Você está na página 1de 108

The Florida State University

DigiNole Commons
Electronic Theses, Treatises and Dissertations

The Graduate School

4-3-2009

Direct Photon Production In Association With A


Heavy Quark
Tzvetalina P. Stavreva
Florida State University

Follow this and additional works at: http://diginole.lib.fsu.edu/etd


Recommended Citation
Stavreva, Tzvetalina P., "Direct Photon Production In Association With A Heavy Quark" (2009). Electronic Theses, Treatises and
Dissertations. Paper 1572.

This Dissertation - Open Access is brought to you for free and open access by the The Graduate School at DigiNole Commons. It has been accepted for
inclusion in Electronic Theses, Treatises and Dissertations by an authorized administrator of DigiNole Commons. For more information, please contact
lib-ir@fsu.edu.

FLORIDA STATE UNIVERSITY


COLLEGE OF ARTS AND SCIENCES

DIRECT PHOTON PRODUCTION IN ASSOCIATION WITH A HEAVY


QUARK

By
TZVETALINA P. STAVREVA

A Dissertation submitted to the


Department of Physics
in partial fulfillment of the
requirements for the degree of
Doctor of Philosophy

Degree Awarded:
Spring Semester, 2009

The members of the Committee approve the Dissertation of Tzvetalina P. Stavreva defended
on April 3, 2009.

Joseph F. Owens III


Professor Directing Dissertation

William Dewar
Outside Committee Member

Laura Reina
Committee Member

Horst Wahl
Committee Member

Marcia Fenley
Committee Member

The Graduate School has verified and approved the above named committee members.
ii

To mum and dad

iii

ACKNOWLEDGEMENTS

I would like to thank my parents for showing me how captivating science can be at an
early age, and also for giving me all the support and encouragement that I needed.
I would also like to thank my friends, for being there for me, whenever I needed a helping
hand, and for creating a home away from home. I would especially like to thank Cecily
Oakley for being a wonderful friend, and of course for her amazing cupcakes that would
provide a good start to the working day, Christelle Castet for being a great roommate and
friend, and someone who was always there to have an adventure with. As well as that I
am thankful to my fellow high energy graduate students, Jose Lazoflores, Edgar Carrera,
Benjamin Thayer, Paulo Rottmann, and Dan Duggan (especially for providing the so needed
experimental results) with all of whom it was great to share an office, as well as many
meaningful (and some not so meaningful, but nonetheless fun) discussions all these years,
also I am indebted to Fernando Febres Cordero and Bryan Field for sharing their knowledge
with me and helping me come to grasps with the subtleties of QCD. And a very special thank
you to Shibi Raj, for being helpful and supportive, and for putting things in perspective, at
times when I needed inspiration.
I want to thank the High Energy Physics Group, Faculty and Staff, for creating a very
pleasant environment in which to work, and also my committee members, with a special
thanks to Laura Reina and Horst Wahl. I am particularly grateful to my advisor Jeff Owens,
first for giving me the opportunity to come back to physics and fulfill one of my goals, to
study high energy physics, and of course for leading me through the intriguing terrain of
next-to-leading-order QCD calculations.
Tzvetalina

iv

TABLE OF CONTENTS

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

vii

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

viii

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

xiii

1. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2. NLO CALCULATION OF THE DIRECT PHOTON AND HEAVY QUARK


CROSS SECTION AT HADRON COLLIDERS . . . . . . . . . . . . . . . . .
2.1 Basic Overview of a Hadronic Cross Section Calculation . . . . . . . . .
2.2 Numerical Methods - Phase Space Slicing Method . . . . . . . . . . . . .
2.3 +Q calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.1 LO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.2 NLO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4 Stability of the Calculation . . . . . . . . . . . . . . . . . . . . . . . . .

8
9
12
14
14
18
32

3. RESULTS . . . . . . . . . . . .
3.1 Tevatron Predictions . . .
3.1.1 Photon Isolation and
3.1.2 Intrinsic Charm . . .
3.2 LHC Predictions . . . . .

.
.
.
.
.

36
38
46
48
50

4. COMPARISON TO DATA . . . . . . . . . . . . . . . . . . . . . . . . . . . .

56

5. MASSIVE VERSUS MASSLESS COMPARISON . . . . . . . . . . . . . . . .


5.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2 Numerical Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

71
72
74

6. CONCLUSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

78

APPENDICES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

80

A. PARTON DISTRIBUTION FUNCTIONS . . . . . . . . . . . . . . . . . . . .

80

APPENDICES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

82

. . . . . . . . . . . .
. . . . . . . . . . . .
NLO Fragmentation
. . . . . . . . . . . .
. . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

B. PHASE SPACE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

82

APPENDICES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

85

C. 2 3 NLO MASSLESS MATRIX ELEMENTS

. . . . . . . . . . . . . . . .

85

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

87

BIOGRAPHICAL SKETCH . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

92

vi

LIST OF TABLES

1.1

A list of the particles comprising the SM. . . . . . . . . . . . . . . . . . . . . .

2.1

A list of all 2 2 LO hard-scattering subprocesses with at least one heavy quark,


Q, in the final state (where Q, can be either a heavy quark or an antiquark, where
appropriate) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

17

2.2

a list of all 2 3 NLO hard-scattering subprocesses . . . . . . . . . . . . . . . .

21

2.3

A list of all 2 3 NLO hard-scattering subprocesses, O(s3 ), with at least one


heavy quark, Q, in the final state (where Q, can be either a heavy quark or an
antiquark, where appropriate) . . . . . . . . . . . . . . . . . . . . . . . . . . .

31

. . . . . . . . . . .

72

A.1 A summary of the power of logs resummed, dependent on the order of the splitting
functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

81

5.1

Subprocesses that contribute to the LO massive calculation

vii

LIST OF FIGURES

2.1

Self Energy of the Photon - an example of a loop diagram . . . . . . . . . . . . .

12

2.2

Compton Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

15

2.3

with the
An example of Leading Order Fragmentation Contributions 1) gg QQ,
photon fragmenting from one of the final state heavy quarks 2) gQ gQ, with the
photon fragmenting off from the final state gluon . . . . . . . . . . . . . . . . .

18

2.4

Virtual Feynman diagrams for the Compton subprocess . . . . . . . . . . . . . .

19

2.5

Example Feynman diagrams for all the possible real subprocesses . . . . . . . . .

21

2.6

Initial collinear singularities for gQ gQ, the final state gluon is collinear to the
initial state heavy quark, or the final state gluon is collinear to the initial state gluon. 23

2.7

Feynman diagrams for the subprocess qQ Qq. . . . . . . . . . . . . . . . . .

27

2.8

diagrams 1) and 2) are s-channel


Feynman diagrams for the subprocess q q QQ,
diagrams, and diagrams 3) and 4) are t-channel diagrams. . . . . . . . . . . . .

29

shown in diagram 1). A photon


Feynman diagram for the subprocess gg QQg
being produced by fragmentation from one of the final states for this subprocess,
diagram 2) from the final state quark, diagram 3) from the final state gluon, and
diagram 4) from the final state anti-quark . . . . . . . . . . . . . . . . . . . . .

2.10 s dependence of N LO (p
p bX), at S = 1.96 TeV, for c = 5 105 when
s > 0.001 , and c = 1 105 , when s < 0.001 . The upper plot shows the
dependence on s of the two-body (dashed line), and three-body (dot dashed line)
contributions, and also the sum of the two (solid line). The lower plot shows
the dependence on the full NLO cross section, with the statistical error included,
= r = f = F , = pT . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.11 c dependence of N LO (p
p bX), at S = 1.96 TeV, for s = 0.01. The upper
plot shows the dependence on c of the two-body (dashed line), and three-body
(dot dashed line) contributions, and also the sum of the two (solid line). The lower
plot shows the dependence on the full NLO cross section, with the statistical error
included, = r = f = F , = pT . . . . . . . . . . . . . . . . . . . . . . . .

2.9

viii

32

33

34

3.1

3.2

3.3

3.4

3.5

3.6
3.7

The LO differential cross section at the Tevatron, d/dpT forthe production of


a direct photon and a bottom quark as a function of pT for S = 1.96 TeV at
LO, with the scale dependence shown, where the three different scales have been
set to be equal = r = f = F , = pT (solid line), = pT /2 (dashed line),
= 2pT (dotted line). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

39

The differential cross section at the Tevatron, d/dpT


for the production of a direct
photon and a bottom quark as a function of pT for S = 1.96 TeV, at NLO (solid
line), and at LO (dashed line), = pT .
. . . . . . . . . . . . . . . . . . . . .

40

The differential cross section at the Tevatron, d/dpT for the production of a direct
photon and a charm quark as a function of pT for S = 1.96 TeV, at NLO (solid
line), and at LO (dashed line), = pT .
. . . . . . . . . . . . . . . . . . . . .

41

Contributions of the different subprocesses to the differential cross section, NLO


(dashed line), Qq qQ, and Q
(solid line), annihilation q q QQ
q

qQ(dotted line), gQ gQ (dot dashed line), gg QQ+LO


(dash dot dotted
QQ,
and QQ QQ (dot dash dotted line), = pT . . . . . . . .
line), QQ

42

A comparison between the differential cross sections, d/dpT for the production
of a direct photon and a bottom quark, and that of a direct photon plus a charm
quark at NLO and LO, charm at NLO (solid line), bottom at NLO (dashed line),
charm at LO (dot dashed line), bottom at LO (dotted line), = pT . . . . . . . .

43

The ratio of the charm and bottom differential cross sections versus pT , at NLO
(solid line) and at LO (dashed line), = pT . . . . . . . . . . . . . . . . . . . .

44

Scale dependence of the NLO differential cross section, d/dpT for the production
of a direct photon and a bottom quark, where the three different scales have been
set to be equal = r = f = F , = pT (solid line), = pT /2 (dashed line),
= 2pT (dotted line). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

45

3.8

Comparison between the differential cross section, d/dpT without isolation requirements and with them, no isolation (solid line) , isolation (dashed line), = pT . 47

3.9

Ratio between the differential cross section d/dpT , with NLO fragmentation
contribution included and the differential cross section with just LO fragmentation
included, no isolation required (solid line), isolation (dashed line), = pT . . . . .

48

3.10 Comparison between the three different charm PDFs at scale Q = 40 GeV,
CTEQ6.6M (solid line), BHPS or CTEQ6.6C2 (dashed line), sea-like or CTEQ6.6C4
(dotted line). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

50

3.11 The differential cross section, d/dpT , for the production of a direct photon and
a charm quark for the three different PDF cases, CTEQ6.6M (solid line), BHPS or
CTEQ6.6C2 (dashed line), sea-like or CTEQ6.6C4 (dotted line), = pT . . . . .

ix

51

3.12 The differential cross section versus the transverse momentum of the photon
d/dpT for the production
of a direct photon and a bottom quark at LHC center

of mass energies, S = 14 TeV, NLO (solid line), LO (dashed line), = pT . . . .

52

3.13 K factor, or the


ratio of the NLO to the LO differential cross section for
pp bX at S = 14 TeV, = pT . . . . . . . . . . . . . . . . . . . . . . .

53

3.14 Contributions of the


different subprocesses to the differential cross section, d/dpT

(dashed
for pp bX at S = 14 TeV, NLO (solid line), annihilation q q QQ
line), Qq qQ, and Q
q qQ (dotted line), gQ gQ (dot dashed line),

QQ,
and QQ QQ (dot dash
gg QQ+LO
(dash dot dotted line), QQ
dotted line), = pT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

54

3.15 Scale dependence of the differential cross section, d/dpT for the production of a

4.1

4.2

4.3

4.4

4.5

direct photon and a bottom quark at the LHC, where the three different scales have
been set to be equal = r = f = F ,for the NLO cross section = pT (solid
line), = 2pT (dashed line), = pT /2 (dotted line) and for the LO cross section
= pT (dot dashed line), = 2pT (dashed dot dot line), = pT /2 (dashed
dashed dot line). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

55

The differential cross section, d/dpT , for the production of a direct photon and a
b quark, + b. In the graph to the left, the solid line is the NLO theory curve and
the circular dots are the data points measured by the D collaboration, for region
one. In the graph to the right, the dashed line is the NLO theory curve and the
squares are the data points measured by the D collaboration, for region one. . .

58

The differential cross section, d/dpT , for the production of a direct photon and a
b quark, + c. In the graph to the left, the solid line is the NLO theory curve and
the circular dots are the data points measured by the D collaboration, for region
one. In the graph to the right, the dashed line is the NLO theory curve and the
squares are the data points measured by the D collaboration, for region one. . .

59

Scale dependence of the differential cross section, d/dpT , for the production of a
direct photon and a c quark, + c, compared to data (circular dots), for region one.
The three different scales have been set to be equal = r = f = F , = pT
(solid line), = pT /2 (dashed line), = 2pT (dotted line). . . . . . . . . . . .

61

Scale dependence of the differential cross section, d/dpT , for the production of a
direct photon and a c quark, + c, compared to data (circular dots), for region two.
The three different scales have been set to be equal = r = f = F , = pT
(solid line), = pT /2 (dashed line), = 2pT (dotted line). . . . . . . . . . . .

62

The NLO differential cross section, d/dpT , for the production of a direct photon
and a c quark, + c, with the use of CTEQ6.6M PDFs (solid line), using the BHPS
intrinsic charm PDF, (dashed line) and with the use of the sea-like model intrinsic
charm PDF (dotted line), compared to data (circular dots), region one. . . . . . .

63

4.6

4.7

4.8

4.9

The NLO differential cross section, d/dpT , for the production of a direct photon
and a c quark, + c, with the use of CTEQ6.6M PDFs (solid line), using the BHPS
intrinsic charm PDF, (dashed line) and with the use of the sea-like model intrinsic
charm PDF (dotted line), compared to data (circular dots), region two. . . . . . .

64

Scale dependence of the NLO differential cross section, d/dpT , for the production
of a direct photon and a c quark, + c, with the use of the BHPS PDFs, = pT
(dashed line), = pT /2 (dotted line), compared to data (squares), region two. . .

65

Scale dependence of the differential cross section, d/dpT , for the production of a
direct photon and a b quark, + b, compared to data (circular dots), for region one.
The three different scales have been set to be equal = r = f = F , = pT
(solid line), = pT /2 (dashed line), = 2pT (dotted line). . . . . . . . . . . .

67

Scale dependence of the differential cross section, d/dpT , for the production of a
direct photon and a b quark, + b, compared to data (circular dots), for region two.
The three different scales have been set to be equal = r = f = F , = pT
(solid line), = pT /2 (dashed line), = 2pT (dotted line). . . . . . . . . . . .

68

4.10 The NLO differential cross section, d/dpT , for the production of a direct photon
and a b quark, + b, with the use of CTEQ6.6M PDFs (solid line), using the
BHPS intrinsic charm PDF, (dashed line) and with the use of the sea-like model
intrinsic charm PDF (dotted line), compared to data (circular dots), region one,
= r = f = F , = pT . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

69

4.11 The NLO differential cross section, d/dpT , for the production of a direct photon

5.1

5.2
5.3

and a b quark, + b, with the use of CTEQ6.6M PDFs (solid line), using the
BHPS intrinsic charm PDF, (dashed line) and with the use of the sea-like model
intrinsic charm PDF (dotted line), compared to data (circular dots), region two,
= r = f = F , = pT . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

70

subprocess. In diagram 1) the two initial


Feynman diagrams for the gg QQ
gluons go into a pair of heavy quarks and the photon is emitted for the final state
heavy quark. In diagram 2) one of the initial state gluons splits into a heavy quark
anti quark pair. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

73

subprocess, in its collinear limit can be written as the Compton


The gg QQ
subprocess times a splitting function. . . . . . . . . . . . . . . . . . . . . . . .

74

The differential cross section, d/dpT for the production of a direct photon and a
bottom quark as a function of pT for S = 1.96 TeV, at massless NLO (solid line),
at massless LO (dashed line), and at massive LO (dotted line), = r = f = F ,
= pT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

75

xi

5.4

The differential cross section, d/dpT


for the production of a direct photon and a
bottom quark as a function of pT for S = 14 TeV, at massless NLO (solid line),
at massless LO (dashed line), and at massive LO (dotted line), = r = f = F ,
= pT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

xii

76

ABSTRACT

In this Thesis we present the Next-To-Leading-Order calculation, O(s2 ), of the inclusive


cross section for a photon and a heavy quark (charm or bottom), p
p/pp + Q + X,
(Q = c, b) at hadron colliders.

We include fragmentation effects through the Next-

To-Leading-Order. This calculation is performed with the use of a phase space slicing
technique so that the effects of experimental cuts can be easily included. We study in
detail the characteristics of this process at both the Tevatron and the LHC. Results for
the ratios of the charm and bottom cross sections are presented and the systematics of the
various subprocesses are compared and contrasted. The theory predictions are compared to
experimental measurements from the D collaboration at Fermilab. A brief overview of the
LO massive calculation is also presented, and compared to the NLO massless case.
We predict that the investigation of this process and our results will be relevant in the
study of heavy quark PDFs at the LHC.

xiii

CHAPTER 1
INTRODUCTION

The idea that the world is composed of indivisible building blocks has been around since
antiquity. The ancient Greeks used the term atom, which means uncuttable, to refer to
these building blocks. This word ended up as somewhat of a misnomer, but the idea stood
the tests of time and experimental observation. And today we still believe that the universe
comprises indivisible building blocks, which we refer to as elementary particles. To describe
the fundamental interactions of these particles, a theory known as the Standard Model (SM)
was developed [1, 2, 3, 4]. It is a quantum field theory (QFT) that incorporates three of the
four known forces: the strong, the weak and the electromagnetic ones. The SM comprises
the particles shown in Table 1.1. They can be divided into two types of particles, fermions
which are spin half particles, and bosons which have integral spin. The fermions, i.e., the
quarks and leptons, form three generations, each generation differing from the previous only
by the mass of the particles in it. The fermions interact with each other via the gauge
bosons which are the mediators of the three forces. They include the eight gluons for the
strong force, the photon for the electromagnetic force and the Z and the W bosons for the
weak force. There is one more particle predicted by the SM known as the Higgs boson. It
is necessary in order for the massive particles in the SM, to acquire their mass. It is the
only particle in the SM that has eluded experimental observation as of yet. Hopes for its
discovery are resting on the new particle collider at CERN, Geneva, [5] known as the Large
Hadron Collider (LHC), which is expected to begin operation in late 2009.
Yet even with the discovery of the Higgs, a few important questions will remain
unanswered by the SM. These range from the Higgs hierarchy problem, to the incorporation
of the fourth force - gravity - in a QFT framework. Also the SM does not explain why there is
more matter than anti-matter, known as the matter anti-matter asymmetry, in the universe.
1

Table 1.1: A list of the particles comprising the SM.


fermions
I II
u c
quarks
d s
e
leptons
e

III
t
b

bosons
vector scalar
g

Z,W

As well as that, the SM does not provide particle candidates to what is mysteriously referred
to as dark matter and dark energy, the need for which comes from the fact that the known
baryonic and radiation matter accounts for only four percent of the mass of the universe,
i.e., its mass is largely unaccounted for. In order to be able to answer these questions, and
create a new theory that incorporates the old one and elucidates these queries, one thing is
clear, a detailed understanding and a thorough check of the SM are necessary. It is the aim
of this dissertation to provide a small part to this understanding.
To test the SM and be able to find out more about the substructure of matter, we
look at collisions characterized by large momentum transfers amongst the particles involved.
Such large momentum transfer collisions probe the short distance structure of the colliding
particles. The investigation of these processes requires effort from both the theoretical and
the experimental communities. Experimentally, these processes are observed at particle
accelerators. The currently highest center of mass energy particle collider is the Tevatron

at Fermilab [6], with a center of mass energy, S = 1.96 TeV. At the predecessors of
the Tevatron, at the beginning of the 1950s and 1960s, a multitude of new particles were
discovered. In order to classify them and bring some sense of order to what was referred
to as the particle zoo, a SU (3) flavor symmetry scheme was proposed independently by
Gell-Mann and Neeman [7, 8], as a means by which to classify the hadronic states, through
the existence of particles named quarks. At the same time the parton model was developed
by Feynman [9], in order to understand the dynamics of high energy particle scattering
processes, with the partons and quarks turning out to be the same particles. As a way
of developing a gauge theory of the interacting colored quarks, Quantum Chromodynamics
(QCD) was proposed, and the combination of QCD and the Electroweak theory was named

the SM. The SM has been deemed a very successful theory, as it predicted the existence of
the gluon, the top and charm quark, the tau neutrino and the W and Z bosons, before they
were observed. The discoveries of these particles along with a number of high precision data
measurements obtained from the Tevatron and the LEP collider at CERN, have confirmed
the SM. However there have been no indications of signals of new physics, and so the last two
decades have been marked by a slow-down of new breakthrough experimental observations.
The LHC is expected to rectify the lack of data that will drive high energy physics in a
direction beyond the SM. It is going to probe energy scales in the TeV range, with a center
of mass energy seven times higher than the one at the Tevatron.
With these long-awaited measurements expected from the LHC in the near future,
calculations in perturbative QCD (pQCD) will prove to be essential in the discovery of new
physics and the discrimination between new physics signals and already observed background
events. As the number of these background events are expected to be much larger than
the searched for signal, it becomes necessary to know them to a high degree of precision.
Furthermore, since there are many predicted possibilities and also unknown physics that can
be discovered a thorough cross-check with the SM is needed.
In order to make theoretical predictions which can be compared to data, a cross section
is calculated. This cross section gives the probability of a certain type of event to occur,
and is computed with the use of perturbative techniques. Since most of the observed SM
events involve the strong interactions, which are described by QCD, it is very fortunate that
we can use perturbation theory to study them. Perturbative techniques are possible to use
in the study of QCD through the existence of asymptotic freedom. Asymptotic freedom is
the term used to describe the fact that at large momentum transfers or short distances the
strong force becomes weaker and we can view the constituents of the hadrons, the quarks and
the gluons (collectively known as partons), as almost free. In Eq. 1.1 the 1-loop leading-log
expression for the dependence of the strong coupling constant s on the scale, Q, is shown
[10],
s (Q) =

2
,
b0 log(Q/)

(1.1)

where b0 = (11 2/3nf ), nf is the number of quark flavors, and is the mass scale below
which non-perturbative effects take over. From Eq. 1.1 we can clearly see that as the energy
scale Q increases the size of the strong coupling decreases. This is due to the fact that we

are dealing with a non-abelian theory, where the gauge bosons, the gluons, can interact with
each other, and, as such, the constant b0 is positive, as long as the number of flavors remains
less than sixteen, nf 16. Thus we can calculate the cross section as an infinite series in
the expansion of the strong coupling constant. However each following term in the series
becomes more and more difficult to calculate, and we have to limit ourselves to just the first
few terms. Some time ago only the first term of the series or the leading order (LO) term was
necessary in order to have a somewhat reliable prediction for the experimental observations.
But now as the experimental measurements are becoming more and more precise, so too
need be the theoretical calculations. The expected order is now the next-to-leading (NLO)
order, or the second term in the infinite expansion series. And with the measurements of
the LHC, next-to-next-to-leading order (NNLO) calculations are most likely to be needed
for certain processes.
One aspect that complicates a higher precision QCD calculation is the existence of
confinement. Even though due to asymptotic freedom the quarks and gluons can be treated
as free in the high energy processes, there are no free quarks or gluons in nature; they form
bound states in hadrons. The description of the behavior of quarks and gluons inside these
bound states is done through the parton distribution functions (PDFs). The PDFs tell us
the probability of finding a parton inside a hadron with some momentum fraction of the
parent hadrons momentum, at a given energy scale. In order to do most QCD calculations,
the PDFs of the partons inside the colliding protons or antiprotons need to be known.
However the PDFs dependence on the momentum fraction is not calculable and needs to
be measured experimentally. Since confinement can pose somewhat of a challenge in the
experimental study of PDFs, a way to get information about the structure of the hadrons,
and probe directly what goes on inside the large momentum transfer or hard scattering
subprocess is to look at particles unaffected by confinement or the strong interactions. One
such particle is the photon, or the carrier of the electromagnetic force. As a carrier of the
electromagnetic force it couples to the quarks charge and is blind to the strong force. Since
we want a photon that comes directly from the hard scattering process, and is not produced
after the partons have hadronized, it should be distinguished from a photon that comes from
the decays of hadrons, e.g. such as 0 , or 0 . We call a photon produced
at the hard scattering level or at the level before hadronization a direct photon.
The study of direct photons is aimed at providing information about the PDFs [11]
4

and also at testing the adequacy of the perturbative techniques used to calculate the hard
scattering subprocesses [12, 13]. It provides a solid testing ground for QCD interactions as the
QFT of the electromagnetic interactions - Quantum Electrodynamics (QED) has been well
tested and understood. Due to the beneficial aspects it provides, direct photon production
has been studied thoroughly since the construction of the SM [14, 15, 16, 17, 18, 19, 20]. It is
the single photon cross section that provides the basic observable for direct photon studies.
The calculation of this cross section involves integrations over the phase space variables of
the accompanying partons, thereby limiting the information which can be obtained about
the underlying subprocesses. Thus the calculation of the direct photon cross section can be
viewed as the sum of many pieces. It is of interest to take this sum apart and investigate
in detail one particular piece of the direct photon calculation. Here we concentrate on such
a piece, namely the study of direct photon production in association with a heavy quark,
where the heavy quark can be either a charm or a bottom quark.
The charm and the bottom quarks were discovered in 1974 [21, 22] and 1977 [23]
respectively. Their discovery offered yet more proof of the validity of the SM and also new
ways of investigating the strong interactions. The charm and bottom quarks have masses
of 1.16 to 1.34 GeV and 4.13 to 4.37 GeV in the MS scheme, correspondingly [24]. They
are known as heavy quarks due to the fact that their masses satisfy mQ  , where is
shown in Eq. 1.1, and has a value of about 200 MeV. As a consequence of the larger energy
required to create the heavy quarks, they are produced in less abundance than the light
quarks, namely the up, down and strange ones, at the Tevatron. However at the LHC due
to the higher center of mass energy of 14 TeV they are expected to play a prominent role.
Generally at colliders it is not just the heavy quark meson that is detected, but it is the
heavy quark jet. Typically, jets are made up of hundreds of particles almost all of which
contain light quarks. A jet containing a heavy quark provides quite a distinctive tag, due to
the longer lifetimes of the charm and bottom quarks. They are distinguished through the
secondary vertex formed as a result of this longer decay time.
The combined study of direct photons and heavy quarks can be used as a further test of
the validity of perturbative QCD and as a way to learn more of the heavy quarks role in
the nucleon. Furthermore final states involving electroweak gauge bosons and heavy quarks
can be components of signals of new physics [25]. From this study we can test the validity
of the approximations we use in our QCD calculations. We will also be able to learn more
5

about the structure of the proton and antiproton by being able to constrain more precisely
the distribution functions of the heavy quarks. Furthermore, an understanding of photon
plus heavy quark production will help us be prepared to identify possible signatures of new
physics.
In this dissertation we have calculated the inclusive cross section for the production of a
photon in association with a heavy quark, p
p/pp + Q + X. There have been previous
studies of this process. In Ref. [26, 27], the NLO cross section is presented for the production
of a direct photon and a charm quark. However photon fragmentation effects, which are
needed for a complete NLO calculation were included at lowest order. The fragmentation
effects occur since a direct photon can be produced not only at the hard scattering level,
but through fragmenting from a quark or a gluon. These fragmentation effects are described
by the photon fragmentation functions (FFs), which give the probability for a photon to
originate from a quark or a gluon, with a certain momentum fraction of the quarks or
gluons momenta. We need to introduce the photon FFs in order to take care of collinear
singularities occurring in the case when a photon is emitted collinearly to a final state quark.
Here we extend these previous efforts by presenting a complete NLO cross section for the
first time.
We also investigate the direct photon plus bottom cross section, and compare it with
the direct photon and charm cross section. Results for this process at the Tevatron and
at the LHC have been obtained. This calculation has been done in the limit of a massless
approximation, i.e., both the charm and bottom quarks are treated as massless. In order to
be able to work in the massless approximation the heavy quarks and photons produced need
to carry a transverse momentum, pT , which is a few times larger than the mass of the heavy
quark mQ , i.e. pT 10 GeV. Since the lower bounds for the values of the transverse momenta
for direct photons and heavy quarks measurable at both the D and CDF collaborations at
Fermilab are above pT 10 GeV, a comparison with a massless calculation is appropriate.
Another important reason for working in the massless approximation is the fact that currently
the most abundantly used PDFs and FFs both treat the heavy quarks as massless in their
evolution. The calculation was performed with the use of the two cutoff phase space slicing
technique [28], so that the effects of experimental cuts can be included. The inclusion of
isolation requirements in the theoretical calculation is also eased by this method. Isolation
requirements need to be imposed in the experimental observation of a direct photon, in
6

order to acquire a reliable measurement of the photons energy, this imposition also helps
ease the differentiation of direct photons from those photons coming from the decays of
hadrons. By isolating the photon we require that it is not surrounded by a large amount
of hadronic energy, thereby reducing the background of secondary photons. This will affect
direct photons produced via fragmentation as they are produced in close proximity to the
partons from which they have been radiated.
The dissertation is ordered as follows: in Chapter 2 a description of the theory and
techniques for the calculation are outlined. In Chapter 3 results for the differential cross
section are shown. Predictions for what process at both the Tevatron and LHC are presented
and a comparison between the differential cross sections is given. The effects of including
the NLO fragmentation terms are shown, as well as the effect of the use of different charm
PDFs on the cross section. In Chapter 4 we present the comparison between the theoretical
calculation and data from D for this process. In Chapter 5 we briefly outline the differences
between a massive and massless theoretical calculation, and in Chapter 6 we summarize and
conclude our findings.

CHAPTER 2
NLO CALCULATION OF THE DIRECT PHOTON
AND HEAVY QUARK CROSS SECTION AT
HADRON COLLIDERS

In this chapter the basics of a NLO QCD calculation are presented. We give specific details
for the calculation of the inclusive massless cross section for the associated production
of a photon and a heavy quark at NLO, which is O(s2 ) in this case, where is the
electromagnetic coupling constant. As mentioned in the Introduction, direct photons are
produced alongside many different particles, and here we have concentrated on one particular
combination. The direct photon cross section constitutes a sum over all these different parts.
In order to calculate the cross section for the associated production of a direct photon and
a heavy quark, we can take that sum apart and retain only those parts from the direct
photon cross section that are relevant to this process. Thus, the general outline of the
calculation follows that of the direct photon production computation. We need to remove
the subprocesses that do not contain heavy quarks in their final state, and, as such, contribute
only to the direct photon production case. Also due to the fact that we have two particles
that we are interested in, in the final state we also need to modify how we treat some
singularities with respect to just the direct photon computation. The reasons for this being,
that when collinear singularities arise in the final state they are treated differently depending
on whether the particle is observed or not. Since the calculation becomes quite complex at
NLO, it is done numerically. The numerical method used to perform this calculation is the
two cutoff phase space slicing method (PSS) [28], where the 2 3 body phase space is
divided into a soft, a collinear and a hard region, more detail about this method will be
given in Section 2.2.

2.1

Basic Overview of a Hadronic Cross Section


Calculation

Any inclusive hadronic cross section for the production of a particle C can be written in
the form:
Z
(AB C + X) =

dxa dxb dzc Ga/A (xa , f )Gb/B (xb , f )


(ab cd..)DC/c (zc , F ), (2.1)

where A and B are the colliding hadron beams. By inclusive we mean that C can be produced
along side all allowed particles for the process. Since we integrate over their phase space, we
are not concerned with what X is.
There are two different components that go into the calculation of the inclusive hadronic
cross section, in Eq. (2.1). One is calculable via perturbation theory and is referred to
appropriately as short distance physics. The other is concerned with nonperturbative effects,
such as hadronization and is referred to as long distance physics. From the uncertainty
principle, xp ~/2, we can deduce that the short distance physics during the hard
scattering occurs through the exchange of large transverse momentum. The further or
previous emission of collinear particles occurs at lower transverse momentum scales, i.e.,
over longer distances. These long distance processes cannot affect what will happen at the
hard scattering level, and we can freely separate the two regions from each other. This
separation of long distance and short distance physics is proved possible by the factorization
theorems [29, 30].
As is shown in Eq. (2.1) the long distance behavior is separated out into the PDFs:
Ga/A (xa , f ) and Gb/B (xb , f ), and the Fragmentation Function (FF): DC/c (zc , d ), both
of which are probability density functions.

Thus Ga/A (xa , f )dxa as mentioned in the

Introduction gives the probability of finding parton a inside hadron A, with a momentum
fraction between xa = pa /PA and xa + dxa of the parents momentum PA , where pa is
the partons momentum. The FF is similar in nature to the PDF. DC/c (zc , d )dzc gives
the probability of parton c to fragment (or hadronize) into hadron C, which has a fraction
zc = PC /pc to zc + dzc of the initial partons momentum. Both the PDFs and FFs are
universal, i.e., they do not depend on the process at hand.

Unfortunately we cannot

compute the PDFs and FFs with the use of perturbation theory, since we need to deal
with energy scales below in order to determine the hadron structure, and at these scales
nonperturbative techniques need to be used. Thus the dependence of the distribution and
9

fragmentation functions on the momentum fraction has to be determined empirically, through


experimental measurements.
Since perturbation theory cannot be used in this energy region, the PDFs dependence
on x and the FFs dependence on z need to be measured experimentally.
In Eq. (2.1), the short distance part of the hadronic cross section is described by the
partonic cross section
(ab cd..). It gives the probability for the hard scattering subprocess
to occur and is calculable via perturbation theory, as an infinite series in the strong coupling
constant s :

sn+i
(n) ,

(2.2)

n=0

where i can be 0, 1, 2, ... depending on the process. At first glance the calculation of

in Eq. (2.8) seems, although toilsome, fairly straightforward, especially if we are limiting
ourselves to the calculation of just a few terms (n = 0, 1, 2). However as soon as we go
beyond LO, i.e., n > 0, divergences appear. There are three different types of divergences
that can appear in a higher precision calculation. These are the ultraviolet (UV), the infrared
(IR) and collinear divergences. The possibility that, for example, a boson can split into a
pair of virtual particles, gives rise to the UV divergences in loop diagrams. Since the virtual
particles need not be on shell, the square of their four momentum does not have to equal
their mass squared. Thus, although the energy and momenta at each of the vertices has
to be conserved, the internal loop momenta of the virtual particles pi and pj can take
on any values as long as pi + pj = p, where p is the initial momenta of the boson, Fig.
2.1. This means that the integral over the virtual particles momenta can run to infinity,
which makes it logarithmically divergent. This fact was very puzzling at the time of the
construction of QFTs, as it was not known how to deal with these large quantities. It
slowed down the progress of the development of QFT and it took about 20 years to solve
this conundrum. The solution was the introduction of the concept of renormalization [31]
independently by Feynman [32] via the path integral formulation, and an operator method
developed by Schwinger [33] and Tomonaga [34]. The two methods were proved equivalent
and merged by Dyson [35]. The basic idea behind renormalization is the notion that we can
redefine the fields, and the constants of the theory. We assume that these appear in the
Lagrangian in their bare form, which is divergent, and is different from the experimentally
measured value. As such we can introduce counter terms for the coupling constants of the
10

theory, Z , and for the wave function of the external fields, Zf ield , which will cancel against
the aforementioned UV divergences. In order for this to happen it is convenient to expose
the divergences as poles. This is done with the use of a regulator. Each way of regularizing a
theory should yield the same results, however the length and the ease of the calculation are
different. The current method used in most calculations is called dimensional regularization
(DR) [36]. In DR the number of dimensions is reduced from 4 to d, where d is no longer
an integer, and is usually given by, d = 4 2, with  being a very small number. Now the
scattering amplitudes are computed in d dimensions where they are no longer divergent, but
where the UV singularities appear as poles in 1/U V . After this procedure the UV poles are
taken care of with the use of renormalization, and we are free to return back to 4 dimensions.
We should mention that while using DR some dimensionless quantities in 4 dimensions such
as the coupling constant can gain dimension, and g r g, where r has units of mass.
There is also some ambiguity as to what is absorbed with the use of renormalization once
the poles have been exposed. There are some finite terms that appear alongside the pole in
the calculations. Instead of carrying those terms around we can redefine the constants of
the theory and absorb not only the pole, but also these finite terms. In our calculation the
MS scheme is used, where one absorbs the following combination,
1
1
(4) exp(E ) ,



(2.3)

with E denoting the Euler-Mascheroni constant.


DR is very useful as it also exposes the IR and collinear divergences as poles in 1/. The
IR divergences occur when the momenta of the virtual particles tend to 0, and also when a
radiated bosons energy tends to zero. We refer to such bosons as soft. The IR divergences
cancel between these two cases. Collinear divergences arise when two massless particles are
emitted parallel to each other. This means that if we label the four momenta of the two
collinear massless particles as pc1 and pc2 , their dot product tends to zero as they become
collinear, since the space angle between them tends to zero:
pc1 .pc2 = Ec1 Ec2 p~c1 .~pc2
cos1

= Ec1 Ec2 |pc1 ||pc2 |cos Ec1 Ec2 |pc1 ||pc2 | = 0.

(2.4)

Once again we can expose the collinear singularity as a pole via DR. Then this singularity
is factored off into the PDF for an initial case collinearity, and into the FF for a final state
11

pj

p
pi

Figure 2.1: Self Energy of the Photon - an example of a loop diagram

collinearity. However in the collinear case there is one more thing that might upset the
stability of the calculation. This can be due to logarithms of the form log(Q/), which
can occur through the integration over the transverse momenta of the collinear particles.
From Eq. 1.1 we can see that the product of the strong coupling constant and the collinear
logarithm, s log(Q/), is of order one. In order to take care of these large logarithms, we
can resum them into the PDFs or FFs. This is done with the help of the Dokshitzer-GribovLipatov-Altarelli-Parisi (DGLAP) [37, 38, 39, 40] evolution equations, the solutions of which
give the scale dependence of the distribution and fragmentation functions.
In the following subsections we will go into detail as to how these procedures apply to
the study of the process at hand.

2.2

Numerical Methods - Phase Space Slicing Method

Due to the complexity of the calculation, it has to be performed numerically. As mentioned in


the Introduction we have used the two cutoff PSS method [28], which is one of the state of the
art methods for performing beyond the LO calculations. Other numerical methods include
the one cutoff PSS [41] and the dipole subtraction method [42]. At LO, we do not encounter
any singularities and so we perform the necessary integrations for the computation of the
cross section with the use of Monte Carlo integration. As we saw in section 2.1, divergences
appear at NLO. At this order there can be 3 particles in the final state. Once we have taken
care of the UV singularities using renormalization, the PSS method provides a way of dealing
with the IR and collinear divergences. To do that in PSS the 2 3 body phase space is

12

divided into a soft region, and a hard region:


Real = H + S .

(2.5)

The soft region is the region where a massless particles energy, such as the gluon tends to
zero, and thus a soft singularity can occur, while we want to keep the hard region free of soft
divergences. The division or slicing of phase space of the two regions is done with the use
of a parameter, called a soft cutoff - s . In order to separate the different regions from each
other, we need to define the Mandelstam variables [43], which are invariant under Lorentz
transformations and are given by
sij = (pi + pj )2 ,
tij = (pi pj )2 .

(2.6)

The soft region is defined as the region where the gluons energy, Eg , is less than the product

of the soft cutoff and the center of mass energy of the hard scattering, s12 ,

0 Eg s s12 /2.

(2.7)

Since there are also collinear singularities that need to be dealt with, the hard region is
further divided into a hard non-collinear and a hard collinear region,
H = HN onColl + HColl .

(2.8)

Here the collinear region is separated with the use of a collinear cutoff - c . With the use of
the Mandelstam variables the collinear region is defined as the region of phase space where
the variables are less than the product of the partonic center of mass energy and the collinear
cutoff,
sij , |tij | < c s12 .

(2.9)

We can use the numerical PSS method after the UV and IR singularities have been
exposed and canceled in n dimensions, then we can move back to 4 dimensions and integrate
over the finite region of phase space that is left. To perform this integration, we use VEGAS,
which is an adaptive multidimensional integration Monte Carlo algorithm [44].
In the next section we will show how the PSS method applies to our calculation. We
have to keep in mind that since the two cutoffs are arbitrary and are not physical, the final
13

result should be independent of them. Even though all three regions, the soft, collinear, and
hard non-collinear region depend on the cutoffs, once the numerical values for each of those
are added, the result should not depend on them.

2.3
2.3.1

+Q calculation

LO

Compton Subprocess
To LO, which is order s , there is only one hard scattering subprocess that can give a
final state photon and a heavy quark. This is the Compton subprocess,
g(p1 ) + Q(p2 ) (p3 ) + Q(p4 ),
where an initial state gluon and an initial state heavy quark scatter into a photon and a
heavy quark, with p1 + p2 = p3 + p4 . This differs from the LO for the fully inclusive direct
photon production, where there is a second subprocess that needs to be included, which is
q q g. As there are no final state heavy quarks produced by it, this process does not
pertain to our calculation.
There are two diagrams describing the Compton subprocess. These are shown in Fig.
2.2. In order to calculate the cross section we need to write the expression for the scattering
amplitude AComp :
AComp = AsComp + AuComp ,

(2.10)

where AsComp and AuComp are the amplitudes for the s and u channel respectively. With the
use of the Feynman rules from Fig. 2.2 we arrive at:
i/q
u(p2 ),g (p1 ), (p3 ),
2
q
iq 0
a
/
= ieeQ igs tij u(p4 ) 02 u(p2 ),g (p1 ), (p3 ),
q

AsComp = ieeQ igs taij u(p4 )


AuComp

(2.11)

where q = p1 + p2 and q 0 = p1 p4 . Here gs is the strong charge, e is the electromagnetic


charge, and eQ is the fractional charge of the heavy quarks, with, ec = 2/3 and eb = 1/3.
The generators of the SU (3) color group are represented by the matrices taij . Using the
following relations:
X

r
r
,p (p3 ),p (p3 ) g ,

14

(2.12)

g(p1 )

(p3 )

g(p1 )

(p3 )

Q(p2 )

Q(p4 )

Q(p2 )

Q(p4 )

Figure 2.2: Compton Scattering

us (p1 )
us (p1 ) = p/1 ,

(2.13)

s0

we get for the matrix element squared, summed over final spins and color and averaged over
initial spins and color, the result:
X

|AComp |2 =

 1  1  1 
4

And making the final substitution, s =

s
t14 
12
gs2 e2 e2Q 32
+
.
8
t14
s12

3
g2
,
4

(2.14)

e2
,
4

we get for the squared matrix element:


 1  s
X
t14 
12
|AComp |2 = 16 2 e2Q s
+
.
(2.15)
3 t14
s12

In order to get the hadronic cross section for the Compton subprocess, we convolute the
P
squared amplitude,
|AComp |2 with the PDFs and integrate over the phase space of the two
1
,
final state particles, dP S2 , as shown in Eq. 2.1, and include a flux factor, 2s
Z
Z
1 X
2
Comp = dx1 dx2 Gg (x1 , f )GQ (x2 , f )
|AComp (x1 , x2 , r )| dP S2 + 1 2. (2.16)
2s

The phase space for the two final state particles is given by the general formula:
dP S2 =

d3 p3
d3 p4
(2)4 4 (p1 + p2 p3 p4 ).
3
3
(2) 2E3 (2) 2E4

(2.17)

Making use of the fact that since the scattering is symmetric about the collision axis the
integration over the azimuthal angle yields 2, and also after some further manipulation
Eq. 2.17 can be written as:
dcos
.
16
In Eq. 2.18, is the scattering angle with respect to the beam axis.
dP S2 =

15

(2.18)

LO Fragmentation
As we can see from the above calculation, the Born term for this process is of order s .
The photon however can be produced not only in a hard scattering, but by fragmenting off
of a parton. As briefly mentioned in the Introduction, the photon fragmentation function
is introduced in order to take care of final state collinear singularities, where the photon is
emitted collinearly to one of the final state partons. One such example is the case when the
photon may be emitted collinear with the final state q, in the qQ qQ subprocess. This
case will give rise to a collinear singularity. This singular contribution can be absorbed into
the photon fragmentation function D/q ., as we pointed out in Section 2.1,
The part of the hadronic cross section for the photon to be produced by fragmentation
will look as follows:

F ragm =

XZ

dx1i dx2i dzG(x1i , f )G(x2i , f )


Fi ragm D/pi (z, F ) + (1 2).

(2.19)

In Eq. 2.19, D/pi (z, F )dz is the photon fragmentation function density, which gives the
probability for a parton pi = u, d, s, c, b, g to fragment into a photon with fraction z to
z + dz of its momenta, at a fragmentation scale F . As was pointed out in Section 2.1, after
factorizing the collinear singularity there still remain large logs coming from the integration
over the angle between the photon and the quark to which it is collinear. These logs are
resummed into the photon FFs by solving the DGLAP equations. The photon couples
directly to the quark via the electromagnetic force. Thus due to the nature of the photons
coupling to the quark, which is pointlike, the photon FFs can be derived as a solution to the
set of inhomogeneous coupled equations shown in Eq. 2.20, [14].
dD/q (z, t)

s
=
Pq (z) +
[D/q Pqq + D/g Pgq ],
dt
2
2
dD/g (z, t)
s
=
[D/q Pqg + D/g Pgg ],
dt
2

(2.20)

where t = ln(Q2 /2 ) and denotes a convolution integral as defined in Appendix A. As a


first iteration to solving the above equations, we can write Eq. 2.20, as
dD/q (z, t)
1
=
[1 + (1 z)2 ],
dt
2 2
dD/g (z, t)
= 0,
dt
16

(2.21)

Table 2.1: A list of all 2 2 LO hard-scattering subprocesses with at least one heavy quark, Q,
in the final state (where Q, can be either a heavy quark or an antiquark, where appropriate)
LO Fragmentation
Subprocesses

gg QQ
gQ gQ
qQ qQ
qQ qQ

q q QQ
QQ

QQ
QQ QQ

where we have substituted for the value of the splitting function Pq (z) = 12 [1 + (1 z)2 ].
Thus we can clearly see that D/q (z, t) is proportional to t. If we go back to Eq. 1.1,
however we see that s 1/ ln(Q/), i.e. s 1/t. As such,
D/pi

(2.22)

Therefore, another class of contributions of order s consists of 2 2 QCD subprocesses


with at least one heavy quark in the final state convoluted with the appropriate photon FF,
as listed in Eq. 2.19, since s2 D/pi is of order s . In Eq. 2.19 i runs from 1 to 7,
since there are seven subprocesses that contribute at this order. In Table 2.1 we list these
appropriate 2 2 QCD subprocesses. We have to keep in mind that we have to have at least
one heavy quark in the final state. As well as that, we should note that since we require that
the photon be isolated, i.e., the hadronic energy by which it can be surrounded is limited,
the direct photon and the heavy quark cannot be both experimentally detected if they are
collinear to each other. Due to this fact the photon cannot fragment from the heavy quark in
the case when there is only one heavy quark in the final state, i.e., gQ gQ, Qq qQ,
Q
q qQ, since we require that both the photon and heavy quark are experimentally
observed. Thus also in the case when there are two final state heavy quarks, the photon can
originate only from the unobserved heavy quark.
Example Feynman diagrams are shown in Fig. 2.3. In diagram 1) the LO QCD subprocess
is shown, with the photon radiating from the final state antiquark. The photon
gg QQ
can be a fragment of either final state quark. However, the contribution to the cross section
17

(zp3)
g(p1)

Q(p1)

g(p2)

g(p2 )

Q(p4)

(zp3)

Q(p4)

2)

1)

with
An example of Leading Order Fragmentation Contributions 1) gg QQ,
the photon fragmenting from one of the final state heavy quarks 2) gQ gQ, with the photon
fragmenting off from the final state gluon

Figure 2.3:

can only come from the heavy quark which does not emit a photon by bremsstrahlung, since
isolation requirements exclude the other case. In diagram 2) the subprocess Qg gQ is
shown. In this diagram, even though the photon can fragment from either the gluon or the
heavy quark, we do not include the case when it is produced by the heavy quark.
The complete LO hadronic cross section is the sum of the two types of contributions,
LO = Compton + LOF ragm .

(2.23)

We can see from Eqs. 2.23,2.19,2.16, that LO , depends on the renormalization scale, r ,
the fragmentation scale, f , and the factorization scale F . We want to reduce this scale
dependence and also obtain a more precise value of , by calculating the NLO contributions,
the calculation of which we describe in the next section.

2.3.2

NLO

To calculate the NLO terms we need to go up by one order of s . Thus, for this calculation
we must consider all possible subprocesses that are of order s2 , and contain a photon and
at least one heavy quark in the final state. At this order there are two types of contributions,
those are the virtual and real ones.

Virtual Corrections
The virtual corrections are calculated as the interference between the LO Born diagrams,
shown in Fig. 2.2 and the virtual diagrams in Fig. 2.4. In the virtual diagrams, just like it
18

Figure 2.4: Virtual Feynman diagrams for the Compton subprocess

is shown for Fig. 2.1, the momenta of the virtual particles are not completely determined
by the external lines. Thus, in this case the virtual momenta can take on values from zero
to infinity. After integration over the virtual momenta, two types of singularities can occur,
ultraviolet and infrared, coming from the upper and lower bound of the integration region
respectively. In order to expose the singularities as poles we evaluate the matrix elements
corresponding to the diagrams in n = 4 2 dimensions. After convoluting them with the
PDFs, the virtual correction can be written as,
Z
V irt = dx1 dx2 Gg (x1 , f )GQ (x2 , f )
V irt + (1 2),

19

(2.24)

where,

V irt =
Vsing
Vf inite
irt (1/U V , 1/IR ) +
irt .

(2.25)

The virtual corrections have been calculated in Ref. [45]. In Eq. 2.25 the singularities
are contained in
Vsing
Vf inite
is finite. The poles resulting from the UV singularities
irt and
irt
are taken care of by renormalizing the coupling constant and fields, and we are left with a
sing
dependence only on the IR singularities, 0 V irt (1/IR ). The IR poles will cancel with soft
singularities coming from the real corrections, while the remaining collinear singularities in
the real corrections part will be absorbed into the distribution and fragmentaion functions.
We proceed to describe the calculation of the real contributions to the NLO cross section.

Real Corrections
The phase space for producing a photon in association with a heavy quark increases now
that we have an extra parton emitted in the final state. And now there are seven possible
hardscattering subprocesses, which are listed in Table 2.2. An example Feynman diagram
of each is listed in Fig. 2.5. The real corrections contribution can be calculated by,
XZ
i
Real =
dx1i dx2i G(x1i , f )G(x2i , f )
Real
+ (1 2),

(2.26)

where i runs through the seven subprocesses listed in Table 2.2, and
Real = 1 |A3 |2 dP S3 .
d
2s12

(2.27)

Here the three body phase space dP S3 in n dimensions is given by:


dP S3 =

dn1 p3
dn1 p4
dn1 p5
(2)n n (p1 + p2 p3 p4 p5 ).
2p03 (2)n1 2p04 (2)n1 2p05 (2)n1

(2.28)

Since we expect soft and collinear singularities to appear at this point in the calculation, we
partition the real phase space into three regions with the use of the PSS method. These three
regions are the soft, the collinear and the finite region. The finite region does not contain
any singularities and we are free to integrate over it with the use of Monte Carlo techniques,
just as we did for the calculation of the LO cross section. It is the soft and collinear regions
that pose a bigger challenge in the real part of the NLO calculation. In what follows we
describe the calculation for these regions.
20

Table 2.2: a list of all 2 3 NLO hard-scattering subprocesses


NLO subprocesses

gg QQ
gQ gQ
Qq qQ
Q
q qQ

q q QQ

QQ QQ
QQ QQ

(p3)
Q(p4)

g(p1)

Q(p1)

(p3)
g(p5 )

5)
Q(p

g(p2 )

gg QQ
Q(p1)

g(p2)

Q(p4)
gQ gQ

(p3)
Q(p4)

q(p1)

Q(p4)
(p3)
5)
Q(p

q(p5)/
q(p5) q(p2)

q(p2)/
q (p2)

q q QQ
(p3)

Qq Qq/Q
q Q
q
Q(p1)

2)
Q(p2)/Q(p

Q(p4)

5)
Q(p5)/Q(p
QQ

QQ QQ/QQ

Figure 2.5: Example Feynman diagrams for all the possible real subprocesses

21

Soft Region
The soft, or infrared, singularities appear when the gluons energy tends to zero. There is
only one subprocess that allows such a configuration, this being gQ gQ, which is shown
in the second diagram in Fig. 2.5. In this case the final state gluons energy tends to zero,
Eg = E5 0. This region is separated from the rest of phase space with the use of Eq. 2.7.
In this limit the calculation of the matrix element is simplified by the use of the eikonal, or
double pole, approximation [46]. In it we set the gluons energy to zero when it appears in
the numerator of the matrix element. We have to retain the terms where Eg appears in the
denominator, as those terms are divergent. As such the matrix element can be written as
the product of the LO matrix element squared and an eikonal factor,
2
|Asof t (gQ gQ)|2 = g 2 2
r |AComp | eikonal .

In Eq. 2.29 the eikonal factor eikonal in general is calculated as,


X pi .pj
Cij .
eikonal =
p
.p
p
.p
i
g
j
g
i,j

(2.29)

(2.30)

In Eq. 2.30 C is a color factor and the sum runs over the number of partons from which the
gluon can be emitted, which in our case are the initial state gluon and heavy quark, and the
final state heavy quark.
There is another approximation, which we can use in this limit. The three body phase
space in Eq. 2.28 can also be greatly simplified in this region, so that it can be written as a
combination of the two-body phase space multiplied by a soft phase space factor,
dP S3sof t = dP S2 dP Ssof t .

(2.31)

With the use of Eq. 2.29 and 2.31 we can write the soft contribution to the cross section as
Z
h (1 )  42  i
s
r
sof t =
Comp eikonal dS,
(2.32)
2 (1 2) s12
with dS given by

Z
1  4  s
dS =
s12
0

s12 /2

dEg Eg12 sin12 1 d1 sin2 2 d2 .

(2.33)

After the integration we can write the soft cross section as a sum of two parts,
sing
f inite
sof t = sof
t (IR ) + sof t ,

22

(2.34)

g(p5)
g(p1)

(p3)

g(p1)
g(p1)

g(p5)
Q(p2)

Q(p4)

Q(p2)

Q(p2)

Figure 2.6: Initial collinear singularities for gQ gQ, the final state gluon is collinear to the
initial state heavy quark, or the final state gluon is collinear to the initial state gluon.

sing
f inite
where sof
t contains the soft poles and sof t is finite.

Collinear Region
The last type of singularities are the collinear singularities. These occur in the region
when two massless particles are emitted collinearly to each other. Depending on their nature
they are treated in different ways. The initial state collinear singularities are absorbed in
the PDFs. The final state ones are either absorbed in the photon FF or when summed
over experimentally unobserved states cancel due to the Kinoshita-Lee-Nauenberg theorem
[47, 48, 49]. Let us as an example analyze the collinear structure of gQ gQ. We label the
partons as shown in Fig. 2.6. Two initial state collinear singularities can occur, in the case
when the final state gluon is collinear to the initial state heavy quark, or when the final state
gluon is collinear to the initial state gluon. The final state heavy quark cannot be collinear
to the initial gluon in our case, since we require that it be experimentally detectable. In
this case the collinear singularity occurs when either t15 or t25 0, and the collinear region
in the PSS method is defined as the region where |t15 | or |t25 | < c s12 . Let us concentrate
on the case where the initial and the final state gluons are collinear. In this limit we can
approximate the momenta of particles g(p1 ), g(p01 ) and g(p5 ) as follows:
p1 = (p, 0, 0, p)
p01 = (zp + p2t /2zp, pt , 0, zp)
p5 = ((1 z)p + p2t /2(1 z)p, pt , 0, (1 z)p)
23

(2.35)

Here z is the fraction of momentum taken by the emitted particle. Using another collinear
approximation the 2 3 phase space, in Eq. 2.28 can be approximated to:
h dn1 p
i
dn1 p4
(4)
3
n n
dP S3coll =
(2)
dzdt15 [(1z)t15 ] ,

(zp
+p
p
p
)
1
2
3
4
2p03 (2)n1 2p04 (2)n1
16 2 (1 )
(2.36)
where the part in the square brackets corresponds to the 2 2 phase space of partons
g(p01 )Q(p2 ) (p3 )Q(p4 ). In this region we can also use another approximation. This is the
leading pole, or collinear, approximation of the matrix element. This approximation comes
about from the use of the collinear kinematics, shown in Eq. 2.35, in the 2 3 matrix
element, which written out in n = 4 2 dimensions takes the form:
X

|A3 (g + Q + Q + g)|2 =

|A2 (gQ Q)|2 Pgg (z, )g 2 2


r

2
,
zt15

(2.37)

where z is the momentum fraction carried by gluon 10 , as shown in Eq. 2.35, and the bar over
the sum denotes summing over final spin and color and averaging over initial spin and color.
0
Here Pgg (z, ) is the unregulated splitting function in d dimensions, Pgg (z, ) = Pgg (z) + Pgg
.
0
With Pgg (z) and Pgg
given by:

Pgg (z) = 2N

h z
i
1z
+
+ z(1 z) ,
1z
z

0
Pgg
(z) = 0.

(2.38)

Putting Eq. 2.37 and Eq. 2.36 together, we get for the collinear part of the partonic cross
section that we are discussing:
h (42 ) i (1 z) t
s
gQQg
r
d
coll,init1
= d
Comp (zs12 , t, u)Pgg (z, )
dz 15 dt15 + (1 2). (2.39)
2 (1 )
z
t15
Looking more closely at the integration over t15 in Eq. 2.39, we get:
Z c s12
1
dt15 t1
= (c s12 ) .
15

0

(2.40)

By performing the integration in Eq. 2.40 we have made the collinear singularity explicit
as a pole in . Using the approximations

1
(1)

'

(1)
,
(12)

and c ' 1  ln(c ), and also

shifting the momentum p1 p1 /z, we obtain for the hadronic cross section:
h (1 )  42  i
s
gQQg
r
dcoll,init1 = Gg/A (x1 /z)GQ/B (x2 )dComp (s12 , t, u)Pgg (z, )
2 (1 2) s12
h 1
i 1 z  dz
+ ln c
dx1 dx2 + (1 2).
(2.41)

z
z
24

Now that the singularity structure is explicit, we need to take care of it. We do that by
introducing a scale dependent PDF, using the MS convention [50], given by Eq. 2.42.
 1 h (1 )  42  i Z 1 dz
s
r
Gg/A (x, f ) = Gg/A (x) +
Pga (z)Ga/A (x/z), (2.42)
 2 (1 2) 2f
x z
where Gg/A (x) is the bare PDF, present in Eq. 2.44, and Pga (z) are the regulated splitting
functions, with Pgg (z) given by:
h z
i  11
1z
1 
Pgg (z) = 2N
+
+ z(1 z) +
N nf (1 z).
1z
z
6
3

(2.43)

dComp = Gg/A (x1 )GQ/B (x2 )d


Comp dx1 dx2

(2.44)

After substituting Eq. 2.42 in Eq. 2.44 and adding together the s pieces we get:
h (1 )  42  ih 1
c (1 z) 
s
gQQg
r
= d
Comp
dcoll,init1
+ ln
Pgg (z, )
2 (1 2) s12

z
1
i
s12 
+
+ ln 2 Pgg (z) dzGg/A (x1 /z, f )GQ/B (x2 , f )dx1 dx2 + (1 2),

f
(2.45)
 2 
 2  

s12
1
+
ln
where we have used the approximation, 1 2r = s12r
, and we have also

2f
f
 
expanded 1z
to order . Here we can explicitly see the 1 poles canceling. However the
z
integration limits over z for the two expressions differ. This is due to the fact that we require
the final state gluon to be hard, which means that E5

s s12
.
2

Since p01 + p2 = p3 + p4 and

substituting p01 ' zp1 , we get that zs12 ' s34 ,

s12 s34
(1 z)
s

E5 =
=
s12 s 12 ,
2
2
2 s12
z 1 s .

(2.46)

Due to this mismatch the final expression for the part of the cross section associated with
the initial gg collinearity is given by Eq. 2.47:
h (1 )  42  ih
s
r
0g/A (x1 , f )
G
2 (1 2) s12

i
 Asc (g gg)
1
+
+ Asc
(g

gg)
G
(x
,

)
f GQ/B (x2 , f )dx1 dx2 + (1 2).
g/A 1
0

(2.47)

gQQg
dcoll,init1
= d
Comp

25

sc
In Eq. 2.47 Asc
1 (g gg) and A0 (g gg) are given by:

11N 2nf
,
6
s12
sc
Asc
0 (g gg) = A1 (g gg) ln 2 .
f

Asc
1 (g gg) = 2N ln s +

(2.48)

0 (x1 , f ) is given by
The modified parton distribution function G
g/A
0g/A (x1 , f )
G

1s

=
x1

dz 0
G (x1 /z, f )Pgg (z),
z g/A

(2.49)

and
 1zs 
12
0

Pgg (z) = Pgg (z) ln c


Pgg
(z).
2
z f

(2.50)

For the other case, when partons Q(p2 ) and g(p5 ) are collinear, we use exactly the same
procedure, and we end up with the following expression:
h (1 )  42  ih
s
r
0Q/B (x2 , f )
G
2 (1 2) s12

i
 Asc (q qg)
1
+ Asc
+
(q

qg)
G
(x
,

)
Gg/A (x1 , f )dx1 dx2 + (1 2).
2
f
Q/B
0

(2.51)

gQQg
dcoll,init2
= d
Comp

Now that we have dealt with the initial collinear singularities let us focus on the possible final
state collinear singularities for this subprocess. Generally if we were to look at the inclusive
cross section for direct photon production, the case when the final state photon is emitted
collinearly from the final state heavy quark would occur. However since we require that both
the photon and heavy quark are experimentally observable we do not observe this singularity
in our calculation. Thus for this subprocess we only have the possibility of the final state
gluon and heavy quark to be collinear to each other. Using the leading pole approximation
for the matrix element as well as the collinear phase space approximation, that we outlined
above, we arrive at the following expression,
gQQg
dcoll,f
Comp
in = d

h (1 )  42  i Aqqg

s
1
r
+ Aqqg
+ (1 2).
0
2 (1 2) s12


(2.52)

One further point to note is the need of a jet definition in this case. Since we have divided
the phase space into a 2 2 and a 2 3 contribution, when the gluon and the heavy
quark are collinear this contributes to the two-body phase space, however when they are
26

1)

2)

(p3)

(p3)

Q(p1)

Q(p1)
Q(p4)

Q(p4)

q(p5)

q(p2)

q(p5)

q(p2)

Q(p1)

Q(p1)
Q(p4)

q(p2)

Q(p4)

q(p5)

q(p2)

(p3)

q(p5)
(p3)

3)

4)

Figure 2.7: Feynman diagrams for the subprocess qQ Qq.

almost collinear this configuration will contribute to the three-body phase space. In this
case the photons transverse momentum will be balanced by the transverse momenta of both
the heavy quark and the gluon. In order to ensure that proper cancelation between the
two and three body phase space occurs a jet definition is introduced, when now the heavy
quark jet momentum is the sum of the momenta of the two partons. And now the photons
momentum on one side is balanced by the momentum of the jet which the collinear gluon
and heavy quark have formed.
In the subprocess described above, as explained we do not observe a collinear singularity
involving the photon. However in all the rest of the NLO subprocesses listed in Table 2.2
such a singularity can occur. We also have to keep in mind that in the case when we have two
final state heavy quarks the photon can be collinear only to the unobserved heavy quark. Let
us take a closer look at the case when a photon is emitted collinearly to the final state light
quark in the subprocess, Qq Qq. There are four Feynman diagrams for this process,
listed in Fig. 2.7. The only collinear singularity involving the photon in our case, is the
27

final state collinear singularity that can occur in diagram 4. In this case the Mandelstam
variable s35 goes to zero. Here following the same procedure as the initial collinear singularity
treatment we reach Eq. 2.53,
QqQq
QqQq
dcoll,f
in =

h e2 i
q

/q (z, F )dz + (1 2).


D

(2.53)

Eq. 2.53 differs from Eq. 2.47, due to the exchange of the strong coupling constant s for the
electromagnetic one , and also due to the fact that here there are no soft collinear terms as
/q (z, F ),
we require the photon to be detectable, i.e. it cannot be soft. The modified FF D
/q , we
is given by the same expression as the modified PDF, in Eq. 2.49. To order for D
get
/q (z, F ) =
D

dy
D/ (z/y)Pq (y),
y

(2.54)

with D/ (z/y) = (1 z/y). After substituting for D/ (z/y), we get for the modified
fragmentation function,
/q (z, F ) = Pq (z).
D

(2.55)

Here Pq (z) follows the form of Eq. 2.50,



s12 
0
Pq (z) = Pq (z) ln c z(1 z) 2 Pq
(z),
F

(2.56)

with the difference that z is now in the numerator instead of the denominator, which comes
out from the fact that we are dealing with a final collinear singularity, instead of an initial
one. The splitting function is given by
1 + (1 z)2
z
0
Pq (z) = z.

Pq (z) =

(2.57)

Substituting Eq. 2.55 and 2.56 into Eq. 2.53, we get,


QqQq
QqQq
dcoll,f
in =

h e ih 
i
s12  1 + (1 z)2 
q
ln z(1 z)c 2
+ z dz + (1 2).
2
F
z

(2.58)

We treat the rest of the final state photon collinear singularities in the same fashion. This
is done by substituting QqQq , and eq in Eq. 2.58 with the appropriate hard scattering
subprocess, and quark charge. We note that there are no photon and gluon collinear
singularities and as such Pg (z) = 0.
28

1)
q(p1)

2)
Q(p4) q(p1)
(p3)

q(p2 )

5) q(p2)
Q(p

Q(p4)
(p3)
5)
Q(p

Q(p4)
q(p1)

q(p1)

(p3)

5)
Q(p
Q(p4)
q(p2)
3)

(p3) q(p2)
4)

5)
Q(p

diagrams 1) and 2) are s-channel


Figure 2.8: Feynman diagrams for the subprocess q q QQ,
diagrams, and diagrams 3) and 4) are t-channel diagrams.

There is one last final state collinear singularity that needs to be taken care of. It occurs
shown in Fig. 2.8. In the t-channel diagrams 3)
in the annihilation subprocess, q q QQ
5 ) are produced by
and 4) in Fig. 2.8, the pair of final state heavy quarks, Q(p4 ) and Q(p
gluon splitting. In this case there is a possibility of a collinear singularity as the spatial angle
between the two quarks goes to zero. If we were to calculate the inclusive cross section for a
production of a direct photon without tagging on the heavy quark, as we do here, this singular
region would be integrated over yielding a two-body contribution dependent on c . This
two-body contribution would be proportional to the subprocess qq g. The contribution
would be added to the one-loop corrections for the qq g subprocess, the poles in  would
cancel and there would be a residual c contribution to the qq g subprocess. This would
cancel against a similar contribution from qq QQ once a suitable jet definition has been
implemented in the calculation. However in this calculation the subprocess qq g does
not enter at LO as it does not contain any heavy quarks in the final state. Since we are
dealing with heavy quarks, we can address this problem by remembering that the pair of
29

heavy quarks cannot be produced unless their invariant mass is equal to or larger than 4m2Q .
Imposing this constraint on the events generated for qq QQ avoids the problem of the
uncanceled c dependence. And as such the situation where the invariant s45 0 is avoided,
by the constraint that s45 4m2Q .
After discussing all the possible collinear singularities and their treatment, we can write
the collinear cross section using the following general expression,
Z
g (x1 , f )GQ (x2 , f ) + Gg (x1 , f )G
Q (x2 , f ))
coll =
dx1 dx2 (G
Comp
XZ
/p (z, F ) + (1 2)
+
dx1i dx2i dzG(x1i , f )G(x2i , f )
Fi ragm D
i
+

i
sing
coll (1/IR ).

(2.59)

Total Cross Section


Now that we have the expressions for the different parts that contribute to the cross
section, we can write the NLO cross section as a sum of the two-body and three-body
components,
2body
3body
N LO = N
LO + N LO .

(2.60)

3body
In Eq. 2.60 N
LO is calculated as shown in Eq. 2.26, with the integration limits imposed

by the PSS requirements Eq. 2.7 and 2.9. The two body cross section is the sum of all the
different pieces that have two particles in the final state,
2body
N
LO = LO + sof t + virt + coll .

(2.61)

We note that the singular parts once added, cancel,


sing
sing
sing
virt
+ sof
t + coll = 0.

(2.62)

The cross section in Eq. 2.60, is finite, however it is not the complete NLO cross section,
as there are NLO fragmentation effects that need to be added to it. We have extended
the calculations of Ref. [26] by adding this contribution. We proceed to describe their
calculation.

30

Table 2.3: A list of all 2 3 NLO hard-scattering subprocesses, O(s3 ), with at least one
heavy quark, Q, in the final state (where Q, can be either a heavy quark or an antiquark, where
appropriate)

NLO Fragmentation
Subprocesses

gg QQg
gQ Qgg
gQ Qq q

gQ QQQ

gq qQQ
Qq Qqg
Q
q Q
qg

q q QQg
QQg

QQ
QQ QQg

NLO Fragmentation
Just as in the LO case, there are fragmentation contributions that need to be taken
into account at this order, so that we have a complete NLO calculation. At this order
we need to convolute all 2 3 QCD subprocesses of order s3 that have at least one
heavy quark in the final state with the photon FF. Thus we get, as in section 2.3.1,
O(s3 ) D/q,g s3 /s = s2 . The subprocesses contributing at this stage are listed
in Table 2.3, the matrix elements for which are computed in Ref. [51]. Let us look as an
An example Feynman
example at the first subprocess in Table 2.3, which is gg QQg.
diagram of this subprocess is shown in diagram 1) of Fig. 2.9. In order to get a final state
that has a photon from this subprocess, the photon can be produced by fragmentation from
either one of the final state partons as is shown in diagrams 2) - 4). As was shown earlier in
this section when we go beyond the LO, divergences start to plague the calculation. Thus
this subprocess can exhibit initial state collinear singularities where either the final state
heavy quark, anti-quark or gluon can be collinear to the initial state gluons. In addition to
that, we can have final state collinear singularities when the final state gluon is collinear to
either the heavy quark or antiquark, and also the two heavy quarks can be collinear to each
other. There is also the possibility of the gluons energy to go to zero, i.e for it to become
31

1)

2)
Q(p4)

g(p1 )

(zp4)

g(p1 )

g(p3)

g(p3)

g(p2 )

5)
Q(p

g(p2 )

5)
Q(p

g(p1 )

Q(p4)

g(p1)

Q(p4)
g(p3)

(zp3)
5)
Q(p

g(p2 )

g(p2)

(zp5)

4)

3)

shown in diagram 1). A photon


Feynman diagram for the subprocess gg QQg
being produced by fragmentation from one of the final states for this subprocess, diagram 2) from
the final state quark, diagram 3) from the final state gluon, and diagram 4) from the final state
anti-quark

Figure 2.9:

soft. All these singularities need to be taken care of. This is done by exactly the same
procedure that was followed for the NLO hard scattering calculation. We remember that
we included the 2 2 QCD subprocesses that contribute to the LO fragmentation part.
Thus when we calculate all the virtual corrections to the subprocesses listed in Table 2.1,
the resulting poles cancel with the soft singularities and the final state collinear singularities,
while the initial state collinear singularities are absorbed into the the PDFs.

2.4

Stability of the Calculation

Once the calculation has been performed we need to make sure that the final result is indeed
independent of the cutoffs. In Fig. 2.10, we present the independence of the cross section
on the soft cutoff s , while we have kept the value of c = 5 105 constant. In the upper
window we show the dependence of the two-body and the three-body parts on the cutoff
s . The two-body provides a negative contribution to the cross section and the three-body
32

200

2>2
2>3
tot

(pb)

100

-100

-200
20
19
18

0.0001

0.001

0.01

0.1

S = 1.96 TeV, for c = 5 105 when


s > 0.001 , and c = 1
when s < 0.001 . The upper plot shows the dependence on s of
the two-body (dashed line), and three-body (dot dashed line) contributions, and also the sum of
the two (solid line). The lower plot shows the dependence on the full NLO cross section, with the
statistical error included, = r = f = F , = pT .

Figure 2.10: s dependence of N LO (p


p bX), at
105 ,

gives a positive one, with the size of these contributions depending on the magnitude of the
cutoff. When the two contributions are added, their sum gives the total cross section, which
is independent of the cutoff as can be seen by the solid line in Fig. 2.10. We should note that
the two-body contribution contains parts which are not dependent on the cutoffs, such as
the LO cross section, but just give a constant shift to the contribution. In the lower window
we show an enlargement of the cutoff dependence of the total cross section. We can see that
the total is constant, while dropping off for the first and last values of s . This means that
33

2>2
2>3
tot

100

(pb)

50

-50

20
19
18

1e-05

0.0001

Figure 2.11: c dependence of N LO (p


p bX), at

0.001

S = 1.96 TeV, for s = 0.01. The upper


plot shows the dependence on c of the two-body (dashed line), and three-body (dot dashed line)
contributions, and also the sum of the two (solid line). The lower plot shows the dependence on
the full NLO cross section, with the statistical error included, = r = f = F , = pT .

we must have a sufficiently small value for the cutoff in order to work with the phase space
slicing method, i.e., s < 0.1. The kink in the graph comes from the change of the value of
the collinear cutoff c , from c = 5 105 to c = 1 105 , this was done in order to keep
the ratio of

c
s

from becoming large, which is what happens in the last point where again

there is an observed drop off.


In Fig. 2.11, we show the collinear cutoff versus the total cross section, while this time
keeping the soft cutoff constant, s = 0.01. As in the previous plot, we see that there is no
dependence on the cutoff for sufficiently small values of c . Again there is a dropoff for the
34

larger values of c , thus we require c < 103 .


We should note that the reason that the two cutoffs need to be kept sufficiently small is
so that the eikonal and the leading-pole approximations remain valid.
At this point since we have a stable cross section we can analyze its behavior. We present
this analysis in the next chapter.

35

CHAPTER 3
RESULTS

In this section we present the results of the NLO calculation from Chapter 2. Firstly we are
going to examine the predictions for this process at the Tevatron N LO (p
p Q). There
are two current experiments at the Tevatron that can provide measurements for this process;
CDF and D . Each of these two detectors has different capabilities in the kinematic ranges
of the particles it can observe. Here we are going to mainly focus on producing theoretical
predictions for the D experiment, as they have recently published measurements for the
inclusive differential cross sections, d3 /(dpT dy dy jet ) for both +b and +c production [52].
As such, the kinematic cuts we have applied reproduce the ones used by the D experiment
at the Tevatron in FermiLab. These kinematic cuts are as follows: the lower bound on the
transverse momentum on the photon is pT > 30 GeV. The corresponding cut for the heavy
quark is pT Q > 15 GeV. Both particles are limited to be in the central region of the detector,
and the restrictions on their rapidities are, |y | < 1, |yb | < 0.8. Here is where the versatility
of using numerical methods such as the PSS is extremely useful, as applying these cuts is
quite simple in the theoretical calculation, as it just requires changes in a few lines of code.
As well as the kinematic cuts listed above, we also need a jet definition. Since partons
hadronize, what will be experimentally measurable is the jet that the heavy quark forms. We
also need the jet definition in order to get a theoretically stable cross section as mentioned
in Section 2.3.2. We use the jet definition of the D detector [53]. Thus if two particles are
produced within a cone of radius R = 0.5 of each other they are merged into a single jet.
p
The radius R is defined as follows, R = (Q pi )2 + (Q pi )2 , with Q , pi being the
rapidities of the heavy quark and the parton that is merged in the jet, and Q , pi being the
azimuthal angles ( in the plane transverse to the proton / antiproton beam direction ) of the

36

two particles. The rapidity of a particle is defined as follows:


=

1  E + pz 
ln
,
2
E pz

(3.1)

where pz is the momentum component along the beam axis. If a particle is massless Eq. 3.1
can be reduced to:
= ln cot(/2),

(3.2)

where is the angle between the particles momentum and the beam axis, making the
rapidity very useful variable to measure experimentally.
Another important issue to be addressed is isolation, since in order to be experimentally
detectable, a photon needs to be isolated. Through this requirement charged particles are
prevented from entering the same cell in the detector as the photon, which could interfere
with the measurement of the photons energy. In order for a photon to be isolated, it
should not be surrounded by hadronic energy more than a certain fraction of its own energy,
Eh =  E in a cone of radius Riso around it [54, 55, 56]. Where the radius of the cone
p
is defined as Riso = ( Q )2 + ( Q )2 . To make sure that the cross section we are
calculating is infrared safe, we also need to impose the isolation requirement on the photons
produced through fragmentation. Thus the momentum fraction z = pT /pT parton , defined in
Section 2.1 also needs to be restricted, by making the condition,
pT j < pT ,

(3.3)

where pT j represents the momenta of the remainder of the particles that come from the
original parton after fragmentation, with pT j = (1 z)pT parton , so that pT + pT j = pT parton .
After substituting the expressions for pT and pT j in Eq. 3.3, we get,
(1 z)pT parton < zpT parton
(1 z) < z
such that, from Eq. 3.4, we get the following condition z >

(3.4)
1
.
1+

The photon isolation

requirements imposed to model the D detection of a direct photon are: R1 < 0.2, 1 < 0.04
and R2 < 0.4, 2 < 0.07 [52].
The PDFs we have used for the numerical calculation are the CTEQ6.6M set [57], unless
otherwise stated. And we have also used a 2-loop s corresponding to s (MZ ) = 0.118. For
37

the photon FFs a set of next to leading log fragmentation functions determined by L. Bourhis,
M. Fontannaz, and J.P. Guillet was used [58]. The values for the charm and bottom quark
masses used are mb = 4.2 GeV and mc = 1.25 GeV [59], and their corresponding charges
are ec = 23 |e|, and eb =

1
|e|,
3

with |e| being the charge of the electron.

After these technical details, we proceed with the numerical results for the Tevatron.
We will investigate how the inclusion of the NLO fragmentation affects the cross section, as
well as the effect of the imposition of isolation. We will also check if the idea that there is
a nonperturbative intrinsic charm component in the nucleon is testable with the currently
used cuts and detectors. In the last section of this chapter we will present results for this
process at the LHC, N LO (pp Q).

3.1

Tevatron Predictions

We start by presenting the LO differential cross section for the production of a direct photon
and a bottom quark, p
p bX as a function of the transverse momentum of the photon,

at the Tevatron center of mass energy, S = 1.96 TeV, Fig. 3.1. There we have presented
the scale dependence of the LO cross section. We have set the three different scales, the
renormalization, r , the factorization, f and the fragmentation one F , to be equal, i.e.
= r = f = F and have varied this scale through pT /2, to 2pT . We hope to decrease
the scale dependence in Fig. 3.1 with the inclusion of the next order in s . We know from the
renormalization group equation that any physical observable, in our case, is independent
of the scale, :
2

d
(Q2 /2 , s ) 0.
2
d

(3.5)

However since is only calculable via perturbation theory, there will be a residual scale
dependence coming from the truncation of the perturbation series, but this scale dependence
will decrease with the inclusion of each following term in the series. And in our case part of
the NLO scale dependence will cancel against the LO scale dependence, thereby reducing it.
Next we proceed to compare the NLO and the LO differential cross sections, which are
both shown in Fig. 3.2, solid line and dashed line respectively. The most striking feature
in Fig. 3.2 is the substantial increase in the difference between the LO and NLO curves
with growing pT . In order to ensure that this trend is not caused by the bottom quarks
charge or PDF, we present the corresponding graph for the production of a photon and a
38

p+p-> +b+X
S =1.96 TeV

LO =pT/2

d/dpT(pb/GeV)

LO =pT
LO =2pT

0.01

0.0001

100

50

150

200

pT(GeV)
Figure 3.1: The LO differential cross section at the Tevatron,
d/dpT for the production of a

direct photon and a bottom quark as a function of pT for S = 1.96 TeV at LO, with the scale
dependence shown, where the three different scales have been set to be equal = r = f = F ,
= pT (solid line), = pT /2 (dashed line), = 2pT (dotted line).

charm quark in Fig. 3.3. Where again we see the same trend - the difference between the
NLO curve and LO curve increases as the transverse momentum of the photon increases.
Generally the ratio between the two orders, is expected to stay of the same order. Thus, it
is essential to figure out where this difference is coming from.
In order to understand the origin of this effect it is necessary to take the process apart
and investigate how the different subprocesses listed in Table 2.2 contribute to the total
cross section. To perform this decomposition we have to make sure that all the singularities
described in Section 2.3.2 are canceled. This calculation is eased through the use of the
39

p+p-> +b+X
S =1.96 TeV

d/dpT(pb/GeV)

NLO
LO

0.01

0.0001

50

100

150

200

pT(GeV)
Figure 3.2: The differential cross section at the Tevatron,
d/dpT for the production of a direct

photon and a bottom quark as a function of pT for S = 1.96 TeV, at NLO (solid line), and at
LO (dashed line), = pT .

PSS method, where we have separated the singularities corresponding to each subprocess,
by placing appropriate switches in our code. How each of the subprocesses add to the cross
section is shown in Fig. 3.4.
In Fig. 3.4, the solid curve represents the NLO differential cross section, and the rest
of the curves show the contributions to the total of the different subprocesses. The dashed
It is apparent from Fig. 3.4 that the
curve shows the annihilation subprocess q q QQ.
effect shown in Fig. 3.2 is driven by it. It overtakes the Compton contribution (dash dot
dotted line, and dot dashed line) and starts dominating the cross section at pT 70 GeV.

40

p+p-> +c+X
S =1.96 TeV

NLO
LO

d/dpT(pb/GeV)

0.01

0.0001

50

100

150

200

pT(GeV)
Figure 3.3: The differential cross section at the Tevatron,
d/dpT for the production of a direct

photon and a charm quark as a function of pT for S = 1.96 TeV, at NLO (solid line), and at LO
(dashed line), = pT .

are added together (dash dot dotted line), since the


Here the LO subprocess and gg QQ
is negative. This negative contribution is what remains after
net contribution of gg QQ
the appropriate collinear terms are subtracted, when using the phase space slicing method.
The gQ gQ (dot dashed line) and Qq qQ, and Q
q qQ (dotted line) subprocesses
contribute to the cross section about equally, with gQ gQ prevailing over Qq qQ
/Q
q qQ at small pT , where the gluon PDF is larger than the light quark PDF, and
then at large pT , Qq qQ, Q
q qQ takes over when the light quark PDFs become
QQ,
and QQ QQ are added together (dot
larger than the gluon PDFs. The QQ

41

p+p-> +b+X
S =1.96 TeV

NLO
qq->QQ
qQ->qQ ; qQ->qQ
gQ->gQ
LO+gg->QQ
QQ->QQ ; QQ->QQ

d/dpT(pb/GeV)

0.01

0.0001

1e-06

1e-08

50

100

150

200

pT(GeV)
Figure 3.4: Contributions of the different subprocesses to the differential cross section, NLO (solid
(dashed line), Qq qQ, and Q
line), annihilation q q QQ
q qQ(dotted line), gQ gQ

and QQ QQ (dot dash


(dot dashed line), gg QQ+LO (dash dot dotted line), QQ QQ,
dotted line), = pT .

dash dotted line), since their role is almost negligible, as the heavy quark PDFs are much
smaller than the light quark and gluon PDFs.
Now let us turn our attention to the annihilation subprocess, since it drives the cross
sections behavior at high pT . We recall from Chapter 2 that there are two channels through
which the annihilation process can be produced, with the corresponding Feynman diagrams
shown in Fig. 2.8. Since the photon couples to the final state heavy quarks in the s-channel,
diagrams 1 and 2, these diagrams are proportional to the heavy quark charge squared e2Q .
Whereas in the t-channel, diagrams 3 and 4, the photon couples to the initial state light
42

p+p
-> +Q+X
S =1.96 TeV

charm NLO
bottom NLO
charm LO
bottom LO

d/dpT(pb/GeV)

0.01

0.0001

50

100

150

200

pT(GeV)
Figure 3.5: A comparison between the differential cross sections, d/dpT for the production of a
direct photon and a bottom quark, and that of a direct photon plus a charm quark at NLO and LO,
charm at NLO (solid line), bottom at NLO (dashed line), charm at LO (dot dashed line), bottom
at LO (dotted line), = pT .

quarks, and thus these diagrams are proportional to the light quarks charge squared, e2q . The
t-channel diagrams begin to dominate as pT grows. In this configuration the photon goes
to one side and the two heavy quarks go to the opposite side, i.e., the photons momentum
is balanced by the sum of the two heavy quarks momenta. Thus in this case the photon
is not in the vicinity of the heavy quarks. Through the s-channel production the photon is
radiated by a heavy quark, and in this case the photon is more likely to be in the proximity
of the heavy quark, in this case this production channel is suppressed due to the fact that
we require the photon to be isolated. Thus at high pT the cross section is largely driven by
43

p+p
-> +Q+X
S =1.96 TeV

charm/bottom ratio

NLO
LO
6

50

100

150

200

pT(GeV)
Figure 3.6: The ratio of the charm and bottom differential cross sections versus pT , at NLO
(solid line) and at LO (dashed line), = pT .

a process that is independent of the heavy quarks charge and also their PDF, since we only
have light quarks in the initial state. As this is the driving subprocess for both the bottom
and charm cross sections, we can expect the difference between the two to decrease as pT
increases.
This indeed is the case as can be seen from Fig. 3.5. There the NLO differential cross
sections for both the charm quark (solid line) and the bottom quark (dashed line) are shown.
As predicted the two curves tend to come closer to each other as the value of the transverse
momentum increases. In Fig. 3.5 we have also shown the curves for the LO differential
cross section for charm (dot-dashed) and for bottom (dotted). The difference between these

44

p+p-> +b+X
S =1.96 TeV

d/dpT(pb/GeV)

=pT/2
=pT

=2*pT

0.01

0.0001

50

100

150

200

pT(GeV)
Figure 3.7: Scale dependence of the NLO differential cross section, d/dpT for the production
of a direct photon and a bottom quark, where the three different scales have been set to be equal
= r = f = F , = pT (solid line), = pT /2 (dashed line), = 2pT (dotted line).

curves however, stays about constant. In Fig. 3.6, we also present the ratio of the NLO
charm and bottom differential cross sections (solid line), i.e.
corresponding ratio for the LO (dashed line),

LO (p
pcX)
LO (p
pbX)

N LO (p
pcX)
,
N LO (p
pbX)

as well as the

. The ratio of the LO cross sections

stays almost constant since the main contribution to the LO cross section comes from the
Compton subprocess. The difference between the charm and bottom LO cross sections arises
from the difference in the charges of the charm and bottom quarks and the relative sizes of
the heavy quark PDFs. The ratio of the two LO cross sections depends on the ratio of the
charm and bottom quark charges squared which is e2c /e2b = 4, and is driven up from that
45

value to about 7 due to the fact that the charm PDF is larger than the bottom PDF.
From Fig. 3.6 we can also see the NLO ratio decrease and tend to one as pT increases, as
expected.
Another consequence of the large dependence of the cross section on the annihilation
subprocess is its scale dependence. Since there is no Born term which involves a q q initial
state, the contributions from the annihilation subprocess start at O(s2 ). As such, the
typical scale compensation between the LO and NLO contributions for this subprocess is
missing and the annihilation subprocess can be thought of as a leading order contribution.
In Fig. 3.7 we show the scale dependence of the cross section, where just like in Fig. 3.1 we
have varied the scale from pT /2 to 2pT . And we can see that unfortunately the dependence
on the scale increases at large pT , where the annihilation subprocess starts to dominate.
The scale dependence increases due to the fact that there are no higher order corrections for
the annihilation process, whose scale dependence will cancel against the annihilations scale
dependence. Thus in order for the scale dependence to decrease we need to add corrections
subprocess.
of order s3 to the q q QQ

3.1.1

Photon Isolation and NLO Fragmentation

As we have seen in Chapter 2, the NLO fragmentation effects need to be included for a
complete NLO cross section. In this section we are going to investigate what effect their
inclusion will have upon the cross section, and also how the imposition of isolation affects
this contribution. As mentioned previously a photon needs to be isolated in order to give a
clear signal at a detector.
In Fig. 3.8 we show the comparison between the differential cross section with the
inclusion of isolation (dashed line) and without it (solid line). As expected without the
inclusion of isolation, there are more events included in the cross section, and so the solid
line lies higher than the dashed line. However with increasing pT , the two curves come close
to each other. This is due to the fact that in that region it is the annihilation subprocess
that contributes the most.
that dominates the cross section. It is the t-channel of q q QQ
And since through this channel the photon is produced such that its momentum balances the
momenta of the pair of heavy quarks, i.e., the photon is not in the vicinity of any partons,
isolation does not affect this subprocess.
Let us take a look at what effect the NLO fragmentation contributions have upon the
46

p+p-> +b+X
S =1.96 TeV

NLO no isolation
NLO isolation

d/dpT(pb/GeV)

0.01

0.0001

50

100

150

200

pT(GeV)
Figure 3.8:

Comparison between the differential cross section, d/dpT without isolation


requirements and with them, no isolation (solid line) , isolation (dashed line), = pT .

cross section, with the inclusion of isolation and without it. Fig. 3.9 shows the ratio between
the full NLO calculation and the cross section with only LO fragmentation included. The
solid line shows this ratio without the restriction of isolation. We can see that the NLO
fragmentation inclusion increases the cross section by up to 30%, at low pT . As the
transverse momentum grows the NLO fragmentation processes are overshadowed by the
annihilation subprocess and the effects of their inclusion are no longer that noticeable. The
dashed line in Fig. 3.9, shows the above ratio with the inclusion of isolation. We can see that
the NLO fragmentation contribution has now decreased to a few per cent. This is due to the
fact that the isolation requirements affect the photon which is produced by fragmentation

47

p+p-> +b+X
S =1.96 TeV

(NLO frag)/(LO Frag) -1

0.4

no isolation
isolation
0.3

0.2

0.1

100

50

150

200

pT(GeV)
Figure 3.9:

Ratio between the differential cross section d/dpT , with NLO fragmentation
contribution included and the differential cross section with just LO fragmentation included, no
isolation required (solid line), isolation (dashed line), = pT .

the strongest, as it is emitted in close proximity to the parton from which it is fragmenting.

3.1.2

Intrinsic Charm

As mentioned in Chapter 2 the PDFs dependence on x is measured experimentally, and their


dependence on the scale Q is obtained from solutions of the DGLAP evolution equations.
Yet at the moment we do not have a clear phenomenological insight about the structure of
the heavy quarks PDFs. In the CTEQ6.6M set of PDFs used in the previous section both
the charm and bottom PDFs are generated radiatively. What this means is that we assume

48

that at scales below the heavy quarks mass, mQ , there is no heavy quark content in the
nucleon i.e. the heavy quark PDF, GQ (x, ) = 0, when < mQ . Thus this is our initial
value for the PDF which we input in the DGLAP evolution equation,
dGQ/p (x, t)
s
=
[Pqq GQ/p (x, t) + Pqg Gg/p (x, t)],
dt
2

(3.6)

i.e. at threshold Eq. 3.6 takes the form,


dGQ/p (x, t)
s
=
Pqg Gg/p (x, t).
dt
2

(3.7)

From Eq. 3.7 we can see that the heavy quark PDF depends solely on the gluons PDF. This
however does not need to be the case, and there is a possibility that there is an intrinsic nonperturbative heavy quark component in the nucleon. The models that study this possibility
[60] concentrate on the charm component of the nucleon, as this component is expected to
be much larger than that for the bottom quark.
The PDFs modeling this possibility are the CTEQ6.6C set of PDFs [60]. There are two
models studied in Ref. [60], these are the BHPS model and a sea-like model. The BHPS
model is a light-cone model [61, 62]. In these types of models the intrinsic charm appears
at large x. Since x

pT

,
S

we expect that the use of this PDF will affect the cross section at

larger momenta. The other model assumes that the shape of the charm distribution follows
that of the light sea quarks, i.e. Gc (x) = Gc(x) Gu (x) + Gd(x).
The difference between these three cases is shown in Fig. 3.10, where the solid curve
shows the CTEQ6.6M or radiatively generated charm scenario, the dashed curve is the
CTEQ6.6C2, or BHPS model, and the dotted curve is CTEQ6.6C4 PDF or the sea-like
model. As mentioned the difference between the BHPS distribution and the radiatively
generated case are most noticeable at large x, whereas the sea-like model is about equally
larger than the CTEQ6.6M PDF at all values of x. The effect of these different PDFs on the
cross section can be seen from Fig. 3.11. The dotted curve shows the cross section generated
with the use of the sea-like intrinsic charm PDF, and it is larger than the solid curve or the
one generated with the use of the radiatively generated charm by about the same amount
at all values of pT . The difference between the radiatively generated charm cross section
and the BHPS charm however is not great at small transverse momentum, but it increases
at large pT , as is expected given the differences between the CTEQ6.6M and CTEQ6.6C2
PDFs at large x.
49

Charm PDFs
Q=40 GeV

CTEQ6.6M
CTEQ6.6C2
CTEQ6.6C4

xc(x,Q)

0.01

0.0001

1e-06

1e-08
0.001

0.01

0.1

Figure 3.10: Comparison between the three different charm PDFs at scale Q = 40 GeV,
CTEQ6.6M (solid line), BHPS or CTEQ6.6C2 (dashed line), sea-like or CTEQ6.6C4 (dotted line).

3.2

LHC Predictions

At this time the future direction of high energy physics will be determined from the
observations that will come from the LHC, at CERN. As such it is imperative to have a
clear understanding of what the Standard Model (SM) processes observed at the Tevatron
are going to look like at the LHC center of mass energies. These very processes will provide
important means of calibrating and understanding the detectors, and ultimately, are likely
to provide significant backgrounds to new physics signals.
Here we present the differential cross section versus the transverse momentum of the
photon at the LHC, and observe how the seven fold increase in c.m energy and also the
substitution of the anti-proton collision beam by a proton beam affects the cross section.
50

p+p-> +c+X
S =1.96 TeV

100

d/dpT(pb/GeV)

CTEQ6.6M
CTEQ6.6C2 BHPS
CTEQ6.6C4 Sea-like model
1

0.01

0.0001

50

100

150

200

pT(GeV)
Figure 3.11: The differential cross section, d/dpT , for the production of a direct photon and
a charm quark for the three different PDF cases, CTEQ6.6M (solid line), BHPS or CTEQ6.6C2
(dashed line), sea-like or CTEQ6.6C4 (dotted line), = pT .

Since it is unlikely that we could apprehend the kinematic cuts that will be applied at the
LHC, we have kept the kinematic cuts that we have used for the D case. The NLO and LO
cross sections are both shown in Fig. 3.12. It is apparent from Fig. 3.12 that the increase of
the difference between the NLO (solid line) and the LO (dashed line) grows much less rapidly
with increasing pT than was the case for the Tevatron. We also show the K-factor for this
case, in Fig. 3.13, which is defined as the ratio between the NLO and LO cross section,
N LO (ppbX)
.
LO (ppbX)

From Fig. 3.13 we can see that the K-factor stays stable and is around 2.

In order to understand the difference between the LHC and the Tevatron curves, we will

51

p+p
-> +b+X
S =14 TeV

NLO
LO

d/dpT(pb/GeV)

10

0.1

0.01

100

50

150

pT(GeV)
Figure 3.12: The differential cross section versus the transverse momentum of the photon d/dpT
for
the production of a direct photon and a bottom quark at LHC center of mass energies,
S = 14 TeV, NLO (solid line), LO (dashed line), = pT .

again look at how the different subprocesses affect the cross section. The contributions of
the different parts contributing to the LHC cross section are shown in Fig. 3.14. From Fig.
3.14 it can be seen that the annihilation subprocess no longer drives the cross section at high
pT , and now it is the LO, and the gQ gQ subprocesses that are the most prominent.
These differences come about for two reasons. As the LHC collides two beams of protons,
instead of the proton and antiproton beams at Fermilab, there are no longer any valence
light antiquarks present. Hence, the relative contribution of the annihilation subprocess
is decreased. Also, because the LHC will ultimately operate at a center of mass energy

which is about seven times larger than that of the Tevatron, lower values of x pT / s are
52

p+p
-> +b+X
S =14 TeV

K-factor

2.5

1.5

100

50

150

pT(GeV)
Figure 3.13: K factor, or the ratio of the NLO to the LO differential cross section for
pp bX at S = 14 TeV, = pT .

probed at the LHC. For the kinematic region shown in Fig. 3.14 the gluon PDF is dominant,
accounting for the continued importance of the gQ initiated subprocesses.
An interesting consequence of this pattern of subprocess contributions is that the
dominant parts are all proportional to the heavy quark PDFs. Such was not the case for the
Tevatron curves, except for the low end of the pT range. Accordingly, heavy quark + photon
measurements at the LHC will have the potential to provide important cross checks on the
perturbatively calculated heavy quark PDFs, as discussed in section 3.1.2. These PDFs are
likely to provide large contributions to other physics signals, either standard model or new
physics, and such checks will be an important part of the search for new physics.

53

p+p
-> +b+X
S =14 TeV

d/dpT(pb/GeV)

100

NLO
qq->QQ
qQ->qQ
gQ->gQ
LO+gg->QQ
QQ->QQ

0.01

0.0001

100

50

150

pT(GeV)
Figure 3.14: Contributions
of the different subprocesses to the differential cross section, d/dpT

(dashed line), Qq qQ,


for pp bX at S = 14 TeV, NLO (solid line), annihilation q q QQ

and Q
q qQ (dotted line), gQ gQ (dot dashed line), gg QQ+LO (dash dot dotted line),
QQ,
and QQ QQ (dot dash dotted line), = pT .
QQ

In Fig. 3.15 we show the scale dependence for both the NLO and LO differential cross
sections. In this case unlike the case for the Tevatron, Fig. 3.7, the NLO scale dependence
decreases and the three curves representing the NLO cross section, the solid, the dashed and
the dotted are almost exactly the same.

54

p+p
-> +b+X
S =14 TeV

100

NLO =pT
NLO =2pT

d/dpT(pb/GeV)

NLO =pT/2
LO =pT
LO =2pT
LO =pT/2

0.01

100

50

150

200

pT(GeV)
Figure 3.15: Scale dependence of the differential cross section, d/dpT for the production of a
direct photon and a bottom quark at the LHC, where the three different scales have been set to be
equal = r = f = F ,for the NLO cross section = pT (solid line), = 2pT (dashed line),
= pT /2 (dotted line) and for the LO cross section = pT (dot dashed line), = 2pT (dashed
dot dot line), = pT /2 (dashed dashed dot line).

55

CHAPTER 4
COMPARISON TO DATA

Surmising how the world works can start with a thought and lead to complex and intricate
theories. In order to differentiate those theories that are just conjectures from those that
describe reality the constraints of experimental data are needed. From this comparison we
can also further investigate the viable theories and be able to make future predictions for
them. In this chapter we are going to present one such comparison between data and theory
for this process, and see what can be learned from it.
Until recently there was no available data for the production of a direct photon and
a heavy quark. In 2005 a measurement with a few points was presented by the CDF
collaboration [63]. However these data points were not enough for an in depth comparison,
as their error bars are quite large, and also their integrated luminosity is only 67 pb1 . The
integrated luminosity over time characterizes the amount of accumulated data in that time
period. The luminosity is given by Eq. 4.1:
L=

1 dN
,
dt

(4.1)

where N is the number of events, and is the measured cross section. As such the unit for
luminosity is pb1 , with 1f b1 = 103 pb1 . Recently as mentioned in Chapter 3 measurements
from the D collaboration have become available [52]. These measurements are based on an
integrated luminosity of 1 f b1 , and the kinematical cuts applied to them are those presented
in Chapter 3.
In Ref. [52] the differential cross section with respect to the transverse momentum of the
photon has been measured for direct photon and bottom production, and for direct photon
and charm production. The measurements in Ref. [52] are separated out into two rapidity
regions. These are region one, where the product of the rapidities is positive, Q > 0,
56

i.e., the two rapidities are simultaneously either positive or negative, > 0 and Q > 0, or
< 0 and Q < 0, and region two where the product of the rapidities is negative Q < 0,
i.e., one rapidity is positive and the other negative, > 0 and Q < 0 or < 0 and
Q > 0. This division of the rapidity intervals is done in order to probe a different range
in the partons momentum fraction x [64]. For example if we only have two back to back
particles produced in an event, for the two momentum fractions x1 and x2 , we have,
pT
x1 = (e + eQ ),
S
pT
+ eQ ).
(4.2)
x2 = (e
S
If we pick the value of the transverse momentum to be pT = 40 GeV, then in region one,
Q > 0 we are probing the following values of x, 0.04 < x1 < 0.1 and 0.017 < x2 < 0.04,
and in region two, Q < 0 the ranges of x is 0.03 < x1 < 0.076 and 0.028 < x2 < 0.066.
In Fig. 4.1 we present the comparison between data obtained from the D Run II, for
the production of a direct photon and a b quark, + b, and the theoretical curves we have
calculated and presented in Chapter 3, Fig. 3.2. In the graph to the left the results for the
rapidity region one are presented. The solid curve is the NLO theory prediction and the
circular dots are the data points. The graph to the right shows the comparison between
theory and data for rapidity region two, where the dashed line is the NLO theory differential
cross section, and the squares represent the data points. Again as in Section 3.1, we have
set the three different scales to be equal to one another, = r = f = F , with = pT .
And the PDFs used for the curves are the CTEQ6.6M set.
We can see that there is really good agreement in Fig. 4.1 between the experimental data
and the theoretical predictions for + b production. Let us now turn our attention to the
direct photon and charm, +c, cross section. In Fig. 4.2 we present the comparison between
theory and data for this process, where the labels on the curves and points, the scale and
PDF set are the same as in Fig. 4.1. In this case we can see that there is still really good
agreement between theory and experiment for the data points that lie below pT = 70 GeV.
However unlike the photon and b quark production in this case we do not observe the good
agreement for all the data points, and the last two experimental points in Fig. 4.2 lie above
the theory curves for both rapidity regions.
In order to figure out where this discrepancy comes from we have to bear in mind that
while the theory predictions for the charm cross section do not entirely fit the data the
57

p+p-> +b+X
S =1.96 TeV

d/dpT(pb/GeV)

Q>0

Q<0

0.01

0.0001

50

100

200
0

150

50

100

150

200

pT(GeV)
Figure 4.1: The differential cross section, d/dpT , for the production of a direct photon and a b
quark, + b. In the graph to the left, the solid line is the NLO theory curve and the circular dots
are the data points measured by the D collaboration, for region one. In the graph to the right,
the dashed line is the NLO theory curve and the squares are the data points measured by the D
collaboration, for region one.

58

p+p-> +c+X
S =1.96 TeV

Q>0

d/dpT(pb/GeV)

Q<0

0.01

0.0001

50

100

200
0

150

50

100

150

200

pT(GeV)
Figure 4.2: The differential cross section, d/dpT , for the production of a direct photon and a b
quark, + c. In the graph to the left, the solid line is the NLO theory curve and the circular dots
are the data points measured by the D collaboration, for region one. In the graph to the right,
the dashed line is the NLO theory curve and the squares are the data points measured by the D
collaboration, for region one.

59

bottom ones do. What this signifies is that if something should be adjusted in the theoretical
predictions it cannot depend on any part of the calculation that is the same for both the
charm and the bottom cross sections. Since if we have to adjust it for the charm cross
section we will have to adjust it for the bottom one as well, and as such we will lose the good
agreement that we observe for the +b cross section. Owing to this we can conclude that the
difference observed in Fig. 4.2 is not due to any part of the cross section that is dependent
on the charge of the heavy quarks, i.e., we can assume that the discrepancy does not arise
in the calculation of the partonic cross section, where the only difference in the computation
of the charm and for the bottom cross section, comes from the exchange of the magnitude of
the two quark charges. Thus the explanation is likely to come from the treatment of the long
distance physics, i.e., through differences in the distribution and fragmentation functions for
the two cases.
Before we look for an explanation of the observed discrepancy we should check to see
how large the theoretical error is. Thus in Fig. 4.3 we have shown the scale dependence of
the cross section, shown by the solid curve, = pT , the dashed line for = pT /2, and
the dotted line for = 2pT . The scale dependence can be considered as the error due to
theory, and the first three data points agree with the theory within this error. However the
last two points again lie well above these error lines. The shaded region around the curve
for = pT comes from the PDF uncertainty. The CTEQ6.6M PDF errors are given by
the average of forty PDFs, representing the allowed variations coming from the global fits.
Thus the shaded region incorporates the values of the cross section with the use of each of
these PDFs. As we can see the dependence on this uncertainty is roughly the same as the
dependence on different values of the scale.
In Fig. 4.4 we show the scale dependence, for region two as was done in Fig. 4.3 for
region one. Again the last two data points continue to lie above the data. Thus we need to
consider an alternate explanation to what causes this discrepancy.
In Section 3.1.2 we described the possibility of the existence of an intrinsic heavy quark
component to the nucleon.

We concentrated on the intrinsic charm component.

We

mentioned that there can be an intrinsic bottom contribution, but that contribution is
negligible. This is due to the fact that we start the evolution of the bottom quarks at
a scale equal to the mass of the bottom quark, which is over three times larger than the
mass of the charm quark. This larger scale does not allow the non-perturbative bottom
60

p+p-> +c+X
S =1.96 TeV

d/dpT(pb/GeV)

Q>0
=pT

=pT/2
=2pT

0.01

0.0001

100

50

150

200

pT(GeV)
Figure 4.3: Scale dependence of the differential cross section, d/dpT , for the production of a
direct photon and a c quark, + c, compared to data (circular dots), for region one. The three
different scales have been set to be equal = r = f = F , = pT (solid line), = pT /2
(dashed line), = 2pT (dotted line).

contribution to become comparable to the one coming from the gluon splitting part of the
DGLAP evolutions. And so the intrinsic bottom component, if it exists, remains insignificant
[65]. As such this is a good possibility to investigate.
In Fig. 4.5 we present the comparison between data and theory with the use of the
regular PDF set, and the ones using intrinsic charm, for the + c production in the rapidity
region one. The solid line is the NLO differential cross section acquired with the CTEQ6.6M
PDF set, the dashed line is the differential cross section calculated with the use of the BHPS
61

p+p-> +c+X
S =1.96 TeV

d/dpT(pb/GeV)

Q<0
=pT

=pT/2
=2pT

0.01

0.0001

100

50

150

200

pT(GeV)
Figure 4.4: Scale dependence of the differential cross section, d/dpT , for the production of a
direct photon and a c quark, + c, compared to data (circular dots), for region two. The three
different scales have been set to be equal = r = f = F , = pT (solid line), = pT /2
(dashed line), = 2pT (dotted line).

IC PDFs, and the dotted one used the sea-like IC model PDFs. As we saw in Fig. 3.10 and
3.11 the increase in the cross section with the use of the sea-like PDF is about constant with
respect to the transverse momenta of the photon, and this is also shown in Fig. 4.5, where
the dotted curve is almost the same distance from the solid one. Even though this line goes
through one of the last two data points, it does not go through the first three data points,
and so it does not describe the data well at all. In Fig. 3.10 and 3.11 we also investigated the
dependence of the cross section on the other intrinsic charm model, which is the BHPS one.
62

p+p-> +c+X
S =1.96 TeV

Q>0

d/dpT(pb/GeV)

CTEQ 6.6M
BHPS IC model
Sea-like IC model

0.01

0.0001

100

50

150

200

pT(GeV)
Figure 4.5: The NLO differential cross section, d/dpT , for the production of a direct photon
and a c quark, + c, with the use of CTEQ6.6M PDFs (solid line), using the BHPS intrinsic
charm PDF, (dashed line) and with the use of the sea-like model intrinsic charm PDF (dotted
line), compared to data (circular dots), region one.

We saw that the difference in the cross sections is not great at small pT , but increases with
larger transverse momentum. This is what we observe in Fig. 4.5 as well, where the dashed
curve moves away from the solid curve as the momentum increases. As such it follows the
data trend, but still remains substantially below the experimental measurements. And as
in Fig. 4.2 the last two data points lie above that value. Thus we can conclude that the
observed difference between the charm data and the theory can not be explained solely by
the existence of the intrinsic charm component used in the BHPS and sea-like PDFs. The
63

p+p-> +c+X
S =1.96 TeV

d/dpT(pb/GeV)

Q<0
CTEQ 6.6M
BHPS IC model
Sea-like IC model

0.01

0.0001

100

50

150

200

pT(GeV)
Figure 4.6: The NLO differential cross section, d/dpT , for the production of a direct photon
and a c quark, + c, with the use of CTEQ6.6M PDFs (solid line), using the BHPS intrinsic
charm PDF, (dashed line) and with the use of the sea-like model intrinsic charm PDF (dotted
line), compared to data (circular dots), region two.

cross section can be further increased with the use of the intrinsic charm BHPS model PDF if
the non perturbative value used for the initial conditions of the DGLAP equations is larger.
However an intrinsic charm component large enough to increase the theory curves by the
amount required is unlikely. Since the intrinsic heavy quark PDFs used are trying to model
the non-perturbative charm component, it is possible that there are different ways to model
this idea [66], which will be able to account for the observed data and theory difference.
In Fig. 4.6 we show the comparison between the data and the differential cross section
64

p+p-> +c+X
S =1.96 TeV

Q<0

d/dpT(pb/GeV)

CTEQ 6.6M
BHPS IC model =pT

BHPS IC model =pT/2

0.01

0.0001

100

50

150

200

pT(GeV)
Figure 4.7: Scale dependence of the NLO differential cross section, d/dpT , for the production
of a direct photon and a c quark, + c, with the use of the BHPS PDFs, = pT (dashed line),
= pT /2 (dotted line), compared to data (squares), region two.

with the use of the intrinsic charm PDFs, for region two, just as was done in Fig. 4.5. As in
region one we observe the same trend as we did in region two.
In view of the above results we cannot deduce that solely the existence of intrinsic
charm is responsible for the observed discrepancy. Presently we cannot conclude if the
disagreement can be fixed by an adjustment in the theoretical calculation or the experimental
measurement. However it might turn out to be a combination of different factors. In Fig.
4.7, we show the scale dependence of the cross section with the use of the BHPS model

65

PDFs, we see that as the scale decreases to = pT /2, the cross section increases and starts
to resemble the data points more, even though it does not pass through all of them. Thus a
combination of varying the scale and using intrinsic charm PDFs seems to describe the data
better, than just one of them.
In Section 3.1 we investigated how the different subprocesses contribute to the cross

section. For Tevatron energies we found that it is the annihilation subprocess, q q QQ


that drives the cross section. It happens to overtake the cross section at pT 70 GeV,
around the value where the difference between data and theory starts. We found out
that this subprocess at high momentum looks the same for both the charm and bottom
cross sections, as it does not depend on either the charm or bottom charge or PDF and
as a result expected the ratio between the two cross sections to decrease, however this
decrease is not experimentally observed. Going back to the calculation of this process we
described the final state collinear singularity that occurs by gluon splitting in 2.3.2. This
singularity was regulated by the requirement that the heavy quarks are massive and there is
a threshold energy needed for them to be produced. By doing so we avoided the use of heavy
quark fragmentation functions. These heavy quark FFs would describe the probability of a
heavy quark to hadronize into a heavy meson. They would also obey the DGLAP evolution
equations just like the photon FF or the PDFs. The collinear singularity described can also
be taken care of by being absorbed into the heavy quark FF. In this case, if any logarithms
of the form ln(pT /mQ ) occur they will be resummed by solving the evolution equations.
How this will affect the two cross sections and if it will consolidate the discrepancy will be
investigated in the near future.
In Fig. 4.8, 4.9, 4.10 and 4.11 we show for completeness the same graphs as described
above for the photon and bottom cross section, + b. For this case there is for both regions
really good agreement between the data and the theory. And all data points lie within the
PDF uncertainty and the scale dependence band. One thing to point out here is that the two
curves that represent the intrinsic charm PDFs cross section are very close to one another
and to the curve calculated with the regular CTEQ6.6M PDFs. This is due to the fact that
there are very few subprocesess that have an initial charm quark in this case. And as such
the dependence on the charm PDFs is minimal, therefore the use of different PDFs affects
the photon and b quark cross section almost inconceivably.

66

p+p-> +b+X
S =1.96 TeV

d/dpT(pb/GeV)

Q>0
=pT

=pT/2
=2pT

0.01

0.0001

100

50

150

200

pT(GeV)
Figure 4.8: Scale dependence of the differential cross section, d/dpT , for the production of a
direct photon and a b quark, + b, compared to data (circular dots), for region one. The three
different scales have been set to be equal = r = f = F , = pT (solid line), = pT /2
(dashed line), = 2pT (dotted line).

67

p+p-> +b+X
S =1.96 TeV

d/dpT(pb/GeV)

Q<0
=pT

=pT/2
=2pT

0.01

0.0001

100

50

150

200

pT(GeV)
Figure 4.9: Scale dependence of the differential cross section, d/dpT , for the production of a
direct photon and a b quark, + b, compared to data (circular dots), for region two. The three
different scales have been set to be equal = r = f = F , = pT (solid line), = pT /2
(dashed line), = 2pT (dotted line).

68

p+p-> +b+X
S =1.96 TeV

d/dpT(pb/GeV)

Q>0
CTEQ 6.6M
BHPS IC model
Sea-like IC model

0.01

0.0001

100

50

150

200

pT(GeV)
Figure 4.10: The NLO differential cross section, d/dpT , for the production of a direct photon
and a b quark, + b, with the use of CTEQ6.6M PDFs (solid line), using the BHPS intrinsic
charm PDF, (dashed line) and with the use of the sea-like model intrinsic charm PDF (dotted
line), compared to data (circular dots), region one, = r = f = F , = pT .

69

p+p-> +b+X
S =1.96 TeV

d/dpT(pb/GeV)

Q<0
CTEQ 6.6M
BHPS IC model
Sea-like IC model

0.01

0.0001

100

50

150

200

pT(GeV)
Figure 4.11: The NLO differential cross section, d/dpT , for the production of a direct photon
and a b quark, + b, with the use of CTEQ6.6M PDFs (solid line), using the BHPS intrinsic
charm PDF, (dashed line) and with the use of the sea-like model intrinsic charm PDF (dotted
line), compared to data (circular dots), region two, = r = f = F , = pT .

70

CHAPTER 5
MASSIVE VERSUS MASSLESS COMPARISON

In Chapter 2 we described the calculation of the NLO massless cross section for the
production of a direct photon and a heavy quark. What is meant by massless is that we set
the quark masses to zero anywhere they would appear in the calculation. We retained them
only in one case, in order to prevent a collinear singularity from occurring. As an example,
the mathematical expression corresponding to the Dirac propagator is:
i(p/ + m)
,
m2 + i

p2

(5.1)

which in the massless approximation after setting the mass to be zero it becomes,
ip/
.
p2 + i

(5.2)

Combining this along with all the other expressions where the mass has been set to zero
greatly simplifies the calculation. As such the terms that contribute to the matrix elements
in a massless calculation are significantly reduced.
In order to be able to work in the massless approximation the heavy quarks and photons
produced need to carry a transverse momentum, pT , which is few times larger than the mass
of the heavy quark mQ , i.e. pT 10 GeV. Since the lower bounds for the values of the
transverse momenta for direct photons and heavy quarks measurable at both the D and
CDF collaborations are above pT 10 GeV, the comparison with the massless calculation
that we showed in Chapters 3 and 4 is appropriate.
Here for completeness we are going to review the differences between the two approaches
and give an outline of the massive calculation, and then present a comparison between the
LO massless and the LO massive cross section. Since at the Tevatron it is the annihilation
subprocess that dominates at the NLO, and it also contributes to the LO massive cross
71

Table 5.1: Subprocesses that contribute to the LO massive calculation


LO Massive
Subprocesses

gg QQ

q q QQ

section, a comparison between the NLO massless and LO massive cross section will also be
shown.

5.1

Theory

The first step to the massless calculation was to recount the subprocesses that appear at
each order we computed. We saw that at the leading massless order, which was O(s )
there was only one hardscattering subprocess that contributed to the cross section. In the
massive calculation the LO subprocesses are different, and in fact the order in s is also
different. In the massive case at LO there are two subprocesses that contribute, these are
and q q QQ,
Table 5.1. These processes contributed at the NLO in the
the gg QQ
massless calculation, and are of order s2 , which means that the LO in this case is one order
higher in s than in the massless case. Thus the massive LO is s2 .
Even though they might look very dissimilar the massive and massless calculations are
just different ways in which we can describe the same process. If we disregard the notion that
there might be any non perturbative heavy quark components to the nucleon, we know that
the only way they come into existence in the proton or neutron, is through gluon splittings.
This is why we can assume that there are no subprocesses that have a heavy quark in the
initial state. And as such the lowest order possible subprocesses that are first to produce a
heavy quark and a photon in the final state, are the ones listed in Table 5.1.
subprocess in Fig. 5.1. In diagram 1) the two initial state
Let us look at the gg QQ
gluons interact and go into a pair of heavy quarks, with the photon being emitted from the
heavy quark. We can also draw the subprocess as in diagram 2), where one of the initial
state gluons splits into a heavy quark anti quark pair, and then the heavy quark and the
72

(p3)
g(p1)

Q(p4)

g(p1 )

g(p2)

5)
Q(p

g(p2)

1)

5)
Q(p

2)

subprocess. In diagram 1) the two initial


Feynman diagrams for the gg QQ
gluons go into a pair of heavy quarks and the photon is emitted for the final state heavy quark. In
diagram 2) one of the initial state gluons splits into a heavy quark anti quark pair.

Figure 5.1:

other initial state gluon go into a final state heavy quark and photon. We know that in the
case where the angle between the pair of heavy quarks tends to zero a collinear singularity
can occur. In this case however since we are retaining the masses of the heavy quarks this
collinear singularity is regulated and the Mandelstam variable, t25 cannot go to zero.
Nevertheless since the production of heavy quarks at high-pT has the potential to generate

logarithms of the form ln(pT /mQ ) as a result of the collinear configuration involving g QQ,
we need a way to resum these logs. As was shown in Section 2.3.2 with the use of the leading
pole approximation, in the case of a collinear singularity we can approximate the matrix
subprocess as:
element squared for the gg QQ
2 s Pqg |A(gQ Q)|2 ,
|A(gg QQ)|

(5.3)

as shown graphically in Fig. 5.2. Since we have to convolute the matrix element in Eq. 5.3,
with the initial state gluon PDFs, we come across the following configuration, s Pqg Gg . This
configuration as shown in Appendix A, is a part of the DGLAP evolution equations. Thus
we can invoke a massless heavy quark PDF, GQ , which will be a solution to the DGLAP
equations through means of which the possible large logarithms will be resummed. With the
existence of this PDF we can carry through the massless calculation, starting with the LO
Compton subprocess, Section 2.3.1.
Thus we have shown that the massless and massive calculations are basically two different
approaches to describing the same process. In the literature the massive approach is known
73

Q(p4)

g(p1 )

g(p1)
s Pqg

(p3)
g(p2)

5)
Q(p

Q(zp2)

subprocess, in its collinear limit can be written as the Compton


The gg QQ
subprocess times a splitting function.

Figure 5.2:

as the fixed flavor number scheme (FFN). The name refers to the fact that we are keeping
the number of quark flavors fixed irrespective of the center of mass energy reached at the
hard scattering. Appropriately, the massless case can be referred to as the variable flavor
number scheme (VFN). In the VFN we describe the proton as composed of flavors whose
number depends on the hard scattering energy scale. If the energy is above the threshold
for production of a given quark, i.e. s 4m2Q , then the quark is taken to be massless and as
a part of the proton. Thus the massless scheme allows us to use the conventional massless
distribution and fragmentation functions [67]. Similarly in Eq. 1.1 describing the evolution
of s the number of flavors, nf changes by one as soon as we pass the energy threshold for
a flavor production.

5.2

Numerical Results

The massive matrix elements for the subprocesses listed in Table 5.1 are obtained by crossing
the matrix elements listed in Ref. [68]. Since there are no divergences at this order, the cross
section is straightforwardly obtained by Monte Carlo integration.
In Fig. 5.3 we present the differential cross section versus the transverse momentum of
the photon, d/dpT for the production of a direct photon and a bottom quark as a function

of pT for S = 1.96 TeV, using the D kinematic cuts, just as was done in Chapter 3.
The dotted line represents the LO massive cross section, the dashed line which is the LO
massless differential cross section, and the solid line which is the NLO one, are as shown
in Fig. 3.2. We note that in Fig. 5.3 the dotted line, starts close to the LO massless
cross section, but as the transverse momentum grows it increases and tends towards the
74

p+p-> +b+X
S =1.96 TeV

d/dpT(pb/GeV)

NLO massless
LO massless
LO massive

0.01

0.0001

100

50

150

200

pT(GeV)
Figure 5.3: The differential cross section,
d/dpT for the production of a direct photon and a

bottom quark as a function of pT for S = 1.96 TeV, at massless NLO (solid line), at massless
LO (dashed line), and at massive LO (dotted line), = r = f = F , = pT .

NLO massless curve. This observation is not unexpected, as in Fig. 3.4 of Chapter 3, we
showed that the subprocess that dominates the NLO massless cross section at high pT , is the
This subprocess is one of the two subprocesses present
annihilation subprocess, q q QQ.
at the leading massless order. This being the case, having the same subprocess drive both
the NLO massless and LO massive calculation we see, as expected, that the cross sections
become similar at large values of the transfer momentum. It would also be interesting to
have a comparison between the NLO massive and massless calculations, however there is no
information available on the NLO massive cross section as of yet.
75

p+p
-> +b+X
S =14 TeV

100

d/dpT(pb/GeV)

NLO massless
LO massless
LO massive
10

0.1

0.01

100

50

150

pT(GeV)
Figure 5.4: The differential cross section,
d/dpT for the production of a direct photon and a

bottom quark as a function of pT for S = 14 TeV, at massless NLO (solid line), at massless LO
(dashed line), and at massive LO (dotted line), = r = f = F , = pT .

In Ref [69] a comparison between the LO massive and LO massless differential cross
sections for the production of a charm quark and a photon at the Tevatron is presented,
in the region below pT < 50 GeV, and there the agreement between the LO massless and
massive cases in this energy range is also observed.
In Fig. 5.4 we investigate the LO massive cross section at LHC energies. There we
present the differential cross section versus the transverse momentum of the photon, d/dpT
for the production of a direct photon and a bottom quark as a function of pT , this time

at S = 14 TeV. The dotted line represents the LO massive cross section, the dashed line
76

is the LO massless differential cross section, and the solid line is the NLO one, just as in
Fig. 5.3. Here we observe that the LO massive curve is no longer is similar to the NLO
massless curve, but in fact is quite comparable to the LO massless differential cross section.
As described in Section 3.2, due to the higher energies and the exchange of an antiproton
beam for a proton beam, the annihilation subprocess is not dominant in this case, and the
NLO massless cross section is driven by subprocesses containing initial state gluons and
heavy quarks. The dominant subprocess in the LO massive calculation in this case is the
and it is comparable to the the massless LO Compton subprocess at the LHC
gg QQ,
energies.
In summary we can conclude that the massless treatment is completely sanctioned, and
it is not necessary to retain the heavy quark masses in the entire calculation.

77

CHAPTER 6
CONCLUSION

In this thesis we have presented the NLO calculation and analysis of the associated
production of a direct photon and a heavy quark. Measurements of this process are available
from the D collaboration and should be available soon from the CDF experiment at
Fermilab. Comparing these measurements with theory provided yet another check and
confirmation of the standard model.

However one of the most important reasons for

performing this calculation is the possibility of learning more about the heavy quarks PDFs.
At this stage we do not have a certain comprehension of the heavy quark structure inside the
nucleon. The radiatively generated heavy quark PDFs that we currently use have worked
well in our presently accessible energy range. It would be enlightning to have conclusive
evidence if there is a non-perturbative heavy quark component to the nucleon, not only for
purely understanding the nucleon better, but also for producing reliable predictions for the
LHC.
At the Tevatron, due to the center of mass energies and the ready availability of both

valence quarks and antiquarks we showed that the annihilation subprocess, q q QQ,
dominates the cross section. As such unfortunately we might not be able to extract much
information about the heavy quark PDFs from this comparison, since the annihilation
subprocess does not provide any dependance on the initial state heavy quark behaviour.
After comparing the data provided by the D measurements to the NLO theory, we saw
a very good agreement for the + b cross section. The + c comparison was not as favorable,
and although there was agreement between the first few data points and theory, the last two
lie above it. The possible explanation which we investigated, that the discrepency is caused
by intrinsic charm, did not increase the theory cross section enough to fit the data. One
further possibility we need to check is if the use of heavy quark fragmentation functions will
78

help reduce the observed difference.


However things look very optimistic for the LHC. Since the colliding beams consist only
of protons, there are no longer any interacting valence antiquarks. Thus the annihilation
subprocess does not play the leading role as it does at the Tevatron. The higher center of
mass energies also increase the contributions from processes that involve initial state gluons
and heavy quarks. Once measurements are available from the LHC we can expect to learn
if our radiatively generated heavy quark scenario is indeed correct or if there is an intrinsic
charm and bottom content inside the protons and antiprotons.

79

APPENDIX A
PARTON DISTRIBUTION FUNCTIONS

The scale dependence of the parton distribution functions is given by the DGLAP equations,
which have the following form:
Z
dGqi /p (x, t)
s 1 dy
=
[Pqq (y, s )Gqi /p (x/y, t) + Pqi g (y, s )Gg/p (x/y, t)]
dt
2 x y
Z
dGg/p (x, t)
s 1 dy X
=
[
Pgq (y, s )Gqi /p (x/y, t) + Pgg (y, s )Gg/p (x/y, t)].(A.1)
dt
2 x y
i

Eq. A.1 is generally written in the shorthand notation:


dGqi /p (x, t)
s
=
[Pqq Gqi /p + Pqi g Gg/p ]
dt
2
dGg/p (x, t)
s
=
[Pgq Gqi /p + Pgg Gg/p ],
dt
2
where denotes an integral convolution,
Z 1

dudvf (u)g(v)(uv z).

(A.2)

(A.3)

In Eq.

A.1 Gqi ,g/p (x, t) are the PDFs and Ppi pj (z, s ) are the Altarelli-Parisi splitting

functions. The splitting functions can be written as a power series in the strong coupling
constant.
Ppi pj (z, s ) = PpLO
(z) + s PpNi pLO
(z) + s2 PpNi pNj LO (z) + ...
i pj
j
The LO splitting functions, [37] are listed in Eq. A.5,
h 1 + z2
i
3
LO
Pqq
(z) = CF
+ (1 z)
(1 z)+ 2
h
i
1 2
LO
Pqg
(z) =
z + (1 z)2
2
h 1 + (1 z)2 i
LO
Pgq (z) = CF
z
h
z
1z
11N 2nf i
LO
Pgg (z) = 2N
+
+ z(1 z) + (1 z)
(1 z)+
(z
6
80

(A.4)

(A.5)

Table A.1: A summary of the power of logs resummed, dependent on the order of the splitting
functions

Splitting Functions
PpLO
i pj

logs resummed
(s ln Q2 )n

PDFs
LL

PpNi pLO
j

s (s ln Q2 )n1

NLL

PpNi pNj LO

s2 (s ln Q2 )n2

NLLL

If we plug in the LO splitting functions in the DGLAP equations, A.1, we sum up logs of
the form (s ln Q2 )n , known as the leading logs (LL), where n runs from 1 to infinity. The
solutions obtained are the LL PDFs. With the use of the NLO splitting functions [70, 71] the
subleading or the Next to Leading Logs (NLL) are resummed, and with the NNLO splitting
functions, the NNLL [72, 73] are resummed, Table A.1.

81

APPENDIX B
PHASE SPACE

2 body Phase Space


The two body phase space is given by the following general equation
dP S2 =

d3 p~2
d3 p~1
(2)4 4 (p12 p1 p2 ).
(2)3 2E1 (2)3 2E2

(B.1)

Using,
4 (p12 p1 p2 ) = 3 (~p12 p~1 p~2 )(E12 E1 E2 ),

(B.2)

d3 p~1 = |~p21 |d|p1 |d,

(B.3)

p21 dp1 d
(E12 E1 E2 ).
16 2 E1 E2

(B.4)

and

we get for the 2 body phase space:


dP S2 =

Let us make the following substitution,


w = E1 + E2 ,

(B.5)

working in the center of mass frame of particles one and two, where p~1 = p~2 , and also
p
p
substituting E1 = p~21 + m21 , and E2 = p~22 + m22 , we get
q
q
2
2
w = p~1 + m1 + p~21 + m22 ,
(B.6)
and
dw = p1 dp1

82

E1 + E2
.
E1 E2

(B.7)

Putting Eq. B.5 and B.7 into Eq. B.4, we get


d p1 dw
(E12 w)
16 2 w
d
p1
=
.
16 2 E1 + E2

dP S2 =

If we work with massless particles then E1 = p~1 , and since E1 + E2 = Ec m, |~p1 | =

(B.8)

Ec m
,
2

and

using that
d = dcosd,

(B.9)

we get for the massless 2 body phase space


dcosd Ecm /2
16 2 Ecm
dcos
=
.
16

dP S2massless =

(B.10)

We can substitute cos for the variable v which runs from 0 to 1 and as such is convenient
for use in Monte Carlo integrations,
cos = 1 + 2v

(B.11)

3 body Phase Space


The three body phase space is given by the following general equation
dP S3 =

d3 p~3
d3 p~4
d3 p~5
(2)4 4 (p1 + p2 p3 p4 p5 ),
2E3 (2)3 2E4 (2)3 2E5 (2)3

(B.12)

which can be simplified to,


dP S3 =

1
d3 p~3 d3 p~4
(E1 + E2 E3 E4 E5 ).
23 (2)5 E3 E4 E5

(B.13)

Moving to the particle 4 and 5 center of mass frame, we have

after some manipulation,

s45 = E4 + E5 ,

E4 + E5
d s45 =
|~p4 |dp4 .
E4 E5
83

(B.14)

(B.15)

Thus,

In the massless case |~p4 | =


dP S3massless

1
s45 ,
2

|~p4 |
d s45

=
dp4 .
E4 E5
s45

(B.16)

and so for the three body massless phase space we get,

1
d3 p~3 1 d s45

= 3
s
s45 )
d
(E
+
E

45
1
2
3
45
2 (2)5 E3 2
s4 5
d3 p~3
1
d45
= 4
2 (2)5 E3
1
= 8 4 |~p3 |dp3 dcosd45
2

(B.17)

Making the subsitution from p3 and cos to the dimensionless variables v and w, which range
from 0 to 1 and are given by,
v =1+
w=

t13
s12

(B.18)

t23
s12 + t13

we get for the 3 body phase space,


dP S3massless =

s12 vdvdwd45
29 4

84

(B.19)

APPENDIX C
2 3 NLO MASSLESS MATRIX ELEMENTS

g(p1 )Q(p2 ) (p3 )Q(p4 )g(p5 )

|A(gQ gQ)|

 t t + s s i
CF h
N
14 25
12 45
= 64

CF
t24 + 2N

8
2
t15
 s2 + t2
t214 + s212
s245 + t225 
34
23
(C.1)
+
+
t14 s45 s12 t25 s34 s45 t23 t25 s34 t14 t23 s12
3

s2 e2Q

5)
g(p1 )g(p2 ) (p3 )Q(p4 )Q(p
5 ) can be obtained by crossing the
The matrix element for g(p1 )g(p2 ) (p3 )Q(p4 )Q(p
one for g(p1 )Q(p2 ) (p3 )Q(p4 )g(p5 ) and is,


h
 t t + t t i
N
14 25
15 24
2 = 64 3 s2 e2Q 3CF
|A(gg QQ)|
CF
s45 + 2N

64
2
s12
 s2 + s2
t2 + t215
t2 + t225 
34
35
+ 14
+ 24
(C.2)
t14 t24 t15 t25 s34 t24 s35 t25 s34 t14 s35 t15
Q(p1 )Q(p2 ) (p3 )Q(p4 )Q(p5 )

1 CF  s12
t24
t25
t15
s45
2
+
+
+

+
2N
t13 t23 t23 s34 t23 s35 t13 s35 s35 s34
t14  s212 + t224 + t215 + s245 s212 + t214 + t225 + s245
+

+
t13 s34
t14 t25
t24 t15
1 (s212 + s245 )(s12 s45 t24 t15 t25 t14 ) 

(C.3)
N
t24 t25 t14 t15

|A(QQ QQ)|2 = 64 3 s2 e2Q

85

2 ) (p3 )Q(p4 )Q(p


5)
Q(p1 )Q(p
2 ) (p3 )Q(p4 )Q(p
5 ) can be obtained by crossing the
The matrix element for Q(p1 )Q(p
one for Q(p1 )Q(p2 ) (p3 )Q(p4 )Q(p5 ) and multiplying it by two, and is,

CF  t15
s45
t25
s12
t24
2
+
+
+

+
N
t13 s35 s35 s34 t23 s35 t13 t23 t23 s34
t14  s212 + t224 + t215 + s245 t215 + t214 + t225 + t224
+

+
t13 s34
t14 t25
s12 s45
1 (t215 + t224 )(s12 s45 + t24 t15 t25 t14 ) 

(C.4)
N
s12 s45 t14 t25

|A(QQ QQ)|2 = 64 3 s2 e2Q

Q(p1 )q(p2 ) (p3 )Q(p4 )q(p5 )

|A(Qq Qq)|

= 64
+

 s2 + t2 + t2 + s2 h
 s
t24
12
12
24
15
45
eQ eq
+
+
N
t25 t14
t13 t23 t23 s34
 t 
 t i
s45 
14
25
(C.5)

+ e2Q
+ e2q
s35 s34
t13 s34
t23 s35

2CF
s2

t15
s35 t13

5)
q(p1 )
q (p2 ) (p3 )Q(p4 )Q(p
5 ) can be obtained by crossing the
The matrix element for q(p1 )
q (p2 ) (p3 )Q(p4 )Q(p
one for Q(p1 )q(p2 ) (p3 )Q(p4 )q(p5 ) and is,

 t
t25
2CF  t215 + t225 + t214 + t224 h
15
eQ eq
+
+
N
s12 s45
s35 t13 s35 t23
 s 
 s i
t24 
12
45
2
2

+ eq
+ eQ
(C.6)
s34 t23
t13 t23
s34 s35

2 = 64 3 s2
|A(q q QQ)|
+

t14
s34 t13

86

REFERENCES
[1] S.L. Glashow. Partial Symmetries of Weak Interactions. Nucl.Phys. 22:579-588, 1961.
1
[2] S. Weinberg. A Model of Leptons. Phys.Rev.Lett. 19:1264-1266, 1967. 1
[3] A. Salam. Weak and Electromagnetic Interactions. Elementary Particle Theory p. 367,
Proceedings Of The Nobel Symposium, 1968. 1
[4] M. Gell-Mann, et al. Advantages of the Color Octet Gluon Picture. Phys.Lett. B47:365368, 1973. 1
[5] http://public.web.cern.ch/public/. 1
[6] http://www.fnal.gov/. 1
[7] M. Gell-Mann. A Schematic Model of Baryons and Mesons. Phys.Lett. 8:214-215, 1964.
1
[8] Y. Neeman. Derivation of strong interactions from a gauge invariance. Nucl.Phys.
26:222-229, 1961. 1
[9] R. Feynman. Very high-energy collisions of hadrons. Phys.Rev.Lett. 23:1415-1417,
1969. 1
[10] M.E. Peskin and D.V. Schroeder. An introduction to quantum field theory. Westview
Press, 1995. 1
[11] P. Aurenche, R. Baier, M. Fontannaz, J.F. Owens and M. Werlen. The Gluon Contents
of the Nucleon Probed with Real and Virtual Photons. Nuclear Physics B297:661-696,
1988. 1
[12] P. Aurenche, J.P. Guillet, E.Pilon, M. Fontannaz and M. Werlen. A New critical study
of photon production in hadronic collisions. Phys.Rev. D73:094007, 2006. 1
[13] P. Aurenche, M. Fontannaz, J.Ph. Guillet, Bernd A. Kniehl, E. Pilon and M. Werlen. A
Critical phenomenological study of inclusive photon production in hadronic collisions.
Eur.Phys.J. C9:107-119, 1999. 1
[14] J.F. Owens. Large-Momentum-Transfer Production of Direct Photons, Jets, and
Particles. Rev.Mod.Phys. 59:465, 1987. 1, 2.3.1
87

[15] H.Baer, J.Ohnemus, J.F. Owens. Next-to-leading-logarithm calculation of direct photon


production. Phys.Rev. D42:61, 1990. 1
[16] H.Baer, J.Ohnemus, J.F. Owens. A Calculation Of The Direct Photon Plus Jet Cross
Section In The Next-To-Leading-Logarithm Approximation. Phys.Lett. B234:127,
1990. 1
[17] E.L. Berger, J. Qiu. Calculation of prompt-photon production in QCD. Phys.Rev.
D44:2002, 1991. 1
[18] L.E. Gordon, W. Vogelsang. Polarized and unpolarized isolated prompt photon
production beyond the leading order. Phys.Rev. D50:1901, 1994. 1
[19] S. Catani, M. Fontannaz, J.Ph. Guillet, E. Pilon. Cross-section of isolated prompt
photons in hadron hadron collisions. J. High Energy Phys 0205:28, 2002. 1
[20] P. Aurenche, R. Baier, M. Fontannaz and D. Schiff. Prompt Photon Production at
Large pT . Phys.Rev. D39:3275, 1989. 1
[21] J.E. Augustin, et al. Discovery of a Narrow Resonance in e+ e Annihilation.
Phys.Rev.Lett. 33:1406-1408, 1974. 1
[22] J.J. Aubert, et al. Experimental Observation of a Heavy Particle J. Phys.Rev.Lett.
33:1404-1406, 1974. 1
[23] S.W. Herb, et al. Observation of a Dimuon Resonance at 9.5 GeV in 400 GeV ProtonNucleus Collisions. Phys.Rev.Lett. 39:252-255, 1977. 1
[24] C. Amsler, et al. Particle Physics Booklet. Physics Letters B667:1, 2008. 1
[25] CDF Collaboration (Anthony Allen Affolder et al.). Searches for new physics in events
with a photon and b-quark jet at CDF. Phys.Rev. D65:052006, 2002. 1
[26] B. Bailey, E.L. Berger, L.E. Gordon. Production of a prompt photon in association with
a charm quark at next-to-leading-order in QCD. Phys.Rev. D54:1896, 1996. 1, 2.3.2
[27] E.L. Berger, L.E. Gordon. Analytic calculation of prompt photon plus associated heavy
flavor at next-to-leading order in QCD. Adv.Ser.Direct.High Energy Phys 5:1-91, 1988.
1
[28] B.W. Harris and J.F. Owens. The Two Cutoff Phase Space Slicing Method . Phys.Rev.,
D(65):094032, 2002. 1, 2, 2.2
[29] J.C. Collins, D.E. Soper, G. Sterman. Factorization for Short Distance Hadron - Hadron
Scattering. Nucl.Phys. B261:104, 1985. 2.1
[30] J.C. Collins, D.E. Soper, G. Sterman.
Phys.Rev. D54:2279, 1996. 2.1

Factorization of Hard Processes in QCD.

[31] B. Delamotte. A Hint of Renormalization. Am.J.Phys. 72:170-184, 2004. 2.1


88

[32] R. Feynman. Mathematical Formulation of the Quantum Theory of Electromagnetic


Interaction. Phys. Rev. 80:440-457, 1950. 2.1
[33] J.S. Schwinger. On gauge invariance and vacuum polarization. Phys.Rev. 82:664-679,
1951. 2.1
[34] S. Tomonaga. On a relativistically invariant formulation of the quantum theory of wave
fields. Prog.Theor.Phys. 1:27-42, 1946. 2.1
[35] Silvan S. Schweber. QED and the Men Who Made It. 1994. 2.1
[36] G. t Hooft and M. Veltman. Regularization and Renormalization of Gauge Fields.
Nucl.Phys B44:189, 1972. 2.1
[37] G. Altarelli and G. Parisi. Asymptotic Freedom in Parton Language. Nucl.Phys.
B126:298, 1977. 2.1, A
[38] V.N. Gribov and L.N. Lipatov. Deep inelastic e p scattering in perturbation theory.
Sov.J.Nucl.Phys. 15:438-450, 1972. 2.1
[39] L.N. Lipatov. The parton model and perturbation theory. Sov.J.Nucl.Phys. 20:94-102,
1975. 2.1
[40] Y.L. Dokshitzer. Calculation of the Structure Functions for Deep Inelastic Scattering and e+ e- Annihilation by Perturbation Theory in Quantum Chromodynamics.
Sov.Phys.JETP 46:641-653, 1977. 2.1
[41] W.T. Giele and E.W.Nigel Glover. Higher order corrections to jet cross-sections in e+
e- annihilation. Phys.Rev. D46:1980-2010, 1992. 2.2
[42] S. Catani and M.H. Seymour. A General algorithm for calculating jet cross-sections in
NLO QCD. Nucl.Phys. B485:291-419, 1997. 2.2
[43] S. Mandelstam. Determination of the Pion-Nucleon Scattering Amplitude from Dispersion Relations and Unitarity. General Theory. Phys. Rev. 112:1344-1360, 1958. 2.2
[44] G. Peter Lepage. A New Algorithm for Adaptive Multidimensional Integration.
Computational Physics 27:192-203, 1978. 2.2
[45] A.P. Contogouris and S. Papadopoulos. The Dominant Part Of Higher Order Corrections For The Subprocess qg q. Mod.Phys.Lett. A5:901, 1990. 2.3.2
[46] S. Catani and M. Grazzini. The soft gluon current at one loop order. Nucl.Phys.
B591:435-454, 2000. 2.3.2
[47] T. Kinoshita. Mass singularities of Feynman amplitudes. J.Math.Phys. 3:650-677, 1962.
2.3.2
[48] T.D. Lee and M. Nauenberg. Degenerate Systems and Mass Singularities. Phys.Rev.
133B:1549-1562, 1964. 2.3.2
89

[49] F. Bloch and A. Nordsieck. Note on the Radiation Field of the electron. Phys.Rev.
52:54-59, 1937. 2.3.2
[50] W.A. Bardeen, A.J. Buras, D.W. Duke, T. Muta. Deep Inelastic Scattering Beyond
the Leading Order in Asymptotically Free Gauge Theories. Phys.Rev D18:3998, 1978.
2.3.2
[51] R.K. Ellis and J.C. Sexton. QCD Radiative Corrections to Parton Parton Scattering.
Nucl.Phys. B269:445, 1986. 2.3.2
[52] D. Duggan, et al. p
Measurement of + b + X and + c + X production cross sections
in p
p colisions at (s) = 1.96 TeV. hep-ex/0901.0739, 2009. 3, 3, 4
[53] G.C. Blazey, et al. Run II Jet Physics. arXiv:hep-ex/0005012, 2000. 3
[54] E.L. Berger, X. Guo and J. Qiu. Isolated prompt photon cross-sections. arXiv:hepph/9610497, 1996. 3
[55] E.L. Berger, X. Guo and J. Qiu. Isolated prompt photon production. hep-ph/9708408,
1997. 3
[56] E.L. Berger, X. Guo and J. Qiu. Isolated prompt photon production in hadronic final
states of e+ e annihilation. Phys.Rev D54:5470-5495 , 1996. 3
[57] P.M. Nadolsky, et al. Implications of CTEQ global analysis for collider observables.
Phys.Rev. D 78:013004, 2008. 3
[58] L. Bourhis, M. Fontannaz and J.Ph. Guillet. Quark and gluon fragmentation functions
into photons. Eur. Phys. J. C2:529-537, 1998. 3
[59] W.-M. Yao, et al. Particle Physics Booklet. Journal of Physics G 33, 2006. 3
[60] J. Pumplin, H.L. Lai, and W.K. Tung. The Charm Parton Content of the Nucleon.
Phys.Rev. D 75:054029, 2007. 3.1.2, 3.1.2
[61] S.J. Brodsky, P. Hoyer and A. Sakai.
Phys.Lett.B93:451-455, 1980. 3.1.2

The Intrinsic Charm of the Proton.

[62] S.J. Brodsky. New results in light-front phenomenology. Few Body Syst.36:35-52, 2005.
3.1.2
[63] A. Gajjar (on behalf of the CDF Collaboration). Di-photon and photon + b/c
production cross sections at Ecm = 1.96- TeV. hep-ex/0505046, 2005. 4
[64] V.M. Abazov et al. Measurement of the differential cross-section
for the production of

an isolated photon with associated jet in p


p collisions at s = 1.96 TeV. Phys.Lett.
B666:435-445, 2008. 4
[65] J. Pumplin. Light-Cone Models for Intrinsic Charm and Bottom.
D73:114015, 2006. 4
90

Phys.Rev.

[66] W.K. Tung. Global QCD analysis and hadron collider physics.
A21:620-628 , 2006. 4

Int.J.Mod.Phys.

[67] J.C. Collins and W.K. Tung. Calculating Heavy Quark Distributions. Nucl.Phys.
B278:934 , 1986. 5.1
[68] R.K. Ellis and Z. Kunszt. Photoproduction and Electroproduction of Heavy Flavors
with Gluon Bremsstrahlung. Nucl.Phys. B303:653 , 1988. 5.2
[69] M. Stratmann and W. Vogelsang. Prompt photon plus charm quark production at p
p
colliders. Phys.Rev. D52:1535 , 1995. 5.2
[70] G. Curci, W. Furmanski and R. Petronzio. Evolution of Parton Densities Beyond
Leading Order: The Nonsinglet Case. Nucl.Phys. B175:27, 1980. A
[71] W. Furmanski and R. Petronzio. Singlet Parton Densities Beyond Leading Order.
Phys.Lett. B97:437, 1980. A
[72] A. Vogt, S. Moch and J.A.M. Vermaseren. The Three-Loop Splitting Functions in QCD:
The Singlet Case. Nucl.Phys. B691:129-181, 2004. A
[73] A. Vogt, S. Moch and J.A.M. Vermaseren. The Three loop splitting functions in QCD:
The Nonsinglet case. Nucl.Phys. B688:101-134, 2004. A

91

BIOGRAPHICAL SKETCH

Tzvetalina P. Stavreva
B.S. in Physics (with honors)
October 2002
Sofia University, St.Kliment Ohridski Sofia, Bulgaria.

PROFESSIONAL EXPERIENCE
Graduate Research Assistant
01/2005 present
High Energy Physics Group, Florida State University Tallahassee, Florida.
Teaching Assistant
08/2004 12/2004
Department of Chemistry, Florida State University Tallahassee, Florida.
Graduate Research Assistant
08/2003 12/2004
Institute of Molecular Biophysics, Florida State University Tallahassee, Florida.

92

PUBLICATIONS

T. Stavreva, J.F. Owens, Direct Photon Production in Association With A Heavy Quark
At Hadron Colliders, arXiv:0901.3791 [hep-ph], http://arxiv.org/abs/0901.3791, Phys.
Rev. D 79, 054017 (2009)

PRESENTATIONS

Direct Photon Production in Association with a Heavy Quark - Overview


and Comparison with Data, High Energy Seminar, Oklahoma State University, 15
January 2009
Direct Photon Production in Association with a Heavy Quark - Comparison
with Data, The Coordinated Theoretical-Experimental Project on QCD (CTEQ)
Meeting, Argonne National Laboratory, 05 December 2008
Direct Photon Production in Association with a Heavy Quark At Hadron
Colliders, DOE visit, Florida State University, 06 October 2008
Direct Photon Production in Association with a Heavy Quark, American
Physical Society April Meeting, St.Louis, Missouri, 12 April 2008
Direct Photon Production in Association with Heavy Quarks, Prospectus
Defense, Florida State University, 20 October 2006

CONFERENCES AND SCHOOLS

The Coordinated Theoretical-Experimental Project on QCD (CTEQ) Meeting, Argonne National Laboratory, 5-7 December 2008
American Physical Society April Meeting, St.Louis, Missouri, 11-15 April 2008

93

Second Annual Dirac Lectures - Cosmology: From Inflation to the Cosmic


Microwave Background, Florida State University, March 26 - 28 2008
First Annual Dirac Lectures - Twistors and Twistor Methods in Higher
Order Loop Amplitudes, Florida State University March 14 - 16 2007
The Coordinated Theoretical-Experimental Project on QCD (CTEQ) Summer School, Rhodes, Greece, 1-9 July 2006
American Physical Society April Meeting, Tampa, Florida, 16-19 April 2005
Electroparamagnetic Resonance Workshop, Cornell University, Ithaca, New
York, 6-8 August, 2004
Biophysical Society 48th Annual Meeting, Baltimore, Maryland, 14-18 February
2004
Southeastern Magnetic Resonance Conference, Tallahassee, Florida, 17-19 October 2003

94

Você também pode gostar