
"standard" life.
The service life in a particular application is calculated from the ratio of the bearing's basic capacity
to the equivalent bearing load, P. To arrive at the equivalent bearing load, the designer first has to
determine the actual axial and radial loads the bearing will carry. These loads, designated Fa and Fro
respectively, are converted to an equivalent load using the equation

P =XFa + YFr

where X and Yare factors dependent on the bearing geometry, and in some instances the nature of the
loading (i.e., the ratio F alFr). The factors are obtained from the bearing design manual.
With the equivalent bearing load, P, and the bearing's basic capacity, C, the designer calculates the
bearing life, designated LIO for 10 percent probability of premature failure, using the equation

LIO = (c/p)n

where the bearing life is in millions of revolutions, and the exponent n is

3 for ball bearings


10/3 for roller bearings

reflecting the difference in stress distribution in the two bearing types.


Research into bearing fatigue life is a continual process, and as a result some manufacturers include
in their design manuals various "life correction" factors. Generally, these take account of actual lubrication
conditions, which can both increase and decrease the theoretical life, and the level of impurities in the
bearing materials.

LUBRICATION OF ANTIFRICTION BEARINGS

In the layout of a line of centrifugal pumps, the choice of the method of bearing lubrication is determined
by application requirements, cost considerations, and in many instances the requirements of the principal
market for the particular pump type.
Application requirements is a broad term, encompassing both the functions the lubricant has to perform
for the particular bearing type, and the dictates of the service for which the pump is to be used. For
example, the application requirements of a low-speed plain bronze lineshaft bushing in a cold-water
pump are distinctly different from those of a high-speed antifriction bearing in a high-temperature
hydrocarbon pump.
For antifriction bearings, the principal functions of the lubricant are to minimize the heat generated at
the various points of sliding friction in the bearing, and to provide elastohydrodynamic lubrication at the
points of true rolling contact. At the same time the lubricant has to serve as a heat transfer medium, protect
the highly finished bearing surfaces from corrosion, and help maintain the bearing housing environment.
Bearings 223

Speaking broadly, antifriction bearings are almost exclusively lubricated with either grease or oil, the
latter by at least five distinct methods. This derives directly from the need to protect the bearing
components from corrosion and abrasive solids. Only in rare instances, hermetically sealed cryogenic
pumps for example, is the nature of the pumped liquid such that it can be used to lubricate antifriction
bearings. And in those rare instances, the bearings are generally of special materials needed for the
service environment.
Grease lubrication (Fig. 11.17) offers simple design, low cost, and reliable operation, and is therefore
widely used in centrifugal pumps. It is ideally suited to low-speed pumps, used either intermittently and
so needing extra protection against bearing corrosion, or in severe service environments, such as mining,
where there is a critical need to exclude the atmosphere from the bearing housing. Vertical pumps are
often grease lubricated because it is easier to prevent leakage of the lubricant down the shaft (Fig. 11.18).

CENTERtNG
JACK-SCREW

Figure 11.17 Typical mounting of ball bearings in


double-suction pump. Pump uses grease lubrication; Figure 11.18 Ball-bearing construction with seal in
the vertical jack-screw is for initial centering of the vertical pump.
rotor in its wearing rings. Seal guards against escape of grease.
224 Bearings

Grease is a suspension of oil in a carrier, usually a metallic soap, and provides lubrication by the
moving parts of the bearing becoming coated with oil as they contact the grease. Because there is
essentially no flow of lubricant through or around the bearing, heat dissipation is limited, and so the
maximum speed at which the bearing can be operated is also limited, generally to some 70 percent of
the limit for oil bath lubrication. The negligible circulation of lubricant has a second effect; it is necessary
to periodically remove the spent grease from the bearing, replacing it with fresh grease to relubricate
the bearing. How frequently this must be done depends on the type of bearing, its size, and the speed
at which it is being operated (see antifriction bearing maintenance).
The need for relubrication imposes some special requirements on the design of housings for grease
lubricated bearings. During relubrication the spent grease must be completely removed from the bearing.
For convenience, this is usually done by arranging the housing to allow fresh grease to purge the bearing
of spent grease, and providing a means of accommodating or expelling the spent grease. Figure 11.17
shows a housing designed to accommodate spent grease (although it will need to be cleaned out at major
overhauls). Expelling spent grease is accomplished with either a normally plugged drain connection or
an automatic grease relief valve (Fig. 11.19).
Grease is classified by stiffness, dropping point (temperature at which the oil separates from the
grease), and carrier compound. The National Lubricating Grease Institute lists nine stiffness grades, 000
to 6, based on tests by ASTM method D217, and there are at least nine generally recognized carrier
compounds, each having a particular dropping point and other characteristics such as moisture resistance,
oxidation resistance, behavior at high temperature, and so on. The choice of grease for a particular
application depends on bearing size, operating speed, operating temperature, and the pump's operating
environment, with high ambient humidity or wind-driven dust dictating special care.
For antifriction bearings a grease of #2 consistency provides a good balance between oil feeding
capability and resistance to churning within the bearing. Larger bearings, those with unsupported grease
depths greater than 1.2 in., usually require a grease of #3 consistency to avoid having the grease slump
into the bearing. Lithium soap is the most widely used carrier, offering a working temperature range of
-29 to 121°C (-20 to +250°F), very good resistance to oxidation, good corrosion protection (with
additives), and good water resistance. For bearing service conditions outside the range of lithium soap
greases, it is best to consult a lubricant manufacturer.

Figure 11.19 Automatic grease relief valve.


Bearings 225

LINE BEAR ING


HOU5 INC

Figure 11.20 Oil bath lubrication with constant-level oiler.

Oil lubrication overcomes the speed limitation inherent in grease lubrication, but at the expense of
added complexity. Just how much added complexity depends on the precise method of oil lubrication,
a choice made on the basis of expected or actual operating conditions.
Oil bath (Fig. 11.20) is the simplest of the methods of oil lubrication. Oil is maintained at a level
nominally 0.3 to 0.5d above the bottom of the lowest rolling element (where d is the diameter of the
rolling element). Lubrication is effected by the passage of the rolling elements thru the oil bath. A
constant level oiler (Fig. 11.20) is the usual means of maintaining oil level. In applications where the
bottle of a constant level oiler is likely to be broken, the bearing housing is provided with a sight gage
to indicate the actual level.
Adding a flinger contributes to oil lubrication in one of two ways. In the simpler arrangement, the
flinger, usually a "lobed" shape, creates an oil mist within the bearing housing, and so increases the
range over which the oil level can vary without impairing lubrication. The more complicated arrangement
uses the flinger to increase the rate of oil circulation in the housing beyond that achieved with oil bath
lubrication. This is necessary as pump operating speeds or pumping temperatures or both increase. In
the usual arrangement (Fig. 11.21) the circular flinger picks up oil from the housing sump and throws
it against the housing wall. The oil then drains into feed canals in the housing, which convey the cooled
oil to the bearing.
As the operating speed of antifriction bearings approaches its upper limit, most of the heat generated
by the bearing is a result of the rolling elements churning the lubricant. To avoid overheating under
these conditions, it is necessary to provide only just as much oil as the bearing needs for lubrication
and heat dissipation. Of the three methods used to do this, oil-ring lubrication is the simplest. In the
classical arrangement (Fig. 11.22), the oil level in the bearing housing is set below the bearings. Oil is
picked up from the sump by an oil ring, and transferred to the bearing by what is effectively "splash
lubrication." An alternative design, necessary for large bearings running at or close to their maximum
speed, is to convey oil to the bearing using feed canals as already discussed for flinger lubrication.
The second method is pure oil mist lubrication. With this approach, the bearing housing is set up for
226 Bearings

Figure 11.21 Flinger lubrication.


Bearing bracket integral with housing; bearing housing jacketed for cooling.

dry sump operation (Le., no oil is retained in the housing), and oil mist is supplied to the housing from
a centralized mist generator (Fig. 11.23). The bearing housing (Fig. 11.24) is designed to ensure the oil mist
passes through the bearing, thereby providing lubrication and cooling. Most of the oil mist accumulates in
the housing (up to 10 percent can be lost to the atmosphere through vents and seals), where it condenses
and is drained out to a collection vessel. Although dependent on an external system, a factor that often
reduces reliability, pure oil mist lubrication has yielded a distinct improvement in bearing lives in refinery
pumps. In many instances, its use has led to simpler overall installations by allowing high-temperature
pumps to operate without cooling water.
Purge oil mist is related to pure oil mist lubrication, but only in that it uses mist supplied from a
central system. The function of purge oil mist is to exclude the atmosphere from the bearing housing.
Lubrication of the bearing is by oil bath, flinger, or oil ring, as already described. The housing differs
from the normal arrangement only by the addition of an overflow drain (Fig. 11.25) to remove the
condensed mist. Purge oil mist has seen great usage in high-humidity environments, where bearing
housings vented to the atmosphere are prone to water contamination of the oil, a circumstance that can
lead to a significant reduction in bearing life (see Armstrong et al. [11.1]).
Forced circulation is used when the service conditions, load, speed, or heat load from the process
require heat dissipation beyond the capacity of pure oil mist. Filtered and cooled oil is supplied to the
bearing housing from a small lubricating oil system, then directed to the bearing's rolling elements by
drilled ports or spray nozzles (Fig. 11.26). After passing through the bearing, the oil is drained from the
housing back to the lubricating oil system. In most cases for centrifugal pumps, the bearing housing is
arranged for dry sump operation, since high speed is generally a factor in the application.
Bearings 227

Figure 11.22 Oil-ring lubrication.


Bearing bracket integral with casing cover; the bearing housing is fan cooled.

G Air supply Dr----+--.&. . .---+ ...... .....,..-~

~ ---- -----t- ---

IReservoir~
I

I
....._.--~
I
I
: Oil mist console
I
1______ - - - - - - - - - - - - - - - - - - -

MR MR MR MR MR SR MR MR SR MR MR MR
MR = Standard mist reclassifier, for light to moderately loaded antifriction bearings. Also used for purge mist
applications.
SR = Directed mist reclassifier. For heavily loaded antifriction bearings and bearings with pitch line velocities
over 10.2 mls (2000 ftlmin).

Figure 11.23 Central system for oil-mist lubrication.


(Courtesy Bloch)
228 Bearings

PURE-MIST (DRY SUMP)

LABYRINTH
O IL RINGS AND
SEAL
FLINGERS REMOVED

MIST-ALL"
DRAIN ASSEMBLY ,/
WITH OVERFLOW VENT

Figure 11.24 Bearing housing with pure oil mist lubrication.


(Courtesy Alemite Division of Stewart Warner Corporation.)

PURGE- MIST (WET SUMP)

d MIST·ALL"
PURGE·M IST
ASSEMBLY
SPRAY FITTING'.,.J-_ _--.

CONSTANT
LEVEL
OILER

MIST·ALL"
/-
OIL DRAIN A$SEMBL Y
LEVEL WITHOUT VENT

Figure 11.25 Bearing housing with purge mist.


(Courtesy Alemite Division of Stewart Warner Corporation.)
Bearings 229

OUTLET FOR
WETSUMP -
(STATIC LEVEL)
MUST BE AT THIS
LEVEL ON ONE
-orr::J.:..:.=.~ . OR BOTH SIDES
~~;1;P"

OIL OUTLET
FOR DRY SUMP -
BOTH SIDES

Figure 11.26 Forced lubrication of anti-friction bearing.


(Courtesy SKF Industries, Inc.)

Lubricating oils are rated by viscosity, viscosity index (a measure of the change in viscosity with
temperature), pour point (lowest temperature at which the oil will pour), flash point (temperature at
which oil gives off flammable vapors), oxidation resistance, and special properties such as extreme
pressure capability or foaming resistance imparted by additives. Of these properties, viscosity is the
most important. The viscosity required for a particular application depends on the bearing's size and
speed, larger bearings running at lower speeds requiring higher viscosities (see Fig. 11.27). Note that
the viscosity determined from Fig. 11.27 is that at the oil's operating temperature. This is typically 3
to 5°e (5 to 20°F) above that of the bearing housing surface. Oil viscosity ratings are quoted today for
a "standard" temperature, commonly 40 0 e or 104°F, therefore a second chart is required to convert the
viscosity at operating temperature to a viscosity rating. Figure 11.28 is one such chart.
As noted in Fig. 11.27, the oil viscosity derived from it is the minimum acceptable. Within limits,
higher bearing lives can be achieved by using a higher viscosity. The limiting factor is increased heat
generated by churning of the more viscous lubricant within the bearing.
Generally solvent refined mineral oils are most satisfactory for temperatures up to 82°e (180°F). At
higher temperatures the oxidation rate (which doubles for each lOoe (18°F) temperature rise) becomes
too high, and if the oil cannot be cooled it is necessarj to resort to synthetic oils. In applications where
very low ambient temperatures are likely, it is important to ensure the pour point of the oil is above the
minimum temperature. This may require a higher-than-normal viscosity index, or providing some means
of warming the oil. For highly loaded bearings operating at high speed, particularly roller-type bearings,
specially compounded oils are generally necessary to realize the best results. The bearing manufacturer
or a reputable lubricant manufacturer should be consulted in such cases.
230 Bearings

1000 r---~-.-----~-r-----r-~--,..-----",,"----, 4600


(mm2/s SUS
(cSt)

500~~--r-----~~--+--~~r----~~~-~2300

200 930 'fii


i
~
:::>
~
Ul

E
....... 100 460
"c:
0

~ ~
.~
>
~
.~

:-s
50 230 c:
:::>

~
en
~
20 100 E

la.
c:(

10 60

Pitch Diameter (mm) ----....J.. ~ dmmm

elm • (bearing bore + bearing 0.0.) + 2


III • required lubricant viscosity for adequate lubrication at the operating temperature
I and II refer to Examptes I and II, respectively, on page 43

Figure 11.27 Minimum required lubricant viscosity for antifriction bearings.


(Courtesy SKF Industries, Inc.)

MAINTENANCE OF ANTIFRICTION BEARINGS

If properly applied and lubricated, antifriction bearings in centrifugal pumps have long life and are
unusually troublefree. Failure can result, however, from the following: (1) use of the wrong type or size
for a particular application, (2) faulty mounting because of improper workmanship in manufacture or
Bearings 231

ApproXimate Temperature Conversions Degrees Fahrenheit

50 90 120 140 160 175 190 210


20000
I ...~
tOooo
~~
5000
3000 , ""' ~

""" " """""


......
2000 '"
~""" ~ ~
l '" .~ "", ~ ~
~

l"-"
1000

500
400 " " ""
""-
"- ~ ~""IiiiI ""'- "-
2300

gii>
"" "" "'" "'"
"- ,-"""
"""
300 1250
~ 200
"-
~
'-'"
""'- IX:
~
f'..
900 .~~

, '-..... """
"' "'"""" ""-"'" "'" """"'"
.!!! 150
~,
~
~'" 700 "'~

~""' ~
E ~~
.§. 100 !oo..'" I".. 470 o c:
()o

"""',
II> 75
i'oo.. ~ ...... """'" ~ 350 ~~
'" ""-0,.,"""
..... 240 .~~

"'" ""-
.l<
50
0
.i !Ioio.' ~ f""'Io..

" l ' "'


.!l ~

" I" "


190
......"" =: .~
40
E ...... ......

..,..'"
~
""'";:.... ~N
"
30 140
()
~5
II>
0 20
~~~~ I.:>t...
100 '§~
&l ~ ~ ~ ~ 1'........ v", 80
e:~
:>
l" ~ ~ L' ~ ~ ~ ~ ~
15 c(OO

,
10 60
8 " " l"-'" ""'"
r.....
I'"""i tp,:"I\ r,;>.~ ~.
~. ~
Ii(.

"" " """"...... ""'.....


6
5 ~ p",
4
-20 -10 o 10 20 30 40 50 60 70 80 90 100 120 150

Temperature, Degree Celsius

NOTE Viscosity classification numbers are


aa:ordng to inlernalional Standard ISO 3448-
1975 for oils having a viscosity index of 95.
Approximate equivalent SAE viscosity grades are
shown in parenthesis.

Figure 11.28 Oil viscosity versus temperature chart.


(Courtesy SKF Industries, Inc.)

during maintenance, (3) improper design of the mounting, (4) improper lubricant or lubricating practice,
(5) entry of water, dirt, or grit into the bearing, and (6) mechanical damage to the balls, rollers, or races.
Pump designers base their selection of bearing type, size, and lubrication to suit the field or fields of
service for which the lines of pumps will be used. Occasionally through a misunderstanding, a pump
will be used for conditions or in surroundings not suitable for its bearing design and consequently suffers
from short bearing life.
The inner race of antifriction bearings must not tum on the shaft; the outer race must not tum in its
housing; and the bearing must be in correct alignment. Antifriction bearings are usually pressed or shrunk:
on their shafts; if thrust loads are involved, they are further held in axial position on their shafts by
shoulders and shaft nuts. If the shaft is undersize, the fit will be too loose, allowing rotation of the inner
race on the shaft with resulting damage to the bearing, the shaft, or both. On the other hand, too large
a shaft diameter can result in expansion of the inner race, causing insufficient clearances between the
balls or rollers and their inner and outer races. Likewise, the mounting must provide sufficient holding
232 Bearings

force through proper gripping of the outer race in the housing to prevent the outer race from turning in
the housing. This force is generally more of a problem with radial bearings than with combined radial
and thrust bearings or straight-thrust bearings because the outer race is clamped between two shoulders
in the housing assembly if thrust is involved. In radial bearings, however, the outer race must be able
to move axially in its housing if temperature changes cause unequal expansion of the shaft and casing.
The fit of the outer race in its housing is therefore in the nature of a push fit. It is also very important
for antifriction bearings to be squarely mounted on their shafts and in their housing and to not be cocked.
A pump designer has to make sure that the casing will not distort unduly when pressure is applied, as
distortion would throw the bearing out of line. Antifriction bearings have close tolerances; pump design
and workmanship must meet them.
Some pumps incorporate means for adjusting the radial position of the bearing housings (Fig. 11.17)
so the rotor can be accurately centered within its running clearances. After replacing the bearings in
pumps so built, it is important to check the rotor centering and correct it if necessary. See the pump's
instruction manual for details on how to do this.
Many failures of antifriction bearings (and other bearings) can be traced to the use of improper
lubricants. Guidance on the correct type of lubricant is given under lubrication of antifriction bearings
earlier in this chapter.
Care should be exercised to prevent water from entering the bearing. If water gets into the housing-
except for small amounts with soda-soap base grease lubricant-the bearing parts are sure to become
rusted and hence fail. Too much cooling of the housing has been known to cause condensation of
atmospheric moisture inside the housing. In liquid-jacketed bearings, the flow of cooling liquid should
be regulated so that the bearing is reasonably warm and the supply cut off when the pump is idle.
Dirt or grit allowed into the bearing will naturally cause damage. As grease makes a good seal against
dust and dirt, grease lubrication is generally preferred if the pump is to be installed in a dusty location.
Mechanical damage to the balls, roller, or races causes early bearing failure. For that reason, proper
mounting and dismounting procedures should be followed.

RELUBRICATION PERIODS

It is not advisable to schedule any fixed time period for adding or renewing the lubricant in an antifriction
bearing but to follow instead the specified period set by the machine manufacturer. The time interval
for grease lubrication is a function of bearing type, size, operating speed, and operating mode. A large
bearing operating continuously at high speed may require additional grease every 2 months, for example;
the grease in a smaller bearing running at moderate speed might last for 4 to 6 months; while a very
small bearing operated intermittently at low speed would likely need additional grease every 1 to 2
years, and then only to offset possible deterioration of the grease itself. Figure 11.29 shows one bearing
manufacturer's recommendation for relubrication interval versus operating speed. This chart is based on
a good quality age-resistant grease and a maximum bearing operating temperature of 70°C (l58°F),
measured on the outer race. The chart interval should be halved for every 15°C (27°F) increase in
operating temperature over 70°C (l58°F), provided the operating temperature is still less than the
maximum allowable operating temperature of the grease. In service conditions where the grease is likely
to be contaminated with water or dust, the relubrication interval should be reduced. How much it should
be reduced can only be determined from experience with the particular service or one very similar.
Oil lubricated bearings may require that oil be added periodically to compensate for losses through
the housing seals and vent. The bearing housing generally has an oil level gage to indicate the level in
the housing. Note that the level should be checked only when the pump is shut down; while it is running,
there is quite a lot of oil in circulation within the bearing. The interval between complete oil changes
Bearings 233

c b a ~ operating hours
o_ft
15000 25000
10000 20000
~~
t"-o...~...;;:..:- ~.f""'ooo.....
r. < . . .; : ~~~ "'::::--.. . .
~nnnn
6000 15000
15000
I......:::...... ~~;::: ..... I':i'oo. . . ~ ~ ........... r-....I'-oo..
4000
10000
2000 • ft. oft
8000
1000 6000

,
800 4000
....:.">
'-.
"" "
~.
-~- ~ ~
3000

" "-
600 ~~

t° ",
" 500
400
2500
2000 " I', " \ \ I\. '\ 1'\ ""-
~\ rY
.~

300 1500 ~ftM i'.. f\. I\~ ~I\ I' I'\.

"
'0 250
': \ \ 'F'\ '\ 1\
..
I!? 1000 ftftM 1\
200
°" 1\ it. ~ ~~W 161 \ \ \
.r:

~
~
150

100
750

500
l~nn

Iftftft
,,~
I~ ~ ~ , ~
\
80 400
~ 60 300 '-
~ 50
rftft
.NU
~ i
r
.M
a: 40 200
30 150 ~M
~

... •
,
~
20 100
,on ~
15 75

10 50 ,M I n r/min

2 3 456789 2 3 456789

100 1000 10000 2QOOO

Scale a: radial bearings


Scale b: cylindrical roller bearings, needle roller bearings
Scale c: spherical roller bearings. taper roller bearings, Ihrust ball bearings: full complement cylindrical roller bearings (0,21,): crossed cylindrical
rolle~ bearings with cage (0,31": cylindrical roller Ihrust bearings, needle roller Ihrust bearings, spherical roller Ihrusl bearings (0.51r)
Scale d: bearing bore ciameter

Figure 11.29 Relubrication interval for grease lubricated bearings.


(Courtesy SKF Industries, Inc.)

is largely a function of the operating conditions and the quality of the oil used. With oil bath, flinger,
and oil-ring lubrication using conventional mineral oil the normal interval between oil changes is 1 year,
provided the operating temperature of the oil is less than 50°C (150°F) and there is negligible contamina-
tion. Higher operating temperatures or service conditions leading to contamination dictate more frequent
changes, for example, a conventional mineral oil operating at 100°C (220°F) should be changed every
3 months. Changing to a synthetic oil for such service temperatures allows a significant increase in the
interval between changes; consult a lubricant manufacturer for specific information. The interval between
oil changes for forced circulation systems also depends primarily on operating temperature and contamina-
tion, but the assessment is complicated by the effects of cooling, storage, and circulation (see maintenance
of oil lubricated sleeve bearings for further discussion). Oil mist lubrication is essentially a "once through"
system, therefore the question of oil change intervals does not arise.

RELUBRICATING PROCEDURE

In relubricating grease-lubricated bearings having housings with drain plugs, the usual practice is to
remove the drain plug and force grease through the bearing until new grease starts to come through the
234 Bearings

drain opening. The machine should then be allowed to run at least 20 min before replacing the drain
plug so that the excess lubricant in the housing can escape.
If it is desired to clean antifriction bearings without removing them from the pump, the following
procedure from SKF[11.2] may be used.

For cleaning bearings without dismounting, hot, light oil at 93 to 116°C (200 to 240°F) may be flushed through
the housing while the shaft or spindle is slowly rotated. In cases of badly oxidized grease and oil, hot, aqueous
emulsions may be run into the housing, preferably while rotating the bearing, until the bearing is satisfactorily
cleaned. The solution must then be drained thoroughly, providing rotation if possible, and the bearing and housing
flushed with hot, light oil and again drained before adding new lubricant. In some very difficult cases an
intermediate flushing with a mixture of alcohol and light mineral solvent after the emulsion treatment may be useful.
If the bearing is to be lubricated with grease, some of the fresh grease may be forced thru the bearing to
purge any remaining c.ontamination. This practice cannot be used unless there are drain plugs which can be
removed so that the old grease may be forced out.
Light transformer oils, spindle oils, or automotive flushing oils are suitable for cleaning bearings, but anything
heavier than light motor (SAB 10) is not recommended. An emulsifying solution made with grinding, cutting
or floor cleaning compounds, etc., in hot water, has been found effective. Petroleum solvents must be used with
the usual precaution associated with fire hazards.

Relubrication of oil lubricated bearings is usually just a matter of draining the old oil and adding the
correct amount of fresh oil. If the old oil is dirty, the bearing and housing should be cleaned using some
approved method, for example, that just given, before the new oil is added.

ANTIFRICTION BEARING CONDITION

For a long time, temperature was used as a indicator of bearing condition, a high temperature meaning
"trouble to come." Provided the measurement is made with a thermometer, and not a "calibrated finger,"
temperature is a useful indication of the conditions under which the bearing is operating, particularly if
the temperature measured is that of the bearing's outer race. In assessing bearing temperature, it is
necessary to establish the stable temperature, evident from a series of readings at say 10- to IS-min
intervals, then compare that to the prevailing ambient temperature and the allowable lubricant temperature.
It is necessary to take account of the ambient temperature because in most installations the bearing's
heat load is being dissipated to the surrounding air. If the bearing temperature will not stabilize, the
pump must be shut down and the cause of the overheating found and corrected. Typical causes of
overheating are lubricant level (or quantity for grease), shaft and housing fits, misalignment, frictional
heat from adjacent seals, rubbing contact between the bearings cage, and shaft or housing shoulders;
but consult the guides published by bearing manufacturers for detailed trouble shooting information. A
higher than "expected" temperature might be the result of a poor initial estimate or an indication of a
serious problem in the bearing's installation or operating conditions, and therefore should be carefully
investigated to determine the root cause. For most centrifugal pumps, an antifriction bearing temperature
(outer race) of 85 to 90°C (85 to 194°F) is the upper limit of "expected."
With modem vibration measuring techniques, it is now possible to learn a great deal more amount
an antifriction bearing's condition than can be gleaned from temperature measurement. As an antifriction
bearing operates, shock pulses are generated by contact between its rolling elements and races. Measuring
these shock pulses and comparing them against reference data shows whether the shock pulses being
generated are too high, and if they are whether bearing condition (rolling element and race damage) or
the nature of the service conditions is the cause. This form of monitoring offers two notable advantages.
Bearings 235

First, it allows a poor operating condition to be detected and corrected before it manifests itself as a
bearing failure. Second, it allows for the orderly replacement of bearings that are nearing the end of
their service life. In both cases, an unscheduled shutdown, with its high risk of associated damage and
plant downtime, is avoided.

MOUNTING AND DISMOUNTING ANTIFRICTION BEARINGS

As the fit between the outer race of an antifriction bearing and its housing classifies as a push fit, the
mounting or dismounting of a bearing in its housing offers little problem. Some housing designs make
it impossible however, particularly when dismantling, to apply the force that is necessary to pull the
bearing out of its housing anywhere except through the balls or rollers. Such a force can easily damage
the bearing.
It is desirable to mount a bearing on its shaft with the equivalent of a press fit. Actually, the bearing
may be pressed on the shaft or shrunk on. Bearings to be shrunk on are first heated in an oil bath or on
a induction heater to about 93°C (200°F) and then slipped into place on the shaft, the inner race being
tapped lightly with a tube over the shaft if necessary. Oil bath heating involves some fire hazard and
the condition of the bath needs to be carefully maintained to avoid the risk of bearing corrosion from
acidified oil. Induction heating avoids these problems. If the bearing is pressed on the shaft, the use of
an arbor press is desirable (Fig. 11.30). The force should be applied to the inner race through a tubular
sleeve or pipe, a ring, or small blocks of equal thickness. If an arbor press is not available, the bearing
can be driven onto the shaft by hammering alternately on opposite points on the circumference of a
tubular sleeve held against the inner race. Care must be taken to keep the bearing from being cocked,
and feeler gages should be used to make sure it is pressed firmly against the shaft shoulder.
Bearings to be dismounted from a shaft must usually be forced off, as the use of heat is seldom
feasible. The technique followed will depend on the design and the equipment available, but a split

f~­
.r ..

Pipe on
inner race
only

~:II-I~'-D" Bearing

Shaft

Figure 11.30 Two methods of mounting anti-friction bearing on a shaft.


236 Bearings

Figure 11.31 Removing anti-friction bearing with an arbor press.

washer is usually employed to bear against the inner race or against a shaft sleeve on which the bearing
has been pressed. A firm, steady pressure is applied through the split washer by an arbor press or a form
of wheel puller (Fig. 11.31). Care must be taken to keep the shaft straight, to avoid damage from cocking.
With proper tools the mounting or dismounting of antifriction bearings is no problem. Improper tools
usually cause damage.

ANTIFRICTION BEARING INSPECTION

After a bearing is removed, it should be dismantled and its components carefully examined to determine
how it failed. This is particularly important if the bearing has failed prematurely, because a prime
objective of the maintenance process in such cases should be to determine the root cause of the failure
and correct it. When examined carefully and the observations compared with reference illustrations
available from the various bearing manufacturers, it is usually possible to state how the bearing failed.
Working from there, the machine condition that provided the circumstances for failure can be identified
and corrected. Of course, if the bearing is essentially destroyed in the failure, finding out why it failed
is much harder and the problem may persist. Avoiding this is a further justification for monitoring
bearing condition using shock pulse measurement, which allows bearing replacement before catastrophic
failure and consequent destruction of evidence.

PLAIN BEARINGS
Plain bearings preceded antifriction bearings, and are nominally simpler because the relative motion is
by sliding rather than rolling. Although most centrifugal pumps made today have antifriction bearings,
Bearings 237

since they are generally the most cost-effective solution, the small proportion of pumps made with plain
bearings spans virtually the entire range of centrifugal pump applications. A better insight into this range
is given by considering the factors governing the choice of plain bearings.

1. Economy-In small, low-cost pumps intended to handle clean liquids, a sleeve-type plain bearing (Fig.
11.2), lubricated by the pumped liquid, offers notable economy of construction.
2. Operating environment-When some or all of a pump's bearings have to operate in the pumped liquid, as
they do in vertical turbine pumps, hermetically sealed pumps, and horizontal multistage pumps with slender
shafts, plain bearings of sleeve design, lubricated by the pumped liquid, offer the most practical solution in
nearly all cases.
3. Configuration-Many pump designs require bearings with small radial dimensions to cause the least obstruc-
tion to hydraulic passages, for example, the liquid end and lineshaft bearings in vertical wet pit pumps. For
these applications, plain sleeve-type bearings offer the smallest radial dimension.
4. Speed-Pumps used for high-pressure services such as boiler feed, hydrocarbon charge, and similar services
run at speeds from 3,600 to 10,000 rpm and have relatively large shafts. For all but the smallest pumps,
the d.,.N factor for antifriction bearings (mean bearing diameter in mm multiplied by rotative speed in rpm)
is too high, thus dictating the use of plain bearings of either sleeve or tilting pad design.
5. Load-In high-pressure multistage pumps and very large pumps, the axial and radial loads imposed on the
pump's bearings generally exceed the capability of antifriction bearings, and thus dictate the use of plain
bearings, usually of segmental or tilting pad design.
6. Preference-Some pump users prefer plain over antifriction bearings, the rationale being that the former
are easier to inspect and maintain and less susceptible to catastrophic failure. Under the influence of such
preferences, purchasers either specify sleeve bearings or devise a bearing selection rule that leads to the
same result.

From these factors it is evident there are wide variations in the design of plain bearings for centrifugal
pumps. To bring some order to the discussion, it is useful to classify the various designs.
A fundamental distinction for all plain bearings is the mode oflubrication. Three modes are recognized:
boundary, mixed film, and fluid film. One way to illustrate the difference between these is to plot the
coefficient of friction, f, against the bearing parameter, ZNIP, where Z is the lubricant viscosity in
centipoise, N is the rotative speed in rpm, and P is the bearing pressure in psi. Figure 11.32 shows a
plot. Boundary lubrication prevails when the film of lubricant between the bearing surfaces is so thin it
allows the high points of the surfaces to come into contact. Bearings operating in this mode have the
highest coefficient of friction (Fig. 11.32), and are therefore only suitable for very low speed or oscillating
motion. At the speeds inherent in centrifugal pumps, the existence of boundary lubrication will result
in rapid bearing wear. Mixedfilm is a transition state between boundary and fluid-film lubrication. Many
successful plain bearings operate with mixed-film lubrication (e.g., internal bearings in horizontal and
vertical pumps), but as Fig. 1l.32 shows, the coefficient of friction, hence the bearing's life, is critically
dependent upon the conditions at the bearing. Fluid-film lubrication results in complete separation of
the bearing surfaces and yields the lowest coefficient of friction (see Fig. 1l.32). Increasing the film
thickness causes the coefficient of friction to rise, a consequence of increasing shear forces in the
lubricant, but the rate of change is relatively insensitive to the bearing conditions. Bearings operating
with full-fluid-film lubrication theoretically have infinite life. In practice wear does occur, although
usually very slowly, as a result of starting and stopping the pump, momentary overloads during operation,
and imperfections in the lubrication.
In centrifugal pump practice, full-fluid-film lubrication is generally achieved by hydrodynamic action,
meaning that relative motion between the bearing surfaces generated the fluid film (Fig. 11.33). Hydrody-
238 Bearings

0.150

.... Mixed
c:
o
n
--w----
1
.;:: ,
LL -.. - - - - Fluid
1
'0 Boundry'
E
II) 1-+-----1
'u
~o
()

0.001
O~----~----~----------------------~
o Bearing Parameter, W
Figure 11.32 Coefficient of friction, 'f: versus ZNIP showing various lubrication modes.

_ _-'I +
~'----~-'-oto-rLoad
Gap

~-M_i
OIL FILM MOVEMENT 1J1LJ.:...
Oil from pump

Figure 11.33 Ordinary cylindrical bearing with oil


film formed by pumping action. Figure 11.34 Principle of hydrostatic lubrication.
Bearings 239

namic lubrication can be achieved only when the load is within the bearing's capacity, a function of its
detail design and the lubricant provided, and there is some minimum relative velocity between the
bearing surfaces. The other means of developing full-fluid lubrication is termed hydrostatic. With this
approach high-pressure lubricant is supplied to the bearing, and the leakage of that lubricant across the
bearing develops the fluid film (Fig. 1l.34). Hydrostatic lubrication is used in large pumps to "lift" the
bearings before starting (line bearings in horizontal pumps and thrust bearings in vertical pumps), and
it prevails to varying degrees in all of a pump's internal running clearances, whether designed as bearings
or not.
The detailed design of plain bearings is influenced significantly by the type of lubricant. From the
bearing parameter ZNIP, it is evident high viscosities make it easier to achieve fluid-film lubrication.
High viscosities are associated with oil, and low viscosities with the pumped liquid, therefore a simple
classification of plain bearing design is to consider oil lubrication and pumped liquid lubrication (or
product lubrication as it is often called). The following discussion takes that course.

OIL-LUBRICATED PLAIN BEARINGS

During normal operation, oil-lubricated plain bearings in centrifugal pumps almost invariably operate
with hydrodynamic lubrication. In fact the term "hydrodynamic bearings" is frequently used when
referring to oil lubricated plain bearings.
In plain bearing terminology, line or radial bearings are generally known as journal bearings. They
are either sleeve type or tilting pad (sometimes referred to as segmental), with sleeve type the more
usual. The design of sleeve bearings varies with speed. Designs for low speed (Fig. 11.35) generally
have a split babbitted bushing with a high LID ratio (length of bearing to journal diameter). The bushing
bore is cylindrical with two axial oil feed grooves at the split (Fig. 1l.36[aD. Bearings of this basic
design are used for journal velocities up to 12 m/sec (40 ft/sec). For higher speeds, the design must give
due consideration to the question of heat dissipation (see lubrication) and bearing stability.

OIL HOLE COVER

SPLIT BEARING
WATER
SH BODY

OIL RING

Figure 11.35 Sleeve type journal bearing for low speed applications.
High LID ratio; oil ring lubrication; spherically seated for self-aligning capability.
240 Bearings

(a)

(b)

(c)

Figure 11.36 Typical sleeve bearing bore, (a) cylindrical with two oil distribution grooves, (b) single-pressure
dam, and (c) tri-lobe.
Bearings 241

Figure 11.37 Sleeve type, babbitted journal bearing for high speed applications.

As the speed of any turbomachine is increased, the rotor weight for a given journal size tends to
decrease. If the bearing design does not account for this reduced loading, there is a risk the bearing will
develop oil whirl, a self-exciting phenomenon that produces a rotating force on the journal at a frequency
just below 50 percent of running speed. Oil whirl can produce one of two results. If the rotor is light
enough and the damping is low enough, it can be forced to respond to the subsynchronous excitation,
and may develop vibration amplitudes equal to the bearing clearance. A more serious problem arises if
the rotor has a lateral critical speed (see Chap. 7) close to the oil whirl frequency. Should this be the
case and the rotor damping is low, the resulting resonant response can easily develop destructive vibration
amplitudes (sufficient to rapidly wear the internal running clearances). Centrifugal pumps, fortunately,
are not particularly susceptible to these problems. This relative immunity derives from the magnitude
of forces and damping produced within their internal running clearances. In spite of this, the design of
sleeve journal bearings for high-speed pumps does incorporate provisions to ensure the bearing loading
is high enough to produce stable operation.
The principal design refinement for high-speed plain journal bearings is a notably lower LID ratio
values on the order of 0.6 being typical, to raise the nominal bearing loading. The bearing itself can be
either a split babbitted bushing (Fig. 11.37) or a pair of precision "automotive"-type inserts in a bolted
retainer (Fig. 11.38). Automotive-type inserts are matching steel half shells with a thin deposit of babbitt
on a copper backing; they offer the advantage of lower parts cost and simple bearing restoration. A
variety of bearing bore profiles is employed to further enhance the bearing behavior at high speed. These
range from the simple and very effective single pressure dam (Fig. 11.36[bD to mUltiple pressure lobes
(Fig. 11.36[cD. Oil groove designs depend on the bearing bore profile.
242 Bearings

Figure 11.38 Sleeve type high speed journal bearing with "precision" automotive type insert.

Tilting pad journal bearings (Fig. 11.39) offer higher capacity than equivalent sleeve bearings, and
have the virtue of being immune to oil whirl. Citing both these advantages, some designers and purchasers
advocate their general use in any pump that requires plain journal bearings. In the majority of cases,
the added complexity of tilting pad bearings is hard to justify. As noted earlier, the bearing loading in
most centrifugal pumps is low, and the capacity of oil whirl to influence the rotor behavior is quite
limited. Where their use is warranted is in large pumps, with both horizontal and vertical axes (Fig.
7.11), In these machines, the loading can be high under some operating conditions, and in vertical-axis
designs there is a risk of oil whirl influencing the rotor behavior when the pump is running dewatered.
A further significant advantage in large machines is the ability to accurately set the bearing clearance,
something that is very difficult with a sleeve bearing, which usually has to be split into more than two
pieces to facilitate installation. Tilting pad journal bearings are installed with a specific "preload,"
meaning that the installed clearance between the journal and pad at the pivot point is less than that for
which the pads were machined. Without preload, there is a risk of unstable bearing operation under light
loading, a result of an unloaded pad tilting the wrong way (leading edge closer to the journal).
Plain journal bearings are sensitive to edge loading caused by angular misalignment between the axes
Bearings 243

Figure 11.39 Tilting pad journal bearing, conventional.


(Courtesy Kingsbury Inc.)

of the bearing and shaft. In older low-speed pumps, where quite large shaft deflections were likely, it
was common practice to spherically seat the bearing bushing (Fig. 11.35). Such sophistication is generally
not necessary in smaller high-speed pumps because the shaft deflection is usually quite low. For these
designs a narrow radial location for the bushing with some provision for rocking (Figs. 11.37 and 11.38)
has proven adequate and is simpler to produce accurately. In large pumps, the need for self-aligning
bearings is still controversial. Many of the more recent designs, however, have resorted to simple pivoted
pads in place of self-aligning spherically supported pads.
Rotors supported in plain journal bearings must be provided with some form of thrust bearing to
locate the rotor axially and to accommodate any axial thrust developed by the pump. One of the following
three arrangements is generally employed to achieve this:

1. Babbitted faces on the sleeve type journal bearing acting against shoulders or collars on the pump shaft
(Fig. 11.40)
2. A separate antifriction type thrust bearing located in one of the journal bearing housings, usually the outboard
housing (Fig. 11.41)
3. A tilting pad thrust bearing incorporated into the bearing housing at one end of the pump, again usually the
outboard end (Fig. 11.42).

The first arrangement is the simplest design. In modem pumps, its use is limited to low-speed designs
with inherently low values of axial thrust, such as large horizontal-axis single-stage double-suction
pumps. And even in these cases, there are many purchasers who will insist on a more substantial thrust
bearing. An antifriction thrust bearing, the second arrangement, is used in medium-speed (journal
velocities up to 12 m/s (40 ft/sec)) pumps with relatively low axial thrust and operating at moderate
pumping temperatures. Within these limits the antifriction thrust bearing can meet the usual life require-
ments and the bearings can be ring-oil lubricated, thereby avoiding the added expense of a separate
lubricating oil system.
244 Bearings

Figure 11.41 Anti-friction thrust bearing with sleeve


Figure 11.40 Simple babbitted thrust bearing. journal bearing.

WATER SLINGER

BEARING COVER

INTERNAL RETURN TO
OIL RESERVOIR
BEARING BRACK

RESERVOIR

ADJUSTING SCREW

Figure 11.42 Section of Kingsbury thrust bearing.


Incorporates sleeve type journal and flood lubricated tilting pad thrust bearing.
Bearings 245

For high-speed pumps or when the axial thrust is or can be high (e.g., when the pump is worn), it
is now standard practice to employ tilting pad thrust bearings of the Kingsbury (or Michell) type. Fixed
thrust bearings such as the "tapered land" design offer a capability between the simple babbitted face
and the relatively complex tilting pad bearing, but the precision necessary in their manufacture tends to
offset the initial cost advantage. Such designs are therefore rarely used in centrifugal pump practice.

TILTING PAD THRUST BEARINGS

The Kingsbury version of this design was first developed to meet the need for a high capacity thrust
bearing for vertical axis turbines and has, over time, been widely applied to other rotating equipment,
including centrifugal pumps. The operating principle is simple. An ordinary cylindrical or sleeve bearing
has a running clearance between the bearing shell and the journal. Because of the relation of the curved
surfaces and the capillary attraction of the oil particles, a "pumping" action takes place that draws a
lubricating oil film into this clearance (Fig. 11.33). If the oil is of correct viscosity, it will resist the
breakdown of the film except at excessive loads. To provide a positive and ample supply of cool oil to
the bearing, a simple gravity device is ordinarily used, although operation at higher speeds resulting in
maximum tendency to heat requires some form of forced feed lubrication. In an ordinary thrust collar
subjected to high pressures and high speeds, the parallel surfaces tend to squeeze out the oil film. The
metal-to-metal contact that results makes this type of bearing unsuitable for heavy loads.
The principle of the Kingsbury bearing can be described as follows: Suppose that a circular collar is
cut into little segments and that each block is suitably supported on its underside so that it may rock
slightly on the point indicated as the suspension point and yet stay in place. When the shaft begins to
rotate, the film of oil tends to be dragged in under the slightly rounded edges of the blocks. As the speed
of the shaft increases, this tendency increases, the block adjusting itself slightly by tipping at a greater
angle, riding up on the oil film as a sled runner rides up upon meeting the surface resistance of snow
underneath (Fig. 11.43). The higher the speed, the greater this tendency for the block to rock forward,
permitting an increased "sledding" action, and the greater the tendency to adjust itself to the increasing
oil film dragged underneath it. Construction details of a typical Kingsbury bearing can be examined
more closely in the sectional assembly shown in Fig. 11.42.
The thrust mounting of Kingsbury bearings used in horizontal pumps is arranged to take thrust in
both directions. Sometimes both loads are approximately equal; other times there may be a major thrust
in one direction and an occasional minor thrust in the opposite direction. In any event, the Kingsbury
bearing is provided with thrust shoes on each side to limit the axial motion of the rotor. The number of
shoes on each side mayor may not be equal, depending on the application.
Conventional tilting pad thrust bearings have symmetrical pads or shoes, meaning that they are
supported at their centre. This design allows equal thrust with either direction or rotation. In their usual
form, symmetrical tilting pad thrust bearings are pressure fed with oil at a specified rate and the oil
"floods" the entire bearing assembly, some of the flow actually passing through the load carrying oil
"wedge," most of it removing heat from the adjacent components. For a given loading the heat generated
in such a bearing is quite high. To reduce the heat load, and hence the size and expense of the lubricating
oil system, the bearing manufacturers have developed more sophisticated designs. The essential features
of these are asymmetric shoes or pads (point of support offset away from the leading edge) and the
introduction of oil close to or at the shoes' leading edges (Fig. 11.44). Figure 11.45 shows the reduction
in oil flow this refinement affords. At the same time, the bearing runs cooler because the cool oil is
being directed to where it's most effective. The one disadvantage of asymmetric bearings is that their
capacity in reverse rotation is reduced. For high speed uni-directional equipment, this is not deemed a
major disadvantage.
246 Bearings

SUSPE~SlO~ POtI\lT

THRUST LOAD

DIR[CTIOI\I OF ROTATION

SHAFT THRUST CO L LAR IN MOT ION

Figure 11.43 Principle of Kingsbury thrust bearing.

Figure 11.44 Kingsbury-type LEG thrust bearing.


(Courtesy Kingsbury Inc.)
Bearings 247

Feet/sec.
100 200 300
Standard thrust bearing

LEG thrust bearing

50 70 90 110
Meters/sec.
Mean sliding velocity

Figure 11.45 Lubricating oil requirements for flood lubricated versus Kingsbury type LEG thrust bearing.
(Courtesy Kingsbury Inc.)

MATERIALS

Tin-based babbitt is the most widely used material for oil-lubricated plain bearings in centrifugal pumps.
Although more expensive than lead based babbitt, it offers greater tolerance of boundary lubrication
under transient operating conditions, has higher corrosion resistance, is easier to bond to steel, is less
prone to segregation, and has better high-temperature properties. Babbitts have relatively low load-
carrying capacity and fatigue strength, and the capabilities decrease with increasing temperature. For
lightly loaded journal bearings, these limitations do not pose a significant design problem, but care is
needed with highly loaded thrust bearings. Good design practice is to keep local metal temperature
below 120°C (250°F). Note that the temperature indicated by temperature detectors will be lower than
this because they are measuring a bulk temperature.
Bearing journals are finished directly into the shaft when it is of carbon or low alloy steel such as
AISI 4140 or 4340. Alloy steels with more than 1 percent chrome and running at over 20 mls (65 ft/s)
pose the risk of "wire wool" damage to the bearing, a phenomenon whereby small embedded chromium
carbide particles "machine" the joumal to produce fine turnings, hence the term "wire wool," which
quickly destroy the bearing. To avoid this, such shafts have the journals hard chrome plated before being
ground to size.
Thrust collars are hardened and tempered low alloy steel.
248 Bearings

PRESSURE RELIEF VALVE SET AT 20 PSI


PRESSURE GAUGE
FEED LINE TEIiW'ERATURE GAUGE
~~~~-f==~~==T==T~~~~COOLER

PRESSURE GAUGE
____ilt.IRVlCE OIL PUMP

OIL TANK
CHECK VALVE OIL PUMP SUCTION

Figure 11.46 Typical forced-feed oil lubrication flow diagram.

PLAIN BEARING OIL LUBRICATION

A ring oiled bearing is furnished with a soft steel or bronze oil ring that rides on the pump shaft through
a slot cut in the middle of the top half of the bearing shell. This ring rotates as the shaft turns and picks
up oil from the reservoir in the bearing housing. The oil is wiped off on the top of the pump shaft, flows
between the bearing bore and the shaft, and is discharged at the ends of the bearing (see Fig. 11.35 and
Fig. 11.40). Lubrication by means of oil rings is fully satisfactory only at relatively low operating speeds.
A provision for automatic circulation of the oil-and if necessary, for cooling it-is an essential feature
of all higher speed (plain) bearings, especially thrust bearings.
In some bearings, the oil circulation is effected by a rotary positive-displacement gear pump directly
connected to the outboard end of the pump shaft by means of a flexible coupling (see Fig. 11.42). The
oil pump takes the oil from a reservoir, located either in the bearing housing itself or separately on the
pump baseplate, and delivers it under pressure through the oil cooler. From the cooler, the oil flows in
part to the outboard thrust bearing, from which it flows into the reservoir located in the lower half of
the bearing housing. It then overflows by gravity from this reservoir into the main reservoir on the
baseplate. This lubricating system is illustrated in Fig. 11.46.
General practice supplies the inboard line bearing of this system with oil under pressure through a
branch line in the discharge from the oil cooler. The oil from the inboard bearing is returned by gravity
through large return lines into the main reservoir. It is essential to provide an adequate pressure drop
from all bearings so that the oil will not overflow because of unsatisfactory evacuation.
Numerous alternative methods exist for supplying the bearing with forced-feed lubrication. For
example, some arrangements use a vertical oil pump driven from the main pump shaft by means of a
worm gear (Fig. 11.47). Other bearings employ the Kingsbury "adhesive lubrication" system (Fig. 11.48).
In this system, oil from the reservoir beneath the thrust bearing is drawn into a bronze ring (Fig. 11.48[a]),
called the "circulator" or "oil pumping ring," which is around the collar. The adhesion of oil to the
collar carries the oil around in the groove in the ring (Fig. 11.48[b]). The oil travels with the collar for
almost a complete revolution. It then meets a dam in the groove and is pushed by the stream behind it
into a port leading to spaces between the two lowest shoes on both sides of the thrust collar. Shaft
rotation carries it to the other shoes, and it finally escapes, above the collar, into a passage leading down
to a cooler. From the cooler it returns to the reservoir. The oil will circulate equally well with the collar
running the other way. When the collar changes direction, the adhesiveness of the oil carries the circulator
Bearings 249

Figure 11.47 Vertical oil pump driven from main pump shaft by worm gear.

with it through a short angle, until the lug at the top of the circulator meets a stop. In either of the two
"stop" positions, oil enters the groove in the circulator by the proper port for the direction of rotation
and is discharged through the middle port.
The Kingsbury "adhesion ring" does not produce sufficient pressure to allow a filter to be included
in the oil circuit. Concern over oil cleanliness in high-speed machines and the desire to avoid a separate

(8) (b)

Figure 11.48 Pumping ring of Kingsbury bearing.


250 Bearings

Operation
1. Arotating drive hub fixed to the pump
shaft causes pressure ring to rotate by
viscous shear forces (VSF). The major
VSF occurs between the flanges of the
drive hub and sidewall of the pressure ring.
2. Surface velocity of the pressure ring,
aided by drive vanes internal to the ring,
accelerate sump oil to a velocity which
generates a static head (or pressure)
when rammed into the pickup tube.
This principle of pressure generation is
well known from Pitot tube gaging in
flowing pipes.
3. Sufficient pressure can easily be
generated to force lubricating oil through
a conventional filter and orificed inlets
to all types of bearings.
4. The pressure ring oil system is not
sensitive to variations in lube oil viscosity
and therefore works well at all
temperatures with a wide range of oil
grades. While fluid friction losses
increase somewhat with viscosity, the
pressure ring velocity also increases.
Therefore, the system is self
compensating.

Figure 11.49 Pitot tube integral oil circulation system. (Patented)

oil pump have led to the development of the "pitot tube" oil circulation device (Fig. 11.49). In this
device, oil is accelerated to a high velocity by hydroviscous action within the pumping ring. The pitot
tube gathers a portion of the high-velocity oil stream, diffuses it to a lower velocity and higher pressure,
then passes it to an oil circuit, which includes a micronic filter. Unlike the Kingsbury adhesion ring, the
pitot tube circulator is unidirectional.
Sometimes the forced-feed lubrication system supplies oil to the driver bearings as well. A typical
system combining pump and driver lubrication is shown in Figs. 11.50 and 11.51. If pumps are driven
by steam turbines or through gears, it is customary to have the turbine or the gear supply oil to the
pump bearings. Such arrangements require reconcilement of the lubricating oil characteristics and of the
operating temperatures established by the manufacturers of the individual pieces of equipment.
The use of oil rings for line sleeve bearings normally supplied with oil under pressure is optional
and not always justified. Their function is basically that of supplying oil to the bearing at the start of
the pump operation, supposedly before the forced-feed system has had the time to do so. It should be
remembered that sufficient oil is generally retained in the bearings to take care of their needs before
forced-feed delivery takes place.
If the normal retention of oil in the bearing or the use of oil rings will not afford adequate protection,
auxiliary oil pumps are called upon. These may be manually operated gear pumps (Fig. 11.52) intended
for use at scheduled intervals when the pump is standing idle. Operation of this auxiliary pump at weekly
or bi-weekly intervals is usually sufficient to keep the oil from draining out completely from the bearings
or the oil piping.
More elaborate lubricating systems incorporate a motor-driven auxiliary oil pump, which is started
before the main pump begins operating. The motor starter controls are interlocked in such a manner that
Bearings 251

1/2 IN. PT CONNECTION FOR PRESSURE SWITCH TO


STOP MOTOR-ORIVEN AUXI~IARY Ol~ PUMP WHEN
SERVICE PRESSURE REACHES 32 PSI AND TO START
PUMP WHEN PRESSURE DROPS TO 10 PSI

RELIEF VAL
SET AT 20

o
RELIEF VAL\IE--~HI
SET AT 30PSI

OIL RESERV'OIR·----j-

AUXILIARY OIL PUMIP--t====~~~~L_~

Figure 11.50 Simple forced-feed system for pump and motor bearings.

the main motor cannot be started until the oil pressure in the system reaches a predetennined value. As
soon as the oil pump driven from the main pump shaft develops sufficient pressure, the auxiliary pump
is shut down by means of a pressure switch. A second pressure switch setting automatically restarts the
auxiliary pump on failure of the regular pump to maintain the desired pressure. This arrangement was
illustrated in Figs. 11.50 and 11.51. The settings for the pressure switch are indicated in the fonner, and
both drawings show the arrangements of oil cooler, oil filter, oil flow indicators, relief valves, and the like.
The lubricating oil system illustrated in Figs. 11.50 and 11.51 is the minimum necessary for the
functional requirements. Many purchasers today specify more elaborate systems in the search for higher
equipment reliability. Specifying API-61O adds approximately 25 percent to the system cost, by requiring
that the reservoir and all oil piping be fabricated from austenitic stainless steel. The next increment is
to specify API-614 [1l.3], which raises the system cost by a factor of 2.0 to 2.5. As an aid to evaluation,
the extra cost comes from the following:

1. Larger reservoir: 8 min retention time at minimum operating level versus 3 min
2. Reservoir and all oil piping fabricated from austenitic stainless steel
3. Duplex full-capacity heat exchangers
4. Dual-pressure system: oil pressure to downstream of the heat exchangers above the maximum available
cooling water pressure
5. More elaborate instrumentation
6. Extensive shop testing.

Whether the additional expense of more sophisticated lubricating oil systems is warranted is a question
only the purchaser can decide. It depends on recent experience with system manufacturers, the capabilities
of those installing and maintaining the plant, and the criticality of the equipment being lubricated.

PRODUCT LUBRICATED PLAIN BEARINGS

The tenn "product lubricated" refers to bearings that are within the pump and therefore lubricated by
the pumped liquid or product, or a liquid other than lubricating oil. The usual attraction of such bearings
~
N

711 ,I
/./~~~~\TER
Ii
Ib t

A
!It:CTIOH A-A

Figure 11.51 Oil piping layout of lubrication system in Fig. 11.50.


Bearings 253

Figure 11.52 Manually operated auxiliary oil pump for forced-feed system.

is a simpler pump. Whether the simpler pump is more reliable depends on the design of the bearings
and their means of lubrication and cooling.
Of the various centrifugal pump classes (see Chap. 1), vertical turbine or diffuser pumps and vertical
wet-pit volute pumps (see Chap. 14) rely exclusively on product-lubricated plain journal bearings. In
the past, some of these bearings were drip oil lubricated, but environmental considerations now preclude
that, so all modern designs use strictly product-lubricated bearings.
Vertical turbine and vertical wet-pit volute pumps have journal bearings in two locations: within the
liquid end or bowl assembly to guide the pump's rotor or spaced up the column to guide the pump' s
lineshaft. With some limitations imposed by pump design, these journal bearings can be either open or
enclosed, meaning the bearings are either exposed to or isolated from the pumped liquid. Because the
bearing arrangement is related to pump design and construction, the two options are dealt with in detail
in Chapter 14.
A wide variety of designs and materials has been advanced for vertical wet pit journal bearings. All
have had the objective of providing bearing life at low initial cost. Not all have been successful, usually
because the design failed to take into account variations in loading, liquid cleanliness, and realistic
bearing material properties. Successful designs have generally met the following guidelines:

l. No reliance on internal seals with a high pressure differential (i.e., more than, say, 0.35 bar 5 psi), to exclude
solids laden pumped liquid from the bearings
2. Design factor of at least 1.5 on the highest expected loading
3. PV (unit pressure in bar [psi] times journal velocity in m/s (ft/min) no higher than 3.50 (10,000) at maximum
design loading
4. Adequate lubrication grooving to allow a high flow of liquid for lubrication and cooling.
254 Bearings

Materials and design techniques are improving and will continue to do so. It is therefore quite likely
that the above guidelines, particularly that for PV, will eventually be too conservative. Noting that most
product lubricated bearings run with mixed film lubrication, a mode in which the coefficient of friction
varies widely with bearing conditions (Fig. 11.32), it should be evident that such a decision can only
be made on the basis of carefully conducted and validated tests. In the absence of such data, the
conservative guidelines will usually prove less expensive in the long term.
For pumps handling clean liquids and having open lineshaft construction (Fig. 11.53) bearings of
PTFE tape (petro coke flour filled), various reinforced plastics, metal impregnated carbon, or rubber
have proven successful. Journal materials are typically that of the shaft, unless the shaft is relatively
soft as is the case with austenitic stainless steels. In these cases, the shaft can be ceramic hard coated
in the journal areas or furnished with ceramic hard coated sleeves. The usual form of rubber bearing is
that known as "cutlass" (Fig. 11.54), which is intended to tolerate some solids and therefore runs with
larger than normal clearances. Since the guidance is not as accurate, designers generally choose this
material only when there is a risk of solids being present.

Figure 11.53 Open lineshaft bearing.


Bearings 255

Figure 11.54 Cutlass rubber bearing.


(Courtesy Johnson Duramax)

When it is known the pumped liquid will contain some solids, and open lineshaft construction is
being employed, most pumps are furnished with Cutlass rubber bearings running on 13 chrome or
ceramic hard coated journals. The tolerance of these bearings to solids is limited, particularly if the
journal surface speed is high. For difficult services, such as primary and secondary steel mill scale pit,
or even cooling tower basin in a dusty environment, the pump really has to have enclosed bearings,
lineshaft and liquid end, to achieve reasonable service line between overhauls. Enclosed bearings (Fig.
11.55) are typically either leaded bronze or reinforced plastic running on 13 chrome steel journals. When
the environment is corrosive to 13 chrome, such as in a sea water application, the duplex stainless steel
or Monel shafting typically used provides an adequate journal.
Horizontal-axis pumps employ product lubricated bearings in two ways. The first and nominally
simpler design has a slender shaft, one whose static deflection exceeds the internal radial running
clearances, and therefore needs additional internal bearings to prevent rubbing contact at the running
clearances. Pump designs following this principle have evolved using various combinations of metal-
to-metal bearings, with alloys, hardness, and surface finish all carefully chosen to reduce the rate of
adhesive wear and the risk of seizure. Most these bearings have only a low-pressure drop across them,
and therefore operate with mixed film lubrication at best. Given this, the life of the bearings, and
subsequently the running clearances, is very much determined by conditions that affect the lubrication
mode and bearing loading, for example, the frequency of starting and stopping the pump, its operating
flow range, the surface finish of the internal bearings, and the pumped liquid properties, particularly
viscosity and cleanliness. Recognizing these difficulties, designers have sought better materials. Filled
Teflon tape has proven moderately successful in small low-head-per-stage pumps (Fig. 11.56), where
speed and loading were within the material's capability. Metal filled carbon has also been employed
with some success in this application. And more recently so, too, have polymers such as polyetheretherket-
one (PEEK).
To be really effective, internal bearings in horizontal pumps need to operate with full-film lubrication.
This means that the bearings must be proportioned to develop a sufficient liquid film thickness by
hydrodynamic action, or be subjected to a pressure differential high enough to achieve the same by
hydrostatic action. For bearings spaced along the pump rotor, the latter is generally easier to achieve.
In pumps whose impellers are opposed to minimize axial thrust, (see Chap. 5) the usual impeller
arrangement has the pump's differential pressure broken down across two internal clearances, each
carrying approximately half the total differential. Being at rotor midspan, the center breakdown bushing
2S6 Bearings

r-
1
~
I-l- ~ ~
-
- -+-
~ ~ ~
--+-
~ ~
i '
..::::t=.
--+- ~ ~ ~
r-r- r"
~

~~ ---..J ~
~
~
;..; ~ ~
~ r---
~ r--- ~
~ r---
~ r--- ~
~ r---

~~
~
~
~ r"

~
L-~ r-

Figure 11.55 Enclosed lineshaft bearing.

(Fig. 11.57) is also a very effective internal bearing, operating with hydrostatic lubrication once the
pump's differential pressure is high enough. Although these hydrostatic bearings are less sensitive to
adverse conditions than the mixed-film lubricated bearings already discussed, their design still needs
close attention to the same operating factors if it is to be reliable.
None of the internal· bearings spaced along the pump rotor, and therefore dependent on the pumped
liquid for lubrication, can tolerate running dry. If such an operating circumstance is thought likely and
the desire is to have the pump survive in an operable condition, the pump design must be based on a
rotor capable of running dry (see Chap. 7).
The second way product-lubricated bearings are employed in centrifugal pumps is at each end of the
rotor in place of the usual oil lubricated bearings. Two considerations can tum the design in this direction.
The first is hermetically sealed pumps (see Chap. 24), a requirement that is increasing as a consequence
of growing environmental concerns. The second, a desire to simplify the pump by eliminating one shaft
seal and the bearing lubricating oil system (Fig. 11.58). Product-lubricated main bearings are generally
designed for hydrodynamic lubrication, with bearing proportions and surface finish selected to ensure
an adequate liquid film with the available lubricant. Material selection has proven critical in the history
of these bearings, which suggests the hydrodynamic condition thought to prevail did not always do so.
This is not too surprising when it is recognized that the viscosity of the usual lubricant, water, is only
1';"8

Figure 11.56 Multistage pump with polymer internal bearings.

N
Ul
-.l
258 Bearings

Figure 11.57 Hydrostatic bearing at midspan of opposed impeller multistage pump.

Figure 11.58 Horizontal multistage pump with internal product-lubricated main bearings.
(Courtesy KSB)
Bearings 259

1/100 that of oil, and the bearing will therefore have a significantly lower minimum film thickness than
an equivalent oil lubricated bearing. Large low-speed designs have successfully employed various
reinforced plastics. Smaller high-speed designs have also had success with plastics, but are today turning
to silicon carbide for the bearing and the journal in the search for longer bearing life.
The lubrication of product-lubricated bearings is just as critical as that of oil-lubricated bearings, yet
it is often not given the same attention, frequently with disastrous results. As a general rule, product
lubricated bearings require a "clean" lubricant to yield tolerable bearing service lives. The one exception
is the Cutlass rubber bearing, and that is by no means a panacea, being limited at nonnal pump speeds
to only low concentrations of solids in the liquid.
By definition, the desired lubricant is the pumped liquid, and when the liquid is clean, that is what
is done. What constitutes clean varies with bearing design and service conditions. With conventional
materials at nonnal speeds, solids concentrations up to 0.5 percent by weight can be tolerated. Resorting
to ultrahard materials (e.g., silicon carbide) has raised the allowable solids concentration for nonnal
speeds to some 2.5 percent by weight. High-speed bearings require the lubricant filtered to at least 15
!lm, even with the use of ultrahard materials.
When the pumped liquid is "dirty," product-lubricated bearings will not survive unless isolated from
the pumped liquid and supplied with a clean lubricant. (This excludes vertical multistage pumps and
horizontal multistage pumps with slender shafts from such services because the pumped liquid cannot
be effectively kept out of the internal bearings; see Chaps. 14 and 7, respectively.) Since the lubricant
generally passes from the bearing into the pump, it has to be compatible with the pumped liquid. Clean
lubricant can be obtained by taking a stream from the pump discharge and cleaning it using a strainer
and a cyclone separator (Fig. 11.59). The virtue of a cyclone separator is that it does not accumulate
solids as does a filter. Installing a strainer upstream ensures the separator inlet does not become plugged
with an occasional large solid. Such a system must be monitored with a flow switch to shut down the
pump on a significant drop in lubricant flow. In a large critical pump, it is desirable to use parallel

"Y" strainer with backflush valve

From
pump
discharge
Cyclone
separator

--"
1 Orifice to
balance flows

CJ
Figure 11.59 Product lubrication from pump discharges through cyclone separator.
260 Bearings

strainers and cyclones, with a differential pressure switch to sound an alann when there is a high pressure
drop. When the pumped liquid cannot be cleaned up with a cyclone separator (low-SG solids or high-
viscosity liquid), clean lubricant is supplied from a central source. If there are several pumps in one
installation, it is often simpler overall to use a central source of clean liquid rather than equipping each
pump with its own cleaning system. Regardless of the source of clean liquid, it is crucial to monitor
lubricant flow to the bearings. Monitoring pressure, although easier, is not a suitable substitute; should
the lubricant path become plugged downstream of the bearing, there will be no flow through the bearing
but the pressure upstream will be normal or higher.

PLAIN BEARING MAINTENANCE

In theory, plain bearings operating with full-film lubrication have infinite life and therefore should not
require any maintenance beyond periodic renewal of lubricant and cleaning. The working practice is
otherwise; rubbing contact during start-up and incidental overloads combined with occasional lubricant
contamination produces some wear in even the most carefully designed and operated bearings. When
the operating conditions are less than ideal, for example, high bearing temperatures, high rotor vibration,
or poor quality lubricant, plain bearings are just as susceptible to premature failure as antifriction bearings.
Plain bearings designed for oil lubrication operate with close to ideal circumstances (full-film lubrica-
tion and a clean, viscous lubricant), and therefore require only periodic dismantling and inspection to
determine actual clearances and the condition of the working surfaces. As a general rule, a journal
bearing should be renewed when its clearance exceeds 150 percent of the original value. If the diametral
clearance is not given in the instruction book, it can be approximated on the basis of allowing 0.001
nun per mm (in. per in.) of journal diameter. Tilting pad thrust bearings and the larger sizes of tilting
pad journal bearings are adjustable so the working clearance can be reset at each inspection. When to
renew or restore the bearing pads is then a matter of allowable wear, and must follow the manufacturer's
recommendation.
If the working surfaces show signs of abnormal damage, for example, pitting, scoring, bearing metal
fatigue, cracking, or corrosion, the precise nature of the damage and its likely cause should be determined.
Comparing the damage to reference photographs provided by some of the manufacturers will often allow
the root cause to be identified, leaving the maintenance engineer to determine what is producing the
root cause. In complex cases, it is often beneficial to retain the services of a specialist familiar with
plain bearings and the type of pump involved.
Babbitt-lined sleeve-type bearings can be restored by melting out the old babbitt, pouring in a new
lining, machining the bore to the required dimensions, then cutting in the oil grooving. When the bearing
has a profiled bore, the machining is more complicated and may be beyond the capability of the usual
maintenance shop. In these cases, it is better to purchase replacement or exchange reconditioned parts
from the manufacturer. Thin wall liners and tilting pads are produced using specialized techniques and
should always be replaced with new parts obtained from the manufacturer.
Damaged journals or thrust collars can be restored by refinishing provided the resulting undersize is
acceptable considering strength, and the availability of undersized bearings or the ability to adjust to
maintain clearances with the undersized part. If there is any doubt, consult the manufacturer. When
undersized parts are not acceptable, journals can be restored by grinding to a smooth surface, chrome
plating, and finish grinding to size. The plating method must avoid hydrogen embrittlement of the shaft,
and the maximum thickness in anyone deposit is 0.40 nun (0.015 in.), ground back to 0.25 mm (0.010
in.) before any subsequent deposit. For example, a journal cleaned up at 1.50 nun (0.060 in.) undersize
would have to be plated and ground a total of three times.
Product-lubricated plain bearings generally do not enjoy operating conditions as good as those of
Bearings 261

most oil lubricated bearings. One immediate consequence of this is a higher wear rate and consequently
a greater need for maintenance. Since most these bearings are located within the pump, periodic inspection
to check clearance and condition is not a simple task. Accordingly, the general practice is to monitor
machine condition, typically vibration, and only open the pump when it is obviously worn. The one
exception to this is those bearings installed at each end of the rotor in the place of conventional oil
lubricated bearings (Fig. 11.58).
The principles for when and how to renew product-lubricated bearings follow those already discussed
for oil-lubricated bearings. New running clearances vary significantly, however, with materials and it
is therefore important to establish what the design values are before assessing bearing condition or
renewing clearances.
Bearing wear should be examined carefully and documented for future reference. Particular care
should be taken with the lineshaft bearings in vertical pumps, since wear patterns can be clues to
fundamental structural problems in the pump. Scoring is usually caused by solids in the lubricating
liquid, and if found should lead to providing cleaner liquid. Damage with the appearance of scoring can
also be caused by running at too high a load or running at too high a temperature. Microscopic examination
of the bearing surface and a section through it will often yield more specific information about the cause
of failure.
Worn bearing elements are almost invariably replaced with new, since the materials used do not lend
themselves to designs such as babbitted bushings. In most cases, worn journals must be either replaced
or restored to new size, as usually there is little or no capacity to accommodate undersize parts. Chrome
plating or hard coating are common means of restoring journals. Plating must follow the rules already
given for oil-lubricated bearings. Hard coating should be a relatively tough ceramic, applied by the ultra-
high-velocity oxy fuel process, then finish ground to new size. Coatings applied by lesser processes run
a high risk of spalling off in service, an event that destroys the bearing.

MAGNETIC BEARINGS
Compared to antifriction and plain bearings, the distinguishing feature of magnetic bearings is freedom
from contact, either rolling or sliding, within the bearing. This means the bearing no longer requires a
lubricant and is relatively immune to its operating environment. These two attributes coincide nicely
with two requirements that are emerging in centrifugal pump development, namely the elimination of
complex oil lubrication systems and the need for highly reliable bearings in hermetically sealed pumps.
Of these two requirements, the latter is likely to have the greater influence because it is associated with
rendering hermetically sealed pumps an entirely viable means of eliminating the emissions produced by
leakage from conventional seals. The viability comes from freeing the pump from dependence on product
lubricated bearings. A third but little publicized need is lower mechanical noise in pumps for military
service. Being contactless and not requiring any mechanical auxiliary system, magnetic bearings offer
a potential solution to that need.
The magnetic bearings now in use are termed "active" and operate as follows: Referring to Fig. 11.60,
a journal of magnetic material is surrounded by a number of electromagnets, all designed to attract the
journal. At each of the electromagnets, there is also a proximity probe to sense the rotor's position
relative to "zero." Signals from each proximity probe are sent to a controller which compares the rotor's
position with "zero" and varies the excitation of the electromagnets to position the rotor within tolerance
of zero. Typically the bearing system is designed to keep the rotor within 0.025 mm (0.001 in) of zero
at the rated load. The power absorbed by the bearing is essentially that required for magnet excitation,
and is lower than that required for lubricating-oil systems or absorbed by high-speed liquid-lubricated
262 Bearings

Electronic
control Bearing
system system

Bearing stator

Power
Reference amplifier

~ signal .--_ _ _-,

Signal
processing

Error
signal
Power
amplifier
Sensor
stator
Sensor signal

Figure 11.60 Magnetic bearing control system.

bearings. The power can be lowered further by accommodating a portion of the load with permanent
magnets, and using the electromagnets for fine control, an arrangement known as "permanent magnet
biasing." Radial bearings (Fig. 11.61) have four or more poles distributed around the journal, positioned
such that the maximum load is shared between poles. Thrust bearings have two poles, one on either
side of a magnetic thrust collar.
Beyond contactless support, active magnetic bearings have two further advantages. The first is direct
measurement of actual bearing loads, a capability afforded by varying magnet excitation to maintain
rotor position. Associated with this, and an inherent feature of the controller, is raising an alarm and
shutting down the pump if the bearing loads exceed capacity. Following on this, it is possible to program
the controller to vary the bearing's stiffness and damping over its working speed range, thereby changing
the rotor's dynamic characteristic to avoid resonant conditions. This second advantage is of limited value
in centrifugal pumps because the liquid effects in internal clearances generally dominate the rotor's
behavior. Depending on the actual pump design, it could be beneficial to pumps required to be capable
of running dry.
The mechanical construction of an active magnetic bearing (Fig. 11.61) has several notable features.
First, the shaft must be magnetic or sleeved with a magnetic journal. Second, when the beining is for
a hermetically sealed pump, the journal, electromagnets, and proximity probes must be canned if the
pumped liquid is corrosive to the materials used for these components. Third, some form of backup
bearing is required to safely bring the rotor to standstill should the magnetic bearing fail, and to support
the rotor when it is at rest and the magnetic bearing is off. The backup bearing is normally inactive and
is designed with a normal clearance equal to half the magnetic bearing air gap. With the usual air gap
at 0.50 mm (0.020 in.), the backup bearing radial clearance is 0.25 mm (0.010 in). This means that
should the magnetic bearing fail, the pump's rotor will drop by some 0.25 mm (0.010 in.), a shift that
Bearings 263

Control Coil

annlng Material

Figure 11.61 Section of canned radial magnetic bearing.


(Courtesy Avcon)

needs to be taken into account in determining the pump's internal clearances to ensure loss of the
magnetic bearing does not result in destruction of the internal running clearances.
At the time of this writing, magnetic bearings have been incorporated in several designs of hermetically
sealed integral motor pumps (Fig. 11.62 shows one such design), and a multistage boiler feed pump.
Their use will increase as capability is demonstrated and as cost comes down, the latter reportedly
awaiting the development of self-programming digital controllers.
The maintenance of magnetic bearings can be likened to that of an electric motor with a variable
frequency controller. The bearing parts themselves should not need any maintenance beyond inspection
when the pump is opened for some other reason. Frequent "letdowns" or known corrosion or erosion
problems with "canned" bearings would be cause to increase the inspection frequency. Existing controller
designs have proven reliable to the point where the only built in redundancy is a battery power supply,
and therefore should only require periodic cleaning and changing of cards when a malfunction develops.
Cleanliness is very important; some of the early installations of magnetic bearings on centrifugal compres-
sors were plagued by poor quality controller enclosures. One fundamental rule to close the discussion:
do not work the controller while the pump is running.

BEARING HOUSINGS
A wide variety of bearing housing configurations is employed for centrifugal pumps, with the choice
being influenced by the type of rotor (overhung or between bearings), the type of casing mounting (foot
264 Bearings

I PUMP VIll..UTE CASE 6 INBOARD BE ARING ROT [R II IlUTBOARD BEARING ROTOR


i! PUMP IMPELLER 7 MOTOR STATOR Ii! IlUTBOARD POSITION SENSOR
3 MOTOR ADAPTOR B MOTOR ROTOR 13 THRUST BEARING ROTOR
4 INBOARD POSITlON SENSOR 9 ELECTRICAL fEEDTHRDUGH 14 THRUST BEARING STATOR
5 INBOARD BEARING STATOR 10 OUTBOARD BEARING STAT[R IS AXIAL POSITiON SENSOR

Figure 11.62 Schematic of canned motor pump with magnetic bearings.


(Courtesy BWIIP International)

or frame), the casing joint (radially or axially split), access to the shaft seal, the pumping temperature,
and the designer's preference based on stiffness and manufacturing considerations. Within the basic
bearing housing configurations, there is a further distinction, namely, whether the housing itself is radially
or axially split. As a general rule, antifriction bearings are installed in radially split housings, and plain
bearings in axially split housings. There are exceptions. Very large antifriction bearings often have
axially split housings to facilitate installation; vertical axis plain bearings will be installed in radially
split housings unless the size of pieces for dismantling dictates otherwise.
In most designs of overhung pumps, both the line and thrust bearings are in a single housing, and
the bearings are antifriction. The housing is therefore radially split and is either connected to the pump
by a bracket or adaptor (Fig. 11.22), or is part of a frame assembly which has the casing attached to it
(Figs. 2.18, 2.20). Since overhung pumps usually have radially split casings, the bracket or adaptor can
easily be designed for "full-circle" support, which is inherently stronger than the alternative "half-circle"
support. Cutouts in the bracket, with suitable compensating reinforcement, allow access to the shaft seal
and ventilation when necessary. Depending on the design of the casing and housing and the pump's
intended service, the bracket or adaptor is either integral with the casing (actually the cover); as Fig.
11.22, integral with the housing (Fig. 11.21), or a separate piece. Having the bracket integral with the
casing ensures it has the same material properties, which is an important consideration in corrosive
services and applications involving very high or very low temperatures. Incorporating the bracket in the
housing allows greater flexibility in the configuration and manufacture of the casing cover. A separate
adaptor allows the greatest design flexibility and can accommodate any special material requirements.
Bearings 265

Against this, there is an additional bolted joint and centering fit involved in maintaining rotor alignment
within the casing.
Between-bearings pumps can have their casing either axially or radially split and are equipped with
either antifriction or plain bearings. As a consequence, they employ a wider variety of bearing housing
configurations than pumps with overhung rotors. When the casing is axially split the bearing housings
are normally "half-circle" connected to the lower or fixed half of the casing. This is the simplest
arrangement for such a casing, and when properly designed has adequate stiffness for the class of pumps
that usually employ axially split casings. The bracket connecting the housing and casing can be integral
with the casing (Fig. 11.20) or part of the bearing housing (Fig. 11.63), with the choice being based
on materials and manufacturing considerations. When abnormal service conditions warrant the added
complexity, pumps with axially split casings are equipped with full-circle bearing housing support. A
common example is pumps for navy combat vessels, in which full-circle support is necessary to survive
high-shock loading. Modem pumps with radially split casings employ bearing housing and connecting
bracket arrangements similar to those previously described for overhung pumps. The only justification
for half-circle-supported bearing housings was easy access to packed box shaft seals when the bearing
housing was axially split for plain bearings. Since packed box seals are rarely used today in radially
split between bearings pumps, there is no reason not to employ the inherently stiffer full-circle support.
Axially split bearing housings are either bolted to a bracket or adaptor or have an integral bracket,
meaning that it is also axially split (Fig. 11.64). The latter arrangement yields a compact design yet
provides good access for rotor, bearing, and seal setting when the upper half of the housing is removed.

Figure 11.63 Axially split pump with half circle supported bearing housing.
Bracket integral with housing.
266 Bearings

Figure 11.64 Full circle supported axially split bearing housing.

The stiffness and strength of bearings housings and their connection to the pump is a critical but
often overlooked requirement. Stiffness has two aspects. First, the bearing housings and their connection
to the casing must be stiff enough to transmit the bearing loads to the casing (or foundation in those
few instances where pedestal bearings are used) while maintaining alignment of the pump's rotor within
the casing. When the bearing housing or adaptor has support built into or attached to it, thus producing
a three point-supported pump (Fig. 11.65), the bearing housing and its connection must now withstand
portion of the piping load as well. Second, the bearing housing connection must be stiff enough to have
its first natural frequency above the highest excitation frequency the pump is likely to produce, which
is usually the vane passing frequency. If the bearing housing's natural frequency is close to an even
moderately intense excitation frequency, the bearing housing will resonate, producing vibration of high
velocity. Although this vibration is generally of very low amplitude, 6 11m (0.00025 in.) or lower, and
does not seem to cause any actual damage, the velocities are well above the limits currently deemed
acceptable, and can therefore pose an acceptance problem. Beyond being stiff enough to accommodate
normal loads, bearing housings need to be strong enough to tolerate some degree of abnormal loading,
such as might be imposed by shaft failure or a similar incident. In the petroleum refining industry there
have been several major fires caused by a cast-iron bearing bracket fracturing under abnormal loading
and allowing gross leakage of a product whose temperature was above the autoignition temperature.
Similarly, there have been instances where a running pump was within a fire but not contributing to it
until thermal shock from fire extinguishing efforts fractured the cast-iron bearing housing. It is for these
two reasons that API-61O now specifies that pumps handling flammable or toxic liquids have steel
bearing brackets, housings, and load-carrying bearing covers.
Depending on the method of lubrication, bearing housings incorporate various features to retain or
Bearings 267

2 + 3

Figure 11.65 "Three-point"-supported pump.

direct the lubricant. Housings for grease lubricated bearings have either space to accommodate spent
grease as it is purged from the bearing (Fig. 11.17) or means of expelling the spent grease, such as a
grease escape valve (Fig. 11.19). With oil bath, flinger, and oil-ring lubrication, the bearing housing has
a sump to hold the necessary volume of oil. Some designs for flinger and oil-ring lubrication go a step
farther and include channels and associated ports to circulate oil from the sump, thru the bearings and
back to the sump (Fig. 11.21). Pure oil mist, oil spray, and force-fed oil lubrication eliminate the need
for a sump, and instead have galleries, often with flow control orifices, to supply the oil to the bearing
and ports to drain it away for return to the lubrication system (Figs. 11.37 and 11.38). The oil from
tilting pad thrust bearings passes to the OD of the thrust collar where it collects in an annulus around
the collar, and is expelled through the oil outlet back to the housing's drain system. Bearing design and
speed influence the location and form of the oil outlet. With conventional pressure-fed flood lubrication,
the oil outlet is in the upper half of the bearing housing, and is usually radial for speeds to 23 mls (4,500
ft/min) (based on the mean collar diameter), tangential for higher speeds (Fig. 11.66). When the oil is
brought directly to the thrust shoes, as in Kingsbury's type LEG bearing, there is no need to flood the
thrust bearing assembly so the oil outlet is in the lower half of the housing. As these bearings are only
used for high-speed applications, the oil outlet is always tangential.
Heat dissipation is a fundamental element of bearing housing design. Although the heat load involved
is usually quite low, so, too, is the capacity of the housing to dissipate heat, which means that the
temperature of the bearings depends on a delicate balance between heat load and dissipating capacity.
The sources of heat load for a typical bearing housing are shown in Fig. 11.67. Some discussion of the
various sources is warranted.

QI,2 The heat generated by the bearings themselves is related to their design and operating conditions.
Antifriction bearings, as generally used in centrifugal pumps, generate most of their heat from the
action of the bearing rolling elements on the lubricant. The only exception to this is large low-speed
bearings whose heat generation is principally a function of bearing load. Plain bearings develop
heat by shearing action in the lubricant film. Figure 11.67 shows labyrinth seals adjacent to each of
268 Bearings

CONVENTIONAL FLOODED
OIL oun.ET LOCATION OIL INLET· OIL INLET"

L.E.G. OIL 0UT1.ET LOCATION

Figure 11.66 Tilting pad thrust bearing oil outlet arrangements.


(Courtesy Kingsbury Inc.)

the bearings. Such seals are not contacting and therefore do not develop any heat. The same is not
true for so-called "positive" seals such as the lip type. These seals function by contact and can gen-
erate significant additional heat when their rubbing speed is high.
Q3 Conduction along the shaft is a factor in high-temperature applications. Many designs include a heat
dissipating thrower (Fig. 11.22) on the shaft between the pump's liquid end and the bearing frame
to draw off some of the heat before it enters the bearing frame. In other instances, the shaft under
the sleeve has been coated with a low-thermal-conductivity ceramic to impede heat transfer. This ap-
proach works as predicted, but the coatings have proved vulnerable during pump maintenance.
Q4 High-temperature applications also raise the possibility of conduction through the bearing bracket or
adaptor. With judicious design, such as an air gap between the faces (Fig. 11.22) to form a thermal
barrier, the heat load from this source can be minimized.
Qs In regions exposed to intense solar radiation, that same radiation can serve as a significant source of
heat into the bearing housing. Often, when this is the case, the bearing housing is equipped with a
sun shield to reduce the heat load from this source.

The means employed to dissipate the bearing housing heat load are:

1. Natural convection-For low-speed applications with lightly loaded bearings at pumping temperatures up
to, say, 120°C (250°F), natural convection to the atmosphere is sufficient to keep the bearing cool.
Bearings 269

Labyrinth
seal ..........

Figure 11.67 Heat balance on bearing housing.


KEY:
QI = line bearing heat input
Q2 = thrust bearing heat input
Q3 = conduction along shaft
Q4 = conduction through bearing bracket
Qs = solar radiation
Q6 = heat lost to convection, natural or
forced
Q7 = heat lost to cooling jacket
QI + Q2 + Q3 + Q4 + Qs = Q6 + Q7.

2. Forced convection-At higher speeds, such as those of pumps driven by two-pole electric motors, and
temperatures above, say, 95°C (200°F), natural convection will not keep the bearings cool. Raising the
velocity of the air over the housing by the use of a shaft mounted fan and the appropriate shroud significantly
increases the heat load the housing can dissipate. With thorough design, including fins on the housing to
increase the surface area, fan-cooled housing can maintain acceptable bearing temperatures for pumping
temperatures to 425°C (800°F) with ambient air of 43°C (110°F).
3. Cooling jacket or coil-Adding a cooling jacket around the bearing housing or inserting a cooling coil into
the housing sump is an alternative to fan cooling. A jacket around the housing (Fig. 11.21) is most effective
when the oil is circulated through the bearings by flingers and internal galleries; it ensures the oil delivered
to the bearings is cooled. Limiting the jacket to under the sump is usual when the housing does not include
oil circulation galleries. With such an arrangement it is crucial that the oil sump be well agitated by the oil
rings or flingers, otherwise the coolest oil will remain at the bottom of the sump where the cooling is being
applied. A principle common to all jacket-cooling arrangements is that the cooling be applied to the lubricant,
not around the bearing. Experience has shown that the latter, an old practice, keeps the bearing outer race
cool but can, by way of differential thermal expansion, lead to loss of internal clearance and consequent
bearing failure. Cooling coils are an alternative to a jacket under the sump, and provided the design is
correct, are probably a more effective approach, although at the expense of greater complexity.
270 Bearings

GREASE

'"

Figure 11.68 "Taconite" bearing housing seal.

4. Forced-oil circulation-For high speeds, high loads, high pumping temperatures, or various combinations
of these, forced-oil circulation is the most effective means of keeping bearings cool. The flow needed depends
on the bearing type and the sources of heat. Antifriction bearings operating at high speed but in a cool
environment require only a very low flow. The same can be said for plain bearings, although the flow is
nominally higher because of the nature of the bearing. When the pumping temperature is high, higher flows
are needed to dissipate the additional heat load.

The final aspect of bearing housings that needs to be discussed is seals. Often overlooked as a minor
detail, the means of sealing bearing housings, both against the ingress of dirt and the like and the egress
of lubricant, can be a major factor in the reliability, and consequent availability, of a pump. A fundamental
principle in all designs is that there be some form of thrower between the bearing housing and any
adjacent pumped liquid seal. If a thrower is not provided or its design is inadequate (e.g., an elastomer
that expands during rotation to give a gap between the thrower and the shaft), leakage from the shaft
seal can pass practically unimpeded into the bearing housing and so contaminate the lubricant. Many
seal designs are employed; only those representing the basic types are discussed here.
For pumps installed indoors, a simple labyrinth machined into the bearing cover (Fig. 11.17) is quite
adequate. The drain from the labyrinth back to the bearing housing is an important feature of this seal;
without it oil will accumulate in the labyrinth and eventually leak out of the housing. Pumps installed
in a moderately severe environment or pumps whose axis is vertical often require a "positive" seal, the
most common form of which is, the lip seal (Fig. 11.18). Lip seals have finite life, very definite limits
on rubbing speed and service temperature, and tend to wear the journal on which they are running.
Given these limitations, they obviously cannot be used for all applications. API-61O precludes their use
in pumps for refinery service. When effective housing sealing is necessary but is outside the capability
of lip seals, a labyrinth plus a thrower offers an effective alternative. The labyrinth is frequently made
renewable (Figs. 11.21 and 11.22) to allow for ready restoration in the event of damage, and to allow
a nonsparking material where needed.
If it is mandatory that the bearing housing be positively sealed from the atmosphere but lip seals
cannot be used, one alternative is the magnetically energized face seal (Fig. 9.24). These seals, used
extensively in the aircraft industry, offer a very effective, albeit moderately expensive, solution to positive
sealing. A second approach, born in the mining industry, is the so-called "taconite" seal. Figure 11.68
Bearings 271

shows a typical design, the essence of which is a grease-filled space between the bearing housing and
the atmosphere. Taconite seals have earned a notable reputation in pumps operating under severe
conditions on mining and mineral processing sites.

BIBLIOGRAPHY

[11.1] E. L. Annstrong, W. R. Murphy, and P. S. Wooding. "Evaluation of Water-Accelerated Failure in Oil


Lubricated Ball Bearings"; Journal of the ALSE (January 1978): 15-21.
[11.2] Bearing Installation and Maintenance Guide, Publication #140-70, SKF USA Inc., King of Prussia, PA,
August 1988.
[11.3] API-614, Lubrication, Shaft Sealing, and Control Oil Systems for Special-Purpose Applications, American
Petroleum Institute, Washington, DC, January 1984
12
Couplings

Centrifugal pumps are connected to their drivers through couplings of one sort or another, except
for close-coupled units, in which the impeller is mounted on an extension of the driver shaft.
Couplings are either rigid or flexible. The choice is determined by the bearing arrangement of the
pump and driver combination. Rigid couplings permit neither radial nor axial relative motion between
the driving and driven shafts, effectively making them a single shaft. Their use is therefore limited
to pump and driver combinations with two or three bearings (Fig. 12.1[a] and [b]). This category
includes lineshaft-driven vertically suspended pumps, which are considered two-bearing machines
with additional precision aligned guide bearings.
A flexible coupling, on the other hand, is a device to transmit torque between the two shafts
while allowing for minor misalignment (angular, parallel, or a combination) between their axes of
rotation. Contrary to some popular perceptions, flexible couplings are not intended to accommodate
gross misalignment between pump and driver. If that is envisaged, the machines should be coupled
with a universal drive shaft. In a general sense, misalignment between axes of rotation imposes
bending on the shaft and additional loads on the coupled machines' bearings, and therefore has to
be kept to a minimum. The purpose of a flexible coupling is to accommodate the minor misalignment
that is either impractical to eliminate or occurs during some transient condition encountered in the
pump's operation. The necessary accuracy of alignment depends on the coupling type, its installed
configuration, and the rotative speed, details that will be dealt with in discussion of the various
coupling types. Unless the size of the pump precludes them, pump and driver arrangements with
four bearings (Fig. 12.1[c]) are almost invariably equipped with flexible bearings. Depending on
their detailed design, three-bearing machines (Fig. 12.1[b]) may be equipped with a flexible coupling
allowing only angular misalignment.
A flexible coupling must also permit some lateral float of the shafts so that the two shaft ends
may move closer together or farther apart under the influence of thermal expansion, hydraulic float,
or shifting of the magnetic centers of electric motors, and so move without introducing excessive
thrusts on the bearings. This aspect of flexible coupling design will be discussed in greater
detail subsequently.

272

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
Couplings 273

J C

(a) Two bearing arrangement:


- coupling transmits torque and bending moment
- no misalignment.
J C
Coupling~
(Rigid) ~

Coupling
)
~L.....----...J~~
(b) Three-bearing arrangement:
- coupling transmits torque and radial reaction
- angular misalignment only (if necessary).

Coupling
)
~L...------'~ ~ --=---I~[p
(c) Four-bearing arrangement:
- coupling transmits torque only
1 - - - = - - 1

- angular and parallel misalignment.

Fig. 12.1 Coupling requirements for various machine bearing arrangements.

RIGID COUPLINGS

Threaded Couplings
In a threaded coupling (Fig. 12.2), the threaded shaft ends are screwed into each end of the coupling
and butted together at the center. When torque is applied, the shafts tighten against each other until the
friction force on the thread faces equals that produced by the torque. The coupling is relieved at the
center to reduce stress concentration, and vented to allow for displacement of lubricant and air as the
coupling is assembled. These simple couplings are widely used to connect lineshaft sections in vertically
suspended pumps of low to medium torque. For higher torque applications, on the order of 7,000 N-m
(62,000 lb. in), the effort necessary to dismantle the couplings after they have been in service usually
dictates some other design.

Clamp Couplings
The clamp coupling (Fig. 12.3) is a typical rigid coupling. It consists basically of a split sleeve
provided with bolts so that it can be clamped on the adjoining ends of the two shafts and form a solid
274 Couplings

Fig. 12.2 Threaded coupling.

connection. Both axial and circular keys are commonly incorporated in the clamp coupling so that the
transmission of torque and thrust is not made solely dependent upon the frictional grip.

Sliding Sleeve Couplings


A sliding sleeve coupling (Fig. 12.4) is made up of a slide fit sleeve that fits over each shaft end and
engages a key in each, thereby allowing the transmission of torque. Axial thrust is transmitted via a
split ring that locks into a groove in each shaft end, and is kept in place by the sleeve. Once in its correct
position, the sleeve is located with two retaining rings, one on each shaft. Sliding sleeve couplings are
often used instead of threaded couplings when difficulty of assembly and dismantling is deemed a
major issue.

Tapered Sleeve Couplings


The tapered sleeve coupling (Fig. 12.5) is a classical design employed for high-torque applications
where the coupling assembly must be tight and of a diameter not significantly larger than the connected
shafts. Each end of the shaft is tapered with a key, usually parallel to the axis, and has a split ring
groove just beyond the taper. The coupling has a corresponding taper in each end, and a threaded
extension for a coupling nut. The coupling is tightened by drawing each coupling nut against its split
ring. Once tightened, each coupling nut is locked to prevent it coming loose during operation.

Flanged Couplings
Quite a wide variety of pumps employ flanged rigid couplings, which may be just a single coupling
(Fig. 7.11) or a double-flanged spacer (Fig. 12.6). All designs transmit torque by way of shear in the
fitted coupling bolts or dowel bushings around the bolts (many also rely on friction between the coupling
faces to transmit a portion of the torque). Bending is transmitted by tension in the bolts. Smaller sizes
employ a premachined rabbet or spigot fit to maintain concentricity. (Because each flange assembly
Couplings 275

CLEARANCE BETWEEN
HALVES TO AllOW
CLAMPING ON SHAFTS

Fig. 12.3 Clamp coupling.


Split lock ring

_ Coupling nut

-Coupling

Fig. 12.4 Sliding sleeve coupling. Fig. 12.5 Tapered sleeve coupling.

Fig. 12.6 Double flange coupling (spacer or extension type).

276
Couplings 277

inherently has some eccentricity, double-flanged spacer couplings cannot maintain the same degree of
assembled shaft straightness as single-flanged couplings.) Larger sizes generally have plain faces, and
are carefully aligned then the bolt holes finished bored to size in situ.

Compression Couplings
All compression couplings rely solely on friction between the coupling and the connected shafts to
transmit torque and axial thrust. This requires that the coupling develop a substantial interference fit
with the shafts, and it is the means of developing the fit that distinguishes the various types of compression
couplings. For smaller sizes, designs using thread driven tapers have proven successful. In a typical
design (Fig. 12.7[a]), the central portion of the coupling is made up of a slotted bushing, bored to fit
the two shafts and taper machined on its outside diameter from the center out to both ends. The two
couplings halves are finish bored to match this taper. When the coupling halves are drawn together by
bolting, the bushing is compressed onto the two shafts, thereby producing the necessary interference fit.

(b)

(8)

Fig. 12.7 Compression couplings; (a) thread driven taper, (b) oil-injection.
278 Couplings

As coupling sizes increase, the radial force necessary for torque transmission is hard to achieve with
thread driven tapers, so shrink fitting or oil injection mounting is employed. Figure 12.7(b) shows a
vertical pump lineshaft coupling design that employs both methods. Each shaft end has a precision
finished stepped fit, and the coupling has corresponding stepped bores in each end. Assembly is carried
out with the shafting vertical, heating the coupling to shrink it on to the shaft ends. When dismantling
is necessary, oil injection is employed to dilate the coupling and push it off the shaft end (the small
annulus formed by the stepped fit serving as a hydraulic jack). Oil injection mounting and dismounting
involves high pressure oil, on the order of 1,000 to 2,000 bar, (15,000 to 30,000 psig), and significant
strain energy, and therefore must be undertaken with due care (see coupling mounting in this chapter).

FLEXIBLE COUPLINGS

A wide variety of flexible coupling designs are used to connect centrifugal pumps to their drivers. Despite
the variations, some order can be brought to the designs by first distinguishing between materials. There
are two classes: elastomer couplings, which employ some form of elastomer or polymer element to
realize flexibility, and all-metal couplings, arrangements relying on the flexure of or sliding between
metal components to achieve flexibility. Within the all-metal class, there is a further distinction. Couplings
that rely on flexure are nonlubricated, those that rely on sliding are lubricated. The net result is three
fundamental classes of flexible couplings. The designs or types commonly used for centrifugal pumps
are tabulated below. Figure 12.8 shows the nominal coverage of each class.

Coupling Class Usual Types


Elastomer Buffer, block, ring, sleeve
All-metal, nonlubricated Disk, diaphragm
All-metal, lubricated Spring-grid, gear

Choosing the type of flexible coupling most appropriate to a particular application requires consideration
of at least the following factors:

1. Torque to be transmitted
2. Rotative speed
3. Attainable alignment
4. Size and weight
5. Operating cycle
6. Type of driver
7. Service environment, i.e. temperature, atmosphere
8. Cost

How the commonly used designs relate to these application factors is dealt with in the discussion of
each design.

Elastomer Couplings
The principal advantages of elastomer couplings are freedom from the need for periodic lubrication,
and, provided the correct type is chosen, tolerance of quite high misalignment at moderate speeds.
Against these advantages, elastomer couplings tend to be larger than equivalent all-metal couplings,
Couplings 279

20,000
15,000
®
10,000
8,000

i-
7,000
6,000 O

::2 5,000 - . , . ,

a. 4,000
a:
I
3,000
® .,
"0
Ql
Ql

C.
m 2,000

CD
1,000

®
500

100 1,000 10,000 100,000 1,000,000 10,000,000


Torque -# in

100 1,000 10,000 100,000 1,000,000


Torque- Nm

Fig. 12.8 Flexible coupling type coverage.


(Based on average rating data-consult manufacturer for specific couplings.)
KEY:
1. Elastomer
2. All-metal, non-lubricated (disk and diaphragm)
3. All-metal, lubricated (spring-grid and gear.)

because the torque is transmitted through a low strength material, and therefore they are limited in speed
and at some point become more expensive, despite the latter's inherently more expensive design.
Compounding that, the size and weight of the larger elastomer coupling can pose problems with rotor
inertia and overhung weight. Finally, the elastomer itself can be sensitive to the operating environment,
thus precluding the use of certain materials or the coupling class in general.

Pin and Buffer Couplings

A pin and buffer coupling is a flexible coupling with pins attached to one half of the coupling; these
projects into the buffers, which are mounted in the half of the coupling on the other shaft (Fig. 12.9).
The buffers are made of rubber or other compressible material to provide the necessary flexibility. The
driving bolts have an easy sliding fit in the bushings; slight longitudinal variations are therefore taken
care of whereas slight errors in angularity are compensated for by the flexibility of the rubber. Because
flexibility is achieved with the elastomer in compression, which means the rubber must elongate since
it is essentially incompressible, pin and buffer couplings have a low tolerance of misalignment.
280 Couplings

Fig. 12.9 Pin and buffer coupling. Fig. 12.10 Lovejoy coupling.

Elastomer Block Couplings

Closely related to the pin and buffer coupling, elastomer block couplings transmit torque via elastomer
blocks in compression between "fingers" located alternately on each half of the coupling. The Lovejoy
coupling shown in Fig. 12.10 is a simple form of elastomer block coupling. Being relatively small, this
design has only four blocks that are made in the form of a cross. As the size increases it is necessary
to use more blocks, and the result is designs such as that shown in Figure 12.11. Elastomer block
couplings are torsionally soft, and the larger sizes are available with blocks of various hardnesses to
allow the coupling to be "tuned" to avoid torsional resonance, a potential problem when the drive is
pulsating. In the same manner as pin and buffer couplings, elastomer block couplings derive their
flexibility from displacement of the elastomer, which means they have high radial stiffness and therefore
low tolerance of misalignment.

Elastomer Ring Couplings

By taking an elastomer ring and putting it in compression during assembly, a practice termed "banding,"
a relatively simple design is produced that transmits torque by further compressing the loaded segments
of the ring, relaxing the compression on the unloaded segments. Figure 12.12 shows a typical design,
this version using radial bolts to provide the initial ring compression. Radial and angular flexibility is
achieved by flexure of the ring, and consequently the design has greater tolerance of misalignment than
Couplings 281

Fig. 12.11 Elastomer block flexible coupling. Fig. 12.12 Elastomer ring coupling.
(Courtesy Kop-Flex, Inc.). (Courtesy Lovejoy, Inc.)

elastomer block couplings. Despite this, it is recommended that the couplings be used in pairs if high
parallel misalignment is expected.

Elastomer Sleeve Couplings

Couplings of this design transmit torque by shear in an elastomer sleeve, an arrangement that offers
significantly higher misalignment capacity than elastomer buffer, block, and ring couplings. The most
common form of elastomer sleeve coupling (Fig. 12.13) has a split convex section sleeve. This confers
the highest torque capacity since the elastomer is acting at a large radius, but does limit the allowable
rotative speed. Resorting to a continuous diaphragm (Fig. 12.14[a]) raises the allowable speed, but at
the expense of added weight and complexity. Employing a concave section sleeve (Fig. 12.14[b]) is a
second means of raising the allowable speed, although in this case, the gain is at the expense of coupling
size since the elastomer is acting at a smaller radius. A third approach is to retain the split convex sleeve
but make it of a stiffer material, a variation that lowers the coupling's misalignment capacity. For
centrifugal pumps, elastomer sleeve couplings are widely used in mining and mineral processing, applica-
tions where pump speeds tend to be low, the service environment generally is not injurious to elastomers,
and the coupling 's negligible maintenance needs and high misalignment capacity offer significant advan-
tages.

All-Metal Couplings

Compared to elastomer couplings, all-metal construction offers smaller lighter designs for the same
torque, greater ultimate torque capacity, higher allowable rotative speeds, and greater tolerance of adverse
service conditions such as high temperature and some forms of atmospheric contamination. None of the
all-metal coupling designs has the misalignment capacity of elastomer sleeve couplings, but several have
greater tolerance than elastomer buffer, block, and ring couplings.
282 Couplings

Fig. 12.13 Sleeve-type clamped elastomer coupling.


(Courtesy Dodge Manufacturing Division, Reliance Electric)

Flexible Disk Couplings


Flexible metal disk couplings (Fig. 12.15) transmit torque by tension in a disk, which is alternately
bolted to each side of the coupling. To raise flexibility without overstressing, the disk is generally a
laminate of thin disks, often of stainless steel for corrosion resistance. Coupling flexibility is achieved
by flexure of the disk. A single disk element has high angular flexibility but is very stiff radially. It is
therefore necessary to use a pair of disk elements to accommodate both angular and parallel misalignment.
Since parallel misalignment is accommodated by an angular displacement at each disk element, the
parallel misalignment capacity of the coupling increases with separation of the disk elements. Because
they do not have internal clearances and do not wear during operation, flexible metal disk couplings can
be accurately balanced and will maintain their balance, thus making them suitable for very high rotative
speeds (see Fig. 12.8).

Flexible Metal Diaphragm Couplings


Instead of transmitting torque at a constant radius, as a flexible disk coupling does, a flexible diaphragm
coupling (Fig. 12.16) transmits torque from one radius to another. Shear stress in the diaphragm at the
smaller radius determines the minimum diameter of the diaphragm. To achieve the required degree of
flexibility, the diaphragm must have a certain ratio of maximum to minimum diameter, which means
that for the same torque a flexible diaphragm coupling has a larger outside diameter than a flexible disk
design. As is the case with flexible disk couplings, a single diaphragm element has high angular
flexibility and high radial stiffness, meaning that a pair of elements is necessary to accommodate parallel
misalignment. Diaphragms are either one piece, profiled, or a laminate of thin plates to provide high
flexibility without overstressing. Some designs rely on the high shear stress at the diaphragm's minimum
radius to serve as a "shear pin" in the event of pump seizure. Such designs need to ensure the transmission
Couplings 283

(a) Continuous diaphragm-type (b) Continuous concave sleeve-type.


(Courtesy Dodge Manufacturing Division, (Courtesy Falk Corporation)
Reliance Electric)

Fig. 12.14 Elastomer sleeve couplings.

Fig. 12.15 Flexible metal disk coupling. Fig. 12.16 Flexible metal diaphragm coupling.
(Courtesy Rexnord) (Courtesy Bendix Fluid Power Division)
284 Couplings

unit (diaphragms plus spacer between them) is retained in the event of diaphragm failure. And if the
environment is hazardous, the coupling parts need to be nonsparking.

Spring-Grid Couplings
In a spring-grid coupling (Fig. 12.17) torque is transmitted by bending in a tempered steel spring
element acting in slots in each of the coupling halves. The slots are shaped such that the span between
the points of contact with the spring in each hub decrease as torque increases, thus increasing the stiffness
of the coupling as torque increases. This provides torsional flexibility and a certain capacity for momentary
overload. Misalignment is accommodated by a combination of flexure of the spring-grid and movement
of the grid within the hub slots. The latter requires that the coupling be lubricated. Because the spring-
grid is very stiff in bending in one direction, the coupling produces high cyclic forces if subjected to
significant misalignment, and therefore should only be used when close alignment can be assured.

Gear-Type Couplings
A gear-type coupling (Fig. 12.18) transmits torque by the mesh of gear teeth cut on the outside
diameter of the hub with internal gear teeth cut into the cover. To allow angular displacement between
the hub and cover axes, without tooth interference, the hub teeth are usually barreled and crowned. In
the same manner as flexible disk and diaphragm couplings, a single gear mesh or engagement can only
accommodate angular misalignment, therefore a pair of meshes is required to accommodate parallel
misalignment. Such a coupling is termed "double engagement," and the greater the separation of the
meshes, the greater the misalignment capacity. Any misalignment of a gear-type coupling produces
sliding in the mesh, which dictates that the coupling be adequately lubricated. Failure to properly lubricate
gear-type couplings results in large forces and moments being imposed on the shafts and bearings of
the coupled machines, thus significantly increasing the risk of premature, even catastrophic, failure. For
low-speed applications, the usual lubricant is grease. At higher speeds, grease tends to separate so oil
is used. If interruption of operation for the purpose of relubricating cannot be tolerated, continuously
lubricated couplings are used. An example of this is an unspared boiler feed pump or one driven directly
from the generator shaft. Figure 12.19 shows a typical arrangement of such a coupling. As a general
rule, the maximum allowable misalignment of a gear type coupling is that which produces a peak sliding
velocity at the teeth of 1.S to 2.4 mls (S to 8 ft/sec). Velocities beyond this will cause rapid wear
regardless of the quality of lubrication. Gear-type couplings can be balanced to operate at high speed,
but wear of the teeth eventually allows the covers to move radially, thus putting the coupling out of
balance. For this reason, today most high-speed pumps are equipped with flexible disk or diaphragm
couplings. The province of gear-type couplings is very high torque severe service applications.

Limited End-Float Travel


Horizontal sleeve-bearing electric motors are usually not equipped with thrust bearings but rather
with babbitted faces or shoulders on the line bearings. The motor rotor, which is allowed to float, will
seek the magnetic center, but a rather small force can cause it to move off this center. This movement
may sometimes be sufficient to cause the shaft collar to contact the bearing shoulders, causing heat and
bearing difficulties.
This effect is particularly noticeable in electric motors of ISO kW (200 hp) and more. As all horizontal
centrifugal pumps are equipped with thrust bearings, it has become the practice to use "limited end-
float" couplings between pumps and motors in this power range to keep the motor rotor within a restricted
location. The motors are built so that the total clearance between shaft collars and bearing shoulders is
not less than 12 mm (O.S in). In tum, the flexible couplings are arranged to restrict the end float of the
Couplings 285

Fig. 12.17 Spring-grid coupling.


(Courtesy Falk Corporation)

~· - -rI
I

L
Fig. 12.18 Fast double engagement gear-type coupling.
286 Couplings

Fig. 12.19 Continuously lubricated gear-type coupling.

COUPUNG SLEEVES

MOtOR
BURING
FACE

Fig. 12.20 Limited end-float coupling.


Couplings 287

motor rotor to less than 5 mm (0.2 in). To keep the gap open between the shaft collar and the shoulders,
one of the following methods is used:

1. For gear or grid couplings-by a "button" at the end of the pump shaft or by a predimensioned plate
between the two shaft ends (Fig. 12.20).
2. For flexible-disk or diaphragm couplings-by the stiffness of the flexible disks themselves, which have
inherent float-restricting characteristics.

Contact between the hubs and the coupling covers prevents excessive movement in the opposite
direction in gear or grid couplings. The stiffness of the flexible disks or diaphragms is the restraining
force in both directions in these types of couplings.

Spacer and Floating-Shaft Couplings


Regular flexible couplings are designed to connect driving and driven shafts with a relatively small
distance between the shaft ends, and therefore have only limited tolerance of misalignment. In some
applications provision has to be made for greater misalignment, or in others the shaft end separation
must be significantly greater to allow dismantling of the pump. Such is the case, for example, with end-
suction pump designs in which the rotor and bearing assembly is removed by withdrawing it axially
toward the driver, an arrangement known as "back pull out." One of the principal objectives of back
pull out design is to allow dismantling of the pump without having to disturb either the pump casing
or the driver (Figs. 12.21 and 12.22). A second common example is the need to be able to remove the
inboard seal of a between bearings pump without having to disturb the pump or its driver (Fig. 12.23).
In either case, the distance between the pump and driver shaft ends has to be enough to allow the
dismantling. For this purpose, an easily removed spacer or extension of sufficient length is necessary.
Beyond providing room for dismantling the pump, an extension or spacer couplings is commonly

Fig. 12.21 Spacer (extension) coupling installed on back pull-out pump.


Fig. 12.22 Back pull-out assembly (bearing frame, casing cover, seal, and impeller) being removed. A spacer
or extension coupling enables the pump to be dismantled without disturbing either the driver or the pump casing.

Fig. 12.23 Between bearings pump with spacer coupling (spacer removed) to allow dismantling inboard
bearing and shaft seal.

288
Couplings 289

Fig. 12.24 Gear-type spacer (extension) coupling.

used for pumps handling hot liquids and therefore subject to thermal expansion and possible misalignment.
Their purpose in this case is to increase the separation of the flexible elements thereby avoiding the
harmful misalignment that would occur within the coupling with minimum separation of the driving
and driven shaft ends. Usually they consist of two single-engagement elements connected by a sleeve
(Fig. 12.24). Elastomer and spring-grid couplings are also furnished in spacer configurations. Most the
elastomer designs have the flexible coupling at one end of the spacer and a rigid coupling at the other
end (Fig. 12.25). With this arrangement, misalignment of the flexible coupling results in a force and

Fig. 12.25 Elastomer spacer (extension) coupling.


(Courtesy Falk Corporation)
290 Couplings

Fig. 12.26 Flexible drive shaft.

moment at the rigid coupling. Depending on the coupling type and the extent of misalignment, these
reactions can be high enough to cause fracture of the rigid coupling. Spacer versions of spring-grid
couplings have the coupling in the middle of the spacer, an arrangement that increases the overhung
weight and the potential for high reactions on the connected equipment shafts.
The floating-shaft coupling consists of two flexible elements connected by a shaft that must be supported
on each end by the flexible elements themselves. Different manufacturers use different approaches as
required by their basic coupling designs. For instance, each of the two couplings may be of the single-
engagement type, may consist of a flexible half-coupling and a rigid half-coupling at each end, or may
be completely flexible couplings with some piloting or guiding construction.
In the smaller horsepower field (below 19 kW [25 hpJ per 100 rpm), "flexible drive shafts" are
commercially available. These use universal joints at each end with a tubular floating shaft and a splined
portion to provide for length variation (Fig. 12.26).
The floating shaft and flexible drive shaft are frequently used in vertical dry-pit pumps, an application
that is discussed with that type of pump.

CLUTCH COUPLINGS

Regular disk clutches are rarely used to connect a centrifugal pump to its driver for two major reasons.
The first is that most clutch designs impose a high additional thrust load on the pump thrust bearing;
the second is that very accurate alignment between the clutch parts is necessary, and this is difficult to
maintain. The two designs of clutch couplings that are commonly used to connect centrifugal pumps to
their drivers are centrifugal and overrunning. A centrifugal clutch coupling (Fig. 12.27) has spring
retained shoes that throw out on rotation to transmit torque by friction with the drum. All but the smallest
of these couplings have "leading" shoes (friction between the shoe and the drum acts to drive the shoe
into the drum) to keep the size and weight of the coupling to a minimum. Overrunning or Sprag type
clutch couplings (Fig. 12.28) employ cam-shaped elements, which lock and transmit torque by friction
when torque is from the driving to the driven machine, but unlock to allow relative motion should the
driven machine tend to overrun the driving. Overrunning couplings of the type shown in Fig. 12.28
require very accurate machine alignment, tend to overheat if run in the overrun mode for extended
periods, and require that the whole unit be shut down to replenish or change the oil in the clutch. When
these limitations cannot be tolerated, a separate, foot-mounted totally enclosed overrunning type clutch
is used (see Fig. 12.29).
Couplings 291

Fig. 12.27 Centrifugal clutch coupling.


(Courtesy Ameridrives International, Centric Clutch Products-formerly Zurn Industries)

Fig. 12.28 Sprag-type clutch coupling. Sprags are kept in contact with members by energizing springs; wedge
tight for one direction of drive and release for the other direction.
(Courtesy Dana Corporation, Formsprag, Warren, Michigan)
292 Couplings

Fig. 12.29 Overrunning-type clutch with separate bearings.


(Courtesy Ameridrives Internationl, Marland Clutch Products-formerly Zurn Industries)

COUPLINGS FOR DUAL·DRIVE

In dual-driven pump installations, it is generally desirable to have one driver idle either to save power
or to save wear. Internal combustion engines, however, cannot be allowed to tum over idle and must
be disconnected. The ideal type of couplings for such units are those that can be readily disengaged
and reengaged.
The simplest means of coupling and uncoupling one driver in a dual drive arrangement is to employ
a disconnect or "cut-out" type of coupling. Figure 12.30 shows a manually operated, gear type version
of such a coupling. The left view shows the coupling in the connected position; the right view, in the
disconnected position. It is a quick and simple operation to release the location pins, slide the sleeve
into or out of engagement, and thereby connect or disconnect the driving unit and the pump.
If time is of extreme importance or if the starting of the standby driver is automatically controlled,
Couplings 293

Fig. 12.30 Fast's gear-type disconnect coupling.

the disconnect coupling needs to be servo operated, a complex arrangement needing care in its operation
and maintenance. The alternative is to use some form of automatic or "free-wheeling" clutch type
coupling. Of these, the simplest is the centrifugal clutch (Fig. 12.27), which is installed with the shoes
in the driven half. The centrifugal action of these shoes can be controlled to any predetermined speed,
and no shoe engagement takes place until this speed is reached. Above this speed, the coupling automati-
cally picks up the load. Centrifugal clutch type couplings are frequently used for dual driven fire pumps,
usually for the engine only, occasionally for both the motor and the engine.
A subtle variation of the dual drive arrangement is having a hydraulic power recovery turbine (HPRT)
driving a pump in tandem through an electric motor (Fig. 12.31). With the process operating at rated
flow, the HPRT produces power and so lowers the electrical energy consumed by the motor. As the
process flow drops, the turbine power also drops, and at some point the turbine starts to absorb power.
When this condition is reached it is necessary to automatically disconnect the turbine. The device
commonly used to do this is a separate foot-mounted overrunning type clutch (Fig. 12.29). An added
advantage of the separate foot mounted clutch in these cases is that it allows a positive separation of
the alternate driver from the train, thereby permitting safe work on it while the rest of the train remains
on-line.

Overrunning
Clutch
I
01 Motor

Fig. 12.31 Drive train of typical charge pump with hydraulic power recovery turbine.
294 Couplings

MOUNTING COUPLINGS

The means employed to mount a coupling hub on its shaft have a significant effect on machine balance,
the reliability of the drive, and the ease of subsequent maintenance. In order of increasing sophistication,
the following means are in common use: slide fit, taper-lock bushing, shrink fit, taper, and oil injection
or keyless.
Slide-fit mounting is convenient but is really only suitable for very small couplings. Torque is
transmitted by a key; the coupling is located axially with a headless set screw driven against the top of
the key. Machine balance is poor, and the clearance fit allows relative motion between the hub and the
shaft, which can lead to fretting corrosion. Taper-lock bushings (Fig. 12.32) offer the convenience of
slide-fit assembly with the integrity of a tight fit. Torque is transmitted by a key. The assembly is
tightened onto the shaft by drawing the taper-lock bushing into the hub with a pair or more of headless
set screws; loosened by moving the set screws to an alternate set of holes (see Fig. 12.32). Because the
maintenance of machine balance is not good and the coupling hubs are heavier, both a consequence of
the bushing, taper-lock mounting is generally limited to low-speed high-torque applications.
A high-integrity coupling mounting, capable of maintaining good machine balance, requires an
interference fit between the hub and shaft. Three means are employed to achieve this result. The simplest
is a cylindrical shrink fit (Fig. 12.33), in which the torque is transmitted by a key, and the coupling hub
is a light shrink fit on the shaft. Shrink fits involve the use of heat for assembly and dismantling. This
is not always convenient-a refinery pump that has to have its coupling removed in the field, for
example-so a tapered mounting (Fig. 12.34) is often used instead. Torque is still transmitted by a key,
but the interference is provided by drawing the coupling hub a predetermined distance along the taper.
In applications where the presence of a key in the coupling to shaft connection is undesirable (stress
concentration, hub size, balance), a keyless mounting employing oil injection is used. The fit may be
stepped cylindrical (Fig. 12.7[b]) or tapered (Fig. 12.35). In both cases, torque is transmitted solely by
friction between the shaft to coupling. The necessary interference is high, on the order of 0.00125 mm
per mm (in per in) of shaft diameter. Cylindrical fits (Fig. 12.7[b]) are assembled by shrink fitting;
tapered fits (Fig. 12.35) by oil injection dilation of the hub, then hydraulic advancement to give the
required interference. Dismounting is by oil injection for both designs. Successfully mounting and
dismounting keyless couplings requires the following:

Withdraw bushing

Fig. 12.32 Taper-lock bushing mounted coupling hub.


Couplings 295

L Interference fit
(otherwise require
secondary device to
lock axially)

Fig. 12.33 Cylindrical shrink fit mounted coupling hub.

1. Fit surfaces must be a ground finish and free of scratches


2. Tapered fits must have at least 85 percent contact
3. Adequate grooving for oil drainage in the fit region of tapered shafts, particularly if assembly is being done
in a cold environment
4. Ensuring the installed fit is correct, by either measuring the diameters before assembling cylindrical fits or
measuring hub advancement as tapered fits are assembled
5. Having the correct tooling-measuring apparatus, lifting devices, oil injection pump and connections, hydrau-
lic nuts, and a restraint for dismounting cylindrical fit hubs (the mounted hub has considerable strain energy
and moves off the last portion of its fit quite suddenly)
6. Taking safety precautions appropriate for the use of high pressure oil--equipment integrity, joint tightness,
venting air before pressurizing.

Taper: typically
0.625-0.750 in/ft on dia

Fig. 12.34 Taper-mounted coupling hub.


296 Couplings

Fig. 12.35 Oil-injection taper mounted coupling hub.


(Courtesy Flexibox International)

As the speed of the pump increases, so does the need for finer mechanical balance of its coupling.
This progresses through four stages. First, certain types are inherently limited in speed because their
construction does not achieve fine balance (see discussion of various types in this chapter). Second, the
precision of manufacture is raised (higher class number in the United States) to improve the balance of
the components. Third, each major component or sub-assembly is dynamically balanced. And fourth,
the entire assembly is dynamically balanced. In the oil industry, couplings for high power, high speed
pumps are generally purchased to API-671 [12.1].

COUPLING GUARDS

Left exposed, the coupling and the adjacent pump and driver shaft extensions pose a significant hazard
to the operator when the pump is running. To eliminate this, couplings and the adjacent shaft extensions
are enclosed in a metal guard, designed to prevent operators coming into accidental contact with the
moving surfaces. The guards are usually made of steel and supported from the pump base or foundation
in the absence of a base. Variations include partial mesh construction for visibility (but beware of small
parts falling through the mesh and being thrown back out at high velocity); a hinged opening for
inspection; heavy duty design (#10 gauge or thicker) to tolerate abuse; nonsparking materials (aluminum,
or bronze where there is concern over the potential for aluminum to spark under some circumstances);
and sealed construction for continuously lubricated couplings.
Couplings 297

COUPLING MAINTENANCE

The greatest contribution to coupling life comes from carefully aligning the coupled machines during
installation, then periodically checking the alignment during routine maintenance (see Chaps. 28 and 31).
Beyond checking alignment, couplings should be periodically inspected for wear of or damage to
their flexible elements. Abnormal wear must be investigated and the cause corrected. The usual causes
are poor alignment and lack of or incorrect lubrication. Damage, such as cracking of flexible disks or
diaphragms, requires careful investigation to ensure there is not an insidious influence, such as an
unexpected operating condition or an occasionally corrosive atmosphere, which could lead to sudden
failure. The damaged parts should, of course, be replaced. After each inspection, lubricated couplings
need to be relubricated.

MAGNETIC CLUTCHES, MAGNETIC DRIVES, AND


HYDRAULIC COUPLINGS

Magnetic clutches, magnetic drives, and hydraulic couplings are not couplings in the strict sense of the
word, as their function is to vary the speed of the driven unit rather than to provide merely a connecting
device between pump and driver.
Magnetic clutches are rarely used to connect a centrifugal pump to its driver because they require
accurate alignment and the few installations that have been made have not been very successful. Their
maintenance costs are also high. The only advantageous application of this device is in accumulator-
tank pumping or similar services for which the demand varies over a wide range. It is now the practice
either to start and stop the entire pumping unit or, if the cycle is too frequent for that, to allow the pump
to operate at reduced capacity during the period of small demand, incorporating a bypass so the capacity
will never fall below a safe value if the demand drops too low for proper operation.
Both hydraulic couplings and magnetic drives are used in centrifugal pumps if variations in operating
conditions warrant the use of variable output speed devices. Although they have approximately the same
overall efficiency as slip ring motors with speed control, they have the advantage of easily producing
any desired output speed, whereas the regular control for slip ring motors permits adjustment of speed
only by steps. A more complete discussion of these devices appears in Chapter 24.

BIBLIOGRAPHY

[12.1] API-671, 2nd Edition, Special Purpose Couplings, 1990. American Petroleum Institute, Washington, DC.
13
Baseplates and Other Pump Supports

For very obvious reasons, it is desirable that pumps and their drivers be removable from their mountings.
Consequently, they are usually bolted and doweled to machined surfaces that in tum are firmly connected
to the foundations. To simplify the installation of horizontal-shaft units, these machined surfaces are
usually part of a common baseplate on which either the pump or the pump and its driver have been pre-
aligned.

BEDPLATES
The primary function of a pump baseplate is to furnish mounting surfaces for the pump feet that are
capable of being rigidly attached to the foundation. Mounting surfaces are also necessary for the feet
of the pump driver or drivers or of any independently mounted power transmission device. Although
such surfaces could be provided by separate bedplates or by individually planned surfaces, it would be
necessary to align these separate surfaces and fasten them to the foundation with the utmost care. Usually
this method requires in-place mounting in the field as well as drilling and tapping for the holding-down
bolts after all parts have been aligned. To minimize such "field work," coupled horizontal-shaft pumps
are usually purchased with a continuous base extending under the pump and its driver (Fig. l3.1);
ordinarily, both these units are mounted and aligned at the place of manufacture.
As the unit size increases so does the size, weight, and cost of the base required. The cost of a
prealigned base for most large units would exceed the cost of the field work necessary to align individual
baseplates or soleplates and to mount the component parts. Such bases are therefore used only if
appearances require them or if their function as a drip collector justifies the additional cost. Even in
fairly small units, the height at which the feet of the pump and the other elements are located may differ
considerably. A more rigid and pleasant looking installation can frequently be obtained by using individual
bases or soleplates and building up the foundation to various heights under the separate portions of
equipment. (Fig. l3.2).
When a baseplate is used, whether it be under both the pump and its driver, or separate bases under
each piece of equipment, it is a fundamental element of the structural connection that maintains alignment

298

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
Baseplates and Other Pump Supports 299

Fig. 13.1 Baseplate under pump and driver, bolted to the foundation.

Fig. 13.2 Pump and driver on soleplates with stepped foundation.


300 Baseplates and Other Pump Supports

MACHINE LOADS
INTO PEDESTALS

FOUNDATION PROVIDES LOADS INTO FOUNDATION


HIGH BENDING AND VIA PEDESTALS,
TORSIONAL STIFFNESS SIDE BEAMS,
TO MAINTAIN ALIGNMENT FOUNDATION BOLTS 80 GROUT

Fig. 13.3 Base supported by "stiff' foundation.

between the pump and its driver. Although this is clearly a very important requirement, it is frequently
sacrificed to cost, invariably at the expense of machine reliability.
Baseplates for horizontal axis pumps fall into two categories, which for this discussion can be termed
"supported" and "suspended." The distinction revolves around how the base is connected to and supported
by the foundation. Supported bases, the more common category, are bolted and grouted to a "stiff'
foundation (Fig. 13.3). Once installed, the function of a supported base is to transmit equipment loads
and drive reaction to the foundation. The remainder of the structural connection to maintain alignment,
bending, and torsion, is then provided by the foundation. Given these functions, supported bases are
designed for

1. Moderate bending stiffness; sufficient for handling as an assembled unit without yielding (taking a perma-
nent set)
2. High pedestal and foundation connection stiffness; sufficient to maintain coupling alignment under the
imposed equipment loads (e.g., piping) and the drive reaction.

It is important to note that torsional stiffness is not a design requirement because the foundation provides
it. This reduces the cost of the baseplate, but increases the installation cost since the baseplate must be
carefully leveled after positioning it on the foundation (see installation in Chap. 28).
A noteworthy consequence of the deliberate lack of torsional stiffness is that precise shop alignment
of the equipment is meaningless; the alignment will change when the unit is moved. Two things are
important in the shop alignment of equipment on supported bases. First, the equipment mounting surfaces
must be carefully leveled before aligning the equipment to mark out the mounting surfaces for drilling
or to make a final check of alignment. Second, the equipment alignment must be accurate enough to
allow precise alignment in the field.
Suspended bases, the second and less common category, are those that do not rely on a foundation
for the bending and torsional stiffness necessary to maintain alignment. They are used in the follow-
ing circumstances:
Baseplates and Other Pump Supports 301

MACHINE LOADS
INTO PEDESTALS

c
BASE

BASE PROVIDES HIGH BENDING a TORSIONAL STIFFNESS


TO MAINTAIN ALIGNMENT

Fig. 13.4 Flexibility mounted baseplate.

1. Simplified installation-the unit is prealigned and can be positioned, connected, and started. Any attachment
to the foundation is nominal and usually three-point to be self-leveling.
2. Minimize structure-borne noise-The unit is suspended above the foundation on resilient mountings (springs
or elastomer pads).
3. Reduce piping loads-The unit is suspended on springs or stilts or free to slide so it can move to accommodate
piping expansion. A typical spring-mounted base is shown in Fig. 13.4.

Compared to supported bases, the additional design requirements for suspended bases are

1. High bending stiffness-sufficient to maintain pump to driver alignment under equipment weight and
imposed loads.
2. High torsional stiffness-sufficient to maintain pump to driver alignment under drive reaction and any torsion
from imposed loads.

These two additional requirements raise the baseplate cost. Both requirements add weight, although not
significantly if well designed, but providing torsional stiffness involves more difficult fabrication. To
achieve worthwhile torsional stiffness, the base must either have a closed cross section (undesirable for
corrosion resistance) or diagonal bracing. Conventional cross-bracing makes little or no contribution to
torsional stiffness, a fact quite evident in Fig. 13.5.
Today, baseplates are furnished in fabricated structural steel, fabricated stainless steel, cast iron, and
reinforced polymer. Regardless of the material, the principal design criterion is stiffness. Structural or
carbon-steel bases generally realize this at minimum cost by a simple arrangement of moderately heavy
pieces. In stainless steel, a more expensive material, lighter, more complex shapes are warranted to
minimize cost. Cast iron is only half as stiff as steel, so (he sections need to be heavier, with the increase
in weight sometimes offset by the ease of producing more complex shapes. Reinforced polymers are an
order of magnitude less stiff than steel, which limits their use in structures designed for stiffness. When
they are used, the sections and configurations are necessarily quite different to those for metals.
The virtue of fabricated baseplates is the flexibility of form allowed to the designer, ranging from a
simple inverted channel (Fig. 13.6) to a complex, suspended skid (Fig. 13.7). Structural or straight carbon
steel is the usual material since it is available in a wide range of shapes and sections, and is relatively
(CI

Fig. 13.5 Effect of section and bracing on torsional stiffness; (a) flat plate, (b) cross-bracing,
and (C) diagonal bracing.

302
Baseplates and Other Pump Supports 303

Fig. 13.6 Small frame-mounted centrifugal pump on inverted channel base.

Fig. 13.7 Offshore water injection pump with pump, driver, and accessories mounted on suspended,
three-point supported skid.
304 Baseplates and Other Pump Supports

SLOPE
- rA

SECTION 'A-A'

DRIP PAN

SECTION 'A-A'
(ALTERNATE SIDE BEAM)

SECTION 'A-A'

Fig. 13.8 Base drain arrangements.

inexpensive. Austenitic stainless steel is used when corrosion is a concern. Frequently pumps handle
liquids that cannot be allowed to fallon or accumulate on the foundation, since they would then pose
a problem of corrosion or combustion. In these cases, the base must also serve as a collector of incidental
leakage. Two arrangements are in use: drip pan and drain rim. Figure 13.8 shows the essential difference.
Drip-pan bases offer an extensive sloping drainage surface but require careful design to ensure structural
integrity and are difficult to fabricate. Drain-rim bases are usually easier to design and fabricate. Most
designs, however, suffer from the disadvantage of the regions where leakage falls being flat, and thus
prone to some leakage accumulation. Bending a "crown" into the top plate overcomes this, although at
some additional cost and fabrication effort. Openings in drip pans and the top plate of drain-rim bases
must be collared or bossed to avoid leakage through the opening.
Cast iron is restricted to small baseplates for standard pumps, where the quantity being produced is
sufficient to justify the pattern expense. Since the base is cast, it is relatively easy to produce a rimmed,
sloping drainage surface between the equipment mounting pads (Fig. 13.9). Reinforced polymer bases
are sometimes used in place of stainless steel when base corrosion is a problem. As with cast iron, this
material is only viable when the quantities are high enough to justify the mold cost. Because the shape
is molded, a rimmed, sloping drainage surface is easily incorporated in the design.
By definition, supported bases must be designed for grouting. If the base is well designed (meaning
that its pedestals do not rely on grout for stiffness), the essential functions of grouting are to

1. Ensure intimate contact between the base underside and the foundation.
2. Provide additional lateral restraint.

A secondary function is to fill voids in or under the base to prevent the accumulation of liquid or debris
or both. At one extreme the base is an open structure designed to be filled with grout (Fig. 13.10); at
the other, a closed structure designed for grouting to the underside of a drip pan (Fig. 13.11). Designs
with drip pans or deck plates require special features to ensure grout can completely fill the void beneath
Baseplates and Other Pump Supports 305

Fig. 13.9 Horizontal centrifugal pump and driver on cast-iron baseplate.

the plate or pan. If the void is not completely filled, there is a risk liquid will accumulate under the
plate, or the plate will "drum" and create unnecessary noise. Figure 13.12 shows the features necessary
for grouting under a drip pan.
Except for very small units, under, say, 225 kg (500 lb), the base generally includes provision for
lifting. In most cases the lift is four point, because rigging to equalize loading becomes complicated
with more than a four-point lift. The lifting lugs are positioned for balance. If equipment obstruction is
a problem, a spreader must be used for the lift.
Since baseplates are designed for stiffness, the volume of welding required in fabricated designs is
not high. Weld extent and size combine to give an actual weld volume greater than that required for
stiffness. Continuous welding is necessary for all joints. Intermittent welding should not be used because
joints so welded are prone to corrosion and subsequent distortion. Weld sizes are those necessary to
develop 50 percent of the plate strength.
As noted in the introduction to this discussion, the intention of a baseplate is to provide precision
surfaces on which to mount and accurately align the equipment. This function can only be realized when
the equipment mounting surfaces of the base are machined. When the fabrication involves extensive

Fig. 13.10 Grout-filled baseplate. Fig. 13.11 Grouted drip-pan baseplate.


306 Baseplates and Other Pump Supports

UNDERPAN STIFFENERS
(WITH CLIPPED CORNER,
GROUT OPENING@ CENTRE)
/
/
+ +
~ ,
I
,
+ l
I
J +
I

~i
I
I
I -$-rn -$-
I c::=:=::J
I
I
I
:~
I

+\
I
+ I \ 1-
'L /
\ /
!SIN. DIA OPENINGS 1/2" DIA DRILLED HOLES
OFFSET IF NECESSARY FOR FOR GROUT VENT
ACCESS BENEATH EQUIPMENT MAX SPACING 36"
COLLARED, 112" HIGH (ci)HIGH POINT IN EACH
'COMPARTMENT'
NO COLLAR UNLESS
SPECIFIED
API-610 REQUIRES
AT LEAST 1-19 SQ. I N.
OPENI NG I N EACH
UNDERPAN 'COMPARTMENT.'
NOT ALWAYS PRACTICABLE
10. OPENING BENEATH EQUIPMENT

Fig. 13.12 Design features necessary for grouting under a drip pan.

welding, it is usual to oven-stress relieve the base before machining. Doing this eliminates the risk of
subsequent distortion as residual stress is relieved over time. The separation of the finished machined
surfaces normally provides for at least 3 mm (0.12 in.) of shims under the driver. Provided the base is
not "sprung" (twisted out of shape) on the machine tool, the machined surfaces in each plane will be
coplanar within 0.15 mm per meter (0.002 in. per foot) of separation, a common specification requirement.
When deemed necessary by the designer or the purchaser, baseplates are furnished with a number of
t:efinements to aid installation and equipment alignment. Typical of these are leveling screws adjacent
to the foundation bolts in supported bases and jacking screws on the pedestals of both supported and
suspended designs.

CENTERLINE SUPPORT

For operation at high temperatures, the pump casing must be supported as near to its horizontal centerline
as possible to minimize the consequences of thermal expansion of the casing. Failure to do this will
result in distortion of the pump if it is three-point supported (Fig. 13.6) and misalignment of the pump
to its driver, both of which can ultimately cause significant damage to the pump. Centerline support is
generally adopted when the pumping temperature reaches 175°C (350°F). The real criterion, however,
is not the temperature, but the thermal expansion as the casing comes up to temperature. Unusually large
pumps therefore require centerline support at temperatures below 175°C (350°F).
Centerline support complicates baseplate design significantly, because the tall pedestals must be
sufficiently stiff to accommodate loads imposed on the pump without any significant change in the
alignment of the pump and its driver. This is particularly difficult with single-stage overhung pumps
Baseplates and Other Pump Supports 307

Fig. 13.13 Centerline-supported overhung process pump.

(Pig. 13.13), which are effectively only 2-point supported. To develop the required stiffness, the pedestals
usually must be closed (box section) and directly connected to the side beam of the base. Units with a
large amount of accessory equipment around them often require baseplates so large the pedestals cannot
be directly connected to the side beams. In these cases, the connection between the pedestals and the
side beams must achieve the same stiffness, which is usually done with judiciously designed lateral and
longitudinal bracing underneath the drip pan or deck plate. Inadequate stiffness in bases for centerline
supported pumps has caused a great of deal of difficulty in the refining industry. So much, in fact, that
API-61O [3.1], the usual industry standard for refinery pumps, now includes requirements for combined
base and pump stiffness, and specifies a simple means of testing to verify the design.
At temperatures significantly above 175°C (350 0 P), say 290 to 315°C (550 to 600 0 P), many designs
have added water cooling to the pedestals, the idea being to remove any heat passed to the pedestals
by conduction and convection from the adjacent casing. Tests to measure the amount of heat gained by
the cooling water show that it is so little as to be of no practical benefit. In the light of this, it is possible
to simplify the installation of high-temperature pumps by eliminating water-cooled pedestals.

SOLEPLATES

Soleplates are cast-iron or steel pads located under the feet of the pump or its driver and embedded into
the foundation. The pump or its driver are doweled and bolted to them. Soleplates are customarily used
for vertical dry-pit pumps and also for some of the larger horizontal units to save the cost of the large
bedplates otherwise required.

HORIZONTAL UNITS USING FLEXIBLE PIPE CONNECTIONS

The foregoing discussion of bedplates and supports for horizontal shaft units assumed their application
to pumps with piping setups that do not impose hydraulic thrusts on the pumps themselves. If flexible
308 Baseplates and Other Pump Supports

Fig. 13.14 Vertical-shaft installation of double-suction single-stage pump.


Casing is provided with mounting support flange .

pipe connections or expansion joints are desirable in the suction or discharge piping of a pump (or
in both), however, the pump manufacturer should be so advised for several reasons. First, the pump
casing will be required to withstand various stresses caused by the resultant hydraulic thrust load.
Although this is rarely a limiting or dangerous factor, it is best that the manufacturer have the
opportunity to check the strength of the pump casing. Second, the resulting hydraulic thrust must
be transmitted from the pump casing through the casing feet to the bedplate or soleplate and then
to the foundation. Usually, horizontal-shaft pumps are merely bolted to their bases or soleplates so
that any tendency to displacement is resisted only by the frictional grip of the casing feet on the
base and by relatively small dowels. If flexible pipe joints are used, this attachment may not be
sufficient to withstand the hydraulic thrust. If high hydraulic thrust loads are to be accommodated,
the pump feet must be keyed to the base or supports. Similarly, the bedplate or supporting soleplates
must be of a design that will permit transmission of the load to the foundation. (For a more
complete discussion of flexible expansion joints, see Chap. 28.)
Baseplates and Other Pump Supports 309

Fig. 13.15 Large vertical-shaft double-suction single-stage pump. Note: outrigger supports for driver,
and tooling for removing and installing front-half casing.
(Courtesy Thompsons, Kelly and Lewis Pty. Ltd.)

BASES AND SUPPORTS FOR VERTICAL PUMPING EQUIPMENT

Vertical-shaft pumps, like horizontal-shaft units, must be firmly supported. Depending on the installation,
the unit may be supported at one or several elevations. Vertical units are seldom supported from walls,
but even that type of support is sometimes encountered.
Occasionally, a nominally horizontal-shaft pump design is arranged with a vertical shaft and a wall
used as the supporting foundation. The regular horizontal shaft unit shown in Fig. 13.9 could be used
for this purpose without modification, except that the bedplate is attached to a wall. For such installations,
it is advisable to lock the pump feet to the bedplate by keys or dowels rather than to rely strictly on the
friction between the pump feet and the pads of the bedplate. Of course, it is assumed that careful attention
will have been given to the arrangement of the pump bearings to prevent the escape of the lubricant.
Installations of double-suction single-stage pumps with the shaft in the vertical position are relatively
rare, except in some marine and navy applications and waterworks installations where floor space is at
a premium. Hence manufacturers have very few standard pumps of this kind arranged so that a portion
of the casing itself forms the support (to be mounted on soleplates). Figure 13.14 shows such a pump,
which also has a casing extension to support the driving motor. As the size of the pumps arranged in
310 Baseplates and Other Pump Supports

this manner increases, so does the need to pay particular attention to the design of the pump casing, the
pump support beneath it, and the motor support above it to ensure the structural stiffness is high enough
to maintain alignment and avoid resonant vibration. For large pumps, typically those for waterworks, it
is sometimes necessary to provide outriggers (Fig. 13.15) to achieve the required stiffness.
A complete discussion of the methods of supporting pumps that are specifically designed for vertical
mounting is given in Chapter 14.
14
Special Designs: Vertical Pumps

Preceding chapters on centrifugal pumps with horizontal-shaft construction should not obscure the
fact that many centrifugal pumps utilize vertical-shafting. Vertical-shaft pumps fall into two separate
classifications: (1) dry pit and (2) wet pit. The former operate surrounded by air, whereas the latter are
either fully or partially submerged in the liquid handled.

Vertical Dry-Pit Pumps


Dry-pit pumps with external bearings include most medium and large vertical sewage pumps, most
medium and large drainage and irrigation pumps for medium and high head, many large condenser
circulating and water supply pumps, many marine pumps, most nuclear reactor primary cooling water
circulating pumps, and an increasing number of vertical in-line pumps for petrochemical and refining
applications. A related design, hermetically sealed dry-pit pumps with integral motors, and hence internal
bearings, is employed for high-pressure circulating services in power generation and hydrocarbon process-
ing (see Chap. 24). Vertical shaft designs are usually justified on the following grounds:

1. Floor space is limited; marine pumps are a particular example.


2. Suction conditions dictate that the pump be mounted at a low level, whereas the installation as a whole
requires that the driver be mounted at a high level.
3. Simplified installation and relative immunity to piping loads; afforded by having the driver mounted directly
on top of the pump.
4. Pump size; large vertical shaft overhung pumps achieve higher efficiency (no obstruction in the impeller
eye) and are more economical to manufacture and install than equivalent horizontal-shaft designs.

Many vertical dry-pit pumps are basically horizontal designs with minor modifications (usually in
the bearings) to adapt them for vertical-shaft drive (see Chap. 13, Figs. 13.14 and 13.15). The reverse
is true of small- and medium-sized sewage pumps; a purely vertical design is the most popular for that

311

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
312 Special Designs: Vertical Pumps

Fig. 14.1 Small vertical sewage pump with


intermediate shafting. Fig. 14.2 Section of pump in Fig. 14.1.

service. Most of these sewage pumps have elbow suction nozzles (Figs. 14.1-14.3) because their suction
supply is usually taken from a wet well adjacent to the pit in which the pump is installed. The suction
elbow usually contains a handhole with a removable cover to provide easy access to the impeller.
To dismantle one of these pumps, the stuffing box head must be unbolted from the casing after the
intermediate shaft or the motor and motor stand have been removed. The rotor assembly is drawn out
upward, complete with the stuffing box head, the bearing housing, and the like. This rotor assembly can
then be completely dismantled at a convenient location.
Vertical-shaft installations of single-suction pumps with a suction elbow are commonly furnished
with either a pedestal or a base elbow (see Fig. 14.1). These may be bolted to soleplates or even grouted
in. The grouting arrangement is not too desirable unless there is full assurance that the pedestal or elbow
will never be disturbed or that the grouted space is reasonably regular and the grout will separate from
the pump without excessive difficulty.
Vertical single-suction pumps with bottom suction are commonly used for larger sewage, water
supply, or condenser circulating applications. Such pumps are provided with wing feet that are bolted
to soleplates grouted in concrete pedestals or piers (Fig. 14.4). Sometimes the wing feet may be grouted
Special Designs: Vertical Pumps 313

Fig. 14.3 Vertical sewage pump with direct mounted motor.

right in the pedestals. These must be suitably arranged to provide proper access to any handholes in the
pump and to allow clearance for the elbow section nozzles if these are used.
Vertical in-line pumps (Fig. 14.5) are also generally single suction but have an elbow-type suction
nozzle to produce a side-suction flange that is in line with the discharge flange. This configuration is
used for a variety of applications ranging from household hot water circulating pumps through hydrocarbon
processing to medium-size pipeline pumps. Small vertical in-line pumps are supported by the piping;
larger sizes have an additional support built into the bottom of the casing.
If a vertical pump is applied to condensate service or some other service for which the eye of the
impeller must be vented to prevent vapor binding, a pump with a bottom single-inlet impeller is not
desirable because it does not permit effective venting. Neither does a vertical pump employing a double-
suction impeller (Fig. 14.6). The most suitable design for such applications incorporates a top single-
inlet impeller (Fig. 14.7).
If the driver of a vertical dry-pit pump can be located immediately above the pump, it is often
314 Special Designs: Vertical Pumps

Fig. 14.4 Vertical bottom-suction volute pumps with lineshaft drive.

Fig. 14.5 Close coupled (extended motor shaft) vertical in-line pump.
Special Designs: Vertical Pumps 315

Fig. 14.6 Vertical double-suction volute pump with Fig. 14.7 Section of vertical pump with top single
direct mounted motor. suction impeller.

supported on the pump itself (see Fig. 14.3). When the driver is so mounted, there are three common
arrangements for the pump and driver shafts:

1. Separate shafts, flexibly coupled-The pump and driver each have their own radial and thrust bearings.
Some designs employ a spacer-type coupling and an open-sided driver support (Fig. 14.8) to allow removal
of the pumps bearing frame and impeller without disturbing the driver.
2. Separate shafts, rigidly coupled (Fig. 2.16)-Usually the pump rotor is supported by the driver's bearings.
When a mechanical shaft seal is used, the coupling is often a spacer type to allow replacement of the seal
without having to disturb the driver.
3. Extended driver shaft (Fig. 14.5)-The impeller is mounted directly on the driver shaft and is supported by
the driver's bearings. If a mechanical shaft seal is used, the pump must be dismantled to replace it.

Although the driving motors are frequently mounted right on top of the pump casing, one important
reason for use of the vertical-shaft design is the possibility of locating the motors at an elevation
sufficiently above the pumps to prevent their accidental flooding. The pump and its driver may be
separated by an appreciable length of shafting, which may require steady bearings between the two units.
It is extremely important that these steadying bearings be rigidly supported and maintained in strict
316 Special Designs: Vertical Pumps

Fig. 14.8 Separately coupled vertical in-line pump.

alignment. The support is generally provided by horizontal structural steel beams tied into the wall
structure, although occasionally a similar vertical support is used. For proper operation of the vertical
shafting, the deflection of the vertical guide bearings under any operating conditions must be kept within
the limits set by the design of the shafting and the operating speed. In small units, a channel located
between the walls of the station usually gives adequate support in all directions. Larger units with larger
reaction loads on the guide bearings may require two channels or beams with lattice bars. Some
installations incorporate reinforced concrete beams in the structure. Naturally, if the design of the building
requires the construction of an intermediate floor, this floor can be used to support the guide bearings.
The most common shafting connecting a small- or medium-size centrifugal pump with its driver
makes use of the universal joint with hollow tubing (Fig. 14.9). The lower section has a universal joint
at both ends whereas the upper sections (if more than one is used) have a guide bearing supporting the
lower end and a universal joint at the upper end. Such shafting compensates for angular misalignment
and, as the lower section incorporates a splined joint, also compensates for any minor discrepancy in
length. If speed permits, shaft sections as long as 3.0 m (lOft.) or more can be obtained. Sections longer
than 3.0 m (10 ft.) are easily sprung and must be handled carefully. As this shafting does not transmit
thrust, both pump and driver must have a thrust bearing.
Although a vertical motor may be mounted directly on soleplates grouted into the floor, a separate
Special Designs: Vertical Pumps 317

stand is sometimes necessary so that the motor may be raised to provide access to the coupling.
Occasionally, removable beams are placed directly across a large opening in the floor to serve as the
motor mounting. This method permits easy access to the pumps for servicing and simplifies lowering
them into place during the initial installation.
A driver supported on a stand above the floor provides access to the flange connection and upper
universal joint for bolting purposes and for relubrication. If the driver uses hollow-shaft rather than
solid-shaft construction, it must be provided with a head shaft guided by a lower bearing to act in the
same capacity. The weight of this shafting (excluding that of the lowest universal joint) is carried on
the motor; provided it is not extremely long, the total weight involved is relatively small and a normal
thrust motor can be used. When shafting weight is a concern, hollow fiber reinforced polymer (FRP)
shafting is an alternative that offers lower weight, albeit at higher cost, than equivalent hollow-steel
shafting. Actually, hollow shafting is more expensive than solid shafting. But the basic universal joint
is so widely used on automobiles and trucks that it is somewhat of a production item, and the increase
in cost it entails over solid shafting is very reasonable.
Units requiring more torque in their intermediate shafting than can be carried by the available sizes
of universal-joint shafting use solid shafting, either with solid or with flexible couplings. If solid or rigid
couplings are used, only one thrust bearing is needed (usually in the driver), and all other bearings are
merely guide bearings. This shafting has the disadvantage of requiring very accurate alignment of all
bearings, a difficult feat for open shafting employing more than three bearings.
Solid vertical shafting using flexible couplings usually consists of several shaft sections (including
pump and driver), each having two or possibly three bearings connected by floating shaft sections and
a piloted or guided flexible coupling at each end, thus acting in effect like a universal joint (see Fig.
14.4). Naturally each section has to have a thrust bearing to carry the weight of the shaft section.
The intermediate shafting for large pumps requiring large shafts is usually of solid construction with
solid flanged couplings that are often forged onto the shaft section (Figs. 7.11 and 14.10).
The size of the shafting used for an installation is initially determined by the torque to be transmitted.
However, if a certain span between bearings is desirable because of existing supports (floors or beams),
a shaft larger than that required by the torque may be necessary so that the operating speed will be
sufficiently below the critical speed. It is thus general practice to have the first critical speed ("first
bending natural frequency" in modem parlance) above the highest operating or runaway speed of the
pump. The critical speed of a vertical solid shaft is a direct function of the diameter and an inverse
function of the square of the span between bearings. Thus if a shaft is to run at twice the speed of
another, it must be twice as large in diameter for the same bearing span, or its permissible bearing span
will be reduced to 70 percent of that permissible with the lower speed.
Bearings for vertical dry-pit pumps and for intermediate guide purposes are usually antifriction
bearings that are grease lubricated to simplify the problem of retaining a lubricant in a housing with a
shaft projecting vertically through it. Typical ball steady bearings used as intermediate shaft steady
bearings are shown in Fig. 14.11. Larger units, for which antifriction bearings are not available or
desirable, use self-oiling babbitt bearings or forced-feed-oiled babbitt bearings with a separate oiling
system (Fig. 14.12 and 14.13). Figure 14.13 illustrates a vertical dry-pit pump design with a single-
sleeve type line bearing. Th~ pump is connected by a rigid coupling to its motor (not shown in the
illustration), which is provided with a line and a thrust bearing.
The supports for the guide bearings of vertical shafting connecting a centrifugal pump and its driver
must be sufficiently rigid. The radial load is usually assumed to be the same as if the unit were in a
horizontal position. With this loading, the deflection of the supports in any direction should not exceed
A in the following equation:
318 Special Designs: Vertical Pumps

MOTOR

MOTOR
STANO

t
STEADY
BEARING
I

PUMP

DISCHARGE

SUCTION

Fig. 14.9 Elevation of vertical pump Fig. 14.10 Elevation of vertical pump with solid
with tabular lineshaft. lineshaft. Motor supports rotating parts.

where d =the deflection in mm


NCI =critical speed of the shafting in rpm.
or for US units

where d = the deflection in inches.


Special Designs: Vertical Pumps 319

Fig. 14.11 Ball bearing used for intermediate shaft guide bearings.
(Courtesy Seal Master.)

This critical speed is usually 125 percent of the pump rotative speed or some value above the possible
runaway speed to allow for back flow through the pump. If beams or channels support the bearings, the
design of the latter naturally depends on the span between them, the radial force, and the permissible
deflection. Small pump installations with short spans usually require a single channel (to which vertically
mounted bearings are most easily attached). Larger units with long spans often require fairly widely
spaced channels or beams with strengthening lattice work (for which a horizontally mounted bearing
resting directly on the beams or on a bridging plate is more convenient). These considerations tend to
make vertically mounted bearings preferable for small units and horizontally mounted bearings preferable
for large units.
Vertical dry-pit centrifugal pumps are structurally similar to horizontal-shaft pumps. It is to be noted,
however, that many of the very large single-stage single-suction (usually bottom) volute pumps that are
preferred for large storm water pumpage, drainage, irrigation, sewage, and water supply projects have
no comparable counterpart among horizontal-shaft units. The basic U-section casing of these pumps,
which is structurally weak, often requires the use of heavy ribbing to provide sufficient rigidity. Some
high-head pumps of this type have been made in the twin-volute design. The wall separating the two
volutes acts as a strengthening rib for the casing, thus making it easier to design a casing strong enough
for the pressure involved (see Fig. 2.9). Another approach is to borrow from water turbine practice and
employ a stay ring, a set of vanes between the impeller and casing, to act as a strut (Fig. 14.14). When
used in a pump, the stay ring is designed as a diffuser (see Chap. 2).
As the size of vertical dry-pit pumps increases, a point is reached where it becomes more practical
to form the casing as an integral part of the pump's foundation, an arrangement known as a "concrete
volute pump." Figure 14.15 shows such a design. The volute shape is produced by either embedding a
fabricated steel shell or pouring around conventional concrete form work. Most designs employ a stay
320 Special Designs: Vertical Pumps

f
; I
1-) I
L- _ .---&.1_ _ -,

Fig. 14.12 Self-oiling steady bearing for large vertical shafting.

Fig. 14.13 Section of large vertical bottom-suction volute pump with single guide bearing.
Special Designs: Vertical Pumps 321

........ ,... ...


····, ...
. ... , ' , .
... .... .. :.,:..': . :" : . :"
', '

:" .'
:.::. :":' :":' :.,', " . .... ... .. .
.:. ':" .:. :.'" ': .. ':' .': ..:.:.'...:.:: ::. ::.::"
", ':... :. ", ", ", ... .-: ..-:..-:...:...:.::..
" ,

, ..... . . . .... ;: . .'. ',. .': .':" ': ,': .


.· ......
. . . . ... ..... . . .
,

, .. , ..... '.
..
,'. , • " ' , ' 0' • " .' " " .'
·. .. . .. .. ., .....
,
. . .. ,. .. . ., ..
'. ' ',' ' . '

.... ....... ..........' ..


,'.'
...... ,': .. ', .' .. '.
' ' . '.'

"."'

-nroe,. 1'>'1;1 ..... ;


......... p..
.., ' ,',' ' , ' , ' , ' , -, ' . ' , ' :
" " " " , ' , ' , ,' .' Rat"" P"'!9 perlonnance .... ..... '-....
", :', ':" :'::'::'a':' "" .:.... . . . Flow 11 m~.
..
' ', . ". ':" :" :" ' , ', " Head BOm
,':' ':',':''':''':':.',.. ':' ........... : ..: ..: ..
", " " "
Speed 4!50rpm
", ....... :" :" :''':' :...:.::.:> .... :.'
':;~~::;~,;~:;~:;::;::.~;:..,~
"'..., . ,. ....,.....,.....,..."T"". -'"'~~
· :', :"
'" '" ....
'"
~~=~R1rt~3
. .. ....:', .
'" "
.
·': . : .': :.' . : :: ~.: ;'".: ",". . ,'.
·':'::" ':"-:" ':" ':'" ':': '" ..... ,'

.., '..'.., .... .. .


' . .. ' . ' , . ' . '
,
. .. . .. ' . ' .
' '
,

.. . ... ... . . .....


' '
. . . .
..' ....:.:.::: :::':::':.':>.,':::: .
.................... : .. ......: .
,"

' .: :.:. , :':.::~: :::: :::: :\ :::: :.:::.:. : :.


:~..., 'o!l'c-IR-o>JIg,·,
: ::
'. '. .. .. .. .. .. '. •. '.
' . ' ... •... : ,:', " . '" ". '.:
. , • . • • ' • • ••
, ' f" " '.

Fig. 14.14 Section of large vertical bottom-suction pump with stay ring. (Courtesy Voith)
....-
-==-

I
LOWER FLOOR
.
'
,
'. ' j
.
TOP OF 8EAR1NG COUAR
. -..;. ......
, OAOUT~
ACCESS8TEP
I-~

PRIMARY CONa:lETE
'fl

----
I

I .......
BOTTOM OF PEDESTAl

.- '.: ~'
.....

I I

1r
.. OIsalAROE

i~
, I

I"@)
n ,

.r-$-J
RUNNING CLEARANCE DIAMETERS
;
PRIIolARY CONCRETE
Impeller to Cover 6.0 mm _1 I
Impeller to Wear Ri1g 2.2 mm on diameter

Fig. 14.15 Section of bottom-suction concrete volute pump.


(Courtesy Thompsons, Kelly and Lewis)

322
Special Designs: Vertical Pumps 323

Fig. 14.16 Wearing ring construction with extended skirt for large vertical pumps.

ring to support the pump's bearings and its driver. Concrete volute pumps have been used quite extensively
for high-flow duties, typically above 15,000 m3/hr (65,000 gpm), such as condenser circulating, water
supply and irrigation, where they can offer a more economical installation than conventional dry-pit
pumps or large "pull out" wet-pit pumps.
Vertical pumps equipped with bottom single-inlet impellers (see Fig. 14.2) have a leakage joint
between the wearing ring hub of the impeller and the suction head. When pumps of this type handle
gritty water, the grit separates out during periods of shutdown and concentrates at or near this joint. As
soon as the pump is started again, this concentration of grit is washed through the leakage joint, causing
wear. Large pumps may resort to a ring construction like that shown in Fig. 14.16, in which the stationary
ring is extended above the suction head to form a pocket for the grit to be deposited in and from which
it can be periodically flushed. A further aspect of large pumps is that they are often started dry or
"dewatered" to reduce their starting torque. When this is the intention, the wearing rings are designed
to be flushed during start-up so that any incidental contact within the running clearance does not cause
damage or seizure (see Chap. 4, Figs. 4.28 and 4.31). These and other refinements are feasible in large
but not in small pumps.

Vertical Wet-Pit Pumps


Vertical pumps intended for submerged operation are manufactured in a great number of designs,
depending mainly upon the service for which they are intended. Thus wet-pit centrifugal pumps can be
classified in the following manner:

1. Vertical turbine pumps


2. Propeller or modified propeller pumps
324 Special Designs: Vertical Pumps

3. Sewage pumps
4. Volute pumps
S. Sump pumps

VERTICAL TURBINE PUMPS

Vertical turbine pumps were originally developed for pumping water from wells and have been called
"deep-well pumps," "turbine-well pumps," and "borehole pumps." As their application to other fields
has increased, the name "vertical turbine pumps" has been generally adopted by the manufacturers. (This
is not too specific a designation because the term "turbine pump" has been applied in the past to any
pump employing a diffuser. There is now a tendency to designate pumps using diffusion vanes as
"diffuser pumps" to distinguish them from "volute pumps." As that designation becomes more universal,
applying the term "vertical turbine pumps" to the construction formerly called "turbine-well pumps"
will become more specific.)
The largest fields of application for the vertical turbine pump are pumping from wells for irrigation
and other agricultural purposes, for municipal water supply, and for industrial water supplies, processing,
circulating, refrigerating, and air conditioning. This type of pump has also been used for brine pumping,
mine dewatering, oil field repressuring, and other purposes.
These pumps have been made for capacities as low as 2.S or 3.5 m3/hr (10 or 15 gpm) and as high
as 6,000 m3/hr (25,000 gpm) or more, and for heads up to 300 m (1,000 ft.). Most applications naturally
involve the smaller capacities. The capacity of the pumps used for bored wells is naturally limited by
the physical size of the well as well as by the rate at which water can be drawn without lowering its
level to a point of insufficient pump submergence.
Vertical turbine pumps should be designed with a shaft that can be readily raised or lowered from
the top to permit proper adjustment of the position of the impeller in the bowl. An adequate thrust
bearing is also necessary to support the vertical shafting, the impeller, and the hydraulic thrust developed
when the pump is in service. As the driving mechanism must also have a thrust bearing to support its
vertical shaft, it is usually provided with one of adequate size to carry the pump parts as well. For these
two reasons, the hollow-shaft motor or gear is more commonly used for vertical turbine pump drive. In
addition, these pumps are sometimes made with their own thrust bearings to allow for belt drive or for
drive through a flexible coupling by a solid-shaft motor, gear, or turbine. Dual-driven pumps usually
employ an angle gear with a vertical motor mounted on its top.
The design of vertical pumps illustrates how a centrifugal pump can be specialized to meet a specific
application. Figure 14.17 illustrates a turbine design with closed impellers and enclosed line shafting;
Fig. 14.18 illustrates another turbine design with closed impellers and open line shafting.
The bowl assembly or section consists of the suction case (also called suction head or inlet vane),
the impeller or impellers, the discharge bowl, the intermediate bowl or bowls (if more than one stage
is involved), the discharge case, the various bearings, the shaft, and miscellaneous parts such as keys,
impeller locking devices, and the like. The column pipe assembly consists of the column pipe itself, the
shafting above the bowl assembly, the shaft bearings, and the cover pipe or bearing retainers. The pump
is suspended from the driving head, which consists of the discharge elbow (for above-ground discharge),
the motor or driver support, and either the stuffing box (in open-shaft construction) or the assembly for
providing tension on and the introduction of lubricant to the cover pipe. Below-ground discharge is
taken from a tee in the column pipe, and the driving head functions principally as a stand for the driver
and support for the column pipe.
Liquid in a vertical turbine pump is guided into the impeller by the suction case or head. This may
SI<><f f"[[l) "'lVE ---':~W~~~9tF~~Z\::t---- SHoFl' TUllE TO<SI()IO SEARING

FIIOONG FOLlCI'II'FR -------/-+Y S~ I NG SO<

'--_--TOP S>WT

~::::;lllllL4~:dlr-- OISOIAAG£ ~
lOP SHAFT TL&: F------SHoFl' CQAJNG
LIMO S>WT -------+-41 ~------ TOP ~ PIP(

f-ff-------ENCL05£O l IN[ SlW'T OfARIIIe;

11n-------- ~ ~WG

--fml--
If-- - - - - - COLUMN PIPE _CEO

SHAFT TI.&: S>tAf TL&: STASI.IZlR

iMf'{ll[ll SHAFT

11------- TOP 80M.. COfoOo£C1OA OfARING

t - - - - - -- SlJll. RING SPlOGER

""TAIN[R .....T£

9OWI.

"If------IMP£U.£R
RI1='------_1NG RING

~------9OWI. SEARING
IW'UI.£R BiISHING

~f_----- ~ING CAP


I--- \ - - - - -- - SUCTIOII >€AD OfAIIONG
~~wM~e--------SUCTIOII HEAD
1t------~<!CToa. PIPE

STRAINER

Fig. 14.17 Section of vertical turbine pump with closed impellers and enclosed line shafting (oil lubrication).

325
;-2\
1 '\

~ /
L

TOP SHIFT
,--..., \I.

PIU)QNG I~ \.D.
5
STUFfIl«i IIOX II£""IHG
t
1'i ~~ r-
PII[-U8IICATlNG PlI'£
~;::~ ~ II ~
-
~
01
I
~
SHAF T COUPI.ING
TOP~PIP[
! I HE SHAFT

I IlEAAING 8AAQ(£T
COl.~ PIP£ COUPI. ING ~ :~ SOiAFT SLEEVE
eEAA II«i II[TAlr«A
~ ". OPEN LIN£5>W'T II!EAAINC<

~I~ UMlI Plp£

I MPELLER SHAFT
AIlING CAP
,
, TOP

~ ~ TOP IIOWL

~~
~'rs !.
IMP£LLER
~ WE AIliNG RING
~ ~ BEAIIING
= IMPELLEA 8I/SHING

~ ~ IlEAAING CAP
SUCTION MUll 8£AIliHG

""'" SUCTION HEAl)


sucnON PIP£

a~

Fig. 14.18 Section of vertical turbine with closed impellers and open line shafting (water lubrication).

326
....... ----~...."""''''i H......\l--~:r-·-- --

~1..._--- -- """-

Fig. 14.19 Section of bowl of vertical turbine pump (closed impellers) for connection to enclosed shafting.

327
328 Special Designs: Vertical Pumps

Fig. 14.20 Section of bowl of vertical turbine pump (open impeller) for connection to open line shafting.
Special Designs: Vertical Pumps 329

Fig. 14.21 Vertical turbine double casing ("can") pump for condensate and hydrocarbon service.

be a tapered section (Fig. 14.19 and 14.20) for attachment of a conical strainer or suction pipe, or it
may be a bellmouth.
Semi open and enclosed impellers are both commonly used. For proper clearances in the various
stages, the semiopen impeller requires more care in assembly on the impeller shaft and more accurate
field adjustment of the vertical shaft position to obtain the best efficiency. Enclosed impellers are favored
over semiopen ones, moreover, because wear on the latter reduces capacity, which cannot be restored
unless new impellers are installed. Normal wear on enclosed impellers does not affect impeller vanes,
and worn clearances may be restored by replacing wearing rings. The thrust produced by semi open
impellers may be as much as 150 percent greater than that by enclosed impellers.
Various applications in power generation and hydrocarbon processing involve pumping from vessels
in which the liquid level is not high enough to provide the NPSH (net positive suction head) required
by a conventional horizontal pump. This difficulty is often aggravated by the need to develop quite a
high head to move the liquid into the discharge vessel. Typical applications are condensate pumps, heater
drain pumps, and pumps transferring "light" hydrocarbons (propane and lighter). Building a pit alongside
the suction vessel to provide additional submergence for a horizontal pump is an expensive solution and
is not always practical. An alternative approach, widely used in modem power plants, is to mount a
vertical wet-pit pump in a tank: (often called a "can") that is sunk into the floor (Fig. 14.21). The length
of the pump has to be such that sufficient NPSH will be available for the first-stage impeller design,
and the diameter and length of the tank must allow for proper flow through the annulus between the
pump and can and then around the tum into the bellmouth. When the pump length necessary to achieve
this poses a problem, the use of a double-suction first stage (Fig. 14.22) allows a shorter but larger
330 Special Designs: Vertical Pumps

1 .Inlet from
condenser
2. On-line
condensate
polishing plant
3. Discharge to
heaters and
feed pump

}<'ig. 14.23 Passout or re-entry type vertical double


casing ("can") pump for condensate service. Fig. 14.22 Section of vertical double casing ("can")
(Courtesy Thompsons, Kelly and Lewis Pty. Ltd.) pump with double-suction volute type first stage.

diameter can. The application of can pumps to difficult services has lead to a number of novel designs.
One notable example is the so-called "pass-out" condensate pump (Fig. 14.23), in which the flow leaves
the pump at an intermediate stage, passes through an on-line polishing plant, then reenters the pump for
the balance of the head addition. A second is the use of low-specific-speed hydraulically balanced tandem
impellers (Fig. 14.24) to develop high heads without imposing intolerable loads on the thrust bearing.

PROPELLER PUMPS

Originally the term "vertical propeller pump" was applied to vertical wet-pit diffuser or turbine pumps
with a propeller or axial-flow impellers, usually for installation in an open sump with a relatively short
setting (Fig. 14.25 and 14.26). Operating heads exceeding the capacity of a single-stage axial-flow
impeller might call for a pump of two or more stages or a single-stage pump with a lower specific speed
and a mixed-flow impeller. High enough operating heads might demand a pump with mixed-flow impellers
and two or more stages. For lack of a more suitable name, such high-head designs have usually been
classified as propeller pumps also.
Although vertical turbine pumps and vertical modified propeller pumps are basically the same mechani-
~ ,,,
rl I'
"-'" -,,'"
...
I I II
L
~ ~~
- r'
I I
I 1
I
...,1. _ -l.,
"1'- -r"
-If -
I
.J

-. (j "

r.
~ ~ l,
1;1= =
~t~~ ~ .. --.
._ t ."
--. -
! ~u..;

I .....
I
..........
f '"
./

~-===:r,;~~
, ;. !l
ll· cW I;u ,· , 11\:" , , ' -. i:4
I! ; II W
werl-· .1
I-
f------oISCloE
if . j .-
iii
IIII-
I l~.
&
~, I- ". :n -;r . '

ffi:-
" D
-[1:
-
~ ~
..~-

''II''
II

IIII'1'1
\l\.
\

\ \

~~ p~ m
Irl

~ lJ jllI
I

~~ ~~!
~r~'
; r..l~
I ~ 1""'"""
I ~J5:i:i ~~ III
I ~t :11
I,'
~
~! M "" .1
--, ~ ~v I
~ ~ 11 1
~ " ~ 1.,
I ... ~
)}: /

~ k~;/~

Fig. 14.24 Vertical double casing ("can") pump with low specific speed hydraulically balanced
tandem impellers.

331
332 Special Designs: Vertical Pumps

1ZZ:I:ZZl2D~-. P-p!hzzzz:zzz:z::o
I
i

Column
--- pipe

eov.r
- pip.

,BearlnQ
/Bearinll
Impeller -----,t-~-"'Il
,haft /80wl
...-seal
Impelle,-- __ _ "

Fig. 14.25 Section of vertical propeller pump with Fig. 14.26 Vertical propeller pump with below-
below-ground discharge. ground discharge. (Courtesy Peerless Pump Co.)

cally and even could be of the same specific speed hydraulically, a basic turbine pump design is one
that is suitable for a large number of stages, whereas a modified propeller pump is a mechanical design
basically intended for a maximum of two or three stages.
Most wet-pit drainage, low-head irrigation, and storm-water installations employ conventional propel-
ler or modified propeller pumps. These pumps have also been used for condenser circulating service,
but a specialized design dominates this field. As large power plants are usually located in heavily
populated areas, they frequently have to use badly contaminated water (both fresh and salt) as a cooling
medium. Such water quickly shortens the life of fabricated steel. Cast iron, bronze, or an even more
Special Designs: Vertical Pumps 333

corrosion resistant cast metal must therefore be used for the column pipe assembly. This requirement
means a very heavy pump if large capacities are involved. To avoid the necessity of lifting this large
mass for maintenance of the rotating parts, some designs (one of which is illustrated in Fig. 14.27) are
built so that the impeller, diffuser, and shaft assembly can be removed from the top without disturbing
the column pipe assembly. These designs are commonly designated as "pull-out" designs.
Like vertical-turbine pumps, propeller and modified propeller pumps are made with both open and
enclosed line shafting. Open line shafting is used only when it is certain the pumped liquid is free of
solids. In condenser circulating applications, the pumped liquid often contains unexpected solids, so
many power plant designers specify enclosed line shafting to be sure of satisfactory bearing life. When

Fig. 14.27 Section of modified vertical propeller pump with removable bowl and shafting assembly.
334 Special Designs: Vertical Pumps

the line shafting is enclosed, lubrication of the bearings is generally with water or a compatible liquid
if the pump is handling other than water. (Environmental concerns have all but eliminated drip feed oil
or grease lubrication in vertical wet pit pumps handling water.) The lubricating liquid comes from either
a separate source or the pump discharge. In the latter case, it is usually necessary to employ a small
booster pump to raise the pressure of the lubricant above that adjacent to the bottom bearing, which is
close to impeller discharge pressure (Fig. 14.27; see product lubricated bearings in Chap. 11 for details
on lubricant filtration and monitoring).
Propeller pumps have open propellers. Modified propeller pumps with mixed-flow impellers are made
with both open and closed impellers.

SEWAGE PUMPS

Except for some large vertical propeller pumps that handle dilute sewage (basically storm water contami-
nated by domestic sewage), vertical wet-pit sewage pumps have a bottom-suction volute design with
impellers capable of handling solids and stringy materials with minimum clogging.
Two configurations are used: suspended and submersible, with most modem installations being
the latter.
Suspended pumps (Fig. 14.28) usually employ an impeller without back wearing rings, a seal of some
form above the impeller to limit the amount of leakage back to the suction pit, and a guide bearing or
bearings separate from and above the shaft seal. Other lower-cost designs incorporate the lower guide
bearing in the stuffing box, which means the seal then has to prevent pumped liquid from entering the
bearing. Experience has shown this is not an effective design; it almost invariably suffers rapid guide
bearing wear. Guide bearings are lubricated with either oil or grease, as in Fig. 14.28, or clear water
(see product-lubricated bearings in Chap. 11). With either approach, the lubrication is not as good as
that in a conventional sleeve bearing, and therefore the bearings will wear relatively faster. Suspended
wet-pit sewage pumps should only be used for services requiring operation for a very limited portion
of the day.
Submersible sewage pumps (Fig. 14.29) are close-coupled pumps with the motor either dry or liquid
filled. In either case, the motor is isolated from the pumped liquid by some form of axial face seal, and
so the pump bearings are always well lubricated. It is this feature that affords greater reliability and
longer service life than the suspended design, and it has led to the almost exclusive use of submersible
pumps for sewage service, where size allows. Liquid end design follows usual nonclogging pump practice
for the impeller and casing. Some incorporate a form of grinding or shredding device upstream of the
impeller to allow the pump to better handle the tramp material found in sewage. Most installations have
the pump on some form of guide rail (Fig. 14.30) to allow removal and reinstallation without having to
drain the suction pit (see Chap. 24 for details on the construction of submersible motors).

VOLUTE PUMPS

Wet-pit volute pumps are used in a wide variety of applications beyond the sewage and sump services
dealt with so far in this chapter.
Single-suction cantilever pumps (Fig. 14.31) are designed so the rotor has its first critical speed at
least 25 percent above the maximum operating speed and does not rely on support from a submerged
bearing. With these features, the design is ideally suited to corrosive and erosive applications where
neither a suspended nor a submersible pump is suitable. Typical applications range from paint lines to
severe sump service to steel-mill primary scale pit. Liquid end construction is either chemical or slurry
pump, depending on the nature of the pumped liquid. If the mounting flange is sealed to the top of the
Special Designs: Vertical Pumps 335

Fig. 14.28 Section of vertical wet-pit sewage Fig. 14.29 Section of submersible sewage pump.
(non-clogging) pump. (Courtesy ITT Flyght AB)

suction vessel, and an effective vapor seal is made where the shaft passes through the mounting flange,
for example, grease injected double labyrinth; this configuration is a practical "sealless" pump.
Large single-suction suspended pumps (Fig. 14.32) have been used for low head condensate service,
and, with the appropriate liquid end materials, abrasive service in mineral processing and steel mills.
The difficulty with the latter application is keeping the pumped liquid out of the submerged bearings;
336 Special Designs: Vertical Pumps

Fig. 14.30 Guide rail system for submersible sewage pump. The pump slides down the guide and automatically
hooks up to the discharge connection. Normally two guide rails are used to ensure proper hook up.
(Courtesy ITT Flyght AB)

unless the lubricating water pressure is greater than that adjacent to the bearing and the seal clearance
close enough to produce a velocity of at least 2.1 mls (7 ft.), the pumped liquid will get into the bearings
and significantly shorten their life.
Double-suction suspended pumps (Fig. 14.33) are used extensively in utility, industrial, refinery, and
mineral processing services where the NPSHA is low, the liquid contains a low concentration of solids,
or the purchaser just wants a single-stage pump because it is simpler. Being double suction, the NPSHR
is 63 percent of that for an equivalent single-suction impeller (same capacity, speed, and suction specific
speed). With both the impeller guide bearings exposed to impeller suction pressure, keeping the pumped
liquid out of the bearings is easier than in a single-suction design. Because the design allows much
lower specific speeds than those used in vertical-turbine pumps (at the expense of pump diameter), it is
possible to develop in a single stage, heads that would require six or seven stages of vertical-turbine
pump. Lineshafts are either open or closed, depending on the cleanliness of the pumped liquid. As these
pumps get larger it becomes necessary to resort to more sophisticated lineshaft couplings than the simple
threaded coupling (see Chap. 12 for further details).
When the head necessary exceeds that attainable with a single-stage volute pump, a common solution
is to make the pump multistage by adding the required number of matching vertical-turbine pump stages
(Fig. 14.22). The vertical-turbine pump stages have to be a special design because the shaft is larger
than normal to accommodate the high first-stage power. This arrangement retains the lower NPSHR
Special Designs: Vertical Pumps 337

$ $
~,.
:= 1=

~
\Vr ~

-
'-- -

, I "",
~
-
i==
i" r?;.
~

]1
.....
III [J

+
dB! 1+

Fig. 14.31 Section of vertical cantilever pump for abrasive service.


(Courtesy Lawrence Pumps, Inc.)

inherent in the double-suction first stage, and reduces the total number of stages needed. It is widely
used for condensate service and oil pipeline boosting.

SUMP PUMPS

The term "sump pump" ordinarily conveys the idea of a vertical wet-pit pump that is suspended from
a floor plate or sump cover or supported by a foot on the bottom of a well, that is motor-driven and
CDOLlJIo .,.".... .1~ ...J • ....-c
~
"'''W'CD nu....... ~,nwl.. ...... J...
~
oc
/ .
~ o
b-0

It
,-
~
"''''lIIl-'''lIld~:~I''''iII
S . ~d ~~U'tJ1D~

Fig. 14.32 Section of single suction vertical suspended mixed flow volute pump for solids handling applications.
(Ingersoll-Dresser Pumps model QMN; patents pending)
Special Designs: Vertical Pumps 339

Fig. 14.33 Section of double suction vertical suspended volute pump.

automatically controlled by a float switch, and that is used to remove drains collected in a sump. The
term does not indicate a specific construction, for both diffuser and volute designs are used; these may
be single-stage or multistage and have open or closed impellers of a wide range of specific speeds.
For very small capacities serviced by fractional hp motors, "cellar drainers" can be obtained. These
are small and usually single-stage volute pumps with single-suction impellers (either top or bottom
suction) supported by a foot on the casing; the motor is supported well above the impeller by some
form of a column enclosing the shaft. These drainers are made as complete units, including float, float
switch, motor, and strainers (Fig. 14.34).
340 Special Designs: Vertical Pumps

Fig. 14.34 Typical cellar-drainer


sump pump. Fig. 14.35 Typical duplex sump pump.
(Courtesy Sta-Rite Products) (Courtesy Economy Photo)

Sump pumps of larger capacity may be vertical propeller or turbine pumps (single stage or multistage)
or vertical wet-pit sewage or volute pumps. If solids or other waste materials may be washed into the
sump, the vertical wet-pit sewage pump with a nonclogging impeller is preferred. The larger sump pumps
are usually standardized but obtainable in any length, with covers of various sizes (on which a float
switch may be mounted), and the like. Duplex units, that is, two pumps on a common sump cover
Special Designs: Vertical Pumps 341

(sometimes with a manhole for access to the sump) are often used (Fig. 14.35). Such units may operate
their pumps in a fixed order, or a mechanical or electrical alternator may be used to equalize their operation.
Most sump pumps are intended to run only occasionally, and therefore employ construction that will
yield several years "life" when operated intermittently. This low-cost construction will not realize an
acceptable service life (in hours) if the pump is run frequently or continuously. When the service involves
pumping more than a few minutes a day, and particularly when the liquid contains abrasive solids, the
more expensive vertical cantilever construction (Fig. 14.31) will prove less expensive to own. Vertical
suspended construction (Fig. 14.28), with clear liquid lubrication to the guide bearings, is le&s expensive
than vertical cantilever but does require the provision of clear liquid and will not realize the same service
life in severe applications.

APPLICATION OF VERTICAL WET-PIT PUMPS

Like all pumps, the vertical wet-pit pump has advantages and disadvantages, the' former mostly hydraulic
and the latter primarily mechanical. If the impeller (first-stage impeller in multistage pumps) is submerged,
there is no priming problem, and the pump can be automatically controlled without fear of its ever
running dry. Moreover, the available NPSH is greater (except in closed tanks) and often permits a higher
rotative speed for the same service conditions. There are two mechanical advantages. First, with the
appropriate arrangement, it is possible to eliminate a liquid shaft seal. Second, the motor or driver can
be located at any desired height above flood level. The mechanical disadvantages are the following: (1)
possibility of freezing when idle, (2) possibility of damage by floating objects if unit is installed in an
open ditch or similar installation, (3) inconvenience of lifting out and dismantling for inspection and
repairs, no matter how small, and (4) the relatively short life of the pump bearings unless the water and
bearing design are ideal. The vertical wet-pit pump is the best pump available for some applications,
not ideal but the most economical for other installations, a poor choice for some, and the least desirable
for still others.

TYPICAL ARRANGEMENTS OF VERTICAL PUMPS

A pump is only part of a pumping system. The hydraulic design of the system external to the pump
will affect the overall economy of the installation and can easily have an adverse effect upon the
performance of the pump itself. Vertical pumps are particularly susceptible because the small floor space
occupied by each unit offers the temptation to reduce the size of the station by placing the units closer
together. If the size is reduced, the suction arrangement may not permit the proper flow of water to the
pump suction intake. This difficulty is compounded by the pump's hydraulic design, with those of higher
specific speed being increasingly sensitive to irregularities in their inlet flow. As many factors are
involved in the design of a suction well and the location of a bellmouth and no simple rules or relations
can be reliably applied, none is included in this discussion. The physical size of the pumps (whether
propeller or volute) rarely affects the design of the suction well, the location of the bellmouth, or the
spacing of the units. These are usually controlled by factors governing the proper flow of the water to
the bellmouth.
Figure 14.36 illustrates an ideal arrangement for a multiple-unit station with dry-pit pumps. It provides
an unrestricted flow on the suction side to all the units. Stations using this arrangement for a group of
vertical volute pumps often have the suction bellmouths and elbows formed right in the concrete
substructure. If dry-pit pumps are installed with vertical bellmouths, adequate clearance must be provided
at the back wall and between the units (Fig. 14.37). This arrangement illustrates a common situation in
342 Special Designs: Vertical Pumps

ELEVATION
Fig. 14.36 Multiple-unit station with vertical dry-pit volute pumps alongside each other in wide suction bay.

Fig. 14.37 Multiple-unit station with vertical wet-pit volute pumps at end of conduit.
Special Designs: Vertical Pumps 343

TOP VIEW

~~. ;. : I~_ ~! . p".~:.-r.

f.~~'
~{.',-
:1".'-=
[(!
~,~\
~}
~;~
~~. £- - .- - -
---
1:~~:

J&~~.~~~(,.,i:.;;:.:; :.: ~} :";i::/:;,.:; :;L;::::';:\'i;:: ~-;-;'~;:;~,,·~.~;;:~~;::t~~_::. ·..;tiXf:;';i /;~:~ :'.-: ;;;;1 , ;',;A~:.~{{;,;~ ;{;.,~:,~;~ '~
. ELEVATION

Fig. 14.38 Multiple-unit station of vertical propeller pumps with suction flow from one end of well.

Fig. 14.39 Pump installation with good intake design.


D = suction bel/mouth diameter.
344 Special Designs: Vertical Pumps

-$- ,
PARTITION WALL

B=2iO(MIN)

f MINIMUM
to RECOMMENDED
Fig. 14.40 Recommended channel and pit design.
KEY:
A = Minimum submergence above impeller cen-
terline, approximately 1.5 to 2.0D depend-
ing on pump cavitation characteristics;
B = minimum width of sump or pit;
C = minimum depth of sump or pit;
D = suction bell diameter (normally same as
bowl diameter).
Cross-sectional area of sump (B x C) shall not be less than ten times the suction bell area (1CD 2/4).
Fig. 14.41 Vertical wet-pit propeller pump with gate valve and flap valve.

Fig. 14.42 Vertical wet-pit propeller pump with siphon discharge.

34S
346 Special Designs: Vertical Pumps

which the suction is located at the end of a conduit the width of which is less than the length of the
suction well. Without a flared section with division walls to guide the distribution of the incoming water
to the various units, the flow would be badly disturbed and the operation of the pumps adversely affected.
A propeller-pump arrangement that is often troublesome (vertical volute-pump arrangements with
suction bellmouths like those in Fig. 14.37 have the same problem) is shown in Fig. 14.38. Unless the
width of the suction well provides sufficient area and unless the locations of the bellmouths permit good
flow, the demand of the units first in line will disturb the flow in more removed units. Very often
installations of this general arrangement require extensive baffling to correct the distribution. Some
stations are made with walls that form individual wells for each pump, a channel to supply these wells
running lengthwise of the station.
Various recommendations have been developed over the years for the dimensioning of intake channels
and approaches. If feasible, an intake like that illustrated in Fig. 14.39 will give excellent results. The
dimensions for the channel width and spacing are given in terms of the suction bellmouth diameter.
Another example of good channel and pit design for vertical turbine pumps is given in Fig. 14.40, which
also indicates recommended clearances between the suction bellmouth and the bottom of the pit and
between the pump, the pit back wall, and the partition walls.
If long discharge lines are involved, valves are required in the piping. Normally, both a gate valve
and a check valve are used (Fig. 14.41). The check valve acts to prevent reverse flow, whereas the gate
valve functions when the unit is shut down for an extended period. In some installations, the gate valve
is omitted, and stop planks or a sluice gate are used. A cone valve that acts both as a check and a stop
valve appears in other installations. The high cost of this valve, however, usually restricts its use to
installations requiring a flow that is started and stopped gradually to prevent water hammer. A few
installations with long discharge lines for single pumps have no valve other than a flap valve at the
discharge end. If the unit is stopped, the water in the discharge line flows back through the pump until
the pipe is emptied.
If the design of an installation or the failure of a check or flap valve to close permits a reverse flow
of water through a pump, the pump acts as a water turbine. The torque developed by the pump as a
turbine will cause reverse rotation in freely rotating drivers like electric motors. Usually it is not sufficient
to cause reverse rotation in internal combustion engines. In motors, the reverse speed that will be attained
will depend both on the net head and the runaway speed of the pump acting as a water turbine. The net
head is then less than the static head because of friction losses. The runaway speed is dependent on the
specific speed of the pump. Higher specific speeds have higher runaway speeds (measured as a percentage
of normal speeds). The reverse speed obtainable in an actual installation is usually below the safe
operating speed of its component parts, and it is not necessary to use a special design.
The use of a siphon discharge eliminates the necessity for valves in the discharge line (Fig. 14.42).
The high point of the siphon must be above high-water level on the discharge to break the siphon and
prevent backflow of the water when the pump is shut down. When a pump operating on a siphon
discharge is started, the usual procedure is to exhaust air from the system by a priming device until the
pump is primed. The pump may then be started to help fill the siphon. The connection to the high point
of the siphon is also provided with a valved opening so that air can be admitted and the siphon broken
when it is desired to stop the unit. It is possible to control the admission of air automatically so that the
valve functions if the unit stops for any reason.
Although siphons with short legs are relatively simple and troublefree both in design and operation,
more care must be taken if they have long legs. Some siphons operate successfully, with legs exceeding
7.6 m (25 ft), but these are primarily limited to circulating systems in power plant installations. The use
of a siphon discharge is desirable in drainage installations for pumping over a levee because it provides
a lower head than would be obtained if the water were discharged at the top of the levee.
15
Special Designs: Self-Priming Pumps
.~--~--- -----

The standard centrifugal pump cannot handle air or vapors. Unless it is located beneath its source of
supply, some means must be found of filling both the pump and its suction piping with liquid, that is,
to prime it. A demand naturally developed, therefore, for a centrifugal pump able to handle appreciable
quantities of air and to reprime itself automatically when located above the water supply. This requirement
is especially important in the construction field because pumps may be used to dewater areas into which
seepage is slower than the pump can handle. A standard pump will operate until it uncovers the entrance
to the suction pipe, get air-bound, and then be unable to reprime itself even after sufficient seepage has
accumulated to prevent further air infiltration.
A true "self-priming pump" is one that will clear its passages of air if it becomes air-bound and
resume delivery of the pumped liquid without outside attention. Therefore, its basic requirement is that
the pumped liquid entrain air (in the form of bubbles) so that the air will be removed from its suction
side. The air must be allowed to separate from the liquid once the mixture of the two has been discharged
by the impeller, and the separated air must be allowed to escape or to be swept out through the pump
discharge. A self-priming pump therefore requires an air-separator, which is a large stilling chamber or
reservoir provided on its discharge side to effect this separation.
Several ways exist of making a centrifugal pump self-priming, the most important being the following:

1. Recirculation from discharge back into suction


2. Recirculation within the discharge and impeller itself.

These two basic methods have many variations; only one example of each will be discussed here.

RECIRCULATION TO SUCTION

A pump made self-priming by this method contains a liquid reservoir either attached to or built in the
casing. The first time the pump is to be started, this reservoir is filled. A recirculating port is provided
in the reservoir, communicating with the suction side of the impeller. As the pump is started, the impeller

347

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
348 Special Designs: Self-Priming Pumps

handles whatever liquid comes to it through the recirculating port plus a certain amount of air from the
suction line_ This mixture of air and liquid is discharged into the water reservoir where the two elements
separate, the air passing out of the pump discharge and the liquid returning to the suction of the impeller
through the recirculating port. This operation continues until all the air has been exhausted from the
suction line. The vacuum thus produced draws the liquid from the suction supply right up to the impeller.
It is essential that the reservoir remain filled with liquid when the pump is brought to a stop. This
is accomplished by incorporating either a valve or some form of trap between the suction line and
the impeller.
A typical self-priming pump operating on this principle is illustrated in Fig. 15.1. The pump housing,
B, consists of a conventional volute and an inlet passage. The inlet has a priming passage with a priming
valve, C, attached to it. This priming valve is a cylindrical rubber tube. The impeller, A, is of a conventional
semi-open design. During priming, the pump body is filled with water. This water is drawn into the
pump housing through the priming valve and discharged from the volute back into the pump body. The
suction created by the impeller draws air from the inlet passage at the same time that it is drawing water
through the priming valve. The air is mixed with the water and discharged into the pump body along
with the water. In the pump body, the air bubbles separate from the water, rise to the surface, and pass
out through the pump discharge while the priming valve picks up water that is relatively free of air.
After the air has been exhausted from the suction piping or hose and water is drawn into the pump,
sufficient pressure difference exists between the pump body and the inlet passage to cause the rubber
priming valve to collapse. The recirculation thereby being stopped, all the water that goes through the

Fig. 15.1 Self-priming pump with valved recirculation to suction.


(Courtesy Homelite Corp.)
Special Designs: Self-Priming Pumps 349

impeller is discharged from the pump body_ A ball check valve is built into the suction line to maintain
the vacuum in the line between operations.

RECIRCULATION AT DISCHARGE

This form of priming is called "volute priming" or "diffuser priming," depending on the design of the
discharge casing. It may be distinguished from the preceding method by the fact that the priming liquid
is not returned to the suction of the pump but mixes with the air either within the impeller itself or at
its periphery. Its principal advantage, therefore, is that it eliminates the complexity of internal valve mecha-
nisms.
A typical "volute priming" self-priming pump is illustrated in Fig. 15.2. An open impeller, A, rotates
within a volute casing, B, discharging the pumped liquid through passage C into the sealing reservoir,
D. When the pump starts, the trapped liquid carries entrained air bubbles from the suction to the discharge
chamber. There, the air separates from the liquid and escapes into the discharge chamber, E. The liquid
in the reservoir returns to the impeller through the recirculation port, F, reenters the impeller, and mixing
once more with air bubbles is discharged through C. This operation is repeated continuously until all
the air has been expelled through E. Once the pump is primed, the uniform pressure distribution established
around the impeller prevents further recirculation, and the liquid is discharged into reservoir D both at
C and at F.
Some sizes of this pump incorporate an externally adjustable recirculation port (Fig. 15.3). The original
clearance between the impeller and the casing can be restored, when these parts become worn, by the
following steps:

Fig. 15.2 Self-priming pump with Fig. 15.3 Adjustable recirculating port of volute
volute recirculation. recirculation pump.
350 Special Designs: Self-Priming Pumps

Fig. 15.4 Self-priming pump with separate motor drive.

Fig. 15.5 Close-coupled self-priming pump.

I. Remove cover nut


2. Turn adjusting stem until recirculating port touches impeller
3. Back off adjusting stem I Y2 turns
4. Replace cover nut.
Special Designs: Self-Priming Pumps 351

Fig. 15.6 Portable engine driven self-priming pump.

This adjustment appreciably extends the usable life of the pump casing. An added advantage of this design
is the ability to use impellers of different diameters in the same casing without losing priming capabilities.
Such pumps are built with as many combinations of drives as ordinary nonpriming pumps. They are
commonly available either with separate drive (Fig. 15.4), close coupled (Fig. 15.5), or engine drive
(Fig. 15.6).

REGENERATIVE PUMPS

One of the "special effect" kinetic pumps, regenerative pumps are self priming provided the casing
retains sufficient liquid to effect a seal between the casing and the impeller. These designs have been
used alone as self priming pumps when their hydraulic characteristics suited the application, and as a
priming impeller acting in series with a conventional centrifugal impeller for other applications. Chapter
16 includes a detailed discussion of regenerative pumps, their characteristics, and typical applications.
16
Special Effect Pumps

Within the classification "kinetic pumps" (see Chap. 1), there is a group termed "special effect" pumps.
These are pumps in which the means of energy addition is still kinetic, the addition of velocity, but that
employ effects other than that of the classical centrifugal pump (see Chap. 2) to do so. At present there
are six distinct types of special effect pumps: regenerative, partial emission, induced vortex, viscous
drag, impact, and reversible.

Regenerative Pumps
The name "regenerative pump" describes a unit with a multi blade impeller that develops head or
pressure by a principle considerably different from that of a centrifugal pump. These pumps have had
a number of other names given to them, for example, "turbulence pumps," "peripheral pumps," "vortex
pumps," and "turbine pumps." The term "regenerative," however, best describes the actual pumping
principle involved.
Principle of operation. Figure 16.1 shows a cross section of a regenerative pump; Fig. 16.2 is an
"exploded" photograph of the same unit. The impeller has a multiplicity of radial vanes cut into its rim
that rotate within an annular chamber. The liquid enters the pump casing and flows to both sides of the
impeller either through a cored passage in the casing or through ports or openings provided for this
purpose in the web of the impeller. This design, in effect, makes the pump a double-suction unit and
balances the axial hydraulic thrust.
At one point of the periphery, there is a separating wall or "stripper" that the impeller passes, in its
rotation, with a very narrow clearance. Passages are provided from the suction into the annular chamber
surrounding the impeller rim, immediately beyond this dividing wall. The liquid is picked up in the
spaces between the impeller vanes and then thrown out again into the annular chamber because of the
kinetic energy it gains from the centrifugal force action in the impeller. The kinetic energy is transformed
into pressure energy as the liquid slows down in the casing.
The manner in which a regenerative pump develops head is illustrated in Fig. 16.3. The liquid enters
the casing and flows to both sides of the impeller, twin passages leading the liquid to the impeller blades
(Fig. 16.3[a]). Each casing is equipped with a dividing wall (or stripper) through which the impeller

352

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
Special Effect Pumps 353

IMPEL L ER KE v RING P",CK ING WAT ER SL INGER

Fig. 16.1 Section of a regenerative pump.

passes with close clearance, as shown in Fig. 16.3(b) (at A). Just beyond this wall in the direction of
rotation, the twin suction passages, which have passed around the sealing wall, come into the impeller
chamber (at B). The impeller blade engages the liquid as it comes out of the suction passage, and
centrifugal force throws the liquid out to the periphery of the impeller (Fig. 16.3[c]). The liquid leaving
the impeller blade has had velocity energy added and leaves the impeller as shown in the vector diagram
(Fig. 16.3[d]). The casing passage causes a gradual reduction of velocity with the accompanying increase
in potential energy (pressure). The pump has thus generated head. The shape of the space between the
impeller vanes imparts a rotating motion to the liquid as it leaves the impeller cavities (Fig. 16.3[e]).
As the rotating motion continues in the annular chamber, the liquid is guided back into the "root" of
the cavities, proceeding circumferentially around the chamber (Fig. 16.3[f]). The cycle is then repeated,
adding energy to the liquid every time it leaves and reenters the impeller. The number of times the
process repeats itself may vary from 2 to 50 depending on the head to be developed by the pump. The
more times the liquid reenters and is discharged from the impeller, the higher the head. When the liquid
finally reaches the discharge side of the separating wall, it flows into the discharge passage and out the
discharge nozzle.
354 Special Effect Pumps

Fig. 16.2 "Exploded" view of regenerative pump.

c d

Fig. 16.3 Development of pressure in regenerative pumps.


Special Effect Pumps 355

Performance characteristics. The perfonnance of a regenerative pump resembles that of a high


specific-speed centrifugal pump in that the head rises very rapidly with a reduction in capacity, as does
power consumption. Typical perfonnance characteristics are illustrated in Fig. 16.4. Both pumps follow
the same laws of speed variation: (1) the capacity varies directly with the speed, (2) the head varies as
the square of the speed, and (3) the power consumption varies as the cube of the speed.
The efficiency of regenerative pumps is considerably lower than that of centrifugal pumps. In the
past, this disadvantage was offset by their ability to develop much higher heads at low flows, applications
that would otherwise have required multistage centrifugal pumps. Typical regenerative applications were
for flows less than 25 m3/hr (100 gpm), with heads up to 150 to 180 m (500 or 600 ft), with a few
special designs capable of capacities to 45 m3/hr (200 gpm) and heads of 360 to 460 m (1200 to 1500
ft). Today, it is more usual to meet such applications with small, mass-produced vertical multistage
centrifugal pumps or single-stage partial-emission (Barske) pumps. For most applications, the fonner
offer lower overall cost, being more efficient than regenerative pumps and generally having a longer
period between the need to renew clearances. The latter have the virtue of large running clearances and
limited sensitivity to increases in the clearances.
Because they are designed only for low flows, regenerative pumps have relatively low NPSHR, and
therefore NPSH is generally not a concern in their application. For the same reason, however, entrained

240

220
\
200 \~
180
\~
t-
IIJ
IIJ
160 ~1b ~
~
La..
140

\
~
ci
« 120

\
IIJ
:t:
.J
« 100
t-
o
t-
80 \ 5

60
\ 4
I'
40
...
~Jf~
·2

~
o 10 20
--
30 40
CAPACITY, IN GPM

Fig. 16.4 Typical performance characteristics of a regenerative pump.


Shut-off head of 102 m (355 ft), 1,730 rpm, and 4.6 m (15 ft) suction lift·
356 Special Effect Pumps

Fig. 16.5 Arrangement of trap in the suction of a regenerative pump.

gas is, because relatively small volumes significantly reduce the pumps' capacity. Given this, many
rating curves are published for various suction lifts, based on lifting water from a reservoir exposed to
the atmosphere.
The regenerative pump can handle viscous liquids up to about 48 cSt (250 ssu); when viscosities
exceed this value, performance falls off very rapidly, and the pump ceases to be practical.
Self-priming features. As long as sufficient liquid remains within the pump to seal the clearance
between the impeller and the separating wall in the casing, the impeller cavities will take up all fluid
present, whether simple liquid or a mixture of liquid and vapor or air. The regenerative pump will
therefore always prime itself-by evacuating the air out of its suction line-provided the installation is
arranged to trap sufficient liquid on shut-down. This condition is usually met by building in a trap in
the pump suction (Fig. 16.5). In addition, an enlargement is provided in the discharge of the pump to
slow down the velocity of the delivered liquid and to permit its separation from any vapor or air.
General application. Because the satisfactory operation of a regenerative pump depends on the
close clearance between the impeller and the separating wall, or stripper, this pump is not too suitable
for handling corrosive liquids or liquids containing abrasive foreign particles. The first may attack the
metal at the running clearance joint to a point that the pump will lose a major part of its capacity through
internal recirculation. Solid particles of the products of corrosion may also build up on the pump surfaces
and cause wear at the running clearances just as grit or other abrasive particles in the liquid would. The
regenerative pump should ordinarily be used to handle clean, clear liquids. To prevent the entry of
foreign material, a 40-mesh strainer is desirable.
Special Effect Pumps 357

The clearance at the dividing wall, moreover, has a greater effect on the effective capacity of this
pump than the clearances at the wearing ring of a centrifugal pump. Regenerative pumps thus require
more frequent maintenance and renewal of internal clearances. An ample margin over the maximum
requirement for pump capacity is recommended. Depending on the pump's construction, clearances can
be renewed either by replacement of parts (side-plates, casing heads, and the like) or by changing the
thickness of the gaskets that determine the relative location of the casing walls and the impeller itself.
Because of the steepness of its head-capacity curve, a regenerative pump operated at excessively low
capacities may develop excessive pressures. Consequently, a relief valve is usually arranged in the
discharge line to bypass some of the capacity back to the suction line whenever the discharge pressure
reaches a predetermined maximum.

Partial-Emission Pumps
Although not in the strictest sense special effect pumps, partial-emission pumps are sufficiently
different in their design and characteristics to warrant distinction from conventional centrifugal pumps.
The partial-emission pump was first described by Barske [16.1], and is therefore frequently so called.
Because of the shape of the impeller, partial emission pumps are sometimes also called "paddle
wheel" pumps.
Principle of operation. Figure 16.6 shows the impeller and casing insert of one design of partial-
emission pump. Energy is added to the liquid in the same manner as in a conventional centrifugal pump.
Where the operation differs is in the amount of energy added, and the means of controlling the pump's
best efficiency capacity. Because the impeller vane angle is very high, close to or equal to 90 deg, the
head developed by partial emission pumps is on the order of 50 percent higher than a low-specific-speed
centrifugal pump with the same impeller diameter and running at the same speed. When run at high
speeds (up to 25,000 rpm), heads up to 2,100 m (7,000 ft) in a single stage have been achieved in partial-
emission pumps intended for process and industrial applications. The term "partial emission" derives

Fig. 16.6 Impeller and casing insert of one design of partial emission pump.
358 Special Effect Pumps

1~ :
Total Head
250
:;
J: 200 m
J: e
40
50
Q.
j 30 J:40 8 2.5
I
Iii 2.0
eU:I
120
::::iE
130 I
Q.
011
~20 4~
J: 1.5 m
Q.
I ~. z 1.0 z

.""
10 ~ .~·Tt
10 2 0.5

0 "" 10 20 30 40
Flow-GPM
50 eo 70 80 90

i i
o 5 10 15 20
M3/Hr.

Fig. 16.7 Typical perfonnance characteristics of a partial-emission pump.

from the nature of the radial flow through the impeller. Unlike higher specific speed impellers, in which
the radial flow is fairly uniform over the periphery of the impeller, partial emission impellers have radial
flow over only part of each passage at the impeller periphery. This is a result of having the ratio of
impeller normal flow area to casing throat area much larger than in conventional centrifugal pumps.
Experience has shown that for small, low specific speed pumps, these designs achieve higher efficiency.
With such an area ratio, the pump's capacity is limited by separation and consequent blockage in the
casing throat. As such, a given pump has a range of best efficiency capacities, each corresponding to a
particular casing throat area. For manufacturing convenience, the casing throat is usually produced by
drilling. Partial emission pumps have open or semiopen (Fig. 16.6) impellers, with relatively large axial
clearances between the casing and impeller, which means the pump performance is not greatly affected
by wear.
PerjormtJ"ce characteristics. With radial or nearly radial impeller vanes, the head characteristic is
effectively flat out to best efficiency capacity, dropping steeply beyond where separation in the discharge
occurs, whereas the power rises continually with increasing capacity. Impellers of "high solidity" (more
vanes of angle less than 90 deg) or other design refinements have constantly rising head characteristics
(Fig. 16.7) and generate less noise, an important factor in high-speed designs. Over a speed range of
2: 1, partial emission pumps follow the affinity laws for variations in speed.
AppUctJtio". For designs at the low end of conventional centrifugal pump specific speeds, partial-
emission pumps achieve equal or higher efficiency, and are therefore a good choice for applications
requiring high head at low capacities. Direct coupled pumps running at 3,600 rpm (Fig. 16.8) are available
for flows to 90 m3/hr (400 gpm) and heads up to 210 m (700 ft). High-speed designs are a viable
alternative to multistage centrifugal pumps in services where the pumped liquid sa, viscosity, and
lubricity are low, factors that reduce the reliability of multistage designs dependent on the Lomakin
effect for rotor stability (see Chap. 7). Motor-gear driven high-speed pumps (Fig. 16.9) are in regular
Special Effect Pumps 359

Fig. 16.8 Horizontal 3,600 RPM partial-emission pump.

Fig. 16.9 High-speed partial-emission pump.


(Courtesy of Sunstrand Fluid Handling Corporation)
360 Special Effect Pumps

use for flows to 90 m3/hr (400 gpm), at heads up to 1,900 m (6,200 ft). In most applications, high-speed
pumps are equipped with inducers to lower the NPSH that must be provided by the system.

Induced-Vortex Pumps

Known also as "recessed impeller," "free flow," or "torque flow" pumps, induced-vortex pumps
employ a means of head generation that makes them particularly useful for pumping liquids laden with
solids or moderate percentages of air or entrained gas.

Principle of operation. Rather than having the pumped liquid pass through the impeller as it does
in a conventional centrifugal pump, an induced-vortex pump has its impeller recessed back out of the
flow path, the degree of recess varying from 50 to 100 percent (Fig. 16.10), depending on the intended
service. Designs with less than 100 percent impeller recess sometimes have axial rotor adjustment to
allow the pump's performance to be "tuned" to a particular service. The recessed impeller imparts head
to the pumped liquid by momentum exchange. With the impeller out of the flow path, the pump's ability
to pass solids is limited by the size of its casing throat.

Performance characteristics. Over the range of specific speeds used for induced-vortex pumps,
typically 1,500 to 2,800, their head and power characteristics are similar to equivalent centrifugal pumps
(Fig. 16.11). Because the means of head addition is indirect, the head produced for a given size impeller
is lower and the power higher, resulting in lower efficiency, usually between 35 and 55 percent. NPSHR
tends to be lower than a centrifugal pump of the same capacity, because the rate of energy addition at
the impeller inlet is lower.

Fig. 16.10 Section of induced-vortex pump.


(Courtesy Met-Pro Corp; FYBROC Division)
Special Effect Pumps 361

100 100

Io ~ Recessed
Standard
J
v - -- .. '0
~
. - . .. ~- ..,- -
~ -.
~
60
-. _. -0- . . - '- 0
~b 60
. . .• 0 -
-
.,
~
,
0- ---
:I:

..---a-
0
-.:J
I--
,,
~

,
40 40

/
20
V ,,
,

--- - .-a--- f..--£I


--
20

~
- . i<>'
--a-- - -0 .' .
----a-- ~
-0 -
, .. - ~
- -

~. - '0 ' -

o o
o 100 200 300 400 500 600 700 800
Capacity (USGPM)

Fig. 16.11 Typical performance characteristic of an induced-vortex pump.


Dashed lines show performance of equal size conventional centrifugal pump; note similar head
but lower power than induced vortex pump.

Application. Induced-vortex pumps are used for pumping liquids carrying large or fibrous solids,
or having a moderate concentration of entrained gas. Such services are common in the food and waste
water industries. Flows typically range from 25 to 900 m3/hr (100 to 4,000 gpm), with heads to 30 m
(100 ft).

Viscous Drag Pumps


First developed in 1910 by Nikola Tesla, viscous drag or laminated-rotor pumps were intended to
effectively handle liquids whose viscosity ranged from 2,000 to 8,000 ssu. Today such applications are
generally handled more efficiently by rotary pumps.
Principle of operation. The "impeller" of a viscous drag pump is a series of coaxial disks, each
with one or more holes in the center, and separated from each other with spacers (Fig. 16.12). Energy
is imparted to the pumped liquid by viscous drag between the disks and liquid. The number of disks
and their spacing is varied, depending on the viscosity of the pumped liquid to maintain a moderate
relative velocity through the impeller, thereby achieving a reasonable efficiency.
362 Special Effect Pumps

I
t
Wearing
ring ' "

........... Concentric
casing

Fig. 16.12 Diagram of viscous drag (Tesla) pump.


Configuration shown is end suction with concentric casing.

140

120

100
~
0

~
c::
(I)
·13 80
==
w
o!S
CD
~ 60
a..
~
CD
J:
40

20

o 20 40 60 80 100 120
Flow-% BEP

Fig. 16.13 Type performance characteristics of a viscous drag (Tesla) pump.


Special Effect Pumps 363

Performance characteristics. Over the usual specific speed range of 700 to 1,000, the head character-
istic of viscous drag pumps ranges from flat to continually rising with decreasing flow (Fig. 16.13),
whereas the power characteristic rises continually with increasing capacity. Because the energy lost to
friction is high, the efficiency is quite low, generally not exceeding 25 percent.
Application. With efficiency well below that of rotary pumps, viscous drag pumps are limited to
pumping abrasive viscous liquids at elevated temperatures, conditions not suitable for rotary pumps.
Because the separation of the impeller disks is relatively close, viscous drag pumps cannot handle solids
of any significant size. Known designs produce flows to 230 m3/hr (1,000 gpm) and heads to 90 m (300 ft).

Rotating Casing (Pi tot Thbe) Pump


Rotating casing pumps, known also as "pitot-tube" or "impact" pumps, are a design that has been in
use, to varying degrees, for the past 80 years. In keeping with regenerative and partial-emission pumps,
they are a special-effect kinetic pump, intended to develop high heads at low flows.
Principle of operation. The principle of operation draws from that of a pitot tube, which when
facing directly into an oncoming fluid stream indicates the total head of the stream at that point. In a
rotating casing pump (Fig. 16.14), liquid enters the casing along the axis of rotation, and has momentum
added as it passes through the enclosed radial vanes of the impeller into the rotating casing. The action
of the rotating casing maintains the high velocity of the liquid stream. High-energy liquid is then drawn
off through the pitot tube and passed to the pump discharge. The head developed by the pump is equal
to the sum of the static pressure created by centrifugal force and the velocity head.

Inlet

Fig. 16.14 Section of "Roto-Jet" pitot tube pump.


(Courtesy EnviroTech PUMPSYSTEMS)
75 Iii
w
~
50 •
:ez
l:
Wr-;PSH REQUIRED 25
1....-----
o
2800

2600 ~-~V-7''''--+--i1
~ t=-i Obtainable
Minimum Rcquired Flow for Safc Pump Operation. L
with Bypass Orifice (0.156" Dia.)(J.962mm) 1 + - - - 1 - - - - 1 - - - ;

2400
I¥'
2200 ..
,....
2000 ..

-
~-
1800 ...

Iii 1600·..
W
~ 1400
::r::
Q 1200
t-=

1000

800

600 -

400
i

~'"
./ :">.
'" i-' 1m, RP"II
i -- DO :'\OT OPERATE
BI:.YONDTHIS UKIo
J
200 .-t - -.........=--t-- 400
- t'">... !1750RP\1 ! ~r-:-:-:=!:;-:,
o "'-~-~-~-~f4!I'~25~R~r~MJ--t---t---tI-~--t~~~+~~~~~i85~4'~K(='RTI'~=U~-t----1 300
t.--:~~ nm1 200
~I:=-:~ ~ ~ n:
~ ::r::
100 ai

o
Perfo manCE Base ~ On VIi ater

o 100 200 300 400 500 600


FLOW·GPM

Fig. 16.15 Performance characteristics of a pitot tube pump.


(Courtesy EnviroTech PUMPSYSTEMS)

Performance characteristics. Design specific speed is typically within the range 60 to 500. The
head characteristic depends on the size of the pitot tube, ranging from continuously rising with decreasing
flow to "drooping" (Fig. 16.15). The power characteristic rises continuously with increasing flow. For
the range of specific speeds covered, efficiency is good, values on the order of 60 percent having been
achieved. NPSHR is typically lower than a centrifugal pump of the same capacity.

364
Special Effect Pumps 365

Application. The hydraulic performance of rotating casing pumps, flows from 1.0-170 m 3/hr (5-750
gpm) and heads to 1,525 (5,000 ft), makes them suitable for the applications also met with regenerative,
partial-emission, or small multistage centrifugal pumps. The advantages offered by this design for such
applications are wide flow rangeability, no critical internal running clearances, smaller size than multistage
centrifugal pumps, and the ability to tolerate running dry. Its disadvantages are erosion when the pumped
liquid contains abrasive solids, and a low tolerance of air or entrained gas in the pumped liquid. Variations
of the design are used for integral lubricating oil pumps (Fig. 11.49).

Reversible Pumps

In the same manner as partial-emission pumps, reversible pumps are not strictly special effect pumps,
but their design, performance, and application are sufficiently different to warrant inclusion in this chapter.

Principle of operation. Head in reversible pumps is developed in the same manner as in


conventional centrifugal pumps. To be reversible, however, several compromises have to be made
in the pump design; specifically the impeller vanes must be radial, the casing concentric, and the
discharge nozzle radial. With this symmetrical construction, the pump performs equally in either
direction of rotation.

Performance characteristics. Unlike partial-emission pumps, whose impeller design is similar,


reversible pumps have a falling head characteristic (Fig. 16.16), a consequence of impeller discharge to
casing throat area ratios used in their design. The power characteristic rises with increasing capacity.
Efficiency at 30 to 40 percent is lower than a centrifugal pump of equal performance because the
design is compromised for reversibility. NPSHR is higher than an equivalent centrifugal pump for the
same reason.

140

120

100

cJ< 80

15<ll
J: 60

40

20

o 20 40 60 80 100 120 140 160


Flow-%

Fig. 16.16 Type head characteristic of a reversible pump.


366 Special Effect Pumps

Application. Reversible pumps are applied to low-energy auxiliary services, for example, cooling
water circulation through an engine, where it is necessary to have the pumped liquid move in one
direction regardless of the driver's direction of rotation. The one known design is specific speed 900,
with a capacity of 11.5 m3Jhr (50 gpm) and head of 12.2 m (40 ft) at maximum speed of 2,000 RPM.

BIBLIOGRAPHY

[16.1] Barske, U.M., and Dr. Ing.; "Development of Some Unconventional Centrifugal Pumps", Proc. Institute of
Mechanical Engineers (Britain), Vol. 174 No. 11, 1960.
17
Materials of Construction
- - . -.. ~ - - - - - - -

Centrifugal pumps are fabricated of almost all the known engineering materials, from simple thermoplastic
polymers through metals ranging from cast iron to the various nickel-based alloys to composites and
ceramics. The conditions of service and the nature of the pumped liquid finally determine which among
this wide range of materials will be the most suitable. A specific choice is based first on past experience
with the same liquid or a similar liquid. When past experience is lacking, material properties and known
performance must be used. Listings of the materials commonly recommended for various liquids can
be readily found in the Standards published by the Hydraulic Institute [1.1], in API-61O [3.1] and in the
catalogs and bulletins of pump manufacturers, particularly those who specialize in centrifugal pumps
for chemical service, the field that presents the greatest variety of material selection problems. Note, in
this connection, that the plant owner is ultimately responsible for the performance of the specified
materials, because the precise nature of the pumped liquid is solely under his or her control. It is the
pump manufacturer's responsibility to furnish the specified materials (provided, of course, they are
mechanically suitable).
The principal service conditions that affect the selection of materials are

1. Operating pressure
2. Pumping temperature
3. Head per stage (affects both the peripheral velocity of the impeller and the liquid velocity in the waterways)
4. Corrosiveness of the pumped liquid (can vary markedly with traces of halogens, halides, or compounds
of hydrogen)
5. Concentration and abrasiveness of any suspended solids
6. Load factor (fraction of time running) and expected life.

In selecting the material for any part of a pump, the material properties to be considered are

I. Strength: tensile, impact and endurance or fatigue.


2. Stiffness

367

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
368 Materials of Construction

3. Thennal expansion and thennal shock resistance


4. Corrosion resistance, considering the effects of velocity and stress
5. Erosion resistance (both abrasion and cavitation)
6. Feasibility of fabrication into the required component.

Table 17.1 gives a qualitative ranking of these properties for the usual pump materials.
In developing general guidelines for material selection, we first concentrate on the materials most
commonly used for individual parts. Table 17.2 summarizes these materials for the three major parts:
casing (plus inner casing where applicable), impeller, and shaft.
Table 17.3 lists the specification and chemical composition of the metals commonly used in centrifugal
pump construction.

CASING MATERIALS

The foremost requirement of the material of a centrifugal pump casing is that it be strong enough, given
the sections employed in the design, to safely contain the maximum working pressure. Next, the material
or combination of materials must be stiff enough to limit distortion of the casing under pressure and
imposed nozzle loads to that which will enable the pump to operate as intended. Finally, the material
must provide an economical balance between service life (determined by loss of wall thickness due to
corrosion or erosion or both) and cost.
Injection-molded polymers are viable materials for the casings of small, mass produced pumps for
low pressure, 1.0 to 1.4 bar (15 to 20 psig), at ambient temperature. Limited strength, low stiffness, and
the cost of molds generally preclude their use above these limits.
Cast iron has higher strength and stiffness than polymers, yet is still economical to produce. For these
reasons it is the material used for the casings of most centrifugal pumps. Its strength and stiffness are
limited, however, which places a practical upper limit on the pressure for which cast iron casings can
be built. This limit varies with size, being around 35 bar (500 psig) for small pumps and falling to 10
bar (150 psig) for large pumps. Cast iron has poor resistance to thermal shock, and therefore is generally
not used at pumping temperatures above 175°C (350°F), nor for services where the pumped liquid is
flammable or toxic. In the latter case, the concern is that should the pump be involved in a fire, attempts
to extinguish the fire may quench and fracture the heated casing.
Ductile iron offers higher strength, stiffness, and thermal shock resistance than cast iron, with the
degree of improvement depending on the grade specified. Many chemical pumps have high-elongation
ductile-iron casings to allow their use for flammable or toxic liquids. Larger pumps have ductile iron
casings, usually of lower elongation, to increase their pressure rating.
Carbon steel is stronger, stiffer, and more ductile than ductile iron. Beyond mechanical properties,
carbon steel has the advantage that it can be welded, thus enabling ready repair in the field, something
that is not feasible with either iron or ductile iron. Given its mechanical properties, carbon steel is the
usual material for medium- and high-pressure casings: in cast form for medium pressure and forged for
high pressure. Carbon steel is the standard casing material for pumps handling flammable or toxic liquids
in petroleum refining service, a practice related to the material's strength and thermal shock resistance.
Because they can be readily repaired by welding, carbon-steel casings are frequently used in mildly
erosive mineral processing services.
Chrome steels offer higher strength than carbon steel, superior corrosion resistance in some services
(e.g., boiler feed) and marginally better erosion resistance. When taking advantage of the higher strength
of chrome steels, the casing design must be able to accommodate the higher deflection that will occur
Table 17.1 Ranking of Properties of Usual Pump Materials

Property Material
-- - -- ------- - --"- - -----.- -"------------ - - - - -- ----- - -----
Iron Ductile iron Hard iron Bronze Steel 13 Chrome 316 Duplex Polymer Ceramic

Strength L L L-W I H I H L-I2 L


Tensile Impact L I L H W 13 H H I L
Endurance L L None H H H H None None
Stiffness L I L L H H H H L L
Thermal expansion L L L I L L I I H L
Thermal shock resistance L I L H H H H H I L
Corrosion resistance I I I H L I H H H H
Erosion resistance: Abrasion L L H L L I I I L H
Cavitation L L L H H H H L L
Ease of manufacture H H I H H I H I I L
I Varies with alloy; aluminum and nickel aluminum bronzes have high tensile strength.
2 Varies with polymer and extent of reinforcement.
l Above nil ductility transition temperature. ~
Code: L = low, I = intermediate, and H = high. 1\
~
~

~
~
1';
g.

$
to>
-.I
=

.iii·~
1:;-
Table 17.2 Common Pump Material Combinations
~

Service Condition Casing Inner casing/liner Impeller Shaft .,~::s


~
Water Fresh Polymer Polymer Steel "g.::s
Fresh Cast iron Bronze Steel
Salt Bronze Bronze Monel
Condensate (water) Steel, Iron Gray iron Chrome steel Steel
Boiler feed Low pressure Iron Bronze Bronze Steel
Medium pressure Chrome steel Chrome steel Chrome steel
High pressure Steel Chrome steel Chrome steel Chrome steel
Hydrocarbon Normal Steel Cast iron Alloy steel
High pressure Steel Steel Chrome steel Chrome steel
Abrasive Steel Hard metal Hard metal Chrome steel
High temperature Chrome steel Chrome steel Chrome steel Chrome steel
Corrosive 317 317 317 Duplex
Water injection Corrosive Duplex Duplex Duplex Duplex
Petrochemical Noncorrosive Ductile iron Cast iron Steel
Corrosive 316 316 316
Corrosive Alloy 20 Alloy 20 Alloy 20
Corrosive Polymer Polymer Stainless
Mineral processing Nonabrasive Steel Chrome steel Steel
Abrasive Iron Rubber Rubber lined Steel
Abrasive Hard metal Hard metal Steel
Abrasive Steel Hard metal Hard metal Steel
Table 17.3 Standard materials and their chemical composition
NOTE: Where the material is in wrought form, the International Standard defined is that relating to bar.
This table is supplied for information only and is not to be used for specification purposes.
(t) The materials defined here are proprietary alloys.

Prod.
Form Standard
Material Type C=Cast C Ni Cr Mo Cu Fe Others
W=
No Orade
Wrought

Grey ASTM
Class 25 C Class 25
Cast Iron A48 At the discretion of the foundry
ASTM
Class 40 C Class 40
A48
Austenitic ASTM 3.0 13.5 1.5 5.5 I
Type I C Type 1 Bal
Cast Iron A436 Max -17.5 -2.5 -7.5
ASTM 3.0 18.0 1.5 0.5
Type 2 C Type 2 Bal
A436 Max -22.0 -2.5 Max
Ferritic
60-40-18 At the discretion of the foundry
S.O. Iron
Si P
60-40-18 ASTM Not 3.0
C 2.5 0.08
Pressure A395 applicable Max
Max Max
Nickel Mn Si
BS 2.8 4.0 8.0 0.5
Chromium NiHard4 C 2D Bal 0.2 1.5
4844 3.2 -6.0 -10.0 Max
White Iron -0.8 -2.2
V
Carbon ASTM 0.30 0.50 0.50 0.20 0.30
WCB C WCB Bal 0.03
Steel A216 Max Max Max Max Max
Max
V
ASTM 0.25 0.50 0.50 0.20 0.30
LCA C LCA Bal 0.03
A352 Max Max Max Max Max
Max
V
ASTM 0.25 0.50 0.40 0.20 0.30
WCA C WCA Bal 0.03
A216 Max Max Max Max Max
~
... Max
~
N

Table 17.3 Standard materials and their chemical composition (Continued)

Prod.
Form Standard
Material Type C=Cast C Ni Cr Mo Cu Fe Others
W=
No Grade
Wrought
Martensiticl ASTM 0.15 1.0 11.5
CA-15 C CA-15 Bal
Ferritic A217 Max Max -14.0
Stainless ASTM 0.06 3.5 11.5 0.4
Steel CA6NM C CA6NM Bal
A4S7 Max -4.5 -14.0 -1.0
Nb N
ASTM 0.07 3.6 15.5 2.5
CB7Cu-l C CB7Cu-l Bal 0.15 0.05
A747 Max -4.6 -17.5 -3.2
-0.35 Max
Austenitic ASTM O.OS 9.0 IS.0 2.0
316 C CF-SM Bal
Stainless A744 Max -12.0 -21.0 -3.0
Steel ASTM 0.03 9.0 17.0 2.0
3l6L C CF-3M Bal
A744 Max -13.0 -21.0 -3.0
N
ASTM 0.03 10.0 16.0 2.0
316L W 316L Bal 0.10
A276 Max -14.0 -IS.0 -3.0
Max
ASTM O.OS 9.0 IS.0 3.0
317 C CG-SM Bal
A744 Max -13.0 -21.0 -4.0
Cb
ASTM 0.08 9.0 18.0
347 C CF-SC Bal SxC
A744 Max -12.0 -21.0
-1.0
Mn N Cb V
ASTM 0.06 11.5 20.0 1.5
XM19 W XM19 Bal 4.0 0.2 0.1 0.1
A479 Max -13.5 -23.5 -3.0
-6.0 -0.4 -0.3 -0.3
-
- - ----
-
Table 17.3 Standard materials and their chemical composition (Continued)

Prod.
Form Standard
Material Type C=Cast C Ni Cr Mo Cu Fe Others
W=
No Grade
Wrought
I
Duplex Nominal analysis
Stainless Ferralium 0.05 N
Steel 255 C 6 25 3 2.5 Bal
255-3SC (t) Max 0.18
Mn Si N
ASTM 0.04 4.75 24.5 1.75 2.75
CD4MCuN C 1B Bal 1.0 1.0 0.10
A890 Max -6.00 -26.5 -2.25 -3.25
Max Max -0.25
Mn Si N W
ASTM 0.03 6.5 24.0 3.0 0.5
CD3MWGJN C 6A Fe 1.0 1.0 0.2 0.5
A890 Max -8.5 -26.0 -4.0 -1.0
Max Max 0.3 1.0
Mn Si N
ASTM UNS 0.04 4.5 24.0 2.9 1.5
255 W Bal 1.5 1.0 0.10
A479 32550 Max 6.5 27.0 3.9 2.5
Max Max 0.25
High Alloy ASTM 0.07 27.5 19.0 2.0 3.0
CN7M C CN7M Bal
Austenitic A744 Max 30.5 22.0 -3.0 -4.0
Stainless Mn Si N
Steel ASTM 0.03 23.5 20.0 6.0 0.75
CN3MN C CN3MN Bal 2.0 1.00 0.18
A351 Max -25.5 -22.0 -7.0 Max
Max Max -0.26
Mn Si N
ASTM 0.02 17.5 19.5 6.0 0.50
254sMo C CK3MCuN Bal 1.2 1.00 0.18
A351 Max -19.5 -20.5 -7.0 -1.00
Max Max -0.24 I

~
~
Table 17.3 Standard materials and their chemical composition (Continued)

Prod.
Form Standard
Material Type C=Cast C Ni Cr Mo Cu Fe Others
W=
No Grade
Wrought I
Nickel W Si
Worthalloy 0.12 22.0 3.0 3.0 10.0
Chromium C (t) Bal 1.0 3.0
55 Max -24.0 -5.0 -5.0 Max
Molybdenum -3.0 -5.0
Nb+Ta Al Ti Co
Inconel 0.10 Min 20.0 8.0 5.0
625 W 3.15 0.40 0.40 1.0
625 (t) Max 58.0 -23.0 -10.0 Max
-4.15 Max Max Max
Ti Al
Incoloy 0.05 38.0 19.5 2.5 1.5 Min
825 W 0.6 0.2
825(t) Max -46.0 -23.5 -3.5 -3.0 22.0
-1.2 Max
Mn W
ASTM 0.02 15.0 15.0 2.0
CW-2M C CW-2M Bal 1.0 1.0
A494 Max -17.5 -17.5 Max
Max Max
Co W Si V
Hastelloy 0.01 14.5 15.0 4.0
C276 W Bal 2.5 3.0 0.08 0.35
C-276(t) Max -16.5 -17.0 -7.0
Max -4.5 Max Max
Nickel V Mn Si Co
Hastelloy ASTM 0.12 1.0 26.0 4.0
Molybdenum C N-12MV Bal 0.20 1.0 1.0 2.5
B A494 Max Max -30.0 -6.0
--0.60 Max Max Max
Nickel Cb
ASTM 0.35 26.0 3.5
Copper Monel C M-35-1 Bal 1.0
A494 Max 33.0 Max
3.0
(1) Al W
K-500 QQ-N-286 0.25 2.0 Co
W Fed'l 63.0 Bal 2.30 0.35
Monel Class A Max Max (1)
-70.0 -3.15 --0.85

(1) Range given for Ni is total for Ni + Co.


Table 17.3 Standard materials and their chemical composition (Continued)

Prod.
Fonn Standard
Material Type C=Cast C Ni Cr Mo Cu Fe Others
W=
No Grade
Wrought
Leaded Sn Zn Pb
ASTM 1.0 84.0 0.03
Gunmetal C83600 C C83600 4.0 4.0 4.0
B-584 Max -86.0 Max
-6.0 -6.0 -6.0
Sn Zn Pb
BS 2.0
LG4 C LG4 Bal 6.0 1.5 2.5
1400 Max
-8.0 -3.0 -3.5
Tin Bronze Sn Pb
ASTM 0.5 88.0
C90700 C C90700 10.0 0.5
B584 Max -90.0
-12.0 Max
Phosphor Sn Zn Pb P
BS 0.5
Bronze PB2 C PB2 Bal 11.0 0.30 0.50 0.15
1400 Max
-13.0 Max Max -0.60
Leaded Sn Zn Pb P
ASTM 1.0 78.0
Bronze C93700 C C93700 9.0 0.8 8.0 0.15
B548 Max -82.0
-11.0 Max -11.0 Max
Nickel AI Mn I
ASTM 3.0 78.0 3.0 I
Aluminium C95500 C C95500 10.0 3.5
B148 -5.5 Min -5.0
Bronze -11.5 Max I

Titanium N H 0 I
ASTM 0.10 0.30 Ti
Grade 2 W 2 0.03 0.015 0.25
B265 Max Max Bal
L - ____
Max Max Max
--- - -- -- --- - - - - - ------ --- --- - - '------- - - '------- - "----- -

~
376 Materials of Construction

at sealing surfaces and locating fits. Chrome steels are weldable, with the 13 chrome, 4 nickel alloy
being considered the most weldable, and 5 chrome, 0.5 molybdenum the least. All require preheating
before welding, and any structural welds require postweld heat treatment.
The choice between a cast or forged casing depends on an assessment of the risk of in-service leakage
versus manufacturing cost. In high-temperature applications, repeated heating and cooling of the pump
can cause the internal shrinkage inherent in cast casings to develop into minor cracking, thereby allowing
an originally pressure-tight casing to leak. For medium-pressure casings, typically 100 bar (1450 psig)
at 230°C (450°F), cast casings are normally used because they are economical and can be repaired by
welding. At higher pressures, forged casings are deemed the better choice because they eliminate the
potential cracking problem. Although forged casings are more expensive to manufacture than cast casings,
the difference decreases with increasing pressure, and is generally negligible for pressures beyond 200
bar (3,000 psig).
At low temperatures, below what is known as the "nil ductility transition temperature" (which varies
with alloy, heat treatment, and section thickness), carbon steel and chrome steels becomes brittle. In
these circumstances, austenitic stainless steel, or in some cases alloys of aluminum, both of which do
not exhibit a "nil ductility transition temperature," are used.
Corrosive liquids pose a problem in that the corrosion resistance of the usual materials is almost the
inverse of their strength and stiffness. Reinforced polymers of the appropriate grades offer extremely
good corrosion resistance. These materials, however, have relatively low strength and stiffness, and
therefore casings made of them have a low pressure rating.
Bronze is produced in a wide range of alloys, from leaded gunmetals, which offer moderate resistance
to seawater corrosion (depending on temperature and contamination), through phosphor bronze, with
good resistance to seawater and mild acids, to nickel aluminum bronze and similar alloys, with excellent
resistance to seawater. All the bronzes have higher strength and stiffness than reinforced polymer, but
across the range of alloys cited, the mechanical properties vary from below that of cast iron to approaching
that of the chrome steels. Because bronze is not used extensively for centrifugal pump casings, most
casings produced are designs based on cast iron. As such, their pressure rating will depend on the alloy
being used. For leaded gunmetals and phosphor bronze, the pressure rating will be lower.
Lined casings afford high corrosion resistance without the expense of having to produce the entire
casing of an exotic or impractical material. Two forms of construction are used: a separate lining
contained by the casing (Figs. 26.42 & 26.43), or a lining bonded to the casing. Of the usual lining
materials, only glass and Teflon® are bonded to the casing. The linings have various properties. Rubber
exhibits high resistance to corrosion and erosion, but is limited by temperature. Ceramic is useful for
corrosive or erosive services at temperatures above the limit of rubber. Graphite, impregnated to make
it impervious, has high corrosion resistance but is soft and cannot tolerate any solids in the pumped
liquid. Teflon® is inert in most liquids but too soft to have useful resistance to erosion; its advantage
over graphite is simpler casing construction. Glass is inert to all liquids, but prone to microcracking
caused by differential thermal expansion at high temperatures, which can lead to corrosion of the substrate.
The pressure rating of lined casings is determined by the material of the structural portion of the casing.
This is usually cast iron or ductile iron, and so the pressure rating is limited to around 17 bar (250 psig).
Austenitic stainless steel covers a wide range of alloys, the most common being type 316 stainless
steel (or ASTM A744 CF8M for castings). The corrosion resistance of these alloys depends on the
particular alloy and liquid. They all rely on a passive oxide film for corrosion resistance. Localized
corrosion will start at damaged areas of this oxide film. An example of how this limits their application
is the high risk of corrosion if exposed to stagnant seawater. The chloride ion initiates pitting and the
stagnant conditions do not provide enough oxygen to re-establish a continuous oxide film. As a class
of materials, austenitic stainless steels have strength and stiffness higher than ductile iron but below
steel. Their elastic limit (yield point) is not well defined, which makes components of these alloys prone
Materials of Construction 377

to plastic deformation (permanent distortion) if stressed close to their yield point (usually based on 0.2
percent offset). With these mechanical properties, casing designs based on steel must be derated for
pressure when produced in austenitic stainless steel. On the other hand, if a casing is required for a
particular pressure rating, its effective sections in austenitic stainless steel must be larger than in steel.
The so-called "higher alloys," those with high nickel or molydenum contents, have excellent corrosion
resistance to a particular liquid, but often poor resistance if one element in the liquid's composition is
changed. For the purposes of casing pressure rating, the mechanical properties of the higher alloys are
similar to those of austenitic stainless steel.
Duplex materials, alloys whose structure is a mixture of two phases, austenite and ferrite, have
excellent resistance to corrosion by sea water and napthenic acid, and can be used for liquids containing
the latter at temperatures up to 260°C (500°F). Although the corrosion resistance of duplex alloys is not
equal to higher-nickel-based alloys for some liquids, they have the advantage of higher strength. Given
this, duplex casings can have pressure ratings close to those of chrome steel while achieving significantly
higher corrosion resistance.
Hard metals for solids-handling pumps range from austenitic manganese steels for dredge pumps
(whose casing material must have sufficient ductility to absorb the impact of large solids), through
Nihard® for high concentration slurries to high chrome irons for high concentration slurries of a corrosive
liquid. Solid casings are constructed entirely of hard metal, generally with replaceable wear plates
alongside the impeller (Fig. 2.20). Hard metals have limited strength and low ductility, therefore the
attainable casing pressure ratings are low. When the required pressure rating is beyond the limit of hard
metal, the casing is constructed by containing a hard-metal liner within a carbon-steel casing. In hydrocar-
bon catalyst slurry service, similar construction is used to achieve high erosion resistance while maintain-
ing the pressure containment integrity required for flammable services.

INNER CASING MATERIALS

Axially split single and multistage pumps of the arrangements shown in Figs 2.13 and 5.9, and double-
casing (barrel) pumps (Figs. 3.14 & 3.15) have an inner casing that serves to collect, diffuse, and guide
the pumped liquid. Depending on the conditions of service, the inner casing can be a lesser material
than the outer casing, the same material, or a higher material. Table 17.2 shows the common combinations
of casing and inner casing materials. In corrosive hydrocarbon applications, there is also good experience
with duplex, 316, and 317 inner casings in carbon-steel casings.

IMPELLER MATERIALS

The impeller of a centrifugal pump is a dynamic component, which in operation is subjected to

1. Stresses caused by centrifugal force


2. Bending stresses produced by pressure differentials and fluctuations within the waterways.
3. High liquid velocities
4. Possible cavitation.

Drawing on the list of fundamental requirements given at the beginning of this chapter, the materials
for impellers must be
378 Materials of Construction

1. Strong enough and stiff enough to withstand the stresses produced by rotation and pressure without fracture
or excessive strain
2. At least as corrosion resistant to the pumped liquid as the casing
3. Significantly more resistant to cavitation erosion than the pump casing
4. As resistant to abrasive erosion as the pump casing
5. Possessed of a thermal expansion rate able to maintain the required fit with the shaft at the pumping temper-
ature.

Injected molded thermoplastic is used for the impellers of small, mass-produced single- and multistage
pumps on water, circulator, and boiler feed service.
Cast-iron impellers are used for hydrocarbon service up to peripheral speeds of 45 rn/sec (145 ft/sec)
and temperatures of 230°C (450°F). For water service, cast iron is really suitable only for small pumps,
up to about 170 m 3Jhr (750 gpm) (typically 4-in. discharge), because in larger pumps the localized
cavitation that frequently occurs has enough energy to cause premature impeller erosion. The same risk
does not exist in hydrocarbon services because the intensity of cavitation is much lower (see Chap. 19).
Bronze is widely used for impellers wherever it is suitable because (1) it is easy to cast in complicated
cored shapes, (2) it produces smooth as-cast surfaces, (3) it is easy to machine, (4) it does not rust, and
(5) its resistance to cavitation erosion, depending on the alloy used, is 3 to 40 times higher than that of
cast iron. Bronze impellers should not be used with cast-iron casings if the pumped liquid is a strong
electrolyte, or if the pumped liquid is basic (pH above 7).
The coefficient of thermal expansion of bronze is 1.4 times that of carbon steel, therefore the clearance
between a bronze impeller and a steel shaft will increase as the pumping temperature increases. At the
same time, the impeller hub will expand axially, thereby applying an axial load to whatever is retaining
the impeller and producing a corresponding tensile load in the shaft. To avoid these difficulties, bronze
is not normally used for pumping temperatures above 120°C (250°F).
Most of the bronze alloys used (leaded gunmetal, tin bronze, phosphor bronze, and leaded bronze)
have low tensile strength and stiffness, which limits the peripheral speed to which they can be used.
The centrifugal stress developed in an impeller and the resulting stretch at the impeller hub can be
significant at the higher peripheral speeds of high-head pumps. For example, a 305-mm (12-in.) bronze
or iron impeller mounted on a 75-mm (3-in.) shaft and rotating at 3,600 rpm will have its bore stretch
by approximately 0.028 mm (0.0011 in.). At a pumping temperature of 120°C (250°F), and assuming
the shaft is steel, a bronze impeller will have its bore increased a further 0.036 mm (0.0014 in.), producing
a total additional clearance of 0.064 mm (0.0025 in.) between the shaft and the impeller, which is
excessive. To avoid the cumulative effect of excessive centrifugal and thermal expansion, the empirical
limit on the peripheral speed of the common alloy bronze impellers handling hot liquids is approximately
59 rn/sec (160 ft/sec), or a head of 114 m (375 ft) per stage.
Aluminum, aluminum manganese, and nickel aluminum bronzes have high tensile strength (approach-
ing that of 13 chrome steel) and excellent corrosion resistance. Provided the foundry has the necessary
expertise, their castability is as good as that of the lower-strength alloys. Aluminum bronzes are used
for seawater, in conjunction with a casing of the same alloy, and for high-head services where the lower
strength alloys will not withstand the centrifugal stress or the bending stress produced in the shrouds
by pressure pulsations within the impeller passages.
Steel impellers do not have the corrosion resistance of cast iron but do exhibit better resistance to
cavitation erosion. They are used to a limited extent in noncorrosive hydrocarbon services. For small
impellers, cast steel costs about the same as 13 chrome, therefore many manufacturers furnish only 13
chrome since it is a superior material.
The chrome steel alloys used for impellers are 13 chrome (CA15) or 13 chrome 4 nickel (CA6NM),
with the latter generally considered the better in terms of castability and weldability. In hydrocarbon
Materials of Construction 379

services where H2S is present, CA6NM is modified to limit the carbon to 0.03 percent so the castings
can be tempered to a hardness low enough to avoid stress-corrosion cracking. Chrome steel is necessary
for high-purity boiler feed water, and is suitable for all boiler feed waters over a pH range of 4.S to
14.0. Both alloys have proven useful in pumping caustic liquors at temperatures up to 120°C (2S0°F).
The nickel bearing alloy CA6NM has the better corrosion resistance, making it a suitable substitute for
cast iron in applications where copper-bearing alloys are not acceptable and cast iron will be prone to
premature cavitation erosion.
Austenitic stainless steel, of which type 316 (CF8M) is the most common, is used for its corrosion
and cavitation erosion resistance. These alloys are suitable for impellers that are mounted with a clearance
fit on an austenitic stainless steel shaft, such as chemical and some water pumps. They are usually not
suitable for impellers in high head multistage pumps at temperatures above 120°C (2S0°F), because
differential expansion with the high strength shafting (see Structural requirements later in this chapter)
will loosen the interference fit needed to maintain rotor balance.
For applications where cavitation erosion cannot be overcome with system or pump design changes,
there are now available proprietary austenitic stainless steels whose cavitation erosion resistance is 40
times that of cast iron. These are high manganese alloys, of high strength and toughness, that work
harden at the surface in the presence of cavitation. They are available as welding consumables for repairs
and castings for new impellers.
Chemical pumps whose casings are fiber-reinforced polymer use impellers of the same material. To
ensure the integrity of the impeller mounting on the shaft, a metal insert is frequently molded into the
impeller, then machined to the required dimensions. The low strength and stiffness of fiber-reinforced
polymer limits the head per stage to ISO m (SOO ft).
By definition, higher alloy chemical pumps are handling liquids that are strong electrolytes, therefore
the impeller, and all the other wetted parts for that matter, are of the same or a similar alloy.
Duplex impellers are used in two circumstances: (1) when the pump's casing or inner casing is a
duplex alloy or (2) when the pump's casing or inner casing is an austenitic alloy but the pump's
temperature or rotative speed or both preclude the use of austenitic impellers due to high-differential
thermal expansion or high strain. As already noted, under casings, duplex alloys cannot be used at
temperatures above 260°C (SOO°F) because of the risk of embrittlement.
Within its limits of temperature and strength, rubber offers outstanding resistance to erosion, therefore
the impellers of rubber-lined slurry pumps are also rubber lined whenever the service conditions allow
it. These impellers are fabricated by molding a thick rubber lining onto a ductile iron or carbon steel
"skeleton." The threaded connection used to mount the impeller on the shaft is machined into the metal
skeleton. Rubber lined impellers are limited to pumping temperatures of 6SoC (1S0°F) for natural rubber
and lOsoC (22S0F) for synthetic rubbers, and heads per stage of 4S m (1S0 ft).
For severe slurry applications (high concentration of solids, sharp or acicular fines, large solids), the
hard-metal slurry pumps used have hard metal impellers, usually of Nihard®, sometimes of high-chrome
iron when the liquid is corrosive. Hard-metal impellers are also used in rubber-lined slurry pumps when
the required head is higher than can be produced with a rubber-lined impeller, or in applications where
the combination of materials has demonstrated better erosion resistance than an all-rubber-lined pump.

WEARING PARTS

Usually a centrifugal pump's wearing parts are its impeller and casing wearing rings, any interstage
bushings and sleeves, the balancing device (if used), the shaft sleeves, and the throat bushings. When
mechanical seals are used, they also have wearing parts, which are specialized and dealt with in Chapter
380 Materials of Construction

Table 17.4 Impeller and Wearing Ring Material Combinations

Impeller/casing wearing ring Impeller material


material combination Iron Bronze Steel 13 Chr 316

Steel/lron L1 L1
Bronze/bronzea L1
13 chrome/17 chromeb L1 L1
316/Stellite on 316U L1
• Generally different alloys to improve galling resistance.
b Hardness difference at least 50 BHN unless softer ring 450 BHN or higher. Some manufactur-

ers use different alloys of 13 chrome steel, for example, types 410 and 420, instead of 13
and 17 chrome.
, One-surface is fusion hard coated for galling resistance. The alternative is to use increased
running clearance.

9. Similarly, the impeller and casing of slurry pumps are wearing parts, and have already been discussed
under casing and impeller materials.
The purpose of wearing rings and important aspects of their mechanical design are covered in Chapter
4. Of particular note for the selection of materials is the very practical requirement of not putting brittle
materials in tension because they are prone to fracture. The choice of materials for wearing rings is
determined first by the impeller material, then by the need for the smallest running clearance consistent
with good galling resistance. Table 17.4 summarizes the usual wearing ring material combinations and
shows which impeller materials they are typically used with. Wearing ring materials always have corrosion
resistance at least equal to that of the impeller. Table 17.4 does not extend to the higher alloys. These
are almost exclusively used in chemical pumps, many of which have semiopen impellers and therefore
do not have wearing rings. In those cases where pumps with closed impellers are used, both the wearing
rings are the same material as the impeller, and usually the running clearances are increased to minimize
the risk of galling. The practice for duplex impellers follows that for 316 stainless steel. The impeller
wearing ring, if used, is duplex, and the casing ring is fusion hard coated duplex.
Hard coating has already been mentioned in connection with improving the galling resistance of
certain wearing ring material combinations. In this circumstance, only one of the surfaces needs to be
hard coated to achieve the desired effect. Fusion hard coating with Stellite or similar materials is the
most common method, and is entirely serviceable, provided the coating is applied to low-carbon alloys
when high corrosion resistance is needed.
In applications involving low concentrations of solids in the pumped liquid, it is often practical to
hard coat both the rotating and stationary surfaces of the impeller's running clearances to reduce the
wear rate. In these cases, it is preferable to hard coat the impeller hub, thereby avoiding the difficulty
of having to safely mount and retain a hard coated impeller ring. The means of coating can be fusion,
plasma transfer arc, or high-velocity oxy fuel spray. When spray coating is used it is important to use
a high-density coating material (e.g., tungsten carbide) to help ensure a good bond to the substrate.
Hard coating is not the only means of improving the abrasion resistance of pump wearing surfaces.
Components of hardenable material can have their wearing surfaces fully hardened to a depth of some
0.5 mm (0.020 in.) by induction or laser hardening, thereby providing better wear resistance while
retaining the general ductility needed for the component function. Alternatively, some materials can
have their abrasion resistance increased by surface conversion, boron diffusion being the most common
today (but not for austenitic or duplex materials, which are embrittled by this process), followed by ion
implantation, then nitriding. The surface conversion processes in use today produce only a relatively
Materials of Construction 381

thin hard region, ranging from 0.01 mm (0.0005 in.) for ion implantation through 0.25 mm (0.010 in.)
for nitriding applied to a suitable alloy.
Interstage sleeve and bushing materials follow the same practice as wearing rings. Balancing device
components, the balancing drum or disk and the matching bushing or head in tandem impeller pumps,
or the sleeves and bushings in opposed impeller pumps, are subject to a high pressure drop per unit
length and therefore have materials chosen for good resistance to high velocity liquid erosion. At the
same time, the running clearance needs to be kept small, so galling resistance is also important. For
small, low-pressure multistage pumps, bronze, Ni-resist, or even iron components have proven serviceable.
Pumps designed for higher pressures generally use 13 chrome versus 17 chrome components, with the
appropriate hardness differences, as a minimum. In applications involving even quite low concentrations
of solids, the rate of balancing device wear can be reduced by hard coating the drum, disk, or sleeves.
Higher concentrations of abrasive solids require even more elaborate treatment, some designs having
resorted to solid tungsten carbide components to achieve a tolerable service life. Corrosive services are
usually accommodated by making both the components of the same alloy as the impellers, then hard
coating the drum, disk, or sleeves for galling resistance.
Shaft sleeves for pumps with packed box seals are bronze, 13 chrome, or hard-coated 316 stainless
steel in order of ascending PV rating (pressure times velocity). The sleeves for mechanical seals are
generally 316 stainless steel because it is suitable for most applications. In high-temperature service,
however, differential thermal expansion is a concern, and the sleeve is the same material as the shaft.
Shaft sleeves for pusher-type mechanical seals (see Chap. 9) are more durable (resistant to fretting
erosion) if hard coated in the region under the dynamic gasket.
Throat bushing materials generally follow those used for wearing rings. When a close clearance is
needed to change the pressure at the seal, hard coating the sleeve under the bushing is desirable to enable
a close clearance without raising the risk of galling in the clearance. An alternative approach is to use
a floating carbon throat bushing.

SHAFTS

Unless a pump shaft is completely isolated from the pumped liquid, a rare and difficult achievement,
corrosion resistance is the first consideration in selecting its material. This is so for two reasons. First,
in general, the strength of typical shaft materials decreases with increasing corrosion resistance. Second,
corrosion, even at low rates of penetration, has a significant effect on endurance strength, an important
factor for a component subject to cyclic bending and torsional stress. Following corrosion resistance,
the next consideration is strength, both impact for shock resistance and endurance to avoid failure by
fatigue. The final consideration, of particular importance for multistage pumps, is dimensional stability,
a property necessary to maintain shaft straightness.
Low-carbon steel is suitable for the shaft in pumps handling a wide variety of liquids at low temperatures
or in noncritical services (see Table 17.2). When these pumps have a packed box shaft seal, the shaft
is usually fitted with shaft sleeves to avoid wear of the shaft. Alloy steels, typically AISI 4140 or 4340,
are employed when a higher strength material is needed, or when the service warrants a higher design
factor. High-speed boiler-feed pumps and pumps handling mildly corrosive water at ambient temperature
or hydrocarbon at 246°C (475°F) and higher, have shafts of 13 percent chrome steel, AISI Type 410.
This alloy is resistant to corrosion by high-velocity water, has strength close to that of the alloy steels,
and has superior high-temperature strength and stiffness. Provided the material used has been adequately
stress relieved, it has good dimensional stability.
Type 316 stainless steel is widely used for shafts in chemical pumps. Its strength is similar to low-
carbon steel, and its dimensional stability is poor, two factors that usually preclude its use for larger
382 Materials of Construction

pumps, where high strength is necessary, or for multistage pumps. Shafts of the higher austenitic alloys
are used only when necessary for their superior corrosion resistance. Most the alloys have mechanical
properties similar to type 316 stainless steel.
Bronze pumps for seawater service use either bronze or Monel shafts depending on the pump's size
and service. The preferred Monel alloy is KSOO, which has strength and stiffness close to that of 13
percent chrome steel.
In medium-temperature, corrosive hydrocarbon service, duplex shafts are used with duplex impellers
to avoid fit relaxation due to differential thermal expansion. Adequate stress relief is necessary to ensure
dimensional stability of duplex shafts at elevated service temperatures.

MATERIAL CLASSIFICATION

The material classification of a centrifugal pump can be determined by the materials of its three principal
components: casing (and inner casing where applicable), impeller, and shaft. This is not to say the
materials of the running clearances are unimportant, but they tend to be determined by the impeller
material and therefore do not have to be included in the classification. Table 17.2 shows a summary of
the common material combinations for the three principal components. API-610 [3.1] includes a detailed
tabulation of the material classes commonly used in the oil industry.

PUMP FITTINGS

During the history of pump development, the expression "pump fittings" has been used rather loosely
to mean two entirely separate things. In the water and industrial markets, it refers to the general

Table 17.5 Materials for Various Classes of Pump Fittings

Part Ref No.' Standard Fitted All Iron Fitted All Bronze Fitted

Casing Cast iron Cast iron Bronze


Stuffing box cover 11 Cast iron Cast iron Bronze
Impeller 2 Bronze Cast iron Bronze
Impeller ring 8 Bronze Steel Bronze
Casing ring 7 Bronze Cast iron Bronze
Stuffing box ring 27 Bronze Cast iron Bronze
Diffuser 5 Cast iron or bronze Cast iron Bronze
Stage piece 109 Cast iron or bronze Cast iron Bronze
Shaft (with sleeve) 6 Steel Steel Bronze or Monel
Shaft (without sleeve) 6-A Chrome steel or steel Chrome steel Bronze or Monel
or steel
Shaft sleeve 14 Bronze Steel or chrome Bronze
steel
Gland 17 Bronze Cast iron Bronze

Typical service from Fresh water Chemical; Salt water


Table 17.2 non-corrosive
I Parts in this list and in Figs. 17.1 through 17.4 are numbered according to the Standards of the Hydraulic Institute [1.1]. This

standard gives stationary parts odd numbers and rotating parts even numbers. The standard was first proposed to the Hydraulic
Institute by Charles J. Tullo, Chief Engineer, Worthington Corporation.
Materials of Construction 383

Fig. 17.1 Section of a double-suction, single-stage pump with shaft sleeves.


Numbers refer to parts listed in Table 17.5.

construction features of the pump, for example, "ball-bearing-fitted pump," or to the combination of
materials used in the pump, for example, "all-iron-fitted pump." In the fire protection market (as in the
expression "underwriter fittings"), it may refer to various pieces of auxiliary equipment such as valves,
gauges, or even tools. Table 17.5 shows the various component materials for the three commonly used
"fitting" classifications, and includes a cross reference to the material classifications in Table 17.2.
Figures 17.1 through 17.4 illustrate the materials used for particular parts in four different pump types.
Beyond the three common pump "fitting" classes detailed in Table 17.4, there are two more worthy
of mention.

Acid-Resisting Pump

An acid-resisting pump is one in which all the parts in direct contact with the pumped liquid are
constructed of materials that offer the maximum resistance to its corrosive action. Typically such pumps
were produced in silicon iron, a highly corrosion resistant but very brittle material, or acid-resisting
bronze. Today, all 316 stainless steel or Alloy 20 construction would be more usual, the choice depending
upon the acid and its concentration and temperature.

Salt Water Pumps

Centrifugal pumps handling salt water may be standard fitted (cast-iron casing with bronze fittings),
all iron or all bronze, or with an iron casing and stainless steel fittings. Although thousands of standard
384 Materials of Construction

Fig. 17.2 Section of a single-suction, single-stage pump without shaft sleeves.


Numbers refer to parts listed in Table 17.5.

Fig. 17.3 Section of a two-stage axially split pump.


Numbers refer to parts listed in Table 17.5.
Materials of Construction 385

Fig. 17.4 Section of an end-suction, single stage pump.


Numbers refer to parts listed in Table 17.5.

fitted pumps are used for this purpose, such construction is not suitable if the sea water is contaminated
(for example, harbour water). Failures are usually caused by galvanic action between the bronze parts
and the cast-iron casing, which results in either the loss of the casing or the bronze parts. Failure in the
latter mode occurs if the casing wetted surface area is large enough to become an effective cathode once
the cast-iron surface is graphitized (see corrosion in this chapter).
An all-iron pump discourages galvanic corrosion, but it may occur nevertheless. A certain amount
of iron dissolution may take place, leaving graphitized areas that act as cathodes to the uncorroded areas
of cast iron. The resulting galvanic action is self-accelerating. To avoid the graphitization (and poor
resistance to cavitation erosion) of cast iron, the impellers and other small pump parts may be made of
stainless steel. The grade of stainless steel should be at least type 316 (CF8M), and that is suitable only
for well-aerated liquid. If the pump is to stand filled with liquid for long periods, austenitic stainless
steel containing at least 6 percent molybdenum or all-bronze construction is necessary.

MATERIAL PROPERTIES

A summary of the qualitative properties of typical pump materials is given in Table 17.1. Structural
requirements, and to some degree corrosion and erosion resistance, are dealt with in the discussion of
386 Materials of Construction

materials for the major components. The following text addresses corrosion and erosion resistance,
thermal expansion, and thermal shock resistance in more detail, then aspects of structural requirements
not dealt with in the discussion of components.

Corrosion
In centrifugal pumps, corrosion differs from general practice in two important respects. First, liquid
velocities in pumps are inherently higher than in pipelines or vessels, so corrosion data based on low-
velocity tests may not be applicable to pump parts. Second, some pump parts (e.g., seals and shafts)
cannot tolerate appreciable penetration or weight loss without failure, thus when corrosion determines
the material selection these parts may have to be made of materials superior to those in the rest of
the system.
Corrosion can be broadly defined as the deterioration of materials by chemical or electrochemical
action. For metals, whose electrons are free to move, the definition can be narrowed to the deterioration
of solids by liquid electrolytes. Nonmetals generally do not have free electrons, therefore any deterioration
is by chemical action alone. The corrosion of metals is fundamentally galvanic. From Fig. 17.5 the
essential requirements are a potential difference between two sites immersed in an electrolyte and
connected with an external electric circuit. The potential difference causes metal loss or oxidation at the
anode (metal ions go into solution and electrons move into the external circuit) and metal deposition or
reduction at the cathode (electrons from the external circuit reduce ions from solution). In pumps, the
prevailing high velocities generally wash away metal deposited at the cathode.
As in electroplating, the rate of anode consumption or corrosion depends on current density. Current
density, in turn, depends on the potential difference and balanced oxidation-reduction reactions. These
dependencies lead to means of limiting corrosion. First, the potential difference can be reduced. Second,
the oxidation-reducation balance can be kept at a very low level. Two means are available to achieve
the latter: cathode polarization, in which the reduction rate limits the balance, and passivation, in which
initial oxidation of the anode renders it essentially inactive. A mechanical factor of great consequence
to the oxidation-reduction balance is the relative size of anode and cathode; a relatively small anode is
susceptible to rapid corrosion. In this connection, materials protected by passivation are vulnerable if
the passive film is perforated, because the unprotected region becomes a small anode.
Corrosion damage to metals is generally identified by 10 types. While in some cases the whole mechanism
is not yet fully understood, these 10 types of corrosion are the products of three basic galvanic cells: unlike
electrodes, which is self-explanatory; stress-induced potential differences; and concentration, potential dif-
ferences created by variations in electrolyte concentration. A brief description of each type follows. For a
detailed treatment of this complex subject, see the Corrosion Engineers Reference Book [17.1 J.

ELECTROLYTE
ANODE CATHODE
(OXIDATION) M+ M+ (REDUCTION)

CORRMO+SION_
M
I __---'c=\=::... . _ PRODUCT OF CORROSION
M++e--M
- +e~

e\ ELECTRICAL
)e-
CONNECTION
POTENTIAL
DIFFERENCE

Fig. 17.5 Fundamental corrosion mechanism.


Materials of Construction 387

Influence of pH

The pH value of a liquid is a quantitative representation of its relative acidity or alkalinity. The value
is based on the concentration of H+ (positive hydrogen) ions as opposed to OH- (negative hydroxyl)
ions in the solution. It is calculated as follows:

1
pH = log ==0----=---,---
H+ concentration

The lower the pH, the more acidic the solution. A solution with a pH value of 7.0 is neutral; values
above 7.0 indicate alkalinity and values below 7.0, acidity. Because pH values are expressed logarithmi-
cally, changes in pH represent more than a direct linear change. For instance, a solution having a pH
of 5.0 is 10 times more acidic than one with a pH of 6.0.
The pH of a given solution varies somewhat with temperature changes, decreasing rather rapidly up
to 150°C (300°F) and remaining fairly constant at higher temperatures. For instance, a solution with a
pH of8.5 at21 °C (70°F) will have a pH of about 7.0 at 150°C (300°F) and 6.8 at 260°C (500°F) (Fig. 17.6).
How pH affects metal corrosion depends on whether the metal oxide is stable in both acid and alkaline
solutions (as it is for noble metals), soluble in acid solutions, or soluble in both acid and alkali solutions.

14
I
I
13
70 DEG F - USUAL LABORATORY
12
w
z
/ TEMPERATURE
:::::i \
~

------
<l
II ~
...J
- ......
10
<l
\ ~
I""

---
....... ........
9
w
:::J
...J
-- .---, '",
.....
r-.., r--........ "'u.._
0
<l 8 ...
:::--
10
> N

:I:
Q. 7 NEUTRAL
"-
........
t- ~
..............
6
r-- -.! I

.... 1'- I

5 -
0
............... I

-
u
<l ....
4 'I'-... :
~ I
I
3
o 50 100 150 200 250 350 350 400 450 500
PUMPING TEMPERATURE, DEG F

Fig. 17.6 Effect of temperature on pH values.


A laboratory test for low pH usually involves contact with atmosphere, and as a result,
the value is about 0.5 high.
388 Materials of Construction

Table 17.6 Galvanic Series of


Metals Commonly Used in
Centrifugal Pumps

Corroded end (anodic)

Zinc

Iron, carbon steel


Chrome nickel iron
13 chrome steel (active)
316 & 317 stainless steel (active)

Aluminum bronze
Brasses
Bronzes
13 chrome steel (passive)
Copper-nickel alloys
316 & 317 stainless steel (passive)

Chrome-nickel alloy 20

Graphite

Protected end (cathodic)

Metal Corrosion Mechanisms

Galvanic: One metal in a multimetal system is preferentially corroded. Corrosion is produced by a


classic galvanic cell; a combination of metals far apart in a galvanic series in the presence of a strong
electrolyte. The series in Table 17.6 gives an approximate idea, based on the corrosion potentials in
seawater, of the interrelation of the metals most commonly used in centrifugal pumps. The active states
of 13 chrome steel and types 316 and 317 stainless steel occur in poorly aerated liquids or in oxygen
deficient regions. Graphite's nobility is of consequence to cast iron pumps handling seawater or brackish
water; once the casing is graphitized, the previously cathodic bronze impeller will be anodic to the
casing, thus subject to corrosion.

General: Corrosion of all surfaces, with greatest metal loss in regions of high velocity. Occurs when
the electrochemical potential of the pumped liquid is sufficient to remove the protective passive film,
and is accelerated where high velocity liquid quickly removes the products of corrosion.

Selective leaching: Only part of the material is corroded. Prevalent in metals whose structure is a
matrix of dissimilar materials (e.g., iron and graphite in cast iron), which in seawater will suffer selective
leaching of the iron until the exposed surface is graphite, a process known as graphitization.
Materials of Construction 389

Fig. 17.7 Corrosion-erosion damage to an impeller.

Corrosion-erosion: Rapid metal loss with the fluted appearance of fine erosion. Process whereby mild
general corrosion is accelerated by particles in the pumped liquid or the high liquid velocities inherent
in high head per stage designs. Figure 17.7 shows corrosion-erosion damage to an impeller.

Crevice: Metal is lost from the surfaces forming the crevice. Caused by a concentration cell in which
the liquid in the crevice has a different concentration than the bulk liquid; usually a lower oxygen content
leading to reduced passivation.

Pitting: Pinpoint penetration of the material, at an accelerated rate within the pits. Initiated at local
variations in the metal; accelerated as the pitting becomes deep enough to form concentration cells.

Stress corrosion cracking: Parts develop cracks at low nominal levels of tensile stress. The stress
may be residual or applied or a combination of both. By a mechanism not yet fully understood, a low
concentration of a corrosive element or compound (e.g., chlorine or hydrogen sulphide), in the presence
of water, promotes cracking at low tensile stress levels. The cracking is typically transgranular (Fig. 17.8).
390 Materials of Construction

Fig. 17.8 Photomicrograph of stress erosion cracking.

Corrosion fatigue: Parts subject to cyclic stress quickly fail catastrophically at stress levels below
the nonnal endurance limit of the material, a consequence of active corrosion reducing the material's
endurance limit.

Intergranular: Corrosion at the grain boundaries of austenitic materials. A direct result of the material
being "sensitized" by slow cooling from above the austenitizing temperature, with resultant chrome
depletion at the grain boundaries (fonnation of chrome carbide; Fig. 17.9).

Microbiological: Localized corrosion under microbiological deposits. A fonn of concentration cell


in which bacteria, either aerobic or anaerobic, alter the bulk environment within the colony.

Beyond the 10 metal corrosion mechanisms just cited above, there is one identified as "corrosion"
that appears not to be the result of electrochemical action.

Fretting corrosion: results from loss of metal from one or both of a pair of contacting surfaces.
Caused by minute, high-frequency movement between the surfaces, it occurs in both wet and dry
environments, and in inert gases, which suggests it is not dependent on electrochemical action.
Materials of Construction 391

Fig. 17.9 Photomicrograph of intergranular carbide precipitation.

Nonmetallic Corrosion Mechanisms


Chemical attack: Softening or swelling of hydrocarbon polymers, caused by the action of the pumped
liquid on the polymer's inter and intramolecular bonds.
Stress cracking: Cracking caused by the combined effects of tensile stress, either residual or imposed,
and a "stress cracking agent," a liquid not normally corrosive to the polymer. As is the case with metals,
even quite low concentrations of the stress-cracking agents will promote failure.

EROSION

Pump components can be eroded by either the action of solids in the liquid, a process termed abrasion,
or the action of collapsing vapor bubbles, a process known as cavitation erosion.

Abrasion
Three mechanisms of abrasion are generally recognized, with each having a distinct effect on the
material choice.

1. Cutting or gouging: The solids contact the pump component surfaces at a relatively low angle (Fig. 17.10)
and thereby remove material by gouging or cutting the surface. The rate of material loss is related to the
392 Materials of Construction

Fig. 17.10 Abrasion-cutting or gouging.

relative velocity between the particles and the surface, and the size and shape of the particles. Cutting or
gouging erosion is the usual mechanism in pumps handling nonsettling slurries. Rubber-lined or hard-metal
components are used to achieve the best resistance to cutting or gouging erosion. Hard metal is required
for high-concentration slurries or slurries of sharp solids.
2. Impact: Large, high-mass solids impinging on the component surfaces at high angles produce fatigue failure
and spalling of the surface. Dredge pumps and pumps handling settling slurries typically suffer impact
erosion. To absorb the energy without spalling, their components are made from strong, ductile metals such
as high-manganese steel.
3. Grinding: Irregularly shaped particles get between the surfaces of the close running clearances and in the
process of passing through remove material by grinding. All pumps handling solids laden liquids are subject
to grinding erosion.

Cavitation
The cause of cavitation and its effect on pump performance are discussed in Chapter 19. Cavitation
itself, the formation of vapor bubbles, does not damage pump components. What can cause damage
(Fig. 17.11) is the subsequent collapse of the vapor bubbles (but not the noncondensible gases) as they
pass into regions of higher pressure. The mechanism is thought to be basically fatigue in nature, a
consequence of high pressures and temperatures associated with bubble collapse.
Corrosion has been suggested as a contributing factor, but data are lacking. In a pump handling
corrosive liquids, however, it is quite conceivable that the erosive action of bubble collapse could aid
corrosion, as in corrosion-erosion, and there could be sufficient temperature rise in the region to accelerate
the corrosion rate.

THERMAL EXPANSION AND SHOCK RESISTANCE

Thermal expansion has two consequences on pump components. First, at elevated temperatures, the
increase in the component's size can be sufficient to require special provisions to compensate for the
Materials of Construction 393

Fig. 17.11 Cavitation erosion in the fillet between the hub and vane underside at the suction side of an impeller.

expansion, or the component can be distorted, overstressed, or cause damage to a connected component.
Provision for up to 13 mm (0.5 in.) of movement in the casings of barrel pumps in hydrocarbon charge
service at temperatures of 400 to 425°C (750 to 800 0 P) is a good example. Second, at even quite low
temperatures, a large difference in the thermal expansion rate of materials used in an assembly can cause
what is termed "differential thermal expansion" to render the assembly unserviceable due to either
loosening or tightening of fits. Bronze impellers on a steel shaft is a good example (see impeller materials).
Thermal shock resistance is a function of a material's thermal conductivity, thermal expansion, and
ductility. Materials of low thermal conductivity and low ductility, ceramics, for example, develop a
significant stress gradient when heated rapidly and are therefore prone to fracture. Several of the chrome
steel alloys in the fully hardened state have a similar characteristic. Ductile materials will not fracture
immediately when subjected to thermal shock, but will suffer low cycle thermal fatigue if the rate of
heating is high enough to produce local stresses beyond the material's endurance limit.

STRUCTURAL REQUIREMENTS

Although the first pass at material selection is almost always based on corrosion and erosion resistance,
the structural features of a pump or the manufacture of the part or both may dictate either a compromise
in the material selection or a change in the pump configuration or its manufacture.
Starting with examples of compromises in material selection, the more usual are

1. Multistage pump shafts: made from one of the duplex alloys or other high strength grades instead of type
316 stainless steel to overcome the low strength and difficulty maintaining straightness inherent in type 316.
2 Impellers: steel, chrome steel, or stainless steel is used instead of iron or bronze because the peripheral speed
is too high for both, the pumping temperature is too high for bronze, or the impellers must be mounted with
a shrink fit (see impeller materials).
394 Materials of Construction

(b)

(8) -T T

~
'-O.3T

Fig. 17.12 (a) Transition in casting section. (b) Coring of a double-suction impeller hub.

3. Interference fit impeller wearing rings: steel or chrome steel is used instead of iron or similar materials to
avoid putting a brittle material in tension with the attendant risk of fracture.
4. Shrink-fit-mounted single suction unbalanced impellers: plain back hubs are used regardless of what has
been done to the front hubs, to avoid putting a brittle material in tension and thereby the risk of it spalling.
Adequate galling resistance of the back hub running clearance is provided by increasing the clearance. When
resistance to abrasive wear is necessary, one of the surface conversion processes is used (see wearing parts).

The strength and pressure tightness of metal castings depends a great deal on the relative uniformity of
their cross sections, a fact that is at the heart of many discussions between pump designers and foundrymen.
The more uniform the cross sections, the stronger the casting, and the lower the risk of internal shrinkage
or tearing. Achieving this generally requires a compromise in the form of extra work in the manufacture
of the casting.
The shape of many of the components of centrifugal pumps is determined first by the hydraulic design
and then by the mechanical requirements, and generally the cross sections are not uniform. To avoid·
the ill effects of this, the designer resorts to gradual changes in section by either adding material (Fig.
17.12[a]) or removing it by coring (Fig. 17. 12[bD. The double-suction impeller (Fig. 17.12[bD is an
interesting case, because for hydrocarbon service the hub must be either cast solid or filled with a bigh-
melting-point solid. The reason is that during operation, hydrocarbon will accumulate in the cored space,
and if the impeller is heated for removal (a practice frequently necessary) there is a risk the hydrocarbon
will bum or explode with consequent personnel injury.
Finally, examples of changes in pump configuration or manufacturing method, include the following:

1. Pumps with lined casings often have the liner constrained within a radially split casing (Fig. 17.13), which
provides the structural strength the liner lacks.
Materials of Construction 395

Fig. 17.13 Hard metal slurry pump with "solid" casing liner.
(Courtesy of Lawrence Pumps, Inc.)

2. The impellers and casings of small, mass-produced pumps for water and industrial service are made by spot
welding together stainless steel pressings. This technique is employed because the necessary metal thickness
of the components is lower than can be produced by the economical casting techniques.
3. Critical-service single-stage pumps, such as nuclear reactor primary coolant circulating pumps, have casings
that are symmetrical and made of wrought material to achieve the highest possible mechanical integrity. A
diffuser within the casing achieves the hydraulic function normally carried out by the casing in single-
stage pumps.

LOAD FACTOR AND SERVICE LIFE

It is obvious that a selection of materials to provide the longest possible service life for a temporary
installation would be very uneconomical. Thus, standard fitted pumps are frequently used for services
in which corrosion or erosion will wear a pump out in a relatively short time, if this pump will no longer
be used after the service is performed. The same reasoning applies to installations in which pumps
operate an extremely small percentage of the time, providing that contact with the pumped liquid during
idle periods does not continue the disintegration process, or else that the pump can be drained and
flushed out.
Plain common sense dictates that materials be chosen on the basis of optimum economic life, that
is, for an initial cost and a cost of part replacement (including the necessary labor) that will yield the
lowest overall total investment during the expected life of the equipment. The materials chosen, therefore,
may often be neither the cheapest nor the most expensive available. If outstanding reliability is desired,
on the other hand, the best materials are none too good, even if a pump is to operate once every 10
years. Although operation ofthe centrifugal pumps on board ships of the U.S. Navy is relatively infrequent
396 Materials of Construction

in peacetime, for example, and not constant even in war, the most rigid material specifications are
enforced because failure of any part of the equipment may prove fatal.

MATERIAL CHOICE AND MATERIALS ENGINEERING PROGRESS

Advances in materials engineering of both metals and nonmetals have had a marked effect on the design
of centrifugal pumps for pumping chemicals, hydrocarbons, corrosive waters (including boiler-feed
water), and slurries.
Looking at chemical services first, there has been a gradual evolution of the stainless steels and the
so-called "high alloys" to achieve better performance. The range of available metals has been extended
to include titanium and zirconium. At the same time, progress has been made with reinforced polymers
such that pumps of high corrosion resistance and adequate pressure containment capability are now
made of polymer alone.
As the quality of crude oil reserves has deteriorated over the past 20 or so years, refiners have been
faced with hydrogen sulphide stress corrosion cracking (of parts in tension), and general corrosion caused
by naphthenic acid. In 1975 NACE [17.2], published guidelines recommending various limits to the
hardness and strength of the commonly used materials to avoid hydrogen stress corrosion cracking. API-
610 has distilled those recommendations into requirements for refinery pump construction. Resistance
to naphthenic acid corrosion is achieved with either one of the duplex alloys, CD4MCu with nitrogen
added being typical, or C08M, the cast version of type 317 stainless steel. Duplex alloys are preferred
wherever possible because their coefficient of thermal expansion is close to that of steel. When CO 8M
is necessary, usually at high temperatures, care is needed in the pump's detail design to compensate for
differential thermal expansion where it occurs.
The term "corrosive waters" covers boiler feed through seawater and brackish water, the last two
being pumped by the oil industry for water injection, a process used in tertiary oil recovery. Boiler-feed
water is discussed first.
Most of the knowledge on boiler-feed pump materials was developed 50 years ago, and is largely
still valid today. The principal development was made in 1944 based on the findings of an investigation
conducted by the Boiler Auxiliary Subcommittee of the Prime Movers Committee of the Edison Electric
Institute. This investigation was initiated to determine the cause of the rapid pump deterioration being
experienced in many high-pressure power plants, a condition that had reached alarming proportions in
the early 1940s. Among the findings made by this subcommittee was the fact that steels containing 5
percent chromium or more (today known as chrome steels) were immune to corrosion by any boiler-
feed water then known.
Table 17.7 is a current general guide to the selection of boiler-feed pump materials. It is more
complicated than earlier guides in that it requires careful attention to feed-water condition. Note that 13
percent chrome steel is recommended, with CA-6NM (the 13 chrome, 4 nickel alloy) being the most
widely used today. Although 5 percent chrome is sufficient to provide immunity to corrosion-erosion
by boiler-feed water, the usual alloy, designated C-5, has poor castability and is difficult to weld. Note,
too, that when austenitic stainless steel is recommended, the whole pump is to be of that material to
avoid difficulties with differential thermal expansion.
Pumping seawater and brackish water at high heads per stage in water-injection service precipitated
the development of several duplex (austenitic ferritic) alloys to overcome the problems of corrosion and
low strength inherent in the austenitic stainless steels used originally, particularly when subjected to
high liquid velocities and the stresses associated with high rotative speeds. The more notable are
Ferralium™ 255 (UNS-S32550) and Zeron™ 100 (UNS-J93380). Ferralium 255 castings are not covered
by an ASTM specification, but are produced by various foundries in the United States and Europe. Zeron
Materials of Construction 397

Table 17.7 Boiler Feed Pump Materials

Temperature O2 Conductivity Materials'


(micromohs/cm)
------~

°C (OF) pH (ppm) I-I 1-2 S-1 S-6 C-6 A-8

:595 (200) 0-14 ~


4.5-14 ~ ~
6-9 ~b ~ ~ ~
6-14 ~ ~ ~
9-14 ~b ~ ~ ~

>95 (200) 0-14 >0.04 :520 ~


4.5-14 ~ ~

>95 (200) 0-14 :50.04 :520 ~


4.5-14 ~ ~

>95 (200) 0-14 :50.04 >20 ~


4.5-14 ~ ~
6-9 ~b ~ ~ ~
6-14 ~ ~ ~
9-14 ~b ~b

, Many chemistries are marginal; avoid selecting less expensive materials when C-6 is indicated. The code for these materials is

Code (API-610) Casing, impeller

1-1 All cast iron


1-2 Cast iron, bronze
S-1 Cast steel, cast iron
S-6 Cast steel, 13 chrome
C-6 All 13 chrome
A-8 All CF-3M

b Head less than 120 m (400 ft) per stage.

100 castings are available as ASTM A890 Grade 6A. Manufacturing methods also took an interesting
tum in some cases, one instance being impellers produced by electrodischarge machining (EDM) from
Ferralium 255 forgings.
Process designers have presented slurry pump manufacturers with slurries of a corrosive carrier liquid
or at temperatures above the operating limit of rubber or both. The result has been a series of high-
chrome hard irons, up to 35 percent chrome, with better corrosion resistance than the lower chrome
alloys. For those applications where hard metal did not have sufficient corrosion resistance, and the
temperature exceeded the limit of rubber, pumps with a ceramic impeller and ceramic-lined casing have
been produced. Silicon nitride bonded silicon carbide is one such material that has been used.
Materials engineering has made substantial progress over the past 20 years, much of it a direct result
of improving the performance of gas turbines. Applicable segments of this work are now gradually being
adopted by the pump industry. One worthy example is the use of hard coatings and surface-hardening
techniques to increase the wear resistance of running clearance surfaces while retaining the necessary
bulk structural properties of the part. This can also lead to simpler pumps. Impellers with hard-coated
hubs, as an example, do not require the added complexity of wearing rings.
Ceramic antifriction bearings running in the pumped liquid have been used for some time in pumps
for cryogenic services such as liquefied natural gas. Similarly, progress is being made in the development
398 Materials of Construction

of high-chrome corrosion-resistant antifriction bearings, raising the possibility of one day being able to
run such bearings in water and similar liquids. Product-lubricated hydrodynamic bearings, typically of
silicon carbide, are now in the proving phase.
Much of the current materials engineering effort is being devoted to ceramics. A new generation of
these materials based on zirconium holds the promise of ceramics with high strength and toughness,
characteristics that will allow their extensive use in tomorrow's centrifugal pumps.

BIBLIOGRAPHY

[17.1] Corrosion Engineers Reference Book, [2nd Edition, 1980] National Association of Chemical Engineers,
Houston, Texas.
[17.2] NACE MR-01-75, Sulfide Stress Corrosion Resistant Metallic Material for Oil Field Equipment, 1975,
National Association of Chemical Engineers, Houston, Texas.
II
PUMP PERFORMANCE
18
Heads, Conditions of Service
Performance Characteristics,
and Specific Speed

In selecting the most suitable centrifugal pump for a given application, the most important information
to be given the manufacturer is the desired capacity and the head against which the pump will be required
to operate while delivering the specified rate of flow.

UNITS OF CAPACITY

The standard unit of capacity for centrifugal pumps varies with the application of the pump as well as
the design standards of the country where the pump is used-gallons per minute in the United States,
occasionally imperial gallons per minute in countries of the former British Commonwealth, and cubic
meters per hour in countries using the metric system. In the United States, units vary with the pump
application as follows: million gallons per day, cubic feet per second, gallons per hour, barrels per day,
barrels per hour, pounds per hour, and acre feet per day. Common equivalent variations in the metric
system are: liters per second, cubic meters per second, and metric tons (tonnes) per hour.
It is a simple matter to convert the various units into gallons per minute (gpm). The equivalents for
most units are incorporated in Table 18.1. For a direct conversion chart, see Fig. 18.1.
The pump capacity required by an installation should be stated in cubic meters per hour (gallons per
minute) at the pumping temperature; any desired or imposed variation in the range of capacities should
also be clearly stated. The proper method of specifying required capacity in preparing an inquiry for
centrifugal pumps is discussed in some detail in Chapter 27.

HEADS

Pumping is the addition of kinetic and potential energy to a liquid for the purpose of moving it from
one point to another. This energy will cause the liquid to do work, such as flow through a pipe or rise
to a higher level. A centrifugal pump transforms mechanical energy from a rotating impeller into the
kinetic and potential energy required. Although the centrifugal force developed depends on both the

401

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
402 Heads, Conditions of Service Performance Characteristics, and Specific Speed

Table IS.1 Capacity Equivalents


For conversion chart, see Fig. 18.1

Various units gpm

1 second-foot or cubic foot per second (cfs) 448.8


1,000,000 gallons per day (mgd) 694.4
1 imperial gallon per minute 1.201
1,000,000 imperial gallons per day 834.0
1 barrel (42 gal) per day (bbVday) 0.0292
1 barrel per hour (bbl/hr) 0.700
1 acre-foot per day 226.3
1,000 pounds per hour (lb/hr) 2.00'
1 cubic meter per hour (m3/hr) 4.403
1 liter per second (lIs) 15.851
1 metric ton per hour 4.403'
1,000,000 liters per day = 1,000 cubic meters per day 183.5
IThese equivalents are based on a specific gravity of I for water at 62°F for
English units and a specific gravity of I for water at 15°C for metric units.
They can be used with little error for cold water of any temperature between
32°F and 80°F. For specific gravity of water at various temperatures, see
Fig. 18.6.

peripheral speed of the impeller and the density of the fluid, the amount of energy imparted per pound
of fluid is independent of the fluid itself. Therefore, for a given machine operating at a certain speed
and handling a definite volume, the mechanical energy applied and transferred to the fluid-joules per
kilogram of fluid or foot-pounds per pound of fluid-is the same for any fluid, regardless of density.
The pump head, or energy in joules per kilogram (foot-pounds per pound), will therefore be expressed
in meters (feet). Barring viscosity effects, the head generated by a given pump at a certain speed and
capacity will remain constant for all fluids. Thus, it is natural to speak. of heads in centrifugal pumps in
terms of meters (feet) of liquid.
Before discussing the various head terms involved in pumping systems it should be mentioned that
(1) heads can be measured in various units, such as feet of liquid, pounds per square inch of pressure,
inches of mercury, and others depending upon the application and the units of measurement of the
country; (2) pressures and head readings can be in gage or absolute units; (3) the difference between
gage and absolute units is affected by the existing atmospheric pressure and thus by the altitude; and
(4) the pressure at any point in a system handling liquids must never be reduced below the vapor pressure
of the liquid.

Conversion of Pressure to Static Head


A column of cold water approximately 10.2 m high will produce a pressure of 1 bar at its base (as
will 2.31 ft produce 1 psi). Thus for water at ordinary temperatures, any pressure can be converted to
an equivalent head by multiplying by 10.2 for bars to meters or 2.31 for psi to feet. For liquids other
than water, the column of liquid equivalent to the basic unit of pressure (1 bar or 1 psi) can be calculated
by dividing the constant (l0.2 or 2.31) by the specific gravity of the liquid.
Figure 18.2 illustrates the effect that specific gravity has on the height of a column of various liquids
for equal pressures. Thus a pump that must handle 1.2 specific gravity brine against 6.90 bar (lOO-psi)
net pressure would be designed for a head of 58.8 m (193 ft). If the pump had to handle cold water
against the same net pressure, the head would have to be 70.4 m (231 ft), whereas a pump handling
Heads. Conditions of Service Performance Characteristics. and Specific Speed 403

CAPACITY ,101 GAllONS (Y S) Pel MlNYTE

I!!
:>
2._ ~
:>

..
Z Z
i i
~ 1,_
'"Z
2
g
.
~ ~
CJ
"
:>
:>
z

..
Z
« «
~ ~
r!I
~ ~
0 0
'"z ~
:> ~
J; J;
t t
~ "«
3 3

CAPACITY 'N GAllONS IU S) 'U ..'Nun;

Fig. 18.1 Capacity conversions.


For more accurate values, calculate from the equivalents shown in Table lS.l.
404 Heads. Conditions of Service Performance Characteristics. and Specific Speed

GASOLINE
(SPECIFIC

---l
GRAVITY 0.75)

WATER
(SPECIFIC
GRAVITY 1,0)
BRINE

---1
(SPECIFIC
GRAVITY 1.2)
-
-

Fig. 18.2 Effect of fluid density on static head.


Comparison of the heights of a column of water. brine. and gasoline needed to
produce 6.9 bar (100 psig) pressure at datum level.

0.75 specific gravity liquid against the same net 6.90 bar (1oo-psi) pressure would require a head of
96.0 m (308 ft).
It is obvious that a pump designed to handle water but applied on brine service would develop a
70.4 m (231-ft) head of brine or 8.28 bar (120-psi) pressure while if it was applied to pump 0.75 specific
gravity gasoline it would develop a 70.4 m (231-ft) head of gasoline or only 5.17 bar (75-psi) pressure.
The equivalents for the conversion of various pressure and head units other than feet into feet of
liquid are indicated in Table 18.2. For quick conversion of pressures and heads into feet of liquid, see
Fig. 18.3.

Gage and Absolute Units


Pressures and their corresponding heads can be expressed either in absolute units or gage units. In
the metric system, pressures are gage unless noted, for instance, 6.90 bar is gage pressure, or 11.0 bar
(abs) is absolute. The same pressures are expressed as 100 psig or 160 psia, respectively. In gage readings
the pressure is given merely in relation to the atmospheric pressure, whereas absolute pressures are gage
readings plus the existing atmospheric pressure. In other words, the pressure is referred to an absolute
vacuum (Fig. 18.4).
To illustrate, assume a person standing part way up a hill, 7.60 m (25 ft) from the bottom. The
elevation or level at which he is standing would be his gage basis of measurement. Points below him
would be negative (-) gage elevations, and points above him would be positive (+) gage elevations.
Thus he would speak of a point 30.5 m (100 ft) up as 30.5 m (100 ft) gage elevation or of one 3.05 m
(10 ft) down as 3.05 m (10 ft) below gage level, corresponding to a vacuum in our problem. If he desired
Heads, Conditions of Service Performance Characteristics, and Specific Speed 405

Table 18.2 Pressure and Head Equivalents


For conversion chart, see Fig. 18.3

Ib/sq in. 2.310 ft of liquid = 2.310 ft of 62°F water


specific gravity'
1 in. mercury (32°F) = 1.134 ft of liquid = 1.134 ft of 62°F water
specific gravity'

atmosphere2 33.95 ft of liquid = 33.95 ft of 62°F water


specific gravity'
kilogram/sq cm = 1 metric atmosphere

32.85 ft of liquid = 32.85 ft of 62°F water


specific gravity'
10.01 m of liquid = 10.01 m of 15°C water
specific gravity'
1 bar 33.51 ft of liquid = 33.51 ft of 62°F water
specific gravity'
10.21 m of liquid = 10.21 m of 15°C water
specific gravity'
1 meter = 3.281 ft
'These equivalents are based on a specific gravity of I for water at 62°F for English units
and a specific gravity of I for water at 15°C for metric units. They can be used, with little
error, for cold water of any temperature between 32°F and 80°F. For the actual specific
gravity of water for temperatures to 220°F, see Fig. 18.6.
2Not used in conjunction with pumps.

to express an elevation measured from the bottom of the hill (absolute datum level, corresponding to
zero absolute pressure or a perfect vacuum), he would add 7.6 m (25 ft) to his gage reading so that the
point 30.5 m (100 ft) above him would be 30.5 + 7.6, or 38.1 m (100 + 25, or 125 ft) above the bottom
of the hill (38.1 m or 125 ft absolute elevation), whereas the point 3.05 m (10 ft) below him would be
-3.05 + 7.60, or 4.55 m (-10 + 25, or 15 ft) above the bottom of the hill (4.55 mor 15 ft absolute elevation).
It is usually feasible to work in terms of gage pressure, but a complicated problem can occasionally
be clarified by working entirely in terms of absolute pressure.

Effect of Altitude on Atmospheric Pressure


For pumps installed at elevations above sea level, it must be remembered that there is a decrease in
atmospheric pressure of about 83 mm of mercury per 1,000 m (1 in per 1,000 ft) of elevation. At an
elevation of 1,220 m (4,000 ft) therefore, the atmospheric pressure is 101 mm of mercury or about 1.4
m of water (4 in of mercury or about 4.5 ft of water) less than that at sea level, with the result that a
centrifugal pump will operate satisfactorily for the same maximum capacities only if the suction lift is
1.4 m (4.5 ft) less than that at sea level. This effect should not, however, lead to the confused notion
that the net positive suction head required for a pump changes with elevation above sea level. It does
not, but the available atmospheric pressure is reduced. For barometric pressures at various altitudes, see
Fig. 18.5.
406 Heads, Conditions of Service Performance Characteristics, and Specific Speed

HEAD IN FEET OF COLD WATER


2 3 4 5 6 8 10 20 30 .a 60 80100 200 300
100 100
80 80

60
SO SO
.a 40

30 30
III III
w
~
« 20 ~
~ ~
0 0
0v", 0V IX
... W
0:" 10 10 ot
"'!
IX __
8 8 "'!
IX--
~z
.........
~v 6 6 --wZ
~v ...
5 5 w
-~
:1:« -'"
:1:«
v:::> 4 4 v:::>
~~ ~~
w
'"
~. 3 3 IX
«w IX
:::>~ :::>Go

:~
"'IX
Got:)
2 2 ~'"
"'~
'" IX
Got:)
:.:0 :.:0
"'=
:I~
v IX 1.0
IX~
:::>:w:
1.0 ~'"
:0 ... 0
:IE 0.8 0.8 :e
...0 0.6 0
...
0.6
'"...
%
0.5 0.5 ~
:I:

~ 0.4 0.4 ~
0.3 0.3

0.2 0.2

0.1 0.1
2 3 4 5 300
HEAD IN FEET OF COlD WATER

Fig. 18.3 Pressure and head conversion chart.


Values are plotted for 18.7°C (62°F) water but can be used for water between DoC (32°F) and 26.7°C (80°F).
For liquids other than cold water, divide the head by the specific gravity (I8.7°C (62°F] water = 1.0) of the liq-
uid at the pumping temperature to get the head in meters (feet). For more accurate values, calculate heads from
the head equivalents in Table 18.2 .
ANY PRESSURE ABOVE ATMOSPHERIC

ABSOLUTE PRESSURE- GAGE


1
GAGE PRESSURE + PRESSURE
BAROMETRIC PRESSURE

VACUUM
(A NEGATIVE
GAGE PRESSURE)

ANY PRESSURE BELOW ATMOSPHERIC

BAROMETRIC
PRESSllRE
ABSOLUTE
PRESSURE

Fig. 18.4 Graphical illustration of atmospheric, gage, and absolute pressures.

215 t:
.......... '"zr
....... 210
'"r
so
r-.... -<:: ...... TEMPERATURE OF
BOILING WATER
205
II<

it
..........
'\. ......... 200 '"'"
-
29
\.. ......... '"
co
21 """""= ....... 115 "'"0
-~ ......... .... 190 ~
27 ......
2S '\. co
...«:>
>- ;
co ~ II<

""-
:>
u
co
25
'":I
Q.

...'"
...'"0:I
24

23
'"'"r
ATMOSPHERIC
PRESSURE i'
""
u 22
~
21

20 ~

19 ~
II
o 2:5 4 5 • T • • 10 1/ 12

ELEVATION ABOVE SEA LEVEL,IN \p00 FEET

Fig. 18.5 Atmospheric pressures for altitudes up to 3,660 m (12,000 ft).

407
408 Heads. Conditions of Service Peiformance Characteristics. and Specific Speed

g
S
IS"
l-
1099
100 20

18
- ...... ~ .!PfClf'O;C I
II. -i"'::~ry J
~ 0.98 16
>= 0.91
j"ooo.
~ I
I-
14
...... ~
~ 0.96 12
~ 095 10 / 7-
Cl
~ 8 -~ ~/ J
~ 6 l/ ~~V
::l
~ 4
~,;' ~?o~'-
LLJ
f 2 POR PRESSURE, ~ ~~~ ~
"A -r"1 I " Of ~~
I'1APORP~
1 IRES~URE'~
40 60 80 100 120 140 160 180 200
TEMPERATURE, DEGREES FAHRFNHEIT

Fig. 18.6 Specific gravity, temperature. and vapor pressure relations for water.

Vapor Pressure
The vapor pressure of a liquid at a given temperature is that pressure at which it will flash into vapor
if heat is added to the liquid or, conversely, that pressure at which vapor at the given temperature will
condense into liquid if heat is subtracted.
For homogeneous or single-component liquids, such as water, the vapor pressure has a very definite
value at any given temperature, and tables (such as steam tables) are available that give the vapor
pressure of such liquids over a wide range of temperatures (see Table 26.1). Certain mixed liquids,
however, such as gasoline, are made up of several components, each having its own vapor pressure, and
partial vaporization may take place at various pressures and temperatures.
In figuring heads for pumps, it is important that pressures expressed in bar (pounds per square inch)
or other pressure units be converted into meters (feet) of liquid at the pumping temperature. Care must
be taken not to use conversion factors applying to other temperatures for such conversions. For example,
the vapor pressure of 100°C (212°F) water is 1.01 bar (abs) or 14.7 psia (standard barometric pressure
at sea level). The equivalent head in meters (feet) of water is 10.34 m (33.9 ft) of 15°C (62°F) water.
As 100°C (212°F) water has a specific gravity of 0.959 compared to a gravity of 1.0 for 15°C (62°F)
water, its equivalent head would be 10.34/0.959, or 10.78 m (33.9/0.959, or 35.1 ft) (Fig. 18.6).

Head Terms
In its elementary form, "head" denotes the distance at which the free surface of a body of water lies
above some datum line; as such, it represents an energy or ability to do work. Energy can also exist as
a pressure. Some consider that static head is the sum of the pressure head and the static head of elevation;
Heads, Conditions of Service Performance Characteristics, and Specific Speed 409

however, these two factors are generally considered separately. In any pumping system, the liquid must
be moved through pipes or conduits that offer certain resistances or, in other words, cause certain
frictional losses. This energy dissipation, or head loss, is called a frictional head whereas the energy
that has been converted into velocity energy is called velocity head. Thus, static heads, pressure heads,
friction heads, and velocity heads may all be encountered in any system. When considering a pump by
itself, "head" is a measure of the total energy imparted to the liquid at a certain operating speed
and capacity.

SYSTEM HEAD

The total head of a system against which a pump must operate is made up of the following components:

1. Static head
2. Difference in pressures existing on the liquid
3. Friction head
4. Entrance and exit losses
5. Velocity head.

Static Head
Static head refers to a difference in elevation. Thus the "total static head" of a system is the difference
in elevation between the discharge liquid level and the suction liquid level (Figs. 18.7 to 18.9). The
"static discharge head" is the difference in elevation between the discharge liquid level and the centerline
of the pump. The "static suction head" is the difference in elevation between the suction liquid level
and the centerline of the pump. If the static suction head is a negative value because the suction liquid
level is below the pump centerline, it is usually spoken of as a "static suction lift."
If either the suction or discharge liquid level is under a pressure other than atmospheric, this pressure
is sometimes considered as part of the static head, but it is often considered separately. The latter practice
usually permits a clearer picture of the system. If the suction supply is taken from a closed vessel and
the liquid level lies above the pump centerline, the difference in elevation of the suction liquid level
and the pump centerline is commonly spoken of as "submergence" instead of "static suction head."

Friction Head
Friction head is the equivalent head, expressed in meters (feet) of the liquid pumped, that is necessary
to overcome the friction losses caused by the flow of the liquid through the piping, including all the
fittings. The friction head varies with (1) the quantity of flow; (2) the size, type, and condition of the
piping and fittings; and (3) the character of the liquid pumped.

Entrance and Exit Losses


Unless it comes from a main under pressure, such as a city water supply, the suction supply of a
pump comes from some form of reservoir or intake chamber. The point of connection of the suction
pipe to the wall of the intake chamber or the end of the suction pipe projecting into the intake chamber
or reservoir is called the entrance of the suction pipe. The frictional loss at this point is called the
"entrance loss." The magnitude of this loss depends on the design of the pipe entrance, a well-designed
bellmouth providing the lowest possible loss.
Similarly, on the discharge side of the system where the discharge line terminates at some body of
liquid, the end of the piping is called the exit. This exit is usually of the same size as the piping, and
410 Heads, Conditions of Service Performance Characteristics, and Specific Speed

TOTAL
STATIC
HEAD

STATIC
STATIC DISCHARGE
SUCTION HEAD

~
HEAD

Fig. 18.7 Static heads.


System with pump suction and discharge vessels at pressures other than atmospheric.

the velocity head of the liquid is entirely lost. The end of the discharge piping is sometimes a long taper
so that the velocity can be effectively reduced and the energy recovered.
Some engineers consider entrance and exit losses as part of the suction and discharge pipe friction
losses. Others prefer to consider them separately to make sure that they are not overlooked. This method
has the additional advantage of clearly showing if either or both losses are excessive.

Velocity Head
Velocity head is the kinetic energy in a liquid at any point, expressed in joules per kilogram (foot-
pounds per pound) of liquid, that is, in meters (feet) of the liquid in question. If the liquid is moving at
a given velocity, the velocity head is equivalent to the distance the mass of water would have to fall in
order to attain this velocity. Thus velocity head can be calculated by the equation:

V2
h=-
v 2g

where hv = the velocity head, in feet


V = the liquid velocity, in feet per second
g = the acceleration due to gravity, or 9.81 rn/s 2 (32.2 ftls 2).

In determining the head existing in a pipe at any point, it is necessary to add the velocity head to
the pressure gage reading, for the pressure gage can indicate only the pressure energy, whereas the actual
Heads, Conditions of Service Perforrrumce Characteristics, and Specific Speed 411

-----~

TOTAL
STATIC
HEAD

~-,L
STATIC
DISCHARGE
HEAD

STATIC
SUCTION
HEAD

Fig. 18.8 Static heads.


Suction and discharge vessels at atmospheric pressure and suction liquid level above pump centerline.

TOTAL
STATIC STATIC
HEAD DISCHARGE
HEAD

ST1TIC
SUCTION
-
cQ.l
- -
--.-- ~.
LIFT
j

-
I......

Fig. 18.9 Static heads.


Suction and discharge vessels at atmospheric pressure and suction liquid level below pump centerline.
412 Heads, Conditions of Service Performance Characteristics, and Specific Speed

head is the sum of the kinetic (velocity) and potential (pressure) energies. Thus, to detennine the actual
suction head or discharge head, it is necessary to add the velocity head to the gage reading.
If the suction and discharge pressures of a centrifugal pump are taken at points at which the velocities
are the same, the velocity head component of each will be the same. The kinetic energy components of
both the suction head and the discharge head will also be equal, and the total head can be determined
by subtracting the suction gage reading from the discharge gage reading.
In high-head pumps, the kinetic energy is relatively small, but in low-head pumps, it is relatively
high. Thus failure to consider the velocity head in detennining heads in high-head pumps will not
appreciably affect the results. For example, consider a pump handling 34Om3/hr (1,500 gpm) with a 6-
in. discharge and 8-in. suction. The discharge velocity head is 1.37 m (4.5 ft), whereas the suction
velocity head is 0.43 m (1.4 ft). If the suction gage showed 2.62 m (8.6-ft) pressure and the discharge
gage showed a 32.17 m (105.5 ft) head, the true total head would be (32.17 + 1.37) less (2.62 + 0.43)
or 30.49 m [(105.5 + 4.5) less (8.6 + 1.4), or 100 ft], whereas the difference in gage readings would be
29.55 m (96.9 ft). Thus the error would be 3.1 percent of the total head. Had this been a pump in which
the discharge gage reading was 305 m (1,000 ft), the true total head would be 303.2 m (994.5 ft), whereas
the difference in gage readings would be 302.3 m (991.4 ft). The error of 0.3 percent is too small to be
of any concern. If this were a pump in which the discharge head was 13.87 m (45.5 ft), however, the
true total head would be 12.20 m (40 ft), whereas the difference in gage readings would be 11.25 m
(36.9 ft), for an error of 7.8 percent.
Whether or not the velocity head can be ignored depends on the desired accuracy of head detennination
and upon the accuracy of the pressure readings that can be made. For the cited 305 m (I,OOO-ft) head
reading, even with an accurate large scale gage it would be impossible for anyone to read the pressure
within 3.05 m (10 ft), a basic error of 1 percent.

SYSTEM FRICTION CURVE

The friction-head loss in a system of pipes, valves, and fittings varies as a function (roughly as the
square) of the capacity flow through the system. For the solution of pumping problems, it is often
convenient to show the relation between capacity and friction-head loss through the system graphically.
The resulting curve is called the "system friction curve," as shown in Fig. 18.10. The detenninations of
friction losses are usually rough approximations at best, for the roughness of the pipe is not known. As
the friction loss will increase when the pipe tuberculates or otherwise deteriorates with age, it is usual
to base the friction loss on constants that have been found from the average of pipe 10 or 15 years old,
thus allowing for friction losses in excess of those that will be obtained when the pipe is new. As a
result, the pump is generally designed for excess head and delivers overcapacity when installed in a
new system or in one that has not suffered from pipe deterioration. (For a complete treatment of friction
loss calculations, see Chap. 20).

SYSTEM·HEAD CURVE

The friction-head losses, pressure differences, and static heads of any system can be graphically related
(Fig. 18.11). The resulting curve is called the "system-head curve." For systems with varying static
heads or pressure differences, it is possible to construct curves for minimum and maximum static heads
or pressure differentials. The capacity that a pump will be able to deliver under varying conditions can
be predicted by superimposing such system-head curves on a pump head-capacity curve (see Fig. 27.3).
Heads. Conditions of Service Performance Characteristics. and Specific Speed 413

SYSTEM FRICTION CURVE

CAPACITY

Fig. 18.10 System-friction curve.

SYSTEM-HEAD CURVE

o
c FRICTION
'"
% LOSSES

TOTAL
STATIC
HEAD

CAPACITY

Fig. 18.11 System-head curve.

Definitions
Explanation of the head terms used with centrifugal pumps should be applicable to all installations
although one or more elements of the total head are usually not involved (because they have zero
414 Heads. Conditions of Service Performance Characteristics. and Specific Speed

values). Except as otherwise noted, the definitions given here are based on the current Standards of the
Hydraulic Institute.

SUCTION HEAD AND SUCTION LIFT

As now defined, the total suction head (h s) is the static head on the pump suction line above the pump
centerline minus all friction head losses for the capacity being considered (including entrance loss in
the suction piping) plus any pressure (a vacuum being a negative pressure) existing in the suction supply.
Rather than express the suction head as a negative value, the term "suction lift" is normally used when
the suction head is negative and when the pump takes its suction from an open tank under atmospheric
pressure. As the suction lift is a negative suction head measured below atmospheric pressure, the total
suction lift (symbol also hs ) is the sum of the static suction lift measured to the pump centerline and the
friction head losses as defined above. (It is sometimes advantageous to express both suction and discharge
heads in absolute pressure, but usually it is more suitable to measure them above or below atmospheric
pressure.) A gage on the suction line to a pump, when corrected to the pump centerline, measures the
total suction head above atmospheric pressure minus the velocity head at the point of attachment. As
suction lift is a negative suction head, a vacuum gage will indicate the sum of the total suction lift and
velocity head at the point of attachment.
The three most common suction supply conditions are illustrated in Fig. 18.12.
System I involves a suction supply under a pressure other than atmospheric and located above pump
centerline; it includes all the components of suction head (h s). If hs is to be expressed as a gage reading
and P s is a partial vacuum, the vacuum expressed in feet of liquid would constitute a negative pressure

II rn

I
B
hs=<-Sl-hfs-h j
h. = S - h,s - hi + p.
- hs = S + his + hi

Fig. 18.12 Suction head determination for three typical examples.


KEY:
P, = pressure other than atmospheric
S = static head
hs = suction head
hI' = total friction loss from A to B
hi = entrance loss at A
-hs = suction lift
The gage reading at B corrected to pump centerline equals the suction head minus velocity head at B.
Heads. Conditions of Service Performance Characteristics. and Specific Speed 415

head and carry a minus (-) sign. If the pressure P s is expressed in absolute pressure values, hs will also
be in absolute pressure values.
A very common installation, II, involves a suction supply under atmospheric pressure located above
the pump centerline. As the suction head (expressed as a gage value) has a P s value of zero, the P s value
can be dropped from the formula.
System III, the most common installation for pumps handling water, involves a suction supply under
atmospheric pressure located below the pump centerline. It is optional whether the suction head is
expressed as a negative suction head or in positive values as a suction lift. As the source of supply is
below the pump centerline (which is the datum line), S is a negative value. It should be noted that the
suction lift formula is the same as that for suction head except that both sides have been multiplied by
-1. A gage attached to the pump suction flange, when corrected to the pump centerline, will register a
partial vacuum or negative pressure. To determine the suction head, it is therefore necessary to add the
velocity head to this negative pressure algebraically, or, if it is desired to work in terms of a vacuum,
the velocity head must be subtracted from the vacuum to obtain the suction lift. For example, if the
gage attached to the suction of a pump having a 6-in. suction and pumping at a capacity of 227 m3/hr
(1,000 gpm) of cold water showed a vacuum of 152 mm (6 in) of mercury (equal to 2.07 m or 6.8 ft
of water), the velocity head at the gage attachment would be 0.61 m (2.0 ft) of water, and the suction
head would be -2.07 + 0.61, or -1.46 m (-6.8 + 2.0, or -4.8 ft of water), or the suction lift would be
2.07 - 0.61, or 1.46 m (6.8 - 2.0, or 4.8 ft of water).
As most centrifugal pump troubles occur on the suction side of the pump, it is a very important part
of pump selection to supply complete information on suction conditions, including all operational
variations. For some complex problems, it is often necessary to superimpose the variation in total suction
head graphically on the suction head limitations of the pump being considered in order to make sure
the pump will be suitable.

NPSH

In the pumping of liquids, the pressure at any point in the suction line must never be reduced to the
vapor pressure of the liquid. The available energy that can be utilized to get the liquid through the
suction piping and suction waterway of the pump into the impeller is thus the total suction head less
the vapor pressure of the liquid at the pumping temperature. The available head-measured at the suction
opening of the pump-has been named "net positive suction head." It is usually indicated by its initials,
NPSH. A complete discussion of Suction Conditions and of the Limitations on Suction Performance is
given in Chapter 19.

Specifying Suction Conditions

The importance of accurately advising a manufacturer of the actual suction conditions for a centrifugal
pump cannot be overemphasized. A pump will be unable to meet its design capacity conditions unless
the suction head can provide enough energy to get the liquid into the pump as previously discussed. If
a cold nonvolatile liquid is to be handled, it is necessary to know whether there will be suction head or
suction lift, and if the latter, what maximum lift can be expected. If the liquid is to be hot or under a
pressure corresponding to or near its vapor pressure, the pump must be installed with head on suction,
and the available submergence must be indicated. For liquids other than water, information on the
pumping temperature and vapor pressure is also necessary. All expected or probable variations in suction
conditions should also be specified.
416 Heads, Conditions of Service Performance Characteristics, and Specific Speed

I
S
-13.0 FT 20.4FT

S
20.4 FT

I n m v
Fig. 18.13 Seven installations of duplicate pumps referred to in Table 18.3.
Pumps have 8 in discharge and 10 in suction and all operate at 454 m 31hr (2,000 gpm), 24.4 m (80 ft) total
head, and 5.6 m (18.4 fl) NPSH.

Examples of NPSH Calculations


The NPSH calculations of seven different installations, with the same hydraulic conditions (gallons
per minute, total head, and NPSH), are shown in Fig. 18.13 and Table 18.3. This illustration and table
show the effect on the physical installation of a pump, for various applications, if the same NPSH is to
be available at the pump.
Installations I and II have the same conditions except for altitude. The reduction in barometric pressure
at 1,220 m (4,000 ft) elevation makes it necessary to raise the liquid level 1.46 m (4.8 ft) lower to obtain
the same NPSH.
Installation III illustrates how a liquid with a high vapor pressure forces a reduction in possible suction
lift. In this case the liquid is considerably lighter than water, and, if it had a vapor pressure equal to
water at 15°C (60°F), it would have been possible to have a static suction lift of 7.68 m (25.2 ft). The
effect of the specific gravity of the liquid on suction conditions is illustrated more clearly in installation
IV. Here, because brine has a specific gravity of 1.2, the atmospheric pressure corresponds to only
6.83 m (28.3 ft) of liquid instead of 10.34 m (34.0 ft) as with cold water (in installation I). As a result,
in order to obtain the same 5.61 m (18.4 ft) of NPSH, the value of S, the static component of the suction
head, can only be -2.38 m (-7.8 ft) instead of -3.96 (-13.0 ft) as with cold water. (Usually in installations
handling gasoline and brine, the pumps are located so that the liquid level is above the pump. Many
such installations have long suction pipes with considerable friction head loss, so that the same suction
conditions indicated here could logically result.)
Installation V is similar to many boiler feed pump installations. When the liquid handled is at a
temperature corresponding to its boiling point at the suction pressure, the suction head available to
overcome friction and provide the required NPSH must be entirely static as Ps - P vp is zero. This is also
demonstrated in installation VI, which shows a typical condition encountered in condensate or hotwell
pumps. Condensate pumps serving surface condensers are generally located on the floor, just slightly
below the liquid level in the hotwell, and pumps of a special design requiring a very low NPSH have
to be used.
Table 18.3(a) Head calculations for pumps in Fig. 18.13 (metric units)
All suction and discharge heads are as corrected to centerline of pumps.
For explanation of head symbols, see Fig. 18.12.

I II III IV V VI VII

Liquid Water Water Gasoline Brine Water Water Water


Temperature, in deg C 17 17 21 0 100 38 85
Specific gravity 1.0 1.0 0.73 1.2 0.959 0.995 0.970
Altitude, in m Sea level 1,220 m Sea level Sea level Sea level Sea level Sea level
Barometric pressure, bar absolute 1.014 0.872 1.014 1.014 1.014 1.014 1.014
P" gage 0 0 0 0 0 714 mm Hg 0
vacuum
P" bar absolute 1.014 0.872 1.014 1.014 1.014 0.062 1.014
P, in m of liquid, gage 0 0 0 0 0 -9.74 0
P, in m of liquid, absolute 10.37 8.90 14.2 8.63 10.8 0.63 10.7
Pvp' bar absolute 0.019 0.019 0.41 0.005 1.014 0.062 0.58
P vp' m absolute 0.19 0.19 5.78 0.04 10.8 0.64 6.10
hr,+hb inm 0.61 0.61 0.61 0.61 0.61 0.61 0.61
S, in m -3.96 -2.50 -2.17 -2.38 6.22 6.22 1.65
NPSH, in m of liquid = 5.6 5.6 5.6 5.6 5.6 5.6 5.6
S-(hr• + hi) + P, - Pvp
h., in m of liquid, gage = -4.57 -3.11 -2.78 -2.99 5.61 -4.13 1.04
S-(hfs + hi) + P,
h... in m of liquid 0.31 0.31 0.31 0.31 0.31 0.31 0.31
h.g = h. - hv.. in m, gage -4.88 -3.42 -3.09 -3.30 5.31 -4.44 0.73
h.g, pressure gage 358 mm Hg 252 mm Hg 165 mm Hg 290 mm Hg 0.50 bar 325 mm Hg 0.069
vacuum vacuum vacuum vacuum vacuum bar
h.t = H + h., in m of liquid 19.8 21.3 21.6 21.1 30.0 19.9 25.4
hVd 1.19 1.19 1.19 1.19 1.19 1.19 1.19
hq = h.t - h vd , in m of liquid 18.6 20.1 20.4 19.9 28.8 18.7 24.2
hq, pressure gage 1.83 1.97 1.46 2.34 2.71 1.83 2.30

~
........
QO Table 18.3(b) Head calculations for pumps in Fig. 18.13 (US units)
All suction and discharge heads are as corrected to centerline of pumps.
For explanation of head symbols, see Fig. 18.12.

I II III IV V VI VII

Liquid Water Water Gasoline Brine Water Water Water


Temperature, in deg F 62 62 70 32 212 100 185
Specific gravity 1.0 1.0 0.73 1.2 0.959 0.995 0.970
Altitude Sea level 4,000 ft Sea level Sea level Sea level Sea level Sea level
Barometric pressure, psi absolute 14.7 12.65 14.7 14.7 14.7 14.7 14.7
p .. gage 0 0 0 0 0 28.1 in. Hg 0
vacuum
p .. psi absolute 14.7 12.65 14.7 14.7 14.7 0.9 14.7
P" in ft of liquid, gage 0 0 0 0 0 -32.0 0
p .. in ft of liquid, absolute 34.0 29.2 46.5 28.3 35.4 2.2 35.0-
PVP' psi absolute 0.275 0.275 6.0 0.Q7 14.7 0.9 8.38
PVP ' in ft of liquid, absolute 0.6 0.6 19.0 0.1 35.4 2.2 20.0
hr, + hi' in ft 2.0 2.0 2.0 2.0 2.0 2.0 2.0
S, in ft -13.0 -8.2 -7.1 -7.8 20.4 20.4 5.4
NPSH, in ft of liquid = 18.4 18.4 18.4 18.4 18.4 18.4 18.4
S- (h,. + h;) + P, - Pvp
h" in ft of liquid, gage = -15.0 -10.2 -9.1 -9.8 18.4 -13.6 3.4
S - (hr. + hi) + P,
hv., in ft of liquid 1.0 1.0 1.0 1.0 1.0 1.0 1.0
hog =h, - hv.. in ft, gage -16.0 ft -11.2 ft -10.1 ft -10.8 ft 17.4 ft -14.6 ft 2.4 ft
hog, pressure gage 14.1 in. Hg 9.9 in. Hg 6.5 in. Hg 11.4 in. Hg 7.2 psi 12.8 in. Hg 1.o-psi
vacuum vacuum vacuum vacuum vacuum
h.t =H + h .. in ft of liquid 65.0 69.8 70.9 69.2 98.4 65.4 83.4
hVd 3.9 3.9 3.9 3.9 3.9 3.9 3.9
hq =h.t - hVd' in ft of liquid 61.1 65.9 67.0 65.3 94.5 61.5 79.5
h.tg, pressure gage 26.4 28.5 21.2 33.9 39.2 26.5 33.4
Heads, Conditions of Service Performance Characteristics, and Specific Speed 419

A comparison of installations V and VII shows how the reduction in temperature below that correspond-
ing to suction pressure affects the required suction conditions.

DISCHARGE HEAD

The discharge head (h d) of a centrifugal pump is the head measured at the discharge nozzle. It is the
algebraic sum of the static head, the friction head losses for the capacity being considered, the exit loss
at the end of the discharge line, and the terminal head or pressure. It can be expressed with absolute or
gage readings in meters (feet) of liquid.
Established practice expresses the discharge and suction heads of a horizontal pump with the pump
centerline as datum. Usually, discharge and suction heads of a vertical pump are given with the centerline
of the discharge as datum. Both heads can be given with other elevations as datum, but it is then necessary
to indicate the datum at which they are measured. This practice is often necessary because the exact
elevation of the pump centerline or discharge centerline have not been determined prior to the purchase
of a pump. When the reading of a gage at the pump discharge has been corrected to the pump centerline,
it will indicate the discharge head minus the velocity head at the point of attachment.

TYPICAL DISCHARGE SYSTEMS

Some typical discharge systems are illustrated in Fig. 18.14.


System I shows a system of pump delivery to an elevated tank in which a pressure other than
atmospheric exists; it therefore includes all the components of discharge head.
System II is similar to I except that atmospheric pressure exists on the discharge-liquid level (typical
of pumps delivering to open reservoirs and elevated tanks). If the discharge head is to be expressed as
a gage reading, Pd equals zero and is therefore not shown in the formula. Should it be necessary to
express the discharge head in absolute values, the atmospheric pressure expressed in feet of liquid must
be added to the discharge head expressed as a gage reading.
Although system III illustrates an overhead tank, it applies to all conditions of "overboard discharge."
The actual useful static head (the distance from the pump centerline to the discharge-water level) is less
than the actual static discharge head, D. It is possible to recover all or part of this difference by
incorporating a siphon leg on the discharge. Although systems IVb and V would theoretically be the
most efficient, it is often desirable not to use a sealed discharge. One reason is to prevent the possibility
of back siphonage when the pump is stopped.
In systems IVa and IVb, the effectiveness of the siphon will depend both on the length of the leg
and the design of the piping. Design differences can make the recovery vary from 0 to 100 percent. For
example, if the pipe in IVa was very large in relation to the capacity, the pipe would not run full and
the actual static discharge head would consequently become the distance to the actual water level in the
loop of the piping. All systems using a siphon leg must be investigated carefully to see what percentage
of recovery can be expected and what loss is to be included in the friction loss (hid)'
As the absolute pressure at any point in a siphon must exceed the vapor pressure of the liquid, it is
theoretically possible to employ a siphon leg nearly 10.3 m (34 ft) long with airfree cold water at sea
level. The water being handled is usually not airfree and a reduction in pressure to below atmospheric
causes separation of this air, reducing the effectiveness of the siphon. Water siphons more than 6 m
(20 ft) high are rarely encountered. It is even questionable if many of those under 6 m (20 ft) are 100
percent effective.
Siphon design must provide for the washing out or removal of entrapped air when operation begins,
420 Heads. Conditions of Service Performance Characteristics. and Specific Speed

I n

A
hd= 0 + hfd + he + Pd

m No
B

llt:j
o
I
rr-T~
I 0 O2

~~l
A

I2:b

B A

Fig. 18.14 Determination of discharge heads for six typical discharge layouts.
KEY:
Pd = pressure deviation from atmospheric
he = exit loss at B
litd =friction loss from A to B (including any siphon losses)
h.rJ = velocity head at A
Heads, Conditions of Service Performance Characteristics, and Specific Speed 421

so that the siphon will be established. Unless the air in the loop can be evacuated, the pump will have
to operate during the starting period against a maximum static component DI (!Vb). In some condenser
circulating installations, this condition results in a starting head much higher than the normal operating
head, and special consideration has to be given to the head-capacity curve of the pump. The same effect
can occur in system I if Pd is a negative pressure that is not established until after pumping has commenced
(as in a barometric condenser).
In systems with variable discharge head, it is usually advantageous to establish the head at various
capacities and prepare a graph showing the variation with capacity. When this graph is related to the
suction head, the resulting chart will indicate the system head.
The proper method of specifying discharge heads in preparing an inquiry for centrifugal pumps is
discussed in some detail in Chapter 27.

TOTAL HEAD

The total head, H, of a centrifugal pump is the energy imparted to the liquid by the pump, that is, the
difference between the discharge head and the suction head. As a suction head lift is a negative suction
head, the total head is the sum of the discharge head and the suction lift. If the discharge head and the
suction head are not determined independently, the total head can be calculated (Fig. 18.15) by determining
the algebraic sum of the static head from supply level to discharge level, H sr, plus all friction losses for
the capacity being considered, hI' plus the entrance hi, and exit, he> losses plus the terminal pressure, Pd,
minus the suction supply pressure, Ps• For complex systems involving both vacuums and pressure, it is
often easier to convert all the vacuums and pressures into absolute pressure values of the liquid being
handled, expressed in meters (feet). (To convert bar to m of liquid, multiply by 10.21 and divide by the
specific gravity of the liquid at the pumping temperature. Similarly, for psi to ft of liquid, multiply by
2.31 and divide by the specific gravity.)
As measured by gages attached to the pump suction and discharge openings, the total head is the
discharge head (the sum of the discharge gage reading corrected to the pump centerline and the velocity

Hst
(0-5)
B

A
H = h d - h,
e Hit + hf+ hl+h.+ (Pd - PI)

Fig. 18.15 Detennination of total head.


422 Heads, Conditions of Service Performance Characteristics, and Specific Speed

head at the point of attachment of the discharge gage) minus the suction head (the sum of the suction
gage reading corrected to the pump centerline and the velocity head at the point of attachment of the
suction gage). As the plus and minus signs of the various elements are easily reversed, and as there are
numerous precautions to be considered in taking gage readings, it is advantageous in any test to follow
the instructions in the Test Code of the Hydraulic Institute.

Outmoded Terminology

Total dynamic head, dynamic suction head, dynamic suction lift, and dynamic discharge head are
outmoded terms. Total dynamic head referred to what is now called total head; dynamic suction head,
dynamic suction lift, and dynamic discharge head were defined as the heads measured by a gage corrected
to the pump centerline, and thus did not include the velocity head element.
Misunderstandings arose if the size of the pump suction and discharge openings were not specified,
resulting in different head values for pumps working under identical conditions if their openings were
not the same. Furthermore, in determining the total dynamic head from the dynamic discharge head and
dynamic suction head it was necessary to correct for any difference in velocity head. The present method
of specifying heads is more satisfactory than the dynamic head method.

HEAD TERMS FOR VERTICAL WET-PIT PUMPS

Vertical wet-pit pumps can be either the volute or the turbine type, the latter covering both propeller
and vertical turbine pumps, which were formerly called deep-well pumps. The special hydraulic and
mechanical problems of vertical turbine pumps have caused them to become virtually independent of
the regular centrifugal pump field with different practices and terminology.
Both volute and propeller wet-pit pumps have been handled primarily by engineers in the regular
centrifugal pump field. With these two types, total head is the discharge head measured at the centerline
of the discharge nozzle, with velocity head included, plus the static distance to the suction water level.
Thus, the loss in the suction bell and further losses of the suction strainer and suction piping, if either
is furnished, as well as the losses in the column pipe and elbow in propeller pumps are charged to the pump.
The following head terminology is used by the National Association of Vertical Turbine Pump
Manufacturers for vertical turbine pump applications.

1. Laboratory head-discharge pressure by gage in feet plus static vertical distance to suction water level in
a test setup using the minimum length of column and shafting for a laboratory test.
2. Total head-discharge pressure by gage in feet plus distance to suction water level. (In case of a closed
suction, the total head is discharge pressure plus distance to centerline of suction gage minus suction pressure
in feet.)
3. Dynamic laboratory head-laboratory head as defined above plus velocity head at the point of the discharge
gage attachment.
4. Total dynamic head-total head as defined above plus velocity head at the point of the discharge gage
attachment minus the velocity head at the point of the suction gage attachment, in case of a closed suction.
(What was formerly called field pumping head in vertical turbine pump terminology is now called total
head.) In most vertical turbine pumps, the velocity head is a very small portion of the head developed by
the pump, and its omission is of little importance.
Heads. Conditions of Service Performance Characteristics. and Specific Speed 423

Pump Characteristic Curves


Unlike positive displacement pumps, a centrifugal pump operating at constant speed can deliver any
capacity from zero to a maximum value dependent upon the pump size, design, and suction conditions.
The total head developed by the pump, the power required to drive it, and the resulting efficiency vary
with the capacity. The interrelations of capacity, head, power, and efficiency are called the pump
characteristics. These interrelations are best shown graphically, and the resulting graph is called the
characteristic curves of the pump. The head, power, and efficiency are usually plotted against capacity
at a constant speed, as shown in Fig. 18.16. It is possible for special problems, however, to plot any
three components against any fourth component. When variable-speed drivers are used, a fifth component,
the operating pump speed expressed in rpm, is involved. Where suction conditions may be critical,the
limit of suction-lift-capacity curve or required NPSH-capacity curve is often shown. Many other
relationships can be shown on the same graph as required for specialized studies, for example, specific
speed plotted against capacity.
The curve H-Q in Fig. 8.16, showing the relationship between capacity and total head, is called the
head-capacity curve. Pumps are often classified on the basis of the shape of their head-capacity curves
as described below.
The curve P-Q in Fig. 8.16, showing the relation between power input and pump capacity, is the

180 90 ~
z
160 H-o~HEAb-cJp4cl ~~ ~ 80
w
u

~
w 140 ~ ............. ~ "\ 70
a::
w
Q.

lI'",'pOINT O~-= ~
'- \.
w ]-
u...
120
j'c]'- ~ MAXIMUM ~
,.,.,- \

"" \,
z 60 >-

- ~~ .K'
EF!,CIENC,( u
z
r 100 'l..!!!
()
4a ~~;t"~\
~,Q '\
50 w
u
0 ,;. ./ \ u...
<f 80
Ij,./ 40 u...

\,
w w
::r
....J 60 30
<f
10-
I
--a::::r
~ I
0 40

L
~ 20
Q.

20 I 10 CD

o 0
a 2 4 6 8 10 12 14 16 18 20 22

CAPACIT Y (a), IN 100 GPM

Fig. 18.16 Typical centrifugal pump characteristics.


Double-suction, single-stage volute pump with 8 in suction and 6 in discharge at 1,760 rpm.
424 Heads. Conditions of Service Performance Characteristics. and Specific Speed

power-capacity curve, but is generally referred to as the power curve, the brake horsepower curve, or
the bhp curve.
The curve T\-Q in Fig. 18.16, showing the relation between efficiency and capacity, is properly called
the efficiency-capacity curve, but is commonly referred to as the efficiency curve.
Usually the graph of a pump characteristic is made for a capacity range from zero to the maximum
operating capacity of the unit. The scales on the graph for head, efficiency, and brake horsepower (bhp)
all have the same zero line at the base of the graph (Fig. 18.16).
In some cases, the curve is made for a limited range in capacity. In other cases, to permit clearer
presentation, the head, efficiency, and power scales are so selected that their zero lines do not coincide,
and sometimes these scales are so enlarged that their full range cannot be shown on the graph.

CLASSIFICATION OF HEAD·CAPACITY CURVE SHAPES

Pump head-capacity curves are commonly classified as follows:

1. Rising characteristic-or rising head-capacity characteristic, meaning a curve in which the head rises
continuously as the capacity is decreased (Fig. 8.17).
2. Drooping characteristic-or drooping head-capacity characteristic, indicating cases in which the head-
capacity developed at shutoff is less than that developed at some other capacities. This is also known as a
looping curve (Fig. 18.18).
3. Steep characteristic-a rising head-capacity characteristic in which there is a large increase in head between
that developed at design capacity and that developed at shutoff. It is sometimes applied to a limited portion
of the curve; for example, a pump may have a steep characteristic between 100 per cent and 50 per cent
of the design capacity (Fig. 18.19).
4. Flat characteristic-a head-capacity characteristic in which the head varies only slightly with capacity from
shutoff to design capacity. The characteristic might also be either drooping or rising. All drooping curves
have a portion where the head developed is approximately constant for a range in capacity, called the flat

...:r~

CAPACITY CAPACITY

Fig. 18.17 Rising head-capacity curve. Fig. 18.18 Drooping head-capacity curve.
Heads. Conditions of Service Performance Characteristics. and Specific Speed 425

o
'"
III
:r

CAPACITY CAPACITY

Fig. 18.19 Steep head-capacity curve. Fig. 18.20 Flat head-capacity curve.

portion of the curve. Other curves are sometimes qualified as fiat, either for their full range or for a limited
portion of their range (Fig. 18.20).
5. Stable characteristic-a head-capacity characteristic in which only one capacity can be obtained at anyone
head. Basically this has to be a rising characteristic (Fig. 18.17 and 18.19).
6. Unstable characteristic-a head-capacity characteristic in which the same head is developed at two or more
capacities. (Fig. 18.18 and 18.21). The successful application of any pump depends as much upon the
intrinsic characteristics of the system on which it is operated as upon the head-capacity characteristic. Most
pumping systems permit the use of pumps with moderately unstable characteristics.

CLASSIFICATION OF POWER CURVE SHAPES

Power-capacity curves are also classified according to shape. Fig. 18.22 illustrates a pump characteristic
with a power curve that flattens out and decreases as the capacity increases beyond the maximum
efficiency point. This is called a nonoverloading curve. When the power curve continues to increase
with an increase in capacity, as in Fig. 18.23, the pump is said to have an overloading curve. The shape
of the power curve varies with the specific speed type. As a result, the power curve may have a very
low value at shutoff (see Fig. 18.22 and 18.23), it may have a high value at shutoff (Fig. 18.24), or any
value in between. Whereas in Fig. 18.23 the power curve is an overloading curve with a decrease in
head and increase in capacity, the power curve in Fig. 18.24 is an overloading curve with an increase
in head and decrease in capacity.
Pumps with nonoverloading power curves are advantageous because the driver is not overloaded
under any operating conditions, but they are not obtainable in all specific speed types of pumps. The
actual range of operating conditions encountered in the operation of a pump determines the range in
power requirements, and the driver size should be selected for the power to be encountered.
426 Heads, Conditions of Service Performance Characteristics, and Specific Speed

CAPACITY

Fig. 18.21 Potentially unstable head-capacity curve .

....
I
-
z
~ 90
...-
""n
a: HEAD_ I
w -~/r't' /
-",vyr--.... '"'-
a.. 80
~
~

««? --- V ---I~


~70 140
w

,
()
iL60 , 120
J~' .'\.
u.
w
50 100
~
~
I ~
~ 40 80
w
u.
z 30 / 60 CL
o I :x:
III
~ 20 40
::c
-' 10 / 20
~
~ 0 V
o 20 40 60 80 100
o

CAPACITY, IN 100 G PM

Fig. 18.22 Characteristics of a pump with a non-overloading power curve with reduction in head.
Heads, Conditions of Service Performance Characteristics, and Specific Speed 427

90

40 I--""" .........
80
~
~
v~/
~
~~('.4P4 1'\
1j7
70
I-
Z 60 1:j30
~,(j
~/7')-
\
I/j
"\ \
UJ
u UJ
0:: LL
50 z
/
UJ

~
c..
.,: cS
u 40 ~20 80
zUJ
u 30
J:
...J
V ~~
I-- \
60
iL:
LL
UJ 20
«
l-
~ 10 / ..,.,.., ~
V ~
40
c..
z
III

10
~ 20

0
I o
0
o 20 40 60 80 100

CAPACITY, IN 100 G PM

Fig. 18.23 Characteristics of a pump with an overloading power curve with a reduction in head.

~o
I-
Z
UJ
(J
90
..... ~ 0:
40 -...!:!.f4 D-C ,.G~~
~
r--.... UJ
Q.

"\
80
.,:
~rzr....... <:v9'Y (J
70 z
I-
UJ
w 30 / -<.... ...... \\ !!;!
(J

iL:

"-
60
LL
............ V LL
~ .......... / 50
UJ
.........
~ K....
0 ~
~
UJ ~O ~

'"
:r 8H 40
...J
~
V r-- ~
.....

V
I- 30
0 ~ c..
I-
10 ~ 20
z

V \ III

10
0 V 0
o 10 20 30 40 50 60

CAPACITY. IN 100 G PM

Fig. 18.24 Characteristics of a pump with an overloading power curve with an increase in head.
428 Heads, Conditions of Service Performance Characteristics, and Specific Speed

MATHEMATICAL RELATIONS OF HEAD, CAPACITY, POWER,


AND EFFICIENCY

The useful work done by a pump is the weight of liquid pumped in a period of time multiplied by the
head developed by the pump and is generally known as the hydraulic power. It is usually expressed as
kilowatts or horsepower and can be calculated from the relations:

. specific gravity
hydraulIc power, kW = QH 368

where Q = pump capacity in cubic meters per hour (m3/hr)


H = total head in meters (m)

or

. specific gravity
hydraulIc power, hp = QH 3960

where Q = pump capacity in gallons per minute (gpm)


H = total head in feet (ft)

The power required to drive the pump is regularly determined in kilowatts or horsepower and is
called the power input to the pump. The ratio of the hydraulic power to the power input is the pump
efficiency. The relation between power, capacity, head, and efficiency is therefore:

QH specific gravity
power, kW = .
efficIency 368

or

h QH specific gravity
power, p = effi·
ICIency 39 60

Type Characteristics
If the operating conditions of a pump at the design speed, that is, the capacity, head, input power,
and efficiency at which the efficiency curve reaches its maximum, are taken as the 100 percent standard
of comparison, the head-capacity, power-capacity, and efficiency-capacity curves can all be plotted in
terms of the percentage of their respective values at the capacity at maximum efficiency. Such a set of
curves represents the type characteristic or 100 percent curve of the pump. Figure 18.25 shows the type
characteristic of the pump whose performance is shown in Fig. 18.16.

Centrifugal Pump Characteristics Relations


A set of relations, known as the affinity laws, allow the performance of a centrifugal pump to be predicted.
for a speed other than that for which the pump characteristic is known. These same relations also allow
prediction of the performance of a pump if the impeller is reduced in diameter (within a limit dependent
upon the impeller design) from the characteristics obtained at the larger diameter.
Heads. Conditions of Service Performance Characteristics. and Specific Speed 429

130

c
120

110 -
-r-- r--t. ~
r--.... ...-

W

0 100
~ """-'
""- ~
",."".

c.... ~ ~t/ ~~
A-

'1/'

Zu 90
4.Z
>w
u- 80
~i'l: / ~ "~ i\.
z~
WI&.
-I&.
~w 70
~/
,..... "~' ~
I&.
I&.a
\oJ:)
60 ~ \ \
...V J
da V
c-
w)( ~
,
XC
a
I
~O
...°c~ ~

-_. ~
40
/
~
z
w
u 30
c
W
A-
I
20

10 /
oI
o 20 40 60 80 120 140 160
PERCENT OF CAPACITY AT MAXIMUM EFFICIENOY

Fig. 18.25 Type characteristic, or 100 percent curve.

When the speed is changed (1) the capacity for a given point on the pump characteristics varies as
the speed; and at the same time; (2) the head varies as the square of the speed; and (3) the brake
horsepower varies as the cube of the speed. These relations take the form of equations as follows:

Q = Q,(n/n,)
H = H,(n/n,)2
P = P,(n/n,?

or

where n =new speed desired, in revolutions per minute


Q = capacity, at desired speed n
H =head, at desired speed n for capacity Q
P = power, at desired speed nat H and Q
n, = a speed, in revolutions per minute, at which the characteristics are known
430 Heads, Conditions of Service Performance Characteristics, and Specific Speed

Q. =a capacity, at speed n.
H. =head, at capacity Q. at speed n.
p. =brake horsepower, at speed n. at H. and Q.
For example, a pump is tested at 1,800 rpm and gives the following results:

Capacity Head Power Efficiency


m3/hr gpm m ft kW bhp decimal

908 4,000 47.9 157 142 190 0.83


795 3,500 56.1 184 138 185 0.88
681 3,000 61.3 201 131 174 0.87
454 2,000 67.4 221 106 142 0.78
227 1,000 69.5 228 80 107 0.54
0 0 70.1 230 57 76 0

To obtain the performance of this pump at 1,600 rpm, the first set of values is corrected to 1,600
rpm, as follows:

Q = 908 (1,600/1,800) = 807 m3/hr =4,000 (1,600/1,800) =3,556 gpm


H = 47.9 (1,600/1,800)2 = 37.8 m = 157 (1,600/1,800)2 = 124 ft
P = 142 (1,600/1,800)3 = 100kW = 190 (1,600/1,800)3 = 134 bhp

Changing the other sets of values yields the following (Fig. 18.26).

Capacity Head Power


m3/hr gpm m ft kW bhp

807 3,556 37.8 124 100 134


706 3,110 44.2 145 97 130
606 2,667 48.2 158 91 122
404 1,777 53.4 175 75 100
202 890 55.1 181 56 75
0 0 55.5 182 40 54

The capacity and head figures for these various points can be calculated on a slide rule with one setting.
In this case 1.8 on the C scale would be set over 1.6 on the D scale, and the new capacities would be
read on the D scale opposite the 1,800-rpm capacities on the C scale. The new heads would be read on
the A scale opposite the 1,800-rpm heads on the B scale. Although it is possible to obtain the cube of
a ratio on a slide rule, errors are often made in this step. Except for shutoff (zero capacity) the bhp can
be calculated from the new head and capacity (at 1,600 rpm) using the same efficiency as for the
corresponding head-capacity at 1,800 rpm. Thus the power for the first point can be calculated as (807
x 37.8)/(368 x 0.83) or 100 kW, and in US units as (3,556 x 124)/(3,960 x 0.83) or 134 bhp. The shutoff
horsepower can only be obtained by using the cube of the speed ratio, as both the capacity and the
efficiency are zero.
These relations for a change in speed can be used safely for speed changes up to 2: 1.
Heads, Conditions of Service Performance Characteristics, and Specific Speed 431

240
HE~D-~APlCITly IAoo IRPJ

-
-r-- r--
220

,
........ ........
-- V~ .. _- .. - I--
.........
200 t"-.....
I- / 'f ~
I It
"-
IU
HEAD-CAPACITY 1600

'"
IU
~ 180 RPf.f
-:-.;.,.
~ r- J
I

0
~
IU
:J:
160
K ........
..........
1\
..J ~ 1\
~
r-....

-
I-
0 140
I-
"\ 9o
120 I-"" ..... ~
,-
"
-
200
. . .V V ".
I'
8o
(i~ ~~
100 O~ ~~~
C
,~ 0 ,.. ~ 7o
180
,.q,O
. /V
~

c)~ ~4.
lc.lc." «.,~ ,,/
c
oo1J" I
160 ~ 0
j l"
140
V/ :?-~~ 1 I-
Z

-
50
/~ i
IU
Il.
:J:
/, V U
a:
120
CD
'/ V J ~
r'" IU
Il.
0 .,:
h ! /
V l.J' '" u
Z

VI V ~~ ~ u
IU

,
100
ro~ 3 0 iL
l'/ V: ~~~ ~
IU

80 JI ./ ~
'"
20
V
60 J V
./
I0
~
~ 0
o 10 20 30 40

CA PACITV. IN 100 G P M

Fig. 18.26 Effects of speed change on pump characteristics.


432 Heads, Conditions of Service Performance Characteristics, and Specific Speed

The diameter of an average impeller can be cut down on a lathe by 20 percent of its original maximum
value without adverse effect. Cutting it down to less than 80 percent will generally result in a significant
reduction in head and consequently a much lower efficiency. This 20 percent limit is approximate, as
some impeller designs can be cut more than this, whereas others cannot be cut more than a small
percentage without adverse effect. Any change in diameter will affect the proportions of the impeller,
and some variations from the theoretical results should be expected when tested.
If an impeller is cut in diameter, it is found that, at the same speed, the characteristics of the pump
will have a definite relation to its original characteristics.
These relations are (1) the capacity for a given point in the pump characteristic varies as the impeller
diameter, and at the same time; (2) the head varies as the square of the impeller diameter; and (3) the
horsepower varies as the cube of the impeller diameter. Expressed as equations, these are

Q = Q.(D/D.)
H =H.(D/D.)2
P = P.(D/D.)3

or

where D. = original diameter


D = cut-down diameter
Q. = capacity with D. impeller
Q = corresponding capacity with D impeller
H. = head with D. impeller at Q.
H = corresponding head with D impeller at Q
p. = power with D. impeller at Q. and H.
P = power with D impeller at Q and H.

Changing only the impeller diameter of a centrifugal pump alters its design slightly, and so the affinity
laws do not exactly predict the new performance, the usual case of reducing the impeller diameter
producing a greater reduction in head and capacity than predicted. To compensate for this, the theoretical
impeller diameter ratio has to be corrected in some way. Fig. 18.27 gives a correction for radial impellers
(Ns up to 2,500) in the form of the required diameter ratio versus the theoretical, both expressed as
percent of the original diameter. At the same time, the power varies as approximately the cube of the
actual impeller diameter ratio, so it doesn't decrease at the same rate as the head and capacity, and
consequently pump efficiency tends to decrease as the impeller diameter is reduced. An example will
help to clarify all this.
Referring back to the tabulation of values of the pump tested at 1,800 rpm (with an impeller diameter
of 14.75 in.), if the impeller is reduced to 14.00 in. in diameter, the actual impeller diameter ratio is
0.949, and from Fig. 18.27 the effective ratio for head and capacity will be 0.940 (read from the theoretical
axis.) Applying these ratios, the first set of values is corrected as follows:

Q = 908 (0.940) = 854 m3/hr = 4,000 (0.940) = 3,760 gpm


H = 47.9 (0.940)2 = 42.4 m = 157 (0.940)2 = 139 ft
P = 142 (0.949)3 = 121kW = 190 (0.949)3 = 162 bhp

The other sets of values yield the following (Fig. 18.28):


Heads, Conditions of Service Performance Characteristics, and Specific Speed 433

Capacity Head Power Efficiency


m 31hr gpm m It kW bhp decimal

854 3,760 42.4 139 121 162 0.82


747 3,290 49.6 163 118 158 0.85
640 2,820 54.2 178 111 149 0.85
427 1,880 59.3 195 90 121 0.76
214 940 61.4 201 68 91 0.52
0 0 62.0 203 48 65 0

These relationships are most commonly used to determine the change in speed, the change in diameter
of an impeller, or the combination of both that is necessary to produce a head capacity curve passing
through a given point. For example, suppose the pump whose characteristics are shown in Fig. 18.26
has to meet the conditions of 681 m3/hr (3,000 gpm) at 54.9 m (180 ft) total head. Since this falls below
the head-capacity curve of the 375 mm (14.75 in) impeller at 1,800 rpm, the desired head capacity is
obtained by reducing the speed or reducing the diameter of the impeller.
If the pump, which is to give 681 m3/hr (3,000 gpm) at 54.9 ft (180 ft) were speeded up, or the
impeller diameter increased so that point on the characteristic became 704 m3/hr (3,100 gpm), the head,
at the same time, would have become 54.9 (704/681)2 or 58.7 m (192 ft). Similarly, if 727 m3/hr (3,200
gpm) were obtained by a further increase in speed or impeller diameter, the head would be 54.9 (727/
681)2 or 62.5 m (205 ft). Plotted as shown in Fig. 18.26, these values form a section of a curve (A).
This intersects the 1,800 rpm (375 mm or 14.75-in. D2) head-capacity curve at 712 m3/hr (3,135 gpm)
and 60.1 m (196 ft), indicating the desired point on that characteristic. To obtain 681 m3/hr (3,000 gpm)
and 54.9 m (180 ft), the required speed can be determined by calculation of 1,800 (681/712) or by 1,800
(54.9/60.1)°.5, both of which give 1,722 rpm.

100

/
a::
wa::
I-W 95 ./
/
WI-
:::2:W /'
e(:::2:
-e( /
0-
a:: ° V
w...J
...Je(
...Jz 90
/
/ "
w-
o..S2
:::2: a::
-0
°LL
~o
5~ 85
/
Oz
w-
a:: /
80 ~
V
./
V
75 80 85 90 95 100
CALCULATED IMPELLER DIAMETER IN % OF ORIGINAL DIAMETER

Fig. 18.27 Correction for theoretical impeller diameter reduction.


434 Heads, Conditions of Service Performance Characteristics, and Specific Speed

1,800 RPM
240
14.75 IMPELLER
TOTAL HEAD
220

14.00
t:L: 200

~ 180
w
::c
160

140
4.75 0.9
EFFICIENCY

200 0.8

180 0.7

160 0.6 >-


U
a.
::c zw
I 140 0.5 U
a: u:::
w L1.
~ 120 0.4 w
a.
100 0.3

80 0.2

60 0.1

0 10 20 30 40
CAPACITY - 100 GPM

Fig. 18.28 Effects of change in impeller diameter on pump characteristics.

If no speed change was desired, it would have been necessary to change the impeller diameter. The
theoretical impeller diameter ratio is the same as the speed ratio, namely 0.957. From Fig. 18.27 the
actual ratio would be 0.965, therefore the impeller diameter would have to be reduced to 375 (0.965)
or 362 mm (14.25 in). Had the new driver run at 1,760 rpm, the 375 mm (14.75 in) diameter impeller
would have given 696 m 3/hr (3,065 gpm) and 57.3 m (188 ft) head requiring, in addition, a cut in the
impeller diameter to 369 mm (14.53 in). In all three cases a new curve that would pass through 681 m31
hr (3,000 gpm) at 54.9 m (180 ft) can be plotted by stepping down a number of the 1,800 rpm and 375
mm (14.75 in) diameter impeller points, the capacities being reduced by the ratio of 681n12, while the
corresponding heads are reduced by the ratio of (681nI2)2, and for the speed reduction only the
corresponding power is reduced by the ratio of (681n12)3. For reduced impeller diameters, noting the
point made earlier in this discussion, the corresponding power is reduced by the cube of the actual
impeller diameter ratio.
Heads. Conditions of Service Performance Characteristics. and Specific Speed 435

CALCULATIONS OF SPEED AND DIAMETER


FOR OFF-THE-CURVE CONDITIONS

The process just described in detail is essentially the application of the affinity laws. It is a powerful
tool for calculating the speed or impeller diameter or a combination of both that a pump must operate
with to meet a head and capacity condition that does not correspond to the performance at the speed
and diameter for which test data are available. A simplified approach to its use is illustrated in the
following example.
Let us assume we are dealing with a known pump performance as given on Fig. 18.29 at a speed of
1,800 rpm. We want to determine the speed at which this pump would have to run so as to deliver 454
ml/hr (2,000 gpm) at a total head of 45.7 m (150 ft). The steps required would be as follows:

1. Since the required condition point is below the known pump performance, select an arbitrary capacity greater
than the 454 m3jhr (2,000 gpm), such that it will be located on the parabola defined by the affinity laws:

Say, for instance, that Q3 = 568 m 3/hr (2,500 gpm), then,

H3 = HI (~:)2 = 45.7 (568/454)2 = 71.5 m (234 ft)

2. Draw the portion of the parabola defining the affinity laws between 454 m 3/hr at 45.7 m and 568 m 3/hr,
71.5 m (2,000 gpm at 150 ft and 2,500 gpm, 234 ft). It can be assumed that this is essentially a straight line.
3. The intersection of this straight line with the head-capacity curve of the pump at 1,800 rpm corresponds to
544 m 3/hr and 65.6 m (2,395 gpm and 215 ft). We can now determine the speed required to meet the desired
conditions of 454 m3/hr (2,000 gpm) and 45.7 m (150 ft):

n2 = 1,800 (45.7/65.6)°.5 = 1,503 rpm

240
H-Q 1800 RPM /-- H: 150_(~ggg)2
220 : 234A ft.

-
intersection --" 1L2 for 2000 gpm. 150 ft.
200 at 215 ft. .' , :1800-J150
215
--: 180
, :1503 RPM

,,
::I:
~
160

14O~____~~____~~____~~~____~~
o 1000 2000 3000 4000
Capacity in G.P.M

Fig. 18.29 Calculation of speed change.


436 Heads, Conditions of Service Performance Characteristics, and Specific Speed

Although the ratio of the capacities could also have been used to calculate the speed, it is easier to read the
head values accurately and taking the square root of them halves any errors in reading. A similar process is used
to determine the impeller diameter required to meet conditions at the same speed and a reduced impeller diameter.
In the example just given, assuming that the original impeller was 375 mm (14.75 in), the new impeller would
have to be cut down to a ratio of (45.7/65.6)°.5 or 83.3 percent theoretically. Applying the correction indicated
in Fig. 18.27, however, we would cut the diameter to 85.5 percent of its original diameter, or 320 mm (12.6 in).

Design Constants
The designing of centrifugal pumps is not an exact science because of the many interrelated factors
whose combined effect cannot be accurately foreseen and thus must be determined experimentally. The
development of centrifugal pumps has been largely a result of the accumulation of data on the performance
of both specific designs in service and of experimental designs, the result of research and experiences
in other hydraulic fields, and the application of this information to the development of new designs. In
analyzing data, centrifugal pump designers use various constants, formulas, and relations, two of which
are of interest to users of centrifugal pumps: (1) model pump relations and (2) specific speed.
Pumps are analyzed and compared basically at their so-called design conditions; that is, at the head
and capacity condition at rated speed at which maximum efficiency is obtained. Thus, for the pump
whose characteristics are shown in Fig. 18.16, the design conditions would be 336 m3Jhr (1,480 gpm)
and 40.2 m (132 ft) total head at 1,760 rpm.

MODEL PUMPS

A model pump has the design features of a full-size unit on a smaller scale. To meet the requirements
of a strict model, all linear dimensions of the model must be in the same proportion as the corresponding
dimensions of the full-size pump. The theoretical relationship of the performance of a model pump can
be easily visualized by considering two pumps identically proportioned, with one having twice the linear
dimensions of the other.
The impeller of the smaller pump will be half the diameter of the larger and will, therefore, have to
run at twice the rotative speed of the larger for the same peripheral velocity and equal design head. The
areas through the waterways of the smaller pump will be one-half squared or one-quarter the areas of
the larger pump. Thus, at equal velocities, the capacity of the smaller pump will be one-quarter that of
the larger pump. Therefore, it is apparent that for the same design head the interrelationship of exactly
similar pumps would be theoretically:

f=~=~=~~
Lb na "'VQb
where f = the ratio or factor of the two pumps
La and Lb = comparable dimensions of the two pumps
na and nb = the rotative speeds of the two pumps
Qa and Qb =the capacities of the two pumps at comparable points on their characteristic curves.

From dimensional analysis the general equations for the performance of model pumps are:

Qb = Q af3(Ilt,/n.,)
Hb = Haf2(nb/na)2
Heads, Conditions of Service Performance Characteristics, and Specific Speed 437

The preceding equations are based on the assumption that the two pumps are proportional in every
way and that the same relative degree of smoothness is obtained in the two pumps. This is difficult to
attain, as the actual smoothness of castings is approximately the same, regardless of size. Thus, the
relative internal smoothness of a larger pump is greater than that of a smaller pump. This is reflected
in the head losses in the pump waterways; the larger pump should produce a higher head than the smaller
pump for points of similar capacity. Inasmuch as part of the liquid pumped leaks through the wearing
rings, this loss may not be in exact proportion in both sizes of pumps, thus affecting the net quantity
delivered. Part of the power input goes into mechanical losses (bearings and stuffing boxes), that are
roughly but not exactly proportional to the pump sizes, resulting in a third discrepancy. Good mechanical
design (especially in the production of a commercial line of pumps) precludes making the shaft, casing
thickness, or thickness of impeller vanes of two pumps in exact proportion to their size factor. A
comparison of the largest and smallest pumps of a closely homologous line of a commercial design,
therefore, will show some difference in performance. The magnitude of this difference will depend upon
the size factor and the actual physical sizes of the two pumps. Centrifugal pump designers are careful,
when making model pumps, to use a size that will be close enough to the full-size pump so that the
results of the model will permit a reasonably close prediction of the performance of the full-size pump.
Model pumps have been used to prove within a reasonable degree of accuracy the performance of
the full-size unit for almost every case involving special large-capacity pumps.

SPECIFIC SPEED

An analysis of the performance of a projected centrifugal pump would be difficult without the progress
achieved in the science of hydrodynamics in the four centuries of its existence. This progress may be
directly credited to the almost universal application of model study, which precludes the necessity of
experimenting upon full size commercial constructions that are too expensive and least convenient for
securing the necessary information. Sir Isaac Newton evolved the theory of dynamical similarity, in
1687, thereby introducing the mathematical background for model investigations.
The application of the Newtonian principle of dynamical similarity has since given rise to the wide
use of models in hydraulic machinery, as well as in other fields of science, and to an extensive knowledge
of the relative performance of models and prototypes.
One such application of the principle of model and prototype relationship has enabled engineers to
predict the performance of centrifugal pumps on the basis of the behavior of other machines, smaller
or larger in size, operating over a wide range of design conditions, but modeled from and similar to
each other.
The principle of dynamical similarity expresses the fact that two pumps geometrically similar to each
other will have similar performance characteristics. In order to afford some basis of comparison among
various types of centrifugal machines, it became necessary to evolve a concept which would link the
three main factors of these performance characteristics-capacity, head, and rotative speed-into a single
term. The term "specific speed" is such a concept. The mathematical analysis used to establish the
relationship between the specific speed of a pump and its operating characteristics does not enter the
scope of this book. In its basic form, the specific speed is a non-dimensional index number which is
numerically equal to the rotative speed at which an exact theoretical model centrifugal machine would
have to operate in order to deliver one unit of capacity against one unit of total head. It is mathematically
expressed as:

-~
Ns - (gH)"3t'4
438 Heads, Conditions of Service Performance Characteristics, and Specific Speed

in which:
N, = specific speed
n = rotative speed
Q = capacity
H = head (head per stage for a multistage pump)
g = gravitational constant, 9.81 m/s2 (32.2 ft/S2) at sea level.

In order for this relation to remain dimensionless, when using English units, the rotative speed would
have to be expressed in revolutions per second, the capacity in cubic feet per second and the head in
foot-pounds per pound or foot. However, since specific speed is used only as an index or type number,
certain liberties are permissible in selecting the units used. Thus, the gravitational constant, g, is dropped
out of the relation, leaving:

The rotative speed is expressed in revolutions per minute. For some time, two units of capacity, gallons
per minute and cubic feet per second, were used in the United States to determine specific speed, but
the gallons per minute basis has been accepted as standard by the Hydraulic Institute and is now the
approved basis. The unit of head is one foot.
In countries using the metric system, specific speed is today designated nq and the usual unit for flow
in pump design is cubic meters per hour (m3Jhr). The unit for head is one meter (m). The mathematical
relationship between values calculated in the two systems of units is

The formula for the specific speed of a pump remains unchanged whether a single- or a double-
suction impeller is used. It is customary, therefore, when listing a definite value of specific speed, to
mention what type of impeller is in question.

Type Specific Speed


The type specific speed, by definition, is that operating specific speed that gives the maximum
efficiency for a particular pump and is the number that identifies the pump type. It should be noted that
this index number is independent of the rotative speed at which the pump is operated, since any change
in speed carries with it a change in capacity in a direct proportion and a change in head varying as the
square of the speed.
The normal range in specific speeds encountered in single-suction impeller designs is from 500 to
15,000. Basically, the lower the specific speed type, the higher the head per stage that can be developed
by the pump.
Normally, the conditions of service for which a pump is sold are relatively close to the maximum
efficiency point, and the specific speed determined from the conditions of service will be a close indication
of the pump type. For example, the true type specific speed of the pump whose characteristics are
illustrated in Fig. 18.16 is 1,740. This pump would normally be applied for a range of conditions between
295 m3Jhr and 42.7 m (1,300-gpm and l40-ft) total head and 363 m3Jhr and 38.1 m (1,600-gpm and
125-ft) total head.

Significance of Type Specific Speed


One of the most important applications of the specific speed concept is the fact that all sizes of pumps
can be indexed by the rotative speed of their unit capacity-unit head model. Thus, the specific speed
Heads, Conditions of Service Performance Characteristics, and Specific Speed 439

concept can be used in such a manner that for homologous designs, the performance of any impeller of
the series can be predicted from the knowledge of the performance of any other impeller of the series.
Because the physical characteristics and the general outline of impeller profiles are intimately connected
to their respective type specific speeds, the value of the latter will immediately describe the approximate
impeller shape in question. As an illustration of this statement, Fig. 18.30 represents a few typical
impeller outlines tied down to their type specific speeds.
The specific speed of a given pump will also definitely be reflected in the shape of the pump
characteristic curves, and, whereas some variations in the shape of these curves can be obtained by
changes in the design of the impeller and casing waterways, the variation that can be obtained without
adversely affecting the pump efficiency is relatively small. Approximate type characteristics for four
single-suction impeller types are shown in Figs. 18.31 to 18.34. Figure 18.35 shows the variation of
head with the specific speed for shutoff, 25, 50, 75, and 110 percent capacity. Figure 18.36 shows the
variation of power with specific speed for the same capacities, while Fig. 18.37 shows the variation of
efficiency with specific speed for these same capacities. The values shown in Figs. 18.31 to 18.37 are
for more or less normal impeller-casing designs and combinations. Variations in the shape of the curve
will be found, depending on the individual design of the pump.
The variation in the shape of the type characteristics between a single-suction impeller with shaft
through the eye and an overhung single-suction impeller is small, therefore Figs. 18.31-18.37 can be
applied to either type. Historical practice for double suction impellers was to design them as two half-
capacity single-suction impellers back-to-back, and so they had a type characteristic approximating that
of a single-suction impeller having a specific speed 70.7 percent or l/-{2 of that of a double-suction
impeller. Modern practice is to design double-suction impellers for the full capacity, changing only the
inlet portion to reflect the different suction arrangement. These designs therefore have efficiency and
type characteristics similar to that of single-suction impellers of the same specific speed. Figure 18.30
also indicates the maximum range of efficiencies obtainable from pumps of different specific speeds.
Low specific-speed impellers have a lower maximum efficiency than medium-specific-speed impellers
because the former have considerably more disk area for a given set of operating conditions and, therefore,
a greater loss in disk horsepower.
The curve of efficiencies in Fig. 18.30 was compiled from data assembled in the late 1940s. More
recent experience gives somewhat different and improved efficiencies, particularly in the range of lower
specific speeds. In addition, the curves in Fig. 18.30 do not take into consideration a number of variables,
all of which affect the efficiency of a pump to a significant degree.
The latest available data on today's commercially attainable efficiencies are provided in two papers
published in 1986 and 1987 [18.1, 18.2]. The data published in these two papers are based on certain
constraints on the following variables:

1. Single-stage pumps only


2. Finish and dimensional fidelity are comparable to precision cast impellers with a I-percent plus or minus
tolerance on all dimensions of the vanes and hydraulic passages.
3. A relative roughness of all hydraulic waterways of the impeller and casing be 0.000020 or better.
4. Standard commercial diametrical clearances of all wearing rings-that is, approximately 0.0015 of the
ring diameter.
5. A suction specific speed (see Chap. 19) not exceeding 8,500 for single-suction overhung impellers or 7,500
for single-suction shaft-through-eye or double suction impellers.
This corresponds to incipient suction recirculation at about 55 percent of best efficiency point (BEP) for
pumps of Ns 500-2,500, rising to about 70 percent for pumps of Ns 2,500-10,000. See Chap. 22 for a
method of estimating the incipient suction recirculation capacity.
440 Heads. Conditions of Service Performance Characteristics. and Specific Speed

0 8 00
8
II')
§. ~
a- § ~ ~ ~ i

-- ---- --
G in;::;.- -- :0:

--
90 90
~
~ I:':-- ~

... ~ r--

--
r<' ~ PAl ~~ iP..-
8 80
I \()A ~l9- ~~ r--
f-
7 !~~ ~ (;\'~ 70
)-
,1G; c..\'+ ...--
~, t Vo ~+
60
/ ~~~,O

50
/ ~i;;
50
V
40
~ J § St J
0
0
~ it
0
0
~

CENTER OF
ROTATION

Fig. 18.30 Approximate relative impeller shapes and efficiency variations with specific speed.
Heads. Conditions of Service Performance Characteristics. and Specific Speed 441

130

120
HEA D-CA PAC I Y

--- .....- V' -V"


>-
u 110
z ~ ...........
w
u 100
Li: E FICIENCY V
u..
w 90 I"
./ /" """""
::e
:::>
::e / /
V
80
x V
«
::e 70 ./
f-
«
>- 60 I/:V BHP

f-
U
« 50
) V
a. f
«
u
40 /J
/
u..
0
f-
/'
z 30
w
u
a:: 20 I
w
a.
10 /
0
V
o 10 20 30 40 50 60 70 80 90 100 110 120

PER CENT OF HEAD,EFFICIENCY, AND POWER AT MAXIMUM EFFICIENCY

Fig. 18.31 Type characteristics for N, = 600 single-suction impeller.


442 Heads. Conditions of Service Performance Characteristics. and Specific Speed

130

120
HEAD-CAPACITY
110 ~
--... ~
100 ~ .......... ~

"
Q

V t:::::
po.
~~ EFFICIENCY V
,.:Z 90 r--.

~V
u'"
zu
"'-
- Lt.. 80 /
~V
ULt..
G:'" 8HP
ti ~ 70 r'
02
cc-
'" ~
~ ~
60 V
::r 2
~ ~ 50 ~~J
u.152
~
za:
'" '"

~
40

30
i
i /
II
i V
20

10
Ii
oV
i

T
I
o 10 20 30 40 50 60 70 80 90 100 110 120

PER CENT OF CAPACITY AT MAXIMUM EFFICIENCY

Fig. 18.32 Type characteristics for N, = 1,550 single-suction impeller.


Heads, Conditions of Service Performance Characteristics, and Specific Speed 443

160

150
u> r-- r--
z
w 140
r---...
U """""" ............
HEAD-CAPACITY
~

'"
LA.
LA. 130
UJ
2 120 ~
:;)
2 "'-
X 110 ~
..
~
2
.- 100 ~
K
~
BHPr -",-
II:
LIJ 90 - ./
Y
~ [\
0
n,
80
V
0
z~ EFFICIENCY
70 i
~
u
z
w 60 V !
U
Li:
LA.
LIJ
50
I(
I
0
~ 40 J
LIJ
%

""0 30
I
.-z
LIJ 20 /
0
cr:
j:t
10

1/
LIJ
Il.

10 20 30 40 50 60 70 80 90 100 110 120

PER CENT OF CAPACITY ATMAXIMUM EFFICIENCY

Fig. 18.33 Type characteristics for N, =4,000 single-suction impeller.


444 Heads, Conditions of Service Peiformance Characteristics, and Specific Speed

300

280

>-
260 1\\
u

"
z 240 ~
w
U HEAD-CAPACI T Y
220 "-
'\~
~
~
w
200 ~

""
-. i"-
'"
~ 100
;:)
V"'"
!)( 180 ~
"
~ / 90
c(

...
~
160 ~ ~/ FFICIENCY
80
>-
u
z
f' / ""
C(
l&.I

a: ~ ~ U
w 140 70 ~

~ ~
~
0
120 ) r---..... ~ 60
l&.I

V ..........
r----. ~.
Q. ~
;:)
0
Z B~ p ~
100 50
1/
C( X
c
~~
0
C( ~
w 80 40
:r
/ t\
~

l&.. ...z
0

...z
0 60
1/ 30
l&.I

w j U
u 40 20 «
a: V ~

t1
l&.I
Q. 20 10

10 20 30 40 50 60 70 80 90
o
100 li D 120

PER CENT OF CAPACITY AT MAXIMUM EFFICIENCY

Fig. 18.34 Type characteristics for N, = 10,000 single-suction impeller.


Heads, Conditions of Service Performance Characteristics, and Specific Speed 445

300 . - 300

.J.
/
2~O 250
I
>- V
u j
z
w
U
20 0
I
I
SHUTOFF
--, 1/ 200
IL.
IL. ! I
/ II'
W
I 25 PER CENT CAPACITY ~ V
~
::>
b( if ~
:E ! / / ./ V-
I V/
)(
«
:E
150 I ./ 150
.l L . / ./
~ 14 0 LL 140
« i lL..l IL I
L
50 PER CENT CAPACITYL
a 13 0 J/ , :/

-
130
« ~ ~. ~
...
-
w ~ ~ j .-A"
J: 12 120
~ io"""" -----75 PER CENT CAPACI!)
Lo-
o I • I
II 0 110
~ j
z
w 10 0 100
u
i

-
a:
w 110 PER CENT CAPACITY
Il. 0
. ! .....-.,1 90

80
r--- r-...... .0

7 TO
- I)
~ I) ~
o~
0 0 0 0 I) 0 ~
o 0 0 o o o o 0
• ... ID 2- ~ ~- 2- i i
SPECIFIC SPEED, SINGLE-SUCTION TYPE

Fig. 18.35 Variation in head values at shutoff, 25, 50, 75, and 110 percent capacity with specific speed.
446 Heads, Conditions of Service Performance Characteristics, and Specific Speed

'00 :5 00

100 I 00

~
!<.,o ~
IS 0
,;.:~v~ t-~ H so
./ ~
, ~ ./ l/~
L.:~
110 PER CENT CAPACITY V ...... ~ ~ ~
10 0 I - - - .- --
f- - - .... - ._.- .....:::
~
i""'"
- I 00
,
-
~5, ~ ¥-ENT CAPA,cITY
~
• 10- 50 PER CENT CAPACITY
10 .....
~
L
10
70

-
,. ~to.P\t!,.....
,• •.oo
.1 ""7
0
1.' pI RCE'L-- .......
I ~\' 1/
" ~-J"O/
..
0
!,...-o i--'"
.,~
.o
, ....
:5 o

.8 ,
1ft

:o
~
.. 8o
.
8o
0 ~
8 8
I 0 0
o

t-
til
• ."

SPECIFIC SPEED, SINGLE -SUCTION TYPES

Fig. 18.36 Variation in power values at shutoff, 25, 50, 75, and 110 percent capacity with specific speed.

/11 o PER CENT CAPACITY


10 0 100

• 0 I
75 PER CENT CAPACI TV
to

I 0 10

-r- -
50 PER CENT CAPACITY

10 TO
1'--0.
10 10

-
0 SO

-
""--
.. 0 l"- t-- IS PER CENT CiPAclTv

t-- ...... )'-.


~
0 :50

I .. o
10
g o o 0 0 0 o
•o •
o
9- ;. o
~
o
o
2
SPECIFlC SPEED, SINGLE - SUCTION TYPE

Fig. 18.37 Variation in efficiency values at shutoff, 25, 50, 75, and 110 percent capacity with specific speed.
Heads. Conditions of Service Performance Characteristics. and Specific Speed 447

6. For single stage pumps with a shaft through the eye, the shaft-to-eye diameter ratio is sufficiently low to
preclude blockage in the fluid passages of the impeller inlet.
7. The discharge recirculation value is not less than nor more than 80 to 90 percent of the maximum
efficiency capacity.
8. A uniform velocity profile of the fluid entering the impeller inlet. This requires an evaluation of the piping
or channel flow at the pump inlet to assure that a uniform velocity profile is achieved at the rated
flow conditions.
9. Pumped liquid is clear water at 66°C (l50°F) or less.
10. Efficiencies for maximum impeller diameters. Cutdown impellers usually result in a two to three point loss
in efficiency.
11. Wet-pit pump efficiencies are based on impellers with no back wearing rings or balancing holes.

The results of analyzing the perfonnance of pumps meeting these constraints are shown in Fig. 18.38
for single- and double-suction pumps and in Fig. 18.39 for wet-pit pumps.
To use these curves requires some measure of the pump speed. The curve shown in Fig. 18.40 relates
speed to capacity, and is based on a n-/Q parameter of approximately 115,000, where n is the rotative
speed in rpm and Q is the rated capacity in gpm. In practice, first enter Fig. 18.40 with the maximum
efficiency capacity of the pump in question and then read off the standard speed from the ordinate. The
capacity of the pump is then corrected by the ratio of the speeds. Next enter Fig. 18.38 or Fig. 18.39,
depending on the type of pump, with this capacity and the specific speed detennined from the original
pump rating. As an example, consider the following conditions of service for a single stage single
suction pump.

1,100 gpm
225 ft. thd
1.750 rpm
1,000 Ns

Enter Fig. 18.40 for speed correction. The corrected speed is 3,450 rpm then

3,450
1,750 x 1,100 gpm = 2,200 gpm

Enter Fig. 18.38 with I,OOONs and 2,200 gpm to detennine the efficiency as equal to 81.9 percent.
The second paper referred to above provided correction factors to be applied to the efficiencies given
in Figs. 18.38 and 18.39 when deviations occur from the constraints established for the basic efficiencies:

1. The effect of changing hydraulic surface roughness from 0.000002 per inch of impeller diameter to 0.00001
per inch of impeller diameter is shown in Fig. 18.41.
2. The effect of increasing or decreasing ring clearances from the stated value of 0.0015 ratio of ring clearance
to ring diameter is shown in Fig. 18.42.
3. The loss of efficiency associated with an increase in the Suction Specific Speed over the 8,500 and 7,500
values chosen as the basic constraint is shown in Fig. 18.43 for an impeller without a shaft through the eye
and in Fig. 18.44 for an impeller with a shaft through the eye and a shaft to impeller diameter ratio of 0.2.
4. Finally, the effect of discharge recirculation values on pump efficiency is illustrated on Fig. 18.45.
448 Heads, Conditions of Service Performance Characteristics, and Specific Speed

96

--
500,000 GPM
94
92
-"~
g:888
1r'000 ~~~
90 ~ ~ """'
~
" 10,000 GPM
88
~ ~ i"""" 5,000 GPM

86
~~ ~ ~ 2,500GPM

84
~ ~~ V-'" ~ 1,000 GPM

82 W 1/ V/ i-""" 750 GPM

80 fHv I.; V L ~ ~ 500 GPM


400 GPM
78
~rl II 1.1 V-" V-" ~ 300 GPM
~
0 r/) 'I II //V
z 76 ,..... 200 GPM
>- 74
V )) .... ~ V
()
zw / If)r/ ~I ' /
(3
72
70
rr; I{j ~V, ~
100 GPM

u:: 68
Vl V. V J ~
V
f/j Vj
~
w /
66
64
rJ.V V
'/ V
62
60
V
58
56
1/
j
54
52 V
50
48
46
500 1,000 2,000 3,000 5,000 10,000

VALUES OF SPECIFIC SPEED Ns = RPM~


H%

Fig. 18.38 Efficiency of single stage end suction and double suction centrifugal pumps.
Heads, Conditions of Service Performance Characteristics, and Specific Speed 449

96
94
92

-
I 00,000 GPM
90 40,000 GPM
88 20,000 GPM
10,000 GPM
86 5,000GPM
84

-
1,000 GPM
82
i'" t--"",
80 r- t- t--"", 500GPM

;:.!!
0
z
78
76
I"'" - l- t--,...
400GPM
300GPM

-
r- ~
>-
<.)
74 -~
200GPM
z
w 72
~ t--....
C3 70 100GPM
u::: 68
LL
w
66
64
62
60
58
56
54
52
50
48
46
500 1,000 2,000 3,000 5,000 10,000

VALUES OF SPECIFIC SPEED Ns= RPM..JGPM


H%

Fig. 18.39 Bowl efficiency of wet-pit centrifugal pumps.


§

1 1.5 2 2.53 4 5 6 7891 1.5 2 2.5 3 4 5 6 7 891 1.5 2 2.5 3 4 5 6 7 891 1.5 2 2.5 3 4 5 6 789 1
10,000 10,000
9 9
8 8
7 7
6 6
5 5
4 4
3
.............
3
2.5 ....... 2.5
2 ~ 2
"" ........
......... ~
1.5 1.5
.......
RPM ........ RPM
.............
1,000 1,000
9 9
8 ="'" ....... 8
7 7
6 ro...... 6
5 r..... 5
4 r--....
4
3 ~ 3
2.5 ........... ~
2.5
........ ......
2 2
.........
1.5 1.5

100 100
1 1.5 2 2.53 4 5 6 7891 1.5 2 2.5 3 4 5 6 7 891 1.5 2 2.5 3 4 5 6 7 891 1.5 2 2.5 3 4567891
100 1,000 10,000 100,000 1,000,000
FLOW-GPM
Fig. 18.40 Chart for speed correction.
Heads, Conditions of Service Performance Characteristics, and Specific Speed 451

10
9
8
7

,
CJ)
I- 6
z \.
0a.. 5
I 4 i\.
>-
0
z 3
w ~

",
0
u:::
u. 2
\
w
Z
w
CJ)
«w
a:
0
1
.9
\ "-
w .8
Cl .7
.6
.5
",
.4
500 1,000 2,000 3,000 5,000

VALUES OF SPECIFIC SPEED Ns = RPM~


H%

Fig. 18.41 Loss of pump efficiency with specific speed for change in relative roughness of waterway surfaces.
Relative roughness increased from 0.000002 to 0.00001 per mm (in) of impeller diameter.

We repeat that the charts on Figs. 18.38 through 18.45 reflect results obtainable with modern pumps.
For the efficiency obtainable from older pumps, it is more prudent to use values from Fig. 18.30.

Approximating Specific Speed Type


From Impeller Outline

Prior to the general adoption of specific speed as a type indicator, the ratio of outside diameter (D 2)
to suction eye diameter (D,) or the reciprocal of this relationship, was generally used for that purpose
(Fig. 18.46). An approximate relation of the DJD, ratio to N, for single-suction impellers is shown in
Fig. 18.47. These values are necessarily approximate, as a true curve would be a fairly wide band. One
reason for this can easily be seen if one considers an impeller for a given set of conditions. The velocity
of the liquid would have to be approximately the same whether the impeller had no hub or a hub
extending into the eye. Thus an impeller with a hub extending into the eye would necessarily have a
larger D, for the same capacity and head or the same D 2• Multistage pumps require a large shaft because
of the power involved, so that the impellers of such pumps would have relatively large hubs. The impeller
would have an abnormally large D, or a smaller DJD, ratio than normal for its specific speed type.
This method of identifying the specific speed fails for axial-flow impellers. Axial-flow impellers fall
into a 9,000 to 20,000 specific-speed range. Their output and, therefore, their specific speed depends on
452 Heads, Conditions of Service Performance Characteristics, and Specific Speed

-10
en
en
0
....J

-5

~
=__ 2 X STD. CL.
x STD. CL.

=,
z _1.5
0
0a.. STD. CLEARANCE
~ 0.5 STD. CL.
0.0015 IN/IN

5
z
<C
CJ

10

1,000 2,000 3,000 4,000

SPECIFIC SPEED

Fig. 18.42 Variation in pump efficiency with specific speed for various running clearances.

the angle and length of the vanes as well as the number of vanes. To predict the characteristics of axial-
flow impellers, a designer would require very detailed information on the impeller and other pump parts.
No simplified guide can be offered for general use for this type of pump.

HEAD AND CAPACITY CONSTANTS

Two design constants can be used to approximate the performance of a centrifugal pump. One expresses
the relation between the impeller peripheral speed and the total head. The second relates the radial
discharge velocity from the impeller (and, therefore, the capacity) and the total head. The formulas for
these constants are:

U2
cp = ...j2gH

and:

in which: U2 = peripheral velocity, in mls (ft/s)


g = gravitational constant 9.81 mls2 (32.2 ft/s2)
Heads, Conditions of Service Performance Characteristics, and Specific Speed 453

12,000
0
w
W
D-
C/)
0
11,000
u:::
C3
W
D-
C/)
z 10,000
0
t= -3 POINTS
0
:::>
C/)
u... 9,000
0 -2.5 POINTS
C/)
w
:::>
....J
~ 8,000

7,OOO...L----.,-------.---------r-------
1,000 2,000 3,000
VALUES OF SPECIFIC SPEED

Fig. 18.43 Loss of pump efficiency with specific speed for various suction specific speeds, hdD2 = O.

Cm2 = radial discharge velocity, in mls (ft/s)


H = total head (per stage), in m (ft).

These relationships can be converted into terms of impeller dimensions that can be measured:

and

C QxlO" Q. . .
m2 = 3,600Az = 0.36A2 m metnc UnIts
C 144Q Q. us .
m2 = 7.48 x 60 x A z = 3.117Az m UnIts

where Dz =outside diameter of impeller, in mm (in)


n = speed, in revolutions per minute
Q = capacity, in m3/hr (gpm)
A z = circumferential discharge area of impeller, cm 2 (in2) (Fig. 18.46)
454 Heads. Conditions of Service Performance Characteristics. and Specific Speed

12,000

o
I±l 11,000
0-
en
(,)
u:::
C3
~ 10,000
en
z
o
i=
(,)
~ 9,000
u..
o
en -3 POINTS
w
=:l
...J
~ 8,000 -2.5 POINTS

-2 POINTS
-1 POINT
7,000

6,000...L..----r--------r----------,----
1,000 2,000 3,000
VALUES OF SPECIFIC SPEED

Fig. 18.44 Loss of pump efficiency with specific speed for various suction specific speeds, hdD2 = 0.2.

(In high-speed Francis vanes, mixed-flow, and axial-flow impellers, the effective discharge diameter
is not D2 but a geometric mean: ""D~ + h'Y2. When there is little difference between D2 and h2' an
arithmetical mean is commonly used. However, for simplicity, constants presented here are calculated
on the basis of D 2 , not the effective diameter that most designers would use in calculating their constants.)
The equations for the design constants can now be replaced by

'" D2n D2n.. .


'f' = 19.1 x 1()3 (2gH)0.5 = 84.6 x 1()3 (H)0.5 m metnc umts
4> = D 2n = D 2n in US units
229 ....j2gH 1,840 -{ii

and

K QQ...
cr =0.36A (2gH)0.5 = 1.60A2 (H)0.5 m metric umts
2

Ker = Q = Q in US units
3.117A2 ....j2gH 25A 2 -{ii
Heads, Conditions of Service Performance Characteristics, and Specific Speed 455

10
9
8
7
en 6
I-
z 5
6
Il..
4
W 3
(!)
z 2 DISCH REC = 1.1 BEP CAPACITY
«
I
DISCH REC = 1.0 BEP CAPACITY
0
>- 0
0 -1 DISCH REC = 0.7 BEP CAPACITY
zw
-2
(5
u::
u.
-3
w -4 DISCH REC = 0.7 BEP CAPACITY
-5
-6
-7
-8
-9
-10
500 1,000 2,000 3,000 4,000
SPECIFIC SPEED

Fig. 18.45 Variation in pump efficiency with specific speed for various ratios of discharge recirculation to
design capacity.

RADIAL TYPE MIXED-FLOW TYPE AXIAL-FLOW TYPE

AXIS OF
ROTATION

r';I>. -Z2~
A,o ••

Fig. 18.46 Dimensional symbols for impellers and formula for determining discharge area A2•

These equations can be further transformed to give head and capacity values directly:

H D2n
= ( 84.6 X103<\1
)2.
III
. .
metrIc umts

D2n)2. US .
H = ( 1,840<\1 III umts
456 Heads, Conditions of Service Performance Characteristics, and Specific Speed

SPECIFIC SPEED DOUBLE - SUCTION TYPES

-I
8
I
0.8
7 0.7
/
6 / 0.6
/
5 / 0.:5
/
1/ 0.4
/
3 ........
*-<1,./ -- 0.3 ~
'u"
...... D.
o..... .....~/O
L4.
0
~
-::-.....' V &/ (f)

3
2 ..... V ./ 0.2
~
~

--
L 0.15
r~
-
.....
/ I":
..... ~
./' .- ~
........
1.0 V ~
0.10
o Q
0.09
O•II 0.08
o.7
a
...
g2· o
~ §
0.01

g
SPECFIC SPEED SINGLE-SUCTION TYPES

Fig. 18.47 Variation with specific speed of DJDI ratio, <I> and Kcr constants.

and

Q = 1.60Ka AzeH)o.5 in metric units


Q = 25Kcr AzeH)05 in US units

Both <I> and Kcr vary with the specific speed type and, to some extent, with the individual impeller
and casing design. They are also affected by the physical size of the pump. Figure 18.47 shows a <I>-N,
relation and a Kcr-N, relation that are representative for normal pump design. Like the DJDJ-N, curve,
these curves show average values; the true values for an individual design will vary somewhat from
those shown.

Approximating Characteristics From


Physical Measurements
Generally, a centrifugal pump user who wishes to determine the performance characteristics of a given
pump in his possession has a large reservoir of information from which these data can be determined:
Heads, Conditions of Service Performance Characteristics, and Specific Speed 457

1. A copy of the order on which the pump was purchased.


2. The nameplate of the pump. This nameplate generally carries the pump shop serial number, the manufacturer's
type designation, and the rated conditions of service, including the operating speed.
3. The pump driver nameplate. If this nameplate is missing, a direct measurement on the driver will give the
operating speed.

If the pump make and serial number are known, it is a simple matter to get the desired information
from the manufacturer. If the make, but not the serial or other identifying number or letter is known,
most manufacturers can identify the pump type and impeller design, if given the following: (1) the
nozzle sizes, (2) a sketch showing the external appearance and dimensions of the pump, and (3) the
major impeller dimensions.
It is seldom that a pump is without both pump and driver nameplates and that no record of the
purchase is available, leaving the physical presence of the pump itself as the only thing certain. But
even in this case it is possible to carry out certain measurements and calculations to obtain the desired
information with some degree of accuracy.
Two separate phases exist in the problem of estimating the performance of a centrifugal pump when
nothing is known except the physical dimensions and proportions. The first phase of the problem is the
theoretical aspect, which gives a reasonable approximation of the pump head-capacity curve at any given
operating speed and the expected power consumption of the pump. This phase requires the application
of the data presented on the preceding pages and in Fig. 18.31-18.37 and 18.46-48 for its solution.
The second phase of the problem concerns the practical aspects of the application of the pump to a
particular service. Is the physical design of the pump suitable for the power and the speed selected? Is
the casing design suitable for the operating pressure? Will the pump operate satisfactorily under the
suction conditions contemplated?

MAXIMUM SPEED LIMITATION

Present-day practice in the United States places a limit on pump speed by limiting the suction specific
speed for a given class of service and pump (see Chap. 19). A much earlier approach was to determine
the maximum rotative speed of an impeller by limiting the peripheral velocity of its suction eye (D 1) to
a certain value, depending on the suction conditions. This velocity can be established exactly, as in the
case of the peripheral velocity of the outside impeller diameter:

· I'
SuctlOn eye ve OClty = 19.1Din. . .
X 10310 metnc umts

· eye ve Ioelty
SuctlOn 229 10 US umts
' = Din. .

A reasonable maximum peripheral velocity for the suction eye is shown in Fig. 18.48 for various
suction conditions.

EXPECTED POWER CONSUMPTION

Determining the possible head and capacity of a pump would be of little value if the power required to
drive the pump could not be predicted. Data on the preceeding pages show the approximate maximum
efficiency for both single- and double-suction pumps that can be obtained with present-day designs.
458 Heads, Conditions of Service Performance Characteristics, and Specific Speed

FEET OF SUCTION HEAD AT SEA LEVEL AND 85- F

10
-20 -I' .10 -5 o .,
~
10
00
>Z
....
_u 0
uw
0(1)
10 ....
...1« .JI
1£11£1
>CL
...I .... ......
~ "
C(w 50
«1£1 10'
W~
I Z .......
CL_ V
ii . V
wllJ

"
CL>
IIJ 40 ./
~Z
=>0
~-
-~
Xu
C(::>
~(I)

50
10
I' 10 10
I' 40

NET POSITIVE SUCTION HEAD. IN FEET

Fig. 18.48 Approximate limit for peripheral velocity of suction eye and required NPSH.

How close an existing pump would approach these values would depend on the individual pump.
Multistage pumps, with abrupt crossover passages from one stage to the next, should be expected to be
less efficient by two to three percentage points.

DETERMINATION OF SAFE POWER INPUT

The shaft of a centrifugal pump is subject to both bending and torsional stresses. Usually, its smallest
diameter is at the coupling and this section is subjected primarily to torsional stress only. Generally,
centrifugal pump shafts are designed not to exceed a torsional stress at the coupling of 48 MPa (7,000
psi) for safe maximum continuous loading. Therefore, in predicting pump performance, a check should
be made to see if the maximum expected power at the selected speed keeps the torsional stress within
the recommended 48 MPa (7,000-psi) limit. If not, the pump is not mechanically safe for operation at
the selected speed. The relation between transmitted horsepower, speed, shaft diameter, and permissible
torsional stress is given by the formula

kW SntP. . .
= 48.6 x H)6 m metric uruts
h SntP. US .
p = 321,000 m uruts
Heads. Conditions of Service Performance Characteristics. and Specific Speed 459

where kW = power in kilowatts


hp = horsepower
S = allowable stress, MPa (psi)
n = rotative speed, rpm
d =shaft diameter at coupling, mm (in)

DETERMINATION OF SAFE OPERATING PRESSURE

Detennination of the safe operating pressure of a given pump requires a very detailed study. Some idea
of a possible maximum can be obtained by examining the size and drilling of the discharge nozzle
flange. This is an indefinite limit, especially in the United States where the "125-lb flange," which is
good up to 175-psi hydraulic operating pressure, is generally used for all pressures below that value.
Many pump designs using such flanges, however, are not good for operating pressures that high.
It is more satisfactory to make a check of safe pressures at the bolting of the casing or casing heads.
In actual design studies, maximum safe bolt stresses are calculated with full knowledge of the areas
subjected to internal pressures, of the maximum expected hydraulic pressure, and of the forces required
to compress joint gaskets. A good approximation can be obtained by limiting the bolting stress at the
root of the threads to 35 MPa (5,000 psi); thus, safe working pressure can be computed as follows:

A,nb 35 x 10 1 • • •
swp = A ill metnc umts
A,nb 5,000. US .
swp = A ill umts

where swp = safe working pressure, bar (psig)


A, = root area of bolt, cm2 (in2)
nb = number of bolts
A = area subjected to hydraulic pressure, cm2 (in2).

For purposes of analysis, assume a 20-year-old double-suction, single-stage centrifugal pump with
an 8-in. discharge and lO-in. suction, both with 125-lb flanges. Various relevant dimensions (see Fig.
18.46) are given below for the impeller:

D2 = 311 mm (12.25 in)


DI = 156 mrn (6.12 in)
W2 = 52.5 mm (2.06 in)
Z2 = 7
S2 = 10 mm (0.38 in)

The shaft diameter is 39.6 mm (1.56 in) at the coupling, increasing to a maximum at the impeller.
The area of the horizontal split is 1290 cm2 (200 in2). The pump casing is held together by 21 qty 19 mm
(0.75 in) studs and bolts. The intended drive is a 60-cycle induction motor.

DJDI = 311/156 = 2.0


Referring to Fig. 18.47, a DJD I ratio of 2.0 indicates an impeller type with specific speed Ns = 1,700
if single suction or Ns = 2,400 if double suction. The head and capacity constants are
Heads, Conditions of Service Performance Characteristics, and Specific Speed

"'=1.06
Kcr = 0.13

Assuming that the pump will be applied to an installation involving a 4.6 m (I5-ft) suction lift at sea
level, handling cold water, the NPSH will be, roughly, 5.2 m (17 ft). From Fig. 18.48, it can be established
that the maximum safe peripheral velocity at the impeller suction eye is about 15.2 rn/s (50 ft/s). Solving
for the rotative speed gives:

= 15.2 x (19.1 x 1()3) = 1 860


n 156 ' rpm
50 x 229
n= 6.125 = 1,860 rpm

Therefore, 1,750 rpm will be the maximum possible rotative speed with a 60-cycle motor.
The pump head can now be calculated at 1,750 rpm:

311 (1750) ]2
H = [ 84.6 x 1()3(1.06) = 36.8 m

H = (12.25 X 1,750)2 = 121 ft


1,840 x 1.06

Using the formula in Fig. 18.46, A 2, the discharge area of the impeller, can be calculated:

A 2-- 52.5 [1t(311)


102
- (7)10] _ 476
- cm
2

= 2.06[1t(12.25) - (7)0.38] = 74.0 in2

The pump capacity is estimated:

Q = 1.60(0.13)476(36.8)°·5 = 600 m 3Jhr


= 25 (0.13)742(121)°·5 = 2,645 gpm

When operated at 1,750 rpm, this pump will deliver 600 m 3/hr (2,645 gpm) against a total head of
36.8 m (121 ft) at its best efficiency point. Its specific speed can be recalculated:

N. = 1,750(600)°.5/(36.8)°·75 = 2,865/1.16 = 2,470 (double suction)


= 1,750(2,645)°·5/(121)°·75 = 2,470

For an older pump, Fig. 18.30 shows that, for this specific speed and capacity, the maximum efficiency
would be somewhat over 85 percent. If 85 percent is used, the power consumption would be

power = 600(36.8)1.0/368(0.85) = 71 kW
= 2,645(121)1.0/3,960(0.85) = 95 bp

Working from these values of capacity, head, efficiency, and horsepower, and using the percentages
in Figs. 18.35 to 18.37 (for a single-suction N. of 2,470 x 0.707 = 1,746), the points in Table 18.4 can
be obtained. Plotting these points, the approximate curve shown in Fig. 18.49 is obtained. (The actual
test curve of the pump has been superimposed on the same graph for comparison.)
Heads, Conditions o f Service Performance Characteristics, and Specific Speed 461

Table 18.4 Characteristics Predicted From Impeller Measurements

Capacity Head Power Efficiency


~-- -~- - - - -- - ~- - -- - - -- -
- - ---- - --
Percent
Percent mlfhr gpm Percent m ft Percent kW bhp of max. 11

0 0 0 117 43.0 142 54 38 51 0 0


25 150 661 63 45 60 46 0.39
50 300 1,322 115 42.3 139 75 53 71 76 0.65
75 450 1,987 110 40.5 133 88 62 84 94 0.80
100 600 2,645 100 36.8 121 100 71 95 100 0.85
110 660 2,915 93 34.2 112 103 73 98 99 0.84

1.0
COI'~5
I"ttEDlCT£O I
----- ACTUAL TUT CUIND

--- -- --
150

~
w 140 - I I
...
w -...;;
~
~
HEAD- CAPACITY

~ 150
Q
c
'"........ .... ,.............. i'-...
w 1 20
........

"-
:J:
oJ
C
I- 1 10
" ,
0
I- ~
100

- --
ItO

eo
[fI1'1C1!NCY _.... ,,/'" -:;:.
-~

10
,. ,. ~
/'" '" ,/

------
.0
I-
Z /
/
/ V 100

7- -- -- --
w
,/' V
50

...-:: --
u
..,
v:- -- ... --
II:
~ eo
... --:::..
40
.e::::
Go
>
U
30 / _Mit
60
~ ,... Go
:J:
e:
III
20 40 GI

10
/; 20

0 I o
o 10 IS 20 30
CAMCITY. IN 100 GPM

Fig. 18.49 Comparison of characteristics predicted from impeller measurements and actual characteristics
determined by test.
Dash lines are predicted curves, and solid lines are actual test curves.
462 Heads, Conditions of Service Peiformance Characteristics, and Specific Speed

To detennine if the pump shaft is suitable for this application:

Safie power -- 48(1750)(39.6)3


48.6 x 106 -
- 107 kW

7,000 x 1,750 X 1.5623 = 145 h


321,000 P

The shaft is obviously safe for 1,750-rpm operation.


It remains to check the pump for safe operating pressure. The 19 mm (0.75 in) bolts and studs holding
the two halves of the casing together have an area of 1.95 cm2 (0.302 in2) at the root of the threads.

. 35(1.95)21 X 101
Safe working pressure = 1290 = 11.1 bar
5,000 x 0.302 x 21 _ 159 .
200 - pSlg

This is approximately 2.6 times the expected shutoff head of 43/10.21 or 4.2 bar (142/2.31 or 62 psi),
when operating at 1,750 rpm. The casing will be strong enough for the intended operation. (The actual
pump design is good for 12 bar (175-psig) operating pressure.)
The foregoing example has demonstrated that it is possible to approximate the perfonnance of
centrifugal pumps other than those of the axial-flow impeller type. Nevertheless, this should be used
with caution and only in the absence of a more reliable method.

Rating Curves and Charts


Rating curves and rating charts were originally intended for pump salesmen to use for making pump
selections. They are now also common in bulletins and other sales literature.
A rating curve for a centrifugal pump of specific design shows in a condensed fonn, the possible
range of applications of that pump, either for a range in speed or for a range in impeller diameter. In
earlier days, some small centrifugal pumps for belt drive were manufactured in lots for stock sale, and
the most efficient operating speed for each installation was obtained by selecting a proper pulley ratio
to give the head and capacity condition desired. A curve showing the head, capacity, and brake horsepower
for such a pump at a number of speeds, could be utilized for detennining the speed necessary and the
power involved, but these rating curves were rarely used. Instead, a table (Table 18.5) was generally
more convenient and pennitted showing a number of pump sizes on the same sheet.
Before the recent energy crisis a few lines of pumps, notably small motor-driven stock units, were
made with several different impeller diameters, of the same or different patterns, that load up various
sizes of motors. A rating curve for this type is shown in Fig. 18.50. With such a line of pumps, a pump
with a 19 kW (25-hp) motor and an impeller that would approximate the results shown on curve (2)
would be furnished if the desired head condition fell anywhere within the zone between the head-capacity
curves (2) and (3) in Fig. 18.50. Thus, for some customers' requirements, such as 57 m 3/hr (250 gpm),
61 m (200 ft) total head, the pump supplied would produce more capacity or head than required. When
installed, this unit would either give excess capacity or excess pressure, depending upon the system,
Table 18.5 Portion of Pump Rating Chart

Total head, in feet


gpm 20 25 30 35 40 45 50 55 60

Size-type 2V2-CF-l 2V2-CF-l 2V2-CF-l 2V2-CF-l 2V2-CF-l 21/2-CF-l 2V2-CF-l 2V2-CF-l 21/2-CF-l
hp 1.6 2.0 2.4 2.8 3.25 3.75 4.2 4.65 5.05
rpm 1,010 1,090 1,170 1,240 1,310 1,380 1,440 1,500 1,550
200
Size-type 3-CF-l 3-CF-l 3-CF-l 3-CF-l 3-CF-l 3-CF-l 3-CF-l 3-CF-l
hp 2.25 2.6 3.0 3.3 3.6 4.1 4.6 5.0
rpm 830 890 945 1,000 1,048 1,095 1,140 1,185
- - - -- - - - - -- - -- - -

Size-type 2V2-CF-l 2V2-CF-l 2V2-CF-l 2V2-CF-l 2V2-CF-l 2V2-CF-l 2V2-CF-l 2V2-CF-l 21/2-CF-l
hp 1.9 2.3 2.7 3.2 3.65 4.15 4.7 5.1 5.55
rpm 1,055 1,130 1,205 1,280 1,345 1,410 1,470 1,530 1,585
225
Size-type 3-CF-l 3-CF-l 3-CF-l 3-CF-l 3-CF-l 3-CF-l 3-CF-l 3-CF-l
hp 2.6 2.9 3.3 3.6 4.0 4.6 5.0 5.5
rpm 850 907 962 1,015 1,065 1,113 1,155 1,205
- - -- - - - -- - ----------- -----------~--------------- -- - -- - -- - - - -- ---- -- --- - ------- ---- -- - - ----------- ---~-------

Size-type 2V2-CF-1 2V2-CF-1 2V2-CF-1 2V2-CF-1 2V2-CF-l 2V2-CF-l 21/2-CF-l 2V2-CF-1 2V2-CF-1
hp 2.2 2.6 3.1 3.6 4.1 4.6 5.1 5.6 6.2
rpm 1,095 1,170 1,245 1,315 1,385 1,450 1,510 1,565 1,620
250
Size-type 3-CF-1 3-CF-1 3-CF-1 3-CF-1 3-CF-1 3-CF-l 3-CF-1 3-CF-l
hp 2.9 3.2 3.6 4.0 4.4 4.9 5.5 5.9
rpm 870 927 980 1,035 1,085 1,130 1,175 1,220

e
464 Heads. Conditions of Service Performance Characteristics. and Specific Speed

unless throttling were employed to increase the frictional head artificially. Today, this practice has
essentially disappeared.
For pumps which are built-to-order with an impeller pattern and diameter individually selected for
the prevalent service condition, a curve showing the range in conditions that can be met by a given
impeller design or by several impeller designs for a given speed is used. These are generally complicated
in appearance because the efficiency that can be obtained varies with the diameter of the impeller. This
variation in efficiency is covered either by isoefficiency curves, as shown in Fig. 18.51 or by figures on
the curves of similar points (Q/Qt = -VH/Ht relation) or lines approximating that relation as shown in
Fig. 18.52. For their proper use, rating curves must also show the required minimum NPSH shown in
Fig. 18.52.
A different chart is required for each motor speed for which the particular pump may be offered. For
unusual conditions of driver speeds not covered by a curve, the use of standard relations for speed
changes permits determining what the pump will do.

EFFECT OF SPECIFIC GRAVITY ON PUMP CHARACTERISTICS

The only effect that specific gravity of liquids with viscosities equal to water has on the operation of a
pump is to vary the power required to drive it. The capacity and head (measured in feet of liquid) are
the same as for water and so is the efficiency. The power input for any capacity is that required with
cold water multiplied by the specific gravity.

EFFECT OF VISCOSITY ON PUMP CHARACTERISTICS

Two of the major losses in a centrifugal pump are through fluid friction and disk friction. These losses
vary with the viscosity of the liquid so that the head-capacity output, as well as the mechanical input,
differ from the values produced when handling water.
It is not practical to present here a complete discussion on viscosity as a property of liquids and on
the effect of viscosity on flow of liquids. The reader can find such a discussion in textbooks on fluid
mechanics. It is necessary, however, to know the three different units that may be encountered describing
the viscosity of a specific liquid:

1. Saybolt seconds universal, or ssu


2. Centistokes-defining the kinematic viscosity
3. Centipoises-defining the absolute viscosity.

Data for the conversion from one to another of these units and relations between the viscosity and
temperature of a number of liquids are given in the Data Section of this book.
Considerable experimental testing has been done in determining the effect of liquid viscosity on the
performance of different centrifugal pumps. Even with extensive data on the effect of viscosity, it is
difficult to predict accurately the performance of a pump when handling a viscous fluid from its perfor-
mance when handling cold water. The Hydraulic Institute has published the charts shown in Figs. 18.53
and 18.54 which permit approximating the characteristics on uniform liquids (not paper stock, slurries,
or the like) of conventional single-stage centrifugal pumps, not of the mixed-flow or axial-flow
types, up to about 8 in discharge. These charts can also be used for multistage pumps if the
Heads . Conditions of Service Performance Characteristics. and Specific Speed 465

r-
I-
,,..
Jo-o.I.

r- r- ~
""'" r"-po."
JII r r"- r--
~ 2~ ,.... ""'"
......
l - t-... r" r-....

,.
-
r"-
'" r--
- ...30 HP MOTOR
~ J-, 1"'"
ci 1 3 L
..... ..... ~

""
<I
I'.
~ 2vv
I""- .....
..J
<I ..... i'oo..
.....
o I""- ..... ~ I-Z5 I-IP MOTOR
.....
r-..
'"
_ r-ZO HP MOTOR
5C)
I~

100 200
CAPACITY. IN GPM

Fig. 18.50 Rating curve of 2V2-in motor mounted pump.


Pump equipped with different impellers that load up several sizes of motors.

I~ ':':-r-~
250 ! ~I!-~~-~il
I.1:-- ;:1
"-. 1-. uFo-
I'-! i'-o ". "-r .':I)
• P- f-I..
rr 1/ r::: r- h
f- .
h V {) I-r:ttil
r-",
P[RCEHTAQES
N / V 7 f'")I V r---~ j j

00 r.L"
~ ",i
Vr" V :
~ Io.i .....

r, ~I'
. . ~ '! /~.; ..
I ",'" i"

""
1 r--1. ) V II l'~ -
/ II V. . . . .L : .. K, 1/ / "" I"v...,
I I r-- "~II ~, /
...... 1 I
J
" 1' ...
V L~ ./ r--.~
50 ..... ~ .... 2 .iC. V V:' v" r~ ...
~ "./ V. . . ~ l.r' i
!~

t"'S ~ '" '" I~


~ ....~
vv
20
CAPACITY, IN 100 G PM

Fig. 18.51 Rating curve of 10 in double-suction single-stage pump.


Revealing the wide range that can be covered by an impeller of single design by machining it to the proper
diameter for a particular service.
466 Heads, Conditions of Service Performance Characteristics, and Specific Speed

I-
W
W
IJ...
Z
cl I f.I '" /
<X
w
:x:
...J I I /, / / ~,() -
5 11 I / ,
~ ~
o
I-

250

20

-
15 ,.... .....

--
,.",.
10 ..... .-,.",.

5
o -
50 100 150 200 250
CAPACITY, IN GPM

Fig. 18.52 Rating curve of 2 in discharge, 3,500-rpm pump.


1.00

.90
Cif .-::
I-- ~
.eo
CIl
0: .70 r-F -=
0
b .60
I
~
Z .50
0
i= "\..<;;
(J
.~
W
0: I-
0:
0 .30
(J l"-
..20 t- I

- r--
.10

.0

10 I~ 20 25 30 40 50 60 10 eo go 100
CAPACITY-GALLONS PER MINUTE (at SEP)

Fig. 18.53 Performance correction for viscous liquids.


Pump design capacities up to 23 m31hr (100 gpm).
(Courtesy Hydraulic Institute)

467
468 Heads, Conditions of Service Performance Characteristics, and Specific Speed

100

90 ~~ ;:::::.. ely
0
< T~ ~ ~
"
C/) w 80 I
cr J: I
I ~ I'
'-O.8.Q.
-O.8.Q.
0 70
I "- ' -I.O.Q.
~ I 1'-12'Q.
U 60
I
~ 100
-r--. i'-.-
>
U
Z 90
- .........
K r-.....
Z w I"'-- ....... I ~9
0 u 80
Li:
10..
~ w 70
't--c '('
I
I \
U
W 0 60
f-
~ \
cr z
c
1'\ \
cr 50 I
0 > I \
U t: 40
U

~ 30
I
I \
u i \
20
~ "(,~ '4 t.t,. ;~ b~~'~ ~
,~ ,~~
~ \~
~ b~
CENTISTOKE'S
\'\ 1\ \ \ \ 1\\ 1\ 1\1\ 1\ ~ ,\ &00
'\ 1\\ \\ 1\ \ \1\ 1\ \ \ i\ \\ \ v
400
300
'\ \ 1\\\ \ 1\ 1\\ \ 1\1\ \ '\ l\\r\\ ~ v-
200
1\ v-:: ~ I--': v
I-- ~
Irsso
1\ \' ~\\ \ I~ \ 1\:\ \~~~ [\; ~ ~ ~ ~ \[\
\
~ ~~~ ~ ~ v- ~I--
40

\ \ \ \ 1\ \~Ps;.- ~ ~ ~ ~ ~
1\ \ \
:;....-

\ \\ \ ~ ~~ ~flI~ ~ ~ ~ \
\
~ ~ ~
\ \ 3 ~ ?(\
~~ ;::;:. ~
\\ \ \

.- ./"':: ~~ ~~
~~~ '\ \\ \ \\ \
W ~ t:/:: ~ ~ tx v~ \ \ .\ 1\\ _\ 1\ \ 1\ \
~ ~ ~ ~ ~ ~ 11\ \ \
&00
W
~
400
300 \\ \\ 1\ \ \ 1\ 1\
Z
200
188 ~ ~~ V \ 1\f\1: 1\ 1\ \ \ 1\\r\\ 1\\ 1\ 1\ l\ 1\
!830 ~ \ 1\1\1"\I I\l\\ \\ l\\' \\ 1\ 1\ \ \\ 1\
a t::Y \
~
4(
W
I'
20
J:
"\
"""'

VISC~I"?, - s~u 0 ~I( p~\ ~~


I
CI!~,:< ~
oc Ibo & ~ \ \~ ~ ~,
L5 2 .]" 5 II 7 II 1110 15 20 30 40 50 110 1011011000

CAPACITY IN 100 GPM

Fig. 18.54 Perfonnance correction for viscous liquids.


Pump design capacities above 23 mllhr (100 gpm).
(Courtesy Hydraulic Institute)
Heads, Conditions of Service Performance Characteristics, and Specific Speed 469

correction factors are selected on the basis of the head per stage, and provided the losses (which
result in heating the liquid) do not cause sufficient increase in the temperature to change the viscosity
of the liquid appreciably. The correction factors in Fig. 18.54 are selected for the head (per stage)
and capacity at which the pump gives maximum efficiency on cold water. For example, a pump
whose maximum efficiency capacity (1.0 x Qn) was 170 m3jhr (750 gpm) at a 30.5 m (100 ft)
total head on water would, on a 1,000-ssu viscosity liquid, have the following characteristics: (1)
A reduction of capacity to 95 per cent of its corresponding water capacity, (2) a reduction in the
head produced at these reduced capacities to 96, 94, 92, and 89 per cent of the cold water heads
at 60, 80, 100, and 120 per cent of normal capacity respectively, and (3) a reduction in the efficiency
to 63.5 per cent of that produced on water for the corresponding capacities. The power required
to drive is determined by calculating by the formula:

kW QH x specific gravity . . .
= 36811 m metnc umts
h _ QH x specific gravity. US .
P- 3,96Ort m umts

in which Q = capacity, m3jhr or gpm (corrected for viscosity)


H =total head, m or ft (corrected for viscosity)
11 =efficiency (corrected for viscosity).

Applying these corrective factors to a pump whose cold water characteristics are identified in Fig. 18.55
by 32 ssu, the approximate performance for 100-,400-, 1,000-,2,000- and 4,000-ssu liquids have been
developed, the values for bhp being calculated on basis of 1.0 specific gravity. Whereas the pump
produced a maximum efficiency of 76 percent when pumping 100 m3jhr (440 gpm) of cold water against
a 40.2 m (l32-ft) total head, it would be expected to produce a maximum efficiency of only 19.7 percent
when pumping 73 m3/hr (321 gpm) of a 4,000-ssu liquid against a 31.2 m (102 ft) total head.
In applying regular cold-water pumps for use in pumping viscous liquids, care must be taken to make
sure that the shaft design is strong enough for the required power, which may be considerably in excess
of the cold-water brake horsepower, even though the specific gravity of the liquid may be less than that
of water.

EFFECT OF AIR OR GASES IN THE LIQUID ON


PUMP CHARACTERISTICS

If as little as 1 percent by volume of air or gases is present in the liquid pumped, the head-capacity
curve is affected very significantly, as illustrated in Fig. 18.56. As this percentage increases, the unfavor-
able effect becomes even more drastic, until at 6 percent by volume for the pump illustrated in Fig.
18.56, we reach a condition when the pump almost ceases to perform satisfactorily.
The presence of air or gases creates a fairly complex phenomenon and, at this moment, cannot be
described by an exact mathematical relation that could be applied indiscriminately to any given pump.
Figure 18.57 illustrates another test of the effect of air content on the performance of a pump. The
deterioration in this case appears to be less severe than in the case of the pump illustrated in Fig. 18.56.
The reason for the difference is that a large number of variables in the geometric configuration of the
impeller and of the casing playa major role in the extent of the deterioration. But the exact relationship
between design and performance still remains an unknown.
470 Heads, Conditions of Service Peiformance Characteristics, and Specific Speed

180

160
0
CI
ILl
l: 140
..J
CI
I-
0 12 0
l-
I-
ILl
ILl
10 0
LL

80

60
0
80

70

60 I-
EFFICIENCY Z
ILl
()
50 Q:
ILl
Q.

~
40 ()
Z
ILl

v 30 (3
~
I&.
/ 20
ILl

.... 5 "v
/

--
>- I
+++-+-+-+-+-+-+-+-+-r-r-r-r-r-~.~OOOSSU~-r~~~+
--
l-
V --, I
~4 0 BHP o
Q:
C) 2000 SSU
~ 30 , ,J..06 SUr-
,v . 'SU
Q.
l:
CD
20
-- 400 SSSU-
'O~2 S~

o 100 200 300 400 500 600


CAPACITY, IN GPM

Fig. 1S.55 Predicted characteristics for a centrifugal pump for liquids of various viscosities.
Heads, Conditions of Service Performance Characteristics, and Specific Speed 471

100
"-
90
I-
W
w
u. 80
ci
«
w 70
:x:
...J

~ 60
0
l-
SO
6%

250 500 750 1,000


CAPACITY, GPM

Fig. 18.56 Effect of entrained air on the head-capacity curve of a centrifugal pump.
(Courtesy Chemical Processing, June 1987.)

25
M ,, O%AIR

10%AIR
.......
r-....
" r\
\.
\
,
20
'" "-
~
~

\
", ,
l

,
15 \
\
\
\

10

9 11 13 M3/H x 1,000

Fig. 18.57 Another example of the effect of entrained air on centrifugal pump performance.
(Courtesy Sulzer)
472 Heads, Conditions of Service Performance Characteristics, and Specific Speed

BmLIOGRAPHY

[18.1] E. P. Sabini and W. H. Fraser. ''The Effect of Specific Speed on the Efficiency of Single Stage Centrifugal
Pumps." Proceedings of the Third International Pump Symposium, Houston, Texas, May 1986.
[18.2] E. P. Sabini and W. H. Fraser. ''The Effect of Design Features on Centrifugal Pump Efficiency." Proceedings
of the Fourth International Pump Symposium, Houston, Texas, May 1987.
19
Suction Conditions and Limitations on
Suction Performance

Every piece of machinery is preordained to have its Achilles' heel. That of the centrifugal pump can
generally be found in its suction. This fact must have been discovered quite early in the commercial
application of centrifugal pumps and, probably, accepted as an inescapable penalty exacted in return for
the advantages that were made available by this newer means of raising water to a desired location at
a reasonable cost. And because most centrifugal pump troubles occur on the suction side of the pump,
it is imperative to understand how to relate the suction capability of a centrifugal pump to the suction
characteristics of the system in which it will operate.
Most difficulties at the suction of centrifugal pumps have as their root cause the phenomenon known
as cavitation: the process whereby the pressure at some point in a pump or turbine falls below the
liquid's vapor pressure thus allowing local vaporization. The resulting vapor can have two effects: a)
its extent can be sufficient to obstruct the machine's waterways leading to a deterioration of hydraulic
performance, or b) the subsequent collapse of the vapor as it moves into regions of higher pressure can
produce noise and erosion of the waterway surfaces in the vicinity of the collapsing vapor. Recognizing
this, it is therefore important in the pumping of liquids to ensure there is always sufficient energy
available at the pump suction to move the liquid from the pump suction nozzle into the impeller
without vaporization to the extent that pump performance deteriorates noticeably or pump parts are
damaged prematurely.
Pump performance is normally presented in the form of curves, such as in Fig. 18.16, showing head
and power versus capacity at fixed speed. The head a pump will produce and the power it will absorb
at various capacities are two of its three fundamental performance characteristics. The third is the net
positive suction head (NPSH) required. This is the energy in meters or feet required at the pump suction
over and above the vapor pressure of the liquid to permit the pump to run without undue deterioration
of performance or expected life. As simple as this definition seems, the practical application of it is
actually quite complicated, with the pumped liquid, impeller energy level, and impeller material having
a profound effect on which of these two criteria determine the NPSH required; see Definition of NPSH
required later in this chapter.
Considering deterioration of performance first, changes in available NPSH do not affect pump perfor-
mance materially, as long as it remains above the value of required NPSH. The characteristics labeled

473

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
474 Suction Conditions and Limitations on Suction Performance

J:
(J)
a.
z
~ HE~. ___ ·~·!
I ~L---AVAILABLE
ou::
UJ •
1
'~IN~UFFICIENT NPSH=
tt AMPLE NPSH N~SH REQUIRED
~r-----:i~::lA~V~A~'LA~B~LE~N;P;S:H~=~R:E:Q~U~/t---~~~____ ~
J: INSUFFICIENT NPSH RED

NPSH REQUIRED

CAPACITY

Fig. 19.1 Effect of cavitation on pump performance.

"Ample NPSH" in Fig. 19.1 show such performance. Lowering the available NPSH to the required value
causes cavitation sufficient to degrade performance and the pump "works in the break" (Fig. 19.1).
Running in this condition, a decrease in system head will produce only a small increase in flow, to the
point where the available NPSH is "insufficient" and the head developed "breaks down". At this condition,
lowering the system head further produces no increase in capacity.
In the early days of pump application, the term NPSH was not used frequently, and the suction
characteristics of centrifugal pumps were generally expressed in terms of "suction lift" and "suction
head," as defined in Chapter 18.

NET POSITIVE SUCTION HEAD

The use of "permissible suction lift" or of "required suction head" has very definite shortcomings. First,
it can only be applied to water, since it refers to the energy of barometric pressure expressed in feet of
water. Second, changes in barometric pressure, whether caused by differences in elevation above sea
level or by climatic conditions, affect the value of these two terms. Finally, changes in pumping
temperature affect these values as well, since they affect the vapor pressure of the liquid. For this reason,
all references to suction conditions today are made in terms of NPSH (net positive suction head above
the liquid vapor pressure).
Both suction head and vapor pressure should be expressed in meters (feet) of liquid being handled,
and must both be expressed either in gage or absolute pressure units. A pump handling 17°C (62°F)
water (vapor pressure of 0.18 m [0.6 ft]) at sea level, with a total suction lift of 0 m (ft), has an NPSH
of 10.34 - 0.18 or 10.16 m (33.9-0.6, or 33.3 ft), whereas one operating with a 4.6 m (15-ft) total
suction lift has an NPSH of 10.34 - 0.18 - 4.6, or 5.56 m (33.9-0.6-15, or 18.3 ft).
A pump operating on suction lift will handle a certain maximum capacity of cold water without
Suction Conditions and Limitations on Suction Performance 475

hj=3FT

1
15 FT

j !
ATMOSPHERIC
PRESSURE

water, BO°F
PVP = 0.5 PSIA

(a) AT SEA LEVEL


NPSHA = (14.7 - 0.5) x 2.31 -15 _ 3
1.0
= 32.B - 15 - 3 = 14.B FT

(b) AT 5,000 FT ABOVE SEA LEVEL


NPSHA = (12.2 - 0.5) x 2.31 -15 _ 3
1.0
= 27 - 15 - 3 = 9.0 FT

Fig. 19.2 NPSHA for suction lift at sea level and at 1,525 m (5,000 ft) above sea level.

cavitation. The NPSH or amount of energy available at the suction nozzle of such a pump is the
atmospheric pressure minus the sum of the suction lift (including friction) and the vapor pressure of the
water. To handle this same capacity with any other liquid, the same amount of energy must be available
at the suction nozzle. Thus, for a liquid at its boiling point (in other words, under a pressure equivalent
to the vapor pressure corresponding to its temperature), this energy has to exist entirely as a positive
head. If the liquid is below its boiling point, the suction head required is reduced by the difference
between the pressure existing in the liquid and the vapor pressure corresponding to the temperature.
It is necessary to differentiate between available NPSH and required NPSH. The former, which is a
characteristic of the system in which a centrifugal pump works, represents the difference between the
existing absolute suction head and the vapor pressure at the prevailing temperature. The required NPSH,
which is a function of the pump design, represents the minimum required margin between the suction
head and the vapor pressure. Figures 19.2, 19.3, and 19.4 illustrate the manner in which available NPSH
at a given capacity should be calculated for a typical installation with a suction lift, for a pump taking
its suction from a tank, and for a pump handling a liquid at the boiling point, respectively.
Both the available and required NPSH vary with capacity (Fig. 19.5). With a given static pressure
or elevation difference at the suction side of a centrifugal pump, the available NPSH is reduced with
increasing capacities by the friction losses in the suction piping. On the other hand, the required NPSH,
being a function of the velocities in the pump suction passages and at the inlet of the impeller, increases
basically as the square of the capacity.
476 Suction Conditions and Limitations on Suction Performance

WATER 80°F
PV=0.5 PSIA
r
10 FT

=4FT
1
NPSHA = (14.7 + 5 - 0.5) x 2.31 + 10 _ 4
1.0
= 44.3 + 10 - 4 = 50.3 FT

Fig. 19.3 NPSHA for pressurized suction.

N-BUTANE
PV = 52.2 ____
PSIA""""'-
SPGR = 0.56
r
10 FT

1
NPSHA = (37.5 + 14.7 - 52.2) x 2.31 + 10 _ 2
1.0
=0+ 10-2=8 FT

Fig. 19.4 NPSHA for liquid at boiling point.

A great many factors, for example, eye diameter, suction area of the impeller, shape and number of
impeller vanes, area between these vanes, shaft and impeller hub diameter, impeller specific speed, the
shape of the suction passages, all enter in some form or another into the determination of the required
NPSH. Different designers may use different methods to produce an impeller that will perform satisfacto-
rily with a specific value of required NPSH. As a result, it is not recommended that users attempt to
Suction Conditions and Limitations on Suction Performance 477

T z

~
B

NPSH = Z + (Ps - Pvp ) - (hfs + h/~


ALL UNITS IN FEET

CAPACITY

Fig.19.5 Available and required NPSH.

estimate required NPSH from the knowledge of just one or two of these factors. Instead, they should
base their selections on the data provided by the manufacturers.

HISTORICAL BACKGROUND

In the early days of centrifugal pumps, limitations on permissible suction lifts were imposed by the
manufacturers on strictly empirical grounds. Because the understanding of the phenomenon of cavitation
was imperfect, these limitations were sometimes overconservative and sometimes quite optimistic. How-
ever, the commercial pressures of a very severe competitive situation seem to have outweighed sound
engineering judgment much too often in the 1920s. The attendant difficulties could not have failed to
478 Suction Conditions and Limitations on Suction Performance

stimulate a considerable amount of head-scratching by centrifugal pump designers. The thinking of the
designers must have been heavily influenced by the experience of water turbine builders. In 1922, at
the Hydroelectric Conference held at Philadelphia, H. B. Taylor and L. F. Moody first presented the
concept of a parameter, Sigma, to facilitate the description of the conditions under which cavitation
occurs. Sigma was defined as

. NPSH
Sigma = (J =- -
H

where H is total head. At about the same time, Dr. Thoma was developing the same concept in Germany
and, therefore, "Sigma" has since then been known to centrifugal pump designers as the Thoma-
Moody parameter.
Means were now available to relate the operating conditions of a centrifugal pump-its capacity,
head, and rotating speed-to the minimum net positive suction head required for satisfactory operation.
But commercial pressures seem to have again outweighed sound engineering judgment much too often
in the 1920s. The number of companies manufacturing centrifugal pumps proliferated without there
necessarily being a corresponding increase of knowledgeable and experienced designers. Spurred on by
the advantage of offering a higher operating speed than the competition or of guaranteeing satisfactory
operation with higher suction lifts, some companies made installations that had disastrously expensive
consequences for user and manufacturer alike. As field difficulties continued to occur, it became evident
that some official guidelines were necessary. An organization for the discussion and solution of technical
problems in the realm of pumping machinery had been formed some years before by the older, larger,
and most reputable pump manufacturers under the name of the Hydraulic Institute. Standards had been
developed and published by this institute to codify sound practices, including testing and guarantee prac-
tices.
At this point, the Hydraulic Institute (HI) appointed a committee to investigate centrifugal pump
suction problems. The committee proceeded to collect information on centrifugal pump installations in
which cavitation troubles had been experienced as well as on satisfactory installations. It was found that
to avoid difficulties for any given total head and suction lift conditions, the specific speed of the pump
should be kept below a certain value. The conclusions of the committee were published in October 1932
in the Hydraulic Institute Standards in the form of charts that became commonly known among centrifugal
pump engineers and users as "specific speed limit charts."
As experience was accumulated on better design than described in these first charts of 1932, revised
charts were prepared and published by the Hydraulic Institute. Ultimately, charts were provided for
several varieties of pump design as follows:

1. Double-suction pumps
2. Single-suction pumps with shaft through the eye of the impeller
3. Single-suction overhung impeller pump
4. Single-suction mixed and axial flow pumps
5. Single-suction hot water pumps
6. Double-suction hot water pumps
7. Condensate pumps with shaft through the eye of the impeller.

Figure 19.6 illustrates one such chart, giving specific speed limits for single-suction overhung impeller
type pumps, such as the ANSI pumps.
For the record, it is important to remember that these charts were strictly empirical. They did not
Suction Conditions and Limitations on Suction Performance 479

o 0
o 0
0
a<t
0
a o
a
a 0000 a a 0 0 a
!D It) ,..,
It) 00'1 (I) ""!D 10 V ,.., (\j
(\j

4,000 A

x~~
~
2
3,500 ,,~/
~ ~
J" f-_J...
~ II

~
d.
~t~~
I? ;$' 11~
-oJ

§I- , --~ I---

~~
0:: 3,000 -;
~ ~
°
!oJ
..J -oJ '
~ ~ I- --:~L
..J " ~ "v ~
!oJ
d.
2 2,500
~
" v.,r" ~I ~ ,,~
,,0 1/
"
u
~I
I I 't?.,'4. t
C> I/~ ., ,,~
Z I ....1
~ ~I
::I: V t1
0:: r- ~ "
...,...."
I 4.1
~ 2,000 1/
~
o ~
1/
z
a
~ let V JL" V
t- ~ u" etl
o i::
~ ~~ ...,
Co;
~/
"v
(/')
I?
I
!oJ 1,500 ~{ I
,,0 ~
..J
C> VII I ,,0
Z
(/') V V ~ 4. " V
I ~ V
V
0::
a
u.
o
W
I V
!oJ
d. V
(/') I II
1,000
~
u.
oW 900
0..
(/')

800

700a a
a 0 a 0000 a a a a a
a a
!Dill
aV 0
If')
a III 00'1 CI) ""!D It) V If') (\j
(\j

TOTAL HEAD,IN FEET(FIRST STAGE)

Fig. 19.6 Specific speed limit chart formerly recommended by the Hydraulic Institute.
Circa 1960; for single-stage. overhung-impeller pumps

indicate that pumps built for the limit allowed were necessarily the best design, nor that pumps built to
lower limits were not more economical in certain cases, nor finally that pumps could not be designed
and built for higher limits. All that these charts were intended to indicate was that for a given set of
head, capacity and suction conditions, a certain maximum rotative speed should give assurance that the
pump would be capable of giving satisfactory service.
480 Suction Conditions and Limitations on Suction Performance

It must be realized that pumps built for the allowed limit are not necessarily the best design for the
intended service and that a lower specific speed type might be more economical. It must also be realized
that the individual pump design limits its application for both maximum head and for suction condition
limitations. For example, using the HI charts, the maximum recommended specific speed for a double-
suction, single-stage pump would have been 1,990 for a 61 m (200 ft) total head and a 4.6 m (15 ft)
suction lift. It does not follow that all double-suction, single-stage pumps of 1,990 specific speed type
are suitable for operation at speeds which will cause them to develop a 61 m (200 ft) total head (at
maximum efficiency); nor that the pump, if suitable for operation at a 61 m (200 ft) total head, is suitable
for operation with a 4.6 m (15 ft) suction lift; nor that a pump of this type operating against a 61 m
(200 ft) total head would on test be found capable of operating on only a 4.6 m (15 ft) maximum suction
lift. These charts were intended to indicate only the maximum rotative speed for which experience had
shown a centrifugal pump could be designed with assurance of reasonable and proper operation for the
combination of operating conditions.
Nothing in these charts suggested that the specific speed indicated corresponds to the point of maximum
efficiency. However, pumps are normally applied for conditions near their maximum efficiency points.
Thus, even though the service conditions do not correspond exactly with the design conditions, the
specific speed value is generally sufficiently close to the specific speed of the design condition.

SUCTION SPECIFIC SPEED

The application of the specific speed limit charts as they were originally developed, however, had a very
important shortcoming: the fact that satisfactory suction conditions were tied directly to the total head
developed by the pump. The perceived flaw was that the performance of an impeller from the point of
view of cavitation cannot be affected too significantly by conditions existing at its discharge periphery,
which, however, are the prime factors in determining the total head that the impeller will develop. This
statement must be qualified to some extent: as the specific speed, Ns, increases, the loading of the suction
region of the impeller vanes extends far enough along the vanes to influence the total head and the
power consumption. Expressed in another way, we can say that as the dynamic portion of the total head
increases, the nature of the flow around the whole vane assumes a greater importance in both the net
positive suction head required and in the total head produced by the impeller. This was recognized much
later than the time when the suction specific speed concept was developed, when the onset of internal
recirculation at the discharge of the impeller was found to trigger off that of internal recirculation at
the suction in the case of higher specific speed pumps (see Chap. 22 for a detailed discussion of
this phenomenon).
Returning to the perceived flaw, the argument was that, if an impeller exhibits certain suction
characteristics, cutting down its diameter within reasonable limits and thus, reducing its head, should
have no influence on its suction capabilities. At the same time, since the total head is changed, a strict
interpretation of the specific speed limits charts would indicate that, unless the suction lift were to be
commensurately altered, the maximum permissible specific speed must be changed. Likewise, to maintain
a fixed value for the Thoma-Moody parameter, a reduction in head by cutting the impeller diameter
should be followed by a proportionate reduction in the net positive suction head.
This inconsistency was finally resolved in 1937 by the development of the suction specific speed
concept. It is essentially an index number descriptive of the suction characteristics of a given impeller.
It is defined as:

n(1J.5
S = NPSH o.75
Suction Conditions and Limitations on Suction Performance 481

where S =suction specific speed


n =rotative speed, RPM
Q = flow per impeller eye, gpm (total flow for single suction impellers; one-half of the total flow for
double suction impellers)
NPSH = required net positive suction head, ft.

As with specific speed, countries using the metric system assign a different designation to suction specific
speed, namely nqn and express Q in m3/hr, NPSH in m. The mathematical relation between values
calculated in the two systems of units is

If one refers to the papers in question,! it will be seen that the specific speed index number can be
derived by considerations of the laws of similarity and by means of dimensional analysis. The fact is
that the steps which led the authors to the solution were much simpler and involved no complex
considerations. We had developed reasonable evidence of the fact that cr, the Thoma-Moody parameter,
appeared to be a function of the specific speed, but the head seemed to refuse stubbornly to disappear
from the relationship. However, we knew that, at least for a certain range of specific speeds, conditions
at the impeller discharge could not be affecting suction conditions.
We suddenly saw that all that was required was a mere algebraic manipulation. If instead of trying
to relate Sigma to the specific speed, we looked for a relation between the specific speed and the 3/4
power of Sigma, the total head disappeared very conveniently from the relation.
To say that this concept was received with unanimous approval would be an exaggeration. In keeping
with any suggestion for a major change from the status quo, there immediately developed a substantial
controversy over the validity of the concept. Fortunately, the controversy was short-lived and the suction
specific speed was soon accepted at the most convenient parameter for describing the suction capabilities
of centrifugal pumps.
The Hydraulic Institute specific speed limit charts have been revised several times since they were
first adopted as a guideline for centrifugal pump suction conditions. Unfortunately, for a long time they
continued to be based on the erroneous concept that the total head developed by the pump plays a part
in determining the maximum permissible rotative speed for a given set of suction conditions, despite
the recognition of the concept of suction specific speed in the Hydraulic Institute Standards.
Finally, in 1983, the 14th edition of the Hydraulic Institute Standards incorporated charts that had
been significantly revised. They are now based on a suction specific speed of 8,500 throughout, that is,
for both sing1e- and double-suction impellers. Thus, the total head factor has finally been eliminated
from these charts. In addition, the recommended values are now expressed in terms of NPSH instead
of suction lifts or suction heads, significantly simplifying their use (see Figs. 19.7 and 19.8).
A final note of caution is necessary. The published Hydraulic Institute Standards suction limitation
charts should be considered as guidelines, in that they represent conditions that have shown satisfactory
operation. As such, they are conservative in certain cases, and a manufacturer may sometimes guarantee
performance at suction specific speeds higher than these Hydraulic Institute ratings, because the manufac-
turer has evidence of successful operation at such ratings. In such a case, the user needs to consider the
reputation of the manufacturer and/or any previous experience with him. If this reputation and experience

t"Some Notes on a New Method of Representing Cavitation Results" by Dr. O. F. Wislicenus, R. M. Watson, and I. 1. Karassik,
presented at the Hydraulic Institute Meeting in New York, December 6, 1937; and "Cavitation Characteristics of Centrifugal
Pumps Described by Similarity Considerations" by O. F. Wislicenus, R. M. Wilson, and I. 1. Karassik, presented at the Spring
Meeting of the ASME at Los Angeles, California, March 23-25, 1938 (ASME Transactions for January 1939).
482 Suction Conditions and Limitations on Suction Performance

Capacity - m3lh
4 6 8 10" 2 4 8 8 10' 2 4 8 8 10' 2 4 6 8 10"
, ii' iii it' , iii i

l00~.-..-.. -. . -.. .~.;-


.. -.. ~
.. ,~ < ..-..~
. .-.. ~ :..~
.; ~ ...-..- .-.. -...~,--~~~~~~----~~~~~~~~~~~~~~~~
.. ;~.;~.<~.,-
80
20
80

. .
- :' .~ ·~··~··~·t,
10

· ··, il TU•1•
8
20 II
Gi
at 4

~ 10 ~ ,:::::::::::::t::::::::t:~:::i: : ::~:::~:: l::~ ::~ :


, •• •• : •• : •.! . :
en
CL
Z
8
. .:'l':::~~:: ~:: 1:: 1:+:1 : :::::::::::::~::::::::~: ~: : :~::::~:::~:: j::~: +: 2
6 ... .... , .~ ..... ! ... ': .. ,t··!··!· ':- ,! .............. !........ ~ ..... ~" .. ~ ... ~ .. !. ,!. '? '

··;r·:r]:!!: ~~~t
4

.8

2 ... r .8

.4
1
10" 2 4 8 8 10' 2 .. 8 8 10' 2 4 8 8 10" 2 4 II 8 10"

capacity U.S. GPM

Fig_ 19.7 Hydraulic Institute recommended maximum operating speeds for single-suction pumps.
(Courtesy of the Hydraulic Institute)

. 6 8 10"
Iii
2 4 8
,
8 10"
I I,
2 4 8 8 11),'
ii'
2 .. 8 8 10"
i i

100
80
20
80

40
10
8
20 II
Gi
at 4
···1·
~ 10 ---- --- _. __ ........ ...................
;r-

en .. ............... . .
,.~ .

CL 8
Z 2

t·l: i:I:
6

4
··::::::::I::::.::.I .. :.:i:...l .
NPSHA va Capacity for speeds
shown at eonatant S = 8500. .8
For use wtth double suctlon pumps.
2 .8
········.. ····1········, .. ·· ., ....,...1-___"_=_::._85:00:'-'-(,6~/=-:-P_::)H7._6~:=·7=6_ .,.....,.... ..,-!
.4
I ~~~~~~~~~~~~~~~~~--~~~~~~~----~~~~~~
10" 2 4 8 8 10> 2 4 8 8 104 2 4 8 8 10" 2 4 8 8 10"
capacity U.S. GPM (total pump capacity)

Fig. 19.8 Hydraulic Institute recommended maximum operating speeds for double-suction pumps.
(Courtesy of the Hydraulic Institute)
Suction Conditions and Limitations on Suction Performance 483

are favorable, the user should have no concern, with but one qualification, over accepting these recommen-
dations. The one qualification is consideration of the effect of designing a pump for low values of NPSH
required on the permissible operating range of this pump (see Chap. 22).

EFFECT OF LIQUID CHARACTERISTICS ON PUMP SUCTION


PERFORMANCE OR NPSH

As has been stated, cavitation occurs when the absolute pressure within an impeller falls below the vapor
pressure of the liquid and bubbles of vapor are formed; these bubbles collapse further within the impeller
when they reach a region of higher pressure. The minimum required NPSH for a given capacity and at
a given pump speed is defined as that difference between the absolute suction head and the vapor pressure
of the liquid pumped at the pumping temperature that is necessary to prevent cavitation.
The fact that a pump is cavitating manifests itself by one or more of the following signs: noise,
vibration, drop in the head-capacity and efficiency curves, and, with time, damage to the impeller by
pitting and erosion. All of these signs are obviously inexact and it became necessary for pump engineers
to agree on certain ground rules to establish some uniformity in the detection of cavitation.
The minimum NPSH required is determined by a test in which the total head is measured at a given
speed and capacity, with varying NPSH available conditions. Preferably this test is conducted in a closed
loop such as is described on Fig. 19.9. The pump takes its suction from a closed vessel in which the
pressure level can be adjusted or varied by the gas pressure over the liquid level, by the temperature of
the liquid, or by a combination of these two variables. This variation in pressure level, in tum, controls
the available NPSH.
Results of such a test, plotted against NPSH appear in a form similar to that in Fig. 19.10. At the
higher values of NPSH, the values of head remain substantially constant. As the NPSH is reduced, a
point is finally reached where the curves break, showing the impairment of pump performance caused
by cavitation. The exact value of NPSH where cavitation starts is, as mentioned before, difficult to

Gas Pressure

Suction
Vessel
Flow
Distributor

Heating or
Cooling Coil

Suction

Fig. 19.9 Pressure and/or temperature control NPSH test with closed loop.
(Courtesy of the Hydraulic Institute)
484 Suction Conditions and Limitations on Suction Performance

a H = 3 PERCENT OF H1

l
!(
HEAD AT CAPACITY Q

:=: t 0
AND AT SPEED N

II:
:f!!

HSo NET POSITIVE SUCTION HEAD

Fig. 19.10 Detennination of required NPSH as presently defined by the Hydraulic Institute.

Pressure Gauge
Thermometer Dampening Valve Dampening Device

Control Valve for


Throttling Suction

Flow Meter if
Located in Suction

Constant
Level
Pump nl~====~
,-, on
f1, Test
,
t~ traightening
Vanes
Booster Pump may be Installed
Discharge Control Valve, or '. - - •
Adjustable Spring Loaded
Back Pressure Valve or
if additional suction pressure Ac:f)UStable Choke Valve
Is required Heat exchanger, if required -r-------I

Retum to sump Dlactwge


- Note:
Position of these
devices may be reversed
in some set ups.

Fig. 19.11 Suppression type NPSH test with constant level sump.
(Courtesy of the Hydraulic Institute)
Suction Conditions and Limitations on Suction Performance 485

pinpoint. To overcome this difficulty, engineers have generally agreed that a drop of 3 percent in the
head developed is taken as evidence that cavitation is taking place. For that particular speed and the
capacity being tested, the NPSH that produces a 3 percent drop in head is stated to be the minimum
required NPSH.
The Hydraulic Institute Standards for centrifugal pump testing [19.1] permits two simpler forms of
cavitation tests. In the first arrangement (Fig. 19.11), the pump takes its suction from a constant level
sump through a throttle valve, which is followed by a section of pipe containing a screen and straightening
vanes. The operation of the throttle valve is used to vary the available NPSH and is therefore often
termed a "suppression test." In the second arrangement (Fig. 19.12), the pump takes its suction from a
relatively deep sump in which the level can be varied to establish the desired available NPSH.
Generally, a "new" pump, that is a brand new design that has not been manufactured before, is always
NPSH tested by the "closed loop" method. If one is dealing with pumps that have been manufactured
before, considerable expense can be saved by using one or the other of the two alternate test methods.
There is a risk, however, that the required NPSH determined by either of these alternate methods will
be higher than that from a "closed loop" test. If the liquid pumped on test contains dissolved air or gas,
some of this will come out of solution at the lower pressure in the suction line and can cause a drop in
head similar to that caused by cavitation (but is actually caused by pumping a two-phase flow (see air
entrainment in Chap. 18). In discussing the various NPSH test methods, the Hydraulic Institute Standards
draws attention to this possibility and its effect. The risk of inaccurate results rises with decreasing
NPSHR, and it is therefore better to test by the "closed loop" method in all cases where the NPSHR is
below 4.5 m (15 ft).
NPSH tests of centrifugal pumps are normally carried out on cold water and both the Hydraulic
Institute Standards curves and pump manufacturers' rating curves indicate NPSH requirements on cold
water. Thus, it might be assumed that the NPSH required by a centrifugal pump for satisfactory operation

PUMP SUCTION
FLANGe~

It IMPELLER EYE ~1It""-"ffi---.ffi-_1 -ttt--.fiIor-----r

10 ft. min.
(3 m)

~
o .. DIAMETER OF PIPE

Fig. 19.12 Suction line for static lift test.


(Courtesy of the Hydraulic Institute)
486 Suction Conditions and Limitations on Suction Performance

Table 19.1 NPSH Requirement by Test, at Design Flow and 3


Percent Drop in Head

NPSH, min.
Temperature accuracy ~NPSH

0.15 m (± 0.5 ft)


Fluid °C COF) (m) (ft) m ft

Water 21.1 ( 70) 3.75 12.3 0.0 0.0


121.1 (250) 3.35 11.0 0.4 1.3
148.9 (300) 2.62 8.6 1.13 3.7
Butane 1.67 ( 35) 2.99 9.8 0.76 2.5
12.8 ( 55) 2.68 8.8 1.07 3.5
32.2 ( 90) 1.07 3.5 2.68 8.8
Freon-ll 29.4 ( 85) 3.11 10.2 0.64 2.1
48.9 (120) 2.56 8.4 1.19 3.9
Pump Ns = 1.600
N = 3,585 rpm.

is independent of the liquid vapor pressure at the pumping temperature. This is actually not true. It is
merely an oversimplification used to illustrate the definition that NPSH is a measurement of the energy
in the liquid at the pump suction over the datum line of its vapor pressure.
At the same time, both laboratory and field tests run on pumps handling a wide variety of liquids
and over a range of temperatures have always shown that the NPSH required for a given capacity and
with a given pump apparently vary appreciably. For example, the required NPSH when handling some
hydrocarbons is frequently much less than that required when the pump handles cold water. Even when
pumping water, there is definite evidence that required NPSH decreases when the water temperature
increases. For example, Table 19.1 illustrates test results of a pump at 3,585 rpm, specific speed of
1,600, and handling different liquids at several different temperatures.
It became evident quite a number of years ago that the reduction in the required NPSH must be a
function of the vapor pressure and of the characteristics of the liquid handled by the pump. In general,
it can be proved that the factors affecting the required NPSH are, at least, the following three:

1. The ratio of the specific volume of the vapor to that of the liquid at the pumping temperature
2. The homogeneity of the liquid
3. The effect of subcooling derived by any partial flashing of the liquid in question.

Because in the case of hydrocarbons this reduction in NPSH could play a most important role in the
relative costs of a refinery installation, most efforts were directed at understanding the phenomena
involved in connection with hydrocarbons. Thus, it was felt that rules could be developed to predict the
effect of liquid characteristics on the required NPSH to take advantage of this phenomenon without the
risk of overoptimistic assumptions. Such rules have been developed by the members of the Hydraulic
Institute and incorporated in its standards. We shall examine these rules, but before doing this, let us
consider the effect of temperature on the required NPSH for water, because this may help us better
understand the effect of other liquids.
Suction Conditions and Limitations on Suction Peiformance 487

PERFORMANCE ON WATER AT VARYING TEMPERATURES

It has been noted for some time that pumps handling hot water seem to require less NPSH than shown
by cold water tests. The theory underlying this effect is fairly simple and need not be discussed in detail
here. It is based on the fact that mild and partial cavitation can take place in a pump without causing
extremely unfavorable effects. The degree of interference with the proper operation of the pump caused
by such minor cavitation will bear a definite relation to the temperature of the liquid handled by the
pump. Remember that when we say that a pump is cavitating we mean that somewhere within the
confines of the pump, the pressure will have fallen below the vapor pressure of the liquid at the prevailing
temperature. Thus, a small portion of the liquid handled by the pump will vaporize and this vapor will
occupy considerably more space within the impeller than the equivalent mass ofliquid before vaporization.
If the pump is handling water at normal temperatures, the volume of a bubble of steam is tremendously
larger than the volume of the original quantity of the water. For instance, at 4.4°e (40°F), 0.45 kg (1
lb) of water occupies 0.0004 m3 (0.016 fe), while steam at the same temperature occupies 69.2 m 3 (2441
ft 3). The ratio of the two volumes is 152,500. The rapidity with which this ratio diminishes as water
temperature increases is illustrated in Fig. 19.13, which presents a plot of the ratio of the volume of
steam to the volume of the equivalent mass of water for temperatures between lOoe (50°F) and 374.1 °e

'\~
100,000

\ \.
10,000

RATIO OF STEAM
TO WATER

1000

~
a:
w
!;t

'"'"
3:
.J
0
>
100 --
::i:
«
w
I-
rJ)

""~
.J
0
>

\
10

TEMPERATURE IN OF
100 200 300 400 500 600

Fig. 19.13 Ratio of steam to water volumes versus temperature.


488 Suction Conditions and Limitations on Suction Performance

(705.4°F) (the critical temperature of water at which steam occupies the same volume as water). At
lOO°C (212°F), 0.45 kg (lib) of water occupies 0.00047 m3 (0.0167 ft 3) and 0.45kg (lib) of steam
0.76 m3 (26.S1 ft3), so that the ratio of volumes is only 1,605-almost 100 times less than at lOoC
(50°F). Thus, the higher the temperature of the water, the more NPSH reduction can be permitted for
the same degree of effect on the pump performance.
There is another important thermodynamic factor that affects the relative cavitation characteristics of
a liquid-the difference in the effect of subcooling caused by any flashing of the liquid. The reduction
in the vapor pressure corresponding to a given change in temperature of the liquid varies for different
liquids as well as for different temperatures of the same liquid. Consider, for instance, what happens
with water at 204.4°C (400°F) and at 26.7°C (SO°F), respectively. The vapor pressure of 204.4°C (400°F)
water is 17.03 bar abs (247 psia). A 0.56°C (1°F) degree drop in temperature to 203.SoC (399°F) will
reduce the vapor pressure to 16.S4 bar abs (244.2 psia), a reduction of 0.19 bar (2.S psi) or 2.3 m (7.5
ft). This, then, is equivalent to instantaneously increasing the available NPSH by these same 2.3 m (7.5
ft). On the other hand, the vapor pressure of water at 26.7°C (SO°F) is 0.035 bar abs (0.507 psia). With
the same 0.56°C (1°F) drop, at 26.1 °C (79°F) the vapor pressure is 0.034 bar abs (0.490 psia), a difference
now of only 0.001 bar (0.017 psi) or 0.01 m (0.04 ft), a completely insignificant contribution to the
available NPSH. Meanwhile, since the available NPSH is calculated based on the liquid temperature
measured at the pump suction (before any flashing and subcooling will have taken place), the apparent
result is that there is a reduction in NPSH required rather than an increase in NPSH available and that
this reduction is greater for water at 204.4 °C (400°F) than at 26.7°C (SO°F).

NPSH FOR PUMPS HANDLING HYDROCARBONS

Pump applications for hydrocarbon processes frequently impose restrictive limitations to the available
NPSH. On the other hand, it was found that variations between the required NPSH when handling
hydrocarbons and that in cold water service were generally in a favorable direction. These two circum-
stances led both pump designers and designers of refineries to direct their efforts at understanding the
phenomena involved and at establishing rules that could be applied to predict the effect of any special
liquid characteristics on the required NPSH of any centrifugal pump.
At first it was thought that these variations were only apparent and that if "true vapor pressures" or
"bubble point" pressures were to be used in the calculations of test NPSH, the discrepancies would
disappear and complete correlation with water test cavitation data would exist. Corrections for NPSH
with hydrocarbons were nevertheless used, as a matter of policy rather than based on accepted theoretical
deductions. It was believed that a reduced NPSH could be justified for the following reasons:

1. Oil companies' specifications generally called for a maximum capacity and head at a minimum NPSH. In
practice, it was unlikely that these two requirements would be imposed simultaneously. In fact, some of the
field conditions are self-regulating-for instance, low capacity occurs at low NPSH, as a result of a reduced
flow in the system. Under these conditions, even if the pump capacity falls off, available NPSH is increased
and equilibrium is eventually attained.
2. The effect of cavitation with hydrocarbons was noted to be not as severe as with water, that is, the head-
capacity curve does not break: off suddenly for two reasons: (a) only the lighter fractions will boil first and
(b) the specific volume of hydrocarbon vapors is very small in comparison with that of water vapor.

Obviously, these facts do not tell the whole story, as many other factors affect the behavior of a pump
handling hydrocarbons with low NPSH. Thus, attempts to arrive at a more reasoned understanding
continued while some interim correction factors of an approximate nature were being used.
Suction Conditions and Limitations on Suction Performance 489

These efforts centered in the accumulation and comparison of many tests, using a variety of pumps
and handling many different hydrocarbons. These tests, in turn, helped generate a variety of correction
curves for NPSH. Some of these charts occasionally led to rather impractical conclusions and additional
rules were then introduced to avoid this situation.
In 1951, the Hydraulic Institute Standards incorporated a conversion chart for hydrocarbons which
has since been updated. It provided an estimate of the NPSH required by a centrifugal pump handling
hydrocarbons of various gravities and vapor pressures in percentages of that required by the same pump
when handling cold water. These curves were derived from an accumulation of experimental data and
did not pretend to be arrived at by analytical means.
The latest revised correction chart is incorporated in the 1983 edition of the Hydraulic Institute
Standards (see Fig. 19.14). To use this chart, enter at the bottom with the pumping temperature in
degrees Fahrenheit and proceed vertically upward to the vapor pressure in psia. From this point follow
along or parallel to the sloping lines to the right side of the chart, where the NPSH reductions in feet
of liquid can be read on the scale provided. If this value is greater than one-half of the NPSH required
on cold water, deduct one-half of the cold water NPSH to obtain corrected NPSH required. If the value
read on the chart is less than one-half of the cold water NPSH, deduct this chart value from the cold
water NPSH to obtain corrected NPSH required.
Because of the absence of available data demonstrating NPSH reductions greater than 3m (10 ft), the
chart has been limited to that extent and extrapolation beyond that limit is not recommended. In addition,
warnings are included in the Hydraulic Institute Standards regarding the effect of entrained air or gases.
This circumstance can cause serious deterioration of the head-capacity curve, of the efficiency and of
the suction capabilities even when relatively small percentages of air or gas are present (see Chap. 18).
The fact remains that there is insufficient correlation at this moment among the many tests cited in
the technical literature. It appears rather probable that the very characteristics of a pump-that is, its
specific speed and its actual design-play some role in the actual reduction in NPSH right along with
the characteristics of the hydrocarbon. This role may be minor, but it probably does exist.
A more exhaustive analysis of the phenomena that take place in a pump handling hydrocarbons is
beyond the scope of this discussion. As a matter of fact, such an analysis would at best be open to
argument, because several somewhat conflicting interpretations still exist with respect to what actually
takes place.
Whether a more rigorous theoretical derivation of NPSH reduction is ever developed is really immate-
rial. The important fact remains that as further experience is gained and more and more experimental
data are accumulated, the validity of correction charts will be even greater. At the same time, we would
like to incorporate here a word of caution: it is probably best to use this correction factor as an additional
safety factor rather than as a license to reduce the available NPSH. This is a personal opinion, but one we
share with a number of rotating machinery specialists of some of the major petroleum and petrochemical
companies, and one that is reflected in the refinery pump specification API-61O [3.1].

NPSH REQUIRED FOR CRYOGENIC PUMPS

It has been frequently claimed by users that the required NPSH curve of cryogenic pumps has a marked
increase as the capacity falls to some 20 or 30 percent of design conditions. The question involves us
with some ambiguity introduced by our semantics. If we consider this statement strictly on its literal
interpretation, the answer is, no, a pump handling cryogenic liquids does not require any greater NPSH
at reduced capacities than would be the case if the pump were to handle water. But if we instead word
the question, Should a pump handling cryogenic liquids be provided appreciably more available NPSH
490 Suction Conditions and Limitations on Suction Performance

Temperature OC
J!!
I/)
a..
·10 o 20 40 60 80 100 200
i i
1000 -----.. .
-... -1- ' : :
-~.-..-....I..------
6000
-.-..~.
600
::=j!..~-·-t--:--Ir
"-r
!

400
•••.. , .•- i f -......-+-It-
I 10 3.0
i
200 -,'-+-+-+-+- 8 2.4

1000
i 8 1.8
c::
o
i ".1:1
I
; 5 1.5 ()
100 ::J
4 1.2 "t:J
600 80 !
3 0.9 a:
80 J:
400
0.6 en
2 a..
40 z
1.5 0.45
200
1.0 0.3
20

100
0.5 0.15
10
60 8

40 8

4 ..-...k~-+-+-++
20 ~. -+-+-+
/ !
! : :
I ,

I i j
2 ··.... -r--r·-!-+--++
I I
; I
10
.!
I I
!,
o 50 100 150 200 300 400

Temperature OF

Fig. 19.14 NPSHR reductions for pumps handling hydrocarbons and high-temperature water.
(Courtesy of the Hydraulic Institute)

than it would require when handling water, if the pump is expected to operate at reduced flows?-the
answer is an unequivocal yes.
The problem of semantics arises from our accepted definition of NPSH, be it required or available.
The only practical means to define NPSH is to refer to conditions prevailing at the pump suction flange,
corrected to the pump centerline (assuming for the sake of simplifying this explanation that we are
Suction Conditions and Limitations on Suction Performance 491

dealing with a horizontal pump). The energy at that location less the vapor pressure of the liquid,
expressed in foot-pounds per pound or feet, is defined as the NPSH available, whereas the NPSH required
by the pump at a given speed and for a given capacity is defined as the NPSH available when the total
head of the pump is reduced by exactly 3 percent. The observed facts are strictly apparent and not real.
What one sees in a test for NPSH of a cryogenic pump is an error in measuring the NPSH available
and not an increase in NPSH required. One does not measure the NPSH required, one measures the
available NPSH and then determines the required NPSH by observing the test values and calling that
NPSH available at which a 3 percent drop in head occurs the required NPSH. The available NPSH is
stated to be equal to the energy over and above the vapor pressure at the pumping temperature at the
pump suction flange.
In most cases, the fact that the vapor pressure at the impeller inlet differs to some extent from the
vapor pressure at the suction flange can be neglected because this difference is negligible. But in the
case of cryogenic pumps this difference can become so large under certain conditions that the effect is
quite dramatic. At or near the best efficiency point, this cannot introduce any significant error, since the
temperature rise in the pump is negligible and the flow past the wearing ring of the first stage is but a
diminutive fraction of the flow into the pump. Thus, the temperature at the eye of the impeller does not
change appreciably from the temperature at the suction flange and the assumed vapor pressure is
essentially correct.
As the capacity is reduced, the temperature rise increases while the leakage flow increases as a
percentage of the suction flow. Of course, the calculated temperature rise takes place as the liquid passes
through the impeller and casing passages. But some of the liquid from the impeller discharge passes
back through the running clearances, where its temperature rises still farther, and mixes with the incoming
flow at the impeller suction. The net effect is that the temperature at the eye of the impeller is no longer
the same as at the suction flange, nor, of course, is the vapor pressure. The effect of the temperature
rise in cryogenic pumps at reduced flows is frequently discussed under the subject of recommended
minimum flows, since it generally dictates the setting of the minimum flow for this type of service.
Consider, for instance, the effect of an increase in liquid temperature of 0.56°C (lOP) on the vapor
pressure of water at 26.7°C (SOOP) and, say, of methane at the usual pumping temperature of -151°C
(-240°F). Por water,

Temperature Vapor pressure


26.7°C (80°F) 0.035 bar abs (0.507 psia)
27.2°C (81°F) 0.036 bar abs (0.526 psia)
~T = 0.56°C WF) Difference = 0.001 bar (0.019 psi)
or 0.012 m (0.04 ft)

whereas for methane,

Temperature Vapor pressure


-151°C (-240°F) 2.28 bar abs (33 psia)
-140°C (-220°F) 4.41 bar abs (64 psia)
~T = 11 °C (20°F) Difference = 2.13 bar (31 psi)
~T = 0.56°C (1°F) Difference = 0.107 bar (1.55 psi)
or, at SG of 0.4 = 2.7 m (8.9 ft)

In other words, an increase in temperature of 0.56°C (1°P) increases the vapor pressure of 26.7°C (SOOP)
water by 0.012 m (0.04 ft) and that of -151°C (-2400P) methane by 2.7 m (S.9 ft).
492 Suction Conditions and Limitations on Suction Peiformance

If we were to imagine that at some low flow the effect described raises the liquid temperature at the
eye of the impeller by 0.28°C (0.5°P) over that at the suction flange, the result is to increase the vapor
pressure by a negligible amount if the liquid is 26.7°C (80 0P) water, but by as much as almost 1.4 m
(4.5 ft) if it is methane at -151°C (-2400P). This increase in vapor pressure is not normally taken into
account when running the NPSH test, and therefore the real NPSH available is 1.4 m (4.5 ft) less than
the apparent NPSH available, if we use the temperature rise we have assumed. Since by definition the
NPSH required is that NPSH available that will not cause a drop in total head of over 3 percent, it
appears that the NPSH required has gone up. But it has not, actually.

EFFECT OF SUCTION CONDITIONS ON PUMP CHARACTERISTICS

The suction limitation of centrifugal pumps is determined by the fact that the impeller cannot impart
energy to the liquid until the liquid is in the impeller between the vanes. Thus, the energy necessary to
overcome the frictional losses up to the entrance of the suction vane ends of the impeller and the energy
necessary to create the velocity required at this point have to come from some outside source. Furthermore,
sufficient additional energy must be available in excess of these and other requirements so that the
absolute pressure at all points is above the vapor pressure of the liquid, to prevent its flashing into vapor.
Figure 18.16 shows the characteristics of a 6-in. pump. If it is operated in a system in which the
available external energy on the suction side could only force 320 m3/hr (1,400 gpm) into the impeller,
and if the total head of the system with this capacity is 30 m (100 ft), the pump will work to pump over
430 m3/hr (1,900 gpm) against this head. As there is insufficient suction head to get more than 320 m3/hr
(1,400 gpm) into the impeller, however, the pressure at that point would be reduced below the vapor
pressure of the liquid and part of the liquid would flash into vapor.
If there is not sufficient available NPSH to permit a pump to develop its normal characteristics,
cavitation will result and the pump will "work in the break." Thus, the characteristics of a centrifugal
pump will vary with the available NPSH. Por the specific pump shown in Fig. 19.15, the characteristics
in solid lines are for 0 suction lift (9.8 m [32 ft] NPSH) whereas with a 6.1 m (20 ft) suction lift (3.7 m
[12 ft] NPSH approximately) the pump follows the 0 suction lift characteristics out to 1,500 gpm when
cavitation starts, evidenced by the pump producing less head. Some increase in capacity results with
further reduction in head until 420 m3/hr (1,850 gpm) is reached, when further reduction in head causes
no increase in capacity. Thus the pump characteristics with a 6.1 m (20 ft) suction lift would be shown
by the solid lines out to 340 m3/hr (1,500 gpm) capacity, and then by the broken line.
The pump illustrated in Fig. 19.15 is of a fairly low specific speed type. In higher specific speed
types, such as the high-speed, Francis screw vane and mixed-flow impeller designs, the operation at
reduced NPSH also reduces the head developed at or near shut-off. With high-specific-speed types this
reduction in head is even more pronounced. Unlike the low-specific-speed types, the higher-specific-
speed types may deliver (with reduced NPSH), with lower total heads, up to a maximum capacity, and
then, as the total head is further reduced, the capacity may be reduced below this maximum, reversing
the head-capacity curve on itself.
Usually cavitation is to be avoided. However, one type of pump, the condensate pump operating on
nonthrottled systems, is especially designed for such operation. Figure 19.16 shows the normal head-
capacity curve, with sufficient NPSH to prevent cavitation, and the system head-capacity curve; Pig.
19.17 shows the layout of the system on the suction side. If the amount of steam being condensed is
equivalent to 11.8 m3/hr (52 gpm), the level in the hotwell will be that which gives 0.31 m (12 in) NPSH
at the suction nozzle, so that the pump is operating in the break at 11.8 m3/hr (52 gpm) capacity and a
16.5 m (54 ft) total head, as dictated by the system head curve. If the amount of steam increases to
equal 16.1 m3/hr (71 gpm), the liquid level in the hotwell will build up until it is 0.46 m (18 in) plus
Suction Conditions and Limitations on Suction Performance 493

100 200

... 90 180
z
w
u
a:
w
Q.

20

10
o
4 8 12 16 20 24 28
CAPACITY, IN 100 GPM

Fig. 19.15 A typical group of curves for a pump operating under varying suction conditions.

100

90

...
W
80

..... 70
1&.1

~
60
0
cI
w 50

%
N

... 40
..J
cI r- ~ N-
>-
... >- >-
...
....0 30 r---i---+--+--f---+t ~ r - U ~ !:;-
a: f ~
20 I----lf----t--+--+-_t_ ucI r - .... u
cI t--~­
-
I I ....
00 '
10 r-~~-+--+-~--+_c1
W -
W t--~---
1&.1
% % %
o I

o 20 40 60 80 100
CAPACITY. IN GPM

Fig. 19.16 Characteristics of a condensate pump operating on a submergence-controlled system.


494 Suction Conditions and Limitations on Suction Performance

SURFACE CONDENSER

WATERlEVEl~~~~
__~ ~_~I __~~=t__-.~~~
__
SUCTION.LINE LOSSES

-r
- +---

NE\
PoSITIVE
HOTWEll SUCTION
HEAD
~--,-_L

Fig. 19.17 Typical installation of submergence-controlled condensate pump.

the friction losses above the pump centerline, and the pump will be delivering 16.1 m3/hr (71 gpm)
against a 17.7 m (58 ft) total head (the intersection of the head-capacity [0.46 m or 18 in] and the
system-head curve). A regular impeller design on such service would be noisy and would show evidence
of cavitation by damage to the vanes after a short time. For hotwell or condensate service, special
impeller designs have been developed with larger suction areas (to operate on low NPSH) and with
special suction vanes to give quiet operation and long life even though cavitating all the time they are
in operation.
In handling liquids containing dissolved gases, the pressure reduction on the suction side of the
impeller vanes can be sufficient to allow the gases to be liberated and so the pump is actually handling
a gas-and-liquid mixture. If the amount of gas liberated is not excessive, the only effect may be a
reduction in capacity output and efficiency. This separation of gas from liquid is often mistaken for
cavitation; it is not. If both cavitation and gas separation occur in a pump, the cushioning effect of the
gas often quiets the cavitation noise. The cushioning effect has sometimes been used to quiet noisy
cavitating pumps by bleeding air into the suction. Although it serves as a temporary expedient, the most
economical solution should be replacement of the impeller by a design suitable for the suction conditions
or a redesign of the pumping system so that the pump has sufficient NPSH to operate on its normal charac-
teristics.

WHEN SUCTION CONDITIONS ARE INADEQUATE

When a system offers insufficient NPSH available for an optimum pump selection, there are several
ways to deal with the problem. Basically, to correct the situation we can either find means to increase
Suction Conditions and Limitations on Suction Performance 495

the NPSH available, means to reduce the NPSH required, or combine both approaches. Under the first
category, we can

1. Raise the liquid level.


2. Lower the pump.
3. Reduce the friction losses in the suction piping.
4. Use a booster pump.
5. Subcool the liquid.

To reduce the NPSH required, we can

6. Use slower speeds.


7. Use a double-suction impeller.
8. Use a larger impeller eye area.
9. Use an oversize pump.
10. Use an inducer ahead of a conventional impeller.
11. Use several smaller pumps in parallel.

Each of these methods presents some advantages and some disadvantages and we shall examine and
evaluate these methods individually.

1. Raise the liquid level-At first glance, this appears to be the simplest solution unless it is impractical
because (a) the liquid level is fixed as in the case of a river, a pond, or a lake; (b) the amount by which
the level must be raised is completely impractical; or (c) the cost of raising a tank or a fractionating tower
is excessive. But frequently it will be found that only a few extra feet may permit the selection of a much
less expensive or much more efficient pump and the resultant savings in first cost, energy, or maintenance
will far outweigh the additional costs incurred by raising the source of the liquid.
2. Lower the pump-Just as in the case of raising the liquid level, the cost of lowering the pump may not
be as prohibitive as one might imagine from past experience, since it may permit the selection of a higher
speed, less costly and more efficient pump. An alternate approach may be to use a vertical pump with the
impeller located below ground level. The penalty for this solution lies in the fact that the pump bearings
may have to be lubricated by the liquid pumped. Although successful bearing designs and materials have
been developed for this purpose, it should be well understood that the pump life cannot compare with the
life obtainable from external bearings either grease or oil lubricated. Thus, one should expect more frequent
scheduled overhauls with this solution.
3. Reduce piping friction losses-This is recommended under any circumstances and the cost of doing so
will be easily repaid by both improved suction conditions and savings in energy.
4. Use a booster pump-This solution is particularly effective in the case of pumps intended for high-pressure
service, where the resulting permissible higher speeds will yield great savings in first costs of the main
pump, higher efficiencies, and frequently a lesser number of stages, which in itself leads to greater reliability.
The booster pump can be selected as a low-speed, low-head pump of single-stage design.
5. Subcool the liquid-This approach increases the available NPSH by reducing the vapor pressure of the
liquid pumped. It is most readily accomplished by injecting liquid taken somewhere from the stream where
it is available at a colder temperature. In many cases, particularly at higher pumping temperatures, the
amount of injected cold liquid is very small. As an example, if we are pumping water at 163°C (325°F),
the injection of only 4 percent of 79°C (175°F) water will subcool our stream to the point that the available
NPSH will have been increased by 6.1 m (20 ft) (see Figs. 19.18, 19.19, and 19.20).
496 Suction Conditions and Limitations on Suction Performance

Q = INJECTION FLOW
T2 =TEMP

CONTROL
VALVE

0= TOTAL FLOW
T3 =TEMP

T3 = (0- Q) T1 + QT2
o
Fig. 19.18 Sub-cooling pumped liquid.

0
wz
~o 45
a:-
50
V / v
wC/)
offl 40
/ / /
lift:
ZW 35 .~ / . . .V / /'
~ 1~V ~ V
_0
:x:W
C/)a: 30 L~~ L
a.:::J
Z!;( 'r>~«7-V / 1~ V V L
,,/
C/)a: 25

-
:::Jw
....I a. l-<U~ ~~ V .....V V j ; fI'" ". V
a.;:E 20
~ ~ ~V ~ V 6° / T
a: w ~
~~, .:;,.- ~
rr
:::JI- 15
C/)w
~ ~ ::::::: ~ ~ ~ ~
W:x: ~ 4° ..,..".
~

::::::~
~I- 10
I-;:E I"""

==== ~ E:::::
--
~
Uo ~ 2° ~~
~
~~
~a: 5 ,
u..u..
W I
220 240 260 280 300 320 340 350
TEMPERATURE IN OF

Fig. 19.19 Effect of sub-cooling on available NPSH with water.

6. Use slower speeds-Once a reasonable value of suction specific speed has been selected, it becomes obvious
that the lower the pump speed, the lower will be the required NPSH. The problem, however, is that a
lower speed pump will be more expensive and less efficient than a higher speed pump selected for the
same conditions of service. Thus, lowering the pump speed will seldom prove to be the most economical
solution.
Suction Conditions and Limitations on Suction Peiformance 497

20
z 18
0
Cii
en 16
w
1I:u.
n. 0
14
WO
oz
Wz 12
II: 0
=>- 1----+-+--b'-7~~A-r-~~~r*~""""~~r__::74 0.07
~G 10 0.06
II:=>
wen 8 ~+-4,~~~~~~~~~~~~~~
n.n. 0.05
::!:::!:
w=> 6 ~~~~~~~~~~~~~~~~~~T-=t~0.~
I-n.
(!)
~~ 4 ~~~~~~::::::r1t:~=r-1'""1:JO.03
!:i
=>
en 2 ~~§~::;~~~Ftf=-t'=::::=r-::::f~ 0.02
w
II: 0

Fig. 19.20 Effect of cold water injection.

7. Use a double-suction impeller-Particularly for larger capacities, whenever a double-suction impeller is


available for the desired conditions of service, this presents the most desirable solution. It is based on the
following considerations: First, if we select the same suitable S value for both single and double suction
impellers such that

where SUbscript I refers to a single-suction impeller and subscript 2 refers to a double-suction impeller.
Since

We can assume first that


(a)

in which case NPSHR 2 = O.63NPSHR h or that


NPSHR 2 = NPSHR 1 (b)

in which case n2 = 1.414 nl' Keeping the pump speed the same in both cases, as in equation (a), we can reduce
the required NPSH by 27 percent if we use a double-suction impeller. Alternatively, with a given required NPSH,
as shown in equation (b), we can operate a double-suction pump at 41 percent higher speed (see Figs. 19.21
and 19.22).
8. Use a larger impeller eye area (Fig. 19.23)-This solution reduces the required NPSH by reducing the
entrance velocities into the impeller. These lower velocities may have little effect on pump performance
at or near its best efficiency point. But when such pumps run at part capacity, this practice can lead to
noisy operation, hydraulic surges, and premature wear. This problem will be discussed in greater detail in
Chapter 22. At this point, suffice it to say that it is a dangerous procedure and should be avoided if possible.
498 Suction Conditions and Limitations on Suction Performance

14

12 ~
,;'
a:W ..L
...J
...J
W
10 V
Q..(f) ./ ~
~a:
OW
wti:i 8 ~
V
9:2: l/
wO
(f)a:
...J~ 6
..L V
al w
=>w V
011.
0 / V
11. 4
0
~
V ASSUME:
~
/ S= CONSTANT
2
...-~ RPM = CONSTANT
Q= CONSTANT
I I I I I

o 2 4 6 8 10 12 14 16 18 20
Hs OF SINGLE SUCTION IMPELLER, FEET OR METERS

Fig. 19.21 Required NPSH of double-suction impeller versus single suction impeller.

9.
Use an oversize pump-Because the NPSH required by a pump decreases as the capacity is decreased, a
larger pump than would otherwise be applied to the service is occasionally selected. This practice is risky
and can lead to undesirable results. At best, the penalty is the choice of a more expensive pump that
operates at a lower efficiency than might otherwise have been obtained (see Fig. 19.24). At worst, the
operation at a lower percentage of the best efficiency flow can lead to exactly the same problems as the
use of excessively enlarged impeller eye areas.
10. Use an Inducer-An inducer is a low head axial type impeller with few blades which is located in front
of a conventional impeller (Fig. 4.17). By design, it requires considerably less NPSH than a conventional
impeller, so it can be used to reduce the NPSH requirements of a pump (Fig. 19.25) or to let it operate at
higher speeds with a given NPSH. The inducer is a very adequate solution for many situations but must
be applied with care, as the permissible operating range of pumps with inducers is generally narrower than
with conventional impellers.
11. Use several smaller pumps in parallel-Obviously, smaller capacity pumps at the same speed require less
NPSH. Although this appears to be a costly solution, this is not necessarily so. In many cases, three half-
capacity pumps of which one is a spare are no more expensive than one full-capacity pump plus its spare.
As a matter of fact, in many cases, two half-capacity pumps may be installed without a spare, since part-
load can still be carried if one pump is temporarily out of service. In addition, if the demand varies widely,
operating a single pump during light load conditions will conserve energy, as we shall see in Chapter 21.

DEFINITION OF NPSH REQUIRED

The traditional definition of required NPSH for centrifugal pumps is that value at which the total head
developed by the pump has deteriorated by exactly 3 percent. This value is often designated NPSH 3•
Simplicity of testing (within reason) lead to the use of deterioration in performance as the indicator of
Suction Conditions and Limitations on Suction Performance 499

2600
ASSUME: J
V
S=CONSTANI
Hs = CONSTANT
2400
Q=CONSTANT
/
2200 I
V
2000 V
J
II:
w
...J
...J
W
a..
1800
'/
)
"
~
uj 1600 /
ci
II:
0
)'
u.
:::ii: 1400
V
/
a..
II:

1200 J
)~

1000 I
/
800
)'
/

600 800 1000 1200 1400 1600 1800


RPM FOR S.S. IMPELLER

Fig. 19.22 Allowable speed of double-suction impeller versus single-suction impeller.

cavitation, and a deviation well beyond the accuracy of measuring total head (generally about 1 percent)
to the measure of 3 percent. In the 1960's, as the heads and capacities of pumps staged a spectacular
climb upward, there developed a growing number of field problems with severe cavitation erosion, even
when the available NPSH exceeded the required NPSH by a generally acceptable margin. As a result
of these problems, pump users started asking themselves and the pump manufacturers whether the
traditional definition of 3 percent head drop needed to be reexamined.
The initial approach, driven by intuition, was to say that the extent of cavitation allowed by 3 percent
head drop is evidently so great it causes erosion in some circumstances, so the solution is to base the
required NPSH on less cavitation, hence a lower head drop. From this came the suggestions of NPSH
required based on 1 percent and 0 percent head drop. Defining NPSH required as that at which the total
head has deteriorated by 1 percent suffers the limitation that the deviation is equal to the accuracy of
measuring the parameter whose deviation is being sought. In other words, the deviation necessary to
500 Suction Conditions and Limitations on Suction Performance

Fig. 19.23 Enlarging impeller eye area to reduce required NPSH.

CAPACITY IN M3 HR.
25 50 75 100 125
80
100
-r---.
300
V
l"oo" ~
A '" 70 ~
o

~ r-- ...... ~ ~

I-
UJ
250 75
(J) ~~
/ "" B
-- r-.... r-..... 60 tl::
UJ
UJ
U. lUi
t;; JIf'
~ 200 50
CI
c:(
:IE
z /
UJ
J: 150
50
L A = DESIRED O,H.
B = BEST EFF. O.H. 30
..J
~ I 8
I-
0
I- 100 J / rn
UJ
cc 6 20 ~
I ./
~ w

-
25 ~
I(
~
~ ~ 4 en
10 ~
/ ~
2
I{ o
o 80 160 240 320 400 480 560 640
CAPACITY IN GPM

Fig. 19.24 Effect of oversizing a pump.


Suction Conditions and Limitations on Suction Performance 501

Dia., in. 3,550 rpm


a>
2
0
6.50 in.
6.20 in. , \
\
.....
(/) <0 \
<D
a> 5.90 in. ~
E 0 1,
-.:t
0
-.:t
0 5.60 in.
C\J
"C
C1l

::M
<Do
C1l
0
0
5.30 in.
a>
2
.....
(/)

.i!l
<D
~ 0
<X>
0
C\J
E
<0

0 ll')
C\J 0 ll')
<0
-.:t
0
M

------- C\J
ll')

6.50 in.
o o
o _ __-:::::::::~:: :::c
,~~~~;;;==~~====~:~5.90
"C
.0" CI)
w, 5.30 in. in. <D a..
bhp a: z
50 100 150 200 gpm
, ,
o 10 20 30 40 m3/hr.
Capacity

Fig. 19.25 Performance of pump with and without inducer.

define NPSH required cannot be separated from the inherent variations in the measured pump head.
Zero percent head drop has no meaning in a definition based on deviation of a measured parameter, and
therefore is not a valid definition. From the point of view of achieving rated performance, it is important
in applying the pump to know the NPSH needed to ensure it produces 100 percent head. This is one
purpose of the NPSH margin, but that value will be a conservative estimate from test data not an attempt
to discern just where the head starts to deviate from 100 percent.
These problems with lowering the defining head drop were compounded by experience that showed
some pumps were suffering severe cavitation erosion even when operating with NPSH well above that
at which there was any discernible deterioration in the total head (Fig. 19.26). This pointed to the need
to adopt another means of investigating the development of cavitation in these pumps. Flow visualization,
the technique used to verify the cavitation performance of model hydro-turbines and pump-turbines,
filled this role. Research by many investigators, [19.2], [19.3], [19.4], [19.5], [19.6] established that the
principal factors leading to rapid erosion despite high NPSH margins were the following:

• Impeller energy level, which is related to the inlet peripheral velocity, VI' Generally impellers with VI of 30
m/s (100 ft/s) and higher are considered "high energy". Specific speed does have an effect, however, and at
high specific speeds the defining limit falls to 25 m/s (80 ft/s).
S02 Suction Conditions and Limitations on Suction Performance

Fig. 19.26 Effect of erosion, noise, and pressure pulsations as available NPSH is decreased.
(Reprinted with pennission from Oil & Gas Journal, Nov. 19, 1984)

• Operating capacity as a fraction of impeller shockless capacity(l).


• Intensity of cavitation associated with suction recirculation, which generally increases with suction specific
speed, S (see chapter 22), but varies for any given S with the detail design of the impeller inlet, particularly
its vane geometry.
• Nature and temperature of the pumped liquid, with pure liquids of high specific gravity, e.g. water and water
solutions, being the most aggressive, mixtures of low specific gravity, e.g. light hydrocarbons, the least.
• Strength and resilience of the impeller material, with high strength and high resilience lowering the erosion rate.

One suggestion that arose out of flow visualization research was to redefine required NPSH as that
at the inception of caviation (Fig. 19.27). This was a philosophically "pure" concept, but one that quickly
ran afoul of practicality when it was realized that the 90 percent or more of the pumps currently giving
good service would need a significant increase in NPSH to satisfy this new definition. Following that
short flirtation with a radical definition, the needs of the pump user began to prevail. These reduce to
two simple questions:

1. How much NPSH should be provided to ensure the pump develops essentially rated total head?
2. How much NPSH should be provided, what materials should be used, and what precautions must be applied
in operation to obtain a reasonable life for the component parts of the pump?

In current application practice, the answers to these questions depend primarily on the energy level of
the impeller (first stage impeller of multistage pumps). The impeller categories used and the corresponding
application practice are:
High energy impellers (VI ~ 25-30 m/s (80-100 ft/s); see earlier discussion). The design or a model

(I) Capacity at which the liquid flow angle equals the impeller vane angle at the eye periphery; usually above pump BEP by
10 percent or more depending on the design.
Suction Conditions and Limitations on Suction Performance 503

., NPSHA

POTENTIAL .... 1_ ., +
EROSION / " 1
NPSHd 1 , • '- -I MARGIN
NPSH 1 " .
"""-- ' __ •
/ '
1/ 1
~
...... ----,,,,.
'_~_--,-+_
iI
NPSHR 1

(2) ....

(1)-~--~
I,
(1) Typical test characteristic
(2) Rise in characteristic observed in some tests
1 1

BEP MAX
FLOW

Fig. 19.27 Desired relationship between available NPSH, NPS~, and NPSH3 over a pump's
operating flow range.

of it is verified by flow visualization to ensure cavitation develops uniformly, and to establish the NPSH
at 3 percent head drop, NPSH3, and at what NPSH the impeller is likely to suffer damage at a rate
compatible with desired impeller life. This value, which is designated NPS~, corresponds in most
practical cases to vapor cavities 10 mm (0.4 in) long. Production tests of the design or others factored
from it are by 3 percent head drop. Such impellers are applied with the available NPSH above NPS~
over the expected operating flow range (Fig. 19.27). The margin over NPSH 3 for well designed impellers
ranges from 100 to 200 percent. In applications subject to transient conditions that can lower the available
NPSH substantially, such as boiler feed, it is necessary to check that this margin is sufficient to prevent
the pump from flashing during the transient (see chapter 26).
Medium and low energy impellers. Pumps that fall into this category-more than 90 percent of
applications-are applied with various margins over NPSH3• The magnitude of the margin is sufficient
to: a) ensure the pump develops essentially rated total head, and b) from experience, avoid serious
cavitation erosion in impellers of moderate suction specific speed (see Chapter 22 for a detailed discussion
of the effect of suction specific speed on impeller erosion). Table 19.2 summarizes typical margins for
both categories of impellers. These margins assume impellers of adequate materials (see Chapter 17).
The lower margin value for each application corresponds to low energy impellers, those whose inlet
peripheral velocity, VI. is 15 m/s (50 ft/s) or lower.
In assessing past experience against these margins, it is important to recognize that many, if not in
fact most, centrifugal pumps normally operate with a greater NPSH margin than that given by the
difference between the NPSH available from the data sheet and NPSH3 for the pump. The additional
504 Suction Conditions and Limitations on Suction Performance

Table 19.2 Typical NPSH Margins

NPSH Margin
Application (% of NPSH3)

Cold water 35-50(1·2)


Hydrocarbon 10(1)
Boiler feed-small 50(3)

High energy 100-200


Notes:
1. 1 m (3 ft) minimum
2. Varies with pump size
3. Pumps to 1,850 kW (2,500 HP) at
3,600 RPM

margin generally comes from conservatism in plant design by using an abnonnally low liquid level, an
abnonnally high liquid temperature if the suction vessel is not saturated, an abnonnally high suction
line friction loss, or a combination of all three. This is not always a conservative approach since it can
lead to a pump that operates poorly all the time rather than cavitating on the few occasions extreme
suction conditions are encountered (see Chapter 22). If we, pump users and manufacturers alike, are to
refine the application of centrifugal pumps in tenns of cavitation, it is important that the data used are
accurate. This means the stated NPSH available should be that under nonnal conditions with the minimum
under extreme conditions given for additional infonnation.
Doolin [19.7] has proposed a general approach to the question of impeller life. By this method, a
series of factors is applied to detennine the relative life of a given impeller in various applications. The
factors considered are:

• Actual margin over NPSH3


• Thermodynamic properties of the pumped liquid
• Corrosiveness of the pumped liquid
• Impeller material
• Operating speed
• Off-design operation
• Suction specific speed
• Pump duty cycle

The NPSH margin is assumed in all cases to be sufficient to allow the pump to produce rated head.
And all the factors assume the suction piping is able to deliver unifonn. swirl-free flow to the impeller
(see Chapter 28).
With a "base" impeller life and a "standard" value of NPSH required, this method can be used to
determine the NPSH that should be made available to ensure the pump impeller achieves its desired life
in any given service. Because the calculation is relative, the "standard" NPSH required can be whatever
is convenient. which today is NPSH3• The difficulty with this approach is detennining the value of the
various "life" coefficients. At the present state of the art, we have only meager data on the effect of the
numerous factors known to affect impeller life, and are hampered in improving those data by the difficulty
of carrying out qualified erosion rate tests. The saving grace is that reasonable data are available for the
effect of water, both cold and hot. on various materials. Since water has a very high potential for causing
Suction Conditions and Limitations on Suction Performance 505

cavitation erosion, the data that are available allow adequate designs for those services. The lack of data
for other liquids may mean that today's applications involving these liquids are overly conservative, but
that is both a safe approach and an opportunity for the future.
Taking account of where we are today and noting the progress being made in cavitation research,
the future holds the possibility of being able to state the required NPSH in terms of a given life, what
we might term NPSHLxo with the value "x" being selected at the design stage taking account of the cost
of impeller replacement versus the cost of providing various amounts of NPSH. In this connection,
Vlaming [19.4] has already proposed 40,000 hours impeller life as a standard and presented means to
estimate the NPSH that must be provided to achieve it. More recently pump manufacturers, [19.8],
[19.9], have reported the development of computer models able to estimate impeller life and the probability
that the estimated life will be achieved. These computer models are based on correlations of cavitation
intensity and impeller erosion rate. Cavitation intensity is, in tum, derived from bubble length and
cavitation noise measurement. Although these models are not yet in wide use, and don't take into account
all the factors involved in the complex phenomenon of cavitation, they do represent a notable step toward
being able to assess impeller life in terms of cavitation erosion in the same way we now treat other
components of turbomachines such as bearings and shafts.

BIBLIOGRAPHY

[19.1] ANSI/HI 1.6, 1994, Centrifugal Pump Test, Hydraulic Institute, Parsippany, NJ, USA.
[19.2] Robert R. Ross, Theoretical Predictions of Net Positive Suction Head Required (NPSHR) for Cavitation Free
Operation of Centrifugal Pumps, United Centrifugal Pumps, San Jose, USA. [or similar paper by Johnson
that preceded this].
[19.3] E. Grist, Net Positive Suction Head Requirements for Avoidance of Unacceptable Cavitation Erosion in
Centrifugal Pumps, I. Mech. E Paper C163n4, London, 1974.
[19.4] D.J. Vlaming, A Methodfor Estimating the Net Positive Suction Head Required by Centrifugal Pumps, ASME
Paper No. 81-WA/FE-32, 1981.
[19.5] P. Cooper and F. Antunes, Cavitation Damage in Boiler Feed Pumps, Symposium Proceedings: Power Plant
Pumps-The State of the Art, EPRI CS-3158, July 1983, pp. 2-24 to 2-49.
[19.6] B. Schiavello, Visual Study of Cavitation-An Engineering Tool to Improve Pump Reliability, EPRllst
International Conference on Improved Coal Fired Power Plants, Palo Alto, California, November 19-21, 1986.
[19.7] J.H. Doolin, Judge Relative Cavitation Peril with Aid of These 8 Factors, Power Magazine, pp. 77-80,
October 1986.
[19.8] J.F. Gulich and S. Pace, Quantitative Prediction of Cavitation Erosion in Centrifugal Pumps, Proc. of the
13th IAHR Symposium on Progress in Technology, Montreal, Canada, September 1986, Paper #42.
[19.9] B. Schiavello and M.R. Prescott, Field Cases Due to Various Cavitation Damage Mechanisms: Analysis and
Solutions, Porch. EPROM Symposium: Power Plant Pumps, June 1991, Tampa, Florida.
20
System-Head Curves

A centrifugal pump must be suitable for operation with the system in which it is used. To select a
suitable pump, the characteristics of the system must be considered. It is usually easy to determine the
characteristics of the system, but, occasionally, a complicated system that requires analysis of each of
its parts is encountered. It is not possible to provide a detailed analysis of every type of problem that
may be encountered when using centrifugal pumps. However, the following discussion of a few typical
examples will acquaint the reader with the general method of solving such problems.
The total operating head for a given capacity through a system is the algebraic sum of the static head
from supply level to discharge level (Hst ); the terminal pressure minus the suction pressure (P rIPs); all
friction losses at this capacity (hI); and the entrance and exit losses (h j and he). These values are expressed
in meters (feet) of the liquid being handled (see Fig. 18.15).
Ideally, the simplest system would have only one static head. In an actual system, there would also
be some friction losses. If there is no static head component and no difference in pressure on the suction
and discharge liquid levels (Fig. 20.1), the head would be entirely frictional.

CAUSES OF FRICTION

The characteristics of the flow of liquid in a pipe vary with the velocity. When the velocity is very low,
the flow is laminar. Under these conditions, the effect is that of concentric cylinders of the liquid shearing
past each other in an orderly fashion. The greatest velocity is at the center of the pipe; the velocity falls
to zero at the pipe walls. With water, laminar flow occurs when the average velocity is very low. As a
result, laminar flow with water is rarely encountered in normal applications. As the average velocity of
the liquid is increased, the flow becomes turbulent. Under turbulent flow conditions, the axial velocity
measured across the pipe diameter is more uniform than in laminar flow; the flow is laminar in an area
adjacent to the pipe walls. The average velocity at which the flow changes from laminar to turbulent is
not absolute; there is a critical range in which the character of the flow may be of either type.
The flow of any liquid is accompanied by two types of friction: internal friction caused by the rubbing
of the fluid particles against one another and external friction caused by the rubbing of the fluid particles

506

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
System-Head Curves 507

DISCHARGE PIPING
ONE a-IN. CHECK VALVE
SUCTION PIPING ONE a-IN. GATE VALVE
ONE 8- x 12-IN. INCREASER
ONE 10- xI2-IN. REDUCER
3,000 FT OF 12-IN. PIPE
ONE 12-IN. LONG RADIUS ELBOW
TWO 12-IN. LONG RADIUS ELBOWS
20 FT TOTAL OF 12-IN. PIPE
THREE 12-IN. 45 DEG ELBOWS

Fig. 20.1 Simple pumping system with head that is entirely friction.

against the pipe walls or against the static layer of liquid adhering to the walls. Energy must be expended
to overcome this friction.
If the flow is turbulent, the friction developed is partly dependent on the roughness of the walls.
Because the interior surfaces of pipes of the same material are practically the same irrespective of
diameter, small pipes are relatively rougher than large ones. Thus, for equal velocities, the larger the
pipe, the smaller will be the friction loss. The roughness of the pipe wall also depends on the material
from which the pipe is made and, after the pipe has been in service, on any change that occurs at the
inner surface.
Numerous pipe friction experiments and studies have been made, and a great number of tables and
charts are available. The Williams and Hazen tables are one of the earlier standards for water and have
been found particularly reliable for cast-iron pipes of 3-in. or larger diameter. These tables are based
on an empirical formula that can be modified to the following form:

hi = 10.45 (QIC)L852(L/(N 87

where hi =loss in head for length L, in feet of water


Q = flow, in gallons per minute
d =inside diameter of the pipe, in inches
C =coefficient of pipe smoothness
L = length of pipe, in feet

The coefficient C is an index of the smoothness of the interior pipe surface (the smoother the pipe
interior, the higher the C value), and the selection of the proper value of this coefficient will determine
the accuracy of the friction head loss calculated for any problem. For new unlined cast-iron pipe, a C
of 130 is the common value, but some new pipes in which the friction head losses indicate C values of
140 or higher have been encountered. Pipes coated on the interior to give a smoother surface naturally
have a higher C value. There are also records of pipes made of rolled metal and very smooth cement
that have C values of 145 to 150 or higher.
50S System-Head Curves

Most pipes deteriorate with age, and thus the C value becomes lower. This decrease in C value or
increase in friction head loss depends upon the material of the pipe, the pipe coating used (if any), and
the character of the water. Therefore, any C value selected for an old pipe represents a pure guess. When
it is necessary to ascertain friction head losses in such pipe, a test should be made, if possible, to find
the friction loss at some known capacity so that the coefficient can be approximated. If such a test cannot
be made, some guide to indicate the average C change with age is desirable. Figure 20.2 shows in chart
form the coefficients that might be expected for cast-iron pipes handling soft, clear, unfiltered water.
Pipes carrying water that has been filtered but not chemically treated have been found to deteriorate
less rapidly than pipes handling unfiltered water. Chemically treated waters have sometimes been found
to produce more corrosion in the pipe than untreated water. Brackish water usually results in increased
tuberculation. Some moderately hard waters have been found to cause a slow rate of deterioration. On
the other hand, they are also capable of depositing calcium carbonate on the interior of the pipe, thus
both reducing its size and increasing its roughness. Smooth cement and cement-lined pipes have been
found to maintain a high C value for many years.

140 I

'NEW'
130 I 30

.
II:
00
S'YEARS ... -,
120 I 20
Zo
Qo
,
-
~-
\) 10 yEARS u~
...: liD I 10 ...i o....
Z
....
2
,J YEiRS ....:;,
~..J

--
~
~~
....0
~
100 26 yJARS I 00 1.0

-.-- -
"""". Q\)
..,..",.. ~ J
~
u II:
~
Z 25 yEARS ~c(
.... ~
I""'"
~Z
N
c( 90 ~ .",...
~ 'liAR"!. 1.2 a. ....
:J:
:J: ~ ".,..,.. ell:'
./
".,.
0
Z ~ io"""" Z
-en
• i""" ~
Z ....
~'ARS
c( 1.4 -:;,
~ ~ :E~

-
<II 1I:c(

"" /
:E 80

l JA~
~>
!! ,/' ",
"""" 1.6
~
~ ./ ~
~
)9-
i"'"""
~\)

V ,
i II:~

"". i-""
1.8 .......
Oz
~ .... 11:
70 ,r ..J ....
2.0 c(~
,/
/
u~
enO
~

60
3
/ 4
V
5
V
6 8 10 12 14 16 1.20 24 3036 4248 60
60
2.5

INSIDE DIAMETER OF PIPE. IN INCHES

Fig.20.2 Change in Williams and Hazen coefficient C with years of service, for cast-iron pipes handling soft,
clear unfiltered water.
System-Head Curves 509

All these possibilities make it difficult to select, with any assurance of accuracy, the proper coefficient,
so that any guide should be used with reservations. In important studies it is often possible to locate a
similar installation and use the head losses obtained on that installation as a guide.

ESTIMATING FRICTION LOSS

When selecting pumping machinery, it is particularly desirable to consider the friction head loss that
may occur when the installation is new, as well as that which may result some years after it has been
in service. Most charts and tables based on the Williams and Hazen formula have been made for a C
value of 100, which is approximately the value expected for pipe that is 15 years old. A C value of 100
is commonly used as a design value; no calculation is made to determine the friction when the installation
is new. This practice tends to distort the problem of compensating for friction losses, and may result in
trouble when centrifugal pumps are used. For example, in a new installation in which most of the
pumping head is friction, the actual friction head would be lower than that allowed for in the selection
of the pump. As a result, the pump would deliver more capacity at some reduced head that would equal
the system head. The increased capacity would depend on the pump characteristics and the increase in
system head with capacity, but might be 15 to 25 percent more than the capacity for which the pump
was selected. Operation at this increased capacity might cause the pump to require more power, thus
overloading the driver. If the installation was such that the available net positive suction head (NPSH)
at the design capacity exceeded only slightly the NPSH required by the pump for this capacity, the
resulting increase in pump capacity with the lower operating head would result in cavitation.
Table 20.1 shows pipe friction losses for 3-in. to 24-in. inside diameter pipes based on the Williams
and Hazen formula with C of 100. The usual values of C for new pipe are:

Smooth, unlined cast iron 130


Asphalted cast iron 140
Cement asbestos 130-140
Very smooth cement or cement-lined cast iron 130-140
Ordinary cement 110-120
Drawn steel or wrought iron 130-140
Riveted steel 90-110

Conversion factors for changing friction values based on C = 100 to other values are indicated on
the right-hand side of Fig. 20.2. For example, with a flow of 700 gpm through a 6-in. pipe, the friction
head loss is 6.23 ft per 100 ft of pipe with C = 100. For C = 130, the conversion factor is 0.613;
therefore, the friction head loss will be 6.23 x 0.613 or 3.83 ft per 100 ft of pipe.
Steel and wrought-iron pipes are used extensively in sizes up to 8 in. and larger with cold water. In
drainage and irrigation work, steel pipe is used almost exclusively with larger sizes fabricated of steel
plate. Cast-iron pipe is now seldom used for water lines in sizes less than 3 in. Most long water lines
are cast iron, although plastic is growing in popUlarity for smaller diameter lines. Cement asbestos is
not used in new lines today. Steel pipe is made with the same outside diameter for a number of different
weights or wall thicknesses. Therefore, the inside diameter will not be the same as the nominal diameter,
and the friction losses for a given capacity must be corrected for such differences. With steel and
wrought-iron pipe, it is more difficult to predict the change in friction head that will result when the
pipe becomes older than it is to predict the changes in cast-iron pipes. In some situations, the pipe
decreases in area due to tuberculation, while in other situations the pipe corrodes and the film is washed
u.
Table 20.1 Velocity and Friction Head Loss in Old Piping
...
~ Friction values apply to cast-iron pipes after 15 years service handling average water. Based on Williams and Hazen's formula with C = 100 .

gpm VI v f v f v f v f v f v f v f v f gpm
f
3-in. ID pipe 4-in. ID pipe

30 1.36 0.534 0.77 0.131 5-in. ID pipe 30


40 1.81 0.910 1.02 0.224 40
50 2.27 1.38 1.28 0.338 0.82 0.114 6-in. ID pipe 50
60 2.72 1.92 1.53 0.475 0.98 0.160 60
70 3.18 2.56 1.79 0.631 1.14 0.213 0.79 0.088 70
80 3.63 3.28 2.04 0.808 1.31 0.273 0.91 0.112 80
90 4.08 4.08 2.30 1.01 1.47 0.339 1.02 0.139 90
100 4.54 4.96 2.55 1.22 1.63 0.412 1.14 0.170 8-in. ID pipe 100
125 5.68 7.50 3.19 1.85 2.04 0.623 1.42 0.256 125
150 6.81 10.5 3.83 2.59 2.47 0.874 1.70 0.360 0.96 0.089 150
175 7.95 14.0 4.47 3.44 2.86 1.16 1.99 0.478 1.12 0.118 175
200 9.08 17.9 5.10 4.41 3.27 1.49 2.27 0.613 1.28 0.151 IO-in. ID pipe 200
225 10.2 22.3 5.74 5.48 3.68 1.85 2.55 0.762 1.44 0.188 225
250 11.3 27.1 6.38 6.67 4.08 2.25 2.84 0.926 1.60 0.228 1.02 0.077 250
275 12.5 32.3 7.02 7.96 4.50 2.68 3.12 1.11 1.76 0.272 1.12 0.092 275
12-in. ID pipe
300 13.6 37.9 7.65 9.34 4.90 3.13 3.41 1.30 1.91 0.320 1.23 0.108 300
350 0.99 0.059 14-in. ID pipe 15.9 50.4 8.93 12.4 5.72 4.20 3.97 1.73 2.23 0.425 1.43 0.144 350
400 1.13 0.076 18.2 64.6 10.2 15.9 6.54 5.38 4.54 2.21 2.55 0.545 1.63 0.184 400
450 1.28 0.094 0.94 0.044 11.5 19.8 7.36 6.68 5.10 2.75 2.87 0.678 1.84 0.228 450
500 1.42 0.114 1.04 0.054 16-in. ID pipe 12.8 24.1 8.18 8.12 5.68 3.34 3.19 0.823 2.04 0.278 500
550 1.56 0.136 1.15 0.064 14.0 28.7 8.99 9.69 6.24 3.99 3.51 0.982 2.24 0.331 550
600 1.70 0.160 1.25 0.076 0.96 .039 15.3 33.7 9.81 11.4 6.81 4.68 3.82 1.15 2.45 0.389 600
650 1.84 0.186 1.36 0.088 1.04 .046 18-in. ID pipe 16.6 39.1 10.6 13.2 7.38 5.43 4.15 1.34 2.65 0.452 650
700 1.99 0.214 1.46 0.100 1.12 .052 17.9 44.9 11.4 15.1 7.94 6.23 4.47 1.53 2.86 0.518 700
750 2.13 0.242 1.56 0.114 1.20 .060 0.95 0.034 12.3 17.2 8.51 7.08 4.78 1.74 3.06 0.589 750
800 2.27 0.273 1.67 0.129 1.28 .067 1.01 0.038 20-in. ID pipe 13.1 19.4 9.08 7.98 5.10 1.97 3.26 0.666 800
900 2.55 0.339 1.88 0.160 1.44 .084 1.13 0.047 14.7 24.1 10.2 9.92 5.74 2.44 3.67 0.825 900
1,000 2.83 0.412 2.08 0.195 1.60 .102 1.26 0.057 1.02 0.034 16.3 29.3 11.4 12.1 6.38 2.97 4.08 1.00 1,000
1,100 3.12 0.492 2.29 0.232 1.76 .121 1.39 0.068 1.12 0.041 18.0 35.0 12.5 14.4 7.02 3.55 4.50 1.20 1,100
1,200 3.40 0.578 2.50 0.273 1.92 .143 1.51 0.080 1.23 0.048 13.6 16.9 7.66 4.17 4.90 1.41 1,200
Table 20.1 Continued

12-in. ID pipe 14-in. ID pipe 16-in. ID pipe 18-in. ID pipe 20-in. ID pipe 6-in. ID pipe 8-in. ID pipe lO-in. ID pipe
1,300 3.69 0.671 2.71 0.316 2.08 0.165 1.64 0.093 1.33 0.056 24-in. ID pipe 14.8 19.6 8.30 4.83 5.31 1.63 1,300
1,400 3.97 0.770 2.92 0.363 2.24 0.190 1.76 0.107 1.43 0.064 15.9 22.5 8.93 5.54 5.72 1.87 1,400
1,500 4.25 0.875 3.12 00413 2.40 0.215 1.89 0.121 1.53 0.073 1.06 0.030 17.0 25.5 9.55 6.30 6.12 2.13 1,500
1,600 4.54 0.985 3.33 00465 2.55 0.243 2.02 0.137 1.63 0.082 1.13 0.034 18.2 28.8 10.2 7.10 6.53 2.39 1,600
1,800 5.11 1.22 3.75 0.578 2.87 0.302 2.27 0.170 1.84 0.102 1.28 0.042 11.5 8.83 7.35 2.98 1,800
2,000 5.67 1.49 4.17 0.703 3.19 0.367 2.52 0.207 2.04 0.124 1.42 0.051 12.8 10.7 8.17 3.62 2,000
2,500 7.09 2.25 5.21 1.06 3.99 0.555 3.15 0.312 2.55 0.187 1.77 0.077 16.0 16.2 10.2 5048 2,500
3,000 8.51 3.16 6.25 1.49 4.78 0.778 3.78 00438 3.06 0.262 2.13 0.108 19.1 22.8 12.3 7.67 3,000
3,500 9.93 4.20 7.29 1.98 5.59 1.04 4.41 0.583 3.57 0.349 2.48 0.143 14.3 10.2 3,500
4,000 11.3 5.38 8.33 2.54 6.39 1.33 5.04 0.746 4.08 0.447 2.83 0.184 16.3 13.1 4,000
4,500 12.8 6.68 9.38 3.15 7.18 1.65 5.67 0.928 4.59 0.555 3.19 0.228 1804 16.3 4,500
5,000 14.2 8.13 lOA 3.83 7.98 2.00 6.30 1.13 5.10 0.675 3.54 0.278 5,000
5,500 15.6 9.70 11.5 4.58 8.78 2.39 6.93 1.35 5.61 0.806 3.90 0.332 5,500
6,000 17.0 11.4 12.5 5.38 9.68 2.81 7.56 1.58 6.12 0.947 4.25 0.390 6,000
6,500 18.4 13.2 13.6 6.24 lOA 3.26 8.19 1.83 6.73 1.10 4.61 00452 6,500
7,000 19.9 15.2 14.6 7.16 11.2 3.74 8.82 2.11 7.15 1.26 4.96 0.518 7,000
7,500 15.6 8.13 12.0 4.24 9045 2.39 7.66 1.43 5.32 0.589 7,500
8,000 16.7 9.16 12.8 4.79 10.1 2.69 8.17 1.61 5.66 0.664 8,000
9,000 18.8 11.4 1404 5.95 11.3 3.39 9.18 2.01 6.38 0.825 9,000
10,000 16.0 7.24 12.6 4.07 10.2 2.44 7.09 1.00 10,000
11,000 17.6 8.63 13.9 4.86 11.2 2.91 7.80 1.20 11,000
12,000 19.2 10.1 15.1 5.71 12.3 3042 8.51 1.41 12,000
13,000 1604 6.62 13.3 3.96 9.12 1.63 13,000
14,000 17.6 7.59 14.3 4.54 9.93 1.87 14,000
15,000 18.9 8.63 15.3 5.27 10.6 2.13 15,000
16,000 16.3 5.82 11.3 2040 16,000
18,000 1804 7.24 12.8 2.98 18,000
20,000 14.2 3.62 20,000
25,000 17.7 5048 25,000
·Velocity, in feet per second.
III
........ lfriction head loss, in feet of water per 100ft of pipe.
512 System-Head Curves

away. With smaller steel piping, it is best to consider what the friction head will probably be when the
pipe is new and to make allowance for an increase in the loss based on local conditions. If a basis of
comparison is not available, an increase in friction of 25 percent with age would be a reasonable allowance.
A general solution for the head loss caused by incompressible flow in piping is given by the Darcy-
Weisbach equation [20.1]

hI =j(UD)(V'l!2g)
where: ht =head loss caused by friction, in meters (feet) of liquid
L = length of straight pipe run, in meters (feet)
D = inside diameter of the pipe, in meters (feet)
V = average liquid velocity, in m/s (ft/s)
g = acceleration due to gravity; 9.S1 m/s2 (32.2 ft/s 2)
1 =friction factor
The value of the friction factor, f, depends on the relative roughness of the pipe, kID, and the Reyn-
olds number

Re = VDlv
where: k = pipe roughness, in m (ft)
v = kinematic viscosity, in m2/s (ft2/s)

For laminar flow, Re < 2,320, the value of / is independent of the relative roughness of the pipe bore,
and is given by the Hagen-Poiseuille equation

1 = 64IR e
The flow in most pumping applications is turbulent with Re > 2,320, for which the value of/is determined
from the Prandtl-Colebrook equation

1/"1 = -2log (2.51/R;.JI + kI3.71D)

For the limiting case of hydraulically "rough" piping, R.(k/D) > 1,300, the value of/can be determined
from the simpler Nikuradse equation

INI = 1.14 - 2log(klD)

Typical values of k are:


k
Pipe material and lorm mm in
Smooth; plastic, glass, copper, brass; up to 0.002 SOJ1in
drawn, extruded, ground finish
Cast iron, cement lined 0.025 0.001
Cast iron, asphalt lined 0.10-0.15 0.004-0.006
Cast iron, unlined, new 0.15-0.25 0.006-0.010
Cast iron, encrusted 1.5-3.0 0.060-0.120
Seamless steel, new up to 0.05 up to 0.002
Welded steel, new 0.05-0.10 0.002-0.004
Galvanized steel up to 0.15 up to 0.006
Welded steel, corroded 0.15-0.20 O.OO6-O.00S
Concrete, new, rendered 0.20-0.S0 0.OOS-O.030
Heavily encrusted, timber, masonry 0.5-2.0 0.20-0.S0
System-Head Curves 513

To allow estimates of piping friction loss without having to repeatedly solve the Prandtl-Colbrook
equation, charts or tables are often used. Table 20.2(a) [20.2] gives friction losses in metric units for
liquids of v = 1.236 mm 2/s (pure fresh water at 12°C or 54°F) and k = 0.10 mm. Friction losses in pipes
with k other than 0.10 mm, are determined by applying the correction factor from Table 20.2(b). Table
20.2(c), drawn from data published by the Hydraulic Institute [20.3], shows friction losses in US units
for liquids of v = 1.130 mm2/s (l2.16xlO-6 ft 2/s; equal to pure fresh water at 60°F) and k = 0.05 mm
(0.002 in or 0.00015 ft).

I~.5 I 2 3 , 5 6
Flowrate 0 In m'lh -
& 10 20 3O~ 60 eo 100 200 JOO
flowa,ele 0
600 800 1000
In m' /h - -
2000 JOOO 5000 10000 20 000 50000
I 00
In

l'. I ~.~ "'

.
60 50
I _I ......
50 50
40 <0
JO
.
t!
~ $
10
y.....
0 2o

::; dl
r---. "'·,6
0 ~IL ... ~ 10
& ~ &
6 ~s

.
5
5 S
$

"!O
<
3 I'll ~
"
r{
<
3
.
E
Qj 1 8.
Co
c, 27- l' 's
... V 111 2 'c,
'0 '0
''/s ;:
"
co
c
!! I
I
~ "-s
II'
I
'"c
!!
E E
§ 0.& 8
"
00&
11,
l',
0.6 0.6
i
;;; ~
Co 0.5 'tl 0oS
i 0,4 11 ~s
~
:t
0
.' =
...,
IS;

".,.
!2 0.1 11$ 1.,/ 0.3 ~
I
0.2 Il' II I
V 0.2 ~
~ 04
...
...
0,1
'J
0'
"!- ... 0-'
oa 0.0&

Q06 0,06
0.os O.OS
o,00. ~ 004

001 001

0.02
J. ~ f' 0.02
l-
'll
-.....
001 0&,01
0.5 3 < 5 6 &10 203040 60 eo 100 200 JOO 600 1001000 2000 JOOO 5000 10000 20000 ~

Flowrale a In m )/h - FlowratftO inm'/h _ _

Table 20.2(a) Head loss, hI' per 100 m of straight pipe.


(Based on Prandtl-Colebrook equation with k = 0.1 mm, turbulent flow, and v = 1.2136 mm2/s
(pure fresh water at 12°C).
514 System-Head Curves

U 2.1

'r-...
2.6
r--..r-... 2.6
I" '
2.4
.........
" I'r-- '-!,I 2.4
r-.... "~~
I""-
,
r--
2.2
~ r--t---.,
2.2
f'.. r- 1"-1'-
i"-. ........ r-- ~..~
2.0 2.0
......... ......
~ r--r--
.......... r- r- "'1-
1.8 t"-. I.e
r... ~~
(; . . . r- (;
U i'"- r-!-.: I "'", U
~
c:
1.&
l"'- I"'-
I.' ~
c:
.~ I'-~ r- .~
U U
~ ~
(; 1.4 1.4 (;
I.) I.)

1.2 " • 0.25 mm 1.2

1.0
II .'0.1 ~~
1.0
k • 0.05 mm

0.1 0.'
o D.gs
1-":0.01 mm_ • ~ , ml,
o• y! 1 m~' 0.s
y. 1m"
,...
o" 25
2 ;0 iO 85
j
eo 100
I 200 )()
15()400 Soo! 00 8)0 1100 mm 2t,a
.6

nominal pipe diameter in mm.

Table 20.2(b) Correction factors for roughness k ± 0.1 mm.


For k = 0.05 to 3.0 mm the differences between the correction factors due to flow velocity can be neglected,
therefore average values are shown. However, for k = O.OJ mm the effect of flow velocity must be taken
into consideration.
Table 20.2(c) Velocity and Friction Head Loss in New Piping
Friction values apply to Schedule 40 (standard weight) steel pipe carrying water.

gpm VI h/ v f v f v f v f v f v f v f gpm

I-in. pipe
(1.049-in. 10) 1V4-in. pipe
(1.380-in. 10) 1V2-in. pipe
0.37 0.11 (1.61O-in 10)
2 0.74 0.39 0.43 0.10 2
3 1.11 0.82 0.64 0.21 0.47 0.10 3
4 1.49 1.37 0.86 0.36 0.63 0.17 2-in. pipe 4
5 1.86 2.08 1.07 0.54 0.79 0.26 (2.067-in. 10) 5
2Y2-in. pipe
6 2.23 2.83 1.28 0.76 0.95 0.35 0.57 0.10 (2.469-in. 10) 6
8 2.97 4.88 1.72 1.29 1.26 0.61 0.76 0.17 3-in pipe 8
10 3.71 7.12 2.14 1.95 1.57 0.90 0.96 0.26 0.67 0.11 (3.068-in. 10) 10
15 (3V2-in. pipe 5.56 15.0 3.21 4.06 2.36 1.87 1.43 0.54 1.00 0.23 15
20 (3.548-in. 10) 7.41 25.6 4.28 6.80 3.15 3.12 1.91 0.92 1.34 0.38 0.87 0.13 20
4-in. pipe
25 0.81 0.10 (4.026-in 10) 5.35 10.3 3.94 4.70 2.38 1.39 1.67 0.58 1.08 0.20 25
30 0.97 0.13 6.43 14.4 4.72 6.60 2.86 1.92 2.00 0.81 1.30 0.28 30
40 1.30 0.23 1.01 0.12 6.30 11.2 3.82 3.35 2.68 1.36 1.73 0.47 40
50 1.62 0.34 1.26 0.18 5-in. pipe 7.87 16.6 4.77 5.00 3.34 2.06 2.16 0.72 50
60 1.94 0.48 1.51 0.25 (5.047-in. 10) 5.72 7.00 4.02 2.85 2.60 0.99 60
70 2.27 0.63 1.76 0.34 1.12 0.11 6-in. pipe 6.68 9.40 4.68 3.80 3.03 1.33 70
80 2.59 0.82 2.01 0.43 1.28 0.15 (6.065-in. 10) 7.62 11.9 5.35 4.95 3.46 1.72 80
90 2.91 1.00 2.27 0.54 1.44 0.18 8.60 14.7 6.02 6.05 3.89 2.13 90
100 3.24 1.24 2.52 0.65 1.60 0.22 1.11 0.09 9.56 18.7 6.70 7.47 4.34 2.58 100
125 4.05 1.82 3.15 1.00 2.00 0.33 1.39 0.13 8.37 11.1 5.41 3.90 125
150 4.86 2.55 3.78 1.39 2.40 0.47 1.66 0.18 8-in. pipe 10.0 15.4 6.50 5.44 150
175 5.66 3.40 4.40 1.90 2.80 0.62 1.94 0.24 (7.981-in.lO) 11.7 20.8 7.58 7.30 175
200 6.48 4.35 5.04 2.40 3.20 0.80 2.22 0.31 8.66 9.18 200
III 225 7.30 5.44 5.66 2.98 3.60 0.97 2.50 0.38 1.44 0.10 9.75 11.6 225
...
III
2.77 0.47 1.60 0.12 10.8 14.0 250
250 8.10 6.59 6.29 3.68 4.00 1.19
!Jl
...
0\

Table 20.2(c) Continued

3Y2-in. pipe 4-in. pipe 5-in. pipe 6-in. pipe 8-in. pipe 3-in. pipe
(3.548-in. ID) (4.026-in. ID) (5.047-in. ID) (6.065-in. ID) (7.981-in. ID) (3.068-in. ID)
275 8.91 7.90 6.92 4.35 4.40 1.43 3.05 0.56 1.76 0.15 lO-in. pipe 11.9 16.9 275
300 9.72 9.30 7.55 5.04 4.80 1.65 3.32 0.66 1.92 0.17 (1O.020-in. ID) 13.0 19.6 300
350 11.3 12.2 8.80 6.85 5.60 2.21 3.88 0.88 2.24 0.23 350
400 13.0 15.9 10.1 8.67 6.40 2.89 4.44 1.12 2.56 0.29 1.62 0.10 400
450 14.6 20.0 11.3 10.9 7.20 3.56 4.99 1.40 2.88 0.37 1.82 0.12 450
500 12.6 13.3 8.00 4.36 5.54 1.72 3.20 0.45 2.03 0.15 500
550 13.9 16.0 8.80 5.17 6.10 2.06 3.52 0.55 2.23 0.18 550
600 15.1 19.1 9.60 6.16 6.65 2.42 3.84 0.63 2.44 0.21 600
650 10.4 7.22 7.20 2.78 4.16 0.73 2.64 0.24 650
700 11.2 8.29 7.75 3.25 4.47 0.85 2.84 0.28 700
750 12.0 9.40 8.31 3.63 4.80 0.97 3.04 0.31 750
800 12.8 10.3 8.87 4.11 5.11 1.11 3.25 0.35 800
900 14.4 13.0 9.96 5.12 5.75 1.33 3.65 0.44 900
1,000 16.0 15.8 11.1 6.17 6.40 1.64 4.06 0.55 1,000
1,100 17.6 19.0 12.2 7.45 7.04 1.98 4.46 0.64 1,100
1,200 13.3 8.73 7.67 2.36 4.87 0.75 1,200
1,300 14.4 10.2 8.31 2.71 5.27 0.88 1,300
1,400 15.5 11.9 8.95 3.10 5.68 1.02 1,400
1,500 16.7 13.2 9.60 3.49 6.09 1.18 1,500
1,600 17.8 15.0 10.2 3.92 6.49 1.31 1,600
1,800 20.0 18.5 11.5 4.99 7.30 1.60 1,800
2,000 12.8 5.96 8.11 1.97 2,000
2,500 16.0 9.00 10.2 2.95 2,500
3,000 19.2 12.5 12.2 4.15 3,000
3,500 22.4 16.6 14.2 5.60 3,500
4,000 16.2 6.90 4,000
4,500 18.3 8.80 4,500
5,000 20.3 10.8 5,000
5,500 22.3 13.0 5,500
6,000 24.4 15.3 6,000

'Velocity, in feet per second.


2Priction head loss, in feet of water per 1()() ft of pipe.
System-Head Curves 517

FRICTION LOSS IN VALVES AND FITTINGS

When liquid flows through valves, elbows, tees, and other fittings, there will be a frictional loss_
Irrespective of the pipe size, these losses in fittings and valves can be expressed as percentages of the
velocity head and may be calculated by the formula:

where hI = head loss, in meters (feet) of liquid


K = constant (depending on the fitting design)
V = nominal liquid velocity, in mls (ft/s)
g = acceleration due to gravity; 9.81 mls 2 (32.2 ft/s 2)

Values of K for common fittings, valves, and other resistances to flow have been determined experimen-
tally.

Type of resistance K value


Globe valve 10
Angle valve 5
Fully open swing check valve 1.5-2.5
Close return bend 2.2
Standard tee acting as elbow 1.8
Standard elbow 0.9
Long-sweep elbow 0.6
45-deg elbow 0.4
Fully open gate valve 0.2

The K values for various types of entrances are shown in Fig. 20.3. The K values for sudden
enlargements and sudden contractions are shown in Figs. 20.4 and 20.5.
There is such a wide variation in the design of check valves that it is impossible to give any general
values of K. In the swing-type valve, the disk is opened by the force of the flowing liquid. Thus, at low
velocities when the disk is not fully open, the flow is throttled and the loss measured in terms of the
velocity head is greater than at higher velocities. Very few data have been published on the loss in swing
check valves. One manufacturer provides a chart for equivalent length of pipe that gives values of K =
2.0, approximately, for valves of 3-in. to 24-in. size. Assumptions of a K value of 2.5 for I-in. valves,
2.0 for 2-in. valves and 1.5 for lO-in. and larger valves should give reasonable friction head allowances.
A check valve using a disk hinged slightly above its center has become quite popular. Published data
on two sizes of this type of valve indicate a K value of about 0.3 at all velocities. Fully open butterfly
valves should have K values in this range.
In many municipal or other important installations of large or fairly large size, a combined check
and stop valve design made on the principle of the plug cock, with the plug rotated by an external
mechanism, is used. This special valve has a straight full-size passage when fully open, and should have
a loss no greater than a section of pipe of the same length.
A flap valve is a form of a check valve used on the end of a pipe. The flap is quite light in most
designs. Some special designs with the flap partially counterweighted have also been made. Flap valves
have very low losses even at low velocities when the disk is not raised very high by the flow. In most
designs, a loss of 0.06 m (0.2 ft) of water irrespective of velocity should be ample allowance. The exit
loss equal to the velocity head at the valve must also be added. Losses in multiported check and foot
518 System-Head Curves

PLAIN END PIPE PROJECTING THROUGH WALL OR VERTICALLY INTO


BODY OF WATER WITH ADEQUATE SUBMERGENCE AND CLEARANCE

K=O.B

PLAIN END PIPE FLUSH WITH WALL


K=0.5

BELL MOUTH PIPE BELL MOUTH PIPE SUSPENDED IN


FLUSH WITH WALL BODY OF WATER, WITH ADEQUATE
K= 0.1 (well rounded) SUBMERGENCE AND CLEARANCE
TO K=0.2(well rounded)
0.25 (slighlly rounded) TO
0.5(slighlly rounded)

Fig.20.3 Various types of piping connections and their K values.

valves vary too widely to make any assumption. This type of construction is now rarely used so that it
is likely to be encountered only in existing installations in which the loss can be detennined by test.
While designs of foot valves with strainers vary widely, a K value of 5 to 15 might be expected.
The value of K varies with the design of any valve or fitting and, in the case of elbows in which part
of the loss is due to the bend and part due to the length of pipe involved, the value of K varies with
the smoothness of the walls. Thus, calculated friction values are approximations, not definite values.
The approximate value of velocity head for any capacity in any size pipe up to NO 600 (24 in) pipe
can be quicldy obtained (Fig. 20.6). This value when multiplied by the appropriate K value gives the
head loss in the fitting. For example, with a flow of 340 m3Jhr (1,500 gpm) through NO 250 (10 in)
pipe, the velocity will be 1.86 mls (6.1 ft/s) for which the corresponding velocity head is 0.18 (0.58 ft).
The loss in any valve, fitting or other resistance can be expressed as the loss in a length of pipe of
the same size as the fitting. The total friction loss involved can be detennined for the total length of
System-Head Curves 519

1.0

-"" I
I I

'" '\1\ ~I : ~=
0.9
v-
0.8

0.7

0.6

:.c 0.5 1\
0.4
\
0.3 1\ \
0.2 \..
'\
'"
0.1

0 i'-..
o 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

dID

Fig. 20.4 Head loss in sudden enlargement of pipe.


Based on assumption that the difference in velocity head is lost.

0.4
"" ~

~
~,

0.3
"'" """-
......

'\.
~
0.2 1\

\
0.1
!\,
1\

o \
o 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
diD

Fig. 20.5 Head loss in sudden contraction of pipe.


Average of data obtained from various sources.
S20 System-Head Curves

25
!O,ooo

20 20,000 20 6
5
10,000 4
15 8,000 15
6,000 3
5,000

----- ---
4.000
2
3,000
I~
2,000

---
1.0
~OOO --_
(/) 800 -_ o.a

'"~6 6
600
500
o.s
0.5
~ 400
~5
5 300 5 OA 2
~ 0 :::>
~
(/) z 0
...J4 4 '"
~
200
8 4
0.3 :J
'"'" 3t u
~ 100 Q..
:::E '"
(/)
0.2
~
0
~ It: ~

'" :!l
(/)
~ 80 C!)
3 Q.. 0.15
0 3 ci 60 3 ~
It:
ct '" 50 ~ ~

~
ct
2t
~
I&J
:::E
40
30
>=
~
(3
'"'"
~
0.10
Z
Ci
~
ct ~ 0.08:
(/)2 0 ct
20 Q.. ~ ~
2 2
'" ct !:: 0.06~
~

'"~ !ill u u 0.05_


It
(/)
~ 10 9 U
0.040
...J
ct 1.5 8 ~ 1.5 ...J
Z
~
I.l.
• 6
5
0.03> '"
0 4
Z 0.02
3
1.0 1.0
2
C9
t 08 Q8 0.01
I
o.a 0.7
t Q6
OS
0.5 o.s
0.4
t 05 Q.3 0.5

Fig.20.6 Velocity of liquids in pipes.

piping, plus the equivalent lengths of all the valves, fittings, and other resistances. Using the K value
of the fitting, the resistance of any fitting expressed in equivalent length of pipe can be approximated
from the chart in Fig. 20.7. For example, the loss in a fully open ND 250 (10 in) gate valve (K = 0.2)
would be approximately the same as the loss in 1.8 m (6 ft) ofND 250 (10 in) pipe. It must be remembered
that any such conversion may result in a somewhat different value for the head loss in any fitting when
compared with the value obtained by the velocity head method. The magnitude of the difference depends
System-Head Curves 521

ilQIE ~~,
10 1,500
48

-IQOO 42
8
800 36
- 600
6 30
500
~E IIlLVE . 5
400
24
300
4 20
- 200 18

150 16
3
14
100
12
80

2 - 60 ~ 10

0
I&J (/)
(!) 50 I&J
~TANI2ARD U;~
AS ELBOW Z
40
LL. / I&J
X
<.>
i= ~ / 8
1.5 t- ~
G: 30
Ii / ~
/
..
0 it:
t- &J'
LL.
/ 6 6 Q.

~ ~
(!) 20 0 it:
Z
X / LL.
1.0 ~
0..
15 ~
/ 5 5 0
~___ STAN~DruS 0..
4
~/ a::
I&J
0.8 W
10 ~ 4 4
~
w
e/!Z 2

JZ-~~
::;)
....J
~ 76 ~ (/)
3t
~
~ 3
ELBOW
0.6
~ / 5
~ ~~ 3 ~
(i5
~ LATERAL
0.5 / 4 W
21
~

/ 3
~
it:
~ 45-DEGREE ELBOW 0.4 / ....J
w
2 2
/ 2 W
/ li;
1.5 It
0.3 / i 1.5
/ 1.0
8Z It
/ 0.8 ~
/
.
II)

I&J" 1.0
~Y.£.~ 0.2 0.6 -- 1.0
0.5 ~

~ ~ •
0.4 ~
0.8
iWo1TT Fm,AL
THOIiLlr 0.15
0.3
~ t
0.7

0.6
0.2

8 STRAIGHT FLOW
THROUGH TEE
0.10 0.15 i- 0.5

Fig. 20.7 Friction losses in fittings expressed in equivalent lengths of straight pipe.
522 System-Head Curves

on the Williams and Hazen C value or equivalent used in detennining the friction head loss per unit
length of pipe.
There is very little loss in taper reducers because a liquid can be accelerated with little loss. For long
reducers there will be a greater loss because of length. In such cases, determine the loss as for a pipe
with a diameter equal to the average diameter of the reducer. There is greater friction loss in increasers.
A taper increaser up to about a 15-deg included angle [Fig. 20.4, (D - dJ/L equals 0.266 or less) will
result in the water following the taper. A taper with over 60-deg included angle (Fig. 20.4, [D - dJ/L
equals 1.15 or more) will have about the same loss as that which would be determined for a sudden
expansion. For tapers between 15-deg and 60-deg included angle, calculate the loss as one-half the loss
determined for a sudden expansion.
This subject of losses in fittings has been discussed primarily for systems handling water. In general,
these methods of determining the head loss through fittings apply as well to systems handling other
noncompressible liquids.

DETERMINING FRICTION HEAD

Using frictional values for C = 100 for the pipe and figuring the losses in fittings and valves in the K-
times velocity-head basis, the head for the system shown in Fig. 20.1 for a flow of 454 ms3fhr (2,000
gpm) would be determined as follows:

Entrance loss (12 in bell not well rounded)-K = 0.5 0.08m (0.25 ft)
ND 300 12-in. long radius elbow-K = 0.2 0.03 (0.10)
6.1m (20 ft) of ND 300 (12 in) pipe (at 1.49 ft loss per 0.09 (0.30)
100 ft of pipe)
ND 250 x 300 (10 in x 12-in) reducer 0.03 (0.10)
ND 200 (8 in) gate valve-K = 0.2 0.16 (0.51)
ND 200 (8 in) swing check valve-K = 1.8 1.39 (4.56)
ND 200 x 300 (8 in x 12 in) increaser-K = 0.30 0.23 (0.76)
915 m (3,()()() ft) of ND 300 (12 in) pipe (at 1.49 ft loss per 13.63 (44.70)
100 ft of pipe)
Two ND 300 (12 in) long-radius elbows-K = 0.2 0.06 (0.20)
Three 45-deg elbows-K = 0.2 0.09 (0.30)
Exit loss (ND 300 [12 in] pipe-1 velocity head) 0.15 (0.50)
Total losses 15.94m (52.28 ft say, 52.3 ft)

Of this 15.9 m (52.3 ft) loss, 13.7 m (45 ft) is loss in the pipe and will vary according to the C value
of the pipe. The remaining 2.2 m (7.3 ft) is the allowance for loss in valves and fittings and will vary
only slightly with the age of the pipe.
By computing the values for various other capacities, we would be able to graph the relation of the
system head to the capacity (Fig. 20.8). Without further analysis it appears that for a flow of 454 m3/hr
(2,000 gpm), a good selection would be a pump with the same characteristics as that in Fig. 20.8.
If the pipe were new, there would have been less friction loss. If the condition of the pipe was such
that C = 130, the pipe friction loss would be 61.5 percent of the friction loss when C =100. For example,
at 454 m3/hr (2,000 gpm) the pipe friction loss would have been 45 times 0.615 or 8.44 m (27.7 ft).
The total friction head, including friction losses in the fittings, would be 8.44 m plus 2.23 or 10.67 m
(27.7 ft. plus 7.3 or 35.0 ft). The resulting system-head curve labeled C = 130 is shown in Fig. 20.8.
System-Head Curves 523

7
SYSTEM HEAp C'= 100

SYSTEM I HEAD C= rio


~ 60 r--I-.

"'- ""'- / /
w
1.L
HEAD-CAPACITY f""". SYSTEM HEAD C= 120
z
0 50
V / SYSTEM HEAD c= 130

«w / /' ~.
:c .i. / /
80 -1 40 ...- / -:-..... l\.

--
~ V / ~
~ 70 ~ L
w
u 60 30
/ E~FICiENCY ~ I--
0:
~ 50
/ J.- ~

>-'" 40 20
/ ,... / ...--:- BHP
u
z 30 :c
CI..
.L "....-
~
u m V 1
ii: 20
LL
W
o/
10

0
oVo 5 10 15 20 25 30
CAPACITY, IN 100 GPM

Fig. 20.8 System head for installation shown in Fig. 20.1 for various roughness factors using an 8-in. pump.

Thus, the pump delivers a greater amount of liquid (527m 3/hr [2,320 gpm]) with less head (14.0 m [46-
ft] total head) and less efficiency (82.5 percent).
If it was known that the water would cause a very slow increase in friction with increasing age of
pipe (reaching a value of C = 110 in 15 years or longer) or if power costs were so high that the friction
was to be kept low by periodic cleaning of the pipe, then it would be advisable to select the pump for
a lower head. Using the system head for C = 110 as the maximum to be encountered and at which 454
m3/hr (2,000 gpm) capacity is desired, the pump would have to be selected for 13.7 m (45 ft) total head.
If the same pump was used with a smaller impeller for 454 m3/hr (2,000 gpm), a 13.7 m (45 ft) head
would yield an initial efficiency of 82.5 percent for the system when the piping had a C value of 130,
and would reach an efficiency of about 84.5 percent when the C value had fallen to 110. If it was desired
to obtain greater economy over the entire operating range, a larger pump would have to be used. The
larger pump would have a IO-in. discharge and a ND 250 (10 in) gate valve. AND 250 (10 in) check
valve could be used with smaller friction losses. A graph of the relation of the system head to the
capacity for this system is shown in Fig. 20.9.
In circulating pumps for surface condensers the head is composed entirely, or almost entirely, of
friction losses. These systems are generally complicated because of the need for less water in winter,
when the water temperature is low, than in summer when the water temperature is high. Generally, two
pumps of equal capacity are used. Both are run during the summer to give the required large capacity
and one is run in winter, when less capacity is needed. In this installation, the head would be made up
of losses through the piping and fittings carrying the capacity handled by each pump and losses through
the piping, fittings, and condenser that carry the combined flow. Thus, the system-head curve for the
operation of one pump would not be the same as the system-head curve for the operation of both pumps.
The losses for the individual pumps are shown by curve PI in Fig. 20.10. If both pumps are running,
twice the capacity flow through the condenser yields the same loss in the individual piping for each
-
SYSTEM HEAD C= 110
/~ I ) I I

-
t;J 50 HEAD - CAPACITY
SYSTEM HEAD C = 120
} I I I
W
u. / SYSTEM HEAD C=130
~ r- r-- L l! V
80 ;j 40 ~ -S 7 r-....
W V // :/7" ~ r-... ~
r /"
"'\
-- "-
70 ~

~ 60
<I
I-
f2 30
/ IL /
W
/ EFFICIENCY
//
I-- ~

-
U
50
V- V
Q:

I--I--
W

-
a. CL BHP
)-'"40 % 20
u
Z
III ~ ~

W 30
Q
u.
~ 20 10
/
I
10 L

o V 5 10 15 20 25 30
CAPACITY, IN 100 GPM

Fig. 20.9 System head for installation shown in Fig. 20.1 using a 10 in pump.
Performance is improved.

40 III I
CO~OEN~ER~ ~a.
o:t
...J::l
PUMP u.a.
~() .():-PUMP ON

+
W I - liC,'I-
30 T ~~ 1-1-/
<1-
Q:~ /
,v

I- c, ....
'I V
/ V
W
W
u.. /
V / '/
20
~
0

/ / '/
<I
W
I
/
10
/ / ~ ~,

1- - --
~'I,-
/ ",---
/
,.~ ~ V

10
" I--
20 30 40
FLOW, IN 100 GPM

Fig. 20.10 Typical condenser installation with two pumps in parallel showing operating conditions for
one- and two-pump operation.
KEY:
SI = System head-one pump running (C + PI)
S2 = System head-two pumps running (C + P2)
C = Friction losses in condenser and common piping
PI = Friction losses in individual pump piping-one pump running
P2 = Friction losses in individual pump piping-two pumps running.

524
System-Head Curves 525

pump. This is shown by curve P 2• The loss in the condenser and in the piping and fittings in which the
flow is the same is shown in curve C. At any capacity, the system head with one pump running (SI) is
the head shown in curve PI plus that shown in curve C. With two pumps running (S2), it is the head
shown in curve P 2 plus that shown in curve C. If the loss in the individual piping is low, curves SI and
S2 are so close that only S2 is constructed and the discrepancy between SI and S2 is ignored.
The system illustrated in Fig. 20.lO indicates no static head, and is based on the assumption that the
full siphon head is recovered. Although siphons up to 7.6 m (25 ft) or more are feasible, full recovery
is rarely obtained. Also, in this system, unless the piping and condenser waterways are primed before
the pumps are started, the pumps will have to fill the piping and condenser before the siphon can be
established. Thus, in the starting cycle, a static head equal to the siphon leg will be encountered just
before the siphon is established. The maximum starting head can be determined by adding the siphon
leg as a static component to curves SI and S2. It is often impossible to obtain a pump that will deliver
sufficient capacity to establish the siphon without impairing the results obtained when the siphon has
been established. Modem practice is to provide priming equipment, so that the siphon loop can be
evacuated, and the siphon established without the necessity of a high starting head.
It is desirable that good efficiency be obtained when two pumps are running as well as when one
pump is running. The system head when both pumps are running can be plotted against the capacity
handled by each pump (Figs. 20.11 and 20.12). The problem is to select a pump for this installation
that will give good efficiency at 363 m3/hr (1,600 gpm) and 7.6 m (25 ft) head as well as at 477 m3/hr
(2,lOO gpm) and 3.5 m (18 ft) head resulting from the intersection of the pump head-capacity curve and
the system curve SI. This generally requires the use of a larger pump than would be used if the pump
were selected only for the 363 m3/hr (1,600 gpm) 7.6 m (25 ft) head condition. The selection of a pump
with 82 percent efficiency at both operating conditions is ideal (Fig. 20.11). If a smaller pump had been

- .........

"
HEAD - CAPACI TY V """" ~Q
1'000.. / ~~,

-~r-.~
10"'" 4,,~+r \
~%~
..
.........
/
~
j"-..... ..~l"'·
, '-
r--...... V-
~.....~-.
I EFFICIENCY
/
..........
~
~7- ,oJ
~.

1
V // >. ~
/
/
~
r-
~I 8HP
10

o V 5 10 15 zo
CAPIlCITY. IN 100 GPM

Fig. 20.11 Ideal pump selection for the system shown in Fig. 20.10 if one- and two-pump operation are
equally important.
S26 System-Head Curves

~EFFICIENCY
10 ".- '"'""'-t:

V ""'" ,,~,
...or 70
HI AO-C ~PACITY
/ ~~~ ""'r-
~ to
II:
~ 50
,:
u
z /
/ - ~~
....... /
/ / ....... ~,
.~
r/],'f,; ~t -
~,?,~
"tP-

"
40

--
III I(
~ J
// ./
S 10 IlL.
i V ~
20 10
I~P
10 7
0
I 5 10 15 20 25
CAPlCITY. IN 100 GPM

Fig.20.12 Results of smaller pump for installation in the system in Fig. 20.10.
This pump shows poor efficiency with single-pump operation.

,
-- -
-. 1
HEA D-clpAclTY 138 F T
200
r--
r-... ....... PUMP
90 .... ~+.... --:----..,...
~
r--....
i"'t
~ ~

--- :::=: :::::. ~


80 :::: ';YSTEM HEAD
1---
IL. 150 / r-... C=IOO
I :
~ 70
z
z
.
EFFICIE /v ....... ~SYSTEM HEJ\D
C=130
138 1FT

-
~ 60 ~
/ STATIC HEAD

-
~ ~
0..
50 100
/ ..-
>40
..J
I ., V BHP

/
<I:
u

--V
~
z 0
w 30 ~
~
<.>
l:::
50 J V o LosSES

--
20
".
0..

f~~ ~
I"'"

-
I.&J :J:
C= ,00
10 m

o
/ I---
10 15 20 25
CAPACITY. IN 100 G PM

Fig. 20.13 Characteristics of pump whose total head is mostly static.


Error in calculating friction element or change in friction with age has little effect on selection.
System-Head Curves 527

selected, more actual capacity would have been obtained with one pump operating but only 74.5 percent
pump efficiency would have been obtained (Fig. 20.12). If, in this installation, both pumps were normally
operated all the time and two units had been installed instead of one larger unit (in order to permit
operation at reduced capacity if one unit is out of service), the selection shown in Fig. 20.12 would
have been preferable to that shown in Fig. 20.11 because the efficiency at 363 m3jhr (1,600 gpm) and
a 7.6 m (25 ft) head is greater and the first cost would be lower.
When there is a static head, or its equivalent in pressure, or both, included in the head, the system
head is the sum of these components plus the friction head. Thus, for the system in Fig. 20.13, the
friction head losses have been determined in curve form and added to the static head, giving the system
head indicated. If the pipe had been new and had a coefficient of C = 130, very little increase in capacity
would have resulted. Consequently, the possible error in determining the friction loss becomes less
important as the percentage of the friction head in the total operating head is reduced.

EFFECT OF VARIABLE STATIC HEAD

At a constant speed, the head developed by a centrifugal pump varies with the capacity delivered by
the pump. Thus, if a pump is to be used in a system in which there is a variation in static head, the
capacity delivered through the system will also vary. The purchaser of a pump for such an installation,
will often calculate the friction at rated capacity, add it to the average static head, and state the sum as
the design head. In addition, he will add the same friction head to the maximum and minimum static
heads and give the resulting heads as the maximum and minimum operating heads that the pump will
encounter. Thus, the change in friction with change in capacity is neglected, and the manufacturer is
handicapped in selecting a suitable pump. For example, in the installation shown in Fig. 20.14 some
purchasers would specify that the pump had to operate over a head range of 19.8 to 29.0 m (65 to 95 ft),
giving a rated capacity of 56.8 m3/hr (250 gpm) at 24.4 m (80 ft) head. The pump in Fig. 20.14 would
appear to be unsatisfactory, although it will actually deliver 79 percent rated capacity at the maximum
static head and 115 percent rated capacity at the minimum static head.
It is often good to know what capacity will be delivered by a pump operating on a system in terms
of static head (Fig. 20.14). For any capacity, the static head will be the total head of the pump minus
the friction loss at that capacity. This can be graphed (Fig. 20.14) to show, for example, that the flow
to the tank will be 55.4 m3/hr (244 gpm) when the static head is 17.4 m (57 ft).

DETERMINING PUMP DELIVERY

If a plant or community is located at some distance from its source of water supply, the demand for
water very often increases over a period of time. It ultimately becomes uneconomical to continue to use
the existing pipe line because of the frictional head loss with increased capacity. If the original line is
in good condition, the usual solution is to install a second line, in parallel with the existing line that
will allow economical pumpage of the desired increased rated capacity. If two pipes are operating in
parallel, the friction head loss in each branch must be the same. The proper approach to this problem
is to plot the relation of capacity to friction head loss for each line and then to determine the relation
of combined capacity to friction head loss for the two by adding together the capacity of each line, when
the head losses are the same, at a number of points. For example, Fig. 20.15 shows the head loss for a
ND 250 (10 in) line and the head loss for a paralleling ND 300 (12 in) line in the form of a graph. With
6.1 m (20 ft) friction head loss, the ND 250 (10 in) line will have a flow of 227 m3/hr (1,000 gpm),
whereas the ND 300 (12 in) line will have a flow of 367 m3/hr (1,615 gpm). Thus, 227 plus 367 or 594
528 System-Head Curves

~
100 ~r.Q 't'>~

90
TOTAL HEAD-CAPACITY " ,'" ·s~~,-.O
S 1 0 ~"'i ~'t'>~ ~\C l~
-.§ T4TtC j ___ ~<;,"'ie "'i <;,"'i
---[-tlf,.4D ------::;:::> ~~~

------
80
.((.~Clry
V----~
~I"""---
- i ~ ~~O
70 :'\~~ ~ "'i~"'i\C ~
_I.~_
"
"AAO ~"'iS
~ 60 f-
w
u..
.--
TANK
Iw -;c-- -- ~ V 1'"
z 50 ,-- tl~ ~ LL~ "-,
o<I: 40 - Illilt: LL~O::!;

w Illj~ ~~~~ _
--
"-
:I:
30 -..,..-
P~P
't'>~~
/ ~
'\.

--
~~\C.~~

- -
20

-
V V
10
V
I I
50 100 150 200 250 300
CAPACITY, IN GPM

Fig. 20.14 Characteristics of pump installation with a variable static head.


Friction head variation with capacity must be considered in the selection.

-f /
30

It:
w 25
I
10-IN plPE LINE

I
(12-'N PIPE LIN!
II
~
«
~
LL
0
20
/
I-

/
w
W ~ 10- AND 12-IN

/
lL
PIPE LINES IN
~
/ PARALLEL

/ / V/
vi 15

V
(/)
0
..J
0
«
w
:r

II
10
z
9 /
I-

/
U
a::

~V
LL 5
/

5 10 15 20 25 30 35 40

CAPACITY, IN 100 GPM

Fig. 20.15 Relationship of capacity to friction head for two pipe lines in parallel.
System-Head Curves 529

m3/hr (1,000 plus 1,615 or 2,615 gpm) will be the combined flow of the two lines with 6.1 m (20 ft)
head loss. The capacity for the two lines in parallel for equal head losses can be determined for a number
of points and a curve showing combined capacity against head loss can be drawn (Fig. 20.15).
In some systems, such as in a water works distribution system, it is desirable to maintain a nearly
constant pressure although the demand varies. To maintain an exactly constant pressure, it would be
necessary to vary the speed of the pump or pumps; but, in most systems, it is rarely necessary to maintain
the pressure exactly and some variation can be allowed. Most electric-motor-driven pumps that maintain
nearly constant pressure are, therefore, driven by constant-speed motors. Thus, the pressure will depend
on the head developed by the pump or pumps operating in parallel at the demand capacity. To produce
a reasonably constant pressure for the full range of demand, it is desirable to select pumps having head-
capacity curves that have a shutoff head of 10 to 20 percent more than the head at design capacity.
Installations of divided flow or branch lines, in which the flow is controlled only to prevent overflowing
of a tank or reservoir, usually involve two or possibly three branches. In an installation that has two
tanks at different elevations (Fig. 20.16), it is obvious that flow will not reach tank A unless the friction
head loss in the ND 100 (4 in) line from point C to tank B exceeds 3.05 m (10 ft), which is the difference
in static head for the two inlets. Considering the branch from point C to tank B, the system head has
no static component and will be frictional head only. This may be calculated (curve B, Fig. 20.17). The
system head for the branch from point C to tank A, has a 3.05 m (10 ft) static component and a friction
component caused by the friction in 152 m (500 ft) of ND 100 (4 in). pipe (curve A, Fig. 20.17). The
capacities for the two branches for equal system heads up to point C can be added together to give the
system head of the two branches as a unit (curve A + B, Fig. 20.17). As the friction loss in the piping
from the supply to point C is the same for both branches, there would be no difficulty in establishing
the friction head curve (curve D, Fig. 20.17). Point C is 15.2 m (50 ft) above the suction supply, therefore
the system head for the common system up to point C will be the friction head plus 15.2 m (50 ft). By

,500 FT OF 4-1N. PIPE

1--,-::::::-- ~.- TANK A


10FT

".c
~ .. 1.000 FT OF 4-IN. PIPE

TANK B

50FT

Fig. 20.16 Pumping system involving two tanks at different elevations.


530 System-Head Curves

130
8s ~8+D+50 A =A+D+50
~

- ~ !J'/
120
(1+8)1 =
110 1 s

V
'"
PUMP H AD- 'IA+8l+D+ 50

100
CAPACITY /
90
j / / "-..
r--....

V/ /
"\
f\
I-
ILl
ILl
80
)V V
--
LL. 70
V
~
~~
0
<t
60
V /QJ f--
~
ILl
/ I
,~Y
:I: 50

J
/
/
V
LV
40
,/'
30 V ......
~
~
V /
/ V

--
".
20
..-"V
V
~ ~ V
10
---V
,., ~
V
./'

2 3 4 5

CAPACITY, IN 100 GPM

Fig.20.17 System head-capacity curves for installation in Fig. 20.16.

adding this value to the head values on curve (A + B), we derive the system-head curve for the entire
system as shown in curve (A + B), (Fig. 20.17).
If the pump used for this service had the head-capacity curve shown in Fig. 20.17, the resulting total
flow would be 81.3 m3Jhr (358 gpm). This capacity would require a head at C of 9.45m (31 ft), as shown
by projecting a line from the intersection of curve (A + B), and the pump head-capacity curve to curve
(A + B). By projecting lines to curve A and to curve B from this point, we see that 43.8 m3/hr (193
gpm) would be going to tank A and 37.5 m3/hr (165 gpm) to tank B.
If the two branches to the tanks are each equipped with a valve actuated by a control that closes the
valve when the tanks become full, at times there would be flow to only one of the tanks. The flow to
tank A only can be determined by constructing its system head curve As (by adding head values on
curves A and D plus the 15.2 m [50 ft] static head) and determining the point at which curve A, intersects
the pump head-capacity curve. In the same way, a curve showing the extent of flow to tank B if the
branch to tank A is shut off can be constructed (curve Bs).
If a pump that developed 18.8 m (61.5 ft), or less, total head at 20.4 (90 gpm) capacity had been
used, the entire flow would have gone to tank B.
In many cases, especially those involving two or more pumps each with its own piping, valves, and
fittings discharging into a common discharge line, the solution of a problem can be simplified by
determining the head each pump will produce for its range of capacity up to the point in the system at
System-Head Curves 531

Fig. 20.18 Pumping system involving two pumps that have considerable individual piping discharging into a
common line.

which their pipes join. A system involving two pumps with separate piping discharging into a common
line is shown diagrammatically in Fig. 20.18. The usual head capacity of pump No.1 (Fig. 20.l8) is
graphed in Fig. 20.l9. The friction loss from suction supply to point C is plotted (curve hf) and the
system head (curve hf + S) for pump No.1 from the suction supply to point C is thus determined. By
subtracting the values on curve hf + S from those on the pump head-capacity curve (H-Q), we derive
curve Hc-Q. Curve Hc-Q indicates the head that will be produced by pump No.1 at point C as a function
of the capacity being delivered.

70

-
He· H-(hr +$)
pu~p ~EAo-tAPAcrrv IH-Q)
60 I'-.
..... f"o '
""- .......
50 ~

HC_Q
hf t S
"- 90\~~
C_

~ 40 -~ ~ ~O
.......... ~c:,

'"z
I&.
r--.... ~ '"
,.",. ~
..- ---
~

- --
-lO
t:J
• ~~

'"
'"% 20 ~

0'
O\~~ C

~ '" "'
10
~V
..- ~ i""" \
..- ~~
\
10 15 20 25
CAPACIT Y, IN 100 GPM

Fig.20.19 System head-capacity curves for installation in Fig. 20.18.


532 System-Head Curves

TYPES OF PUMPING SYSTEMS

Pumping systems are of two types-throttled and unthrottled. In a throttled system, the capacity is
detennined primarily by the demand, and the flow is controlled by throttling the excess head developed
by the pump or pumps. In some systems, boiler feed pumps for example, a throttle valve located in the
discharge line controls the flow. In others, such as city water-supply systems without a standpipe or
reservoir "floating" on the distribution mains, the consumers of the water control pump discharge as
they open or close valves. For an unthrottled system in which pumps discharge into a standpipe or a
reservoir, the flow depends on the head developed by the pumps and on the characteristics of the system.

THROTTLED SYSTEMS

In a throttled system, such as a boiler feed pump installation (Fig. 20.20), the flow is controlled by the
throttle valve, which is usually positioned automatically by the feedwater regulator (valve A. Fig. 20.20).
Fig. 20.21 shows the boiler feed system-head curves superimposed on the head-capacity curve of the
pump. Curve C-B represents the boiler pressure plus the static elevation. Although slight changes take
place in the boiler pressure with changes in load, for the sake of simplicity we shall assume the boiler
pressure to be constant.
When water is supplied to the boiler, the pump operates against a pressure that increases with flow
because of the friction head losses in the piping, fittings, and valves in the line. With throttle valve A
wide open, the system-head curve will be curve C-D (Fig. 20.21). The point at which this curve crosses
the pump head-capacity curve (L) is the rated head and capacity of the pump.
If valve A is partially closed, the friction head increases, and the system-head curve may rise to
position C-E. Further closing of valve A would produce other system-head curves such as C-F or C-G.
If valve A is closed entirely, the pump pressure would go to shutoff (point 1). Thus, the system-head
curve can be varied by opening and closing the throttle valve so that a family of curves is produced.
These curves intersect the head-capacity curve at various points between the fully closed position (1)
and the fully open position (D).
To supply the boiler with a quantity of water, Q. the throttle valve is adjusted until the system-head

THROTTLING
DEVICE "A"
~ BOILER ~

~PUMP

Fig. 20.20 Boiler feed pump installation.


A typical throttled system.
System-Head Curves 533

o
«
w
:I:

CAPACITY

Fig. 20.21 Boiler feed installation system-head curves superimposed on the feed pump head-capacity curve.

curve becomes C-F (Fig. 20.21). This curve crosses the head-capacity curve at K and the head against
which the pump operates is represented by the vertical distance H. The actual head required to deliver
quantity Q to the boiler on the normal curve C-D is represented by HI. As the pump develops a head
H at capacity Q, valve A will have to throttle an excess-head equal to H minus HI (distance h, Fig. 20.21).
When a single pump operates on such a system, the shape of the pump head-capacity curve does not
matter. The two pumps whose characteristics are shown in Figs. 20.22 and 20.23 could be used alone
on a throttled system. If the cost of power were high, the pump shown in Fig. 20.22 would be preferred
because the power it requires at part capacities is slightly lower in this particular case. It is not to be
inferred that a pump with a flatter head-capacity curve will always have lower power requirements at
part capacities. Lower power requirements depend on many factors: individual impeller and casing
designs, ratio of design point to point of maximum efficiency, and the like. A pump with a steeper head-
capacity curve has the advantage in a single-pump throttled system because it is less sensitive-the
throttling valve must be moved through a greater distance as more head is throttled off.
Despite having a drooping head-capacity curve, the pump characterized by Fig. 20.22 could be used
on a single-pump throttled system with general assurance of satisfactory operation provided the difference
between the pump head-capacity and unthrottled system head-capacity curves is greatest at zero flow.
With this provision satisfied, the operating point will always be determined by the throttled system head.
In rare cases, surging has resulted in single-pump throttled systems with pumps having drooping head-
S34 System-Head Curves

- -
240
H~"D-C"PACln
220
..........

- ""
200

90 180
... .......
...
80 III 160
'.J..~/
III

...z 70 ~ 140
~~~'7
~

- -
III
u 60 120 ~/
III
IE %
/
50 -' 100
III

200
A.
,.: 40 ~
0 80
I
~ t-
/ ~
~

...- ...-
III
30 60
......Q V
..ii
100 20 40
r1
III

10 20

0 0
V
10 20 30 40
CAPACITY. IN 100 GPM

Fig. 20.22 Slightly drooping head-capacity curve.

capacity curves and, even less frequently, with pumps having stable (constantly rising) head-capacity
curves. Such problems are usually caused by either high control valve sensitivity or an undetected
discontinuity in the pump head-capacity curve.
Installations of two or more pumps operating in parallel involve piping and fitting losses for each
pump as well as for common piping (Fig. 20.24). Instead of the true pump head-capacity characteristics,
a bead-capacity characteristic measured from B to C should be used in the analysis. In throttled systems
in general, and particularly in high-head systems such as boiler feed installations, the losses in the
individual pump piping are such a small percentage of the total head, that their effect is not noticeable
on a curve drawn to reasonable scale.
Two pumps designed for 795 m 3/hr (3,500 gpm), 60.1 m (197 ft) total head are shown in Figs. 20.22
and 20.23. They have differently shaped head-capacity curves. The effect of the difference in losses in
the individual pump piping in each pump will be ignored in this discussion. The individual head-capacity
characteristics for each pump are shown in Fig. 20.25 as A and B. Their combined characteristics are
labeled C. Curve C is obtained by adding together the capacities of the individual pumps at the same
head. For example, at 64.0 m (210 ft) total head, pump A will handle 670 m 3/hr (2,950 gpm) and pump
B will handle 709 m3/hr (3,120 gpm); thus they will produce 1,379 m 3/hr (6,070 gpm) together at 64.0 m
(210 ft) total head.
If these pumps were installed in a water works plant delivering water into a direct distribution system,
in which the desired main pressure corresponds to a total head of 60.1 m (197 ft), and the demand was
1,363 m 3/hr (6,000 gpm), the two pumps in parallel would produce 64.3 m (211 ft) total head and the
System-Head Curves 535

260

240

220
- ~~-4DI"C-4
~~~
200
"""'""
90

80
I-
w
w
LI-

~
180

160
r,.4y V
......... ~ " ~
I-
Z
W
U
70

60
0C(
w
J:
140

120 ~"
#'7~

a::: ~~V
w ..J
/

---
Q.. C(
50 I- 100
~
u
0
I- /
200 z 40

---
80
!!!
~
LI- 30 60
/ V
6~

/ ,-V
LI-
W
100 20 40
Q..
J: 10 20 /
CD

0 o
/
10 20 30 40
CAPACITY, IN 100 GPM

Fig. 20.23 Stable head-capacity curve.

SUCTION
- B -
C DISCHARGE

lpuMP NQ2\

GATE VALVES~
Fig. 20.24 Simplified piping hookup for two centrifugal pumps operating in parallel.
536 System-Head Curves

210

250 """--
-~ •
"'" "-
-
240
......
~

230 C
-.....
"- y
II. ~
Z V' A ..........
220
-.....
~
'" '""
Q

...
c 210
%
200 ~
~ ~,
~
0 110 \
\ \.,
~

110 'i
110

o 10 20 50 40 eo 10 10
CAPACITY. IN 100 GPM

Fig. 20.25 Individual and combined head-capacity characteristics of pumps in Figs. 20.22 and 20.23.

main pressure would be 64.3 m minus 60.1 or 4.2 m (0.41 bar) (211 minus 197 or 14 ft [6.06 psi])
greater than desired. Of the 1,363 m3Jhr (6,000 gpm), 659 m 3Jhr (2,900 gpm) would be delivered by
pump A and 704 m 3Jhr (3,100 gpm) by pump B. If the demand were reduced to 1,136 m3Jhr (5,000
gpm), the main pressure would be somewhat higher. The pumps would be working against 67.4 m
(221 ft) total head with pump A delivering 516 m 3Jhr (2,270 gpm) and pump B delivering 620 m3Jhr
(2,730 gpm). If the demand were reduced further to 682 m3Jhr (3,000 gpm), the division of pumping
capacity would be 114 m 3Jhr (500 gpm) by pump A and 568 m3Jhr (2,500 gpm) by pump B. Finally,
should the demand fall to 597 m 3Jhr (2,630 gpm), pump B would be delivering all the water while pump
A would be delivering none. Thus, at 597 m3Jhr (2,630 gpm) or less demand, pump A would be backed
off the line by pump B and would be operating at shutoff-a dangerous situation even for a short duration.
If pump A had been operating alone on a demand of 597 m3/hr (2,630 gpm) or less and pump B were
started, pump B would pick up the entire load and back pump A off the line. If pump B had been
operating alone at a demand of 597 m3Jhr (2,630 gpm) or less, and pump A were started, pump A would
be unable to deliver any water to the system. If these pumps were operated in a system in which the
change in demand was relatively slow and in which units were cut out when the demand fell to the
rated capacity with one less unit in service, they should never be allowed to operate in parallel below
795 m 3Jhr (3,500 gpm). For demands less than 795 m3/hr (3,500 gpm), either pump A or B would be
operated alone. In such a carefully supervised situation, the two pumps could be operated successfully
in parallel on the throttled type of system. It should be noted that at 1,136 m%r (5,000 gpm) combined
flow, pump A with 516 m3Jhr (2,270 gpm) flow has 76.8 percent efficiency, requiring 123kW (165 hp)
and pump B delivering 620 m 3Jhr (2,730 gpm) flow has 82.3 percent efficiency, requiring 138 kW (185
hp), a total of 261 kW (350 hp) for the two pumps. If both pumps were the same as pump A, a flow of
1,136 m 3Jhr (5,000 gpm) would have meant 568 m3Jhr (2,500 gpm) per pump with 129 kW (173 bhp)
per pump, or 258 kW (346 bhp) total. If both pumps were the same as pump B, the power would have
System-Head Curves 537

been 133 kW (178 bhp) each or 266 kW (356 bhp) total. The use oftwo pumps that have equal capacities
does not necessarily result in power economy.
Difficulties may be encountered on throttled systems with parallel operation of similar pumps if the
pumps have even moderately drooping head-capacity characteristics (Fig. 20.22). Figure 20.26 shows,
with an exaggerated head scale, the theoretical head-capacity curve of one such pump and of two such
pumps in parallel. Let us assume that the friction losses in the individual pump piping are relatively
small and that they may be ignored. If the demand was 397 m3/hr (1,750 gpm) with one pump operating
alone, the pump would operate against 68.0 m (226 ft) total head and exert a discharge pressure
corresponding to that against the check valve of the second pump. This discharge pressure is greater
than the shutoff head developed by the second pump. If the second pump were started, it would come
up to speed against shutoff and would be unable to establish any flow because the pressure on the
discharge side of the check valve would have been greater than the pump could develop at shutoff. In
some installations with two pumps (Fig. 20.26), various methods are used when it is desired to start the
second pump with the first operating on the top of the curve. One, possibly the most common, is to
throttle a little on the gate valve of the pump that is running so that the net head (B to C, Fig. 20.24)
is less than the shutoff head of the second pump. This, and other techniques, generally require very
experienced manipulation and careful timing.
With the pumps in Fig. 20.26, it is possible to obtain unequal capacities at certain flows even if they
are hydraulically duplicates and operating at the same speed. For example, with a demand of
507 m3/hr (2,230 gpm), one pump could be delivering 114 m3/hr (500 gpm) and the other 393 m3/hr
(1,730 gpm). Actually, it is inadvisable to run two pumps, such as shown in Figs. 20.22 and 20.26, in
parallel for capacities at which the developed head exceeds the shutoff head, in this case below
475 m3/hr (2,100 gpm). First, although the two pumps and their drivers are apparently duplicates, there
will be minor differences in the operating characteristics. This will cause unequal distribution of the
capacity and, sometimes, backing one off the line. Second, a motor-driven pump operating under
apparently stable conditions may have minor speed variations as well as minor variations in hydraulic

240

.....
LLJ
LLJ
230

220 -- ...........
""- J.S.:
-r-- -- ~e-4
~~
lL.
:0, ....... ~c
~
210 ~4r-.
0
eX
LLJ
200 ~0;.
).
~~u.." ~s
:I:
-l
j\/ I'-
eX 190 ~
..... ~..o
0
.....
180

u 10 20 30 40 50 60 70
CAPACITY, IN 100 GPM

Fig. 20.26 Individual and combined head-capacity characteristics of two pumps with slightly drooping head-
capacity curves.
538 System-Head Curves

perfonnance that can result in unequal sharing of the load between the two pumps. This could result in
one pump operating at shutoff.
Thus, for pumps to operate satisfactorily in parallel in a throttled system, it is desirable (1) that they
have stable (steadily rising to shutoff) head-capacity curves and (2) that over the operating head range,
the pumps have approximately the same percentage reduction in capacity, or at least deliver some
capacity. As previously mentioned, the increase in head from design capacity to shutoff should not be
too high, otherwise excessive pressure is developed at part-capacity flows. In a system in which it is
desired to maintain a constant minimum pressure at the pumping plant despite varying demand, the
design heads and shutoff heads of all the units are usually the same, or approximately so. Thus, if flows
are less than the units in service will produce at rated head, the capacity delivered by each pump will
be about the same proportion of the rated capacity. For example, if a 568 m3Jhr (2,5OO-gpm) pump
(pump A) and a 1,136 m3Jhr (5,OOO-gpm) pump (pump B) operate in parallel (Fig. 20.27), and the demand
is 1,249 m 3Jhr (5,500 gpm), the head developed by the two pumps would be 64.3 m (211 ft) (6.4 m
[21 ft] above that desired). The capacity delivered by pump A would be 409 m3Jhr (1,800 gpm) or 72
percent of rated capacity, whereas that delivered by pump B would be 840 m3Jhr (3,700 gpm) or 74
percent of rated capacity.
Usually, two or more pumps with stable head-capacity characteristics, and equal or nearly equal
shutoff heads when operating in parallel in a throttled system, will share the load about equally down
to a system capacity much below the capacity at which one or more pumps would be taken out of service.
When purchasing a new pump or pumps that are to be placed in parallel to existing units, the purchaser
should supply the vendor with the head-capacity characteristics of the existing pumps and infonnation
on the operating pressure, to enable the vendor to select a new pump with suitable characteristics.
Some throttled systems utilize a long transmission line between the pumping station and the point at
which a minimum pressure is to be maintained. In such cases, the reduction in pipe friction in the line
when the flow is reduced will cause increased pressure. If this increased pressure is objectionable,

I- 250
w

--r---.::: "
w

'"
"- 225
~
........... ~~CAPACI Y. BOTH PUMPS

0200 '\. DESIRE ;>


ex ~ "'-
....... ~ H",AD
~ 175 HEAD-CAPAf).~Y

"
"'
;'UMP A HEAD-CAP~~ ITY
PUMP B
150

o 10 20 30 40 50 60 70 80 90 100
CAPACITY. IN 100 GPM

Fig. 20.27 Head produced by two pumps operating in parallel.


System-Head Curves 539

possible solutions are (1) maintaining constant pressure by throttling the excess head with some form
of valve, (2) varying the speed of the pumps so that the required head is developed at the capacity
demand, (3) using one or more booster pumps in series with the pumps operating in parallel so that the
head developed by the pumps in service can be increased in steps as the capacity demand increases, or
(4) installing a number of pumps so that small increments of capacity can be obtained. The proper
solution is usually the one that is economically best and therefore depends, in part, on the cost of power
as well as on the cost of personnel necessary for the operation of the system.

UNTHROTTLED SYSTEMS

For a system in which the flow is not throttled, and in which the capacity is such that the head developed
by the pumping system equals the head necessary to deliver the capacity through the system, it is not
necessary for pumps to have similar characteristics to be operated in parallel. When buying additional
pumps to operate in such a system, many purchasers make the mistake of requiring the additional pumps
to have characteristics exactly similar to those of their existing units-this is not necessary. A system
is shown diagrammatically in Fig. 20.28. The system head beyond point C (Fig. 20.28) is indicated in
Fig. 20.33. The desired pumpage rate is from 568 to 1136 m3/hr (2,500 to 5,000 gpm). Four pumps that
are dissimilar (Figs. 20.29 to 20.32), may be operated in parallel on this system, since the maximum
head against which they will operate is 48.8 m (160 ft). The head-capacity curves plotted in Fig. 20.33
have individual piping and fitting losses deducted.
For unthrottled systems, the most economical pump operation is obtained when there is little variation
in the system head as the capacity changes. In many installations, the friction head is so large a part of
the total head at maximum capacity that pumps designed for specific capacities and heads are better in
the long run than pumps in parallel. Pumps in parallel would operate at poor efficiency at heads other
than the desired head. For example, in an installation in which a flow of 1,420 to 2,360 m3/hr (6,250 to
10,400 gpm) is wanted (Fig. 20.34), three separate pumps designed for 1,420 m3/hr (6,250 gpm) at
42.4 m (139 ft) head, 1,890 m3/hr (8,330 gpm) at 50.6 m (166 ft) head, and 2,362 m3/hr (10,400 gpm)
at 61.0 m (200 ft) head might be the proper solution. In other cases in which the head is practically all
friction, the solution might be a full-capacity pump driven by a two- or three-speed motor. The characteris-
tics of a pump driven by a two-speed (1,200 and 900 rpm) motor operating against a system head that
is entirely friction is shown in Fig. 20.35. When operated at 900 rpm, the capacity would be approximately
three-fourths as much and the pump efficiency would remain almost the same. If pumps are driven by
multispeed motors, the capacities that can be obtained at lower speeds depend on the speeds available

STATIC HEAD

PUMPS
I
Fig. 20.28 System involving four pumps with individual piping but with a common transmission line.
540 System-Head Curves

200
- --. - PUMP HEAD-CAPACITY
I- -...:::: t:"'--
~ ~ .....
\oJ
PUMP HEAD-CAPACITY LESS ...........
~ 150 FRICTION LOSSES TO POINT C
~
ci
<[
.......
.....;:
\oJ
J:
ci 100
I-
oI-

50

FRICTION LOSSES TO PO INT C

o I
5 10 15 20 25
CAPACITY, IN 100 GPM

Fig. 20.29 Characteristics of pump no. 1 in Fig. 20.28.

200
::::::::: ~ PU~P Ht AO-ICAP~CITY
~
~
...........

I-
~
PUMP HEAD-CAPAC ITY LESS ~
~
I'-
150 FRIC T ION LOSSES TO POINT C ~ '-
z
cl
<l
~~
\oJ
J: 100
~~
..J
~
~

50

FRICTION LOSSES TO POINT C


I -
2 3 4 5 6 7 8 9 10 II 12 13
CAPAC ITY, IN 100 GPM

Fig. 20.30 Characteristics of pump no. 2 in Fig. 20.28.


System-Head Curves 541

200
I
.....
I-"-'"
R ~ r--... I
PUMP HEAD - CAPAC IT Y
w
w 150
LL ~~
1",\t\
Z
PUMP HEAD-CAPAC ITY L ESS
a FRIC TI ON LOSSES TO POINT C

\\
<l
w
J:
...J 100
<l
.....
\
0
.....

so
FRICTION1 LOSISES TO ~ O I NTI C

5 10
H---r
15 20 25
CAPACITY. IN 100 GPM

Fig. 20.31 Characteristics of pump no. 3 in Fig. 20.28.

200
r::::::: f=::::: t--...... PU~ P HE1AD - CAPACITY

.....
w
PUMP HEAD- CAPACIT Y LESS ~
...........
r':: ~
w 150
LL
~
FR ICTION LOSSES TO POINT C ~ ........
"-
«0
,-
~'~
w
J:
...J 100
<l
.....
0
.....

50

F~ICT I 9N L9SSE~ TO FOIN! C


I
i
2 4 5 6 7
CAPACITY, IN100 GPM

Fig. 20.32 Characteristics of pump no. 4 in Fig. 20.28.


542 System-Head Curves

200~-+--+--+--4--4--~--~~--+--+--+--4--4-~~~~

O~~ __~__~~__~-=~~__~__~~~-L__~~__~__~~
o 10 20 30 40 50 60 70
CAPACITY, IN 100 GPM

Fig. 20.33 Combined characteristics of pumps in system in Fig. 20.28.


Based on head developed at beginning of common line.

250

..... c.,.~ &'"


t-
':I" ~
/~
200 "- /
..,..,
I-
......... //'.." ~
... ~~~.6(;k
V"
.
~
0 ...........
>< ""1.
i!

..,:t ISO
I'-. / 1-~ ".0 ~

..
"" :90
V
'~ ~G'c'~~

-
..J
-'" ~
'fa '>;.
I--V
oS) ~
".0
l-
e
I- "'~~-,,"1. ~
100 -'9,z
~."p;.~
'0

so
o 10 20 30 40 50 60 70 80 90 100 110 120 130

CAPACITY, IN 100 GPM

Fig. 20.34 System characteristics for head that has a large frictional component.
Different size pumps designed for different heads are required.
System-Head Curves 543

~
....LL.~IOO 4;.~

z90
~~-C4P4C'TY r;. ~"'-
0 1.200 RPM ~~~ ~- ~-
.~ K
-...

"
or
.... 80
pc f""'.. ........... ~ /
%
~~
i)K.
..J 70 ~v 00
or
~ (}«-'1 ~¥
0
60 ,<<t'Z+v
~

"'£40_ «-If ~~
,£- V1.200
100 50 ~P4C'T -L,,<c.<c. BHP "
90'ORi5ii
j~-~-I--
~
z j RPM "
80 .... 40
~

/L
U
IZ:
.... ~~ """ ~.....
~......
,II
./
60 CL 30
CL
,: /

,-
%
m u
z /' BHP
40
.... 20
I: I--
P' I!'- 9)0 RPM
20
Q
10 7' .".

L
LL.
V
0
LL.
&oJ
~
V
2 3 4 6
CAPACITY, IN 100 GPM

Fig. 20.35 System characteristics of an installation that is all friction.


Efficient operation of a pump can be obtained by part-speed operation.

so that it is not always possible to obtain the exact capacities desired. To obtain exact capacities, a
variable speed driver would have to be used.
Frequently, one or more booster pumps are installed, either in the suction line to the main pumps or
in the common discharge line to increase the capacity of existing stations. If conditions in the system
illustrated in Figs. 20.28-20.33 should change so that the maximum demand at times is 1,340 m3/hr
(5,900 gpm), one solution would be to have all four pumps discharge into a 1,340 m3/hr (5,900 gpm)
booster pump that has a total head of 9.2 m (30 ft). Booster pumps are particularly practical when an
increase in head would cause considerable reduction in capacity of the main pumps.
Although the foregoing examples have described systems handling water, the basic principles apply
to systems handling other liquids as well. There are sometimes certain limitations when liquids other
than water are used. In systems handling volatile liquids minimum pressures must be maintained at
every point. These requirements must be checked when analyzing the system.
Pumps with high-specific-speed impeller designs have steeper head-capacity curves than pumps with
low-specific-speed designs. Thus, in systems involving low heads for which a high-specific-speed type
of pump will be used, a greater variation in percentage of total head can be met efficiently than in
systems involving high heads for which a low-specific-speed pump must be used. Occasionally, some
low-head systems will require the use of a low-specific-speed type of pump to accomplish the desired
operation. There are a few systems in which there is a wide variation in head range with no need for a
fixed capacity to be delivered at any specific head. Such a system is met in dewatering a flooded mine.
This is an unusual application of a centrifugal pump because the total head against which the pump
must work varies from approximately zero to a high maximum that occurs when the mine is almost
clear of water. A mine-dewatering pump should be designed not for a single point of head and capacity,
544 System-Head Curves

lZPUM'ING EL ~Mt:NTS 2 PUMPING t: _EWENrs


IN 'UALLEL

CAPACITY. IN 100 GPM

Fig. 20.36 Characteristics of series-parallel type of mine dewatering pump.

but for the greatest possible capacity at all heads within the capacity of the motor. For installations
involving final heads for which multistage pumps are required, the best possible design is a parallel-
series unit (Fig. 20.36). With this unit the dewatering takes place almost twice as fast at the beginning
than it would if the various stages were arranged to pump only in series, but the power expenditure is
the same.

BIBLIOGRAPHY

[20.1] Lewis F. Moody, Friction Factors for Pipe Flow, Transactions of ASME, November, 1944.
[20.2] SIHI Group, Basic Principles for the Design of Centrifugal Pump Installations, 1980, SIHI-HALBERG,
Ludwigshafen, Germany.
[20.3] Engineering Data Book, 2nd Edition, 1990, Hydraulic Institute, Parsippany, NJ, USA.
II
PUMP PERFORMANCE
21
Centrifugal Pumps and
Energy Conservation
~~--~-~ ~--~----------- ----

The high cost of energy and the scarcity of fuels have become a hard fact of life, making it imperative
to examine all energy-consuming processes with a view to improving their overall efficiency. And since
every industrial process that underlies our modern civilization involves the transfer of liquids from one
level of pressure or static energy to another, pumps have become an essential part of all industrial
processes, and in turn, major consumers of energy themselves. It thus becomes even more important to
avoid waste whenever possible and to examine both the selection and operation of our pumps to see
whether we can effect significant savings in energy consumption.

Pump Efficiency

All things being equal, it is natural that the user of centrifugal pumps will look with greater favor on
pumps with higher efficiencies and will favor pumps offerings that might exceed others by as little as
0.5 or 1 point of efficiency. But all things are not always equal, and these small differences in guaranteed
efficiencies may have been obtained at the expense of reliability, either by using smaller running clearances
or a higher head coefficient impeller, or some other feature or configuration that does improve efficiency,
but may at the same time reflect unfavorably on pump reliability.
If we are dealing with a multistage pump, we can increase the efficiency by selecting a higher-
specific-speed design. This requires a lower head per stage, hence more stages, and a longer shaft span.
In turn, this leads to a larger shaft deflection and-unless we choose to be counterproductive and use
larger clearances-to a reduced reliability.

Effect of Specific Speed


This does not mean to imply that improvements in efficiency based on the proper selection of pump
design should be neglected. Thus, we must still consider all the factors that can affect the power
consumption of a centrifugal pump. One of the parameters that is affected by the specific speed is the
maximum efficiency obtainable from pump impellers of different specific speeds and different sizes, as

545

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
546 Centrifugal Pumps and Energy Conservation

Table 21.1 Selections for 454 m 3Jhr (2,000 gpm) and 122 m (400 ft)

Solution Number of stages Speed (rpm) Specific speed Chart efficiency (%) kW (bhp)

1 1 3,550 1,775 0.830 181 (243)


2 2 3,550 2,985 0.855 176 (236)

illustrated in Fig. 18.30. The specific speeds to be used in connection with these curves must correspond
to the maximum impeller diameter. Some reduction in efficiency will generally accompany the cutting
down of the impeller or impellers.
Theoretically, one should always try to use specific speeds in the region of 2,500 to obtain a pump
with the highest possible efficiency for its capacity and head. This is frequently impractical, especially
for small standard-size pumps. Consider, for instance, the case of a pump to be designed for 22.7 m 3/hr
(100 gpm) and a 61.0 m (200 ft) head. At 3,560 rpm, the specific speed will be 669 and the efficiency
can be expected to be 53 percent (see Fig. 18.30). If we were to design this pump for a specific speed
of 2,500, we could probably reach an efficiency of 70 percent, but the pump would have to run at 13,300
rpm. This would require the use of a gear (dropping the overall efficiency to about 65 percent) and a
rather special and expensive type of pump.
Let us consider some typical examples (see Table 21.1) where two alternative selections are examined.
It has been assumed that the design point corresponds to the best efficiency point of the pump selected.
The difference between solutions 1 and 2 appears to be significant enough to warrant consideration of
a two-stage pump, which saves 5 kW (7 hp), equivalent to an annual saving of $2,190 to $2,930 when
evaluated to $0.050 to 0.067 per kW/hr.
Against these savings, one must weigh a number of counterbalancing factors. Among these are
(1) the higher initial cost of the two-stage pump and (2) the fact that a two-stage design precludes the
use of a simpler overhung, end-suction pump with a single shaft seal.
It may still be that the higher-specific-speed pumps will show sufficient savings to justify their
selection. But are these savings really there? The answer to this question depends entirely on the expected
operating capacity range of these pumps. The shape of the power consumption curve varies considerably
with the specific speed of the pump in question, as illustrated in Fig. 18.36. Before we can decide which
of the possible solutions is best from the point of view of energy consumption, we must examine the
power consumption not only at the design point but also over the entire range of capacities the pump
will encounter.
Let us, for instance, expand our analysis of the case examined in Table 21.1. In addition to the bhp
at the 100 percent design capacity, we shall compare the power consumption of the single-and two-

Table 21.2 Comparison of Power Consumption at Part Loads for Selections for 454 m3Jhr (2000 gpm) and
122 m (400 ft)

Percentage of Power as percentage Gain or loss


Solution Specific speed design flow of design (hp) kW (bhp) compared to solution 1

1 1,775 100% 100% 181 (243) Base


1 1,775 75 89 161 (216) Base
1 1,775 50 76 138 (185) Base
2 2,985 100 100 176 (236) 5kW gain
2 2,985 75 95 167 (224) (6kW loss)
2 2,985 50 86 152 (203) (13kW loss)
Centrifugal Pumps and Energy Conservation 547

Table 21.3 Comparison of One- and Two-Stage Pump Selections in Tenns of Yearly Operation at Various
Loads (Conditions: 454 m 3Jhr [2000 gpm], 122 m [400 ft])

Operating time Advantage in kWhr


Capacity in percentage One-stage Two-stage
of design flow In % In hrs pump pump

100% 30 2,628 13,140


75 60 5,256 31,536
50 10 876 11,388
Total 100 8,760 42,924 13,140
Net savings in favor of one-stage pump = 29,784 kWhr.

stage pumps at 75 percent and 50 percent flow (see Table 21.2). Instead of saving energy at all flows,
the two-stage pump uses 6kW (8 hp) more at 75 percent flow and 13 kW (18 hp) more at 50 percent.
To establish the real energy balance between the two solutions, we need to predict the subdivision of
operating hours at various loads. If we assume that this subdivision will correspond to that shown in
Table 21.3, we find that the most efficient pump is not the best selection and that the single-stage pump
will save 29,784 kWhr yearly over the two-stage pump. Obviously, the final answer will always depend
on the load factor that will prevail in a given installation.

WHERE CAN WE SEARCH FOR REAL SAVINGS?

A single-stage, double-suction, 750 kW (1,000 hp) pump may be worth something on the order of
$20,000. The operating cost of such a pump may be as much as $300,000 per year and a 1 percent
reduction in power consumption is worth $3,000/year. Notice that we did not sayan increase in efficiency
of 1 percent, but rather a reduction in power consumption. The reason that we did not refer to efficiency
is that essentially the centrifugal pump is a mature product and that there is very little more to be
squeezed out of design improvements. On the other hand, it can be demonstrated that substantial savings
in power consumption are generally available by exercising reasonable restraint in selecting the margins
added to both capacity and total head when specifying pump service conditions. By substantial, is meant
savings of as much as 10, 15, and sometimes even 20 percent.
For a 750 kW (1,000 hp) pump this means up to 150 kW (200 hp), savings of up to $60,000 per
year or present-day worth of between $200,000 and $400,000. This is 10 to 20 times as much as the
initial cost of the pump itself. Certainly, there are greater savings of energy available in the area of
reasonable selections of conditions of service than in an elusive chase for a 1 or 2 point improvement
in pump efficiency.
Savings can also be achieved in two-pump installations, by running a single pump whenever it can
meet the required demand.
Finally, restoring internal clearances is one more factor in effecting energy savings.

EFFECTS OF OVERSIZING

One of the greatest sources of power waste is the practice of oversizing a pump by selecting design
conditions with excessive margins in both capacity and total head. It is strange on occasion to encounter
548 Centrifugal Pumps and Energy Conservation

a great deal of attention being paid to a I-point difference in efficiency between two pumps, while at
the same time potential power savings are ignored through an overconservative attitude in selecting the
required conditions of service.
Consider that we are not primarily interested in efficiency, we are interested in power consumption.
Pumps are designed to convert mechanical energy from a driver into energy within a liquid. This energy
within the liquid is needed to overcome friction losses, static pressure differences, and elevation differences
at the desired flow rate. Efficiency is nothing but the ratio between the hydraulic energy utilized by the
process and the energy input to the pump driver. And, without changing the ratio itself, if we find that
we are assigning more energy to the process than is really necessary, we can reduce this to correspond
to the true requirement and therefore reduce the power consumption of the pump.
It is true that some margin should always be included in the rated capacity to account for wear at
the internal clearances, which will eventually reduce the effective pump capacity, and to accommodate
system upsets. How much margin to provide is a fairly complex question, because the rate of wear
varies with the type of pump, the liquid handled, and the severity of the service, while the margin needed
for system upsets depends on predictions of the resultant flow swings. We shall examine the effect of
wear later when we analyze the savings in power consumption that can be realized from restoring internal
clearances to their original values.
A centrifugal pump operating in a given system will deliver a capacity corresponding to the intersection
of its head-capacity curve with the system-head curve, as long as the available NPSH exceeds the
required NPSH (Fig. 19.1). To change this operating point in an existing installation requires changing
the head-capacity curve, the system-head curve, or both. The first can be accomplished by varying the
speed of the pump (Fig. 18.26) or its impeller diameter, whereas the second requires altering the friction
losses by throttling a valve in the pump discharge (Fig. 20.21). In the majority of pump installations,
the driver is a constant speed motor and changing the system-head curve is used to change the pump
capacity. Thus, if we have provided too much excess margin in the selection of the pump head-capacity
curve, the pump will have to operate with considerable throttling to limit its delivery to the desired value.
If, on the other hand, we permit the pump to operate unthrottled, which is more likely, the flow
into the system will increase until that capacity is reached where the system-head and head-capacity
curves intersect.
Let us examine a concrete example, for which the maximum required capacity is 613 m 3Jhr (2,700
gpm), the static head is 35.1 m (115 ft), and the total friction losses, assuming 15-year-old pipe, are
18.3 m (60 ft). The total head required at 613 m 3Jhr (2,700 gpm) is therefore 53.4 m (175 ft). We can
now construct a system-head curve and this is shown on curve A in Fig. 21.1. If we add a margin of
about 10 percent to the required capacity and then, as frequently is done, we add some margin to the
total head above the system-head curve at this rated flow, we end up by selecting a pump for 681 m3Jhr
(3,000 gpm) and 61.0 m (200 ft) total head. The performance of such a pump, with a 375 mm (14.75 in)
impeller, is superimposed on the system-head curve A in Fig. 21.1.
The pump develops excess head at the maximum required capacity of 613 m 3Jhr (2,700 gpm) and if
we wish to operate at that capacity, this excess head will have to be throttled. Curve B is the system-
head curve that will have to be created by throttling. If we operate at 681 m 3Jhr (3,000 gpm), the pump
will take 131 kW (175 bhp) and we will have to drive it with a 150 kW (200 hp) motor. If we operate
it throttled at the required capacity of 613 m 3Jhr (2,700 gpm), operating at the intersection of its head-
capacity curve and curve B, the pump will require 124 kW (166 bhp).
The pump has been selected with too much margin. We can safely select a pump with a smaller
impeller diameter, say 359 mm (14.12 in), with a head-capacity curve as shown in Fig. 21.1. It will
intersect curve A at 640 m3Jhr (2,820 gpm), giving us about 4 percent margin in capacity, which is
sufficient. To limit the flow to 613 m3Jhr (2,700 gpm) we will still have to throttle the pump slightly
and our system-head curve will become curve C. The power consumption at 613 m3Jhr (2,700 gpm) will
Centrifugal Pumps and Energy Conservation 549

240
H P 80 R. 8

0 2.20
4. 51 II. II PI /
c(
i J.l h i" ~ J P A
W
r.-- l/ 1/
...r
M

""
J:
200
,,., ~ ~
ELI ER r- r-.... ~ _I
..J

..... 1
/ n
lVlt
~ ~~
12 / t)c; .~
~ .~
l/.:/
I- 180
w .1/ ~
~

..' "..
w ~ ~

'" ,
LL
160 ~
r/' .~
r" ~
I' I;' I..... ]\
L ~

.~ .;~ " 10 ·

-- --
140 . ~
r...... .10' I'
~ ~.
.-
.. ,- .. - -
90
I ".
WI ~
".,..

. ~
:::TlI TIl"'
' ;;;;;; 1"'. ~
;,#' .-~ ......
200 80
~~
~ ",.

-
180
['y«J ~ ~<f: i"o' " 70

I§....
~ 1&.«)
", """'" ~ ~
160 ~
60 >-
~I/, ~~ .0 ~ ..... ~ U

.,
Z
,.: rtl lA ~~ ry
w
140
,"
I-' 50 §
ra
Q..
..:-
J: ~J 'It .1::>~ ./ LL
LL

""
OJ
II. ...Q ~
W
~.

120 ~ ~-.> 40 <ft


JrL ~ ...,L. """.:
\'1"

IIJ
..,
I"""
100 l......-
,..... ~ ~ 30

..,""
rJ,
80 ..... ~ ~
20
~'Ii
60 /' 10
IJJ.
U o
o 1,000 2,000 3,000 4,000
CAPACITY IN GPM

Fig. 21.1 Various system-head curves superimposed on pump head-capacity curves.


KEY:
A = System head with margin on capacity and head
B = System head with throttling to operate at required capacity, pump with 14.75
in impeller
C = System head with throttling to operate at required capacity, pump with 14.00
in impeller
D = System head with "new" pipe.
550 Centrifugal Pumps and Energy Conservation

Table 21.4 Comparison of Pumps with 375 mm (14.75 in) and 359 mm (14.72 in) Impellers, with the System
Throttled to 613 m3/hr (2700 gpm)

Impeller-mm (in) 375 (14.75 in) 359 (14.12)


System head curve B C
Power-kW (bhp) 124 (166) 112 (150)
Savings-kW (bhp) 12 (16) or 9.7%
-% 9.7

now be only 112 kW (150 bhp) instead of the 124 kW (166 bhp) required with our first overly conservative
selection. This is a very respectable 10 percent saving in power consumption. Furthermore, we no longer
need use a 150 kW (200 hp) motor, and a 132 kW (175 hp) motor will do quite well. The saving in
capital expenditure is another bonus from not oversizing. Our savings may actually be even greater than
we have shown. In many cases, the pump can be operated unthrottled, the capacity being permitted to
run out to the intersection of the head-capacity curve and curve A. If this were the case, a pump with
a 375 mm (14.75 in) impeller would operate at approximately 715 m3jhr (3,150 gpm) and take 132 kW
(177 bhp). Ifa 359 mm (14.12 in) impeller were to be used, the pump would operate at 640 m 3/hr (2,820
gpm) and take 114 kW (153 bhp). We could be saving over 13 percent in power consumption. Tables
21.4 and 21.5 show these savings.
Our real margin of safety is actually even greater than just indicated. The friction losses we used to
construct the system-head curve A were based on losses through 15-year-old piping. The losses through
new piping are only 0.613 times the losses we have assumed. The system-head curve for new piping is
that indicated as curve D in Fig. 21.1. If the pump we had originally selected (with a 375 mm
[14.75 in] impeller) were to operate unthrottled.. it would run at 818 m3/hr (3,600 gpm) and take 140
kW (187 bhp). A pump with only a 359 mm (14.12 in) impeller would intersect the system-head curve
D at 734 m 3/hr (3,230 gpm) and take 121 kW (162 bhp), with a savings of about 14 percent. As a matter
of fact, we could even use a 352 mm (13.88 in) impeller. The head-capacity curve would intersect curve
D at 704 m 3/hr (3,100 gpm) and the pump would take 113 kW (151 bhp). Now, the saving over using
a 375 mm (14.75 in) impeller becomes 19 percent (see Table 21.6).
There is obviously no question that important energy savings can be made if, at the time of the
selection of the conditions of service, reasonable restraints are exercised to avoid incorporating excessive
safety margins into the rated conditions of service. But what of existing installations in which the pump
or pumps have excessive margins? Is it too late to achieve these savings? Far from it. As a matter of
fact, it is possible to establish more accurately the true system-head curve by running a performance
test once the pump has been installed and operated. A reasonable margin can then be selected and three
choices become available to the user:

Table 21.S Comparison of Pumps with the System Unthrottled on Curve A

Impeller-mm (in) 375 (14.75) 359 (14.12)


Flow-m3/hr (gpm) 715 (3,150) 640 (2,820)
Power-kW (bhp) 132 (177) 114 (153)
Savings-kW (bhp) 18 (24)
-% 13.6
Centrifugal Pumps and Energy Conservation 551

Table 21.6 Effect of Different Size Impellers in System with New Pipe and Resulting Savings New Pipe
(Unthrottled Operation, Curve D)

Impeller-mm (in) 375 (14.75) 359 (14.12) 352 (13.88)


Flow-m3/hr (gpm) 818 (3,600) 734 (3,230) 704 (3,100)
Power-k W (bhp) 140 (187) 121 (162) 113 (151)
Savings-kW (bhp) 19 (25) 27 (36)
-% 13.6 19.3

1. The existing impeller can be cut down to meet the more realistic conditions of service required for the instal-
lation.
2. A replacement impeller with the necessary reduced diameter can be ordered from the pump manufacturer.
The original impeller is then stored for future use if friction losses are ultimately increased with time or if
greater capacities are ever required in the future.
3. In certain cases, there may be two separate impeller designs available for the same pump, one of which is
of narrower width than the one originally furnished. A replacement narrow impeller can then be ordered
from the pump manufacturer. Such a narrower impeller will have its best efficiency at a lower capacity than
the normal width impeller. It mayor may not need to be of a smaller diameter than the original impeller,
depending on the degree to which excessive margin had originally been provided. As in case 2, the original
impeller is put away for possible future use.

VARIABLE SPEED OPERATION

Although by far the greatest majority of motor-driven centrifugal pumps are operated at constant speed,
some installations do take advantage of the possible savings in power consumption provided by variable
speed operation to meet the exact required conditions of service without throttling. Wound-rotor motors
were once frequently used for this purpose, but the practice today is generally to interpose a variable-
speed device such as a magnetic drive or a hydraulic coupling between the pump and the electric motor.
Alternatively, one can use a variable-voltage, variable-frequency motor control, which has the great
advantage of essentially maintaining a constant motor efficiency regardless of its operating speed, within
the range of speeds required for normal centrifugal pump operation.
For instance, if the system-head curve when the installation is new corresponds to curve D (Fig.
21.1), if variable frequency drive is used, and if we wish the pump to handle exactly 613 m 3Jhr (2,700
gpm) and 46.6 m (153 ft) head, we could use the pump with a 375 mm (14.75 in) impeller and run it
at 87.5 percent of design speed; we could also use a 359 mm (14.12 in) impeller at 92.2 percent of
design speed. In the former case, the pump would take 90 kW (120 bhp) and in the latter, because the
smaller impeller has lower efficiency, 93 kW (124 bhp). Either compares favorably with the 124 kW
(166 bhp) taken by a constant speed driven pump with a 375 mm (14.75 in) impeller at the same 613
m 3Jhr (2,700 gpm).
The savings with a variable-speed device such as a hydraulic coupling are somewhat reduced. The
power input to the variable-speed device can be calculated from the following relation:

motor bhp = pump bhp x motor spee: + fixed horsepower losses of variable speed device
pump spee

Magnetic drives as well as hydraulic couplings generally have a slip of about 3 percent at full load and
additional fixed losses of about 1 percent. Thus, the efficiency of such a variable-speed device at rated
552 Centrifugal Pumps and Energy Conservation

240
Ht-q loo
" . 'Ill

0 220

.
« ~~ A
w ....... .......
J: ~
...J 200
~ '"':C ~
~
~
w
w
180
./
.' ~ • '" 0
V r\.
u.
160 V ."'1"'" I\.
.•V' ~
I..... I'

--
10·

-'
140 ~

..
~

I...... .~

- .... 90
~ ~ IA IC ......
- .---
~ E~ 0 1.".00
.;:: ":. _. I-
".
-- i ""- .- ~

-
.~ ~
200 'N 80
~
180 V N~ ~
," ~
70

""""
,~

I"'''
V b'r- ~ 1..00'"
160 I ,,~
~
~ IJ: V 'r-~~ ~ V'
60 >-
0 ~ V ()
z
Q. 140 I.. ~ I lf
''' I C/)
II V If
50
w
Q
J:
CIl J ..tV' .i.P./ u.
u.
w
120 I ~~
I"'" 9,."" ,
~'
40 *-
J V .Q,~ '/.,
100 I i/
'fI"
.~ 'l- W 30

80 )~
~
"" q~ ~O
. ~~) v~

20
If IL
6o J 10
I
't
o
o 1,000 2,000 3,000 4,000
CAPACITY IN GPM

Fig. 21.2 Effect on power absorbed of running pump in Fig. 21.1 at variable speed.
Note that power absorbed by the pump does not include power losses in the variable speed drive.

speed is on the order of 96 percent and decreases further as the speed is reduced. In the sample just
given above. the power consumption at 613 m3Jhr (2.700 gpm) with a 375 mm (14.75 in) impeller would
be approximately 104 kW (139 bhp) at the motor. We would still be saving 20 kW (27 bhp) by using
a variable-speed device.
But even here it is not necessarily too late to achieve these power savings in an existing constant-
speed installation. In many cases. it is possible to modify the installation by acquiring a magnetic drive
or a hydraulic coupling and installing it between the pump and its constant-speed motor or by retrofitting
Centrifugal Pumps and Energy Conservation 553

Table 21.7 Power Savings at Variable Speed

Total Head Power at Total head- Speed for Power for


Capacity at 100% Speed 100% Speed Curve A Curve A Curve A
m3Jhr (gpm) m (ft) kw (bhp) m (ft) % Rated kW (bhp)

636 (2,800) 63.4 (208) 127 (170) 55.5 (182) 94.7 110 (148)
545 (2,400) 65.2 (214) 117 (157) 49.7 (163) 88.5 86 (115)
454 (2,000) 67.4 (221) 106 (142) 45.1 (148) 83.0 67 (90)
363 (1,600) 68.7 (225) 96 (128) 41.5 (136) 78.6 52 (69)

the existing electric motor with variable-frequency controls. To decide whether it is advisable to make
such a modification, it is necessary to plot the actual system-head curve, to calculate the speed required
at various capacities over the operating range, and to determine the motor horsepower input to the
variable-speed device over this range. The difference between this horsepower and the pump power at
constant speed represents potential power savings at these capacities. It then remains to assign a predicted
number of hours of operation at various capacities and then calculate the potential yearly savings in
kilowatt-hours. These savings are translated into cost savings and can be used to determine whether the
cost of the modification to variable-speed is justified. If it is, one can proceed with the modification. If
not, the use of smaller impellers or of narrower impellers is still available to us.
The following shows an example of the calculations involved. Figure 2l.2 gives the full speed
performance of a pump designed for 681 m 3/hr (3,000 gpm) and a total head of 6l.0 m (200 ft).
Superimposed on this curve is the system-head curve calculated for 15-year-old pipe (curve A). Curve
D shows the system-head curve with new piping. Savings are calculated only for curve A, even though
higher savings would be available in the early years of installation. The maximum capacity that is
expected to be required is 636 m3/hr (2,800 gpm). Savings at various capacities are shown in Table 2l.7
and are based on an assumed time of operation at these capacities, total yearly power savings are shown
in Table 2l.8.
There are several ways of evaluating these savings. For instance, at $400 or $540 per kW-year, the
savings will amount to between $11,240 and $15,170. At a figure of $4,000 in present day worth per
kW, it would pay to spend up to $112,500 for the extra cost of variable frequency drive to achieve these
savings. Expressed differently, since this application requires alSO kW (200 hp) motor, it would pay
to go to variable frequency as long as the extra cost did not exceed about $750 per kW, which is
obviously several times the expected extra cost involved.
In general, one can predict that the average savings will run to between 15 and 20 percent of the

Table 21.8 Horsepower Savings Based on Assumed Time of Operation

Savings in Power from Equivalent Power


Capacity Operating at Variable Speed Percentage of Savings
m3Jhr (gpm) kW (bhp) Operating Time kW (hp)

636 (2,800) 17 (22) 40 6.6 (8.8)


545 (2,400) 31 (42) 30 904 (12.6)
454 (2,000) 39 (52) 20 7.8 (lOA)
363 (1,600) 44 (59) 10 404 (5.9)
--~-,---

Total Savings 28.1 (37.7)


554 Centrifugal Pumps and Energy Conservation

total HP of an installation. This means that the break-even point for the EXTRA involved in providing
a variable frequency drive for a typical installation is on the order of 15 to 20 percent of the present
day worth per kW. With this present day worth running from $2,700 to as high as $5,400 per kW, it
would appear that most installations can be shown to save enough through variable-speed operation to
warrant the use of variable-frequency drive.

RUN ONE PUMP INSTEAD OF TWO WHENEVER YOU CAN

Many installations are provided with so-called "half-capacity" pumps, with two pumps operating in
parallel to deliver the required flow under full-load conditions. If the service on which these pumps are
installed is such that the required flow varies over a considerable range, important power savings may
be available through improved operating practices. This refers to the fact too often both pumps are kept
on the line even when the demand drops to a point where a single pump can carry the load (see Fig.
22.3). Operators who keep both pumps running at all times claim that they do so for safety reasons
because, if one pump were to fail, they could still supply a portion of the required flow. Until the pump
that has failed can be restored to service, the process served by the pumps can be made to operate at
reduced load or, if a standby pump is available, it can be brought into service without complete
flow interruption.
We examine this question of safety later. Let us first examine how important may be the savings
resultant from operating a single pump whenever it can meet the required demand. This can be best
demonstrated by reference to a so-called 100 percent gpm curve, as shown on Fig. 18.25. To simplify
our analysis, let us neglect the question of capacity or pressure margins and imagine that the pumps
carry their full load with throttling valves wide open. We shall also assume that these pumps are operating
at constant speed. Thus, full load conditions are met by running two half-capacity pumps, each running
at its 100 percent capacity point, and each taking 100 percent of its rated brake-horsepower.
If we want to reduce the flow to half-load and still maintain both pumps on the line, it will be
necessary to throttle the pump discharge and create a new system-head curve. Under these conditions,
each pump will deliver 50 percent of its rated capacity at 117 percent rated head, much of which will
have to be throttled. Each pump will take 72.5 percent of its rated power consumption.
We could, instead, stop one of the pumps and carry the half-load requirement with a single pump
running at 100 percent of its rated capacity. The discharge would be throttled considerably less than if
both pumps were left running. The power consumption would be 100 percent of the rated value for a
single pump. Thus, the total power consumption of two pumps operating under half-load conditions
would be 145 percent of that required if a single pump were to be kept on the line. As a matter of fact,
a single pump could carry much more than the capacity corresponding to half-load, since the head-
capacity of a single pump intersects with the system-head curve anywhere from 60 to 70 percent load,
depending on the exact shape of the system-head and the head-capacity curves.
There are other benefits to be had from such an operating practice. First, if we assume 8,500 yearly
operating hours for the process served by these pumps, of which 20 percent takes place at flows of 50
percent of maximum flow or lower, each pump will operate for only 7,650 hours a year instead of 8,500,
extending the calendar life of all the running parts by over 11 percent. But more important than this
straight arithmetical effect is the fact that pumps which frequently operate at reduced capacities do not
have as long a life as pumps operated closer to their best efficiency point. Thus, running only one pump
whenever it can handle the required flow will add much more life to each pump than just the arithmetical
difference in operating hours.
Returning to the safety argument, the usual concern is the risk of tripping a unit should a single
operating pump fail suddenly. This concern is really only valid for steam turbine driven pumps, which
Centrifugal Pumps and Energy Conservation 555

take some time to accelerate from turning-gear speed compared to the few seconds needed to establish
flow with a motor driven pump when it is started across the line. A similar capability can be realized
with a steam turbine driven pump by idling the standby pump at say 20-25 speed (enough to establish
bearing lubrication and reduce the risk of seizure), which then allows it to be rapidly accelerated if
needed. What should be realized with steam turbine driven pumps is that they're usually high energy
density machines and so one of the consequences of prolonged operation at low flows, often as low as
30 percent of best efficiency capacity, is premature wear at the running clearances and premature erosion
of the hydraulic components. Running two pumps when one will carry the load therefore means both
are being worn rapidly, which is not conducive to safety. Compounding that, if both are running
and there is an operating mishap, there is a reasonable probability that both pumps will be damaged
simultaneously, which would then trip the unit and cause lost production while both pumps were repaired.
The evaluation of any energy conservation program consists of a comparison between the amount of
energy saved and the cost of the implementation of the program. The remarkable fact is that the cost
of implementation in this case is exactly zero, or no more than the cost of the time required to explain
to the operators how this method of running a parallel pump installation saves power.

RESTORING INTERNAL CLEARANCES

As mentioned earlier, the rate of wear of internal clearances depends on many factors. First, it should
be obvious that it increases in some relation to the differential pressure across the clearances. It also
increases if the liquid pumped is corrosive or abrasive. On the other hand, the rate is slowed down if
hard, wear-resisting materials are used for the parts subject to wear. Finally, wear can be accelerated
very rapidly if momentary contact between rotating and stationary parts occurs during the operation of
the pump.
As running clearances increase with wear, a greater portion of the "gross" capacity of the pump is
short-circuited through the clearances from the higher-pressure discharge side to the lower-pressure
suction side and must be repumped again. The effective, or "net," capacity delivered by the pump against
a given head is reduced by an amount equal to the increase in internal leakage. Although in theory the
leakage varies approximately with the square root of the differential pressure across a running joint and
therefore with the square root of the total head, it is sufficiently accurate to assume that the increase in
leakage remains constant at all heads. Figure 21.3 shows the effect of increased leakage on the shape
of the head-capacity curve of the pump. Subtracting the additional internal leakage from the initial
capacity at each head gives a new head-capacity curve after wear has taken place.
As noted before, we must compare the cost of restoring the internal clearances after wear has taken
place with the value of the power savings we can obtain from operating a pump with original size
clearances. This cost is relatively easy to determine by obtaining prices on new parts and estimating the
cost of the required labor. But the savings are not the same for every pump. Both analytical and
experimental data have indicated that leakage losses vary considerably with the specific speed of a pump.
Figure 21.4 shows the relationship between the leakage losses of double-suction pumps and their
specific speed.
Let us examine a few typical cases. The pump illustrated in Fig. 21.1 when fitted with the maximum
375 mm (14.75 in) impeller, has its best efficiency at 772 m3Jhr (3,400 gpm) and 57.3 m (188 ft) head.
Its specific speed is

N = 1,800 (3400)°·5 =2 070


S (188)0.75 ,
556 Centrifugal Pumps and Energy Conservation

o
i1i
::r:

INCREASE
IN LEAKAGE -
1
1-

CAPACITY

Fig. 21.3 Effect of wear at internal running clearances on pump head-capacity curve.

From Fig. 21.4, we can estimate that its internal leakage losses-when the pump is new-are of the
order of 1.4 percent. Thus, when the internal clearances will have increased to the point that this leakage
will have doubled, we can regain approximately 1.4 percent in power savings by restoring the pump
clearances. But if we are dealing with a pump designed for 41 m3Jhr (180 gpm) and 76.2 m (250 ft)
head at 3,550 rpm, the specific speed is

10
I-
:::l .
c..

\
~ 8
c:
w
~
0

\
Il..
u.. 6
0
"e
0

en
w
(/) 4

'~
(/)
0
-l
W
C)
2
~ -............
-----
«
w
-'
0
0.5 1.0 1.5 2.0 2.5 3.0 3.5
SPECIFIC SPEED x 10- 3

Fig. 21.4 Variation in leakage losses with specific speed for double suction pumps.
Centrifugal Pumps and Energy Conservation 557

N = 3,550 (180)°5 = 755


s (250)075

Such a pump will have leakage losses of about 5 percent. If the clearances are restored after the pump
has worn to the point that its leakage losses had doubled, we can count on 5 percent savings.
Thus, restoring clearances of the lower-specific-speed type pumps gives us greater returns in terms
of the reduction of leakage losses. In addition, lower-specific-speed pumps generally have higher heads
per stage than higher-specific-speed pumps and, all else being equal, wear is increased with higher
differential heads. Thus, one will generally find more reasons to renew clearances of high-head pumps
and be rewarded by more savings from doing so.

ADDITIONAL AREAS TO INVESTIGATE

There are a few additional suggestions that can be made to save energy in the use of pumps and
pumping stations:

1. Weigh cost of larger piping, valves, and heat exchangers versus cost savings in pumping.
2. Select low-pressure-drop-control control valves.
3. Do not specify pumps to be rated at some arbitrarily set percentage of BEP.

Finally, there is a fourth area of investigation where additional power savings may be available for a
given pump installation. Many processes require pumps to handle quite viscous liquids, but the opportunity
of substantially reducing the viscosity of the liquid by the addition of heat energy before the liquid
reaches the pump is frequently overlooked. In many cases, however, the heat energy that would have
to be added is far less than the additional energy that would have been required by the pump to overcome
the higher system friction losses and to compensate for the lower efficiency of the pump when pumping
a higher-viscosity liquid. Furthermore, even allowing for radiation losses, many processes will utilize
that heat energy at some point downstream from the pump, so that the added heat energy would be
largely conserved.
22
Pump Operation at
Off-Design Conditions

Theoretically, as long as the NPSH available is greater than the required NPSH, a centrifugal pump is
capable of operating over a very wide range of capacities. As explained earlier, the exact operating
capacity is determined by the intersection of the pump head-capacity curve with the system-head curve.
This operating capacity can be changed by altering either one or both of these curves: varying the pump
speed will alter the head-capacity curve, whereas throttling the pump discharge will alter the system-
head curve. But at any given speed, the performance of a centrifugal pump is at its optimum at only
one capacity point, and that is the capacity at which the efficiency curve reaches its maximum; at all
other flows, the geometric configuration of the impeller and of the casing no longer provide an ideal
flow pattern. Our definition of "off-design" conditions must therefore be any condition wherein a pump
is required by the force of circumstances to deliver flows either in excess of or below the capacity at
best efficiency.
The treatment of off-design operation of centrifugal pumps starts with the recognition of the various
events that take place within such pumps as their flow is varied from BEP. Once the events are understood,
it is then necessary to quantify their effect on the pump, and from that determine the allowable flow
range of the pump for a given set of operating conditions. As a guide to the following discussion, Fig.
22.1 shows in a qualitative manner the events that take place in low to medium specific speed pumps
as their flow is varied.

OPERATION AT HIGH FLOWS

There are two circumstances that might lead to the operation of a pump at flows in excess of its best
efficiency or even of its design point. The first of these occurs when a pump has been oversized by
specifying an excessive margin on total head. Under this circumstance, the pump performance and its
relation to the system-head curve might look as in Fig. 22.2. The head-capacity curve intersects the
system-head curve at a capacity much in excess of the real required flow.
If the process is mass-flow controlled, such as in steam power plants or refineries, the pump will be
throttled back to the required capacity. In these cases, the consequence of specifying an excessive head

558

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
Pump Operation at Off-Design Conditions 559

HIGH
TEMP. HIGH
RISE ROTOR
I FORCES SUCTION
1 RECIRCULATION
I DISCHARGE
TOTAL I RECIRCULATION
HEAD
NPSHR
STEEPENS
I IINLET
CHOKED

BEP FLOW

Fig. 22.1 Events at off-design operation of low and medium specific speed centrifugal pumps.

H-Q

!="o

.w
u..I
u.......J

0..1--
IO
all--
a:: =HEAD AT DESIRED
I--I--...;!~-~-= -- -- -- (§§ CAPACITY
-0:- - - - - --
w=
a:=
w==
CiS
0=

CAPACITY

Fig. 22.2 Effect of overstating pump head.


560 Pump Operation at Off-Design Conditions

~_____H~-Q~(nN~O
__P_UM~PS~/N~P~~~R
I ~LLEL)
N::o
7,0

.-' -.-
~~

----------- a.1 ::i:,


I~
~I '::::>
w, Ia.
~I ,W

0' I~
eI
w,
'0
Ix

~I
a::: , ::i:

CAPACITY

Fig. 22.3 Two- and one-pump operation.

margin is higher than necessary power consumption (which can be significant; see Chapter 21). When
the process is not mass-flow controlled, as is the case in most water works applications, the pump is
allowed to run against the system and so will operate at the higher capacity shown in Fig. 22.2. Unless
sufficient NPSH has been made available, the pump may suffer cavitation damage and, of course, the
power consumption will be higher for pumps with a specific speed under about 4,500.
The second circumstance occurs when two or more pumps are nonnally used in parallel and one of
the pumps is taken out of service because the demand has decreased. Figure 22.3 describes the operation
of two such pumps. Whenever a single pump is running, its head-capacity curve intersects the system-
head curve at flows in excess of the design capacity. This is called the "run-out" point. Here, again, the
available NPSH and the size of the driver must be carefully selected to satisfy the conditions that will
prevail at the run-out point.

OPERATION AT LOW FLOWS

The most frequent cause of operation of a pump at reduced flows is, of course, the reduction in demand
of the process served by the pump. However, it can also happen that two pumps operating in parallel
may be unsuitable for this service at reduced flows and one of the pumps on the line may have its check
valve closed by the higher pressure developed by the stronger pump.
Operation of centrifugal pumps at reduced capacities leads to a number of unfavorable conditions
that may take place separately or simultaneously. Some of the consequences that must be anticipated or
circumvented are

1. When operation at reduced flow is required by the characteristics of the process served by the pump, its
effect can be obviated somewhat by using variable speed drive or by using several pumps for the total
required capacity and shutting down pumps sequentially as the total demand is reduced.
Pump Operation at Off-Design Conditions 561

2. If the pump has a single volute casing, its rotor will be subjected to a higher radial thrust as flow decreases.
A pump that is expected to operate at lower-than-design flow must therefore have a shaft that is stiff enough
to keep deflection within limits and bearings with enough capacity to accommodate the higher load.
3. As the capacity is reduced, the temperature rise of the pumped liquid increases. To avoid exceeding permissible
limits, a minimum flow bypass must be provided. This bypass, which can be either manual or automatic,
will also protect the pump against the accidental closing of its check valve while the pump is still running.
4. At certain flows below that at best efficiency, all centrifugal pumps are subject to internal recirculation, both
in the suction and discharge areas of the impeller. This can cause hydraulic surging and erosion of the
impeller similar to that caused by classic cavitation, although taking place in a different area of the impeller.
S. High-specific-speed pumps have power curves that rise with reduced capacities. Unless the driver size has
been selected with this fact in consideration, it may be overloaded when operating capacities are reduced.
6. As previously mentioned, if the liquid contains an appreciable amount of entrained air or gas, and if the
pump capacity is reduced below a certain minimum, the pump can become air-bound.

Each of these effects may dictate a different recommended minimum operating capacity. Obviously, the
final decision must be based on the greatest of the individual minimums. We shall now examine them
in greater detail except for the first effect, which has been treated extensively in Chapter 21, and for the
effect of radial thrust, covered in Chapter 2.

TEMPERATURE RISE EFFECT

The thermodynamic problem that arises from the operation of a centrifugal pump at very low flows is
caused by the heating up of the liquid handled by the pump. The difference between the brake horsepower
consumed and the water horsepower developed represents the power losses within the pump itself, except
for a very small amount lost in the pump bearings. These power losses are converted into heat ana
transferred to the liquid passing through the pump.
If the pump is operating against a completely closed valve, the power losses are equal to the shutoff
brake horsepower and, since no flow takes place through the pump, all this power goes into heating the
small quantity of liquid contained within the pump casing. As this process occurs, the pump casing itself
heats up and a certain amount of heat is dissipated by radiation and convection to the surrounding
atmosphere. If the amount of heat added to the liquid is small, it can be transmitted through the casing
with a low differential in temperature between the liquid in the casing and the outside air. If the power
loss is very high, however, the liquid temperature might have to reach an exceedingly high value, far
in excess of the boiling temperature 'at suction pressure, before the amount of dissipated heat equals that
generated in the pump proper. Operation of the pump under such conditions would have disastrous effects.
The rate of heating the liquid in the pump casing for a given power loss depends on both the volume
of water contained in the casing and the surface area of the casing that can dissipate heat. For practical
reasons, dissipation of heat by radiation can be ignored, and the temperature rise can be determined
from the formula

AT'I . kP o
L.1 ,mm = WpCp + WwC w

where .:iT/min =Temperature rise, in degrees Celsius (Fahrenheit) per minute


Po = power at shutoff, in kW (bhp)
k = Conversion constant, 14.34 for kW to Kilocal per minute (42.4 for brake horsepower to BTU
per minute)
Wp = Net weight of pump, in kilograms (pounds)
562 Pump Operation at Off-Design Conditions

Cp = Specific heat of pump metal (approximately 0.13)


Ww = Net weight of liquid in pump, in kilograms (pounds)
Cw = Specific heat of liquid (1.0 if liquid is water).

Because the temperature rise may be so rapid that there is no time to transmit the heat from the liquid
to the pump casing, it is generally safer to neglect the casing altogether. The formula is then simplified to

Il.T/min = /cPo
WwCw

For example, if the pump handles water (Cw = 1.0) and contains 45.4 kg (100 lb) of liquid, and if the
power at shutoff is 75 kW (100 hp), the water temperature will increase at the rate of 23.6°C (42.4°F)
per minute. Operation at shutoff under these conditions is very dangerous. But with a low-head, high-
capacity pump that contains 2,770 kg (5,000 lb) of water and that takes the same amount of power at
shutoff, the rate of temperature increase will be only 0.47°C (0.85°F) per minute-hardly serious if the
operation against a closed discharge valve is not prolonged.
If liquid is flowing through the pump, conditions become stabilized and the amount by which the
temperature at the discharge will exceed the suction temperature can be calculated for any given flow.
The heat transmitted to the liquid in the pump is equivalent to the product of the difference between
the shaft power (Ps) and the water or hydraulic power (Ph) by a constant to convert mechanical energy
to thermal (860 for kWhr to Kcal-2,547 for hp-hr to BTU). The temperature rise can therefore be
calculated from the formula:

Il.T = (Ps-Ph )860


(Flow in kg/hr)C w

for IlT in °C, or for IlT in OF from

Il.T = (Ps-Ph)2,547
(Flow in Iblhr)Cw

A more convenient formula relates the temperature rise to the total head and to the pump efficiency:

Il.T = ( total head in meters) ( _1_ )


427Cw 11-1

for IlT in °C, or for IlT in OF from

Il.T = (tOtal head in feet)


778Cw
(_1
11-1
)
This formula can be used to plot a temperature-rise curve directly superimposed on the performance
curve of a centrifugal pump. In looking at temperature rise from the aspect of minimum allowable pump
flow, the fundamental objective is to avoid destructive flashing within the pump. The temperature rise
of concern is therefore that of the liquid that is returned to suction pressure (or close to it) within the
pump. This is generally referred to as the "total temperature rise" and is conservatively estimated as
Pump Operation at Off-Design Conditions 563

For !1T in °C, or for f)T in of from

H
Total IlT = 778Cw(,rt)

for !1T in OF and H in feet. The head, H, in these equations is the head that is broken down to
suction pressure within the pump across the wearing rings or balancing device. This varies with pump
configuration and is generally the following:

Pump Configuration Head Broken Down to Suction Pressure

Single stage Total head of the pump


Multistage, opposed impellers Typically half the total head(1)
Multistage, tandem (in-line) impellers Total head of the pump
'Depends on number of stages and whether first stage is double suction.

Figure 22.4 shows the characteristic of total temperature rise plotted directly on the performance
curve of a centrifugal pump, in this case a tandem impeller boiler feed pump designed to handle 125
m3Jhr (550 gpm) of 121°C (250°F) feedwater against a total head of 549 m (1,800 ft). As shown, the
temperature rise increases very rapidly at lower flows. This is caused by the fact that the losses in the
pump increase as the mass flow to absorb them decreases. For example, Figure 22.4 shows that at 11

Ii: 22
TOTAL HEAD
0
0
T"" 20
r
Cl
80 «w 18
:::I: EFFICIENCY___--_~
....J
70 « 16 50
b
I- 14
60 40 \f0-
r
50 12 30m
:.e
0
ii:
~40 200...
::t
u.. W
II.. I-
w 30 10
TOTAL TEMPERATURE RISE
20 400 o
a.. L-~~~-------B~H~P~--------------------l
10 ai 200
0 0
0 2 3 4 5 6 7
FLOW - 100 USGPM.

Fig. 22.4 Centrifugal pump performance curve.


Total temperature rise is also indicated.
564 Pump Operation at Off-Design Conditions

m3/hr (50 gpm) the total temperature rise is 13°C (23°F), whereas at the rated capacity of 125 m3/hr
(550 gpm) it is only 1.8°C (3.2°F).
The maximum permissible temperature rise varies over a wide range, depending on the pump type
and the pumped liquid. As already noted, the fUndamental objective to avoid destructive flashing within
the pump. With homogeneous liquids such as water, flashing must be avoided altogether. Current practice
with water is to have a 0.3 bar (5 psi) margin over flashing at all points of low pressure within the
pump, and an upper limit of 25°C (50°F) on the total temperature rise itself. With multifraction liquids,
such as most hydrocarbons, some flashing across the balancing device of multistage pumps can be
tolerated. This is particularly so for light hydrocarbons, which produce both a high temperature rise and
quite a deal of auto-refrigeration when flashing does occur. With our expanded knowledge, factors other
than temperature rise generally establish the minimum recommended flow for most pumps, as will be
seen later.

INTERNAL RECIRCULATION

The subject of internal recirculation had, until very recently, been little understood outside of a small
number of pump designers. At certain flows, generally below that of best efficiency, all centrifugal
pumps are subject to internal recirculation in the suction and discharge areas of the impeller. The result
of these two phenomena is a great increase in pressure pulsations (Fig. 22.5). Note that the capacities
at which the recirculation at the suction and at the discharge arises are not necessarily coincidental.
Internal recirculation at the suction is most frequently the cause of field problems.
That abnormal flows could occur in and around impellers under certain conditions was first suggested
by Smith [22.1] (reported by Gibson [22.2] in 1908). Stewart [22.3], in 1909 measured prerotation
upstream of an impeller. Carrard [22.4], addressing impeller flow separation in 1923, suggested that
continuity would dictate separation at the channel inlet as the through flow was reduced. Fisher and
Thoma [22.5] visualized recirculating flow in 1932. Stepanoff [22.6], dealing with centrifugal pump
performance in 1948 ascribed the departure of the actual power curve from the theoretical to the onset
of "secondary flow" within the impeller.

w
a: INCIPIENT DISCHARGE RECIRCULATION
::::>
C/)
C/)
w
a: INCIPIENT SUCTION RECIRCULATION
a..
~
SUCTION PRESSURE

CAPACITY

Fig. 22.5 Effect of internal recirculation on pressure at pump suction and discharge.
Pump Operation at Off-Design Conditions 565

For centrifugal pumps, abnormal flows in and around the impeller remained largely a hydraulic
curiosity until the late 1950s and early 1960s. Around this time, the introduction of higher head per
stage designs resulted in very unfavorable behavior whenever these pumps were operated at reduced
flows. On frequent occasions, the impellers of these pumps showed premature erosion in their suction
areas and, sometimes, near their discharge tips. In some extreme cases, catastrophic failures occurred,
including the destruction of impellers and volutes or diffusers. Although the majority of these problems
may have been acutely noted in the area of such equipment as multistage boiler feed pumps, equally
unwelcome difficulties were being encountered with pumps on less severe services. For example, a
number of single-stage high-head pumps in the medium range of heads per stage exhibited similar
problems, although possibly of a lesser magnitude.
It was in the 1960s that these symptoms were first associated with the phenomenon of internal
recirculation in the impeller, either at the suction or the discharge and, sometimes, at both locations.
Minami et al. [22.7] provided insight into the connection between suction recirculation and cavitation
type erosion of the impeller. Various other authors [22.8, 22.9, 22.10] offered qualitative explanations
for the causes of recirculatioIl. In 1972 Fraser [22.11] developed a proprietary mathematical model
of the relationships that exist between the geometric configuration of an impeller and the onset of
internal recirculation.
A wider circulation was given to this information in articles published in 1975 and 1976 [22.12].
Finally, in 1981, Fraser [22.13, 22.14] published two papers dealing quantitatively with recirculation.
These two papers represented a substantial contribution to the art of centrifugal pump design and
application. They established a simple means of calculating the onset of recirculation for known impeller
geometry, the means of estimating same when the geometry is not known, guidelines for pump operation
based on recirculation flows and, finally, suggestions for correcting or accommodating existing recircula-
tion problems.
Fraser presented the first published means of calculating the onset of recirculation. In some quarters,
the initial reaction to this work was the argument that the complex, three-dimensional flow prevailing
within an impeller could not be modeled in so simple a manner. From a strictly scientific point of view,
that argument held some truth, but it overlooked the essential point-whether the method gave results
good enough to identify, correct, and ultimately avoid problems caused by recirculation. Field experience
has shown the method satisfies the essential criterion and has thus transformed criticism into efforts
aimed at improving the accuracy of calculation, refining application guidelines, and extending pump
rangeability. Of note in this connection is the work of Oshima [22.15], Bosman and Bhrabian [22.16],
Palgrave [22.17], and Gopalakrishnan [22.18].

INTERNAL RECIRCULATION AT THE SUCTION

One of the means used to reduce the required NPSH is to increase the eye diameter of the impeller,
thereby reducing the entrance velocities (see Fig. 19.23). This, in tum, increases the peripheral velocity
of the impeller at the eye and, at some capacity, the distortion of the velocity triangles causes the flow
at the outer eye diameter to reverse itself and flow back out of the impeller. The exact flow at which
suction recirculation takes place is dependent on the design of the impeller. The important fact that must
be remembered is that the larger the impeller eye diameter, and the larger the areas at the impeller
suction relative to its overall geometry, and, therefore, the lower the required NPSH at a given capacity
and speed, the higher will be the capacity at which recirculation takes place in percentage of the capacity
at best efficiency (Fig. 22.6).
Internal recirculation causes the formation of very intense vortices (Fig. 22.7) with high velocities at
their core and, consequently, a significant lowering of the static pressure at that location. This, in tum,
566 Pump Operation at Off-Design Conditions

u.:
IJ..
W
ci
~
::I:
-.J

~
::i
en
a...
Z L-________________________

CAPACITY

Fig.22.6 Effect of NPSH on recirculation flow.

ENTERING
FLOW --.

Fig.22.7 Section through a single-suction impeller indicating the recirculation of liquid at the inlet during oper-
ation at low capacities and the recirculation vortices fonned at the impeller discharge at low capacities.
Pump Operation at Off-Design Conditions 567

Fig. 22.S Diagnosis of cavitation damage.

leads to intense cavitation accompanied by severe pressure pulsations and noise and can be damaging
to the operation of the pump and to the integrity of the impeller material. Note, incidentally, that the
location of the material damage is an excellent diagnostic tool in identifying whether the cause is classic
cavitation or internal suction recirculation. If the damage is on the visible side of the inlet impeller
vanes, the cause is classic cavitation. If the damage is to the hidden side (pressure side) of the vanes
and must be seen with the help of a small mirror, the cause is suction recirculation (see Figs. 22.8, 22.9,
and 22.10).

Fig. 22.9 Front view of impeller damaged by recirculation.


568 Pump Operation at Off-Design Conditions

Fig. 22.10 Cavitation damage caused by suction recirculation.


The damage is on the upper (non-visible) side of the inlet vanes and can be seen with a small mirror.

Field problems caused by this phenomenon obviously did occur from time to time, but it was in the
early 1960s that they became intensified. Two factors led to a greater incidence of problems:

1. More pressure was exerted by the users to have the pump designers reduce required NPSH values. This
was achieved by increasing the impeller suction eye areas, bringing the onset of internal recirculation closer
and closer to the best efficiency capacity.
2. Higher heads per stage and larger pump capacities were increasing the energy levels of individual impellers,
intensifying the unfavorable effects of internal recirculation.

Information on the explanation of the phenomenon was first released in 1972 in a limited circulation
paper and later, more widely, in an article [22.12]. For certain obvious reasons, the mathematical solution
was considered proprietary and was not published until 1981. Guidelines were given in [22.12] suggesting
that on water applications suction specific speed (S) values not be allowed to exceed a range of 8,500
to 9,500 (US units) so as to avoid field problems if operation at significantly reduced flows was
contemplated. These guidelines were based on the qualitative reasoning that in high S impellers, the
onset of suction recirculation occurred close to BEP and with such intensity that the NPSH to avoid
damage, NPSlI.!, was generally above the available NPSH (Fig. 22.11). Lowering the value of S lowered
both the capacity at which suction recirculation started and its intensity, such that in normal low S
designs the available NPSH was more likely to be above NPSH d (Fig. 22.12). Following this reasoning,
the allowable minimum flow of a high S design was determined by the risk of premature impeller
erosion, and so became the incipient suction recirculation capacity, whereas the low S design could
safely be run down to a much lower capacity determined from mechanical or thermal considerations
(Fig. 22.13). Subsequent work on cavitation erosion in impellers has substantiated this qualitative
reasoning and allowed NPSHd to be calculated for various designs (see Chapter 19).
Pump Operation at Off-Design Conditions 569

1
cl
~ :8 1
~ i'
~
~
-:i
jlj
c

g -__ -8'1 ~

II.. NPS
. . .Hav.a.I.la.b.le~~................-,. .~~,~~\~-~c+:....~~
..

NPSH required

R.OW

Fig. 22.11 Relation between NPSH, suction recirculation capacity, and cavitation erosion for a high suction
specific speed (low-NPSH) design.

~
~ no cavit3tion

~1"""·~"~~~~~~~"-r....~NPS~~H~.va~i~~b~le~. . . . .
~ \ ) , increaed NPSH required
V 1 at center of vortex
\ ,durlns suction

\:
\ , recircmtlon

·_ _------~--iNPS~H required

R.OW

Fig. 22.12 Relation between NPSH, suction recirculation capacity, and cavitation erosion for a moderate
suction specific speed (higher NPSH) design.
570 Pump Operation at Off-Design Conditions

I
----'-UNSTABLE--I
I REGION
, PUMP A

::t PUMPB
0' NPSHR
«
UJ
I

25% 100%
CAPACITY Q IN %

Fig. 22.13 Comparison of safe zones of operation for normal and high suction specific speed impellers.

Fraser's paper [22.13] gave exact fonnulas from which to calculate the flow at which internal
recirculation would start at the suction, once certain geometric data were known about an impeller.
Fraser [22.14] provided close approximation curves for the event that these data were not readily available
(see Figs. 22.14,22.15, and 22.16). Note that these curves are representative of general design practice.
In the 12 years since they were published, more has been learned about impeller design. As an example,
for a multistage pump of 9,500 suction specific speed (US units), Figure 22.14 puts the suction recirculation
capacity at 82 percent of BEP, yet there are today designs of such impellers whose suction recirculation
capacity has been verified by flow visualization at 45 percent of BEP. The curves are therefore helpful
to the pump user for assessing general designs, but will be overly conservative for "state-of-the-art"
designs. When the latter are being considered, the design should be assessed using the method proposed
by Fraser, a computer routine for calculating impeller inlet flow that has been validated by flow visualiza-
tion, or by flow visualization itself.
There seems to be some discontinuity in the curve for the suction recirculation data given for specific
speeds of 500 to 2,500, and those between 2,500 and 10,000 in Figs. 22.15 and 22.16. This discontinuity
is created by the desire to simplify the presentation of the data. Theoretically, there would be a set of
curves for every discrete value of specific speed, and thus, a three-dimensional plot would be necessary
to present the facts· in the most accurate manner possible. The reason why a specific speed of 2,500 was
chosen as the breaking point is that this provides for conditions when the ratio of the eye diameter to
the outlet diameter exceeds 0.5 (for double suction impellers), at which ratio suction and discharge
recirculation occur simultaneously, the capacity being detennined by the higher of the two values (see
the Appendix). In nonnal designs, discharge recirculation occurs at a higher capacity than suction
recirculation and therefore establishes the recirculation capacity for the impeller. If data are available
to pennit the more exact calculations given in the Fraser's paper, his fonnulas should be used instead
of the curves (see the Appendix).
A ready means to confinn the value of the suction recirculation flow consists of locating a probe at
the eye of the impeller, with the impact portion of the probe directed into the eye. During nonnal flow
the reading is essentially the suction pressure less the velocity head at the impeller eye (Fig. 22.17). As
Pump Operation at Off-Design Conditions 571

100

LL
0
90 W
CJ
«~
z
w~

80 °w 0LL
0:-1
a..>-
1 o
Zz
Qw
~-
«$:2
-ILL
70 ::JLL
OW
O:~ 500 TO 2,500 Ns
-Cf)
Ow
Wee
0:
z
60 0
i=
°::J
Cf)

50
0 o o o o o
0 o o o o o
0 o o o o o
cD ex> o C\I ..,f cD

SUCTION SPECIFIC SPEED-S AT BEST EFFICIENCY


FLOW SINGLE SUCTION OR ONE SIDE OF DOUBLE SUCTION

Fig. 22.14 Suction recirculation capacity as a function of suction specific speed and hub-to-eye diameter
(hi /D I ) ratio.
Specific speed: 500-2500.

the output of the pump is reduced and suction recirculation occurs, the flow reversal at the impact head
probe will show a sudden rise in pressure, as shown in Fig. 22.18. This is as close as one can get to
measuring the onset of suction recirculation without going to the extremes of providing an experimental
pump in transparent materials and obtaining a visual observation of the phenomenon.
In all of this discussion it must be noted that S, the suction specific speed of the pump, must always
be calculated for the conditions corresponding to the capacity at best efficiency for that pump. The
guaranteed conditions of service mayor may not correspond to this best efficiency flow-they rather
seldom do. It must also be calculated on the basis of the pump performance with the maximum impeller
diameter for which it was designed. This constraint becomes obvious when one considers that the internal
recirculation at the suction occurs because of conditions that arise in and around the impeller inlet,
conditions that are not necessarily affected by cutting down the impeller diameter. This cutting down
will move the best efficiency point to a lower flow value, but will not reduce the flow at which suction
recirculation will occur.
Liquid characteristics do not affect the flow at which internal recirculation takes place. But they do
have a profound effect on the severity of the symptoms and on the extent of the damage. The reason is
S72 Pump Operation at Off-Design Conditions

II
100
V J
V

90
IL
0 / / V
I-
z
1&.13=
U o l.
V ~
r~
~
/
80
a:...J
~IL
f
~ .~I( W' ~ ~c.

fj V'
1>-
ZU
OZ
-1&.1
~"/ ~ QO
~)
.1(

1--

/ /'
ct~ V

70
...JIL
;:)IL
UI&.I /
V /
a:
-I-
uen
I&.Iw
a: en / 500-2500 NS

60
Z
0
j::
V / /
/ V
U

/
;:)
en
/
50 /
/ V
o o o o o o
o o o o o o
o o o o o o
to en o N <t to

SUCTION SPECIFIC SPEED-S AT BEST EFFICIENCY FLOW


SINGLE SUCTION OR ONE SIDE OF DOUBLE SUCTION

Fig. 22.15 Suction recirculation capacity for single- and double-suction and multistage pumps.
Specific speed: 500-2500.

exactly the same as in the case of the NPSH required to prevent the symptoms and damage caused by
"classical" cavitation, as described in Chapter 19. Thus, internal suction recirculation will always be
less damaging when the pump handles hot water and particularly when it handles hydrocarbons than
when it handles cold water.
It would seem that now that means are available to calculate the onset of internal recirculation, users
and designers of centrifugal pumps should be in the position to establish sensible limits for minimum
operating flows. To do so, it would be necessary to establish some sort of guidelines between minimum
flows and recirculation flows. But this is just the problem. Setting up such guidelines is a fairly complex
task for a number of reasons. Operation of a pump at flows below the recirculation capacity can have
a number of consequences, typically abnonnal noise, high vibration, premature bearing failure, and
premature impeller erosion. Whether these consequences are severe enough to dictate that the minimum
flow be above the recirculation capacity depends on a variety of factors, such as

1. Size of the pump, that is capacity, head, and power


2. The design of the impeller, which is indicated by the value of S
3. Liquid characteristics
4. Materials of construction
5. Length of time the pump operates below certain critical capacities
6. What the pump user considers "normal" pump behavior.
Pump Operation at Off-Design Conditions 573

v V
100

/
90 IL.
/ V /
V V
0
r-
z:::
1&.1 0
~,~ ~ AO~ L
V
~ cj ~
u..J
ffilL.
80 a.. ~/ ~
V :! ~ ~
IU
>-
ZZ

vr
o~ ~,.~

70
ij2
..JIL.
:;,IL.
ul&.I
V / V

V
~r-
UIF)
1&.11&.1
ireD V V
./
2500-10000NS

60
Z
Q
I-
U
V
:;,
IF)

50
o o o o o o
o o o o o o
o o o o o
12 eD en
., II
o N

SUCTION SPECIFIC SPEED-S AT BEST EFFICIENCY FLOW


SINGLE SUCTION OR ONE SIDE OF DOUBLE SUCTION

Fig. 22.16 Suction recirculation capacity for single- and double-suction and multistage pumps.
Specific speed: 2,500-10,000.

normal flow
Fig. 22.17 Pressure indicated by pitot facing impeller eye at flow above recirculation capacity.
574 Pump Operation at Off-Design Conditions

recirculation flow

Fig. 22.18 Pressure indicated by pitot facing impeller eye at or below recirculation capacity.

The last factor has, in the past, made the choice of minimum flow guide-lines a very subjective one.
With growing general awareness of what constitutes normal pump behavior, a product of pump owners
seeking to lower maintenance expense on their pumps, the subjectivity is being replaced with specific
limits on allowable vibration at minimum flows and desired impeller life.
It can easily be demonstrated that there is a definite relationship between acceptable minimum flows
in percentage of recirculation flow and the suction specific speed, both by intuitive reasoning and by
actual field experience. Note that for the same speed and capacity, a higher suction speed design requires
a larger impeller eye diameter than does a lower suction specific speed design. This means that when
the backflow starts at the eye diameter of the impeller, the peripheral velocity of this backflow will be
higher with the higher suction specific speed design. It follows that the higher backflow velocity represents
a higher energy level, and that all the symptoms of operation in the recirculation zone will be intensified
with the higher suction specific speed design. One observation, however, can give some guidance to
users with respect to the effect of the S value on what constitutes an acceptable minimum flow. Referring
to Fig. 22.14, and assuming a hub to eye diameter (hlD t ) ratio of 0.45, an impeller with an S value of
14,000 will have its suction recirculation occur at about 100 percent of its best efficiency flow. If we
are dealing with a fairly substantial pump handling cold water, we should be most reluctant to consider
running such a pump below this 100 percent. On the other hand, a pump with the same hub-to-eye-
diameter ratio and an S value of 8,000 will have a recirculation flow of only 56 percent of best efficiency
flow and we need have no hesitation running it when necessary at as little as 25 percent of best
efficiency flow.
Some guide-lines, vague as they may appear to be, are
1. Unless there is a compelling reason to do so, do not specify NPSH values that result in S values much above
9,000 for water or above 11,000 for hydrocarbons.
2. When dealing with relatively small pumps, say 75 kW (100 hp)-per stage in multistage pumps, the effect
of suction recirculation is not apt to be as significant as for larger pumps.
3. Pumps handling hydrocarbons can be operated at lower flows than equivalent pumps handling cold water.
4. The risks of operating at flows much below the recirculation flow can best be determined after the pump is
in operation. Provision should therefore be made to increase the minimum flow bypass if there is a suspicion
that too optimistic a decision as to this minimum flow has been made at the time the pump was selected.
5. When the pump is never expected to operate at flows below its design condition, higher S values can be
used without concern for unfavorable effect from internal suction recirculation.
Pump Operation at Off-Design Conditions 575

r-..
h~ cu~e,13 JmJs iJ paral/ei
...... -- r--.. r--..
~9
r""""'-
~ <I
PI./.
-~ ~$ ./~<
\\ r'\.
""
system-head
~

-
curve

1 """,,-
...- ......... ~
.-r
--~
~
- static head

design point

capaCIty

Fig. 22.19 Operation of three condenser circulating pumps in parallel.

Point 5 will be the case, for instance, with constant speed condenser circulating pumps operating in
parallel, since it is not the practice to throttle their discharge. This situation can best be understood by
reference to a set of curves that describes the operation of three condenser circulating pumps operating
in parallel superimposed on the system-head curve (see Fig. 22.19). Note that whenever only one or
two pumps are operating, the head-capacity curves intersect the system-head curve at flows in excess
of the design capacity. As a matter of fact, this situation is even further accentuated whenever the system-
head curve has been drawn up pessimistically high. In a case such as this, one can certainly use S values
considerably higher than when a pump may be called on to operate over a wide range of capacities to
the left of its design condition.
The only precaution that must be taken is not to apply this approach if the pumps discharge through
several heat exchangers in parallel, some of which may, on occasion, be bypassed or taken out of service.
This is described in a curve showing two cooling tower pumps discharging through 10 heat exchangers
(see Fig. 20.20). Whenever, let us say five of these heat exchangers are bypassed, the system-head curve
steepens, and the operating capacity is reduced to much below the design conditions.
There is a "post-factum" modification that, in a number of cases, has been used quite successfully
to reduce and even sometimes to eliminate the unfavorable effects of suction recirculation. It consists
of retrofitting pumps with stationary casing rings the apron of which extends inwardly of the impeller
eye diameter (Fig. 22.21). If preferred, such rings can instead be rotating and mounted on the impeller.
These rings are commonly referred to as "bulk-head rings." They prevent the recirculation vortex from
extending axially beyond the plane formed by the apron. Of course, since they do increase the required
NPSH, bulk-head rings can only be resorted to if there is sufficient margin in the available NPSH.
Despite what is now known about recirculation, it is an unfortunate fact that there is still some
resistance to using safe and sound guidelines, both in choosing suction specific speed values and in
setting minimum flow restrictions. In this last connection, one might sometimes suspect that the minimum
recommended flow for one and the same value of suction specific speed seems to vary inversely with
576 Pump Operation at Off-Design Conditions

I !
-
I I I I I
system-head curve r--
Ir with 5 heat exchangers t - -

- ; - I--.
hV 1 I
- -q. 2 PlJrnps
s~stem-head
curve with

-
.......
........ / r-r--- ./ "" 1 0heat
'" ).
.......
,/ V ...... ........ exchangers

- ""-
~
V
i,..- ...... ...... static
head

capacity

Fig. 22.20 Operation of pumps in parallel with ten and with five heat exchangers.

Fig. 22.21 "Bulk-head" ring construction.


Used in 1939 by Oscar H. Dorer of Worthington to eliminate unfavorable effects of excessively large impeller
eye diameter.

a manufacturer' s desire to obtain an order. This is not a sound way to ensure a safe. reliable. and
trouble-free installation. Such practice should be discouraged.

INTERNAL RECIRCULATION AT THE DISCHARGE

Recirculation at the impeller discharge (Fig. 22.7), is a less frequent cause of problems than suction
recirculation, but when problems do occur they can be quite severe. Once discharge recirculation is
Pump Operation at Off-Design Conditions 577

established, the vortices associated with it produce low pressure regions with the potential for caviation
erosion, and random pressure pulsations that act both within the impeller between its shrouds and in the
spaces between the impeller and the adjacent casing. The phenomenon is known to produce four effects.
The first is a discontinuity in the head characteristic (Fig. 22.22) caused by the change in flow
regime as discharge recirculation is established. As shown in Fig. 22.22 there is some hysterisis in the
establishment of this discontinuity; the capacity at which it is evident will be lower with decreasing
pump flow, and it will persist to a higher capacity with increasing flow. This discontinuity can present
an operating problem if its magnitude is enough to create an unstable (multiple point in this case)
intersection between the pump and system head characteristics. Should that happen, the correction is to
redesign the impeller.
In operation, the next effect of discharge recirculation is what is termed axial instability or shuttling
of the pump rotor, caused by pressure pulsations between the impeller and adjacent casing walls. Rotor
shuttling can be seen, and is therefore more distressing, in pumps whose thrust bearing has a relatively
large axial clearance. This is inherent in tilting pad thrust bearings, and can occur in anti-friction bearings
when the build-up of tolerances between the bearing outer race and its housing components allows large
endplay. Whether rotor shuttling is visible is not as important as the magnitude of the instantaneous
forces imposed on the thrust bearing and the section of shaft connecting the impeller to the bearing.
The magnitude of these forces depends on the intensity of the pressure pulsations, which are related to
the differential pressure and the hydraulic design, and the annular area of the impeller shrouds. Given
these factors, the forces tend to be most severe in low specific speed pumps. In severe cases, the forces
can be high enough to cause premature thrust bearing failure and even fatigue failure of the shaft.
Visible rotor shuttling can be corrected by reducing the endplay in anti-friction thrust bearing assem-

HEAD CURVE FOR INCR --


EASING F-L--'::~-­
Ow

z
0
~
-I
I- ::J
Z 0
0
w a:
«w ii: (3
(3 W
I ~ a:
0 w
w (!J
I- a:
::J «
!l. I
:::!: 0
0 en
0 (5

FLOW

Fig. 22.22 Discontinuity in head-capacity curve caused by discharge recirculation.


578 Pump Operation at Off-Design Conditions

PROJECTIONS
FROM CASING
WALL TO REDUCE
AXIAL UNBALANCE -----fS4-----.--+---.

Fig. 22.23 Projections from casing wall to minimize problems caused by discharge recirculation.
Close clearance with impeller shroud reduces pressure pulsations acting over impeller shroud;
see also Fig. 4.35.

blies (being careful to allow for thermal expansion), or changing the diameter of one wearing ring to
produce a thrust bias large enough to keep the rotor in one axial position. This does nothing, however,
to reduce the fluctuating axial forces acting on the rotor. To do that, it is necessary to reduce the intensity
of the pressure pulsations acting over the impeller shroud. Generally this is done placing what is in
effect an "orifice" between the impeller discharge and the regions between the impeller shrouds and
casing walls. Various arrangements of close clearances at or near the impeller shroud O.D. (Figs. 4.35
and 22.23) are employed to achieve this effect.
The third effect of discharge recirculation is cavitation erosion of the fillets between the pressure
(upper) side of the vanes and the shrouds (Fig. 22.24). The fact that these effects (for both suction and
discharge recirculation) are a form of cavitation is illustrated by the observation that in a high pressure
multistage pump, physical damage is generally limited to the first stage. Subsequent stages have a greater
margin over the liquid vapor pressure and the effect is either minimized or eliminated. Cavitation erosion
of the fillets weakens the impeller shrouds and leads ultimately to a piece of the shroud breaking out,
after which the impeller is badly out of balance.
As well as causing rotor shuttle, the pressure pulsations at the impeller discharge produce fluctuating
forces on the impeller shrouds. If the pulsations are intense enough, the shrouds of impellers in brittle
materials, e.g. cast iron or Niresist™, will fracture, leading again to a badly out of balance impeller.
When the impeller material is more ductile, such as bronze or chrome steel, the shrouds are spread apart
(bowed) between the vanes (Fig. 22.25).
Pump Operation at Off-Design Conditions 579

- -
-- ~
~
- .
-~ ~-- ..

Fig. 22.24 Cavitation erosion caused by recirculation at impeller discharge.

Fig. 22.25 Bowed shrouds caused by recirculation at impeller discharge.


580 Pump Operation at Off-Design Conditions

Fig. 22.26 Erosion of first-stage impeller vane tips caused by running with close clearance to stationary vanes_

Fig. 22.27 Erosion of first-stage twin volute corresponding to impeller in Fig_ 22.26.
Pump Operation at Off-Design Conditions 581

The ideal cure for either cavitation erosion or shroud damage at the impeller discharge is to not
operate with discharge recirculation, which means either raising the operating flow of the pump or
changing the impeller design. If this cannot be done, changing materials can help lower the rate of
cavitation erosion or reduce the risk of shroud fracture. Finally, if shroud bowing is the problem,
increasing the thickness of the shrouds will help.
Damage to the impeller and pressure pulsations caused by discharge recirculation should not be
confused with the effects of the so-called "vane-passing syndrome." This is a phenomenon caused by
the interaction between the moving impeller tips and the stationary volute tongues or diffuser vanes. A
hydraulic shock occurs when the impeller vanes pass these stationary parts; the magnitude of the shock
and the resulting pressure fluctuations increase with the impeller tip velocities and with the pump size,
while the frequency is a multiple of the pump speed and of the number of impeller vanes.
Because the unsteady flow caused by the wake of the vortices shed from the vanes persists for some
distance, the major factor in the magnitude of the shock resides in the gap between the vane tips and
the volute tongues or diffuser tips. On the other hand, an excessive gap will act unfavorably on the
pump efficiency. Opinions differ somewhat about the minimum recommended gap, but it should not
fall below approximately 4 to 6 percent of the impeller diameter.
Figures 22.26 and 22.27, show the impeller and twin-volute of the first stage of a high-pressure
multistage pump that was subject to the vane passing syndrome because of an excessively close gap.
Instead of damage to the vane tips near the shrouds, it occurs near the central part of the vane. At a
point where the pressure would normally be 500 to 600 psi above the vapor pressure, the marked
cavitation erosion indicates that the shock pressure of the impeller vane tip passing close to the twin
volute. tongues was momentarily reducing the local pressure to less than the vapor pressure. This
cavitation-erosion occurred only at the first stage. Pressures reached at the higher stages prevented the
shock pressure from reducing the local pressures to values below the vapor pressure. This does not
mean, however, that hydraulic shock and pressure pulsations were not being generated. The cure for the
vane passing syndrome was to increase the gap by cutting back the volute tongues.
Tables 22.1 and 22.2 summarize the comparisons between classical cavitation and internal recirculation
at suction as well as between the vane passing syndrome and internal recirculation at the discharge respec-
tively.

Table 22.1 Comparison of Classical Cavitation and Internal Recirculation at Suction

Classical cavitation Internal Recirculation at suction

Occurs at high flows Occurs at low flows


Symptoms disappear Symptoms disappear as capacity is increased
As capacity is decreased or
As NPSH available is increased
Damage to impeller on visible side of vanes Damage to impeller on invisible side of vanes

To eliminate To eliminate
Increase available NPSH or Operate at higher flows or
Operate at lower flow or Get a redesigned impeller
Obtain new impeller with lower required
NPSH
582 Pump Operation at Off-Design Conditions

Table 22.2 Comparison of Vane Passing Syndrome and Internal Recirculation at Discharge

Vane-passing syndrome Internal Recirculation at discharge

Occurs over entire range of capacity Occurs at low capacities


Frequency is multiple of pump speed and number Frequency is random
of impeller vanes
Damage greatest at center of vane tips at periphery Damage greatest on vane tips near shrouds
Can be eliminated or reduced by increasing gap To eliminate
between impeller and collector Operate at higher flows or
Get a redesigned impeller

POSSIBLE OVERLOAD OF HIGH-SPECIFIC-SPEED PUMPS

Pumps with low to medium specific speeds have brake-horsepower curves that decrease with reduction
in capacity (see Fig. 18.32). Thus, operating such pumps at lower than design flows will not overload
the driver. On the other hand, if we deal with high-specific-speed pumps, such as one whose characteristics
are illustrated on Fig. 18.34 (specific speed = 10,000), the break-horsepower increases with a reduction
in capacity. For instance, were such a pump to be operated at 50 percent of its rated capacity by means
of throttling the discharge, the brake-horsepower would increase to almost 140 percent of the rated
horsepower. If the motor size had not been selected with this end in view, it would be overloaded.
Obviously, sizing drivers of high-specific-speed pumps so that they are not overloaded over the entire
range of operating capacities down to 0 percent would be uneconomical. Thus, the minimum permissible
flow is dictated by the selection of the driver horsepower.

EFFECT OF ENTRAINED AIR OR GASES ON MINIMUM FLOWS

The fact that when entrained air or gas enters a centrifugal pump along with the pumped liquid, the
pump performance is affected unfavorably is discussed in Chapter 18. As the percentage of volume of
air or gas increases, the flow at which it can be swept out of the pump increases and therefore so does
the permissible minimum flow (see Fig. 18.56).

MINIMUM FLOW BYPASS

For one or more of the reasons just discussed, various types and sizes of pumps cannot be operated
below a certain minim flow. That flow is determined by the manufacturer based on the pump design
and the particular application. If the flow from the pump to the system must on occasion be less than
Pump Operation at Off-Design Conditions 583

t CONTROL VALVE

ORIFICE

VALVES

Fig. 22.28 Arrangement for minimum flow bypass.

the minim allowable for that pump, some means must be provided to ensure the pump always operates
at or above its minimum allowable flow. This is accomplished by installing a bypass in the discharge
line from the pump, located upstream of the check and gate (block) valves (Fig. 22.28) and leading to
some lower pressure point in the installation where the heat absorbed by the pumped liquid during
operation at low flows can be dissipated. Chapter 28 provides a detailed discussion of the design of
minimum flow bypass systems.

DISTINCTION BETWEEN "CONTINUOUS" AND


"INTERMITTENT" OPERATION

On occasion, manufacturers make a distinction between "continuous" and "intermittent" operation when
recommending minimum flows for their pumps. But how do you distinguish between the two? Of course,
one might select some quite arbitrary apportionment of time as a differentiation between the two: for
instance, any operation of a pump at some specific rate of flow (or below it) that takes place for less
than 25 percent of the time might be termed "intermittent," whereas operation for more than 25 percent
of the time would be "continuous." The problem created by such a definition is that if a pump were to
operate for 3 continuous months of each year at flows below some critical value, and the remaining 9
months at flows well above this value, one would find it difficult to accept the idea that the pump is
operating "intermittently" at subcritical flows for these 3 months. It would be equally difficult to assume
that a pump is operating "intermittently" at subcritical flows if this were to take place regularly for 6
consecutive hours out of each 24 hour day.
We can clarify this situation somewhat if we consider that there are several separate end-results
created by operation below some critical flow:
584 Pump Operation at Off-Design Conditions

1. Excessive radial thrust, leading to shaft or bearing failure


2. Excessive temperature rise
3. In the case of internal recirculation, hydraulic pulsations and mechanical vibrations, leading to possible
mechanical failure of the pump components, including that of the impeller, of the bearings and/or of the
seals (these failures can be either progressive or instantaneous)
4. Still in the case of internal recirculation, cavitation type damage to the impeller (this will always be progressive)
5. Overload of drivers in the case of high-specific-speed pumps
6. Air- or vapor-binding in the case of pumps handling liquids with appreciable amounts of entrained or
dissolved air or gas.

These effects must be examined separately when sizing the minimum flow bypass. Minimum flows
dictated by effects that can lead to very rapid failure or deterioration of performance must be assumed
to be discrete numbers, and operation below these flows should never be permitted. On the other hand,
operation below the minimum flow dictated by cavitation type damage to the impeller need not cause
an immediate destruction of the impeller, but will shorten impeller life. Therefore operation below this
flow can be permitted to occur intermittently, but not continuously.
Another approach to the question of "continuous" versus "intermittent" operation, one that is still
evolving, is the concept of applying what can be termed "life factors." Figure 22.29 shows service life
(period between the need to open the casing and renew running clearances or replace major parts),
expressed as a fraction of that at BEP, versus percentage of BEP flow for 4 classes of pumps. This is
a tentative chart based on the experience of the authors and their colleagues. Its objective is to allow a

100

90

80

70

60
LIFE CD 1 & 2 STAGE API PUMPS
J MAX :s; 30 KW (40 HP)
50
® 1 & 2 STAGE API PUMPS
:s; 450 KW (600 HP)
40
@ MULTISTAGE PUMPS
:s; 3000 KW (4000 HP)
30
@ MULTISTAGE PUMPS
20
:s; 18,500 KW (25,000 HP)
SINGLE STAGE PUMPS
:s; 15,000 KW (20,000 HP)
10

0
0 10 20 30 40 50 60 70 eo 90 100

F\..OW - J BEP

Fig. 22.29 Tentative chart of pump life factor versus capacity for 4 classes of pump.
Pump life is the period between having to renew internal running clearances.
Pump Operation at Off-Design Conditions S8S

pump user to judge more accurately than could be done in the past, the effect on total pump life of
operation at various off-design flows. As an example, consider a 15,000 kW (20,000 hp) boiler feed
pump that spends, on average, 40 percent of its running time at 100 percent of BEP, 25 percent at 60,
20 percent at 40, and 15 percent at 25, the minimum bypass flow. Drawing from Fig. 22.29, the following
estimate of total life can be made.

Operating Flow % BEP Time at Fraction of Total Service Life % Maximum Time x Life

100 0.40 100 40


60 0.25 81 20
40 0.20 51 10
25 0.15 12 2
Total 1.00 72

By this estimate, the boiler feed pump should achieve about 72 percent of its expected life if operated
over the load profile assumed. Beyond the utility of this approach in estimating the effect of various
load profiles, it gives the plant engineer a numerical argument for changing operating practices to raise
expected pump life, the desired change in this example being to not run two pumps when one will carry
the load.

BIBLIOGRAPHY

[22.1] J. A. Smith. "Notes on Some Experimental Researches on Internal Flow in Centrifugal Pumps and Allied
Machines." Engineering 74 (December 5, 1902):763.
[22.2] A. H. Gibson. Hydraulics & Its Application. 1st edition. Constable, 1908.
[22.3] C. W. Steward. "Investigation of Centrifugal Pumps." Bulletin of the University of Wisconsin, no. 318,
Engineering Series, vol. 5, no. 3, (1909):308-310.
[22.4] A. Carrard. "On Calculations for Centrifugal Wheels." La Technique Modern. T.x.U. no. 3 (February 1, 1923).
[22.5] K. Fischer and D. Thoma. "Investigation of the Flow Conditions in a Centrifugal Pump." ASME Transactions.
HYD-54-8, 1932.
[22.6] A. J. Stepanoff. Centrifugal and Axial Flow Pumps. 1st edition, New York: Wiley, 1948.
[22.7] S. Minami, et al. "Experimental Study on Cavitation in Centrifugal Impellers." Bulletin ASME 3, no. 9
(September 1960).
[22.8] Polikovskiy and Levin. The Operation of Pumps and Air-Blowing Machines at Lower Feeding Conditions.
Tep Loenergetika, 1965.
[22.9] F. Schweiger. "Flow in Centrifugal Pumps working at Part Capacity." NEL Report no. 245, 1966.
[22.10] R. C. Worster. "An investigation of the Flow in Centrifugal Pumps at Low Deliveries." BHRA RR770, 1963.
[22.11] W. H. Fraser. Worthington Pump, Private Communication, 1972.
[22.12] Bush, Fraser, and Karassik. "Coping with Pump Progress: The Sources and Solutions of Centrifugal Pump
Pulsations, Surges and Vibrations. Pump World (Summer 1975 and March 1976).
[22.13] W. H. Fraser. "Recirculation in Centrifugal Pumps." Proceedings ASME Winter Annual Meeting. Novem-
ber 1981.
[22.14] W. H. Fraser. "Flow Recirculation in Centrifugal Pumps." Proceedings Texas A&M 10th Annual Turboma-
chinery Symposium. December 1981.
S86 Pump Operation at Off-Design Conditions

[22.15] M. Oshima. "Inlet Flow & Aspects of Cavitation in Centrifugal Impellers." Proceedings Texas A&M 11th
Annual Turbomachinery Symposium. December 1982.
[22.16] C. Bosman and D. Bhrabian. "Calculation of Stalled Flow in a Centrifugal Impeller." Institute of Mechanical
Engineers, Computational Methods in Turbomachinery, Paper no. 667/84, April 1984.
[22.17] R. Palgrave. "Operation of Centrifugal Pumps at Partial Capacity." Proceedings of BPMA 9th International
Technical Conference. Coventry, England, April 1985.
[22.18] S. Gopalakrishnan. "Minimum Flow Criteria." Proceedings of the Pacific Energy Association Workshop.
October 1986.
Pump Operation at Off-Design Conditions 587

Appendix!

DISCHARGE RECIRCULATION:

1. Calculate discharge vector diagram angle ~2:

sin~2 = 1tX ;2
2X
B refer to Figs. Al and A2
2

2. Determine Cm.JU2 from Fig. A3


3. The flow in m3Jhr at discharge recirculation is equal to

(1)

and in gpm to

(1)

D2
Single
Suction
F2 = B~ X W2 X No. Vanes Fz = B2 X W z x No. Vanes
FI = BI X W 1 X No. Vanes x 2 FI = BI X WI X No. Vanes

Fig. Ai

IAppendix is extracted from ref. [22.13].


588 Pump Operation at Off-Design Conditions

~~j3; ·
SIn (l
~2 -
_ C~2
-
_
-
F2
--=-=---:=-
Uz >" w,
13,
~ nx~x~ U
Inlet Vector Diagram '
S. Il _ V. _ 1.273 F,
In 1-" - W, - D~-h~

Fig. A2

0.14

0.12
/
/
/
/
0.10 /
/
0.08
/
/
0.06 /
~
/
0.04 /
/

0.02
10 15 20 25 30
DISCHARGE DIAGRAM ANGLE JJ2 - DEGREES

Fig. A3 Discharge diagram angle B2 in degrees


Pump Operation at Off-Design Conditions 589

SUCTION RECIRCULATION

D
Case 1 where D: < 0.5

1. Calculate inlet vector diagram angle ~I:

.
SIn R
1-'1 = 1.273F 1 fi F·
D2 h2 re er to IgS.
1- 1
Al and A2

2. Determine VelU I from Fig. A4


3. The flow in m3/hr at suction recirculation is equal to

(2)

0.32

0.30

0.28 I
0.26 /
0.24 /
V
0.22
Ve
U,
0.20
V
0.18 /
0.16
..........::; ~ ~\O~
0.14 L ~~
0.12
II"
/
0.10 /
/
0.08
10 15 20 25 30 35 40
INLET DIAGRAM ANGLE PI - DEGREES

Fig. A4 Inlet diagram angle BI in degrees


590 Pump Operation at Off-Design Conditions

Case 2 where D1/Dz > 0.5


Whenever DdD2 is equal to or greater than 0.5, the suction recirculation is either equation I or 2,
whichever is greater.

Nomenclature

B2 =width of the impeller waterway at the discharge diameter mm (in)


Cm2 = meridional velocity at the impeller discharge m/s (ft/s)
DI =eye diameter of the impeller mm (in)
D2 =discharge diameter of the impeller mm (in)
FI = area between the vanes at the impeller inlet normal to the average meridional velocity
mm2 (in2)
F2 =area between the vanes at the impeller discharge normal to the average meridional velocity mm 2 (in2)
hI =shaft diameter through the impeller eye mm (in)
VI = peripheral velocity of the impeller eye m/s (ft/s)
V 2 = peripheral velocity of the impeller discharge diameter m/s (ft/s)
Ve =axial fluid velocity in the impeller eye m/s (ft/s)
~2 = discharge vector diagram angle-degrees (see Fig. A2)
~I = inlet vector diagram angle-degrees (see Fig. A2) m (ft)
NPSH =net positive suction head.

EXAMPLES

Example 1
Double suction pump

Speed 1750 rpm


Flow 792 m3/hr (3500 gpm)
Head 67.06 m (220 ft)
NPSH 5.18 m (17 ft)
Ns 2,100 (1812)
SS(Y2 cap) 10,140 (8744)
D2 393.7 mm (15.5 in)
B2 44.45mm (1.75 in)
F2 19355 mm 2 (30 in2)
DI 182.5 mm (7.185 in)
hI 73 mm (2.875 in)
FI 19355 mm2 (30 in2)

Sin ~2 = 7tX ;2
2X
B = 0.352 (20.61°)
2

~~ = 0.084 from Fig. A3 (1)

Flow =m1.69
x B2 x rpm Cm2
X 106 x u; =598 m /hr (2636 gpm)
3
Pump Operation at Off-Design Conditions 591

Sin ~I = ~(~~~) = 0.44 (26.1°)


~: = 0.172 from Fig. A4
x rpm Ve
Flow = DI(Dt-hD
6.75 x 106 x U I = 227.5 m /hr(2) = 455 m /hr (2000 gpm)
3 3
(2)

D\D2 = 0.463 is less than 0.5 therefore the flow at suction recirculation is equation (2) or 455 m3/hr
(2000 gpm).

Example 2
Single Suction Pump

Speed 1750 rpm


Flow 397 m3/hr (1750 gpm)
Head 30.48 m (100 ft)
NPSH 5.18 m (17 ft)
Ns 2,685 (2315)
Ss 10,140 (8744)
D2 301.6 mm (11.875 in)
B2 47.6 mm (1.875 in)
F2 12258 mm2 (19 in2)
DI 165.1 mm (6.5 in)
hi 0 (0)
FI 10967 mm2 (17 in2)

Sin ~2 = 1tX ;2
2X
B
2
= 0.271 (15.66°)

~~ = 0.062 from Fig. A3


Flow =m1.69
x B2 x rpm Cm2
X 106 x U2 = 277 m /hr (1223 gpm)
3
(1)

Sin ~I = ID·2;3h~1 = 0.511 (30.66°)


1- I

~: = 0.175 from Fig. A4


FIow = DI(Dt-hD x rpm Vi = 204 3/hr (900 ) (2)
6.75 x IQ6 x UI m gpm

DdD2 = 0.547 exceeds 0.5 therefore the flow at suction recirculation is the greater of equation (1)
and (2) or 277 m3/hr (1223 gpm).
III
CONTROLS, DRIVERS,
and PRIMING
23
Controls

Centrifugal pumps are much simpler to control than either reciprocating or rotary pumps. They owe this
ease of control to the flexibility of their characteristics, which enables them to adapt well to the varying
requirements of the systems in which they can be applied. To understand the fundamentals of centrifugal
pump control, however, it is necessary to have a clear conception of the relationship between the pump
performance characteristics, the characteristics of the system in which the pump operates, and the
operating conditions of head and capacity.

PUMP AND SYSTEM CHARACTERISTICS

As shown in Chapter 20, a centrifugal pump installation consists of the pump and the system in which
the pump is to operate. The characteristics of the system are determined by varying conditions of flow
and the resulting heads, whereas those of the pump indicate the ability to produce flow in the system.
The superimposition of the head-capacity curve of the pump over the system-head curve will indicate
the operating conditions in the system, since a pump will always operate at those conditions of head
and capacity that correspond to the intersection of the two curves.

FUNDAMENTAL CONTROL FUNCTIONS

In most installations, pumps are operated continuously in a system for relatively long periods. Often,
the only interruption of operation occurs as a result of some mechanical defect in the system or the
pump itself, or because of a scheduled examination or overhaul. Nevertheless, a certain amount of
flexibility in operating conditions always exists, and sometimes rather extreme variations may be encoun-
tered. The fundamental functions of centrifugal pump controls are, therefore, directed toward permitting
the pump to meet the required variations in operating conditions, including the complete absence of
delivery demand.
For simplification, centrifugal pump controls can be divided into two main groups according to function:

595

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
S96 Controls

1. Controls that completely interrupt flow or cause it to resume


2. Controls that vary the operating conditions-either the pump capacity, the total head, or both.

INTERRUPTION CONTROLS

The reasons for starting or stopping pump delivery will vary markedly; the following list must therefore
be considered representative, rather than exhaustive. Pump delivery may be stopped if

1. The source of supply has been drained either completely or to the desired level, as (generally) with sump
pumps, mine-gathering pumps, or tank-car emptying pumps.
2. The vessel into which the pump is delivering has been filled to the desired level. A typical example is the
filling of a reservoir tank for service or for drinking water supply.
3. The pressure of the system into which the pump is delivering has reached its required magnitude, typical
of hydraulic systems with dead-weight or air-bottle accumulators. However, pump design capacity and
accumulator size generally should be selected to assure that maximum and average demand are such that
the pump capacity remains within reasonably narrow limits in the neighborhood of its design value.
4. A batch process served by the pump has been completed. A typical example is the use of centrifugal pumps
in a steel rolling mill descaling system, in which delivery through the descaling nozzles is interrupted between
consecutive steel sheets.
5. A system is served by a battery of several pumps operating in parallel, and the required rate of flow has
decreased to a point where the delivery from one or more of the pumps can be reduced to zero, to permit
the remaining pumps to operate at a more economical point of their characteristic curve (as described in
Chap. 20).
6. Mechanical trouble has occurred in the pump, its driver, or in some part of the system-overloading of a
driver, overheating of an electric motor or of the pump, overspeeding of a steam turbine, burst piping in
the system, and the like.

Pump delivery must be resumed when the conditions described above are reversed, as in an increase in
supply level, reduction of delivery level, reduction of system pressure, or resumption of the batch process.
Controls that interrupt flow or cause it to resume can be subdivided into two separate classes: (1)
controls to start or stop the driver and (2) controls to open or close a valve in the path of the flow. Each
of these two classes of controls can be further divided into source of impulse and controlling medium.
These subdivisions are related to the primary function of the controls, as the impulse source is created
by the reason that dictates interruption or resumption of flow. Drivers can be started or stopped by
various controls, as discussed below.

Pushbutton Controls for Electric Motors


Pushbutton controls are widely used because this method requires only that the operator be informed
of the necessity for flow interruption or resumption.

Float Switches
Float switches can be applied to maintain certain predetermined maximum or minimum levels in
tanks or reservoirs into which liquids are discharged, or from which liquids are removed by electric-
motor-driven centrifugal pumps. In such an application, a float switch makes an electrical contact when
the reservoir level reaches a predetermined height, and the pump motor starts. When the reservoir level
reaches the other predetermined height, the contact is broken, and the pump motor stops. The same
Controls 597

function can also be performed by level-sensing electrodes connected through relays that will make or
break electrical circuits.

Pressure Switches
Pressure switches are similar in principle to float switches but are operated by changes in pressure
and are used to maintain pressure or vacuum, within certain selected limits, in a system or in a closed
tank. Generally, the pressure to be controlled is applied to a diaphragm actuating the switch. When the
system or tank pressure reaches a predetermined value, the diaphragm pressure is changed, an electrical
contact is made, and the motor starts. When the pressure reaches the other predetermined value, the
contact breaks and the motor stops.
Pressure switches can sometimes be used interchangeably with float switches, as they can be made
to maintain reservoir or tank levels. One method of employing pressure switches uses the static head of
the fluid in the reservoir as a source of varying pressure. In another method, a quantity of air is trapped
in a closed tank; the pressure of this air varies directly with the liquid level.

Power Consumption Controls


A power consumption control can be used to stop a unit when the driver has been overloaded for
any number of causes. This type of control can also be applied to stop the unit when the pump has lost
its prime and is operating dry. In this case, an inverse-current relay will stop an electric motor as soon
as the input current (that is, the power consumption) has dropped to 50 percent of the normal operating
minimum value.

Thermostat Control
Thermostat controls are used in a number of centrifugal pump applications, especially in process
industries, in which certain definite temperatures are to be reached or maintained by means of heat
transfer, recirculation, or similar method. In such cases, a thermostat control can be applied in the same
manner as a float or a pressure switch; however, temperature levels are substituted for either static levels
or pressure differences.

Overspeed Governors
Overspeed governors are applied mainly to steam turbine drivers, although they can also control
internal combustion engines or water turbines used to drive centrifugal pumps. Governors are intended
to protect the driver, the pump, and the system in which the pump is operating. The pump and the
system could be damaged if the driver possesses sufficient power to drive the unit at high enough speeds
to cause excessive pressures. A typical use is that of the overspeed governor that acts on the inlet valve
of a steam turbine and is generally set to trip out the turbine, should the operating speed exceed 5 to
10 percent more than the design operating speed.

Complete Interruption of Flow Without Stopping Unit


Many installations require interruption and resumption of flow without stopping the unit. This may
be necessary because variations in demand are too frequent to warrant bringing the unit to a stop; because
starting the driver is too lengthy and complex a process, once the unit has stopped; or because interruption
of flow by stopping the driver does not prevent possible injurious effects, such as reverse flow through
the pump. Flow can be interrupted or resumed by positioning a valve without affecting driver operation
as follows:
598 Controls

1. Discharge valve manual controls-Some installations require that the unit be operating at all times, as a
safety measure, even though no demand exists for flow delivery. A gate, globe, or plug valve located in the
pump discharge can be manually adjusted to cut off the flow entirely. This is a frequent arrangement when
several units are operating in parallel and when one or more of the units is installed for standby duty. This
form of control requires a bypass to prevent the pump from overheating when it is operating at shutoff. In
many cases, especially with across-the-line electric motor starters, the unit can be put on the line more
quickly by starting the motor than by manipulating a valve in the discharge. Therefore, when the time
element is important, it is preferable to start and stop the motor, rather than manipulate the discharge valves,
which, if performed too quickly, may cause water hammer.
2. Discharge line check valves-Check valves are almost universally used in centrifugal pump discharge lines.
A check valve acts to protect the pump, its driver, and the suction portion of the system against possible
damage caused by reverse flow, if the external head in the discharge system exceeds the head generated by
the pump, or if the pump comes to a standstill and develops no head whatsoever. Reverse flow through a
pump would cause it to operate as a water turbine running opposite to the normal direction of rotation;
therefore, if the driver is not suitable for this reverse operation, it could become seriously damaged. Even
if the pump or its driver were not harmed, excessive pressures could be imposed on the suction piping and
fittings. A check valve is fully automatic in its operation, closing when adverse conditions occur and reopening
immediately on the resumption of normal conditions.
3. Float control valves-Variation in suction or discharge levels can be transmitted to a mechanically or
electrically operated valve in the pump discharge so as to fulfill basically the same function as a float switch.
Generally, this type of valve is used for throttling action, but it can be provided with a mechanism to open
or close the valve completely.
4. Pressure control valves-Pressure control valves bear the same relation to pressure switches as float control
valves bear to float switches. One specific application, however, differs somewhat in its fundamental function,
that is, the use of pressure control valves as relief valves. Many installations require relief valves as protection
against either the building up of excessive pressure through operation at very reduced flows, or the overloading
of a driver in the same range of operating conditions when the power consumption curve rises sharply with
the head developed.
5. Thermostat control valves-Thermostat control valves utilize temperature variations to open or close discharge
valves instead of operating on the driver itself.

The last three control applications are relatively rare, except where they perform the added function
of modifying the pump operating conditions over a wide range and where they, incidentally, are permitted
to travel to the full closed position if externally imposed conditions require it. This phase of their
operation is discussed in detail in Chapter 22 and 28. It is important that a centrifugal pump should
never be permitted to operate against a completely closed discharge and that a bypass recirculation line
be provided whenever a valve positioning control may fully interrupts flow delivery.

VARIATION OF OPERATING CONDITIONS

As many reasons exist that may require a variation in the operating conditions as those that may require
interruption or resumption of flow. A few possible reasons are

1. The need to maintain a constant level, either in the suction or discharge reservoir, regardless of the variation
of supply or demand. A typical example of this requirement is the boiler feedwater system, where a constant
boiler level must be maintained, regardless of the fluctuation of the load imposed on the boiler.
2. The need to maintain a constant rate of flow with varying differences in static elevation between the source
of supply and the ultimate delivery. Such conditions occur most frequently when a centrifugal pump takes
its suction from a body of water subject to tidal variations.
Controls 599

3. The need to maintain a constant pressure in a system, or a constant pressure margin over some specific
pressure, regardless of the total flow into the system.
4. The need to maintain a constant temperature in a batch process by means of cooling or heating through heat
exchangers, regardless of temperature variations, in the cooling or heating medium.

Since the operating conditions of a centrifugal pump are established by the intersection of the head-
capacity curve at operating speed and the system-head curve, as described previously, it follows that
variation of the operating conditions can be obtained by either one of two methods (1) mpdification of
the pump head-capacity curve, or (2) modification of the system-head curve.

MODIFICATION BY SPEED VARIATION

The pump head-capacity curve is modified by variations in the pump speed except in some very rare
applications where the head-capacity curve is modified through mechanical changes of the interrelation
of internal pump parts while the pump is in operation. Pump speed can be varied either by modifying
the speed of the pump driver itself or by applying a variable speed power transmission mechanism, such
as a hydraulic coupling or a magnetic drive between the pump and its driver. The controlling mechanism
itself is generally unaffected by the choice between these two forms of speed change. This choice is
generally dictated by the choice of the type of driver best suited for the installation and the relative
economy of the several possible driver combinations. If electric motors are used, variable speed operation
can be obtained with dc motors or ac slip-ring-wound rotor motors. Squirrel-cage induction motors can
be operated at variable speeds through frequency variation. Steam turbines and internal combustion
engines lend themselves most conveniently to speed variation and, therefore, are seldom connected to
the driven pump by means of variable speed transmission mechanisms.

Methods of Controlling Speed


Pump operating speeds can be controlled by means of

1. Manual control-Manual control may require a rheostat for motor-driven units, or a throttle valve control
for steam turbines and internal combustion engines.
2. Flow control-Flow control is generally used when it is necessary to maintain a constant flow despite
variations in the system-head curve. The control speeds up the pump whenever the intersection between the
head-capacity and system-head curves falls below the desired value and slows down the pump speed when
the intersection rises above the desired value.
3. Float or level control-Float or level control principles of operation are similar to float switches, except
that instead of a start-and-stop action, they perform the continuous function of slowing down and speeding
up the pump to maintain the desired level.
4. Pressure control-Pressure control is exemplified by constant pressure steam turbine regulators that maintain
whatever operating speed is necessary to produce a constant discharge pressure, regardless of flow variations.
5. Temperature control-Temperature control functions to increase or decrease the pump delivery to maintain
a constant rate of heat exchange, despite temperature variations between the cooling medium and the liquid
to be cooled.

MODIFICATION BY VALVE POSITIONING

If it is impossible or impractical to change the pump operating speed, it is necessary to alter the shape
of the system-head curve, by varying one of its components, in order to alter the operating conditions.
Controls

HEAD-CAPACITY CHARA I D IC
OF PUMP CTER/ST/C

DESIGN HEAD

TERMINAL PRESSURE
PLUS STATIC ELEVATION

o SYSTEM HEAD WITH


<t THROTTLING VALVE
w FULLY OPEN
:::c >-1
1-.

~
~I
a

CAPACITY

Fig. 23.1 Effect of throttling valve used to vary pump capacity.

In this case, neither the tenninal pressure nor the static level difference can be expected to pennit
variation, and the only possible solution lies in varying the friction head loss in some part of the system.
An artificial source of friction loss, such as a valve, must be introduced. This friction loss is controllable
through valve positioning. This positioning, in turn, is controlled by the quantity (capacity, pressure,
level, or temperature) to be controlled. Figure 23.1 illustrates the operation of a centrifugal pump at
constant speed and the effect of a throttling valve used to vary pump capacity. The design capacity is
obtained with the throttling valve fully open, while curves A to D represent the system-head curves
corresponding to several positions of the valve, as it is closed. Therefore, the pump operating conditions
will correspond to the intersection of the head-capacity curve with curves A to D, or any intennediate
curve, depending on the required capacity and consequently on the position of the throttling valve.

Methods of Valve Positioning


The positioning of the throttle valve can be controlled by

1. Manual control-Manual control is the most widespread application and is used when operating conditions
are expected to require infrequent changes. The pump operator progressively closes or opens a gate or globe
valve in the pump discharge. Throttling the pump suction is seldom recommended as equally satisfactory
or better results can be obtained by controlling the discharge without running the risk of pump cavitation.
2. Flow control-The throttling valve may be automatically operated to maintain the flow at its desired
magnitude when the static head is reduced. An uncontrolled system-head curve would intersect the pump
head-capacity curve at an increased capacity (Fig. 23.2).
3. Float or level control-Float or level control operation is similar to that directed at varying pump speed,
except that the level is maintained by artificially altering the system-head curve. The boiler feedwater
Controls 601

o
<t
LU
::x:

CAPACITY

Fig. 23.2 Effect of throttling valve automatically positioned to maintain required flow under reduced static head.

regulator is a typical example of this application. Its operation is exactly as shown in Fig. 23.1 with the
terminal pressure being equivalent to the boiler pressure.
4. Pressure control-The pressure regulator operates in a manner similar to that of the flow or level regulator,
except that a constant discharge pressure at the pump is maintained by throttling off the excess pressure as
capacity demand varies. Conversely, the valve opens up if the pressure falls below its desired value.
5. Temperature control-The most common application of temperature control is in cooling or refrigerating
systems, when the type of driver does not permit speed variation.

Although these controls are intended to provide several dissimilar services, such as starting and
stopping the driver, modifying the pump head-capacity curve, or modifying the system-head curve, they
are all based on a group of similar impulse sources: flow, level, pressure, or temperature. This similarity
is illustrated in Table 23.1, which shows an analysis of the fundamental control functions.

ANALYSIS OF CONTROL ELEMENTS

The description of the fundamental functions of centrifugal pump controls leads to a logical subdivision
of controls into two distinct classifications, namely, corrective and protective controls. These two classifi-
cations are practically self-defining.
602 Controls

Table 23.1 Fundamental Control Functions

Pushbutton on motor
Float switch
Start and stop Pressure switch
of driver Power consumption
Thermostat
Interruption and Overspeed governor
resumption of flow
Discharge valve manual control
Check valve
Valve {
Float or level control
positioning Pressure control
Thermostat

Manual control
Driver Flow control
{
speed Float or level control
Modification of pump control Pressure control
head-capacity Temperature control
curve by speed
Modification variation Power Flow control
of operating trans- { Float or level control
conditions mission Pressure control
control Temperature control

Manual control
Modification of system- Flow control
{
head curve by Float or level control
valve positioning Pressure control
Temperature control

Corrective controls are obviously directed at changing some relationship within the pump system to
compensate for changes in the conditions imposed upon the system. For example, when the boiler water
level, which must be maintained constant, approaches the desired height, the delivery of the boiler feed
pump must be reduced.
Protective controls, on the other hand, are directed at protecting the pump or the system against
certain harmful combinations of conditions or, in the event such conditions have arisen, at eliminating
them in the shortest possible time. Thus, the bypass control in a boiler feed pump discharge, intended
to prevent operation of the pump at flows so reduced that overheating or accelerated wear will occur,
is a protective control. This classification is important because a great many features of pump control
will depend on whether corrective or protective controls are involved. Of course, once in a while, the
same control can be applied for either corrective or protective purposes.
For example, if a centrifugal pump is to be stopped when sufficient water has been delivered to a
reservoir, the float switch stopping the pump is a corrective control, as no specific harm will be done
if the pump continues to run a little while longer. If, however, the same float switch is intended to
prevent pump operation beyond the point where the water in the suction vessel has been completely
drained, so that the pump will not run dry, the control is a protective one.
A mechanism or a process is easier to understand when it has been separated into its component parts
Controls 603

before it is considered as a whole. This holds true for centrifugal pump controls, which will be broken
down and discussed in their functional parts, before being combined again. Pump controls can be logically
separated into four individual functional parts:

1. Measuring element
2. Impulse element
3. Relay element
4. Power element.

Measuring Element
All control problems encountered in the operation of centrifugal pumps can be reduced to the problem
of balancing flows, pressures, temperatures, or combinations of two or more of these. As a result, the
measuring element must determine some force or forces set up in the pumping system that change in
magnitude as the quantity to be controlled varies. If this quantity is fluid flow, the force can be created
by a pressure differential across an orifice. If liquid levels are to be controlled, the difference in static
head between the level to be controlled and some fixed arbitrary reference level provides this force. If
a certain pressure or pressure difference is to be controlled, this pressure in itself provides the force to
be measured. The main concern is to select the measuring element and its location in the pumping cycle
so that it gives the simplest and most reliable indication of the quantities to be balanced.

Impulse Element
The impulse element is that portion of the control that does most of the thinking, deciding when the
measured variable has reached a predetermined value, or when it is properly balanced against some
other variable with which it must remain in a certain correlation.
If the control function is such that variation in the quantities to be balanced is to be constantly
accompanied by the desired change in valve setting, pump speed, or similar changes, the impulse and
measuring elements are integrated. For example, if the pump capacity is to be reduced by throttling in
some proportion to the reduction in the supply reservoir level at the pump suction, the force set up in
the measuring element by the magnitude of the level acts directly as the impulse, which will ultimately
be transmitted to the throttling valve control.
If control of operation and change in relationship is to occur only when the measured quantity reaches
some predetermined magnitude, the impulse function becomes divorced from the measuring element
and becomes operative only when the forces involved are balanced at the desired value. As an illustration,
it may be assumed that a valve in the pump discharge is to be closed off whenever the suction level
drops to some set value. The measuring element records the suction level at all times and balances a
force proportional to this level against a selected force representing the desired minimum. As long as
the measured force exceeds the predetermined force, the impulse is inoperative. On reaching a balance
between the two forces, the impulse mechanism sets certain controls in motion to close the valve as desired.
Amplification of forces. Sometimes the force set up through quantity variation is very small and
cannot be used directly to actuate any control mechanism. An amplifying feature must then be incorporated
into the control mechanism. This amplifying feature, however, might be more logically considered as
part of the relay than of the impulse element.

Relay Element
The relay of the impulse mayor may not exist as a mechanism, depending on whether this portion
of the pump control is automatic or not. In other words, if the function of the impulse element is to ring
604 Controls

an alann bell and warn an operator when certain conditions are detected by the measuring element, the
function of the relay element will be accomplished when the operator walks over to the valve the alann
has indicated must be throttled. However, in the majority of control mechanisms, the function of the
relay element is to transmit the decision reached by the impulse element to the power mechanism which
actually operates the control. This function can be discharged in many ways and can be performed
mechanically, hydraulically, pneumatically, or electrically.

Power Element
The power element, the last of the four functional parts of a complete pump control, is that element
which changes some relationship in the combination of pump and pumping system, such as a change
in valve setting, a reduction or increase in pump speed, or stopping or starting the pump. The source of
the power used in this element may be either mechanical, hydraulic, pneumatic, or electrical.

HOW FOUR CONTROL ELEMENTS ARE INTERRELATED IN PRACTICE

It has been assumed so far, in the interest of simplicity, that the four elements of a centrifugal pump
control can easily be separated and examined. Actually, the four elements and their functions are
frequently interlaced, introducing an element of uncertainty into the analysis. For example, consider a
constant pressure pump regulator installed in the discharge line of a centrifugal pump, as shown in Fig.
23.3. The pump is operated at constant speed, and, therefore, the discharge pressure will vary with the
quantity delivered by the pump. Assume further that the pump capacity is varied by some other control
not pertinent to the present analysis. In the valve itself, the valve disk is mounted on the valve stem,
the position of which is controlled by a diaphragm. The control chamber above the diaphragm is connected
to the pump discharge line at point B, where a constant pressure should be maintained. The balancing
force, corresponding to the desired discharge pressure, is provided by a spring. Any increase in the pump
discharge pressure at B will act against the force exerted by the spring and cause gradual valve throttling.
The valve will become partially closed, increasing the frictional losses through it, and, therefore, increasing
the difference between the pressures at A and at B, until the pressure at B corresponds to the desired
value and balance has been reestablished. The pump performance is illustrated in Fig. 23.4.
In this particular case, the measuring element is the diaphragm, which reacts to variations in the

CONSTANT PRESSURE
REGULATOR

• A •
QUANTITY VARIED
BY CONTROL AT END
CENTRIFUGAL PUMP OF DISCHARGE LINE

Fig. 23.3 Constant-pressure regulator installed in pump discharge.


Controls 605

THROTTLED BY
CONSTANT PRESSURE
REGULATOR

PRESSURE LOSS
THROUGH
CONSTANT
PRESSURE
REGULATOR
I&J WIDE OPEN
a::
:>
(f)
(f)
I&J PRESSURE
a:: AT B
n..

CAPACITY

Fig. 23.4 Perfonnance of constant-pressure pump in Fig. 23.3.

pressure to be controlled. The impulse is provided by the spring, which is calibrated to correspond to
the desired pressure at B. The relay does not exist in this application, while the power element is, in
the last analysis, a variable one. When the pressure at point B exceeds the desired value, the force
required to close the valve partially and create additional artificial friction is provided by the controlled
element itself. Conversely, when the pump delivery is increasing and the pressure at point A decreases,
carrying with it a decrease of pressure at point B, the spring intervenes to provide the power required
to reopen the valve. The power element could also be described as consisting of the difference between
the hydraulic and spring forces, respectively. The actuating force is positive for valve closure when the
hydraulic force exceeds the spring force and negative for valve opening, in the reverse case. Regardless
of the interpretation given to this phase of the control analysis, it is apparent that measuring, impulse,
and power elements are definitely interrelated in this particular application.

INFLUENCE OF OPERATING MEDIUM

The same example can serve to illustrate certain differences that arise from the selection of the operating
medium, either for the power or for the impulse element. In the control shown in Fig. 23.3, it was
assumed that the desired discharge pressure at point B was represented by a valve spring. The same
function, however, could have been fulfilled either by a weight or by some present hydraulic or pneumatic
pressure, to be introduced on the side of the diaphragm opposite the chamber where the pressure to be
controlled is applied. While both spring-loading and weight- or pressure-loading have their advantages,
it is necessary to visualize the basic differences between the two types to determine the proper type for
606 Controls

any given application. The important characteristic of spring-loading is that the spring power is not
constant, but varies with compression. As the pump delivery decreases, therefore, the spring is compressed,
exerting more pressure in counteraction to the controlled pressure. Since the discharge pressure at B is
balanced by the spring action, the pressure increases with a decrease in pump capacity, although this is
not shown in Fig. 23.4 to avoid confusion.
IT the maximum movement of the spring is very slight, the valve-loading will be practically constant.
IT, however, the spring is to be operated in a wide compression range, spring-loading is not well suited
to provide constant loading and weight- or pressure-loading is preferred.

AUTOMATIC AND MANUAL CONTROLS

Although a great number of centrifugal pump controls are automatic in their operation, and although
attention is generally directed to such controls, in many instances a choice must be made between manual
and automatic controls. This choice is considerably complicated by factors more psychological than
technical. Automatic controls often lead to neglect and indifference on the part of the operators. Should
a control failure occur, the operator may be unable to correct any resulting harmful effects, either through
lack of knowledge, or simply because he might be so dependent on the proper operation of the control
that he would not be present where his attention becomes necessary. However, human response is never
as rapid as mechanical response and requires an excessively attentive attitude, all of which may lead to
an equal or greater incidence of failures than in the case of automatic operation. The decreased operating
personnel requirements, furthermore, more than overbalance the greater initial financial investment in
the automatic control.
No overall rule or preference can be established in the choice between manual and automatic operation.
First, it is very important to realize that the differentiation between the two types is not as sharp as some
engineers may be led to believe. Between the two extremes of fully manual and fully automatic control,
a large number of intermediate arrangements exists, all of them semiautomatic in some phase of the
control mechanism. Referring to the divisions outlined previously, and separating control mechanism
into measuring element, impulse, relay, and power element or positioning, it appears that each of these
components can be manually or automatically operated. To illustrate this, a typical control operation,
involving the throttling of a centrifugal pump discharge in response to level changes in the reservoir at
the pump discharge, is described in Table 23.2. In addition to the fully manual and fully automatic
controls five other methods of control are shown, in which some of the elements are operated manually
and others automatically. Since there are four separate component control steps and two different methods
of operation, that is manual or automatic, for each step, there will be 16 different possible combinations
altogether. One of these will be fully manual, a second fully automatic, and fourteen semiautomatic. It
is obvious that some of these combinations will be unpractical. For example, it would be illogical to
have automatic operation of all steps but the measuring element. However, the combinations described
in Table 23.2 indicate that a rather complex chain of reasoning is necessary to determine the most
suitable combination and that to base the ultimate decision on a broad preference for manual or automatic
controls is decidedly shortsighted.

FACTORS IN ANALYZING CONTROL METHOD

Some of the most important questions indicating the general character of the analysis that must precede
a choice between manual and automatic controls are presented below, but not necessarily in order of
relative importance.
Table 23.2 Combinations of Manual and Automatic Operations

Fully manual Semiautomatic Semiautomatic Semiautomatic Semiautomatic Semiautomatic Fully automatic

Manual: Manual: Operator Automatic: Direct Automatic: Float in Automatic: Float in Automatic: Direct Automatic: Float in
Operator measures level reading gauge reservoir reservoir reading gauge reservOlr
<)
....
;::l measures in reservoir with glass indicates registers level registers level glass indicates registers level
'"«S
<)
level in rod level in reservoir level in reservoir
:::E reservoir
with rod
Manual: Manual: Operator Manual: Operator Automatic: Float Automatic: Float Manual: Operator Automatic: Float
I
;;.,
Operator realizes that sees that level switch rings bell switch rings bell sees that level switch makes
.$ realizes that level has has increased has increased contact on rise
<)
~ level has increased in reservoir
I increased
Manual: Manual: Operator Manual: Operator Manual: Operator Manual: Operator Manual: Operator Automatic:
<) Operator walks over to walks over to hears bell and hears bell and walks over to Impulse from
"3'"~ walks over to push-button valve walks over to walks over to push-button float switch
valve station valve push-button station transmitted
.5 station electrically
Manual: Automatic: Motor- Manual: Operator Manual: Operator Automatic: Motor- Automatic: Motor- Automatic: Motor-
I Operator operated valve closes valve closes valve operated valve operated valve operated valve
....<) closes valve throttles pump partially and partially and throttles pump throttles pump throttles pump
~ partially and delivery throttles pump
0 throttles pump delivery delivery delivery
j:l..,
throttles delivery delivery
pump
\ delivery

?
~
1:;

§
608 Controls

Frequency of Operation

Is the control operation required infrequently or at frequent intervals, possibly constantly? The starting
up of a standby pump on failure of the main pump in a non-critical transfer service, an operation that
is apt to occur very rarely, is less in need of automatic controls than, for instance, the maintenance of
a constant pressure in the discharge header, when variations in the flow delivery require constant
repositioning of the regulating valve.

Expectancy of Operation

Is the control operation to occur at regularly scheduled intervals or is the requirement sudden and
unexpected? For example, starting up additional boiler feed pumps in a power plant before an expected
increase in load at certain times of the day can readily be performed manually. On the other hand, the
start-and-stop operation of a transfer pump which delivers into a storage tank depends wholly on the
varying rate of usage from the tank and, therefore, cannot be accurately predicted as to time of occurrence.

Urgency of Operation

Must the controlled operation take place instantly on the occurrence of certain conditions, or can it
take place at leisure? Basically, the difference is between corrective and protective controls. In the case
of protective controls, it is almost imperative that the measuring and impulse elements be automatic.
Certain corrective controls must operate almost instantly, as, for example, centrifugal boiler feed pump
controls intended to maintain a constant level in high-pressure boilers. Other corrective controls can rely
on operator judgment and, hence, can be manual.

Difficulty of Operation

Is the application of the control suitable for manual operation or are the forces required to set the
control in motion excessive? Taking two extremes for illustration, it is obvious that while a I-in. valve
can readily be throttled manually, a 24-in. valve can be opened or closed much more easily if it is
provided with automatic electric drive.

Facility of Detection

Is the quantity to be measured subject to accurate determination by human senses or is it impossible


to obtain sufficiently accurate measurements except by mechanical means?

Economics of Personnel Attendance

Is an operator always available where needed, or does the manual operation of the control require
additional personnel not needed otherwise? The best example of this question is the application of
automatic controls to start and stop numerous cellar-draining or flood-protection pumps.
A large number of other questions may arise that apply to individual cases and that must be considered
in the ultimate decision between manual, semiautomatic, and fully automatic control operation. It is a
popular misconception that once automatic operation is provided, all manual control is precluded. Not
only can most automatic controls permit manual operation, but the installation must always be provided
with a means to switch from automatic to manual operation, at least as far as the measuring, impulse,
and relay elements are concerned.
Controls 609

PILOT DEVICES

A large number of automatic controls are electrically operated, so that although the measuring element
may be mechanical or hydraulic, the impulse and relay elements are based on electrical contracts and
transmission or interruption of electrical currents. Such automatic controls are principally directed at the
interruption or resumption of flow, based on level or pressure fluctuation. Because the electric currents
involved in the operation of the pump drivers are generally excessive for the control apparatus, the latter
serve as pilot devices. Thus, electric motors may use 2,300-volt current, whereas control equipment may
use 110-V current to actuate the 2,300-V starters.

TYPICAL CONTROL APPLICATIONS

This chapter is not intended as an exhaustive catalogue of all the forms and types of control equipment
that may be encountered in centrifugal pump applications. Instead, its purpose is directed at understanding
the various functions of pump controls and the effect the operation of these controls may have on the
performance of centrifugal pumps. Therefore, a complete explanation of the functioning of one or two
typical centrifugal pump controls will make the application of other types of control more easily under-
stood.
The boiler feedwater level regulator represents one of the most typical applications in centrifugal
pump practice, that of maintaining a constant level regardless of demand fluctuation.
To understand the problems involved in the maintenance of a constant boiler water level, it is essential
to remember the statement made earlier in this chapter, namely, that a centrifugal pump operates on a
system-head curve. In other words, it operates at a capacity and pressure corresponding to the intersection
of its head-capacity curve with the system-head curve. Therefore, in a given feedwater system, if the
static pressure (boiler drum pressure plus static elevation) and the friction losses remain constant, the
boiler feed pump will deliver a constant capacity into the boiler.
This, however, will not permit the maintenance of a constant boiler water level, inasmuch as the
steam outflow from the boiler is not constant, but depends on the turbine load. It becomes necessary,
therefore, to introduce an artificial source of friction loss, which is controlled by changes in the boiler
drum water level. This artificial source of friction loss is provided by the feedwater regulator. Although
the operation of such a regulator could be handled manually in a few isolated cases, automatic operation
of this control is required in all but a few extremely small installations.
Consequently, the feedwater regulator consists basically of two elements (1) a throttling valve and (2)
a controlling mechanism, which determines the setting of this valve, depending on boiler level fluctuations.
Figure 23.1 is typical of the operation of a centrifugal boiler feed pump at constant speed and illustrates
the effect of the feedwater regulator throttling valve at several different positions corresponding to
various flow requirements of the boiler. It has been assumed in this case that the boiler pressure remains
constant at all loads. Depending on the boiler load and, consequently, on the position of the feedwater
regulator throttle valve, the pump discharge capacity will correspond to the intersection of the head-
capacity curve with curves A to D or with any intermediate curve.
The feedwater regulator represents the simplest but most important of boiler feed pump controls.
Without it, boilers would have to be fed manually by an operator who would watch the boiler gauge
glass and open or close a valve as he saw the level fall or rise. Of course, if the pump were to run at
variable speed, the throttling required from the feedwater regulator would be reduced. But the choice
of a control depends on the degree of refinement desired for the solution of the problem at hand. Variable
speed operation of feed pumps is one such refinement; it is described later. For direct comparison between
610 Controls

the simpler and more complex forms of control, Fig. 23.5 shows a feedwater system where the pumps
are operated at constant speed and where the only control is a feedwater regulator.
Many feedwater regulators are available today. The specific manner in which boiler level fluctuations
are transmitted to the throttling valve, as well as the design of the valve itself, vary with manufacturer.

DIFFERENTIAL OR EXCESS PRESSURE REGULATORS

The constant boiler level feedwater regulator can be compared to an orifice in a pipeline through which
the feedwater flows. The quantity of flow through this orifice depends on its area as well as the pressure
drop across it. As the water level in the boiler drum rises or falls, the regulator valve moves so that the
valve opening will provide the correct rate of feed to the boiler. However, if the pressure drop across
this valve is allowed to vary, the flow will vary even though the valve area remains constant. Although
these variations in flow will eventually result in changing the boiler level, and consequently the area of
valve opening, they will have an undesirable effect on the boiler operation since the flow should vary
only as required for load changes.
Since the discharge pressure of a centrifugal pump operated at constant speed increases from 10 to
20 percent between its design operating condition and extremely low flows, while the required system
head decreases as flow decreases, the excess pressure generated over and above the system head require-

BOILER
FEED
PUMP
SUCTION DISCHARGE

- FEEDWATER
REGULATOR BOILER

'"::>
It:
DROP THROUGH
(I)
(I) FEEDWATER
'"a.
It:
REGULATOR
AT FULL LOAD

CAPACITY

Fig. 23.5 Constant speed boiler feed pump controls.


Feedwater regulator only.
Controls 611

ments increases appreciably at light loads. The excess pressure must be absorbed (throttled) in the
feedwater regulator valve; this results in a wide variation in pressure drop across the valve (see Fig. 23.5).
Another disadvantage of an increasing pressure drop across the feedwater regulator is that since the
valve ports are designed for maximum boiler capacity, the valve will operate through a very small
portion of its travel when the boiler is operating at reduced ratings, with a resulting sacrifice of accuracy
of control.
To eliminate the disadvantages of a varying pressure drop across the feedwater regulator, differential
or excess pressure regulators are used. The function of these regulators is to maintain a more or less
constant pressure drop across the feedwater regulator valve either by

1. throttling the excess pressure generated by the pump (this is used with constant speed pumps) or
2. decreasing the pressure generated by the pump at various loads by varying the pump speed.

A comparison of Figs. 23.5 and 23.6 will show how an excess pressure regulator maintains a uniform
pressure drop across the feedwater regulator valve. Figure 23.6 illustrates the hookup of this refinement
in pump control and the pressure and capacity relations in a feedwater system so controlled.
Basically, a differential pressure regulator consists of a throttling valve located ahead of the feedwater
level regulator valve. The throttling valve is actuated by a metal bellows subject to, and influenced by,

BOILER
FEED
PUMP
SUCTION DISCHARGE

.. EXCESS FEEDWATER
PRESSURE REGULATOR BOILER
REGULATOR

UJ
0::
:::> DROP THROUGH
(f)
(f)
UJ FEEDWATER
0:: REGULATOR
Q.
AT FULL LOAD

>-
zt:
(!)O
u;~
UJe:(
°0

CAPACITY

Fig. 23.6 Constant speed boiler feed pump controls.


Feedwater regulator and excess-pressure regulator.
612 Controls

the differential pressure across the feedwater valve, as the two sides of the bellows are connected to the
upstream and downstream sides of the feedwater valve respectively. As the pressure drop across the
feedwater regulator valve increases, the differential pressure regulator closes until the pressure drop is
again restored to its initial value.
The differential pressure regulator valve should be located as far from the pump and as close to the
feedwater regulator as possible, to minimize the pressure variations caused by the friction losses in the
intervening piping and closed heaters. However, this is not always possible and the valve is often located
almost immediately following the boiler feed pump discharge.
In some cases, instead of maintaining a constant pressure drop across the feedwater regulator, a
constant discharge pressure is maintained at the boiler feed pump. Instead of being actuated by a difference
of pressures across a bellows or a diaphragm, the regulator is influenced by the pump discharge pressure
on one side and a constant pressure set up by a spring or an air-loading mechanism, this constant pressure
corresponding to the desired value at full load.

CONTROL OF VARIABLE SPEED PUMPS

When a centrifugal pump is driven by a variable speed motor or, especially, by a steam turbine, it
becomes possible to apply the differential pressure control to even greater advantage. Figures 23.7 to
23.9 show three different arrangements driven by a steam turbine, all employing a boiler feed pump
with a variable speed governor. In all cases it has been assumed that the boiler drum pressure is not
constant but increases with the load to maintain a constant steam pressure at the superheater outlet. The
increase in drum pressure compensates for increased superheater losses at increasing steam flows.

DISCHARGE

BOILER

CONSTANT-PRESSURE
VPUMP GOVERNOR
STEAM SUPPLY

o
<{
UJ
J:

CAPACITY

Fig.23.7 Variable speed boiler feed pump controls.


Constant discharge pressure.
Controls 613

PUMP I/-- FEEDWATER :


I PILOT LINES REGULATOR I
BOILER

: EXCESS-PRESS~RE :
REGULATOR 1 ' " - - - -.......
I
---------- sfiAM SUPPLY

FEEDWATER
REGULATOR
DROP AT
FULL LOAD

EXCESS PRESSURE
I---~_-- CONSTANT AT ALL
LOADS

MAXIMUM DRUM PRESSURE

Fig.23.8 Variable speed boiler feed pump controls.


Constant excess over boiler pressure.

The three arrangements differ only in the location of the pilot lines leading to the governor. Figure
23.7 illustrates a constant pressure governor applied to the steam turbine driving the boiler feed pump.
Only one pilot line is provided from the discharge of the boiler feed pump, and the discharge pressure
remains constant at all loads.
In Fig. 23.8, which shows a constant excess pressure governor, the pilot lines lead to the pump
discharge and also to the boiler. The pump speed is varied in such a manner that the difference between
the pump discharge pressure and the boiler drum pressure remains constant at all loads. The feedwater
regulator pressure drop is not constant but, instead, throttles the difference between the pump discharge
pressure at variable speed and the pressure representing the system-head curve, less the feedwater
regulator drop.
In Fig. 23.9 the pilot lines lead to the two sides of the feedwater regulator valve, and the pump speed
varies to maintain a constant pressure drop across this valve.
All these controls can be arranged to operate separate valves in the steam line ahead of the turbine
governor valve or to operate the turbine governor valve itself.
Similar controls can be applied to vary the speed of electric-motor-driven centrifugal pumps, if these
are slip-ring-wound rotor motors, rare today because of their greater cost, or induction motors equipped
with variable frequency drive. Another common solution today is the use of hydraulic couplings or of
magnetic drives. The control of these variable speed devices is similar to that of variable speed turbines,
in that the same impulse elements are used to transmit the necessary signal to the particular speed-
varying mechanisms of the drives.
Although the control of centrifugal boiler feed pumps cannot be termed as typical of every kind of
pump control that may be encountered, certain similarities will occur in most cases. Therefore, an
614 Controls

PUMP FEEDWATER REGULATOR

SUCTION DISCHARGE

BOILER

TURBINE

PUMP GOVERNOR STEAM SUPPLY

PRESSURE DROP ACROSS


FEEDWATER REGULATOR
CONSTANT AT ALL LOADS

CAPACITY

Fig. 23.9 Variable speed boiler feed pump controls.


Constant pressure drop across Jeedwater regulator.

understanding of what may be accomplished in this particular field of application will prove helpful in
solving most pump control problems.
The controls described here, however, are basically corrective controls. Protective controls are charac-
terized by certain problems and certain features, which make them considerably different.

PROTECTIVE CONTROLS

By their very nature, protective controls are automatic in their operation, since their main purpose is to
complement the work of the operating personnel and perform functions that would normally be performed
manually, if the personnel were on hand and aware of the need. Most protective controls are Intended
to bring the pump and its driver to a stop whenever certain harmful operating conditions arise. However,
a few of these controls cause some alteration in the operating conditions instead.
In general, the impulse for the operation of the protective control is selected from the conditions to
be avoided or eliminated. A typical case of a protective control used to bring a unit to a stop is that of
a check valve provided with a switch which can control the electric motor driving the pump. Such a
check valve is intended to protect the pump against damage caused by loss of water during operation.
While the pump is discharging water, the check valve flap is in a raised position and the switch is held
Controls 615

closed. If for any reason the flow ceases (due to lack of water at die suction, for instance), the valve
flap falls and opens the switch, stopping the pump automatically before it is damaged by running dry.
Since under certain conditions a centrifugal pump may be permitted to operate against a closed valve
in its discharge, the simple check valve is a protective control device by itself, as it prevents reverse
flow through the pump whenever the pump discharge pressure falls below the pressure in the header
into which the pump is discharging. However, it must be remembered that a pump operating against
shutoff will become damaged through overheating in most cases, and, therefore, every check valve
installation must be checked from this point of view, to assure that it will not operate at shutoff for any
length of time. Wherever danger of such conditions exist, a manually or automatically operated bypass
of some sort is operated.
Protection against loss of prime, provided by the check valve and switch described above, could be
incorporated into the motor starter itself. In such an arrangement, an inverse current relay is used to
interrupt the electric circuit whenever the power consumption, and therefore the current input to the
motor, falls to 50 percent of the minimum value for normal operating conditions. For instance, if a pump
has a maximum of 500 bhp and a minimum (at shutoff) of 280, the inverse current relay will stop the
pump if the power drops to 140 bhp, indicating that the pump has lost its prime and is in imminent
danger of running dry. A time relay must be incorporated in this control, to permit starting the pump.

Driver Protective Controls


The various forms of motor controls which provide for protection against undervoltage, overload,
frequency change, and other motor conditions are not centrifugal pump controls and will be left out of
this discussion except for a few remarks. Motor overload can easily be caused by pump operating
conditions, therefore, provision must be made for this protection if driver overload is possible in the
particular system. The most important consideration in the application of motor protective controls is
that the operator should know when one of these controls has stopped a motor. A bell or a signal light,
therefore, will allow detection of faulty operating conditions as soon as they occur, and they can be
corrected immediately.
In many cases, the driver and its pump should be started again when normal conditions resume;
therefore, the interrupting device is provided with features to reestablish current. In other cases, however,
such as when a pump must be primed manually before starting, the contactors must be prevented from
closing again until the circuit is reestablished manually by pressing the starting pushbutton.

Automatic Operation of Standby Pumps


Another example of a protective device is the automatic operation of standby boiler feed pumps in
a power plant. The exact moment of a pump or drive failure is unpredictable and can occur when the
operators are at some remote portion of the station, unable to start the standby unit without a dangerous
delay. Most protective controls directed towards starting standby pumps are based on pressure conditions
in the main feedwater header. This header pressure is transmitted to one side of a spring-loaded diaphragm,
the minimum permissible header pressure providing the selected value of spring compression. The
diaphragm-operated valve is held in a closed position as long as the header pressure remains in excess
of the predetermined minimum. If the header pressure falls below this value for any reason, the valve
opens. This value is generally used as a pilot valve and can be made to operate any suitable mechanism
that will start one or more of the standby units.

Protection Against Operation Below Recommended


Minimum Flow
The reasons for establishing certain minimum flows and the means used to prevent pump operation
below these flows are discussed in detail in Chapters 22 and 28 respectively.
616 Controls

The most commonly encountered example of this type of protective control is the minimum-flow
bypass on boiler feed pump installations. However, this control performs as important a function in any
installation of centrifugal pumps where the system cycle may introduce the possibility of operation at
extremely reduced flows or against a completely closed discharge. However, when light flow operation
occurs at extremely frequent intervals, as in a steel mill descaling installation, it may be more practical
to maintain the bypass valve constantly in the open position or to provide a three-way valve so that
when the main flow is shut off, the bypass flow is automatically established.

Protection Against Reverse Flow


The use of check valves in the pump discharge to prevent reverse flow in the pumping system has
already been mentioned. The application of foot-valves to pumps operating with a suction lift clearly
belongs in the same class of protective pump controls, although it serves the additional duty of retaining
water in the pump casing, allowing the pump to remain primed at all times.
However, ordinary check valves have the characteristic of closing very quickly, possibly causing
excessive pressures in the system through the creation of water hammer. It is therefore desirable in
certain cases to install slow-closing power-operated valves such as cone valves or butterfly valves. These
valves are generally hydraulically or pneumatically operated and use either the hydraulic pressure in the
pumping system itself or an external source of liquid or air under pressure for their source of power.
These protective controls can be applied not only to prevent reverse flow, but also to protect the system
against situations arising from the rupture of the pumping line, preventing the loss of pumped fluid.
This function is performed by making the valve control responsive to excessive velocities in one direction
only, or in either direction, depending on the particular features of the installation.
The automatic siphon breaker is another protective device used to stop reverse flow through a pump
under certain conditions. It may be applied where the pump discharges into a siphon system, and where
power failure and the resulting pump stoppage would be followed by the loss of water from the discharge
reservoir through the siphon unless siphon action is interrupted. The siphon breaker valve is located at
the top of the siphon and is maintained in a closed position by means of an energized solenoid. Power
failure deenergizes the solenoid and opens the valve, admitting air into the system and breaking the
siphon action.

Relief Valves
A few pumping systems may become subject to excessive pressures or even to water hammer under
certain operating conditions. It is usual, in such cases, to provide protection both for the pumps and for
the system in some form, such as air chambers, surge tanks, or relief valves. The exact nature of the
protective controls depends on all the features of the system into which they are incorporated and must
be solved individually. As a result, it would be misleading to indicate any overall rules or suggestions
on the application of such controls. However, relief valves can be used to protect certain separate portions
of the system against pressures that may be suitable for the remainder of the system but excessive to
that portion. One typical example of this application is the use of relief valves in the suction line of a
pumping system provided with both check valves.in the discharge and foot-valves in the suction. Foot-
valves are seldom suitable for the pressures that may exist in the discharge piping and may need protection
against possible failure or leakage past the discharge check valve.
The surge suppressor used in some discharge lines is a reverse form of relief valve, that is, a relief
valve that opens far enough in advance of the occurrence of excessive pressures to afford full protection
against water hammer. This valve is arranged to open on an excessive drop in discharge pressures and
close very slowly as the pressure is again built up. When the check valve in a discharge line closes
suddenly, the excessive pressures caused by water hammer are preceded by a sudden drop in pressure
Controls 617

immediately beyond the check valve, and by the time the surge or rise in pressure takes place, the surge
suppressor valve is fully open and ready to relieve the situation.

TECHNIQUES OF CONTROL APPLICATION

The knowledge of the fundamental control functions, of the pump characteristics that these controls are
intended to modify, of the component elements of controls, and the understanding of some typical
applications are not a sufficient guarantee of perfect handling of all control problems which may be
encountered in the installation of centrifugal pump systems. In addition, a certain familiarity with the
technique of control application is needed, a familiarity best acquired through experience. However, it
is practical to acquire some knowledge of this technique analytically. The following, therefore, describes
some of the basic principles of the technique of control application. Because the majority of pump
controls are based on throttling either the main delivery stream itself, or some part of the stream in a
take-off branch, this description will be devoted primarily to the application of manually, automatically,
and semiautomatically controlled throttling valves.

Control Characteristics
All control apparatus must have the following characteristics:

1. Accuracy
2. Sensitivity
3. Speed
4. Power

These characteristics are built into the control apparatus by the manufacturer. However, if the engineer
applying the controls does not utilize these characteristics, then all the care the designer exercised to
insure that the control possesses such characteristics will have been wasted.
Accuracy. Accuracy is possible only if the variable to be controlled is measured at a point in the
system where a true reading can be obtained. Such precautions sound obvious, but, unfortunately,
pressure or flow measurements are often attempted at locations where the reading is definitely distorted.
Measurements can also be distorted if the measuring element or the piping next to it is dirty, clogged,
or incrusted. For example, flow measurements by means of venturi meter or orifice should be made at
a point where the flow of liquid in the piping system has been properly straightened. Control manufacturers
generally recommend the length of straight piping required ahead of the measuring element; these
recommendations should be carefully followed.
To be accurate, controls must respond to the slightest deviations from the desired value of the variable
being controlled. To obtain such sensitivity, it is necessary to consider the application thoroughly. For
instance, if the controlling or measuring element is excessively oversized in relation to the quantity to
be measured, all sensitivity is lost. In some cases where the range of operating conditions varies within
wide margins and where sensitivity is required under all conditions, it may be practical to split the
quantity to be measured into two portions. Two measuring elements would integrate the total quantity
for the upper range of its value, and a single element would be used whenever the total quantity is
reduced to less than one-half of its maximum value.
It is preferable, although not always practical, that the greater accuracy of regulation occur in that
portion of the operating range where sensitive control is most desired.
Flow-meter control of a boiler feed pump bypass presents a typical example of the analysis required
618 Controls

to ensure sensitivity in the proper range of the quantity being measured. This control is intended to open
the bypass valve whenever the feed pump flow falls to dangerously low values. The impulse used to
operate the control consists of the differential pressure across a metering orifice. This differential is
highest at the maximum flows and lowest in that particular range where it must operate the bypass
control. The size of the metering orifice must, therefore, be selected to compromise between two
contradictory requirements:

1. The differential at low flows must be sufficient to provide sensitivity in the bypass operating range.
2. The differential at maximum flows must not be excessive, or the resulting head loss may become an adverse
factor in the boiler feed pump installation.

This differential varies with the square of the feedwater flow, and the solution lies in locating the
orifice so that excessive differentials at high flows will not affect operation. This explains the frequent
recommendation to place the orifice in the boiler feed pump discharge rather than in the suction piping
where the resulting head loss would rob the pump of an appreciable portion of the available net positive
suction head.
Sensitivity. Sensitivity of corrective pump controls is affected by valve size as much or more as by
proper selection ofthe measuring element. Valve manufacturers frequently point out that the proportions
and contours of the fluid passages must be given special attention to avoid excessively turbulent flow
and ensuing erosion. Although this phase of the control problem belongs in a treatise on valve design
rather than on the application of controls to centrifugal pumps, the relation between the valve lift, the
area of the opening, and the resulting flow do play an important part in the successful application of
valves. If the valve size selected is too small for the range of flow capacities encountered, friction losses
will be higher than necessary and valve wear will be excessive. The penalty for undersized valves,
therefore, will be an unjustified increase in power consumption because of the increased total head
required in the pumping system and in valve maintenance.
Just as a valve may be too small for its intended service, it may also be too large. Since the area of
a valve opening must be made commensurate with the quantity of liquid due to pass through the valve,
the valve would be almost"c1osed under most operating conditions. This may lead to wire-drawing and
will certainly cause a decrease in sensitivity. Consider, for example, the valve curves in Fig. 23.10. If
a selected gate valve is considerably oversize, that is, if the maximum flow in the system corresponds
to approximately 50 percent of the maximum permissible flow through the valve, then the valve stem
would have to be held at approximately 23 percent of its maximum travel. A reduction of 50 percent
in demand (or to 25 percent of the possible maximum flow through the valve) would call for the
movement of the valve stem to a position at 20 percent of its travel, a reduction of only 3 percent of
the total travel from its former position, corresponding to a negligible angular displacement of the control
wheel. Such a situation is even further magnified in the case of a globe valve and explains the reason
for the unpopularity of the latter for throttling service.
It is apparent, then, that to fulfill the requirements of accurate and sensitive regulation throughout
the full operating range of the controlled equipment, valve design should be chosen to give substantially
equal increments of flow for equal increments of valve-stem travel, regardless of valve position. Such
a design, usually called a "straight line characteristic valve" is illustrated in Fig. 23.10 and is gener-
ally available.
The disadvantages of valve oversizing do not apply to fixed, nonthrottling valves used for on or off
service. In this case, oversizing will reduce friction losses, at the penalty of increasing initial costs to a
point of diminishing returns.
Situations may arise, however, that call for the use of valves much too large for normal operating
conditions, yet properly sized when certain emergency increased flows occur. Valves installed in parallel
Controls 619

100
~~ /

---
~!
~~v/
90
~ / -.J.~'-'/
0 Q) ,,~ ~/ /
..J
IJ.
80 - 0/ (:)~ ",.J..<I. -<-,c, /
t-C::, ~,'? /
~
70 IA".O~/ ....~
:::l In~ ~c,
~
c,+~~/
60 I ;/
X
I
\,\~~~
<t
50
~
A". A
IJ.
0 40 9,0/ I
~ "
,:?-"\ P' -.1
~ "A...." \'.Il-\()......
c;~
....Z" 30 '-
=~/ :,.....-
u

....0:Il.
20
f-
/ ,., ,;'

'/ I
10
/ /

o ---
10 20 30 40 50 60 70 80 90 100
PER CENT VALVE - STEM TRAVEL

Fig. 23.10 Flow characteristics of different types of valves.


(Courtesy Bailey Meter Co.)

present definite advantages whenever such wide variations in service demands are encountered. Only
one valve is open at very light flows; additional valves are opened when the demand increases.
Speed. Except for those applications where the impulse is transmitted manually, speed of impulse
transmission is generally the problem of the control manufacturer. However, in many instances, the user
may handicap the manufacturer of controls in this respect by his insistence on a particular operating
medium that precludes suitable speed of transmission under the prevailing conditions and may introduce
a time lag that ultimately defeats the entire purpose of the automatic control application.
Power. Electrically operated valve controls are generally available in the semiautomatic or in the
fully automatic types. In both cases, an electrically operated valve consists of a reversing torque motor,
which drives the valve stem through a train of gears. The motor control is arranged to stop the valve
in any position required to obtain the desired control. In semiautomatic operation, the impulse is manual
and is controlled by the operator at a pushbutton station. In a large installation, requiring the manipulation
of a great number of valves, full control of each valve can be provided by a centrally located battery
of pushbuttons. Such control permits the installation of the valves where best plant design and not
accessibility considerations dictate. When fully automatic operation is required, the measuring element
is made to transmit the necessary impulse to the control panel where, in tum, an electric circuit operates
the main valve motor in the desired direction.

Applied Techniques

Some suggestions on the techniques of control application may appear self-evident, and yet attention
must be directed to certain pitfalls. A typical example of an error in application would be the use of the
system hydraulic pressure to operate certain protective controls for the very centrifugal pump activating
this system. If the function performed by the protective control is not necessary should the pump be
suddenly stopped, the use of the system pressure for the power element of the control is acceptable, at
least from the point of view of the safety of the supply. If, however, the protective control is required
620 Controls

to function either on the stoppage of the pump or during the starting and stopping period, there will be
no system pressure to be used as the power element.
Another problem is a tendency toward "overengineering" in control selection, as in the selection of
any other equipment. Overengineering refers to the selection of equipment that is actually too good for
the intended job, and of controls designed to perform a more difficult function than required in the given
instance. This leads to unnecessary expense and, at times, to excessive outage and maintenance which
could easily be avoided by the application of less complex mechanisms.
In considering automatic control, it should be remembered that in most cases automatic control is
directed at supplementing, not supplanting, proper instrumentation and careful observation by the opera-
tors. An indicating or recording instrument should always be installed in proximity to the point in the
centrifugal pumping system where control action is applied. This instrument, which registers flow, level,
pressure, or temperature, depending on which variable should be controlled, is especially important in
the case of fully automatic controls, as it will indicate at a glance whether the control operates properly;
it can even be arranged to warn the operator of any failure in the functioning of the control.

Selection of Controls
It is necessary, when selecting or ordering centrifugal pump controls, to give complete information
to the control manufacturer. Some data most essential to the contemplated installation are as follows:

1. Service or control function-Specify what is to be accomplished.


2. Nature of fluid handled-The character of the fluid will have an effect on the selection of the type and
the size of the required valving. If the liquid is corrosive, indicate complete analysis. Also, if the viscosity
of the liquid differs appreciably from that of water, indicate average or, at least, the range of viscosities.
Specify liquid temperature and whether it is constant or variable.
3. Capacities of the flow being controlled and values of the initial and final pressures, if pressures are
regulated-If steam flow to a turbine driving a centrifugal pump is to be controlled, full steam conditions
must be indicated.
4. Character of system operation-Is it constantly under load or is the operation intermittent?
5. Manual or automatic control-If automatic, which elements of the control must be automatic and which
can be manually operated? Is remote control indicated or can it be self contained?
6. Required accuracy of regulation.
7. Required range of adjustment.
8. Type of service-Is dead-end service involved for the regulating valve or is the flow continuous and small
leakage, as in double-seated valves, permissible?
9. Control safeguards-Should failure of control place the regulating valve into fully open or fully closed
position? This question is especially important in the case of protective controls.
10. Structural details-Position of valve (vertical or horizontal), and type of connections for valve nozzles (if
flanged, state desired facing and drilling).
11. Range-For all of the data, indicate the full operating range of all variable factors, such as capacities,
pressures, and temperatures, including definite maximum and minimum values.
24
Drivers

During the history of centrifugal pumps, probably every form of prime mover and source of power, with
some form of intermediate transmission when necessary, has been used to drive them. The prime movers
commonly used today and the reasons for their choice are:
Electric motor. The alternating current (ac) induction motor is simple, easy to install, and reliable,
and is therefore the usual first choice of prime mover for any installation where there is an adequate ac
electrical supply system. Alternating current motors are fixed speed unless of wound rotor construction
or connected across a variable frequency power supply. Direct current motors are used only for spe-
cial applications.
Steam turbine. These are generally limited to (1) standby or emergency drives, (2) installations where
steam has to be throttled, (3) applications where variable speed operation is required and steam is
available, or (4) drives where the power required exceeds the economic capability of electric motors,
typically a consideration above 7,500 kW (10,000 hp).
Internal combustion (IC) engine. Widely used for emergency drives for fire pumps, and in agricultural,
industrial, or oil production installations where there is no electrical supply system. Occasionally used
when there is a source of fuel available that would otherwise go to waste, e.g. methane from a sewage
treatment plant. Ie engines allow variable speed operation over a limited speed range.
Gas turbine. Used extensively in oil production installations, where fuel is readily available and the
electrical supply system is limited or nonexistent, and when the power required exceeds the economic
capability of electric motors. Mechanical drive gas turbines generally can be operated over a limited
speed range.
Hydraulic turbine. Usually limited to supplementary drives for power recovery from throttling a high
pressure liquid stream; typically either petroleum refining or water supply applications.
The choice of rotative speed is limited with many types of drivers and often does not match the speed
of the ideal pump for the service conditions. For continuous service units, a speed-increasing or
-decreasing device is warranted. For standby units, which are used occasionally, a compromise pump
design can often be employed, saving the cost and complication of an intermediate transmission or
electrical control at some sacrifice in pump performance and, in some cases, with some increase in
pump cost.

621

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
622 Drivers

SIZE OF DRIVERS

Local conditions will determine the load permissible for different drivers. For example, the design of
an electric motor, as well as the temperature and density of the surrounding air, and the variations in
available electric current will determine the safe load that can be carried by the motor.
The driver should be capable of carrying the loads that will be imposed on it over the entire range
of operating conditions of the pump under the most adverse operating conditions. The meaning of the
phrase "most adverse conditions" cannot always be interpreted literally. Sound economics will not justify
the selection of a squirrel cage induction motor for a rating under conditions that might last only a few
hours per year. For example, if an unusually high ambient temperature might occur for 20 or 50 hours
a year, it may have to be disregarded. Thus, if the temperature is 50°C instead of 40°C for a limited
number of hours, economics will justify selecting the motor on the basis of the 40°C ambient temperature.
Theoretically, this would shorten the thermal life of the motor insulation somewhat, but, practically, it
may make little change in the actual life of the motor because so many other conditions also determine
the operating life of the motor.
If the driver is variable speed and the service conditions require operation over a wide head range,
necessitating a wide power range as well, the pump speed is selected to keep the maximum power
required within the available driver size.
The power required to drive a centrifugal pump from shutoff to maximum capacity varies with its
specific speed type and its individual design (see Chap. 18). In the lower-specific-speed types, the power
requirement is minimum at shutoff and increases to maximum at, or near, maximum capacity. With this
type of pump, operation at heads below normal will result in an increased power requirement and may
impose an overload on the driver.
High-specific-speed types, however, have the opposite power requirements, and a reduction in head
causes a reduction in power. In a high-specific-speed pump, an increase in operating head will cause an
increase in power consumption and may overload the driver.
In many industries, petroleum refining, paper and mineral processing being three examples, pumps
are required to be capable of a head increase between 5 and 10 per cent of rated head. When head
reserve is required, the driver sizing must have sufficient margin to provide the additional power should
the pump be rerated.

Electric Motors
If the brake horsepower of a centrifugal pump exceeds the safe operating load of the motor, which
depends on the type used, the motor may be damaged or burned out. The shape of the power characteristic
curve and the system-head characteristics will determine if pump operation may exceed the safe loading
of its electric motor. Careful attention must also be paid to the shape of the speed-torque curve of the
motor and the voltage supply of the power system. A motor with ample speed-torque characteristics at
rated voltage may not be capable of handling the load at some reduced voltage.

Steam Turbines
A steam turbine driving a centrifugal pump cannot be overloaded. With a fixed throttle position, a
steam turbine develops a constant torque; the speed of the unit will always be at a point where the
turbine torque equals the pump torque. If the turbine is equipped with a constant-speed governor, the
governor will throttle the steam supply until-at the required speed-the developed torque and the
torque demand are balanced. With an increase in torque demand, the governor will open the throttle
valve until the point is reached where the torque demand starts to exceed the available torque. Since
the governor can no longer provide additional steam to further increase the available torque, the steam
Drivers 623

turbine slows down until the torque demand equals the available torque at some speed below the
design speed.
When the turbine is equipped with a constant- or excess-pressure regulator, the effect is identical to
that described for a constant-speed governor; the only difference is the source of the impulse that tends
to maintain the required torque balance until external conditions occur to interrupt the functioning of
the governor.
A turbine driving a centrifugal pump should be capable of carrying the maximum load that will be
imposed on it by the pump under the most adverse steam conditions. These can be determined fairly
accurately in large units, but less so in smaller units. As wear causes increased leakage through the
wearing rings, the pump capacity will be reduced. It is desirable, therefore, to increase the speed of a
turbine-driven pump slightly to restore the desired capacity. Generally, small pumps are not checked as
carefully as large pumps, and the leakage may become more pronounced before the running clearances
are restored to their original value. In small stream-turbine-driven units, the turbines are generally selected
for some excess power (roughly 5 percent over maximum expected loading), whereas in large units,
turbines are selected for a rating equal to, or just slightly in excess of, the maximum loading.

Internal Combustion Engines


Power ratings of some internal combustion engines are given for the load they can carry continuously
at the rated speed, whereas in other cases, ratings are given for the load the engine can develop on test
with fully open throttle. In the latter category, the power rating may include the power required by the
auxiliaries normally attached to the engine, such as the circulating pump or the fan. Automotive type
engines are generally rated for their maximum developed power on dynamometer test while stripped of
all auxiliaries (Fig. 24.1). Usually, these engines can safely be operated with a continuous loading of
75 to 80 percent of their developed power.
Most automotive engines are of the high-speed type and have a very limited life when used for
continuous operation at full load. Medium-speed engines with a longer life and low-speed engines, which
have a very long life with a minimum of maintenance, are also available.
Although internal combustion engines rated on the basis of the load they will carry constantly can
carry an overload and although engines rated on the basis of maximum developed power can be operated
while loaded to a higher percentage of that developed power, prolonged overloaded operation will
generally result in mechanical failure.
The engine best suited for any particular application must therefore be selected with due consideration
to the service requirements.

Gas Turbines
The maximum continuous power rating of mechanical drive gas turbines is determined by the allowable
temperature of their power blading, and is usually specified as an allowable exhaust gas temperature,
which can be easily measured. Gas turbines are sensitive to inlet air temperature and density, therefore
the installed power rating depends on the expected ambient air temperature and the installation's elevation
above sea level.

DRIVE ARRANGEMENTS

In the simplest and most common arrangement, the driver is a single machine (e.g., an electric motor
or a steam turbine). More complex arrangements involve a "train" of two or more machines in series,
usually the driver plus some form of transmission, to allow variable-speed operation or achieve a speed
624 Drivers

Fig. 24.1 Gasoline engine.


This type is often employed as a fire pump driver and for other standby applications.

not available with direct motor drive or both. A second arrangement is multiple drivers such as two half-
capacity motors, or a motor plus a hydraulic power recovery turbine. In applications where dependability is
critical (e.g., pumps for fire protection), dual drives, usually motor and engine, are often used. Details
of the components of these various drive arrangements are discussed later.

VARIABLE· SPEED DRIVES

When varying operating conditions exist in a centrifugal pump installation, variable-speed operation
may be desirable; drivers that economically allow variable-speed operation are preferred for such installa-
tions. A wide choice of drivers suitable for variable-speed operation is available: steam turbines, internal
combustion engines, gas turbines, wound-rotor motors, or, in certain cases, dc electric motors. Synchro-
nous or squirrel cage induction motors can be used in series with a variable-speed transmission, such
as a hydraulic coupling or magnetic drive, or they can have the frequency of their power supply modulated
to vary their speed, an arrangement termed "variable-frequency drive" (VFD). In low-power applications,
where friction-type variable-speed drives were once used, VFD is now the usual method of varying
speed. Its use for higher-power drives is increasing as the cost of electronic components continues to fall.

Advantages of Variable Speed


Many centrifugal pump applications require operation at varying capacities and total head. Since the
operating conditions of a centrifugal pump are determined by the intersection of its head-capacity curve
with the system-head curve, the only way to vary operating conditions is to alter the system-head curve
by throttling if the pump is operating at constant speed. It may often be more practical and more
Drivers 625

economical to change the intersection point described above by varying the operating speed and, therefore,
the head-capacity curve of the pump.
Many factors enter in the evaluation of the most economical method of operation under such conditions.
Whereas considerable pump horsepower can be saved by varying the speed of a steam turbine or of an
engine, the efficiencies of wound-rotor motors and of hydraulic and magnetic transmissions are practically
proportional to the ratio of their operating speed and full load speed. Both the actual torque and the
ratio of speeds must be taken into consideration. For example, if there is a 20 percent speed reduction
at 100 percent of full load torque, the slip losses are 20 percent. But if the speed reduction is 20 percent
and the torque at reduced speed is only 50 percent of full load torque, the slip losses are only 10 percent.
Therefore, the actual power consumption as well as the increased cost must be considered in the evaluation
of a variable drive against a constant speed drive.
A description of both hydraulic couplings and magnetic transmissions appears in a later portion of
this chapter.

DUAL DRIVES

The pump industry defines a dual-driven pump as one that can be driven by either of two drivers. The
two drivers are generally of different types; one is for regular use, such as an electric motor, and the
second is a standby type such as a gasoline engine, which is used if the power supply for the regular
driver fails (Fig. 24.2). The pump selection may have to be compromised by the available drive speed.
Dual-drive units are generally arranged with a driver on each end of the pump so that the unused
driver can be disconnected. It is then necessary to use a double-extended pump shaft; end-suction pumps
cannot be arranged in this manner. Some dual drives are arranged with a double extended-shaft motor,
with both driving units on one end, in which case end-suction type pumps can be used.

Fig. 24.2 Dual-driven unit for water works.


Pump is normally electric-motor driven. The gasoline engine driver is used if power fails.
626 Drivers

Fig. 24.3 Multistage pump (left) driven by motor and hydraulic power recovery turbine (HPRT)
with an overrunning clutch.

TANDEM DRIVES

In a few installations, the pump is driven by two drivers in tandem, with each carrying part of the load. The
most common installations of this type involve an electric motor and a steam or hydraulic turbine without
governors, or with governors set at a higher speed. When induction motors are used, the speed of the unit
varies slightly with changes in operating conditions to reach a balance in which the power developed by
the two drivers equals the power required by the pump. The speed change is quite small, being on the order
of 1.0 to 1.5 percent from no-load to full-load speed in motors 150 kW (200 hp) and larger.
The hydraulic turbines used in tandem drives usually recover energy from a liquid stream whose
pressure must be lowered. Depending on the process, the high-pressure liquid stream may not always
be available, or the energy from it may vary. In these cases, common in petroleum refining, the motor
driver is sized to provide 100 percent of the pump's power requirement, and there is an overrunning
clutch installed between the turbine and the rest of the unit (Fig. 24.3) so the turbine cannot act as a
brake should its power output fall to zero. Turbines so arranged are generally equipped with a mechanical
overspeed trip to prevent runaway under any circumstance.
Occasionally, when the pump power is very high, tandem motors are used to keep the size of the
motor and the associated electrical apparatus within reason.

BELT DRIVES

Before electric drive became almost universal for centrifugal pumps, many isolated pumping plants were
steam operated and used steam engines for drivers. In many cases the speed of the pump could not be
Drivers 627

matched to the slower engine speed, and the engine drove the pump through a flat belt with different
diameter pulleys to give the correct pump speed. Flat-belt-driven pumps are seldom used now except
in small sizes and for irrigation pumps that are driven from tractor power-take-offs.
Multi-V-belt drives are frequently used for centrifugal pumps. The most common application is on
slurry pumps used for mineral processing. These pumps have either rubber-lined or hard metal impellers
whose diameter cannot be easily changed, so a multi-V-belt drive offers a convenient means of adjusting
their performance to match the system requirements. The one mineral processing application where belt
drives of any form cannot be used is bauxite refining by the Bayer process. The caustic environment
inherent in this process lowers the coefficient of friction causing belts to slip. Other applications of
multi-V-belt drive include driving pumps in agricultural installations or cases where the pump is an
auxiliary mechanism, such as a cooling water pump for internal combustion engines.
Belt drives require first that the bearings of the coupled machines be capable of handling the radial
load produced by the drive. Second, provision must be made for adjustment of the belt tension and for
sufficient slack to allow the belts to be put on the sheaves. The pump location is usually fixed because
of its suction and discharge piping. This adjustment must therefore be accomplished by moving the
driver or, if that is impractical, by introducing an adjustable idler pulley.

GEAR DRIVES

The optimum range of pump speed often does not match the optimum or even practical range of driver
speed. These circumstances prevail in many steam-turbine-driven low-head pumps (Fig. 24.4); motor-
driven high-head pumps when a short, stiff rotor is required (Fig. 24.5); and large pumps driven by
internal combustion engines (Fig. 24.6). For such installations, gears are used to provide the desired
pump speed, and are very reliable provided they are correctly rated and installed.
Gears connecting horizontal-shaft drivers to horizontal-shaft pumps are usually double-helical, single-

Fig. 24.4 Large-capacity series pumping unit driven by condensing steam turbine through a reduction gear.
Small pump, driven at end of main pump, supplies circulating water to condenser serving the turbine.
628 Drivers

Fig. 24.5 High-pressure high-speed boiler feed pump.


Step-up gear permits a 1.450-rpm motor to drive pump at 3.800 rpm. through a fluid drive that gives
variable-speed control.

Fig. 24.6 Pump installation driven by slower speed diesel engine through a step-up gear.
Drivers 629

reducing, or -increasing gears. Right-angle gears are often used to connect horizontal-shaft drivers to
vertical-shaft pumps. These drives were developed for engine-driven, vertical turbine pumps and are
capable of carrying a large downward thrust load on their vertical output shafts. Most of these gears are
the hollow-shaft type favored in the vertical turbine pump field.

The Selection of Electric Motors


For Centrifugal Pump Drive
Drivers for centrifugal pumps must be selected carefully so that the unit will be satisfactory. When the
drive is to be a steam turbine or an internal combustion engine, full details of the requirements are
usually given to the manufacturer of the driver. Frequently, less care is exercised with electric motors,
resulting in the use of motors with unsuitable electrical or mechanical constructions. This misapplication
is due mainly to the fact that so-called general purpose motors are offered by the various motor
manufacturers for ratings of 150 kW (200 hp), or less, and 450 rpm, or more, without restriction to a
particular driver application. The user must realize the requirements of the application to ensure that the
proper motor is employed (Figs. 24.7 and 24.8).

ELECTRICAL SYSTEMS

In dc power supply systems the flow of electricity is in one direction and the power in kilowatts is volts
x amperes/l,OOO. Direct current systems have certain limitations and disadvantages that limit their use;
therefore most power systems are ac.

Fig. 24.7 Small-size motor-mounted pump.


630 Drivers

Fig. 24.8 Motor-driven centrifugal pumps for water works.

Time in electrica l degrees

Time in seconds

Fig. 24.9 Alternating current cycle.


Relation between time in seconds and time in electrical degrees for a frequency of 60 Hz.
Drivers 631

In ac systems, the flow of current reverses or alternates periodically. The voltage and the current vary
in amount and direction with time. The current and voltage can reach their maximum at the same time,
or the peak of the current can be ahead of, or behind, the peak of the voltage. The measure of this lead
or lag is called the power factor, and it can range from 0 to 1.0 or 0 to 100 percent. The power in
kilowatts is

volts x amperes x power factor x KIIOOO

where K is 1.0 for single-phase systems, 2.0 for two-phase, four-wire systems, and 1.732 for three-
phase systems.
An ac cycle is completed when the current or voltage rises from zero to peak, falls to zero, rises to
peak in the opposite direction, then falls to zero again (Figure 24.9). The number of cycles occurring
per second is the frequency, and is known universally by the term hertz; abbreviation Hz. A number of
different frequencies were once used around the world, common values being 25, 30, 40, 50 and 60 Hz.
Today only two frequencies, 50 and 60 Hz, are in wide use. The United States, Canada, much of Central
and South America, and North East Asia have 60 Hz systems. Europe, Africa, the Middle East, South
East Asia, and Australasia have 50 Hz systems.
Alternating current systems are classified by the number of phases; single two- or three-phase systems
are used, with single- or three-phase systems the most common. Two- and three-phase systems are
known as polyphase systems.
Tables 24.1(a) and (b) list the standard voltages used for 60 and 50 Hz ac systems around the world.
The voltages most commonly used are shown in bold type. The table also lists the normal range of
motor sizes that can be obtained with these voltages.
The speed of an ac motor is a function of the frequency of the supply and the number of poles in
the motor's stator, always an even number as the poles must be in pairs. The synchronous speed of a
motor in revolutions per minute is

RPM = 120 (frequency)/(number of poles)

The 25-, 50-, and 60-Hz synchronous speeds are shown in Table 24.2.

Table 24.1(a) Standard Voltages and Corresponding Normal Motor Size Ranges-IEC Practice
Boldface indicates the voltages most commonly used.

Reasonable motor sizes,


in kilowatts
Supply Voltage Minimum Maximum

Single phase alternating current 110 1.0


220-240 3.7
440 7.5

Two or three-phase alternating current 380


415--440 0.1 300
600 0.1 300
3,300
6,600 110 5,000
11,000 150 7,000
1,000 20,000+
632 Drivers

Table 24.1(b) Standard Voltages and Corresponding Nonnal Motor Size Ranges-U.S. Practice
Boldface indicates the voltages most commonly used.

Reasonable motor sizes, in


horsepower
Supply Voltage Minimum Maximum

Direct current 115 30


230 200
550-600 0.5
Single phase alternating current 11 0-115-120 1.5
220-230-240 10
460-550 5 10
Two or three-phase alternating current 11 0-115-120 15
208-220-230- 200
240 500
440-550 40
2,300 75
4,000-4,160 400
6,000-6,600

Table 24.2 Synchronous Speeds for 60-, 50-, 40-, 30- and 25-Cycle Motors

Motor poles 60 cycles 50 cycles 40 cycles 30 cycles 25 cycles

2 3,600 rpm 3,000 rpm 2,400 rpm 1,800 rpm 1,500 rpm
4 1,800 1,500 1,200 900 750
6 1,200 1,000 800 600 500
8 900 750 600 450 375
10 720 600 480 360 300
12 600 500 400 300 250
14 514 428 343 257 214
16 450 375 300 225 187
18 400 333 267 200
20 360 300 240
22 327 272 218
24 300 250 200
26 277 231
28 257 214
30 240 200
32 225
36 200

DIRECT CURRENT MOTORS

Direct current motors are available in three types: shunt wound, series wound, and compound wound,
although the commercial stabilized shunt-wound dc motor has some compound windings. The general
purpose type of dc motor is not suitable for centrifugal pump drive as the full load speed of such motors
may be 5 to 7Y2 percent above or below their rated speeds. If the actual speed of the motor is below
Drivers 633

rated speed, the pump will not produce its rated conditions, while if the actual speed of the motor is
above rated speed, the pump will deliver excess head or capacity with increased power requirements,
possibly imposing a dangerous overload on the motor. Direct current motors intended for centrifugal
pump drive, therefore, must be specifically ordered, and a design is required that will provide a hot full
load speed within plus or minus 3 percent of the rated full load speed.
This requirement can generally be provided only in the case of larger dc motors. With small dc
motors, a small fixed resistor is necessary to obtain the speed desired for pump drive.
The speed of a regular shunt-wound, dc motor can be increased with little change in efficiency by
the use of a field rheostat. The motor speed also can be reduced by introducing resistance in the armature
circuit. The latter method of speed control causes a considerable loss of efficiency.
With the advent of economical, high-power electronics, development is proceeding on new forms of
dc motors. One example uses electronic commutation of the stator field to produce torque in a permanent
magnet rotor. By varying the speed of commutation, the rotor speed is also varied. Prototypes of such
motors have delivered 150 kW (200 hp) at 18,000 rpm. It is motors such as these that will likely drive
the centrifugal pump of tomorrow (see Chap. 33).

ALTERNATING CURRENT MOTORS

Single-Phase Motors
Single-phase, ac motors are available in the following types: split-phase, series (universal), repulsion,
repulsion-induction, and capacitor. Generally either the repulsion-induction or capacitor-type motor is
used for centrifugal pump drive.

Squirrel Cage Motors


Squirrel cage induction motors are the simplest polyphase electric motors and are the most widely
used driver for centrifugal pumps. They have wound primary (stator) windings and squirrel cage secondary
(rotor) windings (Fig. 24.10). The secondary windings develop the circulating current necessary to

15

Item Description
1 Stator Core
2 Stator Yoke
3 Stator Coils
4 Bearing HOUsing
5 Rotor Shaft
6 Inner End Cap
7 Bail Bearing
8 Shaft Seal
9 Main Terminal Box
10 Grid Cover
11 Fan Inlet Baffle
12 Ex1ernal Fan
13 Fan Housing
14 Interna I Fan
15 Rotor Core

Fig. 24.10 Section of typical squirrel cage motor, totally enclosed fan-cooled (TEFC) enclosure.
(Courtesy Siemens Energy & Automation, Inc.)
634 Drivers

produce torque by induction from the rotating stator field. Mechanical construction permitting, squirrel
cage motors can be run in either direction. They are reversed by interchanging any two leads of a three-
phase motor, or the two leads of one phase of a two-phase motor.
The speed-torque characteristic of squirrel cage induction motors varies with the resistance of their
secondary winding, thereby providing the means of designing the motors for various speed-torque
characteristics. Figure 24.11 shows the four characteristics adopted by NEMA [24.1] as their standard.

TORQUE TERMINOLOGY
BREAKDOWN TORQUE
200 STARTING TORQUE (PULL-OUT)
(LOCKED ROTOR
Cl

9
BREAKAWAY)

....J FULL-LOAD
....J
::::l TORQUE
LL.
~ 100 +-------~------~------------_\
I
W
::::l
oa::
~
o 20 40 60 80 100

300

Cl 250
«
9
:::l200
::::l
LL.
~
I
w 150 I~_ _""'"
::::l
oa::
o
I- 100

50

I I
o 20 40 60 80 100
SYNCHRONOUS SPEED - %

Fig. 24.11 Typical speed-torque characteristics of squirrel-cage induction motors standardized by NEMA.
(After Westinghouse Electric Corp.)
Drivers 635

Centrifugal pumps usually do not require motors with high starting torques, therefore the motors normally
used are NEMA design B (normal torque, low starting current) or the equivalent NEMA design A motors
(normal starting torque, normal starting current) can be used, but usually require some form of reduced
voltage starter to keep the current inrush within the limits set by the public utilities. This combination
of equipment is more expensive than a design B motor with a full voltage starter. In a few instances,
typically when the power supply has limited capacity, it is necessary to use NEMA design C motors
(high starting torque, low starting current). These installations employ an autotransformer starter with
a tap as low as 50 percent of the available voltage, to limit the starting current. With this low voltage,
the starting torque of a normal torque motor may not be sufficient to start some pumps, such as vertical
turbine pumps, which have higher than normal starting torque. Using a high-starting torque motor in
these cases offsets the effect of the low-starting voltage.
Squirrel cage induction motors normally have only a single nominal speed. Special versions are
available with a number of discrete speeds. In one version, the stator winding is wound so that by
external switching the number of poles can be changed. In this way, the winding of a 6-pole (1,200 rpm
at 60 Hz) motor can be reconnected to make it a 12-pole (600 rpm at 60 Hz) motor. Mechanical and
electrical problems limit the combination of poles and therefore the choice of available speeds. Commer-
cial single-winding motors, for example, are limited to ratings in which the lower speed is half the higher
speed, such as 3,600 and 1,800 rpm, 1,800 and 900 rpm, 1,200 and 600 rpm, and similar combinations.
The second version of mUltispeed motor has two separate stator windings, each wound with a different
number of poles. One speed is obtained with one set of windings, another speed with the second set.
The motor can be built so the number of poles in one or both windings can be changed by external
switching, thus obtaining a three- or four-speed motor, as for example, a 1,800/1,200/900/600 rpm motor.
Constant horsepower and both constant-torque and variable-torque mUltispeed motors are available.
The power required by a centrifugal pump varies with the cube of the speed, and, consequently, the
torque varies with the square of the speed, therefore a variable-torque multispeed motor functions well
as a centrifugal pump drive. Three- and four-speed motors are seldom used as centrifugal pump drives
because the speed range is very wide and the resulting capacity and head at reduced speed usually too
low to be useful. They can be employed when the pump is used in a circulating system in which the
head is entirely friction. In systems with a small static head component it is sometimes feasible to utilize
a drive with three or four speeds by coupling two single- or two-speed (two windings) motors together,
or on either end of the pump. With the advances in frequency modulation (see variable-speed drive),
the use of multispeed motors is hard to justify unless the pump has distinctly different duties that can
be met at the fixed speeds available.

Wound-Rotor Motors
Wound-rotor motors have both primary and secondary windings. The primary (stator) winding is the
same as that of a squirrel cage motor. The secondary (rotor) winding is connected to slip rings so an
external resistance can be introduced into the secondary winding circuit for starting, or for speed
regulation. The resistance reduces the current in the secondary winding, thereby changing the torque-
speed characteristics of the motor (Fig. 24.12). The major field of application of wound-rotor motors is
for installations where low-starting-current or variable-speed operation is necessary. When wound-rotor
motors are used for variable-speed operation, the resistors must dissipate the heat resulting from the
current flowing through them, or, in an electronic arrangement known as "slip recovery," convert the
surplus rotor power to line frequency and put it back into the grid. Commercial speed controls cut the
resistance into the rotor circuit in steps so the resulting speed change is also in steps. For fine adjustment
of speed, a control with a large number of contact points or a liquid slip regulator is required. Where
very fine adjustment of speed is required, variable speed drive by frequency modulation is now the
preferred approach.
636 Drivers

ZERO
EXTERNAL
RESISTANCE

o
RESISTANCE
9
...J
STARTE ...J
::::>
LL
cfl.
I
w
0123456 ::::>
oa:
~

SYNCHRONOUS SPEED - %

Fig. 24.12 Basic connection diagram and typical speed-torque characteristics of wound-rotor induction motor.
(After Westinghouse Electric Corp.)

The starting characteristics of wound-rotor motors are one of their major advantages. With proper
resistance in the rotor circuit, the current demand throughout the entire starting cycle can be kept very
low. The standstill line current specification at rated torque is generally within 130 percent of rated
current. (This current difference is due to the power factor being lower at standstill than at rated load
and speed.) Wound-rotor motors, therefore, are desireable when the power supply is limited, or the size
of the supply line to the motor is limited.

Synchronous Motors
Synchronous motors have a wound stator connected to the ac supply and a rotor with wound poles
(salient pole construction) connected to a dc circuit. The direct current is usually supplied through slip
rings from a separate source of supply or from a dc generator, called an excitor, which may be attached
to the end of the motor (Fig. 24.13). An alternative design, known as a "brushless synchronous motor,"
avoids the use of slip rings by having an ac generator on the end of the shaft (Fig. 24.13), whose ac
output is converted to dc by a solid state rectifier, also mounted on the motor shaft, then directly connected
to the field coils. The virtue of this design is the elimination of slip rings, which produce sparking (not
allowed in many environments) and are a high-maintenance item. This design of motor runs at its
synchronous speed regardless of the load, hence its name.
Because they inherently have very low starting torque, synchronous motors use an auxiliary squirrel
cage winding built into the rotor for starting duty. The motor is started as a squirrel cage motor, and
once the speed has reached approximately 95 percent of synchronous, the field (rotor) current is applied
when the position of the rotor matches the ac flow in the stator and the motor accelerates to synchronous
Drivers 637

STATOR LAMINATIONS

AMORT I SSEUR
WI NDING ---r---I-I~~

BAUSHLESS
EXCIT ER

BASEPLATE

Fig. 24.13 Section of open, pedestal-type, brushless synchronous motor.


(Courtesy Dresser-Rand, Motor and Generator Division.)

speed. The timing of the application of the field current is determined electronically by the automatic
controller, which greatly simplifies the starting of synchronous motors.
The use of synchronous motors for centrifugal pump drives is justified by the energy cost savings
from their higher power factor and higher efficiency. These savings do not offset the motors' higher
capital cost until the speed is below 600 rpm or the power greater than approximately 0.75 kW (1.0 hp)
per rpm (Fig. 24.14).

Heat Dissipation
It has been said that electric motor design is about heat dissipation. The electrical losses in a motor
go into heat that must be dissipated to the surrounding atmosphere or to a circulated coolant, either air
or liquid. The amount of heat generated depends on the load, whereas the operating temperature of the
motor's windings depends on the load and the surrounding air or coolant temperature (Fig. 24.15). The
allowable operating temperature of motor windings depends on the class of insulation applied to them
(see insulation).
The rate of heat dissipation from motors is affected by the form of enclosure, the frame material and
surface area, the velocity of air moving over or through the enclosure, the temperature of the ambient
air and its density. The first four of these factors are dealt with in the discussion of enclosures. The
remaining two are discussed here.
Motors are rated on their maximum temperature rise above an assumed ambient air temperature of
40°C (l04°F), when carrying rated load. If the ambient temperature is lower than 40°C, a greater
temperature rise can be tolerated in the motor. Since the temperature rise of the motor's windings is
very nearly proportional to the load squared, the load that can be carried by the motor without exceeding
638 Drivers

SYNCHRONOUS
~ 3
0
0
0

I
CJ SYNCHRONOUS
z OR
~
a: 2 INDUCTION
a:
~
0
~

INDUCTION

3600 1800 1200 900 720 600 514 450 400 360

SYNCHRONOUS SPEED - RPM.

Fig. 24.14 Guidelines for synchronous motor selection.


Below 600 rpm, or power greater than approximately 0.75 kW (1.0 hpj per rpm, synchronous motors are a
better selection than squirrel cage induction motors because higher cost can generally be offset by higher
power factor and efficiency.
(After Power, Special Motors Report, June 1969, and Mitsubishi.)

RISE
TOTAL
TEMPERATURE

I AMBIENT

Fig. 24.15 Motor temperature and total temperature.


Drivers 639

the allowable winding temperature is greater. If the surrounding air temperature is higher, however, the
maximum load the motor can safely carry is lower. Similarly if a motor is rated for a temperature rise
lower than the maximum allowable for its class of insulation and 40°C ambient air, it can be operated
at a higher load without the allowable winding temperature being exceeded.
As the altitude increases, the lower air density reduces the rate of heat transfer from the motor.
Standard motors are designed for operation at altitudes up to 1000 m (3,300 ft) above sea level. Above
1000 m (3,300 ft) the safe loading of the motor is reduced. Motors designed for high altitudes are
available at an additional cost.

Insulation
The purpose of insulation is to electrically isolate the motor's windings from each other and from
the rest of the machine. Because the windings necessarily operate at some temperature above ambient
to dissipate the heat generated within them, the insulation has to withstand that temperature. There are
presently four classes of insulation universally employed in electric motors. Each of these classes has
a maximum temperature that it can withstand and still provide a practical insulation life. The classes
and their maximum operating temperatures are given in Table 24.3.
The choice of insulation class depends on the motor's energy density, higher density requiring a
higher allowable insulation temperature, and the means of heat dissipation (see enclosure). Many users
take this a step farther and build in a margin on insulation temperature by specifying a class of insulation
but limiting the allowable temperature to that of the next lower class (e.g., class F insulation with class
B allowable temperature).
Electric motor application includes a "service factor" which ranges from 1.0 to as high as 1.5. The
implication is that the motor has the thermal capacity to operate at its nameplate power times the service
factor. This it will do, but with what is termed a "safe insulation temperature," which is above the
maximum operating temperature for its class. Continuous operation at this power would therefore result
in shorter insulation life, and so the purpose of the service factor is to provide capacity for short-term
operation beyond nameplate power.

Enclosure
An electric motor's enclosure serves first to either protect the motor from its environment or the
environment from the motor, then to provide the means of heat dissipation. Looking at the environment
first, the motor has to be protected from water and dust. Reciprocally, when the environment contains
gases, vapors or combustible dusts, the motor cannot be allowed to be a source of ignition (see hazard-
ous locations).
There are many variations of enclosure design. For centrifugal pumps six are in broad use. The six
fall into two broad categories: open and totally enclosed. With an open enclosure the motor is ventilated

Table 24.3 Insulation Classes and Maximum Operating


Temperature

Insulation class Maximum operating temperature (0C)

A 90
B 130
F 155
H 180
640 Drivers

with ambient air (Fig. 24.16); therefore most of the heat generated by it is dissipated directly to the
ventilation air. The variations in open enclosures are distinguished by the degree to which they protect
the motor from its operating environment. Totally enclosed motors are designed so there is no free
exchange (although there may be leakage) between the air circulating within the motor and the ambient
air. In all these designs, the heat generated by the motor is finally removed by some form of heat
exchange with the ambient air or a liquid coolant. The variations are in the means of cooling. The six
widely used enclosures and their distinguishing features are

1. Open drip-prooj(ODP)-Open enclosure (Fig. 24.16), whose air openings are designed so liquid drips or
particles falling on the motor at angles from vertical to 15 deg off vertical will not affect the operation of
the motor. Open drip-proof motors are usually used only in relatively clean indoors installations.
2. Weather-protected type I (WP I)-Open enclosure with internal ventilation circuits designed to minimize
the entrance of rain and the like into the electrical parts of the motor.
3. Weather-protected type 1/ (WP /I)-Similar to WP I but with the ventilation circuits designed to discharge
ingested rain and the like without having it enter the internal ventilating passages that lead directly to the
electrical parts of the motor (Fig. 24.17).
4. Totally enclosed jan-cooled (TEFC)-The motor frame does not have air openings so there is no free
exchange of air between the inside and outside of the motor. A shaft mounted fan circulates ambient air
over the outside of the frame to dissipate heat (Fig. 24.10). TEFC motors are widely used for outdoor
installations up to the limit of their cooling capacity.
5. Totally enclosed air-to-air cooled (TEMC)-A totally enclosed motor equipped with an air-to-air heat
exchanger to cool the air within the frame. Circulation of the air within the frame is by internal shaft mounted
fans. Ambient air is circulated through the heat exchanger by either an external shaft mounted fan (Fig.
24.18) or a separate fan. TEAAC motors have higher thermal capacity than TEFC motors, and are used
for larger motors installed outdoors.

Sleeve
Bearing
Detail

12
13

Item Description Item Descrlpllon


1 Rotor Assembly 9 Ball Bearing
2 Stator Assembly 10 Rotor Shaft
3 Stator Yoke II TermInal Bo.
4 Beanng Housing 12 Dust Cap
5 Air Deflector 13 Dust Washer
6 End Cap 14 Stator Coils
7 Shaft Seal IS Yoke Baffle Ring
11 8 Set Screw

Fig. 24.16 Section of open drip-proof (ODP) motor.


(Courtesy Siemens Energy & Automation, Inc.)
Drivers 641

'f :
...................... .1. .................... ..
:::::::::::::::::::L.!... C::::·.:·.::·.:·:.:~

Fig. 24.17 Section of weather-proof II (WP-II) motor.


(Courtesy Siemens Energy & Automation, Inc.)

6. Totally enclosed water-to-air cooled (TEWAC)-Conceptually similar to TEAAC, but using one or more
air-to-water heat exchangers instead of air to air. TEWAC enclosures are typically used for large motors
installed outdoors and indoors when noise is a major consideration.
7. Explosion proof (ExP)-A special version of totally enclosed motor designed to withstand an internal
explosion of a gas or vapor and to prevent ignition of the gas or vapor surrounding the motor. The means
of cooling are the same as for other totally enclosed motors. The specific requirements for explosion proof
motors vary with the authority having jurisdiction over the installation, therefore it is important to clearly
identify the authority to the motor manufacturer when inquiring and ordering the motor.

Special Construction
Severe operating conditions, noise limitations, or an unusual drive arrangement often require special
motor construction. The more common requirements are

1. Special insulation-In drip-proof motors, insulation for increased resistance to moisture, conducting dust,
or abrasive dust.
2. Tropical treatment-For motors installed in a tropical environment, insulation suitable for increased moisture,
treatment of all internal surfaces against fungus, insects and corrosion, and screens over all openings.
3. Noise reduction-Various treatments ranging from unidirectional fans to silencing enclosures to lower the
noise generated by air cooled motors installed in buildings or urban areas.
4. Double-extended shaft-An additional shaft extension at the normal nondrive end for driving another pump
or allowing dual drive in tandem.
642 Drivers

Item De.crlptlon
1 Top Cover
2 Stator Coils
3 Bearing Bracket
4 Air Deflector
5 Bearing Cap
6 Split Sleeve Bearing
7 Oil Guard
8 Rotor Shaft
9 Oil Ring
10 Stator Yoke
11 Rotor Fan
12 Rotor
13 Main Terminal Box
14 Snap Ring
15 Fan Housing
16 Grid Cover
17 Fan·External

13

Ball Bearing and Rolier Bearing Detail

Fig. 24.18 Section of totally enclosed air-to-air cooled (TEAAC) motor.


(Courtesy Siemens Energy & Automation, Inc.)

Hazardous Locations

Electric motors are frequently installed in areas where the atmosphere contains or may at times
contain flammable or explosive gases, vapors or dusts. There are now a set of universally accepted area
classifications used to designate the hazardous atmospheres and set the maximum temperature of any
surface operating in each (see Table 24.4). Within each class there are two divisions that afford some
latitude in the choice of motor enclosure:
Class J, division 1. The atmosphere is or may be hazardous under normal operating conditions,
including during normal maintenance. An explosion proof enclosure is mandatory for all class I, division
1 areas.
Class J, division 2. The atmosphere may become hazardous only under abnormal conditions. Generally
motors of standard enclosures can be installed in division 2 areas provided the motor normally does not
generate sparks. Under this provision standard open or totally enclosed squirrel cage motors are acceptable,
whereas designs employing open slip rings or commutators are not.
The two divisions in class II follow the same principle as class I: when the hazard is normally present,
a dust-ignition proof enclosure must be used; when it is present only under abnormal circumstances, a
standard totally enclosed motor is allowable but with a limit on its external temperature.
In all cases involving hazardous locations, the plant designer determines the area classification,
and the motor manufacturer, using information provided by the pump manufacturer, determines the
appropriate enclosure.
Drivers 643

Table 24.4 Hazardous Atmospheres

Maximum
temperature
Class Group (OC;oP) Atmosphere

I A 280/536 Acetylene
B 280/536 Hydrogen, manufactured gas
I C 160/320 Ethyl ether vapor
D 215/490 Acetone, acryolnitrile, alcohol, ammonia, benzine,
benzol, butane, dichloride, ethylene, gasoline, hexane,
lacquer, solvent vapors, naphtha, natural gas,
propane, propylene, styrene, vinyl acetate, vinyl
chloride, or xylenes
II E 165/329 Metal dust including magnesium and aluminum or their
commercial alloys
II P 165/329 Carbon black, coal or coke dust
II G 165/329 Flour, starch or grain dust

Mounting, Shaft, and Bearings

Most the motors used to drive centrifugal pumps are horizontal, foot-mounted, with a cylindrical
shaft extension suitable for either direct coupling to the pump or belt drive. The bearings of these motors
are generally antifriction, grease or oil mist lubricated, with one of the bearings acting as a thrust bearing
(usually the drive-end bearing to minimize shaft end movement caused by thermal expansion). Sleeve
journal bearings are used when required by the operating conditions (speed and load) or purchaser's
specification. Sleeve bearing motors normally do not have a thrust bearing, so the pump has to axially
locate the motor's rotor by the use of a limited-end-float coupling. Small- to medium-size motors have
the bearing housings incorporated into the end of the frame (Fig. 24.10). Large motors generally have
pedestal bearings (Fig. 24.13).
Vertical motors are flanged and are usually mounted directly on the pump. Their shafts are either hollow
(Fig. 24.19) or solid (Fig. 24.20). Hollow-shaft motors are rigidly coupled to the pump via an extension of
the pump shaft, which passes through the motor to a rigid coupling at the top of the motor. Solid-shaft motors
are either rigidly or flexibly coupled to the pump depending on the thrust bearing arrangement. In the United
States, the normal practice is to rigidly couple the pump to the motor and to have the motor accommodate
the pump's thrust load whenever possible. Such motors therefore have a high-capacity thrust bearing, anti-
friction up to the limit of load and speed, and tilting pad hydrodynamic beyond that. The thrust bearing is
better located at the top of the motor (Fig. 24.19) where it is accessible, heat dissipation is better, and there
is less risk of fire in pumps handling flammable liquids in the event the thrust bearing fails. Outside the
United States, vertical pumps are usually flexibly coupled to the motor. The motors therefore are solid shaft,
and have thrust bearings sized only for the weight of the motor's rotor. In this arrangement, the pump has
a separate high-capacity thrust bearing. Antifriction bearing lubrication is by grease, oil occasionally, or oil
mist. Hydrodynamic bearings are oil lubricated.
Some pumps are close coupled to their motor (Fig. 24.20). In this arrangement, the motor's shaft
must be stiff enough to support the pump's impeller and any associated hydraulic loads, and its bearings
have to have the capacity to accommodate the sum of the pump and motor loads. When the pumping
temperature is above 1000 e (212°F), the bearing adjacent to the pump and motor's windings are subject
to additional heat transferred from the pump. Horizontal motors for close coupling are often foot mounted
with an additional mounting flange for attaching the pump.
644 Drivers

Fig. 24.19 Vertical hollow-shaft motor.


Used/or vertical turbine pumps and many other vertical wet-pit pumps. (Courtesy Westinghouse Electric Corp .)

ELECTRICAL CHARACTERISTICS OF MOTORS

Electric motors are made for standard voltages depending upon their sizes; for example, 30-hp, 1,800-
rpm, three-phase, 60-cycle motors are regularly offered for 208, 220,440, and 550 V. A motor can be
operated successfully, but not necessarily in accordance with standard guarantees, at voltages 10 percent
above or below its rated voltage (at the same frequency in the case of ac motors). Motors wound for
special voltages can be obtained at additional cost.
The frequency of the power supply varies in some ac installations, thereby causing a variation in the
operating speed and resulting in a variation of the operating characteristics of the pump. In such cases,
the maximum loading on the motor under the most adverse conditions must be checked carefully to
insure safe operation. Standard motors are usually designed to operate satisfactorily at 5 percent above
or below rated frequency. The effects of voltage and frequency variations on motor characteristics are
illustrated in Table 24.5. In a number of water works installations where variable pump output is required,
the frequency of the current is deliberately varied in order to vary the motor speed and thus the
characteristics of the pump.
Some ac motors are wound for use at either of two voltages, for example, 220 or 440 V, depending
Drivers 645

Fig. 24.20 Close-coupled pump for process service. (Courtesy BWIIP International)

on how the leads are connected. This construction is standard for some sizes and combinations of voltages
but is special for others. The voltage ratios in these dual-voltage motors are necessarily limited to certain
fixed values.
By using special designs and special materials, at increased cost, induction motors of more than 150
kW (200 hp) and synchronous motors can be built to operate more efficiently. With units that are to be
operated more or less constantly, the saving in power costs will, in many cases, warrant the increased
first cost of premium efficiency motors. If power costs are low, however, the increased first cost of
premium efficiency motors may not be justified.

STARTERS FOR ELECTRIC MOTORS

A starter for an electric motor connects the motor electrically to the source of supply. Manual starters
move the starter contactors manually; automatic or magnetic starters move them automatically, by means
of magnets, on receiving an impulse from a pilot device.

Direct Current Motor Starters


A knife switch can serve as a starter for a dc motor, but this would impose full voltage across the
motor terminals with a high current inrush. Although this is feasible with small motors, it is generally
desirable to have an adjustable resistance, a rheostat, connected in series with the motor armature to
~

Table 24.S Approximate Effects of Variations in Voltage and Frequency on A-C Motor Characteristics I ~
~
;::
Voltage Frequency
Characteristic 110 Percent 90 Percent 105 Percent 95 Percent

Torque
Starting and maximum running Increase 21 percent Decrease 19 percent Decrease 10 percent Increase 11 percent
Speed
Synchronous No change No change Increase 5 percent Decrease 5 percent
Full load Increase 1 percent Decrease 1.5 percent Increase 5 percent Decrease 5 percent
Percentage slip Decrease 17 percent Increase 23 percent Little change Little change
Efficiency
Full load Increase 0.5-1 point Decrease 2 points Slight increase Slight decrease
3/4 load Little change Little change Slight increase Slight decrease
1/2 load Decrease 1-2 points Increase 1-2 points Slight increase Slight decrease
Power factor
Full load Decrease 3 points Increase 1 point Slight increase Slight decrease
3/4 load Decrease 4 points Increase 2-3 points Slight increase Slight decrease
1/2 load Decrease 5-6 points Increase 4-5 points Slight increase Slight decrease
Current
Starting Increase 10-12 percent Decrease 10-12 percent Decrease 5-6 percent Increase 5-6 percent
Full load Decrease 7 percent Increase II percent Slight decrease Slight increase
Temperature rise Decrease 3-4°C Increase 6-7°C Slight decrease Slight increase
Maximum overload capacity Increase 21 percent Decrease 19 percent Slight decrease Slight increase
Magnetic noise Slight increase Slight decrease Slight decrease Slight increase
'Courtesy General Electric Co.
Drivers 647

Fig. 24.21 Multistep rheostat used as dc motor starter.


(Courtesy General Electric Co.)

limit the initial current. As the motor accelerates, the resistance is cut out until full voltage is applied
to the armature and the motor is running at full speed. If this resistance is used solely for starting, it is
called a starting rheostat (Fig. 24.21). If the motor is run at reduced speed by leaving some of the
resistance in the armature circuit, the rheostat must be designed to dissipate the heat energy used up in
the resistance, thereby making the rheostat more expensive.

Alternating Current Motor Starters


Starters for single-phase, ac motors are usually some form of an across-the-line starter. Where this
type cannot be used because of the limited line capacity, one of the following special starters is used
to reduce the voltage initially applied to the motor terminals and to increase it in steps until full voltage
is reached.
Squirrel cage motor starters. Three forms of starters are in broad use for polyphase squirrel cage
motors: full voltage or across-the-line, part winding, and reduced voltage. Within the category of reduced
voltage starters are primary resistor, star (wye)/delta and autotransformer. The characteristics of these
various starters are

1. Full-voltage-The simplest, least expensive, and most commonly used form of starter. With full voltage
applied across the motor terminals, the starting current is on the order of six times full-load current for the
motor designs typically used for centrifugal pump drives. In installations with limited line size or grid
capacity, the high starting current may not be acceptable.
2. Part winding-An economical means of lowering the starting current for low-inertial loads such as centrifugal
pumps. The starter initially energizes half of the winding to produce a starting torque typically one-half to
two-thirds the normal value while drawing about two-thirds the full-voltage starting current. Because there
is no interruption in the applied voltage when the motor is switched to full winding, the "transition" current
surge is minimal. Unless the motor is dual voltage, part winding starting requires special winding connections
that must be specified when the motor is ordered.
3. Primary resistance-The voltage across the motor's terminals is lowered by introducing a resistance into
each phase for starting. After a fixed time, the resistors are short circuited to apply full voltage to the motor's
648 Drivers

terminals. Although not as efficient as other methods of reduced voltage starting, primary resistor does lower
the starting current and does not have a high transient current surge. Primary resistance starters can be one,
two, or three step depending on the allowable current fluctuations during starting.
4. Star/delta-By starting with the windings connected in star (wye), the starting voltage is reduced to 58
percent of line voltage, which reduces the motor's starting torque and stating current to one-third of the
full-voltage values. The motor is connected in delta once it has accelerated. Open transition star/delta starters
are less expensive than closed, but produce a higher transient current in the switch from star to delta.
5. Auto transformer-Widely used because of their efficiency and flexibility, auto transformer starters generally
have taps to allow 80, 65, or 50 percent of line voltage to be applied to the motor terminals for starting.
Which tap to use depends on a compromise between starting current and acceleration time. The usual tap
for motors driving centrifugal pumps is 65 percent, which produces 42 percent of full-voltage torque while
lowering the starting current to 45 percent of that for full voltage. To avoid a large transient current surge
during the switch to full voltage, closed transition switching should be used.

Wound-rotor motor starters. Starters for wound-rotor motors have a primary switch and resistance
for the three leads of the wound rotor. In starting, full voltage is applied to the stator with the maximum
resistance cut into the wound-rotor circuit. As the motor speeds up, the resistance in the wound-rotor
legs is cut out until, when all resistance is cut out, the motor is running at full speed. With the proper
choice of resistance and steps, the wound-rotor motor can be started with very little increased current
demand and is, therefore, very valuable when the power supply is limited. The resistance used in the
wound-rotor circuit can be designed either for starting or for both starting and speed-regulating use.

Synchronous-Motor Starters
Starters for synchronous motors are basically the same as those used for squirrel cage motors except
for the addition of a mechanism to apply the direct current to the motor field circuit when the motor
comes up to the pull-in speed.

Manual and Automatic Starters


Except for the action of protective devices built into some manual starters, motors controlled by these
starters are started and stopped manUally. With automatic starters, motors are started or stopped either
by momentary- or maintained-contact pilot devices, or manually by pushbuttons.

Protective Devices
Protective devices in electric starters guard against overload and undervoltage. Heating elements are
often incorporated in starters to protect against damage from overload. A sustained excess current will
cause the contactors to trip out (Fig. 24.22). The mechanism must then be reset manually before the
contacts can be reengaged.
After being started, a three-phase squirrel cage motor will continue to run on one phase if the other
two phases are disconnected. If the same load is carried on the motor, the current at single-phase operation
will be greatly increased and the motor may be seriously damaged. Some form of overload protection
should be incorporated in any motor installation.
Two methods are employed to protect the motor against low voltage. One is called undervoltage
protection and is used with momentary-contact pilot devices. In case of undervoltage, the holding coil
in the control releases, and the contactors are opened. The contactors will not reengage when full voltage
is resumed, however, as the opening of the contactors has opened the holding coil circuit. The other
Drivers 649

Fig. 24.22 Magnetic full-voltage starter with magnet to close contacts.


Heater elements cause overload relays to open contacts if current becomes excessive.
(Courtesy General Electric Co.)

method is called undervoltage release and is used with maintained-contact pilot devices. In this form,
undervoltage causes the holding coil to release the contactors, which then open. The motor stops, but
the control circuit does not open. When the voltage is reestablished, the control will close the motor
line contactors and the motor will start. In some applications, undervoltage release is dangerous and
undervoltage protection should be used. If the pilot device is a maintained-contact type, undervoltage
protection can be used by adding relays to the control circuit.
The electrical installation should provide protection against a possible short circuit as well as providing
overload and undervoltage protection for the motor. Most starters (both high- and low-voltage) have an
interrupting capacity of ten times the rated motor current. However, most high voltage circuits can
produce short-circuit currents far in excess of this value. For both personnel and property safety, equipment
should provide adequate protection in case of a short circuit.

Starting Torque Characteristics of


Centrifugal Pumps
The proper selection of a driver and, in the case of motors, of its control, requires consideration of the
starting torque characteristics of the machine to be driven. Centrifugal pumps have favorable starting
torque characteristics and rarely cause concern; however, when the pump motor is connected to a line
of limited capacity, the starting torque characteristics of the pump may have to be analyzed in detail.
650 Drivers

DEFINITION OF TORQUE

Torque is defined as "the moment of a system of forces that causes rotation" and is usually expressed
in tenns of force at a radius, Newtons at 1 meter or Newton-meter (Nm) in SI metric units and pounds
at 1 foot or pound-foot (lb-ft) in US units. (Small torques in US units are often expressed in ounce-
inches or in pound-inches to avoid minute fractions.) Figure 24.23 illustrates the moment involved in a
l-lb force acting at a I-ft radius, using as an illustration a wire wound on a I-ft radius cylindrical drum
with a l-lb pull on the wire. Torque is independent of rotation; it is the moment tending to cause or
causing the rotation. If, in Fig. 24.23, the wire were being pulled off the drum at a rate which would
cause one revolution of the drum in 1 minute, the distance through which the l-lb force would have
acted in the minute would be equal to the circumference of the drum or 27t x 1 ft. The work done,
therefore, would have been 27t x 1 ft x 1 Ib or 27t ft-lb. It can be shown that

21tTn Tn
hp = 33,000 =5,250
or

T = 5,250 hp
n

where T = torque, in pounds-feet


n = speed, in revolutions per minute
hp =horsepower (33,000 foot-pounds per minute).

Similarly, in metric units a force of 1 N acting on a drum of 1 m radius would produce a torque of 1
Nm and the work done would be 27tNm per revolution of the drum. From this the power is

21tTn Tn
kW = 60,000 = 9,550
to give torque

T= 9,550kW
n
where T =torque, in Nm
kW = 60,000 Nm/minute

ACCELERATING TORQUE

To start a load and bring it up to its running speed, the driver must deliver more torque than required
by the driven machine at any speed up to its normal running speed. The difference between the torque
required by the load and the torque developed by the driver is called the net accelerating torque.
Acceleration ceases when the torque required by the driven machine equals the torque developed by th.e
driver. If the required torque exceeds the developed torque, the unit will slow down until the required
torque again equals the torque developed by the driver. If the torque of the load at any reduced speed
is greater than the torque developed by the driver, the driver will stall or stop. It is, therefore, a basic
Drivers 651

Fig.24.23 Torque of 1 Ib at I-ft radius-lIb-ft.

requirement that the driver torque must be greater than the torque required by the load at all speeds
between rest and full speed if the driver is to start the load and accelerate it to full speed.
The time required to start a load and bring it to full speed depends on two factors: the margin of
torque available, that is, the net accelerating torque, and the mass of the rotating parts to be accelerated.
This mass has a flywheel effect; the greater the flywheel effect, the longer the time required to accelerate
the rotating parts with the same accelerating torque. The time required to bring a load from one speed
to another can be calculated in metric units from the formula

where t = time, in seconds


J =mD2/4, mass moment of inertia of all rotating parts and liquid, referred to driver speed, in kgm 2
m =mass, in kg
D = diameter of gyration, in m
nl =original speed, in rpm
n2 =final speed, in rpm
T =average available accelerating torque, in Nm

For US units the calculation is made using

WK2(n2 - nl)
t= 308xT

where WK 2 =flywheel effect of the rotating parts of the driver and load, referred to driver speed, in Ibft2
T =average accelerating torque available, in Ibft
The most adverse type of load to be accelerated, therefore, is one that requires a high torque to drive
it at all speeds from rest to full speed and has a large flywheel effect, such as a reciprocating compressor.
Centrifugal pumps, however, have very low J (WK2; called WR 2 in the past) values and low starting
torque characteristics, the latter varying considerably with the type of pump and the method of starting.
For this reason the high-starting torque motors necessary for machines with large inertia and high torque
values are not required for centrifugal pump applications. There are, however, many cases where the
652 Drivers

starting current inrush of electric-motor-driven pumps must be kept to a minimum and the pump torques
during the starting cycle must be known to make the proper motor and control selections.

DEFINITION OF TERMS

Motor manufacturers use the following definitions of torque terms (see also Fig. 24.11). "Full load
torque" is the turning moment exerted by a motor at rated speed and load. "Starting torque," "locked-
rotor torque," or "static torque" is the minimum torque developed by the motor for any angular position
of the rotor at standstill with rated voltage and frequency applied to the motor terminal. "Pull-up torque"
of an ac motor is the minimum torque it develops during the period of acceleration from rest to full
speed with rated voltage and frequency.
"Pull-out torque," "maximum running torque," or "break-down torque" of an ac motor is the maximum
torque an induction motor will develop with rated voltage and frequency, without an abrupt drop in
speed, or that torque a synchronous motor can sustain at synchronous speed.
"Pull-in torque" for a synchronous motor is the maximum constant torque under which the motor
will pull its connected inertial load into synchronism at rated voltage and frequency, when field excitation
is applied.
Electrical manufacturers usually express the various torque values of a motor as percentages of full-
load torques. Knowing the horsepower and speed, these can be easily converted to Nm (lbft) if necessary.
Torque values are also given on the basis of rated voltage and, in case of ac motors, on the basis of
rated frequency being supplied to the motor. If reduced voltage is used for starting, the torques are
reduced proportionally to the square of the applied voltage. Thus, 0.8 or 80 percent applied voltage
reduces the developed torques to 0.64 or 64 percent of their full-voltage values. There is one exception
to torque varying as the square of the applied voltage. The pull-out torque of a synchronous motor varies
directly with the applied voltage.

ANALYSIS OF TORQUE

Theoretically, any torque, no matter how small, would start a centrifugal pump from rest; in practice,
this is not quite true because of friction in the bearings and shaft seals. In the case of the former,
specifically the case of babbit sleeve bearings, the oil between the shaft, and the bearing is squeezed
out, resulting in metal-to-metal contact. In shaft seals, there is static friction between the packing and
shaft of packed box seals, or between the seal faces of axial face (mechanical) seals. The magnitude of
this breakaway torque varies with the design of the pump bearings (very large pumps having "jacking
oil" to develop an oil firm between the shaft and bearing before starting), and the type and condition of
the shaft seals. The torque required to start from rest can be safely taken to be less than 20 percent of
the torque required to drive the pump at design conditions. Usually 15 percent is used for horizontal
sleeve bearing pumps and 10 percent for ball bearing pumps. It is also impossible to make any general
rule as to how quickly the adverse conditions causing this breakaway torque disappear when the pump
begins to rotate. It is safe to assume that the breakaway torque decreases as the speed increases, until
it reaches practically zero at 10 to 20 percent speed.

Low-Specific-Speed Pumps
Most centrifugal pump applications involve relatively low specific speed pumps, which are usually
started against a closed gate valve. Figure 24.24 illustrates the characteristics of such a pump. If the
Drivers 653

200

,c;>.l 'r '(


yl.~ t-.........
~
80
15 0
...J:j A - APACITy
/ ~
:Y' I"-
~ 70
r-..... ...........
I-

U
10.1
10.1
"-
/ ~
560
0. ~
0 I ,.,.....- ~ \
>-' 50 ~
100
'\
<l

/ ¢' / '
U
:z ILl
1:
w
G 4
V/ \
.J
<l
I-
~
0
I-
5 o~fo""'
1
/
0.
1:
III
10

0 0 1
o 10 20
CAPACITY. IN 100 GPM

Fig,24.24 Constant speed (1,770 rpm) characteristics of 6 in discharge, 8 in suction (6 x 8)


double-suction pump.

pump were operated against a closed gate valve, the power required to operate it at any given speed
would be equal to the shutoff horsepower at rated speed times the cube of the ratio of the given speed
to the rated speed. Therefore, the torque required at that given speed, with the pump operating at shutoff
(against a closed gate valve), varies with the square of the ratio of the two speeds (Fig. 24.25). The
broken line in Fig. 24.25 indicates the effect of initial friction (15 percent of normal torque at zero speed
and zero torque at 10 percent of normal speed) and the solid line directly above it indicates the resulting

100
/
/
,/'
,..... /
L V
o '\.
~"
-~
o 10 II 20
SPEED. IN 100 RPM

Fig. 24.25 Torque curve of pump of Fig. 24.24 when started against closed gate valve.
654 Drivers

total torque-speed curve. This total torque is the torque required to drive the pump and provides no
excess for acceleration.
This same pump could be installed in the system in Fig. 24.26 and started against a check valve with
an open gate valve. During the starting cycle, the pump will operate against shutoff until it reaches the
speed at which the shutoff head developed by the pump exceeds the static head S, and flow will start
through the pipe. Some excess head is required to establish flow through the pipe line, but this can be
ignored in this analysis of starting torque, and it can be assumed that the pump capacity and head during
the rest of the starting cycle follow the system-head curve. In case the power required to drive the pump
between shutoff and rated conditions is maximum at rated conditions, the necessity of an excess head
to accelerate the water in the pipe line would result in less torque being required. It would only have
an adverse effect in the case of high specific speed pumps (Fig. 24.27) when the pump power requirements
are maximum at shutoff. These high-specific-speed pumps are basically applicable to relatively low
heads that would not normally be encountered in a system with a long pipe line and, therefore, seldom
involve the problem of accelerating a mass of water.
To help clarify the resulting power and torques involved, Fig. 24.24 has been redrawn, and the speed,
power, and torque required for operation against the system head from shutoff to 327 m3/hr (1,440 gpm)
has been shown in Fig. 24.28. In starting the pump under these conditions, the torques would be the
same as shown in Fig. 24.25 until the unit reached the speed of 1,427 rpm, when the pump develops a
shutoff head of 30.5 m (100 ft)--equal to the static head. In this installation, as the speed is increased
further, the pump will deliver water through the system until it reaches a 327 m 3/hr (1,400-gpm) capacity
at full operating speed. The resulting speed-torque curve of the pump operating in this installation is
shown in Fig. 24.29.
If the same pump is used on a system entirely made up of frictional head, with the same operating
conditions (318 m 3/hr [1,400 gpm], 39.3 m [129 ft] total head) as described previously, and the pump
is started against a quick-opening check valve with the gate valve open, flow would start immediately.
For all practical purposes, again ignoring the acceleration of the water column, the flow would be directly
proportional to the speed, the head would be proportional to the square of the speed, the power would
be proportional to the cube of the speed, and the torque would be proportional to the square of the speed.
On that basis the torque-speed curve would be as shown in Fig. 24.30.

S. 100 FT

{
.c7 P

- I
Fig. 24.26 Application for pump characterized by Fig. 24.24.
Drivers 655

50

46

160 80 40 ~

140 70 35 \ /
'f'"
1\
120 1-60
z
1-30 \'1\ ~ EFFICIENCY \
~ "- /
LU

\
LU LU
100 ~50 II. 25
z
~~
LU
Q.
0 HEAD-CAPACITY \
8 >=" 40

-\-
Q. ~20
u
/~~ ~~
J:
m z J:

60 ~30
u ct 15
oJ
.- ~ --
u:II. j ~S-Ts"TEM ~
0
I- HEAD
0
I- 14-FT STATIC ~
40 LU 20 10
/ \r-..
10 5 J
20

o o
o
V 2 4 6 8 10 12 14 16
\
18 20
CAPACITY, IN 1,000 GPM

Fig. 24.27 Constant speed characteristics of 30 in discharge propeller pump at 700 rpm.
I ,800
ly
c,' 1,700

200 ~ g:
~

~
1,600
'<..~\)
90 t.1'!I r\ I,500

80 ~
S~~
'JJ~
t.~
~C'<
-..... I ,400
150
~
/
I-
z 70 I-
~

~k
W W
U
60
W
LL
rI I\.Q L..-----' ~
K
0:: 200
W
~ ~
b~~ V \

11~
• 50 ci 100 l-
>-
u
~ 40
«
w
I
/ 'Or\9 y-<V~\) ~<;>\) '\ ,
LL
CD
..J 150
U
LL
30
..J
«
l- / ~~
s""' s-<" ~~ \ ~
LL 0 <iO~ ",.J.. uS
50
~
W I- :::>
20 'O~ «:t 0
a:: 100
I-- ~<v
~
0
~
I 10 ",O~~ I-

V..- V'"
aJ

0 o 50
o 5 10 15 20
CAPACITY, IN 100 GPM

Fig. 24.28 Analysis of 6 x 8 pump in Fig. 24.26.


656 Drivers

200

150
I
m
...J
I
l-
II
II:
~ 100 I
w
~
o /
a:
o
I- /
50 /)
v
'"
./

v p

o
o
- ~
5 10 15 20
SPEED, IN 100 RPM

Fig. 24.29 Torque characteristics of 6 x 8 pump when starting against check valve in system shown
in Fig. 24.26.
200

/
/
150
V
/
I-
lo- /
CD
..J

100
V
/
~
.
w
;:)
0
a:
0
l- /
V
SO /
V
/

"V
V
5 10
SPEED, IN 100 RPM
15

Fig. 24.30 Torque characteristics of 6 x 8 pump when started in wholly frictional system.
20
Drivers 657

8
50

40
-- ~.II
C4P4
~~
.......... ~ -.......
I-
Z
UJ
u 70 I-
l? r----.. ~ t'\.
0::
UJ
UJ / I'--- ........... \

-- / ~ --
UJ U.
60 z 3 0

-.\
Cl.
~~
>-'" o<t () ~
u
50 <,.' ~
z tj
r--- r-- h
UJ
UJ
U I _~.e- ~
iL 40 <i 2
u.
UJ
30
I-
oI- "'\
Cl.
:r 20 0 / \,
/
III

10
0 oV
o 10 20 30 40 50 60 70 80
CAPACITY, IN 100 GPM

Fig. 24.31 Constant speed characteristics of 16 in volute pump with mixed-flow type impeller requiring same
power at shutoff as at reduced heads.

Using this same method of analysis, approximate torque-speed curves can be established for any
centrifugal pump if the pump characteristics and method of starting are known. Figure 24.31 shows the
characteristics of a medium specific speed pump, which requires the same horsepower at shutoff and at
the point of maximum efficiency. The speed-torque curve for this pump will be practically the same
whether it was started against a closed gate valve or against a check valve. In the latter case it would
make little difference whether the system was all static, all friction, or part static and part friction (Fig.
24.32). Curve A shows the torque resulting when starting against a closed gate valve and also when
starting against a check valve in an entirely frictional system that would have a head of 7.62 m (25 ft)
at 1447 m3jhr (6,370 gpm) flow. If the pump were operated against a system consisting solely of a 7.62
m (25 ft) static head, the torque-speed curve would be like curve B. For any system with combined
friction and static heads, the torque-speed curve would be some curve falling between curves A and B.

Propeller-Type Pumps
The propeller-type pump now used extensively for low-head applications is a high specific speed
type and may have a high-speed mixed-flow or true-axial flow impeller. The typical characteristics of
one of these designs with an axial-flow impeller are shown in Fig. 24.27. The shutoff horsepower is
approximately 2.5 times that required at maximum efficiency. Whereas Fig. 24.27 shows only a flat
portion in the head-capacity characteristics, many of these axial-flow designs, especially those over 9,000
specific speed, have an unstable portion in their head-capacity curves. For these two reasons, pumps of
this type are applied where the maximum head will be below any head in the unstable range, and the
pumps are either installed on a siphon system or started against a check or flap valve. This propeller-
type pump might be one of several installed as shown in Fig. 24.33 with maximum discharge level at
elevation 6.40 m (21.0 ft) and minimum suction level at elevation 1.52 m (5.0 ft). The pumps would
658 Drivers

300 /
/
/
250
~

kJ
A ___

I- II
/t-
l1. 200
I
m B
..J
~
W /
:::>
0
a:
150
KJ
/
0
I-

100
/
/
/
50
/
\ /
o \V
o 2 3 4 5 6 7 8 9 10
SPEED, IN 100 RPM

Fig.24.32 Torque characteristics of 16 in volute pump shown in Fig. 24.3\.


Started against closed gate value or wholly frictional system (A) and started against theoretical 7.6 m (25 ft)
static-head system with no friction component (B).

probably be controlled by float switches, which would stop all pumps when the suction level was pumped
down to elevation 1.52 m (5.0 ft). The various pumps would normally be started at different water levels
in the suction well. Assuming that one pump was arranged to start when the suction level reached
elevation 2.13 m (7.0 ft), that pump, if maximum discharge water level prevailed, would operate against
a 4.3 m (14 ft) static head.
Two possibilities exist with such an installation. The pump might be empty above elevation 2.13 m
(7.0 ft), or, if the pump had been in service only shortly before, it might be completely filled with water.
If the pump is filled, the impeller, when it is started, will not deliver any water until after it reaches a
speed at which the shutoff head exceeds the static head. The resulting torque curve would be as indicated
in Fig. 24.34. If the pump above elevation 2.13 m (7.0 ft) is empty, the impeller will immediately begin
to fill the column pipe with water. Without a detailed study of the time necessary for the motor to reach
Drivers 659

21.0 FJ

7.0 FT

5.0FT

Fig. 24.33 Installation of 30 in propeller pump of Fig. 24.27 showing water levels.

various speeds, it would be difficult to predict the torque-speed curve of the pump. The curve would,
however, have torque values lower than those shown in Fig. 24.34 until the column pipe, discharge
elbow, and discharge pipe were filled, at which time the curves would be identical.
The high-specific-speed, axial-flow pump has the most adverse starting torque characteristics of any
centrifugal pump type but offers no serious problem for any normal type of drive when starting against
a check or flap valve with a relatively short discharge line. This type of pump, however, can offer a
more serious problem when installed in a siphon system in which the establishment of the siphon depends
on the flow washing out the entrapped air. If the pump described in Fig. 24.27 and 24.33 has its discharge
higher than elevation 5.49 m (18.0 ft) with a siphon discharge, the pump column will fill until water
begins flowing over the low point of the siphon as the pump speeds up in the starting period, and, as
pumping continues, with a proper design of siphon, the air will be washed out, the siphon established
and the head will drop to that of the system. Here again the required pump torque at various speeds
would be tied in with the time that the motor would require to reach the various speeds. The pump
torque would be affected by the time, which affects the water level in the column, the operating head,
and, therefore, the pump torque. With a relatively high siphon leg, it is possible to have a torque curve
rising to a peak at a speed less than rated speed. In an extreme case with a very short starting period,
660 Drivers

500
/

400 1/
f---
/
~
lJ..
I /
I/
ID 300
~
Z
W
::J
0

/
a: 200
0
~

100 /
L
v
o Iv
o 2 3 4 5 6 7 8
SPEED, IN 100 RPM

Fig. 24.34 Torque characteristics of 30 in propeller pump when starting against 4.3 m (14 ft) static head with
pump full of water.

the required pump torque when the unit reaches rated speed could be greater than at rated conditions
because flow would not have been established. Such an unusual case would require the combined study
of the pump and motor torques in order to establish the approximate starting torque of the pump.
Vertical turbine pumps of the type used in deep-well and open-pit, high-head installations usually
have impellers of the mixed-flow type and are built in a number of stages. Some designs have shutoff
horsepower lower than required at rated capacity, whereas a few designs have shutoff horsepower greater
than required at rated capacity.

Cone-Type Stop Valves


Many installations employ cone-type stop valves, which function as a combined check and gate valve.
In practically all installations of motor-driven pumps, the time required for the motor to bring the unit
to full speed after the shutoff head reaches a value equal to the head on the discharge side of the valve
is very short. The travel of the valve is so small during this interval that the starting of the unit is, in
effect, the same as against a closed gate valve. This same effect could result, in whole or in part, with
other types of slow-opening check valves. In such installations, however, it is best practice to select and
analyze the starting cycle as for an instantaneously opening check valve and for no inertia effect in the
hydraulic system. Systems in which propeller pumps are used should employ quick-opening check valves
or flap valves or have a siphon discharge.

Specialized Applications
Centrifugal pumps require a very low torque to start from rest, and the torque required to drive them
does not reach a maximum value until maximum speed is attained. Situations in which the starting
Drivers 661

torque characteristics of a centrifugal pump is of any concern, therefore, are rarely encountered. The
situations in which starting torque characteristics will be of concern are as follows: (1) when the limited
size of the transmission line supplying a motor driving a pump makes a detailed study of the electrical
installation necessary to select the proper type of motor and motor control and (2) when synchronous
motors are to be used, and the pull-in torque of the motor has to be compared with the torque that will
be required by the pump.
Squirrel cage motors. Normally, a squirrel cage motor develops much greater torque at lower speeds
than required by a pump. As the pump has a small J (WK2) value (flywheel effect in kgm2 or Ibft2), the
time required for the unit to reach full speed is very short. For the low-specific-speed pump discussed
earlier (but for I,755-rpm rated speed), Fig. 24.35 shows the starting torque characteristics of the pump:
(1) curve A for starting against a closed gate valve, (2) curve B for starting against a check valve on a
system having a static head component, and (3) curve C for starting against a check valve on a system
involving all friction head. All the torques are shown in percentages of the rated motor torque. Figure
24.35 also shows the torque developed by a 45 kW (60 hp), I,800-rpm squirrel cage motor plotted on
the same basis both for 100 and 80 percent voltage. The excess torque produced by the motor over that
required to drive the pump at the lower speeds is very apparent.
Synchronous motors. In the case of synchronous motors, the starting characteristics are dependent
upon auxiliary windings that are built so they will bring the motor and its connected load up to 95
percent rated speed, at which speed the load is transferred to the regular windings.

2Z0r--'---'---.--~--r--'---r--.---r-~

ZOO~-+--~--r--+--;---~-+?~
~~~--1
~V 1\
..... 18~--+---~~~-+--~--~~~_~-4l~\~--4
J MDTOR TORQUE ~I / '"
a FULL VOLTAGE ,..V
a:
o
.-
a: 140 ~~r-~---+--·-+--~--~--+---~~--~
o.-
o 120 1---+-~f---+---+-+--+-----=-1'-V'
--H\r--+-I---i
~ MOTOR TORQUE 80 ~ /" \
Q PER CENT VOlT~""""'-
W
~
IOOt:::t:::::J===r=-t"""""'-r - I I -t1j1P"-----1
0::
.-Z 80r-~r-~---+---r--;---~--+---~~--_1

W
u
~~7
a:
~ 4"
~v l V~
20 r-~f--f---+----7'.c;.;r-:ti "to~Q\) :,...0'

o~ --"~
o 2 4 6 8 10 12 14 16 18 20
SPEEO, IN 1
0 0 RPM

Fig. 24.35 Starting torque characteristics of double-suction pump of approximately 1,700 specific speed.
Three different applications compared against torque characteristics of a normal torque squirrel cage induction
motor for both 100 and 80 percent voltage.
662 Drivers

100

w 80
::;)
Ii.. 0
00:: 80
1- 0
zl-
we
U w 40
0::1-
w<t
a.. 0:: 20

a
70 80 90 100

PER CENT OF RATED SPEED

Fig. 24.36 Partial curve of starting torque characteristics of double-suction pump of approximately
1,700 specific speed.
Showing torque requirements at 95 percent speed for three different applications.

The pull-in torque that must be developed by a synchronous motor driving a pump is, therefore, the
torque required by the pump at 95 percent speed. Using the same illustration of the low-specific-speed
pump as in Fig. 24.35, but on the basis of 1,800 rpm rated speed instead of 1,755 rpm, the starting
torque characteristics from 70 to 100 percent rated speed would be as shown in Fig. 24.36. At 95 percent
speed, curve A indicates 39 percent torque, curve B indicates 77 percent torque, and curve C indicates
81 percent torque needed to drive the pump. The synchronous motor used to drive this pump must,
therefore, be good for at least 39 percent pull-in torque if the pump is to be started against a closed gate
valve (curve A), at least 77 percent pull-in torque if the pump is to be started against a check valve on
a system involving a 30.5 m (100 ft) static head (curve B), and at least 81 percent pull-in torque if the
pump is to be started against a check valve on a wholly frictional system (curve C). Some additional
leeway would be allowed as the pump might produce a slight excess over the rated conditions.
Check valves. Occasionally the check valve of a centrifugal pump will fail to close when the unit is
shut down. When the system on which it operates has a static head component, or when other pumps
are delivering into the same discharge header, water will flow back through the pump, causing it to
operate as a water turbine. The pump will then run in the reverse direction unless prevented. The speed
at which it will run will be the speed at which the torque developed as a water turbine equals the torque
required to rotate the driver at that reverse speed. As electric motors require very little torque to drive,
the speed reached with motor-driven units will approach the runaway speed of the pump when operating
as a water turbine at the net head available under the resulting system conditions. With a net head equal
to the design or rated head as a pump, this runaway speed will vary with both the specific speed type
and the individual pump design. It might be expected to be between 100 and 150 percent of rated
pump speed.
In some installations, such as in a water works pumping plant with a number of other units in service,
the net head available to drive the pump as a water turbine will be equal to its rated head as a pump.
In other installations, such as a single pump delivering through a pipeline to a reservoir, the pumping
head is the static head plus friction head loss, whereas under reverse flow, the net head causing the
Drivers 663

w
:::> 120
0
i""'"
0::
0
100 / ~~
I-
0
w
/1' ""- ~
80
'\
I-
<l
I
0::
60 J J
/ \
lL.
0
I-
Z 40 J
U
w
/
1/ 'rV
0:: 20

/
w
a..
o
120 100 8060 40 20 0 20 40 60 80 100
PER CENT OF RATED SPEED
REVERSE FORWARD

Fig.24.37 Torque characteristics of double-suction pump of approximately 1,700 specific speed, from runaway
speed in reverse direction to rated speed in forward direction.
Head measured from suction to discharge being equal at all times to rated total head as a pump.

pump to run as a water turbine would be the static head minus the friction loss at the resulting capacity.
In cases where the net head is less than the pumping head, the runaway speed will be reduced.
When reverse flow occurs through a motor-driven pump, because of the failure of a check valve to
close or other reasons, it is dangerous to apply power to the motor to try to arrest its reverse runaway
speed and restore it to full speed in the forward direction. The nature of the risk varies with the method
of starting the motor. If starting is across-the-line, the pump and motor will be subjected to a significant
torsional shock, and may suffer severe structural damage. If by partial winding or reduced voltage, the
peak reverse torque developed by the pump may be more than the motor can develop and therefore
reverse rotation will continue until the starter switches to full voltage. At this point one of two things
will happen: torsional shock will cause structural damage or the high current being drawn by the motor
as it continues to slow the pump will trip the overload in the starter. Figure 24.37 shows the high torque
that must be overcome by the motor to carry one specific design of pump from runaway speed in the
reverse direction to full speed in the forward direction.

COMPARISON OF PUMP AND MOTOR TORQUES

The torque characteristics of a 45 kW (60 hp), I,800-rpm normal-starting-torque, squirrel cage motor
for full voltage and reduced voltage (80 percent tap of the autotransformer) is shown in Fig. 24.38
together with the pump torque curve from Fig. 24.25. With reduced voltage, the initial current inrush
is reduced to 64 percent of the full voltage inrush, but it is still possible for the motor to reach 1,770
rpm before the torque of the motor equals the pump torque. The point at which full voltage is applied
would depend on the timing of the starting cycle. In this case, if full voltage is applied any time after
the pump reaches 1,340 rpm, the second inrush as full voltage is applied would not exceed the initial
664 Drivers

r-- --...
600
CURRENT WITH
r--. r-..... FULL VOLTAGE

'" "
500

400 "-
!'....

300 r--- CURRENT ~-....... ""'" ........


~

200
80 PER CENT VOLTAGE
i'-.. r-.... \ 1\
"- ~
100
~
o
o 2 4 6 8 10 12 14 16 18 20

400

V
MOTOR TORQUE AT
300 r---C'ULL VOLTA::"'
V
./
v ~
l-

1\
,
V
ll..

en
-J
V
\
Z 200

w
. ~
~

;j MOTOR TORQUE AT
~ 80 PER CENT VOLTAGE
o
I-

100
V
V V

.kd:::.---
PUMP TORQUE
~

o ~
o 2 4 6 8 10 12 14 16 18 20
SPEEO,IN 100 RPM

Fig. 24.38 Torque-speed and percent current-speed curves of 45 kW (60 hp), 1,800-rpm, 3-phase, 60-cycle
squirrel cage motor both for full and 80 percent reduced voltage.
Drivers 665

inrush. With automatic reduced voltage starters, the application of full voltage is controlled by timing
relays. In most cases the timing relays are set so that the motor reaches the maximum speed possible
on reduced voltage (1,770 rpm in the example) before full voltage is applied. This is an advantage
because in order to apply full voltage, the motor has to be switched from the reduced voltage to full
voltage. As a result, the current drops to zero and a second inrush, the magnitude of which depends
upon the motor speed, results when full voltage is applied. In the case illustrated in Fig. 24.38, therefore,
the second inrush, if full voltage is applied at 1,000 rpm, will not be an increase in demand of 188
percent (490 percent less 312 percent) but the full 490 percent.
If, instead of starting the pump against a closed gate valve, it is started against a check valve in a
wholly frictional system with a resulting torque curve as in Fig. 24.30, the reduced voltage would be
able to accelerate the unit until 1,730 rpm is reached. The torque demand of this pump, when started
in this manner, is of the same magnitude as the torque demand of a high-speed Francis-vane impeller
or of a medium-speed, mixed-flow impeller starting against either a gate valve or check valve (Fig.
24.32), and is close to the same magnitude as a high-speed, mixed-flow or propeller pump (Fig. 24.34).
Centrifugal pumps of other specific speed types, therefore, offer starting problems only slightly more
difficult than those described above.
It is very unusual to start a squirrel cage motor driving a centrifugal pump on the 65 percent voltage
tap of a reduced voltage autotransformer starter. The use of such a low reduced voltage would be feasible
only with a pump having a shutoff horsepower that is considerably less than the horsepower required
at rated capacity and which is always to be started against a closed gate valve (Figs. 24.24 and 24.25).
With 65 percent voltage, the 45 kW (60 hp) motor illustrated above would accelerate this pump to
approximately 1,700 rpm, resulting in an inrush of 150 percent current when full voltage was applied.
In the case shown in Fig. 24.30, starting the motor on 65 percent voltage would only allow it to accelerate
the pump to about 1,200 rpm with a resulting second current inrush of 440 percent when full voltage
was applied.
As previously described, the starting of a wound-rotor motor involves applying current to the stator
winding with resistance connected into the rotor windings. The amount of resistance cut into the rotor
winding and the amounts of resistance cut out in each successive step in the starting cycle determine
the amount of torque available at various speeds as well as the current demand. The number of steps used
will depend on the limit of current allowed. Figure 24.39 illustrates the torque and current characteristics of
a 45 kW (60 hp), 1,800-rpm slip-ring motor with a four-step starter proportioned for the pump torque
requirements shown on the same figure. The first step results in a 337 Nm (249 Ibft) torque at zero
speed with 150 percent of rated current. This first speed allows the motor to accelerate the pump to
1,465 rpm. If the second step is cut in when that speed is reached, the motor torque would become 263
Nm (194 lbft) with the current stepping up to 106 percent. This step would result in accelerating the
pump to 1,685 rpm, when the third step would be cut in with a 263 Nm (194 lbft) torque and 108
percent current. The third step would accelerate the pump to 1,775 rpm, whereupon the fourth step (no
resistance in the rotor circuit) would be cut in, resulting in a 244 Nm (180 lbft) torque and 100 percent
current. An ideal starter is not always obtainable and it may be necessary to purchase a starter with
more steps than theoretically required to meet a limitation.
The modem synchronous motor has an auxiliary squirrel cage motor built into it to start the
motor and bring it and its connected load up to about 95 percent of the synchronous speed. To
pull the motor into step, the load has to be transferred from the squirrel cage element to the regular
synchronous motor element. With present-day controls, which incorporate proper timing devices,
this no longer requires expert handling. Except for some differences in relative values, the starting
torque and starting characteristics of a synchronous motor are similar to those illustrated in Fig.
24.38 for a regular squirrel cage motor.
666 Drivers

500

400 ~
\4
300

200
.......
~
" '\ \ 1\

100
I--
~ ~ "- f\\
~ ...........
~
0 2 4 6 8 10 12 14 16 18
450

400
, \
'I

350 \ 1\
\~ \4
'" ""2
t-
~
"'-
I 300
m
...J
"-
z 1\ \
250
-.........
w
.
::> 200
~ t--....
\
1\ 1
K \ \
0 ............
a:::
0
t- 150

100
~
~ \\ \
~ ~ ~
I--

50

~
o 2 4 6 8
~
pu~~
10
I
12
~

14
""-
N
16 18
SPEED. IN 100 RPM

Fig.24.39 Torque-speed and percent current-speed curves of 45 kW (60 hp), 1,800-rpm, 3-phase, 60-cyc1e
wound-rotor motor with four-step starter.

Submersible Motors
The submersible motor was first developed some 50 years ago as an alternative to line-shaft drive for
vertical turbine pumps installed at great depths. The main advantages of a submersible pump and motor
unit for such applications are the following:
Drivers 667

1. Elimination of long lineshafts and numerous bearings


2. Elimination of a pump house for the protection of motors located at the surface
3. Freedom from motor noise in applications where noise from a surface unit might be objectionable
4. Ease of installation
5. Freedom from periodic lubrication and adjustment.

The first attempts to provide a motor that could be close coupled to the pump and installed at the pumping
level involved the use of air bells, in which the motor housing was kept filled with air under pressure,
while the bottom of the housing was open to the water. This construction was only partially successful;
eventually motors were developed that were able to run submerged in the water, the only connection
required being a water-proof electric power supply cable.
Since their original development, two distinct classes of submersible motor have evolved. One is an
extension of the original concept: various designs intended to operate submerged in the pumped liquid.
The other employs the principle of having the pumped liquid within the motor to produce a hermetically
sealed or so-called sealless pump. Such pumps are generally dry pit installations, and are used for
applications where there is concern over the integrity of conventional shaft seals (see Chap. 8).
There are now five basic designs of submersible motor. Their use for the two classes of pump
application, submersible and hermetically sealed, is shown in Table 24.6. Except for dry stator, air filled
motors, all these designs depend on the liquid within the motor to lubricate and cool their bearings.
When the liquid is the pumpage, this places definite limits on the allowable solids content to prevent
bearing erosion, and on the minimum margin over vapor pressure to ensure there is no flashing within
the bearings. A number of prototype designs have employed magnetic bearings (see Chap. 11) as a
means of overcoming this limitation. To avoid confusion over the differences in construction for the
two applications, the motors commonly used for each are discussed separately.

Table 24.6 Basic Submersible Motor Designs and Common Uses

Motor classification Pump type commonly used for


Submersible Hermetically sealed

Wet-stator; pumped liquid


Wet -stator; clean liquid filled
Wet -stator; barrier liquid
Dry-stator; air filled
Dry-stator; canned

MOTORS FOR SUBMERSIBLE PUMPS

Wet-Stator Types
In the simplest wet-stator design, no attempt is made to prevent the pumped liquid from contacting
the motor windings. The conductors of the stator windings are coated with insulation impervious to the
pumped liquid. The pumped liquid cools the motor and lubricates its bearings. Because submersible
pumps can be exposed to waters of varying quality, including waste water and sewage, the wet-stator
motors made today for submersible pumps are usually designed to exclude the pumped liquid from the
motor housing by having the motor filled with a clean liquid.
The clean liquid used is either a solution of water and glycol or a refined nonhygroscopic dielectric
668 Drivers

Corrosion resistant slailless


steel outpul shaft

Mechanfcal seal ---~


with sand slinger

Twin. OYefsized water --~


lubricaled guide bealings

Oynamicaly
balanced rotor

Corrosion resistant stailless


steel stator tube _ _ _ __ _ _ _-!

Water cooled,
open design (net canned).
wet wound

~ustable. sen aligning


thrust becwlng

Preuure compenaatlng
diaptngm

Fig. 24.40 Water-glycol filled submersible motor.

mineral oil. The water glycol solution protects the internal ferrous parts from corrosion, provides a clean
liquid for the bearings, and is not impaired by a small amount of leakage of pumped liquid into the
motor housing. Oil protects against corrosion, protects the insulation against breakdown, provides excel-
lent lubrication for the bearings, but is sensitive to water leakage into the motor housing. Liquid filled
motors (Fig. 24.40) have a mechanical shaft seal to prevent the filling liquid from leaking out and the
pumped liquid and foreign matter from entering. To keep the differential pressure across the seal low,
the seal is located adjacent to the pump's suction or the pump is designed so the sealed pressure is
Drivers 669

suction pressure. In some designs, the differential pressure is lowered still further by having a pressure-
compensating arrangement to keep the pressure in the motor casing equal to that adjacent to the shaft
seal (Fig. 24.40). With either filling liquid, an impeller within the motor circulates the liquid through
the stator, and the motor's heat is finally dissipated to the pumped liquid passing over the outside of
the motor casing.
Wet-stator motors are available in sizes up to 1,500 kW (2,000 hp), and special designs can be
obtained in larger sizes. Wet-stator motors tend to be long and slender, with a 5 to 1 ratio of rotor length
to diameter quite common. Two factors account for this. First, with the rotor running in liquid, the
viscous drag increases as the fifth power of the rotor diameter, therefore a smaller diameter rotor helps
minimize this parasitic loss. Second, most submersible motors are intended for pumps in well or borehole
service, which favors long, slender motors. As a result of the combined effect of viscous drag and rotor
proportions, wet-stator submersible motors have lower efficiencies and power factors than normal motors.
On the other hand, their small diameter rotor has lower inertia resulting in a higher rate of acceleration
to full speed.
The installation of a vertical turbine pump with a submersible motor is illustrated in Fig. 24.41. The
motor is located at the bottom of the pumping unit. The complete unit is suspended at the upper pump
flange and lowered into the well. Consecutive lengths of piping are added until the unit is lowered to
the required depth. As each length of pipe is added, the power supply cable is paid out and clipped to
the piping.
A variation of the submersible well pump is illustrated in Figs. 24.42 and 24.43. In this arrangement,
the pump and motor are capable of operating in the horizontal position. This allows the pump and motor
unit to be suspended horizontally within a pipeline to serve as a pipeline booster pump. The entire unit
is manufactured with flanges at each end and becomes an integral part of the line.

Dry-Stator Types
With the use of submersible pumps for waste water, sewage, and slurry service, applications in which
the a long, slender motor is not necessary, several manufacturers have developed dry-stator motors to
allow a stiffer rotor and performance comparable with normal motors. These motors are air filled, with
an oil barrier contained between axial face shaft seals (Fig. 24.44) to exclude the pumped liquid from
the motor casing. The conductivity of the barrier oil is monitored so that in the event of pumpage leaking
into the barrier oil, the pump can be shut down before its motor is damaged. Heat dissipation from these
motors is to the pumped liquid passing over the outside of the motor casing.

MOTORS FOR HERMETICALLY SEALED PUMPS

Wet-Stator Types
Wet-stator motors running in the pumped liquid are the simplest arrangement and are used whenever
liquid cleanliness will allow them. The design of their rotor and stator is similar to that for conventional
submersible pumps. The motor casing and stator lead seals are different in that they now form part of
the pump's pressure boundary or pressure containment, and therefore often have to be designed for a
high pressure.
Wet-stator motors running in the pumped liquid are used extensively for boiler circulating pumps
(Fig. 24.45). The pump's suction conditions in this application, typically on the order of 185 bar (2,650
psig) and 375°C (7lO0F), preclude the use of a shaft seal. In this arrangement, free interchange between
the liquid passing through the pump and that in the motor is restricted by a "thermal throttle," thereby
allowing the external heat exchanger to cool the liquid in the motor to the temperature needed to dissipate
670 Drivers

Fig. 24.41 Installation of vertical turbine pump with submersible motor.

--------

•.
Fig. 24.42 Horizontal pipeline booster pump with submersible motor.
(Courtesy Layne & Bowler.)
Drivers 671

Fig. 24.43 Installation of submersible pump in piping layout with bypass system.
(Courtesy Layne & Bowler.)

Fig. 24.44 Section of the lower end of an air filled motor showing the barrier oil chamber. The inner oil seal is
located in the middle of the oil chamber and the outer seal is below it.
(Courtesy ITT Flyght AB.)

the motor's electrical losses without exceeding the allowable insulation temperature. At the same time,
the cooler water has sufficient viscosity to serve as an adequate lubricant for the motor's hydrodynamic
bearings. An impeller at the bottom of the motor circulates the cooled water through the motor and
external heat exchanger. Motors of this design have been built for working pressures to 320 bar (4,640
psig), design temperatures to 475°C (890°F), power to 2,500 kW (3,350 hp) and motor voltages to 6,000 V.
A second version of wet-stator motor running in the pumped liquid is the design used to drive pumps
handling cryogenic liquids (Figure 24.46). In this application a hermetically sealed pump is considered
672 Drivers

<D

<D <D

<D BOILER WATER


-APPROXIMATELY 370°C
(700°F)

® BOILER WATER
®- -APPROXIMATELY 63°C
(145°F)

® COOLING WATER
-APPROXIMATELY 46°
(115°F)

Fig. 24.45 Section of hermetically sealed boiler circulating pump with wet-stator motor.

simpler than the elaborate shaft seal necessary to provide the same degree of security. This design differs
from the motors for boiler circulating pumps in its means of cooling and the type of bearings. Because
the pumped liquid temperature is as low as -176°C (-285°F), adequate cooling is provided by the
pumped liquid passing over the outside of the motor casing, and the motor is considerably smaller than
normal for its power. The bearings are antifriction because liquefied gases are poor lubricants, and are
made of ceramic material to withstand the low temperature.
In applications where the pumped liquid contains abrasive solids, it cannot be allowed to enter the
motor casing lest it damage the product lubricated bearings and the insulation on the windings. At the
conditions that warrant a hermetically sealed pump, one practical means of excluding the pumped liquid
from the motor casing is to inject a small flow of clean barrier liquid, sufficient to dissipate the motor's
heat, at a pressure above that in the pump adjacent to the motor. A close clearance bushing between the
pump and motor (Fig. 24.47) keeps the injected flow velocity high enough to ensure the pumped liquid
does not enter the motor. The injection flow is typically provided by a small plunger pump.
Drivers 673

Fig. 24.46 Section of single-stage "in-tank" Fig. 24.47 Section of hermetically sealed
removable cryogenic liquid pump hydrocarbon reactor circulating pump with wet-stator
with balancing drum. motor and external clean liquid injection.
(Courtesy J.e. Carter Co., Inc.) (Courtesy BWIIP International.)
674 Drivers

Dry-Stator Types
Integral motor hermetically sealed pumps are used extensively in the chemical industry to handle
liquids whose leakage to the atmosphere cannot be tolerated under any circumstance. Many of these
liquids are corrosive and so cannot be allowed to come into contact with either the motor's stator or
rotor. To achieve this, both the rotor and stator are isolated from the pumped liquid by a thin layer of
corrosion resistant metal (Fig. 24.48), a construction known as "canned." If the motor is subjected to
any significant pressure, the stator can has to be backed up where it passes under the end windings (Fig.
24.48). An important safety feature of "canned" motors is the use of the motor casing, including the
stator lead seals, as a means of secondary containment in the event the primary containment, the stator
can, is ruptured.
Although it is protected from corrosion, the rotor of a "canned" motor still runs in the pumped liquid
and therefore has the same geometry constraints as the rotors of wet-stator motors. With the thickness
of the cans increasing the effective air gap, the efficiency of "canned" motors is lower still than that of
wet-stator motors. Depending on the pumping temperature, the motor's heat is dissipated partly to
pumped liquid circulated through the motor, partly to air passing over the outside of the motor casing.

Fig. 24.48 Section of canned motor pump for corrosive chemical service.
(Courtesy of Chempump, a Division of Crane Pumps and Systems, Warrington, PA.)
Drivers 675

As pumping temperatures increase, a thermal barrier is introduced between the pump and the motor,
and the stator is filled with oil to improve heat transfer to the motor casing wall. At still higher
temperatures, it is necessary to either circulate the stator coolant through an external heat exchanger, or
to resort to ceramic insulation, which has an allowable surface temperature of 450°C (850°F). Standard
design canned motors for chemical pump service are available up to 180 kW (240 hp) at 3,600 rpm.

Variable-Speed Devices
As mentioned earlier in the chapter, many centrifugal pump installations could benefit from variable
speed operation of the pump, but the pump drivers (squirrel cage electric motors, for example) can
operate only at constant speeds. In such cases, some form of variable-speed device is used to allow
variable-speed operation of the pump. The devices are some form of slip coupling, hydraulic or magnetic,
a hydrostatic transmission, or a controller to vary the frequency of the power supply to the motor, an
arrangement known as variable-frequency drive.

HYDRAULIC COUPLINGS

All of the many mechanical designs of hydraulic couplings are composed of four basic elements: an
impeller, a runner, a casing, and a means for changing the amount of oil contained within the unit during
operation so as to obtain an adjustable speed output while the speed input remains constant.
The impeller and the runner are mounted facing each other on two separate shafts with no mechanical
connection between the two. The power is transmitted from the impeller to the runner entirely by a
vortex of oil circulating between the two. The impeller, mounted on the driven shaft that is connected
to the driver, acts as a centrifugal pump and imparts energy to the oil, which is accelerated radially
outward. The runner is mounted on a shaft connected to the driven equipment and acts as a reaction
turbine, absorbing the kinetic energy in the oil as it passes from the periphery inwardly toward the center
of rotation. The circulation of oil between the impeller and runner is called the working circuit. A casing
is attached to the impeller to confine most of the oil to the working circuit. An adjustable scoop tube
can withdraw a variable amount of oil from the working circuit to provide speed variation.
The construction of a hydraulic coupling is illustrated in Fig. 24.49. The entire rotating assembly is
enclosed in a welded steel housing, which is split on the horizontal centerline. The bottom portion of
this housing serves as a support for the bearings in which both the driving and driven shafts rotate. It
also contains an oil reservoir from which the lubricating and operating oil is supplied to the unit. A
centrifugal pump, gear-driven from the input shaft, picks up oil from this reservoir and sends it through
an oil cooler, from which it returns to the tank as well as to the bearings.
The oil in the working circuit rotates in an annular ring by virtue of the centrifugal force imparted
to it by the impeller. Connections are provided between the working circuit and the scoop tube chamber.
Therefore, the amount of the oil in the annular ring can be varied by varying the oil level in the scoop
tube chamber. When the tip of the oil scoop tube is extended toward the periphery, the amount of oil
in the working circuit is reduced and the output speed decreases. Withdrawing the scoop tube toward
the center permits the working circuit to fill with oil and the slip decreases, increasing the output speed.
The hydraulic coupling is ideally suited for variable speed operation of variable torque machines such
as centrifugal pumps, because the pump horsepower varies with the cube of the speed, whereas the
676 Drivers

Fig. 24.49 Hydraulic coupling.


(Courtesy American-Standard, Industrial Division.)

hydraulic coupling efficiency is almost inversely proportional to the speed. The power input to the
hydraulic coupling can be calculated from the relation: Input brake horsepower = (pump bhp at variable
speed) (input speed/pump speed) + fixed losses.
With the working circuit of oil there is a slight slip which may vary from I to 3 percent, depending
on the power output and the size of the hydraulic coupling. The pump speed, therefore, will be from
99 to 97 percent of the driver speed. The fixed losses are the bearing and windage losses and run to
approximately 1 percent.
The effect of using a hydraulic coupling with a centrifugal pump is illustrated in Fig. 24.50, which
shows a typical performance of a boiler feed pump with the capacity, head, efficiency, power, and speed
expressed in percentages of their values at the design-rated condition. The system-head curve has been
superimposed on this figure, and the required rpm for the system-head curve and the pump brake
horsepower when operated at variable speed are also shown. The system-head curve at 100 percent rated
flow is only 95 percent of the rated head, reflecting the margin of safety that might be added to the
pump head in the calculations.
If the hydraulic coupling has a slip of 3 percent at rated conditions and the fixed losses are 1 percent,
the motor speed will, therefore, be approximately 103 percent of the pump rated speed. The input brake
horsepower, when operating at variable speed is calculated as shown in Table 24.7.
A comparison of the power input listed above with the pump brake horsepower at the same flows,
when operated at constant speed, shows the advantages of using hydraulic coupling. For example, at 60
percent flow, the power consumption is 79 percent at constant speed. It is only 58 percent at variable
speed, and, including the hydraulic coupling losses, the input horsepower is only 68 percent.
Drivers 677

HEAD-CAPAClr~
110 AT Co
NST4.N ~
T r S~e~
....-
- "~
100 :E
>- ~~~
3,;.. ~
a..
a:
u 90 ASSUMED SYSTEM HEAD .....
z 0
UJ
~V UJ
u
Ii:
80
t~f?~~ ~~"'¢1
100
a:
I-
<l

,.
u...
fOR
90 u...
UJ
70 vO RPM 0
0
z I-
120 <l 60 "'/ z
80 UJ
u
SPEEO __ I--
0
.l...1
W
a.. <l
a:
:z: 100 UJ
:z: 50
C~;"-V
UJ
CD a..
0
80
0
40 r.:f ~"I'

- r~ ~--
UJ UJ
l- I-
<l <l
a: 60 a: 30 41. ~ E. S ",E
u... u... ~fI.\~e\": cu~
0
40
0
I- 20 tJ.-'"
r--- BHP
ON S'lS"I'E
~"I' '" E~O

--j
I-
Z Z
UJ UJ
u 20 u 10
a:
UJ
a.. 0
a:
UJ
a.. o /
o 20 40 60 80 100 120
PER CENT OF RATED CAPACITY

Fig. 24.50 Performance curve of variable speed centrifugal boiler feed pump.

Table 24.7 Calculation of Input Brake Horsepower

Input brake
Pump brake Pump speed, horsepower,
horsepower of in in
Pump variable speed, percentage percentage
capacity, in in percentage of rated of rated
percent of rated power speed power

0 32 85 40
20 38 85.2 47
40 46 86.7 55.7
60 58 89 68
80 74 93 83
100 95 98 101
678 Drivers

HYOROVISCOUS DISK PACK


SPEED Sl:NSOR

RUGGED WELDED STl:El HOUSING


INClWES INTEGRAl 01.. SUMP

Fig.24.51 Section of horizontal hydroviscous drive, rated 3,750-15,000 kW (5,000-20,000 hp).


(Courtesy Philadelphia Gear Corporation.)

HYDROVISCOUS DRIVES

An alternative fonn of hydraulic drive in which torque is transmitted by viscous friction between an
assembly of closely spaced disks running in an oil bath. The drive's output speed is varied by changing
the disk spacing. Heat generated by slippage between the disks is removed by forced circulation of the
oil bath, and dissipated through a heat exchanger to a coolant. Compared to kinetic-type hydraulic drives,
the advantage of a hydroviscous drive is that with the disk assembly locked (all disks in tight contact),
there is no slip through the drive. Typical construction is shown in Fig. 24.51. Other than zero initial
slip, the application of hydroviscous drives follows the same rules as for kinetic hydraulic drives.

MAGNETIC DRIVES

One of the different designs of variable speed magnetic drives is illustrated in Fig. 24.52. It consists
essentially of two rotating members: One, called the ring, is driven by the driving motor, the other,
called the magnet, is coupled to the driven machine.
The ring revolves at the same speed as the driving motor; the magnet is free to revolve within the
ring, and is separated from the ring by an air gap. The magnet poles are energized (or excited) by direct
current through collector rings in the shaft.
Drivers 679

COtU'",,.1 uno
-'"

Fig. 24.52 Variable speed magnetic drive.


(Courtesy Dresser-Rand Motor and Generator Division.)

The difference in speed between the ring and the magnet causes the ring to cut the magnetic flux
produced by the magnet, which induces currents in the ring. The induced currents produce poles in the
ring, which react with and drag around the poles of the magnet, thus causing the magnet to revolve.
The torque is transmitted magnetically through the air gap between the ring and the magnet and is
varied gradually by varying the excitation of the magnet, thereby varying the speed of rotation of the
magnet. A magnetic amplifier rectifier supplies the excitation, providing automatic speed control.
Magnetic drives are available in several forms of mechanical construction to fit various methods of
mounting and for different conditions of alignment. For centrifugal pump drive, the most common forms
are the self-contained unit, with either pedestal-type or bracket-type bearings; they are connected to the
motor and to the pump by flexible couplings at both ends as in Fig. 24.53. Other forms are available
for vertical pump drive, as in Fig. 24.54, which shows a vertical squirrel cage induction motor and an
adjustable speed magnetic drive operating a vertical centrifugal pump. The motor is rated 75 kW (100
hp), 1,180 rpm. The magnetic drive provides adjustable speed from 1,180 to 350 rpm.
The losses in a magnetic drive are essentially similar to the losses of a hydraulic coupling, namely,
slip losses and fixed power losses. The input to the drive, therefore, is basically the same as if the
variable speed was obtained by means of a hydraulic coupling.
680 Drivers

Fig. 24.53 Horizontal volute pump motor driven through variable speed magnetic drive.

Fig. 24.54 Vertical centrifugal pump with vertical motor and magnetic drive.
(Courtesy Dresser-Rand, Motor and Generator Division.)
Drivers 681

REUEF[==~~~~~~~~~
VALVE .. ..
PRIME
MOVER

HYDRAULIC
PUMP
HIGH PRESSURE
FEEDER LINE
Court.yal:
MWI, MovIng W_lndualr1e8 Corp.

HYDRAULIC
MOTOR

WATER PUMP

Fig. 24.55 Flow diagram of hydrostatic transmission.


(Courtesy MWI Corp.)

HYDROSTATIC TRANSMISSION

In a number of applications, notably marine, a hydrostatic transmission is used to drive cargo, ballast,
and fire pumps. The elements of a hydrostatic transmission (Fig. 24.55) are a variable displacement
hydraulic pump, driven by an electric motor or other convenient driver, a fixed-displacement hydraulic
motor close coupled to the pump, interconnecting hydraulic piping, and a heat exchanger to dissipate
the losses in the hydraulic pump and motor. The speed of the centrifugal pump is varied by changing
the flow rate of hydraulic oil in the circuit. Hydrostatic transmissions are relatively complex (variable
displacement hydraulic motors have many moving parts), but have the virtues of high efficiency, on the
order of 80 percent, over a wide speed range, and can have the prime mover, motor, or engine located
some distance from the pump.

VARIABLE-FREQUENCY DRIVE

With the rapid development of solid state electronic devices over the past 20 years, it is now economically
viable to modulate the frequency of the power supply to a conventional squirrel cage motor. This
development allows a simple, readily available driver to operate with stepless speed variation from zero
to the mechanical limit of its rotor. And so it is now feasible to select a pump to run at some speed
between the available motor speeds, 2,400 rpm, for example, or for a speed above 3,600 rpm, without
having to add another device to the mechanical drive train, and without paying a substantial penalty in
extra energy consumption,
The variable frequency controller is located between the motor and the electrical supply mains (Fig.
682 Drivers

DIVERTER

SMOOTHING
REACTOR
INPUT BREAKER ~ OUTPUT BREAKER
,------ :00' ,------'------,
INCOMINGAC OUTPUT
AC-DC DC-AC
FIXED RECTIFIER INVERTER LINE
FREQUENCY FILTER
'--____----' (UO'; L -_ _ _--"

VARIABLE ACPOWER
ELECTRONIC CONTROL MODULE

Fig. 24.56 Basic schematic and controller for variable frequency drive.
(Courtesy Ansaldo-Ross Hill.)

24.56). When the purpose of VFD is to vary the speed from a normal motor speed, the controller is
wired so mains frequency is applied to the motor until the speed has to be varied, thus avoiding the
electrical losses in the controller when running at rated speed. For the powers typical of centrifugal
pump drives, VFD can be considered 95 percent efficient from full load down to 25 percent. The losses
in the controller are dissipated to ambient air in small drives and to a circulated coolant in larger drives.
There are several variations in the electrical design of variable-frequency controllers. These have an
effect on the form of the voltage or current supplied to the motor. Designs that produce high "spikes"
in the current can dictate that the motors used with such controllers be rated for 110 percent rated power
to ensure the motor does not overheat. Similarly, variations in the electrical design affect the extent of
electrical harmonics reflected back onto the electrical supply system, and in tum, the cost of eliminating
these if the utility, or the balance of the plant system, cannot tolerate their effect.
Evaluating a variable-frequency drive generally follows the principles used for slip-type drives, the
notable difference being that drive efficiency is largely independent of the output speed. To illustrate

Table 24.8 Energy Input to VFD for Boiler Feed Pump


in Fig. 24.50
See Table 24.8 for pump power and speed.

Equivalent input to VFD, in


Pump capacity, in percent percentage of rated power

o 34
20 40
40 48
60 61
80 78
100 100
Drivers 683

the difference, Table 24.8 shows the equivalent input power to the VFD (Le., pump powerNFD efficiency),
for the boiler feed pump whose performance is shown in Fig. 24.50. Compared to the data for a slip-
type variable-speed drive (see Table 24.7), VFD offers lower energy consumption over the entire speed
range, a result of the high slip ratio (pump speed 95 percent of driver speed) at 100 percent capacity in
this example. Variable frequency drives for pumps have been furnished for powers up to 7,500 kW
(10,000 hp).

HERMETICALLY SEALED DRIVES

Beyond the various forms of submersible motors discussed earlier in this chapter, there are two types
of magnetic drive employed to provide a hermetically sealed pump. These drives are similar in principle
to the magnetic drive used for variable speed (see the earlier discussion) but with important differences.
Figure 24.57 shows a magnet drive hermetically sealed centrifugal pump. The driving or outer magnet
ring, supported by a separate bearing frame or the driver rotor, passes a rotating magnetic flux through
the containment can (pressure boundary) to the driven or inner ring mounted on the pump rotor. Two
drive principles are used: eddy current and synchronous. In eddy current drives (Fig. 24.58), inner ring
magnets are created by induced eddy currents, the product of relative motion (slip) between the two
rings. Slip increases with the torque transmitted. Synchronous drives (Fig. 24.59) have matching perma-
nent magnets in the inner ring, thus driven speed equals driving speed whilever the torque required is
within the drive's capacity.
The choice of magnetic drive type has been influenced significantly by advances in magnet technology.
At one time, the synchronous drive's advantages of constant speed and lower heat load to the pumped
liquid were generally not enough to offset the lower weight, lower cost, and higher starting torque of
eddy current drives. With the development of higher permeability magnetic materials, the size of synchro-

Fig. 24.57 Hermetically sealed pump, magnetic drive type.


684 Drivers

ROTATING OUTER MAGNET ASSEMBLY

MILD STEEL/COPPER
TORQUE RING (DRIVEN) ~__"""'J

CONTAINMENT SHELL

Fig. 24.58 Eddy current magnetic drive.


(Courtesy The Kontro Co ., Inc.)

OUTER MAGNET RING


/

Fig. 24.59 Synchronous magnetic drive.


(Courtesy The Kontro Co ., Inc.)

nous drives has decreased, and they are now the normal design for all but high temperature (above
400 oP-200°C) and non-critical, low cost applications.
Both types of drive incur losses beyond those in conventional pumps: disc friction from rotation of
the inner ring in the pumped liquid, magnetic drag between the outer ring and the can when it is metallic,
and, in eddy current drives, the power loss produced by slip across a constant torque device.
The traditional can material has been metal, with as low an iron content as possible, for example
Drivers 685

Hastealloy C, to minimize eddy current drag. Designs to overcome this problem include a laminated
metal can and cans from various fiber reinforced polymers.
The rotor of magnetic drive pumps is supported by product lubricated bearings, so the design is
sensitive to abrasive solids in the pumped liquid, and dry running from either loss of suction or vaporization
within the drive at points of high temperature. Advances in bearing materials are improving the drive's
tolerance of these difficulties, but the best results will still be obtained when the pumped liquid is clean,
the pump is protected against running dry, and the circulation of pumped liquid through the drive ensures
there will not be any internal flashing. In connection with the last two points, motor current monitoring
(see Chap. 30) provides excellent protection against dry running, and drives in which the cooling liquid
flow is returned to pump discharge, using a small auxiliary impeller, are less prone to internal flashing.

BIBLIOGRAPHY

[24.1] National Electrical Manufacturers Association (NEMA).


25
Priming

A centrifugal pump is primed when the waterways of the pump are filled with the liquid to be pumped.
The liquid replaces the air, gas, or vapor in the waterways. Removal of the air, gas, or vapor may be
done manually or automatically, depending on the type of equipment and controls used. To one familiar
with positive displacement pumps of the reciprocating and rotary types only, it might seem strange that
a centrifugal pump cannot prime itself. Positive displacement pumps, if properly sealed, will pump air
as well as liquid and will, therefore, exhaust any air in the suction line. Centrifugal pumps will also
pump air (contrary to commonly held opinion), but because of the low density of air the actual pressure
developed when pumping it is very small.
For example, a pump that develops a 61 m (200 ft) head with water would likewise develop a 61 m
(200 ft) head with air, but a head of 61 m (200 ft) of air is equivalent to a vacuum of approximately
75 mm (3 in), measured in terms of a water column. Such a vacuum is obviously insufficient for
normal priming.
When first put in service, the waterways of a centrifugal pump are filled with air. If the suction supply
is above atmospheric pressure, this air will be trapped in the pump and compressed somewhat when the
suction valve is opened. Unless the suction pressure is sufficiently high, however, the air will not be
compressed enough to permit the suction waterways and the eye of the impeller to be filled with water,
and the pump will not be primed; therefore, the air must be removed. With a positive suction head on
the pump, priming is accomplished by venting the entrapped air out of the pump through a valve provided
for the purpose.
If the pump takes its suction from a supply located below the pump itself, the air in the pump must be
evacuated, either by some vacuum-producing device, or by providing a foot valve in the suction line so that
the pump and suction piping can be filled with liquid, or by providing a priming chamber in the suction line.

FOOT VALVES

Foot valves were very commonly used in early installations of centrifugal pumps. Except for certain
applications, their use is now much less common. A foot valve (Fig. 25.1) is a form of check valve

686

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
Priming 687

Fig. 25.1 Foot valve.

installed at the bottom or foot of a suction line. Like an ordinary check valve, it allows flow in one
direction only-toward the pump. When the pump is stopped and the ports of the foot valve close, if
the valve seats tightly, the water cannot drain back to the suction well.
Unfortunately a foot valve does not always seat tightly, and the pump occasionally loses its prime.
The rate of leakage is generally small, however, and it is feasible to restore the pump to service by
filling and starting it promptly. This tendency to malfunction is increased if the water contains small
particles of foreign matter like sand, and foot valves should not be used for such service. Another
disadvantage of foot valves is their unusually high friction loss. To keep friction loss within reasonable
limits, a foot valve with a large port area is necessary. Unfortunately, the size of the pipe connection is
not a true indication of the area of the valve ports.
A typical installation of a pump with foot valve is shown in Fig. 25.2. The pump can be filled with
water through a funnel attached to the priming connection or from an overhead tank or any other source
of water. If a check valve is used on the pump and the discharge line remains full of water, a small
bypass around the valve permits the water in the discharge line to be used for repriming the pump when
the foot valve has leaked. When a pump equipped with a foot valve is primed, provision must be made
for filling all the waterways and for venting out the air. The actual point at which water is introduced
into the waterways is of no significance but is usually at the priming connection provided on the top of
the pump casing (Fig. 25.2).
When the discharge head is low, the discharge check valve is often eliminated and the foot valve
acts as a check valve to prevent reverse flow through the pump. As most foot valves are designed for
relatively low pressures, such use is not generally practicable with systems having a high discharge
688 Priming

PET COCK PIPE FROM


FOR VENT~ ~ PRIMING SUPPLY

DISCHARGE

-+--;.-- --
GATE VALVE

FOOT VALVE

Fig. 25.2 Installation using foot valve.

pressure. In such an installation, a discharge check valve is necessary, but there is always the danger
that the check valve will not seat as tightly as the foot valve. The foot valve would then be subject to
the static discharge pressure. In a system in which the water can drain out of the pump through the
discharge line when the pump is shut down, it is necessary to have a discharge gate valve. This valve
must be closed before the pump is stopped and opened after it is started.

PRIMING CHAMBERS

A priming chamber in its simplest form is a tank with an outlet at the bottom that is level with the pump
suction nozzle and directly connected to it. An inlet located at the top of the tank connects with the
suction line (Fig. 25.3). The size of the tank must be such that the volume contained between the top
of the outlet and the bottom of the inlet is approximately three times the volume of the suction pipe.
Leakage of air when the pump is shut down may cause the liquid in the suction line to leak out, but
the liquid iIi. the tank below the suction inlet cannot run back to the supply. When the centrifugal pump
is started, it will pump this entrapped liquid out of the priming chamber, creating a vacuum in the tank.
The atmospheric pressure on the supply will force the liquid up the suction line into the priming chamber.
Precautions must be taken to prevent vortexing at the outlet of the chamber and to prevent siphoning
when the unit is shut down. It is usually more advantageous to buy a commercial priming chamber with
proper automatic vents and other features than to attempt to make one's own design. The use of priming
chambers is restricted because of their size to relatively small pumps. A well-known design of a priming
chamber with automatic features is shown in Fig. 25.4. This type of priming chamber is more satisfactory
for most installations than a simple tank (Fig. 25.3).
SUCTION
LINE
-
- -
----

rT1
I

I
I ~~RGE

Fig. 25.3 Simple priming tank.

" ........... --_.-_.

Fig. 25.4 Commercial priming tank.


When the pump starts (left), it draws liquid from the lower tank and discharges it through the upper tank. With-
drawal of liquid causes a partial vacuum in the lower tank, and liquid flows from the well into the tank and then
to the pump. When pump stops (right), liquid from the upper tank runs back into the pump and the lower tank,
by gravity, keeping the pump ready for starting. (Courtesy Valve and Primer Corp.)
690 Priming

TYPES OF VACUUM DEVICES

Almost every commercially made vacuum-producing device can be used with systems in which pumps
are primed by evacuating the air. Formerly, water- and steam-jet primers had wide application, but with
the increase in the use of electricity as a power source motor-driven vacuum pumps have become popular.
The wet-type vacuum pumps are best for manually controlled, electrically driven units because no
damage will result if slugs of water are carried over into the vacuum pump. When dry vacuum pumps
are employed, some protective device must be interposed between the centrifugal pump and the vacuum
pump to prevent water entering the vacuum pump. The dry vacuum pump is used extensively for central
priming systems.

Ejectors
Priming ejectors work on the jet principle, using steam, compressed air, or water as the operating
medium. In steam or compressed-air ejectors, the steam or air is discharged at high velocity into the
throat of a venturi tube carrying with it some of the surrounding air and forcing the combined mixture
out of the tail piece of the venturi (Fig. 25.5). A high vacuum can be produced by both steam and water
ejectors. Water ejectors are similar to steam ejectors except that multijet nozzles are used when the
actuating water pressure is low.
A typical installation for priming with an ejector is shown in Fig. 25.6. Valve VI is opened to start
the ejector and then valve V2 is opened. When all the air has been exhausted, water will be drawn into,
and discharged from the ejector. When this occurs, the pump is primed, and valve V2 is closed. Valve
VI is then closed.
An ejector may be used for a number of pumps if it is connected to a header through which the
individual pumps are vented through isolating valves.

Dry Vacuum Pumps


Dry vacuum pumps, which may be of either reciprocating or rotary type, cannot accommodate mixtures
of air and water. When they are used in priming systems, liquid must be prevented from entering them.

Wet-Vacuum Pumps
Any rotary, rotative, or reciprocating pump that can handle air or a mixture of air and water is
classified as a wet-vacuum pump. The most common type used in priming systems is the Nash Hytor
pump (Figs. 25.7 and 25.8).
This is a centrifugal displacement pump consisting of a round, multiblade rotor revolving freely in
an elliptical casing that is partially filled with liquid. The curved rotor blades project radially from the
hub, and form, with the side shrouds, a series of pockets and buckets around the periphery. The rotor
revolves at a speed sufficient to throw the liquid out from the center by centrifugal force, resulting in
a ring of liquid revolving in the casing at the same speed as the rotor, but following the elliptical shape
of the casing. This condition alternately forces the liquid to enter and recede from the buckets in the
rotor at high velocity.
The complete cycle of operation in a given chamber is illustrated in Fig. 25.8. At point A chamber
(1) is filled with liquid. Because of centrifugal force, the liquid follows the casing, withdraws from the
rotor, and pulls air in through the inlet port, which is connected with the pump inlet. At (2) the liquid
has been thrown outward from the chamber in the rotor and has been replaced with air or gas. As rotation
continues, the converging wall (3) of the casing forces the liquid back into the rotor chamber, compressing
the air trapped in the chamber and forcing it out through the discharge port, which is connected with
Priming 691

VENTURI TAIL PIECE

Fig. 25.5 Steam jet primer.


(Courtesy Schulte & Koerting Co.)

the pump discharge. The rotor chamber is now filled with liquid and the cycle is repeated. Each revolution
includes two cycles.
If a solid stream of water circulates in this pump in place of air or an air and water mixture, the
pump will not be damaged, but will require more power in order to drive it. For this reason, in automatic
priming systems using this type of vacuum pump, a separating chamber or trap is provided so that water
will not reach the pump.
Water needed for sealing a wet vacuum pump can be supplied from a source under pressure-a city
water supply or some adjacent tank. When water is taken from a pressure supply, the amount of the
flow is regulated by a throttle valve. The flow is controlled by a separate shutoff valve so that the
adjustment of the throttle valve need not be disturbed. The shutoff valve can be manually or solenoid
operated. If solenoid operated, its operation is connected with the motor control.
692 Priming

--~~~~~s=~
/
STEAM. AIR. OR
WATER SUPPLY

--
Fig. 25.6 Arrangement for priming with an ejector.

Fig. 25.7 Partially disassembled Nash pump.


(Courtesy Nash Engineering Co.)

If no water under pressure is available, or if it is desired to maintain an independent system, the


vacuum pump can be mounted, as shown in Fig. 25 .9, on a base containing a reservoir. The sealing
water is taken from this reservoir and the pump discharges its mixture of air and water into it. This type
of installation is very desirable in locations where freezing may occur for the sealing can be done with
a solution of antifreeze.
Fig. 25.8 Operating principle of Nash pump.
(Courtesy Nash Engineering Co.)

Fig. 25.9 Nash priming pump with sealing liquid reservoir base.
(Courtesy Nash Engineering Co.)
694 Priming

Fig. 25.10 Portable pumping unit with hand primer.

Hand Primers

Hand primers were at one time quite commonly used to prime centrifugal pumps but are seldom used
in modem practice. A hand primer is a hand-operated displacement pump (Fig. 25.10). A common type
was the old-fashioned pitcher pump, which was often mounted on the priming connection on top of the
casing with a shutoff valve interposed.

Tightness of Suction Lines

Sometimes a pump that has been operating on a lift and that does not have a foot valve will hold its
prime when shut down for a considerable period. This is the result of a tight suction line and shaft seals
that are in good condition and properly adjusted in packed boxes. Such units can be restarted because
they are fully primed. The duration of the time interval in which a pump loses its prime is a measure
of the tightness of the suction line and the condition of the shaft seals. If a pump loses its prime quickly,
the suction line should be checked for leaks and the shaft seals should be inspected.

CENTRAL PRIMING SYSTEMS

If there is more than one centrifugal pump to be primed in a station, one priming device can serve all
the pumps. Such an arrangement is called a central priming system. Figure 25.11 illustrates the method
of piping for a central priming system. If the priming device and the venting of the pumps is automatically
controlled, the system is called a central automatic priming system.
Priming 695

• TO VACUUM====~==================~~===========
PROOUCER
~HAND OPERATED
.r VALVE

Fig. 25.11 Connections for a central priming system.

VACUUM-CONTROLLED AUTOMATIC PRINTING SYSTEM

A vacuum-controlled automatic priming system consists of a vacuum pump exhausting a tank. The pump
is controlled by a vacuum switch and maintains a vacuum in the tank: of 50 to 150 mm (2 to 6 in) of
mercury, above the amount needed to prime the pumps with the greatest suction lift.
The priming connections on each pump served by the system are connected to the vacuum tank: by
automatic vent valves and piping. The vacuum tank: is usually provided with a gage glass and a drain
in order to detect water in the tank as the result of leakage in a vent valve. In such instances, the tank
may then be drained.
This type of system (Fig. 25.12) is the most commonly used type of central automatic priming system
and has been made with both wet and dry motor-driven vacuum pumps. In the vacuum-controlled
automatic priming system, manually controlled valves for priming have been replaced by automatic vent
valves and operation of the vacuum pump is automatically controlled by the vacuum switch.
The vacuum tank:, pump, and controls of a well-known make of this type priming system are shown
in Fig. 25.13. The vacuum pump is a belt-driven dry type mounted with its motor on top of the vacuum

.. TO VACUUM
PRODUCER

AUTOMATIO
/VENT VALVE
~
VAOUUM
SWITOH
OONTROLLING

NOOUG/
VACUUM

VAOUUM
TANK

Fig. 25.12 Connections for a central automatic priming system.


696 Priming

Fig.25.13 Vacuum tank, vacuum pump, and control for automatic priming system.
(Courtesy Valve and Primer Corp.)

tank. A vacuum switch on the side of the tank controls the operation of the vacuum pump. The circuit
in this switch is closed if the vacuum falls below a certain value (for example, 400 mm [16 in] of
mercury) and is opened if the vacuum reaches a higher value (for example, 500 mm [20 in] of mercury).
Contained in this switch is an unloader to assure that the vacuum pump starts with atmospheric suction
pressure and atmospheric exhaust pressure. As the unloader closes, the vacuum pump exhausts through
a check valve (shown next to the vacuum switch in Fig. 25.13) from the vacuum tank and builds up the
vacuum until the circuit in the switch is broken, thus stopping the motor. The pipe connection for the
vacuum line running to the various automatic vent valves is not shown in the illustration.
A gage glass on the tank shows the presence of water in the tank in case of leakage of any of the
vent valves. Above the gage glass is a vacuum relief valve, set to admit air into the tank if the vacuum
in the tank reaches a value above the upper limit of vacuum normally maintained. This protects the
vacuum pump against operation on a very high vacuum if the vacuum switch fails to open when the
upper limit of vacuum is reached.
Priming 697

AUTOMATIC PRIMING SYSTEMS USING EJECTORS

Several automatic priming systems use water ejectors. Figure 25.14 shows an installation of one type
that uses water from the discharge line of the pump. The flow of water to the ejector (Fig. 25.15) is
controlled by a piston valve that is spring- and water-pressure actuated. This valve is governed by a
float chamber (Fig. 25.16) mounted on the suction line.
Air is drawn from the top of the priming chamber until the water rises in it. The rising water lifts
the float which, by mechanical linkages, closes the connection to the suction of the jet. A second valve,
which when open relieves the pressure on the top of the piston of the water supply valve and allows it
to open, is also closed. The closing of the pressure relief valve in the priming or float chamber equalizes
the pressure on both sides of the piston and the spring closes the valve, thus shutting off the water
supply to the jet.
The same manufacturer has a unit for use in installations where water under sufficient pressure is not
always available (Fig. 25.17). It consists of a jet supplied with water under pressure by a small electric-

Fig. 25.14 Installation of centrifugal pump served by automatic water jet priming device.
698 Priming

CONTROL LINE TO
PISTON VALVE
PRESSURE
RELEASE
VALVE

FLOAT
CHAMBER

Fig. 25.15 Water jet primer with piston valve to Fig. 25.16 Float chamber to control primer in
control water supply. Fig. 25.15.
(Courtesy Skidmore Corp.) (Courtesy Skidmore Corp.)

motor-driven pump mounted on a reservoir tank and arranged so that the water is recirculated. The float
chamber (Fig. 25.18) uses a switch that is actuated by the float to control the motor of the supply pump.
The installation of the float chamber is illustrated in Fig. 25.19.
A third type, Figs. 25.20 and 25.21, made by another manufacturer, is basically similar to the first
type except that the water supply to the jet is controlled by a solenoid-operated valve. This is, in tum,
controlled by electrodes in the priming chamber, so that the jet operates if the level of the water in the
priming chamber falls and the longer electrode is uncovered. The jet continues to operate until the water
level rises and the shorter electrode comes into contact with the water. The priming chamber is mounted
directly on the pump to be primed (Fig. 25.21) and is connected by a special venting arrangement to
the top or tops of the suction waterways of the pump.

CENTRAL PRIMING SYSTEMS USING TWO VACUUM PUMPS

Most central priming systems are provided with two vacuum pumps (Fig. 25.22). This is desirable for
two reasons: First, a breakdown of one does not put the system out of operation and, second, the capacity
Priming 699

FLOAT SWITCH
¢==~
TO

TANK

I
CONNECTION
FROM FLOAT CHAMBER
Fig. 25.18 Float chamber to control vacuum
Fig.25.17 Vacuum producer using water jet with producer shown in Fig. 25.17.
motor-driven centrifugal pump to supply power water. (Courtesy Skidmore Corp.)

of a vacuum pump is selected for the nonnal leakage to be encountered after the pumps served by the
system have been primed. When one of the units served has been out of service and it is desired to
reestablish its prime, the amount of air to be removed is relatively large and the priming time would be
quite long if a single vacuum pump were used. Furthennore, there would be a reduction in the vacuum
in the priming system particularly when another pump is being primed. This would be objectionable if
anyone of the units in service had accumulated air and the vent valve had opened to vent this air.
It is usual practice to have the control of one vacuum pump switched on at some predetennined
vacuum and the control of the second switched on at a slightly higher vacuum. Thus if the capacity of
one pump is sufficient to prevent the vacuum in the system from falling, the second pump need not
start. Generally, the controls are equipped with switches so that the sequence of starting can be changed
and the use of the two pumps can be equalized.

STANDBY UNITS

A station having a central automatic priming system, using an electric-motar-driven vacuum pump, may
have standby units with auxiliary drive in case of electrical power failure. These standby units are usually
kept fully primed by the central priming system. If these units are manually started and controlled, a
standby, manually controlled vacuum producer is generally installed. Thus in a vacuum-tank type of
priming system, with gasoline-engine-driven standby pumps. a gasoline-engine-driven vacuum pump
would also be installed and the operator would manually start and stop it as required to maintain the
required range of vacuum in the vacuum tank.
Standby priming equipment that is automatically controlled presents a more complicated problem.
700 Priming

z
o
~
u
~ ~~:::::, ~-.--t~:::::::L--'----:
VACUUM END ELEVATION
CONNECTION
TO PUMP

Fig. 25.19 Connections for float chamber shown Fig. 25.20 Priming device using water jet with
in Fig. 25.18. electrode and solenoid valve control.
(Courtesy Skidmore Corp.) (Courtesy DeLaval Steam Turbine Co.)

One solution used in units that are driven by internal combustion engines is to connect a continuously
running .vacuum pump directly to a relief valve on its suction line or to an unloader in the form of
diaphragm valve controlled by the discharge pressure of the centrifugal pump. In either case, proper
vent valves and check valves have to be incorporated into the priming piping. For units driven by steam
turbines, automatically controlled steam jet primers can be used. Most systems of this nature are especially
designed for the requirements of the installation.

POINT OF AIR REMOVAL

Air is usually exhausted from a centrifugal pump that is being primed through the high point of the
volute. Generally, it is possible to start a centrifugal pump if the liquid covers the eye of the impeller.
The remaining air entrapped in the casing will be driven out by the flow of liquid through the pump.
On some units, the priming is accomplished solely from the high point in the suction waterways. On
Priming 701

Fig. 25.21 Centrifugal pump equipped with priming device in Fig. 25.20.
(Courtesy DeLaval Steam Turbine Co.)

other units, notably those of large size, air is exhausted both from the high point of the volute and the
high point or points of the suction waterways.

AUTOMATIC VENT VALVES

Automatic vent valves consist of a body containing a float that actuates a valve located in the upper
part of the body. The bottom of the body is connected to the space to be vented. As air or gas is vented
out of the valve, water rises in the body until the float is lifted and the valve is closed. A typical vent
designed principally for vacuum priming systems is illustrated in Fig. 25.23. The valve is provided with
auxiliary openings on the lower part of the body for connection to the auxiliary vent points in the system.
When one or more of these vent points is at higher pressure, such as at the top of the volute of the
pump, an orifice is used to limit the flow of liquid. This constant flow of liquid from the discharge back
to the suction means a constant loss of effective capacity. When a unit is used more or less constantly,
a separate valve is desirable.
In some systems it is necessary to prevent reverse flow of air through the valve. This can be done
702 Priming

Fig.25.22 Vacuum tank type of priming unit with two vacuum pumps.
(Courtesy Valve and Primer Corp.)

by adding a check valve between the vent and the vacuum source in the vent piping. One make of
automatic vent valve is designed so that a ball can be inserted to act as a check valve.
Ip. addition to their use with automatic priming systems, automatic vent valves are employed alone,
without connection to a vacuum source, to vent air from the passages of pumps and pipes that are under
greater than atmospheric pressures. They are used frequently to vent air from pumps that have a positive
suction head, thus ensuring that the pump is always primed.

SELF·CONTAINED UNITS

Today, centrifugal pumps are available with various designs of priming equipment that render them self-
contained units. Some have automatic priming devices, which are basically attachments to the pump,
and become inactive after the priming is accomplished. Other units, which are self-priming pumps,
incorporate a hydraulic device that can function as a wet-vacuum pump during the priming period. For
stationary use, the automatically primed type is more efficient. The self-priming designs are usually
more compact and are better for portable or semi portable use.
Priming 703

SECTION A - A

Fig. 25.23 Automatic vent valve.


(Courtesy Nash Engineering Co.)

SELF·PRIMING UNITS

In self-priming units, the pump design incorporates a chamber or other device that recirculates entrapped
air. The air can be removed from the suction waterways and discharged into the discharge line by means
of an ejector. When the presence of air in the discharge line is objectionable, an automatically primed
unit would be preferred. Since the self-priming type was discussed in Chapter 15, this discussion will be
restricted to priming methods and systems that may be used in conventionally designed centrifugal pumps.
Figure 25.24 illustrates an interesting combination of a priming tank and an ejector. In effect, the
regular centrifugal pump is made into a self-priming unit. It consists of a regular centrifugal pump with
a priming tank that is smaller than would be required if a priming tank: alone was used. For proper
functioning there must be a check valve in the suction line, but none in the discharge line. The discharge
line must be vertical, and the volume of the vertical discharge pipe must equal the volume of the priming
tank plus the suction piping to the check valve. When first installed, the pump, the priming tank:, the
suction line back to the check valve, and the discharge pipe are filled with water level with the top of
the priming tank:. Two electrodes are installed in the tank and incorporated in the control of the motor.
Thus, the unit cannot start unless the water covers the top electrode and will stop if the water falls below
the lower electrode.
When the pump is started, the water is drawn from the priming tank: and the suction line between
the pump and the check valve. Thus a vacuum is established, and air is drawn through the check valve
from the balance of the suction line (Fig. 25.24). At the same time the stream of water leaving the drain
nozzle traps the air in A and carries it into the pump, which discharges it with the water. When the
water in the priming tank is pumped out, the lower electrode is uncovered and the control stops the
pump. The suction line check valve closes and water from the discharge flows back through the pump,
704 Priming

Fig. 25.24 Self-priming unit using small priming tarue


(Courtesy Barret-Huentjens Co.)

forcing the air in the priming tank (and in the suction line up to the check valve) through the vent pipe
and its check valve into the discharge pipe, until the upper electrode is again submerged. The pump
starts again, and more air is drawn from the suction line beyond the check valve. When the suction line
has been evacuated sufficiently and water partly fills the horizontal run, the water forms a wall at B that
aids in carrying air into the pump. As the horizontal run of suction pipe fills, the flow of water through
the pump increases until normal capacity is reached. The flow of water past C draws the air from the
priming tank until it is again filled with water. After the prime has been established and all air has been
exhausted from the priming tank, the priming tank becomes inactive. This device permits priming of a
system in which the suction line has a large volume and the priming tank a small volume. The number
of cycles needed to prime the system depends on the volume to be evacuated.
The manufacturer of this unit also makes a separate primer that operates on the same principle for
use with regular centrifugal pumps.

AUTOMATICALLY PRIMED UNITS

Many automatically controlled priming systems for attachment to centrifugal pumps have been developed
in recent years. A pump equipped with such a device is called an automatically primed pump.
A rotary wet vacuum pump, either connected directly or driven by a separate motor, is used in most
automatically primed, motor-driven pumps. With the direct-connected vacuum pump, the controls that
Priming 705

Fig. 25.25 Two views of automatically primed pump with direct-connected vacuum pump.

function when the centrifugal pump is primed open the vacuum pump suction line to the atmosphere,
enabling it to operate without a load. With the separately driven vacuum pump, the controls that function
when the centrifugal pump is primed stop the vacuum pump.

Direct-Connected Vacuum Pump Unit


A typical unit with a direct-connected vacuum pump is shown in Fig. 25.25. The centrifugal pump
is mounted on one end of the motor and the vacuum pump, which runs all the time, is mounted on the
opposite end. The suction of the vacuum pump is connected to the vent on the suction of the centrifugal
pump and to a pressure-operated valve that is closed when the centrifugal pump is not generating pressure.
When the unit is started, the vacuum pump exhausts the air from the centrifugal pump and its suction
line until the centrifugal pump is primed. When the centrifugal pump becomes primed, it generates
pressure and the pressure-operated valve opens, allowing outside air to flow into the suction of the
vacuum pump. The branch of the vacuum pump that is connected to the centrifugal pump suction vent
has a check valve to prevent back flow of air when the pressure-operated valve opens, as well as a
strainer to prevent dirt or foreign material from being carried over to the vacuum pump. The connection
from the discharge waterways of the centrifugal pump to the pressure valve is also connected to the
vacuum pump suction line through an orifice. This provides a small flow of sealing water for the vacuum
pump. A check valve is required in this line in front of the branch.
Another type of direct-connected vacuum pump unit is illustrated in Fig. 25.26. This type utilizes a
float valve on the vent on the suction of the centrifugal pump. The vacuum line has a relief valve set
to open on high vacuum, so that the vacuum pump is not required to pull against a closed suction. This
unit incorporates a sealing liquid reservoir also. Other types of direct-connected vacuum pump units
might include a float-actuated valve that shuts off the vent of the pump from the vacuum pump and
opens the suction line of the vacuum pump to the atmosphere, or a pressure switch controlling a solenoid
valve in the vent pipe either by a vacuum relief valve or three-way solenoid valve.

Separately Driven Vacuum Pump Unit


An automatically primed pump using a separate motor-driven vacuum pump is usually electrically
connected to enable the main pump motor and the vacuum pump motor to start at the same time. The
706 Priming

Fig.25.26 Automatically primed pump with direct-connected vacuum pump.

Fig. 25.27 Automatically primed pump with separately driven vacuum pump.
Priming 707

suction of the vacuum pump is connected to vents at the high point of the suction waterways of the
main pump and evacuates the air in the waterways until the pump is primed. When the pump is primed,
it actuates a pressure switch in the circuit of the vacuum pump motor. This switch opens to stop the
pump. A check valve must be located in the line from the vacuum pump to the suction vent connection
to prevent back flow of air when the vacuum pump is idle. Sealing water is usually provided from a
reservoir in the base.
With this and some other automatic priming systems, it is necessary for the centrifugal pump to run
empty during the time required for priming. This can be done if the clearance between the rings is ample
and the size of the vacuum pump is sufficient to prime the centrifugal pump in less than two minutes.
In normal service there is generally some water trapped in the volute of the centrifugal pump-this aids
in lubricating the rings during the priming period. A unit with this type of priming is illustrated in
Fig. 25.27.

SYSTEMS WITHOUT VENT VALVES

When a pump is primed from the suction side, it is possible to have the vent piping from the pump to
the vacuum producer in the form of an inverted loop of sufficient height to prevent the water or liquid
being pumped from being carried over into the vacuum producer. Theoretically, a loop 10.4 m (34 ft)
above the water level is correct for water. However, if there is leakage, air bubbles in the water will
lower the net density of the column and a loop extending 15.2 m (50 ft) or more above the water level
would be necessary to prevent liquid from being carried over when the vacuum produced is high.
Generally, a system incorporating such a high loop is difficult to accommodate in a station, and automatic
vent valves and other devices may be used more advantageously.
Some systems use risers without vent valves; the vacuum producer is controlled by a vacuum device
that keeps the water level in the risers within a range corresponding to the vacuum range. If there is
considerable variation in the suction level, the height of the riser has to be high enough to match the
water level in the riser resulting from the maximum suction water level and the maximum vacuum.
Naturally, the minimum vacuum must produce a water level in the riser slightly above the pump with
the minimum water level in the suction supply.
Another system that does not involve the use of a vent valve is illustrated in Fig. 25.28. This system
can also be used as a central priming system to serve pumps taking suction from the same source.

SYSTEMS FOR SEWAGE PUMPS

A pump handling sewage or similar liquids that contain stringy material can be equipped with automatic
priming. However, a special type must be used. One system (Fig. 25.29) uses a tee on the suction line
immediately adjacent to the pump suction nozzle, with a vertical riser mounted on the top outlet of the
tee. This riser is blanked at the top, thus forming a small tank. Two electrodes are located in the tank
at different levels. The top of the tank is vented to a vacuum system through a solenoid valve. This
solenoid valve is controlled electrically, closing if the liquid level reaches the top electrode and opening
iJ the liquid level falls below the level of the lower electrode. A third electrode, installed at a still lower
level but slightly above the pump suction, is used to control the pump motor, stopping the motor if the
level of liquid falls to the point at which this third electrode is uncovered.
The automatic priming system described previously (see Fig. 25.27) that uses a separate motor-driven
vacuum pump controlled by a pressure switch actuated by discharge pressure has also been successfully
used on sewage pumps by incorporating an inverted vertical loop in the vacuum pump suction line. This
708 Priming

VALVE
VACWM PUIF STOPS
WtEN WMrER REACHES
THIS LEVEL
TO SUCTION OF

~-----II------A=~~ MOTOR DRIVEN
VACUUM PUMP

~
"'CUUM PUMP STARTS TANK SWITCH
WHEN WATER FALLS CONTROLLING
TO THIS LEVEL MOTOR OF
VACUUM PUMP

-
PRIMING SYSTEM WITHOUT VENT VALVES
USING ELEVATED TANK WITH MAINTAINED
WATER LEVEL

Fig. 25.28 Priming system without vent valves using elevated tank with maintained water level.

prevents the sewage from being carried over into the vacuum pump because the pump shuts down before
the liquid reaches the top of the loop.

SYSTEMS FOR AIR-CHARGED WATERS

Some types of water, particularly from wells, have considerable dissolved gas that is liberated when the
water is pumped with suction lift. In such installations an air-separating tank (also called a priming tank
or an air eliminator) should be used in the suction line. One type using a float-operated vent valve to
permit the withdrawal of the air or gas is shown in Fig. 25.30. Another arrangement that is often used
utilizes a float valve mounted On the side of the tank which controls directly the starting and stopping
of the vacuum pump. With this arrangement there is danger of frequent, repeated starting and stopping
of the vacuum pump unless the air-separating tank is relatively large and the capacity of the vacuum
pump is just a little larger than the amount of air to be evacuated.
When sand is present as an impurity in the water, the air-separating tank can be made to function as
a sand trap (Figs. 25.31 and 25.32).

SYSTEMS FOR UNITS DRIVEN BY GASOLINE OR DIESEL ENGINES

Automatic priming is desirable for any centrifugal pump installation in which air leakage in the suction
line might occur. In a unit driven by a diesel engine, an automatic priming system using motor-driven
pumps can be used if a reliable source of electric power is available in the station.
Priming 709

-ri-r Ir==tp4~~ ________________-L~~~~~


: ~ . : . E:~:-:-':_llc:
'_;c..··~·",-:·-··---------------,-,-,·-,->
·
\
\ \
--..rr.~~E~~~
;' ...... -- '.
,.t.(':.~.----- -- -------
. ..
"" -""t'~ -
itl;~~" :~-I)i
Hi ._. - .,.~ ..,.,
DISCHARGE
• ... ---- - --------
'''CHECK'"GilE~
VALVE VALVE

SEWAGE CHAMBER

ELEVATION

Fig. 25.29 Priming system without vent valves using riser with maintained water level.
KEY:
A = solenoid-operated air valve
B = electrode unit
C = suction submergence chamber
D = suction to vacuum pump
E = sewage pump
F = vacuum pump--motor controlled by electrode relay
G = main pump motor starter
H = induction relay_
(Courtesy DeLaval Steam Turbine Co.)

An auxiliary gasoline-engine-driven vacuum pump for emergency use might be desirable in case of
electric power failure. If a reliable source of electric power is not available, a direct-connected wet
vacuum pump with controls similar to those used in motor-driven automatically primed units is very
satisfactory (Fig. 25.33).
The choice of the priming device for a gasoline-engine-driven centrifugal pump depends on the size
of the pump, the required frequency of priming and the portability of the unit. Most portable gasoline-
engine-driven pumps are used for relatively low heads and small capacities, for use in pumping out
710 Priming

Ca.!MI!:CTIOII

Fig. 25.30 Air-separating chamber for 4- to 12-in. suction lines.


(Courtesy DeLaval Steam Turbine Co.)

excavations and ditches, for example. Self-priming pumps of various types are most satisfactory for this
service and are preferable to regular centrifugal pumps.
Some relatively small-capacity high-head portable gasoline-engine-driven pumps (notably units made
for auxiliary fire pump service), utilize the vacuum in the intake manifold of a gasoline engine as a
means of priming or keeping the pump primed. The rate at which the air can be drawn from the pump
is relatively low, so that many of these units use foot valves. They are initially primed by filling the
pump through a funnel or by means of a hand primer. When the vacuum in the manifold is utilized,
water must be prevented from being drawn over into the manifold.
For larger volume low-head portable units, a good priming unit is a wet-vacuum pump belt driven
from the main shaft by a tight and a loose pulley (Fig. 25.34). The wet-vacuum pump can then be
stopped when the pump is primed.
For permanent installations, a separate wet-vacuum pump driven by a small gasoline engine is
generally preferred.
Priming 711

Fig. 25.31 Installation with automatic priming equipment utilizing combined air-separating and sand chamber.
(Courtesy DeLaval Steam Turbine Co.)

SYSTEMS FOR STEAM-TURBINE-DRIVEN UNITS

The most logical means of priming a steam-turbine-driven pump is with a steam air ejector. If the turbine
is a condensing unit, the vacuum existing in the condenser can be used with an automatic vent valve to
keep the pump primed automatically. However, if there is considerable leakage, the regular steam air
ejector serving the condenser may not be able to handle this additional volume of air and the condenser
performance will be affected. A regular steam condenser should not be used as a source of vacuum for
priming unless the suction lines are tight and the condenser is served by an oversize ejector. With this
arrangement care must be taken to prevent leakage of water from the pump into the condenser in order
to avoid contamination of the condensate. This requires a constant check of the automatic vent valve to
insure that it seats tightly.
With a noncondensing steam-turbine-driven centrifugal pump, a semiautomatic priming system (Fig.
25.35) can be used if there is danger of air leakage into the suction line.
712 Priming

AUTOMATIC VACUUM SWITCHES


~~m:~ ::11: ~~~~~ ~~ :~ ::: ~-F9-:::;___

SWITCHBOARD . SUCTION 'ROIl WEllS


rOR INCOMING LINE
CIJ~RENT,MOTOR CON-
TROLS,ANO INOliCTION SECTION A-A
RELAYS

PLAN

Fig.25.32 Drawing of installation in Fig. 25.31.


(Courtesy DeLaval Steam Turbine Co.)

Fig. 25.33 Series pump with attached vacuum pump and controls to keep unit primed when in operation.
Primim! 713

Fig. 25.34 Pump driven by a gasoline engine with belt-driven priming pump using tight and loose pulley.

STEAM JET PRIMER


CHECK VALVE

\ /
-
TO _5TE DIAPHRAGM STEAM VALVE
V CONTROLLED BY WATER
PRESSURE

-.-rER LINE: TO 'taLVE V

WATER "OT "'LYE SPRING Oft


RANGE OF _TER SUPPLY """"""--- WEIGHT LOADED DlAftH-
LEVEL IN RISER - - - - RAGM TYPE GONTROLLED
PPE TO DRAIN BY VACUUM IN RISER
PIPE. CONTROLS WAnR
PRESSURE ON "'LVE Y

Fig. 25.35 Automatically controlled steam jet primer controlled by vacuum in riser.
714 Priming

TIME REQUIRED FOR PRIMING

The time required to prime a pump with a vacuum-producing device depends on (1) the total volume
to be exhausted, (2) the initial and final vacuum, and (3) the capacity of the vacuum producer over the
range of vacuums that will result in the priming cycle. The actual calculations for determining the time
necessary to prime a pump are complicated. To permit close approximations, jet primers are usually
rated in net capacity for various lifts. It is necessary to divide the volume to be exhausted by the rating
to obtain the approximate priming time. Unless such a simplified method of calculating the required
time is available, the selection of the size of primer should be left to the vendor of the equipment.
Central automatic priming systems are usually rated for the total volume to be kept primed. The time
initially required to prime each unit served by the system is not usually considered, as the basic function
of the system is to keep the pumps primed and in operating condition at all times.

USE OF RELIEF VALVES

The power required by a vacuum pump increases with the magnitude of the vacuum. If the vacuum
pump is allowed to operate uncontrolled after the vent valves have closed, it would continue to exhaust
air from the piping and vacuum tank (if they are used) until it had established its maximum vacuum.
This would result in an increased load on the driver. To avoid this difficulty, most priming systems of
an automatic or semiautomatic type using power-driven vacuum pumps incorporate a relief valve in the
suction line of the vacuum pump. This valve is adjusted to admit air to the line if the vacuum exceeds
a value above that needed for priming.

PREVENTION OF UNPRIMED OPERATION

Various controls may be used to prevent the operation of a pump when it is unprimed. These controls
depend on the type of priming system used. For most installations, a form of float switch in a chamber
connected with the suction line is used. If the level in the chamber or tank is above the impeller eye of
the pump, the float switch control allows the pump to operate. If the liquid falls below a safe level, the
float switch acts through the control to stop the pump or to prevent its being started. In the system
illustrated in Fig. 25.29, three electrodes are placed in the liquid. When the level drops below the longest
electrode, the pump stops or, if stopped, will not start.
Since a great number of automatic priming devices and systems are available, care should be taken
to use the type or variation best suited to the application. This discussion and the accompanying
illustrations do not, of course, cover all makes and modification for every specific application.
An automatic priming system will often allow units to be operated with excessive air leakage into
the suction lines. This is a poor practice because it requires the operation of the vacuum producer for
a greater part of the time.
IV
SERVICES and SELECTION
of PUMPS
26
Services

Centrifugal pumps are used wherever any quantity of liquid must be moved from one place to another.
Centrifugal pumps are found in such services as steam power plants; water supply plants; sewage,
drainage, or irrigation; oil refineries, chemical plants, and steel mills; food processing factories, and
mines; dredging or jetting operations; hydraulic power service; and almost every ship, whether driven
by steam or diesel engine. Although these pumps have much in common, they are varied to meet the
special requirements and particular needs of each service. Although it would be a monumental task to
present all the requirements of each different type of application and every individual particularity that
distinguishes pumps on different services, this book would not be complete without a general discussion
of these requirements as they apply to some of the most common applications.

Steam Power Plants


Power is produced in a steam power plant by supplying heat energy to the feedwater, changing it into
steam under pressure, and then transforming part of this energy into mechanical energy in a steam engine
or steam turbine to do useful work. The feedwater, therefore, acts merely as a conveyor of energy. The
basic elements of a steam power plant include the heat engine, the boiler, and a means of getting water
into the boiler. Modem power plants use steam turbines as heat engines and centrifugal boiler feed
pumps for feeding water to the boiler, except in very small installations.
The basic cycle is improved by connecting a condenser to the steam turbine exhaust, thereby increasing
the pressure drop through the turbine and using more of the energy in the steam. The condenser also
recovers the feedwater, almost entirely eliminating makeup. The cycle is further refined by heating the
feedwater with steam extracted from an intermediate stage of the main turbine. This results in an
improvement ofthe cycle efficiency, provides de aeration of the feedwater, and eliminates the introduction
of cold water into the boiler and the resulting temperature strains on the latter. The combination of the
condensing and feed heating cycle (Fig. 26.1) requires a minimum of three pumps: The condensate
pump, which transfers the condensate from the condenser hotwell into the direct contact heater; the

717

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
718 Services

ITEAM

DIRECT BOILER
CONTACT
HEATER

CONDENSATE
PUMP

BOILER FEED
PUMP

Fig. 26.1 Simple steam power plant cycle.

boiler feed pump; and a circulating pump, which forces cold water through the condenser tubes to
condense the exhaust steam. The cycle illustrated in Fig. 26.1 is very common and is used in most small
steam power plants. A number of auxiliary services not illustrated in Fig. 26.1 are normally used, such
as heater drain pumps, service cooling water pumps, oil-circulating pumps, ash sluice pumps, and pumps
for various applications in flue gas desulfurization.
As the size and number of central steam-electric generating stations grew, the desire for improvements
in operating economy dictated many refinements in the steam cycle. These refinements created new or
altered services for power plant centrifugal pumping equipment. Some of these refinements involved a
steady increase in operating pressures, with 165-bar (2,400-psi) turbines becoming the normal for
subcritical central stations. Starting in the 1960s many plants operating at supercritical steam pressures
were put into service, one an experimental unit operating at 310 bar (4,500 psi), the remainder at 240
bar (3,500 psi). Operating experience since then has shown that supercritical plants achieve high availabil-
ity when run as base load units, but not on swing service. Given this, most of the central station plants
now being built are designed to operate at 165 bar (2,400 psi) to allow greater operating flexibility.
Other refinements were directed toward a greater utilization of heat through increased feed heating,
introducing a need for heater drain pumps--equipment with definite problems of its own. The introduction
of forced or controlled circulation as opposed to natural circulation in boilers created a demand for
pumping equipment of an entirely special character.
The problem presented by the introduction of multiple-stage heaters between the condensate pump
and the boiler centered in the choice between direct-contact heaters and closed heat exchangers. From
the thermodynamic point of view, the direct-contact heaters have certain advantages. A separate pump
Services 719

STEAM

BOILER

BOILER FEEO PUMP

Fig. 26.2 Open feedwater cycle with one deaerator and several closed heaters.

would be required after each direct-contact heater, however, whereas the use of a group of closed heaters
pennits a single boiler feed pump to discharge through these heaters and into the boiler. The average
power plant is based on a compromise system: one direct-contact heater is used for feedwater deaeration,
and several closed heaters are located upstream and downstream of the direct-contact heater and the
boiler feed pump (Fig. 26.2). Such a cycle is known as an "open cycle." The major variation of this
method is the "closed cycle," where the de aeration is accomplished in the condenser hotwell and all
heaters are of the closed type (Fig. 26.3).
In the last 20 years two new concepts have arisen to supplement the traditional steam power plant.
The first and earlier is cogeneration, a process in which steam is supplied simultaneously to a process
and to a conventional steam power plant, the objective being to take advantage of a source of low-cost
fuel. Initially small, cogeneration unit ratings have now risen to 350 MW. A feature of most cogeneration
plants is two boilers: low pressure for process steam and high pressure for power generation. These two
boilers can be fed by two separate boiler feed pumps, or by a single pump with a low-pressure take-off.
The second concept, which came into prominence in the mid-1980s, is a power plant design known
as combined cycle. In this design a combustion gas turbine (Brayton cycle) drives one generator, while
a small steam turbine (Rankine cycle), using steam raised by a heat recovery boiler at the turbine's
exhaust, drives another. The steam segment of the plant follows the cycle shown in Fig. 26.1. Combined
cycle units rated up to 250 MW are now in service, with the steam plant generating around one-third
of the total output. The advantage of combined cycle units is higher thermal efficiency, on the order of
55 percent, which makes them more economical than conventional steam power plants in regions where
the cost of fuel is high.
720 Services

c=::::l STEAM
_WATER

STEAM ,'URBINE

BOILER

BOILER FEED PUMP

Fig. 26.3 Closed feed water cycle.

BOILER FEED PUMPS

A listing of conditions of service of boiler feed pumps should include not only the pump capacity,
suction conditions and feedwater temperature, and discharge pressure, but also such data as the chemical
analysis of the feedwater, the pH at pumping temperature, the variation in suction pressure and temperature
if any, and all other pertinent information that may reflect on the hydraulic and mechanical design of
the boiler feed pumps. Whenever the power plant designer has any doubts, he should submit a complete
layout of the feedwater system and of the heat balance diagram to the boiler feed pump manufacturer.
After studying this layout, the manufacturer will often be able to suggest an alternate arrangement of
the equipment that will result in more economical operation, in decreasing the first cost of the installation,
or even in improving the life of the equipment, thereby reducing the eventual maintenance expense.

Boiler Feed Pump Capacity


Once the maximum flow to the boiler has been determined, the boiler feed pump total capacity can
be established by adding a margin to this maximum flow to cover boiler swings and the eventual reduction
in effective capacity from wear. This margin varies from 20 percent in small plants to 8 percent in the
larger central stations.
Heat balance calculations as well as turbine and boiler guarantees are always expressed in kilograms
per second in metric units (pounds per hour in US units), whereas it is customary to define the service
to be performed by centrifugal pumps in terms of cubic meters· per hour (gallons per minute) against a
head of so many meters (feet). To convert kg/s to m3/hr, use the formula: m3/hr = 3,600(kg/s)/(density).
Similarly, to convert lb/hr to gpm, use the formula: gpm = Ib/hr(500 x SG). Table 26.I(a) lists values
of the density of water over the temperature range 0.01 to 374.15°C. Values of the specific gravity of
water for different feedwater temperatures are given in Fig. 26.4.
The total required capacity must then either be handled by a single pump or divided between several
Services 721

Table 26.1(a) Vapor Pressure &


Density of Water from O.OI°C to
374.15°C (Adapted from table
xl.4.320. Sulzer Centrifugal Pump
Handbook)

e Celsius temperature
p Pressure
p Density

e p p
°C bar kg/m)

0.01 0.006112 999.8


10 0.012271 999.7
20 0.023368 998.3
30 0.042417 995.7
40 0.073749 992.3
50 0.12334 988.0
60 0.19919 983.2
70 0.31161 977.7
80 0.47359 971.6
90 0.70108 965.2
100 1.0132 958.1
110 1.4326 950.7
120 1.9854 942.9
130 2.7012 934.6
140 3.6136 925.8
150 4.7597 916.8
160 6.1804 907.3
170 7.9202 897.3
180 10.003 886.9
190 12.552 876.0
200 15.551 864.7
210 19.080 852.8
220 23.201 840.3
230 27.979 827.3
240 33.480 813.6
250 39.776 799.2
260 46.940 783.9
270 55.051 767.8
280 64.191 750.5
290 74.448 732.1
300 85.917 712.2
310 98.697 690.6
320 112.90 666.9
330 128.65 640.5
340 146.08 610.3
350 165.37 574.5
360 186.74 528.3
370 210.53 448.3
374.15 221.20 315.5
722 Services

Table 26.1(b) Temperature-Vapor Pressure Relationship of Water


At sea level, the saturation pressure or vapor pressure (PS/G) = vapor pressure
(PS/A-14.7)

Temperature, Temperature, Temperature,


degF Vapor pressure, psia degF Vapor pressure, psia degF Vapor pressure, psia

32 0.088 232 21.58 400 247.3


35 0.100 236 23.22 405 261.7
40 0.122 240 24.97 410 276.8
45 0.148 244 26.83 415 292.4
50 0.178 248 28.80 420 308.8
55 0.214 252 30.88 425 325.9
60 0.256 256 33.09 430 343.7
65 0.306 260 35.43 435 362.3
70 0.363 264 37.90 440 381.6
75 0.430 268 40.50 445 401.7
80 0.507 272 43.25 450 422.6
85 0.596 276 46.15 455 444.3
90 0.698 280 49.20 460 466.9
95 0.815 284 52.42 465 490.3
100 0.949 288 55.80 470 514.7
105 1.102 292 59.36 475 539.9
110 1.275 296 63.09 480 566.1
115 1.471 300 67.01 485 593.3
120 1.692 304 71.13 490 621.4
125 1.942 308 75.44 495 650.6
130 2.222 312 79.96 500 680.8
135 2.537 316 84.70 505 712.0
140 2.889 320 89.66 510 744.3
145 3.281 324 94.84 515 777.8
150 3.718 328 100.3 520 812.4
155 4.203 332 105.9 525 848.1
160 4.741 336 111.8 530 885.0
165 5.335 340 118.0 535 .923.2
170 5.992 344 124.4 540 962.5
175 6.715 348 131.2 545 1003.
180 7.510 352 138.2 550 1045.
185 8.383 356 145.4 555 1088.
190 9.339 360 153.0 560 1133.
195 10.385 364 160.9 565 1179.
200 11.526 368 169.2 570 1226.
204 12.512 372 177.7 575 1275.
208 13.568 376 186.6 580 1326.
212 14.70 380 195.8 585 1378.
216 15.90 384 205.3 590 1431.
220 17.19 388 215.3 595 1486.
224 18.56 392 225.6 600 1543.
228 20.02 396 236.2
Services 723

TEMPERATURE, IN DEGREES FAHRENHEIT

0.960

0.96!!

0.970 0.1J1~1111

0.97!!

0.980

...
II>

0.98S - ._.0'""lf.IHIfIflHF!lHIH
'"
n
;;
;:;
0.990

o.st!!

1.000 o.t!K@l!ilillll
0.870

0.890
300 3SO
ISO 200
TEMPERATURE, IN DEGREES FAHRENKEIT

Fig. 26.4 Specific gravity of water at temperatures from 30 to 750°F.


Curve A (from 30 to 200°F) is from u.s. Bureau of Standards' circular no. 19; curve B (from 200 to 350°F).
curve C (from 350 to 550°F). and curve D (from 550 to 750°F) are from Goodenough Steam Tables . All curves
are based on specific gravity of water at 62°F = 1.000.

duplicate pumps operating in parallel. The number of pumps is determined primarily by the intended
operating mode of the plant. Industrial and cogeneration plants are expected to swing over a wide load
range and therefore generally use at least two half-capacity pumps. Central stations intended for base
loading normally use single full-capacity pumps for units rated up to 350 MW and two half-capacity
pumps for larger units. Exceptions to this practice occur as follows: Some engineers prefer to use multiple
pumps even for units smaller than 350 MW, particularly when swing loading is expected; and steam-
turbine-driven boiler feed pumps designed for full capacity have been used for units as large as 1,300 MW.
In most cases, a spare boiler feed pump is included. In central stations, however, it has become
common to eliminate spare pumps when two half-capacity pumps are used and, in some cases, even if
a single full-capacity boiler feed pump is installed. Units rated 350 MW and larger generally have a
startup pump, designed for between 15 and 50 percent capacity, to avoid prolonged operation of a large
pump at low flow.

Suction Conditions

As discussed in Chapter 19, the net positive suction head or NPSH represents the net suction head
at the pump suction, corrected to the pump centerline, over and above the vapor pressure of the feed water.
724 Services

If the pump takes its suction from a deaerating heater, as in Fig. 26.2, the feedwater in the storage space
is under a pressure equivalent to the vapor pressure corresponding to its temperature. Therefore, the
only energy available at the first-stage impeller over and above the vapor pressure is the static submergence
between the water level in the storage space and the pump centerline, less the friction losses in the
intervening piping. Tables 26.1(a) and (b) give the relationship between water vapor pressure and temper-
ature.
The required NPSH is independent of the operating temperature by virtue of its definition (Chap.
19). Practically, this temperature must be taken into account when establishing the recommended submer-
gence from the de aerator to the boiler feed pump. A margin of safety must be added to the boiler feed
pumps required NPSH to protect them against the transient conditions that follow a sudden reduction
in load for the turbogenerator. 1 At the same time, for large units employing high-speed boiler feed
pumps, it is necessary to ensure the NPSH margin is sufficient to prevent premature impeller erosion
in the event the pumps are operated for extended periods at low flow (see Chaps. 19 and 22). Table
19.2 shows typical margins for the two broad classes of boiler feed pumps. These are for guidance only;
it is still necessary to check the margin against the characteristics of the pumps' suction system and
first-stage impeller.
Whereas the previous discussion applies primarily to the majority of installations where the boiler
feed pump takes its suction from a deaerating heater, it holds as well in the closed feed cycle (Fig. 26.3).
The discharge pressure of the condensate pump must be carefully established so that the suction pressure
of the boiler feed pump cannot fall below the sum of the vapor pressure at pumping temperature and
the required NPSH.
Normally this is not difficult; however, some installations break up the total pressure to be generated
by the boiler feed pump between two separate pumps and locate some of the closed heaters between
these two pumps. The first of these is called a "primary feed pump" or a "booster condensate pump,"
and the second is called the "secondary feed pump" or the "main feed pump." In the event the primary
pump is operated at variable speed, great care must be taken to meet the requirement given above. The
controlling factor for this speed variation is the flow to the boiler and the controlling mechanism, generally
an excess-pressure regulator (Chap. 23), which varies the speed of the pump in question. This control
is unaffected by the relationship between the temperature and the pressure at the suction of the main
feed pump, and conditions can arise where the speed regulation lowers the suction pressure at the feed
pump dangerously close to, or even below, the required minimum. In such installations, therefore, it is
absolutely necessary to provide some form of lower limit control to the speed of the pump, subject to
variation. This lower limit is determined on the basis of always assuring sufficient NPSH to the main
feed pump.

Discharge Pressure and Total Head


To determine the discharge pressure for which the boiler feed pumps are to be designed, it is necessary
to obtain the sum of the maximum boiler drum pressure and of all the frictional and control losses that
occur between the boiler feed pump discharge and the boiler drum inlet. The calculations of the frictional
losses must, of course, be based upon the maximum capacity previously determined. The required
discharge pressure will generally vary from 115 to 125 percent of the boiler drum pressure.
The net pressure to be generated by the boiler feed pump is the difference between the required
discharge pressure and the available suction pressure. This must be converted into a total head, using
the formula: total head, in meters = (net pressure, in bar x 1O.21)/specific gravity, or in feet = (psi x
2.31)/specific gravity.

lFor a complete treatment of this problem, see I. 1. Karassik et ai., Centrifugal Boiler Feed Pumps under Transient Operating
Conditions. Paper 53-F-32, ASME Fall Meeting, Rochester, N.Y., 1953.
Services 725

Head-Capacity Curve Slope


In the range of specific speeds nonnally encountered in multistage centrifugal boiler feed pumps, the
rise of head from the point of best efficiency will vary from 10 to 25 percent. Further, the shape of the
head-capacity curve for these pumps is such that the drop in head is very slow at low capacities,
accelerating as the capacity is increased (see Fig. 24.50 and 22.4).
If the pump is operated at constant speed, the difference in pressure between the pump head-capacity
urve and the system-head curve must be throttled by the feedwater regulator. Thus, the higher the rise
of head toward shutoff, the more pressure must be throttled off and, theoretically, wasted. Also, the
higher the rise, the greater the pressure to which the discharge piping and the closed heaters will be
subjected. It is not advisable, however, to select too Iowa rise to shutoff because too flat a curve is not
conducive to stable control, as a small change in pressure corresponds to a relatively great change in
capacity, and a design that gives a very low rise to shutoff may result in an unstable head-capacity
curve, difficult to use for parallel operation.
When several boiler feed pumps are to be operated in parallel, they must have stable curves and
equal shutoff heads. Otherwise, the total flow will be divided unevenly and one of the pumps may
actually be backed off the line after a change in required capacity occurs at light flows.

Driver Horsepower
A boiler feed pump is selected for a given capacity and pressure and generally it is not expected to
operate at any capacity beyond the design condition. In other words, in contrast to a standard general-
service pump for which the power consumption may increase with a decrease in head and, consequently,
an increase in capacity, a boiler feed pump has a very definite maximum capacity because it operates
on a system-head curve made up of the boiler .drum pressure plus the friction losses in the discharge.
If, as it should be, the design capacity of the pump is chosen as the maximum capacity that can be
expected under emergency conditions, there can be no further increase under any operating conditions
since the pressure requirement corresponding to an increased capacity would exceed the design pressure
of the pump. Even when the design pressure includes a safety margin, the boiler demand does not exceed
the design capacity and the feed water regulator will impart additional artificial friction losses to increase
the required pressure up to the pressure available at the pump.
When two pumps are operated in parallel to feed a single boiler, the situation is somewhat different.
If one of the pumps is taken off the line at part load, the remaining pump could easily operate at capacities
in excess of its design, since its head-capacity curve would intersect the system-head curve at a point
lower than the design head (Fig. 26.5). In such a case, it is necessary to detennine the pump capacity
at the intersection point; the horsepower corresponding to this capacity will be the maximum expected.
It is not necessary to select a driver so large that it will not be overloaded at any point of the boiler
feed pump operating curve. Electric motors used on boiler feed service, however, generally have an
overload capacity of 15 percent, and it is usually the practice to reserve this overload capacity as a safety
margin and to select a motor that will not be overloaded at the design capacity. Exceptions occur in the
case of very large motor sizes. For example, if the pump power is 2300 kW (3,100 hp), it is logical to
use a 2250 kW (3,000 hp) motor that will be overloaded by about 3 percent rather than a considerably
more expensive 2600 kW (3,500 hp) motor.
Because steam turbines are not built in definite standard sizes but can be designed for any intennediate
rating, they are generally selected with about 5 percent excess power over the maximum expected
pump horsepower.

General Structural Features


As described in Chapter 3, boiler feed pumps used for discharge pressures under 165 bar (2,400 psi)
"rf' now !!enerally designed with axially split casings (Figs. 3.4 through 3.7). An alternative to axially
726 Services

CAPACITY

Fig. 26.5 Method of detennining maximum pump horsepower for two boiler feed pumps operating in parallel.

split casings, favoured by some purchasers for its lower first cost, is the segmental or ring section casing
(Fig. 3.13). This arrangement trades first cost for maintenance cost (see Chap. 3), and therefore is really
better suited to low pressure industrial applications. Pumps with radially split double casings (Figs. 3.14,
3.15, and 3.17), whose virtue is a single, symmetrical high-pressure joint, are used for discharge pressures
above 165 bar (2,400 psi).
The merits of single-suction over double-suction impellers are discussed in Chapter 4, and the means
used to balance the axial thrust of multistage pumps with single-suction impellers are described in
Chapters 5 and 6. The single-suction impeller design has been almost universally adopted for the reasons
presented in those chapters. The exception is the double-suction first-stage impeller used in some very
large capacity boiler feed pumps. The selection of materials for the boiler feed pump casings and internal
parts is described in Chapter 17.

High-Speed, High-Pressure Boiler Feed Pumps


Boiler feed pump design practices and the use of operating speeds above 3,600 rpm for most designs
are easier understood after examining the effect the use of increasing operating steam pressures had on
boiler feed pumps. As pressures increased from 85 to 125, then to 165 and even 240 bar (1,250, 1,800,
2,400, and 3,500 psi), the total head that had to be developed by the pump rose from somewhere around
1,220 m to as high as 2,130 and 3,660 m (4,000 ft to 7,000 and 12,000 ft). The only way to achieve
these higher heads at 3,600 rpm was to increase the number of stages. The pumps, therefore, had to
have longer and longer rotors, which threatened to interfere with the long uniterrupted life between
overhauls to which steam power plant operators were beginning to become accustomed.
To decrease the frequency of maintenance requirements, the rate of wear at the internal running
clearances must be decreased. To evaluate the probable wear rate of several different designs, however,
the causes of wear must first be considered. Wear occurs not only because of the erosive action of the
leakage thru the internal running clearances, but also--to a greater measure-through contact and rubbing
between the two metal parts (a process known as "adhesive wear"), even if this contact is infrequent
and too slight to cause galling and immediate seizure. In other words, the two prime factors affecting
the life of a given high-pressure boiler feed pump, provided the hydraulic design is suitable for the
Services 727

Table 26.2 Comparison of Design Data for High and Low-Speed High-Pressure
Boiler Feed Pumps

Data Low-speed pump High-speed pump

Number of stages 9 4
Pump speed, rpm 3,465 9,000
Impeller diameter, mm (in) 350 (13.8) 213 (8.4)
Bearing span, mm (in) 1,220 (87.4) 1,308 (51.5)
Average shaft diameter at impellers, 121 (4.75) 78 (3.06)
mm (in)
L4/d 2 x 109 mm 2 (U/d2 x 106 in2) 1.66 (2.58) 0.48 (0.75)

pump's intended operating flow range, are the rotor deflection and the internal clearances. The greater
the margin between these two, the less will be the chance of accidental contact and the longer the life
of the equipment. The deflection of a large shaft rotor varies approximately as the fourth power of the
bearing span and inversely as the square of the shaft diameter, as shown in Chapter 7. The logical
solution, then, is to reduce the bearing span by reducing the number of stages. By 1953, experience had
indicated that stage pressures could be increased from the 17 to 24 bar (250 to 350 psi) range commonly
used at 3,600 rpm to as high as 55 or 69 bar (800 or 1,000 psi). More recently, designs developing 97
bar (1,400 psi) per stage have successfully been put into service.
In tum, these higher heads per stage could better be achieved by increasing the speed of rotation
rather than by increasing the impeller diameters. The beneficial effect of high-speed operation on the
shaft deflection can be seen from a comparison of two pumps, each designated for 313 m 3/hr (1,380
gpm) at 2,000 m (6,550 ft); see Table 26.2. This comparison indicates that the high-speed pump will
have less than one-third the deflection of the low-speed pump. A more dramatic comparison is afforded
by Fig. 26.6, which shows the rotors of the two pumps side by side. If the diametral running clearances
are kept the same for both designs, the frequency of momentary contact will be materially reduced, if

3465 rpm

1380 GPM
6550 FT. HEAD

Fig. 26.6 Rotor of motor-driven and high speed boiler feed pumps for same performance.
728 Services

not eliminated altogether. As a result, the wear at the running clearances of the high-speed pump will
be due solely to the erosive action of the leakage. These are the reasons why high-speed boiler feed
pumps are now used in most central stations operating at high pressures.

Medium Pressure Boiler Feed Pumps


The boilers used for industrial applications, cogeneration, and combined cycle power plants, operate
at lower pressures than central stations, typically 40 to 100 bar (600 to 1,500 psi). At the lower pressures,
the number of stages needed to develop the required head are few enough to allow a rotor whose static
deflection is less than the internal clearances, and therefore the pumps usually follow the previously
presented design philosophy for high-pressure pumps.
As the pressure rises, however, so do the number of stages required, yet the capital cost constraints
placed on these plants usually preclude the cost of the drive arrangement needed for high speed operation.
The result is pumps of up to 14. stages whose rotors have static deflections far greater than their internal
clearances. The use of these designs is justified on the basis that when the pump is running its rotor is
supported by the Lomakin effect (see Chap. 7) and under favorable conditions they will achieve between
50,000 and 75,000 hours between overhauls. The dangers with these are, first, frequent start/stop operation
or wide flow swings or both reduces the period between overhauls, often to as little as 8,000 hours, and,
second, if the pump is accidentally flashed or otherwise run dry it will be severely damaged.

Boiler Feed Pump Drives


The majority of boiler feed pumps in small and medium size steam plants are electric motor driven
(Fig. 26.7). In the past, steam-turbine-driven standby pumps were installed as a protection against the
interruption of electric power supply, but this practice has disappeared in central steam stations and is
encountered rarely even in industrial plants.
A trend away from electric motor drive in large central steam stations for units in excess of 150,000
kW has occurred in recent years. The first departure was to steam turbine drive (Fig. 26.8) because

1. An independent steam turbine increases plant capability by eliminating the auxiliary power required for
boiler feeding.
2. Proper utilization of the exhaust steam in the feedwater heaters can improve cycle efficiency.
3. In many cases, the elimination of the boiler feed pump motors may permit a reduction in the station auxiliary
voltage, with an appreciable reduction in cost.
4. Driver speed can be matched exactly to the pump optimum speed.
5. A steam turbine provides variable speed operation without an additional component such as a hydraulic cou-
pling.

During the late 1950s, several U.S. utilities built plants with full-capacity boiler feed pumps driven
from the exciter end of the main turbine generator (Fig. 26.9). Pump speed was varied with a hydraulic
coupling or a magnetic drive between the generator and the pump. A gear was incorporated in the drive
train when a high-speed pump was used. What was gained in eliminating a driver was lost in unit
flexibility, and main shaft drive fell from popUlarity after a few years.

CONDENSATE PUMPS

Condensate pumps take their suction from the condenser hotwell and discharge to either the deaerating
heater in open feedwater systems (Fig. 26.2), or the suction of the boiler feed pumps in closed systems
Services 729

Fig. 26.7 Electric-motor-driven boiler feed pumps in a central steam station.

Fig. 26.8 Steam-turbine-driven boiler feed pump in a central steam station.


730 Services

Fig. 26.9 Installation of main turbine-generator-driven boiler feed pumps.


Hydraulic coupling and step-up gear are used between generators of cross-compound unit and
two half-capacity boiler feed pumps.

(Fig. 26.3). These pumps, therefore, operate with a very low pressure at their suction, that is, from 25
to 75 mm (1 to 3 in) Hg absolute. The available NPSH is obtained by the submergence between the
water level in the condenser hotwell and the centerline of the condensate pump first-stage impeller.
Because of the desire to locate the condenser hotwell at as Iowan elevation in the plant as possible and
to avoid the use of a condensate pump pit, the available NPSH is generally extremely low, of the order
of 0.6 to 1.2 m (2 to 4 ft). The only exception occurs when vertical-can condensate pumps are used,
because they can be installed below ground elevation and, therefore, higher values of submergence can
be obtained. Friction losses on the suction side must be kept to an absolute minimum. The piping
connection from the hotwell to the pump, therefore, should be as direct as possible, of ample size, and
have a minimum of fittings.
Because of the low available NPSH, condensate pumps operate at relatively low speeds, ranging from
1,750 rpm in the low range of capacities to 880 rpm or even less for larger flows.
It is customary to provide a liberal excess capacity margin above the full load steam condensing flow
to take care of the heater drains that may be dumped into the condenser hotwell if the heater drain
pumps are taken out of service for any reason.

Types of Condensate Pumps


Both horizontal and vertical condensate pumps are used. Depending on the total head required,
horizontal pumps may be either single or multistage.
Services 731

Figure 26.10 shows a single-suction, single-stage pump with an axially split casing that is used for
heads up to approximately 30 m (100 ft). It is designed to have discharge pressure on the shaft seal.
The suction opening in the lower half of the casing keeps the suction line at floor level. An oversize
vent at the highest point of the suction chamber permits the escape of all entrained vapors that will be
vented back to the condenser and removed by the air-removal apparatus.
Multistage pumps are used for higher heads. A two-stage pump is shown in Fig. 26.11, with the
impellers facing in opposite directions for axial balance. By turning the impeller suctions toward the
center, both shaft seals are kept under positive pressure to prevent leakage of air into the pump. For
higher heads and larger capacities, a three-stage pump as shown in Fig. 26.12 may be used. The first-
stage impeller is a double-suction type and is located centrally in the pump. The remaining impellers
are single suction types and are also arranged so that both shaft seals are under pressure. Two liberal
vents connecting with the suction volute on each side of the first-stage double-suction impeller permit
the escape of vapor.
Beginning in the late 1950s a trend developed toward the use of canned vertical turbine pumps (Fig.
14.21) for condensate service. The chief advantage of these pumps is that ample submergence can be
provided without building a dry pit. The first stage of this pump is located at the bottom of the bowl
assembly, and the available NPSH is the distance between the water level in the hotwell and the centerline
of the first-stage impeller. As plant sizes have increased, some designers have combined a double suction
volute type first stage with several vertical turbine stages (Fig. 14.22) to develop the same head in a
shorter, although larger diameter, pump. Yet another refinement has been the development of a so-called
"pass-out" condensate pump (Fig. 14.23), which raises the condensate to a partial pressure then passes
it out to an in-line polishing plant, from which the condensate is returned to the pump and raised to the
final pressure.

Fig. 26.10 Single-stage, horizontal condensate pump with axially split casing.
732 Services

Fig. 26.11 Two-stage. horizontal condensate pump with axially split casing.

Although condensate pumps are designed so that their shaft seals are under pressure, air leakage can
still take place into a particular pump, either when partial cavitation of the first-stage impeller reduces
its discharge pressure below atmospheric (in all designs except vertical canned) or when a pump is idle
but the suction valve is left open. To prevent this air leakage, special provisions are always included in
condensate pump shaft seals. Packed boxes are provided with seal cages, with water for gland sealing
taken from the condensate pump discharge manifold beyond all check valves. Mechanical seals are used
in one of two configurations: single or double. Single seals are equipped with throat bushings and sealed
with condensate in the same manner as packed boxes (although the sealing water flow needed to develop
at least 0.35 bar [5 psig] at the seal is much higher). Double seals have condensate circulated between
the seals, with an orifice at the outlet to maintain around 0.70 bar (10 psig) between the seals.

Fig. 26.12 Three-stage. horizontal condensate pump with axially split casing.
Services 733

Materials
Although condensate pumps handle essentially the same water as the boiler feed pumps-very pure
and unbuffered-its low temperature makes it considerably less corrosive. In most small plants, therefore,
standard fitted pumps are used with cast-iron castings and bronze internal parts. Larger units generally
use water treatment that leaves the condensate slightly basic, and it has become common practice to use
condensate pumps with cast-iron casings and 13 chrome, or 13 chrome 4 nickel, internal parts.

Condensate Pump Regulation


When a condensate pump operates in a closed cycle, its capacity is controlled by the same means
that are used to control the delivery of the boiler feed pump. In other words, the two pumps can be
considered as a combined unit insofar as their head-capacity curve is concerned. This head-capacity
curve intersects the system-head curve, and flow is varied either by throttling in the boiler feed pump
discharge and altering the system-head curve or by varying the speed of the boiler feed pump and altering
the combined head-capacity curve of the two pumps.
In an open feedwater system, several means can be used to vary the condensate pump capacity with
the load:

1. The condensate pump head-capacity curve can be changed by varying the pump speed (used very infrequently).
2. The condensate pump head-capacity curve can be altered by allowing the pump to operate in the "break"
(Fig. 19.16 and 19.17).
3. The system-head curve can be artificially changed by throttling the pump discharge by means of a float control.
4. The pump can operate at the intersection of its head-capacity curve and the normal system-head curve. The
net discharge is controlled by bypassing all excess condensate back to the condenser hotwell.
5. Methods 3 and 4 can be combined so that the discharge is throttled back to a predetermined minimum, but
if the load, and consequently the flow of condensate to the hotwell, is reduced below this minimum, the
excess of condensate handled by the pump is bypassed back to the hotwell.

Operation in the "break," or "submergence control," as it has often been called, has been applied
very successfully in a great many installations. Condensate pumps designed for submergence control
require specialized hydraulic design, correct selection of operating speeds, and limitation of stage pres-
sures. The pump is operating in the break (cavitates) at all capacities. However, this cavitation is not
severely destructive in nature because the energy level of the fluid at the point where the vapor bubbles
collapses is insufficient to create a shock wave of a high enough intensity to inflict physical damage on
the pump parts. If, however, higher values of NPSH were required-as with vertical-can condensate
pumps-operation in the break would result in a rapid deterioration of the impellers. It is for this reason
that submergence control is not applicable to can-type condensate pumps.
The main advantages of submergence control are its simplicity and the fact that the power required
for all operating conditions is less than with any other system. Disadvantages occur when the pump is
operating at very light loads, however, as the system head may require as little as one-half of the total
head produced in the normal head-capacity curve. In this case, the first stage of a two-stage pump
produces no head whatsoever, and, if the axial balance was achieved by opposing the two impellers, a
definite thrust is imposed on the thrust bearing, which must be selected with sufficient capacity to
withstand this condition. In addition, no control is available to provide the minimum flow that may be
required through the auxiliaries such as the ejector condenser.
The condensate pump discharge can be throttled by a float control arranged to position a valve that
increases the system-head curve as the level in the hotwell is drawn down. It eliminates the cavitation
in the condensate pump, but at the cost of a slight power increase. Furthermore, the float necessarily
734 Services

, S IPHON VALVE

CONOENSATE
RECIRCULATION
TO MAIN CON O.

\
CONOENSATE
I /
OUTLET

;
,--- -- -
......... .. .. .. ~ --
I

I
I
I

CONDENSATE
INl..ET

Fig. 26.13 Thermostatic control for condensate recirculation.

operates over a narrow range, and the mechanisms tend to be somewhat sluggish in following rapid load
changes, often resulting in capacity and pressure surges.
When condensate delivery is controlled through bypassing, the hotwell float controls a valve in a
bypass line connecting the pump discharge back to the hotwell. At maximum condensate flow, the float
is at its upper limit with the bypass closed and all the condensate is delivered to the system. As the
condensate flow to the hotwell decreases, the hotwell level falls, carrying the float down and opening
the bypass. Sluggish float action can create the same problems of system instability in bypass control
as in throttling control, however, and the power consumption is excessive because the pump always
operates at full capacity.
A combination of the throttling and bypassing control methods eliminates the shortcomings of excessive
power consumption. The minimum flow at which bypassing begins is selected to provide sufficient flow
through the ejector condenser.
A modification of the bypassing control for minimum flow is illustrated in Fig. 26.13, which shows
a thermostatic control for condensate recirculation. With practically constant steam flow through the
ejector, the rise in temperature of the condensate between the inlet and outlet of the ejector condenser
is a close indication of the rate of flow of condensate through the ejector condenser tubes. Therefore,
an automatic device to regulate the rate of flow of the condensate can be controlled by this temperature
differential. A small pipe is connected from the condensate outlet on the ejector condenser back into
the main condenser shell. An automatic valve is installed in this line and is actuated and controlled by
the temperature rise of the condensate. Whenever the temperature rise reaches a certain predetermined
figure, indicating a low flow of condensate, the automatic valve begins to open, allowing some of the
condensate to return to the condenser and then to the condensate pump, which supplies it to the ejector
condenser at the increased rate. When the temperature rise through the ejector condenser is less than
the limiting amount, indicating that ample condensate is flowing through the ejector condenser, the
automatic valve remains closed.
Services 735

HEATER DRAIN PUMPS

Condensate drains from the closed heaters must be returned to the feedwater cycle to avoid wasting
water and the heat content of these drains. Two basic methods are available for this: (1) drains can be
flashed to the steam space of a lower pressure heater or (2) drains can be pumped into the feedwater
cycle at some higher pressure point.
When it leaves the hotwell of a closed heater, condensate is at saturation temperature; if it is passed
to a region of lower pressure, part will flash into steam and part will remain as cooler liquid. Piping
each heater drain to the next lower pressure heater is the simpler mechanical arrangement and requires
no power-driven equipment. This "cascading" is accomplished by an appropriate trap in each heater
drain. A series of heaters can thus be drained by cascading from heater to heater in the order of descending
pressure, the lowest being drained directly to the condenser.
This arrangement, however, introduces a loss of heat since the heat content of the drains from the
lowest pressure heater is dissipated in the condenser by transfer to the circulating water. It is generally
the practice, therefore, to cascade only down to the lowest pressure heater and pump the drains from
that heater back into the feedwater cycle, as shown in Fig. 26.14. Because the pressure in that heater
hotwell is low (frequently below atmospheric even at full load), heater drain pumps on that service are
commonly described as on "low pressure heater drain service."
In an open cycle, drains from heaters located beyond the de aerator are cascaded to the deaerator.
Although the deaerator is generally located above the closed heaters, the difference in pressure is sufficient
to overcome both the static and the friction losses. This difference in pressure decreases with a reduction
in load, however, and at some partial main turbine load it becomes insufficient to evacuate the heater
drains. Unless a pump is used to "boost" the drains back up to the de aerator, they have to be switched
to either a lower pressure heater, or even to the condenser, with a subsequent loss of heat. To avoid

c:::::::::J STEAM

- - ~}~~\..rWATER
_WATER

STEAM

BOILER

LOW PRESSURE HEATER


DRAIN PUMP

Fig. 26.14 Typical arrangement for heater drain pumps.


736 Services

these complications, a "high-pressure heater drain pump" is generally used to transfer these drains to
the deaerator. Actually this pump has a "reverse" system head to work against-at full load, the required
total head may be negative, whereas at light loads the required head is at its maximum.
Heater drain pumps, especially those in high-pressure service, are subject to more severe conditions
than boiler feed pumps encounter:

1. Suction pressure and temperature are higher.


2. Available NPSH is generally extremely limited.
3. Transient conditions during sudden load fluctuations are the same as for boiler feed pumps, but are more severe.

Types of Heater Drain Pumps


In the past, heater drain pumps were always designed as single-stage or multistage horizontal pumps,
depending on total head requirements. In the single-stage type, end suction pumps of the heavier "process
pump" type were preferred for both low-and high-pressure service (see Figs. 1.9 and 13.13). Recently,
however, the vertical-can pump (Fig. 14.21) has been frequently applied on heater drain service, as well
as condensate service.
As previously described, the advantages of the vertical-can pump are lower first cost and additional
NPSH obtained because the first-stage impeller is lowered below floor level in the can. The vertical-
can pump has several disadvantages, however. A horizontal heater drain pump is more easily inspected
than a can pump. The external grease- or oil-lubricated bearings of the horizontal pump are less vulnerable
to the severe operating conditions during swinging loads than the water-lubricated internal bearings of
the can pump.
Although these disadvantages should not exclude the vertical-can pump from consideration, it is wise
to limit its application to low-pressure service and give preference to the horizontal pump for high-
pressure applications.

Materials
The same considerations that dictate the choice of materials of the boiler feed pump should be used
for heater drain pumps. In other words, low-pressure heater drain pumps can use cast-iron casings and
bronze internals if Table 17.7 allows it or if no evidence of corrosion-erosion has been uncovered. On
high-pressure service, 13 chrome steel internals are generally mandatory, and 13 chrome steel casings
should be used if the feedwater is unbuffered.

Regulation
Heater drain pumps should never be designed to operate on submergence control, because the required
NPSH is considerably higher than that needed for condensate pumps so cavitation is very unfavorable
to pump life. Capacity is generally controlled by a throttle valve in the discharge of the heater drain
pump, actuated by the heater hotwellievei. Unlike a condenser hotwell, the heater hotwell contains very
little storage. The control must be carefully arranged to prevent pumping the hotwell dry and steam
from binding the pump. Just as in the case of boiler feed pumps, some minimum flow must be made to
prevent an abnormal temperature rise. However, automatic bypass recirculation would not be economical,
and it is best to allow a continuous bypass equal to the minimum recommended flow.

Installation
Heater drain pumps handle water at or near saturation pressure and temperature. Like condensate
pumps, heater drain pumps should be adequately vented to the steam space of the heater. Because heater
Services 737

drain pumps and especially those on low-pressure service may operate with suction pressures below
atmospheric, it is necessary to ensure the pressure at the shaft seal is maintained above atmospheric to
prevent air leakage into the system; see condensate pumps.

CONDENSER CIRCULATING PUMPS

Four distinct pump configurations are now used for condenser circulating service:

1. Horizontal double-suction volute


2. Vertical dry-pit volute, both mixed flow and radial flow
3. Vertical wet-pit
4. Concrete volute.

To a great extent the choice of pump configuration is determined by the required total flow and head
and the source of the cooling water. There is, however, still latitude for economic evaluation and
engineering judgment in selecting the most appropriate configuration for a particular installation. That
may involve consideration of the number of pumps, suction conditions, performance characteristics, the
cost of installation, the likely frequency of repair, and the ease of carrying out repairs.

Pump Capacity
The total required cooling water flow is determined first by the size of the power plant, with fossil
fuel plants requiring up to 130 m3/hr (575 gpm) of cooling water per megawatt of output. Nuclear plants
using water-moderated reactors operate at a higher working fluid mass flow rate than the equivalent
fossil fuel plant, and so have a higher heat rejection rate. Nuclear plants therefore require about 35
percent more cooling water than an equivalent fossil fuel plant.
How many pumps are used to deliver the required total flow is a function of the plant's intended
operating range and its size. A plant designed for swing loading would typically divide the flow between
at least two pumps to allow pumping energy to be saved at low plant loads. When the range of ambient
temperature allows it, some plant designers have taken this a step further, and chosen three one-third
capacity pumps, with the intention of running two pumps during winter when the ambient temperature
is low, and three during the warmer summer months, an approach that lowers energy consumption during
the winter. Even in a base loaded plant, where a single pump would be acceptable, the physical size of
the pump configuration that best suits the other characteristics of the installation may dictate that the
flow be divided between two or more pumps.

Suction Conditions
In an ideal installation, the condenser circulating pumps are located below the minimum reservoir
level so they do not need priming. Just where this level should be relative to the pump depends on the
configuration and is discussed later for each configuration. Beyond avoiding the need for priming, the
pump must be far enough below the minimum reservoir level to (1) provide sufficient NPSH to allow
a sound design for the flow and head required (see Chap. 22) and (2) provide sufficient submergence
over the suction inlet or suction bell to prevent vortexing (see Chap. 28).
For all pump configurations except vertical wet-pit, the pump's performance is not unduly sensitive
to the design of the suction piping or inlet duct provided care is taken to ensure uniform flow to each
side of double-suction impellers (see Chap. 28) and a relatively uniform velocity distribution into the
eye of single suction impellers.
738 Services

Vertical wet-pit pumps, being high specific speed, are very sensitive to distortions of the flow into
the impeller. As a consequence the wet pit must be correctly designed to ensure the installed pumps
achieve rated performance. Pump spacing needs particular attention because the small floor space occupied
by each pump offers the temptation to reduce the size of the installation still further by placing the
pumps closer together. If this is done, the suction arrangement may not permit the proper flow of water
to each pump's inlet. A discussion of the pump arrangements recommended for wet-pit installations is
presented in Chapter 14.

Pump Head
Condenser circulating pumps are normally required to work against low to moderate heads. Plants
cooled by circulation from a body of water (river, lake, or the sea), have heads of 8 to 15 m (25 to 50
ft) depending on the elevation of the power plant and the variation in water level. Wet tower cooling
involves higher heads, from 15 to 28 m (50 to 90 ft). Care should be exercised in calculating the system
friction losses, which include the condenser friction. If more total head is specified than is actually
needed, the resulting driver size may be unnecessarily increased by an appreciable amount. For example,
an excess of only 0.3 or 0.6 m (l or 2 ft) in an installation requiring only 6 m (20 ft) of head represents
an increase of 5 to 10 percent in excess power costs.

Performance Characteristics
The range of specific speeds used for condenser circulating pumps, typically 2,500 to 10,000, produces
a great variation in the shape of the head-capacity and the power-capacity characteristics. For specific
speeds 2,500 and 10,000, respectively, the head rise from BEP to shutoff varies from 30 to 180 percent
(Fig. 18.35), while the power at shutoff varies from a 30 percent drop to a 110 percent rise (Fig. 18.36).
Although the shape of these curves can be varied by changing the impeller and casing waterway designs,
the variations which can be obtained without a significant loss of efficiency are relatively small.
Depending on a multitude of factors, such as the amount of static and friction head, and the relative
frequency of operating a single pump instead of two pumps in parallel, there may be some advantages
to a flatter or a steeper head-capacity curve. But the steeply rising power curve is always a disadvantage,
since it can introduce electrical supply problems by dictating a driver rating significantly higher than
needed at the pump's rated flow. From Figs. 18.38 and 18.39, it is evident that the efficiency of large
single- and double-suction pumps of specific speed 3,000 to 4,000 is up to 4 points higher than the bowl
efficiency of similar capacity wet-pit pumps. Allowing for column and elbow losses, this difference rises
by a further 1 to 2 points. The import of this is a trade-off between installed cost and operating cost,
which tends to favor operating cost as the required head, hence the power, increases.

Horizontal Double-Suction Volute Pumps


Horizontal, axially split, double-suction volute pumps (Fig. 26.15) have for many years been the
preferred construction when the required flow would allow them. Their specific speed ranges from 2,500
to 3,000, which gives them a rising head-capacity curve, a falling power-capacity curve and high
efficiency. A relatively stiff, low-speed rotor supported in oil-lubricated bearings produces high reliability,
while accessible bearings and seals and the axially split casing make for easy dismantling if repairs are
required. For plants cooled by circulation from a body of water, horizontal, double-suction pumps have
two disadvantages. First, they usually require some means of priming, unless local conditions permit
installing the pump in a dry pit so the lowest water level is at or above the high side of the impeller
eye. Second, for the low heads required they have to run at low speed, and so higher capital cost tends
to outweigh their lower power consumption. Given these limitations, horizontal, double-suction pumps
Services 739

Fig. 26.15 Installation of double-suction, horizontal circulating pumps.

CODE:
CD VERTICAL WET - PIT ® VERTICAL DRY - PIT
® CONCRETE VOLUTE (3) HORIZONTAL AXIALLY SPLIT
100 30 ...,.......----------,1..-----------------,
® .
M
OR
(3)
I ®
~ FT 20

_ ._._._1. _._._._._._ ..
w
::r::
..J 50
~
~ 10- CD
OR

®
20 40 60 80 100 120
1000 M3/HR

100 200 300 400 500


1000 USGPM
TOTAL REQUIRED FLOW

Fig. 26.16 Current practice for selecting type of condenser circulating pump.
740 Services

are now normally used only in wet-tower-cooled plants for total cooling water flows up to 45,000 m3/
hr (200,000 gpm); see Fig. 26.16). The upper limit of flow per pump is around 22,500 m3/hr (100,000
gpm), so multiple pumps in parallel are required for higher total flows.

Vertical Dry-Pit Volute Pumps


As the size of power plants increased in the 1950s and 1960s, many designers turned to medium- to
high-specific-speed vertical pumps, of either the dry-pit type, which operates surrounded by air, or the
wet-pit type, which is either fully or partially submerged in the water pumped. The choice between these
two types was somewhat controversial; very strong preferences were expressed in favor of one or the
other. The controversy resolved to inherently longer mean time between repair for the dry pit type versus
lower total installed cost for the wet pit type.
With refinements in the design of wet-pit pumps, which are discussed later, the use of vertical dry-
pit pumps is now limited to wet-tower-cooled plants with total cooling water flows up to 45,000 m 3/hr
(200,000 gpm). Two pump designs are used: radially split, single-suction mixed-flow (Fig. 14.13) or
axially split, double-suction volute (Fig. 14.6). The performance characteristics of the two designs are
similar, with the single-suction mixed-flow pump having a steeper head-capacity characteristic and a
flatter power-capacity characteristic. Both designs are considered equally reliable and will achieve
comparable mean times between repair, so the choice between them comes down to ease of maintenance
versus the cost of installation. Ease of maintenance favors the radially split design, whose rotor is
removed through the top of the casing, without having to disturb the driver when it is mounted at a
higher level (Fig. 26.17), the usual practice to avoid accidentally flooding the motor. Installation cost
usually favors the double-suction pump because its required NPSH is lower for a given capacity, and
the suction piping or passage configuration is simpler. The rotor can still be removed without disturbing
the driver, but special tooling is required (Fig. 13.15).

Vertical Wet-Pit Pumps


For condenser circulating service, the term "wet-pit pump" normally implies a diffuser pump. Vertical
wet-pit double suction volute pumps (Fig. 14.33) are used for condenser circulating service in smaller
plants, and are identified as such when used.
Vertical wet-pit pumps are high-specific-speed designs, typically 6,000 to 10,000, and therefore have
a very steep head-capacity characteristic and a rising power-capacity characteristic. The head-capacity
characteristic is often unstable, that is it has a dip at some capacity below BEP (Fig. 18.21), which is
not a concern provided the dip occurs at a flow below the pump's lowest operating flow, and the lowest
head is above the system head. As already noted in the general discussion of performance characteristics,
the rising power-capacity characteristic affects the sizing of the pump's driver and its electrical supply.
When provided with sufficient submergence, vertical wet-pit condenser circulating pumps can operate
at higher rotative speeds than conventional horizontal pumps or vertical dry-pit pumps, whether single
or double suction. This aspect has contributed to a cycle in their popUlarity. At first, it rose because the
pump and its driver were smaller and cost less. After some field experience with premature erosion of
impellers and suction bells, because the suction specific speed was too high (see Chap. 22), and structural
vibration problems, brought about by not appreciating that the combined stiffness of the pump and its
foundation determine the installed pump's natural frequencies, their popUlarity waned until it was evident
these problems were fully understood. This experience, coupled with improvements in the design of
water-lubricated bearings, has made wet-pit-type pumps the usual choice for plants cooled by circulating
from a body of water, where the lower head required means lower pumping power (Fig. 26.16).
Wet-pit pumps (Figs. 14.25 to 14.27) employ a column to support the bowl assembly. The column
is furnished with open line shaft bearings, lubricated by the pumped water, or, preferably, with an
Services 741

Fig. 26.17 Vertical dry-pit, mixed flow condenser circulating with extended shafting to
motor at higher elevation.

enclosed line shaft and bearings, lubricated by clean, fresh, filtered water from an external source. Even
in the latter case there is some danger of contamination for the lubricating water from seepage into the
shaft enclosure during shutdowns. A water lubricated bearing is never the equal of an oil lubricated
bearing even with the best design and care. Consequently, wet-pit pumps can be expected to incur higher
maintenance costs.
Larger power plants are generally located near centers of population and, as a result, often have to
use badly contaminated water-either fresh or salt-as a condenser cooling medium. With such water,
a fabricated steel column and elbow would give short life, and cast iron, bronze, or even a more corrosion
resistant material must be used. This results in a very heavy pump when large capacities are involved.
When the bowl assembly, line shaft or discharge head requires maintenance, the entire pump must be
742 Services

removed. This involves first disconnecting the discharge flange, then the use of a large mobile crane
(unless the pumps have an overhead crane for maintenance) to lift the pump assembly out of the pit.
To avoid the necessity of lifting the entire pump when the internal parts require maintenance, some
designs (Fig. 14.27) are built so that the impeller, diffuser, and shaft assembly can be removed from
the top without disturbing the column pipe assembly, although the driver must still be removed. These
designs are commonly designated as pull-out type.

Concrete Volute Pumps


In large power plants with wet-tower cooling, the large flows and higher head (Fig. 26.16) required
for condenser circulation no longer favor wet-pit pumps because of the size of the pump and the
significance of the pumping power. The design that is commonly used instead is the concrete volute
pump (Fig. 14.15). First built in France in 1917, the concrete volute pump is similar in configuration
to the vertical dry-pit mixflo pump, but with its casing formed in the foundation. Specific speeds are
typically 4,500 to 6,000, so the pumps' head-capacity and power-capacity characteristics are also similar
to those of the dry-pit mixed-flow pumps. Concrete volute pumps are also used in large stations cooled
by circulating seawater, to either avoid the cost of large casings in corrosion-resistant metal or to allow
the use of a single, unspared pump for the total required flow, of which there are now examples delivering
115,000 m3/hr (500,000 gpm).
As used for condenser circulating service, concrete volute pumps have the motor located on a separate
floor above the pump (Fig. 14.15). The drive is either direct, which requires a large, low-speed motor,
or through a reduction gear to allow the use of a smaller, higher-speed motor. Removing the pump's
rotor involves dismantling the lineshaft, then lifting the rotor, complete with guide bearing and head
cover, up through the casing.

Materials
The materials used for condenser circulating pumps range from gray iron through various alloys of
bronze to duplex (austenitic ferritic) stainless steels. Material selections depend on the type of pump
and the quality of the water it handles. Table 26.3 shows the materials commonly used for the major
components of condenser circulating pumps of all four configurations. Where more than one material
is listed, the final choice depends on the size of the pump and the actual site conditions.

Drivers
Horizontal pumps are driven by a conventional squirrel cage motor (see Chap. 24). For both dry-
and wet-pit vertical pumps, the axial thrust and weight of the pump rotor are normally carried by a
thrust bearing in the motor, and the driver and driven shafts are connected through a rigid coupling. The
higher rotative speeds of wet-pit pumps acts to lower the cost of electric motors. Against this, the
considerably higher axial thrust of wet-pit pumps raises the size and cost of the thrust bearing that must
be provided. Vertical dry-pit double-suction volute pumps and most European installations of both dry-
and wet-pit vertical pumps incorporate the thrust bearing in the pump. In these cases the motor has a
thrust bearing large enough to carry the weight of the motor's rotor, and the motor is flexibly coupled
to the pump. Most installations of vertical dry-pit pumps have the motor located above the potential
flood level, and are connected to the pump via a lineshaft (Fig. 26.17).

Condenser Circulating Arrangement for


High·Static Heads
Along midwestern rivers such as the Ohio, the Mississippi and the Missouri, where flood stages are
as much as 24.4 m (80 ft) above low water, it is customary to place the condensers in deep pits to
Services 743

Table 26.3 Typical Materials of Construction

Water quality
Component Freshwater (clean) Sea water (pure) Seawater (polluted)

Impeller and wearing rings Aluminum bronze Aluminum bronze, Duplex stainless
duplex stainless steel
steel
Shaft (protected with shaft sleeve) Carbon steel 13 chrome steel 13 chrome steel
Shaft (exposed)1 13 chrome steel Duplex stainless Monel
steel, Monel
Shaft enclosing tube l Carbon steel Ni-resist Aluminum bronze
Lineshaft bearings Leaded bronze, Leaded bronze, Leaded bronze or
cutless rubber cutless rubber cutless rubber
Casing and cover Gray iron, ductile Ni-resist Ni-resist
iron
Bowl and suction belP Gray iron Ni-resist Aluminum bronze
Column and discharge elbow l Carbon steel Epoxy-coated Ni-resist
carbon steel, Ni-
resist
IFor vertical wet-pit pumps.

maintain a normal pumping head and at the same time keep the turbine-generator above flood stage.
This is very expensive construction to which is added the additional heavy expense of a long and large
exhaust connection piece between turbine and condenser.
An arrangement of circulating water pumps has been devised for such an installation that eliminates
the costly condenser pit construction and permits locating the condenser directly beneath the turbine as
in the conventional type of plant with only a slight increase in the power required for pumping.
This method is shown in Fig. 26.18, which is a diagrammatic sketch of the circulating water circuit
of a steam power plant constructed on the Ohio River. The elevations, as shown, correspond to those
of the plant; the amount of water involved is 12,500 m3Jhr (55,000 gpm).
The arrangement, as shown, consists of two circulating pumps operating in series, each at full capacity.
The primary pump is a propeller-type, motor-driven pump; the secondary or booster mixflo pump is
mounted directly upon and driven by a hydraulic turbine.
Advantage is taken of maximum siphon effect by the discharge drop leg from the condenser. From
the sealing or control well all of the circulating water is passed through the hydraulic turbine. Thus, 75
to 80 percent of the static head from control well level to river level is available for useful work in the
hydraulic turbine-driven booster pump.
With its high-starting-head characteristic, the propeller pump is able to deliver enough water to the
condenser for starting purpose and bypasses the mixflo pump. When there is enough water in the control
well, the hydraulic turbine is started and the propeller pump discharges directly to the mixflo automatically,
and the two pumps operate in series at full capacity.
The booster pump functions until the river rises high enough to make the hydraulic turbine ineffective,
at which point the motor-driven propeller pump operates alone, giving full capacity under the reduced
head. The system described above is the most feasible for installations with considerable fluctuations in
water level.
Another application of this same idea would apply to a contemplated power plant where it would be
desirable, for reasons of lower construction cost, investment, accessibility, or geographical location, to
-..I
t

[,mu ..

~.-- , I "- ~

RIo'-' Ow.
$

£ KfflU!"E

Fig. 26.18 C~>odenser circulating water circujt.


Circuit utilizes hydraulic turbine-driven booster pump to recover static head beyond possible siphon effect.
Services 745

build at an elevation considerably above a source of circulating water supply, even though flood conditions
do not exist. In such an installation, the two pumps could be designed to operate in parallel, each pump
discharging separately against the total head. The total capacity requirements would be proportioned
between the two pumps, with all water being returned through the hydraulic turbine.
If flood hazards do not exist, horizontal pumps and drivers can be applied, providing other conditions
are suitable.

BOILER CIRCULATING PUMPS

The forced circulation or "controlled circulation" boiler, as it is frequently called, has certain advantages
over the conventional natural circulation boiler in the opinion of some power plant designers. The major
advantages they claim for controlled circulation are

1. Use of smaller diameter tubes


2. Freedom of tube circuit arrangement and distribution
3. Reduction in number and size of downtakes and risers
4. Lower over-all boiler weight and, consequently, lower structural support weight
5. Greater flexibility of operation.

A negative factor is that the forced circulation must be accomplished by pumps that are subject to
most severe conditions of service. The circulating pumps take their suction from a header connected to
several downcomers, which originate from the bottom of the boiler drum. They develop a pressure
equivalent to the friction losses through the various tube circuits operating in parallel (Fig. 26.19). Thus,

FEED WATER REGULATOR

BOILER FEED DISCHARGE


ECONOMIZER
STEAM

BOILER DRUM FORCED CIRCULATION


BOILER

BOILER CIRCULATING
PUMP

Fig. 26.19 Forced or controlled circulation cycle.


746 Services

in the case of a boiler operating at 131 bar (1,900 psig), the boiler circulating pump must handle
feedwater at approximately 332°C (630°F) under a suction pressure of 131 bar (1,900 psig). Such
a combination of high suction pressure and high water temperature at saturation impose very severe
conditions on the circulating pump shaft seals, making it necessary to develop special designs for
this part of the pump.
The net pressure to be developed by these pumps is relatively low, ranging from 2 to 7 bar (30 to
100 psi), at most, so they are all single-stage designs. Because of the difficult conditions imposed on
the shaft seals, it has always been preferable to use single suction, overhung pumps with a single shaft
seal. This construction has a disadvantage, however; as described in Chapter 5, the high suction pressure
develops a very high axial thrust, which can amount to several tonnes (tons) even with a hydraulically
unbalanced impeller (no back wearing ring or balancing holes).
A second difficulty arises from the fact that the forced circulation pumps have relatively limited
NPSH, whereas their capacity may be from four to eight times the maximum boiler feeding capacity.
Although the low design net pressures permit the use of operating speeds much lower than for boiler
feed pumps, the suction problem can become relatively difficult. The pumps can take advantage of the
fact that required NPSH for satisfactory operation is reduced at high temperatures (Chap. 19) even
though they are started before the water in the boiler drum has reached this high temperature and,
therefore, before this beneficial effect on required NPSH has taken place. Remedies for this situation
will be discussed later.
Two general types of construction are used for boiler circulating pumps: (1) conventional single-
suction, overhung pumps with various shaft seal and thrust bearing modifications and (2) submersible
motor pumps of either the wet- or dry-stator type (Chap. 24).
In the lower pressure range-up to 35 or 40 bar (500 or 600 psig)-a construction as shown in Fig.
26.20 may be used. The pump is of the same general type as is used on high-pressure heater drain
service. To lower the axial thrust as much as possible, the impeller is hydraulically unbalanced, which
exposes the shaft seal to close to pump discharge pressure. Older designs of these pumps were sealed

Fig. 26.20 End-suction, boiler circulating pump with mechanical seal for low-pressure range.
Services 747

with an injected breakdown bushing and packed box (see Chap. 8), with the injection coming from the
boiler feed pump discharge, a point of higher pressure and lower temperature. Part of the injection flow
went into the pump, while the rest passed through the breakdown bushing and was bled to a lower
pressure. In this way, the pressure and temperature at the shaft were brought within the capabilities of
a packed box. Axial face (mechanical) seals capable of withstanding the pressure within the pump are
now well proven, and so the sealing of low-pressure boiler circulating pumps has become less complicated.
The temperature at the seal is lowered by either injection from the boiler feed pump discharge through
a heat exchanger, or by self-circulation through a heat exchanger using a pumping ring built into the
seal. The latter is the simpler and more efficient arrangement when the nominal peripheral speed at the
seal is high enough to allow it.
More sophisticated designs are required for the higher pressures from 125 to 190 bar (1,800 to 2,800
psig; Fig. 26.21). The sealing of the shaft is accomplished by two floating-ring type breakdowns and a
mechanical seal. Boiler feedwater is injected at a point between the upper and lower stack of floating-
ring seals at a pressure approximately 3.5 bar (50 psi) above the pump internal pressure. Part of this
injection flows into the pump interior. The rest leaks past the upper stack of seals and is bled to a region
of low pressure in the cycle. Older designs used a jacketed packed box to control leakage to the
atmosphere. The modem practice is to use a mechanical seal, as shown in Fig. 26.21.
Manufacturers who build both the injection-sealed type and the submersible circulating pump say
that there are advantages to both types; the former is somewhat lower in first cost, but the latter eliminates
the injection and bleed-off system.
This injection requirement may pose a problem in certain installations. If the boiler feed pumps
operate at constant speed, the discharge pressure always exceeds the boiler drum pressure by a very
substantial margin, a margin that is throttled in the feedwater regulator. If the boiler feed pump operates
at variable speed, however, and the feedwater regulator is eliminated, there is no appreciable excess
pressure between the feed pump discharge and the boiler drum at low flows. Consequently, it becomes
necessary either to provide a small booster pump for the injection flow, or to install a valve in the feed
pump discharge. The valve creates sufficient artificial friction to provide the excess pressure necessary
to permit the injection flow to take place, even at low flows to the boiler.
As described previously, the available NPSH may not be sufficient at start-up, when the water in the
boiler is cold. Therefore, certain installations include two-speed motors, so that a lower NPSH is required
at start-up while operating at the lower of the two speeds. There is an added advantage to this arrangement.
Under normal operating conditions, the feedwater will be at boiler saturation temperature and, therefore,
will have a specific gravity of as low as 0.60; at start-up, however, the specific gravity will be 1.0. The
power consumption on cold water, therefore, would be 65 percent higher than in normal operation if
the pump operates at the same speed, and a much larger motor would be required. If a two-speed motor
is used, however, and the pump is operated at lower speed when the water is cold, the motor need be
no larger than the maximum horsepower under normal operating conditions.
The pump can be protected against low available NPSH by another means that is not only especially
effective against the possibility of fluctuation in boiler pressure and temperature, but also avoids possible
long-term damage from operating too close to cavitation range. This method consists of subcooling the
feedwater entering the boiler circulating pump and increasing the available NPSH by lowering the vapor
pressure of the water. This subcooling is accomplished by allowing all or a portion of the boiler feed
pump discharge to mix with the circulated flow ahead of the circulating pump suction (Fig. 26.22).
Figure 26.23 gives the temperature depression achieved by this mixing in terms of both the temperature
difference between the downcomer and the injected feedwater, and the ratio between the injected and
total flows. The effective increase in NPSH for a given temperature depression and at a given boiler
drum saturation temperature is plotted in Fig. 26.24. For example, if it is assumed that 30.5 m (100 ft)
748 Services

Fig. 26.21 Vertical, injection-type, boiler circulating pump for high pressures.

surplus NPSH is desired for an installation involving a 131 bar (1,900 psig) boiler (saturation
temperature = 332°C [630 0 P]), the required temperature depression is only 1.I°C (2°P). If injection
feedwater is available at 166°C (330 0 P), the difference between saturation and injection temperatures
is 166°C (300 0 P). The ratio of injected flow to the total circulation flow need be only 0.0067, or
less than 1 percent.
Services 749

T AM

STE AM SUPPLY
TO HEATER

~ORCEOCIRCUL AT ION
BOILER

HEATER

...::E
...:::>
.., PUMP
....
..,'o"
III
Z

z
o
<>
::IE
oa:
"'-

BOILER FEEO PUMP


\PROPORTIONING VALVE

Fig. 26.22 Arrangement of boiler circulating cycle with subcooling.


Feedwater is injected directly into circulating pump suction.
~~~~~~I/
~~V~~~L/
-v'-I~~ IOO
<!)
90 w
I-+-+--+-+--+-
Q:%
J;l 17 ¥
' .,0/ V
V
/ ' r""
V
LV
80
0
z
Z
1-----11'----'+-+-+-+V¥+- /-+IY o~Y
I -+- ./ V 70
0
en
V - ~ I- 0 ,*"'-1--+---+--::::01 60
vl/ V /V ~~ vV' 50
III
w
II::

/ 1/ /V V /V ~ ~v
11.
W
0
40
//VV / V /v,..-..... ~I9J9ost;:~ 30
w
II::

~ ~ ~ v V ~J.- ~ ~ t::-~r'Q.0!- 1:::-


:::>
20 ~
II::
W
10 11.
::IE
w
0 I--
o 50 100 150 200 250 300 350
BOILER TEMPERATURE MINUS TEMPERATURE OF INJECTION

Fig. 26.23 Temperature depression.


dT = r (TI - T2)
where
dT = temperature depression;
r = qlQ;
q = injected capacity, in pounds per hour;
Q = total flow, in pounds per hour;
TI = boiler temperature, in degrees Fahrenheit;
T2 = temperature of injection, in degrees Fahrenheit.
750 Services

2 20
.v 1/ / I
I J
If
/ /
'i / J / I /
20
~ / /
.
w
z
80
V~ I / I / L
/ ,;( I :; VI J
V
14
1i
II I L I;' I V )

'Ij I J ~~ V V
12_
V
..... .I :..~ /
-
. .A I V ~ / V :,V /
/ /
~
V

/ / V / / V V'

-... ~ ~V
/'
/
~

P'
,/'

--- -
./
/ V
". .."
~
... ~
~
I--"'
~

0
400 450 500 550 600
SATURATION TEMPERATURE IN BOILER IN oF.

Fig. 26.24 Surplus NPSH obtained by sub-cooling.


In terms of temperature depression and boiler saturation temperature.

FLUE GAS DESULPHURIZATION

As the consumption of electrical energy has increased, so has the volume of SOx (oxides of sulfur)
emitted to the atmosphere from the combustion of sulfur bearing fossil fuel. Atmospheric SOx combines
with water to produce acid rain, which has been identified as an environmental problem. Starting in the
late 1960s, various technologies have been developed to remove or "scrub" most of the SOx from power
plant flue gas. Now it is mandatory in many parts of the world to equip new and existing fossil fuel
fired power plants with flue gas desulfurization.
Of the various systems in use, the most common is single-loop wet limestone scrubbing (Fig. 26.25).
In this system a limestone slurry is pumped into the absorber tower where it is sprayed into the flue gas
stream. The SOx reacts with the limestone (calcium carbonate) to produce calcium sulfite and calcium
sulfate, which is collected as a slurry in the bottom of the absorber tower. In most modem systems, air
is injected into the slurry to fully oxidize the calcium sulfite to calcium sulfate (gypsum). Finally, the
products are drawn off as a slurry, dewatered, and either disposed of as limestone-gypsum sludge in a
landfill, or cleaned to produce building quality gypsum. A more complicated two-loop system, incorporat-
ing a separate quencher loop, is used when the fuel has a high chlorine content. This loop concentrates
Cl- which is then drawn off as hydrochloric acid.
Wet scrubbing uses two groups of pumps: 1) slurry pumps of various sizes for absorber recirculation,
quencher recirculation (when used), and general slurry service and 2) chemical pumps for clarified
Services 751

(f) ."" ••••


~::~ .... r.;::::::f----~~~::t. =-s...-:'.•
.~, I

Si8~k
,=--=...F3::===t:~: From ESP

From
Abs rber
Mo ule
1+------ Hydroclone
Oxidation Air
Compressor
To Dewatering
From
Limestone Absorber
Prep Bleed
Pump
From
Filtrate Return

Fig.26.25 Wet scrubbing, single loop system.


(Courtesy Babcock & Wilcox)

recycle and general water service. Of these, the most significant pumps are those for absorber and
quencher recirculation. As the size of wet scrubbers has increased so has the size of the recirculation
pumps, with capacities now extending to 15,000 m3/hr (65,000 gpm). Their basic configuration is a
horizontal, end suction, overhung pump, but the size and service requires special considerations in casing,
impeller, and bearing frame design. The service is severe in that the pumped liquid has an acidic pH,
ranging from 5.5 down to below 2 depending on the chlorine content of the fuel, a relatively high
chloride content, and contains abrasive solids. This has lead to two approaches to pump construction:
rubber (chloro-butyl or HypalonTM) lined casings with either rubber lined or duplex stainless steel
impellers in the United States (Fig. 26.26); duplex stainless steel casings and impellers in Europe. Shaft
seals are generally mechanical, designed for operation without flushing, to avoid the need for an auxiliary
service. Circulation around the seal for heat dissipation is allowed by an open seal housing, the bore of
which is often hard coated to prevent local erosion.

SPECIAL NUCLEAR POWER APPLICATIONS

With only a few exceptions, the nuclear power plants in service today have either pressurized water
reactors (PWRs) or boiling water reactors (BWRs). The critical distinction between these is the means
of transferring the reactor's heat to the power plant's working fluid. In a PWR, the reactor gives up its
heat to a primary circuit, which then heats the working fluid via a heat exchanger known as a steam
generator (Fig. 26.27). A BWR, on the other hand, has the working fluid pass through the reactor (Fig.
26.28), where it is heated and vaporized by direct contact with the reactor's fuel rods. The claimed
advantage of a BWR is easier control, which is rational because the thermal inertia of the system is
lower. Against this advantage, there is the distinct disadvantage that the working fluid is mildly radioactive
752 Services

Fig. 26.26 Absorber recirculation pump, rubber lined casing with duplex stainless steel impeller and wear plate.

and therefore all the equipment it passes through, much of which is outside the "nuclear island" must
be shielded.
In an oversimplified view, the nuclear energy steam power plant only differs from the conventional
power plant in the type of fuel it uses. Setting aside the means of reactor heat exchange, the "secondary
cycle" or "balance of plant" (consisting of turbogenerator, condenser, auxiliaries, and boiler feed pumps)
is not very different from its counterpart in the conventional steam power plant. To the extent that there
are differences, they are a desire for even greater equipment reliability and a preference for an absence
or minimum of leakage to avoid any possibility of radioactive contamination.
Although there is little difference in the arrangement of the "secondary cycle," the much lower steam
pressures of nuclear plants, 50 to 70 bar (725 to 1,000 psi), mean that the mass flow rate for a given
electrical output is significantly higher. The boiler feed pump (often termed the steam generator feed
pump for PWRs and the reactor feed pump in BWRs) is, in turn, designed for a higher flow and lower
head than those for an equal size fossil fuel steam power plant. Most designs are single stage double
suction pumps (Fig. 26.29), running at 7,000 and 5,000 rpm depending on the size of the plant.
In PWRs the heat output from the reactor is not varied over a wide range, so the temperature in the
"secondary" side of the steam generator rises as station load is reduced. The result of this is that the
steam generator's system head is essentially constant for all loads (even rising with decreasing load in
some older reactors). To ensure control, the feed pump head characteristic must be steep enough to
produce a well defined intersection at all capacities down to the minimum flow (Fig. 26.30). BWRs do
Services 753

~---,
I
3 I

I Reactor 8 Feedwater tank I Reactor 7 Condensate cleanup


2 Turbine 9 Booster pump 2 Turbine 8 Low-pressure
3 Generator 10 Feed pump 3 Generator feedwater heater
4 Condenser II High-pressure 4 Condenser 9 Reactor feed pump
5 Cooling water feedwater heater 5 Cooling water 10 High-pressure
pump 12 Steam generator pump feedwater heater
6 Condensate pump 13 Reactor recirculation 6 Condensate pump II Reaetor recirculation
7 Low-pressure pumps pump
feedwater heater 14 Pressurizing system
Fig. 26.27 Schematic of a pressurized water Fig. 26.28 Schematic of a boiling water
reactor (PWR). reactor (BWR).
(From Sulzer Centrifugal Pump Handbook, (From Sulzer Centrifugal Pump Handbook,
Elsevier Science Ltd, 1989, Oxford) Elsevier Science Ltd, 1989, Oxford)

not have this characteristic, so the system head for their reactor feed pumps is similar to that in conventional
steam power plants.
Most nuclear power plants operate with a closed cycle (Fig. 26.3), consequently the NPSH required
by the feed pump has to be provided by a booster pump downstream of the condensate pump, or the
condensate pump must develop the necessary head.
In addition to the pumps in the "secondary circuit," nuclear power plants have a number of auxiliary
pumps, many of them classified as "safety related." Safety-related pumps are all those associated with
dissipating heat from the reactor core or spent fuel storage, under either normal or emergency operation.
The usual services are summarized in Table 26.4. The design, installation, and operation of safety related
pumps must conform to rules issued by the government in the country of installation, for example the
Nuclear Regulatory Commission (NRC) in the United States. These rules in tum cite design standards
such as Section III of the ASME Boiler and Pressure Vessel Code, Rules for the Construction of Nuclear
Power Plant Components [26.1]. Typically the design involves meeting certain proven construction
requirements, and submitting detailed structural analyses, including the effect of earthquakes, for the
pressure boundary parts. To simplify structural analysis and allow volumetric examination, the pressure
754 Services

Fig. 26.29 Turbine-driven single stage double suction steam generator feed pump.

HEAD - CAPACITY
@ CONSTANT SPEED
H
~------~=-----~--------'
~ l PUMP
_ HEADRISE

FRICTION
LOSSES
SYSTEM HEAD CURVE
STEAM GENERATOR
PRESSURE

Fig. 26.30 System head curve for pressurized water reactor (PWR) steam generator feed pump.
Services 755

Table 26.4 Safety Related Pumps

Function Pump Service

Circulating, spraying, or injecting liquid in primary circuit Reactor coolant circulation


Charging (PWR)
Residual heat removal
Safety injection (PWR)
Low pressure core spray (BWR)
High pressure core spray (BWR)
Reactor core isolation cooling (BWR)
Standby liquid control (BWR)
Service water
Containment spray (PWR)
Circulating coolant through the closed circuit cooling system, and Component cooling
through the heat exchanger to the ultimate heat sink. Auxiliary feedwater (PWR)
Circulating liquid through the spent fuel storage pool. Spent fuel pit cooling
Pumping fuel for emergency power generation. Diesel fuel transfer

boundary parts are generally designed as simple, symmetric shapes. Once installed, any emergency
pumps are subject to periodic operational testing and visual inspection to verify that they are functional.
The reactor coolant pump that circulate liquid through the reactor (and then through the steam generator
in a PWR), are large, critical pieces of equipment. Por PWRs the coolant pumps are designed for high
flow at low head, the typical operating pressure is 170 bar (2,500 psig) at 350°C (650 0 P), and the power
is about 5,500 kW (7,000 hp). In the usual design of BWR, the coolant recirculation pumps actually
provide driving liquid to jet pumps located inside the reactor, so the flow is lower and the head higher.
The operating pressure is lower, typically 120 bar (1,700 psig) at 300°C (575°P), while the power per
pump is about the same. In principle the design problems with reactor coolant pumps are similar to
those of medium pressure boiler circulating pumps, but are compounded for reactor coolant pumps
because the reliability of pressure containment is critical and leakage to the containment from the shaft
seal cannot be tolerated. The size of most reactor coolant pumps dictates a conventional motor, and the
normal configuration is a vertical, rigidly-coupled pump (Pig. 26.31), with two or three axial face seals
in series, each taking a fraction of the total pressure drop. The small flow that is necessary to achieve
staged sealing is bled back to low pressure point in the system. Some European designed BWRs have
been equipped instead with a higher number of smaller sealless pumps similar in construction to that
shown in Pig. 24.45. The liquid end of these pumps is inserted into the bottom of the reactor vessel and
there are no jet pumps.
Por liquid-metal-cooled reactors, the ~esign of the coolant circulation pumps is made more difficult
by the pumped liquid being either liquid sodium or sodium potassium at 580°C (1,076°P). The pumps
used for this service are typically vertical, separately coupled, with the casing and impeller located some
distance below the pump's mounting plate (Pig. 26.32). The pump's elevation relative to the reactor is
such that there is a free surface of liquid metal between the impeller and the shaft seal. The liquid end
of these pumps is inserted into the bottom of the reactor vessel and there are no jet pumps. An inert
gas blanket is maintained on the free surface to prevent any reaction with the sodium, and so the shaft
seal has only to retain the gas whose pressure is normally 10 bar (140 psig). The lower end of the pump
shaft is guided by a liquid metal lubricated hydrostatic bearing located next to the impeller.
756 Services

~~~~~~ii---Motor Half Coupling

• Primary Loep Waler


@ Temper.lur• • Spacer
Cooled Prlmlry loop Wiler
Cempgnenl COOIIIIII Waler

M~C!::J~zt.---·PL,mp Hall Coupling

111-'_----Shaft

_ • •1-. Co",", WI,", 0uII0I

.1'11~...._'ntegral Heat
Exchanger

Mechanical
I:::::::=I--- Seals

mpeller

PRIMARY NUCLEAR PUMP


802().1

Fig. 26.31 Closed-coupled reactor coolant circulating pump.


(Courtesy BWI/P International)
Services 757

Fig. 26.32 Coolant circulating pump for liquid-metal cooled reactor.


(Courtesy KSB)

Water Works
The supply of water to industry and residential users is one of the major fields of application for
centrifugal pumps. Both ground water from shallow or deep wells and surface water from rivers, lakes,
or artificial reservoirs are used as a source of supply. Except for some well water, which may be
sufficiently clear to require little treatment, most raw water must be processed to remove suspended
758 Services

matter and bacteria. Thus, raw water is transported by centrifugal pumps from its source to a purification
plant. The amount of treatment required for the raw water will depend on the use to which it will be
put and the character of the impurities it contains. Raw water may be subject to any or all of the following
treatments: Coagulation, sedimentation, filtration, activation, chlorination, fluoridation, and softening.
Because pumps that transfer raw water from a surface supply to a purification plant usually operate
against a low discharge pressure, they are known as "low-service" pumps.
Pumps that discharge into distribution systems after purification must generate higher heads and are
known as "high-service" pumps. The distribution system consists of large-size transmission pipes, which
feed into a network of medium- and small-size piping.
The following are the types of centrifugal pumps most frequently used in the water works field:

1. Horizontal-shaft, end-suction, single-stage, volute pump, either foot mounted back-pull-out (Fig. 2.14) or
motor-mounted (Fig. 1.19).
2. Horizontal-shaft, double-suction, volute, side-discharge, side- (Fig. 2.15) or bottom-suction (Fig. 2.26) pump,
with casing axially split. Occasionally, a bottom-suction, bottom-discharge pump is used.
3. Horizontal-shaft, single- or double-suction, multistage (usually two, or possibly three, stages) volute pump.
4. Series units, consisting of two, or occasionally three, horizontal-shaft, double-suction, volute pumps, with
casings axially split, connected in series with separate or common driver (Fig. 2.27 and 2.28).
5. Vertical-shaft, single-suction, single-stage, dry-pit, volute pump (Fig. 2.9 and 14.13).
6. Vertical-shaft, wet-pit, single-stage or multistage turbine (Fig. 14.17 and 14.19) or propeller pump (Fig.
14.25 and 14.26).

The horizontal, end-suction, single-stage, volute, frame- or motor-mounted pump is an ideal design
for small to medium capacities. This pump has not found universal acceptance in the water works field,
possibly because many pumps of this type have been light-duty lines made for intermittent service and,
therefore, not suitable for water works service. Currently, most manufacturers offer substantial designs
as well as light-duty lines in this construction. These frame- or motor-mounted substantial designs are
volume built, usually with impeller designs that load up standard motor sizes. They are standardized
production line units and, in most cases, are built in lots for stock. As a result, the manufacturers cannot
tailor the impellers for specific service conditions, nor can they afford to make special tests to demonstrate
the pump performance. Nevertheless, this type of pump has considerable application for small-capacity
water works service and can be obtained at reasonable cost, provided it is purchased without voluminous
specifications or special guarantee requirements.
In general, the horizontal-shaft, double-suction, volute pump is the most commonly used type for
medium capacities up to 75k m3jday (20 mgd) and for heads up to 92 m (300 ft) or more. At this capacity
the vertical-shaft, single-suction, single-stage pump sometimes makes a more economical installation,
when all factors are taken into account. In the 75k to 150k m3/hr (20 to 40 mgd) capacity range, local
conditions may favor either type, but over 150k m3jday (40 mgd) the probability is that the vertical
pump would be the most economical.
The vertical turbine pump is universally used for pumping from deep wells. There has been some
tendency to apply this type to shallow wet pits in place of the conventional horizontal-shaft or vertical
dry-pit types. Although the vertical turbine pump installation has a lower first cost in many cases, the
horizontal-shaft or vertical-shaft, dry-pit pump has a much lower maintenance cost and, therefore,
is favored.
Unless desalination is involved, most of the energy used in supplying water is that consumed by the
pumps. Given this circumstance, the evaluation of pumps for water works applications generally gives
considerable weight to energy consumption, frequently to the extent that the capital cost of the pumps
becomes relatively insignificant. Because the energy the owner pays for is directly related to the overall
Services 759

kilowatt-hour absorbed by a pump's motor, the evaluation of energy consumption needs to take account
of the load profile the pump is expected to operate against. Not doing this and instead evaluating efficiency
at rated capacity runs the risk of selecting a pump that actually consumes more energy per year than
one that has higher peak efficiency (see Chap. 21).

Sewage
For centuries, inland rivers and coastal waters afforded a convenient disposal of various untreated wastes,
such as domestic sewage and storm water drainage. As technological civilization progressed, industrial
process wastes of many kinds, drainage of contaminated mine waters, and liquids such as brine from
oil fields were added. The resulting pollution made these river waters unfit for consumption and some
coastal waters extremely unpleasant to the nearby population. For a number of years, therefore, local
and national government bodies have exerted greater and greater efforts to pass legislation that will
force communities and industries to treat wastes before ultimate disposal.
Today, municipal sewage treatment prevents excessive pollution of water caused by the discharge of
raw domestic sewage and liquid wastes from cities and towns. In the same manner, individual industries
are treating their wastes, either voluntarily or under the pressure of legislation.
The degree of sewage treatment in a particular installation depends primarily on the amount of
pollution to be disposed of and the size of the stream flow into which wastes are discharged. Partial
treatment may involve merely screening of the wastes and chlorination, followed by some sedimentation.
Intermediate treatment may include the addition of chemical precipitation, as well as some filtration.
Finally, complete sewage treatment may also include multistage high-rate filtration, trickling filtration,
and an activated sludge process. For a complete discussion of various types of sewage treatment the
reader is referred to the numerous textbooks available on the subject.
All of these processes require the use of pumping equipment to move liquids-either raw or treated
sewage-from one part of the process to another, and finally to the ultimate disposal of the effluent.
Pumps suitable for pumping sewage must be capable of passing the solids contained in the sewage.
Raggy and stringy solid material gives the most trouble and, except for very large pumps with stationary
covers over the shaft, makes it undesirable to use pumps in which the shaft projects through the suction
waterways. For the same reason, diffuser vanes with thin edges that will catch rags and stringy material
are also undesirable. Thus, almost all sewage pumps are of the end-suction, volute type (Fig. 14.2) and,
in general, the impellers are of a special non-clogging design (Fig. 4.12). Most applications of sewage
pumps are for low heads up to 21 to 24 m (70 to 80 ft) maximum in the smaller capacity range and up
to 12 to 15 m (40 to 50 ft) maximum for large capacities, so that single-stage pumps can always be used.
The standard water closet bore is 64 mm (2.5 in) in diameter and non-clogging impellers are specified
to pass objects up to at least 64 mm (2.5 in) in diameter. If screens with finer mesh than 64 mm (2.5
in) or comminutors that macerate and cut up foreign matter in the raw sewage are used ahead of the
pumps, some specifications reduce the requirement to 50 mm (2 in) or even 38 mm (1 Y2 in), to take
advantage of the resulting increase in pump efficiency.
Vertical sewage pumps are preferred because they can be located so that priming is unnecessary (Fig.
26.33). The centerline of the impeller is located at an elevation below the suction supply. Wet-pit,
submersible pumps (Fig. 14.29), up to the limit of their capacity, are the preferred construction in current
design practice. Beyond that, dry-pit construction offers greater reliability though at higher capital cost.
The same general type of end-suction pump with special non-clogging impellers is used for secondary
and activated sludge pumping in the sewage treatment pumps as well as for sludge recirculation. For
recirculation or transfer of settled sewage that contains neither large solids nor stringy material, both
760 Services

Fig. 26.33 Vertical, bottom-suction sewage pump with motor supported directly on pump.

wet- and dry-pit vertical pumps are used, with a trend to axial-flow, wet-pit propeller pumps when the
total heads are very low.
Both wet-pit, axial- or mixed-flow propeller pumps and dry-pit, vertical or horizontal volute pumps
are used for storm water pumpage, whichever type best suits the individual problem.
A trend may exist toward the elimination of multiple sewage pumps or stations in favor of single
Services 761

pumps of larger capacity, because maintenance is costly with many small installations spread over a
large area. In contrast to this is a growing interest in small package-type sewage disposal plants with
capacities as low as 6 to 24 m3jhr (25 to 100 gpm) for small public installations such as schools.
The majority of sewage pumps are electric motor driven, although some gasoline engine or diesel
engine installations may be encountered. Where wide variations in pumping capacities and total heads
are encountered, variable speed operation is becoming more and more popular. The variable speed is
obtained through the use of magnetic or variable-frequency drives, and the operation is generally made
fully automatic, responsive to level control.

Drainage and Irrigation


Drainage pumping is employed either to recover low-lying lands otherwise suitable for farming, or to
maintain areas subject to infiltration flooding in relatively dry condition. In many cases this latter function
is essentially similar to flood control of areas lying below water level that are protected by levees.
Capacities of drainage and flood control pumps are generally high, although the total head is relatively
low-frequently as low as 1.5 to 3.0 m (5 to 10 ft). Vertical and horizontal, mixed-flow and axial-flow
pumps are used on this service. When higher heads, up to 15 m (50 ft), are encountered, mixed-flow
pumps only are used.
Drainage pumps are subject to the same problem as sewage pumps; the water may contain a consider-
able amount of foreign matter, including small branches, vegetable matter, trash, abrasive sand, and
mud. Although a bar screen ahead of the pump suction or coarse mesh strainers on the bellmouth of
the suction piping can eliminate some of these solids, it is still necessary to use the same type of non-
clogging impellers as for sewage pumps. When large-capacity pumps are involved, however, the water-
ways between the vanes of a normal impeller will pass sufficiently large solids. Similarly, vertical-
turbine pumps or propeller pumps with diffusers are undesirable because the debris carried in the drainage
water could catch on the diffuser vane edges.
Ditch-type or flooded-type irrigation presents the same problems as drainage, namely, heads are low
and the water handled may be very dirty and even contain abrasive sand or mud. Capacities of pumps
on irrigation service, however, are limited to a lower range than on drainage service. Sprinkler system
irrigation, on the other hand, demands a practically clean water supply. Therefore ordinary general-
service centrifugal or vertical-turbine pumps are used, depending on the source of supply and the hydraulic
requirements. Pumps on this service may require total heads from 30 to 60 m (100 to 200 ft) and higher.
In some cases, irrigation pumps are installed in permanent locations whereas in others pumps are
portable or semiportable. Permanent pumps can be driven by electric motors (Fig. 26.34) or diesel or
gasoline engines. Portable pumps are mostly gasoline or diesel engine driven and are mounted on skids,
trailers, or trucks.
Except for wet-pit, submerged-type pumps, irrigation pumps must be provided with some form of
priming device (Chap. 25). Figure 26.35 shows a diesel-engine-driven irrigation pump mounted on a
portable skid; a jet primer, which uses the engine exhaust, is employed to prime the pump. A clutch
between the pump and the engine permits the engine to idle with the pump disconnected while it is
being primed.
Unless there is a considerable amount of sand or grit in the water, the manufacturer's standard
materials for ordinary cold water service are used. Special materials for parts subject to the erosive
action of sand may be justified for the larger pumps; this may include nickel cast-iron casings, chrome-
steel wearing rings, chrome-steel shaft sleeves, and sometimes chrome-steel impellers.
762 Services

Fig. 26.34 Horizontal 42 in propeller type pump on drainage service.


Electric-motor driven through a Multi-V drive.

Fire Pumps
Even though the centrifugal fire pump is purchased with the hope that it will never be operated, it is a
very sound and profitable investment. Not only is it important to provide a fire protection system, but
the reduction in fire insurance premiums will generally amortize the initial cost of the installation in a
few years.
Centrifugal fire pumps are classified in two groups from the point of view of function. If the installation
to be protected is located where municipal fire protection is not available, fire pumps are installed to
provide this protection by themselves; in other cases, inadequate water pressure supply is augmented
by booster fire pumps. The latter is probably the more frequent case, and most large buildings in urban
areas include such booster fire pumps.
Because of the interest insurance companies have in the satisfactory protection obtained from fire
pumps, the size and type of this equipment, as well as many details of its installation, are subject to the
approval of several governing bodies connected with insurance companies. These pumps, therefore, must
be built in accordance with the requirements of the Underwriters Laboratories, Inc., the National Board
of Fire Underwriters, or the Associated Factory Mutual Fire Insurance Companies, Inc. In addition, an
installation may have to be locally inspected and approved. These underwriters' organizations issue
standards on approved equipment; they should be consulted for specific details not given here. In general,
these standards specify both the hydraulic characteristics that must be met before acceptance and the
auxiliary equipment that must be furnished with the installation.
Services 763

Fig. 26.35 Diesel-engine-driven, mixed flow drainage pump with exhaust jet-primer mounted on
portable skid-type base.

Fire pumps are classified in the standards as low-pressure service (2.8 to 6.9 bar [40 to 100 psig))
and high-pressure service (6.9 bar [100 psig] and above). Both horizontal and vertical pumps are
acceptable for fire service. Horizontal pumps are limited to a total suction lift of 4.6 m (15 ft) when
delivering 150 percent of rated capacity. The standards specify a permissible range of heads at shut-off
and at 150 percent of rated capacity and provide standard capacity ratings. Approved ratings range from
114 to 568 m3/hr (500 to 2,500 gpm). Special ratings of 45 to 114 m3/hr (200 to 500 gpm) are also
listed for limited-service fire protection. These are subject to special consideration in each individual
case. Vertical-shaft, turbine pumps are listed in the underwriters' standards for capacities between 114
and 568 m3Jhr (500 and 2,500 gpm).
Where the power supply is adequate and reliable, electric-motor drive is preferable for fire-pump
service (Fig. 26.36). The electrical equipment must comply with the provisions of the National Electrical
Code, except as modified by the standards described above. Steam turbines and internal combustion
engines may be used to drive fire pumps where other sources of power are unreliable or not available
(Fig. 26.37).
Any number of the following accessories may have to be supplied with a fire pump installation,
depending on the particular local requirements:
764 Services

1. Automatic air release valve


2. Circulation relief valve (for horizontal pumps)
3. Eccentric tapered reducer at suction nozzle (for horizontal pumps)
4. Hose manifold with hose valves
5. Suction and discharge pressure gages
6. Priming connections (for horizontal pumps operating with suction lift)
7. Relief valve and discharge cone
8. Splash shield between pump and motor (for horizontal pumps driven by open motors)
9. Test valves with piping connections
10. Umbrella cocks (for horizontal pumps)
11. Suction strainer (for vertical submerged pumps).

Chemical Industry
The chemical industry presents the widest variety of pumping problems and different liquids to pump
of any industry today. Raw materials in liquid fonn are generally delivered to a chemical plant in tank
cars and must be transferred into reservoirs. They are later transferred to the plant and pumped from
one part of the process to the next. The semifinished and finished products undergo many changes in
chemical composition, consistency, temperature, viscosity, and corrosiveness.
Centrifugal pumps are used in approximately 90 percent of all the applications that involve the
handling of corrosive liquids in chemical plants. The most compelling reason for this preference lies in
the much lower cost of centrifugal pumps made of special alloys.
The most important factor in the selection of pumps for chemical service is the selection of materials
to withstand the corrosive or corrosive-abrasive liquids that have to be pumped. As described in Chapter
17, centrifugal pumps have been made of every imaginable material, including graphite, rubber, plastics,
glass, and ceramics. The Standards of the Hydraulic Institute recommend a very extensive listing of
materials and their suitability for pumps handling a variety of chemically active liquids. Although these
recommendations are followed exactly, as little as I percent variation in the constituents of an alloy
may alter the rate of corrosion with a given liquid by as much as several hundred percent. Variations
in foundry techniques may also have the same effect without any change in the alloy composition.
Finally, the unexpected presence of even minute amounts of alien components in the liquid pumped, a
slight increase in temperature, or the presence of dissolved gases may seriously alter the ultimate life
of a given material. In many cases, therefore, the operators who are using the equipment are the only
ones who have a finn basis for the proper selection of the material to be used.
Most manufacturers specializing in centrifugal pumps for chemical service maintain extensive corro-
sion data files which are the result of many years of field experience and corrosion testing. As a result,
it is generally sound policy to give the manufacturer as much infonnation on the liquid as possible even
when a particular material is being specified.
The extreme severity of chemical-service applications has resulted in the acceptance of very different
criteria of satisfactory life as compared to most centrifugal pump uses. A life of six months to two years
for an impeller, for example, may be quite acceptable in many cases, although this would not be tolerated
on boiler feed service.
Because so many applications require pumps made of special alloys or special materials, designers
tend to use the simplest possible configurations and construction. The complexity and cost of an effective
shaft seal reinforces this tendency, with the result that most chemical pumps are single stage, overhung
Services 765

Fig. 26.36 Motor-driven centrifugal fire pump with underwriters' approved fire fittings.

Fig. 26.37 . Diesel-engine-driven fire pump undergoing shop test.


766 Services

impeller, a configuration that has only one shaft seal. Impellers are usually open or semiopen (Fig. 2.22).
Horizontal pumps are end suction, foot mounted, with their casings designed for "back-pull-out" of the
bearing frame complete with casing head, shaft seal and impeller (Fig. 2.14). Since the mid-1960s, the
minimum performance and overall dimensions of end suction chemical pumps has been standardized,
starting with the American Voluntary Standard (AVS) which has become ASME B73.1 [9.3] for horizontal
chemical pumps. The European standard for similar pumps is ISO 2858 [26.2]. Reduced floor space and
higher tolerance of piping loads has lead to the widespread use of vertical in-line pumps for chemical
service. All three of the usual configurations of this design are used: (1) close coupled with the impeller
mounted directly on the motor shaft (Fig. 14.5), (2) rigidly coupled which has a rigid coupling between
the motor and pump shaft to allow seal replacement without removing the motor (Fig. 2.16), and (3)
separately coupled, an arrangement in which the pump's rotor is supported by a separate bearing frame
and the motor is flexibly coupled to the pump (Fig. 14.8). For chemical service, vertical in-line pumps
are covered in the United States by ASME B73.2 [26.3].
Some of the liquids pumped in the chemical industry are either very difficult to seal or cannot be
allowed into the atmosphere under any circumstance. Hermetically sealed or so called "sealless" pumps
are used for these applications. Two configurations are employed: magnetic drive (Fig. 24.57) and canned
motor (Fig. 24.48). Magnetic drives are simple, have tolerable efficiency up to around 75 kW (100 hp)
at 3,600 rpm, but require a dry-running axial face seal to provide secondary containment. Canned motors
achieve higher efficiency than magnetic drives, and with the appropriate design of the stator lead seals
provide high-integrity secondary containment without moving parts. With only two exceptions, the
bearings that support the pump's rotor in either design are product lubricated, so bearing life is sensitive
to solids in the liquid and to flashing of the liquid. The exceptions are magnetic bearings for canned
motor pumps (see Chap. 11), or inert gas filled drives or motors, an innovation now being promoted by
some manufacturers.
For capacities beyond the capability of single suction, overhung pumps, the single-stage, double-
suction pump is used. Most applications are for general service, cooling water circulation, or the handling
of liquids that have the same corrosion and abrasion characteristics as water. With increasing plant sizes
and their inherently higher capacities, there are instances, however, where it is necessary to use single-
stage, double-suction pumps for corrosive service. They are more expensive than an equivalent single
suction pump because (1) the casing is harder to cast in corrosion resistant alloys (long, thin sections)
and uses relatively more metal and (2) most designs have "between bearings" rotors, which require two
shaft seals.
This discussion will be limited to general remarks on certain specific features of centrifugal pumps
intended for chemical service that can assist the user in evaluating any particular design. Many special
design refinements can be furnished that would prove to be an asset to a plant with well-trained mechanics,
but these same refinements may prove to be a disadvantage if certain adjustable or complicated features
are improperly used. As a general rule, therefore, the simplest construction will prove to be best.
Sealing the pumped liquid has been and remains a difficult aspect of the design and operation of
chemical pumps. Although significant advances have been made in fluid sealing since this book was
first published, growing concern for the environment has led to lower allowable leakage from shaft seals,
which has continued to tax the technology. At the same time, process conditions have in many cases
become more severe, thereby further taxing the technology. The net result is that chemical pumps are
now either equipped with some form of axial face (mechanical) seal or are hermetically sealed (see the
preceeding discussion and Chap. 24).
Four factors account for the fact that the axial face (mechanical) seal has all but replaced the packed
box for pumps handling chemical process liquids: (1) negligible leakage rates; (2) the ability, with
multiple seals, to isolate the pumped liquid from the atmosphere; (3) less need for auxiliary services;
and (4) no need for periodic adjustment. This is not to say, however, that axial face seals are a panacea;
Services 767

Fig. 26.38 Fiber-reinforced polymer chemical pump for chemical service.

they still have to be properly selected, installed and operated if they're to function as intended. For a
detailed discussion of the operation of axial face seals and the control of their operating environment
see Chapter 9.
For services where the pumped liquid contains abrasive solids but is not deemed environmentally
hazardous, the hydrodynamic or "expeller" seal (Fig. 8.26) offers a viable alternative to a mechanical
seal. Chapter 8 describes the operation and limitations of this seal design.
Most chemical pumps have all-metal liquid ends. For some applications, however, it is necessary or
beneficial to use alternative constructions. Polymers of the appropriate compound offer excellent resistance
to a wide range of the liquids used in chemical processing, often surpassing the performance of metals.
With due attention to strength and stiffness, including the use of fiber reinforcement when necessary,
polymer can be used for the liquid end without the need for metal constraints (Fig. 26.38). Casings are
typically rated for pressures up to 15 bar at 120°C (225 psig at 250°F) and are able to withstand normal
piping loads.
In highly corrosive services at pressures or temperatures above the capability of polymers, it is often
preferable to use a pump with a metal liquid end that has been lined with a chemically inert metal, such
as polytetraftuoroethylene (PTFE) (Fig. 26.39). The lining is preformed and attached by bonding.
Figures 26.40 and 26.41 show the sectional and external views of an elbow-type propeller pump
frequently used in evaporator transfer service or as an external circulator pump. Agitators use a modifica-
tion of this design, eliminating the casing. The pump manufacturer furnishes the shaft, the propeller,
and a stuffing box; the user mounts this assembly in a tank where mixing and agitation is required,
locating the propeller inside a draft tube.
Rubber lined pumps are widely used for chemical service because they offer excellent resistance to
768 Services

Fig. 26.39 PTFE lined chemical pump.


(Courtesy The Duriron Company, Inc.)

Fig. 26.40 Section of elbow-type propeller pump.


(Courtesy Lawrence Pumps, Inc.)
Services 769

Fig. 26.41 Elbow-type propeller pump for corrosive circulating service.


(Courtesy Lawrence Pumps, Inc.)

abrasive corrosion and erosion. The pump shown in Fig. 26.42 has a casing that is radially split at the
discharge centerline, so the two halves of the rubber casing liner can be installed. Common lining
materials are natural rubber, neoprene, and polyurethane. The impeller is lined with the same material.
An alternative casing design, with an asymmetric discharge that puts the casing liner joint behind the
impeller discharge, is shown in Fig. 26.43.

Petroleum Industry
The pumping requirements of the petroleum industry may be divided into three general categories,
as follows:

I. Production includes well pumping, gathering, and waterflooding, as well as various auxiliary services such
as fire pumps, and contractors' pumps.
2. Transportation includes transportation of the crudes, the refined oils, and gasoline. The transportation may
be by pipeline, loading to tank cars or tankers, and cargo unloading.
3. Refining includes pumping for all processes and auxiliary plant services such as cooling.
770 Services

Fig. 26.42 Section of rubber lined pump with two-piece casing liner and closed impeller.

Fig. 26.43 Section of rubber lined pump with asymmetric casing liner and semi-open impeller.
(Kestner Model KRL; Courtesy Kestner Eng. Co. Ltd [UK])
Services 771

PRODUCTION

The only type of centrifugal pump used in oil-well pumping is the submersible pump described in
Chapter 24.
Waterflooding, or tertiary recovery, is the injection of water into the structure surrounding wells that
have stopped flowing as primary production wells. The repressuring of exhausted areas permits recovery
of large quantities of oil, otherwise out of reach. In the past, most waterflood installations only required
relatively low flows and so reciprocating pumps were commonly used. As the flows required have risen,
high-speed multistage centrifugal pumps operating at discharge pressures as high as 380 bar (5,500 psig)
have become commonplace. The pumps are axially split or radially split double casing, depending on
the working pressure and the location of the pump. Many waterflood installations are on offshore
platforms, where the cost of floor space often dictates the use of axially split casings to unusually high
pressures. Because of this, there are special design axially split casings in service at pressures as high
as 275 bar (4,000 psig). Most waterflood pumps are driven by natural gas fired engines through a step-
up gear or directly by gas turbines. Figure 26.44 shows a large 17,000 kW (23,000 hp), gas-turbine-
driven water flood pump.
Saltwater, either produced or from the sea, is the most widely used medium for waterflooding, any
produced freshwater being conserved for irrigation or other purposes. Most oil field saltwater is brackish
and corrosive to bronzes and copper alloys. When produced salt water is deaerated and kept oxygen-
free by a gas blanket in any intermediate storage tanks, pumps of carbon steel with chrome steel internals
are satisfactory. Aerated saltwater is extremely corrosive and requires pumps of duplex stainless steel

Fig.26.44 Gas-turbine-driven 17,000 kW (23,000 hp) water injection pump.


772 Services

to achieve reasonable service lives. Erosion caused by fine silica sand in the water can compound the
corrosion problem, and in extreme cases has forced the use of solid ceramic (tungsten carbide) wearing
parts to achieve a tolerable period between the need to renew running clearances. High-speed centrifugal
waterftood pumps are also prone to premature cavitation erosion of the first stage impeller, even when
the available NPSH is 2 to 3 times that required at 3 percent head drop, because they are handling cold
water at very high values of NPSH. To avoid operating problems, it is important that the available NPSH
be sufficient, and that the impeller design has been flow visualization tested to verify that when operating
with the available NPSH, there is no potentially erosive local cavitation on the inlet vanes (see Chap. 19).
Oil is usually produced as a mixture of oil, water, and natural gas. These used to be separated before
being pumped to their respective storage vessels. The desire today is to pump the mixture as produced
from various wells to a central separation and processing plant. This involves two-phase pumping, with
the gas volume at suction conditions sometimes as high as 90 percent. To date twin-screw-type rotary
pumps have been used almost exclusively for this service, but there are now special centrifugal pump
designs being advanced.

TRANSPORTATION

Since the middle 1920s, centrifugal pumps have almost completely replaced reciprocating pumps for
pipeline transportation service. Pipelines are used to transport crudes, as well as refined or finished
products, including gasoline. Pipelines for motorized transport and aviation gasoline serve an important
strategic purpose in Europe, the Middle East, and Africa. In certain cases, the same pipeline may serve
to deliver several different products at different times.
Pipeline stations are installed at reasonable intervals, but still the total heads involved are high enough
to require multistage pumps. In the past, pipeline pumps were frequently built in six, eight, and even
ten stages. Stations were installed where two such pumps may have operated in series, as well as in
parallel. Series operation made it necessary to provide casings suitable for the high pressures developed
and shaft seals capable of withstanding the high suction pressure of the second pump in the series. As
pipeline flows increased, installations were made of pumps with fewer stages, partly because the larger
flows allowed the pumps to develop higher heads per stage. Figure 26.45 shows the sectional view of
a two-stage pump design frequently applied to pipeline service. Pumps of this type are normally furnished
with mechanical seals, generally tandem (Fig. 9.15) to minimize the risk of a major spill should the
primary seal fail in a remote installation. Figure 26.46 is a motor driven two-stage pipeline pump. When
bearing cooling is required and water is not available, as is common in the Middle East and many other
regions of the world, the bearings are either forced air cooled or cooled with the pumped liquid.
The trend today is to still larger pipelines (sometimes obtained by "looping" or paralleling an existing
line), outdoor stations, and automatic remote control. Pump drives are electric motor, engine, or gas
turbine depending on the availability of electricity and the size of the pump. When engine or gas turbine
driven, the fuel may be either natural gas or crude oil.
A vertical, "in-line," single-stage volute pump for pipeline booster service is shown in Figs. 26.47
and 26.48. The design shown has its impeller shaft rigidly coupled to the driver, and is similar in basic
configuration to the type VC vertical in-line chemical pump produced to ASME B73.2. Smaller pumps
are light enough to be supported by the pipeline itself, so the casing has no feet. Larger sizes need a
foundation and therefore have supporting feet as an integral part of the casing.
Pipelines terminate either at refineries or bulk-loading stations for tankers, or at storage tank stations
from where the crude or refined product is later loaded to tank trucks or tank cars. Pumps used at these
terminal points may be either single-stage or multistage and either horizontal or vertical. Vertical pumps
can be used for "stripping," that is transferring oil or gasoline from the very bottom of the tanks.
Services 773

Fig. 26.45 Section of two-stage volute pump with axially split casing.

Fig. 26.46 Motor-driven pipeline pump with axially split casing.


774 Services

Fig. 26.47 Section of single-stage, single-suction, vertical in-line booster pump.


(Courtesy BWIIP International.)

REFINING

This discussion will not present a detailed catalogue of all the pumping applications encountered in
refining petroleum. Engineers in this industry are quite familiar with these applications, and a complete
textbook would be necessary to acquaint others with them. The range of products handled in a refinery
is extensive; specific gravities as low as 0.35 to as high or higher than water; viscosities from lower
than water to values so high that rotary pumps must be used; and temperatures from cryogenic to as
high as 450°C (850°F). These products are pumped at flow rates from I to 8,000 m3/hr (5 to 35,000
gpm); and at pressures from as low as a fraction of a bar absolute (a few psia) to 215 bar (3,100 psig)
for hydrocarbon, 290 bar (4,200 psig) for water. Finally, the product may be as inert as water or extremely
corrosive and require pumps of higher alloy stainless steels.
The petroleum industry has always tended toward a high degree of standardization in its pumping
equipment. This can be attributed to the fact that pumps are spread throughout the refinery, therefore
their failure can have a significant effect on the refinery's operation and the safety of its operators. With
the pumps subject to a wide range of services, many prone to change depending on the crude slate, there
is a distinct advantage to standardizing, to the extent possible, on proven pump configurations, design
details, and materials of construction. In the 1940s, each oil company developed its own standard
Services 775

Fig. 26.48 Installation of vertical in-line, pipeline booster pump.


(Courtesy BWIIP International.)

specifications for various services in the refinery, and manufacturers developed lines of pumps that could
readily fit those specifications. This activity lead to the development of a combined specification, prepared
by the American Petroleum Institute and designated API-61O, the first edition being issued in 1954.
Since then seven further editions have been issued, the most recent, the 8th edition, in April 1995. Most
oil companies issue a supplement to API-61O to cover any special requirements they have for design,
materials, shop inspection, tests, and preparation for shipment. Although the later editions of API-61O
have covered the preparation of proposals and the submittal of order documents, most oil companies
issue a separate specification covering these and other commercial matters.
For most applications where a single stage pump can be used, the preference is for end suction (Figs.
1.9 and 13.13) or top suction (Fig. 2.17), centerline supported pumps, with overhung, single suction
impellers. Modem designs of these pumps have API-61O or better piping load capacity, high shaft
stiffness (see Chap. 7), large-diameter seal housings, high-capacity bearings, and the ability to operate
to 450°C (850°F) with only forced air cooling of the bearing house. Shaft seals are exclusively mechanical,
the type and arrangement varying with the service. Seals mounted directly in the casing head or cover
(Fig. 26.49), are currently out of fashion but offer the advantage of a shorter shaft overhang, hence a
stiffer rotor, and eliminate the need to make connections to the shaft seal through openings in the bearing
bracket, a construction legacy from packed box seals. Most manufacturers standardize on a small number
of bearing frames, each size of which is suitable for a large number of liquid ends.
Radially split, single stage, centerline supported double suction pumps with the impeller mounted
between bearings (Fig. 26.50) are necessary when the capacity is higher than can be handled by a single
suction pump. This limit is generally determined by the NPSH required for a single suction pump being
776 Services

Fig. 26.49 Section of overhung process pump with internal mechanical seal.
Note short impeller overhang.

more than the available NPSH less a reasonable margin (see discussion later in this section), and therefore
varies with plant design. Many oil companies also require a single stage between bearings pump when
the head exceeds 200 m (650 ft) and the power absorbed is greater than 200 kW (270 hp).
For heads higher than can be developed in a single stage at the prevailing two-pole motor speed, the
usual solution is a two-stage, single-casing, centerline supported pump. These pumps used to have
overhung rotors, but most the designs were marginal for rotor stiffness with a packed box shaft seal,
and therefore chronically unreliable when converted to mechanical seals. Given this experience, the
prevailing design today has a between bearings rotor (Fig. 5.12). The first stage is either single suction
as shown in Fig. 5.12 or double suction, depending on the pump's design flow.
A wide range of materials is employed in the construction of refinery process pumps. The minimum
standard for "on plot" applications is carbon-steel casings with cast-iron internal parts. Higher temperatures
dictate the use of 13 chrome internals and finally 13 chrome casings. The presence of naphthenic acid
at temperatures above 230°C (450°F) requires molybdenum bearing austenitic or duplex stainless steels.
Other acid services are handled with stainless steel, 25 chrome, 20 nickel steel, or Monel overlaid carbon
steel depending on the acid and its condition. API-610 includes a guide to the selection of materials for
various services.
Suction conditions are one of the most important aspects of the application and operation of refinery
pumps for two reasons: (1) in many applications the pump must handle hydrocarbons at temperatures
corresponding to the boiling point and (2) it is expensive to locate the vessels from which these pumps
take their suction at such an elevation that ample NPSH is available. The hydraulic design of refinery
Services 777

Fig. 26.50 Section of single-stage, double-suction, radially split process pump.

pumps, therefore, was directed at achieving very low values of required NPSH. As process unit sizes
increased, however, it became evident that in large pumps the hydraulic designs requiting the lowest
values of NPSH were also very limited in flow rangeability. When these pumps were run at flows below
design, they produced noise and vibration, which often lead to premature failure of the shaft seal,
bearings, shaft, or impeller (see Chap. 22). To avoid these difficulties, the general practice now is to
limit the suction specific speed of pumps for hydrocarbon service to 12,750 (m3/hr basis) or 11,000
(gpm basis).
As described in Chapter 19, the NPSH required for a given capacity and a given pump is frequently
lower for hydrocarbons than for cold water. The Hydraulic Institute [1.1] provides a chart (Fig. 19.14)
that allows the correction of cold water tests for application on hydrocarbons. It should be noted, however,
that the industry does not use the so called "hydrocarbon correction" as a matter of course. Instead it
is kept as an additional margin to cover unforeseen changes, such as an increase in unit throughput.
Normal application practice is to provide an NPSH margin equal to 10 percent of the required NPSH
or 1.0 m (3 ft), whichever is greater.
In some applications, usually debottlenecking, the available NPSH is too low for conventional horizon-
tal centrifugal pumps. One solution to this problem is to use a pump with an inducer (Fig. 4.17). Correctly
designed inducers can achieve suction specific speeds of 17,500 to 19,750 in metric units (15,000 to
17,000 in U.S. units) while retaining flow range ability equivalent to an impeller of suction specific speed
12,750 (11,000). Designs of higher suction specific speed have narrower allowable flow ranges, a
limitation that needs to be respected in the operation of the pump. Another solution, employed when
the available NPSH is below the capability of an inducer or the inducer's allowable flow range is too
narrow, is the use of vertical can pumps (Figs. 14.22 and 14.23).
778 Services

Hot oil charge pumps now develop pressures up to 214 bar (3,100 psig). At temperatures to 230°C
(450°F) and specific gravities 0.7 and higher, pumps with axially split casings are used for pressures up
to 100 bar (1,450 psig). For higher pressures, higher temperatures, or lower specific gravities, the usual
construction is radially split double casings. The double-casing pumps used in this service are the same
basic configuration as those for boiler feed service (Figs. 3.14 and 3.15). In hot oil charge service,
however, the choice between these two configurations is currently quite controversial. It is generally
known as the "volute" versus "diffuser" pump debate. These terms derive from the fact that most pumps
with opposed impellers and axially split inner casings use "volute" collectors, and those with a balancing
device and a radially split inner casing, "diffuser" collectors. Many refiners argue that compared to
"diffuser" pumps, "volute" pumps don't suffer axial thrust problems and are easier to overhaul. The two
weaknesses observed in "volute" pumps are wear at the center bushing and internal leakage. Chapters
3, 5, and 6 deal with the relative merits of the means of splitting the inner casing and balancing axial
thrust. The discussion is not complete, however, without taking into account rotor stiffness (Chap. 7).
Much of the experience with hot oil charge pumps has been with pumps of many stages, small shafts,
and therefore slender rotors. The time between overhauls of many of these pumps is quite short, from
1 to 3 years depending on the application. In such cases the ease of overhauling pumps with an axially
split inner casing is an advantage, unless the inner casing or casing has been damaged by internal leakage.
When the rotor stiffness is raised by using fewer stages and a larger shaft, the practice in modem
"diffuser" pump design (Fig. 1.11), the time between overhauls increases and freedom from internal
leakage (reliability) becomes more important than the ease of overhaul (maintainability), thus pointing
to a radially split inner casing as the better choice.
Hot oil charge pumps differ from boiler feed pumps in several important respects, all to do with
higher temperatures, up to 480°C (900°F). Internal joints are arranged to not "gum up" with hydrocarbon;
differential thermal expansion is avoided, particularly in radial fits, and where it cannot be avoided,
special means are provided to accommodate it; the casing is free to expand as it comes up to temperature,
and to do so without significantly changing the driver-to-pump shaft spacing; shaft seals are mechanical
and usually suitable for the pumping temperature, even when injected with cool oil to help extend their
face life. Many advances have been made in hot oil charge pump design, with the result that good
examples are now realizing 50,000 hours between overhauls. With attention to the shaft seals, 75,000
hours is achievable.
Decoking is another special service in refineries where high-pressure, multistage pumps are used.
Water discharged through special nozzles under high pressure is used to cut the coke inside coking
drums. Decoking jet pumps for pressures to about 100 bar (1,500 psig) had axially split casings with
integral volutes (Figs. 3.6 and 3.7). As drum sizes increased and the necessary cutting pressures
rose, pumps with radially split double casings became necessary. The most common configuration
is a large-shaft pump with tandem impellers and a balancing device in a radially split inner casing.
The shaft in modem jet pumps is large enough to allow rotors of up to 10 stages without the static
deflection exceeding the minimum radial clearance at rotor midspan. Experience has shown that this
degree of rotor stiffness achieves significantly longer periods between overhauls than do more flexible
rotors. With a low specific speed hydraulic design, a lO-stage jet pump can develop up to 235 bar
(3,400 psi) direct driven at 3,575 rpm. For higher pressures, the pump is driven at higher speeds
using motor/gear, steam turbine, or variable-frequency motor drive. Because they handle cold water,
the first stage impeller of jet pumps should be designed for a suction specific speed of 10,500 to
11,000 in metric units (9,000 to 9,500 in U.S. units) to avoid rough operation at low flows and
premature erosion of the impeller.
Both electric motors and steam turbines are used to drive refinery service pumps. Steam turbines
used to be the preferred driver because they allowed process steam to be raised economically at a high
pressure, then dropped to the required process pressure across back-pressure turbines. At the same time,
Services 779

it afforded a separate source of motive power, which improved refinery availability. With a swing in
the economics of steam generation, and the development of cross wired electrical systems in refineries,
steam turbines can no longer be justified unless the power and speed are high, variable-speed operation
is necessary, or a nonelectrical standby pump is deemed mandatory.

Pulp and Paper Mills


The manufacture of pulp and paper is a complex series of processes, each of which employs a great
number of centrifugal pumps. The paper industry is a separate industry from pulp making, as many mills
making paper do not make pulp. Pulp manufacturing consists of separating cellulose fibers from any
foreign matter present in the raw material and preparing these fibers into a pulp suitable for the manufacture
of paper, paper board, rayon, cellophane, explosives, and a variety of other products.
Pulp can be manufactured either by a strictly mechanical process or by breaking up wood or other
vegetable matter and cooking the broken-up material with chemicals in a digester. The products of
chemical pulp making may be an alkaline pulp (soda or sulphate) or an acid pulp (sulphite), depending
on the chemicals used and on the desired end product. Thus, a pulp mill uses pumps to handle liquids
such as clear water, acids, caustics, white liquor (the cooking liquor in the digester), black liquor (the
spent liquor separated from the stock after cooking), and green liquor (chemicals recovered from the
recovery furnace and dissolved in water). Therefore, in addition to general-service pumps, which handle
clear water, pulp mills use a variety of specialty pumps, generally similar to chemical pumps.
In paper mills pumps must handle stock with a consistency ranging from 0.1 percent to as much as
14 percent, as well as clear water. Centrifugal pumps with semi-open impellers (Fig. 2.23) are used for
stock up to 6 percent consistency. Higher consistencies require screw pumps or centrifugal pumps with
a non-clog impeller and a special booster or screw feeder that helps the high consistency stock enter
the impeller without separating or dehydrating. The shaft seal for stock pumps can be a packed box, an
expeller, or a mechanical seal. When a packed box seal is used (Fig. 2.23), the seal cage is located at
the bottom of the box. An external source of clear sealing water prevents any of the fibers in the stock
from entering the seal and shortening the life of the packing. If a mechanical seal is used, it is a
configuration suitable for solids laden liquids (Fig. 9.27), so its compression unit will not become clogged.
The friction losses of stock vary widely with the stock consistency. Friction losses of groundwood
stock and of SUlphite stock are detailed in the standards of the Hydraulic Institute, and this information
should be used in estimating head requirements of paper stock pumps.

Textiles
Textiles is another industry that requires the handling of a wide variety of liquids in addition to clear
water. General-service pumps are used for water or cold noncorrosive liquids. Chemical pumps are
widely used to handle liquids such as acids, alkalis, acetates, and dyes. One very important requirement
in textile manufacturing is the complete assurance of the absence of foreign matter that could affect the
coloring of the end product. As a result, stainless alloys are generally used-more for eliminating any
corrosion products whatsoever than for increasing the life of the pumping equipment.
780 Services

Natural and Synthetic Rubber


Centrifugal pumps are used to handle a variety of liquids such as acids, sodas, hydrocarbons, latex, and
solvents. The problems of pump selection for these applications are similar to those encountered in the
chemical industry.

Food and Beverage Processing


Pumps handling foods or beverages must be of a special sanitary design. The main requirements are a
complete absence of corrosion and ease of dismantling for cleaning. Most food-processing pumps are
made of stainless steels, glass, or porcelain, although some all-bronze pumps are used, especially in the
beverage industry. The designs must be free of any pockets where particles or liquids could accumulate
and where bacteria might breed. Neither oil nor grease lubricant can be permitted to enter the pump,
and, as a result, special packing must be used in the stuffing boxes or mechanical seals applied.

Mining
Two general categories of application of centrifugal pumps are encountered in mining operations. The
first involves the dewatering of different types of mines, such as coal and metal ore. The second category
includes all applications where pumps are used in the processing of the mined product, such as in
leaching operations, coal washing, transfer of precipitates, and waste disposal.
In almost all cases mine-dewatering pumps handle corrosive or abrasive waters against relatively
high heads. Multistage pumps, built of corrosion-resistant materials, ranging from all-bronze to chrome
or nickel stainless alloys are, therefore, used most frequently. Vertical turbine pumps have become a
favorite style in recent years as they require very little space. In general, the difficulty of introducing
pumps into their ultimate location within the mine calls for special construction features. In some cases,
it is necessary to dismantle the pump, transport the parts into the mine, and reassemble the pump when
it is in place.
As mines are always in danger of becoming flooded, special sinking pumps, mounted on small trucks
for portability, are employed. These units can use the mine tracks to move from place to place. Frequently
the pumps must be capable of operating in an inclined position. These pumps may also be mounted in
cages that are lowered into the shaft as dewatering progresses. Lengths of discharge piping are added
each time the pump is lowered further.
As described in Chapter 20, the problem of dewatering a flooded mine is unlike most applications
of centrifugal pumps in that the pump must work against a total head that varies from practically zero
to the maximum that occurs when the mine is almost clear of water. A dewatering pump designed to
work as two two-stage pumps in parallel, initially, and one four-stage pump when the total head
requirement exceeds that developed by the two-stage combination is shown in Fig. 26.51. Its hydraulic
performance was shown in Fig. 20.36. The pump with its 450 kW (600 hp), 1,450-rpm motor is mounted
in a welded channel frame. A cable sheave is mounted in the frame to permit the unit to be suspended
Services 781

Fig. 26.51 Mine dewatering pump arranged to operate as either two- or four-stage unit.
782 Services

vertically. The frame and sheave are capable of supporting the discharge pipe full of water weighing
up to 34,000 kg (75,000 lb).
Pumps used for coal washing, mineral processing, and leaching generally handle either water contami-
nated with a considerable amount of solids, or high-concentration aqueous slurries. These pumped liquids
are always abrasive to some degree depending on the nature of the solids and their concentration.
Frequently the mixture is also corrosive, which complicates the selection of materials. Slurry pumps are
generally rubber lined when ever the service conditions (head; type of solids and concentration; tempera-
ture) allo~ because this construction (Figs. 26.42 and 26.43) is economical and affords excellent corrosion
resistance. For conditions beyond the capability of rubber lined pumps, the usual sequence of construction
choices is hard metal impeller in a rubber lined casing; hard metal impeller and casing (Fig. 2.20);
manganese steel impeller and casing. The final material choice, manganese steel, is for services handling
large solids, such as dredging, where the pump materials must absorb significant shock without spalling.
In general, operating speeds are much lower than for pumps handling clear water, and as the severity
of slurry services increases, the pump speed usually decreases further. Common shaft seal options are
(1) packed box with clear water injection to keep the slurry out of the packing, (2) hydrodynamic or
expeller seal with a packed or lip type auxiliary seal (see Chap. 8), (3) a mechanical seal designed for
slurry service (see Chap. 9).

Construction
Centrifugal pumps used in construction work include units for drainage or jetting. Drainage units are
generally small, engine-driven, portable units, self-priming, if horizontal. Jetting pumps may develop
pressures from 5 to 10 bar (75 to 150 psi) and are usually designed as portable, engine-driven units.

Steel Mills
The steel industry uses a large number of centrifugal pumps for a wide variety of services, such as
cooling of rolls and furnaces, general water supply service, hydraulic-descaling, scale-pit recirculation,
and many others.
Service water pumps handle clean water at pressures up to 10 bar (150 psig) and for capacities up
to 2,250 m3/hr (10,000 gpm) or more. Standard horizontal or vertical, general-service pumps, both single-
stage and multistage are used on this service. In a few cases, where river water is contaminated by
industrial wastes and is quite corrosive, special materials may be required, but, generally, standard fitted
pumps are used.
Various cooling services may use pressures up to 17 bar (250 psig). Here again, unless special water
conditions exist, standard fitting pumps are satisfactory.
Hydraulic descaling is the process in which scale is removed from metal by using the impact of a
jet directed at the metal surface through special nozzles (Fig. 26.52). In the 1930s, when descaling was
first developed, pressures of 70 bar (1,000 psig) were considered quite sufficient. Since then, pressures
have risen through 125 bar (1,800 psig) and 160 bar (2,300 psig) to the current maximum of 205 bar
(3,000 psig). A typical hydraulic descaling system is illustrated in Fig. 26.53. For lower pressure systems,
preference has been given to multistage pumps with axially split casings (Figs. 3.6 and 3.7). Higher
Services 783

PIPE
HREAO

STANDARD TYPE STRAINER TYPE SHUTOFF TYPE

Fig. 26.52 Descaling nozzle.

pressure systems have generally employed pumps with radially split double casings (Figs. 3.14 and
3.15). Some mills, concerned with first cost, have used axially split casings for high pressures. Others,
concerned about the cost of energy, have used very large vertical plunger pumps. Because of their long-
time familiarity with gears, steel mills often used four-pole motors and step-up gears to drive pumps at
3,600 rpm and higher, years ahead of steam power plants taking up the same practice.
Descaling pumps are designed either to pump directly to the system or to be assisted by accumulator
air bottles. An accumulator system has two advantages: (1) smaller pumping equipment may be used,
the accumulator is recharged during the time when water is not required, and (2) the system has some
"capacitance," which reduces the severity of flow changes in the pump, an important aspect for the
larger pumps now in use. The cost, however, is considerably higher than that of a direct pumping system,
and a thorough analysis must be made in the average and peak demands and the means of varying pump
flow between rated and bypass before choosing between the two systems.
Primary scale-pit pumps recover the water used for descaling and pump it to the mill water treatment
system for further settling, then filtering and cooling before it is return for reuse. Pumps in this service
therefore handle water containing considerable amounts of scale. Large cantilever wet pit pumps without
any bearings below the pit's high water level (Fig. 14.32), are commonly used for this service; their
liquid ends are typically hard iron to provide adequate erosion resistance. Agitation nozzles below each
pump prevent scale buildup on the pit floor from blocking the pump's suction.

Hydraulic Presses
Applications for hydraulic presses occur in the metal-forming industry and in many other industrial
processes. Installations include such applications as forging, die-forming, extrusion, injection-molding,
and diecasting. Unless a large number of individual operations are served by a single pump, the tendency
is to reduce the size of the pumping equipment by using accumulator systems, as in the de scaling process.
The same type of pumps used in descaling service are employed for the higher pressures, whereas two-
and four-stage pumps with axially split casings are used for the lower range of pressures.
The water handled by these pumps is generally pure and non-corrosive, with occasionally a slight
admixture of water-soluble oil. Corrosion problems are, therefore, seldom encountered, and materials
are selected for their suitability for the particular pressure requirements.
784 Services

7$
I !
!I'IItAOllt

~- -
- . - .! ' -

Fig. 26.53 Hydraulic de scaling installation.


KEY:
A = pump and motor
B = accumulator
C = air compressor
D = gage glass
E = two-way automatic valve
F = stop valve
G = nozzles and header
H = pressure regulators
I = automatic stop valve
J = automatic strainer
Ie = stop valves
L = bypass
M = stop valve
N = check valve
o = automatic bypass
Q = stop valve
R = check valve.
S = stop valve
T = bypass for starting
U = orifice nipple
V = relief valves
W = bypass
X = orifice nipple
Y = stop valve (lock type)
Z = drain valve.
Services 785

Air-Conditioning
Centrifugal pumps used in air-conditioning include general water supply units, chilled water pumps to
circulate water from the chillers through the air-cooling coils, air washing pumps in some installations,
and circulating pumps for the refrigerating condenser. The water handled is clean, and pressures in most
cases such that single-stage, standard, fitted pumps can be used. Depending on size of the unit, end-
suction, foot-mounted (Fig. 2.14), close-coupled (Fig. 1.19), or axially split casing pumps (Fig. 2.15)
are used. Some vertical-turbine pumps are used for the water supply application. The availability of
mechanical seals in recent years has made this construction of value in some cases because of the reduced
amount of overseeing required.

Refrigeration
Brine pumping is the one special application required in refrigeration systems, cold storage, or ice
making. Both sodium chloride and calcium chloride brines are used, the latter much more often than
the former because it has a lower freezing point and is less corrosive. The specific gravity of the brine
solution that the pump has to handle varies between 1.05 and 1.28, depending on the minimum temperature
maintained in the expansion coils.
The viscosity of brine solutions increases with specific gravity and with decreasing temperatures to
a degree that it definitely affects pump efficiency. If this effect is not considered, a strong probability
exists that a motor driver selected on the basis of water efficiency will be overloaded. It is necessary,
therefore, to establish the viscosity of the brine at the minimum pumping temperature and determine
the efficiency correction factors as described in Chap. 18 (Figs. 18.53 and 18.54).
All-iron pumps are used for calcium chloride brine, and all-bronze pumps are preferred for sodium
chloride brine. Special attention is required for the bearing construction; since the temperature of the
brine is below that of freezing water, adequate seals must be used to prevent the moisture present in
the air from condensing inside the bearings. In some installations, it is necessary to circulate warm water
through the cooling jackets of the bearings to avoid lowering the temperature of the lubricants excessively.

Heating
Building heating systems require hot water circulating pumps and, in some cases, condensate return
units. The first are generally small end-suction, close-coupled pumps built for that special purpose by
several manufacturers. Regenerative pumps (Chap. 15) are also used for this purpose, as they are suitable
for the range of low capacities and high heads encountered. There is a tendency to supply these pumps
with mechanical seals to reduce maintenance.
Various condensate return units are available on the market. They consist of a condensate returns
receiver and one or two small, close-coupled centrifugal pumps (Fig. 26.54). An automatic float switch
starts and stops the pump, which discharges the returns back into the boiler.
A definite trend exists toward high-temperature heating applications, with temperatures ranging from
115 to 205°C (240 to 400°F). A high-temperature, hot-water system consists of a hot-water boiler, an
expansion drum, a pumping system and the heat-using system. Two different arrangements are used:
786 Services

Fig. 26.54 Condensate return unit.

(1) The single-pump system (Fig. 26.55), where the same pump or pumps are used to provide circulation
through the boilers and to distribute the hot water through the system, and (2) the separate-pump system
(Fig. 26.56), where two separate sets of pumps are used, one for boiler circulation and one for hot-water
circulation to the heat users. In either case, the pumps take their suction from the expansion drum, which
is used to pressurize the system and to store hot water.
Because of the relatively high operating temperatures and suction pressures, end suction pumps with
only one shaft seal are employed. Their construction is similar to high-pressure heater drain pumps or
low-pressure boiler circulating pumps (Fig. 26.20). Shaft seals are mechanical, with cooling by self-
circulation through a heat exchanger (API Plan 23; Fig. 9.34), or where that is not possible because the
nominal seal surface speed is too low, by circulation from the pump discharge through a heat exchanger
(API Plan 21; Fig. 9.33).
Services 787

EXPANSION DRUM

BOILER BOILER BOILER


NO. NO. NO.
I 2 :!>
SPARE

HEAT CONSUMERS

Fig. 26.55 High-temperature, hot-water (HTHW) heating system.


Single-pump system.

SYSTEM CIRCULATING PUMP

BOILER
NO.
I
BOILER
NO.
2
BOILER
NO.
3
\
BYPASS VALVE
SPARE

HEAT CONSUMERS

Fig.26.56 High-temperature, hot-water (HTHW) heating system.


Separate-pump system.

Process Condensate Returns


In a number of processes that reclaim condensate returns, a certain amount of air returns with the
condensate and has to be adequately vented to avoid air binding the return pump. One special design
is illustrated in Fig. 26.57. No return tank is used, as the unit is self-contained. The centrifugal pump
draws water from the priming loop and discharges it at high velocity through a nozzle. The mixture of
hot condensate, steam, and air from the returns flows into the low-pressure area surrounding the nozzle
and is drawn into the mixing tube where it is combined with the circulating water. This additional
amount of condensate discharges an equal amount through the air separator, either directly to the boiler
788 Services

Fig. 26.57 Self-contained condensate return centrifugal pump.


(Courtesy Cochrane Corp.)

or to the reservoir from which a boiler feed pump delivers it to the boiler. The fins that surround the
priming loop act to subcool the liquid-vapor mixture in it and reduce the submergence required at the
pump suction by reducing the vapor pressure of the condensate entering the pump.

Marine Pumping
The wide range of requirements for pumping services aboard ship calls for a correspondingly extensive
selection of types and sizes of centrifugal pumps. The most important applications for marine service
Services 789

are listed in Table 26.5. Although it is impractical to discuss the requirements of each of these applications
in detail, some of them will be described in this chapter.
There are, first, certain general considerations that apply to all marine services. Centrifugal pumps
for ship use must operate at or near the safe maximum speed at which a particular condition can be
met, as the use of the highest rotative speed permitted by the suction conditions will result in the smallest
pump and driver and, therefore, the lowest weight and smallest volume unit.
Marine service requires the greatest possible degree of reliability. Standards have been developed to
guarantee acceptable designs, and numerous agencies have prepared specifications for centrifugal pumps
in this service. Some of these agencies are

1. American Bureau of Shipping


2. U.S. Maritime Commission
3. U.S. Coast Guard Marine Inspection Service
4. U.S. Navy Department Bureau of Ships
5. U.S. Navy Department Bureau of Yards and Docks
6. Army Transport Service
7. U.S. Corps of Engineers
8. Lloyds
9. Bureau Veritas

In most cases, special inspection procedures made by a customer's representative or by a navy


inspector are required at the point of manufacture. Witnessed performance tests are generally required.
Type approval of a particular pump is generally required in the case of pumps for U.S. Navy vessels.
This involves a thorough analysis of the design as well as extensive tests at navy laboratories.

Table 26.5 Index to Centrifugal Pump Applications in Marine Service

Atmospheric drain Elevator


Automatic priming Engine cooling
Ballast Evaporator system
Bilge Fire
Boiler feed Flushing
Booster Fresh water
Brine circulating Gasoline
Butterworth system General service
Caisson gate dewatering Gun cooling
Cargo Heater drains
Condensate Hot-water circulating
Condensate return Ice-water circulating
Condenser circulating Jetting
Damage control Make-up feed
Diesel fuel oil Potable water
Distiller plant Refrigeration brine circulating
Drainage Refrigeration condenser circulating
Drinking water Sanitary
Drydock Wash water
790 Services

Fig. 26.58 Two-stage navy boiler feed pump connected to steam turbine by flexible coupling.

BOILER FEED PUMPS

With the exception of some of the latest ships, which operate at higher steam pressures and require
double-casing pumps, most marine feed pumps are multistage pumps with axially split casings. "They
are primarily designed for steam~turbine drive, which permits the desirable higher speed operation and
variable speed control.
Although some standard commercial designs adapted for marine service are occasionally used, special
boiler feed pumps are generally developed by manufacturers for this service, especially for navy applica-
tion. Figure 26.58 shows a two-stage boiler feed pump connected to a steam turbine through a flexible
coupling. Forced feed lubrication for pump and driver bearings is provided by means of a submerged
oil pump. Figure 26.59 shows a sectional view of a three-stage navy boiler feed pump.
Figure 26.60 shows a three-stage boiler feed pump with its steam turbine. A single shaft, supported
in three bearings, is used for pump and turbine, thus eliminating the connecting coupling.
A four-stage, double-casing, barrel-type boiler feed pump used on navy vessels operating with high-
steam pressures is shown in Fig. 26.61.
A different type of boiler feed pump is shown in Fig. 26.62. It is a turbine-driven unit of Monobloc
construction, designed so that there is no stuffing box between the pump and the turbine. Leakage
from the pump proceeds into a mixing or quenching chamber, from which it is fed back into the
feedwater system.
Some vertical, multistage boiler feed pumps have been installed aboard ships to save space. One such
unit is shown in Fig. 26.63.
Except for low-pressure service and some commercial applications, boiler feed pump casings are
made of chrome or chrome-nickel steel. Impellers, wearing rings, diaphragms, shaft sleeves, and shaft
nuts; that is, all parts in contact with the feedwater, are usually made of chrome steel or Monel metal.
Services 791

Fig. 26.59 Three-stage navy boiler feed pump.

Fig. 26.60 Turbine-driven, three-stage navy boiler feed pump with common turbine and pump shaft.
792 Services

Fig. 26.61 Four-stage, double-casing, barrel-type navy boiler feed pump.

CONDENSATE PUMPS

Condensate pumps are almost always arranged for vertical operation. Figure 26.64 shows a vertical,
single-stage, single-suction condensate pump with an internal water-lubricated bearing, an external ball
thrust bearing, and a mechanical seal. Figure 26.65 shows a vertical, two-stage, single-suction condensate
pump. The inlet of the first-stage impeller points upwards to facilitate venting. Figure 26.66 shows a
photograph of this pump, with the removable casing half lifted and the motor sectionalized for display.
Condensate pumps must always be located close to the condenser hotwell, to minimize the effect of
the ship roll. Casings are generally made of bronze or Monel while impellers are bronze or monel.

CONDENSER CIRCULATING PUMPS

Commercial designs of horizontal, double-suction, single-stage pumps with axially split casings are
frequently adapted for vertical service on marine application. Figure 26.67 shows such a pump with
external ball bearings and a flexible coupling. This is essentially a low-speed pump and requires more
head room than the mixflo or propeller pump would under identical operating conditions. This type of
pump has also been built with an internal grease- or water-lubricated bearing and a rigid coupling; the
motor or turbine thrust bearing is designed to take the weight of the pump rotor and any unbalanced
axial hydraulic thrust.
Vertical mixflo pumps are frequently used for marine condenser circulating service. They operate at
higher speeds than double-suction pumps. Still higher rotative speeds can be used with propeller pumps.
Services 793

Fig. 26.62 Turbine-driven, Monobloc boiler feed pump without stuffing boxes.

Figure 26.68 shows a vertical propeller pump of a special short design with a water-lubricated bearing
and a rigid coupling, and Fig. 26.69 shows a photograph of the same pump driven by a geared-down
steam turbine.
Materials generally used for condenser-circulating service are: bronze for the casing, impeller, and
wearing rings; Monel metal for the shaft; bronze or Monel for the shaft sleeves; and leaded bronze for
the water-lubricated bearings.

CARGO PUMPS

Centrifugal cargo pumps are designed to pump large volumes of crude oil, fuel oil, and other liquid
hydrocarbons, as well as salt water ballast, to give fast cargo unloading, so essential to profitable
tanker operation.
794 Services

Fig. 26.63 Vertical, multistage boiler feed pump.

Fig. 26.64 Vertical, single-stage, single-suction Fig. 26.65 Vertical, two-stage, single-suction marine
marine condensate pump. condensate pump.
Services 795

Fig. 26.66 Condensate pump shown in Fig. 26.65, with front-half casing removed and rotor sectioned.

Figure 26.70 shows a typical single-stage, double-suction pump used in this service. Two large flanged
vent connections on top of the casing suction spaces vent off vapors and air that would otherwise cause
the pump to lose suction when the liquid level in the cargo tanks lowers. This feature permits pumping
down each cargo tank to the lowest possible level with the high-capacity centrifugal cargo pumps, leaving
a minimum of cargo to be stripped.
Vertical volute and turbine pumps are also widely used in cargo service (Chap. 14).

DRYDOCK PUMPS

Drydock service requirements are very special, in that the head varies from a low value, made up
principally of frictional head at the beginning of pumping, to a higher head composed largely of static
head at the end of the pumping period. The head-capacity curve of drydock pumps should, therefore,
be quite steep, to provide the minimum of capacity reduction with this increase in head and a minimum
of reduction in efficiency. As the characteristics of mixflo pumps are well-suited for this requirement,
most of the drydock pumps are this type (Fig. 14.13). Other pumps used as main drydock pumps are
the vertical turbine pump, or the vertical propeller type (Fig. 14.26).
All of these general types of pumps may be used for graving docks or in floating drydocks, with
special construction in each case to suit the installation. Graving docks generally have a dry-pit pump
room on one side of the dock, with the main pumps located on the floor that is approximately at the
same level as the dock floor. A flared suction bell below the pump opens into the suction well connected
796 Services

Fig. 26.67 Vertical, double-suction marine condenser circulating pump with axially split casing.

to the dock by means of a large culvert. The pump is connected to the motor near the ground surface
level by a long shaftipg with guide bearings.
Floating drydocks invariably require submersible pumps located in the flooded sections of the dock.
The pump shafting is enclosed in a cover pipe to protect the shaft, bearings, and couplings from water.
It is usually desirable to have a split section of cover pipe adjacent to the pump, enclosing a removable
short piece of shaft. This permits dismantling the pump without disturbing the intermediate shafting.
Services 797

Fig. 26.68 Vertical, propeller-type, condenser Fig. 26.69 Condenser, circulating pump shown in
circulating pump. Fig. 26.68.

Pumped Storage
A pumped storage installation is, in a manner of speaking, a hydraulic storage battery. In a pumped
storage system, electric power is used to lift water into a reservoir during off-peak periods. Later, the
water is drawn from the reservoir to drive a hydraulic turbine that provides incremental power during
peak load periods. The electricity used for pumping comes from adjacent base-loaded steam power
plants, either fossil fuel or nuclear. In many installations, special storage reservoirs have been built so
798 Services

Fig. 26.70 Horizontal, single-stage, double-suction cargo pump.

MOTOR CONTROl.. CENTER

MAX POOL ANE


ELEVATION 642.0 FT

rl'tv~~~~
INTAME GA............,...,~-H":"i
MOTOR GENERATOR

ORAFT TUllE GATE SLOT

TRASH RACK AND RAKE SLOT


W. ELEVATION !\&!I.O n
NORMAL ELEVATION 5M.0 n
~ ELEVATION 548.0 FT

Fig. 26.71 Reversible pump-turbine unit.


Tuscarora Pump Generating Plant, Niagara Falls, NY. (Courtesy Voith.)
Services 799

water pumped during evening and weekend off-peak periods can be utilized in the daytime for generating
additional electricity.
Several arrangements are possible. Older installations used a separate motor-driven pump and a
hydraulic turbine driven generator, or a tandem unit, which consists of a motor-generator, a hydraulic
turbine, and a centrifugal pump. When delivering power, the motor-generator acts as a generator, driven
by the turbine to produce electricity; the pump runs dewatered or is disconnected and remains idle. At
off-peak load periods, the motor-generator operates as a motor driving the pump, which delivers water
to the upper storage reservoir. During pumping, the turbine is dewatered with air under pressure and
absorbs negligible power. Modem installations use reversible pump-turbines. These are single turbo-
machines able to operate as either a pump or a turbine. The impeller runs within a casing that has
adjustable guide vanes (wicket gates) and a stay ring. It operates as a pump rotating in one direction,
driven by the motor, to deliver water to the upper storage reservoir. Rotating in the opposite direction,
it operates as a hydraulic turbine, drawing water from the upper storage reservoir, and the motor becomes
a generator. Figure 26.71 shows one of 12 reversible pump-turbines installed at the Tuscarora Pump
Generating Plant, Niagara Falls, N.Y. It is rated 21,000 kW (28,000 HP) at 23 m (75 ft) as a turbine,
25,000 kVA as a generator, 28,000 kW (37,500 HP) as a motor, and 96.4 m3/sec (3,400 ft 3/sec) at 26
m (85 ft) as a pump. Its operating speed is 112.5 rpm. Since Tuscarora, significantly larger pump-turbines
have been put into service, the largest being at Bath County in Virginia (turbine output 380 MW rated,
508 MW maximum), while the highest head is 701 m (2,300 ft) at Chaira in Bulgaria.

BIBLIOGRAPHY

[26.1] ASME Boiler and Pressure Vessel Code, Section III, Rules for the Construction ofNuclear Power Components,
1995, American Society of Mechanical Engineers, New York, New York.
[26.2] ISO 2858, End-suction centrifugal pumps (rating 16 bar~Designation, nominal duty point and dimensions,
2nd Edition, 1975, International Organization for Standardization, Geneva, Switzerland.
[26.3] ASME B73.2M, Specification for Vertical In-Line Centrifugal Pumps for Chemical Process, 1991, The
American Society of Mechanical Engineers, New York, New York.
27
Procuring Centrifugal Pumps

"Procuring" for this discussion means getting what is needed. For machinery such as centrifugal pumps,
this usually involves four steps:

1. Deciding on the purchasing arrangement


2. Preparing the specification and bid request
3. Evaluating the bids and selecting the equipment
4. Placing the order and verifying its progress.

In the following sections, each of these steps is addressed with the emphasis on what is necessary for
the successful procurement of centrifugal pumps.

Purchasing Arrangement
Reduced to the essential, there are two arrangements for purchasing centrifugal pumps and, for· that
matter, most other equipment and services: specify the pump and select the manufacturer from a number
of specific bids, or select the manufacturer then jointly specify and select the pump. For convenience
these arrangements can be termed "specify and evaluate" and "select and negotiate," respectively. Which
arrangement is used depends on the fiscal objectives of the purchaser, or in the case of some government
instrumentalities, leglislation that dictates how the purchaser must proceed.
Under the first arrangement, "specify and evaluate," the purchaser prepares a specification, issues
inquiries to a number of qualified manufacturers (or, in some government business, issues a public
invitation to bid), then evaluates the resulting bids against certain criteria to select the manufacturer.
Depending on the extent of the specification and how the bids are evaluated, this arrangement generally
achieves the lowest purchase price, but often at the expense of compatibility between the plant design
and the pumps (the plant design is usually well advanced before the pumps are selected), and always

800

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
Procuring Centrifugal Pumps 801

at the expense of lead time from plant design to ordering pumps. The risk in using this arrangement is
that over the long term the "total cost of ownership" can be higher than it need be, because the
pumps might not be the optimum selection for overall energy consumption and mean time between
repair (MTBR).
The second arrangement, "select and negotiate," has the purchaser first select a manufacturer, by
either a sample bid or an assessment of overall competence, then the pump specification and selections
are jointly developed in parallel with plant design. Pricing is determined as each selection is made, based
on a negotiated price list for standard designs and options. Using this arrangement, compatibility between
the plant design and the pumps is ensured, and the lead time to ordering pumps, hence the receipt of
engineering information, is significantly shorter. Compared to "specify and evaluate," the net result of
"select and negotiate" is a shorter project cycle, and generally a lower "total cost of ownership."
Another way of looking at the two purchasing arrangements is to consider who determines the
fundamental pumping arrangement. Under "specify and evaluate," this is done by the plant designer,
who then must be well versed in the design and application of centrifugal pumps to determine the
optimum selection. The alternative approach, "select and negotiate" puts the onus for determining the
optimum pump selection on the pump manufacturer. As such, the pump manufacturer becomes the plant
designer's pump specialist, thus providing the best environment for developing compatibility between
the plant design and the pump selection. Against this, it can be argued that a good plant designer will
seek opinions from various manufacturers before finalizing the specification for a difficult or novel pump
application, a process that therefore affords the best of both worlds. That it may do, but the probity of
taking a good solution from one manufacturer and allowing others the opportunity of bidding for it is
questionable. And in recognition of this, it is unlikely that a manufacturer will advance his best solutions
unless he is certain they'll be kept confidential.
The following discussion will center on what is required to purchase by the "specify and evaluate"
arrangement since it is the more complicated. "Select and negotiate" contains the same elements, but
because the pump specification and selection is a joint effort, they are not developed in as formal a manner.

Essential Data Required


In ideal circumstances, the data required to select a centrifugal pump are those that describe its function:
liquid, conditions of service, required redundancy, required service life, and intended drive arrangment.
In practice, the responses to such a specification can be very hard to evaluate, so most specifiers include
at least some specific design requirements to try to ensure they get the type and construction of pump
they want. The specific requirements range from citing an industry standard to an elaborate treatment
of various aspects of design, manufacture, and testing.
It is impossible to present an exhaustive survey of each one of the factors to be studied when preparing
a formal pump inquiry; experience and intimate contact with the particular installation are the only
guides to complete knowledge. Nevertheless, it is possible to outline the essential data required to allow
the manufacturer to select a centrifugal pump for any given installation intelligently (Fig. 27.1). Finally
a fundamental point in preparing the inquiry is to state what is required in the simplest possible manner;
complexity defeats the objective.
The following represents the required essential data:
1. Number of units required
2. Nature of liquid to be pumped
a. Fresh- or saltwater, acid or alkali, oil, gasoline, slurry, or paper stock
b. Vapor pressure of liquid at pumping temperature
802 Procuring Centrifugal Pumps

c. Specific gravity
d. Viscosity conditions
e. Amount of any suspended foreign matter present, and size, nature, and abrasive quality of solids-if
liquid is pulpy, consistency either in percent or in pounds per cubic foot of liquid.
f. Chemical analysis, including pH value, predicted variations from analysis, impurities, oxygen content,
past history, and scaling tendency, if any

FORM A 1 CENTRIFUGAL PUMP DATA SHEET


PUMP IIZE AND MDDEL _ _ _ _ _ _ _ _ _ BRG. FflAME _ _ _ _ __ SERVICE _ _ _ _ _ _ _ _ _ •_ _ __
NO. PU. . . REaD. _ _ _ ND. MOTORS REO·D. _ _ _ ITEM ND. _ _ _ _ ND. TURBINEI REaD. _ _ _ ITEM NO. _ _ _ _ __

OPIIIAT_ CONDiTICIIII- EACH PUMP PEIlFORMANcE

LIQUIDISLURRY PERFORMANCE CURVE NO. _ _ _ _


MAX. _ _ _ RATED _ _ _ RPM _ _ _ _
PT."F NOIIM US GPM AT NORM _ _ _ NPSH IWATER' _ _ _
sP. GR." NORM PT. TOTAL HEAD, FT RATEO EFF. _ _ _ _ " 8HP RATEO _ _ _
VAP. PRESS." NORM PT. PSIA _ _ _ SUCT. PREss. PSIG MAX, _ _ _ RATEO _ _ _ MAX. 8HP RATED IMPELLER
VIS• ., NORM PT SSu NPSHA,FT MAX. HEAD RATED _ _ _
CORR.lEROS. CAUSED 8Y pH HYD. HP MAX. DISCH. PRESS. PSIG _ _ _
DRIVER HP TO 8E SELECTED FOR MAX. S.G. • MAX. VISCOSITY _ _ _ _ MIN. CONTINUOUS GPM _ _ _

_TESTS

o -.wIT. PERF. o WIT. PERF.


CDNaTIIucnDN - 0 ANSI 8n.1 o ANSI 873.2 o OTHER o NONWIT. HYDRO. o WIT. HYDRO.
o NONWIT. NPSH o WIT. NPSH
PUMP TYPE: 0 HORIZ. 0 VERT. IN·LINE 0 COUPLED MOTOR SHAFT 0 CRADLED MNT.
o o o NONWIT. VI8RATION o WIT. VIIR.
CASE HORIZONTAL MOUNT: FOOT
o
CENTER LINE
o o o DISMANTLE. INSPECT AFTER TEST
VERTICAL MOUNT:
sPLIT: o
RADIAL
MOTOR SHAFT
AXIALo
RIGID COUPLING
TYPE VOLUTE: o
OTHER
SINGLE ODOU8LE
o OTHER:
PRESS: o
MAX. ALLOW. _ _ _ PSIG _ _ _ "F o
HYDRO TEST _ _ _ PSIG
CONNECT: ODRAIN o GAGE SUCTION o GAGE DISCHARGE PUMP MATERIALS
IMPELLER DIA. RATED _ _ _ MAX. _ _ _ IMPELLER TYPE
BEARINGS TYPE: RADIAL THRUST CASING
LU8E: DOlL DOlL MIST GREASE o o
GREASE FOR LIFE IMPELLER
COUPLING: MFR. MODEL _ _ _ GUARD _ _ _ OILER _ _ _ WEAR RINGS
DRIVER HALF MTD. BY: o
PUMP MFR. DRIVER MFR. o PURCHASER o SHAFT/SLEEVE
STUFFING BOX COVER: STANDARDo JACKETED o SEAL ONLY o GLAND
PACKING: DMFR .• TYPE SIZE/NO. OF RINGS GASKETS
LANTERN RINGS: DYES ONO 8ASE PLATE
MECH.SEAL: DMFR.'MOOEL MATERIAL CODE COUPLING GUARD
oBALANCED o
UNBALANCED o
SINGLE OINSIOE OUTSIDE o OTHER:
OOOUBLE o
BACK TO BACK TANDEM oFACE TO FACE o
AUXlLlAIlY ....NG .IIE FIG. NO. _ _ _ _ FOIl elIDE. INII'ECTION o
NOT REQUIRED
DIN PROCESS o
FINAL
o STUF. BOX PLAN NO. o C. W. PIPING PLAN NO. _ DAYS NOTIFICATION REQUIRED
TOTAL COOLING WATER REO'D .. GPM o SIGHT F.I. REO'D.
o PACKING COOLING INJECTION REO'D.. TOTAL GPM PSIG BOUND _,F!CATION IIEOUIII-.nl
EXTERNAL SEAL fLUSH FLUID GPM PSIG
SEAL QUENCH PLAN SEAL QUENCH FLUID

..
MFR.
TYPE
DIlIVEII:

_--
o MOTOR o TURBINE o OTHER
RPM _ _ _ fRAME _ _ _ VOLTSIPHASE/HEIITZ
BEARINGS _ _ _ _ _ _ _ SERVICE FACTOR
INSULAnON
PROVIDED BY

_ : FL _ _ _ LA _ _ _
ADDITIONAL IIEOUIIIIMEN_INTI

LUBE TEMP. RISE "C _ _ ENCL


INLET PRESSURE _ _ EXHAUST PRESS. _ _ STlAM TEMP. _ _ WATER RATE _ _
OTHER

CUSTOMER/ USER
LOCATION
CUSTOMER P.O. NO.
ITEM NOIS •• EQUIP. NO.st.
FACTORY ORDER ND,S'. PUMP SERIAL NO'S'.
ISSUED BY DATE
REV. _ _ _ DATE

Fig.27.1 Data sheet listing information required to select and build a pump,
Procuring Centrifugal Pumps 803

3. Required capacity, as well as minimum and maximum amount of liquid pump must deliver
4. Suction conditions
a. Suction lift or suction head
b. Constant or variable suction conditions
c. Length and diameter of suction pipe, fittings and valves involved
5. Discharge conditions
a. Static head description-constant or variable
b. Friction head description and how estimated
c. Maximum and minimum discharge pressures against which pump must deliver liquid
6. Type of service-continuous or intermittent
7. Pump installation-horizontal or vertical position (if vertical, type of pit-wet or dry)
8. Type of driver and characteristics of power supply
9. Location of installation
10. Local environmental and safety requirements
11. Applicable industry standards
12. Special requirements or marked preferences with respect to pump design, construction, or performance
13. Protection
14. Space, weight, or transport limitations
15. Scope of supply

NUMBER OF UNITS

The number of units is important primarily to increase pump reliability; standby units are often necessary,
especially in cases where the life of the pump may be jeopardized by the severity of the service. It is
important to determine whether one or more units can be operated in parallel, as the hydraulic performance
of the individual units may need adaptation for this purpose.
The choice between the use of a single pump and the installation of several pumps in parallel for the
full demand is influenced by the expected load factor. When the demand is more or less constant, the
tendency is to select a pump for the full demand, adding a safety margin as provision against pump
wear. If, on the other hand, the load is of a fluctuating nature, two or more pumps may be operated in
parallel. At periods of light load, one or more pumps can be taken off the line for more efficient operation.
The capacity of the individual pump is selected with this type of operation in view. For instance, if the
demand remains at 65 percent of the maximum capacity most of the time, two pumps, each designed
for about 70 percent of maximum flow, could be installed. One pump alone would be able to carry the
load most of the time. When the total required flow exceeds 65 percent of maximum, the second pump
is put on the line to share the load.
Exceptions to this general practice are as follows:

When the total demand is too low to be divided efficiently between two pumps, a single pump may have to be
used, regardless of the nature of the load factor.
When the efficiency of operation is inconsequential-intermittent service or turbine-driven pumps used to provide
exhaust steam for process work or for heating-low first cost will generally determine the number of pumps to
be used in the installation.
When the maximum demand is too great to permit the use of the most economical pump or of the most economical
operating speed, the demand may be split up between two or more units, regardless of the nature of the load factor.
804 Procuring Centrifugal Pumps

If more than one pump is used to provide the maximum demand and if a partial reduction in available
capacity is permissible, the installation may not require a spare pump. Enough spare parts must then be
stocked to restore full available capacity rapidly. If, however, no reduction in available capacity is
permissible, a spare pump must be provided, which can be put into service immediately upon failure or
stoppage of one of the main pumps. Because this substitution must provide against the interruption of
the priine mover as well as for pump failure, a different source of power is frequently used to drive the
spare pump. Thus, if the normal operation is based on motor drive, the spare pump may be driven by
a steam turbine or an internal combustion engine. The switch over to the spare pump can be accomplished
automatically when the pressure in the discharge header decreases.

NATURE OF THE LIQUID

Type. To a certain extent the nature of the liquid pumped determines the types of pumps most frequently
used for the service involved. The widest range of selections is available for water service, but this must
be divided into fresh and salt water. The selection of materials for the saltwater service varies greatly
from the point of view of first cost, length of life, and customer preference and experience. Apparently
insignificant impurities in the liquid may be a highly important factor in the selection of the proper
materials. Similarly, the nature of the liquid to be pumped will greatly affect not only the pump material,
but possibly even the mechanical construction best suited to the purpose, depending on whether the
liquid is an acid, alkali, or oil. For example, if the pump is to handle highly corrosive liquids, not only
must the pump be built of corrosion-resisting material, but the acid must be prevented from leaking out
into the atmosphere through the shaft seal.
Temperature. The temperature of the pumped liquid is a very important factor. A standard line of
general-service pumps has definite temperature limitations. Higher temperatures may dictate the use of
special materials, additional cooling, or special mechanical features such as centerline support of the
casing. For extremely low temperatures, such as in brine service, a non-flammable liquid, the use of a
nickel cast iron may be warranted in preference to standard gray cast iron, in order to obtain a more
refined crystalline structure and thus prevent fractures.
Any wide variations in the operating temperatures should be known, as they will affect the specific
gravity and viscosity range of the liquid handled.
If the liquid is water, the vapor pressure can be easily determined from steam tables. If some other
liquid is involved, the vapor pressure at the pumping temperature must be carefully noted, because it
figures importantly in determining whether or not the prevailing suction conditions are satisfactory.
Specific gravity. The specific gravity must be known in order to determine the power consumption
at the design conditions and to select the proper driver size. A pump inquiry frequently expresses the
required discharge or net pressure in bar (pounds per square inch), which must be converted into meters
(feet) of the liquid handled. If a net pressure of 6.90 bar (100 psi) is specified, the total head is 74.4 m
(231 ft) of fresh cold water, 58.7 m (193 ft) of brine with a specific gravity of 1.2, or 98.9 m (308 ft)
of gasoline with a specific gravity of 0.75 (Fig. 18.2).
Viscosity. When the viscosity of the liquid handled differs from that of water, the pump capacity,
head, and power consumption are all affected appreciably, so that correction factors are necessary.
Solids content. The size and nature of solids suspended in the liquid will determine the type of impeller
and pump arrangement best suited for the application, and the materials of whi<;:h the pump will be
constructed. Low concentrations of abrasive solids are generally handled with an open impeller, and,
where necessary, special materials. Nonclogging impellers are required for large solids or solids of a
stringy nature. To pump high concentrations of abrasive solids, such as in mineral processing, slurry
Procuring Centrifugal Pumps 805

pumps are necessary, either rubber lined or hard metal depending on the size, shape, and concentration
of the solids.
Chemical analysis. Special consideration should be given to the chemical analysis of the liquid if its
corrosive or electrolytic properties are not readily apparent from the description of the liquid itself. Thus,
pH (measure of acidity or alkalinity) of the water should always be stated if indications show that the
water is not neutral.

REQUIRED CAPACITY

The capacity required for the installation should be stated in m3jhr (U.S. gallons per minute) at the
pumping temperature. Any expected variation in the range of capacities should be clearly stated, because
centrifugal pumps do not permit as much flexibility in capacity variations without affecting pump
efficiency as other types of pumps, such as reciprocating steam pumps. Furthermore, it is generally
preferable to have the maximum efficiency of the pump occur at or near the normal capacity conditions.
Any centrifugal pump can run occasionally at much more than its rated capacity, but this may not
always be practical or permissible. An increase in capacity means a decrease in the head generated; this
may prevent operation of the pump at emergency overloads if excess capacity was not included in the
design and if the pump operates on a system-head curve, since the friction losses which make up part
of the required head will increase with the capacity. Emergency over-capacity pumping may also be
prohibited if the prevailing suction conditions leave no margin over those required for the normal rated
capacity. Finally, if the power consumption increases with the capacity, as it does with the majority of
centrifugal pumps, operation at capacities greater than originally expected may seriously overload the
pump driver.
Minimum operating capacity information is also of great importance. In certain cases, operation at
extremely reduced capacities, even for very short periods, presents a definite danger and must be avoided.
In other instances, the only disadvantage from operating at reduced capacities is poor economy, and a
thorough analysis of the problem may result in the installation of additional smaller units that would be
operated during periods of light loads.

SUCTION CONDITIONS

Correct suction conditions for centrifugal pumps are of paramount importance (Fig. 27.2). Unless the
available net positive suction head (NPSH) is greater than that required by the pump selected at the
capacity involved, the pump will not be able to meet its design conditions. In addition, there is the risk
that the resulting cavitation will damage the pump. How much the available NPSH should be above that
required depends primarily on the liquid pumped, the size of the pump, and its impeller design (see
Chap. 19).
If cold liquids are handled, it is necessary to know whether there is head on the suction or the pump
will operate with a suction lift, and if the latter, what the maximum lift will be. If the liquid is hot or
under a pressure at or near its vapor pressure, the pump must be installed with head on the suction and
the available submergence must be described. In all cases, it is advisable to determine separately the
static difference between the liquid level and the pump centerline, and the friction and entrance losses
in the suction piping. If these losses have not been determined, it will generally suffice to describe
accurately the suction layout, listing all lengths, pipe sizes, and valves.
806 Procuring Centrifugal Pumps

Fig. 27.2 Suction lift curve.


Capacity and efficiency fall below normal when suction lift exceeds a given value.

DISCHARGE CONDITIONS

The discharge head for the design conditions should be stated with the understanding that it is generally
composed of static elevation (or pressure) and frictional losses in the discharge piping. Any variations
in the static head must be known in order to detennine the maximum and the minimum head against
which the pump is to operate (Fig. 27.3). Specifying an excessive total head has actually the same effect
as specifying excessive capacity. Since a centrifugal pump will always operate at the intersection of its
head-capacity curve and of the system-head curve, a pump which develops an excess of head will, unless
artificially throttled, deliver an excess in capacity as its curve will intersect the system-head curve at a
greater flow.
By breaking up the discharge head into static and friction head, excessive friction losses can be
detennined. If the piping to be used is too small, the required pump and its driver will be more expensive
than necessary and the cost of operation greater than if the proper size pipe were used.
When the total cost involved is high and justifies an extremely detailed analysis, it is possible to
detennine the most economic size of pipe by plotting the sum of the amortization of initial pump, driver,
and piping cost plus the operating cost against pipe sizes.
In certain special applications, the pump may be required to develop discharge pressures in excess
of the design conditions for short periods. A typical example is a condenser water circulating pump that
is to be installed in a system utilizing a siphon effect. The head to be developed consists only of the
frictional losses through the piping and through the condenser itself and is, therefore, fairly small.
However, to enable the condenser circulating pump to deliver water through the condenser, the system
must be primed, that is, water must be discharged over the siphon. The design of the pump, therefore,
must pennit it to develop a head at shutoff considerably higher than the nonnal operating head.
In some condenser centrifugal pump installations, the head required before the siphon is established
might so affect the required characteristic curve that the nonnal operating head occurs at a point of the
curve where the pump efficiency is relatively poor. In a siphon system, however, it is generally possible
to eliminate the higher starting head requirement by incorporating a vacuum-producing device. This
device is connected to the top of the siphon, so that the air at the high point can be evacuated and the
Procuring Centrifugal Pumps 807

90

80

70
~

60
z
I&j
u
a:
W
50 11.
~
U
Z
0 W
U
i&:
30 II.
W

80 'd 0
If
60 I I0

If
0
o 10 20 30 40
CAPACITY, IN 100 GPM

Fig. 27.3 Static head curve.


When static head changes considerably, pump may have to operate over wide capacity range.

siphon established with the pump producing a head no greater than that required for normal operation.
In many cases, a little study or the installation of relatively inexpensive additional equipment will often
simplify a problem of this nature.

TYPE OF SERVICE

The expected load factor of the contemplated installation will play an important part in the selection of
the best pump to be used. As described previously, the type of service will affect the number of units
that will best meet the capacity requirements. If the service is intermittent, it is unnecessary to use the
most efficient pump available; the selection is generally made on the basis of lowest first cost. A pump
intended for continuous service, however, should be selected for efficiency, reliability, and long life.
808 Procuring Centrifugal Pumps

POSITION OF INSTALLATION

Whereas the majority of centrifugal pumps are horizontal units, conditions occur to make a centrifugal
pump with a vertical axis of rotation more suitable. Axial-flow propeller pumps and certain mixed-flow
type pumps are frequently arranged for vertical operation. This results in a very compact installation
and permits submerging the impeller so that the pump is always primed. At the same time, the motor
driver can be placed above the pump at a level sufficiently high to avoid accidental damage to the driver
if the water level rises excessively. Large and small sewage pumps are frequently installed in this manner.
Vertical pumps can be arranged either for wet- or dry-pit installations (Figs. 14.1 and 14.27). Since
the pumps used for these two categories are quite different, it is very important to determine which of
the two installations is preferable.

DRIVER TYPE AND POWER SUPPLY

Although centrifugal pumps are usually driven by electric motors, steam turbines, internal combustion
engines, or gas turbines, other prime movers and intermediate transmissions are used. A centrifugal
pump, its drive, and the method of operation must be compatible for the pump to perform as expected.
It is therefore important to specify the type of driver intended, the direction of rotation if unidirectional,
and where applicable, the power supply and any fluctuations that must be taken into account in sizing
the driver (e.g., reduced voltage for electric motors). Centrifugal pump drivers are described in detail
in Chapter 24.

LOCATION OF INSTALLATION

The geographical location of the installation has a great bearing on proper pump selection and on
its maintenance.
The elevation above sea level affects the pump, as there is a decrease in atmospheric pressure of
about 25 rom (1 in) of mercury per 300 m (1,000 ft) of elevation. At an elevation of 1,220 m (4,000
ft), therefore, the atmospheric pressure is 100 m (4 in) mercury or about 1.4 m (4.5 ft) less than at sea
level, which means that a centrifugal pump can handle 1.4 m (4.5 ft) less suction lift than at sea level.
The geographical location must also be taken into consideration when recommending spare parts, as
a pump that is to operate in an out-of-the way place should be supplied with sufficient spares to avoid
interruption of service if parts become worn and cannot be readily replaced.
The immediate surroundings of a pump will affect its accessibility after installation. A pump located
in a cramped, dirty and wet, or poorly lighted position will be neglected by the operators, will not give
satisfactory service, and will be difficult to inspect, dismantle, and repair.

LOCAL ENVIRONMENTAL AND SAFETY REQUIREMENTS

In many locations around the world, there are now in place stringent requirements dealing with the
allowable noise and emissions from equipment such as centrifugal pumps. At the same time, the prevailing
authorities may have safety codes dealing with such matters as electrical apparatus and the guarding of
rotating equipment that have to be complied with. If these requirements are to be met, they must be set
out specifically in the inquiry. Resorting to a catchall clause such as, "the equipment shall comply with
Procuring Centrifugal Pumps 809

all applicable local, state, and federal rules and regulations," may be legally tidy but will not ensure that
the equipment built and shipped to the site meets the requirements.

APPLICABLE INDUSTRY STANDARDS

For many classes of service (e.g., chemical and petroleum refining), industry standards exist that cover
the generally agreed requirements of centrifugal pumps for those services. Their use saves a great deal
of effort on the part of the purchaser in specifying pumps, and does the same on the part of the
manufacturer in engineering and fabricating the pumps. The latter is so because each manufacturer has
available designs and production engineering for pumps that meet these standards. A direct result of
preengineered pumps is lower cost, therefore it is to the purchaser's advantage to use pumps to an
industry standard wherever they are suitable. And in those cases where special features are required,
these should be issued as a supplement to the applicable industry standard so the pump manufacturer
can easily see what is special and the extent of the changes. See Chapter 26 for the industry standard
applicable to various centrifugal pump services.

SPECIAL REQUIREMENTS

The pump manufacturer must become familiar with any special requirements and preferences of the
personnel who are going to operate the centrifugal pumping equipment. Some of these preferences may
be based on insufficient knowledge of modern practices, or they may originate from experience with
pumps operating under the same conditions the new pumps will have to meet. This information may be
valuable to help the pump manufacturer in furnishing equipment that will give the longest and most
reliable service.
Customer recommendations, however, should be limited to experience with pumps operating under
similar conditions, rather than a list of his preferences, which may result in the purchase of very special
equipment. The manufacturer's standard construction is preferable to specially built units, both from the
point of view of the original cost and of obtaining repair parts later.
It is also necessary to take the life expectancy of the installation into consideration. If the design and
materials of construction are selected for a life span much in excess of the life expectancy of the process
or installation for which the equipment is intended, the original cost of the equipment will be out of
proportion. In other words, it is a short-sighted policy to purchase equipment with a 12-month life for
an installation designed to last 15 years; it is just as wasteful to buy a pump to last 20 years if only a
temporary 6-month use is expected.

PROTECTION

Although the location of the plant provides a clue to the degree of special protection needed for
various components to operate correctly (e.g., tropical proofing of electric motors for service in tropical
environments), there is more to this subject than operation. The first need, one often critical to the
successful startup of the plant, is the degree of protection necessary to survive transport to the plant site,
site storage until installed, and the period between installation and startup. To be sure the pumps are
operable when needed, the inquiry should state means of shipment, period and type of site storage, and
means of protection after installation.
810 Procuring Centrifugal Pumps

SPACE, WEIGHT, AND TRANSPORTATION LIMITATIONS

Sometimes pumps must be installed in very cramped quarters. Vertical pumps may be preferred in such
cases, as they need only a small fraction of the floor space required by a horizontal pump of equal
capacity. A pump operating at the highest speed compatible with the conditions of service will also
reduce space requirements. Although the use of close-coupled pumps (Fig. 1.19) was introduced primarily
from consideration of first-cost economy, the application of such pumps presents definite space-saving
advantages. Finally, in a number of cases, the use of horizontal pumps with bottom suction (Fig. 2.26)
may simplify considerably the problem of spacing and suction piping accommodation.
The weight of a pump usually does not matter, but aboard ships, for instance, light weight is a deciding
factor. For this reason, it is the practice in navy and merchant marine installations to use extremely high-
speed units of special design, operating at speeds up to 10,000 rpm (generally turbine-driven).
The facilities available for transportation should be outlined if there is anything unusual about them.
Sometimes it is a long journey from the factory where a pump is built to the place where it is needed
and the transportation facilities may be poor. For transportation over such routes, lightweight pumps
must be supplied, preferably of sectionalized construction.
In other cases a pump must be brought to its ultimate destination through tunnels or shafts of limited
dimensions, and the pump selection must be made with such restrictions in mind. All such limitations
must be carefully listed, or a fully built pump transported part way may prove to be undeliverable.

SCOPE OF SUPPLY

With the increasing sophistication of the plants and processes that use centrifugal pumps, the pumps
themselves have become more complex, often incorporating quite elaborate auxiliary systems for shaft
sealing, lubrication, monitoring, and controlling operation. To avoid items being missed, which may
delay installation and consequently plant startup, it is necessary to clearly specify the scope of supply.
Typically this will cover drive train components, couplings, base, piping and valves, and the cited
auxiliary systems. For clarity, it is better to specify the terminal points of piping and systems the customer
has to connect to by the use of a diagram (Fig. 27.4) rather than relying on words.
Although not technical in nature, the proposed commercial terms of a purchase order can have a
significant effect on the supply of equipment. It is therefore necessary that the inquiry include the fol-
lowing:

1. Delivery required
2. Tenns of payment and escalation if applicable
3. Perfonnance guarantees
4. Warranty required
5. Penalties and reciprocal bonuses
6. Default provisions

Inquiries for Specific Services


A knowledge of the factors outlined above is a prerequisite to an intelligent analysis of the problem
involved and to its solution. The factors are not equally important, however, for all the various services
tiMES :
1. LEGEND
E t><I : GAH VALVE
: CLOlt VIII"vl
, .. M
N : CHICIt V."VI
I I m ITlNt 7IUIP
IQJ I SIOIn' OLUI
I ., '""I I7IUIINIR
YafDOR
~~f.! :! ~
I>-- PLUG POll IAJIf POI"" IIMI)!

'( : DRIP rvNNtL


~ DrIOlA.. I i!
LOCAL MOIJIWD PUSIUlI
• TO DRAIN 1:ASING vt'" t ® CAUG£
PUIICIWID LOCAL MOUNI'ID tlMPDATVIIl
0""'" ~ @ OlllUG£
~ Ml:ClWlJCAL 'rACllCltftD
~r-'J. ®
TI Z. llOLAftCli n.uoa NIl MDUtMD ....
W£LDID CDlNECTIOlll MI utD.
IUC'I'IOM
:to IF tim ..... II MDI 11"-¥lNtlNO,
~a!v;;;;- ~
§ "HI CAlIIIO ftM II ..". RIOUI. . . .
•• CDIIIIIe!'ICIIt POll .... CASI........'"
I
IJIG DMl'" LIQUID .HALL • • • U....IID n VIllDOll,
.BtI .PlCIrIID.
I, POll 'nI .,..l1li lIP eul~.&.......~ 'IMP
II IllCUIAItY ONLY . . . . ~"IC
loP CASINO DRAIN I"AM' II IPICIF11D.
tJ I. CIOVI..- llIAIoIo • IIOCID-w" OIL
lilLA' 'tIPlL.~' cmllWll.
AGIIIID . , ~D.
y
aD COOLI.. _fElt ..",..,
I~
I! CIIOIdIlO WAt'tI M'I'UIII
I~
nTSIUIAlo P'UmlI"

OUDICII ItEM (.."


QC
......
• VDIDCII I I'U"CHMD
Fig.27.4 Piping and instrument diagram (P&ID) showing terminal points.
812 Procuring Centrifugal Pumps

under which centrifugal pump applications can be classified. Some representative services most commonly
encountered and the most important points to be examined for each of these are described next.
As most pumps are constant-speed motor-driven, these lists have been prepared for this type of drive,
with one exception. If a variable-speed drive is used, the operating speed can be adjusted to allow the
unit to cover a much wider range of conditions than is feasible with a constant-speed drive. Except for
the one item, namely, turbocentrifugal pumping units, these lists do not describe information needed to
make driver selections. If drivers are to be purchased with the pumps, full information as to type to be
furnished must be included in the inquiry.

WATER WORKS SERVICES

Raw Water-Source to Treatment Plant

1. Operating head range. Indicated in the form of system-head curves for minimum, normal, and maximum
static heads and supplemented by information as to the duration or importance of minimum and maximum
head pumpage.
2. Capacity rating at design head. When local conditions require obtaining certain variations in capacities with
variation in total head, it is advisable, before issuing a formal inquiry, to determine from one or more
manufacturers, if the desired characteristics can be obtained with constant speed. If not, it may be necessary
to use a variable-speed drive.
Centrifugal pump guarantees are limited to one head-capacity condition with very few exceptions, such
as large-capacity, specially built units. Even in such cases th!! manufacturers expect a reasonable tolerance
on head or capacity for any secondary guarantee point.
3. Suction conditions. Detailed information how these are interrelated with total head or capacity.
4. Character of water. Dissolved air or gases and foreign material such as grit or leaves; also chemical
contamination, if any.
5. Type of installation. Horizontal or vertical with full data on local conditions.

Treated Water-Direct Service Into Mains

1. Rated capacity at design head with data on operating range of capacity either alone or in parallel with other
units. If the new unit or units are to operate in parallel with existing units, the characteristics of these existing
units should be given.
2. Rated head.
3. Suction conditions.
4. Type of installation.-horizontal or vertical.

Treated Water-Direct Service Into Mains With Stand


Pipe or Reservoir on Line

1. Capacity at design head. Information as to desired capacity at other heads if head varies with rate of pumpage
or for other reasons.
2. Rated head. Head variation data, if any, described in detail.
3. Suction conditions.
4. Type of installation.
Procuring Centrifugal Pumps 813

Treated Water-Long-Distance Pumpage Through


Pipeline to Distribution System or Reservoir
Such systems necessarily involve considerable friction heads which vary with capacity. Unless pump-
age is to be at a constant rate to a distribution reservoir, full data on the system should be given with
information as to what is to be accomplished.

Treated Water-Booster Services


Pumps for booster service may either take water from mains under nearly constant pressure and
discharge it into a distribution system in which a higher pressure is needed or may take water from
mains when the pressure falls below the minimum permitted and add more pressure for the area served
by the pump. Without a stand pipe on the system, pumps for the first service should generally have a
reasonably small head rise so that excess pressure is kept to a minimum. Depending on local conditions
and the inclusion of a standpipe, pumps intended for the second type of installation may require anything
between a reasonably flat and a very steep head-capacity characteristic. The purpose, therefore, must be
clearly outlined in detail if the manufacturer is to make the proper selection of pumps to be used for
booster service.

Treated Water-Thrbocentrifugal Pumping Units


Turbocentrifugal pumping units, that is, steam turbine, reduction gear, centrifugal pump, circulating
pump, condenser, condensate pump with interconnecting piping and auxiliaries, involve various items
that must be carefully selected individually to make a proper operating whole. This type of unit is seldom
purchased, and the manufacturer must be given very complete information for this type of unit, both
for the hydraulic data and the physical location in which the unit will be installed. The data needed is
summarized as follows:

1. Rated capacity and head. Range of capacity and heads to be met with data on suction conditions for
these ranges.
2. Steam pressure and superheat. Range in variation for both.
3. Water temperature range. Temperature specification at which duty guarantees are to be made.
4. Discharge head for condensate pump.
5. Drawing showing space available for unit, location of basement, operating floor, source of suction supply
or location of suction main, and point to which discharge is to be connected.

SEWAGE

1. Operating head range. Whenever two or more pumps are involved, a curve showing the system head-capacity
characteristics. For operating head range specification, reduction in friction head losses with decreased
capacity, resulting from an increased static head and the reverse with increased capacity, resulting from a
decreased static head must be considered.
2. Capacity at average or design head with limitations, if any, at other heads. Unnecessary restrictions of
capacities at other than design heads may require special designs with unnecessary high cost.
3. Suction conditions. Full information on how they vary with total head, capacity, or number of units in service.
4. Type of pump installation-horizontal or vertical If vertical in dry pit with shafting between pump and
motor or gear, elevation of pump pit floor, centerline of suction (if fixed by existing construction), and motor
supporting floor must be described. If the vertical distance is such that steady bearings may be required for
814 Procuring Centrifugal Pumps

the vertical shafting, location of bearing supporting beams or of floors, if fixed by some local conditions,
must be described. For vertical wet-pit pumps, distance from supporting floor to pit bottom must be stated.
This, with information as to water levels in the pit, will permit the manufacturer to select proper length of pump.
5. Size of solids. The maximum size of solids to be expected in straight domestic sewage is 65 mm (2.5 in.)
in diameter, as could be flushed down a water closet. Unless the sewage is screened or comminuted, it is
desireable that pumps on straight run sewage service be capable of passing 65 mm (2.5 in.) spheres. On
storm water or combined domestic and storm water systems, larger solids can be expected. It is usual now
to specify that the large pumps generally used for such applications be able to pass alSO mm (6 in.) sphere.
The large passages necessary in the impellers of these pumps compromizes their hydraulic performance and
rangeability to some extent.
The specific information required for drainage and irrigation installations is generally the same as for sewage instal-
lations.

BOILER FEED SERVICE

1. Required capacity in cubic meters per hour (gallons per minute) or in kilograms per second (pounds per
hour), including maximum, minimum, and normal. If expressed in cubic meters per hour (gallons per
minute), the conversion from kilograms per second (pounds per hour) should include density (specific
gravity) corrections. The total pump capacity installed, exclusive of spare equipment, should exceed the
maximum steaming capacity of the boilers by 8 to 20 percent. This margin is intended to provide against
boiler swings as well as against the eventual reduction of pump capacity through wear before internal parts
are renewed.
2. Required intermediate flow or bleed-off flows for reheater and superheater attemperation. These flows to
be expressed in the same units as the pump's suction capacity (item 1 above).
3. Temperature of the feedwater and possible variations.
4. Suction conditions. Variations in suction pressure and information on the minimum available NPSH. If the
pump takes its suction from a direct contact feedwater heater, the NPSH (available energy over and above
the vapor pressure) is the static submergence between the water level in the storage space and the pump
centerline, less the friction losses in the suction piping. If, however, the pump takes its suction from the
discharge of some other pump such as the condensate or a booster pump with closed heaters located
between them, the available NPSH is the difference between the suction pressure and the vapor pressure
at pumping temperature, converted into meters (feet) of water at the pumping temperature.
5. Discharge pressure. If not available, boiler pressure, as this will permit estimation of the required discharge
pressure. Experience shows that the required discharge pressure will vary from 115 to 125 percent of the
boiler pressure. For a more accurate determination, it is necessary to know the maximum drum pressure,
the static elevation of the drum above the pump floor, the friction losses in the discharge piping and in
the closed heaters downstream of the pump, and the losses across the feedwater regulator.
6. Pressure at the intermediate flow and reheat and superheat attemperation bleed-offs, if required (see item
2 above).
7. Chemical analysis of the feedwater. Description of the pH value, conductivity, and dissolved oxygen content
if the analysis is not complete. Information, if available, on the past experience with boiler feed pump
materials handling the same feedwater.
8. Type of shaft seal and means of cooling.
9. Type of regUlation.
10. A complete layout of the feedwater system and heat balance should be supplied to the pump manufacturer,
if possible.
Procuring Centrifugal Pumps 815

CONDENSATE SERVICE

1. Capacity in cubic meters per hour (gallons per minute) or in kilograms per second (pounds per hour),
maximum, minimum, and normal. If condensate pumps discharge through some auxiliary cooling heat
exchangers, such as air ejector aftercondensers or oil coolers, specify the minimum flow required as this
may necessitate the inclusion of automatic recirculation controls.
2. Suction conditions. Vacuum in hotwell, minimum available submergence, and expected variations.
3. Discharge pressure. Maximum and normal. If prevailing conditions are complex, supply a system-head curve.
4. Type of system--open or closed. In other words, does the condensate pump discharge into a direct contact
heater or surge tank or does it deliver the condensate into the suction of a boiler feed pump?
5. Type of control contemplated. Will the capacity be automatically regulated by variations in hotwell level
through operation in the break, will the discharge be throttled by means of a float-regulated valve, or will
the float control regulate the bypassing of a certain quantity of condensate back into the condenser hotwell?

CONDENSER CIRCULATING SERVICE

1. Rated capacity. In case of parallel operation of several or more units, desired capacity with fewer units
in service.
2. Rated head. In case of parallel operation of several or more units, detailed information so that heads at
reduced flows can be determined.
3. Suction conditions. Detailed information if conditions will result in increased capacity at reduced head oper-
ation.
4. If the pump will operate on a siphon system, the method of establishing the siphon is either to evacuate the
entrapped air or to fill the siphon with the pump. If the pump must fill the siphon, the static head from the
suction supply to the top of the siphon must be specified. (Usually it is better to establish the siphon by
evacuating the entrapped air.)
5. Character of water. Information as to materials of construction found suitable for existing units.
6. Type of installation. Horizontal or vertical. If the latter, whether wet or dry pit with information as to
local conditions.

INDUSTRIAL WATER SUPPLY

The requirements for water supply to industrial plants vary widely with small to large capacities and
low to very high heads handling all types of water-fresh, salt, brackish, often contaminated with
chemicals. The data described earlier under water works services and condenser circulating service will
serve, generally, as a proper guide.

CHEMICAL SERVICE

1. The nature of the liquid. Acid or alkaline, concentration of the solution, and impurities present in the liquid
to be handled, if any. This last item is of paramount importance, as experience has shown that presence of
various impurities has a marked effect on the relative resistance to corrosion of various pump materials.
816 Procuring Centrifugal Pumps

2. Amount of solids in suspension and their abrasive qUality.


3. Pumping temperature and vapor pressure or boiling point. Possible or expected variation.
4. Specific gravity and viscosity at pumping temperature.
S. Capacity. Maximum, minimum, and normal.
6. Discharge conditions.
7. Suction conditions. Method used to prime pump if there is a suction lift. Possible change in pump location
to arrange for operation under submergence if suction lift is impractical or to reduce positive suction head
if sealing stuffing box is impractical.
8. Type, configuration, and materials of shaft seal. Quench fluid if required for single seal; buffer or barrier
fluid respectively if seal is tandem or double.
9. Past experience with various materials or combination of materials handling this liquid. In many cases,
dissimilar materials of the reservoir from which the liquid is drawn and of the pump itself set up a galvanic
action which may be harmful to one or the other material, thereby requiring proper isolating precautions.

PETROLEUM SERVICES

General and Cold Oil


1. Nature and complete description of liquid to be handled. Specific gravity over range of pumping temperatures.
2. Capacity. Maximum, minimum, and normal.
3. Suction conditions with special consideration to NPSH.
4. Discharge conditions.
S. Viscosity and variations if any.
6. Corrosive and abrasive properties of liquid.
7. Type of service relative to pump control.

Hot Oil
The information listed under general and cold oil, plus the following:

8. Range of pumping temperatures.


9. Vapor pressure at pumping temperature.
10. Correction of capacity on hot-to-cold operating range. Resulting effect on power consumption with reference
to possible driver overload.
11. Type, configuration, and materials of shaft seal. Quench fluid if required for single seal; buffer or barrier
fluid respectively if seal is tandem or double.

Volatile Fractions
The information needed for ethanes, propanes, butanes, gasolines, kerosenes, and other light fractions
is the same as listed under general and cold oil, plus the following:

8. Range of pumping temperatures.


9. Vapor pressure at pumping temperature.
10. Shaft seal requirements as described earlier.
11. Local practice with reference to protection against fire hazards (such as firewalls and special driver enclo-
sures).
Procuring Centrifugal Pumps 817

Pipeline Service
The information listed under general and cold oil, plus the following:

8. Graphical representation of required pressure gradients for entire pipeline.


9. Possible variations in capacity or in the nature of products handled.
10. Method of operation-manual or automatic control
11. Shaft seal requirements as described earlier.
12. Local practice with reference to fire hazard.
13. Local preference for series, single, or multiple units.

PROCESS AND PAPER PULP SERVICE

The information needed for sugar, brine, brewery, and paper stock service is as follows:

1. The nature of the liquid.


2. Acid or alkaline.
3. Amount of solids in suspension carried, if any. How solids are carried-floating or through agitation and
flow. Characteristics of the solids-inert or subject to dewatering or swelling. Consistency of the solids
and past experience in pumping the liquid.
4. Pumping temperatures.
5. Specific gravity and viscosity at pumping temperature.
6. Capacity. Maximum, minimum, and normal.
7. Discharge conditions.
8. Suction conditions.
9. Sanitary precautions to be observed. Local preferences for specific details or materials of construction.

MINE DEWATERING

1. Nature of mine water. Chemical analysis, presence or absence of suspended materials, abrasive quality of
the solids in suspension.
2. Capacity requirements.
3. Discharge pressure. Static head and friction losses.
4. Suction conditions.
5. Location of installation. Importance of space factor. Size limitations imposed by transportation through the
mine shaft.
6. Special considerations for the drivers. Description of special insulation or enclosure necessary for the
electric motors, if any.
7. Previous experience with centrifugal pumps handling this mine water. What materials stood up best, which
ones proved unsuitable?

HYDRAULIC PRESSURE SERVICE

1. Liquid used. Water or oil.


2. Capacity. Maximum, minimum, and normal. Rate of capacity change. Rapid fluctuations may require
special precautions.
818 Procuring Centrifugal Pumps

3. Total head conditions. Static head, friction losses, pressures to be maintained.


4. Suction conditions.
5. Type of pressure system. Method of pump discharge-directly to press or nozzles, or through an accumulator.
A sketch of the installation is very helpful.
6. Type of control. Characteristics of accumulator, if used. Maintains a constant pressure automatically and
pump is shut down during the no-load operation, or pump operates at no flow. If the latter, arrangements
used to protect pump against overheating, that is, control of bypass recirculation.

SLURRY PUMPING

1. Nature of slurry. Carrying liquid, temperature, and SG; concentration and type of solids, size distribution,
and SG.
2. Required flow.
3. Discharge pressure. Static head and friction losses.
4. Suction conditions.
5. Pumping arrangement-single, in parallel, or in series. If in series, whether grouped or spread along the line.
6. Materials of construction. If any past experience handling the slurry, provide details of materials used
and results.
7. Type of shaft seal.
8. Drive arrangement and any special considerations for driver construction.

Evaluating bids
Bid evaluation typically takes one of three courses. All conclude by making a selection based on lowest
evaluated cost, but the course taken by each to develop this cost is quite different. In making the
evaluation, all use a tabulation to allow easy comparison of each manufacturer's offer (Fig. 27.5). The
arrangement of the purchaser's data sheet usually anticipates the bid tabulation. And in some cases
today, the transmission of data between manufacturer and purchaser is taking place by electronic data
interchange (EDI) , using neutral files which allow each party to display the data in the format most
convenient to them.
The simplest course in evaluating any set of bids is to determine whether all meet the specification,
upgrade or reject any that do not, a process known as "bid conditioning," then make the selection based
on purchase price alone. For pumps mass-produced to an established industry specification, this is an
acceptable approach provided past experience shows that the specification ensures pumps of acceptable
quality and durability, and that manufacturer support is not diluted by having many brands of the same
class of pump in one plant.
For engineered pumps, those whose basic design has been tailored to the particular application, it is
prudent to compare the fundamental aspects of the pumps offered. Accordingly, in addition to "bid
conditioning," the evaluation should also compare at least the following:

1. Pressure rating versus operating pressure, both at pumping temperature.


2. Performance characteristics: NPSH margin at rated and specified maximum capacity (note that a large margin
may indicate high S; see Chaps. 19 and 22), head rise to minimum allowable flow, power at rated and
maximum specified capacity.
0 /ssU/50 ,c:OR CJ.lIiNT APPROVAL ~ J<)Er. ;IX" YAY. .&:fP' 51". X'O(
II!V. D!lClll'Tl0N INI 011 CHK SU~ .~TL APPRovaLS DATI
~ NOTES: ENC1~Cl£D ITEM$ ARE UNDESIRAllE.
A CHECK , INDICATES COM'LIANCE WITH S'ECt
~
~
~ I 2 3 4 S 6
~ IIMU'ACTUItEIt M-M PRC$PcR F. ElJRliKA
I:> '-UXO/i'IISS0C. I!
" _ 0' .100[0 " "
'ROPOSAL NO. A/·3,-,.,73 t.A 6:14 P437~-S,c
.-
~
X
DATE: 0' 'ROPOSAl 1- '../·7/ /I-Ci'-7/ 11-/~'7/
-< "'.·5. III)OEL NO. 4X1..3"1H 16-3':<NR 3.;J.-//C/XI SPEC
~ N T,"E 75T6. 'lSTG. 6STG.
-Q..:J::::l ~ MeTOl'? M,4/o1UFACTURlER eI..EC.PRo/).
s.§~ ~ C/I ApO/T/ON-S
--
'" " N
LJ t?AlE SET t"'r A<.-r£5~o.R/£:; 11177" REI¥ 'D 1/oI(!/... . NtJT R£(j'CJ REt) '[)
0 ~ ~i C/oiE 5£T ~'pAIi'£ ,PARTS lAIC/.. __ til,
550. - /A/t.°L. .,
(")~&-...I
g iJ. ~~ ON£" S£T .:;-P/.. :.·/Al.. TOOL.:::; Npr PEQ'o +/80 I /O/PTI?EQ'D NO .
Q."'" (, C;-<
~ ~ .,
EREe/ln;; .:-/IPEPl··"':;/O// '''':;.540 -/,/30 ..~54o(3;
~ ...... cn fTl l:. tll 0
FEATURES (PlRIlNIT) If/OPAY:S- ~j),JjV5)\1) (/00"''''5) BID
~ ~ 8 :::0 '\ ~ ~ --
""1 VOl .sAV/I/G~ 7'0 el,..IMINA7£ VP -:J./J?u:,r 3.4-40';>
< I?t:QV/R,-b <3,6':>0> ALT.QIIOTE'
::;:t'1']"g. SAVING:; !4-'/7,,,:./ ::","0.1..'7"£1.. 8A~f'/icL
0 <':/i330> <2840> C~
-.:e.. 0 0
C/I ~';''''' FO{" HY'.o, 7£S;/N6 INC~ .,.;JOO INet...
;:; ~: g: 0 '\J - " - - ', INS7"/)Ll. COIIPL/I/r; ,/ 0/ 1-3:10
<:> ;:s 0-
~ ,/ +350 0 IN<:.t...
~ ~ " " TEMP. ;;;Vc..'"r/oA/ S'rRAll/eR
:... s " 1/·/3Cr Jsr:; TG. I~\#..c:: ,/ flSOO (:9 INC/..
1:>'Og" n 11-13 Cr .s,j~~ 'TIOI.I BeLL +3.140 OeCLINeO I-:/,~O
~. ~ [ ~ ~ "
Z TO BID
~ ~ g' ENEF'V PENALTY @ "/~7/.HP~I?I/N" +/.:> 650 BAS~ ""/4.:17
~.
-< -- -- -- --- --
~ TOTAL
." API)I TlOt/5 (PEtlALTY (rwo UMrs -.33,10 a ,8. 0'70 t/8.374
~
IJI C ' •• Cl Fe.:: S#IP~/)'/6 .001/.:0:- (TWO) #':;093.;1S iI';;3'9760 $,:2/4 78~
~ :);: ~ AoolTlONS AS ITEMIZED MOVE ~.3..IOO 8090 /_~374 ~
'"1 COMPARATIVE COST AT SOURCE ,)4~ 4;s :N7,8'::D ':>33/$6
..
"-
~.
;:s.. C)
~~ ><.
I

X.
X.
TRANSPORTATION COST(£STIM.oI1T£D) •
COMP'AIIATIV£ COST AT DESTINATION •
5.430
A7,8C"!:"
6,580
';;S44'ln
7~40
.:J403<P~
-{ 8IDDE.R.
To .E S"r,
~.
'1::s f"\)
~ TOTAL. Cl)ST ! 2 UNITS) ;)4713S'::- ';"S44-,-0 ~4o. 3'Yre.
"- ~
SHU"ING 'OINT 5£R!?A UMPQA LAII£VleW
" 0J
~
1:;' SHIPPING ,",IGMT .LrOTA/"liJCt....PRIVI!Ii"S 137 <",,/.IT 4:1 ......~/".# 131..:70,;) ., '?
'" DELIVERY. ·/Uo. ,;FTER PO /:lo 14 I~ ;2- ""

;:s u~ .... ''''' ;1/0#£ NOI.'£. AIr.'.", ,&"/RM ~
'"1:1
,. R[COINENG: Fi-
0 ~ RUlON L.OW€ST EVALUA7£D "RIC~ MEETS ::;1"5.;. Dt::t...IV/iP·- ,-/."!"""!cIC,.e"-D' 0-
'ZZ)"", r:"f;".~ ~ '-'F DEC. 3 . C;J L.ETr£R OF e·£"" 8 . (2)/..ETTE.R .,?,c Nov.:l9 ~lJOtIBtE/)I"OIi'EVA"14TIOI/Ct:JMPAR/SP/o/
~'"
'"

QC
...
\C
820 Procuring Centrifugal Pumps

~P!
><~ NlTM IWSP£R FL'/ SPEC
PERFORMANCE
~
~. i RATED t:LOW
EFFleIENe't'__@ IDqOO GPM
D£5IGM FLC;V
GPM

GPM
0/
7'.5%
~

83·0%
.,.
7.3.0%
.r
0'?a?GPNI
-
B400GPU
~i
~ v'

..
DESIGN ','f':I/P, op ~ 0/ ...-- /I.:l0~
SPEC- GRAVITY @l I/:l o~ O.qql 0/ 0·q91 -
:c DIF~ !-IEAD@ 8400 GPM
'" 1080'TDH

~ ./
PRESS. @ MIN. FLoW PSIA 710 670@ {,IO (i,75PSIA
MIN. coN,. FLOW GPM BoO 1700 Boo
~~ FLOW@ RuN CuT GPM 10,000 <l40o I/'SOO 9400GPM
I./PS# RcQ. @ 9400 GPM 15' ;11.4' /5'

~! NPSH REG: @ DESIGN (8400GPM. 14.5'


B;.IP @., 900GPM HP ;1700
/7.6'
;1600(::>')
14.S'
,;165'0 -
BHP @ <1400GPM ;.IP :;<100 .;;.880 3100
SPeED RPM /180 1180 /180 ~/I':>O<:>
"'
0
IS' CRITICAL SPeED RPM NS >9500 ;tI5 -
AlP 0;:' STAc-i£!> 7 q 6 -
~! WeIGHT R07. PARr LBS
REV£R5E FLOW RPM@IOO%IIGAO <1,)5%
1-/5 3700
I~SORPM
3100
<-I.;JS?(
-
-

D£S/GAI
SUCTiON: SIZE /P'':.,4/t1GE RATING .;14,,/1$0# 30,,/150# ,)4 "/!so' ;14',oREF.
/)lscH4li'GE: S'I.ZE/AAI'/(;E RAT/~ /8-/.300- 1~'/.300'" /6"/-*'0.# /6 N..AI"E,c.
DMM(O.o.) OF SHELL. 4;1# 44# 4,;1'" -
LIiNGTI-I 0;:' SH£/..L. 18!6'" I ::I;1'-Sy';''' ,;1o'-B'"
-l
WEIGHT;!)F PUMP!'c,"lICH) LaS 43,000 46,SOO .fQooo -
~ We/GIlT OF MoroR J.BS :>.6.00" ,;1.3.0t:70 .:J.6.ooo -
~VCTIO,/tl: D£S/6,/t1 P""£ss PSI /.S'x.+W',su:r. .so ~sxAM/f.S'Ut:7: -
~,'I: PlSCIIANG£: /)£S/~N "'''''£5S. PSI
Sl/C770N: reST PRt£;SS. PSI
/. S xS/I(/r-()~
/.SrAMr..fI/~r.
b50
75
t$XSIIVT-oFF
/.S~.fIk'T.
-
-
D/$CHARG£:T£Sr PIi'£SS. PSI /.S • .sNIIr·orF 97S /.5xSllur-aFF ~MAX.S/KII'.(Yj
I-
~
~
"
\l

to 0
o AS PER LETTER, MARCil ~/"70; INcl?e~s£ IN AVAILA8LE SUCTION
tiEl/I> S l" 8 J!7: ENA81..ES NTllfro SlIoRTcAi U;NGHr OF 8~""R£L.
~~
~~I @ AO./LlST€D ,0
FROM ,#1'-6" ro /8'-6": .
p.x,>osPtE/i? LtErreR .IIA72C/l6, /970
NOTES: ENe I RClEO ITE!.IS ARE UNOES I RABlE
'~I A CHECK'; INDICATES COMf'lIAI(CE WlTtI SPECS
o~,
::1

I SW.:~.::~f!Y 0:= eIDS .!c:; ___ ._


:;;L]{"~x..)(.
fr.i:';'~-
P.O.NO.
CONDENSATE PuMPS
M-23 0
XYZ POWER COMPANY SHT.2 0~4

Fig.27.5 Continued
Procuring Centrifugal Pumps 821

~.

"'><.S.. AlIT"" PROSPER ':/-.1 SPEC


~' C,ISE :NIT. OF svcno.v 8£'-'.. 1/-/!JCrS7 C.I.U) /N.~C"ST 1/-/3Cr !iT
k~r. pF IVFAR RIAI~.5 eRZ C.Z. eRZ
--
~~ AUT£ij'llU. OF CASE -4- 48C.r. /I-.-I8C.I. A-4ec.I. -
~f
STb. WAJ.L T#/t:'KAlESS /MIA/. /·/?"'8" /Q/I~~" /"'L~
V.#NES ~ /V/lT£RWAYS srD ,/. AS CAST STOv'
/VEARRINGS sucr/pls('#. -/,/ all -:__
~
CLEARANCE ()'o:J-aoM lII IC/) AlS a~·o.O:U· -
tc /Vl?FlU/?: MAr. 1~/srG. /DTNeRS 11-130> ST/8IIZ [.f·J»oS1J2/.PZ 1-1!Jt).ST/lIIlz /I-IJCrsrbz
• rYPE IVSTG. /OTHeR5 D/CL./SEMNI ENCL.
/:)/AM. IgST(fj. ~SI6N/MifIt'. /'1:67;tJ¥I. ~-JfJ~' /9'1Ii:r20";l,
£~LfsEMI'E
--
~~ 01N£R SR;. J:)G5ItSIV/iI#./MIN. ""-IJO_ ,~~
"¥7.70~'
BAlANCE -IIYPR. /srAr. /bYN.
'"-/0/ _/0//0/ " - .- R£Q'o.

~I WEA-~ RINISS S"i/CT./PISt:/I.


WR.:l JoB F7:l
-/ ,/
1;50
all-
'so
-/v'
'/.;150 . c-._-:.-...-
JI'/I),1?, TN/?uST MAX. bO/IIIK 40._eI 74700 Cl) 440e>0

!
}/YI)I? TIIHusr MAJ(. UP #DIVE.
vELOCITY-EYE «SIGN/MAY. /S~/~3 l'1/1fFPS /~/.:11
~4(iO'-"!) NONE --
-
I~ !
;
BEAHIAIfi: MATERIAL BRz BHz 8HZ ,aRZ
80~ LUIS. 7 LID SELP/.-;.4 l.5aFfts® S£V/:I.4 S£LF
-
II
(jR()()I/tEI)/rYP8 ,,/.1iuV6 "IJPURN. ,,/SI.£EVE
5CR.''''- S"6LF/.:l •SElF/loO (3) .s6i.F/~ seLF
._..,
J.u8E /1./0
tVfOOVED/rYPE
.t.1I8£ /4h
'" /suEvel ,/ IJDt/RN. "'/SlEEV£
.s£~ /01 ISELM.o C~ SELF/.;z
-
.sELF
-
",,,0
q/?OO/IEt> /TYP£ '"/sa"EVE I .-/ JDtlHM v/~V£
THRDTTLe 1.1J8E/4,,'D/U.IfJV. ,sElF. ';/ '" :SELI'/UI(3JJ- ~u~/v ~EL~
SIlN7: MAr. PUMP Ei.EMENT S5410 S5416 S5410 SS~S£RIt5

MAT. Cbl.LlMN 55410 C.ST. s5410 -


DlI/M. /i.ON/$ "£ARI1I6 SPAN S'/48~ S16"/4S' S"/48~ -
'"
COM81NEO STJlfESS MAJ(.
SURFAC£ FINISII ¥ ITEl/lfls.
ID.OOO H,S4oPSi: laooo
3.:1.
-
6RLJ
~ ;:J~ 3~RMS
.sURFAC£ @ snIFF". Box .3.;2 J/.:I.5"RMS -fJ.;2 GRD
~ SUEW.STWT.64t/_aLANO "/0/ I '-10/ ,//t/ I~QlHalb.
~ PH6SS <§) sruFF. sox ISPSI>sucr.II5-2f)PSI>SIXT. ISPSI>SIK:r.

""- :S,EEVE MAr. S$-4.:10 AI.LOY·sr. S5·4::10 -


§
\!
(f) AS PER PRosPeR LE7TER MARCH"" 1970. (sP6C./i'EQUESTeO 1I·13C,..ST
~ SUC7,01l BELL 01 1.1.1' .sr'G.IMPEt.L£R M.4TEHI.4t..)
~ H/6/1 oD/IIIN 7"HRUST.
3 SPE.C. RECOMMENDS A MIN. RATIO OF.:2. P/ir)sP6R DGC(.INES TO
~
::;j UP/$,I;'AO£ SU8./ceT RATIO.
~ NOTES: ENCIRCLED ITEMS ARE UNDESIRABLE
A CHECK I INDICATES COMPLIANCE WITH SPECS
:>
OW
c<.

SU!.!:-J.:'.RY Of BIDS JG<l No ){ j( x )(


COAlD£N5A TE PUMPS P.O.NO. REV.

M-23 0
XYZ POWER COMPANY SHr.~OF4

Fig. 27.5 Continued


822 Procuring Centrifugal Pumps

'",

~i MTM PROSPER F/.J SPEC


COUPL/J16S: IloIrUWAI- $P4Io1/TYPE -/K£VGD 14'5~AI.(I -IK£YEI> IIDrSPG~D.
~ Dli'IVG.: SOI.ID/FLAIoI6ED v'lv v/.- v/v
-
~i
AA/UST ",. V' ,/

MAKE/SIze MTM/iI$' PIi05P£RIIS}1, FL.I! -S" -


Si./Cr.lJARRE/..: WALL/WANTED 3hi" / v ~'/"" %'/"" -
~I IIff.VeLOt:lTY DEfiGN/MAIf. -/6 3~/4 - /~(2) -
&5£ PAIi'T F~N6E - S7~57~:214 - -
-
Jj

Ii GASNET UIIT./TYP/i. -/F£l/T"



-/li'OtlND - /.Ilr1VAlD
MIITERIAL FAlJ. ST. C.St. .t=AB.ST. -
NOZZ.LE HEAP: AlET AREA MoT.lMlfREL -- $.;1" - -
~~
~!
A44.t. /)lAM. BARRe-I..
BASE
FAST/iAlIIoIISS: MAT. INTEliWl/1..
-ST
46"
$'.;1 ~
-
1'/0>.
-
-
ST
-
-
EK7'ERAlAI.. 5T ,4-193 'B7 :JT -
• PGRFMMAM:"G TIiST IN S#DP v NO v RIiClX>
0
FACSIMILe TES7'/SAME PUMFTYPE #OTREGD v v )/oTREiQ'/>
.,,-
-
CO
~!
NEC#. 7E5T1NSIID~fdllYAS5EM. v MU'/I./?e1T.
WELPIJ./6 RcQVIREMENrs 7'0 ASAIE v (.oJ) 0/
tt'tJMP/...oFNAr£Ii'IAI- TO /lSTM ,/ CO v (P) ,1ST""

-J
V. k'.eY,DRIVIiAiCOt/PJ..INGS
~ -IFf" J'?.EcOMM'£A/P£n
~ (6) H£L:OMMENt>E() RANGe 4'~~FPS
(3180H!1- AssEM. ONLy' FfOrA7'IiD BY
~ III1Nt> DNI.. V
~
(4) WE./../J£RS /oIOTCObC OVAI..IFIEO
k @ PIIY.s.~t!IIEM. PHDPERTYONI.. v
~ ~) COMP/..IES H!ITIlItSTM pKOPE~,
~ lies /JtlT DOcS Nor SuPPLY
"
'J MILL r£~T REPORTS.

'~"
.~
~~ NOTES: ENCIRCLED ITEMS ARE UNDESIRABLE
A CHECK' INDICATES COMPLIANCE WITH SPECS
05
SUMMARY OF BIDS IJOB"o XX XX
RLV.
P',O.NO.
COAIDENSA T£ Pt./M~
M-23
0
XYZ POWER COMPANY $#T-40F4

Fig. 27.5 Continued


Procuring Centrifugal Pumps 823

3. Materials of construction.
4. Driver rated power.

The principle behind this limited analysis is that there is no advantage to the purchaser in having pumps
that exceed the specification. This approach rests on the assumption that a specification can ensure pumps
meeting it will achieve or exceed the desired mean time between repair (MTBR). For those users who
are continually seeking to increase the MTBR of their pumps, meeting a minimum standard is not
enough; they want to purchase the pump with the lowest "total cost of ownership."
Determining the "total cost of ownership" involves taking the simple bid evalution outlined above
and adding to each bid the estimated cost of installation, operation, and maintenance. These costs are
sometimes weighted to take account of the circumstances of a particular installation (e.g., in waterworks
and pipeline applications), the cost of energy can be as high as 80 percent of the "total cost of ownership,"
because the pumping energy is the whole process and the pump designs typically used are inherently
very reliable. At the other extreme, maintenance cost and the risk of unscheduled shutdowns are often
far more important than energy consumption in process applications, where the service is often severe
and the cost of lost production is significantly greater than any other consideration.

Installation
The cost of installation normally takes account of the following:

l. Providing the necessary NPSH, motive power supply, and auxiliary services
2. Piping complexity needed to stay within the pump's piping load capacity
3. Size and complexity of any bypass system that is needed
4 The extent of auxiliary services piping required
5. Means of getting the unit into position, leveling and grouting its base, and aligning the pump and its driver.

In many cases, the cost of providing NPSH, motive power, and auxiliary services is similar for all
selections and so can be left out of the evaluation. Note, too, that a selection that requires significantly
less NPSH than the others may in fact cause operating problems (piping vibration) and pose a high
maintenance cost because its suction specific speed is too high (see Chap. 22). Other aspects of the cost
of installation are dealt with in Chapter 28.

Energy Consumption
In operating a pump, the user pays for the total energy consumed. Given this, it is important to
evaluate the power consumed by the pump operating over its expected load profile, then factor that for
the planned life of the plant (see Chapter 21). The only differences in energy consumption that should
be evaluated are those that differ by more than the accuracy with which the energy consumption can
be measured. Noting that the uncertainty of pump efficiency is relatively high (2 to 5 percent depending
on the test code), it is evident that small differences in energy consumption are not significant. In
assessing energy consumption it is important to take account of any permanent bypasses. The provision
of these is quite common in small to medium size pumps to ensure the pump does not run below its
minimum allowable flow while avoiding the complexity of an automatic bypass system. The energy lost
across them can be quite high, often dominating the evaluation of energy consumption.

Maintenance
Long considered an intangible issue, maintenance cost is now accepted by many pump users as a
cost than can be estimated for various designs with acceptable certainty. The usual means of accounting
824 Procuring Centrifugal Pumps

for maintenance cost is to apply various multipliers to the basic purchase price. These multipliers are
determined from the following:

1. Hydraulic fit. A broad term based on the fraction of best efficiency point (BEP) at which the pump is
selected, the fundamental idea being that selections further from BEP will wear more rapidly. In practice,
the variation in wear rate with fraction of BEP varies with pump size and design. Recognizing this, the
more sophisticated analyses include suction specific speed (Fig. 27.6) and pump power (Fig. 27.7).
2. Rotor stiffness. In all classes of pumps, it has been established that high rotor stiffness increases pump
MTBR. This has been most obvious in overhung pumps, probably because they make up most of the
population. One approach to evaluating overhung pumps is to assign a maintenance cost factor to each
design offered based on relative rotor stiffness (Fig. 27.8). Another approach is to only accept those designs
whose rotor flexibility factor, U/D 4, falls below an allowable maximum value. This value decreases with
increasing pump size; see Chap. 7. A similar approach can be used for between bearings pumps using the
factor K = (WU/D4)O.5 (see Chap. 7), except that now the allowable values decrease with increasing pump
rotative speed. For I and 2 stage pumps to have rotors whose first dry bending critical speed is at least 1.2
times their running speed, K for their rotors cannot be higher than about 40 percent of the "dry-running"
values in Fig. 7.9. Acceptable values of K for pumps of more than 2 stages depend on the rotor design
philosophy deemed appropriate for the application, a matter dealt with in detail in Chap. 7.

NORMAL FLOW RATED FLOW


MAXIMUM MAXIMUM

1-- PENALTY
%
CLASS OF HYDRAULIC FIT
-IGOODI-
FAIR
NO PENALTY
EXCELLENT_ RUNOUT
FLOW
141210 8 6 4 2 0 I
\ 1\ \ ' ~ L\ , ~ 1
L/-rt
I

!0~ ~l\-
"~\ ~~ ~\~ rt~
t-- S-4~00

,
l.L-rI
,
1'--- t-- 5pOO
,

~ r-- 6boo iJ--, V


~ ~\r\' t\- r-- 7600 J VJ
I~ kir\\ f\-\ r-- L~
-
NbT 8600
A CEF TABl
E't ~ \\ ~ ~ I

9pOO ...... ~

'~ ~rY ~ 10,pOO J-1


\-\ ~M~
,
11,boo_ V~
\ \\\ ~ ~ \ ,,
~
\ \ \
12,boo,..j ~r
I I
10 20 30 40 50 60 70 80 90 100 110 120 130 140 150
FLOW - % BEP OF RATED IMP

Fig. 27.6 Chart for evaluating suction specific speed and hydraulic fit.
(Courtesy Fluor-Daniel.)
NOTES:
1. Suctlo,", lP9clflc 'pHd (nu) Is defined u : Allume • ,ingle;"uctIOn w ... r ~ump. At tho b ... ·.lI1cl.ncV
I I I I 1111 1111 11 1I111I HHH1II1IlIi11111 11111 111111 11111111111111 1• poInt (BEP'. Ullng ma" lmym diameter I mp. II .'~ the
NU · RPM~ perfor m.nee curve st.t•• a f low of 650 GPM ana. NPSHR
(NPSHRI·75 requ irement of 1S fee~ Th. pump run. et 3510 RPM .
GPM • V.S. G.llon't*pet"·mlnute, m ........ rQd It bet-t M-t lel.nev flOw «BEP) for Calcu lating Suctlon-Speclf lc·Sp88'd :
,n ...mlxlmum dl,me, .. Imp.II.,. Usa on.half of the bel1-e'ffc:l."cy ·
N .. . (35701~ 10.415
flow fol'" • doubll suction Imp.HI.' ,
(181 '5
NPSHR • N.t posltfvl luetlon heed requl,ed at bln-efflclency· flow for tn.
At 650 GPM BEP f l ow. the allowabl. min i mum flow II 29"
mu: Imum dl.mlter ImP4t'.t.
of BEP. or 31g GPM.
'2. "umpt. wrth Nil valu . . .bo", 11,000 .r. p ..tlcularlv ""Iceptlbl. to low f low
r.clrcul.tlon. S" De.lgn P,.ctlce G·121.' for guldaru:: .. 2. Anum •• doubl.,ucdon hydrOc.. rbon pump w i th SEP flow
014000 GPM. NPSHR II 16 loet and tho lpeed 1.1780 RPM .
3. Allo U.I "W.tet Minimum Continuous Flow" for WI'.' .olution'S. acid,
CMJltlCS.. ,nd .Iurrl .... U .. ·'Hydroe.rbon Minimum Continuous FloW" ,1,0 N .. . (1780m.51 (40001 • 9 951
'or non-wa,e... baMd 'norOlnfc liQuids lueh .. anhydrous ammOn II.
70
l161 '5 •
C. Thll chart dOl. not apply to high speed, Integrally ... r.ct pump,- U ..
The . IIOw.bl. m i nimum flow II 42" of eEP or 1680 GPM.
manuft<;tur.,', r.tommanded mlnlm'-lm flow Ir'\lte~.

I I I I I 1111111111
II 11 11 1' 80

3: ~
o
-J 70 -J
50 u..
u.. 0..
0.. W
W III
III
u.. u..
40 11 111 1 1 1 1 1111~111 11"11~lmllllll"I~~mllllll_ II ' 11:Uffi~~_"IIII I "I" ll lm~m~11 III 1 1 1 I 1 1111 11 1111 11 11111 60 o
o ;I!
;I!
(1)1 (
Z3: 3:
alo 30 I I " I " IIIlln 1111 111 ~IIHjjjj 0:: 0
W-J
O::-J 11,1111111111111111111 50 .... u..
«u.. «II)
g~ 3::;:)
0::0 o
:;:)
0:;:)
Z
>z
:x:-
20'U=HilUW11tut.~~mllll ll "mlll"""~ ff=1 1 111 mn l lll l"_ItI~I~II~m~~II~~UiU:t1:l 1 1 1 1 11 1 1 1 11111 40 t=
....
Z
Z
o o
u u
:!: :!:
:;:)
:;:)
:!: :!:
Z
Z
... ... c... 0. ..... tiD '0 .... ... ...,0. . . . 00'0 ....
:!: '" '" .. '" '" i
BEST EFFICIENCY FLOW FOR MAXIMUM DIAMETER IMPELLER - GPM
(ONE·HALF FLOW FOR DOUBLE SUCTION IMPELLER)

Fig. 27.7 Chart for minimum continuous flow versus suction specific speed.
~ (Courtesy Chevron.)
826 Procuring Centrifugal Pumps

~~ PUMP
2.0 NOT
a: /~~, ACCEPTABLE
w
-w
~() 1.8
-z
,
~c(
::::>z 1.6
~w
I
w .....
()~
/ SFF' - L3/D 4 FOR
-c(
1.4 / PUMP BEING

/
a:~
Q.a: EVALUATED
90
cau. j SFF L - LOWEST L3/D 4
1.2

/
OF PUMPS BID

1.0
1.2 1.4 1.6 1.8 2.0
SFF'/SFF L

Fig. 27.8 Chart for evaluating overhung pump rotors stiffness.

3. Bearing capacity. Following serious problems with short antifriction bearing life in overhung pumps in
chemical and refinery applications, pump users introduced comparison of bearing capacity (basic load rating)
as another factor in assessing maintenance cost. Comparison of bearing capacity alone can be misleading,
because the lever ratio of the rotor has a direct effect on the actual radial load applied to each bearing (Fig. 27.9).
4. Interchangeability and product support. A high degree of interchangeability of parts with equipment already
installed reduces inventory costs. Harder to quantify but nevertheless of notable value is support from the
manufacturer for the supply of parts, carrying out major repairs, and providing technical assistance in
resolving operating problems .

....0 - - - - - - - - - - L - - - - - - - - - - - -
a::
~ ....0------- L1 - - - - - -
...J
W
0..
~

-$=--
--=----
-eJ

-Lt-L--
If! EQUIVALENT RADIAL LOAD @ LINE BEARING
FrL = Frj (L/L2)
EQUIVALENT RADIAL LOAD @ THRUST BEARING
FrT = Frj (L1/L2)

Fig. 27.9 Effect of overhung rotor geometry on bearing loads.


Procuring Centrifugal Pumps 827

Ordering Equipment
Once the evaluation is complete, the customer will normally select the equipment and place it on order.
A critical aspect in placing the order is to do so with enough time to allow manufacture of the equipment
and delivery to the site on or before the date it is required. With increasingly complex purchasing
procedures, it is now not unusual to have order placement delayed without any extension of the site
need date. This practice is not conductive to procuring first-class equipment.
Promises of shipment made by pump manufacturers date from the receipt of complete manufacturing
information; therefore, to avoid delays in the preparation of drawings and in the issuance of specifications
to the shop for manufacturing, it is necessary that the purchaser's order contain all required information.
In addition to the complete conditions of service for the pump as well as the driver, the order should
specify the desired direction of rotation, a statement as to whether manufacturing should proceed without
awaiting approval of elevation drawings, and other pertinent information. For example, if the pump is
to be driven by a steam turbine equipped with an excess-pressure regulator, and if the turbine is to be
ordered by the pump manufacturer, the order should specify the location of the pilot lines from the
regulator and the value of the excess pressure to be maintained.
If the order for the pump requires the pump manufacturer to furnish the driver, he will order it and
obtain a certified elevation drawing from the driver manufacturer before proceeding with the combined
elevation drawing. If the driver is ordered separately, the purchaser must expedite the receipt of the
driver print necessary for the preparation of the combined drawing. The proposal submitted by the
manufacturer usually includes a bulletin describing the construction features of the equipment as well
as an outline drawing of the pump in question, as shown in Fig. 27.10. Although the dimensions given
on that drawing are not to be used for final construction purposes, they are sufficiently detailed, not only
to determine the desired direction of rotation, but also to prepare a preliminary layout of the installation.
The significance of the direction of rotation varies with the class of pump and its driver. Lines of
standard, single-stage, end-suction pumps generally have only one direction of rotation, whether close-
couple mounted on a motor (Fig. 1.19) or driven by a separate driver. Since most modem designs of
this class of pump have centerline discharge, the manufacturer keeps pattern and inventory costs down
without affecting nozzle location. Most single- and multistage axially split pumps can be built for either
direction of rotation, so it is important to specify the required direction in the purchase order. Without
this information, the driver order often cannot be finalized, because while motors and steam turbines
can be built for either direction of rotation, most cannot be easily changed once built (modem electric
motors usually have unidirectional fans to reduce power loss and noise). Radially split single and
mutistage pumps are usually designed for one direction of rotation, although some ranges include sizes
that can be built in either direction, either by their construction (single-stage pumps) or two sets of
pattern equipment. When dealing with large, complex pump and driver trains, the second detail to be
settled between all parties should be the direction of rotation of all the components.
In addition to showing the general dimensions of the equipment and the location of the baseplate
foundation bolts, an outline elevation drawing may also depict a great number of details such as the
location and size of all auxiliary cooling or lubricating piping, and space requirements for dismantling.
The complexity of the information that will be contained on an outline drawing reflects the severity of
the conditions of service involved and, therefore, the complexity of construction of the pump itself.
The construction of equipment for small- or medium-size pumps intended for general service is
usually standard, and unless some special features have been discussed with the pump manufacturer
before placement of the order, it is practical to release the pump for manufacturing without formal
approval by the customer.
828 Procuring Centrifugal Pumps

COUNTER - CLOCKWISE ROTATION DETERMINED


FROM OUTBOARD END OF PIJMf'

~------------------H--------------------~

r-- B --+oo-----C

"(f' PIPE TAP DRAIN


~--------------N-----------------~.I

.....
r ~~
iffi
iii
I
• I! G. . . I
. . . FLANGE DIMENSIONS.

.....
... ...0
~~
...0
.. ........ ..d... .... ..... .,
i:t A B C 0 E F H J I( L M N 0 RR IL

~i !!:I
as
0 0 ..
!:! c
'" ID c U IE
.. '!! e .. ... :l
. ;d .....
0 "'0 ~ .. u
... ~ 0 :II .. "z
liZ ~5 :>a z><
ell
~& e ..
is ... iii
ZGI ..",
10 LN-18 10 12 12)- 21 19 201i 26; 251 191 92\l 141- 19t 26t ~ 81t i 13~
10 16 i
aID

14! 12 Ii
10 LN-22 10 14 14 25 19i 221 29i 29* 221 103* 17 221 30 37t 86l 15i-
f 12 19 i- 17 12 I·f
14 21 I 18' 12 11
16 23t I 21f 16 1.1

18 25 If 22f 16 11
"
20t 991- 14t 83t t
32i 25, 114t• 17
12 LN-17 12 16 14 24 21 22i 28, 261 23 30 33 14t
22i 24i 311
20 27t It 25 20 Iii
12LN-21 12 18 15 28~ 2~ 31'.- 36 94' t 16f

12 LN-32 12 20 18 32 30 29~ 38i 441 30 1481 25 30 40 40 130 f 23


• APPROXIMATE ONLY. DEPfNDlN6 ON SIZE OF MOTOR.

Fig.27.10 Typical tabulated elevation drawing.


Used in proposals for double-suction, single-stage pumps with axially split casings.

Frequently, the construction of the pump itself may be of a standard nature, but approval may be
required for the general arrangement, including the baseplate construction. In such an event, the order
should authorize the manufacturer to proceed with construction of the pump but withhold approval for
the manufacture of the baseplate.
Placing an order with a large specification and various classes of approval is a complex matter. That
complexity is compounded by the paper handling process within the manufacturer's organization. To
help avoid the delays that can occur in these circumstances, the procurement process should include the
following two steps:
1. Before the order is placed: a pre-award meeting to be sure all the requirements of the order are understood.
When the order is large or unusually complex, the manufacturer's contract engineer should attend this meeting.
Procuring Centrifugal Pumps 829

2. Within 4-6 weeks after the order is placed: a post-award meeting at the manufacturer's plant to ensure that
the order is in work, resolve any questions that have arisen, and review and settle all issues related to
inspection and testing.

Simpler orders do not warrant these two steps, but most do warrant a call, within 4 weeks of order
placement, to the plant making the equipment, to verify that the order is being processed. This draws
from an axiom in the construction industry that says "An order goes wrong in the first 4 weeks."
V
INSTALLATION, OPERATION,
MAINTENANCE, and DIAGNOSTICS
28
Installation

Except for small domestic pumps such as hot water circulators and cellar drainers, the successful
installation of a pump has two phases: designing the installation and physically installing the pump and
associated equipment in the field. Projects involving large pumps or a large number of pumps, such as
power plants, refineries, and large water or sewage installations, have, of necessity, a great deal of the
installation design finalized before even writing the specification.
The general information contained in this chapter and in Chapter 29 should be supplemented by the
specific instructions prepared by the manufacturer for the centrifugal pump in question. These instructions
are shown on the outline drawing for the pump, or in a supplement to the outline drawing, and are
included in the instruction book that manufacturers ship with the pump or send to the customer when
the pump is shipped.

INSTRUCTION BOOKS

An instruction book contains direction for the following phases of pump use:

1. Installation for maximum service with minimum wear and expense.


2. Adjustment and operation for optimum performance.
3. Maintenance and repair when necessary.

Because instruction books are intended to help keep the machinery efficient and reliable at all times, they
should always be available to the following personnel: (1) construction men responsible for installation, (2)
operators who use the equipment and make periodic checks or tests, (3) maintenance men who repair
and service the equipment, and (4) engineers who determine the proper use of the apparatus. If the books
contain a parts list and instructions for ordering repair parts, a copy should be available to personnel
responsible for ordering repair parts.
The number of instruction books needed, therefore, depends on the extent to which these functions

833

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
834 Installation

are departmentalized. If one man is in charge of installing, operating, and maintaining a centrifugal
pump, he only needs one copy. If each function is performed by a separate department, however, each
department should have at least one copy of the instruction book.
Too often, sufficient instruction books are requested, but are never transmitted to the proper personnel
once they nave been received. Later, when a department needs the books, they may be difficult to locate
in a short time.
A request for abnormal quantities of instruction books adds to the cost of providing this material,
costs that are passed on to the customer. On the other hand, books requested after the pump is delivered
may be more expensive, particularly if they are prepared for equipment deviating from standard. It is
wise, therefore, to order the exact number of books required when the pump is ordered. It is equally
important that the books be given to the personnel who will need them and not be kept locked away.

Installation Design

PROVISION FOR TOLERANCES

In this age of computers, there is a tendency to take the exactness with which dimensions can be
calculated and try to transfer that to the installation of equipment such as pumps. Despite the widespread
use of computers in their manufacture, pumps are still an assembly of machined castings and fabrications,
and therefore have a certain tolerance inherent in the location of their connections and foundation bolts.
At the same time, prefabricated piping inevitably has some tolerance in its dimensions. The same is true
for the location of cast-in foundation bolts. Given these realities, it is necessary to design into the
installation some means of accommodating the tolerances so the installation can proceed as planned.
Failure to do this complicates the physical installation of the equipment, which drives up the installation
cost and can delay plant startup. The latter, if it occurs, can significantly increase the cost of a project.

GENERAL RULES FOR PUMP LOCATION

Pumps should be installed in easily accessible, well lit, dry, and clean locations whenever possible.
Doing this helps ensure they will be inspected regularly, will not suffer damage to pump or driver from
accumulated dirt or moisture, and will be easy to maintain when necessary.
If a motor-driven unit will be operated in a damp, moist, or dusty location, the proper motor must
be selected. Pumps and drivers designed for outdoor installation are specially constructed to withstand
exposure to weather.
Sufficient room must always be provided for dismantling the pump; that is, enough headroom must
be provided above horizontal pumps with axially split casings to allow the casing upper half to be lifted
clear of the rotor. Similarly, horizontal multistage pumps with radially split casings must have sufficient
clearance at their outboard end to allow removal of the inner element without canting it (Fig. 3.16). For
large pumps with heavy casings and rotors, a traveling crane or facilities for attaching a hoist should
be provided over the pump location.
When pumping equipment must be used at levels where flooding is possible, two courses of action
can be followed: (1) a vertical wet-pit pump may be used or (2) auxiliary wet-pit drainage pumps must
be provided as insurance against damage to the main equipment.
Installation 835

In normal installations, pumps should be located as close as possible to the source of liquid supply.
When convenient, the pump centerline should be placed below the level of the liquid in the suction
reservoir. The manufacturer's recommendations for suction conditions should always be followed.

FOUNDATIONS

Most horizontal centrifugal pumps are mounted on what are termed "supported" baseplates (see Chap.
13), designs that are intended to be supported by a "massive" foundation. The purpose of the foundation
in these cases is to provide the bending and torsional stiffness necessary to maintain alignment between
the pump and its driver, and to absorb any normally imposed loads and shocks. The foundations are
usually concrete, with a mass at least three times that of the supported equipment mass, and sufficient
surface area to not exceed the allowable ground pressure. Foundations are generally cast separately from
the floor to avoid the influence of any adjacent floor movement. Equipment mounted on "massive"
foundations is always grouted, therefore it is necessary to provide a group space, typically 50 mm (2
in.), between the top of the rough cast foundation and the underside of the equipment.
Very large equipment is often mounted directly on the foundation to avoid the cost of a baseplate.
In these cases, soleplates should be provided under the pump and driver feet. In this manner, the alignment
can be corrected with shims, and the equipment can be removed and replaced without difficulty if
necessary. The fundamental design of the foundation for soleplate-mounted equipment is the same as
that described above for equipment mounted on a "supported" baseplate.
The space required by the pumping unit and the location of the foundation bolts are determined from
the outline drawing supplied by the manufacturer. Each foundation bolt (Fig. 28.1 [aD should be surrounded

(8) (b)

SLEEVED
FOUNDATION
BOLT

"""r-+-..,..-Y -.4--\-- - GROUT; PLACED


AFTER EQUIPMENT
IS POSITIONED

Fig. 28.1 (a) "Cast-in" foundation bolt. (b) Grouted foundation bolt.
836 Installation

by a pipe sleeve, three to four diameters large than the bolt, to allow some bolt movement after the
foundation is poured and to allow the bolt to stretch freely as it is tightened. If there is some doubt
about the accuracy with which cast-in foundation bolts can be positioned, another approach is to cast
pyramid holes in the foundation for the foundation bolts (Fig. 28.1 [bJ), then grout the bolts once the
equipment is set in position. Grouted foundation bolts still have to be sleeved so they can stretch
during tightening.
When a unit is mounted on steelwork or some other structure, either the structure or the base must
be designed to provide the bending and torsional stiffness necessary to maintain equipment alignment.
Chapter 13 discusses the design of such bases.
Resiliently mounted bases (see Noise control) have to be designed to provide the bending and torsional
stiffness necessary to maintain equipment alignment (Chap. 13), so the foundation design has to consider
only unit weight plus whatever forces piping displacement produces. Three-point-supported bases, used
at present on off-shore platforms and some land-based installations to lower erection costs, are also
designed to maintain equipment alignment (Chap. 13). Their foundation design, however, has to take
full account of the imposed piping loads since the connection to the foundation has high vertk:al and
lateral stiffness.

MOUNTING OF VERTICAL WET-PIT PUMPS

Vertical wet-pit pumps are generally supported on a reinforced concrete structure. The stiffness of this
structure has a profound effect on the natural frequencies of the installed pump. The structure's design
therefore needs to ensure that the stiffness is high enough to maintain adequate separation, typically 10
percent, between running speed and close natural frequencies. Failure to respect this requirement can
result in resonant vibration of the installed pump.
A curb ring or soleplate must be used as a bearing surface for the support flange of a vertical wet-
pit pump. The mounting face must be machined to an even, flat surface since the curb ring or soleplate
will be used in aligning the pump.
If the discharge pipe is located below the support flange (below-ground discharge), the curb ring or
soleplate must be large enough to allow passage of the discharge elbows during assembly. A rectangular
ring (Fig. 28.2) should be used.
If the discharge pipe is located above the support flange of the pump (above-ground discharge), a
round curb ring of soleplate should be provided (Fig. 28.3). Sufficient clearance should be left on the
inner diameter to allow passage of all sections of the pump located below the support flange, such as
manhole openings on the side of the casing. A typical method of arranging a grouted soleplate for a
vertical pump is shown in Figure 28.4.
If the discharge is below ground (Fig. 14.26), and a Dresser-type coupling is used, it is necessary to
design for the moment that can be imposed on the structure. Generally, the pump casing must be securely
attached to some rigid structural members with tie rods. If vertical wet-pit pumps are very long, some
steadying device is required regardless of the location of the discharge or the type of pipe connection.
Here again, tie rods can be used to connect to the unit to a wall, or a small clearance can be provided
around a flange to prevent excessive displacement of the wet-pit pump in the horizontal plane.

SUCTION SYSTEM

The principal considerations in the design of the suction system are


Installation 837

NPSH available
Vessel or sump configuration
Piping length and size
Exit from the vessel or sump
Piping configuration
Priming, if required, and venting
Need for a temporary permanent strainer

TAPPED HOLES
FOR PUMP TO FOUNDATION
CURB RING BOLT HOLES
CONNECTION

GROUT

CRILL IH) REAM TAPPED HOLES FOR


FOR TAPER PINS PUMP TO CURB
AT ASSEM8LY RING CONNECTION

FOUNDATION GROUT
BOLT HOLES

SHIMS --tLo4ollllN I
I
AND I 1
WEDGES ~'--J<---'1 . r ~~'-'
I, \...... --1
. . . --J::--;
Fig. 28.2 Curb ring details, rectangular type for Fig. 28.3 Curb ring details, round type for
below-ground discharge vertical pump. above-ground discharge vertical pump.

SHIMS AND WEDGES

OF FOUNDATION
'-TOP OF FOUNDATION---J

Fig. 28.4 Grouting form for vertical pump soleplate.


838 Installation

Chapters 19 deals with the calculation of NPSH available, and how much NPSH should be available
ensure pump hydraulic performance and avoid cavitation erosion. Chapter 22 discusses the effect of
required NPSH, hence available NPSH, on centrifugal pump rangeability, and therefore the allowable
minimum flow. The various means of priming are discussed in Chapter 25. Chapter 14 provides guidelines
for the layout of vertical wet-pit pump installations. This section therefore deals with the layout of
the suction piping for horizontal and vertical dry-pit pumps. Suction strainers are discussed in the
following section.

Importance of Suction Piping


Generally, the layout of the piping system precedes not only the delivery of the pumping equipment
to the site, but even the preparation of the specifications for this equipment and its purchase. The greatest
amount of attention should be devoted to the layout of the suction piping, because it is the most vulnerable
portion of the piping system. Few problems are ever created by improper discharge piping: excessively
high velocities in this portion of the system will result only in increased friction losses and wasted
energy, whereas the choice of too Iowa velocity will make the piping and valves more costly than
necessary. An inadequate suction piping layout, on the other hand, can lead to disastrous effects on
pump performance.
A great many installations show evidence that the user's engineers have exercised considerable
forethought to provide adequate NPSH and to select pumps in such a manner that low-flow operation
will not create undue problems arising from internal recirculation at the suction. Too often, however,
having taken these precautions, they throwaway all the advantages they have thus gained by neglecting
the most elementary precautions in the layout of the suction piping leading to the pumps.
In designing the installation of medium to large pumps, there is merit to reviewing the proposed
suction piping layout with the pump manufacturer early in the project so any necessary changes can be
easily incorporated. Note, in this connection, that the pump manufacturer will not accept responsibility
for the suction piping design, unless design is part of the contract, but will review it and provide comment.

Suction Piping Length and Size


The suction piping should be as short and as direct as possible, taking due account of thermal expansion
in high-temperature services, and sized for a low velocity, typically 1.5-2.5 mls (5-8 ft/s). At the pump,
the suction piping should be at least the same size as the suction nozzle, and is usually one pipe size
larger, occasionally more, to meet the given velocity limit. If a long suction line is required, the pipe
size should be increased to reduce friction losses, with two qualifications. First, in very long suction
lines, the inertia of the liquid can be so high that the suction head cannot accelerate it as quickly as the
pump can accelerate the liquid in the discharge system. Unless such systems have some means of slowing
the discharge acceleration rate (e.g., a slow opening valve or slow acceleration of the pump's driver),
the liquid in the suction piping is likely to separate. This phenomenon, known as "column separation,"
can cause cavitation (or even vapor binding), then produce very high peak pressures as the suction
system catches up and the column closes. Second, in boiler-feed pumps for open-cycle power plants
difficulties can arise during transient conditions of sudden load change if the suction piping volume is
excessive. For a detailed explanation of this phenomenon and its avoidance see Karassik [28.1].

Protection Against Vortex Formation


Under certain conditions, vortices may form in the reservoir from which a pump takes its suction.
This can happen not only when the liquid surface is open to the atmosphere (as in Fig. 28.5), but also
when the liquid is contained in a closed tank (Fig. 28.6) the upper portion of which is filled with vapor
Installation 839

SUCTION
PIPE

V
t S

V
..-
_'~I_+---'-
II
I I
I I

Fig. 28.5 Submergence over suction pipe opening Fig. 28.6 Submergence over suction pipe opening in
from reservoir open to atmosphere. a closed reservoir.

of the liquid itself, such as in a feedwater deaerator or in a fractionating tower. As a matter of fact, the
case of closed tanks is the more insidious of the two possibilities. When a pump takes it suction from
an open reservoir, it is generally possible to observe the vortex formation if it occurs and to quickly
diagnose the problem. A vortex in a closed tank is hidden, and therefore it can escape detection.
To prevent the formation of vortices, it is necessary to provide a certain minimum submergence over
the suction pipe opening; this minimum submergence increasing rapidly with the velocity at the opening
itself. Note that this problem is completely independent of NPSH requirements. It arises when liquid
more or less at the rest is forced to accelerate at the point of entrance to a reduced area, such as the
inlet to a suction pipe. As the submergence over the suction pipe opening decreases, the tendency to
develop a vortex increases, until at some value of this submergence, a vortex forms. As the submergence
decreases further, an empty core develops at the center of the vortex, air or vapor is drawn down into
this core and ultimately enters the pump in quantities sufficient to air- or vapor-bind it.
Figure 28.7 gives conservative values for the minimum submergence over the suction pipe opening.
The advantage of reducing the velocity at the inlet is quite evident and explains why a bellmouth
connection (as in Fig. 28.8) is preferable, as the velocity at the opening can be easily reduced by a factor
of 4 or even more, thus permitting the pump to operate satisfactorily with much less submergence.
In cases where the minimum submergence necessary to prevent vortexing cannot be provided, even
by the use of a bellmouth, it is possible to lower the necessary submergence by incorporating a "vortex
breaker" in the outlet from the suction vessel. Figure 28.9 shows various arrangements. In using these,
it is important to appreciate that they will not prevent vortexing, that their function is to lower the
submergence at which vortexing will occur. Data on how much the submergence can be lowered are
meager. Figure 28.9 gives values typical of four- or eight-vaned "vortex breakers," with the qualification
that the submergence must not be less than 2d over the outlet.

Avoidance of Air Pockets


When the source of the liquid is below the pump centerline, the piping to the pump suction flange
must slope constantly upward from the source to the pump, as shown on Fig. 28.1O(a). If the line has
a high point ahead of the pump, as in Fig. 28.1O(b), it is possible for air or vapors to accumulate in the
pocket thus formed. At best, this will reduce the effective cross section of the pipe, increasing the liquid
840 Installation

14

12

10
~
ill 8
~
CJ) 6

O~------~------r-----~-------r------r--
2 4 6 8 10
V (FT/SEC)

Fig. 28.7 Minimum submergence required over a plain pipe exit to prevent vortexing.

1--------

Fig. 28.8 Suction bellmouth arrangements.


Installation 841

- --
(a)
d-i ~
d/2
t hJ - -

-:±--
.
-_.- ' .
.
- - ---
-
~

~ ~:-
- .- -

--- - - -
----.--- --4d -
::::
--- --::
=~-
::-:":
_- ::
--.- -- -"- -
~ -

- ..=:- - .
......
-:-~--
--
tJ
-.

--
.. .
- -

-
-
STORAGE
TANK

-1
- -

d/2 d ~

(b) 4
MIN ~ CURVE ~ 2d
l-
lL
W
::::E U
::>z
W
_Cl
::::E
za: 2
-w
::::E::::E
CD
::>
C/)

0 2 4 6 8 10
VELOCITY @ INLET FT/S

Fig.28.9 (a) Various "vortex breaker" arrangements. (b) Typical minimum submergence
with four- or eight-vaned "vortex breaker."
842 Installation

QSUCTION QSUCTION
I
- A I R POCKET-

(a) Recommended (b) Not recommended

Fig. 28.10 Suction piping slope: (a) correct, (b) incorrect.


Proper sloping of suction line is required to avoid formation of air pockets.

velocity and the friction losses in the suction line. At worst, the accumulation of air or vapor will block
the suction line completely, causing air binding of the line and consequent interruption of pumping.
Between these two extremes is the risk of air or vapor being washed into the pump where it can become
trapped in the impeller eye, thereby causing the pump to lose capacity, or in extreme cases cease
pumping altogether.
Pumps with the suction source above the pump centreline generally have their suction piping rising
continually back to the suction vessel. When the suction piping has to fall below the pump centreline
part way along the run from the suction vessel to the pump, it is must then rise continually in both
directions from the low point (Fig. 28.11).
Even when the proper slope is provided in the suction line, the installation may be vulnerable to an
accumulation of air at some high point such as, for instance, in the clearance space in the body of a
valve installed with its operating shaft vertically upward. It is therefore recommended that gate valves
be installed with their stems horizontal rather than vertical.

Reducers
With the suction piping typically one to two pipe sizes larger than the pump's suction nozzle (see
Suction piping length and size), it is necessary to use a reducer at or close to the pump suction connection.
At the same time, reducers accelerate flow, an action that tends to remove flow distortions, so they can
also serve to reduce some of the ill effects of elbows in the suction piping (see Elbows).
When installed horizontally, reducers are a special case of the need to avoid the possibility of forming

Fig. 28.11 Correct suction piping layout with suction source above pump and low point in suction line.
Piping must rise continuously either side of low point.
Installation 843

an air pocket in the suction line. Figure 28.12 is adapted from the Hydraulic Institute Standards and
illustrates the fact that eccentric reducers installed with the flat side up (as in Fig. 28.12[a]) should be
used, rather than concentric reducers (as in Fig. 28. 12[b]) or eccentric reducers with their flat side down.
What the standards imply, but do not state explicitly, is that this recommendation is directed at installations
where the liquid is located below the pump centerline or where the piping comes from below the pump.
If the liquid is located above the pump and the piping falls continuously to it, the choice of horizontal
reducer installation, so far as forming air pockets is concerned, is immaterial. The choice should then
rest on the following considerations:

• Eccentric reducers installed fiat side up (Fig. 28.12[a]) will give satisfactory results in most cases, and allow
the same installation rule to be applied regardless of the adjacent piping arrangement.
• Eccentric reducers installed fiat side down (Fig. 28.13) are preferred by many plant designers for two reasons:

1. No low points in the suction piping, which could trap settled solids. Installing reducers in this way is
important when there is not an elbow immediately upstream of the reducer.
2. Can assist the venting of any vapor formed at the pump suction, if the downward liquid velocity in the
suction piping does not exceed the rate at which bubbles will rise through the liquid.

To avoid abrupt changes in velocity at the pump suction, reducers connected directly to the pump's
suction nozzle should be conical (constant convergence angle and with an included angle of no more
than 10 deg) rather than contoured as are most forged fittings. For the same reason, reducers should be
limited to a change in diameter of one pipe size, such as 10 x 8 in. or 24 x 20 in. If the suction line is
more than one size larger than the suction flange, two reducers should be installed in series. When other

eccentric reducer
(liquid source
below pump)

desirable

concentric reducer

undesirable desirable
Fig. 2S.12 Reducers at pump suction: (a) eccentric Fig. 28.13 Eccentric reducer at suction with fiat side down.
fiat side up and (b) concentric. Used only when suction source is above the pump.
844 Installation

considerations require the use of contoured reducers, they should be located three to five diameters from
the suction flange. Reducers installed in vertical pipe runs are always concentric.

Elbows
An elbow or series of elbows close to the suction of a centrifugal pump can have a serious effect on
its performance; how serious depends on the pump's suction specific speed, specific speed, whether the
impeller is single or double suction, and the form of the suction nozzle. The general effects of elbows
are discussed now. Following that are guidelines for specific pump configurations and designs.
As liquid passes through an elbow, or the branch of a "tee" in a manifold, it tries continuing to flow
in the direction it was following just ahead of the tum. The net result is that both higher velocities and
higher pressure occur on the outside of the tum, disturbing the uniform velocity that may have existed
just ahead of the elbow (Fig. 28.14). Incidentally, this is not a contradiction of Bemouilli's law, which
applies only to individual streamlines. The phenomenon of this velocity and pressure distribution is
caused by a crowding of the streamlines toward the outside of the elbow as a result of inertia.

Two Elbows in Series


When two elbows or tees are located close together and in planes at right angles to each other, higher
velocity flow streamlines are made to migrate from the outside of the first fitting to the outside of the
second, creating a circumferential swirl in the flow from the second (Fig. 28.15). Left to its own devices,
this swirl will persist for 50 to 100 diameters downstream from the second tum.
If the pump suction is close to the second tum, the swirling liquid is swept into the pump and enters
the impeller with a significant amount of swirl or prerotation. Since impeller designs are based on either
zero prerotation or a specific amount introduced by the suction nozzle, the pump performance will
deteriorate. If the prerotation is in the same direction as the impeller's rotation, a reduction in head will
take place. If the prerotation is in the opposite direction, a slight increase in head will occur. But in
either case, the pump efficiency will be lowered, the operation will be noisy, and the impeller may suffer

r
d

Fig. 28.14 Change in flow distribution through an elbow.


Installation 845

swirl after
seoond bend

slow moving fluid from inside


of first bend moving around
wall by shortest route to the
inside of the seoond bend

Fig. 28.15 Swirl produced by two elbows in series.

cavitation erosion. As a general rule, the severity of these effects will increase with specific speed,
because the proportion of head developed by hydrodynamic action also increases with specific speed,
reaching 100 percent in axial flow or propeller pumps, and hydrodynamic head is directly affected by
velocities at the impeller inlet.

End Suction Pumps


End suction pumps of low to medium specific speed and suction specific speed up to 11,600 (10,000)
are not particularly sensitive to a single elbow close to their suction flange (Fig. 28.16[a]). The elbow
should be long radius, one size larger than the pump's suction nozzle, and followed by a reducer. Where
possible the reducer should be concentric (Figs. 28.12[b] and 28.16[a]) to ensure uniform velocity of
the flow into the suction nozzle.
End suction pumps of specific speed 3,200 (2,750) or higher, or suction specific speed over 11,600
(10,000) are sensitive to even minor flow distortion at the impeller inlet. A concentric reducer is therefore
mandatory for pumps taking suction from above, and the closest elbow should be at least five diameters
upstream of the reducer (Fig. 28.17[a]). In those cases where such pumps are installed with the suction
source below the pump, it is necessary to use an eccentric reducer, installed flat side up, but with at
least five diameters of straight pipe downstream of the reducer (Fig. 28.17[b]) to reduce the velocity
distortion produced by the reducer.
Figures 28.16(b), 28. 17(a) and 28.17(b) show the piping layout proportions to be followed when there
are elbows in series upstream close to the pump suction.

Side or Top Suction Pumps


The sensitivity of side or top suction pumps to an elbow at the suction depends on whether the
impeller is single or double suction. Single-suction impellers are not particularly sensitive because the
elbow-type suction nozzle accelerates the inlet flow and therefore tends to remove distortions in its
velocity distribution.
Double-suction impellers, on the other hand, are very sensitive to the orientation of an elbow close
to the pump's suction nozzle. The typical sketch in Fig. 28.18(a) illustrates the correct way to install an
846 Installation

(8) (b)

D, . SUCTION
~ ~ SUCTION + 1 SIZE

J
ECCENTRIC

Fig. 28.16 Suction piping layouts for end-suction pumps, suction specific speed up to 10,000 (US units).
(a) Preferred arrangement of concentric reducer and long radius elbow when suction source is above the pump
and straight pipe run cannot be provided. (b) Preferred arrangement of eccentric reducer and two elbow
in series when suction source is below the pump.

elbow at the suction of such pumps (the plane of the elbow at right angles to the pump's axis of rotation)
and has been adapted from the Standards of the Hydraulic Institute, in which it has appeared as far back
as the 1930s. Despite this, however, there are untold numbers of installations made as in Figure 28.16(b),
which is described as "undesirable" but really should be called "unacceptable."
The sensitivity of double-suction pumps arises from the fact that when the orientation of an elbow
near their suction is not correct (Fig. 28.18[b)), the flow distortion produced by the elbow causes an
unequal flow to either side of the impeller. This unequal flow can cause the following serious problems:

• Cavitation erosion in one side of the impeller. With one side of the impeller handling more than half the pump
flow, the NPSH required by that side will be greater than shown on the pump's rating curve for the pump
flow. Given this, there is a risk the high flow side of the impeller will suffer cavitation erosion if the available
NPSH is close to the pump's required NPSHR at rated flow.
• Noise, vibration, and impeller erosion on the pressure side of the impeller vanes from suction recirculation.
If the side of the impeller that is being starved operates at flows within the zone of internal recirculation at
the suction (see Chap. 22), the pump can suffer the ill effects of suction recirculation even though the total
flow to the pump may appear to be beyond the range where this phenomenon could occur.
• Reduction in pump performance. The secondary flow established within the impeller as the unequal flows mix
can cause a drop in pump head without a corresponding reduction in power. The pump efficiency is consequently
lower and the energy consumption therefore higher.
• Thrust bearing failure. With a different flow to each side of the impeller, the axial momentum forces produced
by the liquid turning through the eye will not be balanced. Of itself, the axial thrust thus produced is generally
not enough to cause any distress. Should the imbalance in flow also cause cavitation or suction recirculation
on one side or the other of the impeller, the resultant fluctuating axial thrust can have peak values high enough
to cause premature thrust bearing failure.
Installation 847

(8)
D1 = SUCTION
D2 ~ SUCTION + 1 SIZE
~
+
t

~2'\0 Dz
(b)

) . / ECCENTRIC

Fig.28.17 Suction piping layouts for end suction pumps, specific speed 2,750 and higher,
suction specific speed 10,000 and higher.
(a) Suction source above pump. (b) Suction source below pump.

path of water

suction
d••lrabl.

Fig. 28.18 Elbow at suction of double-suction pump: (a) preferred orientation (plane of elbow normal to axis
of rotation.) (b) undesirable orientation (elbow in plane parallel to axis of rotation.)
848 Installation

Typical layouts for piping close the suction of side and top suction pumps are shown in Figures 28.19
and 28.20. The layouts include the proportions to be followed when there are elbows in series close to
the pump's suction.
If for whatever reason the installation of a pump with a double-suction impeller cannot accommodate
an elbow at the suction with the proper orientation (Figs. 28.18[a], 28.19, and 28.20) tolerable performance
can be achieved with an elbow parallel to the pump's axis of rotation by locating the reducer immediately
downstream of the elbow, then providing a straight pipe run of at least 10 diameters to the pump's
suction (Fig. 28.21).
In the compounded case when it is not possible to employ the correct suction piping layout (Figs.

0 1 = SUCTION
O2 ~ SUCTION + 1 SIZE

1~401

I/, ECCENTRIC, FLAT SIDE UP

(a)

1
(b)

Fig. 28.19 Suction piping layout for pumps with side suction nozzle (single- or double-suction impeller)
and two elbows in series.
(a) Suction source below pump. (b) Suction source above pump.
Installation 849

Fig. 28.20 Section piping layout for pumps with top suction nozzle (single- or double-suction impeller) and
two elbows in series.

D1 = SUCTION
~ <?: D1 + 1SIZE

Fig. 28.21 Tolerable alternate suction piping layout for double-suction pump with last elbow in plane parallel
to axis of rotation.

28.l8[a], 28.19, and 28.20) and there is not enough room for the tolerable piping layout shown in Figure
28.1, one possible solution is to use one of the two available devices that reportedly will produce an
almost uniform velocity at the outlet of an elbow. The first is a special elbow (Fig. 28.22) whose cross
section is designed to correct the velocity distortion produced by a standard elbow. The second is an
upstream guide vane designed to develop a compensating distortion of the flow entering an elbow,
thereby a uniform velocity distribution of the flow out of it. Experience with these devices is mixed,
850 Installation

Fig. 28.22 Special elbow for double-suction pump.

which suggests they have to be carefully applied to achieve the desired effect. Of particular note in their
application is the fact that neither device can compensate for a distorted inlet flow.

Multiple Pumps
The suction piping manifold shown in Figure 28.23 effectively has two elbows in planes at rights
angles to each other close to the pump suction, which can cause high prerotation or swirl at the impeller
inlet. To avoid this, it is necessary to either follow the piping layout proportions shown in Figures 28.16,
28.17,28.19, and 28.20, or place the two elbows in the same plane, as shown on Figure 28.24. With
the two elbows in the same plane, the second turn tends to rectify the nonuniform velocity profile created
by the first elbow and the liquid issues from this second elbow with an almost uniform velocity profile
and without a swirl. Actually, two 45-deg elbows will give an even better velocity profile than two 90-
deg turns.
If space limitations interfere with the arrangement shown on Figure 28.23, the pumps could be
alternated on opposite sides of the manifold, without affecting the desirability of this suction piping
arrangement. Another arrangement, which is used frequently in large waterworks installations, is shown
on Figure 28.25.

SUCTION STRAINERS

Without some precautions in the design of the installation and in operation during plant startup from
new or after a major revamp, the passage of foreign material from the system into the installed pumps
Installation 851

UNDESIRABLE

Fig. 28.23 Typical, but undesirable, arrangement of three pumps installed for parallel operation.

PREFERRED

Fig. 28.24 Preferred arrangement for three pumps installed for parallel operation.

can be a major source of additional expense and delay in getting the plant on line. Foreign material
entering centrifugal pumps can have three effects. Large material can block internal waterways, causing
a drop in hydraulic performance. If the blockage is in the impeller, the rotor will be out of balance,
perhaps seriously enough to cause premature shaft seal failure. Fine material, such as grit and scale,
will pass into internal running clearances, where it can cause rapid wear (by grinding abrasion) in a
852 Installation

discharge header

Fig. 28.25 Acceptable piping arrangement frequently used for large water pump installations.

running pump, and seizure as the pump runs down to rest. The same fine material can also cause
premature shaft seal failure, and rapid casing erosion in regions where centrifugal force tends to raise
the local solids concentration.
Normally the preventive measures to be taken during plant start-up would be discussed in Chapter
29, but because two of the principal measures, the provision of suction strainers and the means of
circulating liquid for flushing, are part of the suction system design, they are discussed in this chapter.
The actual method used in cleaning the suction piping varies considerably in different plants. The
essential ingredients in all cases, however, are forced circulation of liquid through the system and the
use of temporary strainers, located at strategic points, to collect foreign matter before it enters the pump.
Forced circulation is by either a low-head pump if one is installed in the system, or a temporary "cleanup"
pump. Generally, the cleaning starts with a thorough flushing of the suction vessel. It is preferable to
flush all the piping with hot water, as more dirt and mill scale can be loosened at higher temperatures.
Some plants use a hot phosphate and caustic solution for this purpose.
Temporary screens or strainers must be installed in the suction line as close to the pump as possible.
The choice of the size of openings is difficult to make because the foreign matter likely to cause trouble
falls into two categories: pieces large enough to become lodged in the impeller or casing, and particles
fine enough to enter the running clearances, which typically range from 0.15 to 0.35 mm (0.006 to 0.014
in.) per side. High concentrations of very fine particles pose a special problem, since some solids will
probably settle in the clearances and cause damage when the pump is started. If an 8-mesh screen with
0.025-in. wire is used, the openings are 2.5 mm (0.1 in.), small enough to remove pieces that could
become lodged in the pump, but too large to remove particles small enough to cause difficulties in the
pump clearances.
The safest solution is to use a strainer with a 40- to 60- mesh screen and flush with the pump
stationary, until the strainer remains essentially clean for half a day or longer. After that a coarser mesh
can be used if a higher rate of circulation is necessary. It is very important to turn the pumps by hand
both before and after flushing to check whether any foreign matter has washed into the clearances. If
the pump is stiff to turn after flushing, it must be cleared before operation.

Types of Strainers
Although conventional flat screen strainers (Fig. 28.26) placed between two flanges are used very
frequently to catch foreign matter and are inexpensive, they are not really economical. Considerable
expense is involved in removing, cleaning, and reinstalling such strainers, and pipe fitters must be on
duty to break and make the joints.
A more economical form of temporary strainer is the inverted cone or "witch's hat" (Fig. 28.27),
Installation 853

2 ' 2 MESH - -·

12' 12 MESH

Fig. 28.26 Flat screen type suction strainer.

--++-t--l~ FLOW

4 x4 WIRE MESH 20 x20 WIRE MESH

Fig. 28.27 Inverted cone-type suction strainer.

which is again installed between two fianges. Compared to a fiat screen, this design has a much larger
surface area, and because the water fiow impinges on the outside of the cone at a relatively acute angle,
the strainer is to a certain extent self-cleaning.
An even more sophisticated design, typical of that used for a permanent strainer, is shown in Figure
28.28. It has a sloping screen, which affords a lower pressure drop, and therefore longer periods between
cleaning than the conventional fiat screen strainer. The design shown is equipped with a blow-down
valve for removing accumulated foreign matter. Thus, in many cases, dirt accumulation can be dropped
out by opening the blow-down valve while the system is running. To allow back-washing of the screen,
this strainer is provided with a capped port downstream of the screen. With this provision, it is possible
to clean the strainer in an outage of 10 min instead of the 4 hr or so required for the fiat screen or
inverted cone strainers.
If the size of the free-straining area is properly selected to minimize the pressure drop, the start-up
854 Installation

EASILY CLEA EO WITH


AIR OR WATER LANCE

Fig. 28.28 Strainer with blow-down and back-wash.


(Courtesy Leslie Co .)

strainer may be left in the line for a considerable period before the internal screen is removed. Alternatively,
the entire unit can be removed and replaced with a spool piece. The strainer is then available for the
next initial start-up.
Unless the system is thoroughly flushed before starting the pump, the use of a fine mesh screen may
cause trouble. For example, 40-mesh screening with O.015-in. wire leaves only O.25-mm (O.OlO-in.)
openings, which would clog rapidly with any significant concentration of fine solids in the liquid.
Pressure gauges must be installed both upstream and downstream of the screen, and the pressure drop
across it watched very carefully. When the pressure drop across the screen is around two times the drop
with the screen clean, the pump should be stopped and the screen cleaned.

DISCHARGE PIPING

Piping size is determined by a balance between the capital cost of the piping and the energy expended
overcoming the friction loss in the piping. With pump discharge nozzles typically sized for a nominal
velocity of 7.5 mls (25 ftls), the piping size is usually at least one size larger than the pump discharge.
The piping configuration depends on where the liquid is to be discharged, and the amount of thermal
expansion that has to be accommodated to keep piping loads within the pump's capability.
Generally both a check and stop valve (usually a gate type) are installed in the discharge line. The
check valve is placed between the pump and the stop valve and protects the pump against reverse flow
in the event of unexpected driver failure. The stop valve is used when priming the pump, or for isolating
it for inspection and repairs. When the discharge run is short and there is only one pump installed, the
Installation 855

check valve can be deleted. If isolation is not required, such as in an irrigation application, the stop
valve can also be deleted.
Stop valves are either manually operated or motorized for remote control. Manually operated valves
that are difficult to reach should be fitted with a sprocket rim wheel and a chain, or bevel gears and a
lineshaft connected to a floor-mounted hand wheel.

PIPING LOADS

Unless a pump's piping is flexibly connected to each nozzle or the pump is resiliently mounted, it is
subjected to some loading from the piping. As a result, the design and therefore the cost of the piping
are affected significantly by the pump's piping load capacity. Piping loads beyond the capacity of the
pump have three possible effects:

1. The nozzles can be stressed to the point where they bend or break.
2. The casing can be distorted so much that there is rubbing contact within the running clearances, or misalign-
ment through the seals and bearings, or both.
3. Distortion of the pump's feet and baseplate can be enough to produce gross coupling misalignment in
separately coupled pumps, with the consequent risk of failure of the coupling, shaft, bearings, or seals. The
piping load capacity of a centrifugal pump therefore depends on its configuration, casing material, internal
clearances, and baseplate.

Most centrifugal pumps will accommodate moderate piping loads, but it must be appreciated that
"moderate" in this case is well below the loads based on allowable pipe stress. Medium and large axially
split double-suction pumps with cast-iron casings have the least capacity and should have their piping
independently supported (Fig. 28.29) to minimize the loads imposed on the pump's nozzles. Industry
standards for chemical pumps (AS ME B73.1-1991 [9.3]) and refinery pumps (API-61O[3.1]) include
allowable piping loads that pumps meeting these standards must be able to accommodate. The loads in
API-61O are based on overhung pumps, which are effectively two-point supported (see Chapter 2).
Between bearings pumps, or even overhung pumps with an elongated suction nozzle and long casing
feet, are four-point supported, and so generally capable of being designed for loads up to four times the

ECCENTRIC
REDUCER

Fig. 28.29 Correct method of installing suction and discharge piping for medium and large
axially split double suction pumps.
856 Installation

I ,&OG Le
DISCHARGE
NOZZLE
,.....""""--11,... 1,800 LB

Lf----r-....... 3,200 lB
60.000 IN.L ~~~:
!,ZOOll

3,200 I I

~
l8
1,600l1

30,000 IIII lB
3,200 lB
o,ooOIN.lB 1.600 LB

Fig. 28.30 Diagram of allowable piping forces and moments for multistage pump with side suction
and discharge and an axially split casing.

API allowable. Double casing or barrel pumps of up to say 10 stages, have high stiffness casings and
well-separated feet, properties that allow them to withstand very high piping loads. The objective of
having the pump accommodate higher piping loads is lowering the cost of the connected piping. In these
instances, the pump manufacturer supplies the user with information on the maximum allowable piping
loads. Typical diagrams are illustrated in Figure 28.30 for an axially split multistage pump with side
suction and discharge nozzles and in Figure 28.31 for a radially split double-casing pump with top
suction and discharge nozzles.

EXPANSION JOINTS

Expansion joints are sometimes used in the discharge and suction lines of centrifugal pumps, to avoid
transmitting any piping strains to the pump, whether these strains are from expansion during handling
of hot liquids, misalignment of the piping, or any other cause. On occasion, expansion joints are formed
by looping the pipe, as is customary with steam piping. More often, expansion joints are of the slip joint
or corrugated diaphragm type. They eliminate piping stresses, but introduce an entirely new problem,
namely, a reaction and a torque on the pump and its foundations. Unless they are applied correctly,
therefore, expansion joints can cause greater problems than they alleviate. Precautions must be observed
in the design of the piping and positioning of the expansion joints so that reactions due to flow and
pressure conditions are absorbed by the strategic placing of anchors, hangers, and bolts controlling
movement. Normally, pumps, their drivers (steam turbines), and baseplates will withstand certain limited
reactions; however, special study and calculation of these reactions is important.
When analyzing expansion joints, it is necessary to remember that action equals reaction. For example,
if an expansion joint is located vertically ahead of the suction elbow (Fig. 28.32), the downward force
Installation 857

Z Z

Y
Y A'
J ~ I Z-~ --l
1// I
Z- -;K'- --z ~

Z z r/: y y

I~
y

x y
x y

t ~-.-
y y
- - - ----FI"I-Z

Fig. 28.31 Allowable piping forces and moments for multistage pump with top suction and discharge
and a radially split double casing.
Allowable forces, in kN (lb)
X Y Z
Suction 70.2 (15,800) 70.2 (15,800) 49.8 (11,200)
Discharge 85.8 (19,300) 70.2 (15,800) 49.8 (11,200)
Total 85.8 (19,300) 70.2 (15,800) 49.8 (11,200)

Allowable moments, in kNm (lblt)


X-X axis Y-Yaxis z-z axis
Suction 31.0 (22,900) 61.5 (45,400) 30.5 (22,900)
Discharge 29.3 (21,600) 58.5 (43,200) 29.3 (21,600)
Total 31.0 (22,900) 61.5 (45,400) 30.5 (22,900)

will be the product of the pressure (2.07 bar or 30 psig) and of the area on which this pressure acts in
the expansion joint. This area can generally be approximated by assuming that the pressure is applied
over an area corresponding to the mean diameter of the expansion joint element or corrugated bellows.
In the case illustrated in Figure 28.32, if the pipe is ND 200 (8 in), the expansion joint would have 337
mm (13.25 in) OD and 219 mm (8.62 in) ID in the corrugations. The effective area would be approximately
600 cm2 (93 in2) and the downward force 2.07(600)/lO2 or 12.4 kN (30 x 93 or 2,790 lb). Some engineers
feel that this force should be further corrected by the spring constant of the expansion joint, which is
3.5 kN for 13 mm (800 lb for V2 in) expansion of each corrugation. However, the displacements are
never of the order of magnitude where this factor would be significant, and, therefore, it can be neglected.
This 12.4 kN (2,790 lb) force, acting downward, applies a couple to the pump and may twist the pump
on its foundation if no provision has been made to counteract the force.
858 Installation

SUCTION$=$~ 00 MIG .--_ _.

EXPANSION JOINT

Fig. 28.32 Expansion joint in a suction line.

Similarly, when an expansion joint is located in the 6-in. discharge pipe in Fig. 28.33, the 6.9 bar
(100 psig) pressure is acting over the approximately 342 cm2 (53 in2) effective area of the expansion
joint in the discharge pipe, and the horizontal reaction against the pump is 23.6 kN (5,300 lb). If there
is no expansion joint, the pump and the pipe will pull against each other, and if the flange bolts hold,
there is no stress on the pump, its foundation bolts, or the foundation. But with an expansion joint
interposed between the pump and the piping, the 23.6 kN (5,300 Ib) act on the pump and tend to pull
it off the foundation bolts. Depending on the size and number of these bolts, the stress on the foundation
bolts may become excessive.
Expansion joints are not generally recommended for high-head pump installations because the reaction
forces rise rapidly with increased sizes and pressures. With a ND 600 (24 in) pipe, an effective corrugation
area of 3,874 cm2 (600 in2), and a 11.0 bar (160 psig) pressure, the reaction force would reach 427 kN
(96,000 Ib or 48 tons). The casing would have to be almost entirely sunk and anchored in concrete to
hold it in place.
A vertical pump with side discharge (Fig. 28.34) will have a side force equal to the discharge pressure
times the effective expansion joint area, or

100PSIG

EXPANSION
JOINT

Fig. 28.33 Expansion joint in a discharge line.


Installation 859

1.38 bar x 5,800 cm2 = 80.0 kN in metric units

and

20 psi x 900 in2 = 18,000 lb in US units

The vertical force upward on the elbow is balanced by the hydraulic thrust, as far as the plate support
is concerned. Therefore, the only downward force on the supports is the dead weight of the unit. On
the other hand, the thrust on the motor bearings is the dead weight plus the hydraulic thrust.
Figure 28.35 illustrates some problems introduced by the use of expansion joints. The high-head
(67-99 m [220 to 325 ft)) vertical pumps in this installation were provided with a sleeve-type flexible
pipe connection on the downstream side of the cone-type discharge valves. The valves, the pumps, and
the pipe fittings between them were subjected to a high hydraulic reaction. To transmit this reaction to
the foundation, the 45-deg elbows were substantially anchored and the pump casings were securely tied
into the foundations by casting the concrete up around part of the pump casting as illustrated. On the
largest, highest head pump this was still insufficient anchorage, and the flexible pipe connection had to
be replaced with flanged fittings.

VENTING AND DRAINAGE

Provisions for venting are required when (1) the pump operates with a suction lift and has to be primed
(see Chapter 25), (2) the pump has a positive suction head but is not "self-venting", or (3) there is a
risk of accumulating gas or vapor at the pump's suction during operation. A valved vent connection in
the high point(s) of the casing is the usual practice in cases 1 and 2. Vents are piped to a drain system
if that is permissible; to a point of safe disposal in the case of flammable or toxic liquids, or to the
suction vessel when the suction is close to or under a vacuum. The last requirement is necessary to

MOTOR

HYDRAULIC
THRUST

Fig. 28.34 Expansion joint in the side discharge piping of a vertical pump.
860 Installation

Fig. 28.35 Expansion joints in an installation of vertical pumps.

avoid "air-in" leakage while the pump is being filled with liquid. In case 3, common with pumps handling
condensate in steam power plants and light hydrocarbons in refineries, the venting must be continuous
and is therefore piped back the pump's suction vessel (Fig. 28.36). Vent connections can usually be
avoided when the suction pressure is on the order of 15 bar (220 psig) and higher. At such suction
pressures, the volume of air or vapor remaining in the casing once the pump is at suction pressure is
generally not sufficient to cause any startup or operating problems.
With pumps classified as "self-venting," it is only necessary to vent the discharge piping upstream
of the check and stop valve. Such pumps have a casing configuration that allows the removal of sufficient
gas or vapor through the discharge nozzle to ensure the pump is primed and will therefore pump when
started. In most designs, this means the impeller eye is submerged. The advantage this confers is the
elimination of an additional vent on the pump casing. End suction pumps with a centerline discharge
(Fig. 2.14) are considered self-venting at any suction pressure above atmospheric because the volute
tongue or casing cut-water is located well above the top of the impeller eye. End suction pumps with
top tangential discharge (Figs. 1.9 and 13.13) are effectively self venting when the suction pressure is
above 1.0 bar (15 psig).
Drain connections, either plugged or valved, are provided in the casing when it is necessary to remove
liquid still in the casing before opening the pump. Drains are piped to a point where the liquid can be
disposed of or collected for reuse, depending on whether or not it is worth reclaiming.
Installation 861

OPEN VENT LINE


BACK TO SUCTION
VESSEL VAPOR
SPACE

Fig. 28.36 Suction nozzle vent to prevent vapor or gas accumulation.

WARMUP PIPING

Three operating circumstances require the installation of warmup piping to ensure the temperature of
standby pumps is kept close to operating temperature: the risk of pump damage from thermal shock,
transient differential thermal expansion, and thermal distortion. Thermal shock is generally not a concern
with the usual pump materials at operating temperatures below 175°C (350°F). At higher temperatures,
and at even lower temperatures with materials prone to thermal shock, the pump casing should be slowly
warmed up at a rate no greater than 40°C (75°F) per hour until the temperature is within 55°C (lOO°F)
of pumping temperature. Transient differential thermal expansion depends on the detailed design of the
pump. It can produce overstressing at radial or axial casing fits, overcompression of mechanical seals,
or in the worst cases, contact between the rotor and casing. Thermal distortion is a risk in any standby
pump when the pumping temperature is more than 55°C (lOO°F) above ambient. The principal cause
of difficulty is minor leakage past check or stop valves.
Circulation of warmup flow is typically accomplished with one of three arrangements. If the pump
operates with a positive pressure on its suction, the pumped liquid can be permitted to drain out through
the pump and a casing drain connection to some point in the cycle at a pressure lower than suction
pressure (Fig. 28.37). Alternatively, the flow can be taken from the discharge header, circulated through

TO POINT OF
1---L_\j--t>C)oI~ LOWER PRESSURE
FLOW
CONTROL
ORIFICE

FLOW FROM SUCTION HEADER, THRU DRAIN CONNECTION


BACK TO POINT OF LOWER PRESSURE.

Fig. 28.37 Wann-up flow from suction header, through drain connection back to point of lower pressure.
862 Installation

IF REQ'D, FLOW ~ 2".


FLOW CONTROL ORIFICE,

RATED; DAVIDSON [1.5J

FLOW FROM DISCHARGE HEADER, THRU DRAIN


CONNECTION BACK TO SUCTION.

Fig.28.38 Warm-up flow from discharge header, thru drain connection back to suction.

the casing, and out into the suction header (Fig. 28.38). The need for an orifice to regulate the warmup
flow with circulation from the suction depends on the actual differential pressure and flow required. An
orifice is usually required with circulation from the discharge header. The third arrangement is employed
in applications where the system is brought up to temperature before the pump is started. Warmup flow
is generated by a small circulating pump taking its suction from the pump's suction vessel, discharging
through the pump and back to the suction vessel (Fig. 28.39).
The warmup data and the three piping arrangements shown are typical. Depending on the pump's
design, it is sometimes necessary to pass flow through more than one region of the casing to properly
warm the pump, therefore making the piping more complicated. Always consult the pump manufacturer
for specific warmup recommendations, then follow the recommendations.

BYPASS SYSTEM

A bypass system is necessary when the pump's startup or operating mode can cause it to operate at
flows that will cause rapid degradation of the pump, system over-pressure, driver overload or air-binding.
With respect to rapid degradation of the pump, the minimum allowable flow is determined by one of
the following: (1) the risk of premature impeller erosion; (2) mechanical considerations such as vibration,
rotor deflection, and bearing loads resulting from hydraulic forces developed within the pump; or (3)
the risk of flashing within the pump as the temperature rise increases with decreasing flow. Chapter 22 gives
a detailed treatment of all the factors that influence the minimum allowable flow of centrifugal pumps.
In some cases, a permanent bypass is used to raise the pump's gross flow to a value high enough for
continuous operation. This arrangement wastes energy, but might be justified in applications requiring
a low flow at very high heads, or when a pump is going to run at a low flow for say a year until the
system or process requires a higher flow.
Once the minimum allowable flow has been determined, there are five fundamental points to be
observed in designing the system:
Installation 863

/:,...-L-L-+-...,
f-L-....L..-----..,,,....---'
~TCSG

-- - - \---- - - -I

CASING INSULATED, TOP & BOTTOM

Fig. 28.39 Wann-up by forced circulation from suction vessel, through pump and back to suction vessel.

Bypass Flow Return


Nonnally, the bypass flow is returned to the pump's suction vessel (Figs. 28.40 and 28.41) where
the heat it gained in passing through the pump and flow control orifice can usually be dissipated without
significantly raising the temperature at the pump suction. On occasion, with pipeline pumps for example,
it is not practical to return the bypass flow to the pump's suction vessel. The solution in these cases is

THESE VALVES MUST BE GOOD FOR


FULL PUMP DISCHARGE PRESSURE
AND MUST ALWAYS BE FULLY OPEN

-
OR FULLY CLOSED.
TO BOILER

fl~
X w
Z ~
~ ~

IJ-- ~~
-
0.....J
\.JI--I- - - - - I 17 ~
7.";~~I~t=r--:BOILER 0..
~--~ FEED
PUMP

Fig. 28.40 Manual bypass with flow back to pump suction vessel.
864 Installation

ANTI-FLASH ORIFICE
/

FLOW CONTROL ORIFICE


/
FLOW SWITCH OR CONTROLLER
'\
s:------ ___ J ,-_I

FLOW MEASURING ORIFICE . / '

Fig. 28.41 Boiler feed pump with automatic bypass to suction vessel.
Flow measurement at pump suction. Bypass may be on/off or modulating.

I. .
·f------BYPASS'LOOP'------... -I
FLOW CONTROL ORIFICE

~YPASS FLOW SWITCH


VALVE /
/ .--- ---5
-{)- )'----'
D

S FLOW MEASURING ORIFICE

Fig. 28.42 Automatic bypass to suction piping.


Used when it is not practical to return bypass flow to suction vessel. Bypass is automatic on/off
with flow measurement at pump discharge.
Installation 865

to return the flow to some point in the pump's suction piping (Fig. 28.42). In doing this, it is important
to recognize that when the pump is operating on the bypass alone, all its power goes into heating the
liquid in the bypass "loop". The plant designer's task is to make the "loop" large enough to ensure the
temperature in it stays below the allowable over the longest time the pump is expected to run on the
bypass alone. This is done by locating the point where the bypass returns to the suction piping far enough
upstream of the pump to provide the necessary liquid mass.

Piping Size
Make the piping large enough for at least 25 percent more than the stated minimum flow in case
field experience or a change in operating conditions requires that the flow be increased.

Flow Control
Since the bypass is necessarily returned to a point of pressure lower than that at the pump discharge,
there must be some form of flow control element in the bypass line to limit the flow to the required
minimum. And of equal importance, there must be one flow control element per pump. It is often
tempting, for potential cost savings, to provide one flow control element for two pumps on the rationale
that only one pump will be running on the bypass at any given time. Experience has confirmed that this
rationale is false; that inevitably a time comes when both pumps are running on the bypass together.
Because the head characteristic of most centrifugal pumps is relatively flat at low flows, and the variation
in head from one pump to the next is greater than it is around design capacity, one pump invariably
produces more head and therefore drives the other pump to zero flow. The pump running at zero flow
quickly fails from mechanical damage or overheating.
Flow control elements range from a simple plate orifice, usually limited to 7 bar (100 psi) pressure
drop, through bar-stock orifices (Fig. 28.43) for intermediate pressure drops, to multiple pressure reducing
orifices for high pressure drops. Multiple pressure reducing orifices may be a separate component (Fig.
28.44) or integral with the bypass valve (see later discussion). The liquid velocity at the discharge of

l-is IN. ~ IN. ~IN. ~IN. IN. 3Ja


1250 ,---r-r----y-,...---.--r--r----r--r---.....",----,

I::
~
~ 7~ ~~~_+--~~~-~~-~~~~

o
; ~ ~~-h~--~--h~-~~-r-~r-~~&;~
250 HH~+-+~__:~-_+__ _+_-~.I. .AND
. . ~:~
IN. FOR \.i4 IN.,
o ~_~__~_~_~_~_~~~~~.~M~~~lk~
o ~ ~ ~ ~ ~ ~ M
CAPACITY. IN GPM

Fig. 28.43 Bar-stock orifice; selection chart.


866 Installation

••••
SECTION
A-A
SECTION
B-B
SECTION
C-C
SECTION
D-D

Fig. 28.44 Multiple pressure-reducing orifice.


Designed to avoid erosion from fluid flow at high differential pressures.

PIPE PLUG

TEE OR
PIPE COUPUNG

02
1--
.,0
NI') TO HEATER OR
SOURCE OF SUPPLY

__- --COUPLING OR FLANGE CONNECT10N


~-----~/

ORIFICE - STAINLESS STEEL ROO

PIPE SIZE

Fig. 28.45 Piping arrangement for bar-stock type bypass orifice.


Installation 867

bar-stock orifices can be high enough to cause impingement erosion of an elbow immediately downstream
of the orifice. If the piping must bend immediately downstream of the orifice, the arrangement shown
in Figure 28.45 reduces the risk of erosion. The pipe plug facing the orifice discharge should be stainless
steel to raise its erosion resistance.

Bypass Control
The bypass control parameter must always be flow, whether the valve is manually or automatically
operated, since this is the aspect of pump operation the bypass is intended to protect. Periodically, the
suggestion is made that discharge pressure is a simpler means of monitoring a pump's flow. This is not
really the case for the following reasons: (1) discharge pressure is affected by suction pressure, (2) the
pump's pressure rise varies with manufacturing variations and the condition of internal clearances, and
(3) when the head characteristic is relatively flat, a small difference in pressure represents a significant
change in flow.
At its simplest, the bypass valve is manually operated (Fig. 28.40) and is either fully open or fully
closed. Such an arrangement requires that the operator monitor pump flow and open or close the bypass
valve as necessary. For small to moderate bypass flows, the on-off operation can be automated by using
a flow switch to monitor pump flow (Figs. 28.41 and 28.42). The switch typically acts on the differential
pressure across some form of primary element, such as a plate orifice, installed in the pump suction or
discharge piping. An alternative approach is to use what is commonly termed an "automatic recirculation
control" or "ARC" for short. These devices (Fig. 28.46) are a vertical check valve with an integral
bypass valve and multiple pressure reducing orifice. The position of the check valve piston, which varies
with flow, is followed by a lever mechanism that opens the bypass when the discharge flow falls to the
minimum allowable.
As bypass flows increase, a point is reached where there is a risk of damage to the system from
hydraulic shock if the bypass valve is fully opened quickly. This difficulty is overcome by modulating
the bypass flow, using a flow controller instead of a switch, so the gross flow through the pump is always
at or just above the minimum allowable. The valves used for this (Fig. 28.47) are variable orifice devices
in which the orifice flow area increases as the valve plug is moved farther away from the seat.

Anti-Flash Orifice
Most bypasses require an anti-flash orifice at the point of entry into the vessel to which the bypass
is returned (Figure 28.41). Failure to provide this orifice can cause flashing in the latter stages of the
pressure reducing device, with consequent rapid cavitation erosion of that device.

BALANCING DEVICE LEAK-OFF

Whenever a multistage pump is equipped with a balancing device to counteract the axial thrust developed
by a group of single suction impellers all facing in the same direction (tandem impellers; see Chap. 6),
it is necessary to return the balancing leak-off into the cycle at a point where the pressure is equal to
suction pressure, or in some cases, below suction pressure. In this case, the pumped liquid is subject to
a definite temperature rise at low flow operation, and the balancing device leak-off itself undergoes a
further temperature rise as it flows through the balancing device because of the degradation of pressure
energy. This leak-off, if returned directly to the pump suction, will mix with the incoming liquid, thereby
raising the temperature of the mixture and lowering the NPSH available at the first stage impeller. If
the pump is handling liquid at or near its boiling point, it is therefore safer to return the leak-off to a
point in the cycle where the pressure is equal to suction pressure but where the added heat can be
868 Installation

Fig. 28.46 High pressure Automatic Recirculation Control (ARC) valve.


Incorporates a check valve, flow sensing device, bypass control valve, and pressure letdown device into a single
body. Specifically designed to protect high energy centrifugal type pumps, such as boiler feed pumps.
(Courtesy Yarway Corporation.)

dissipated. For example, in the case of a boiler feed pump taking its suction directly from a deaerating
heater with a relatively low NPSH margin, the balancing device leak-off should be returned directly to
the heater rather than to the pump suction proper. The same applies to hydrocarbon pumps taking suction
from a vessel at saturated conditions or handling light liquids such as propane, ethylene, or ethane. Note,
too, that in light hydrocarbon service, there is some autorefrigeration as the leak-off is "expanded" across
the balancing device, and so the temperature rise is lower than calculated from simple heat input. Under
these conditions, however, there is a certain volume of vapor in the balancing leak-off flow, which is
better piped back to the suction vessel where it can be recondensed.
The cases where the balancing leak-off flow is returned to a point in the cycle lower than suction
pressure are those where there is a suction booster pump between the suction vessel and the multistage
pump, and it is necessary to lower the pressure at the multistage pump's seals. This is common in boiler
feed pumps with condensate injection seals (see Chap. 10).
There are four important aspects to the design of balancing leak-off piping. First, the piping must be
the size shown on the pump outline drawing, and it must be piped to the point specified on the drawing.
Second, there should be not more than one valve in the line, a stop valve for isolation when necessary,
and that valve should be locked open and its being verified as open made a hold point in the pump's
start sequence. (If the pump is started with the balancing leak-off line closed, the pump's thrust bearing
will fail immediately, with the risk of consequent damage within the pump, and the balancing leak-off
Installation 869

Fig. 28.47 Modulating bypass control valve with stacked plate type orifice.

piping may rupture. In some regions, the balancing leak-off line would have to be rated for pump
discharge pressure back to the stop valve, or equipped with a full-flow relief valve.) Third, an antiflash
orifice is often necessary to prevent flashing in the balancing device and the line back to the suction
vessel (see Chap. 10 for a detailed discussion). Four, for boiler pumps in cold climates, provisions must
be made to prevent freezing if the balancing leak-off line passes outside the turbine house. (There have
been instances of a balancing leak-off line freezing when a boiler feed pump was worked on in winter.)
Multistage pumps with opposed impellers (Fig. 3.14) also have what is often termed a balancing
leak-off flow. Whether this flow is associated with balancing axial thrust or just lowering the pressure
at the adjacent shaft seal depends on the detail design of the pump (see Chapt. 5). When the manufacturer's
outline drawing states that this flow must be returned to the pump's suction vessel or some other point
of lower pressure, the same design precautions necessary as in the balancing leak-off from a tandem
impeller pump. The one difference is that the total temperature rise of the liquid is lower because the
head being broken down is lower.

INSTRUMENTATION

A number of instruments are essential to maintain a close check on the performance and condition of
operating centrifugal pumps. At the least, two pressure gauges should be provided, one connected to
the pump suction and the other to the discharge. Pressure taps are generally provided in the suction and
discharge flanges for this purpose. The gauges should be mounted where they can be easily read.
When the pump is motor driven, the starter should include an ammeter, which provides an indication
of pump power. Variable-speed pumps need a tachometer so their running speed can be checked.
More elaborate instrumentation is warranted for pumps in critical or severe services (see Chap. 30
for a detailed treatment of the instrumentation needed to monitor such pumps).
In many installations, additional instrumentation is added for process control. The instrumentation
870 Installation

needed depends on the process and the means of control, aspects of the installation design that are
beyond the scope of this text.

FIREWALL EXTENSIONS

When a pump handles a flammable volatile liquid such as gasoline, there is danger of an explosion
unless an explosion-proof driver is used. An alternative approach to an explosion-proof driver is to
isolate the driver from the hazardous environment by locating the pump and its driver in separate rooms,
with a firewall between them. This requires the use of a shaft extension with a seal so the driver can
be connected to the pump through the wall. Various types of firewall extensions have been used, some
of which are shown in Figure 28.48. These firewall extensions have a packed-box-type seal at each end.
The space between the seals is usually filled with a light grease to form a vapor tight seal.
Modem plants have most of the pumps handling flammable and volatile liquids located outdoors, so
the use of firewall extensions is generally limited to marine applications.

NOISE CONTROL

Noise is objectionable sound, and is detected by the human ear as small fluctuations in ambient air
pressure. The magnitude of these fluctuations is termed sound pressure level (SPL). It is expressed as
the logarithm of the ratio of the measured pressure to a reference pressure, in this case 2 x 10-5 N/m2,
and is known as decibels. The human does not respond equally to all sound frequencies, so a weighting

Gr111~
PUMP

Fig. 28.48 Firewall extensions.


Installation 871

is applied to approximate the response. For almost all industrial sound measurement, the weighting used
is "A." Sound measurements made with this weighting are reported as dB(A).
The noise in the vicinity of an installed centrifugal pump is a function of (1) the noise generated by
the pump and its driver equipment, (2) noise radiated from the pump's suction and discharge piping
(and bypass and balancing leak-off if installed), (3) amplification by resonance of panels in the pump
unit or adjacent structures, and (4) reflection from the adjacent structural surfaces. Of these four sources,
the only one the pump manufacturer can exert any control over is the pump and its driver equipment.
Given this circumstance, a pump manufacturer can provide some assurance of the noise the pump and
its driver will generate, but cannot guarantee the noise level in the vicinity of an installed pump. This
means that the design of the installation must take into account noise control if the pump is likely to
generate high noise or the surrounding environment is sensitive to noise. The design of the installation
also has to consider the effect of structure and liquid borne noise, originating at the pump, but manifesting
itself as vibration, thence airborne noise, elsewhere in the system.
The noise emitted by a centrifugal pump is related to the following factors:

1. Power level
2. Rotative speed
3. Clearance between the rotating (impeller) and stationary (volute or diffuser) vanes
4. Hydraulic design
5. Vibration
6. Operating capacity relative to BEP
7. Upstream turbulence
8. NPSH margin

Of these items 2 through 4 are directly under the pump manufacturer'scontrol, and items 5 and 6 can
be influenced significantly. Item 1 is a function of the conditions of service and can only be affected
greatly by changing the number of pumps used to produce the required flow. Items 7 and 8 are very
much under the control of the plant designer. It is now appropriate to discuss each of these groups in
more detail, starting with the pump and dealing first with some details of sound measurement.
SPL varies with distance from the source, whereas sound power level, designated PWL, does not.
As a consequence acoustic engineers prefer to work in PWL. Sound is therefore measured in SPL and
is converted to PWL using the relationship

PWL = SPL' + 10 10g(S/So)

where SPL' is the mean value of the sound pressure level measurements made at points on the imaginary
surface around the machine, corrected for extraneous noise and room reverberation, S is the area of the
imaginary surface in square meters, and So is a reference surface area of 1 rn2•
In the absence of more definitive data from the pump manufacturer, an estimate of the SPL' a pump
is likely to produce operating at BEP, in dB(A), 1 m from the casing surface and 1.5 m above the floor,
can be made from the following correlations:
For one and two stage pumps

SPL' = 62.3 + 10 log Power

and for multistage pumps (three stages and higher)

SPL' = 59.3 + 10 log Power


872 Installation

where Power is the pump power at BEP in kilowatts.


Operation away from BEP will increase the noise emitted; by approximately 2 dB at 50 percent of
BEP, 4 dB at 25 percent.
When the predicted airborne noise from the pump is higher than can be tolerated, there are three
options available:

l. Reduce the power of the pump by increasing the number of pumps used to produce the required flow. For
example, changing from a single IO,OOO-kW multistage pump to two 5,OOO-kW pumps will lower the noise
emitted by about 3 dB from 99 to 96 dB(A).
2. Apply high-density insulation, what is termed an "acoustic blanket," to the pump's casing and nozzles. This
typically yields a reduction of 4-5 dB, and is normally fitted during pump manufacture. In terms of cost
benefit, it is generally a better solution than using smaller pumps unless there are other justifications for
doing so.
3. Enclose the pump, and possibly its driver equipment, with an acoustic housing. Substantial reductions in
airborne noise emitted to the vicinity are possible with this approach, as much as 30 dB, but at the
expense of access for maintenance and the need to wear hearing protective during routine inspections of
the running pump.

Whether to enclose the driver equipment depends on the nature of the sound it emits. The dominant
component of noise generated by two-pole electric motors is cooling air unless the motor enclosure is
CACW (TEWAC). Cooling air noise can generally be reduced significantly with special treatment of
the motor's enclosure. With four-pole motors, the dominant component of noise tends to be magnetic,
usually at two times line frequency. In these cases, some form of acoustic enclosure is often necessary.
Many high energy pumps are driven through a step-up gear. Depending on the pitch line velocity of the
gear and the precision with which the gear and pinion are made, the noise generated can be high enough
to warrant enclosure. Steam turbines of the sizes used for pump drives are not noted for high noise
levels, and an "acoustic blanket" is usually sufficient treatment. Engines usually have to be enclosed to
meet modem industrial noise limits.
Upstream turbulence, or more accurately the lack of it, is extremely important in lowering the noise
emitted from pumps. It is so for two reasons. First, turbulent flow into the impeller is likely to produce
local cavitation and consequent air and liquid borne noise. Second, pumps inherently produce turbulence
in the action of developing head. When that turbulence is added to turbulence produced upstream of a
pump, the result is higher air and liquid-borne noise at the pump discharge. Laying out the suction piping
correctly, already dealt with in detail earlier in this chapter, is the first and most important step in
reducing noise from this source. Insulating the suction and discharge piping is a means of further reducing
noise from this source, or reducing noise caused by a poor original design.
NPSH margin can reduce pump noise by suppressing the cavitation often produced in an impeller
whose suction specific speed is too high (see Chap. 22), or the local cavitation produced when the suction
piping produces too much upstream turbulence, as just discussed.
Vibration of the pump and its driver equipment is a function of the machines' design, manufacture,
installation, and operation. The first two and the last of these are outside the scope of this section.
Installation, as it relates to design, is covered in foundations, while the important aspects of physically
installing the equipment are dealt with later.
When the pump installation is in an area that is very sensitive to noise, such as in a building or close
to a residential community, it is necessary to minimize the transmission of structure and liquid-borne
noise. If this is not done effectively, there is a good chance some element or elements of the system
will have a natural frequency at or close to a frequency in the noise being generated by the pump, and
so will become resonant. Resonant vibration is a source of airborne noise. The usual means of limiting
Installation 873

structure- and liquid-borne noise transmission is to isolate, to the extent possible, the pump from its
supporting structure and connecting piping.
Mounting the pump unit on springs (Fig. 13.4) is the most effective way of reducing the transmission
of its inherent vibration to the structure. Note in this connection, that any rotating machine, no matter
how well designed and balanced, will generate some vibration. When resonance exists and there is
negligible damping available, it takes very little energy to excite the resonant component. Note, too,
that the baseplate for units mounted on springs must be correctly designed if it is not to cause more
trouble than it corrects by allowing dynamic misalignment of the pump and its driver (see Chap. 13).
Installing flexible connectors in the pump's suction and discharge piping tends to attenuate pressure
pulsations produced by the pump. The degree of attenuation depends on the flexibility of the connector,
which varies, in tum, with the pressure that has to be contained. As a result, the higher the system
pressure, the lower the attenuation. The other aspect of flexible conilectors is the reaction unrestrained
connectors exert on the pump and piping (see Expansion joints in earlier in this chapter).

Installation
The following sections deal with physically installing the pump, its driver, and any accessories. Because
it is relevant to what has to be done in the field, the section begins with an outline of how the pump is
prepared for shipment.

PREPARATION FOR SHIPMENT

The general procedure in preparing pumps for shipment after manufacture is practically identical with
all manufacturers. If the pump has been shop tested, it is drained and its internals inhibited, generally
with a water displacing inhibitor. Bearings are inhibited with an oil-soluble compound. Small piping is
cleaned, inhibited, and sealed. Flanges and exposed machined metal surfaces are cleaned and treated
with an anticorrosion compound such as grease, vaseline, or heavy oil. To prevent damage and the
ingress of dirt, dust, moisture, or foreign material during shipment and erection, all nozzles, pipe flanges,
and tapped pipe openings are closed with blank flanges (of wood, plastic, or metal) or with screwed-in
metal plugs. The pump's rotor is blocked or strapped to prevent movement. Protective guards are installed,
if necessary.
When long-term outdoor storage is intended, containers of moisture-absorbing desiccant are suspended
inside the pump and its accessories. The equipment is then plastic wrapped, leaving the bottom open,
and enclosed in waterproof boxing. On occasion, for storage in tropical environments, some purchasers
have specified that the boxing include a manifold of oil mist piping to each bearing housing and the
pump casing so these regions can be purged while the pumps are in storage.
Usually the driver is delivered to the pump manufacturer, where it is assembled and aligned with the
pump on a common baseplate. The baseplate is drilled for driver-mounting, but the final dowelling is
left for the field, after final alignment. When size and weight permits, the unit is shipped in the assembled
state; that is, pump and driver on the baseplate. Sometimes, however, large drivers are shipped directly
to be mounted in the field. Then the baseplate should be drilled at the job site.
Sometimes it is impractical to ship the pump mounted on its baseplate, as when the suction and
discharge nozzles are located vertically downward and extend for an appreciable distance below the
baseplate, or if size or weight limitations exist. These pumps must be shipped separately or turned upside
down and temporarily secured to the baseplate.
874 Installation

CARE OF THE EQUIPMENT IN THE FIELD BEFORE USE

The equipment should be inspected and checked against the shipping manifest immediately on receipt
of the shipment. Any damage or shortage should be reported to the transportation company's local agent.
If the pumping equipment is received at the site before it can be used, one of the following courses
of action should be taken:

1. When the equipment has not been prepared for site storage but will be installed within say 1 month, it
should be immediately stored in a dry location. The protective flanges and finishes should be left on the
pumps. The bearings and couplings must be carefully protected against sand, grit, and other foreign matter.
Depending on the design and the manufacturer's specific instructions, the pump rotor should usually be
turned by hand at frequent intervals to prevent rusting or binding. Sometimes a rotor will become slightly
bound in storage. To free it, the thrust bearing should be removed or dismantled, and oil throwers, guards,
or other parts that may restrict axial movement should be loosened. Moving the rotor axially a few times
will free it at the wearing rings or wherever it has become bound at the internal clearances.
2. Equipment that has been prepared for site storage and will be installed within the specified storage period,
typically 6 months, should be stored as stated in the purchase specification.
3. When, for whatever reason, pumping equipment is going to be stored at the site for a longer period than
originally intended, the manufacturer should be consulted for specific recommendations. The essential matters
to be addressed are storage location, inhibiting the pump internals and bearings, closing openings, and
whether to tum the rotor periodically.

When the pump is finally installed, the protective coatings applied to external machined surfaces need
to be removed. It is not usually necessary to remove the oil-soluble inhibitor used in the bearings.

LEVELING AND ALIGNMENT

When a unit consisting of pump, base, coupling, and driver is assembled at the factory, the baseplate is
placed on a flat, even surface, the pump and driver are mounted on the baseplate, the pump is centered
and dowelled, and the machines are aligned, laterally by moving the driver, vertically by shimming
under the driver. Unless the baseplate has been specially designed (see Chap. 13), this factory alignment
cannot be maintained with sufficient accuracy to allow starting and operating the unit without realignment
in the field.
The first step in field alignment is to level the baseplate. The importance of this fundamental step
cannot be overemphasized, because if it is not done, the alignment of the machines will be difficult if
not impossible. As obvious as this may seem, there are still far too many instances of field alignment
problems whose root cause is failure to level the baseplate.
With units of moderate size it is usually not necessary to remove the pump or driver from the baseplate
to level it. Once the foundation has been prepared (surface roughened, locations of packers ground for
70 percent contact), place the unit over the foundation and support it on packers (short strips of steel
plate) and shims or wedges (Fig. 28.49[a]) close adjacent to the foundation bolts, or with the leveling
screws in the base's mounting flange (Fig. 28.49[bD. Set the pump's horizontal axes in the correct
position using site survey points, then check that there is sufficient space between the underside of the
base and the top of the foundation for grouting, typically 25-75 mm (1-3 in.). Remove the coupling
bolts (if not already removed before shipment). Clean the projecting machined surfaces of the equipment
pads, and using a precision machine level proceed to level the base until the equipment pads are level
Installation 875

l---~=?
PACKERS

(0) WEDGES

............ L.EVELLING
~ SCREW

(b) LEVELLING SCREW

Fig. 28.49 Means of leveling bases: a) wedges, b) leveling screws.

within 0.012 mm in 25 (0.0005 in per in) in both directions. Leveling is accomplished by shimming
under the base or adjusting the leveling screws, then retightening the foundation bolts.
Some small pumps are mounted on bases that not machined. These are leveled by checking that the
pump is level, using the shaft extension or a flange as a reference surface, then adjusting the baseplate
shimming to bring the coupling into coarse alignment.
Once the base is leveled, check the shaft separation between pump and driver, being careful to set
the rotor of each machine in its normal running position. Pump rotors typically do not have more than
0.40 mm (0.015 in.) of endplay. The rotors in large sleeve bearing motors, however, can have 12 mm
(0.5 in.) or more of endplay, and so have to be set at their magnetic center before checking shaft
separation. Magnetic center is usually located by a pointer on the motor's inboard bearing housing. The
running position of these rotors is controlled by a limited-end-float coupling.
The objective of coupling alignment is to align the axes of rotation of the two coupled machines.
This is best achieved by rotating the shafts of both machines so measurements at each of the four angular
positions (top, bottom, and sides) are made at the same point on the coupling hubs. In doing this, the
effects of errors in coupling hub squareness or concentricity, or shaft straightness are eliminated. Figure
28.50 illustrates the principle and shows the measurement methods typically used to achieve coarse
alignment.
Accurate alignment follows the principle of rotating both the shafts while improving the accuracy of
the measurements. The "rim and face" method (Fig. 28.51) replaces the feeler gauge and straight edge
with dial indicators to take measurements of axial and radial position as the shafts are rotated. With the
buttons resting on the other hub peripherary and face, the dials are set at zero and a mark made on the
hub at the point where the buttons rest. For any check (top, bottom, or sides) both shafts must be rotated
by the same amount so all readings of the dial indicators are made with the buttons on the mark. The
dial readings indicate whether the driver must be raised, lowered, or moved to either side to correct
876 Installation

. -----
SCRIBED LIN> /;-~
~/ 9"0· I "90•

l~)
:t,
1 '
90· I ...::-

~90.
,-··7 .... ,,~
~ . " .... _ -""1"

._- ---\
r ..... /'

.\
I

SCRIBED LINE ~
IN ROTATED POSITION

Fig. 28.50 Principle of coupling alignment.


Measure at 4 places, turning both shafts together.

- 1 - -- 1
f

CHECK FOR ANGULAR MISALIGNMENT CHECK FOR PARALLEL MISALIGNMENT

DIAL INDICATOR MEASURES MAXIMUM LONGITUDINAL VARIATION DIAL INDICATOR MEASURES DISPLACEMENT OF ONE SHAFT CENTER
IN HUB SPACING THROUGH 360· ROTATION. LINE FROM THE OTHER.

1. Attach dial indicator to hub, as with a hose clamp; 4. Reset pointer to zero and repeat operations 1 and
rotate 360· to locate point of minimum r.ading on 2 when either driven unit or driver is moved dur-
dial; then rotate body or face of indicator so ing aligning trials.
that zero rtading lin .. up with pointer.
5. Check for para 11 e 1 mi sa 1 ignment as shown. Move
2. Rotate both half couplings togtther 360·. Watch or shim units so that parallel misalignment is
indicator for misalignment reading. brought within the maximum allowable variations
for the coupl ing style.
3. Driv.r and driv.n units will be lined up when
dial indicator rlading comls within maximum allowa- 6. Rotate couplings several revolutions to make sure
ble variation for that coupling style. Refer to no "end-wise creeplt in connected shafts is
specific installation instruction sheet for the measured.
coupling being installed. NOTE: If both shafts
cannot be rotated together, connect dial 7. Ti ghten all locknuts or capscrews.
indicator to tht shaft that is rotated.
8. Recheck and tighten all locknuts or capscrews after
several hours of operation.

Fig. 28.51 Rim and face dial indicator method of checking coupling alignment.
(Courtesy Rexnord)
Installation 877

VIEW TOWARD PUMP


DIAL ON MOTOR HUB
ANGULAR MEASURED @ 6.0 IN RADIUS

o o

+0.005 -0.002 +0.001


+0.015

+0.020 -0.001

MOTOR 0.010 LOW MAX ANGLE = 0.014 DEG.


0.005 TO LEFT

Fig. 28.52 Example of rim and face alignment readings and interpretation.

parallel misalignment, and whether one end has to be moved more than the other to correct angular
misalignment. As an example, the alignment readings repeated in Figure 28.52 show that the angular
alignment is good (maximum error 0.014 deg), and that the motor is 0.250 mm (0.010 in) low and 0.12
mm (0.005 in) to the left.
Other alignment methods such as "reverse dial indicator" and "laser measurement" (Fig. 28.53)
achieve the same alignment accuracy, but with the aid of computer calculation routines can do so in
much less time.
When the unit outline drawing specifies shaft movements due to thermal expansion of one or both
of the coupled machines when at operating temperature, the "cold" alignment must include an offset to
compensate for this. Provision should then be made to check the coupling alignment when the machines
are at operating temperature. Traditionally this was done in a so-called "hot alignment" check, which

Fig. 28.53 Rotalign® laser shaft alignment system.


(ROTALlGfV® is a registered trademark of Prueftechnik AG. Illustration courtesy of Ludeca. Inc.)
878 Installation

--r-.,- --. ----


2.0 ,, I

l i I
1.8
1\ ,, I
I
,
1.6 1
I I

I I
I

c:: 1.4 1\, I

j I
I
'E
1.2
: [\' Unacceptable
~ : \
~ : I
1.0 I
: '" ""-
i
l I

\ : I
~
t 0.8
~
r
I
I
.......
"'" .......
r-.. 100.

"
0.6 I r- ..... ~
I
Acceptable
0.4 i
I
I ~
.......
r- ro- '"'-....
~
..... ~ ....
I I

,I I
0.2 , I Good
I

I I I I I
o 1 2 3 4 5 10 20
rpm x 1,000

Fig.28.54 Piotrowski coupling alignment recommendation [28.2].

involved shutting down the unit from full temperature and quickly checking coupling alignment. A more
effective approach is to measure machine movement from known cold reference positions and compare
these with the offsets incorporated in the "cold alignment." Both mechanical and laser techniques are
now available for doing this.
How accurate the alignment has to be is a function of coupling type, separation of the flexible
elements, and rotative speed (see Chap. 12). The values quoted by the coupling manufacturers are the
maximum the coupling can tolerate without suffering rapid deterioration or producing high reactions on
the connected equipment. Generally these values can be reduced by factors ranging from 0.5 to 0.1
without becoming impractical, therefore leading to lower reactions on the connected machines and longer
coupling life. Piotrowski, cited by Bloch [28.2], has recommended practical alignment accuracy in mils/
in. of flexible element separation versus rotative speed (Fig. 28.54).
The alignment must be rechecked after the suction and discharge piping have been connected to
verify that there is negligible effect from any imposed piping loads.

GROUTING

Ordinarily, the baseplate is grouted before the piping connections are made and before the final alignment
check. The purpose of baseplate grouting is to prevent lateral shifting of the base, to increase its mass
to reduce vibration, and to provide continuous support between the foundation and the baseplate.
Installation 879

Grout can be cementitious, basically one part pure portland cement and two parts building sand, or
epoxy resin. Most modem cementitious grouts are "nonshrink" and are furnished as premixed dry
materials needing only the addition of water, following the manufacturer's instructions, before being
ready for placement. Epoxy grouts are furnished as two parts, resin and hardener, for mixing in the field
just prior to use.
When using cementitious grout, the top of the rough foundation should be well saturated with water
before grouting. A wooden form is built around the outside of the baseplate to contain the grout, either
tight against the lower edge of the base or a small distance from it. For convenience in getting the grout
under the base, tin plate funnels are used at several points around the edge. Grout is added until the
entire space under the base is filled (Fig. 28.55). The grout holes in the base serve as vents to allow the
air to escape. A stiff wire is used through the grout holes to work the grout and release any air pockets.
The exposed surfaces of the grout should be covered with wet burlap to prevent cracking caused by
drying too rapidly. When the grout has sufficiently cured (about 48 hr) so the forms can be removed,
the exposed surfaces of the grout and foundation are given a smooth finish. When the grout is fully
cured (72 hr or more), the foundation bolts should be retightened and the coupling alignment checked.
In broad terms, the practice with epoxy grout is similar to cementitious, the major differences being
preparation of the underside of the base (it must be painted with a compatible material), preparation of
the foundation surface (it must be clean and dry), and the ability to rapidly place and work the grout
because of the high curing rate. Further practical guidance on epoxy grouting is given by Monroe [28.3].
Whether to remove the packers and shims or wedges or the leveling bolts after grouting has long
been a controversial topic. The key element in the differing opinions seems to be the background of the
erector. Those whose background is reciprocating machines are adamant that the packers, wedges, and
leveling bolts should be removed. This rests on ample evidence that for reciprocating machines, which
subject the foundation to significant shaking forces, it is mandatory to remove the packers and the like
to ensure that the machine base has uniform support.
On the other hand, many erectors whose background is turbomachinery see no need to remove the
packers. This, too, rests on practical experience; in this case, the general finding being that leaving the
packers in place does not cause any harm. To quite a degree, this rests on whether the grout shrinks
and the piping loads imposed on the pump. High grout shrinkage or high piping loads could lead to the
base being supported only at the packers (grout, whether cementitious or epoxy, is significantly less stiff
than steel), with a consequent risk of base deflection between.
To ensure there is no risk, the safer course is to remove the packers or leveling bolts after the grout
has cured and fill the voids with fresh grout.

Fig. 28.55 Application of cementitious grout.


880 Installation

PIPING

The installation of a pump's piping must reproduce, as closely as practical, the loads used in the design
of the pump and piping (see Installation design). Failure to do this can result in high piping loads and
the problems that often accompany that.
Except when "cold spring" is used in the design to reduce piping loads when the system is at operating
temperature, piping should align with the pump's nozzles without having to be forced into position. This
means there must be some provision for field cutting and welding to accommodate the inevitable variations
in the installed location of the pump nozzles, and in the dimensions of prefabricated piping spools (see
Provision for tolerances). It also means there must be some means of adjustment in any piping supports
located close to the pump. The practical test for correctly aligned piping is to check that the pump and
piping flanges are parallel when the piping is correctly positioned but not bolted up.

DOWELLING OF PUMP AND DRIVER TO THE BASEPLATE

Most pumps are dowelled to their baseplate during assembly in the factory, so they can not be subsequently
moved out of center during field erection, and so create an alignment problem. In the case of large
single-stage and all multistage high-temperature pumps, the dowelling is more elaborate because it has
to be strong enough to withstand high piping loads while allowing for thermal expansion.
Modem practice for multistage pumps is to dowell the casing at the inboard end, which is usually
the suction end of the pump, and leave the outboard end free to move. The rationale for this is to
minimize the change in shaft separation between the pump and its driver as the pump comes up
to operating temperature. Figure 28.56 shows the dowelling practice for a modem high-temperature
multistage pump.

CLEARANCE

EJ

DOWEL

Fig. 28.56 Dowelling of high-temperature multistage pump.


Installation 881

The driver is not dowelled in the factory because it has to be moved to achieve final alignment in
the field. In low-temperature services, the driver should be dowelled after the pump's piping has been
connected and the coupling alignment checked. For high-temperature services or when the driver has
significant expected shaft rise, such as with a steam turbine, dowelling the driver should be delayed
until after the unit's "hot alignment" has been checked and, if necessary, corrected.

BIBLIOGRAPHY

[28.1] I. J. Karassik et ai., "Centrifugal Boiler Feed Pumps Under Transient Operating Conditions," (Paper 53-F-
32), ASME Fall Meeting, Rochester, New York, 1953.
[28.2] H. P. Bloch, Use Laser-Optics for Machinery Alignment, Hydrocarbon Processing, pp. 33-36, October 1987.
[28.3] P. C. Monroe, "Pump Baseplate Installation and Grouting," Proc. 5th Pump Symposium, Houston, May
10-12,1988, pp. 117-125.
29
Operation

PRIMING

Centrifugal pumps should almost never be started until they are fully primed, that is, until they have
been filled with the liquid pumped and all the air has escaped. The exceptions involve self-priming
pumps and some special large-capacity, low-head, and low-speed installations where pump priming
before starting is not practical and the priming is almost simultaneous with the starting. Priming of
centrifugal pumps is described in Chapter 25.

WARMUP

Pumps that handle hot liquids should be maintained at approximately operating temperature when idle.
A constant small flow through the pump will provide for this. Many arrangements are available for this
warmup procedure. In some cases, flow goes from the open suction, through the pump and out through
a warm-up orifice downstream of the pump casing (Fig. 28.37). The drains from the warm-up valve are
returned the pumping cycle at some lower pressure point than the pump suction. In other cases, flow
goes through a jumper line around the discharge check valve, through the pump, and into the common
suction header (Fig. 28.38). An alternative for double suction and multistage pumps is to take the flow
from downstream of the discharge check valve and pass it through the casing at multiple points then
back to the suction header (Fig. 29.1). Yet another approach, often employed for very high temperature
services, is to use a separate pump to circulate liquid, as it is being heated, from the suction vessel,
through the pump, and back to the suction vessel (Fig. 28.39). The exact arrangement to be used should
be recommended by the pump manufacturer.
Some pump designs are capable of starting up cold in an emergency, whereas others should never
be exposed to this sudden shock; the pump manufacturer should be consulted in each particular case.
Some general considerations are as follows:

882

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
Operation 883

FROM . . - - - - - - - - . LEGEND:
LOW PRESSURE
F. W. HEATERS _ PATH OF
FEEDWATER
c:> PATH OF
WARM UP
DEAERATOR WATER
STORAGE TANK


r-----Jj\---:::¢:I=..---Jj\--r---- _ TO BOI LER

IDLE PUMP OPERATING PUMP

Fig. 29.1 Typical warm-up arrangement for two pump installation.

Impeller Mounting

When a pump is started cold, its impellers warm up more rapidly than the shaft. If the impellers are
mounted with an interference fit sufficient to stay tight despite the temperature difference between shaft
and impellers, starting the pump cold will not have any injurious effects. Impellers mounted with a light
interference fit may loosen during a cold start and have been known to move out of position if located
by a shoulder or split-ring in the direction of normal hydraulic thrust only. Clearance fit impellers are
located by spacers or extended hubs and so cannot move out of position, but they may become abnormally
loose, hence likely to cause rotor unbalance, until shaft and impeller temperatures equalize.

Double-Casing Pumps
With double-casing pumps, such as used for high-pressure boiler feed service (see Figs. 3.14 and
3.15), the relative position of the inner assembly and of the outer casing must be examined carefully to
see whether or not cold starts will lead to difficulties. When hot water is suddenly admitted to a cold
double-casing pump, the relative expansion of the outer casing barrel and of the inner element goes
through two separate and distinct phases. At first, the inner element, which is much lighter than the
barrel and which is in more intimate contact with the hot water, expands at a considerably faster rate
than the outer casing itself. To simplify the analysis, it can be assumed that the inner element reaches
its final temperature before any appreciable temperature change has taken place in the outer casing.
Then, as the pump continues to operate, the outer casing heats up and reaches its own final temperature
at some later time. If the casing barrel is not lagged, the temperature on its external surface may be
somewhat lower than the internal temperature, but this is negligible.
A particular example concerns an 8-stage boiler feed pump, designed to operate at 160°C (320°F),
with an inner element of 13 percent chrome stainless steel, and an outer barrel of forged SAE-1020
steel. The coefficients of expansion of the two metals in question can be considered to be the same and
884 Operation

equal to 11.7 x 10-6 mm/mmoC (6.5 x 10-6 in/inOF). The length of the inner element within the barrel
(dimension A in Fig. 29.2) is 1,067 mm (42 in).
If, when the pump is at rest on standby service, the metal temperature is permitted to fall to 49°C
(120°F), the sudden admission of 160°C (320°F) feedwater will cause the inner element to expand 1.39
mm (0.055 in) with relation to the casing barrel when it reaches the final temperature. Ultimately, the
outer casing will also come up to temperature and will expand by an equal amount, nullifying the initial
expansion of the inner element.
The effect of this initial relative expansion, followed by a return to the initial relative position of
inner element and outer casing, will have different effects on different designs, depending on whether
or not the construction of the double-casing pump permits free movement of the inner element within
the barrel. If it does, the events which take place will have but little effect on the unit. If, however, the
inner element is constrained within fixed limits, it becomes necessary to interpose some form of compress-
ible gasket between the inner element and either the barrel or the discharge head. (Sometimes these
gaskets are incorporated at both points.) These gaskets must absorb the difference in expansion just
calculated. The reliability and life of the compressible gaskets determines whether or not the unit can
safely be started cold.
Some consideration should also be given to the effect of the warmup method selected. It is true that,
in some cases, the pump will be subjected to a certain amount of distortion because heat may flow
unevenly to various parts of the pump. When warmup has been properly completed, however, this
distortion should disappear; the distortion will not have affected the pump while it was at rest. A careful
analysis is necessary to check that a pump started up cold does not undergo this type of distortion as it
is coming up to temperature, since interference at the running joints or misalignment at the bearings
will be a possible cause of trouble. Although some pumps can be started up cold, whereas others should
not, all pumps handling hot liquids will profit from starting warmed up if the warmup operation insures
a thorough and even distribution of heat to all parts of the pump.

Fig. 29.2 Double-casing pump.


Operation 885

There is one further consideration in cold starting double casing pumps in hot service. When the
application involves very high temperatures or very high pressures or a combination of the two, the
stress induced in the barrel by differential thermal expansion following a cold start can be high enough
to initiate cracking at points of stress concentration. That is obviously an extremely dangerous condition
and therefore to be avoided. With today's finite element analysis (FEA) techniques, a manufacturer can
investigate the capabilities of a particular barrel and make reliable recommendations.

FINAL CHECKS BEFORE STARTUP

After a pump has been installed but before it is started for the first time, the following checks should
be made:

1. The system in which the pump will operate has been cleaned or the appropriate strainers are in place to
protect the pump if system cleanup is to done with the pump running (see Chapt. 2S).
2. All the auxiliary piping is connected as shown on the pump drawings.
3. The coupling alignment, cold, is within acceptable limits.
4. The bearings are clean and correctly lubricated. Pumps with anti-friction bearings that have been standing
a long time without protection (oil mist or long-term inhibiting) should have the bearing housings dismantled
and bearing condition verified. If the bearings are pressure lubricated, the lubricating oil system must have
been flushed with the bearings removed and the oil temperature cycled from 40-S0°C (lOO-IS00P) until
the system would run 4 hours without accumulation in the filters (no increase in pressure drop). When
cleaning bearings and housings, waste should not be used as lint can find its way into the lubricant; clean
rags are superior for this purpose.
5. The driver's direction of rotation is correct. This should be checked with the driver disconnected, particularly
when the driver is an electric motor, as the power leads may have been incorrectly connected. (The procedure
should be repeated each time the motor is disconnected for maintenance purposes or when there is a possibility
that the leads have been disconnected at the starter.) Generally an arrow is inscribed on or attached to the
pump casing to show the correct rotation. The least effect of reverse rotation on a centrifugal pump is a
significant reduction in performance (Pig. 32.4); the worst is substantial mechanical damage such as can
happen in smaller vertical turbine pumps that have threaded lineshaft couplings.
6. The pump's rotor is free to turn. When the pump operates at high (or low) temperatures, its rotor must be
free to turn at both ambient and operating temperature. If the rotor is bound or tight to turn, the pump must
not be run until the cause of the trouble is found and corrected.

STARTING AND STOPPING PROCEDURES

The steps necessary to start a centrifugal pump will depend on its type and the service on which it is
installed (Fig. 29.3). Many installations require steps that are unnecessary in other installations. For
example, standby pumps are often held ready for immediate starting, particularly centrifugal boiler feed
pumps. The suction and discharge gate valves are held open, and reverse flow through the pump is
prevented by the check valve in the discharge valve.
The methods used to start pumps are greatly influenced by the performance characteristics of the
pump in question, that is, by the shape of its power-capacity curve. High- and medium-head pump (low
and medium specific speeds) curves rise from the shutoff condition to the normal operating capacity
condition (see Fig. 18.31 and 18.32); therefore, these pumps should be started against a closed discharge
valve in order to reduce the starting load on the driver. The use of a check valve in the discharge line
886 Operation

Check ,GOlfe
I valve
Pressure- vOtIve \ .
reducing
orifice \
,, Recircu-
\
\
, \
,
D,schar~e
/

,
I
IOtlion
line \
\ ,,
I

W"r",-up
Isolorfin!J valve.
Conf:O,
vOtIve
, VOtIve.,
,. I

Sue/ion

Pump

"-Sfut:"ing-hDIf~ ......
cooling fIV"ter .... , ,
Outboard-hearin!! coolin!! WOtterJ

Fig. 29.3 Connections and auxiliary services to a centrifugal pump.


Of importance in establishing starting and stopping procedures.

is equivalent to a closed valve for this purpose, as long as another pump is already on the line. The
check valve will not lift until the pump being started has come up to a speed sufficient to generate a
head high enough to lift the check valve from its seat. Precautionary measures against pump overheating
from operation at shutoff may have to be employed in certain cases by installing a recirculation bypass.
The power consumption curve of low-head pumps (high specific speed) of the mixed flow and propeller
type has the opposite characteristic, rising sharply with a reduction in capacity, as shown in Fig. 18.34.
Such pumps, therefore, should be started with the discharge valve wide open, against a check valve if
required, to prevent back flow.
Assuming that the pump in question is motor driven, that its shutoff horsepower does not exceed the
safe motor horsepower, and that it is to be started against a closed gate valve, the starting procedure is
as follows:

1. Prime the pump, opening the suction valve and closing the drains to prepare the pump for operation.
2. Open the valve in the cooling water supply to the bearings.
3. Open the valve in the cooling water supply if the stuffing boxes are water cooled.
4. Open the valve in the sealing liquid supply if the pump is so fitted.
5. Open the wannup valve of a pump handling hot liquids if the pump is not normally kept at operating
temperature. When the pump is wanned up, close the valve.
Operation 887

6. Open the valve in the recirculating line if the pump should not be operated against dead shutoff.
7. Start the motor.
8. Open the discharge valve slowly.
9. Check the operation of the shaft seal. With mechanical seals, the most common type today, this means
verifying that seal leakage is within limits (generally not visible; for hydrocarbons measured as gas
concentration in the vicinity of the seal), and that any auxiliary systems (such as API Plans 21, 23, 52,
53, or 54) are functioning correctly. If the pump is equipped with packed box seals, observe the leakage
rate and adjust the sealing liquid valve or gland for proper flow to ensure the lubrication of the packing.
If the packing is new, do not tighten up on the gland immediately, but let the packing run in before reducing
the leakage through the stuffing boxes.
10. Check the general mechanical operation of the pump and motor.
11. Close the valve in the recirculating line once the flow to the system is above the minimum allowable.

If the unit is to be started against a closed check valve with the discharge gate valve open, the steps
would be the same, except that the discharge gate valve would be opened some time before the motor
is started.
In some cases, the cooling water to the bearings and the sealing water to the seal cages is provided
by the pump itself. This eliminates the need for steps 2 through 4 in the starting procedure.
The procedure for stopping a pump also depends upon the type and service of the pump. Generally,
the procedure to stop a pump that can operate against a closed gate valve is as follows:

1. Open the valve in the recirculating line.


2. Close the gate valve.
3. Stop the motor.
4. Open the warmup valve if the pump is to be kept up to operating temperature.
5. Close the valve in the cooling water supply to the bearings and to water-cooled stuffing boxes.
6. If the sealing liquid supply is not required when the pump is idle, close the valve in this supply line.
7. Close the suction valve and open the drain valves, as required by the particular installation, or if the pump
is to be opened up for inspection.

If the pump does not permit operation against a closed gate valve, steps 2 and 3 are reversed. Many
installations permit stopping the motor before closing the discharge gate valve.
Generally, the starting and stopping of steam-turbine-driven pumps require the same steps and the
same sequence as for motor-driven pumps. As a rule, steam turbines require warming up before starting
and have drains and seals that must be opened and closed before and after operation. The operator
should, therefore, conscientiously follow the steps outlined by the turbine manufacturer in starting and
stopping the turbine. This is also true in the case of internal combustion engines and gas turbines used
to drive pumps.

AUXILIARY SERVICES ON STANDBY PUMPS

Standby pumps are frequently started up from a remote location, and several methods of operation are
available for the auxiliary services, such as the cooling water supply to the bearings or to water-
cooled stuffing boxes. The choice among these methods must be dictated by the specific circumstances
surrounding each case. The most logical methods are the following:
888 Operation

1. A constant flow can be maintained through the bearing jackets or oil coolers and through the stuffing box
lantern rings, whether the pump is running or standing still on standby service.
2. The service connections may be opened automatically whenever the pump is started up.
3. The service connections may be kept closed while the pump is idle and the operator instructed to open them
within a short interval after the pump has been put on the line automatically.

Method 1 wastes the cooling water and can be harmful. The necessity of regulating the amount of
cooling water to the pump bearings is frequently overlooked, and, generally, the error is to overcool
rather than to supply insufficient cooling water. Many ball bearing failures are due to the bearing being
almost refrigerated so that the resulting condensation on the cold inside walls of the bearing housing
mixes with the lubricating oil or grease. Rusting and pitting of the balls leads to obvious trouble. The
outflowing cooling water temperature should not fall much below 40 to 46°C (lOS to 115°F).
Cooling water is frequently available at temperatures as low as 16 to 21°C (60 to 70 0 P), and if allowed
to flow through the bearing housing of an idle pump installed in a warm or moist atmosphere, it may
lead to bearing troubles. While the pump is not running, no heat is generated at the bearings and the
bearing housings will be maintained at exactly the temperature of the cooling water.
Sometimes cooling or sealing water to the pump stuffing boxes must be maintained whether or not
the pump is running. Some examples are pumps that handle liquids corrosive to the packing or liquids
that may crystallize and deposit on the shaft sleeves.
If the stuffing boxes are equipped with water-cooled jackets, leaving the connections open at all times
may be wasteful, but presents no particular danger.
In method 2, individual water supply lines can be equipped with spring-loaded pressure control valves
as illustrated in Pig. 29.4. The pressure side of the diaphragm is connected to the pump discharge by
means of a pilot line, so that the valves will open as soon as the pump starts and develops pressure.
If the standby pump is motor driven, solenoid-operated control valves can be installed in the water
cooling supply lines. The valves remain closed under spring action as long as the solenoid is deenergized
and open as soon as the motor is put on the line, energizing the solenoid.
Valves, whether controlled by pressure or by solenoids, should be supplied with locking devices. The
operators can lock them in the open position, when they have time to attend to a pump that has been
started automatically.
If operators are available near the pump location, and if the pump is of such a design and on such a
service that it can be operated a few minutes without cooling or sealing water supply, method 3 may
be the most suitable.
The discussion of cooling water is largely relevant to pumps installed some time ago. Modem designs
generally use cooling water for lubricating oil systems and mechanical seal heat exchangers only. The
operating practice with standby pumps in these cases is to have cooling water to the lubricating oil
system whenever the auxiliary oil pump is running, and to have cooling water to the seal heat exchangers
whenever the pump is on warm-up, the latter to avoid thermal shock to the seal when the pump is
started. This is not without its difficulties, however, and in some boiler feed pump installations it has
been necessary to add a small circulation pump to each of the mechanical seal plan 23 loops to ensure
the shaft is not bowed by thermal stratification of the liquid in the seal housings.

OPERATION OF CENTRIFUGAL PUMPS AT OFF-DESIGN FLOWS

Centrifugal pumps are generally selected for a given capacity and total head when operating at rated
speed. These characteristics are referred to as rated conditions of service and, with few exceptions,
Operation 889

SUCTION DISCHARGE

PILOT LINE
-----------,----~
I I
I I

,
I I
I
I COOLING WATER TO
BEARING OIL
COOLER

SPRING-LOADED DIAPHRAGM
PRESSURE CONTROL VALVES
RETURN FROM STUFFING BOXES

RETURN FROM
OIL COOLER

Fig.29.4 Automatic control for cooling water supply.

represent the conditions at which the pump will operate most of the time. Pump efficiency should be at
its maximum for these conditions of service, and pumps are so selected wherever possible.
Often, however, pumps are required to operate at capacities and heads differing considerably from
the rated conditions. Examples are applications for steam power plant services, where boiler feed,
condensate, and heater drain pumps may be called on to deliver any flow to the boiler from full capacity
to zero, depending on the load carried at the moment by the turbogenerator. Condenser circulating pumps
are subject to somewhat lesser variations, but nevertheless these pumps may have to run against widely
varying total heads and, therefore, at varying capacities. General service pumps in a variety of applications
may also be called on to operate over a very wide range of flows. It is very important, therefore, that
the user of centrifugal pumps acquaint himself with the effects of operating pumps at capacities and
heads other than the rated conditions and with the limitations imposed upon such operation by hydraulic,
mechanical, or thermodynamic considerations. The effect of operation at off-design flows and the
precautions to be taken against the unfavorable effects of such operation are discussed in detail in
Chapter 22.

RUNNING A PUMP DRY

The question of running a centrifugal pump dry is a very complex one and certainly controversial. To
begin with if, as in certain cases, the pump design is such that support for the shaft is derived in part
from the hydrodynamic effect of wearing rings or interstage bushings intended to act as intermediate
bearings, running the pump dry cannot be permitted. This peremptory statement is based on the fact
890 Operation

that since no liquid will be flowing through the so-called internal bearings, the auxiliary support will
not be provided and contact will take place at all internal clearances.
But even if the pump design is such that no intermediate support is required for the normal operation
of the pump, we must consider that it is unlikely that the pump could pass instantaneously from normal
operation to running dry. After all, dry-running conditions could be expected to occur only if cavitation
and flashing takes place in the pump. This means there would be a transition from liquid filled to vapor
filled, during which the rotor would be subjected to higher radial loads than normal with consequently
greater deflection. Whether a pump can withstand the transient conditions imposed on it depends on the
stiffness of its rotor, the magnitude of its running clearances versus the rotor's static deflection, the
rotor's dynamic characteristics when the pump is dry, and the ability of the running clearances to survive
a "touch" without galling. The majority of pumps cannot economically meet all these requirements, and
so as a general rule serious risks are taken if a pump is permitted to run dry.
There are three exceptions to this rule. The first is a particular design of automatic priming used with
large, low-head pumps. The pump is started dry when the vacuum pump is started and runs dry for not
more than two minutes, at which time the priming has been completed and the pump goes into normal
operation. To ensure successful operation under these conditions, the clearances at the wearing rings are
made slightly larger than in the normal design.
The second is large overhung and between bearings pumps that are started dry to lower their inertia,
and so lower starting current drawn by their electric motor driver. These machines have high stiffness
rotors, and their wearing rings are equipped with water injection to provide lubrication and cooling while
the pump is running dry.
The third is a class of multistage pumps whose rotor stiffness is abnormally high, and whose rotor
dynamics with the pump dry are similar to those of centrifugal compressors. Figure 7.9 shows the
demarcation for this class of pump, and there is field experience supporting the demarcation. Pumps
falling into this class generally are no more than 6 stages, have unusually large shafts, and have generous
internal running clearances with at least one surface grooved.
Under abnormal conditions, such as a sudden load reduction in an open cycle power plant or a sudden
fall in fractionator liquid level in a refinery, a pump sometimes becomes vapor bound. The question
then is whether to keep the pump running until suction is restored or shut it down. If the pump is a
design that relies on liquid effects in its internal running clearances to support the rotor, then it should
be shut down because there is contact at the clearances and damage is being done. On the other hand,
designs whose rotors are stiff enough to not have contact at the clearances may in fact suffer less damage
if left running. This seeming contradiction rests on the fact that pumps generally seize on rundown when
there is very little torque available to overcome any galling in their clearances. With the pump having
survived the transient, it is now running without contact at the clearances and so is better left running
until suction is restored.

THROTTLING AT THE PUMP SUCTION

Throttling the suction of a centrifugal pump causes a reduction of the absolute pressure at the inlet to
the impeller. This can be made to result in a reduction of capacity by forcing the pump to operate "in
the break," and reducing the delivered capacity by altering the shape of the head-capacity curve. Such
operation is harmful to the pump, unless, as in the case of a condensate pump, it is specifically designed
for it. Pump efficiency is reduced when operated "in the break," but most important, erosion and premature
destruction is caused by the cavitation induced when the suction is throttled.
Pump capacity can be reduced simply and safely by throttling the discharge. In this manner, artificial
Operation 891

friction losses are introduced by throttling, and a new system-head curve is obtained that intersects the
head-capacity curve at the desired flow.
Throttling at the suction is permissible only when the suction pressure exceeds the minimum require-
ments by a large margin, such as the case of the second pump of a series unit. The effect, however, is
not to reduce capacity by operation in the break, but rather by the reduction of the total net head generated
by the series unit. This causes the head-capacity characteristics and the system-head curve to intersect
at a lower rate of flow.

RESTARTING MOTOR-DRIVEN PUMPS STOPPED BY POWER FAILURE

If a check valve protects a pump against reverse flow after a power failure, there is generally no reason
why the pump should not be restarted once current has been reestablished. The type of motor control
used will determine whether or not the pump will start again automatically once power is restored.
Starters are made with low-voltage protection, with low-voltage release, or without either. Starters with
low-voltage protection will deenergize under low-voltage conditions, or following power failure, and
the units they control must be restarted manually. Starters with low-voltage protection can only be used
with momentary-contact pilot devices and cannot be used with maintained-contact pilot devices, such
as float switches, unless auxiliary relays are incorporated in the controls.
If the starter does not incorporate low-voltage protection, resumption of power will always cause the
unit to start again automatically. Because pumps operating on a suction lift may lose their prime during
the period when power is off, starters should be provided with low-load protection for such installations.
This does not apply, of course, if the pumps are automatically primed, or if some protection device is
incorporated so that the pump cannot run unless it is primed.

OPERATION OF CENTRIFUGAL AND RECIPROCATING PUMPS


IN PARALLEL

Although centrifugal pumps often can be operated in parallel with reciprocating pumps, the general
performance of the centrifugal pump will be affected both mechanically and hydraulically by the pulsations
of the reciprocating pump. A triplex pump would have the least effect on the operation of the centrifugal
unit whereas actual troubles can be experienced if attempts are made to parallel a centrifugal pump with
a single-acting, single-cylinder pump. It is important not to use a common suction line for a centrifugal
and a reciprocating pump, especially if they operate under a high-suction lift.
A large number of problems encountered in the operation of centrifugal pumps cannot be discussed
here. The pump manufacturer should be consulted if difficulties are encountered or if the choice between
two different solutions of an operating problem seems difficult.

MONITORING PUMP PERFORMANCE

To predict when pump internal clearances need to be renewed, it is necessary to monitor its performance,
that is, the pump capacity, total head, and power consumption. Periodic monitoring can also alert the
operator to sudden and rapid changes in performance that can be caused by conditions other than the
gradual wear of the internal clearances.
Procedures for monitoring pump performance are discussed in detail in Chapter 30. It is important
to understand that a field test of a centrifugal pump cannot, and need not, be as accurate as a test carried
892 Operation

out on the manufacturer's test stand. Accuracy is less important than repeatability of the results. It is
more important for the user to know how the pump is performing over its life than to ascertain that, as
received, the pump met its guarantees. Thus, flow and head readings taken during a field test could be
in error, but still serve their purpose, as long as the error is always made in the same direction and is
of the same order of magnitude.
30
Monitoring and Performance Testing

Monitoring and perfonnance testing both have the broad objective of learning more about a pump's
operation and condition. In detail, however, their objectives are different. Monitoring is the gathering
of infonnation pertinent to (1) the safe operation of a pump and (2) establishing an optimum time for
overhaul. Testing, on the other hand, is used primarily to verify the operation and perfonnance of
emergency pumps, and occasionally to detennine the perfonnance of an installed pump or the system
in which it is operating. For this reason, the two topics are discussed separately.

Monitoring
Expanding on the given definition, monitoring involves gathering data on the operation of a running
pump, then comparing those data with either preset acceptance limits or values of the same parameters
when the pump was new. Violating a preset limit usually either sounds an alann to initiate further
investigation or trips the pump's driver when there is a risk of the pump being seriously damaged. A
sudden change in trend is cause for further investigation to detennine what caused the change and when
correction will be necessary. Monitoring for changes in trend is the key to what is tenned "on condition
maintenance" (see Chap. 31).

EXTENT OF MONITORING

The extent of monitoring ranges widely. At the very least, an operator takes and records periodic
measurements of operating and perfonnance data, which are then evaluated manually. At the most,
continuous on-line monitoring apparatus gathers a wide range of operating and perfonnance data, then
analyses them using a computer based "expert system." The grading between these is largely the extent
of data gathered continuously and how they are analyzed. What is justified depends on the assessment
of four factors:

893

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
894 Monitoring and Performance Testing

1. Criticality of the service


2. Consequence of pump failure
3. Severity of the service
4. Sensitivity of the pump design to damage.

These factors also have a distinct influence on the choice of pump and pumping arrangement for a given
service, so it is evident service conditions, pump selection, and monitoring are all interrelated.
The best way to assess the extent of monitoring needed for a given installation is to develop a
numerical grading using the four factors listed above. Before setting down a suggested scheme for that,
it is beneficial to review the qualitative aspects of the factors.
Criticality of the service refers to the importance of the pumping intended being done when required.
It includes the consequences of an unexpected cessation of pumping. Transferring water from one
reservoir to another is generally not critical because it's usually done with large capacities in reserve.
Injecting water into the core of a nuclear reactor, on the other hand, is critical since it is the final means
of maintaining the core at a safe temperature following a major loss of coolant from the system.
Consequence of failure is determined primarily by what is likely to happen if the shaft seal fails.
This ranges from serious, when it risks possible damage to the pump bearings, to critical, when there
is a high risk of fire or unacceptable damage to the environment, which includes the release of toxic
liquid or vapor. When gross leakage from the shaft seal has critical consequences, it influences both the
selection of the pump and seal and the extent of pump monitoring.
Severity of the service usually requires a lot of care to judge accurately, there being numerous instances
of seemingly benign services turning out to be severe, at least in their effect on the pumps installed.
Table 30.1 provides some guidance for this assessment, but is not exhaustive because it cannot take into
account all the factors and their effects in a particular application.
Sensitivity of the pump's design to damage is generally thought to mean that the susceptibility of a
pump design to failure increases with its rotative speed. Although there is no doubt increasing the rotative
speed of pumps raises their specific energy (power per unit of rotor weight), comparison with centrifugal
compressors, which achieve high reliability, says that such designs are not inherently less reliable than
their low-speed brethren. What does need to be recognized is that as the specific energy is increased, a

Table 30.1 Severity of Pump Application

Severity of application
Operating parameter Benign (low) Severe (high)

Pumping Below minus 100°C (150°F);


temperature above 250°C (475°F)
NPSH margin At least 10 percent of Low in water or high energy
NPSHR with I-m (3-ft) applications (U I > 30 mls
minimum in [100 ft/sec]); see NPSH
hydrocarbon applications
Vapor binding Negligible risk High risk
Abrasive solids None present or expected Present in high concentration
for class of pump
Flow range Narrow; <10 percent of Wide: > 40 percent of rated
rated
Flow fluctuations Rare; low rate of change Frequent; high rate of change
Number of starts Low; less than 1 per month High; 1 or more per day
Monitoring and Performance Testing 895

trend driven by economics, so is the need for more careful design and manufacture. Looking at the
difficulties that have arisen in the evolution of turbomachinery design, it is evident that experimentation
and refinement generally only took place once problems caused by extrapolation manifested themselves.
With that said, it is true that the sensitivity of turbomachines to mechanical damage increases with
rotative speed (since the forces caused by unbalance vary as the speed squared), therefore it makes sense
to monitor the rotor vibration of an unspared, high-speed pump because relatively minor damage to the
rotor can significantly increase the vibration, thereby raising the risk of sudden seal or bearing failure.
Conversely, there is little to be gained by going to such lengths with a low speed, spared water pump,
where gross rotor damage would be necessary to produce dangerous rotor vibration.
Assigning possible grades of 1, 2, and 3 to the four factors, the overall grading for a particular
application can be determined by summing the grades for the individual factors. With the overall
grading determined, Table 30.2 can be used as a guide to the extent of monitoring warranted for a
particular application.
To illustrate the use of this grading system, consider four examples: a spared intermittent operation
water supply pump; a spared, high-temperature hydrocarbon reflux pump; a spared boiler feed pump in
a swing loaded station; and an unspared, continuous operation hydrocracker charge pump. Table 30.3
summarizes the gradings for each of these examples against the four factors for determining the extent
of monitoring. From these gradings the water supply pump warrants periodic monitoring to check its
performance and to be sure its seals are operating correctly, because gross seal leakage risks contaminating
the bearings. The reflux pump needs more attention because of the risk of fire should the seal fail.
Although the boiler feed pump is spared, the severity of operating in a swing loaded plant justifies

Table 30.2 Extent of Monitoring for Various Classes of Pump Service

Extent of monitoring
Grading Periodic Continuous

~5 Measurement of
(Class 1) bearing temperature
bearing housing vibration
suction and discharge pressure
flow
power
6-8 As Class 1 plus Measurement of
(Class 2) VOC leakage from seal bearing housing vibration
AF bearing vibration
9-10 Measurement of Measurement of
(Class 3) VOC leakage from seal pump suction condition
minimum flow
seal injection
lubricant supply
casing temperature
bearing temperature
vibration
seal condition
balancing leak-off flow
~11 As Class 3 As Class 3 plus
(Class 4) pump flow
speed
896 Monitoring and Per/ol7TUlnce Testing

Table 30.3 Example of Grading Pump Applications

Application and grading


Factor Water supply HC reflux BFpump HC charge

Criticality of service 1 1 2 3
Consequence of seal leakage 1 3 2 3
Severity of service 1 2 3 2
Pump sensitivity to damage 1 1 3 3
Total grading 4 7 10 11

extensive, continuous monitoring of operation to ensure the plant does not have to reduce load because
one of the feed pumps is unexpectedly out of service. The unspared charge pump presents a still greater
risk because leakage from its seals could cause a fire, and an unexpected shutdown puts the reactor at
risk, unbalances the refinery's operation, and incurs a high cost in lost production. This pump, therefore,
warrants continuous monitoring of its operation and performance to reduce the risk of sudden failure in
service and to determine the extent of overhaul necessary at each unit turnaround.

MONITORED PARAMETERS AND INSTRUMENTATION

The instrumentation needed for monitoring a particular pump is determined by two factors: the extent
of monitoring deemed necessary and the type of pump involved. Figures 30.1, 30.2, and 30.3 illustrate
the point, showing Class 2 monitoring for a simple process pump and Class 3 monitoring for a sealless
pump and a high-pressure barrel pump. This leaves the machinery engineer with having to specify what
will be done with the output of the monitoring instrumentation. When there is some uncertainty over

V1 - BEARING HOUSING VIBRATION

Fig. 30.1 Simple process pump with class 2 monitoring.


Monitoring and Performance Testing 897

@- ---------------,

Fig. 30.2 Hermetically sealed pump with class 3 monitoring.


Key:
LS 1 = Liquid level at pump section
11 = Motor current (or kW)-alternate to LSI.
TSI = Can temperature
VI = vibration; bearing housing of magnetic drive; casing
or rotor of canned motor
V2 = Rotor axial position (canned motor)

the necessary extent of monitoring and its use, it is better to strive for simplicity over complexity; too
much instrumentation with its output poorly used is almost as bad, if not in fact worse, than too little
instrumentation with its output well used.
Table 30.2 sets down the parameters normally measured for the four classes of monitoring recom-
mended. These parameters can be divided into two categories: those that will affect the pump's operation,
and are therefore likely to cause a problem, and those that indicate a problem has already developed.
Generally the first category is made up of system conditions, the state of auxiliary systems and pump
flow, whereas the second is the pump's operating characteristics, principally temperatures and vibration.

OPERATING CONDITIONS

Suction Conditions
With around two-thirds of all centrifugal pump problems having their origin in the pump's suction
conditions, there is a substantial benefit to ensuring the suction conditions are correct.

1. Suction pressure-Most installations have a local gauge to indicate suction pressure; some supplement that
with a transmitter and remote indicator in say a control room. In a system where there is a high risk of
losing suction pressure, such as one with low capacitance, a switch is used to continuously guard against
too Iowa suction pressure.
2. Strainer pressure drop--Following sound engineering practice, any strainer in the suction, either temporary
or permanent, should be equipped with at least some means of measuring the pressure drop across it. This
Lubricating oil

Fig. 30.3 High-pressure barrel pump with class 3 monitoring.


Key to Symbols:
TI = Inboard journal bearing temperature
T2 = Outboard journal bearing temperature
T3 = Inner thrust shoe temperature
T4 = Outer thrust shoe temperature
T5 = Barrel temperature, top2
T6 = Barrel temperature, bottoM
VI and V2 = Inboard shaft, X and Y displacement
V3 and V4 = Outboard shaft, X and Y displacement
V5 and V6 = Rotor axial position
KPI = Keyphasor®
PSI = Oil pressure switch
TSI = Oil temperature switch
TS2 = Inboard seal temperature switch
TS3 = Outboard seal temperature switch
QI = Balancing leak-off flow
PI = Suction pressure
PS2 = Suction pressure switch
J Pumps with tandem impellers only 2When pumping temperature >150°C (300°F)
Monitoring and Performance Testing 899

can be as simple as two pressure gauges, one on each side, for local indication, or as complicated as
differential pressure transmitter for remote indication and sounding an alarm.
3. Suction level-The liquid level in the suction vessel is monitored instead of suction pressure when the normal
suction level is close to that which will allow vortexing, which risks the pump becoming vapor bound.
Alternatively, in installations with several pumps taking suction from a single vessel, the liquid level at the
pump suction is monitored to ensure the pump is not run dry. Monitoring suction level is advisable for
pumps with product lubricated bearings, which will be seriously damaged if run dry. Sealless pumps and
multistage pumps with slender rotors are two notable examples of these.
4. Specific gravity-In process applications, the liquid pumped can be a mixture drawn from a number of
sources. On occasion the pump may be sensitive to the specific gravity of the mixture, the usual risk being
flashing from the introduction of a light fraction into a high temperature stream. Monitoring the specific
gravity of the mixture at the pump suction and sounding an alarm on a preset low value is one means that
has been employed to protect against this.
5. NPSHA-For a liquid of known vapor pressure characteristic, feeding suction pressure and temperature into
a programmable logic controller (PLC) allows the NPSHA to be calculated and compared to a preset
minimum, sounding an alarm or initiating corrective action, such as quenching the suction, in the event
NPSHA falls below the preset value.
6. Motor current-Motor-driven pumps operating with a suction lift or prone to running dry for some other
reason, can be protected from dry running by monitoring the current drawn by their motor and tripping the
motor if the current falls below a preset value.

Minimum Flow
For all but Class 4 monitoring, the concern with pump flow is that it not fall below the value that
will cause rapid damage to the pump or risk a serious accident.

1. Pump discharge temperature-A low- to medium-energy centrifugal pump will show a gradual rise in the
temperature of the liquid in its casing when run at zero flow, the rate of temperature rise depending upon
the power being absorbed and the volume of liquid in the casing (see Chap. 22). In applications with multiple
discharge paths, where there is a risk of the pump inadvertently being run at zero flow, a switch sensing
the temperature of the liquid in the discharge region of the pump's casing has been used to trip the pump
in the event a preset temperature was exceeded. That temperature is the highest tolerable for the lightest
liquid being pumped unless a mechanical constraint or thermal inertia dictates a lower temperature. For the
switch to function correctly, it is important that its sensor be located in flowing liquid.
2. Minimum flow-High-energy pumps are equipped with an automatic bypass control system (see Chaps. 22
and 28). Pump suction flow is the preferred control parameter for these systems. The usual instrumentation
is a differential producing primary element, such as an orifice, and a controller that either switches or
modulates the bypass valve.
3. Bypass flow established-Those pumps that require automatic bypass control systems can run for only a
few seconds at zero flow without boiling the liquid within them. To ensure the bypass has opened when it
should, some plants have added an orifice and differential pressure switch in the bypass line. The control
logic with this arrangement is that if the differential pressure switch is not showing flow established within
say 5 sec of the signal to open the bypass valve, the pump is tripped.

Seal Injection
Pumps with externally flushed mechanical seals (API Plan 32; see Chap. 9) or condensate injected
breakdown seals (see Chap. 10) are dependent on the availability of adequate seal injection.

1. Pressure-When the operation of the seal depends upon injection from an external source, it is important
that the available pressure exceed that within the pump, adjacent to the seal, by some minimum value. The
900 Monitoring and Performnnce Testing

available pressure can be read from a locally mounted gauge or remotely using a transmitter and indicator.
When deemed necessary, a switch can be incorporated to sound an alarm if the available pressure falls
below the minimum allowable value.
2. Filter pressure drop--Any filter in the seal injection system must have some means of indicating the pressure
drop across it (see suction conditions earlier for instrumentation details).
3. Flow-With axial face seals, the important aspect of injection from an external source is the flow to the
seal. The value of this flow is low, typically ranging from 0.15 to 1.5 m3/hr (0.5 to 6 gpm), and is usually
monitored by a variable orifice type flow meter. The flow meter can incorporate a switch to sound an alarm
if the flow falls below a preset minimum value. In critical applications, established seal injection flow could
be a prerequisite for starting the pump's driver.
4. Temperature-Monitoring seal injection temperature is only warranted in instances where the liquid is being
taken from a point of variable temperature or is being cooled. In those cases, the monitoring instrumentation
would take the same form as that for injection pressure (see earlier).

Lubricant Supply
Pumps with oil-bath lubrication generally have only a level gage for local indication that they have
adequate oil. When forced-feed lubrication is employed, it is necessary to monitor the supply to prevent
a failure through lack of lubrication.

1. Pressure-Most pumps with forced-feed lubricated bearings use pressure upstream of an orifice adjacent to
the bearings to determine whether lubricant is available and thereby control the pumps' operation (see Chap.
11 for details of typical forced-feed lubrication controls).
2. Filter pressure drop--When filters are used, monitoring the pressure drop across them is mandatory (see
Chap. 11 for details).
3. Flow-In pump designs where there is no orifice upstream of each bearing, or where there is a risk of
blockage in the flow path downstream of the first bearing (as there is in vertical wet-pit pumps with enclosed
lineshafts), it is necessary to monitor lubricant flow instead of pressure. In oil lubrication systems, the control
functions are the same as for pressure monitoring. When water from an external source is the lubricant, a
single flow switch is used to allow the driver to start when lubricant flow is established and to trip the driver
if lubricant flow falls below a preset value.

PUMP CONDITION

Casing Temperature
Monitoring casing temperature is warranted when there is a risk of thermal distortion or thermal
shock, or when it serves as an indirect means of indicating the pump is being run at too Iowa flow.

1. Differential top to bottom-When the pump temperature is 150°C (300°F) or higher, it is advisable to monitor
the temperature difference, top to bottom, of horizontal multistage pump casings (Fig. 30.3) to ensure the
pump is not started with the casing seriously distorted. The usual practice is to not allow the driver to be
started until the temperature difference is within 22 to 28°C (40 to 50°F) depending on the manufacturer's
specification. Generally it is adequate to measure the casing temperature at two points: mid span top and
bottom. The usual temperature detectors are thermocouples or resistance temperature detectors (RIDs).
2. Differential to pumped liquid-At pumping temperatures 150°C (300°F) and above, most manufacturers
recommend that standby pumps be kept "warmed up" to close to the operating temperature. This is accom-
plished with a warm-up system of some form (see Chap. 28). To ensure the pump is not subjected to a
severe thermal shock, it is common practice to caution against starting the pump unless the difference
Monitoring and Performance Testing 901

between the pump's casing temperature and the temperature in the suction vessel is within a given limit,
55°C (100°F) being a typical value. This is not to say the pump cannot be started when the temperature
difference is greater provided the casing differential temperature is within the limit. There are many examples
of pumps that have been occasionally started when their casings were on the order of 110°C (200°F) below
the temperature of the liquid in the suction vessel. The risk is in frequent starting with high differentials
between the casing and the pumped liquid, a practice that has lead to thermal cracking in barrel pump casings.
3. Total temperature-Monitoring the temperature of a pump's casing provides an indirect means of detecting
when the pump is being run at too Iowa flow, since the casing's temperature will eventually rise above
that of the pumped liquid. Because the thermal inertia of conventional pump casings is high, detecting low-
flow operation by casing temperature measurement is only useful in sealless pumps. These designs have a
relatively thin pressure boundary that quickly reflects a rise in the temperature of the liquid in the pump.
High casing temperature in sealless pumps is generally cause to trip the pump since it is a lagging indicator
of incorrect operation.

Bearing Temperature
Three means are used to monitor bearing temperatures: oil temperature in the bearing housing or
drain from it and bearing temperature by either a contact type detector or an embedded detector. The
difference between them is practicality and their response to overheating in the bearing. Bearing tempera-
ture is used to alarm on high temperature, trip on what is termed "high-high" temperature. The values
of these temperatures vary with the means of monitoring.

1. Oil temperature-A thermometer or temperature detector in the oil reservoir of pumps with antifriction or
ring-oiled sleeve bearings, provides a lagging indication of overheating in the bearing. The lag in response
is too slow to protect against sudden overheating in the bearing but will warn of a problem that develops
slowly. In pumps with forced-feed lubricated bearings, obtaining a useful measurement of the temperature
of the oil leaving the bearings is difficult because the detector either does not have enough immersion to
reliably measure the temperature, or the detector is immersed in a large volume of oil that responds only
slowly to changes in bearing temperature.
2. Contact type detectors-In pumps equipped with antifriction bearings, a detector in contact with each
bearing's outer race provides the best indication of the actual temperature within the bearing. The response
of the detector is improved by having a copper heat sink and a means of ensuring the detector tip is in
contact with the bearing (Fig. 30.4), both to improve thermal conductivity between the bearing and the detector.
Contact-type detectors have been used in hydrodynamic bearings, with similar construction precautions, but
are now generally considered superseded.
3. Embedded detectors-The most accurate and rapid indication of the temperature in hydrodynamic bearings
is obtained with embedded detectors (Fig. 30.5). These designs bring the detector tip close the bearing's
working surface, and in the case of tilting pad thrust bearings, locate the detector where the temperature is
normally the highest.

Contact and embedded temperature detectors are usually either thermocouples or resistance temperature
detectors. API-670 [30.1] is a comprehensive industry standard covering the specification and installation
of these instruments.

Vibration
Always a good indicator of machine condition, vibration measurement in conjunction with modem
analysis techniques has become an indispensable tool for monitoring rotating machinery to both know
more about its condition and gain insight into the cause of trouble as it become apparent. Measurement
methods ranges from periodically taking readings of overall bearing housing velocity with a simple
hand-held instrument to continuous readings of shaft to housing displacement with the data being fed
902 Monitoring and Performance Testing

TEMPERATURE BABBln LAYER


DETECTOR

SPRING
BEARING
SHOE

LEADWIRES

Fig. 30.4 Contact-type temperature detector.


(Courtesy Kingsbury, Inc.)

A
CCW

~__ --7~'"

ROTATION

SECTION A-A

Fig.30.5 Embedded-type temperature detector installed in tilting pad thrust bearing shoe.
(Courtesy Kingsbury, Inc.)
Monitoring and Performance Testing 903

back to an analyzer that automatically determines likely causes of trouble if it detects a sudden change
in trend. Whether to measure vibration and how to do so if measurement is required depend on the class
of monitoring warranted (see the preceding discussion) and the construction of the pump. Because
centrifugal pumps handle liquids, their vibration is materially affected by the nature of the flow through
them. One consequence of this is that vibration tends to vary with pump flow, the usual trend being an
increase in vibration as flow is raised above or lowered below design or best efficiency point (BEP). In
fact some specifications base the minimum allowable continuous flow on a certain rise in vibration above
that at BEP. The significance of this to monitoring is that vibration data should always be related to
some measure of pump flow, and if the pump is variable speed, its rotative speed.

1. Bearing housing-A pump with antifriction bearings has a stiff connection between the shaft and the bearing
housing that supports it. This means the transmissibility of the forces acting on the shaft to the housing is
high, and therefore it is possible to obtain a good indication of those forces, hence the condition of the
pump, from measurements of bearing housing vibration. Measurements are made with a seismic transducer,
either an accelerometer or a velocity transducer, which produces an output signal proportional to the absolute
motion of the surface it is in contact with or attached to. The transducer's output is displayed in velocity
units mm/sec RMS or in./sec equivalent peak, the latter being 1.414 times the RMS reading. Being RMS
or RMS-based, either of these readings gives a useful measure of the effective energy being applied to the
pump's shaft. Choosing between an accelerometer or a velocity transducer depends primarily on the desired
frequency range (accelerometers have the higher but pumps need not be evaluated beyond two times the
vane passing frequency), and the space available for installation (velocity transducers are larger). For
continuous monitoring, the transducer is attached to the bearing housing with a threaded connection. In most
cases only one transducer is installed in each bearing housing (Fig. 30.6), in the direction of lower stiffness.

(a) (b)

PERMANENT
MAO.HET - -t_-I-...Y

Fig. 30.6 Seismic vibration sensor: a) cross section, b) installed on bearing housing.
(Courtesy SKF.)
904 Monitoring and Performance Testing

A~

VIEW A-A
( AS VIEWED FROM DRIVER)

Fig. 30.7 Proximity probes 90 degrees apart in bearing housing.

2. Antifriction bearing "noise" -When lubrication is inadequate or there are minor defects in the rolling
elements or races, antifriction bearings exhibit very high frequency "noise." With appropriate transducers,
this noise, if it is present, can be detected and then its significance judged from charts provided by the
various instrument manufacturers. Periodic measurement of antifriction bearing "noise" is now a well-
established technique for avoiding unexpected failures of these bearings.
3. Shaft to housing-When a pump's rotor is supported in hydrodynamic bearings, the inherently lower bearing
stiffness means that the rotor can develop significant motion before it is noticeable in bearing housing
vibration. Given this, pumps with hydrodynamic bearings, particularly those of high energy density (high
power-to-weight ratio for the rotor or low rotor-to-casing weight ratio), are equipped with means to measure
the motion of the shaft relative to the bearing housing. The usual means are proximity probes, electronic
devices that produce, in conjunction with their driver, an output proportional to the gap between the probe
tip and the target surface. Generally two probes, 90 deg apart, are installed at each end of the rotor (Fig.
30.7). This arrangement allows measurement of the rotor's displacement in any direction, and the shaft's
orbit can be derived to aid analysis. Adding a probe to produce a once-per-revolution signal, what is termed
a Keyphaser®, provides a direct speed measurement and enables the phase of vibration at each end of the
rotor to be determined, which is necessary for any serious analysis. Finally, in pumps with hydrodynamic
thrust bearings, two additional probes are often used to monitor the rotor's position and thereby detect a
thrust bearing failure. Two probes are used for dual voting to guard against spurious trips. Figure 30.8
shows the location and orientation of the proximity probes used in Class 3 or 4 monitoring of a between
bearings pump.
4. Absolute shaft motion--On occasion it is desirable to measure the absolute motion of a shaft mounted in
hydrodynamic bearings. In these instances, the relative motion of shaft to housing determined by proximity
probes is converted to absolute by subtracting the bearing housing motion obtained from a seismic transducer
located in the region of the proximity probe (Fig. 30.9).
Monitoring and Performance Testing 905

Fig. 30.8 Typical proximity probe locations for a critical multistage pump rotor.
(Courtesy Bentley Nevada Corp.)

Shaft absolute vibration


(shalt relative minus casing absolute)

t~"
.!!! Shaft relative vibration
"0
> ~~~~.~~~~-

~ Shaft absolute vibration


o'5 Casing absolute vibration

~SO'2

Time ~

Fig. 30.9 Measuring absolute shaft motion.


(Courtesy Power Magazine)
906 Monitoring and Performance Testing

For specific details on the selection and installation of vibration transducers, consult the manufacturers
of this equipment and API-670, a standard covering vibration measurement (and temperature measure-
ment; see earlier) in the hydrocarbon-processing industry.

Shaft Seal

The extent to which a pump's shaft seal can be monitored varies with the type of seal, and then with
detailed variations of each type. Packed box seals do not really allow any worthwhile monitoring, a
disadvantage largely offset by their inherent need for periodic adjustment. Axial face (mechanical) seals
often depend on either pressure or temperature control to provide the environment they need to operate,
and therefore it is prudent to monitor this in critical applications. Breakdown-type seals with condensate
injection have long been fully automated, and therefore can be readily monitored.

1. Pressure-When the pressure at the shaft seal is raised or lowered by injection or bleed-off across a breakdown
bushing (see Chaps 8 and 9), that pressure will vary as the breakdown bushing clearance increases with
wear. Monitoring the pressure at the seal with a gage or transmitter provides a forewarning that the seal's
environment is deteriorating, and therefore the opportunity to avoid an unscheduled shutdown caused by
seal failure.
2. Temperature-In shaft seals where the temperature at the seal has to be kept below the pumping temperature
and the means of doing this depend on a heat exchanger that is part of the pump unit, as it is in API plans
21 and 23, the temperature of the flow to the seal faces should be continuously monitored, and an alarm
sounded if the temperature rises above the allowable value. The monitoring device can be a temperature
switch or a transmitter with the appropriate secondary instrumentation. Failure to follow this precaution will
likely result in premature seal failure, particularly if water is used for cooling and the water is hard (prone
to scaling).
3. Leakage-Current practice with shaft seals is to use tandem or double mechanical seals (see Chap. 9) when
leakage of the pumped liquid is hazardous. Tandem seals enable the detection of primary seal leakage by
either a rise in pressure at the secondary seal or a rise in liquid level in the buffer liquid reservoir if one is
used. By definition, double seals have between them a barrier fluid at a pressure above the sealed pressure,
so a fall in barrier fluid pressure indicates leakage across either the inner seal (to the pumped liquid) or the
outer seal (to atmosphere). With either seal design, buffer or barrier fluid pressure or level or both is
monitored to warn of leakage should it occur.
4. Drain temperature-Breakdown-type seals with condensate injection use seal drain temperature as the control
parameter. Because aberrations in seal drain temperature can cause significant difficulties in other aspects
of pump operation, seal damage and high O2 in the feed water to name two, continuous monitoring of this
parameter is warranted. As with the sensor used for injection control, the sensor for monitoring must be
installed so as to not minimize its thermal inertia otherwise it will not detect critical changes in seal
drain temperature.

Balancing Leak·OtT Flow

In multistage pumps whose rotors have tandem impellers and a single balancing device, measuring
the leak-off flow from the balancing device gives a direct indication of the condition of that running
clearance, and an indirect indication of the condition of the rest of the pump's running clearances since
they tend to wear at the same rate. Balancing leak-off flow is obtained by measuring the differential
pressure across an orifice or flow nozzle in the balancing leak-off line. Readings can be taken periodically
and compared with the value when the pump was new, or the device measuring the differential pressure
can be set to give an alarm when the flow has increased by, say, 50 percent.
Monitoring and Performance Testing 907

VIBRATION

100
FLOW - t BEP
Fig. 30.10 Typical variation in vibration with pump flow.

Pump Flow and Speed

As machine monitoring has become more sophisticated, it has been appreciated that pump flow, as
a fraction of its design or best efficiency flow, has a direct effect on vibration (Fig. 30.10), and in some
cases bearing temperature, therefore the assessment of a pump's condition should be based on operating
data taken at the same nominal flow and speed. The means of measuring flow and speed are dealt with
in field testing. To get good results it is not necessary to determine the other two aspects of a pump's
performance, head, and power, and therefore determining these is left to field testing if that becomes nec-
essary.

ACTION ON MONITORED PARAMETERS

The preceding section dealt with many of the actions usually taken on deviations in the monitored
parameters of operating conditions and pump operation. For a convenient ready reference all the monitored
parameters and the usual actions are summarized in Table 30.4. The list is not exhaustive, and not all
rotating machinery engineers agree with all the actions, but it is a starting point.

Performance Testing
As noted at the beginning of this chapter, field performance testing is used primarily to verify the
operation and performance of emergency pumps, and occasionally to determine the performance of an
installed pump or the system in which it is operating. With respect to verifying the integrity of emergency
and safety related pumps, the following material is typical only, and does not supersede any requirements
stipulated by local, state, or federal government regulations with respect to specific installations.
The quantities to be measured in the field testing of a centrifugal pump depend upon its application
and the extent of monitoring in normal operation. Table 30.5 compares an emergency pump with one
that has Class 3 monitoring.
908 Monitoring and Perj'omumce Testing

Table 30.4 Action on Monitored Parameters

Parameter Allow start Control Alann Trip

Operating conditions
Suction
pressure 6. 6. 6.
strainer flP 6.
level 6. 6. 6.
SG 6.
NPSHA 6.
loss of prime 6.
Discharge temperature 6. 6.
Minimum flow
Bypass flow not established 6.
Seal injection
pressure 6. 6.
filter flP 6.
temperature 6.
Lubricant
pressure 6. 6. 6.
flow 6. 6. 6.
filter flP 6.
temperature 6.
Pump condition
Casing
6.T top to bottom 6.
6.T to suction temp 6.
total temperature 6. 6.
Bearing temperature 6. 6.
Vibration 6. 6.
Shaft seal
pressure 6.
temperature 6.
leakage 6.
drain temperature 6.
Balancing leak-off flow 6.

TEST ARRANGEMENTS

Although it is possible to obtain reasonable test data from an improvised setup, the best results are
obtained from installations whose design has taken field testing into account when that was a requirement.
Of the quantities that have to be measured during a field test, flow rate poses the greatest difficulty.
Pumps that operate during normal conditions need a means of flow metering incorporated in the discharge
line, or the suction line if there is sufficient NPSHA (Fig. 30.11). High-energy pumps will have some
form of flow metering to control their minimum flow bypass system, which may be sufficient for field
testing. Emergency pumps generally cannot be run through the emergency system, and therefore have
to be provided with a bypass (Fig. 30.12). The means of flow metering can be incorporated in the bypass
line or, as before, in the suction line.
High-energy pumps are provided with a minimum flow bypass to allow their startup and shutdown
Monitoring and Performance Testing 909

Table 30.5 Test Quantities to be Measured

Quantity Symbol Emergency pump Monitored pump

Speed N ~ ~
Suction P, ~ ~
Suction temperature T, (1) ~
Discharge pressure Pd ~ ~
Differential pressure M ~ ~
Flow Rate Q ~ ~
Power P (2) ~
Balancing leak-off flow Qb (2) ~
Vibration
displacement Vd ~
velocity Vv ~
Lubricant level or pressure ~
Bearing temperature Tb ~

(I) Pumps nonnally tested with cold water.


(2) Optional quantity; measurement not required by usual "codes."

FLOW METER
- May be in common discharge
line if only one pump run
Note need for straight run
upstream

n
. . --_..J
r ... _-.,

U
Alternate location of
flow meter provided
NPSHA sufficient

Fig. 30.11 Test arrangement for normally running pumps.

without risk of rapid wear of sudden failure. In most installations, these bypasses are sized for a flow
too low to allow meaningful evaluation of a pump's mechanical operation and hydraulic performance.
There are two reasons for this. First, at flows well below design (best efficiency point) various hydraulic
phenomena typically produce high and erratic pump vibration (Fig. 30.10). Comparative data gathered
under these conditions can have inherent variations greater than the deviations allowed by the "codes"
that govern safety related pump inservice testing (ASME Section XI, Subsection IWP [30.2]). Second,
the usual cause of centrifugal pump performance deterioration, wear at the running clearances, has very
little effect on the performance at low capacities (Fig. 30.13). The import of this is that a series of tests
910 Monitoring and Performance Testing

ORIFICE
Back anti-flash

FLOW METER
BYPASS MUST BE FROM POINT
UPSTREAM OF PUMp·S CHECK &
BLOCK VALVES

PRESSURE BREAKDOWN DEVICE


(if necessary)

--..,
&'---..1

U
Alternate location of
flow meter

Fig.3O.12 Test arrangement for pumps used only during plant emergency.

at low capacity may not detect a significant drop in perfonnance at rated flow. To provide meaningful
results, bypasses intended for perfonnance testing need to be sized for flows on the order of 60 percent
of best efficiency flow.
By the "codes" for testing safety related pumps flow metering has to be accurate within ±2 percent.
The other major code for field testing, ASME PTC 8.2 [30.3], requires similar accuracy for its Class B
tests. Accuracy of this order dictates care in the selection and installation of the primary element and
associated instrumentation. Table 30.6 offers some general guidance; consult Fluid Meters: Their Theory
and Application [30.4] for a detailed treatment.
In assessing the accuracy of various flow meters, it is worth noting the distinction between accuracy
and precision. Accuracy refers to the deviation from a "known" value, whereas precision is the variation
between readings of the same value. Since field testing usually involves the comparison of subsequent
quantities with a reference quantity, precision is clearly more important than accuracy. The latter must,
of course, comply with any applicable "code" requirements.
Pressures should be measured using calibrated, standard test gauges or transducers, in preference to
the normal service gauges. Tapping fonn and location should follow the guidelines shown in Fig. 30.14.
Gauges should be damped with a needle valve to the point where there is still minor motion of the
pointer. The measurement of pressures below 1.5 bar (20 psig) requires particular care to obtain accurate
data. It is generally easier to and more accurate to determine a pump's differential pressure by measuring
suction and discharge pressures, then subtracting suction from discharge.
Tests run on any liquid other than cold water should include measurement of the liquid temperature
Monitoring and Performance Testing 911

Table 30.6 Flow Metering

Type Comment

Plate orifice Low-cost primary element; expensive


installation (requires longest upstream
straight pipe run); high head loss; "noisy"
signal at high pressure differentials
Venturi High-cost primary element; less expensive
installation (shorter upstream pipe run);
low head loss; low signal "noise"
Elbow Low-cost primary element; calibration
required for accuracy to "code"
requirements
Magnetic l High-cost instrument; low installation cost;
reported high accuracy over a wide flow
range
Ultrasonic l High-cost instrument; low installation cost;
multipath instruments reportedly accurate
over a wide flow range with a distorted
flow profile
lConsult manufacturer or specialist literature for details.

HEAD
f6 H
...... ,I~
*I .........

J 1.-6 L

POWER

60 100
FLOW - % BEP

Fig. 30.13 Effect of wear at internal running clearances on pump head capacity and power characteristics.
912 Monitoring and Performance Testing

Table 30.7 Power Measurement

Method Comment
Motor current Indicator only; precision depends on quality of
ammeter
Motor kilowatt absorbed Requires test grade kilowatt meter and associated
electrical apparatus; accurate for direct drive
pumps provided motor calibration is by
dynamometer, that is, wire to shaft
Torque plus speed Requires special coupling spacer and corresponding
instrumentation; strain gage-type reportedly
accurate within 2 percent; laser measured phase
shift type reportedly accurate within I percent

at the pump suction. Depending on the compressibility of the liquid being pumped and the pump's
differential pressure, it may also be necessary to measure the temperature at the pump's discharge. And
if the flow meter is in a bypass, downstream of the flow control device (as in Fig. 30.11), the temperature
at the inlet to the flow meter should be measured as well so the flow rate can be corrected to suction
conditions. Temperature measurements are made with a calibrated thermometer or temperature detector
(thermocouple or RTD).
Power measurement is usually not required in the periodic testing of emergency pumps. When it can
be, or at least some indicator of power can be, it is valuable data for the diagnosing the cause of pump
deterioration. Three approaches are discussed in Table 30.7.
Speed is measured most easily using photoelectric or magnetic pulse counters, or taken from the
vibration measuring system's Keyphasor® when the pump is so equipped. All these methods have an
accuracy well within the safety related pump test "code" requirement of ±2 percent. As noted in most
the test "codes," speed measurement is mandatory for variable speed pumps. It does no harm, however,
to measure the speed of motor driven pumps as well, because on occasion the cause of performance
problem is in the motor or its electrical connections.

ANALYSIS OF RESULTS

When performance testing is carried out to verify the integrity of emergency and safety related pumps,
the relevant "code" for the particular service states how the test data are to be analyzed and the action
to be taken in the event of significant deviations. These codes generally limit the allowable change in
vibration and head developed at a given flow. A moderate increase in vibration or a moderate reduction
in head are typically cause for more frequent testing, whereas a large increase in vibration or a large
reduction in head (which means less flow available) dictates that the pump be taken out of service and
its performance corrected.
Performance testing carried out to help determine why a pump is not performing as it should is most
meaningful when it produces several points of flow, head, power, and speed. Corrected to constant speed
and then compared to earlier test results, the values of head and power at some capacity close to BEP
can quickly give an indication of what is wrong with the pump. The three principal circumstances and
their causes are:
Monitoring and Performance Testing 913

LSd' 0.12-0.25 ins OIA

!\-; d

4
d

r 5 - 10 DIAS
STRAIGHT , P

NOTE: With set-up shown, ~ P lower than


pump t::. P by head loss across
reducers.

Fig. 30.14 Pressure measurement tapping location and form.

HEAD -.- ............. ..... .....


..... ......
...... .,

_.-.-.-._.-
_.-.-:..0 0 0
POWER <.r""

FLOW

- •- SHOP OR PREVIOUS TEST


- 0 - FIELD TEST

Fig. 30.15 Typical effect of problem at pump suction or impeller obstruction.


Head and power at given flow are low.
914 Monitoring and Performance Testing

HEAD -.-.- ..... ..... .....


..... .......
.,

__ -~-Lr"'"

POWER 0--

FLOW

- •- SHOP OR PREVIOUS TEST


--0- FIELD TEST

Fig. 30.16 Typical effect of obstruction in pump casing or discharges piping upstream of pressure gauge.
Head is low, power is normal at given flow.

1. Head low and power high (Fig. 30.13). Typical of high internal leakage through worn running clearances
or across internal seals. The impeller is absorbing more energy than it should for the given flow but is
producing less head, indicating that the impeller flow (as distinct from the pump flow) is higher than it
should be.
2. Head and power low (Fig. 30.15). When both the head and power are low, it is evident the impeller is
absorbing less energy than it should, thus indicating a problem at the pump suction. The usual problems
are poor suction piping layout (Chap. 28), an obstruction in the suction piping or impeller, or air leakage
into the pump.
3. Head low and power equal (Fig. 30.16). In this case the impeller is absorbing the energy it should, hence
producing the correct impeller head, but more head than normal is being lost between the impeller outlet
and the point of pump discharge pressure measurement. The usual causes of this are an obstruction in the
collector (diffuser or volute) or in the piping between the pump and the discharge pressure gage.

The third purpose of performance testing, assessing wear at the internal running clearances, enables
pump overhaul at the optimum point on the basis of energy versus overhaul cost. In this case data are
used to determine the additional cost of producing rated capacity over some period, say a year, then
that cost is compared the cost of restoring the internal clearances to their new values. This assumes
pump vibration does not increase significantly as the internal clearances increase with wear, an assumption
that is not always correct.

BIBLIOGRAPHY

[30.1] API-670, Vibration, Axial-Position, and Bearing Temperature Monitoring Systems, 3rd Edition, 199?, American
Petroleum Institute, Washington, DC.
Monitoring and Performance Testing 915

[30.2] ASME Boiler and Pressure Vessel Code, Section XI, Subsection IWP, Inservice Testing of Pumps in Nuclear
Power Plants, 1990, American Society of Mechanical Engineers, New York, New York.
[30.3] ASME PTC 8.2, Power and Test Code, 1991, American Society of Mechanical Engineers, New York, NY.
[30.4] Fluid Meters: Their Theory and Application, 6th edition, 1971, American Society of Mechanical Engineers,
New York, N.Y.
31
Maintenance

To provide an adequate return on the capital invested in it, thereby allowing further investment, a process
must run at or above its planned production rate and at or below planned cost. The objective of maintenance
is conservation of the equipment to help achieve this. Looking at rotating machinery as a class, its effect
on process output and operating costs is associated with availability, energy consumption, and the cost
of maintenance.
Of the rotating machines used in or various plants and processes, the second most common is the
centrifugal pump. Given the extent of their use and the variety of services to which they are put,
centrifugal pumps can have an effect on the financial output of plants that far outweighs the capital
spent on them. When that happens, the pumps causing lost production, high energy consumption, high
maintenance costs, or some combination of all three become serious maintenance problems.
Maintenance involves monitoring equipment to verify its condition, making adjustments as necessary,
carrying out periodic preventive maintenance, restoring equipment to new condition as required, and
correcting equipment whose service life or performance is impairing plant performance. Chapter 30 has
dealt with monitoring and performance testing, therefore this discussion covers the activities associated
with adjustments through correction.
Given the variety in types, sizes, component parts, and design of centrifugal pumps, this discussion
of maintenance can only be general in nature. The manufacturer's instruction should always be studied
carefully before developing a maintenance program and before attempting to service any particular pump.

MAINTENANCE PROGRAM

What is necessary in terms of staffing, skill level, and machine tools and equipment for centrifugal pump
maintenance in a given plant depends on the likely rate of pump wear, the difficulty of pump repair,
and the availability of parts and tools to quickly carry out the repairs. Assessment of the following four
factors will provide further guidance to determining what is necessary:
1. Nature of the plant or process. Water pumping is generally relatively benign; slurry pumping is almost
invariably very aggressive.

916

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
Maintenance 917

2. Complexity of the pump types employed. Small single stage pumps are the simplest; high speed (hence
high energy density) multistage the most complex.
3. Spare parts availability. Parts for one of a large manufacturer's mass produced pump lines would typically
be readily available; those for a special design often are not, and therefore on-site restoration may be necessary.
4. Proximity of manufacturer's service facilities. A well equipped service facility close to the plant reduces
the need for major investment in machine tools and equipment for the plant's maintenance shop; being a
long way from such facilities raises the need for major machine tools and equipment at the plant site.

Once the maintenance facility and staffing is decided, the next critical issue is to establish a good
equipment record system. The value of this cannot be overemphasized. Although this may not be evident
at plant startup, it is critical to the effective management of spare parts (avoiding unnecessary inventory)
and to the quick correction of troublesome machines. The data that need to be kept in this system are:

• Equipment description and purchasing history


• Manufacturer's drawing list
• Manufacturer's parts list
• "As built" clearances and materials
• Manufacturing quality control documents
• Startup data (of a tie to same in the monitoring system)
• Preventive maintenance history
• Overhaul history, including any changes made

PREVENTIVE MAINTENANCE

The procedures covered by this term are carried out as needed or periodically to keep machines running
at their optimum. In order of decreasing frequency, the usual procedures for modem centrifugal pumps are:
Seal adjustment: packed box and some forms of mechanical seals for slurry service need adjustment
to take up wear and minimize leakage. Not adjusting such seals results in high leakage and can lead to
major damage to the seals or bearings or both if it is allowed to persist.
Lubrication: bearings require periodic replacement of their lubricant. In some designs the assembly
has to be dismantled, cleaned, and re-Iubricated, usually with grease. Other designs use oil, either in the
bearing housing or from a force feed system, which has to be changed. The frequency of lubricant
replacement depends on the machine, the service, and the environment in which it is operating.
Cooling: various services require cooling of the shaft seal or bearings or both, by either liquid, usually
water, or air. Depending on the condition of the coolant, and in some cases the cooled liquid, periodic
cleaning of the heat exchange surfaces is necessary to maintain adequate cooling.
Wear: some pumps with an axial clearance at the impeller (see Chap. 4) can have their impeller
position adjusted to compensate for wear. This would be carried out whenever the pump showed an
obvious drop in performance, or perhaps periodically until a trend was determined.
Drive: coupling alignment tends to change with time, so periodic checking and correction is warranted.
Similarly, V belt drives require retensioning to compensate for stretch. Failure to attend to the drive can
result in serious damage to both the drive and the connected equipment.
As noted, the timing of these preventive maintenance procedures depends on the pump service and
operating environment. In the absence of better information, the following can be used as a start then
918 Maintenance

modified to reflect experience. Note that under adverse conditions, it may be necessary to shorten the
typical timing between some of these procedures.

As needed Seal adjustment


Every 6 months Lubrication, drive
Every 12 months Cooling, wear

The preventive procedures given do not cover the pump's driver; see the manufacturer's instructions
for what is to be done in this connection.

OVERHAUL

The age-old question concerning overhauls is When? There are three approaches? One is to dismantle
the machines periodically, regardless of evident condition, and replace any parts that look dubious. The
argument for this practice is that it essentially assures the machine will run trouble-free until the next
overhaul. Provided the examination of the parts is sufficiently thorough and the period consistent with
component life, this approach is valid. An example of it is aircraft maintenance. When examination of
the parts is not so thorough and the period is chosen randomly, periodic overhaul can result in a sense
of false security. Component failure by fatigue is the usual shortcoming. At the other extreme, the
practice can unnecessarily raise the cost of maintenance.
The second approach is to "run it till it breaks." When the service is not critical or the pump is fully
spared (and the spare is immediately available), and the added cost of consequent damage in the
breakdown can be tolerated, such an approach can be lived with. There are, however, a lot of "ifs," and
it is questionable whether the total cost of this approach is really as low as claimed. Compounding that,
the cavalier attitude inherent in this approach often carries over into the maintenance process, leading
to breakdowns more frequently than ought to be the case.
A refinement of the second approach yields the third. By seeking to learn as much as feasible about
a machine's condition while it is running, it is possible to develop a reasonably accurate profile of its
condition and, using a little judgment to determine when it should be taken out of service for overhaul.
This practice is generally ~own as "on condition maintenance." The data used to establish the condition
profile are behavior and performance, hence the importance of measuring and recording these data at
startup. Behavior, i.e. vibration, temperature, and noise, are useful data, but are often lagging indicators
of trouble. When behavior is combined with performance, the picture is more complete and usually
clearer. See Chap. 30 for a detailed discussion of monitoring to determine pump condition.
Of the three approaches, the third is the most favored since it promises tolerable availability and
reliability at reasonable expense.
Under an "on condition maintenance" program, a pump will be taken out of service for overhaul in
one of the two following circumstances:

1. One or more aspects of its behavior has increased beyond acceptable limits or has changed suddenly by a
significant amount (see Chap. 32).
2. Its hydraulic performance has deteriorated to the point where plant production is being affected, or the cost
of overhaul will be quickly paid back by the resultant reduction in energy consumption (see Chap. 30).

A sudden change in operating behavior is usually a result of misadventure, such as foreign material
passing through the pump, accidental dry running, or operating conditions that precipitated the failure
of a component within the pump. In some cases, such misadventure can cause catastrophic failure,
Maintenance 919

leading to the pump's immediate shutdown. In these cases the consequential damage is often extensive,
and returning the pump to new condition involves much more than just restoring its running clearances.

Dismantling the Pump


Patience and planning are fundamental to successful pump overhaul. There are more spectacular
approaches, but the spectacle tends to wear off when the overhaul has to be repeated to correct errors.
The first step is to isolate the pump from the system and the driver from its power supply. This
should be done using a proper "tag-out" procedure to ensure there is no risk of accidentally putting
liquid to the pump or starting its driver while it is being worked on. The pump should then be drained
to the extent possible.
Before the pump is dismantled, check and record its coupling alignment and shaft spacing. Remove
auxiliary piping (seal, lubricating oil, and cooling water) and instrumentation to the extent necessary to
allow dismantling of the pump. If the complete pump is to be removed, disconnect the suction and
discharge piping, noting any major strain and its effect on coupling alignment.
Dismantle the coupling, bearings, and shaft seals as required by the particular design to allow the
casing to be opened. Remove any casing dowels and all the nuts of the casing bolts. Horizontal pumps
with axially split casings should have the casing upper half lifted straight up to prevent damage to
internal parts (Fig. 31.1). Using slings positioned at "quarter points" on the rotor (to minimize the
bending moment), lift it vertically to avoid damage to the impellers, wearing rings, and other internal
parts. Similar principles applies to vertical pumps with axially split casings, except that now the front

Fig. 31.1 Pump with axially split casing.


920 Maintenance

casing half must be moved away horizontally (Fig. 13.15), and the rotor lifted from the upper end of
its shaft.
Horizontal pumps with radially split casings have their inner element or cartridge removed laterally.
In the case of overhung pumps of the back-pull-out type, a spacer coupling is provided between the
pump and its driver. The length of the spacer is such that the pump's cartridge (assembly of bearing
frame,· casing cover, shaft seal, and impeller) can be removed without excessive canting (Fig. 12.22).
Single and multistage between bearings pumps have their inner element or cartridge withdrawn from
the outboard end of the pump (end opposite the driver). When the pump is multistage, inner element or
cartridge withdrawal generally involves the use of a carriage on wheels (Fig. 3.16) or a cradle without
wheels bolted to the casing.
Dismantle the pump's rotor, inner element, or cartridge in an orderly manner. Mark the various parts
to ensure correct reassembly. Examine all individual parts and important joints carefully. Visual inspection
is good for wear, as least qualitatively. Measure critical parts, recording the dimensions on a form
suitable for the pump type and determine the "worn" running clearances. Inspect dynamic parts by NDE
such as liquid penetrant or magnetic particle. Do the same for pressure containment parts if there are
any suspicious results from visual inspection. Look for and note signs of corrosion or abrasive or
cavitation erosion. Take photographs of NDE findings and any corrosion or erosion to supplement the
dismantling report.

Cause of Damage
Compare the pump's condition with what could be considered "normal" for the service. This is a
difficult task because many factors influence pump wear. There are, however, reasonable classifications
of "normal," and it is worth the trouble to try to get a better indication of whether what has been
encountered is to be expected or not. The "not" is the important part because that says there is scope
for improvement, which may well constitute a cost saving for the plant. Finding the root cause of
abnormal damage will require quite a deal of dedication, but it is worth the effort to avoid having to
needlessly overhaul a pump time and time again. In a very general sense, Table 31.1 can serve as a
guide to "normal" pump damage and the likely cause of abnormal damage. The guide is based on liquid
condition. See Chap. 32 for further listings of possible causes and Bloch [31.1] for a detailed treatment
of the whole subject.
Whenever possible, correct the cause of abnormal damage. In many instances the cause is relatively
simple, requiring only perseverance to find it. At the same time, inquire into progress if what is considered
"normal" damage is proving hard to tolerate. Since the pump was built, there may have been progress
in a direction that would help limit the damage. In many cases, the damage being encountered is a
consequence of the pump not being operated as originally intended, a fact that can be discerned by a
careful examination of the parts (see Chap. 32) and a thorough investigation of the actual operating
conditions. In these cases, the manufacturer is usually able to provide new design parts to either achieve
a better hydraulic fit or raise the pump's mechanical integrity or both, and so extend its service life
or MTBR.

Maintenance of Specific Parts


Maintenance of the following specific parts is described in the chapters listed.
1. Casings-Chapters 2 and 3
2. Impellers and wearing rings-Chapter 4
3. Shafts and shaft sleeves-Chapter 7
4. Shaft seals-Chapters 8, 9, and 10
5. Bearings-Chapter 11
Maintenance 921

Table 31.1 Guide to Pump Damage

Liquid Condition Damage

Clean Normal
Clearances eroded, minimal damage to other hydraulic parts, shaft seal, and bearings.
Abnormal
• Abrasive erosion: solids in the liquid
• Corrosion erosion; mild corrosion
• Cavitation erosion: insufficient NPSH or operation at low flows
• Adhesive erosion (scoring at clearances): high rotor forces, pump distortion, running dry
• Short seal life: unsuitable seal, poor seal environment or seal angularity
• Short bearing life: low flow operation, lubricant contamination from high seal leakage or
atmosphere, insufficient cooling, worn clearances, undersize bearings
• Cracking: corrosion fatigue, overstress of thermal or mechanical origin, manufacturing
errors (thin sections or sharp corners), resonance
Corrosive Normal
Clearances eroded, minimal metal loss from other parts unless pump intentionally
sacrificial or being used at limit of material.
Abnormal
As for clean liquid.
Abrasive Normal
Abrasive erosion of all hydraulic parts, severity depending on service conditions; seal
similar.
Abnormal
Similar to clean liquid after allowing for abrasive service, need to distinguish between
damage from abrasives and other causes.

Returning a pump to new condition is usually done by a combination of replacing and restoring parts.
Some parts, badly eroded impellers for example, cannot be easily restored and are therefore routinely
replaced. Other parts, such as wearing rings, can be restored to new condition by turning to "clean up"
and then installing a corresponding undersize part to restore the clearance to its new value. When there
is a choice between restoring and replacing, the following factors need to be considered:

• Cost of replacement versus restoration


• Availability of replacement parts
• Integrity of restored parts
• Effect of restoration on performance

In some instances the integrity of restored parts can be superior to that of replacements. Such cases
usually reflect a "short cut" in the original equipment purchase, and could probably be rectified if the
manufacturer were made aware of the field problem being encountered.
Rebuild the pump to original specifications unless there are very sound and well researched reasons
to do otherwise. The term "well researched" should include consultation with the manufacturer (see
Cause of damage). On many occasions, users of equipment have made well engineered changes and
realized a notable improvement in service life. There are, unfortunately, just as many instances where
exactly the opposite has been the case. If there is a rule to be followed in refining machine design, it
is: "no one has a monopoly on ideas."
922 Maintenance

Reassembling the Pump


As the pump is assembled, a building record of important fits, clearances, and movements should be
completed; see Figs. 31.4 and 31.5 for typical impeller and wearing ring dimension record sheets. Doing
this serves as a checklist and as the base record for comparison when the pump is next dismantled.
Impellers must be mounted on the shaft so they will rotate in the correct direction, always away from
the curvature of their vanes (Fig. 31.2).
There are 4 fundamental rules for assembling rotors: they are:
1. Slide fit impellers must be assembled onto a clean, burr-free shaft lubricated with molybdenum disulfide
powder (not grease) or equal. If the impeller becomes more than hand tight, it should be removed by heating
the back shroud (the shrouds and center hub of double suction impellers) to expand the bore. Do not use
mechanical force; doing so will likely result in the loss of both the impeller and shaft.
2. To avoid a bent rotor, the faces of all abutting joints (Fig. 31.3) must be square to the shaft axis. If the
faces are not square, the rotor will bend either when its shaft nuts are tightened, or in service when hydraulic
forces impose high loads on the abutting joints.
3. Rotors with shrink fit impellers should be assembled vertically to reduce the risk of bowing the rotor from
differential cooling of the hubs.
4. The assembled rotor should be checked for runout. When the rotor has abutting joints, its runout should be
checked with the shaft nuts both loose and tight to ensure it will remain straight. (It is to avoid this difficulty
that many designs of multistage pump have their impellers independently located with a step on the shaft
or a split ring; see Chap. 7).

When the pump is multistage or single stage between bearings running above 3,800 rpm, the dynamic
balance of the rotor must be checked if any major rotor component has been repaired or replaced. A
generally applicable balancing standard is

Fc ~ O.lW
where: Fc = centrifugal force produced by unbalance (N or lb.)
W = rotor weight (N or lb.)

In terms of the balancing grades given in ISO 1940 [31.2], this standard allows G.6.3 up to 1,500
rpm, G2.5 to 3,800 rpm, and G1.0 to 9,400 rpm.
For radially split pumps, assemble the remainder of the inner element or cartridge, following the specific
instructions given by the manufacturer. The cartridge for overhung pumps is relatively straightforward.
Multistage pump inner elements, whether radially or axially split, are somewhat more complex. Complete
cartridges for multistage pumps require a jig to simulate the casing for setting the rotor position and
locking the assembly for installation.
Install the rotor or element. Close the casing, taking particular care with the split-joint gasket of axially
split casings (see Chap. 2), and the bolt tightening sequence of both axially and radially split casings.
In designs that allow for it, check the rotor's radial and axial position. The radial position is checked
by determining whether the rotor is centered in its running clearances. Variations beyond limits are
corrected by adjusting the position of the bearings (Fig. 11.17). The correct axial position can simply
be in the center of available endplay or a specific distance from one position. Adjustment is generally
made either in the thrust bearing or by a shim on the shaft. Once the rotor's axial position is set, the
sleeve of cartridge mounted mechanical seals (Fig. 9.25) can be locked in place.
When the whole pump has been removed, reinstall it taking the same care with piping strain and
coupling alignment as required for the original installation. In those cases where whole pumps are
exchanged and by definition not returned to their original position, piping strain can be a problem. The
problem arises because pump casings have, of necessity, a tolerance in their dimensions. When nozzle
Maintenance 923

DIRECTION
OF ROTATION
..- sa

Fig. 31.2 Direction of impeller rotation.

ABUTIING JOINTS

Fig. 31.3 Four-stage pump rotor.


924 Maintenance

UNIT N'!
1='UMJiONt

rr~1 __
'4('1) __
!
:u
I
LL~
-I

~"TAGe; 'A' '~ ·C:


IF·
'aND

~.D

4'"
STH
&....

---
R&:M"A~.:

ic.KECKeO,
IMPELLER
ICAT.:

Fig. 31.4 Record card for impeller dimensions.

locations are determined by castings, the tolerance can be quite large. Pumps that are overhauled by
restoring or exchanging an element or cartridge (assembly of the rotor plus stationary parts) avoid
this problem.
Put the overhauled pump back into service using the start-up sequence set out in Chapter 29, Operation.
Lest it be overlooked, add the overhaul data and the new baseline conditions to the equipment
record system.

SPARE AND REPAIR PARTS

To ensure that a pump can be quickly put back into service in the event of an unexpected shutdown,
an adequate stock of spare parts must be maintained at the plant. The size of this stock of parts depends
on the severity and criticality of the pump service, the uniqueness of the pump design, and the number
of pumps in the plant using the some or all of the same parts. A typical list of spare parts for a single
pump is given in Table 31.2. It is often advisable to stock a complete spare cartridge or element (rotor
plus stationary parts). This allows the flexibility of changing the cartridge or element when necessary,
then overhauling the worn assembly at leisure.
Maintenance 925

UNIT N~

~ ~UMPN!
1ICt1"l - r--

- .. -
:11, :<

~~ " ..c:a .....o N .

tl!'A<".a .,.: ." "",l:_m- -"T\.Mt"


,-

.-......
'l'lllO

ST..

WeARINQ, R. IN ~ .
ICHECKED :
)=-n-.,

Fig. 31.5 Record card for wearing ring dimensions.

Table 31.2 Typical Recommended List of Spare Parts

1. A set of casing wearing rings.


2. A set of impeller wearing rings (if used).
3. A set of interstage bushings for multistage pumps.
4. An impeller, stock bored, for multistage pumps.
5. A diffuser for multistage pumps (if used).
6. A spare balancing device made up of rotating and stationary parts (if used).
7. A set of mechanical seals (if used).
8. Two sets of shaft sleeves and 4 sets of packing (if used).
9. A set of bearings (if anti-friction) or bearing bushings (if sleeve).
10. A thrust collar and a set of thrust shoes if a Kingsbury type thrust bearing is used.
11. A complete set of gaskets and shims.
926 Maintenance

Spare parts should be purchased at the time the order for the complete unit is placed. Depending on
the contemplated method of wearing ring overhaul, the spare rings are ordered to the same size as those
used in the new pump or bored undersize (for casing rings) and turned oversize (for impeller rings).
When hardened casing rings are used, spare rings are often furnished in specific undersizes. With this
arrangement, running clearances are restored by turning the worn impeller hubs to the next undersize
and installing the appropriate casing rings. (The balancing device, if used, should be similarly sized).
The pump serial number and size as stamped on the pump nameplate should always be given when
ordering spare and replacement parts after the pump has been received in the field, so the manufacturer
can identify the pump and furnish repair parts of the correct size and materials. Most centrifugal pumps
are of standard design and a great number of combinations are made for each size of casing, using
different impeller sizes and designs. Without an identification number, the pump manufacturer would
not be able to furnish correct parts, even though the type and size of pump might be known.

BIBLIOGRAPHY

[31.1] H.P. Bloch, and F.K. Geitner, Machinery Component Maintenance and Repair, Gulf Publishing Co., Houston,
Texas 1985.
[31.2] ISO 194011, Mechanical Vibration-Balance Quality Requirements for Rigid Rotors-Part 1 : Determination of
Permissible Residual Unbalance", 1986, International Organization for Standardization, Geneva, Switzerland.
32
Diagnostics of Field Problems

The operation of a centrifugal pump may be affected by hydraulic or mechanical troubles. Hydraulic
troubles may cause a pump to fail to deliver water altogether; or, the pump may deliver an insufficient
capacity, develop insufficient pressure, lose its prime after starting, or consume excessive power. Mechani-
cal difficulties may appear at the shaft seals and bearings, or produce vibration, noise, or overheating
of the pump.
It is important to note that there is often a definite connection between these two kinds of difficulties.
For instance, increased wear at the running clearances must be classified as a mechanical trouble, but
will result in a reduction of the net pump capacity-a hydraulic symptom-without necessarily causing
a mechanical breakdown or even excessive vibration.
The secret of successful and rapid diagnosis lies in the ability to select as comprehensive a list of
causes as possible and eliminate as many of these as can be eliminated by simple means. Then, the
remaining list should be examined in the descending order of probability. To facilitate this process, it
is of great advantage to classify symptoms and causes separately and to list a schedule of potentially
contributory causes for each symptom, as shown in Tables 32.1 and 32.2. The cure for each trouble is
almost always self-evident.
There is a reason why all possible causes should be examined: although a given cause will generally
produce a very specific effect, an effect can have anyone of several causes. For instance, if a pump is
rotating in the wrong direction, the head-capacity curve of the pump will be affected adversely to a
considerable degree. But if the symptom (the effect) is a major deterioration of the head-capacity curve,
the cause may be anyone or even several of the following:
1. Speed lower than design speed
2. Wrong direction of rotation
3. Reverse mounting of a double-suction impeller
4. Excessively worn internal clearances
5. Pump or suction piping not completely filled with liquid
6. Available NPSH is insufficient
7. Inlet of suction pipe insufficiently submerged, causing the formation of vortices and entrainment of air
into the pump

927

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
928 Diagnostics of Field Problems

8. Foreign matter in the impeller


9. Viscosity of the liquid exceeds design viscosity
10. Defective casing gasket, causing internal leakage.

As a matter of fact, this list is not complete, and Table 32.2 gives another 18 possible causes that
can lead to this one effect.

Table 32.1 Check Chart for Centrifugal Pump Problems

Possible cause of trouble (Each number is defined in


Symptoms Table 6.2)

1. Pump does not deliver liquid 1, 2, 3, 5, 1~ 1~ 13, 14, 16, 21, 22, 25, 30, 32, 38,
40
2. Insufficient capacity delivered 2, 3,4,5,6,7, 7a, 10, 11, 12, 13, 14, 15, 16, 17, 18,
21,22,23,24,25,31,32,40,41,44,63,64
3. Insufficient pressure developed 4,6, 7, 7~ 10, 11, 12, 13, 14, 15, 16, 18,21, 22, 23,
24,25,34,39,40,41,44,63,64
4. Pump loses prime after starting 2,4,6,7, 7a, 8,9,10,11
5. Pump requires excessive power 20,22,23,24,26,32,33,34,35,39,40,41,44,45,
61, 69, 70, 71, 71a
6. Pump vibrates or is noisy at all flows 2,16,37,43,44,45,46,47,48,49,50,51,52,53,
54, 55, 56, 57, 58, 59, 60, 61, 66, 67, 78, 79, 80, 81,
82, 83, 84, 85
7. Pump vibrates or is noisy at low flows 2,3,17,19,27,28,29,35,38,77
8. Pump vibrates or is noisy at high flows ~3, 1~ 11, 1~ 13, 14, 15, 1~ 1~ 18,33,34,41
9. Shaft oscillates axially 17, 18, 19,27,29,35,38
10. Impeller vanes are eroded on visible side 3, 12, 13, 14, 15, 17,41
11. Impeller vanes are eroded on invisible side 12,17,19,29
12. Impeller vanes are eroded at discharge near center 37
13. Impeller vanes are eroded at discharge near 27,29
shrouds or at shroud/vane fillets
14. Impeller shrouds bowed out or fractured 27,29
15. Pump overheats and seizes 1, 3, 12, 28, 29, 38, 42, 43, 45, 50, 51, 52, 53, 54, 55,
57,58,59,60,61,62,66,77,78,82
16. Internal parts are corroded prematurely 65
17. Internal clearances wear too rapidly 3,28,29,45,50,51,52,53,54,55,57,59,61,62,
65,66,77
18. Axially split casing is cut through wire-drawing 63,64,65
19. Internal stationary joints are cut through wire- 53, 63, 64, 65
drawing
20. Packed box leaks excessively or packing has short 8,9,45,54,55,57,66,68,69,70,71,72,73,74
life
21. Packed box: sleeve scored 8,9
22. Mechanical seal leaks excessively 45,54,55,57,58,62,74,75,76, 76a, 77, *
23. Mechanical seal: (a) faces are damaged, (b) sleeve 45,54,55,57,58,62,74,75,76, 76a, 77, *
damaged, (c) metal bellows fail
24. Bearings have short life 3, 29,41,42,45,50,51,54,55,58,66,77,78,79,
80, 81, 82, 83, 84, 85
25. Coupling fails 45,50,51,54,67

*Also see instruction book for specific mechanical seal in question.


Diagnostics of Field Problems 929

Table 32.2 Possible Causes of Trouble

A. Suction troubles
1. Pump not primed
2. Pump suction pipe not completely filled with liquid
3. Insufficient available NPSH
4. Excessive amount of air or gas in liquid
5. Air pocket in suction line
6. Air leaks into suction line
7. Air leaks into pump through stuffing boxes or through mechanical seal
7a. Source of sealing liquid has air in it
8. Water seal pipe plugged
9. Seal cage improperly mounted in stuffing box
10. Inlet of suction pipe insufficiently submerged
11. Vortex formation at suction
12. Pump operated with closed or partially closed suction valve
13. Clogged suction strainer
14. Obstruction in suction line
15. Excessive friction losses in suction line
16. Clogged impeller
17. Suction elbow in plane parallel to the shaft (for double-suction pumps)
18. Two elbows in suction piping at 90 deg to each other, creating swirl and prerotation
19. Selection of pump with too high a suction specific speed
B. Other hydraulic problems
20. Speed of pump too high
21. Speed of pump too low
22. Wrong direction of rotation
23. Reverse mounting of double-suction impeller
24. Uncalibrated instruments
25. Impeller diameter smaller than specified
26. Impeller diameter larger than specified
27. Impeller selection with abnormally high head coefficient
28. Running the pump against a closed discharge valve without opening a bypass
29. Operating pump below recommended minimum flow
30. Static head higher than shut-off head
31. Friction losses in discharge higher than calculated
32. Total head of system higher than design of pump
33. Total head of system lower than design of pump
34. Running pump at too high a flow (for low-specific-speed pumps)
35. Running pump at too Iowa flow (for high-specific-speed pumps)
36. Leaky or stuck check valve
37. Too close a gap between impeller vanes and volute tongue or diffuser vanes
38. Parallel operation of pumps unsuitable for this purpose
39. Specific gravity of liquid differs from design conditions
40. Viscosity of liquid differs from design conditions
41. Excessive wear at internal running clearances
42. Obstruction in balancing device leak-off line
43. Transients at suction source (imbalance between pressure at surface of liquid and vapor pressure at suction
flange)
Continued
930 Diagnostics of Field Problems

Table 32.2 Possible Causes of Trouble (Continued)

B. Mechanical Troubles---(Jeneral
44. Foreign matter in impeller or impeller damaged
45. Coupling misalignment
46. Foundations insufficiently rigid
47. Loose foundation bolts
48. Loose pump or motor bolts
49. Inadequate grouting of baseplate
50. Excessive forces and moments from piping on pump nozzles
51. Improperly mounted expansion joints
52. Starting the pump without proper warm-up
53. Mounting surfaces of internal fits (at wearing rings, impellers, shaft sleeves, shaft nuts, bearing housings,
etc.) not 100 percent perpendicular to shaft axis
54. Bent shaft
55. Rotor out of balance
56. Parts loose on the shaft
57. Shaft running off-center because of worn bearings
58. Pump running at or near critical speed
59. Use of too long a shaft span or too small a shaft diameter
60. Resonance between operating speed and natural frequency of foundation, of baseplate or of piping
61. Rotating part rubbing on stationary part
62. Incursion of hard solid particles into running clearances
63. Use of improper casing gasket material
64. Leakage at internal seals from mismatched or damaged surfaces
65. Pump materials not suitable for liquid handled
66. Hydraulic unbalance
67. Lack of lubrication of certain couplings
C. Mechanical troubles-Sealing area
68. Shaft or shaft sleeves worn or scored at packing
69. Incorrect type of packing for operating conditions
70. Packing improperly installed
71. Gland too tight, resulting in no flow of liquid to lubricate packing
71a. Mechanical seal over-compressed
72. Excessive clearance at bottom of stuffing box, causing packing to be forced into pump interior
73. Dirt or grit in sealing liquid
74. Failure to provide adequate cooling liquid to water-cooled stuffing boxes or mechanical seal heat exchangers
75. Incorrect type of mechanical seal for prevailing conditions
76. Mechanical seal improperly installed
76a. Seal housing face not square to shaft axis
C. Mechanical troubles-At bearings
77. Excessive radial thrust from single volute, assymetric double volute, eccentric diffuser, or diffuser stall
78. Excessive axial thrust caused by excessive wear at internal clearances or by failure or excessive wear of
balancing device (if such is used)
79. Wrong grade of grease or oil
80. Excessive grease or oil in antifriction bearing housings
81. Lack of lubrication
82. Improper installation of antifriction bearings (damage during installation, incorrect assembly of stacked
bearings, use of unmatched bearings as a pair. etc.)
83. Dirt getting into bearings
84. Moisture contamination of lubricant
85. Excessive cooling of water-cooled bearings
Diagnostics of Field Problems 931

In many plants, as little as 10 percent of the pump population is responsible for up to 80 to 90 percent
of the failures. One has to identify these "bad actors" and concentrate diagnostic efforts and preventive
maintenance on them. One should also try to find similar pumps and similar services in other plants and
compare the accumulated experience with these other installations. This is particularly important because
there are problems that are specific to certain services and applications. Such, for instance, as the
unfavorable effect of transient conditions on boiler feed pump operation whenever a sudden load drop
takes place in steam-electric power plants.
The remainder of this chapter is devoted to an analysis of some of the most important trouble spots-
an analysis that may be helpful in diagnosing and preventing these troubles.

PUMP NOISE

Pump noise often gives an experienced maintenance man a definite indication of the source of trouble.
If a pump produces a crackling noise, the source of trouble is probably at the pump suction. This type
of noise is usually associated with "cavitation"-the condition existing in flowing liquids when the
pressure at any point falls below the vapor pressure of the liquid at the prevailing temperature. Some
of the liquid flashes into vapor, and bubbles of the vapor are carried along with the liquid. If this happens
in the suction area of a centrifugal pump or within the entrance of the impeller, the bubbles are carried
into the impeller and undergo an increase in pressure and, therefore, condense. This process is accompanied
by a violent collapse of the bubbles, possible pitting and erosion of the impeller vanes, and a definite
crackling noise. Of course, the presence of vapor within the liquid pumped causes a reduction in the
pump capacity. Cavitation, therefore, is a direct result of insufficient pressure at the pump suction-
operation with insufficient net positive suction head (NPSH).
This diagnosis can be checked. For instance, throttling the pump discharge will reduce pump capacity
and possibly restore pump operation to a range in which sufficient NPSH is available at the pump
suction. If this step eliminates the crackling noise, the diagnosis is correct, and corrective measures will
consist of either increasing the NPSH for the normal range of operating capacities or replacing the
existing impeller with one that can operate with the existing NPSH when that cannot be increased.
A rumbling noise in the discharge waterways of the casing is usually caused by operation of the
pump at part-load capacities when the pump is not hydraulically suitable for such operation, or by
operation of the pump at capacities well in excess of those for which it was designed. A thorough
discussion of this problem is presented in Chapter 22.
Water hammer is caused by a sudden change in the velocity of flow of a column of liquid and, in
general, is serious only when long pipe lines are involved. Water hammer can be avoided by starting a
pump against a closed gate valve and then opening the valve slowly afterward. This procedure causes
a very gradual increase in the velocity of the liquid in the pipe line. During the stopping cycle, it is
necessary to shut the gate valve slowly and stop the pump after the valve has been fully closed. This
method fails when a unit or units are stopped suddenly by a control or power failure. Additional provision
for the control of water hammer must be provided in installations in which the resulting pressure due
to water hammer can reach a dangerous level.
In such installations, slow opening and closing check valves are generally employed to increase the
flow gradually when the pump is started and to decrease the flow gradually when the pump is stopped.
In other installations, a special protective valve is employed. This valve opens wide quickly with the
drop of pressure that is part of a water hammer cycle, and then closes slowly to throttle the resulting
back flow. Sometimes air-charged surge tanks or air chambers have been used satisfactorily.
Because a motor-driven pump will stop delivering water almost instantaneously when the power is
cut off, water hammer is a more serious problem in this type of installation. In a few very important
932 Diagnostics of Field Problems

RotMIng

-fO,
UnlNl18nce

-3X -2X
R_
-1X
I I
+1X
F_rd
+2X +3X

011 Whirl

'0 I,
3

4 2
I
-3X
I
-2X -1X
I +1X +2X +3X
R_ FOI'WIIrd

-fO
011 Whip

I
-IX
I
-2X -1X
II ,
+1X +2X +3X

~ FOI'WIIrd

Rub

~8 I
-3X -2X

~
,I II
-1X
I
+1X +2X

FOIWWII
+3X

MIsalignment

-8J I
-3X
I I I
-2X -1X
II
+1X +2X +3X

R_ F-.d

Fig. 32.1 Spectra and orbits of common vibration causes.


(Courtesy Bentley Nevada Corp.)
high-head installations, the stopping time has been lengthened by adding a flywheel to the unit. Although
the actual time it takes for the unit to stop pumping is lengthened very little, this slower stopping cycle
helps greatly in minimizing the water hammer.

PUMP VIBRATION

A sudden change in vibration or vibration rising above an alarm level are both indicators of distress
within the pump. To learn the likely cause of the distress, the first step is to determine the dominant
frequencies within the overall vibration. Modem vibration analyzers do this readily and produce a printed
trace of the spectrum. Still more sophisticated analyzers, in conjunction with "x" and "y" probes and a
once-per-revolution marker signal, produce a complete spectrum (forward and reverse motion relative
to rotor rotation) and a plot of the motion or orbit of the rotor at each pair of probes (Fig. 32.1).
Once the dominant frequencies have been determined, reference to a table of typical causes of vibration
at various frequencies, expressed as fractions or multiples of running speed, generally narrows the likely
cause to one or two possibilities. At that point, it is usually possible to decide whether it is safe to continue
running the pump until the next planned shutdown, or to shut it down immediately for rectification. Table
32.3 provides guidance on diagnosing pump vibration spectra. For additional detail on analyzing vibration
see Karassik et al [32.21], Booser [32.2], and Sanks [32.3] for pumps specifically, and Bloch and Geitner
[32.4] for machinery in general.
Some additional comments are warranted on several of the causes cited in Table 32.3.

Table 32.3 Vibration Diagnosis

Vibration Frequency
(relative to shaft rpm) Cause

Low, random Recirculation within the impeller


Cavitation
0.10-0.15 Rotating diffuser stall
0.40-0.49 Oil whirl in hydrodynamic bearings
0.5 Internal rubbing
0.7-0.9 Stall at the impeller inlet
Hydraulic excitation of rotor natural frequency
1.0 Rotor mechanical unbalance from: poor initial balance, foreign material in impeller,
damaged impeller, b.ent rotor
Hydraulic unbalance
Running speed at or close to rotor natural frequency
Internal rubbing
Coupling misalignment
2.0 Loose parts on the rotor
Coupling misalignment
Vane passing effect from twin-volute when gap is too small
Z (# of impeller vanes) Vane passing syndrome (insufficient clearance between impeller vanes and collector
vanes); also evident sometimes during operation with discharge recirculation
Odd Structural resonance
Pressure pulsations
Transmission from adjacent equipment
High multiples Defective anti-friction bearings

933
934 Diagnostics of Field Problems

Diffuser stall. With the conservative designs used in most centrifugal pumps, diffuser stall is not
often encountered (see Chap. 2). When it is, the result is generally rapid wear but relatively even wear
of the rotating and stationary surfaces that make up the running clearances.
Oil whirl. Because the forces developed in hydrodynamic bearings are usually relatively small com-
pared to those in pump running clearances, oil whirl in centrifugal pumps is very rare indeed. When it
does occur, however, it usually indicates bearing failure is imminent.
Internal rubbing. In compressible flow turbomachines, internal rubbing shows at 0.5 and 1.0 times
running speed, both forward and reverse, with a corresponding "figure 8" orbit (Fig. 32.1). Depending
on what was causing the rubbing, the same characteristics could be evident in a centrifugal pump.
Hydraulic excitation. Subsynchronous vibration at 0.7-0.9 of running speed is one of the current
controversies in centrifugal pump design. That the rotor is excited at its natural frequency is no longer
in serious doubt. The question is what provides the excitation? One theory says that it is provided by
destabilizing forces developed by the angular rotation of liquid in internal running clearances. The other
attributes the excitation to rotating stall at the inlet of the impeller, a phenomenon that has been reported
quite widely in the literature; see Marscher [32.5]. Both these sources of excitation occur at constant
fractions of running speed, and so are not classical instabilities which are self-exciting at a given
frequency. Field experience supports this; in variable speed pumps subsynchronous vibration has been
seen to exist over only part of the pump's operating flow range (Fig. 32.2), meaning that the separation
between the exciting frequency and the rotor natural frequency increased to the point where there was
no longer a response, or the cause of excitation ceased because the pump flow was now closer to design.
Unbalance. Most the causes of rotor unbalance are fairly straight forward mechanical issues, although
great care in design and manufacture is necessary to achieve the very small mass eccentricity inherent
in ISO balance grade G 1.0. Hydraulic unbalance is produced when an irregularity in the impeller causes
a pocket of liquid or vapor to rotate with the impeller, thereby hydraulically changing its mass distribution.
An obstruction in one or more of the waterways, such as a small fin across the waterway, does this very
effectively and can be difficult to find in low specific speed impellers. Resonance is rare too, since
centrifugal pumps with well balanced rotors and clearances within 2 times their new value have low

,1,,-- WORN CLEARANCES


VIBRATION
(DISPL)
,,
"1:
I I ,
' y NEW CLEARANCES
,,~
~

C I ) ROTOR VIBRATION

PUMP SPEED

SUBSYNCHRONOUS
FREQUENCY .--.~ EXCITATION
(CPM) 1 sC-- - ROTOR NATURAL
FREQUENCY (WET)

FLOW

Fig. 32.2 Subsynchronous vibration from hydraulic excitation of rotor natural frequency.
Diagnostics of Field Problems 935

synchronous excitation and high damping, and therefore are generally not very sensitive to running at
or close to their running speed. The exception is slender shaft multistage pumps with worn running
clearances (see Chap. 7).
Misalignment. Generally characterized by vibration at 2 times running speed, but actually produces
vibration at both 1 and 2 times, and a "figure 8" or elliptical orbit (Fig. 32.1).
Vane passing. As already discussed in Chap. 22, the vane passing syndrome proved very destructive
as pump heads rose from 250 m (800 ft) per stage to 550 m (1,800 ft) and then to 760 m (2,500 ft).
After many instances of impeller and diffuser fracture, the problem was corrected by changes in the
vane number combinations (impeller to collector), larger clearances between the rotating and stationary
vanes, and in some cases modifying the geometry of the stationary vanes. Today high vibration at vane
passing frequency is relatively rare except when the frequency is close to a bearing housing natural
frequency. In these cases, measurements of bearing housing vibration will show velocities at vane pass
frequency well beyond normal limits (see Vibration limits Table 32.4), while the rotor vibration (pump
with hydrodynamic journal bearings) at the same frequency is negligible. If it is judged that the bearing
housing vibration really is a problem, the solution in these cases is to change the frequency of excitation
or the natural frequency of the bearings housing. It is worth noting that the effect of vane number
combinations on the intensity of pressure pulsations at vane passing frequency is to some degree counter
intuitive; the worst combinations are those with a difference of 1, 7 impeller vanes and 8 diffuser vanes
being a well documented example, whereas even number combinations, for example 6 impeller vanes
and 2 diffuser vanes, are quite tolerable. The critical aspect seems to be subsequent reinforcement of a
pulse rather than simultaneous pulses; see Domm & Dernedde [32.6].

Vibration Limits
Vibration limits are still a somewhat controversial subject in the application, operation, and mainte-
nance of centrifugal pumps, but the evidence is now pointing to lower allowable limits producing longer
MTBR. In the absence of better information from practical experience, the values given in Table 32.4
are a conservative guide for vibration measured on the bearing housings of horizontal and vertical dry
pit pumps. When the pump has hydrodynamic journal bearings, shaft motion is generally greater than
that evident from bearing housing vibration, and so for a complete picture it is necessary to also measure
shaft vibration. Figure 32.3 provides guidance on shaft vibration for horizontal and vertical dry pit
pumps. In applying these limits, note that they are for installed pumps operating relatively close to BEP.
Shop tests of new or overhauled pumps are likely to show higher bearing housing vibration because the
test set-up usually does not have the same stiffness as the installation. Off-design operation usually
produces higher vibration (at what fraction of BEP depends on the size and design of the pump), and
the upper and lower limits of a pump's operating range are often defined as those capacities at which
the vibration is 1.3 times that at BEP.
The values in Table 32.4 are correlations from vibration whose principal component was at running
speed, and the running speeds were 2 pole motor speeds. There are data showing that the average bearing

Table 32.4 Vibration Limits; Measurement on the Bearing Housing


Units are mmls (in/s) RMS

Schedule for Shutdown


Good Acceptable Poor Repair Immediately

< 2.0 2.0-4.0 4.0-5.5 5.5-9.0 > 9.0


(0.08) (0.08-0.16) (0.16-0.22) (0.22-0.35) (0.35)
936 Diagnostics of Field Problems

150 6.0 .,..._ _ _ _ _ _ _ _....... ~

5.0
100 4.0
80 3.0
~ 60
I 2 . 0 + - - - -.......~
~ 40
I-C/) C/)
zZ GOOD
wO ~1.0
::E a: 20 0.8
wU 0.6
U-
:5::E 0.5
a.. 0.4
C/)
Ci 0.3
10
0.2

0.1 ;---------~--~--~--~~~~~----_r--~~
2 3 4 5 6 7 8 910 15 20 25
ROTATIVE SPEED-1000 RPM

Fig. 32.3 Vibration limits; measured on shaft.

housing vibration of pumps over the speed range 1,500 to 4,500 rpm, expressed in RMS velocity,
increases with pump speed to about the 0.33 power of the speed ratio. Similarly, as the frequency of
vibration increases, there are data showing that higher velocities can be tolerated. An example of this
is the many pumps that have ron for years without consequence to seal or bearing life with bearing
housing vibration of 7-11 mrn/s RMS (0.28-.43 in/s RMS) at vane passing frequency.
The shaft vibration limits in Fig. 32.3 are based on API-61O, 8th Edition [3.1], for new and overhauled
equipment and data provided by Bloch [32.4] for the classification of higher levels. In this classification,
"poor" is roughly the equivalent of "schedule for overhaul" in Table 32.4.
Vertical wet pit pumps have lower mounting stiffness than horizontal and vertical dry pit pumps, and
their shafting has higher ronout (long shaft sections) and generally less guidance. One consequence of
these inherent construction features is higher vibration than horizontal and vertical dry pit pumps.
Measurements on the pump thrust bearing housing or motor mounting flange if the pump thrust is taken
by the motor, are typically 65 percent higher. Measurements on the shaft can be 12 percent higher and
the upper limit of each class on Fig. 32.3 is 75, 125, and 200 JlIll (3, 5, and 8 mils) for "good",
"acceptable," and "poor" respectively.

AIR LEAKAGE INTO A PUMP

If a pump operates with a suction lift, air will sometimes leak into the pump through the stuffing boxes
unless lantern rings are installed in the boxes and liquid under positive pressure is piped to the lantern
rings. Air may also leak under the shaft sleeves. Sometimes the suction piping itself is not quite air-
tight and air enters it to collect at the top of the casing; at other times, the water handled by the pump
is saturated with air and this air may be liberated in the pump.
If a limited amount of air is allowed to leak into the suction side of a centrifugal pump, the air will
be carried through the pump. This ability of a centrifugal pump to carry a small amount of air through
its passages is utilized in some installations involving a pneumatic tank that adds air to the tank. If air
leaks into the suction passages and passes through the pump, there is a reduction in the amount of water
Diagnostics of Field Problems 937

handled and the power required. If the amount of air allowed to leak into the suction side is increased,
a point at which the pump will be unable to carry the air through with the water is reached, and the
pump will lose its prime.
Sometimes a pump that operates satisfactorily under full discharge will lose its suction when throttled
to a lower rate of flow. This condition may be caused by the presence of air leakage into the pump. At
full discharge capacity, the velocities in the pump casing are sufficiently high to wash the air into the
discharge pipe, thus constantly purging the air from the pump casing. When the capacity is reduced, the
lower velocities are insufficient to wash out the air that accumulates in the pump. This air accumulates
within the central portion of the impeller and is prevented from being vented out by the ring of water
that is thrown off by the impeller. The ability of the impeller to generate total head is reduced by the
presence of the air in its central portion. This air eventually prevents all further pumping action and the
pump loses its prime.
A pump may deliver its normal rated capacity when it is started but gradually slacken off until it
handles a fraction of its normal flow. This behavior indicates an air leak. Air is accumulating within the
pump and reducing the effective capacity. It is possible to check the accuracy of this diagnosis after
stopping the pump and letting the air rise to the top of the casing and of the suction volutes. If the pump
operates under a suction head, opening the casing and suction vents to the atmosphere will clear the air
out and the pump capacity should be restored after the pump is restarted. If the pump operates under a
suction lift, opening the vents will obviously not reprime the pump, and it should be reprimed by whatever
means are normally used for this purpose.
Some noisy pumps operating under cavitating conditions can be quieted by admitting a very small
quantity of air into the suction. This effect is produced by the fact that the small amount of air bubbles
cushions the "implosions," which follow the recondensation of the vapor bubbles produced during
cavitation. However, this is strictly an after-the-fact correction and is not recommended as the best
solution of the problem. Instead, the pumping system should be redesigned to permit the existing pump
to operate without cavitation or a more suitable impeller should be obtained to accept the existing suction
conditions. Admission of air should be used only as a last resort if no other solution is practical. The
amount of air necessary to quiet the pump varies with the capacity, and a variation in capacity due to
vibration in operating head may require a different amount of air. Furthermore, the presence of air will
generally decrease the efficiency of the pump and may accelerate corrosion. In addition, the presence
of air in the pumped water is frequently undesirable. The fact that a small amount of air leaking into
the suction line will quiet certain noisy pumps should be used as a test to determine if the cause of the
noise is on the suction side, but the admission of air is not a remedy.

SYSTEM TROUBLES

Some system troubles are relatively easy to correct. For instance, if the motor leads are improperly
connected, the pump will rotate in the wrong direction. As soon as this condition is diagnosed, the motor
leads can be reversed and the installation made ready for proper operation. When a centrifugal pump is
operated in a reverse direction, its performance characteristic is completely abnormal and highly inefficient
(Fig. 32.4). The motor of a newly installed pump should be tested for correct rotation when the motor
is disconnected from the pump. This is particularly important for vertical turbine pumps and for other
pumps that have shafting sections joined by screwed couplings. With this type of shafting, incorrect
rotation, even for testing, would result in one or more couplings becoming unscrewed and would require
drawing the pump out to reconnect the joint. Damage to some of the pump parts might even result.
Another cause of major loss of capacity and increase in power consumption is the incorrect reassembly
of a pump. As mentioned in previous chapters, centrifugal pump impellers have backward curved vanes,
938 Diagnostics of Field Problems

-NORMAL ROTATION
---PROBABLE EFFECT
OF REVERSE
ROTATION

250

90 HEAO-CAPAC/r
-..:X- E.FFICIENCY i
....
--- '"
80 200
r--..... ,I
~

70
V
I- I-
W
,/ ~~/ ............ ~

"
Z
w
(,) 60
w
~ 150
/ v;;;;'~'
a:
w z /
9~'
~It. /
Cl. 50 a ~"7
~
~ 40
c(
w
100
1.°'C'i!
,,-4.0 Iv~
,~
200
L~C~)-
:J:
W ~
_ ~~~ / 4-
.J
U c(
150
Li: 30 I-
BHP
I
-
~ 0 ....
~ .".--
W I- Cl.
20 50 ,....- 100 ffi
.-{ .r.,E.~~ ~
-~ ~ 50
10

0 ~ -~.~
5
V
10
\
15 20 25 30
o
CAPACITY, IN 100 GPM

Fig.32.4 Effect of incorrect rotation on perfonnance of a 6-in. double-suction, single-stage pump at 1,750 rpm.

that is, they revolve away from the curvature of their vanes (Fig. 32.5A). If a pump is dismantled
completely and then reassembled, an error in mounting a double-suction impeller is especially easy to
overlook. Such an impeller is usually symmetrical about its central axis and will therefore fit on the
shaft even if it is turned in the wrong direction (Fig. 32.5B). The hydraulic performance is the same as
would be obtained in a normal casing with correct rotation, but with forward curved vanes. If an impeller
is reversed on the shaft, the effect is not the same as running the pump in reverse rotation (Fig. 32.5C).
The probable performance of a pump rotating in the proper direction both with a properly mounted
impeller and with a reversed impeller is shown in Fig. 32.6. The loss in head and capacity and the
increase in power consumption when the impeller is mounted in reverse is much less severe than with
a correctly mounted impeller that is running in reverse.
Occasionally, the nonretum or check valve used to prevent reverse flow through a pump when the
unit is shut down, fails to function, and remains open. If this fault occurs, there is a back flow of water
through the pump causing the pump to become a water turbine and, if the driver does not act as a brake,
the unit rotates in the reverse direction at the runaway speed of the pump. This runaway speed depends
on two factors: the type of the pump and the net head available. With an effective net head equal to the
design pumping head, the runaway speed varies from a maximum of approximately 175 percent of rated
speed with low-head, high-specific-speed types of propeller pumps, to even less than rated speed for
high-head, low-specific-speed types. Most pumping systems involve both a static and frictional compo-
nent. Thus, when reverse flow occurs, the net head, which is now the static head minus friction head,
is usually much less than the normal pumping head (which is static head plus friction head). This lower
Diagnostics of Field Problems 939

....... OIREC T ION OF ROTA TlO ~

, L-- -D L.--fl
~ ~- ~.­ ~.~ ~ ~~
.'--i--- ---U '--------U .-----.---.-U
CORRECT ROTATION CORRECT ROTATION REVERSED ROTATION
IMPELllR MOUNTED IMPELLER REVERSED IMPEUER MOUNTED
CORRECTLY CORRECTLY

A B c

Fig. 32.5 Pump assembly and rotation.

-NORMAL ROTATION
---PROBABLE EFFECT
OF REVERSED
MOUNTING OF
IMPELLER

250

90
~p"'CI1"'{ _
~tl~c. _r-- J
80 200 ......
...z ...w ~. . ,.. ~c.Yr::\f"'... t--.... l/ ~
70 \~~

~ ~ fJ' I'..
~

W ILl

~ '"""
I,)
I<. ~~ ,,~
a: 60 ~ 150
ILl
Q.
6 j
V J.~f~;;'e,ENCY
...;;., ~ ~fi
Q
(j

~
>-- 50
U
Z
~
ILl
X I ~/,.
"t.;
~X,
ILl 40 100 200
I
~
/'
~'k1
u ~

1/ "
i:;: l-
...
0
.- 150

- -:r
I<. 30
V --
ILl ,BHt"
J- ~\! ~
Q.
20 50 100 ~
.L ~ ~- ~

10

0 o
o
~'
.- 1---'--
5
V

10 15 20 25 30
50

o
CAPACITY, IN 100 GPM

Fig. 32.6 Effect of reversed mounting of impeller on performance of a 6-in. double-suction, single-stage pump
at 1,750 rpm.
940 Diagnostics of Field Problems

net head available at the time of reverse flow causes the runaway speed to have a lower value than
would result with a net head under reverse flow equal to the design head.
An operating speed that is too low or too high is generally the cause of faulty operation only in
turbine- or engine-driven units. Correction of this trouble may sometimes be more involved than the
correction of wrong direction of rotation for a motor, but it is not too difficult to effect. It is necessary
to remember, however, that if the pump driver is subject to speed variation, the first step in diagnosing
a variation of operating conditions from expected value is to determine the operating speed and see if
it corresponds to the design.
The problem of parallel operation of centrifugal pumps has been discussed in Chapters 20, 21, and
22. If it is attempted to operate centrifugal pumps that, by virtue of their characteristics, are unsuitable
for parallel operation, the choice of corrective measures is limited. In some conditions, it may be found
that as long as parallel operation is restricted to the higher capacity range of the various pumps, no
difficulties arise. It becomes necessary, then, to maintain only enough pumps on the line to insure their
individual operation in this higher range of capacities. If high-head pumps are involved and, especially,
if these pumps handle liquid at or near the boiling point, it is absolutely essential to keep the protective
recirculation bypass lines open at all times to eliminate the hazard of a check valve slamming on one
of the pumps and causing the unit to bum when an unforeseen reduction in total capacity occurs. Manual
control of all the units provides a second solution. This procedure is both tedious and uneconomical.
The third and last solution consists of revamping the pumps in question and replacing the hydraulic element
with parts that will correct the shape of the characteristic curves and thus make parallel operation feasible.
A break in the discharge line of a centrifugal pump is not a very frequent source of trouble. If the
discharge line break occurs near the pump, the pump will operate against a very low head and, as a
result, will deliver nearly its maximum capacity. In some installations, this capacity will produce a
dangerous overload on the driver. Should the break occur at some distance from the pump, a larger
frictional component may be involved-especially if the static portion of the total head is low-and the
capacity of the pump will be limited.

FOREIGN MATTER IN THE PUMP

The presence of foreign matter in the impeller or casing of centrifugal pumps can lead to serious trouble
and should, unless the pumps are specially designed to handle such material, be carefully prevented.
Strainers or screens should always be located in the suction line if foreign matter may be expected in
the liquid handled by the pump. Even under normal operation with the presence of foreign matter
unlikely-as in boiler feed pumps-welding beads, bolts, nuts, and other objects may find their way
into the pump suction in the early stages of operation because the pump is piped without precautionary
flushing of the suction piping. Temporary screens or strainers should, therefore, be used in all pump
installations. After the pump has been in successful operation for a while, the screens may be removed.
If the suction eye of one side of a double-suction impeller becomes completely clogged, the pump
capacity is reduced by approximately one-half. A complete blocking of one side does not impose an
excessive load on the thrust bearing because there is little difference in pressure on the two eye areas.
A partial blockage of one or both sides, however, often results in throwing the rotor out of balance and
may impose an additional load on the bearings.

MISALIGNMENT AND DAMAGE TO INTERNAL PARTS

Misalignment usually means misalignment of the coupling between the driver and the pump, but can
also refer to misalignment of the rotor within the pump itself. Coupling misalignment generally has one
Diagnostics of Field Problems 941

of two causes; either 1) the pump and its driver were not properly aligned in the first place or 2) piping
loads or drive reaction are changing the alignment once the unit is running. There is really no excuse
for the first cause; see Chap. 28. In the strictest sense, there is no excuse for the second either, but now
there are several factors that can affect the result; namely:

1. Piping design, fabrication, and installation must keep operating piping loads within the pump's capacity
(see Chap. 28). If the operating loads are much greater than allowable, the resultant displacement of the
pump shaft at the coupling can easily be great enough to produce serious coupling misalignment.
2. The intended means of supporting the base must be clearly set down in the specification, and then the
foundation and base must be designed accordingly (see Chaps. 13 and 28).

High coupling misalignment accelerates the wear rate of the coupling and can lead to premature
coupling failure. It also produces reactions on the connected machines. These, in tum, can cause premature
failure of bearings, shaft seals, and on occasion, the shaft. The difficulty with the shaft seal comes from
either distortion of the shaft or of the bracket connecting the bearing housing to the pump casing, each
of which effectively changes the alignment of the rotor within the pump. In a similar manner, excessive
piping loads can distort the pump casing to the extent that the rotor is no longer correctly aligned within
the pump, leading to rapid wear at the running clearances, and premature shaft seal and bearing failure.
Rarely, a pump may develop mechanical troubles, when it is first being started, because of rubbing
of rotating parts, a bent shaft, damaged internal parts, defective casing gaskets, a rotor out of balance,
and the like. Most of these defects should be discovered prior to shipment during the test at the
manufacturer's plant; they can, however, develop after a pump is dismantled and reconditioned in the
field. No criticism of maintenance personnel is meant here; but it must be recognized that personnel
who engage in such maintenance work at infrequent intervals cannot be expected to be as familiar with
the necessary operations and precautions as personnel regularly employed in such work.
First on the list of the difficulties that may occur in field reassembly of rotors is contact between
rotating and stationary parts of the rotor. Precautions must be taken to eliminate the possibility of the
occurrence of such contact. If a pump is started before it has been thoroughly checked for the concentricity
of the rotating element with the stationary parts the possibility of a damaged or even seized rotor
is increased, particularly if the pump is stainless steel fitted, because such materials gall easily on
making contact.
Other practices that may, under some conditions, cause the binding of a rotor in its internal clearances
are starting a pump that handles high temperature liquids before warming it up thoroughly and, also,
running it at a much reduced speed (say one-half or one-fourth of normal) a number of times and then,
by cutting the power, letting it run down to a stop several times. The pump may seize, because if any
of its internal parts are rubbing even very slightly, heat is generated, and if the pump is not up to speed,
there is not enough liquid flowing past the internal clearances to keep the contacting elements cooled.
If the power is cut off when the pump seizes, the rotor will lock solidly in the wearing rings while
rolling to a stop. It is therefore recommended that when a pump is started up cold or only partly warmed
up, the unit should be brought up to full operating speed and kept going at all times. If, through faulty
internal alignment, a slight momentary contact does take place, the chances are that the pump will not seize.
If the axial contacting faces of impeller hubs, shaft sleeves, and shaft nuts are not 100 percent
perpendicular to the bores of these parts-such alignment is difficult to attain in the field without very
accurate machining facilities-excessive tightening of the shaft nuts will lead to bending of the shaft
and to potential contact of the rotating parts. For this reason, a reassembled rotor should always be
checked for concentricity and true running before being inserted into a pump casing.
In a pump that has suffered rapid wear at its running clearances, careful inspection and documentation
of the wear will often provide a clue as to the cause. Careful inspection means determining not only
942 Diagnostics of Field Problems

Table 32.5 Causes of Wear at Running Clearances

Wear Pattern at Running Clearances


Stator Rotor Likely Cause of Wear

Even all around Wear on one side; same angular Rotor has run bowed; either
position bent or from unbalance
Even all around Wear on one side; different Rotor has run in 2nd or higher
angular positions mode shape; resonance
Wear on one side; same Even all around Rotor has run deflected by
angular position stationary force; radial thrust
Wear on one side; different Even all around Rotor has run with initial
angular positions rubbing contact;
misalignment within stator
Even all around Even all around Rotor has run with whirl at
subsynchronous speed;
diffuser stall or hydraulic
excitation

the magnitude of the worn clearances, but also the presence and angular location of any eccentric wear
of both the rotating and stationary surfaces. These data may then be interpreted using the general rules
in Table 32.5. The longitudinal pattern of wear caused by deflection will be greatest at the clearance
farthest from the line bearing in overhung pumps, and at midspan in between bearings pumps. The same
is true for rotors whirling in the first bending mode. Wear caused by rotor misalignment depends on
whether the misalignment was roughly parallel (even wear), at one end (single cone), or at both ends
(double cone).

SHAFT SEAL

Shaft sealing problems are the most frequent cause of pump failure. This is not to say that shaft sealing
techniques have not made significant advances in the past 30 odd years, they have, but that improvement
has only kept pace with more severe sealing conditions and improvements in other aspects of centrifugal
pump design. Shaft sealing is difficult because the functioning of whatever device is used depends on
the design of the device itself, the design of the pump with respect to the seal, and the actual operation
of the pump, which is not always what was intended when the pump was designed. In the case of the
packed box seal, these difficulties are compounded by the need for periodic inspection and adjustment.

Packed Box Seal


The conditions that contribute to packed box seal difficulties are:

1. Shaft running off center because of excessive wear in the bearings, a bent shaft, or misalignment. This
condition can be readily checked-first by disconnecting the coupling and rechecking the alignment, and
secondly by mounting an indicator on the pump casing in the vicinity of the stuffing box to detennine
whether the shaft revolves concentrically.
2. Shaft or shaft sleeves worn and scored at the packing. A routine examination of these parts will reveal
whether they must be renewed or repaired.
Diagnostics of Field Problems 943

3. Shaft vibration due to unbalance in the rotor, cavitation, operation at extremely light flows or beyond the
recommended maximum capacity, or instability in parallel operation.
4. Plugging of the water seal connection or improper location of the seal so that no sealing liquid can enter
the stuffing box. The presence of dirt or grit in the sealing liquid will similarly cause stuffing box difficulties
by scoring the shaft or shaft sleeves.
5. Excessive tightening of the gland with resulting absence of the leakage that lubricates the packing. Hourly
and daily observation of the pump operation, together with the knowledge that some leakage is necessary
for proper stuffing box operation, will prevent troubles from this cause.
6. Failure to provide suitable cooling through water-cooled stuffing boxes if the pump is so fitted.
7. Excessive clearance between the bottom of the stuffing box and the shaft or shaft sleeve, which causes the
packing to be gradually pushed into the pump interior. This condition can arise when the shaft or shaft
sleeves are repaired by grinding them down excessively instead of replacing them or building them up to
original dimensions.
8. Packing not properly selected for pressure, temperature, or rubbing speed conditions.
9. Packing not properly inserted into the stuffing box because the individual rings are too short and the gap
between ring ends is excessive, or because the ring joints are not staggered.

Table 32.6 provides a diagnosis of problems from the appearance of the worn stuffing box packing.

Mechanical (Axial Face) Seal

Most designs of mechanical seal rely on boundary lubrication of the sealing interface (see Chap. 9).
They are therefore sensitive to the following:

1. Pressure at the seal


2. Temperature at the seal
3. Margin over vapor pressure (unless a gas seal)
4. Nominal surface speed at the sealing interface
5. Seal face material combination
6. Heat dissipation (which is related to margin over vapor pressure)
7. Cleanliness of the sealed liquid
8. Angularity of the shaft through the seal
9. Motion (vibration) of the shaft at the seal.

Table 32.6 Diagnosis From Appearance of Stuffing Box Packing

Symptom Cause

Wear on one or two rings next to packing gland; other Improper packing installation
rings okay
Wear on OD of packing rings Packing rings rotating with shaft sleeve or leakage
between rings and ID of box (wrong packing size or
incorrectly cut rings)
Charring or glazing of inner circumference of rings Excessive heating, insufficient leakage to lubricate
packing, or unsuitable packing
ID of rings excessively increased or heavily worn on Rotation eccentric
part of inner circumference
944 Diagnostics of Field Problems

In cases of abnormally short seal life, the first step in diagnosing the cause is a careful examination
of the damaged seal faces, dynamic gaskets or bellows, and compression element parts. Examination of
the seal faces should be microscopic if necessary. Using the findings from the examination, one of the
guides issued by the various mechanical seal manufacturers should be consulted to determine what has
happened at the seal. The root cause is often not in the seal itself but in some aspect of the control of
its environment, the pump's condition or design, or the operation of the pump. Finding the root cause
is generally painstaking work. It is worth the effort, however, to avoid the cost in maintenance expense
and possibly lost production of frequent unscheduled pump shutdowns. As a closing comment, note that
many users seek to correct a mechanical seal problem by installing a more sophisticated seal, only to
find after many more failures that the problem was really something simple such as the location of a
seal flush connection to effectively remove heat from the seal faces.

Breakdown Seals
Premature failures of breakdown seals, either a sudden increase in leakage or seizure onto the rotor,
are usually caused by one of the following:

1. Flashing in the seal from loss of seal injection


2. Differential thermal expansion from temperature gradients within the seal
3. Operating misadventure producing radial rotor excursions large to cause rubbing within the seal.

Chapter 10 deals with the design of reliable condensate injection systems, and the details that have to
paid attention to with respect to differential thermal expansion.
Operating misadventure is generally flashing the pump (see Chapter 22).

BEARINGS

Bearing troubles probably occur in centrifugal pump installations as frequently as packing and shaft
sleeve difficulties. The following list of contributory causes is relatively lengthy and should be thoroughly
analyzed if bearing replacement becomes a consistent problem:

1. Excessive radial load caused by low flow operation or some mechanical failure inside the pump.
2. Excessive thrust load caused by high suction pressure (end suction pumps), inlet flow distortion (double
suction pumps), or badly worn running clearances, the last particularly in multistage pumps.
3. Excessive grease or oil in the housing of an antifriction bearing. This defect will cause overheating and
failure of the bearing.
4. Excessive heating of the bearing due to improper dissipation of heat or a failure of the cooling medium supply.
5. Lack of lubrication.
6. Dirt or foreign matter in the lubricant, often from shaft seal leakage.
7. Improper installation of antifriction bearings (damage during assembly, incorrect assembly of stacked
bearings, use of unmatched bearings as pairs, and so forth).
8. Rusting of antifriction bearings because of water in the bearing housing.
9. Excessive cooling of water-cooled bearings, resulting in condensation of atmospheric moisture in the
bearing housing.
10. Misalignment of the unit, bent shaft, or severe vibration, which impose excessive loads on the bearings.
11. Insufficient clearance to accommodate thermal expansion.
12. Rolling element "skidding" in the unloaded row of a duplex bearing under high axial load.
Diagnostics of Field Problems 945

It should be noted that most antifriction bearing manufacturers issue excellent treatises on their
equipment, in which the subject of bearing failures is thoroughly discussed. They provide photographs
of the various types of failures, thus enabling the user to analyze his own specific problem and to
diagnose his troubles by visual comparison. Because antifriction bearings are widely used in centrifugal
pumps as well as in other equipment, it is recommended that the maintenance personnel have such
literature available and become familiar with the various symptoms of antifriction bearing failure.

CORROSION AND EROSION

Among the most insidious and most costly of the troubles that may affect centrifugal pumps is the failure
of the materials used in the construction of the pump to withstand the corrosive or erosive properties
of the liquids handled. Unfortunately, this problem is extremely involved and cannot be treated extensively
in a book of this character. Some general information on the subject is presented in Chapter 17. For
more detailed information, the reader may consult the extensive literature available. It should be noted,
in passing, that the present trend in material selection is to be conservative-that is, to choose materials
more resistant to corrosion and erosion than may be necessary, in order to avoid expensive breakdowns
and interruption of service.

BIBLIOGRAPHY

[32.1] Karassik, LJ. et al., Pump Handbook, 2nd Edition, 1986, McGraw Hill Inc., New York, New York, pp. 2.294-
2.303.
[32.2] Booser, E.R., CRC Handbook of Lubrication and Tribology, Volume III: Monitoring, Materials, Synthetic
Lubricants and Applications, 1994, CRC Press, Inc., Boca Raton, Florida.
[32.3] Sanks, R.L., et aI., Pumping Station Design, 1996, Butterworth-Heinemann, Boston, Massachusetts.
[32.4] Bloch, H.P. and Geitner, F.K., Practical Machinery Management for Process Plants, Vol. 2, Machinery
Failure Analysis and Troubleshooting, 1983, Gulf Publishing Co., Houston, Texas.
[32.5] Marscher, W.D., Subsynchronous vibration in boiler feed pumps due to stable response to hydraulic forces
at part-load, 1988, Institute of Mechanical Engineers, Proc. Part-Load Pumping Symposium, London.
[32.6] Domm, U. and Demedde, R., Uber eine Auswahlregel fur die Lauf- und Leitschaufelzahl von Kreiselpumpen,
1972, KSB Technische Berichte, FrankenthaI, Germany.
VI
DEVELOPMENT
33
The Centrifugal Pump of Tomorrow

We felt that this book would not be complete if we failed to try to predict what the future holds for the
centrifugal pump. And to do so would first require to examine the past, and to present the story on "how
we got to where we are." Only then could we turn to predictions of future developments. In looking at
the past, we arbitrarily selected a span of 50 years, although we might add that this choice was dictated
by two separate and yet related circumstances.
The most important of these two reasons is that it was approximately 50 years ago that a great upsurge
took place in the technology of most of the processes that are served by centrifugal pumps. For instance,
it was the moment that steam power plant operating pressures started to move up to the level of 85 bar
(1,250 psig) and above. This imposed discharge pressure requirements from 100 to 125 bar (1,500 to
1,800 psig). It was also the moment when the centrifugal pump started displacing positive displacement
pumps in great numbers in such areas as the chemical plants and the refineries.
The second reason is much less dramatic and yet reasonably important. This 50-year period covers
almost the entire span of the career of one of the authors in the field of centrifugal pumps. Thus, all
that we recount as to the past took place during this career and by the accident of fortune, we can rely
on the personal experience of that author in our analysis of the evolutionary steps that were taken to
develop today's centrifugal pumps.

THE ORIGIN OF TECHNOLOGICAL TRENDS

But before we examine the past and attempt to predict the future, we should reflect briefly about what
are the circumstances which lead to the evolution and change of technical products or processes. We
can classify these circumstances into four very distinct categories:

1. Changes in the industry served by the particular product or process in question. An example of this category
would be the need to develop a boiler feed pump for higher pressures that could only arise after the
development of steam turbine generators and boilers operating at these higher pressures.

949

I. J. Karassik et al., Centrifugal Pumps


© Chapman & Hall 1998
950 The Centrifugal Pump of Tomorrow

2. Ideas that originate within the industry that provides the product or the process. These ideas may be the
result of a desire to obtain an improved competitive position or the natural desire of one or more engineers
to develop a "better mousetrap." This is probably the greatest source of improvements and examples are
too numerous to single out anyone of them.
3. Improvements and progress that take place in peripheral or different disciplines. The centrifugal pump came
into its own as a replacement of the steam pump only with the emergence of electric power and the
development of suitable and economically attractive electric motors. Thereafter, its growth in popularity
became intimately related to further improvements in electric motor design. And it was not until the large
two-pole, 3,550 rpm squirrel cage induction motor became available in the 1930s that high-pressure pumps
operating at those speeds could be built.
4. External political or economical circumstances which change the economic justification of a technological
product or process. A prime example of this category would be the oil crisis of the 1970s. The subsequent
meteoric climb of oil prices created the need of placing a greater emphasis on efficiency and energy conservation.

If we wish to expand on this examination of the effect of these external circumstances, we should
mention that they played an important role during the early development of the centrifugal pump and
probably through the 1940s, by creating some marked differences in the design philosophies between
Europe and the United States. And here we are not referring to the possible effect of the difference in
current frequencies, but rather to the fact that labor costs had always been much lower in Europe than
in the United States, whereas the cost and availability of raw materials were more favorable in the United
States. The European designer had, through necessity, learned to spare materials even at the cost of
increasing labor man-hours, both in production and in maintenance.
A typical result of these differences was the fact that the adoption of the double-casing barrel type
high pressure pump in Europe lagged the United States by about 20 years. In the interim, European
designs continued to use the segmented-casing construction which, of course, requires more labor both
in the original construction phase and in maintenance. Similarly, European pumps frequently omitted
the use of oil lubricated thrust bearings when they were provided with hydraulic balancing devices of
the disk or drum-and-disk type.
But as the economic differences between Europe and the United States started to disappear by the
1960s, so did any major differences in design philosophy.
It is obvious that the continuous evolution of the centrifugal pump, as is the case for that of all other
technological equipment, has benefited from the invigorating effect of competition that characterizes the
Western system of free enterprise.

THE LAST 50 YEARS

The centrifugal pump of 1938 was far from being a primitive machine. As a matter of fact, even then
it was a relatively mature product and some engineers had begun to doubt whether many extraordinary
changes could be expected in the future. Of course, they were wrong, because these 50 years have been
quite fruitful in producing significant breakthroughs in centrifugal pump technology We cannot cover
a complete listing of all these breakthroughs and we have limited ourselves to speaking of such matters as
1. The double-casing barrel type high pressure pump (1934-1938) (Fig. 1.11)
2. The development of the concept of suction specific speed (1937-1938) (see Chap. 19)
3. The condensate injection stuffing box (1952-1953) (Fig. 10.9)
4. The vertical pump with double-suction impeller and double-volute (1953) (Fig. 14.33)
5. The high-speed, high head per stage pump (1954) (Fig. 1.17) .
6. The understanding of the phenomenon of internal recirculation (1968-1972) (see Chap. 22)
7. The commercialization of the inducer (1970) (Fig. 4.17).
The Centrifugal Pump of Tomorrow 951

Note that many of these innovations did not spring full-blown from the engineering departments of the
manufacturers. Their commercialization had to wait on occasion for the right moment in the technological
requirements of the user. As an example, the inducer is not a recent invention: U.S. patent no. 1,586,978
was granted to Oscar H. Dorer (of Worthington) on June 1, 1926, for his disclosure of the inducer filed
in February 1921. But it remained somewhat of a curiosity until the 1950s when it was broadly applied
to the liquid oxygen and hydrogen pumps used for rocked propulsion systems. The inducer became
commercialized for chemical and refinery services in the 1960s, and by 1970 several pump manufacturers
were offering it as a standard option for their lines of process pumps.
We should also mention that once the Depression and war years were over, major pump companies
started to broaden their experimental and analytical research work. This led to a number of important
advances in the state of the art. For instance, the seminal work of A. J. Stepanoff (of Ingersoll-Rand)
and of V. Salemann (of Worthington) on the effect of liquid characteristics on cavitation permitted a
more precise prediction of NPSH requirements for a variety of conditions other than when handling
cold water.
Another development that should be mentioned is the experimental work carried out by Agostinelli,
Nobles, and Mockridge (of Worthington) on the radial thrust in volute pumps that led to a better
understanding of this phenomenon and thereby gave pump designers the ability to predict its magnitude
under a variety of alternative treatments.
In a similar manner, it was the work of Warren H. Fraser (of Worthington) in the 1960s that led to
his development by the early 1970s of the mathematical relationships that exist between the geometric
configuration of an impeller and the onset of internal recirculation at the suction and at the discharge.
With the understanding of the manner by which internal recirculation created serious field problems,
guidelines were developed that would permit users to avoid these problems during operation at low flows.
The temptation to list every single improvement in centrifugal pump construction or application is
overwhelming. Still, one cannot stop to examine each one of these. We list only a few more items that
saw their beginnings in the past 50 years:

1. The availability of a wide range of stainless steels


2. The development of plastics suitable for pump construction
3. The development of commercialization of mechanical seals
4. The availability of slurry pumps capable of handling very high concentrations of abrasive solids.

The Immediate Future: The Next Ten Years


Among the several trends that will emerge in the immediate future, the one that we want particularly
to single out here is the development and commercialization of variable frequency motor drives, because it
will have a most dramatic effect on the application of centrifugal pumps and on their future standardization.

POWER CONSERVATION

It is improbable that optimum efficiencies attainable today for any given size and type of centrifugal
pump will improve to any significant degree. Thus, any power savings to be achieved will have to come
from a reduction in power losses in the system, that is, from eliminating excessive margins imposed by
the system's designer. These excessive margins require throttling the pump discharge and result in
952 The Centrifugal Pump o/Tomorrow

unrecoverable power losses. Additionally, power savings can be achieved as mentioned above by a more
frequent use of variable speed drive. This subject has been treated in great detail in Chapter 21.

THE FUTURE OF MECHANICAL SEALS

The development and commercialization of mechanical seals has been nothing short of phenomenal,
particularly within the last 25 years. And the use of double seals has led to the prevention of even the
slightest of mists of toxic or flammable liquids to the surrounding atmosphere. But there is a new problem
looming on the horizon: the inert sealing liquids injected between the two seals are frequently hydrocarbons
themselves-inert and nonflammable though they may be. And the growing severity of the regulations
that are being imposed on industry by various national and local government agencies will soon prohibit
the emission of these inert vapors. What shall the pump industry do in the face of this prohibition?
There are three possible solutions:

1. When necessary to eliminate vapor emission into a space where it would create a supposed health hazard
for the operating personnel, the pumps could be installed in a sealed containment structure. Vapors would
be withdrawn from this containment by some mechanical means.
2. We could use some form of a "triple seal," with water forming the barrier against vapor emission at the
second or intermediate seal.
3. We would have to develop economical, reliable and efficient hermetic pumps.

We are afraid that the first solution is too complicated and impractical. The second solution would
likewise be rather complicated and would significantly increase the shaft span or overhang of centrifugal
pumps. Hence, we believe that the future belongs to the hermetic pump. But what will it look like? We
shall speak of this when we examine the long-range future of the centrifugal pump.

THE UPGRADED MEDIUM-DUTY PUMP

Several chemical process and petrochemical companies have noted with some dismay that the mean
time between failures of their standard single-stage ANSI type pumps is far from being satisfactory. It
is true that some of the causes of this short fall in life expectancy are directly traceable to improper
operating and maintenance practices. Still, the industry feels that pump manufacturers can contribute
materially to the reduction of operating and maintenance costs by providing an intermediate design
between the two standards presently used, that of the ANSI type pumps and the API-610 construction.
Recommendations have been made by a number of users that a new intermediate standard be created,
the so-called "upgraded medium-duty" centrifugal pump. Its range of Services would extend to 20 bar
(300 psi) and 175°C (350°F). Such pumps would incorporate the following features:

1. Oversized stuffing boxes, to accommodate more sophisticated and more reliable mechanical seals
2. Magnetic bearing housing seals, to reduce or eliminate lubricant contamination
3. Duplex angular contact thrust bearings, mounted back to back
4. Dramatically reduced shaft deflections
5. One-piece steel bearing frame and adapter
6. Oil-mist lubrication, preferably.
The Centrifugal Pump of Tomorrow 953

Whether such a new standard is warranted or not is debatable. Our opinion is that it may be preferable
to incorporate these various improvements in construction into the ANSI standards in the form of
individual options, with customers being given the choice as to how many of them they wish to purchase
as extras. Time will tell whether the "all-or-nothing" or this "options" approach will prevail. Frankly,
it does not matter that much; the customer will benefit from either one of them.

SOME THOUGHTS ON MATERIALS OF CONSTRUCTION

By the middle 1940s there took place a marked increase in the number of inert materials (such as carbon,
plastics, and cermets) used for pump construction. Simultaneously, a constantly growing panoply of
stainless steels came to be developed, partly in the hope of providing an "all purpose" corrosion and
erosion resistant metal, partly directed to various very specific applications. These new stainless steels
were intended to be used for the constantly growing number of applications for very corrosive chemicals
that created conditions that could not be met with any of the older types of stainless steels then available.
Somewhat later, the problem came to be solved occasionally by the use of coatings--either nonmetallic
such as vinyl plastics, resins, neoprenes, or baked-on formaldehydes, or by glass and other ceramics or
cermets. Most recently there have been various designs of highly corrosion resistant pumps made entirely
of fibber reinforced polymer.
But the search for better or more economical solutions to the materials problem continues. Thus,
even in the proximate future, we will see a growing number of alternative solutions. Large pumps will
depart from the traditional approach of using cast iron or cast steels and will shift toward the use of
cast/weld fabricated construction. In some cases, clad metals can be used very effectively. And we will
find more reliable and better means of coating the wetted areas of our pump casings and of other
components with wear and corrosion resistant materials: we can no longer afford to expend our natural
resources by using pump components made entirely of such materials.

Long-Range Developments
And what of the long-range future? We must realize that all the easy improvement have been incorporated
in today's centrifugal pumps and that a few harder ones will be a fact in another 10 years. History tells
us, however, that our task will not be finished by then and that further evolutionary progress will still
be taking place. Certainly we are not so presumptuous as to claim that we can predict where the course
of our research will take us or where external circumstances will force us to seek new frontiers. But we
can at least mention a few of the directions in which some of these developments will take us and some
of the milestones we should reach.
Improvements in materials technology will permit us to go to heads per stage of 1,800 m (6,000 ft)
or higher. Boundary layer control will lead to improved efficiencies in the lower range of specific speeds.
There is much too much use of expensive and scarce materials in services where they are needed only
at the boundary surface where they are in contact with the pumped liquid.

WORLDWIDE STANDARDIZATION

A most important source for future improvements is the growing possibility of worldwide standardization.
It has been an accepted belief that the major obstacle to worldwide standardization of ratings and even
954 The Centrifugal Pump of Tomorrow

geometric dimensions of centrifugal pumps has been the conflict between the use of U.S. units in our
part of the world and that of SI units everywhere else. It seems to us that this is merely an oversimplification
of the problem. In our opinion, the major obstacle is the existence of two different current frequency
standards, that of 60 and 50 cycles. And here is where the recent development of a commercial version
of variable-frequency drive will come to our rescue.
We have already stated that the use of variable frequency eliminates the forced selection of speeds
dictated solely by the electrical characteristics of the territory where the pumps are to be installed. All
that remains for us to do is to come to an agreement as to the speed or speeds at which the standard
pump ratings will be established.

WEAR AT THE INTERNAL RUNNING CLEARANCES

Two distinct trends are evident in the design of running clearances. When the pumping temperature and
differential pressure allow them, polymers of various types, for example PEEK (polyetheretherketone),
are being used to allow closer running clearances, hence some improvement in pump efficiency, while
lowering the risk of damage or seizure in the event of incidental contact within the clearances. For
service conditions beyond the capability of polymers, hard coating techniques developed for gas turbines
and compressors are being employed to lower the wear rate and raise the galling resistance of running
clearances. Hard coatings are already being applied directly to impeller hubs (except those that are
dilated when the impellers are shrunk on the shaft), thereby eliminating impeller wearing rings and
simplifying the pump. One can predict that the centrifugal pump of the future may dispense with
renewable wearing parts altogether as it may become more economical to deposit the hard coating
directly on the stationary surfaces. If the thickness of the deposit is such that 5 to 10 years will elapse
before it is completely worn off, it would be an acceptable life span to the user.

CAVITATION EROSION RESISTANCE

Materials such as high manganese austenitic stainless steels have shown significantly higher resistance
to cavitation erosion than the materials commonly used today. The combination of design to lower the
risk of cavitation erosion (see Chap. 19) and materials of much greater cavitation erosion resistance
promise both faster, smaller pumps and longer impeller life.

BEARINGS

There is much promise in a variety of approaches to a bearing solution wherein the fluid being pumped
acts as the lubricant. This, of course, is not entirely new, but great advances are still possible in improving
the life and reliability of such bearings. It may well be that a combination of pumped-liquid lubrication
and of magnetic bearings is the course to study next.
Magnetic bearings (see Chap. 11) are already being employed in centrifugal pumps. The designs are
active, using feedback from proximity probes to vary electromagnet excitation and so keep the rotor in
its correct position. Before they are widely used, however, standard designs able to run in the pumped
liquid are needed and the control systems have to be simplified.
And it is the development of such bearings that will ultimately lead us to the hermetic pump of the
future. We do not mean to imply that we have no hermetic pumps today. But they are still not as reliable
or as efficient as would be necessary for them to obtain the share of the pump market that this concept
The Centrifugal Pump of Tomorrow 955

merits. We can visualize the evolutionary process beginning with the perfecting of activated bearings.
Then, these bearings will be moved into the interior of the pump, located probably at such positions as
the wearing rings. Later, pancake-type motors will be located in the same position, surrounding the
activated magnetic bearings. Presto! Here will be the hermetic pump that we spoke of when we deplored
the fact that today's mechanical seals will not meet tomorrow's demands for pollution-free environment.
Prototype single stage hermetic pumps incorporating some of these principles have already been built,
but none yet incorporate all of them in the form described. Adapting the same principles to multistage
pumps will raise another problem: how to develop the power required in a well proportioned, high-
speed, hermetic motor.
An interesting observation on all of this is that the centrifugal pump of the future will incorporate a
great deal of electrical apparatus: the electric motor will be integral; the bearings will be electronic and
integral, and the motor controls will interface directly with the plant controls.

CONCLUSIONS

We cannot imagine the discovery of a new principle that would supplant the centrifugal pump in the
task of moving liquids from one point to another. It follows then that our task in the search for
improvements of this pump will never be finished. And as exciting as the last 50 years have been, we
firmly believe that the next 50 years will be even more so. One of us wishes he could be part of the
team that will be working on the program we have outlined here; the other looks forward to participating
in it.
Data Section

957
958 Data Section

Index of figures and tables in the text that are of special interest

Atmospheric pressures up to 3,660 m (l2,OOO ft), Fig. 18.5, NPSH available, recommended margin over NPSH required,
407 Table 19.2,504
Boiler feed pump materials, selection of, Table 17.7, 397 NPSH available, recommended values for double-suction
Capacity conversions, Fig. 18.1,403 pumps, Fig. 19.8,482
Capacity equivalents, Table 18.1,402 NPSH available, recommended values for single-suction
Efficiency of single stage end suction and double suction pumps, Fig. 19.7,482
centrifugal pumps, Fig. 18.38, 448 NPSH required, correction for pump hydrocarbon and hot
Efficiency (bowl) of wet-pit centrifugal pumps, Fig. 18.39, water, Fig. 19."14,490
449 NPSH surplus obtained by sub-cooling, Fig. 19.19,496
Frequency and voltage effects on a-c motor characteristics, pH values, effect of temperature on, Fig. 17.6,387
Table 24.5, 646 Piping connections and their K values, Fig. 20.3, 518
Friction losses in fittings expressed as length of straight pipe, Pressure and head conversion, Fig. 18.3, 406
Fig. 20.7, 521 Pressure and head equivalents, Table 18.2, 405
Galvanic series, Table 17.6,388 Pressure, atmospheric to 3,660 m (12,000 ft), Fig. 18.5, 407
Hazardous area classification, Table 24.4, 643 Pump selection and construction data sheet, Fig. 27.1, 802
Head and pressure conversions, Table 18.3,406 Relative roughness, effect on pump efficiency, Fig. 18.41,
Head and pressure equivalents, Table 18.2, 405 451
Head loss and velocity in new piping, Table 20.2(c), Specific speed and impeller shapes, Fig. 4.8, 66, Fig. 18.30,
515-516 440
Head loss and velocity in piping, k = 0.1 mm, Table 20.2(a), Speed correction for pump capacity, Fig. 18.40,450
513 Synchronous speeds for a-c motors, Table 24.2, 633
Head loss correction factor, k ± 0.1 mm, Table 20.2(b), 514 Temperature depression, Fig. 26.23, 750
Head loss and velocity in old piping, Table 20.1, 510-511 Viscous-liquids, correction chart for pump handling, design
Head loss in sudden contraction of pipe, Fig. 20.5, 519 capacities to 23 m3Jhr (100 gpm), Fig. 18.53,467
Head loss in sudden enlargement of pipe, Fig. 20.6, 519 design capacities about 23 m3Jhr (100 gpm), Fig. 18.54,
Impeller profile variation, indicating approximate range of 468
specific speeds, Fig. 4.8, 66 Voltage and frequency variation effects on a-c motor charac-
Impeller shapes and efficiency variations with specific speed, teristics, Table 24.5, 646
Fig. 18.30, 440 Voltage and corresponding normal motor size ranges, IEC
K values of several piping connections, Fig. 20.3, 518 practice, Table 24.1(a), 631
Liquids, their velocity in pipes, Fig. 20.6, 520 U.S. practice, Table 24.1(b), 631
Materials, chemical composition, Table 17.3,371-375 Water, its specific gravity, temperature, and vapor pressure
Materials, common combinations for various pump services, relations, Fig. 18.6, 408
Table 17.2,370 Water, specific gravity from 30 to 75O"F, Fig. 26.4, 722
Materials for various classes of pump fittings, Table 17.5, Water, vapor pressure and density from 0.01 to 374.15°C,
382 Table 26. 1(a), 721
Motor characteristics affected by voltage and frequency Water, temperature-vapor pressure relationship, Table
variation, Table 24.5, 646 26.1(b), 722
Motor size ranges and corresponding standard voltages, IEC Wearing ring clearances, Fig. 4.41,86
practice, Table 24.I(a), 631 Williams and Hazen coefficient C variations, Fig. 20.2,
U.S. practice, Table 24.1(b), 631 508
Data Section 959

TABLE A.l CONVERSIONS, CONSTANTS, AND FORMULAS

VoluUle and Weight Power and Torque-(Continued)


1 U. S. gallon =8.34 IbsXSp Gr
bhp = gpmXHead in feetXSp Gr
1 U. S. gallon =0.84 Imperial gallon
1 cu ft of liquid =7.48 gal 3960 X efficiency
1 cu ft of liquid =62.32 IbsXSp Gr
Specific gravity of sea water = 1.025 to 1.03 bhp = gpm X Head in psi
1 cu meter = 264.5 gal 1714 X efficiency
1 barrel (oil) = 42 gal
Navy formula to determine lip rating of motor:
Capacity and Velocity
1 gpm = 449 cu ft per sec Hp = ohp ( 1.05 + h1.35 )
o p+3
Ibs per hour
gpm = 500XSp. Gr. efliciency of pump
gpm = 0.069 X boiler Hp where ohp is the output horsepower or water horse-
gpm =0.7Xbbl/hour=0.0292 bbl/day power work done by the pump which is determined
gpm = 0.227 metric tons per hour by:
1 mgd = 694.5 gpm gpm X Head in feet X Sp Gr
V =gpm XO.321 grm X 0.401) ohp 3960
area in sq in 02
gpm X Head in psi
V=y2gH or ohp 1714
gpm = gallons per minute
Sp Gr = specific gra\·ity based on HpX5252
Torque in lbs feet
water at 62°F rpm
H p = horsepower bhp = brake horsepower
bbl=barrel (oil) =42 gal rpm = revolutions per minute
mgd = million gallons per day of 24
hours Miscellaneous Centrifugal PUUlP ForUlulas
V = velocity in ft /sec
D = diameter in inches
. y'gpmXrpm
g = 32.16 ft /sec /sec Specific speed = N ,
H = head in feet • H~~
Head where H = head per stage in feet
Head in feet = Head in psi X2.31
Sp <.ir 1840 KUVH
Diameter of impeller in inches =d
1 foot water (com. fresh) = 1.133 inehes of rpm
mercury
1 psi =0.0703 kilograms per sq centimeter where Ku is a constant varying with impeller type
1 psi =0.068 atmosphere and design. Use H at shut-off (zero capacity) and
Ku is approx. 1.0
H= V2
2g
psi = pounds per square inch dl gpml y'HI ~bhpi
At constant speed: - = - - =-= = - =
Power and Torque d2 gpm2 yH2 ~bhp2
1 horsepower = 550 ft-lb per sec
= 33.000 ft-Ib per mill At constant impeller diameter
=2545 btu per hr
rpml gpml y'HI ~Bhpi
=745.7 walts - - = - - = ---==- = - =
=0.7457 kilowatts
960 Data Section

TABLE A.2 MEASUREMENT CONVERSION

Anl.trom Unit ..... 10-10 me'er Ho..., .•••....... 33,000 rt Ib pot" ruinnle Ounc •............ t37.5 grains
1,'10.000 micron 42.41 Btu per minute 0.911 troy ouucoa
O.00393i millinnl.h~ or un 1.014 cheval 28.35 ....._
inch ":"-'6wat ...
Ounc. (Fluid) . . • . .• • 1.805
cubic ioel...
Atm ••ph.'•... .... 14.7 po..."l=- (Jo;uglitlh) Hund.od ••I.ht 29.573 miUiD ......
1-'.223 POlllldM (RumM"") 'Iriti,", .....•..... 112 pounds Ounc. (Fine) ....... Troy ounce
50.80 kiloJrrali. 480graill8
Itu (8rm". Thermal Unit) 7i8 foot PlMIUdli 31.104 ....._
0.2930 watt. hour Inch . ... " ......... 39.540 .1..-2 waye Ifln.LlIIS of
0.252 c:tlorie red ray or cadmium Plod (F...." fool) . .... 12 Paris inch..
25.-1 millimet.E!rs 0.325 meter
Calo,I •. ........... 1 kiloglllll of' WllW" raiMed PInt ............... 0.4732 liter
1 de!Uf'.e CfOnt.igrade .Ioul•.... . . 1 watt MCOnd
16 fluid ounces
3.97 "tu lCJIo••am. .. . ... 2.2046 pound. Pound Avol.dupol.: 16 ouuces
Centar •. (square met.lIO.76-1lK1llore r~t 35.274ou..... 7.000 grain.
15432.36 l1'aius 454 .........

p..
Centlllleter . ....... 0.3937 inch 0.0011 shorl ton U54 kiJosram
0.00098 101lM' ton 14.58 troy oune..
Ch.val (F,ond! hp.) . •. 0.986 h'",,"power Pound CubIc
K••••am p •• CubIc 'oot ............ 16.02 ki ........m per cubic
CI.cula. Mil .. " .... Area or circle wI..... Met••.. " ....... 0.062' II.. per cu fl
diameter is one mil or meter
1/1.000 inch KII•••am Pound p •• SCI In ... 2.31 reet head or water at
0.00000078;) MIllON! ineh pO. SClH•• 1.00 ap ..
C.ntlm.t••. .1 t:!:!:) 11M 1_ sq in 0.0703 kilogram per sq
CubIc C.ntlm.t•• centimeter
(milliliter) 0.061 cubic inch Kllo••a ..
p •• SClua •• Pound p •• SCI 't. .. ~.88 ki........ m per square
Cubic 'oot . ........ 1,728 cubic incht'!r4 M.t... . . . . . .. . .0.205 !be per oq rt meter
7.48 gallons Quart. " . . . . .. . ... 2 pinta
60 pints Kllemet ••......... 1,000 meten 14 plloll
8,'10 bushel 0.621 mile 8.9~ liter
62.32 lbe water (6:2°«.,) Quart (SrWish quarlOl"
1,000 ouncp.s oj' wH.l~r,
Kllo.edt. . .... " .• 1.34 honepowor
44.257 rt Ib per minute hundmw.ight) ...... 28 pound.
approx.
56.87 Btu per minute 12.70 ki ............
0.028 cubic meier
liters Qulnt.I ........... 100 ..i ........ms
~8.32 Lit ••.. .... 1.000027 cubic decimeter
220.~ pounds
1.057 quart
CubIc Inch ......... 16.39 cubic centim~' .... 0.264 881100 St••o . . . . . . . . ..... 1 cubic meter
61.02 cubic inch.,. SClua•• C.ntlmet•• 0.155 square inch
CubIc M.t••....... 35.315 cubic reet .035 cubic reet
1.308 cubic Y8rd~
33.8147 Ouid ounc... SClUa•• 'oot ....... 0.093 """"'" meter
1.14 square inches
CubIc Ya.d ........ 2i cubic reel 270.518 Ouid drams
0.765 cubic 111tt:It'!r SCI.ar. Inch . ...... 6..&52 square centimeters
Lit•• p •• S.calld ... 2.12 co rt per minule
o.-&n U. S. Gal per miu SClua•• K.omet••.. 0.386 square mile
Dedm..er . ........ 3.937 inches SClua•• M.t.r( _ _110.764squ&re reet
' • •t. . . . . .. . .. . ... 12 inches
Mete.. . .. . .. . . . . . .39.37 incheo 1.196equareyaM
3.28 feet
00185 meter l.09yuda
Stlua•• Mil ........ 0.000001 equare Inch
0.000645 sq oentimet.or
'oot Pound ........ 0.1364 kilogram..... ' .... Metric 'on ........ 2204.6 pounds ..ua•• Mil •....... 640 ......
1.1023 short to...
Gallon . ........... 231 cubic inches 3.097,600 """"'" yards
4quarta MIcron ............ 0.001 miDi_ 2.59 """"'" kiIometen
8 pints 10.000 Anptrom IInibl Stlva.. MIIR...to•.. 0.00155 square mcb
3.785lit.ers 39.37 miWODtha of au iu
128 ftuid ou ...... Stlva•• Y.rd ....... 0.836 oquare meter
8.33 pounds or w ...... Micro............. 1/1000 millipam Stono (Srll_) ••••••• 14 pounda
6.35ki\ocrama
Gallon p.. Mlnut. 419 cubic reet per -.ond MR ........ " ...... 0.OOlinch
25.4 _ _
0.227 metric totw per huur
'on ("""') . . . . . . .... 2.000 pounds
907 ki\otlruM
'on
0.0254 miDimel«
• •on ,Iri,i, .. '.,..;ot) 277.3 cubic inches (long) •••••••••• 2.240 pounda
1.201 U. S.....00... Mlle ............... 1.760 yards
1.016 ki\ogra_
10 U. water at 15°C.
5.280 reet 270 pIIono
1.61 kilometen
4.S46liten
, •• ,.r Hou. (metric).&.4 pIIono per minute
. . . . .0 • . . . . . . . . . . 0.015-& grain
Groin . ............ 1/7000 lb ayuir.lnpuill '.nn. (_i<l ....... 1.000 ~
O._ .....m _ . . . . . .. . . . . . . 1.000027 COl _timet ... 22OU2pounds
0.0610 cubic inch
.ro............... 15.43 pallbt Ya.d .............. 3reet
0.0353 ounce MyrI.moto•....... 10.000 . . - 36incheo
0.0022 pound 6.213~ miIeo 0.914_
Data Section 961

TABLE A.3 FRICTION LOSS IN FInINGS FOR INTERPOLATED INTERMEDIATE VISCOSITIES


It should be noted that this table is only an approximation. Very little reliable test data 0/1 losses
in valves and fittings for laminar {low are available. Figure 20.7 is on page 521.

3-30 gpm 30-50 gpm 50-JOO gpm JOO-250 gpm 250-J ,000 gpm
ssu ssu $lU SSU $$U

Use full value from Fig. 20.7 when


viscosity is 100 200 300 400 500
Use '% value from Fig. 20.7 when
viscosity is 1,000 2,000 3,000 4,000 5,000
Use Y2 value from Fig. 20.7 when
viscosity is 10,000 20,000 30,000 40,000 50,000
Use ',4 value from Fig. 20.7 when
viscosity is 100,000 200,000 300,000 400,000 500,000
Use actual length of valves and fit.
tings when the viscosity exceeds 500,000 500,000
962 Data Section

TABLE A.4 FRICTION LOSS FOR VISCOUS LIQUIDS

Loss in pounds per square inch per 100 feet of new Schedule 40 steel pipe based on specific grav-
ity of 1.00. (For a liquid having a specific gravity other than 1.00, multiply the value from the table
by the specific gravity of that liquid. For commercial installations, it is recommended tllat 15 per cent be
added to the values in this table. No allowance for aging of pipe is inclllded.)

Pipe VISCOSITY-SA YBOLT SECONDS UNIVERSAL


GPl\f Size
100 200 300 400 500 1000 1500 2000 2500 3000 4000 5000 6000 7000 8000 9000 10,000 15,00II
-
% 1.2 23.6 35.3 47.1 59 118 177 236 294 353 471 589 706 824 942 ..... ...... ......
3 ~ 3.7 7.6 U.S 15.3 19.1 38.2 57 76 96 liS 153 191 229 268 306 344 382 573
1 1.4 2.9 4.4 5.8 7.3 1'1.5 21.8 29.1 36.3 43.6 58 73 87 101 U6 131 U5 218
- --64 - -
96
-- --
159
-- -- -- - ---- - - - - -
319 382 ,1-16 510 573
- -956
~ 6.1 12.7 19.1 25.5 31.9 127 191 255 637
5 1 2,3 4,9 7.3 9,7 12.1 2".2 36.3 48,5 61 73 97 121 145 170 19·' 218 2'2 363
lU 0.77 1.6 2.4 3,3 4.1 8,1 12,2 16,2 20.3 2,1,3 32,5 40,6 48.7 57 65 73 81 122
- -- - - - - - - -- - - - - - - -- -- --- - -- -I -
~ 8,5 17,9 26,8 35,7 44.6 89 134 178 223 268 357 ·146 535 713 803 892 ...... 62'
7 1 3,2 6,8 10,2 13.6 17 33,9 51 68 85 102 136 170 203 237 271 305 339 509
lU 1.1 2.3 3.4 4.5 5,7 11.4 17 22,7
- -- - - -- -- - - -
28.4 34.1 45.4 57 68 80 91
- - - ---- - - -
102 IH
-727170

10
1
lU 4iJ 9.7 14,5 19.4 24.2
3,3 4,9 6,5 8,1
48,5
16.2
73
24.3
97 121 145 194
32,5 40.6 48,7 65
242 291
81 97
339
II4
388
130
·\36
146
485
162 2\3
- - -727- -......
1% 0,84 1.8 2,6 3,5 4.4 8,8 13,1 17,5 21.9 26.3 35 43.8 53 61 70 79 88 131
14,5 21.8 29,1
- -- --
36,3 73
-- - - -- -
363 436
--
581 654
1 11 109 145 182 218 291 509
15 lU 2,8 4,9 7,3 9.7 12,2 24,3 36,5 48,7 61 73 97 122 146 170 195 219 2t:'l 365
1~ 1.3 2,6 3,9 5,3 6,6 13,1 197 26,3 32,8 39.4 53 66 79 92 105 lI8 131 197
I---- - - - -- - - - - - - - ,\85 -- - --.....
-I -

20
1

1~
18
lU 4,9
2,3
+.T
3,5
29,1 38,8 48,5
9,7 13
5,3 7
16.2
8,8
97 145 194
32.5 48,7 65
17,5 26,3 35
242
81
291
97
43,8 53
130
70
388
581 678
162 195 227
88 105 123
775 872
260
140
292
158
......
,
325
175
487
263
2 0,64 1.3 1.9 2,6 3.2 6,4 9,6 12,9 16,1 19,3 25,7 32,1 38,5 45 51 58 6i 96
- - - --- -- -- -- -- -- - - -- - ---- - - - - -
1~ 3,5 4.4 6.6 8.8 11 21.9 32,8 43,8 55 66 88 IIO 131 153 176 197 219 328
25 2 1 1.6 2.4 3,2 4 8 12,1 16,1 20,1 24,1 32,1 40.2 48,2 56 M 72 80 121
-- -- --
0,4 0,79 1.2 1.6 2 4 5,9 7,9 11,9 11,8 15.8 19,7 23.7 27.6 31.6 35.5 39.5 59
- 2~
-- - - -26,3- -39.4
5,3 7,9 10.5 13,1
-- - -66- - - - - -- - - - -- 394
1~ 5 53 79 105 131 158 184 210 237 263
30 2 1.4 1.9 2,9 3.9 4,8 9,6 14,5 19,3 24,1 28,9 38.5 48,2 58 67 77 87 96 145
2% 0.6 0.95 1.4 1.9 2.4 4,7 7,1 9.5 11,8 14,2 19 23.7 28.4 33.2 37,9 42,6 47.4 71
r--- --- - - -- - - - - - - --- - -- -- -- -- -I -
1~ 8.5 9 10,5 14 17,5 35 53 70 88 105 140 175 210 245 280 315 350 526
2 2,5 3,9 5,1 6,4 12,9 19.3 25,7 32.1 38.5 51 64 77 90 lQ3 116 129 193
~
40
1.1 1.9 2,5 3,2 6.3 12.6 15,8 19 25,3
31.6 37,9 57 63 95 44,2
I---
2~
- -669.5--88
-43.8 -- 51
- --- - - -- - - -- - - - -- -I -
50
1~ 12,5
2 3.7
1'1
4 rtT 17,5 21.9
6,4 8 16.1 24.1
110
32.1
131
40,2
175 219 263 307 350 394
48,2 64 80
438 657
96 112 129 145 161 241
-58 - -- ------ -- - - -193-I289-
1.6 1.7 2.4 3,2 4 7.9 11.8 15,8 19.7 23.7 31.6 39.5 47.4 55 63 71 79 118
- 2~

60
2
2%
5
2,2
5,8
2.4 +.tJ 7,7
3.8
9,6
4.7
19.3
9,5
28,9
14.2
38,5
19
48.2
23,7 28.4
77
37,9
135 154 173
96 116
47.4 57 66 76 85 95 142

- 3
2~
0,8
2.8
0,8
3,2
1.2 1.6
3.4 4,4 5,5
2
--
4
-- 6
11.1 16,6
8
22.1
9.9
27,6
11.9
33,2
15.9
44,2
19,9 23.9 27,9 31.8 35.8
55 66 77 88 100 111 166
39.8 60
I-

I---
70 3
4
1
0.27 rHt
4,2
1.4 1.9 2.3
0.47 0,63 0.78
5,1 6,3 12.6
4,6
1.6
7
2,4
9.3
3,1
25,3
11.6
3.9
13.9
4,7
18,6
6,3
-51
23,2 27.8 32,5 37.1 41.7 46.4 70
7.8 9.4 11 12.5 H.l 15.6
I~

I!:~
2~ 3.6 19 31.6 37,9 63 76 88 101 114 126 190
80 3 1.3 1.4 2,1 2,7 5.3 8 10,6 13,3 15~9 21.2 26,5 31.8 37,1 42.4 47.7 53 80
4 0.36 O.:ui 0.72 0,89 1.8 2,7 3.6 4,5 5.4 7,2 8.9 10.7 12,5 H.3 16.1 17.9
I--- I~

rK fHJ
2~ 5.3 6.1 6.4 8 15.8 23.7 31.6 39,5 47.4 63 79 95 111 127 142 158 237
100 3 1.9 2.2 3.3 6.6 9,9 13.3 16,6 19,9 26.5 33,1 39.8 46,4 53 60 66 99
4 0.52 0,57 0.89 1.1 2,2 3..1 4,5 5,6 6,7 8,9 11.2 13,4 15.6 17.9 20.1 22,3 33.5

TURBULENT
~-~FLO~~W--~.~~4---------------LAMINAR FLOW---------------+.

Cojl)'right 1955 by the HydrauTic Illstitule.


Data Section 963

TABLE A.4 (CONT.) FRICTION LOSS FOR VISCOUS LIQUIDS

Loss in pounds per square inch per 100 feet of new Schedule 40 steel pipe based on specific grav-
ity of 1.00. (For a liquid having a specific gravity other than 1.00, multiply the value trom the table
by the specific gravity of that liquid. For commercial installations, it is recommended that 15 per cent be
added to the values in this table. No allowance tor aging of pipe is included.)

VISCOSITY-SAYBOLT SECONDS UNIVERSAL


GPM Pipe
Size
100 200 300 400 500 1000 1500 2000 2500 3000 4000 5000 6000 7000 8000 9000 10,000 15~
I-

Ig.nrt+
I-- I-
3 2.7 3. I 3.2 4 8 11.9 15.9 19.9 23.9 31. 8 39.8 ·17.7 56 61 72 80 119
120 4 0.73 0.81 1.3 2.7 4 5.,1- 6.7 8 10.7 1:1.1 16.1 18.8 21A 21. I 26.8 40.
6 .098 0.11 0.21 0.26 0.52 0.78 1.0 1.3 1.6 2.1 2.6 3.1 :1.6 1.2 4,7 5.2 7.H
'---- -

ttb r++
3 3.4 4 4.3 4.6 9.3 13.9 18.6 23.2 27.8 37.1 16..1 56 65 jJ. 81 93 139
1.f.O 4 0.95 1.1 1.6 3.1 4.7 6.3 7.8 9.4 12.5 1;'.6 18.8 21.9 ?-
_.) 28.2 31.3 16. ~
6 0.13 O. IS 0.21 0.30 0.61 0.91 1.2 1.5 1.8 2.1 3.0 :1.6 ·1.2 ·1.9 5,3 6. I 9.1
- 1-
160
3
4
6
4.4
1.2
0.17
5 5.7 5.7
1.4 1.4 IA .8 rH-
10.6 15.9
3.6 5.4
0.18 0.21 u.28 0.35 0.69 1.0
21.2
7.2
1.4
26.5
8.9
1.7
31.8
10.7
2. I
12.1
14:3
2.8
5:1 M 7·j.
17.9 21.5 25
85 95 106 159
28.6 :12.2 35.7 5,1
:L5 4.2 4.9 5.5 6.2 6.9
I-- --- - I~
3 5.3 6.3 7 7
H- 1l.9 17.9 23.9 29.8 35.8 47.7 60 72 81 95 107 119 179

I--
180 4
6
--
1.5
0.2
- - --
1.8 1.8
0.24 0.24 rHt 2 4 6
0.39 0.78 1.2
8
1.6
-26.5- -
10.1
2
12.1
2.3
16. I
3.1
20.1 21.1 28.1 32.2 36.2 40.2 60
:1.9 4.7 5.5 6.2 7
- - - - - - - -- - - - - - - - -I -
7.8 II.7
3 '6.5 7.7 8.8 8.8 8.8 13.3 19.9 33.1 39.8 53 66 80 93 106 119 133 199
200 4 1.8 2.2 2.2 2.2 2.2 4.5 6.7 8.9 11.2 13.4 17.9 22.3 26.8 31.3 35.7 ·10.2 4·1.7 67
6 0.25 0.3 0.3 0.87 1.3 1.7 2.2 2.6 3.5 ·L3 5.2 6.1 6.9 7.8 8.7 13
-- -2.6- -3.2- - - - - - - - - - - -- - - - - - - - -I -
! U.3' U.4J
I--
250
4
6
3.5 3.5
0.36 0.43 0.45 rH, 5.6
1.1
8.4 11.2 H
1.6 2.2 2.7
16.8
3.3
22.3
4.3
27.9 33.5 39.1 44.7 50
SA 6.5 7.6 8 .. 7 9.8
56
10.8
84
16.3
I--
8 .095 0.12 0.12
-4- -3.7- ~ 8:t~ 0.18 0.36 0.54 0.72 0.9 1.1 1.5 1.8 2.2 2.5 2.9 3.3
- - - - - - - - - - - - - - - - - - - -I - 3.6 5.4
4.3 5 5 5 6.7 10. I 13.4 16.8 20.1 26.8 33.5 40.2 47 51 60 67 101
300 6 0.5 0.6 0.65 0.65 0.65 1.3 2 2.6 3.3 3.9 5.2 6.5 7.8 9.1 lOA 11.7 13 19. ~
8 0.13 0.17 0.17 I U.18 U.2~ 0.43 0.65 0.87 1.1 1.3 1.7 2.2 2.6 3 3.5 3.9 4.3 6.5
I-- -- -- -3.5- - - - - - - - - - - - - - - - - - -I -
6 0.82 I 1.1 1.2 1.2 1.7 2.6 4.3 5.2 6.9 8.7 10.4 12.1 13.9 15.6 17.3 26

I-
400 8
10
--
0.23
0.08
-1.2-
0.27 0.29 0.29
0.09 0.1 0.1 I Hi 0.58
0.23
0.87
0.35
1.2
0.47
1.5
0.58
1.7
0.7
2.3
0.93
2.9 3.5 4.1 4.6 5.2
1.2 1.4 1.6 1.9 2.1
-4.3- - - - - - -- -- - - - - - - - - -I -
5.8
2.3
8.
3.5
6 1.5 1.6 1.8 1.8 2.2 3.2 5.4 6.5 8.7 10.8 13 15.2 17.3 19.5 21.7 32.1
500 8 0.33 0.39 0.44 0.47 0.47 0.72 1.1 1.5 1.8 2.2 2.9 3.6 4.3 5.1 5.8 6.5 7.2 10.1
10 0.11 0.14 0.15 0.15 O.IE 0.29 0.44 0.58 0.73 0.87 1.2 1.5 1.8 2 2.3 2.6 2.9
I-- -- - - -2.2- -2.3- I~
600
6
8
1.8
0.47
2.4 2.6
0.57 0.62 0.67 0.6 ~
0.87
3.9
1.3
5.2
1.7
6.5
2.2
7.8
2.6
10.4
3.5
13 16 18.2 20.8 23.4
4.3 5.2 6.1 6.9 7.8
26
8.7
39
13
10 0.16 0.18 0.2 0.22 0.2: 0.35 0.52 0.7 0.87 1.1 1.4 1.8 2.1 2.4 2.8 3.3 3.5 5.
I--- - - - - - - - - - - - - - - - - - - - -- - -- - -I -
r¥-
6 2.3 2.7 3 3.2 3.5 4.6 6.1 7.6 9.1 12.1 15.2 18.4 21.2 24.3 27.3 30.3 45.1
700 8 0.6 0.71 0.82 0.89 0.93 1.5 2 2.5 3 4.1 5.1 6.1 7.1 8.1 9.1 10.1 IS.
0.25 0.27 0.3 0.3
-6.9- --
10 0.2 0.41 0.61 0.82 I 1.2 1.6 2 2.4 2.9 3.3 3.7 4,1 6.1
I--- -- - - - - - - - -- - - -- - -- - - -- - -I -
6 2.8 3.5 3.7 4 4.2 4.8 5.2 8.7 10.4 13.9 17.3 20.8 24.3 27.7 31.2 34.7 52
800

-- --
8
10
0.78
0.26
-3.5- -4.3- -4.6-
0.94 1 1.1 1.2
0.3 0.34 0.38 0.4 rH .4
1.7
0.7
2.3
0.92
2.9
1.2
3.5
1.4
4.6
1.9
5.8 6.9 8.1 9.3 10.4
- - - - - - - -- - -I -
2.3 2.8 3.3 3.7 4.2
-11.7- -15.6- -19.5
11.6
4.7
17.3
7
6
900 8 0.95
5.0 5.2
1.1 1.3 1.4 1.5
6
r-4-
2
7.8
2.6
9.8
3.3 3.9 5.2 6.5
23.4 27.3 31.2 35.1
7.8 9.1 10.4 11.7
39
13
58.~
19.5
I--
10 0.32
-- - --- -- 0.37 0.43 0.46 0.5 ~ 0.79 1.1 1.3 1.6 2.1 2.6 3.1 3.7 4.2 4.7 5.2
-
7.9
8 1.1 1.4 1.5 1.6 1.8 1.9 2.2 2.9 3.6 4.3 5.8 7.2 8.7 10.1 11.6 13 11.5 21.7

r+h
1000 10 0.38 0.45 0.5 0.55 0.6 0.87 1.2 1.5 1.8 2.3 2.9 3.5 4.1 4.7 5.2 5.8 8.7
12 0.17 0.2 0.22 0.24 0.25 0.43 0.58 0.72 0.87 1.2 1.5 1.7 2 2.3 2.6 2.9 4.3

TURBULENT
~.~---~--~~~.----------LAMINAR FLOW--------~
FWW

CO/Jyright 1955 by the Hydmlllic Institl/te.


964 Data Section

TABLE A.4 (CONT.) FRICTION LOSS FOR VISCOUS LIQUIDS

Loss in pounds per square inch per 100 feet of new Schedule 40 steel pipe based on specific grav-
ity of 1.00. (For a liquid having a specific gravity other than 1.00, multiply the value from the table
by the specific gravity of that liquid. For commercial installations, it is recommended that 15 per cent be
added to the values in this table. No allowance for aging of pipe is included.)

VISCOSITY-SA YBOLT SECONDS UNIVERSAL


GPl\1 Pipe
Size 20,000 25,000 30,000 40,000 SO,OOO 60,000 70,000 80,000 90,000 100,000 125,000 ISO,OOO 175,000 200,000
- ~
2 19.3 24.1 28.9 38.5 48.2 58 67 77 87 96 120 145 169 193 482
3 2% 9.5 11.8 14.2 19 23.7 28.4 332 37.9 42.6 47.4 59 71 83 95 237
3 4 5 6 8 9.9 11.9 13.9
29.8 15.9 99 17.9 19.9 24.9 34.8 39.8
- -2 --32 - -40 - - - - - - - - - - - - - - -- -- - - - - --- - -I -
48.2 64 80 96 112 129 145 161 201 241 281 321 803
5 2% 15.8 19.7 23.7 31.6 39.5 47.4 55 63 71 79 99 118 138 158 395
3 6.6 8.3 9.9 13.3 16.6 9.9 23.2 26.5 29.8 33 41.4 49.7 58 66 166
I - - - -- - - - - - - - - - - - - - - - -- - - - - - - - ---- - - -I -
2 45 56 67 90 112 135 157 180 202 225 281 337 393 456 ... "

7 2% 22.1 27.6 33.2 4·i.2 55 66 77 88 100 III 138 166 194 221 553
- --
3 9.3 11.6
- -- -- - - - - -
2% 31.6 39.5 47.4 63
13.9
79
--
95
18.6 23.2 27.8 32.5
- - - ---- -- - - -------- - -
III 126 142 158 197 237 276 316
37.1
790
41.7 46.4 58 70 81 93 232

10 3 13.3 16.6 19.9 26.5 33.1 39.8 "6.4 53 60 66 83 99 116 133 331
- - - - - - - - - - - - - - - - ---- - - - - - - - ---- - -
4 4.5 5.6 6.7 8.9 11.2 13.4 15.6 17.9 20.1 22.3 27.9 33.5 39.1 44.7 112
-
2% 47.4 59 71 95 118 142 166 190 213 237 296 355 415 474 .......
15 3 19.9 24.9 29.8 39.8 49.7 60 70 80 89 99 124 149 174 199 497
- - - - - - -- - - - - - - - - --- - - - - - ---- - - -I -
4 6.7 8.4 10.1 13.4 16.8 20.1 23.5 26.8
168 30.2 33.5 41.9 50 59 67
!- -3 - -
26.5 33.1 39.8 53 66 80 93 106 133 119 166 199 232 265 663
20 4 8.9 11.2 13.4 17.9 22.3 26.8 31.3 35.7 40.2 44.7 56 67 78 89 223
- -133- - - - -
6 1.7 2.2 2.6 3.5 4.3 5.2 6.1 6.9 7.8 8.7 10.8 13 15.2 17.3 43.3
I-- -3 - -33.1- -41.4
- -49.7--66 --
83
--
99
--
116 149 166
- - -------- -I -
207 249 290 331 828
25 4 11.2 14 16.8 22.3 27.9 33.5 39.1 44.7 50 56 70 84 98 112 279
2.2 2.7 3.3 4.3 5.4
9.8 6.5
10.8 7.6
13.5 8.7 16.3 21. 7
- -63 - - -- - - - -80 -- -- ---------- 19
- - - - -- - --- - - -
64
39.8 49.7 60 99 119 139 159 179 199 249 298 ....... 348 398
30 4 13.4 16.8 20.1 26.8 33.5 40.2 46.9 54 60 67 84 101 117 134 335
6 2.6 3.3 3.9 5.2 6.5 7.8 9.1 10.4 11.7 13 16.3 19.5 22.7 26
- -3 - -53 --66 --80 --
106
--
133
- - - ---
160 186
~--

212
-
239
- ---
265
-
331
--
398
--
464
--
532
-I65-
40 4 17.9 22.3 26.8 35.7 44.7 54 63 72 80 89 112 134 156 179 447
6 3.5 4.3 5.2 6.9 8.7 10.4
15.6 17.312.1
21.7 2613.9
30.3 34.7 87
I---- - -- - - - - - - - -
22.3 27.9 33.5 44.7 56 67
- ---
78
- - - - - - --- ---- ------ --- -
101
4 89 112 140 168 196 223 559
SO 6 4.3 5.4 6.5 8.7 10.8 13 15.2 17.3 19.5 21.7 27.1 32.5 37.9 43.3 108
- -- - - - -- - - - - - -- - - - --- ------ -
8 1.5 1.8 2.7 2.9 3.6 4.3 5.1 5.8 6.5 7.2 9 10.8 12.6 14.5 36.1
- --
-4 - -26.8 33.5 40.2 54 67
--
80 94 107 121 134 168 201 235 268 670
60 6 5.2 6.5 7.8 10.4 13 16 18.2 20.8 23.4 26 32.5 39 45.5 52 130
-4 --31.3- -- - --- - - - -- ------ -
8 1.7 2.2 2.6 3.5 4.3 5.2 6.1 6.9 7.8 8.7 10.8 13 15.2 17.3 43.4
- - -- --
39.1 46.9 63
-- -- -- --
78 94 110 125 141 156 196 235 313 274 782
70 6 6.1 7.6 9.1 12.1 15.2 18.4 21.2 24.3 27.3 30.3 37.9 45.5 53 61 152
-6 --6.9- --
8 2 2.5 3 4.1 5.1 6.1 7.1 8.1 9.1 10.1 12.6 15.2 17.7 20.2 51
!- - - - - - - - - --- --- ------ - - - ------- -I -
8.7 10.4 13.9 17.3 20.8 24.3 27.7 31.2 34.7 43.3 52 61 69 173
80 8 2.3 2.9 3.5 4.6 5.8 6.9 8.1 9.3 10.4 11.6 14.5 17.3 20.2 23.1 58
f----
10 0.93 1.2 1.4 1.9 2.3 2.8 3.3 3.7 4.2 4.7 5.8 ,---2- 8.2 9.3 23.3
6 7.8 9.8 11.7 15.6 19.5 23.4 27.3 31.2 35.1 39 48.7 59 68 78 195
90 8 2.6 3.3 3.9 5.2 6.5 7.8 9.1 10.4 11.7 13 16.3 19.5 22.8 26 65
- -6 - --
10 1.1 1.3
- - - r - - - - --- - -
- - - --17.3
8.7 10.8 13
1.6
21.7 26
2.1
30.3
2.6
34.7 39
3.1 3.7 4.2 4.7
-- ---
5.2
43.3
6.6
54
7.9
:---
65
9.2
------
76 87
I-
10.5
217
26.2

100 8 2.9 3.6 4.3 5.8 7.2 8.7 10.1 11.6 13 14.5 18.1 21.7 25.3 28.9 72
10 1.2 1.5 1.8 2.3 2.9 3.5 4.2 4.7 5.2 5.8 7.3 8.7 10.2 11.6 29.1

~~~-----------------------LA]fINAR FLOW------------------------~~

Copyright 1955 by the Hydraulic Institute.


Data Section 965

TABLE A.4 (CONT.) FRICTION LOSS FOR VISCOUS LIQUIDS

Loss in pounds per square inch per 100 feet of new Schedule 40 steel pipe based on specific grav·
ity of 1.00. (For a liquid having a specific gravity other than 1.00, multiply the value from the table
by the specific gravity of that liquid. For commercial installations, it is recommended that 15 per cent be
added to the values in this table. No allowance for aging of pipe is included.)

Pipe V1SCOSITY-8AYBOLT SECONDS UNIVERSAL


GPM Size 20,000 25,000 30,000 40,000 50,000 60,000 70,000 80,000 90,000 100,000 125,000 150,000 175,000 200,000 500,000
- 6 10.4 13 15.6 20.8 26 31.2 36.4 41.6 46.8 52 65 78 91 104
1-
260
120 8 3.5 4.3 5.2 6.9 8.7 10.4 12.1 13.9 15.6 17.3 21.7 26 30.4 34.7 87
~ --
10
6
-12.1- --
1.4
- - - - - - ---42.5
1.8 2.1 2.8
15.2 18.2 24.3 30.3 36.4
--48.5- -55- -61- -76--91--106-- -
3.5 4.2 4.9 5.6 6.3 7 8.7 10.5 12.2 14
121
34.9
I-
303
140 8 4 5.1 6.1 8.1 10.1 12.1 14.2 16.2 18.2 20.2 25.3 30.4 35.4 40.5 101
r-- -6 - --
10 1.7 2
-- - -- -- - - - - - - - - -- - -
- - - - - -- -
2.4
13.9 17.3 20.8 27.7 34.7
3.3 4.1 4.9
41.6
5.7
48.5 56
6.5
62
7.3 8.1
69
10.2
87
12.2
104
14.3
121
16.3
139
40.7
347
160 8 4.6 5.8 6.9 9.3 11.6 13.8 16.2 18.5 20.8 23.1 28.9 34.7 40.5 46.2 116
t - - - - ---- - - - - -- -- - - - -- - - -- -- - - - - - -
10 1.9 2.3 2.8 3.7 4.7 5.6 6.5 7.5 8.4 9.3 11.6 14 16.3 18.6 46.6
6 15.6 19.5 23.4 31.2 39 46.8 55 62 70 78 98 117 137 156 390
180 8 5.2 6.5 7.8 10.4 13 15.6 18.2 20.8 23.4 26 32.5 39 45.5 52 130
t - - ---- - - -- - -- - -- - -- - - -- - - - - -- - - --
10 2.1 2.6 3.1 4.2 5.2 6.3 7.3 8.4 9.4 10.5 13.1 15.7 18.3 21 52
8 5.8 7.2 8.7 11.6 14.5 17.3 20.2 23.1 26 28.9 36.1 43.4 51 58 145
200 10 2.3 2.9 3.5 4.7 5.8 7 8.2 9.3 10.5 11.6 14.6 17.5 20.4 23.3 58
12
8
1.2
7.2
----
1.5
9
- - -- - - -25.3
1.7 2.3
10.8 14.5 18.1 21.7
2.9
- -32.5
--28.9 - ---
3.5
36.1
-45.2
4.1
--54--63- -72- -181
4.6 5.2 5.8 7.2 8.7 10.1 11.6 28.9

2SO 10 2.9 3.6 4.4 5.8 7.3 8.7 10.2 11.6 13.1 14.6 18.2 21.8 25.5 29.1 73
t - - - --
12
8
1.5
8.7
1.8
-10.8 2.2
--13 - 2.9
- -I -
3.6
17.3 21.7
--
4.3
26
-30.4 - -39- ---
- -34.7
5.1 5.8
43.4
-54- -65--76- -87- -217
6.5 7.2 936.2 10.9 12.7 14.5

BOO 10 3.5 4.4 5.2 7 8.7 10.5 12.2 14 15.7 17.5 21.8 26.2 30.6 34.9 87
r-- -- -11.6--14.5- -- -- - -527.8 - -72
- -588.7 - -8713- -10115.2
--11617.4
- -28943.4
12 1.7 2.2 2.6 3.5 4.3 5.2 6.1 7 10.9
8 17.3
-- 23
I-
28.9 34.7
-40.5
--46.2
400 10 4.7 5.8 7 9.3 11.6 14 16.3 18.6 21 23.3 29.6 34.9 40.7 46.6 116
t-- - -
12
8
--2.3
- -- -- - - -- -51--58--65- ---
2.9 3.5 4.6 5.8
14.5 18.1 21.7 28.9 36.1 43.4 72
-90- -
7
108
--12620.3- -14523.2-,36158-
8.1 9.3 10.4 11.6 14.5 17.4

sao 10 5.8 7.3 8.7 11.6 14.6 17.5 20.4 23.3 26.2 29.1 36.4 43.7 51 58 146
12
8
2.9 3.6
1-
17.3 21.7
4.3 5.8
-26 --34.7 7.2
--43.4
-
8.7
I-
52
-6110.1--6911.6- -7813- ---
14.5
87
-
108
--13021.7-
18.1 25.3
152
28.9
173 434
72

600 10 7 8.7 10.5 14 17.5 21 24.4 27.9 31.4 34.9 43.7 52 61 70 175
- --
12
8
3.5
I- - --30.3--40.5 -51 - --
4.3
20.2 25.3
5.2 7 8.7
61
10.4
-71--81- ---
91
--- -126--152--177- --
12.2
101
13.9 15.6 17.4 21.7 26.1 30.4 34.7
202
I-
87
506
700 10 8.2 10.2 12.2 16.3 20.4 24.4 28.5 32.6 36.7 40.7 51 61 71 82 204
- -8 - --
12 4.1
- - -6.1- --
5.1
- - - - - - - - - - ---- -- - - - - -1578
8.1 10.1 12.2
23.1 28.9 34.7 46.2 58 69
14.2
81
16.2
93
18.2
104
20.3
116
25.3
145
30.4
173
35.5
202 231
40.5 101

800 10 9.3 11.6 14 18.6 23.3 27.9 32.6 37.3 41.9 46.6 58 70 82 93 233
- --
12
-26 -32.5--39 --52 --65 --78 - -91- -
4.6 5.8 7 9.3 11.6
--
13.9
--130--
16.2
--195--228- --
18.5 20.8 23.1
116
'650
28.9 34.7 40.5 46.3

-
8 104 117 163 260
900 10 10.5 13.1 15.7 21 26.2 31.4 36.7 41.9 47.1 52 66 79 92 105 262
--
12
8
-5.2
- - 6.5
28.9 36.1
- -, - -- - - - - - - - - --- - -- - - -
7.8
43.4
10.4
58
13
72
15.6
87
18.2
101
20.8
116
23.4
130
26.1
145
32.6
181
39.1
217
45.6
253
--52
289
-723
130

1000 10 11.6 14.6 17.5 23.3 29.1 34.9 40.7 46.6 52 58 73 87 102 116 291
12 5.8 7.2 8.7 11.6 14.5 17.4 20.3 23.2 26.1 28.9 36.2 43.4 51 58 145

~4------------LAMINARFLOW------------~~

Copvright 1955 bv the Hvdratllic [nstitute.


966 Data Section

TABLE A.S VISCOSITY OF COMMON LIQUIDS

VISCOSITY
Liquid ·Sp Gr at 60 F
SSU Centistokes At F

Freon 1.37 to 1.49 @ 70 F .27-.32 70

Glycerine (100%) 1.26@68F 2,950 648 68.6


813 176 100

Glycol:
Propylene 1.038 @68F 240.6 52 70
Triethylene 1.125@68F 185.7 40 70
Diethylene 1.12 149.7 32 70
Ethylene 1.125 88.4 17.8 70

Hydrochloric Acid (31.5%) 1.05 @68F 1.9 68

Mercury 13.6 .118 70


.11 100

Phenol (Carbolic Acid) .95 to 1.08 65 11.7 65

Silicate of Soda 40Baume 365 79 100


42 Baume 637.6 138 100

Sulfuric Acid (100%) 1.83 75.7 14.6 68

FISH AND ANIMAL OILS:


Bone Oil .918 220 47.5 130
65 11.6 212

Cod Oil .928 ISO 32.1 100


95 19.4 130

Lard .96 287 62.1 100


160 34.3 130

Lard Oil .912 to .925 190 to 220 41 to 47.5 100


112 to 128 23.4 to 27.1 130

Menhadden Oil .933 140 29.8 100


90 18.2 130

Neatsfoot Oil .917 230 49.7 100


130 27.5 130

Sperm Oil .883 110 23.0 100


78 15.2 130

Whale Oil .925 163 to 184 35 to 39.6 100


97 to 112 19.9 to 23.4 130
MINERAL OILS:
Automobile Crankcase Oils
(Average Midcontinent Paraffin Base):
SAE 10 ··.880 to .935 165 to 240 35.4 to 51.9 100
90 to 120 18.2 to 25.3 130

SAE20 ··.880 to .935 240 to 400 51.9 to 86.6 100


120 to 185 25.3 to 39.9 130

SAE 30 ··.880 to.935 400 to 580 86.6 to 125.5 100


185 to 255 39.9 to 55.1 130
·U nJesa otherwlIIe noted.
··Depends on origin or percent and type of solvent.

Revised 1958. Copyright ]955 by the Hydral/lic Institllte.


Data Section 967

TABLE A.S (CONT.) VISCOSITY OF COMMON LIQUIDS

VISCOSITY
Liquid *Sp Gr at 00 F
S8U Centiatokell AtF

SAE40 ··.880 to.935 580 to 950 125.5 to 205.6 100


255 to 55.1 to 130
SO 15.6 210

SAE50 ··.880 to .935 950 to 1,600 205.6 to 352 100


80 to 105 15.6 to 21.6 210

SAE60 ··.880 to .935 1,600 to 2,300 352 to 507 100


105 to 125 21.6 to 26.2 210

SAE70 ··.880 to .935 2,300 to 3,100 507 to 682 100


125 to 150 26.2 to 31.8 210

SAE lOW ··.880 to .935 5,000 to 10,000 1,100 to 2,200 0

SAE 20W ··.880 to.935 10,000 to 40,000 2,200 to 8,800 0

Automobile Transmission Lubricants:


SAE 80 ··.880 to.935 100,000 max 22,000 max 0

SAE 90 ••.880 to.935 800 to 1,500 173.2 to 324.7 100


300 to 500 64.5 to 108.2 130

SAE 140 ··.880 to.935 950 to 2,300 205.6 to 507 130


120 to 200 25.1 to 42.9 210

SAE 250 ••.880 to .935 Over 2,300 Over 507 130


Over 200 Over 42.9 210

Crude Oils:
Texas, Oklahoma .81 to .916 40 to 783 4.28 to 169.5 60
34.2 to 210 2.45 to 45.3 100

WyouUng, !&ontana .86 to.88 74 to 1,215 14.1 to 263 00


46 to 320 6.16 to 69.3 100

California .78 to.92 40 to 4,840 4.28 to 1,063 00


34 to 700 2.4- to 151.5 100

Pennsylvania .8 to .85 46 to 216 6.16 to 46.7 00


38 to 86 3.64 to 17.2 100

Diesel Engine Lubricating Oils (Baaed on


Average !&idcontinent Paraffin Baae):
Federal Specification No. 9110 ··.880 to.935 165 to 240 35.4 to 51.9 100
90 to 120 18.2 to 25.3 130

Federal SpecIfication No. 9170 ··.880 to .935 300 to 410 64.5 to 88.8 100
140 to 180 29.8 to 38.8 130

Federal Specification No. 9250 ··.880 to .935 470 to 590 101.8 to 127.8 100
200 to 255 43.2 to 55.1 130

Federal Specification No. 9370 ··.880 to .935 800 to 1,100 173.2 to 238.1 100
320 to 430 69.3 to 93.1 130

Federal Specification No. 9500 ··.880 to .935 490 to 600 106.1 to 129.9 130
92 to 105 18.54 to 21.6 210
·Unless otherwise noted.
··Depends on origin or percent and type of solvent.

Revised 1958. Copyright 1955 by the Hyrlmlllic Institllte.


968 Data Section

TABLE A.S (CONT.) VISCOSITY OF COMMON LIQUIDS

VISCOSITY
Liquid °Sp Gr at 60 F
SSU Centistokes AtF
Diesel Fuel Oils:
No.2D 00.82 to .95 32.6 to 45.5 2 to6 100
39 I to 3.97 130

No.3D 0·.82 to .95 45.5 to 65 6 to 11.75 100


39 to 48 3.97 to 6.78 130

No.4D **.82 to.95 140 max 29.8 max 100


70 max 13.1 max 130

No.5D .0.82 to .95 400 max 86.6 max 122


165 max 35.2 max 160

Fuel Oils:
No.1 00.82 to .95 34 to 40 2.39 to 4.28 70
32 to 35 2.69 100

No.2 00.82 to .95 36 to 50 3.0 to 7.4 70


33 to 40 2.11 to 4.28 100

No.3 00.82 to.95 35 to 45 2.69 to .584 100


32.8 to 39 2.06 to 3.97 130

No. SA .0.82 to .95 50 to 125 7.4 to 26.4 100


42 to 72 4.91 to 13.73 130

No.5B 0·.82 to .95 125 to 26.4 to 100


400 86.6 122
72 to 310 13.63 to 67.1 130

No.6 ··.82 to .95 450 to 3,000 97.4 to 660 122


175 to 780 37.5 to 172 160

Fuel Oil- Navy Specification 00.989 max llO to 225 23 to 48.6 122
63 to ll5 11.08 to 23.9 160

Fuel Oil - Navy II 1.0 max 1,500 max 324.7 max 122
480 max 104 max 160

Gasoline .68 to .74 .46 to .88 60


.40 to .71 100

Gasoline (Natural) 76.5 degrees API .41 68

GaaOil 28 degrees API 73 13.9 70


50 7.4 100

Insulating Oil:
Transformer, switches and ll5max 24.1 max 70
circuit breakers 65 max 11.75 max 100

Kerosene .78 to.82 35 2.69 68


32.6 2 100
Machine Lubricating Oil (Averap
Pennsylvania Paraftin Base):
Federal Specification No.8 ··.880 to .935 ll2tol60 23.4 to 34.3 100
70 to 90 13.1 to 18.2 130
·UnI_ otherwise noted.
··Depends on origin or pereent ad type of eoIvent.

Rel'ised 1958. Copyright 1955 by the Hydraulic Institute.


Data Section 969

TABLE A.S (CONT.) VISCOSITY OF COMMON LIQUIDS

VISCOSITY
Liquid ·Sp Gr at 60 F
SSU Centistokes AtF
Federal Specification No. 10 ··.880 to .935 160 to 235 34.3 to SO.8 100
90 to 120 18.2 to 25.3 130

Federal Specification No. 20 ··.880 to .935 235 to 385 SO.8 to 83.4 100
120 to 185 25.3 to 39.9 130

Federal Specification No. 30 ··.880 to .935 385 to 550 83.4 to 119 100
185 to 255 39.9 to 55.1 130

Mineral Lard Cutting Oil:


Federal Specification Grade 1 140 to 190 29.8 to 41 100
86 to 110 17.22 to 23 130

Federal Specification Grade 2 190 to 220 41 to 47.5 100


110 to 125 23 to 26.4 130

Petrolatum .825 100 20.6 130


77 14.8 160

Turbine Lubricating Oil:


Federal Specification .91 Average 400 to 440 86.6 to 95.2 100
(Penn Bue) 185 to 205 39.9 to 44.3 130

VEGETABLE OILS:
Castor Oil .96@68F 1,200 to 1,500 259.8 to 324.7 100
450 to 600 97.4 to 129.9 130

Cbina Wood Oil .943 1,425 308.5 69


580 125.5 100

Coooanut Oil .925 140 to 148 29.S to 31.6 100


76 to 80 14.69 to 15.7 130

Com Oil .924 135 28.7 130


54 8.59 212

Cotton Seed Oil .88 to.925 176 37.9 100


100 20.6 130

Linaeed Oil, Raw .925 to .939 143 30.5 100


93 18.94 130

Olive Oil .912 to .91S 200 43.2 100


115 2U 130

Palm Oil .924 221 47.8 100


125 26.4 130

Peanut Oil .920 195 42 100


112 23.4 130

RapeSeed Oil .919 250 54.1 100


145 31 130

RoeinOil .980 1,500 324.7 100


600 129.9 130

·Uu. otbenriIe DOted.


··DepeDda on oriciD or ~t and type of IOlvent.

Revised 1958. Copyright 1955 by the Hye/Tal/lic Institute.


970 Data Section

TABLE A.S (CONT.) VISCOSITY OF COMMON LIQUIDS

VISCOSITY
Liquid ·Sp Or at 60 F
SSU Centistoke AtF

Rosin (Wood) 1.09 AVI. 500 to 20,000 108.2 to 4,400 200


1,000 to 50,000 216.4 to 11 ,000 190

Sesame Oil .923 184 39.6 100


110 23 130

Soja Bean Oil .927 to.98 165 35.4 100


96 19.64 130

Turpentine .86 to .87 33 2.11 60


32.6 2.0 100

SUGAR, SYRUPS, MOLASSES, ETC.


ComSyruPII 1.4 to 1.47 5,000 to 500,000 1,100 to 110,000 100
1,500 to 60,000 324.7 to 13,200 130

GlUCOl!e 1.35 to 1.44 35,000 to 100,000 7,700 to 22,000 100


4,000 to 11,000 880 to 2,420 150

Honey (Raw) 340 73.6 100

Molasses "A" 1.40 to 1.46 1,300 to 23,000 281.1 to 5,070 100


(Firat) 700 to 8,000 151.5 to 1,760 130

Molasses "B" 1.43 to 1.48 6,400 to 60,000 1,410 to 13,200 100


(Second) 3,000 to 15,000 660 to 3,300 130

Molasses "C" 1.46 to 1.49 17,000 to 250,000 2,630 to 5,500 100


(Blaekatrap or final) 6,000 to 75,000 1,320 to 16,500 130

Sucroee SolutioDS (Sugar SyruPII):


OOBrix 1.29 230 49.7 70
92 18.7 100

62 Brix 1.30 310 67.1 70


111 23.2 100

64Brix 1.31 440 95.2 10


148 31.6 100

66Brix 1.326 650 140.7 70


195 42.0 100

68 Brill 1.338 1,000 216.4 70


275 59.5 100

70 Brill 1.35 1,650 364 70


400 86.6 100

72Brix 1.36 2,700 595 70


640 138.6 100

74 Brix 1.376 5,500 1,210 70


1,100 238 100

76 Brix 1.39 10,000 2,200 70


2.000 440 100

Revised 1958. Copyright 1955 by tile Hydralllic Institute.


Data Section 971

TABLE A.S (CONT.) VISCOSITY OF COMMON LIQUIDS

VISCOSITY
Liquid 'Sp Gr at 60 F
SSU Centistokes AtF
TARS:
Tar-Coke Oven 1.12+ 3,000 to 8,000 600 to 1,760 71
6SO to 1,400 140.7 to 308 100

Tar-Gas House 1.16 to 1.30 15,000 to 300,000 3,300 to 66,000 70


2,000 to 20,000 440 to 4,400 100
,
Road Tar:
Grade RT-2 11.07+ 200 to 300 43.2 to 64.9 122
55 to 60 8.77 to 10.22 212

Grade RT-4 1.08+ 400 to 700 86.6 to 154 122


65 to 75 11.63 to 14.28 212

Grade RT-6 1.09+ 1,000 to 2,000 216.4 to 440 122


85 to 125 16.83 to 26.2 212

Grade RT-8 1.13+ 3,000 to 8,000 660 to 1,760 122


ISO to 225 31.8 to 48.3 212

Grade RT-10 1.14+ 20,000 to 60,000 4,400 to 13,200 122


2SO to 400 53.7 to 86.6 212

Grade RT-12 1.15+ 114,000 to 456,000 25,000 to 75,000 122


500 to 800 108.2 to 173.2 212

Pine Tar 1.06 2,500 559 100


500 108.2 132

MISCELLANEOUS
Corn Starch Solutions:
22 Baume 1.18 ISO 32.1 70
130 27.5 100

24 Baume 1.20 600 129.8 70


440 91i.2 100

25 Baume 1.21 1,400 303 70


800 173.2 100

Ink-Printers 1.00 to 1.38 2,500 to 10,000 550 to 2,200 100


1,100 to 3,000 238.1 to 660 130

Tallow .918 Avg. 56 9.07 212

Milk 1.02 to 1.05 1.13 68

Varnish - Spar .9 1,425 313 68


6SO 143 100

Water - Fresh 1.0 1.13 60


.55 130
·Unless otherwise noted.

Revised 1958. Copyright 1955 by the Hydraulic Institute.


972 Data Section

TABLE A.6 VISCOSITY CONVERSION

The following table will give a comparison of various viscosity ratings so that if the viscosity is given in terms
other than Saybolt Universal, it can be translated quickly by following horizontally to the Saybolt Universal
column.

leeonde .I....tic leeonde


..,lIolt "ecoalt, ..,lIolt Seconde Seconde Degre.. Degr..e Seconde Seconde Seconde Seconde Seconde Seconde
... lver.. 1 Cantleto" Furol leelwood I ledtooocI 2 £ngler "rile, Parlin Parlin Parlin Perl in Ford Ford
eeu • alf (Standard) (Ad.1 ral t,) Cup 17 Cup '10 Cup liS Cup '20 Cup 13 Cup ,.

--- -- - -- - - -- ---
)1 1.00 29 1.00 6200
)5 2.56 )2.1 1.1E 2420 - - -
40 4.)0 36.2 5.10 1.)1 1440 - - - - -
50 7.40 - 44.) 5.8) 1.58 8)8 - - - - - -
60 10.) - 52.) 6.77 1.88 618 - - - - - -
-
-- - - - -
70 1).1 12.95 60.9 7.60 2.17 48)
80 15.7 1.3.70 69.2 8.44 2.45 404 - - - - -
90 18.2 14.44 n.6 9.)0 2.73 )48 - - - - - -
--- --
100 20.6 15.24 85.6 10.12 ).02 )07 - - - -
150
200
)2.1
4).2
19.)0
23.5
128
170
14.48
18.90
4.48
5.92
195
144
-
40
-
- -
- -- --
250 54.0 28.0 212 2).45 7.)5 114 46 - - - -
)00 65.0 32.5 254 28.0 8.79 95 52.5 15 6.0 ).0 30 20
400 87.60 41.9 ))8 )7.1 11.70 70.8 66 21 7.2 ).2 42 28
500 110.0 51.6 42) 46.2 14.60 56.4 79 25 7.8 ).4 50 )4
600 1)2 61.4 508 55.4 17.50 47.0 92 30 8.5 3.6 58 40
700 154 71.1 592 64.6 20.45 40.) 106 35 9.0 ).9 67 45
800 176 81.0 677 73.8 23.)5 )5.2 120 )9 9.8 4.1 74 50
900 198 91.0 762 8).0 26.)0 )1.) 1)5 41 10.7 4.) 82 57
1000 220 100.7 896 92.1 29.20 28.2 149 43 11.0 4.5 90 62

---
1500 »0 150 1270 1)8.2 4).80 18.7 65 15.2 6.) 1)2 90
2000 440 200 1690 184.2 58.40 14.1 86 19.5 7.5 172 118
2500 550 250 2120 230 7).0 11.3 108 24 9 218 147
3000 660 300 2540 276 87.60 9.4 - 129 28.5 11 258 172

---
4000 880 400 »80 368 117.0 7.05 172 37 14 »7 230
5000 1100 500 4230 461 146 5.64 215 47 It1 425 290
-
6000 1)20 600 5080 55) 175 4.70 258 57 22 520 )50
7000 1540 700 5920 645 204.5 4·03 300 67 25 600 410
8000 1760 800 6770 737 233.5 ).52 - 344 76 29 680
--
465
9000 19110 900 7620 829 26) ).13 )87 86 32 780 520
10000 2200 1000 8460 921 292 2.82 4)0 96 )5 850 515

-- --
15000 3300 1.3700 4)8 2.50 650 147 5)
~
1280 860
20000 4400 18400 584 1.40 860 203 70 1715 1150

• K' . V· • (. . k) Absolute viscosity (in centipoi_)


mematic ISCOSlty m centisto es = S'fi G .
peel C ravlty

Above 250 SSU, use the following approximate conversion:


SSU = Centistokes x 4.62

Above the range of this table and within the range of the viscosimeter, multiply their rating by the following factors
to convert to SSU:

Viscosimeter
Saybolt Furol
---
Factor
10.
Viscosimeter
Parlin cup lIS
Factor
98.2
Redwood Standard 1.095 Parlin cup 120 187.0
Redwood .Admiralty 10.87 Ford cup 14 17.4
Engler. Degrees 34.5

Copyright 1955 by the Hydraulic Institute.


Data Section 973

TABLE A.6 (CONT.) VISCOSITY CONVERSION

The following table will give a comparison of various vil!Cosity ratings so that if the vil!Cosity is given in terms
other than 5aybolt Universal, it can be translated quickly by following horizontally to the Saybolt Universal
column.

SecOlld, Approl. Second,


Approl. Approl.
lin ..atic Second, Second, Second, Second, Second, Second, Second, Second. Pratt
Sa,bolt Second, Sardner
Uni". ..al
Vi.co.i t,
MIIc Kolt
Zahn Zahn Zahn Zahn Zahn o.•• ler D_ler Sto... r and
Cent i ,totee.· Cu, n Cup II Cup " Cu, '5 Cup '1 Cu, flO 100 , . La.bert
"u Michael lubbl. Cu, 'I
Loed "F"

31 1.00 - - - - - - - - - - -
35 2.56 - - - - - - - - - -- --
40 4.30 -- - - - - - - 1.3 -
50 7.40 - - - - - - 2.3 - 2.6 -
10.3 - - - - - - - - -
-
60 3.2 3.6
70 13.1 - - - - -- - - 4.1 - 4.6
8l 15.7 - - - - - - 4.9 - 5.5 -
90 18.2 - - - - - - - 5.7 - 6.4 -
100 20.6 125 - 38 18 - -
-
- 6.5 - 7.3 -
--
150 32.1 145 -A 47 20 - 10.0 1.0 11.3 -
ax> 43.2 165 54 23 - - 13.5 1.4 15.2 -
250 54.0 198 A 62 26 - - - 16.9 I."' 19 -
300 65.0 225 8 73 29 - -- - 20.4 2.0 23 -?
400 87.0 270 C 90 37 -- - 27.4 2.7 31
500 110.0 320 0 - 46 - -- 34.5 3.5 39 8
600 132 370 F - 55 - - 41 4.1 46 9
700 154 420 G -- 63 22.5 - - 48 4.8 54 9.5
800 176 470 -H 72 24.5 -18 - 55 5.5 62 10.8
900 198 515 - 8l 27 -13 62 6.2 70 11.9
1000 220 570 I - 88 29 20 69 6.9 77 12.4
1500 330 805 M - - 40 28 18 103 10.3 116 16.8
2000 440 1070 Q - - 51 34 24 137 13.7 154 22
2500 550 1325 T - - 63
" 29 172 17.2 193 27.6
3000 660 1690 U - - 75 48 33 206 20.6 232 33.7
4000 88D 2110 V - - - 63 43 275 27.5 308 45
w - - -
--
5000 1100 2635 77 50 344 34.4 385 55.8
6000 1320 3145 X - - - 65 413 41.3 462 65.5
7000 1540 3670 - - - - 75 481 48 540 17
8lOO 1760 4170 Y - - - - 86 550 55 618 89
9000 1981 4700 -Z - - - - 96 620 62 695 102
10000 220D 5220 - - - - - 690 69 770 113
15000 3300 7720 Z2 - - - - - 1030 103 1160 172
20000 4400 10500 Z3 - - - - - 1370 137 1540 234

* K'lDematic
. V' . C' • k)
l8C08lty ID centisto es = Absolute S
viscOlity (in centipoisea)
" c Gravi'ty
peel

Above 250 SSU, ll8C the following approximate conversion:


SSU = Centistoke! x 4.62
Above the range of this table and within the range of the viscosimetcr, multiply their rating by the following factors
to convert to 5SU:
Viscosimeter Factor
Mac Michael 1.92 (approx.)
Demmler*l 14.6
Demmler *10 146.
Stormer 13. (approL)

Copyright 1955 by the Hydraulic Institute.


974 Data Section

TABLE A.7 THEORETICAL DISCHARGE OF NOZZLES, IN US GALLONS PER MINUTE

Velocity
Head oC dill- DUMl.'TER OF NOZZLE IN INCHES
eltarJ!e.
r...t per
Lbe. Feet
-- ---
second 'Ai
!1t )-0
-- -- -- -- -- - - --
~ 7.i
H ~ ~'i Ji .,4 2 27.i 1 1)-0 1~ 1~ 1~ 3 3_1'2 4 ,
4', '
-- -- -- -- - - - - - - -
2~ 2~ .\ 6
-
.I)':l
;--
10 23,1 38,68 0.37 1.43 3.30 5.110 13.2 23.6 3&.8 53,2 72.2 91..4 119 146 178 212 28tI
- - - - - ---- --
373 478 5110 715 850 1160 1512 1912 2360 2800 3400
15 34,7 47,25 0.45 1.31 4.02 7,23 1&.2 23.7 45.0 85.1 83.4 118 14& 181 218 260 354 4&11 688 723 880 1040 1420 1852 2844 2800 3520 4160
20 46,2 64.55 0.52 2.09 4.88 8,35 18.7 33.4 52.0 75,3 102 134 189 209 252 300 409 5S4 678 835 1013 1200 1040 213& 2700 3340 4072 4600
25 57.8 60,99 0,68 2.33 5.23 9.33 20,9 37.2 68.2 84.1 114 149 139 233 282 338 457 597 768 933 1138 1250 1828 2390 3024 3730 4682 .\400
30 69,3 88.52 0.84 2.68 5.71 10.2 22.8 40.9 89.7 92,2 125 18i 207 258 309 388 501 854 828 1022 1240 1480 2000 2816 3312 4OS8 4960 5920
35 80,9 72.16 0,89 2,78 6.1& 11,0 24.7 44,2 69,8 99.6 135 177 223 276 334 397 641 707 895 1104 1340 15110 2160 2828 3560 4416 5360 6360
40 92.4 77.14 0,74 2,95 6,60 11.8 28.4 47.2 73,6 106 144 189 239 295 357 425 578 755 95& 1180 1430 1700 2320 8020 3824 4720 5720 6800
45 104.0 81,83 0,78 3,13 6,99 12.5 23.0 50.2 78.1 113 153 300 253 313 378 450 613 801 1014 1252 1520 1800 2450 3200 4056 5000 6060 7200
50 115,5 86,26 0.82 3,30 1,37 13,2 29,5 52.8 82,3 119 161 211 287 330 399 475 846 845 1069 1820 1800 1900 2580 3388 4276 5230 8400 7801
65 127.1 110,46 0.86 3,46 7.73 13,8 30,9 55.4 86,3 125 169 221 230 346 418 498 673 866 1122 1385 1680 2000 2720 3544 4486 5540 6720 8000
60 138.6 94.49 0,110 3,62 8.08 14.5 32,3 57.8 110,1 130 177 231 293 362 437 520 708 925 1171 1448 1755 2080 2840 3700 4684 5784 7020 8320
65 150,2 98.35 0.94 3.77 8.40 15.1 33.6 60.2 93,8 136 184 241 305 377 455 642 737 963 1219 1506 1830 2185 2950 3850 4676 6024 7320 8680
70 161.7 102,06 0.97 3,91 8,73 15.6 34.9 62.5 97.4 141 191 250 316 391 472 562 785 999 1285 1681 1895 2250 3060 4000 5060 6244 7680 9000
75 173.3 105.65 1.01 4.04 9.03 16,2 36.1 84,6 101 146 198 259 ' 327 404 486 682 792 1034 1309 1618 1960 2330 3170 4136 523& 8484 7840 9320
80 184.8 109.11 1.04 4.18 9.33 16,7 37.8 86,6 104 150 201 287 333 418 504 601 81810&8 1352 1889 2030 2404 3230 4272 8400 8876 8120 9616
85 196,4 112.46 1.07 4,31 9.62 17.2 33.5 68.8 107 155 210 275 348 431 520 &20 843 1101 1391. 1720 2080 2480 3380 4400 5676 6880 8320 9920
110 207.9 115,72 1,10 4.43 9.89 17.7 39,6 70.8 110 160 217 233 358 443 535 837 8871133 1434 1770 2150 2540 3470 4532 5736 7080 8600 10160
95 219.5 118,89 1,13 4.56 10,2 18,2 40,7 72,3 113 184 223 291 368 456 MO 655 8911184 1474 1820 2200 2&20 3560 41156 6896 7280 8600 10480
100 231.1 121. 98 1,16 4,67 10.4 18,7 41,7 74.6 116 168 228 299 378 4&7 584 672 9141191. 1512 1888 2260 2700 3650 4776 6046 7484 IIOlO 10800
105 242.6 125,00 1.19 4.78 10,7 19.1 42,8 76.5 119 172 234 306 337 473 573 688 927 l224 1649 1912 2320 2760 3750 4696 8200 7846 9280 11040
110 254.2 127.91. 1,22 4.110 10.9 19.6 43.8 78.3 122 177 239 313 39& 4110 691 705 959m3 1688 1957 2380 2820 3840 5012 6844 7828 9520 11230
115 285.7 130.82 1.25 5.01 11,2 20,0 44.8 80.1 125 181 245 320 405 501 605 720 980 1281 1621 2002 2430 2380 3920 5134 8484 8008 9720 11520
120 277.3 133.89 1.27 5.12 11.4 20.4 45.7 81,8 127 184 250 327 414 512 818 736 1001 1306 1656 2044 3460 2950 4004 5232 8834 8176 9920 11800
125 238.8 136,38 1,30 5,22 IU 20.9 48,7 83.5 130 188 255 334 422 522 830 751 1022 1335 1891 2088 2540 3000 4100 5340 6784 6844 10160 12000
130 800.4 139.03 1,33 5.32 11.9 21.3 47.6 85.1 133 192 260 341 431 532 843 7&& 1042 1362 1724 2128 2580 3070 4160 5446 6896 8512 10320 12230

The actual quantity dilcbarpd by • noule will be Ie. thaD above table. A well tapered IIDOOIh DOllie may be uoumed to Ii...bove 94 per <en! .r the values in the tables.
Index

A Automatic vent valves, 701-702


Automatically primed units, 704-707
Absolute pressure, defined, 404-405
Axial-face seals (see Mechanical seals)
Abrasion, designs to minimize, 76-77, 335-336, 782
Axial-flow impellers, 7, 64
types of, 391-392 (see also Erosion)
Axial flow pumps, 9 (see also Propeller pumps)
Acid-resisting pump, 383
Axial shuttling of rotor, causes of, 577
Aeration, effect on corrosion, 376, 385, 388
Axial thrust, balancing of, 91-115
Affinity laws, defined, 428-430
of double suction impellers, 91
application of, 430-436
liquid acceleration, effect on, 92
Air-conditioning, pumps for, 785
liquid momentum, effect on, 92
Air entrainment, effect on pump characteristics, 469
of mixed and axial flow impellers, 96-97
Air leakage into pumps, 141, 174,732,936-937
in multistage pumps, 97-103
Alignment, equipment, 875-878
of overhung impellers, 94-96
cold,877
residual, 108, 114-115
effect of piping loads on, 40, 878
in single stage pumps, 91-97
at factory, 300, 873-874
of single suction impellers, 91-97, 111-113
hot check, 877-878
Axially split casings, 8, 28, 34-35, 44-55
method of, 875-878
distortion of under pressure, 39, 45-49
periodic check of, 917
flanges, split-joint, 39, 42, 46-49
tolerances for, 878
gasket, split-joint, 39, 42, 46-49
All-bronze fitted pump, 382
interstage crossovers, form of, 51
All-iron fitted pump, 382
maintenance of, 42-43
Alloy 20®, 383
materials for, 54
Altitude, effect on atmospheric pressure, 405
multistage, design of, 51-55
effect on electric motor cooling, 639
piping loads, effect on, 54
Ambient conditions, effect on motors of, 637-639
pressure limit of, 46
American Petroleum Institute Standard API-61O, 46, 775
support, location of, 55
Antiflash orifice, 867
wearing rings, mounting in, 80, 82
Antifriction bearings (see Bearings, antifriction)
Ash-sluice pumps, 718
Atmospheric pressure, variation of with altitude, 405
B
Automatic and manual controls, 606
Automatic priming system, 695-696 Back-pull-out pump, 34
Automatic recirculation (bypass) control, 867 Balanced mechanical seals, 171

975
976 Index

Balancing axial thrust, 91-115 diagnosis of operating problems, 944-945


Balancing device leak-off, point of return, 105 future designs, 951-952
monitoring flow through, 907 loads on, 211, 213
piping for, 867-869 effect of rotor geometry, 213, 826
Balancing disk, 108-109 magnetic, 261-264
Balancing drum, single diameter, 105-108 plain, 236-261
stepped, 11 0 journal, oil lubricated, 239-245
Balancing, dynamic, standard for, 922 materials, 247
Balancing radial thrust, means of, 23 maintenance, 260-261
Ball bearings (see Bearings; antifriction) mode of lubrication, 237
Barrel type pumps, 55-59 (see also Radially split double monitoring lubricant supply to, 900
casing pumps) oil lubrication of, 248-251
Barske impeller; see Partial emission pumps oil whirl in, 239
Baseplates, 298-310 product lubricated, 251-259
bending stiffness of, 300, 301 sleeve (see journal)
centerline support type, 306-307 thrust, 243-247
design criteria for, 300-301 tilting pad journal, 242
function of, 298 temperature of, monitoring, 90 1
grouting, provisions for, 304 types, 213
jacking screws for, 306 Bearing bracket, 35
leakage collection, designs for, 304 design requirements, 264-265
leveling, method of, 874-875 Bearing housings, 264-271
need for, 300, 874 adjusting radial position of, 232
provision for, 306 vibration of, monitoring, 903
lifting, provisions for, 305 Bearing span, 123, 125
machining of, 305-306 Bellows seal, 176, 178
materials for, 301 Belt drive, 626-627
piping loads, effect of, 300, 307, 855-859 Boiler circulating pumps, 96, 745-748
stress-relief of, 306 Boiler feed pumps, 720-728
supported type, 300 capacity of, 720-722
suspended type, 300-301 condensate injection sealing of, 195
spring mounted, 301, 873 data for inquiry, 814
three-point support of, 301, 303 materials for, 396
torsional stiffness of, 300 minimum flow bypass for, 583, 862-867
vertical pumps, designs for, 309-310 monitoring of, 895-8%
welding of, 305 NPSH available to, 722-723
Bearings, 211-271 regulation of, 609-614
antifriction, 214-236 standby, 722
angular contact, 215-218 total head required, 723
bal1,214 variable speed operation of, 676-677, 682
condition of, 234 warmup of, 861-862, 882-885, 900-901
cylindrical roller, 218 Booster pumps, 543, 668, 743, 772
deep-groove ball (Conrad), 215 Borehole pumps (see Vertical turbine purnps)
double-row ball, 215 Bottom suction, 9, 37
double-row angular contact, 218 Bowl assembly, 324-329
life of, 222 Brake horsepower; see Power
lubrication of, 222-229 Breakdown seals, 195-210
maintenance of, 230-236 condensate injection sealing of, 195
monitoring "noise" of, 904 closed cycle, effect on, 206-208
roller, 214 controls for, 201-204
self-aligning ball, 215 drains from, 199-201
spherical roller, 218-220 open cycle, effect on, 204-206
tapered roller, 220-221 supply for, 196-199
types of, 214 floating ring type, 210
arrangements of, 212 justification for, 195
capacity, evaluation of, 826 other applications of, 208
design considerations for, 211 Bronze pumps, 382
Index 977

Building record, pump assembly, 922 Classification of centrifugal pumps, 7-10


Bypass, minimum flow, 581-583, 602 by axis of rotation, 8
design of, 602, 862-867 by casing types, 6, 9
by drive arrangement, 9
c by impeller types, 6
by nozzle location, 8
Can pumps, 329 by rotor construction, 9
Canned-rotor motors, 673 by specific speed, 64, 438-439
Capacity, conversion factors for, 402-403 Clearances of wearing rings, 85-86
units of, 40 I Close-coupled pumps, 6, 272, 643
Capacity regulation, 492-494, 532-544 Closed impellers, defined, 67
Cargo oil pumps, 793-795 Clutch, centrifugal, 290-291, 293
Casing, construction for open impeller, 35 overrunning, 293
distortion by piping loads, 40, 54 Coatings for abrasion and corrosion resistance, 77,
double, defined, 55 380-381
forged, justification for, 376 Cold spring, of piping, 880
functions of, 18 Collector, pump classification by, 7
hand-holes, 39 function of, 18
inner, 55 diffuser, 7, 19-20,24-28,44
design comparison, 56-57 types, 7, 18
maintenance of, 59 volute, 7, 18-28,44
materials for, 377 Column pipe, 324
lined, 376-377,767 Combined balancing disk and drum, 109-110
maintenance of, 42-43, 59-61 Compression couplings, 277-278
materials for, 54, 368-377 Concrete volute pumps, 319-323, 742
maximum allowable working pressure, 55 Condensate injection sealing (see Breakdown seals)
mechanical features of, 8, 39-40 Condensate pumps, 313, 728-734, 815
mounting, 39-40 dry pit type, 313, 731
nozzle locations, 36-38 materials for, 733
multistage pumps, 12,44-61 "pass-out" type, 330
nozzles, locations of, 36 regulation of, 733-734
safe working pressure, estimate of, 459 shaft seal for, 732
segmental (see Radially split multistage "ring casing") submergence control of, 492-494
"solid", 28 vertical can pump for, 329, 339, 731
split-joint, choice of, 28-30 Condensate return units, 787-788
temperature of, monitoring, 900-901 Condenser circulating pumps, 523-527, 737-744
Casings, for multistage pumps, 44-61 capacity required, 737
for single-stage pumps, 18-43 configurations used, 737
(see also Axially split casings, Radially split casings) data for inquiry, 815
Caustic liquors, materials for, 379 materials for, 742
Cavitation, effects of, 84, 392, 473, 565 suction conditions for, 737-738
defined,473 Constant-pressure governors, 613
Cellar drainers, 339, 608 Constant-pressure regulators, 604-605
Centerline support, justification for, 39-40 (see also Constant-speed governors, 622
Baseplates) Construction pumps, 782
Central automatic priming system, 695-700 Control elements, 601-605
Central priming systems, 694 Control functions, 595-601
Characteristic curves, 423-428, 439 Control and operating medium, 605
Charge pumps, hot oil, 777-778 Controls, typical applications of, 609-617
Check valves, uses of, 598, 614-615, 686-687 Control valves, 617-620
losses in, 517 (see also Corrective controls; Pressure controls; Protective
Chemical pumps, 764-769, 815-816 controls)
Chloride, effect on corrosion, 376 Conversion tables, 959, 960
Circulating pumps (see Boiler circulating pumps, Circulator Conversions, of capacity units, 40 1-402
service, Condenser circulating pumps) of head units, 402-404
Circulator service, 313, 378, 769 of pressure units, 402-404
Clamp couplings, 273-274 of viscosity units, 464
978 Index

Cooling water pumps (see Condenser circulating pumps) Cracks, in casing, 54, 59
Corrective controls, 602, 609-614 in impeller, 84
Corrosion, 42, 55, 83, 386-391, 945 in shaft, 130
pH, influence of, 387 Criticality of service, defined, 894
chemical attack, of polymer, 391 Critical speed, defined, 116 (see also Rotor, critical speed)
concentration cell, 389 Curves (see Characteristic curves; Head-capacity curves; Ef-
corrosion-erosion, 389 ficiency curves; Power curves; Rating curves; System-
corrosion-fatigue, 390 head curves)
crevice, 389 Cut down of impellers, 432-436
endurance strength, effect of on, 381 Cut-water (see Volute tongue)
fretting, 390 Cyclone separator, 144, 189,259
galvanic, 385, 388
general, 388
graphitization, 42, 388
D
intergranular, 390 Damage, pump, assessment of, 920
limiting rate of, 386 Damping within running clearances, effect on rotor dynamics,
microbiological, 390 119-120, 124
nature in pumps, 386 Decoking jet pumps, 778
pitting, 389 Deep-well pumps (see Vertical turbine pumps)
selective leaching, 388 Deflection of shaft (see Shaft deflection)
stress corrosion cracking, 389 Descaling pumps, 782-783
stress cracking, of polymer, 391 Design constants, 436-461
Coupling alignment (see Alignment, equipment) Dewatered, mode of starting pump, 321
Couplings, 272-297 Dewatering pumps, 543-544, 780-782
all-metal flexible, 281-289 Differential pressure governor, 613
cut-out (see Couplings, all-metal, disconnect) Differential pressure regulator, 610-612
diaphragm, 283 Diffuser, 7, 19-20,24-28,44
disconnect, 292 efficiency compared to volute, 26
disk,283 radial forces in, 24
extension (see spacer) stall in, 25
floating shaft, 285 within single volute, 26
gear, 285 Diffusion, defined, 4, 18
limited-end float, 285-289 Diffuser pump, 7, 18, 324, 778
spacer, 285-289 Diaphragm, 51 (see also Stage piece)
spring-grid,285 Discharge, above and below ground, 9, 324
clutch type, 289 Discharge head, 419-421
elastomer flexible, 278-281 Discharge nozzle flanges, design of, 54-55
block,280 Discharge nozzle, location of, 36-38
pin and buffer, 279 Discharge, of nozzles, theoretical, 974
ring, 280 Discharge recirculation (see Internal recirculation at impeller
sleeve, 281 discharge)
flexible, 272, 278-289 Discharge piping, design of, 854-855
misalignment capacity of, 272 Dismantling, complete pump, 919-920
need for, 272 Displacement pump, comparison with centrifugal, 3
selection of, 272, 278 Dissolved air or gas, effect on pump performance, 485, 494
flexible drive shaft, 289 Double casing pumps, 55-59, 883-885
for dual-drive, 292-293 Double mechanical seals (see Mechanical seals, double)
hubs, mounting of, 294-296 Double suction impeller, 7, 63, 91
magnetic, 297 NPSH required by, 481, 497
maintenance of, 297 Double volute, 23-24
rigid, 272, 273-278 Double wearing rings, 72
clamp, 273-274 Dowelling of pumps and drivers, 873, 880-881
compression, 277-278 Drainage pumps, 761-762
flanged, 274 Drawing approval, 827-828
sliding sleeve, 274 Drive ammgements, 623-629
tapered sleeve, 274 Drivers, 621-685
threaded,273 factors affecting size of, 622-623
Coverage, centrifugal pumps versus displacement, 6 Drum, balancing (see Balancing drum)
Index 979

Dry operation of a pump, 321, 728, 889-890 Exit losses (see Friction losses)
means of preventing, 714 Expansion joints, installation design with, 856-859
Dry-pit pumps, 9, 311-323, 738-740 Expeller (see Hydrodynamic seal)
Dual drive, 625
Dynamic balancing, standard for, 922
F
Feedwater cycles, 717-719
E Feedwater level regulators, 609-614
Efficiency, pump, 428 Fire pumps, 762-764
estimating from specific speed, 439-452 Firewall extensions, 870
impeller mounting, effect on, 63 Fittings, pump construction, 382-385
prerotation, effect on, 32 piping, friction losses in, 517-522
significance of in energy consumption, 545 Fibre-reinforced polymer, 376, 379, 767
suction nozzle form, effect on, 31 Finite element analysis (PEA), 45
Efficiency curves, 424, 439 Fixed packing, 157-158
Ejectors, 690 Flat elbow type suction nozzle, 31, 37
Elastic seal rings, 53 Flanges, casing split-joint, 39, 42, 46-49
Elbow type suction nozzle, 31 raised face use of, 54-55
Elbows, losses in (see Friction losses) suction and discharge nozzles, 54-55
orientation at pump suction, 844-850 Flexible couplings (see Couplings)
Electric motors, 621, 622, 629-645 Flexible pipe connectors, use of, 873
a-c, 621, 632-643 effect on baseplate design, 307
polyphase, 633 Flexible shaft, defined, 117
single-phase, 632 Flexible drive shaft, 290
squirrel-cage, 632-635 Float controls, 598-600, 733-734
synchronous, 636-637 Float switch, 596, 602
wound-rotor, 624, 635-636 Floating shaft couplings (see Couplings, all-metal flexible,
d-c, 621, 624, 632 floating shaft)
electrical characteristics of, 644-645 Flow control, 598-600
enclosures for, 639-641 Flow, margin, 548, 720
heat dissipation from, 637-639 measuring of, 899, 907-910
insulation, classes of, 639 minimum, operating, 812-818
mechanical features of, 643 permissible, 560-585
selection of, 629-645 normal, 812-818
starters for, 645-649 rated, 805, 812-818
submersible, 334, 667-673 Flow regulation, defined, 3
temperature rise in, 637 Flue gas desulphurization, systems for, 749
Electrical systems, 629-631 pumps used in, 750
Element, inner, defined, 51, 57 Fluid drives (see Hydraulic couplings)
dismantling of, 920 Flushing before plant startup, 852, 885
installation of, 59 Food and beverage pumps, 361, 780
reassembly of, 922 Foot-supported pumps, 40
removal of, 59 Foot valves, 686, 688
End-float travel, 284-287 losses in, 518
End suction, 31 Foundation bolts, 835-836
End suction pumps, 9, 32 Foundations for pumps, 835-836
Endurance strength, effect of corrosion on, 381 Frame mounted pumps, 34, 39
Energy consumption, savings in, 547 Francis-vane impellers, 64, 454
evaluation of, 823 Friction losses, 506-522
Entrained air or gas, effect on pump performance, 469, 485, in elbows, 517, 521
489 entrance, 409-410, 518
Entrance losses (see Friction losses) exit, 409-410
Environmental and safety requirements, 808-809 in fittings, 517-522
Equipment record system, 917 in piping, 506-516
Erosion, 42, 55, 391-392 Darcy Weisbach equation for, 512-516
Euler's equation, 3 Williams and Hazen data for, 507-511
Excess-pressure governors, 613 in valves, 517-522
Excess-pressure regulators, 610-612 for viscous liquids, 962-965
980 Index

G Horsepower, brake (see Power)


Hot oil pumps (see Charge pumps, hot oil)
Gage pressures, 404-405
Hot water circulating pumps, 785-786
Galvanic corrosion (see Corrosion, galvanic)
Hydraulic balancing devices, 97, 104-1l5 (see also
Gas turbine drive, 621, 623-624
Balancing disk, Balancing drum, Combined balancing
Gasket, compensator, 59
disk and drum)
split-joint, 39, 42, 46--49
materials for, 381
Gasoline engines (see Internal combustion engines)
modifications to, II 0-1l1
Gear drive, 627-629
residual thrust with, 108, 1l4-1l5
Glass, corrosion resistant linings of, 376
Hydraulic couplings, 599, 613, 624, 673-677
Governors, constant pressure, 613, 623
Hydraulic fit, evaluation of, 824
constant speed, 622
Hydraulic Institute Standards, friction losses for viscous
differential pressure, 613
liquids, 962-965
excess pressure, 6 I 3
recommended NPSH, 481
overspeed, 597, 626
viscosity of common liquids, 966-971
Graphite, pumps constructed from, 376
viscosity conversion, 972-973
Graphitization, 385
viscosity corrections for pump performance, 464--469
Grouting, base, 878-879
Hydraulic turbine, 621, 626, 799
base design for, 304
Hydraulic power recovery turbine (HPRT), 626
cementitious, 879
Hydraulic presses, pumps for, 783, 817-818
epoxy, 879
Hydrocarbon processing, pumps for, 774-779
foundation bolts, 836
Hydrodynamic bearing, 214, 237
Hydrodynamic seal, 158-160
auxiliary seal for, 160
H
expeller for, 159
Hand primers, 694 power absorbed by, 160
Hard coating, of running clearance surfaces, 77, 83, 380-381 suction pressure limit of, 159
Hazardous areas, classification of, 642-643 temperature rise in, 160
Head, discharge, 419--421 tolerance of solids, 159
friction, 409, 412 Hydrogen sulfide, stress corrosion cracking caused by, 389
pump, defined, 402 Hydrostatic bearing, 1l4-1l5, 239, 256
static, 409 Hydrostatic pressure tests, 55
suction, 414--415 Hydrostatic transmission, 679-681
system, 412 Hydroviscous drive, 677-678
total,421--422
units of, 402
velocity, 410
I
Head-capacity curves, 3, 19-20, 423--425, 533-538 Impeller diameter change, effect of, 432--436
discontinuity in, cause of, 577 calculation of, 435--436
drooping, 20, 424, 533, 534, 537 influence of collector type on, 20
flat, 424, 533 Impellers, 62-72
rising, 424, 538 action of, 3, 62
stable, 425, 534, 538 axial flow, 7, 64, 96-97, 451--452
steep, 424, 533 balancing of, 84
unstable, 425 closed, 67, 329 (see also enclosed)
Head and pressure conversion, 402, 406 cut down of, 432--436
Head terms, 408--412 double-suction, 7, 63, 91
outmoded,422 enclosed (see Impeller, closed)
for vertical wet-pit pumps, 422 life, estimates of, 504-505
Heads,401--422 fabricated, 71
Heater drain pumps, 735-737 Francis-vane, 64, 454
Heating systems, 785-786 high energy, defined, 68
Hermetically sealed drives, 683--685 induced vortex, 68
Hermetically sealed pumps, 136--137, 31l (see also large eye design, 499
Sealless pumps) maintenance of, 83
motors for, 669--673 manufacture of, 71-72
use in chemical industry, 673, 766 materials for, 368, 377-379, 396
Horizontally split casing (see Axially split casing) mixed-flow, 7, 64, 96-97, 454
Index 981

mounting of on shaft, 127-129 erosion produced by, 578


nonclogging, 68 head-capacity characteristic, effect on, 577
one-piece, cast, 71 rotor axial shuttling caused by, 577-578
open, 64 shroud bowing or fracture caused by, 578
overhung, 62 at impeller suction, 565-576
plain vane, 64 calculation of onset, 568, 589
propeller, 7, 64 cause of, 565
radial-flow, 7 effects of operation with, 568, 572
rubber-lined, 379 erosion produced by, 565
screw-vane (see Francis vane) modification to lower onset of, 575
semiopen, 64-67, 329 pump size, effect of, 574
shaft-through-eye, 62 testing for, means of, 569
single suction, 7, 63, 91-98,111-113,481 liquid characteristics, effect on, 571
for slurry service, 68 Iron, cast, 368, 376, 378, 381
split and staggered vanes, 70 chrome, 377, 379
type, selection of, 72 high silicon, 383
vane tip clearance with collector vane(s), 68-70, 579 Irrigation pumps, 761-762
In-line impellers (see Tandem impellers)
Inboard end, defined, 39, 212
Induced-vortex pump, 360-361 J
Inducer, 70, 499
Joints (see Expansion joints)
Industry standards, use of, 809
Inner casing (see Casing, inner)
Inner element (see Element, inner)
Inquiries; preparation of, 801-818
K
Installation, 833-881 Kinetic pumps, defined, 3, 9
design of, 834-873 Kingsbury type thrust bearings, 245
balancing device leak-off piping, 867-869
bypass system, 862-867
discharge piping, 854-855 L
expansion joints, 856-859
Leakage, allowable, influence on pump design, 136-137
firewall extensions, 870
L-type wearing rings, 73
foundations, 835-836
Level control, 596, 598, 599, 600
instrumentation, 869-870
Limitations of space and weight, 810
noise, control of, 870-873
Lineshafting, hollow, 317
piping loads, 855-856
critical speed, required separation margin, 319
provision for tolerances, 834
bearings for, 319
pump location, 834-835
bearing supports for, 3 I 9
suction strainers, 850-854
solid, 317, 319
suction system, 836-850
Liquid, nature of at shaft seal, 137
vent and drain piping, 859-860
Liquid end, 6
warmup piping, 861-862
Liquid level control (see Level control)
installing equipment, 873-881
Liquid, nature of, 804-805
alignment, method of, 875-878
Load factor, 807
dowelling of pump and driver, 880-881
and number of pumps, 803-804
grouting, 878-879
Location; pump (see Pump location)
leveling on foundation, 874-875
Lomakin effect, defined, I 19
piping fit-up, 880
effect of on rotor dynamics, 120
storage in field, 874
form of running clearances, influence on, 119
Instruction books, 833-834
running speed, influence on, 119
Instrumentation, selection of, 869-870
Lubrication, bearings, 222-229, 248-251, 259-260
Intake channels, 341-346
Internal combustion engine, 621, 623, 624
Internal-combustion-engine-driven pumps, priming of,
708-710
M
Internal recirculation, 564-580 Magnetic bearings (see Bearings, magnetic)
at impeller discharge, 576-580 Magnetic clutches, 297
calculation of onset, 587-588 Magnetic drives, 599, 613, 624, 678-679, 683-685
982 Index

Maintenance, 916-926 for impellers, 377-379


of bearings, 230-236, 260-261 for inner casings, 377
of casings, 42-43, 59-61 leaded gunmetal, 376
dismantling pump, 919-920 low carbon steel, 381
equipment record system for, 917 metals commonly used, 371-375
cost, evaluation of, 823-826 Monel·, 382
of impellers, 83-84 nickel aluminum bronze, 376
inspection of parts, 920 Nihard*, 377, 379
overhaul,918-924 Ni-resist, 381
on condition, 918-919 phosphor bronze, 376
on failure, 918 polymers, injection molded, 368, 378
periodic, 918 polymers, reinforced, 376, 379, 767
of mechanical seals, 194 properties affecting selection of, 367, 369
of packed box seals, 156-157 rubber, 376, 379
preventive, 917-918 temperature limit of, 379
program for, 916-917 head limit for impellers, 379
reassembling pump, 922-924 service conditions affecting selection of, 367
building record for, 922 for shafts, 381-382
restoration of parts, 921 surface hardening, 380
of shafts, 129-130 surface conversion, 380
of shaft sleeves, 134 Teflon*lining, 376, 767
of wearing rings, 84-90 titanium, 396
Marine pumps, 788-798 for wearing parts, 379-381
Materials, 367-398, 950 zirconium, 396
Alloy 20*, 383 Mechanical seals, 161-194
alloy steel, 381 angular misalignment, effect of, 179
aluminum, 376 arrangements of, 171-174
aluminum bronze, 378 double, 172, 189
aluminum manganese bronze, 378 external, 171
austenitic manganese steel, 377 internal, 171
austenitic stainless steel, 376, 379, 380, 381, 385 staged,l72
effect of tensile properties on casing design, 377 single, 172
temperature limit in high head multistage pumps, 379 tandem, 172
bronze, 376, 378 auxiliary seals for, 184
speed limit for impellers, 378 axial rotor movement, effect of, 174
temperature limit for impellers, 378 balance of, hydraulic, 169-171
carbon steel, 368 barrier fluid for, 172, 191
for casings, 368-377 bellows type, 176, 178
cast iron, 368, 376, 378, 381 breakdown bushing for, 189
risk of premature cavitation erosion, 378 buffer fluid for, 172, 191
speed limit for impellers, 378 capabilities of 161, 165, 178
temperature limit for casings, 368 classification of, 176
temperature limit for impellers, 378 compression unit, 176-179
ceramic, 376 cyclone separator, use with, 189
chrome steel, 368, 378, 380, 381 dynamic gasket, 176
sour service, modification for, 379 environment for, 185
classification by major components, 368, 382 face loading in, 163, 165
ductile iron, 368, 376 heat generated by, 163
duplex alloys, 377, 379, 380, 382 housing for. 174, 175
embrittlement, temperature limit to avoid, 377, 379 hydrostatically loaded, 179
glass lining, 376 hydrostatic type, 168
graphite liquid end parts, 376 interface, lubrication of, 163, 165, 167, 168
hard coating, for galling resistance, 380 pressure drop across, 163, 165
for wear resistance, 380-381 vaporization across, 165
high chrome iron, 377, 379 justification for, 161-162
high manganese austenitic stainless steel, 379 leakage from 161, 163
higher alloys, 377, 379, 382 liquid boiling point, margin below, 165, 185
Index 983

liquid flow to, 179, 181, 185 margin over required, 502-504
maintenance of, 194 means of raising, 495-499, 747
materials for, 183-184 NPSH required, 473
magnetically loaded, 179 characteristic of, 475
mounting of, 179 by cryogenic pumps, 489-492
operation of, 192-194 definition of, 473, 499-505
pressure at, raising, 185, 187 hydrocarbon correction for, 488-489
lowering, 187 impeller diameter, effect on, 480
principles of, 162-168 impeller mounting, effect on, 62
pump condition, effect on, 194 liquid nature, effect on, 483-492
pump design, effect on, 174-175 means of lowering, 495-499
pusher type, 176 prerotation, effect on, 32
quench fluid for, 192 suction nozzle form, effect on, 31
rotor deflection at, effect of, 174 tests to determine, 483, 485, 500
secondary sealing of, 176 (see also dynamic gasket) Nash vacuum pump, 690-692
slurry, designs for, 189 Natural frequency(ies), 116, 117, 120,319
temperature at, lowering, 185 Nature of liquid pumped, 804-805
throat bushing for, 181, 189 Nil-ductility transition temperature, 376
vapor pressure, margin over, 165 Noise control, 870-873
Mine dewatering pumps, 543, 780, 817 Noise, of motors, 640-641, 872
Minimum operating flow, 560-585 of pumps, 71, 870-873, 931-932
bypass for (see Bypass, minimum flow) Nomenclature, of bearings, 214-215, 239, 243, 261-262
continuous versus intermittent operation, 583 of casings, 7-8, 18-40, 51, 55-56
entrained air or gas, effect on, 581 of impellers, 62
for high specific speed pumps, 580 of pump parts, 11-13
impeller design, effect on, 564-581 of wearing rings, 72-75
liquid characteristics, effect on, 574 Non-clogging pumps, 334,759-760
operating mode, effect of, 583-585 Nozzle locations, 36
operation below, effects of, 584 Nozzles, flow through, 974
temperature rise, based on, 563 Nuclear power plants, cycles, 750-753
Mining pumps, 780-782 reactor coolant circulating pumps, 755-757
Mixed-flow impellers, 7, 64, 96-97, 454 safety-related pumps, 753-755
Mixed-flow pumps, 9, 737, 740, 742 reactor feed pumps, 751-753
Model pumps, 436-437 steam generator feed pumps, 751-753
Modulating bypass valve, 867 Number of pumps and load factor, 803-804
Monitoring, 893-915
action on monitored parameters, 907
extent necessary, 893-896
o
parameters and instrumentation, 896-907 Off-design operation, defined, 558
pump hydraulic performance, 891-892 effects as flow varied, 558
field testing, 907-915 Oil production, pumps for, 771-772
analyzing results, 912-915 Oil well pumps, 771
Motor characteristics and electrical variations, 644-645 Open impellers, 64
Motor starters, 645-649 Operating data, 891-892, 896-907
Motors (see Electric motors) Operating problems, diagnosis of, 927-945
Motor mounted pump (see Closed coupled pump) Operation of pumps, 882-892
Multistage pumps, 7 in the "break," 474, 492-494, 733
axial thrust in, 97-103, 104-115 gas bound, 889-890
casings for, 44-61 at high flows, 558-560, 889-890
rotor design for, 119, 123-127 at low flows, 560-561, 888-890
Multi-V-belt drive, 627 observation of, 893
(see also Dry operation, Parallel operation, Series
operation, Shutoff operation)
N Opposed impellers, axial thrust balancing with, 97-103
NPSH available, 416-418, 475 residual axial thrust in, 99
defined,415 stage arrangements, 10 1-103
influence on operating capacity, 473-474 (see also Axial thrust, in multistage pumps)
984 Index

Ordering pumps, 827-829 Pipeline pumps, 8, 339, 772-774


Outboard end, defined, 39, 212 Piping, balancing leak-off, 867-869
Overhaul, checks before removing pump, 919 bypass, 862-867
complete, 919-924 discharge, 854-855
on condition, 918-919 drain, 859-860
periodic, 918 friction losses in (see Friction losses, piping)
on failure, 918 installation of, 880
Overhung pump, 12,213 loads from, 40, 855-856
bearing arrangement, 212 effect on baseplate design, 300, 307
Oversize pump, effect on required NPSH, 499 suction, 838-850
effect on energy consumption, 547-551 vent, 859-860
Overspeed governor, 597, 626 warmup, 861-862
Pitot tube pump, 250 (see also Rotating casing pump)
p Plastic, use in centrifugal pumps, 301, 368, 378, 767
Power consumption controls, 597, 899
pH values, 387 Power, characteristic curves, 425-427
Packed box seal, 136, 139-157 hydraulic, 428
abrasive solids in pumped liquid, design for, 144 input, 428
adjustable gland for, 139, 154-156 measurement of, 912
materials for, 155 Power plant (see Steam power plant)
smothering type, 155 Prerotation, 32
solid, 155 Pressure, absolute, 404-405
split, 154, 155 atmospheric, 405
adjustment of, 156 gauge, 404-405
bottoming ring for, 139, 140 measuring of, 910
capabilities of, 141 units of, 404-405
cooling of, 145 vapor, 408
cyclone separator, use with, 144 Pressure controls, 597-599, 601, 604, 610-614, 623
leakage from, 139, 142, 156 Pressure distribution, over impeller shrouds, 92, 106
maintenance of, 156-157 Pressure and head conversion, 402, 405, 406
packing for, 139, 150-154 Pressure-reducing devices, within pump, 148
causes of wear of, 157 Pressure-reducing orifices, 865
seal cages for, 141, 144 Preventive maintenance, 917-918
sealing liquid for, 142, 144 Priming, 686-714 (see also Self priming pumps)
shaft (or sleeve) runout, effect of, 154 automatic, 695-700
shaft (or sleeve) surface under, 154 of internal-combustion-engine driven pumps, 708-710
stuffing box for, 139 of sewage pumps, 707-708
throat bushing for, 139, 140, 144 of steam turbine driven pumps, 711
Packing for stuffing boxes (see Packed box seal) time required for, 714
Packless stuffing boxes (see Breakdown seals) Priming chambers, 688
Paintline pumps, 336 Procuring pumps, 800-829 (see also Purchasing)
Paper stock pumps, 779, 817 Propeller pumps, 7
Parallel operation, 20,499, 534-538, 539 arrangements of, 341-346
of centrifugal and reciprocating pumps, 891 as external circulators, 767-769
practice to minimize pumping energy, 554-555 torque-speed characteristics of, 655-657
Partial emission pumps, 357-360 vertical, 330-334, 339
Pedestals, 39, 300, 306, 313 pull-out design, 333
Performance characteristics, 423-472 stages, usual number of, 332
of induced vortex pumps, 360 Protective controls, 614-617
of partial emission pumps, 358 Pollution, influence on pump selection, 136-137
of regenerative pumps, 355-356 Pulp and paper pumps, 779, 817
of rotating casing pumps, 364 Pump characteristics, curves of, 423-428
of viscous shear pumps, 363 estimating from physical measurements, 456-461
Performance testing, in field, 907-915 Pump development, 946-952
analyzing results, 912-915 Pump fittings, 382-385
Petroleum industry, pumps for, 769-779, 816-817 Pump life, estimate of, 584-585
Pilot devices, 609 Pump location, 808, 834-835
Index 985

Pump-out vanes, 67, 68, 77, 94 Rigid couplings, 272, 273-278


Pump-turbine, 4, 799 Rigid shaft (see Shaft, stiff; see also Rotor, stifO
Pumped storage, 797-799 Roller bearings, 214
Pumping, defined, 3 Rotating casing pump, 363-365
energy savings in, 547 Rotation, checking motor, 885
Pumping systems, types of, 532 of pump, 38-39, 827 (see also Reverse rotation)
throttled, 532-539 Rotating wearing rings, mounting of, 82-83
unthrottled, 539-544 Rotor assembly, defined, 57
Pump speed, measurement of, 907, 912 Rotor, balancing of, 922
Pump support, number of points of, 40 between bearings, 122, 123-127,212
effect on piping load capacity, 40 cantilever (see Rotor, overhung)
Purchasing, arrangement for, 800-801 critical speed of, defined, 116
essential data required, 801-810 bending, 122, 123
evaluating bids, 818-826 "dry," 120, 122, 123
inquiries for specific services, 810-818 "wet," 120
ordering equipment, 827-829 separation margin, 120
torsional, 120
deflection, static, 118, 119
Q dynamic, 117, 118, 120
Quench fluid (see Mechanical seals) (see also Gland, factors related to, 122, 124 (see also Shaft
smothering) flexibility factor)
at shaft seal, 122
design choice, multistage pumps, 125-127
R
dynamics, 114, 115, 116
Radial-flow impellers, 7 flexible, defined, 117
Radial thrust, 20-28 forces acting on, 44, 118, 125 (see also Radial thrust)
Radially split casings, 8, 28, 32-34, 39, 44-45 hydraulic stiffness, 119-120, 123 (see also Lomakin
double, 55-59 Effect)
maintenance of, 59-61 hydraulic support of, 107 (see also Lomakin Effect)
multistage "ring casing" type, 55 maintaining balance of, 56, 57
single, 56 monitoring vibration of, 904
Raised face flanges, use of, 54-55 overhung, 122-123, 212
Rating curves, 461-464 design practice, 122
Reassembly, of pump, 922-924 stiff, defined, 117
of inner element, 922 stiffness, mechanical, 117-119, 122-125
of rotor, 922 evaluation of, 824
Recessed impeller pump (see Induced-vortex pump) Rubber lined pumps, 750, 769, 782
Recirculation, internal, 564-580 Runaway speed, with reverse flow, 346
minimum flow, bypass for, 862-867 Running clearances, axial, 79
Record of inspection and repairs, 917 future developments, 951
Refinery pumps, 774-779 hard coating of, 77, 83, 380-381
Refrigeration pumps, 785 at impeller periphery, 79
Regenerative pumps, 351, 352-357 radial, 72-79
Regulation of capacity (see Capacity regulation) (see also Wearing rings)
Regulators; constant pressure, 604-605
differential pressure, 610-612
excess pressure, 610-612
s
feedwater level, 609-614 Salt water pumps, 383-385
Relief valves, 616 Seal cages, 141
Repair parts (see Spare and repair parts) Seal rings, elastomer, 53
Repair record, 922 Seal, shaft (see Shaft seal)
Resonance, 122 Sealless pumps, 136-137 (see also Hermetically
Return channel, 23 sealed pumps)
Reverse flow, through pump, 346 "canned" motor, 673
Reverse rotation, protection against, 616 integral motor, 669-673
Reversible pump, 365-366 magnetic drive, 683-685
Reversible pump-turbine, 4, 799 vertical cantilever, 335
986 Index

Seawater, materials for, 376, 378, 385 types, selection of, 137-139
stagnant, effect of on corrosion, 376 varying pressure at, 67, 148-149
Selection of pumps, 801-826 Shaft span (see Bearing span)
Self-priming pumps, 347-351 Shipment, preparation for, 873
recirculation at discharge, 349-351 Shutoff, head at, 424-425, 439,
recirculation to suction, 347-349 operation at, 537, 561-562, 653
regenerative type, 351 Side discharge, 37
Self-priming units, 703-707 Side suction, 37
Semiclosed impellers (see Semiopen impellers) Single-stage pumps, 7
Semi open impellers, 64-67, 329 Single-suction impeller, 7, 63, 91-98, 111-113,481
Separately-coupled pumps, use of, 32 Single volute, 20
Series operation, 539, 543-544, 743, 772, 780 Sigma factor, 478
Series units, 7, 40 Siphon breakers, use of, 616
Severity of service, defined, 894 Siphons, 346,419, 525, 656
Sewage pumps, 334, 759-761, 813-814 Sleeve bearings, 236-261 (see also Bearings, plain)
priming of, 707-708 Slurry pumps, 782, 818
Shaft deflection, 118-119, 122-123,317-318 (see also Soleplates, 307, 836
Rotor, deflection, static) Solids, pumping of, 68
Shaft sleeves, 128, 130-135 Sound (see Noise)
materials for, 133, 381 Space and weight limitations, 810
maintenance of, 134 Spare and repair parts, 808, 924-926
sealing of, 13 2 Special effect pumps, 352-366
Shaft induced-vortex, 360-361
chrome plating of, 130, 247 partial emission, 357-360
design requirements for, 116 regenerative, 352-357
flexibility factor of, 123, 124 reversible, 365-366
heavy (see Shaft, large, defined) rotating casing, 363-365
impeller mounting, methods of, 127-129 viscous-drag, 361-363
individually located, 129 Specific gravity, effect of, on casing design, 30, 45
impeller overhang, defined, 123 on power, 428
large, defined, 123 on pump characteristics, 464
maintenance of, 129 Specific heat, pumped liquid, 137, 562
materials for, 128, 370, 381-382 Specific speed, 437-452
for vertical pumps, dry pit, 317-319 defined, 437-438
wet-pit, 324, 334 effect on energy consumption, 545-547
safe torque, estimate of, 122, 458 effect on impeller sensitivity to inlet flow, 30, 341
slender, defined, 123 effect on pump characteristics, 439-446
"solid," 127, 129 effect on pump cost, 126
strength of, 122 effect on pump runaway speed, 346
straightness of, 129 effect on running clearance leakage losses,
welding of, 130 555-557
Shaft orientation, effect on casing design, 30 relationship to impeller profile, 64, 439, 451-452
Shaft power (see Power, input) significance of, 438-446
Shaft seal, 136 Specific speed limit charts, 481
allowable leakage from, 136, 137 Speed, calculation of for off-curve condition, 435-436
breakdown, 195-210 with reverse flow, 346
environment at, 137 Speed change, effect of, 428-431
fixed packing, 157-158 Speed-torque curves, 652-667
function of, 136 Split casing, defined, 28
future of, 949 Stable curves, 425, 534, 538
hydrodynamic, 158-160 Stage-pieces, 51
monitoring, injection to, 899-900 fit within axially split casing, 53
performance of, 906 Standard fitted pump, 382
mechanical, 161-194 Standardization, 6, 950-951
operating problems, diagnosis of, 942-944 Standby pumps, 887-888
packed box, 139-157 Starting pumps, 885-887
pressure at, 100, 103, 137 Static head, 409
temperature at, 137 Stationary wearing rings, mounting of, 53, 80-82
Index 987

Stay ring, 321 System head curves, 412, 522-544, 609


Steam power plant, cycles, 717-719 for complex systems, 527-532
pump services for, 720-757 variable static head, effect of, 527
Steam turbine driven pumps, priming of, 711 varying friction head, effect of, 522-523
Steam turbine, use of, 621, 622, 624
controls for, 612-614
Steel mill pumps, 782-783, T
hydraulic descaling, 782
Tandem drives, 626
primary scale pit, 336, 783
Tandem impellers, defined, 97
Stock pumps, 779, 817
Temperature, effect on casing design, 30, 45
Stop-pieces, 32
at shaft seal, 13 7
Stopping pumps, 887
at pump suction, measurement of, 910-912
Storage plants, 797-799
Temperature control of condensate injection, 597, 599, 601
Storage of pumps, 874
Temperature rise, 561-564
Stuffing boxes, integral, 35, 147
calculation of, 562
separate, 35, 147
effect of axial thrust balancing on, 563
(see also Packed box seal)
recommended maximum, 563
Submergence, of pump (see Suction head)
Tesla pump (see Viscous drag pump)
to prevent vortexing, 838-839
Textile manufacture, pumps used, 779
Submergence control (see Operation of pumps
Terminal points, diagram for specifying, 810
in the "break")
Thermostatic control, 734
Submersible motors, 334, 667-673
"Three-way joint" in axially split casing, 53
dry-stator, 668, 673
Throttled pumping systems, 532-539
wet-stator, 668-673
Throttling pump suction, 890-891
Suction conditions, effects of, 415, 473-505
Thrust, axial (see Axial thrust)
effect on pump characteristics, 492-499
radial (see Radial thrust)
history of problems caused by, 477-480
Thrust bearings, 96-97,99,107,212,215-221,243-245
monitoring of, 897-899
Titanium, 396
Suction head, 414
Tongue-and-groove anti-rotation lock, 53, 82
Suction lift, 414
Top discharge, 37
Suction nozzle, 30-31
Top suction, 37
arrangements of, 31
Torque, conversion to head, 3
flange for, 55
definition of, 650-652
location of, 36-38
motor, 634-635, 662-667
Suction piping, 838-850
pump, 652-667
configuration of, 839-842
Torque-speed curves, 652-667
design considerations for, 836
low specific-speed pumps, 652-655
elbows, orientation of, 844-850
propeller pumps, 655-657
length of run, 838
Total dynamic head, 422
reducers, orientation of, 842-844
Total head, 421-422, 506
size of, 838
Toxicity, pumped liquid, effect on pump selection,
venting of, 859-860
136-137,673
Suction recirculation (see Internal recirculation at
Tramp material, in pumped liquid, 335
impeller suction)
Troubles, diagnosis of, 927, 945
Suction specific speed, 337, 480-483
Turbine pumps (see Vertical turbine pumps)
cavitation erosion, influence on, 501
Turbines (see Hydraulic power recovery turbines; Steam
defined,480, 571
turbines; Water turbines)
design value, selection of, 568, 574, 575
Turbulence pumps (see Regenerative pumps)
pump operation, influence on, 568
Twin volute, 23, 44, 319
Suction strainers, 850-854
Two-phase pumping, oil production, 772
Suction well design, vertical pumps, 341-346
Type characteristics, 428
Sump pumps, 336, 339-341
abrasive solids, design for, 340-341
life of, 340
Supports (see Base plates and pump supports)
u
Surface speed, at shaft seal, 137, 141, 161 Unbalance, rotor, allowable, 922
Surging, causes of, 534 Unbalanced mechanical seals, 169
System friction curve, 412 Underwriter fire pumps, 762-764
988 Index

Units of, capacity, 401 for sewage service, 324


head, 402 submersible, 335
Universal drive shafts, 316 (see also Couplings, flexible for sump service, 339
drive shafts) suspended, 335
Unstable conditions, 19 Vertically split casings (see Radially split casings)
Unstable curves, 424-425 Vibration, causes of, 71, 116-117, 120,572,933-936
Unthrottled pumping systems, 539-544 resonant, 116
Utility service, refineries, pumps for (see Steam power at other than running speed, 117, 934
plants) at running speed, 934-935
vane-passing, 579, 935
v Viscosity, of common liquids, 966-971
effect on pump characteristics, 464-469
"V" belt drive, 627 conversion of units, 972-973
Vacuum producing devices, 690-694 Viscous drag pump, 361-363
Vacuum pumps, 690-693 Viscous liquids; friction loss for, 962-965
Valves; friction losses in, 517-518 lowering energy consumption by preheating, 557
(see also Automatic vent valves; Check valves; Controls; pumping of, 464-469
Foot valves; Relief valves) Voltages, standard, 630
Vane passing syndrome, 579 Volute casing, 18
Vanes, impeller, form of, 64 Volute pump, 7, 18,44, 324, 339, 778
clearance with collector vane(s), 68, 579 Volutes, double, 23
closure, 64 modified, 24
split and staggered, 70 single, 20-23, 44
Vapor pressure, 408 twin, 23, 44
Variable frequency drive (VFD), 599, 613, 624, 681-683 Volute tongue, 42
Variable speed opeation, advantages of, 624 Vortexing, 737, 838-839
drives for, 624, 673-683 Vortex breaker, 839
effect on energy consumption, 551-554, 676-677, 682 Vortex pumps (see Regenerative pumps)
Velocity head, 410
Vent valves (see Automatic vent valves)
Venting centrifugal pumps, 55, 700-701, 859-860
w
Vertical dry-pit pumps, 311-321, 740, 808 Warm up, need for, 861, 882-885
driver mounting on, 313 piping for, 861-862
driver mounted separately, 315 Waste water pumps, 361 (see also Sewage pumps)
single suction, 312-313, 319 Water cooled seal housings, 145-146
Vertical in-line pumps, 313 Waterflood pumps, 4, 771
Vertical pumps, 9, 77, 311-346 Water flushed rings, 76, 77
arrangements of, 341-346 Water horsepower (see Hydraulic power)
suction well design for, 341-346 Water injection (see Waterflood pumps)
valves for, 346 Water turbine (see Hydraulic turbine)
Vertical turbine pumps, 7, 324-329, 339 Water works pumps, 758-761, 812-813
can type, 329 Wear plates, use of, 35
foundations for, above ground discharge, 836 Wear of running clearances, adjustment for, 79
below-ground discharge, 836 allowable, 85
hydraulically balanced impellers for, 329 based on energy consumption, 555-557
line shaft adjustment, need for, 324 effect on axial thrust, 100, 114-115
thrust bearing, loads on, 324 effect on performance, 555
effect of impeller type, 329 effect on rotor dynamics, 120, 124
Vertical volute pumps, 336-339 Wearing parts, defined, 379-380 (see also Wearing rings)
cantilever type, 336, 341 materials for, 379-381
double-suction suspended, 337 Wearing rings, 72-83, 94, 97, 329
multistage, 339 casing, 74
Vertical wet-pit pumps, 311, 321-341 clearances, 85
above-ground discharge, 9, 324 dam-type,77
below-ground discharge, 9, 324 double, 73
enclosed line shafting for, 324, 333 flat, 72, 73
lubrication of bearings in, 336 flushed, 77
open line shafting for, 324, 333 grooved (see serrated)
Index 989

impeller, 82-83 serrated, 75, 76, 119


L type, 73 single, 72, 73
labyrinth,74 stationruy,80
maintenance of, 84 stepped, 75, 76
materials for, 83, 85, 380 Weight and space limitations, 810
mounting of, rotating, 82-83 Wet-pit pumps, 9, 321-341, 808, 836
stationruy, 53, 80-82 Williams and Hazen friction data, 507-511
nozzle type, 73, 74
renewable, 72
retention of, 80-83
z
ports for inspection of, 77 Zirconium, 396
sealed, 82 carbide of, 183

Você também pode gostar