4.." 1. .‘a W “-7 1—.11’"! IT"; J Fir." .r' T '21:: :'I I" N I: [2.13 .. ;,;_. II‘IIpfi‘h . ,3 "TI‘ ' """.’.,‘."I" ~I ',_ .I::'I n .."’. I ‘c 5‘.- “1... ‘ ‘ ~"‘ I‘M‘ha ..' “I ,:II'II:‘*.”‘ ”{3‘ '1 III“. MIMI“ "7‘ 3“.“12“! ' t I342: 033:, It: I; «III . ' L . , ~ . I 3531" ”‘3‘“: F‘I IQI‘II M‘, I ”I-r ' IN: Filu gfng'lv‘ .. :III :' 2-4 2 WM i --L . . . WI 59-17 -L'Iv 31'1" i"~‘.I"‘.‘E~"«dI‘. .113“. .3 :III\y"n|IIffi1{€II:‘('_ :I I‘ Ifiigfl‘ I").l|\"“. ‘il' 5‘.“ IIiz' II.) I mi l ‘L'Ig “[1 m I: '10 'IIzI I I l "..'.L-I.B‘1“,1'.:1'-'fi‘:}£“l:i‘:rrfirrh§d<'% 31. l I,” IEII‘J I)“ ”..‘.IL I‘lbu-I“ ‘I‘II': .EIIQP 1.2:“ L‘II'I‘. J \IJ‘I 3“ 633?“? Tenn c 'HII "‘i“~‘:‘.'1“'..‘.!‘3 ‘E'E'I‘IT‘W‘I'I‘IIE'II’ ‘II' “ N": I??? III “"‘“ 12‘3". taEI‘IJl I I.I‘;B“‘[ fix-1| _'|I .'“ 14““ "II 1.)}: I“; .‘ ,._ !.-|II-“I'-I;,1PI:If'I‘:;‘I SH???“ “A“: . ‘l‘L‘:I”-~I‘I~‘.‘,‘|Efi1-;'I ..' {flu-‘5‘, 2””: High“? filo‘I: ”I: “II 131‘: "’I' trn‘a 5:4,; 1:1,. I . ~ . 'I‘I \I\ :iftdfl‘r‘mu‘d I}: {II} I-I ‘IL Il‘aifllm ‘ 'IIIY‘IILI"I.:'J 9'13} “ lo 1‘1!” $111." k)- y! F? é‘fl'l‘gzfiY o ' L I ' ‘19": I‘ III ’5 II": ELEIII?‘ " Iréf 1* I IL L 'I" ““I’I‘I v A .{I 333:! I‘QIIIII“ 1“.“III’H‘ 17“ If]: “inf?!“ a"? ' Is'I-t .r! : l‘nfi‘h 13‘. {‘fifim . .7. Ea‘égfii‘ . ' r’: M;- 4 III *‘ Jl-LI' I.III~I-I wings} I: “3:5“ '- II, :IILILI; I: :1; '4? a I I\- III) ' IIJEii f‘hé:- {UHF 47‘!“ “I ‘4 _ pf. \ gig“: I“! .M. a. , I I III .- ‘Py. “1‘.II‘...I. _ IV . :k.bg; -r 394%: . (I. ‘ "2 y?" . ii; ‘ :I‘hIILgII ”“11 .3? I ' ifh$1¥ ""“g ‘14:)”. ’h" I I do, "'3 ‘I’ 1:1 .. III :I. I. II {’II.‘ _, , GJIIII‘ “"t' . I‘ SIIIJ‘. 31- III I. III‘II. .I I I $18,451 “If. 0;}?123 61"153‘ ' I: I‘ in 3h' .I :5: .‘ ‘ W h": gig; y: .‘ ‘3‘; ._ I '3‘ .II II}. :A I. II: I‘M“ LI III‘I“-“‘“" ' "" ‘ III .ztwfi ’5‘“? =: . \l .‘I‘ .I"‘ :7 ‘IL " ‘ ‘ “ ."|’lc"3-~ . 319‘.“ _' I. -‘-7.,II:II'.. . "Edi \ . IIZI'JA‘IIHK‘IL 1 ' II' "13;“; fl“; no?".!‘|‘.}’, I: {III‘I mm," .ggI‘IIIIiI-‘IIEI v emI- I“ ‘ as» " (LEI: H: II I "III .,. II~.--a...I‘_ . I_.<;'- . _ "‘OI’-I~“ J -" :LL I "‘W ""l ‘ ii“ " :3‘5‘2‘|“""'H" ”mm I3'3‘I‘3‘lgim L :1 LI“? '2" ." ‘ g3. .‘ II ' ' ~"I II..- '."‘ III: .Irv.v;~,:' ' 7‘"! I ' ‘~ "III“? ' ‘ “ V. ' II I ' l‘?‘ Ivcv I. .. {III n” ' I ' 1““ '1 Ir . ... 5“] ‘L"“::.-. L! ‘j‘I-I I'VIIF‘I I» my“ ‘5‘; 3‘“ r ”‘ngrpI ‘fi- WE ‘I‘I’ILII' I W. ‘1: ”‘3’; II'Ii fin a - ~.- 2"“. ' LA...- .01. ‘12.”, fiIW: I II" I‘, III I “fl fizjg .- :v...‘ M”? ..... ;.-'-“ -. ii II'IM III vs.‘ .‘II‘ "’ {I 5:“. “v I _ ' A ‘_A -" ‘— .'.\\L.\- .2??? '3 3“-" ‘\\ . _.. 7‘: 3.I4_.-‘ THESIS Illllllll'illllllllllllllllllillllllll 31293 01716 3522 This is to certify that the dissertation entitled INCLUSIVE DIRECI‘ PHUION PRODUCTION IN THE CENTRAL AND EDWARD RAPIDITY REGIONS IN PROI’ON:ANI'IPROION COLLISIONS AT A CENTER OF MASS ENERGY OF 1800 (EV presented by Steven A. Jerger has been accepted towards fulfillment of the requirements for Ph 1) degree in maize— flewzciée. Majgr professor Date [0 /Zq [Cr] v MSU is an Afflrmatiw Action/Equal Opportunity Institution 0-12771 LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINE return on or before date due. DATE DUE DATE Due DATE DUE mo Wow-p.14 INCLUSIVE DIRECT PHOTON PRODUCTION IN THE CENTRAL AND FORWARD RAPIDITY REGIONS IN PROTON-ANTIPROTON COLLISIONS AT A CENTER OF MASS ENERGY OF 1800 GEV By Steven A. Jerger A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Physics and Astronomy 1 997 ABSTRACT INCLUSIVE DIRECT PHOTON PRODUCTION IN THE CENTRAL AND FORWARD RAPIDITY REGIONS IN PROTON-ANTIPROTON COLLISIONS AT A CENTER OF MASS ENERGY OF 1800 GEV By Steven A. Jerger A study of isolated direct photon production in proton-antiproton collisions at a center of mass energy 5/} = 1800 GeV is reported, as measured at the DO Detector at the Fermilab Tevatron. Cross sections for the central (OS Inl $0.9) and forward (1.6SI'nI $2.5) rapidity regions are presented as a function of photon ET (15 GeV S ET S 150 GeV ), and compared with a next-to—leading order QCD calculation. In the central region, the data and theory are consistent in both shape and normalization; however, in the forward region the data are consistently above the theory, especially in the ET region below ~ 30 GeV. A preliminary measurement of the correlation between the rapidity of the photon and that of the leading jet in the event shows qualitative agreement between the data and the theoretical prediction. To David and Janice Jerger iii ACKNOWLEDGEMENTS My thanks to my friends: Gian, Kate, Terry, etc... there are too many people who have helped me get through life to make a long list and risk leaving someone out. If you think your name belongs here, past or present, it does. Thanks. Thanks to Anne Keman, Steve Wimpenny, and John Ellison at UC Riverside for giving me the chance to get started down this path as an undergraduate, and to Dave, Kevin, and Mike for their teachings which helped me through grad school (and life). Thanks to Chip, Scotty, Packy, Dean, Bruce, and especially all the kids of the Waubonsie Valley and Neuqua Valley bands - you gave me another home where I could go for a little sanity or insanity, whichever I needed. Hope this explains protons &... stuff. Thanks to the direct photon group, especially Sal Fahey, who paved the road and looked out for me; and Steve Linn for reminding me I am allowed to sit at the big table with the grownups. Most of all I thank John Womersley, for his guidance and endless patience - while I doubt I could ever match it, he has always defined for me the level of excellence for which I should strive. Finally, I give my appreciation to my advisor, Bernard (1 Pope, for his wisdom and his uncanny ability to remind me of what’s really important. I will always remember a big green shamrock on my desk one stressful morning, a symbol of the unfailing support he has given me through thick and thin times for both of us. Thanks, Bernard. iv TABLE OF CONTENTS LIST OF TABLES ix LIST OF FIGURES xi 1 Introduction 1 2 Theory 3 2.1 The Standard Model ............................................... 3 2.2 QCD Interactions ................................................. 7 2.3 Direct Photon Theory ............................................. 13 2.3.1 Direct Photon Diagrams ..................... . ................ 13 2.3.2 Bremsstrahlung Component and Isolation ....................... 14 2.4 Variables for Hadron Collider Physics ................................ 17 3 Experimental Apparatus 20 3.1 Accelerator ..................................................... 20 3.2 Detector ........................................................ 23 3.2.1 Tracking ................................................. 25 3.2.2 Calorimetry ............................................... 34 3.2.3 The Muon System .......................................... 39 3.3 Data Acquisition and Processing .................................... 3.3.1 The Trigger System ........................................ 3.3.2 Event Reconstruction ....................................... 4 Event Selection and Efficiencies 4.1 Data Set and Monte Carlo .......................................... 4.1.1 Direct Photon Triggers ...................................... 4.1.2 Candidate Sample .......................................... 4.1.3 DOGEANT Monte Carlo .................................... 4.2 Acceptance ..................................................... 4.3 Cut Efficiencies .................................................. 4.3.1 Tracking Considerations ..................................... 4.3.2 Selection Cuts ............................................. 5 Background Estimation 5.1 Introduction ..................................................... 5.2 Background Estimation Techniques .................................. 5.2.1 Method .................................................. 5.2.2 Background Monte Carlo .................................... 5.2.3 EMl Calibration ........................................... 5.3 Purity Calculation ................................................ 6 Cross Sections and Theoretical Comparisons 6.1 Cross Section Calculation .......................................... vi 40 40 43 48 49 49 50 52 55 59 59 59 63 63 65 65 67 69 71 79 79 6.2 Comparison with Theory .......................................... 84 7 Photon + Jet Rapidity Correlations 93 7.1 Introduction ..................................................... 93 7.2 Rapidity Correlation Measurement ................................... 94 7.2.1 Motivation ................................................ 94 7.2.2 Data Selection ............................................. 95 7.2.3 Technique and Candidate Sample Behavior ...................... 97 7.2.4 Background Behavior ....................................... 97 7.2.5 Theoretical Prediction and Comparisons ........................ 99 7.3 Conclusions .................................................... 102 8 Conclusions 106 A Triggers and Luminosity 108 A.1 Triggering .................................................... 109 A.1.l Level 1 and Level 2 Requirements ........................... 109 A.1.2 Trigger Tum-On and Efficiency ............................. 111 A2 Integrated Luminosity Determination ............................... 117 B Tracking Considerations 121 3.1 The Tracking Efficiency ......................................... 122 B.2 Track Overlap Probabilities ....................................... 128 B.2.1 Golden Overlap Method .................................... 128 B.2.2 Z Overlap Method ........................................ 131 vii B.3 Overall Efficiency .............................................. C EMl Calibration C.1 Offset Estimation and Origins ..................................... C.2 Cross Checks of the Offset Estimate ................................ C.3 Effects of Pre-EMl Material in the Monte Carlo ....................... C.4 Final Calibration Determination ................................... C.5 Summary ..................................................... D Comparison with Run 1A Results D.1 Corrections to the 1A Analysis .................................... D.2 Effects of Changes in Background Modeling ......................... D.3 Conclusions ................................................... E Purity Fits F Monte Carlo Composites REFERENCES viii 132 133 133 135 136 140 143 144 144 147 150 152 160 173 2.1 2.2 2.3 4.1 4.2 5.1 5.2 5.3 6.1 6.2 6.3 A.1 A2 A3 B.1 B.2 B.3 LIST OF TABLES The three generations of leptons ......................................... 4 The three generations of quarks. ........................................ 4 Vector bosons and the forces they mediate. ............................... 5 Description of triggers and associated integrated luminosities. ............... 50 Number of events surviving the application of photon cuts ................... 54 Neutral meson decays to photon states ................................... 64 CC Purity points, variations, and errors. ................................. 74 EC Purity points, variations and errors ................................... 74 Central region cross section points and errors. ............................ 82 Forward region cross section points and errors. ........................... 83 Contributions to the total systematic error on the 39-42 GeV cross section bins. ...................................................... 84 Description of triggers and associated filters ............................. 109 Integrated Luminosities for the four single photon filters ................... 118 Comparison of luminosities calculated using Equation A.4 with those taken from the Production database .................................... 120 Tracking efficiencies and errors in the central and forward regions. .......... 126 Tracking efficiencies at different instantaneous luminosities. ............... 127 Overlap probabilities from golden candidate sample ....................... 129 ix B.4 Variation of the overlap probability with isolation and golden cuts, varied independently using the EM1_GIS_HIGH trigger. ....................... 130 CI Additive constants used to calibrate the Monte Carlo log(EMl/E) distributions ...................................................... 143 2.1 2.2 2.3 2.4 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 3.10 4.1 4.2 LIST OF FIGURES Diagram picturing different aspects of a QCD interaction. ................... 9 Leading order diagrams for direct photon production ....................... 13 Examples of higher order direct photon diagrams .......................... 15 Photon production through Bremsstrahlung .............................. 16 The Fermilab Tevatron Collider. ...................................... 21 The DO Detector. .................................................. 24 The D0 Tracking System ............................................. 25 Side schematic of the DO Vertex Detector. .............................. 27 Cross section of the first TRD layer. ................................... 30 End view of a section of the Central Drift Chamber ........................ 31 An exploded view of the Forward Drift Chamber .......................... 33 The DO Calorimeter (cutaway view) .................................... 34 Diagram of the layers within a calorimeter cell ............................ 36 Side view of a calorimeter quadrant. Lines of pseudorapidity are shown to demonstrate the projective tower geometry ............................... 37 H-matrix x2 distributions from test beam electrons, test beam pions, and electrons from W decays .......................................... 53 Comparison of data (points) and Monte Carlo (histograms) W electrons in the central region ................................................. 56 xi 4.3 4.4 4.5 4.6 5.1 5.2 5.3 5.4 5.5 5.6 6.1 6.2 6.3 6.4 6.5 6.6 7.1 Comparison of data (points) and Monte Carlo (histograms) W electrons in the forward region ................................................ 57 The Run 18 vertex 2 distribution ....................................... 58 Selection cut efficiencies measured using PYTHIA Monte Carlo photons, and corresponding fits. ................................................. 6O Instantaneous Luminosity profiles for the EM1_GIS (left) and EM1_GIS_HIGH (right) filters ....................................................... 61 Comparison of EMl energy fraction left by 10 GeV MC photons and Ito’s ...... 66 Log(EMllE) distributions for W electrons from data (points) and Monte Carlo (histograms) ....................................................... 7O Log(EMllE) for Monte Carlo and data between 33—39 GeV in the central region. The solid line is the MINUIT fit, which results in a measured purity of 0.467 (P2). ..................................................... 72 Purity vs. ET for the central (left) and forward (right) regions. ............... 75 Variations in the CC (left) and EC (right) purity fits, using the nominal (solid), exponential (dotted) and straight line (dashed) fit functions .................. 76 Effect of a plus 16 (dotted) and minus 16 (dashed) shift in the EC EMl offset, relative to the nominal (solid) purity curve ............................... 78 Inclusive isolated photon cross sections in the central (top) and forward (bottom) regions. .................................................. 81 (Data-Theory)/Theory for the central (top) and forward (bottom) regions ...... 86 (Data-Theory)/Theory using purity determination bins ...................... 87 Variation of the theoretical prediction with a change in scale for the central (top) and forward (bottom) regions ..................................... 89 Comparison of the theoretical prediction with CTEQ2M and CTEQ4M PDFS. ........................................................... 91 Change in the theoretical cross sections from that used in Run 1A. ........... 92 Diagram of the final state topology when colliding partons have similar (left) or very dissimilar (right) momentum fractions ............................ 94 xii 7.2 7.3 7.4 7.5 7.6 A.1 A2 A3 A4 A.5 8.1 B] B.2 C.1 C2 C3 C4 Signed If“ distribution of the leading jet for golden photon candidates. The shaded band indicates the rapidity bin of the photon; shown at left are the mean values of signed 11’“ in each bin. ................................. 98 Signed 11"" distribution of the leading jet for background (electromagnetic jet) candidates. ...................................................... 100 Signed 11"“ distribution of the leading jet in a NLL QCD prediction from Owens, et al ...................................................... 101 Mean signed If“ versus mean InYI for candidate, background, and theory samples. ........................................................ 103 Mean signed nie’ versus mean mill for candidate, background, and dijet samples with E]- between 45 and 55 GeV ............................... 104 7 GeV Level 1 Trigger efficiency for the central and forward regions. ....... 112 6 GeV Level 2 Filter efficiency for the central and forward regions .......... 113 14 GeV Level 2 Filter efficiency for the central and forward regions ......... 114 25 GeV Level 2 Filter efficiency for the central and forward regions ......... 115 40 GeV Level 2 Filter efficiency for the central and forward regions ......... 116 Matching of luminosity-corrected filters (CC only). ...................... 119 Z mass distributions before and after track requirement .................... 125 Log(EMI/E) distributions for Z electrons. Note that 1% EMl corresponds to -2.0 on this scale. ............................................... 129 Values of EC EM] offset produced by the purity fitting routine for several Er bins .................................................... 136 Offset measured from W electrons versus ET, using a) full W Monte Carlo; b) single particle Monte Carlo ....................................... 137 Fraction of energy deposited in each EM layer for several pre—EM] estimates (simple model), data, and Monte Carlo ................................. 139 Change in percent of total energy in each layer between different material estimates, and between data and Monte Carlo ............................ 140 xiii C.5 D.1 D2 D3 D4 E.1 B.2 E3 E4 E5 E6 E7 F.1 F2 R3 R4 F.5 Values of EC EMl offset produced by the purity fitting routine for EC North (top) and South (bottom) samples ..................................... 142 Ratio of the Run 1A inclusive cross sections to the Run 1B results in the central (top) and forward (bottom) regions .............................. 145 Effect of refitting the Run 1A central region Purity/Efficiency to allow for the proper maximum value at higher ET values. ............................ 147 Left: Run 1A CC purity points and resulting fit (solid line), along with the 1B CC purity fit (dashed line). Right: 1A purity points and fit (solid line) using the 1B Monte Carlo, along with the 1A fit (dashed line, same as solid line on left) .......................................................... 149 Ratio of the “corrected” Run 1A inclusive cross sections to the Run 1B results in the central (top) and forward (bottom) regions. .................. 151 CC purity fits for ET bins of 10-12, 12-15, 15-21, and 21-25 GeV. .......... 153 CC purity fits for ET bins of 25-30, 30-33, 33-39, and 39—54 GeV. .......... 154 CC purity fits for ET bins of 54-60, 60-70, 70-88, and 88-100 GeV. ......... 155 CC purity fits for the ET bin of 100-150 GeV ............................ 156 EC purity fits for ET bins of 10-13, 13-21, 21-25, and 25-30 GeV ............ 157 EC purity fits for ET bins of 30-33, 33-39, 39-54, and 54-60 GeV ............ 158 EC purity fits for ET bins of 60-75, 75—90, 90-150 GeV .................... 159 Comparison of data (points) and composite Monte Carlo (histogram) EM fraction distributions in the central region. .......................... 161 Comparison of data (points) and composite Monte Carlo (histogram) Hmatrix x2 distributions in the central region ............................ 163 Comparison of data (points) and composite Monte Carlo (histogram) isolation ET distributions in the central region. .......................... 165 Comparison of data (points) and composite Monte Carlo (histogram) EM fraction distributions in the forward region. ......................... 167 Comparison of data (points) and composite Monte Carlo (histogram) Hmatrix x2 distributions in the forward region ........................... 169 xiv F .6 Comparison of data (points) and composite Monte Carlo (histogram) isolation ET distributions in the forward region. ......................... 171 XV Chapter 1 Introduction This thesis presents a measurement of the production of direct photons in proton- antiproton collisions, using the DO Detector at the Fermilab Tevatron collider. The rate at which single, isolated photons are produced in such collisions is considered over a range of photon energies, and for two separate ranges of photon polar angle. In addition, the structure of direct photon events is considered by measuring the angular correlation between the photon and the other particles in the event. In both cases, the experimental results are compared with theoretical predictions. Direct photon measurements have historically played a role in the modeling of the Strong force which binds nuclei together (for a review of early direct photon experiments, see [1]), and this thesis continues that tradition by extending the measurement to the highest photon energies available in the world. As such, this thesis serves as an extension to and improvement upon previous DO direct photon measurements [2][3]. The structure of the thesis is as follows: in Chapter 2, an introduction to basic particle 2 physics is given, along with a brief explanation of the theory of the Strong force and of the theoretical calculation of direct photon production. Chapter 3 describes the DO Detector, while Chapter 4 describes the way photons are identified in the detector, as well as the efficiencies with which photons are identified. Chapter 5 presents a discussion of the backgrounds to the direct photon signal, and describes the method by which the level of photon signal was determined. Finally, Chapter 6 presents the results of the measurement of the direct photon production rate, while Chapter 7 presents results from a study of direct photon event structure. The results are summarized in Chapter 8. The reader will note that extensive use is made of Appendices. This is done in order to avoid excessive detail during the description of the measurement; however, it also serves to partition detailed subjects, so that the Appendices may be more useful to those who may use this thesis as a reference. Chapter 2 Theory 2.1 The Standard Model In order to set the framework to which the study of direct photons contributes, we briefly consider the known elementary particles and the forces which govern their interactions. This set of particles, their antiparticle counterparts, and the physical laws they obey are known as the Standard Model. As its name implies, the Standard Model is a stable and consistent structure which has developed over time, and which explains most of the fundamental behavior of particles and forces quite well. In the Standard Model, all matter is composed of point-like particles which have half- integer values of spin. These are classified as fermions, and can be further divided into the lepton (Table 2.1) and quark families (Table 2.2) [4]. Both the lepton and quark families can be grouped into three “generations,” each generation having identical properties except for mass. The reason for these generations is still unknown - most matter in the 4 Table 2.1: The three generations of leptons. Particle Symbol Charge Mass (MeV/cz) electron e -1 0.51 1 electron neutrino v, 0 < 5.10 E-6 muon It -1 105.6 muon neutrino v1; 0 < 0.17 tau 1: -1 1777 tau neutrino v.t 0 < 24 Table 2.2: The three generations of quarks. Particle Symbol Charge Mass (MeV/cz) up u +2/3 2-8 down d -1/3 5-15 charm c +2/3 1000-1600 strange s - 1/3 100-300 top I +2/3 180000 bottom b -1/3 4100-4500 universe appears to be composed of particles from the first generation: up and down quarks combine to form protons (uud) and neutrons (uud) that make up the nuclei of atoms; and electrons orbit the nuclei, giving atoms their chemical and material properties. Governing how these particles interact are the four known forces: the Electromagnetic, Weak, and Strong forces and the Gravitational force. Each of these forces is mediated by a spin 1 particle, classified as a boson; the bosons are listed in Table 2.3. Particles with 5 Table 2.3: Vector bosons and the forces they mediate. Boson Symbol Force Charge (M12377?) photon 'y Electromagnetic 0 0 W W Weak :|:l 80300 Z Z Weak 0 91187 gluon g Strong 0 0 graviton ? Gravity ? ? electric charge interact through the Electromagnetic force and its carrier, the photon; Weak interactions proceed through exchange of W or Z bosons. In very high energy interactions, where the mass of the W or Z becomes negligible, the Electromagnetic and Weak force are actually the same; therefore, the two forces are treated theoretically as two aspects of one Electroweak force. While the Gravitational force was known before any of the other forces, there is as yet no well-developed quantum theory of Gravity, and therefore no standard Gravitational boson. The Strong force acts between quarks - for example, it is the Strong force which holds the nucleus of an atom together. The Strong force is mediated by the gluon, and introduces a new type of quantum number called color, analogous to Electromagnetic charge but with three facets rather than the one of electromagnetism. Thus, quarks can be “red”, “green” or “blue” (or antired, antigreen, or antiblue), and gluons carry these characteristics in color- anticolor pairs in order to exchange color between two quarks. Note that this is markedly different than in Electromagnetism, where photons do not carry charge themselves. Also, 6 only colorless particles are allowed to exist free in nature - therefore, quarks only exist in quark-antiquark states called mesons (in which color charges cancel) and in three quark states (such as protons or neutrons) called baryons (in which red, green, and blue form a colorless combination). Note that these groupings also eliminate the possibility of states with fractional charge. The more general term hadron is used to refer to any particle composed of quarks, whether meson or baryon. Historically, the framework of the Standard Model was laid out during the 1960’s and 1970’s [5], with experimental evidence accumulating in the following decades which confirmed the model [6]: 0 experimental evidence of point-like scattering centers inside nucleons, which led to the association with quarks (1969) [7] 0 observation of weak interactions through the exchange of a Z boson (1973) [8] 0 observation of the charm and bottom quarks (1974, 1977) [9] 0 observation of jet structure and gluon radiation (1979) [10] 0 direct observation of the W and Z bosons (1983) [11] 0 observation of the top quark (1995) [12] Increasingly precise measurements of particle properties and interactions have followed these discoveries, as physicists seek not only to confirm the building blocks of the model, but also to refine and extend our understanding of how those building blocks interact. 2.2 QCD Interactions While the theory of the Electromagnetic force, known as Quantum Electrodynamics (QED), is now well understood and well tested experimentally, the equivalent theory for the Strong force, Quantum Chromodynamics (QCD), is less completely understood, and is thus a very active theoretical field. This thesis studies the production of direct photons in proton-antiproton (i.e., hadron-hadron) collisions - interactions involving the Strong force. Thus it is important to consider the workings of QCD in more detail. If we consider a proton-antiproton collision, we might imagine a very complicated picture. We have essentially a bag of three quarks interacting with another bag of three (anti)quarks, all the while exchanging gluons as necessary for the quarks to stay bound together. While these gluons generally have only a small amount of the momentum of the proton, they are plentiful. In addition, because of the quantum nature of QCD, some number of quark-antiquark pairs (called sea quarks) are being created and then annihilated from the vacuum in addition to the 3 valence quarks. How do we keep track of all that goes on between all these different particles? The answer lies in one of the more curious aspects of the Strong force. Because the gluons carry color, and thus couple to themselves, the strength of the Strong force (expressed as the coupling constant of the Strong force, us) actually gets weaker as the energy of the collision gets higher (or alternatively, the distance scale gets shorter). This is the reverse of everyday experience, such as in Electromagnetism or Gravity where increased energy (shorter distance) means a stronger effect. This can be roughly imagined by considering that since the gluons can couple to themselves, they can emit other gluons 8 which increase the color force. Thus the quicker and shorter the interaction, the less chance there is for gluons to radiate, and the smaller the strength of the field. Using this fact, known as asymptotic freedom, we are therefore free to consider the collision between a proton and an antiproton as an interaction between one component quark or gluon (generically called a parton) of each. Also, we can use the methods of perturbation theory to calculate the interaction in orders of us, since at higher energies as becomes small enough to decrease the importance of higher orders. Generally we calculate a cross section (0'), measured in units of area, which indicates how likely it is that a given interaction will occur; this is analogous to the classical cross section, where a larger surface area with which to collide means a larger chance of a collision. In order to give an overview of the process, consider the sample interaction shown in Figure 2.1, in which a quark and a gluon (or more generally, two partons) collide to produce a photon and a quark (or parton). Looking first at the center of the figure, we see the “hard scattering” part of the calculation. Here we consider just the two partons scattering off each other - to calculate the cross section for photon production perturbatively, we consider all the possible outcomes involving a photon and partons. To leading order, this is simply a photon plus a Single parton. Notice, however, that we might also have a gluon radiated off the final state parton (shown as “jet 2”); this requires another strong interaction, and hence another factor of as. We must continue to sum such possible higher order terms, as well as sum over all the different quark/gluon types which the “partons” might specifically be, and integrate over all possible parton momenta in order to produce a total cross section. However, because of the complexity of the calculation, it is not possible to sum to all Bp—9 y+lot+X ee- Parton Distributions Hard scattering Non-perturbative, must be High momentum transfer experimentally determined, Perturbative QCD but are universal noon Q's-f: '. \ was -. ,< . r ' " 1.. .1 ll . , u l -‘, ..‘. , ‘ wk '.‘ | r i. v" photon no fragmentation —) direct probe of the hard scattering process jet2 let 1 Fragmentation (quark, gluon) —> jet Figure 2.1: Diagram picturing different aspects of a QCD interaction. 10 orders - as a result, infinities in the momentum integration no longer cancel, leading to unphysical results. To resolve this problem, a mathematical method known as renormalization is used, which neatly removes the infinities, but at the price of introducing an arbitrary momentum scale ”R (the renormalization scale). It is important to note that 11R is an artificial, non-physical parameter which appears in the result of the calculation only because we cannot complete the calculation to all orders. We choose this scale to be roughly the order of the produced photon momentum, representative of the scale of the interaction; however, we expect that the dependence of our result on uR will be minimized if we have carried the calculation to high enough order. Given this method of calculating the hard scattering, we can expand our consideration to the left side of Figure 2.1 - namely, if we collide a proton and antiproton, what iS the likelihood that two partons of a given momentum will be available to interact? These directly affect the perturbative calculation, since we need to know how often each parton combination that we enumerated above might occur. Also, we need to know what fraction x of the proton’s momentum each of these partons is likely to have, so that we can integrate over momentum properly. Unfortunately, these are exactly the quantities which we cannot calculate in perturbative QCD, as they deal with the relatively low energy structure of the proton. In order to continue the calculation, then, we must make one further assumption. We assume that the part of the calculation which specifies the parton densities and momenta can be separated from the hard scattering part - this is known as factorization. Given this assumption, we can replace the non-perturbative portion of the theory with experimentally ll determined functions which describe the quark and gluon content of the proton - the parton distribution functions (PDFS). These PDFS can then be considered universal - they can be applied to any process which requires knowledge of the content of the proton. However, to implement this separation in the calculation, we must use a mathematical method similar to that used in renormalization - hence we introduce another (unphysical) factorization scale up, on which the PDFS are dependent. As in the case of ”R, we choose up approximately equal to the photon momentum to set the scale of the interaction; however, it is important to note that the two scales need not be exactly the same, as they have independent origins. With an understanding of the PDFS and the perturbative hard scattering, we can finally consider the right side of Figure 2.1. In enumerating the outcomes of a hard scattering, we must consider what happens to the outgoing partons, as we know they cannot remain free because they are not colorless. Here again the nature of the Strong force comes into play: because the Strong force grows stronger with distance, as the outgoing parton gets farther away from the interaction the color force between it and the parent hadron increases. Ultimately, enough energy builds up to create one or more quark-antiquark pairs from the vacuum - these quarks then combine to form colorless hadrons. This process is called hadronization or fragmentation, and it results in a collimated “jet” of particles being observed as the final signature of an outgoing parton. Because of this fragmentation process, there is a finite probability of obtaining any given particle from a parton fragmentation which can be expressed in terms of fragmentation functions; these functions are also not calculable in perturbative QCD and therefore must be modeled using experimental data. However, like the PDFS, they can be considered universal, and 12 independent of the nature of the hard scattering process. Combining these aspects of the calculation, then, we can express the QCD cross section for the process A + B —> C + X (where X indicates that we do not explicitly keep track of what, other than C, is produced) with the following expression: om + B —> C + X) = 2 [dxadxbdcha/Aoa, uF)Fb/B(xb, #1:ch /C(zc) (2.1) ach X Gab —) cX(xa, xb: NR9 lip) where a and b are the partons which interact from hadrons A and B; c is the parton which gives rise to C; the F are the parton distribution functions which describe the likelihood of getting a and b (with momentum fractions xa and xb) from A and B; D is the fragmentation function which describes the probability of getting C from c with some fraction 26 of c’s momentum; and 6' is the partonic hard scattering cross section, implicitly summed to some order in as. Note that in the specific case of photon production (as illustrated in Figure 2.1), the final state photon is free of the complication of fragmentation (D c H = l ). The photon is seen in the final state exactly as it was produced, directly from the parton level interaction. This is an important advantage of the study of direct photon production. In contrast, the study of jet production suffers from ambiguities in reconstructing a parton from the spray of hadrons; also, because of the fluctuations in hadronization, the measured energy of jets can be uncertain because the exact particle composition of any given jet is not known. 13 2.3 Direct Photon Theory The study of direct photons thus provides experimentalists with a unique and useful tool with which to test QCD. The mechanisms by which they are produced, as described by QCD theory, imply a direct sensitivity to the gluon distribution of the proton. Also, the characteristics of the final state particles that emerge in a direct photon event, including the ease and precision with which one can identify and measure a photon experimentally, provide a practical advantage. 2.3.1 Direct Photon Diagrams Consider the examples of leading order direct photon production shown in Figure 2.2. We GLUON COMPTON SCATTERING QUARK-ANTIQUARK ANNIHILATION 9 q 5 r q 7 q 9 Figure 2.2: Leading order diagrams for direct photon production. 14 see that photons can be produced in quark-antiquark annihilation, in which the direct photon is produced along with an outgoing gluon. Another process is the QCD analogue to Compton scattering, in which a quark scatters from a gluon, producing a photon in the final state along with the scattered quark. Because of the relative abundance of low x gluons in the proton, this process is the dominant form of direct photon production for EV T below ~ 100 GeV; above this, the (typically) larger momentum fractions of quarks are necessary to produce such an energetic photon, and the annihilation diagram becomes more prominent. Because of the Compton dominance, we expect a measurement of direct photon production to have sensitivity to the gluon content of the proton. Some examples of higher order photon production diagrams are shown in Figure 2.3. These typically involve additional gluon radiations, and result in characteristic signatures of a photon plus two or more jets. Current direct photon QCD calculations include only up to the two jet level (order (132), known as next-to-leading order (NLO), whereas the experimental measurement includes all orders unless an explicit cut is made, and is therefore termed inclusive. However, if terms beyond NLO are small, the NLO prediction should be a good approximation to the inclusive cross section. 2.3.2 Bremsstrahlung Component and Isolation As shown in Figure 2.4, there is another source of single photons: bremsstrahlung from an outgoing quark in a dijet event. In this case, the photon is not produced directly from the interaction vertex, and is therefore not really a “direct” photon - however, the existence of this production method affects the way direct photons are measured and modeled Figure 2.3: Examples of higher order direct photon diagrams. l6 BREMSSTRAHLUNG DI -0| Figure 2.4: Photon production through Bremsstrahlung theoretically, as described in the next section. This method of production can be described using the probability that a quark fragments primarily to a photon - D q H. In order to reduce the occurrence of the bremsstrahlung case, an isolation criterion is imposed on the photon: typically, a cone around the photon is examined and required to have less than a certain amount of energy. Since a photon produced through bremsstrahlung tends to be very collinear with the quark it radiates from (and thus the jet resulting from the quark’s subsequent fragmentation), the requirement that the photon be isolated removes all but the very largest angle radiations. In the theoretical model, a bremsstrahlung/fragmentation component is added into the calculation and then (partially) removed by the imposition of an isolation requirement matching that of experiment. This component contributes a large (up to 50%) fraction of the total (um-isolated) rate, and remains significant in the ET region below ~ 25 GeV even after the isolation Criterion is applied [13][14]. 17 Another important reason to include an isolation criterion is the possibility that a jet might be misidentified as a photon. In the process by which a parton fragments into observable particles, if most of the parton energy remains with a Single particle (for 0 example, a 1: meson) and that particle produces photons as its decay products (e.g., no -—) w), the multiple photons can be indistinguishable from a single photon. While only one in 103-104 jets fragments in this way, the dijet production cross section is 103-104 times larger than the photon cross section; therefore this background to the direct photon signal is of the same magnitude as the signal itself. The experimental challenges and solutions involved in removing this background are a major part of any direct photon analysis, and are described in Chapter 5. 2.4 Variables for Hadron Collider Physics It is necessary to define a set of quantities which are used in the study of high energy hadron-hadron collisions. While the standard phase space variables of energy (E), 9 momentum (p), polar angle (0), and azimuthal angle (4)) can be used, the nature of hadron colliders makes the introduction of additional quantities necessary. An important feature of hadron colliders is that while the hadrons can be given equal and opposite momenta, there is inherent uncertainty in the momenta of the partons which interact. Unlike positron-electron (e+e') colliders, where the two colliding particles and their momenta can be exactly defined, the quark and gluon substructure of the proton implies that we cannot predetermine which partons will interact, and what momentum 18 they will have (as discussed in Section 2.2). Therefore in the lab (detector) frame, the momenta along the beam axis (defined to be the z direction) will usually not sum to zero, and the system will be Lorentz boosted. In the transverse (x-y) plane, however, we expect that the final state momenta will cancel, as the initial state transverse momenta are negligible. Therefore, we define the transverse momentum and energy: ’ 2 2 pTE px +py (2'2) 2 2 ET; /E —pz (2.3) which become important variables for collider physics. Note that if the particle mass m « E , or for massless particles such as the photon, ET = pT = EsinO (2.4) The Lorentz boost also motivates an alternative to the polar angle 0, which is not Lorentz invariant. Instead, we define the rapidity (y): 1 E +pz 2]n(E-pz) (2.5) y which transforms under a Lorentz boost as y —-> y + [constant], thereby preserving the shape of the rapidity distribution under the boost. Again, if a particle’s mass is much less than its energy, we can simplify to form the pseudorapidity (11): n = —ln tang) (2.6) Thus a pseudorapidity of n = 0 corresponds to 0 = 90° from the beam (2) direction, 19 increasing to oo (—oo) as the angle grows nearer to the +z (-z) direction. Since the condition m « E is a good approximation for high energy jets and photons, the pseudorapidity is used almost exclusively, and is often referred to as simply the rapidity. Finally, since we are interested in measuring a cross section, we consider the rate (R) at which events with a cross section 0' ( with units of area) occur: R = OL (2.7) where L is the luminosity in units of inverse area per second. The luminosity is a measure of beam flux, and is tied to the performance of the accelerator and its ability to focus the colliding beams densely enough to achieve significant event rates for high energy physics 0-24 processes. Cross sections are typically measured in barns (1 barn = 1 cm2), with rare events having cross sections of several picobarns (pb) or less; therefore, modern colliders such as the Fermilab Tevatron provide luminosities of up to ~1031 cm'zs’l. Integrating over the amount of time the accelerator makes this luminosity available to experiments gives a useful measure of how much data has been taken - the integrated luminosity. For the 1994-1995 Tevatron Run 1B, from which the data in this analysis was taken, a total of ~ 100 pb—1 of data was accumulated. Chapter 3 Experimental Apparatus 3.1 Accelerator The Fermilab Tevatron Collider, located in Batavia, IL, accelerates protons and antiprotons to energies of 900 GeV, making a total of 1.8 TeV of energy available as the particles collide. This makes the Tevatron the most powerful accelerator in the world, and thus, in the particle physics tradition of giving Nature as much energy as possible to see what She will create from it, the premier facility at which to study elementary particles and the forces by which they interact. While the Fermilab accelerator might be described simply as a huge ring of magnets almost 4 miles in circumference, there are actually several different, increasingly sophisticated stages of acceleration which make the final collision possible (Figure 3.1). Cockroft-Walton generators produce H' ions which have a small initial kinetic energy; these are then fed into a series of accelerating electric fields known as a linear accelerator, 20 21 TARGET HALL ANTIPROTON SOURCE Figure 3.1: The Fermilab Tevatron Collider. and then stripped of their electrons. The remaining protons are then accelerated further in a small “booster” synchrotron, consisting of a ring of accelerating cavities alternated with magnets to bend the particles, which are also grouped into bunches at this stage. The bunches of protons, by now at an energy of 8 GeV, are then extracted into the Main Ring, so named because it was once itself the final stage of acceleration, giving the protons an energy of 150 GeV. Housed in the same underground tunnel as the Main Ring, however, is its successor the Tevatron, which uses superconducting magnets to produce a magnetic field strong enough to bend even particles of 900 GeV kinetic energy into the necessary circle. 22 The production of antiprotons takes place by directing some of the protons circulating in the main ring into a target, which produces (among other particles) a small amount of antiprotons. These are carefully selected out, condensed into bunches, stored in a small “accumulator” accelerator, and finally, when a sufficient number have been produced, injected into the Main Ring to circulate in the opposite direction to the protons. This points out a major advantage of colliding protons with antiprotons: the same apparatus can be used to accelerate both sides of the collision, since the two particles are identical but for the opposite charge, which will draw the particles in opposite directions in an electromagnetic field. The two bunches of particles pass through each other without collisions because of the low density of particles within the bunches. The antiproton bunches are then fed into the Tevatron and ramped up in energy to 900 GeV. The Tevatron operated during Runs 1A (1992-93) and Run 1B (1994-95) with six bunches of each particle circulating within the accelerator, with collisions occurring at two separate interaction points corresponding to the locations of the two detectors (DO and its sister experiment CDF). At the collision points, focusing magnets are used to squeeze the particle beams down to densities which make collisions sufficiently likely (and thus produce a useful luminosity). The arrangement of particles into bunches means that the collisions take place at definite time intervals as the bunches cross each other - this allows the detector to synchronize with this time cycle so that the readout electronics have sufficient and uniform time intervals to process one bunch crossing and prepare for the next. During Run 1 this time between beam crossings was 3.5 us; the average number of collisions per bunch crossing varied as a function of the luminosity from ~05 to over 3.0 interactions/crossing. 23 In practice, the accelerator runs in cycles called “stores” - at the beginning of each store, the particles are injected into the Tevatron bunch by bunch, the beam is “scraped” (particles with very off-center orbits are removed), and collisions begin. At this point the luminosity of the collider is at its highest - as the store progresses, luminosity decreases as particles are lost by attrition from collisions and unstable orbits. After about 8-15 hours the available luminosity is too low to be useful, and the store is ended by dumping the particles into a “beam stop,” and a new store is readied. 3.2 Detector The DO detector is a multi-purpose apparatus, designed with the physics possibilities of the Tevatron in mind. The goal of studying high-PT phenomena and high-mass states such as the W and Z bosons, as well as the search for the top quark, led to design emphasis on electron and muon identification and measurement; high-PT jet energy measurement; and accurate accounting of missing transverse energy. The detector is shown in Figure 3.2. It consists of three main sections: the central tracking detectors, which lie close to the bearnline and detect ionization left by particles as they travel from the interaction point outward; the liquid argon sampling calorimeter, which stops most particles and measures their deposited energy; and the muon detection system, which surrounds the calorimeter and provides identification and momentum measurement of muons. A full description of these systems can be found in [15]; a brief description is presented below, with particular emphasis on the tracking detectors and calorimeter, which are the systems used in this analysis. Figure 3.2: The DO Detector. 25 3.2.1 'h'acking The DO Tracking System (Figure 3.3) consists of three subsystems, each providing [EH m M @ I 1“ fl ii All (I) O Central Drift Vertex Drift Transition Forward Drift Chamber Chamber Radiation Chamber Detector Figure 3.3: The D0 Tracking System. different information as the distance from the interaction point increases. The Vertex Detector (VTX), which surrounds the beampipc over the range -2.0 < n < 2.0 , provides fine resolution of tracks very close to the interaction point. The Transition Radiation Detector (TRD) surrounds the vertex detector, covering the range -1.2 < T] < 1.2, and 26 provides discrimination between electrons and charged pions. The Central and Forward Drift Chambers (CDC/FDC) give a measurement of charged tracks further away from the interaction point, where they are more physically separated; the CDC surrounds the TRD, while the FDC extends the tracking coverage down to n = 3.1. The Vertex Chamber The DO Vertex Detector, like the CDC and FDC, is a drift chamber, designed to detect the ionization left behind as a high-energy charged particle passes through a medium and Coulomb scatters with atomic electrons. As the name implies, this type of detector consists of a “chamber” containing a medium to be ionized (usually a gas, which allows for easy movement of the ionized electrons), and a scheme by which an electromagnetic field is applied to the medium in order to “drift” the ionized electrons to collection points, these points being organized to provide spatial information on the path of the charged particle. The Vertex detector (Figure 3.4) consists of 3 cylindrical chambers, concentrically layered such that the inner radius of the first layer is just outside of the beampipc at r = 3.7 cm, and the outer radius of the last layer is at 16.2 cm. The actual walls between each chamber are constructed of carbon fiber support tubes, and the chambers are filled with a mixture of carbon dioxide (95%) and ethane (5%). Each of the layers is segmented in azimuthal angle (p, with the inner layer segmented into 16 cells and the outer layers segmented into 32. As seen in Figure 3.4, the cells are defined by rows of wires, which run the length of the cylinder (in the z direction) under tension, and which provide the carefully shaped electric field. Traces on the carbon fiber tubes provide a cathode which 27 Figure 3.4: Side schematic of the DO Vertex Detector. 28 coarsely shapes the field; fine Shaping is provided by the cathode wires, and the grounded grid wires serve as anodes to attract ionized electrons to the adjacent sense wires, which read out the accumulated charge. Note that the cell positions in the three layers are staggered in order to improve resolution in the r-(i) plane, which is typically ~50 um. The track position can be determined in the z coordinate by comparing the amount of charge read out at either end of the sense wires, obtaining a z coordinate via charge division. Measuring the 2 position of tracks is crucial to reconstructing the location of the primary interaction vertex, as well as any secondary vertices due to multiple interactions; knowledge of these vertex positions is used to fix the exact direction of final state particles and therefore measure their production angles and transverse energy components. While the Vertex Detector was designed to achieve a 2 resolution of ~1 cm, the high track multiplicity environment near the interaction point at Tevatron luminosities has hampered this performance. As a result, the position of the primary vertex is determined with the aid of the Central Drift Chamber. The Transition Radiation Detector The Transition Radiation Detector relies on the fact that a charged particle moving at highly relativistic speeds will radiate photons (as X-rays) when it transits between regions with different dielectric constants. The amount of radiation emitted depends on the speed of the particle, which in turn will depend on the energy and mass of the particle. For example, an electron at a certain energy will travel much faster than a 1t+ meson with the identical energy, since the 1t+ is almost three hundred times more massive than an electron - and as a result the electron will on average emit more transition radiation. It is this 29 difference which the TRD is designed to exploit. The TRD, like the VTX chamber, is made up of 3 separate cylindrical units. Each contains foils of polypropylene, 18 pm thick, separated by 150 pm in a volume of nitrogen gas (Figure 3.5). Since the probability of radiation is about 1% per boundary interface [4], there are 393 foils in each unit to insure maximum efficiency. After the foils in each unit there is a drift chamber, which serves to convert the X-rays and read out the ionization signal. Since the drift chamber uses a Xe(9l%)/CH4(7%)/C2H6(2%) gas mixture, the drift chamber in each unit is separated from the nitrogen environment of the foil layers by two layers of mylar, between which dry C02 gas is circulated to prevent any leakage of gases between the two volumes. While this analysis does not explicitly use the information read out from the TRD, its presence affects the detection of direct photons. At normal incidence (0 = 90°) the TRD presents 8.1% of a radiation length to a photon; this is most of the material which lies between the interaction point and the calorimeter. As a result, some fraction of photons will convert to an electron-positron pair, and will thereafter leave a pair of ionization tracks. Averaged over the region 09 S. n S 0.9, the probability for such a conversion is roughly 10%, while in the regions corresponding to 1.6 S lnl S 2.5, where non-normal incidence on the TRD results in greater material depth and the endplates of the VTX and TRD present additional material, the probability rises to about 38% [16]. Therefore, a Significant fraction of real photons and jets decaying to photons will convert before the CDC/FDC, and will behave as an e+e' pair thereafter. 30 CROSS-SECTION 0F TRD LAYER 1 OUTER CHAMBER SHELL 70pm GRID WIRE ALUMINIZED MYLAR ./ 65mm 2 o ______ _}_.__ __———o——. ——— 0 4mm 2 21r/256 mmt. o J o ______ — — — — — — — — o— — ..4 _ _ o RADIATOR STACK CONVERSION . STAGE 0] N2 30pm ANODE WIRE] / 100nm POTENTIAL WIRE 23am MYLAR WINDOWS HELICAL CATHODE STRIPS Figure 3.5: Cross section of the first TRD layer. 3 1 The Central Drift Chamber Surrounding the TRD is the Central Drift Chamber, which like the VTX and TRD has a cylindrical shape. As seen in Figure 3.6, the CDC is segmented in azimuthal angle (32 Figure 3.6: End view of a section of the Central Drift Chamber. cells in all) and is four layers deep, with alternating layers staggered for better coverage. While the physical construction of the cells differs from that of the VTX, with Kapton- covered Rohacell “shelves” defining the physical cell borders, the configuration of sense wires centered in the cell along with surrounding grounded guard wires is similar. In addition to the sense and guard wires, two Teflon coated delay lines run in grooves cut into 32 the inner and outer shelves of each cell. Signals induced in the delay lines by the nearest sense wires are read out at each end of the CDC, and the comparison of arrival times at the two ends yields a measurement of the 2 position of the ionization track. The CDC uses an Ar(92.5%)/CH4(4%)/COZ(3%) gas mixture. The Forward Drift Chambers At smaller angles (closer to the beamline), rather than extending the central detectors, it becomes advantageous to place separate detectors perpendicular to the bearnline, so that particles will again impact normally on the detector surface and the z direction extent of the detector can be minimized. The Forward Drift Chambers thus have a different geometry than the CDC: each detector (one on each side of the central region) consists of three disk-like layers (Figure 3.7) oriented perpendicular to the beamline. The first and third layer are constructed of four quadrants of Six stacked rectangular cells each, so that each cell covers a different range of angle 0. These two “O” layers are rotated by 45° with respect to each other. The middle layer is very different, composed of 36 cells which are like pie wedges, providing discrimination in azimuthal angle q). As in the CDC, though, both “8” and “4)” layers utilize multiple anode and sense wires running along the long dimension of each cell, and several deep in z, to detect the ionization signal. The gas mixture used is also the same as in the CDC. Also, in the “O” layers, a single delay line similar to those in the CDC provides a determination of the location of the track along the length of the rectangular cell. 34 3.2.2 Calorimetry The DO Calorimeter System (Figure 3.8) measures the energy of most particles by END CALORIMETER Outer Hadronic (Coarse) Middle Hadronic (Fine & Coarse) I .\ ,\ \. '.§? 4 . 3%; \\ ”F /// ”Ix / l \ ' CENTRAL " CALORIMETER \ . ’ 7' Electromagnetic r‘~ v“- eff/t § Inner Hadronic ' Fine Hadronic (Fine a Coarse) Coarse Hadronic Electromagnetic Figure 3.8: The DO Calorimeter (cutaway view). Stopping them completely within the calorimeter volume, thus transferring the (kinetic) energy of the incident particle to the material in the calorimeter. Accordingly, the calorimeter contains layers of very dense materials (plates of uranium and Steel), within which the particles shed their kinetic energy, interspersed with layers of an ionizing 35 medium from which a signal can be easily measured (liquid Argon). This is therefore a “sampling calorimeter” - since only the ionization left within the Argon layers is read out, only some fraction of the energy (about 10%) is sampled. The total energy deposition is then inferred from this sample by calibrating with particles of known energy in a test beam, and deriving "sampling weights" which multiply the raw energy read out in each layer so that the sum exactly represents the true particle energy. We further expect that readout charge will be proportional to particle energy, so that the calorimeter response will remain linear with increasing energy. Practically, this layering is achieved by enclosing the dense plates within a cryostatic vessel filled with liquid Argon, so that the Argon-filled gaps between plates become the readout layers. An example of a calorimeter cell implemented this way is shown in Figure 3.9. As can be seen in Figure 3.8, the calorimeter is separated into one Central and two Endcap sections (North and South), housed in separate cryostats to allow easier access to the interior of the detector. The Central Calorimeter (CC) covers roughly the range Inl S 1.0 and the Endcap Calorimeters (EC) extend the coverage to Inl z 4. The CC is further segmented radially into 3 sections (Figure 3.10). The Electromagnetic (EM) calorimeter is the section closest to the interaction, and it is read out in 4 separate layers (EMl-EM4) which correspond to 2.0, 2.0, 6.8, and 9.8 radiation lengths, respectively. Most electromagnetic showers from photons and electrons are contained almost completely within these 4 layers, and thus it is the most important part of the calorimeter for this analysis. Beyond the last EM layer, and designed to stop the more penetrating hadronic showers, are the three layers of the Fine Hadronic (FH) section 3 O\ ( Absorber Plate Pad Resistive Coat Liquid Argon D N 010 Insulator JL 1 Gap— 7/ \. 7/ / i % __ _ _/// _L c <-——- U3_it_0_e_li —--> o Figure 3.9: Diagram of the layers within a calorimeter cell. which include 1.3, 1.0, and 0.9 nuclear interaction lengths. Finally, the Coarse Hadronic (CH) section, which uses copper as an absorber rather than uranium, provides one more depth segment of 3.2 nuclear interaction lengths to contain the few hadronic Showers which penetrate beyond the EM and FH sections. The depth segmentation of the EC is similar to that of the CC, with the difference that the hadronic calorimetry is actually broken up into 3 separate sections (concentric about the beampipc) called the Inner (IH), Middle (MI-I), and Outer (OH) sections, with the 1H and MH sections containing both fine (uranium) and coarse (stainless steel) layers of depth sirrrilar to the PH and CH in the central region. The OH section contains only coarse layers of stainless steel, as it lies in a region of overlapping coverage with the FH section 37 0.0 IETA Erik-366mm,». n o Figure 3.10: Side view of a calorimeter quadrant. Lines of pseudorapidity are shown to demonstrate the projective tower geomeU'y. of the Central Calorimeter, and thus serves essentially as an extension of the Central Coarse Hadronic layers. Also, the EC EM layers have slightly different thicknesses than in the CC, comprising 2.0, 20,79, and 9.3 electromagnetic radiation lengths. Figure 3.10 also shows that the individual cells of the calorimeter are arranged projectively in towers of An XAO = 0.1 x0.l so that their centers point back to the midpoint of the detector, which allows good shower Shape resolution. In addition, to provide better measurement of electromagnetic shower position, the third layer of the EM 38 calorimeter (EM3), where EM showers typically leave the majority of their energy, is more finely segmented (An x Ari) = 0.05 x 0.05 ). To calibrate the calorimeter modules, sections of CC and EC calorimeter assemblies were studied in a test beam [17][18]. Electrons and charged pions with well-measured energies ranging from 2 to 150 GeV were scanned across the modules to Simulate various angles of incidence. These studies found that the response of the calorimeter was linear above ~ 10 GeV, with a resolution for electrons of 15%/(JE), and 50%/(JE) for pions. To finely calibrate the EM calorimeter, calorimeter response was measured in situ by reconstructing the decays Z —) e+e' , J/\|t —> e+e' , and Ito —> w [19]. Since the masses of these particles are well established (91.187 GeV, 3.096 GeV, and 134.97 MeV, respectively) and their decay products span a large energy range, exact calibration of the response versus energy is possible. The Intercryostat Region To accurately measure energy deposited in the gap between central and end cryostats, two additional detector systems are installed. The first, termed massless gaps, consist of single calorimeter readout cells placed in the region between the cryostat wall and the edge of the PH calorimeter section (in the CC) or the MH and OH sections (in the EC). These essentially make the regions between the calorimeter modules and the cryostat walls into an extra readout cell, albeit with very little mass to stop particles. In addition to the massless gaps, scintillating counter arrays called the Intercryostat Detectors (ICD) are mounted in the space between the central and endcap cryostats. Each ICD array consists of 39 384 tiles of An x All) = 0.1 x 0.1 , read out by phototubes, and covering an area on the inside face of the EC cryostats roughly corresponding to the MH and OH sections. While they are of no aid in identifying electromagnetic showers, the ICD and massless gaps provide enough information on hadronic showering between calorimeters to assist in the proper reconstruction of energy deposition between cryostats. 3.2.3 The Muon System While the calorimeter stops and absorbs most particles within its volume, there are two particles it will not contain: neutrinos and muons. Neutrinos, due to their extremely small cross section for interaction with matter, cannot be detected at all, and the presence of a neutrino in an event must be inferred by summing the energies in the detector and determining the amount and direction of “missing” transverse energy (ET ) needed to make the total transverse energy sum vectorially to zero. Muons, however, will leave a trace of their passage via ionization - thus the outermost component of the D0 detector, the Muon System, is designed to identify muons and measure their momenta and position. The Muon System consists of three sets of proportional drift tube chambers (PDTS) along with a toroidal iron magnet, implemented in 5 separate pieces, maintaining fields of ~2 T. The first set of PDTS, called the A layer, is mounted before the toroid and consists of four planes of PDTs; the B and C layers are mounted outside of the toroid, and consist of three planes of PDTS each, with the two layers separated by 1m. The A layer, with its extra plane of PDTS, detects muon hits and allows them to be associated with tracks in the Central Tracking Chambers; the B and C layers track the muon trajectory after it has been 40 bent by the magnetic field in the toroid, and thus allow the momentum of the muon to be determined from the magnitude of the bend. 3.3 Data Acquisition and Processing 3.3.1 The Trigger System Since the accelerator produces beam crossings every 3.5 1.1.5, the detector must evaluate a possibly interesting collision 285,000 times per second. At the other end of the timescale is the rate at which a complete readout of the detector - close to 150,000 channels - can be written to tape for storage: this is limited to 2-5 Hz. Condensing 285 kHz of potential collisions down to 2-5 Hz of interesting events which comprise the final data sample is done using the DO Trigger System, which consists of three successive levels. The Level 0 Trigger The first level of triggering is the Level 0 Trigger, which signals that an inelastic collision occured in the beam crossing and this event is therefore worth examining by the higher levels of the trigger. The Level 0 system consists of two planes of rectangular scintillation counters, mounted 90° relative to each other to form a checkerboard-like pattern of coverage; one such pair is mounted on the front surface of each of the endcap calorimeters. When both counters register hits after a beam crossing, there is a >99% probability that an inelastic collision occurred; the rate of coincident bits is also an independent measure of the beam luminosity, and can provide feedback to accelerator operations. The timing difference between the two counters can be used to give a rough 41 measurement of the z of the interaction vertex, and can also indicate the presence of multiple interactions - all of this information can be utilized in the higher trigger levels. The Level 1 Hardware Trigger Once it is established that a collision has occurred, the Level 1 triggering system examines the calorimeter and the muon system for evidence of significant energy deposition. In order to do this within the 3.5 us deadline, the Level 1 system uses fast, hardware-based algorithms to evaluate the event. In the Muon system, 200 electronic readout boards known as Module Address Cards (MACS) receive from their assigned subset of muon chambers a list of cells from all planes which registered hits. These hits are processed by the MACS into a single “coarse centroid,” representing the best quick estimate of the muon position. Centroids output by the MAC cards are then combined over the A, B, and C layers to give a rough momentum measurement; this information is then used by the Level 1 system to compare with trigger requirements for position and minimum momentum within 3.5 its of the interaction. If needed, the system can initiate additional processing known as the Level 1.5 muon trigger system: this uses “fine centroids” generated by the MACS and can produce a more accurate measurement of the transverse momentum, but requires another 3.5 us and thus incurs a deadtime penalty for the entire detector. In the Calorimeter, a fast readout samples the rising edge of each cell’s signal pulse, and sums these signals into An = All) = 0.2 trigger towers, keeping separate account of EM and hadronic energies. Using lookup memories and the 2 position from Level 0, these tower energies are processed to provide the Level 1 system with EM, hadronic, and total 42 ET values for each tower, as well as the missing ET resulting from a sum over all towers. These quantities are then compared against the trigger requirements. The average rate of events passing at Level 1 is about 100 Hz. The Level 2 Software Digger The final level of triggering is provided by the Level 2 system, which consists of 50 VAXstation 4000 computers (termed “nodes”) connected to the rest of the system by high speed data cables. Events which pass the Level 1 system initiate a complete digitization and readout of the detector, and this information is sent to the first available Level 2 node, which then runs software-based algorithms to evaluate the event in much more detail than at Level 1. This digitization and processing requires about 200ms; however, this is well under the average time between events sent to a node from Level 1. To increase the speed of the readout, calorimeter cells are subjected to zero suppression, in which cells with less than 20' fluctuations in energy from their normal, quiescent readout level are zeroed and suppressed, therefore saving the need to read out an energy value. In practice, the three levels of triggering are carefully tuned and matched to each other to produce an optimum output stream of events. The Level 1 trigger system, having received a positive Level 0 result, can evaluate up to 32 separate sets of requirements (termed “triggers”) on energy in one or more towers, and the position of those towers. These triggers are carefully selected to correspond to general types of physics events. Should any Of these requirements be met, the event is passed on to a Level 2 node, where up to 128 sets of requirements (termed “filters”) can be made. Each filter is matched to a Specific Level 1 trigger, and serves to identify a more detailed and Specific type of final 43 state. In addition, each trigger and each filter can be “prescaled” to pass only 1 of N events identified as matching its requirement - in this way, the passing rates can be kept within limits, and triggers/filters corresponding to common event types can be controlled to ensure that rarer triggers/filters are not lost to bandwidth limitations. 3.3.2 Event Reconstruction Once the event has passed the online event selection system, it is then stored on 8mm tapes. However, in its raw form it consists of only the signals read out from each detector cell, information from the triggering process, and information about running conditions at the time it was taken. To put the data in a useful form for physics analysis, the DO Reconstruction program (DORECO) was developed to serve three main functions: 0 Apply calibration information to the raw data to correct all detector hits 0 Apply algorithms to the detector hits to identify and describe objects within the detector: photons, jets, muons, etc. 0 Reduce the event to a manageable size for later storage and analysis Reconstruction of the raw data was done using a farm of SGI and IBM Unix machines, though the program can be run on multiple platforms and can be rerun on processed data in order to apply new techniques or calculate new quantities. The DORECO program itself is a large assembly of code packages which perform all the necessary functions of calibrating and interpreting the output of each detector; the relevant sections for this analysis are described below. The output of the program is a file using the ZEBRA data 44 structure, in which information is organized in “banks” which are referenced by pointer, and the ultimate size of the output file can be controlled by which banks are written to the file. DO maintains several levels of datafiles, each carrying successively less information but allowing successively more case in processing and disk storage. The Z TRAKS package This package is reponsible for the reconstruction of tracks within the tracking chambers; however, it also includes the determination of the event vertex using tracks reconstructed in the CDC. Tracks are found in the CDC by reconstructing hits in the r—(i) plane and aligning them first within each layer, forming what are called “segments.” Starting with the outer layer, these segments are then compared, and if 3 of the 4 layer segments are sufficiently aligned a track is confirmed, after which the entire track is refit, a 2 position calculated, and a final cut made on the quality of the track. Tracks in the FDC are determined in a similar way; however, because the forward region tends to have many more tracks from the breakup of the colliding particles, full reconstruction of the entire FDC is very time-consuming and is usually not done. Instead, only the region between the vertices and physics objects such as electrons or photons is reconstructed - the tracking “road,” which is usually A0 x All) = 0.2 x 0.2. To determine the vertex, reconstructed tracks within the CDC are pointed back to their z-intercept point, and the points from all such intercepts are put into a histogram. A cluster finding algorithm is run to identify clusters of points, and those with at least three points are resolved into separate vertices, with the primary vertex assigned as the one with the 45 most points. The mean and error of the points in each cluster are then taken as the mean and error on the vertex position. If there is only one cluster, even with only one point, it is taken as the vertex; if there are no CDC tracks (perhaps 5% of the time), the FDC is used instead, in which case the entire FDC is reconstructed. The error on the vertex position can range from 6mm for an event with many tracks, to 2 cm for an event with only one CDC track, to 10cm if the FDC tracks must be used. For each found vertex a fit is then done using only those tracks pointing to that vertex and constraining them to originate from a single point, resulting in the final mean and error of the vertex 2 position. Note that determination of the x-y coordinate of the vertex is done using VTX readings during data- taking - the x-y dimensions of the beamspot are small and usually vary only a few microns from run to run, so that an average value can be used over a certain run period. The CAHI TS package The CAI-[ITS package is responsible for the bookkeeping of all the energies of each calorimeter cell, the application of calibration corrections, and the mapping of cells to physics coordinates. Its output, basically a corrected bank of all cells in the calorimeter, is then used by object finding packages such as those relevant to this analysis, CAPHEL and CAJETS. The CAPHEL package As implied by its name, this package searches the calorimeter to find photons and electrons. It employs a clustering algorithm which finds towers ( 0.1 X 0.1 ) in the EM calorimeter above a certain ET threshold, then associates neighboring towers with this 46 tower until no neighboring towers above a 50 MeV threshold are found. The package then calculates quantities for this cluster (such as its total energy, position, etc.), making several ZEBRA banks full of information, and also associates with it any tracks coming from the primary vertex within the tracking road between the vertex and the cluster, making links to the corresponding tracking ZEBRA banks. The CAJE TS package The CAJETS package is Similar to the CAPHEL package, in that it must search the calorimeter to find jets. Since the definition of a jet, and which particles in the event should be associated with it, is somewhat ambiguous, several different algorithms are used. The one relevant to this analysis is the fixed cone definition, using a radius of 0.7 in An x Ad) Space. In the first stage of the algorithm, a “precluster” is found similarly to the electron/photon case: the cell with the highest ET in the calorimeter is used as a starting point, and cells with ET greater than 1 GeV and within 0.3 in n or (i) are grouped together around it; these cells are then removed from consideration and the remaining highest cell has a precluster built around it; and so on until all cells with more than 1 GeV ET are in a precluster. Then for each precluster, cone clustering is applied: starting with the highest ET precluster, the ET weighted centroid of the cluster is found, and all towers within a radius of 0.7 in An x All) are assigned to the jet. The jet axis is recalculated using these towers, and the process is then iterated until the jet axis becomes stable. If the resulting jet has 47 ET> 8 GeV it is kept, and it is compared against other jets found for any Shared towers. Should any two jets share towers, an algorithm is run which either splits them, assigning contested cells to the nearest jet, or merges them, depending on how much ET is shared between the two. Split or merged jets then have their axis recalculated a final time. Chapter 4 Event Selection and Efficiencies In order to measure the cross section for direct photon production, we must first define what we expect a direct photon to look like in our detector, and place requirements on all photon candidates in order to select these characteristic events and reject non-signal events. AS outlined in Chapter 3, this can be done at the trigger level, selecting events which look like direct photon events, and then refined further using offline cuts to impose even stricter requirements. The goal of background rejection may lead to cuts harsh enough to reject even some real direct photon events; therefore, we must also estimate the number of direct photon events lost so that the total number measured can be corrected to equal the total number produced. 48 49 4.1 Data Set and Monte Carlo 4.1.1 Direct Photon Triggers Because the direct photon production rate falls steeply with ET, multiple triggers/filters must be used to acquire data in each ET region and prescaled accordingly to achieve comparable statistics over the measurable ET range. The direct photon candidate sample was obtained using the four Single photon triggers and filters summarized in Table 4.1. The two lowest ET filters, GAM_6_ISO_GAM and GAM_14_ISO_GAM, were taken in Special runs, when the lurrrinosity and event rate were lower and the full bandwidth of the trigger system was devoted solely to photon triggers; therefore, usually only the 6 Gev filter was prescaled. The two highest filters, EM1_GIS and EM1_GIS_HIGH, were taken as part of global running, where only a part of the bandwidth was devoted to photons. The EM1_GIS filter was therefore prescaled, with the EM1_GIS_HIGH filter also prescaled at very high luminosity running where high event rates taxed the trigger system’s capability. Total integrated luminosities for each filter are also shown in Table 4.1, and carry a 5.4% uncertainty. Inefficiencies due to triggering do occur; however, to avoid this complication, filters were used in the cross section measurement only where fully efficient. Detailed descriptions of the triggering algorithms, triggering efficiencies and the determination of integrated luminosity for each filter are included in Appendix A. 50 Table 4.1: Description of triggers and associated integrated luminosities. L1 L2 L . i Level 1 Trigger Threshold Level 2 Filter Threshold “31:9: ‘3' (GeV) (GeV) EM_1_2_GAM 2.5 GAM_6_ISO_GAM 6 13.644 EM_1_7_GAM 7 GAM_14_ISO_GAM 14 225.29 EM_1_HIGH 14 EM1_GIS 25 14577 EM1_GIS_HIGH 40 94657 4.1.2 Candidate Sample The standard cuts applied offline to define a good photon candidate are : 0 0 S Inl S 0.9 (1.6 S Inl S 2.5) in the central (forward) regions. These cuts define the region of phase space which will be measured in the cross sections. 0 0 < IIETAI S 9 (16 S IIETAI S 25) in the central (forward) regions. The quantity IETA is the integer index of each 0.1 x 0.1 calorimeter tower in pseudorapidity (see Figure 3.10). This cut restricts the candidates to be well within calorimeter boundaries. - vertex 2 position IZm | < 50 cm to retain the benefits of projective geometry. 0 Exclude candidates within 10% of a module edge in 4) (CC only). Energy reconstruction near these edges is not well understood. 51 no track pointing to the candidate from any found vertex. EEM E > 0.96 , where EEM (Em) is defined as the EM (EM + hadronic) energy in a tot core cone of 0.2 around the candidate. This rejects hadron showers, which leave significant energy past the electromagnetic layers of the calorimeter. [ET-(R = 0.4) —ET(R = 0.2)] < 2 GeV where R = A/Ar]2 + A1112 and the ET within a cone of radius R refers to the sum of EM and hadronic layers. This isolation requirement rejects most neutral mesons, which are typically part of a jet and therefore are accompanied by additional energetic particles. H-matrix 12 < 150. The H-matrix algorithm is a measure of shower shape in the calorimeter, and uses the total energy, energy deposition in each layer, each energy in a 6x6 array of cells in EM3, and the vertex 2 position as 41 separate, characteristic observables of a Shower. These are combined in a covariance matrix N 1 ,, - ,, _ Mij = IT, 2 (x, -xi)(xj -xj) (4.1) n=l where x? is the value of observable i for candidate n and 3c,- is the mean value of the observable i for a sample of “idea ” electromagnetic showers (test beam and 52 Monte Carlo electrons were used to tune the matrix). A x2 distribution x2 = 2(xf—£,)H,j(xf—ij) (4.2) it (H = M.1 ) can then be formed to measure how well each candidate compares to the expected shower shape. An illustration of the H-matrix discriminating power between 25 GeV test beam electrons and pions is Shown in Figure 4.1 [20]; the distribution for electrons from a sample of W—r ev decays (~ 25-40 GeV) is included for comparison. Figure 4.1 demonstrates the effectiveness of a cut at x2: 150 in selecting electromagnetic showers. The first four requirements are classified as acceptance cuts; the last 3 are known as the photon selection cuts, as they provide the bulk of the photon identification and discrimination from backgrounds. To illustrate the effect of each of these cuts on the data sample, and to Show the number of events each filter contributed to the analysis, the number of raw events satisfying the physics 1] requirement in each region is shown in Table 4.2. 4.1.3 DOGEANT Monte Carlo In order to calculate cut efficiencies and estimate background from neutral mesons, several different types of simulated (Monte Carlo) events were generated [21] using the PYTHIA event generator [22]. These were generated including PYTHIA modeling of the number of events 53 100 electrons 80 pions O w-9ev 60 40 20 Figure 4.1: H-matrix x2 distributions from test beam electrons, test beam pions, and electrons from W decays. Table 4.2: Number of events surviving the application of photon cuts. Filter # before cuts # after # after 8:13;“ (after 11 cut) acceptance track cut cuts CC 28878 20087 11850 4628 GAM_6 EC 25370 21264 6214 2777 CC 5820 4170 2327 1346 GAM_14 EC 6839 5579 1485 890 CC 79469 56558 27622 13823 GIS EC 82091 68580 15036 7496 CC 79499 57490 26988 12283 GIS_HIGH EC 58602 5057 l 9323 3861 spectator event, and simulated using a DO-Specific implementation [23] of the GEANT detector simulation package [24] with fully detailed geometry. In order to model uranium noise effects, multiple interactions, and “pile-up” (which occurs when a cell is read out before the calorimeter electronics are fully recovered from energy deposition in a previous beam crossing) the Monte Carlo was merged with non-zero suppressed zero bias data, taken without any trigger requirement at Level 0 or above. The combined files were then zero suppressed and reconstructed as in the data. The Monte Carlo was put through several iterations of this overlay process to produce 5-10 times the number of final events; this smooths the effects of noise and multiple interactions for a given number of Monte Carlo showers. Since the statistical power of the Monte Carlo sample is limited by the number of actual PYTHIA events generated, however, statistical errors were kept equal to that due to 55 the actual number generated before the overlay process. Single photon events were generated over the range 10-200 GeV (segmented into four parts for roughly equal statistics in each), resulting in a falling ET spectrum closely resembling the data. A large sample was generated for background studies (Chapter 5) in which only the region around the photon was simulated in DOGEANT to save processing time; a smaller sample in which the entire event was Simulated was used for efficiency and physics studies. A sample of full W and Z events was also generated for comparison to data. As a check on the ability of the Monte Carlo to model the data, a sample of electrons from Monte Carlo W -—) ev events was compared to electron candidates from data W—) ev events taken from the direct photon trigger sample. The results are shown in Figure 4.2 (CC) and Figure 4.3 (EC). Shown are the EM fraction, the H-matrix x2, and the isolation ET; the Monte Carlo (histograms) Show good agreement with the data (points), though the isolation ET distribution has a slightly larger tail in the data than in the Monte Carlo. 4.2 Acceptance The acceptance A corrects for the effects of the IETA, vertex position, and o module cuts on events with physics 1] in the central or forward range. The Run 18 vertex 2 distribution is shown in Figure 4.4. This distribution was found to be stable during Run 1B and between different photon triggers, and is described by a Gaussian peak of mean -0.3 cm and width 28.5 cm. The vertex position acceptance was 56 £1 _ £1 — § — E : or: iii L E; E; '- 111411111 _ ...- L*:‘:_:Jllll 0.94 0.96 0.98 1 1.02 0 50 100 150 200 EMFraction x' i m . a: T E : i z P— L +++ r r + ++ .— + L- lll11111l111111111111 —1 0 1 2 3 4 lsolatiorrli1 Figure 4.2: Comparison of data (points) and Monte Carlo (histograms) W electrons in the central region. 57 3 : 3 § L 5’ 111 E m ‘g :— ii 2 §_ 2 0.94 o 50 100 150 200 x) 3 a > In :3 E O 2 11JJIILLL — 1 0 1 2 3 4 Isolation IiT Figure 4.3: Comparison of data (points) and Monte Carlo (histograms) W electrons in the forward region. 58 Events _IIIlllllIl—LIFTIIIIFIIIllllllII—IT lllllllillllllll ll — 100 —50 O 50 100 Vertexz (cm) Figure 4.4: The Run 1B vertex 2 distribution. determined simply by measuring the number of events lost from the vertex distribution after the cut. This was measured separately for all four filters, and was found to be very consistent; the acceptance for this cut using all four filters is 0.913 with a negligible error. The IETA acceptance was measured using the Owens, et.al. Monte Carlo program [25] to generate photons with the theoretically modelled 1] distribution in each region. Each event was smeared with a Gaussian vertex distribution to mimic the data distribution, a 50 cm cut was placed on the vertex position, and the calorimeter IETA of the resulting candidates was calculated geometrically. While this should give the appropriate acceptance for real photons, the calculation was also done assuming a flat distribution in physics 1], and the difference taken as a systematic error. The resulting IETA acceptance is 0.917 i 0.007 (0.930 :1: 0.007) in the CC (EC). 59 Combining the vertex and IETA acceptances with the exact 11) module acceptance of 0.80 in the CC, the total acceptance is then A = 0.670 i 0.007 (0.849 i 0.007 ). 4.3 Cut Efficiencies 4.3.1 Tracking Considerations By vetoing candidates with a track, real photon candidates are lost in two ways: ' A photon might convert to an e+e' pair in the material in front of the CDC/FDC; the fraction of such conversions whose track is accurately reconstructed will be removed from the sample. 0 A charged track produced as part of the underlying event might overlap the photon closely enough to lie within the tracking road, in which case the photon will be removed from the sample. The effect of the charged track cut is considered in detail in Appendix B; the total efficiency due to both sources is found to be e k = 0.828 i 0.013 (0.554 1: 0.030). [70C 4.3.2 Selection Cuts Selection cut efficiencies were determined using full Monte Carlo photon events which were put through the trigger simulator, and are shown in Figure 4.5. The efficiency dips at very low ET due to the Hmatrix and EM fraction cuts, as the Shower shape changes and the 60 a . C1, 1' ............................................................................................................ g _ 0;868 -; 0 0002E,-exp— (0 24E,— 0 008) t 1 ,_... ....................................................... LIJ '- 3 r- . . . . E . . . . 0.9 ... ........ l ..,- ........ .§,,,,. .......... ,,.,. .......... :, .......... Z ‘ “LI—LC " 3- I‘ i :fi—ii E E _ - 1T2! 21* 737 a w : : :—— 0.8 p”... _..........a_...........}..........: ..... +..:...........:\..........g ........... : .......... 3 .......... r- ! ' ' E I I p—jy—4 I 0.7 L ...... ...L .......... .......... C ‘ ' §CC Cut Efficiency; 0'5 ..‘- .......... .......... " I 2 2 I I Z Z 2 3 I i- 1 . - . . 0.51-11l111111111L11L11111111111111111l11111 0 20 40 60 80 100 120 140 160 180 200 (GeV) 5‘ r: 1.1 .93 .9 t 1 LL! 0.9 0.6 0.7 0.6 t- . . . . 0_5b11111g1111111111111111111111l O 20 4O 50 80 100 120 140 (GeV) Figure 4.5: Selection cut efficiencies measured using PYTHIA Monte Carlo photons, and corresponding fits. roughly constant contribution of noise from the PH] layer becomes an increasingly larger fraction of the total energy. The isolation cut causes a fairly linear decrease with increasing ET, as outlying edges of the shower become more energetic relative to the fixed 2 GeV cut. The selection cut efficiencies were thus fit to the form —- bE a—e( ”Chara, (4.3) with the values of the parameters a, b,c, and d as shown in Figure 4.5. 61 Because the global running data were taken over a range of instantaneous luminosities, the effect of instantaneous luminosity on the selection cut efficiencies was studied using full event Monte Carlo overlaid with SE30 and 19E30 zero bias data. Initial studies using Monte Carlo W electrons overlaid with SE30, 14E30, and 19E30 zero bias data indicated that the selection cut efficiency dropped linearly with increasing instantaneous luminosity, resulting in a 10-20% lower efficiency at the highest luminosities. Since the instantaneous luminosity of the data set was peaked at ~ 6E30, however (Figure 4.6), with only a small £9 -— Mean 6.861 :3 I Mean 7.669 5 : RMS 2.526 5 : nus 3.548 > _ > .T' Lu _ m : :— EM1_GIS E— EM1_GIS_HIGH :_ E- : E g : ' 1 4 1 1 1 1 1 1 1 1 1 1 J 1 : 1 1 1 1 1 1 1 1 I 1 1 L 1 1 1 1 O 5 10 15 20 O 5 10 15 20 Instantaneous L (xE30) Instantaneous L (x1330) Figure 4.6: Instantaneous Luminosity profiles for the EM1_GIS (left) and EM1_GIS_HIGH (right) filters. tail extending to the highest luminosities, the average efficiency was expected to be close 62 to that at SE30, allowing selection cut efficiencies to be determined using Monte Carlo with only the SE30 overlay. This expectation was checked using the photon Monte Carlo: for several ET bins above 30 GeV, the 5E30 and 19E30 efficiencies were measured and used to define a linear parametrization of the efficiency vs. luminosity. Using the instantaneous luminosity profiles of the EM1_GIS filter (or EM1_GIS_HIGH, depending on the ET of the bin), a weighted efficiency was then constructed for each ET bin by summing over the luminosity bins, weighting each with the parametrized value of 838,. The resulting weighted efficiencies were found to agree with those derived from the SE30 Monte Carlo to within 2.5% or less; an uncertainty of 2.5% is therefore placed on the fitted selection cut efficiencies derived from the 5E30 Monte Carlo. To cross-check these efficiencies, a sample of Z events was selected and a background—subtracted selection cut efficiency calculated for Z electrons between 30-50 GeV. The resulting efficiencies were 0.865 i0.011 (0.902 i0.018) for electrons with mean ET of 42.7 (36.4) GeV. The fitted efficiencies evaluated at these points, 0.859 :t 0.021 (0.880 i 0.022) , are in good agreement with the Z efficiencies. Chapter 5 Background Estimation 5.1 Introduction The main background to the direct photon signal comes from jet fluctuations. While most jets consist of many particles, and are thus easily distinguishable from a single photon, a small fraction (10'3 - 10'4) fragment such that a single particle carries most of the momentum of the parent parton. Several neutral mesons decay to produce two or more photons, including the lightest (and therefore most commonly produced) neutral meson, the no, which decays to two photons with a branching ratio of 99% [4]. Thus non-direct photons are common in hadronic collisions, and because jet production rates are a factor of 103 larger than that of direct photons, the number of jets which fragment to a single meson and then decay to photons - referred to as electromagnetic (EM) jets - are comparable to the level of direct photon signal. Because of this, the level of background in the direct photon data sample must be carefully measured and subtracted. 63 64 Table 5.1: Neutral meson decays to photon states. Particle (($37?) decay products branching ratio 1:0 0.135 w 0.99 11 0.547 w 0.39 11 0.547 31:0 0.32 n 0.547 n°n+1r 0.23 KOL 0.494 211° 0.31 (0 0.781 1107 0.09 Further complicating matters is the fact that, at the energies considered in this analysis (above ~10 GeV), the individual photons from meson decays cannot be separately identified given the spatial resolution of the D0 calorimeter - the decay products coalesce into a single shower which is difficult to distinguish from a single photon. This can be seen by considering that the two photons arising from a 1:0 decay have an opening angle distribution peaked at the minimum [26] 0 - = — (5 .1) which for a 10 GeV 1:0 results in a minimum separation of 2.5 cm at the first layer of the D0 calorimeter. Since a typical cell size in the calorimeter is ~7 cm, it is clear that most no decays will not be identifiable as two distinct photons. While Table 5.1 identifies several neutral mesons which decay to photons, the no and 11 are the main particles considered in this analysis, as they are the two neutral mesons which are most commonly produced. Other backgrounds are expected to be further 65 suppressed in the data sample due to small branching ratios (e.g., the (1)) as well as the photon selection cuts. For example, while a significant fraction (~30%) of K? mesons decay to 216’, at lower values of ET the four resulting photons are sufficiently spread out that they usually fail the photon selection cuts. At higher ET the showers coalesce and the passing rate is higher; however, due to the long lifetime of the K2 , at higher E;- a significant fraction do not decay before they reach the EM calorimeter, and thus create a hadronic shower which also fails the photon cuts. As a result, the expected ratio of K? to 11:0 in the data sample is less than 5%. 5.2 Background Estimation Techniques 5.2.1 Method While background subtraction on an event-by-event basis is impossible, there is a statistical method which exploits the difference between single photons and multiple photon clusters as they traverse the detector. At the energies considered in this analysis, photons begin showering in the calorimeter by converting to an electron-positron pair (7 —-) e+e') near a nucleus. The probability for this to occur is dependent on the amount of material the photon passes through; the deeper into the calorimeter the photon gets, the more likely a conversion becomes. Consider the probability PZ—m that a single photon will convert in or before the first 66 calorimeter layer. For a no meson decaying to two photons, the probability that at least one of the photons will convert is then greater than in the single photon case: 2 PTEIMI = ZPIYE‘MI -(PIIM1) (5-2) In other words, multiple photon showers are more likely to begin earlier in the calorimeter, since there are more independent chances for a photon to convert early. As an example of this, consider Figure 5.1 [2], which shows the clear difference in the fraction of total 350 H —— MC Photons ........... MC Pizeros 300 -* 250 ~ 200 ; . ...... O 1 J 1 l 1 1 1 l 1 l 1 l 1 l 1 I 1 l O 0.025 0.05 0.075 0.1 0.125 0.15 0.175 0.2 0.225 0.25 EM 1 / E Figure 5.1: Comparison of EM] energy fraction left by 10 GeV MC photons and 0 1: ’s. 67 energy left in the first electromagnetic calorimeter layer (EMl) by single Monte Carlo photons and no’s. While a significant fraction of photons leave little or no energy in BM], relatively few no’s do so. This difference provides discriminating power with which to estimate the fraction of background present in the direct photon candidate sample. The data distributions can be compared to those of Monte Carlo photons and neutral mesons to determine the relative fraction of each present in the data. As a discriminant variable, we use the logarithm of the fractional EM] energy E log(EMl/E) = loglO(EE—MI) (5.3) total with the denominator Em“, calculated as the sum of the layers: 5 Em = 2 E. (5.4) n = l (where n = 5 represents the PH] layer). This quantity emphasizes the region of small fractional EM] energy where the greatest difference between photons and mesons is seen. 5.2.2 Background Monte Carlo Previous analyses [16] [2] used single 11:0 and 11 mesons, generated at discrete energies, D¢GEAN Ted, and with the appr0priate noln production ratio, as a model of the EM jet background. While this model provided the primary constituent of an EM jet, and reproduced the data EM] distributions fairly well, it was found that the single particles did not recreate some other distributions such as the isolation ET [16]. In general, the Monte 68 Carlo mesons were found to be more isolated than data showers, which is to be expected since real jets produce a spectrum of particles, and in EM jets, the meson that carries most of the momentum is usually accompanied by softer particles. Further evidence that a more detailed background model was desirable came from the fact that when considering the n0 and 11 backgrounds separately, comparison to data favored the 1] even though the n0 is expected to be produced and selected more often. Since the 11 Monte Carlo included the 3n0 and n0n+n' decay states as well as the W state, this indicated a preference for more complex, multi-particle showers. In order to provide a better background model for this analysis, a large number of PYTHIA QCD dijet events were generated in a falling ET spectrum over a range similar to the Monte Carlo photons (Section 4.1.3). Just as in the data, these events were then “filtered” to find EM-fluctuated jets: at the particle level, n0 and n mesons were selected by requiring . [ET(R = 0.4) —ET(R = 0.2)] < 2 GeV where R = Janz + 11112 around the meson 0 No charged particle within a cone of R = 0.2 0 Energy within a cone of R = 0.2 > 60% of parent parton energy Only events passing this filter (~0.l-0.01%) were then run through DGGEANT; in addition, only a cone of R = 1.0 around the parent meson was simulated to save 69 processing time. It is important to note that even with the particle level filter, a major constraint on Monte Carlo production was the availability of processing power. This has also been a factor in the past, and is in fact one reason why previous analyses were limited to single particle Monte Carlo. The Monte Carlo events used in this analysis were generated on a farm of DEC 3000/400 Alpha workstations at the Supercomputer Computations Research Institute in Tallahassee, FL; a fully detailed DflGEANT event required approximately one hour of CPU time to generate. Also, while the particle level filter was intended to avoid unnecessary simulation of events which would fail the offline cuts, it was necessarily loose enough to ensure a reasonably realistic EM jet sample. As a result, the fraction of DQGEANTed EM jets which passed the photon cuts (and could therefore be used in modeling the data sample) was still only about 30%. As in the case of the photon Monte Carlo, events were overlaid with non-zero suppressed zero bias data to add detector and multiple event effects. 5.2.3 EMl Calibration Because this method of background estimation relies on accurate modeling of the data by Monte Carlo, it is important to verify that the Monte Carlo reproduces the data distributions. Such agreement was demonstrated in Section 4.1.3; however, the log(EMl/E) distributions must also be examined. These distributions are shown in Figure 5.2 for W electrons from data and from Monte Carlo. While the central region Monte Carlo (left) shows good agreement in shape and peak with the data, the forward region (right) shows a slight offset - the Monte Carlo 70 CC + EC Normalized Events llllllllllllllllllllllTTlllllllllllllllllllllll Normalized Events llIlTrlllIIIIIITTTIIIIIIIIllllllllllllll l l L I I l l 1 cl.“ 1 1 1 1 1 1 l - 4 -— 3 —- 2 — 1 0 - 2 — 1 0 log(EMl/ETOT) log(EMl/ETOT) llllllll I 4:. I (A Figure 5.2: Log(EMllE) distributions for W electrons from data (points) and Monte Carlo (histograms). showers have a slightly smaller fraction of energy deposited in BM] than the data. This difference directly affects the background estimation, since a shift in the Monte Carlo distributions represents a shift in the position of the data between the extremes of the Monte Carlo jet and photon behavior. To ensure that the Monte Carlo best models the data, this offset must be corrected. The reasons for the difference are considered in detail in Appendix C, where it is determined that the offset is consistent with a difference in EMl sampling weight between the Monte Carlo and the data. Since the sampling weight is simply a multiplicative factor applied to the raw energies read out from EM], the Monte Carlo can be properly calibrated by applying another multiplicative factor to the EM] energy, or equivalently an 71 additive offset to the log(EMl/E) distributions. Using the W electron data, the offsets were derived separately for the North and South endcap calorimeters, and were found to be 0.147 i 0.013 (North), and 0.094 :1: 0.013 (South) in the log(EMl/E) variable. For completeness, a very small central region offset (0.014 i 0.005) was also applied. 5.3 Purity Calculation An example of the log(EMl/E) distributions for central region data, Monte Carlo photons, and Monte Carlo EM jets between 33 and 39 GeV is shown in Figure 5.3. The Monte Carlo distributions are binned in 50 bins across the range —4.0 E 1- 8 ~ 8 1 4: T— .r: I. m 0.6 _ but 06 _ 0.4 ‘- 0.4 l 1— 1— 0.2 :- 02 :_ - r O —4111’ l 1 1111111 0 11111 1 11111111 10 102 10 102 ET(GeV) El.(GeV) Figure 5.5: Variations in the CC (left) and EC (right) purity fits, using the nominal (solid), exponential (dotted) and straight line (dashed) fit functions. as the error on the fit to the nominal form, augmented in quadrature by the difference between the nominal and the two alternate forms. However, because this fit dependence 77 affects the lowest ET points by a large fraction of their value, and because the low ET region suffers from a lack of well determined purity points, it is difficult to extrapolate into this region. We therefore restrict the use of the purity to values of ET above 15 GeV, corresponding to the first measured point in the EC and the first statistically significant point in the CC. Additionally, for the forward region, the uncertainty due to the EMI offset was estimated by varying the nominal North & South offsets by i1 0 , and fitting the purity for each case. Since an increase or decrease in the offset directly shifts the purity points up or down, this results in an overall shift of the purity function (Figure 5.6). The uncertainty due to the offset was calculated as P —P + P —P Goff“; = I +10 nomlzl nom -lol (59) where the P are the purity functions (as in Equation 5.6) for each case. This was then added in quadrature with the other uncertainties to arrive at the final uncertainty in the forward region. 78 .0 on Photon Purity .0 \1 0.5 0.4 0.3 O O) IIIUIIYIIIT'IIIUIIIIT‘TYUII'YIIIYYIIII .‘h. " ..‘. o 0 2 v ' , . ‘a a 0.1 VWTIYIHF. ‘. 0 l 1 1 l l l I 1 I 1 l l l l 6 7 8 910 20 30 40 50 60 7080 100 Edam Figure 5.6: Effect of a plus 16 (dotted) and minus 10' (dashed) shift in the EC EMl offset, relative to the nominal (solid) purity curve. Chapter 6 Cross Sections and Theoretical Com- parisons 6.1 Cross Section Calculation The differential cross section for isolated direct photon production (measured as a function of photon ET, for specific regions of photon rapidity), is given by do _ NP dE _ LAETAnAesele (6.1) 8 trig track where 0 N is the number of photon candidates in each bin of ET and 11 0 P is the photon purity (as determined in Chapter 5) 0 L is the integrated luminosity over which the data were taken 79 80 ° AET is the transverse energy range spanned by each bin 0 AT] is the rapidity range spanned by each bin (both the central (0 S Inl S 0.9) and forward (1.6 S In] S 2.5 ) region measurements cover a range of An = 1.8) 0 A is the geometric acceptance for photons (as calculated in Chapter 4) 0 838, is the efficiency of the photon selection cuts (as calculated in Chapter 4) am“ is the efficiency of the charged track veto (as calculated in Chapter 4) Note that there is an uncertainty associated with the ET binning due to the electromagnetic energy scale, which has a measured uncertainty in the central region of approximately 0.2% [19]. Since the direct photon falls steeply (of order ET'S), the resulting error on the central region cross section is taken as 1%. While the forward calorimeter is expected to perform equally well, its calibration has not yet been verified to the same precision, resulting in a 4% energy scale uncertainty on the forward cross section. Using the components of Equation 6.1 as described previously, we can form the central and forward region cross sections; these are plotted versus photon ET in Figure 6.1. The data points and errors are summarized for the central region in Table 6.1 and for the forward region in Table 6.2. Statistical errors on each point were calculated as the Poisson error J1?! . Systematic uncertainties from each component were added in quadrature to arrive at a total systematic uncertainty (plotted on each point in Figure 6.1 as horizontal 81 $1055 304;: . |n|<0o9 g. NLooco.CTEQ4M.u=E¥ .~103F :— 49 E 2. '6 : g! 1 I O N I 1 I rntrl C 1. 10 F ‘1! E n : 1. . ..H- 1 :- ‘. : ‘1‘ i- ~‘~ . “~.‘ -1 “t. 10 :— “a. :1 1. l. 1 1 1 1 _L_1_ l l 1 1 1 1 1 1 1 #LLJ L l I l l L l l l 20 40 60 80 100 120 E, (hm g : DQPreliminary 210‘? ‘ 3 f X. I 1.6<|n|<2.5 r- ,_ 103.E.— a? NLO QCD,CTEQ4M,u=E; : I \ I ‘- Nb1025— ‘g. '5 : ~!. I g. 10 &- ‘9 '9 >- ‘.\ 1 E'— ‘.‘ E ‘l‘ - I —1 ‘l. 10 .— x l. E “l. -2 a 10 E— 311111111...1...111.1...I.$411.. 20 40 60 80 100 120 f 40 E, (GeV) Figure 6.1: Inclusive isolated photon cross sections in the central (top) and forward (bottom) regions. 82 Table 6.1: Central region cross section points and errors. ET bin Mean E Cross Section Statistical Purity Systematic (GeV) T (pb'l) Error (%) Error (%) Error (%) 15-18 16.3 0.161E+04 6.2 22.3 23.3 18-21 19.3 O.779E+03 9.7 18.3 19.6 21-24 22.4 O.343E+03 3.8 15.5 16.9 24-27 25.4 O.198E+03 5.4 13.2 14.8 27-30 28.4 O.133E+03 6.9 11.2 13.1 30-33 31.3 O.775E+02 9.4 9.7 11.9 33-36 34.4 0.539E+02 1.5 8.4 10.8 36-39 37.4 O.388E+02 1.8 7.4 10.0 3942 40.4 O.27OE+02 2.2 6.5 9.4 4245 43.4 O.206E+02 2.6 5.9 9.0 45-48 46.4 0.150E+02 3.1 5.4 8.6 48-51 49.4 0110sz 3.7 4.9 8.4 51-54 52.4 O.824E+01 4.3 4.6 8.2 54-60 56.8 O.550E+01 1.5 4.2 8.0 60-65 62.3 0.37213+01 2.0 3.9 7.8 65-70 67.4 O.244E+01 2.6 3.7 7.7 70-75 72.4 0.171E+01 3.1 3.6 7.7 75—80 77.4 0.125E+01 3.7 3.6 7.7 80-85 82.5 O.967E+OO 4.3 3.7 7.7 85-90 87.4 O.731E+00 5.0 3.9 7.8 90105 97.0 0.444E+00 3.8 4.4 8.1 105-120 111.4 O.209E+OO 5.6 5.4 8.7 120-150 132.3 0.829E-Ol 6.5 7.0 9.8 Table 6.2: Forward region cross section points and errors. 83 E T bin Mean Cross Section Statistical Purity Systematic (GeV) ET (pb'l) Error (%) ”123' Error (%) 15-18 16.3 O.217E+04 8.0 16.2 18.7 18-21 19.2 O.961E+03 12.6 14.0 16.8 21-24 22.3 0.461E+O3 4.6 12.3 15.4 24-27 25.3 O.243E+03 6.6 11.0 14.4 27-30 28.4 O.139E+03 9.0 9.9 13.6 30-33 31.5 O.865E+02 11.6 8.9 12.9 33-36 34.4 0.510E+02 1.9 8.2 12.4 36-39 37 .4 0341sz 2.4 7.5 12.0 39-42 40.4 0235sz 2.9 7 .0 11.7 42-45 43.4 O.165E+02 3.5 6.6 1 1 .4 45-48 46.4 0.107E+02 4.4 6.3 11.3 48-51 49.4 0.656E+01 5.7 6.1 11.2 51-54 52.4 0.533E+01 6.4 6.1 11.1 54-60 56.7 0.318E+01 2.3 6.1 11.2 60-66 62.7 O.164E+01 3.3 6.4 11.3 66-72 68.8 O.907E+OO 4.5 6.8 11.6 72-7 8 74.9 O.493E+OO 6.2 7 .4 11.9 78-84 80.8 0.325E+00 7 .7 7.9 12.3 84-93 88.1 O.137E+00 9.9 8.6 12.8 93-102 97.2 0.713E-01 13.9 9.5 13.3 102-120 109.0 0.305E-01 15.2 10.5 14.1 120-150 129.9 0.447E-02 31.6 11.9 15.2 84 bars); as an example, the systematic errors on the central and forward 39-42 GeV bins are shown in Table 6.3. Table 6.3: Contributions to the total systematic error on the 39-42 GeV cross section bins. E scale 41%) (.70) 2.41%) 8%" L(%) P(%) Totat(%) CC 1.0 1.0 3.5 1.6 5.4 6.5 9.4 EC 0.8 4.0 3.4 5.4 5.4 7.0 11.7 It should also be noted that the results of this analysis differ from those found in previous D0 analyses [16][2][3]. The differences, which are mainly due to improvements in the modeling of the detector and the EM jet backgrounds, are discussed in Appendix D. 6.2 Comparison with Theory Also shown in Figure 6.1 is a next-to-leading order prediction from Baer, Ohnemus, and Owens [25] using the CTEQ4M parton distributions [31]. For this calculation both the renormalization and factorization scales were set to the ET of the photon. The Dfl photon and jet energy resolution functions were applied to the theory; this resulted in a negligible change. Also, an isolation cut was applied at the parton level to match that applied to the data. 85 In order to show more clearly the comparison between data and theory, the quantity Data — Theory Theory (6.2) is plotted in Figure 6.2; statistical errors are indicated on the points. The grey band indicates the combined systematic errors, some elements of which (such as the luminosity) are correlated across the ET range. In the central region, the data lie slightly above the theoretical prediction, but are consistent given the systematic error. In particular, the overall shape of the distribution agrees well, although a slight increase in cross section is seen in the very lowest points. Since the systematic error (mainly from the purity) becomes large at low ET, this shape deviation is not necessarily significant. In the forward region, the data are consistently higher than the theoretical prediction, especially in the lowest ET bins where the data are almost a factor of two above the prediction. Also, a difference in shape between data and theory is evident in the region below ~ 40 GeV. In order to check the effect of fitting the purity points to a functional form, the cross sections were also calculated using the same coarse binning as in the purity determination - each cross section point was then adjusted by the corresponding purity value for that bin, rather than the fitted value. The resulting (Data-Theory)/Theory plots are shown in Figure 6.3. The systematic and statistical uncertainties are plotted separately on each point; the systematic uncertainty is the error on the purity in each bin (see Figure 5.4), augmented in the forward region by the EMI offset uncertainty (Equation 5.9). A comparison of Figure 6.2 and Figure 6.3 indicates no significant difference in shape, and in particular 86 “2 >3, : I‘ll<°-9 D0 Preliminary - : NLO QCD,CTEQ4M,11=E} .935 r 1 ._ 0.5 '- ++ + a...“ 0 ¢ 0 _ ......... 91. -1 ............ 9 ....... 9 3.? ........... 1 ......... 1 ........ —O.5 l _, // /////////// /////‘//6/// Mia/WM/M/i/MV/M O 100 120 140 1:; (GeV) «2 £5. : 1-60l1 ‘1}10‘ ‘ ‘ EHGeV) 88 demonstrates that the low ET excess in the forward region is not an artifact of the purity fit. A low ET increase in cross section over that predicted by theory has been seen by the CDF collaboration [32] in a measurement limited to the central region only (0 S Inl S 0.9 ). The CTEQ collaboration, in an analysis of direct photon results from that and past experiments [33], concluded that there is a shape difference between experiment and theory which cannot easily be accounted for in the parton distribution functions. Recent theoretical work has included attempts to account for this apparent excess through soft gluon emission [34]. On the other hand, some theorists argue that the current theory is compatible with experiment given the uncertainty due to the choice of theoretical parameters [13], or that theoretical problems in implementing the isolation criteria may play a role [35]. In Figure 6.4 we show the variation of the theoretical cross section with a reasonable change of the renormalization and factorization scales, to that of twice or one half the value of the photon ET. The resulting shift in the cross section is less than 10%, and is flat over the ET range; however, it must be pointed out that the two scales were varied in tandem, while some of the effects noted in [13] arise from separate variation of the scales. It should also be noted that the imposition of an isolation requirement must be matched to the particular experiment to which the theory is compared - therefore, theoretical problems in implementing the cut may have different effects when comparing to different experiments. While it is beyond the scope of this thesis to present a detailed examination of these and other theoretical issues, it is important when comparing data and theory to consider that the theoretical treatment of direct photons is an area of much current debate and 89 «0.5 ___ £i‘14 ; 013<|n|< 0.96 to! - isolation cut: [ET(R = 0.4) — ET(R = 0.2 )1 < 2 GeV H-matrix x2 < 150 The photons were required to have ET above 45 GeV in order to select a region where the photon purity is enhanced. The sample was further subjected to the “golden” photon cut, which requires less than 1% of energy deposited in the EMI layer, and which is expected to increase the purity of the photon sample since multi-photon backgrounds are, on average, expected to shower earlier than direct photons. In addition to the above cuts on the photon, standard jet quality cuts were applied to the leading jet: 0 E9“ > 2OGeV 0 At least 5% but not more than 95% of jet energy in the EM layers 0 No more than 40% of jet energy in the coarse hadronic layers (to prevent lost energy due to punch-through) - Ratio of highest to next highest energy calorimeter cell < 10 (to eliminate the 97 effects of spurious signals) 7 .2.3 Technique and Candidate Sample Behavior The sample was binned in 5 regions of the photon pseudorapidity lnYI (0.0-0.3, 0.3- 0.6, 0.6-0.9, 1.5-2.0, 2.0-2.5). The "signed nie’" of the leading jet in the event was then formed: Signed 11"“ = Signmy) . nj" (7.1) This quantity indicates the degree to which the leading jet is on the same side of the final state (in rapidity) as the photon. The signed ni“ distribution for the candidate sample is plotted in Figure 7.2. The data show a tendency for the leading jet to follow the photon candidate forward (though not fully), Confirming the expected behavior. Also shown in Figure 7.2 are the mean values of the signed nj“ distribution for each bin. 7 .2.4 Background Behavior Since the candidate sample is not background subtracted, it is important to gauge the effect of the jet component of the distribution. To investigate this, candidates in the raw photon sample were anti-selected - i.e., selected especially to fail the photon cuts. This was done by requiring: EEM] E > 0.01 (anti-golden) to! 0 [ET-(R = 0.4) — ET(R = 0.2)] > 2 GeV (anti-isolated) DO Preliminary Mean 17,, 0.07 0.19 0.26 0.59 0.61 98 IlrllTll l ’5’ ..fi- 1111.1?[111111111111 1,. “3‘!- :1131‘11111111111‘1114.“ 11111111 -fi- 1.LJ_I-1P+11ILL11I1111 p— 7‘3" IIIIIIII I ’? 114,1!I'fil’1flL-‘LL11111111 11151 I1111111‘1fi‘1_.1._1_11 -*__#_-*— -fi- —*‘ 11 iii-11111 1:11 ljfiy-LLLI 1 -3?— mimi‘i'hmlmr “that:- .1 . -fi- IJLLIJ1 JlllLLLll llllIIllTl_llllTTIIl .fi- 1111 71*“1111111111 4.5- 1111I1111L12:1*17I_1_L11 -4 —3 -2 —1 0 77.1 1 2 3 0.0<|77,|<0.3 0.3 150 (poor shower shape) with the acceptance and missing Er cuts retained as in the golden photon case. While there is a limit to how "bad" a candidate can get, due to the trigger and reconstruction quality cuts already placed on raw photon candidates, this "bad" photon sample is expected to be mostly composed of jets which have fluctuated to electromagnetic energy. The signed If“ distributions for this sample are shown in Figure 7.3, and indicate a much smaller tendency for the leading jet to follow the "bad" photon forward. This is expected, since dijet production reflects processes such as g g —> gg , in which the interacting partons tend to have similar average momenta. 7 .25 Theoretical Prediction and Comparisons Using the Monte Carlo program of Baer, Ohnemus, and Owens, a NLL QCD prediction for the leading jet behavior in direct photon events was obtained (Figure 7.4). Energy resolution smearing and a photon isolation cut matching that used on the data were applied to the prediction, which used the CTEQ2M parton distributions. Note that the 00 Preliminary Mean 77,. 0.03 -0.01 -0.02 0.14 0.18 100 11 111111111111 — 43*: 373 _ -‘A'- o‘er-“‘1" _ _ _*_ ‘ . ‘fi' #- -*- _*_ _. -fir— “A'- 1 11.1.4.1??1111111111119311111111111m1111 L- 4,. $31. _ raga- __ “A" -*- _ -fi- -*- _ ..*. -*— lilllIll 11.1.4.1'9'1—111111111111 xii-i qr 1‘91: 1.1111 111111-1-4—1-11 .fi. 0*. &- 11.1.1_I'*l'111111111111 11«1*5-11111111'1*‘I..1_1_11 : -*~ &Zfi" 1:": - —3‘r— 41*], :— -4— .4 : -k- * f1 -— “*- 1111I1131» {ll—lfi'I_1_.Lll .fi- 11 *1'1111111111LL _ _*_.*_ 1- .fi- afi- *’ ‘1‘:— '_ g. _ _fi_ 1? ' -*— .1 - 4‘1"- 11111L111L4T1-11L11 -4-3-2-101 77.1 234 0.0\ 0 C .22 .2 1 I I": c o . ...... 1.... H— ” l 0.8 is l. 0.6 o 0.0 s n a; 0.9 . > I I 1.6 s 17 :E 2.5 0.4 0.2 0 1 l +L 1 1 l o 5 10 15 20 25 Er Figure A2: 6 GeV Level 2 Filter efficiency for the central and forward regions. 114 Efficiency _._.. '01 —O— I I t l 0.6 r T O 0.Cl§n§09 0.4 0.2 Er Figure A3: 14 GeV Level 2 Filter efficiency for the central and forward regions. trigger efficiency was evaluated using a special “mark-and-pass” run in which the photon special run filters were set to pass all events regardless of the filter evaluation. In this way an unbiased source of candidates was obtained; however, in such a sample the number which pass offline selection cuts is very small, and the trigger efficiency determination suffers from low statistics. Efficiencies for the 14 GeV filter were determined using events which passed the 6 GeV filter, since it is fully efficient by 14 GeV and provides a sample of events which should pass the 14 GeV filter. The 25 GeV filter efficiencies were determined from the 14 115 Efficiency 5 0.8 I 0.6 t ooosnsos l16§n§25 0.4 0.2 + Figure AA: 25 GeV Level 2 Filter efficiency for the central and forward regions. GeV filter in the same way, making use of a special run filter which is identical to the 25 GeV EM1_GIS filter used in global running. The 40 GeV filter efficiency was determined using global data, checking the outcome of the 40 GeV filter on events which passed the 25 GeV filter. Inspecting these efficiency curves, the ET above which each trigger is completely efficient was determined to be 10 GeV for GAM_6_ISO_GAM, 21 GeV for GAM_14_ISO_GAM, 33 GeV for EM1_GIS, and 54 GeV for EM1_GIS_HIGH. 116 Efficiency d I T 06 - C) OJ)§17§(19 1J i ll 1|i§17§12‘i l 0A 02 l O Illl llll llll lfilillll Llll llll l O 10 20 30 4O 50 60 70 E: Figure A.5: 40 GeV Level 2 Filter efficiency for the central and forward regions. 117 A.2 Integrated Luminosity Determination The integrated luminosity for each filter was determined using the D0 Production Database, which stores information on beam and trigger conditions for each run. The instantaneous luminosity L is determined as RLO measured = 6:0 (A3) L where RLO is the rate at which the Level 0 Trigger indicates inelastic collisions, and GL0 is the cross section subtended by the Level 0 counters. Further adjustments are made to account for multiple interactions and other effects; the final luminosity calculation has an overall uncertainty of 5.4% [39]. Production Database utilities integrate the instantaneous luminosity readings during each run and adjust appropriately for any trigger prescales in effect for that run in order to arrive at the proper total luminosity for which the filter was active. This luminosity calculation is only valid for events which had a Level 0 vertex found within 100 cm. of the center of the detector; therefore a cut was imposed on the direct photon events to enforce this requirement (less than 2% of the events were removed due to this cut). Luminosities for each filter were calculated for each run used in this analysis; for the two lowest filters this represents approximately 30 dedicated special runs, while for the two higher filters this spans the thousands of runs collected during 1994-1995. The resulting total integrated luminosities are shown in Table A.2. The spectrum of good (passing all photon cuts) photon candidates passing each of the filters, corrected only by the relevant integrated luminosities and without any background 118 Table A.2: Integrated Luminosities for the four single photon filters. filter Luminosity (nb'l) GAM_6_ISO_GAM 13.644 GAM_14_ISO_GAM 225.29 EM1_GIS 14577 EM1_GIS_HIGH 94657 subtraction, is shown in Figure 8.1 (CC only shown). It is expected that each of the filters should line up with the next, since correcting for the calculated luminosity and prescales should reconstruct the equivalent of a single filter allowed to populate the entire ET region. The agreement appears good; however, a more quantitative check of this can be obtained by comparing the number of events observed by one filter to that of the next higher in an ET region where both filters are fully efficient. We would then expect that Nlower X Lhigher Nhigher Lzower = (AA) The results of such a comparison are shown in Table A3 and confirm that the trigger regions agree given the statistical uncertainties. O u dN/LdndP, O 10 119 I IIIIIIU l I Tillllll l IIIIIIII D 9 O O O o 0 gam 6 iso gam o A gam 14 iso gam o 0 eml gis °o * em1 gis high :- o : O ' A " A : 3" _ . “if A I I IIIIII a 3+ .fi . §_ :1 , Kn — i : O . N . , l? , <3 t fly llllllllllllllllll‘llllllllllllllllIIIWIIIJJI O 10 20 3O 4O 50 6O 70 80 90 ET Figure 8.1: Matching of luminosity-corrected filters (CC only). 120 from the Production database. Table A3: Comparison of luminosities calculated using Equation A.4 with those taken Filter (CC) LPDB (nb'l) Lea!c (nb'l) Stat. Err GAM_6_ISO_GAM 13.644 12.43 1.2 GAM_14_ISO_GAM 225.29 222.8 14.3 EM1_GIS 14577 14724 367 Filter (EC) LPDLmb'l) Lglc (nb'l) Stat. Err GAM_6_ISO_GAM 13.644 14.27 1.3 GAM_14_ISO_GAM 225.29 252.3 23 EM1_GIS 14577 15970 730 Appendix B Tracking Considerations Since the selection of direct photon candidates requires the absence of a matching track in the CDC/FDC, it is important to determine not only the efficiency of these detectors, but also the ways in which real direct photons might be misidentified due to tracking considerations. Specifically, " A photon might convert to an e+e‘ pair in the material in front of the CDC/FDC; only the fraction of such conversions whose track is accurately reconstructed will be removed from the sample. ’ A charged track produced as part of the underlying event might overlap the photon closely enough to lie within the tracking road, in which case the photon will be removed from the sample. The fFaction snack of real direct photons remaining in the sample after the tracking cut is 121 122 then 8mm], = (1 -Pt)(1 -\II) (B.1) where p is the probability that a photon will convert before the tracking chambers, measured at 0.095i0.00985 (038610.041) in the CC (EC) [16]. The tracking efficiency t is the probability that a track present in the detector will be reconstructed. The track overlap probability w is the probability that a reconstructed track from the underlying event will be found in the tracking road of a good photon. It should be noted that in the data set used in this analysis, EM clusters are matched to tracks from any primary vertex, although this was found to add only 4% (2%) to the number of raw candidates with tracks. B.l The Tracking Efficiency The method used to measure the tracking efficiency t is similar to that employed in Previous tracking efficiency studies at DC [40][41]. A sample of Z —> e+e' events is selected as follows: ' Select the EM2_GIS_GAM filter, which requires two EM clusters with ET above 20 GeV at Level 2; one cluster is required to pass shape and isolation cuts, the other only shape cuts. ' Require 0 jets in the event to avoid electrons which have overlapped a jet; this 123 ensures a similar topology to photon events where the jet is generally in the opposite hemisphere from the photon. Require that the reconstructed E7 of both EM clusters be above 25 GeV. Require that both EM clusters satisfy the direct photon 7], 11"“ cuts to eliminate the ICD region. Exclude EM clusters in the central region which are within 10% of a module edge in (1). This selects a fairly clean sample of electrons from Z bosons. Both electrons are then checked against the following strict requirements: 11 < 0.9 at least one track in road Track match significance 0', < 5 (10) in the CC (EC) EM fraction > 95% Isolation fraction < 10% Hmatrix x2 < 100 Passed Level 2 shape, isolation, and ET cuts 124 Those electrons passing these requirements are considered “tagged” as good Z electrons; the other electron is then considered unbiased and can be used in measuring the tracking efficiency. If both electrons in the event are tagged, then both are used as unbiased. Standard photon selection cuts (minus the track cut) are then applied to each unbiased candidate. For each passing candidate, an entry is made into a histogram of the invariant mass of the two electrons (separately for CC and EC candidates). A further requirement is then made that candidates in the CC (EC) have a track with 0', < 5 (10) , and two more histograms are filled for these candidates - we then have histograms of the mass spectra for events before and after the tracking cut, for both the central and forward regions (Figure B. 1). The tracking efficiency can then be determined by calculating how many electrons from real Z —-) e+e' events were found to have passed the track requirement. To establish the level of background in the sample, we define a region S as the signal region, between 86 and 96 GeV/c2 - this region should contain mostly real Z events with some background events. We define two equal regions on either side of the mass peak as B] (61 - 71 GeV/c2) and 82 (111 - 121 GeV/c2), which should contain only background events. Using the technique established in [40][41], we assume a linear background distribution, approximating the level of background in the signal region S as NB, +ng2 NB: 2 (B.2) 250 200 150 100 50 300 250 200 150 100 50 125 10 101 :— Entries was __ noon 90.90 _ Rus 3.794 , L LIL-d] l l 1 1 1 60 80 1 00 1 20 1 40 CC candidates passing track cut -— ID 100 "’ Entries 2396 : than 90.63 .. RMS 9.134 ll E l... V 60 80 100120 140 ‘ CC Candidates before cut 100 80 6O 40 20 140 120 100 80 60 4O 20 :“ game. :37 L C C 60 80 100 1 20 1 40 EC candidates passing track cut g 3.... 3° _ than 92.71 __ nus 3.535 C L ’- 1 L11] ml. 1 l l I l . 60 80 1 00 1 20 1 40 EC Candidates before cut Figure B. 1: Z mass distributions before and after track requirement. 126 in which case the tracking efficiency becomes where the subscript T indicates the number in each region which have been tracked. t— _ NTS‘NTB NS-NB (B3) The resulting tracking efficiencies are shown in Table B.l, along with estimates of the background in each sample. Systematic error due to the background estimation is Table B. 1: Tracking efficiencies and errors in the central and forward regions. . Stat. Sys. Background Efficrency err Error ( % ) CC 0.830 0.0096 0.0029 2.66 :1: 0.4 EC 0.767 0.016 0.0037 2.54 :1: 0.6 estimated by taking the difference in efficiency with and without background subtraction. Note that because the definition of “tracked” applied here includes a cut on the track match significance, the resulting efficiencies are lower than in [40] [41]. This is necessary because, since we are considering the possibility of overlaps separately, we wish to extract the probability that an existing track which is directly pointing to the cluster (i.e., from conversions) will be seen by the tracker. Loosening the definition of “tracked” to mean simply finding a track in the tracking road would doubly count the effect of overlapping tracks. 127 The mean luminosity of the runs used is 6.6X1030cm-zs_l (6.6E30); runs ranged from 0E30 to 20E30. The change in tracking efficiency with increasing instantaneous luminosity is shown in Table B.2; in the central region, the efficiency decreases as the Table B.2: Tracking efficiencies at different instantaneous luminosities. Luminosity . Stat. Sys. Background Range Efficiency Error Error (%) CC 0.898 0.028 0.0066 1.7 :1: 1.2 0 - 3E30 EC 0.780 0.059 0.0043 2.0 :1: 2.0 CC 0.860 0.016 0.0022 3.0 :1: 0.8 0 - SE30 EC 0.814 0.026 0.0077 2.8 :1: 1.1 CC 0.822 0.013 0.0064 2.8 :1: 0.5 5 - 10E30 EC 0.742 0.021 0.0009 2.7 :t 0.7 CC 0.792 0.034 0.0040 1.4 :1: 0.6 10 - 15E30 EC 0.786 0.054 0.0067 0.8 :1: 1.0 luminosity goes up, which is consistent with previous studies and reflects the fact that vertex finding becomes poorer as the average number of vertices increases, causing the tracking road to be drawn inaccurately. This effect is less noticeable in the EC, due to the higher average number of tracks and the larger z region which the tracking road intercepts at smaller angles. The lowest bin, between 0E30 and 3E30, has a mean luminosity of 2.1E30, which is close to the average luminosity of the photon special runs from which 128 data below 33 GeV are taken. B.2 Track Overlap Probabilities B.2.l Golden Overlap Method In order to estimate the probability w of a random track from the underlying event to overlap a photon, a “golden” photon sample was selected from candidates which satisfied the standard photon cuts (without applying the no track cut), and which further had less than 1% of the photon energy in the first layer. This golden requirement is > 98% efficient at rejecting electrons from Z —) e+e' events (Figure B.2), since charged particles begin depositing energy as soon as they enter the calorimeter, whereas photons must convert to an e+e' pair first. Any tracks in the tracking road pointing to a golden photon candidate are therefore presumed to be overlaps, rather than a real track associated with the calorimeter cluster, and the overlap probability is obtained by observing how many golden candidates have tracks. This is similar to a technique used for the Run 1A analysis. The results of this method are shown in Table B3. The overlap probabilities obtained are quite consistent when evaluated for samples obtained from the special runs (6 and 14 GeV Level 2 thresholds), the global EM1_GIS filter (25 GeV threshold), and the global EM1_GIS_HIGH filter (40 GeV threshold). While this method has the advantage of using a subset of the direct photon sample, and thus matching the luminosity & trigger profiles and E7- range of the overall sample, one possible source of uncertainty in this method comes from the fact that the sample is not 350 300 250 200 150 100 50 129 IIITTTIIIIIIIIIIIIIIIIIIIIIFIIII IIII CC Z candidates 180 160 140 120 100 80 60 4O 20 "r E. E. C E- :1_Lll_l_.l_L [111111—11 —4 —3 —2 —1 O log(EMllE) ECanndidates Figure B.2: Log(EMllE) distributions for Z electrons. Note that 1% EM] corresponds to -2.0 on this scale. Table B.3: Overlap probabilities from golden candidate sample. - olden golden Stat Trigger NTg N w error CC 169 1558 0.108 0.008 Special Runs EC 435 2129 0.204 0.009 CC 564 5810 0.0971 0.004 EM1_GIS EC 1481 7105 0.208 0.005 CC 707 7184 0.0984 0.004 EM1_GIS_HIGH EC 1408 6485 0.217 0.005 130 pure photons. While the golden photon cut does boost the purity of the sample by preferentially selecting single photons over multiple photon backgrounds (which tend to convert earlier), the remaining background may unfairly add overlaps which are not from the underlying event but from jet fragmentation, which will not affect real photons. One might hope that by making the isolation or golden cuts more restrictive, a better purity might be obtained, and the overlap probabilities would approach a stable value. This was found not to be the case, as shown in Table 3.4 - requiring progressively less isolation ET Table B.4: Variation of the overlap probability with isolation and golden cuts, varied independently using the EM1_GIS_HIGH trigger. 12:33“ W EStat. EMl cut 1! Stat. (GeV) rror (%) Error 2.0 0.098 0.004 .1 1.0 0.098 0.004 1.5 0.093 0.004 0.8 0.093 0.004 CC 1.0 0.083 0.004 0.5 0.081 0.004 0.5 0.070 0.007 0.3 0.069 0.005 2.0 0.217 0.005 1.0 0.217 0.005 EC 1.5 0.193 0.005 0.8 0.205 0.006 1.0 0.151 0.006 0.5 0.183 0.006 0.5 0.099 0.010 0.3 0.157 0.007 or less energy in EMl progressively lowered the overlap probability (and removing the isolation cut greatly increased the probability). This is understood when it is considered 131 that the isolation ET is in fact itself a direct measurement of underlying event activity, especially when a 2 GeV out has been imposed to reduce the isolation ET to approximately the level of underlying event. Similarly, the energy in BM] is correlated with the amount of underlying event overlapping the candidate. The level of overlap is therefore dependent on the choice of cuts for these two variables. Since the isolation cut used here is the standard cut used in this analysis, it is clear that it should give the appropriate level of overlap. The cut on EMl is less obvious; however, the level of electron rejection observed with a 1% cut appears to be a reasonable balance between allowing real associated charged tracks and overly minimizing the underlying event effect. As an estimate of the systematic error arising from these choices, the variation in w from varying the isolation cut from 2.0 - 1.5 GeV and from varying the EMl cut from 1% to 0.8% were added in quadrature. The resulting overlap probability, obtained by averaging the numbers of the two global triggers and combining systematic and statistical errors in quadrature, is then 0.0978 :1: 0.010 (0.213 t 0.026) for the CC (EC). 13.2.2 2 Overlap Method As a cross check, an alternative estimate was obtained using the Z —-> e+e' sample. Using the same method used to determine the tracking efficiency t, an efficiency was determined for a Z electron to have an additional track in its tracking road. This probability was found to be 0.108 i 0.008 (0.261 i 0.018) (errors statistical only). The analysis was also performed on a sam le of Herwig Z —> e+e' events, usin late-level P g P 132 GEANT, in order to estimate the probability that a Z electron would produce an extra track through bremsstrahlung and conversion, etc. After correcting for this effect, the final overlap probability using the Z sample is then 0.107 :l:0.008 (0.244i0.018) with statistical error only. This agrees well with the result obtained by the primary method using golden candidates. B.3 Overall Efficiency Using Equation B.1, we can then construct the overall efficiency for photons to be lost due to tracking considerations: 8 k = 0.828 :1: 0.013 (0.554 :1: 0.030) (B.4) (me While this efficiency has been calculated using global run data, with a mean instantaneous luminosity of 6.6E30, substitution of the tracking efficiency t calculated in the 0E30-3E30 bin does not change the result within errors. This is due to the stability of the overlap probability between the global and special runs, and also to the constancy of the conversion probability p and its dominant error. Therefore the values calculated in Equation B.4 can be used over the entire range of the photon ET spectrum. Appendix C EMl Calibration The observed offset in log(EMI/E) between EC Monte Carlo and EC data (Figure 5.2) directly affects the purity measurement, since a Shift in the distribution of the Monte Carlo represents a shift in the position of the data between the extremes of the Monte Carlo jet and photon distributions. In order to ensure that the Monte Carlo best models the data, this offset must therefore be understood and corrected. C.1 Offset Estimation and Origins The most direct way to correct the Monte Carlo is to determine the offset between the Monte Carlo and data W electron peaks, and then apply this additive offset to the Monte Carlo photon and jet distributions used to estimate the photon purity. The offset estimated in this way is 0.120 1 0.011 . This additive offset in log(EMI/E) is equivalent to a multiplicative correction of ~30% to the EMl energy; or more 133 134 accurately, to the sampling weight applied to the uncalibrated energy read out from EM]. The sampling weights for the EC EM layers were determined using electrons in a test beam study of EC calorimeter modules [18]; a plate level test beam Monte Carlo, the precursor to the Monte Carlo used in this analysis, was also used in the study. It is apparent from this study [28] that 0 When different methods were used to determine the sampling weights (in order to find the scheme which optimized the energy resolution), the EMI sampling weight varied much more than other layer weights. This effect was dismissed as negligible to the goal of the study, however, since the energy in BM] is a small (~4%) fraction of the total energy of the shower. 0 The EMl sampling weight varied depending on the amount of material upstream of the calorimeter. An attempt was made to approximate the amount of material which would be present in the real detector; however this was modeled only with a steel plate of uniform thickness. 0 The EMl sampling weight varied up to a factor of 3 between the testbeam data and testbeam Monte Carlo. It is also important to note, as pointed out in [29], that the sampling weights determined so as to optimize energy resolution may in fact not reproduce the exact energy in any given layer. Because the energies in each layer are correlated, the ideal weight for EMI may depend on the behavior of EM3, for example. The authors of [29] estimate that this might cause up to a 100% difference in the EM] energy when comparing Monte Carlo and data. 135 However, one expects that Since the sampling weights are simply multiplicative factors applied to each layer for all energies, the calibration of the Monte Carlo EMl weights relative to the data will provide, if not the actual energy in BM], a well-determined quantity which allows direct comparison of Monte Carlo and data to extract a background fraction. C.2 Cross Checks of the Offset Estimate As a qualitative check, the photon purity fitting routine (described in Section 5.3) was modified to allow the signal and background distributions to shift by an offset; the routine then minimized the x2 to find not only the best purity estimate, but also the best offset estimate in each ET bin. The result is shown in Figure C.l. This method is not exact, as it relies on the fitter to determine dual quantities and their errors, but it does indicate that the purity fits prefer the Monte Carlo photon and jet distributions offset by roughly the amount determined from W electrons, with no obvious ET dependence. To check more rigorously the possibility that the offset might be ET dependent, the Monte Carlo and data W samples were binned in ET and compared (Figure C.2a); within the limited statistics, there appears to be no ET dependence. To provide better statistics, an alternate sample of plate level Monte Carlo W electrons was obtained. These were generated using ISAJET as Single particles distributed with W electron kinematics [30], and therefore cannot be compared with full data events to derive an absolute offset; however, the dependence of the offset with ET can be studied. The analysis was repeated with this sample (Figure C.2b), and again no obvious dependence on ET was observed. 136 *5 0.3 — E Z O 0.2 l d] 0.1 1 Cl C] [3 D D a E3 [#3 0 L —o.1 :- L11 1 1 Li 1 1 1 lllgjl 1 ll 1 l I l l I l l l l 1 Lil l l_Ll l 10 20 50 40 50 60 ET(GeV) Figure C. 1: Values of EC EMl offset produced by the purity fitting routine for several ET bins. C.3 Effects of Pre-EMl Material in the Monte Carlo While the offset appears consistent with a mismatch in sampling weight between data and Monte Carlo, there is also the possibility that the DOGEANT simulation of the detector has mismodeled the amount of material in front of the calorimeter. If the amount of material were greater in the real detector than in the Monte Carlo, electromagnetic showers would tend to start earlier, develop further by the time they reach EMl, and thus leave more energy in BM]. The application of the offset in this case would still help correct the Monte Carlo, but there would be an additional, uncorrected effect on the probability that a photon converts to e+e' before the EMI layer, which is precisely the discriminant we hope to exploit. 137 OJ 02 '- Ql — 25 30 35 4O 45 50 55 03 OJ f1 r T I I I I t T Figure C.2: Offset measured from W electrons versus ET, using a) full W Monte Carlo; b) single particle Monte Carlo To examine this possibility, it is necessary to understand how extra material would affect the longitudinal energy deposition of an electron shower. A model of this energy deposition was created using the formula [4] dB (1:1)“ ‘6’” E = Eob 1"(a) 1 (8.1) where t is distance in radiation lengths (X0), and a and b are constants determined, in this case, for 35 GeV electrons passing through uranium. Given the known ECEM layer 138 thicknesses of 2, 2.6, 7.9, and 9.3 X0, we can then integrate Equation 8.1 in each layer, assuming different initial radiation lengths before EMl, and observe the effect on all layers. An initial estimate of 2.0 X0 before EMl was taken as a benchmark value (the design estimate quoted in [18] is 2.31X0). In order to better calibrate the model, a set of “sampling weights” was derived using the benchmark estimate so that the benchmark layer profile matched that of the data, and applied to all estimates: W _ Fndata n _ (8.2) 1“"20x0 This weighting was meant to add in the effect of “optimized” sampling weights as are present in the data; however, since the quantities of interest here are relative shifts between different pre-EMl estimates, the weighting had little effect on the results. Figure C.3 shows the fractional energy deposited in each layer for several values of pre-EMl material. Note that an increase in material results in increased energy in the first two layers, and a decrease in layers 3 and 4 (and in layer 5, though this is not distinguishable in Figure C.3 due to the small fraction of energy deposited there). Also plotted are the data (identical to the 2.0X0 points) and Monte Carlo fractional energies. Figure C.4 shows the percent (of the total Shower) change for each layer between several combinations of pre-EMl estimates, and also between the Monte Carlo and the data. For example, going from 2.0X0 to 1.5X0 results in a loss of ~1% of the total shower energy from EM], and a gain of ~1% in EM3. Note that a difference of 0.5X0 results in a change in EMl fractional energy similar to that seen between the data and the Monte Carlo; this difference is consistent between assumptions of 1.5-2.0X0 and 3.5-4.0X0 , implying that 139 - (material assumed before EMl) A 4.0x, .0 \l l 3.0)(0 [>130- C] O 2.0x, / Data I MC .0 U! 1 Fractional Energy 7’ , 0.4—- 0.3— 0.2— _ i C] e A 0.1— A : [I] 0” Lil illlLleliiiiliiLJ‘ O 1 2 3 4 5 Layer Figure C.3: Fraction of energy deposited in each EM layer for several pre-EMI estimates (simple model), data, and Monte Carlo. the observed changes are fairly stable over the range of realistic pre-EMl estimates. The data/Monte Carlo EM3 difference, however, is not consistent with a 0.5X0 change, but with a larger change of ~1.0X0. Most importantly, whereas the model shows that a decrease in material results in an increase in the EM4 and F111 energies, the data/Monte Carlo difference shows a decrease in those two layers. While the model in Equation 8.1 is fairly basic, these results indicate that the difference in EMl is not wholly due to a failure of the Monte Carlo to account for pre-EMl material correctly. Also, since 0.5X0 is roughly 140 50.04 — 13 2.0XOtol.5XO 9 I 8 __ O 4.0X0l0 3.5x° l 50.03 - A 3.0x,1e 2.0xo g l A a: “0.02 — a Data to MC .5 o 8 9 U . 1 ~— 0 E00 0 0 __ ................................................................... 121-- ' a —0.01 — g : i l t O )— -0.02 — - A —0.03 — -0.04 e 1 l l J 1 1 i L L 1 1 l J l 1 l l t J I l o 1 2 3 Layer Figure CA: Change in percent of total energy in each layer between different material estimates, and between data and Monte Carlo. 20-25% of the expected pre-EMl material, and corresponds to roughly 1cm of steel, it is unlikely that the DOGEANT simulation would exclude this much material. C.4 Final Calibration Determination In order to arrive at a final correction with which to calibrate the Monte Carlo EMl distributions, the offset was examined separately in the North and South endcap calorimeters. While the same sampling fractions are applied in both the North and South, 141 in calibration studies the two endcaps have been found to differ by about 2% in total response, and are separately calibrated offline. While the Monte Carlo sample was found to be consistent between the two endcaps, the data W EMl distribution peaks were found to differ by a statistically significant amount, resulting in different offsets in the two regions: 0.147 :1: 0.013 (North), and 0.094 i 0.013 (South). In order to achieve the best calibration of the EM] energy deposition, the cryostats were therefore corrected separately. This difference was again qualitatively checked by separating the Monte Carlo photon/jet and data photon candidates into North and South samples, and rerunning the purity fitter to attempt to fit the offset separately in each cryostat. While the results of the fits (Figure C.5) show the effects of the reduced statistics, a difference can be seen which is consistent with a greater offset in the North endcap than in the South. Note that since the EMI layer energy is such a small fraction of the total energy, the difference in total response between the two endcaps may, depending on its origin, affect the EM] significantly. Also, while the overall offset seems unlikely to be caused by conversion material differences, and there is no expected asymmetry in the detector between the North and South, a slight difference may also contribute to the relative shift. For completeness, the CC W EMl distributions were examined in similar detail to the EC, and the data and Monte Carlo peaks were found to differ by only 0.014 i 0.005. While this is a small difference compared to the EC offset, to maintain a consistent calibration method the CC Monte Carlo was adjusted by this factor, which resulted in a slight (~ 5%) increase in the CC cross section. OJ —OJ OJ -OJ 142 : l.— Ei’ North I- E: Z [:1 C.3 2— D D 9 e 4 :1 i +— :l i 1 l i l l l l l l 1 l l l l l l l i l l l l i 1 l l 1 i l l l_l I l 10 20 3O 40 50 6O 70 80 ET(GeV) E- South C a i] a 9 D [:1 D ED ; cl- e b-l I l 1 I l i 1 l l l l l l L L l L L L 1 i l l l l i 1 l l 1 l J_l l I l 1 10 20 30 4O 50 60 ET (GeV) Figure C.5: Values of EC EMl offset produced by the purity fitting routine for EC North (top) and South (bottom) samples. 143 C.5 Summary The observed offset in log(EMl/E) distributions between Monte Carlo and data W electrons has been investigated, and was found to be consistent with a variation in EMl sampling weight between data and Monte Carlo. Also, the offset does not appear to be consistent with incorrect modeling of pre-EMl material in the Monte Carlo. In order to calibrate Monte Carlo photons and EM jets for use in determining the photon purity, we derive from the W electron comparisons a set of additive constants in log(EMl/E) Table C. l: Additive constants used to calibrate the Monte Carlo log(EMl/E) distributions. Calorimeter log(EM llE) correction CC 0.014 i 0.005 ECN 0.147 :1: 0.013 ECS 0.094 i 0.013 (multiplicative constants in BM] energy). These were determined separately for the North, South, and Central calorimeters (Table C. 1), and applied to the photon and EM jet Monte Carlo log(EMl/E) distributions. Appendix D Comparison with Run 1A Results To compare this analysis with the Run 1A inclusive photon analysis [3], the cross section was rebinned to match that of [3] and the ratio of 1A/ 1B was formed (Figure D. 1). In both the central and the forward regions, the Run 1A results are found to be significantly below the Run 1B measurement. It is therefore important to consider the sources of the disagreement. D.1 Corrections to the 1A Analysis To investigate this difference, the Run 1A analysis was reviewed, and several discrepancies were found. First, the calculated acceptance for the central region was found to be higher in the 1A analysis than in the 1B. This was found to be attributable to the IETA acceptance, which was found in the 1A analysis [16] to be 97.5% compared to the updated value (Section 4.2) of 91.7% for identical IETA cuts - a difference of 6%. Note 144 145 be: r WWWVVV’W \$8 L b 1.6 L 18 la systematic error 1.4 - 1.2 >- : O 0.6 — 0.4 F + : 1A 10 systematic error % 0.2 — 7/ 0 0 20 40 60 80 100 120 140 13; (GeV) 2 7 ba / / % ‘ r $8 — é 3 é 1.6 h- 18 la systematic error 5 1.4 - TIII 0.6E wfimfijflifl] 1A 10 systematic error 0.2 IT‘TII \\\.\\\\\\\ ~ . / 3 N)" 40 so 80 100 120 140 Elm Figure D. 1: Ratio of the Run 1A inclusive cross sections to the Run 1B results in the central (top) and forward (bottom) regions. 146 that while the 1A vertex was more off-center than the IB vertex, in fact this should serve to lower the acceptance rather than increase it. It was verified that the actual cut used in the 1A analysis was 0 < IIETAI S 9; however, the acceptance quoted in the 1A analysis is more consistent with the looser cut 0< IIETAI S 10, which had been used in earlier versions of the 1A analysis. Applying the IB acceptance to the 1A cross section would raise the 1A result and bring it closer to the 1B. Also, since the Production Database (which stores all run and luminosity information) has been improved since the 1A analysis was completed, the integrated luminosities for the triggers and filters used in the 1A analysis were recalculated. For the lowest ET filters, the new numbers are 3.2% greater than those quoted in [3], which is consistent with the shift expected from a re-analysis of the DO luminosity calculation [39]. The highest filter, however, used above ~ 30 GeV, was found to be a factor of 2.7% lower than the old, implying a 5.9% discrepancy. It was found that luminosity values for a minimum-bias filter were used for the 1A analysis rather than those calculated explicitly for the photon filters, due to Production Database difficulties existing at the time. Correcting the 1A luminosity would increase the 1A cross section above ~30 GeV by 2.7%, bringing it closer to the 18 result. Another discrepancy is related to the fact that the 1A analysis fit a function to the efficiency-corrected purity (purity divided by efficiency), rather than simply the purity. Though the maximum value of the purity/efficiency quantity is therefore approximately 1.23, the fit was made assuming a maximum of 1.0, as would apply to the purity alone. Figure D.2 shows the effect of refitting the 1A central region purity points to allow for the 147 Run IA Central Region Purity/Efficiency --------- PRL fit Refit to max = 1.23 Photon Purity r 0.8 — 0.6 - O.4 -— + l. 0.2 ‘- {K 1 l I l l l 1 l l 1 l l 6 7 8 910 20 30 4O 50 60 70809000 E, Figure D.2: Effect of refitting the Run 1A central region Purity/Efficiency to allow for the proper maximum value at higher ET values. proper maximum; while only the region of ET > 70 GeV is affected, it is in this region that the 1A central cross section was found to deviate below the theory by 20-40%. Refitting the purity in the forward region yields no significant change, however, as the forward purity/efficiency was lower and therefore less sensitive to the maximum allowed value. D.2 Effects of Changes in Background Modeling The remaining discrepancies are consistent with the change in the jet background 148 modeling between the 1A and 1B analyses, and the statistical improvement in the number of purity points. In the forward region, of course, the dominant change is the calibration of the EMI distribution between Monte Carlo and data, which was not performed in the 1A analysis, and therefore causes the 1A result to be lower than the 1B result. This also produces a difference in the central region, though much smaller (~ 5%). It is expected that the “messier” nature of the PYTHIA EM jets compared to the single particle backgrounds used in the 1A analysis will tend to increase the purity of a given sample: the messier jets leave slightly more energy in EMl, increasing the difference between signal and background, as well as data and background. Since the difference between the data and the signal behavior remains the same, this results in a higher purity. In the 1A analysis, it was found that when the fitter was allowed to compose the background sample with no restriction on the 1] to 1:0 ratio, it chose a much higher fraction of us than than the expected value. This indicated that the fitter preferred more complex, multiphoton showers when trying to reproduce the data, as the n decay modes include significant 31:0 and 7:01t+1t' states, whereas the 1:0 decays almost exclusively to w. This, along with examination of quantities such as the single particle isolation ET distributions, gave evidence that the single particle Monte Carlo did not perfectly approximate EM jet behavior, and (along with an increase in available processing power) led to the generation of PYTHIA jets as an improvement for the 1B analysis. To demonstrate the improvement in background modeling, a “composite” sample was created by combining the 13 Monte Carlo photon and jet candidates in the fractions indicated by the 1B purity, and then plotting other quantities of interest (EM fraction, H- 149 matrix x2, and isolation ET) and comparing to the data distributions. These comparisons are presented in Appendix F; one example of interest is the composite isolation ET distribution. There is generally good agreement between the Monte Carlo and data; when compared to a similar study done in the 1A analysis [16], the improvement from the 1A analysis can be seen. In another study [43], the Run 1B PYTHIA Monte Carlo was used in place of the Run 1A single particle Monte Carlo to refit the 1A candidate sample and rederive purity points in the CC. Figure D.3 (left) shows the original 1A purity points (without the efficiency 0.0 '- 0.6 '- 0.4 0.2 V I r ' I r Y V ' A All I l A A A j 078910 20 30 40 506070801m El- 678910 20 30 40 506070801“) El! 4 A L A I l l A L LL A4 A A A l l Figure D.3: Left: Run 1A CC purity points and resulting fit (solid line), along with the 1B CC purity fit (dashed line). Right: 1A purity points and fit (solid line) using the 1B Monte Carlo, along with the 1A fit (dashed line, same as solid line on left). 150 correction), along with a fit to these points (solid line) and, for comparison, the 1B purity fit (dashed line). Figure D.3 (right) shows the new purity points determined using 1A data with 1B Monte Carlo, along with the resulting fit (solid line). Note that the rederived purity points and resulting fit are about 10-20% higher than the 1A result (the 1A fit is plotted again as a dashed line, for reference). This indicates that the single particle Monte Carlo used in the Run 1A analysis may have underestimated the purity, resulting in a lower cross section. D.3 Conclusions In order to estimate the combined effect of these differences, a rough correction of the Run 1A cross sections was made by applying the above factors. To estimate the effect of the EMI calibration on the 1A data, the Run 18 cross section was calculated without the calibration and a rough parametrization of the difference versus ET was employed. The resulting ratio of “corrected” 1A to 1B is shown in Figure D.4. While only a reanalysis of the Run 1A data will provide a more exact comparison, the corrected 1A result and the 1B result appear consistent. 151 .5 L 18 la systematic error e 11),)le .............. _ 9 (5 0.8 :- 0. g. i 1 1A 10 systematic error 0.2 — ., i 7% 0 0 20 I 40 60 i 80 10 120 140 EINGCV) 2 7 7 '5! \ ' I x / $8 " / ‘ / e / 2 v I % 1.6 Z. 18 la systematic error é a f 1.4 i— I? : t e 1 '— """"""""""" quQ--H-E§n# """" """rrtu ---------------- : [389%] in til 0.8 — [:3 0.6 L #1 T 0.4 :- 2 : 1A 10 systematic error / 0.2 ~ 2 // é . ,l , / / ¢ // r 0 0 20 40 so 80 ' I 100 ' 120 140 E: (GeV) Figure D.4: Ratio of the “corrected” Run 1A inclusive cross sections to the Run 1B results in the central (top) and forward (bottom) regions. Appendix E Purity Fits In the purity fits which follow, the mean ET of the sample is given by the fit parameter P1, the fit purity as the parameter P2. For clarity, statistical errors are not included on the Monte Carlo points. 152 153 :xr/nd1156.4 7 40 {/ndf6226 / 32 -P1 10.885: 0.0000E+00 *‘P1 13.2311: 0.0000E-1-00 "P2 0.112521: 0.1552 P2 0.14351 0.4419E-O1 “ o 1665 cc data .:'-~.. 5”- 789 cc data E-O 5513 CC MC 75 '1 o .1 :0 5848 CC MC 75 E :1 1180 cc MC jets .. .. ' :a 731cc MC jets .. O 1.1 II >- Milu' '- 0 '1 .000. 1‘ r r ' 3 I . I + ; I O. ‘ + .. 11 [j i- 0”. ‘u -‘ 1‘11. - I' II .- .‘}... . II . l . '11 D ! :7- 0 O l 1 1 1 l l l 1 1 l 1 l l 1 '- 1 1 1 1 l —4 -3 -2 —1 0 -4 -3 -2 —1 W313...) WW {/11813185 / 31 /ndf47.03 / 40 :P1 17.131: 0.0000E+00 PF’l 22.71 :1: 0.0000E+00 :PZ 0.2447i 0.1461 P2 0.3947i 0.4699E-01 C o 368 cc data :‘o 811 00 data ......“ :0 4333 cc MC'ys "In: E. 8119 cc MC'ys .. . __ Cl 429 CC MC jets ,' ‘ : El 1608 CC MC jets 1‘ O f‘r ... F "‘ F .....“' ' I 11 E . :— 0 . .. .. it . I I T 11 .. . ~ I P . 0 0‘ ' . ttl'm'l'. " 11 P U . l D . .0... Q ' ’1 A, l I 1 l l 1 J 1 l l l l l I 1 fl 1 l 1 l l l J l I l 41 -4 —3 —2 —1 o -4 -3 -2 -1 W ”W Figure B]: CC purity fits for ET bins of 10-12, 12-15, 15-21, and 21-25 GeV. 154 {#1015268 / 37 -x'/n01345.8 / 46 ___P‘l 27.17: 0.0000E+OO P1 31.411 0.0000E+OO ”P2 0.35055 0.5878E-01 5P2 0.37173: 0.1581E-01 7 o 422 cc data 5 0 7137 cc data E O 5709 CC MC 73 ~"3' g... r O 2907 CC MC 73 " D . ' . u " . _ 721 CC MC jets . i + _D 1029 CC MC 18.150 0‘11 . I : ... 4." = l— .O 1'. : .- I F o. I + . l' r t O I l- ’ fl ’- Cl 11 I" D III -4 -3 —2 —1 0 -4 -3 -2 -1 4711612795 / 45 'f/ndr320.4 / 44 P1 35.634: 0.0000£+00 791 44.463 0.0000900 42 0.4671 :1: 0.15666-01 ~92 0.45971 0.1618E-01 O 7887 CC data I 6703 CC MC 73 a 2884 cc 11c jets I TWIIII I T I IIIIII E ' O 5936 CC data I 15536 CC MC 78 Figure B.2: CC purity fits for ET bins of 25-30, 30-33, 33-39, and 39-54 GeV. 155 _ /n81285.3 / 44 5171 56.753: 0.0000£+00 ~92 0.55803: 0.1763E-01 I 0 4447 cc data — o 3168 CC MC 7s D1129CCMC' I I IIlIlIIl l Ef/ndeBJfi / 44 -91 64.263: 0.0000£+00 P2 0.64381: 0.20266-01 : O 3888 CC data 0 4549 CC MC 7 C] 1630 CC MC' I IIIIIII I —4 —3 -2 -1 o -4 -3 —2 -1 10813113...) will...) ; /nd1155.0 / 43 {/11616893 / 39 :91 77.391 0.00001:+00 P1 93.41 :1: 0.00001:+00 _P2 0.72611 0.2206E-01 on: 3P2 0.82631 0.4498E-01 Cl .. — o 2543 cc data : o 653 cc data .. .. DD 0 11313CC MC'ys -0 6053CCMC75 ~- 1:1 :- , ' Cl : C] 3842 CC MC jets ' C] 1 168 CC MC jets : _ Cl I l I IIII] a“: J) I IIIIIII Figure B.3: CC purity fits for ET bins of 54-60, 60-70, 70-88, and 88-100 GeV. 156 1711615525 / 40 P1 115.7:h 000001300 2 0.78872t 0.3379E-01 E o 752 cc data -""D :I7655CCMC'ys ’ D — D 1019 CC MC jets. “ II .. I. D a r 9 Cl 1 111111! \J V a Cl 13 . O _+ I Figure E.4: CC purity fits for the ET bin of 100-150 GeV. 157 f/nd138.14 / 34 131 15.34: 0.0000E+00 :P2 0.4824: 052925-01 : I 484 EC data CPD - I 7439 EC MC 78 ‘ ..'}. * a 352 EC MC jets " . t- .. . . .0 :— I '. . E . I-hu . : 'I I I I I ' ‘. "InllIIllll'D l‘ : Cl 1 ' l L 1 1 1 l i l #L L i l l I —4 -3 -2 -1 0 bill/B...) {/11613435 / 34 J/ndf41.81 / 31 P1 22.70: 0.0000£+00 :91 27.11 : 0.0000E+OO :2 0.5196: 0.5979E-01 :92 0.6029: 0.6701E-01 ” I 559 EC data I I 257 EC data E o 7064 EC MC 75 - o 4480 EC MC 75 .. - a 621 EC MC jets ‘3 . a 336 EC MC jets .. g I . .." 0 ' ... EFF] . ...— L7. :— 0... E .0. D + D C Q. j. 0 l I l- _ D : a - --' - .. O " . 11 a - .11 l 13 717111L11111L111 L1 1111111111171¥1 -4 —3 -2 — 0 -4 —3 -2 -1 0 10018.18...) Figure E.5: EC purity fits for ET bins of 10-13, 13-21, 21-25, and 25-30 GeV. 158 _ /ndf172.8 / 40 :91 31.41 : 0.0000E+OO :92 0.6142: 0.1535E—01 E I 4253 EC data - o 2132 EC MC 73 'l :1 TTIIII 8 Lifiln E l I l l l l DIJILLLll -2 =*I/ncu 144.5 / 42 :91 _ I 4471 EC data r O 4575 EC MC 75 I TTTTIII‘I 35.59 :t 0.0000E+OO P2 0.57131: 0.1725E-01 l I l 1 l 1 l l l -3 -1 0 —4 -3 —2 —1 o MEN MEN :x'/ndt 74.43 / 40 _ /ndf82.42 / 39 ~P1 44.06: 0.0000E+OO -131 56.69: 0.0000E+00 ”P2 0.6057: 0.2316E-01 92 0.6954: 025606—01 _ I 3025 EC data 0 I 1826 EC data 0 9038 EC MC 75 ' y. 1828 EC MC 75 D E0 2272 EC M ED 395 EC MCj D : ~ ad] : ' DD 0 ' 1:1 1:1 .. .0. DD _ ' D o O O r.‘ DEED :— 0”.++ D :. o I CC . -. fl [3 a C 0 ¥ G + a l l l l I l l l 1 J l l l l L l l l 1 l l l i J_ 1 l I l l l l l . l J— -4 —3 -2 -1 o —4 -3 -2 —1 “(Ea/E...) Figure E.6: EC purity fits for E7- bins of 30-33, 33-39, 39-54, and 54-60 GeV. 159 _{/ndf72.47 / 41 Ef/ndf3427 / 33 P1 65.62: 0.0000E+OO :P1 60.47: 0.0000E+OO "P2 0.6124: 0.3060E-01 PP2 0.6655: 0.5143E-01 __I 1538 EC data 1:: i I 360 EC data 0% EC 4497 EC MC'ys a I 11859 EC MC 75 D :0 934 EC MC' . ”a 1342 EC MC' '3 1— D I D e _ f + b : s»; D I " t L- U I :1 U E f _ 69:13 1‘ _ D _ . CED D b s D D U ‘ DD I D D a E 5' l 1 D] 1 1m 1 l l I l l L l I h l I ID 1 l J l I l l l l l l l —4 J" -2 —1 0 — - —2 —1 WM WW ji/natseso / 29 _P1 101.6: 0.0000E+OO P2 0.9574: 0.6767E-01 » I 137 EC data I 7321 EC MC 75 ..'. __u 541 EC MC 'et . De " D ; :3 T :1 .— ' D D _ 6.1- D"- - ' g I II D :- d3 +- l l l E] l 1 l 1 J l 1_ l L i —4 — -2 -1 0 ”lain/Eu.) Figure E.7: EC purity fits for ET bins of 60-75, 75-90, 90-150 GeV. Appendix F Monte Carlo Composites Following are comparisons of data (points) to a composite of Monte Carlo photons and jets (histogram) combined using the purity calculated for each bin. Bins are the same as for the purity calculation in each region. 160 161 0 Data 10-12 Cefi o 061012-150ev o 061615-21Cev - MC MC 1— r— p.- p— p- p— h— p— — u— _ IIIIIIIIIITT lllII—rllll 0.95 1 0.95 I Data 21 -25 Gev I Date 25-30 Gev 1— - MC 0 Data 310-33 GOV Tlilllllll lrfllllllI—Tlllllll 0.95 1 EMFrIctlon EMFracflon I Data 3.3-39 GOV I Data 54-60 GeV ITIIIIIIIIII aflIIIIIIIIIIIII P 0.95 Figure F. 1: Comparison of data (points) and composite Monte Carlo (histogram) EM fraction distributions in the central region. 162 2 0 Data 60—70 C v : o 0616 70-66 GeV :‘ 0 Data 66-10 v T— — _ : E T :— t‘ _ E — I. C I _ L— __ t— l A .95 0.95 1 0.95 1 EMFnction EMFncflon I Data 100—150 GeV 3 TIIIITIIIIHIIIHIITITI O Figure F.1 (cont’d) _ r o Data10-126ev p— - MC 1 l l 1 Ii 1 l l I l L 0 50 1m ‘50 Ema-bx: I Date 21-25 GOV - MC I Data 33-39 GOV - MC 163 I Data 12—15 Gev — MC I Data 25-30 GOV - MC O Data 39-54 GOV — MC I Data 15-21 Gev — MC I Data 30-33 GOV I Data 54-60 GOV — MC Figure F2: Comparison of data (points) and composite Monte Carlo (histogram) Hmatrix x2 distributions in the central region. 164 I Data 60-70 GeV - MC I Data 70—88 GeV - MC i1 50 100 150 o 50 100 150 O Data 100-150 Gev - MC Figure F2 (cont’d) llllll7 I Data 88-100 GeV - MC h"I Data 10-12 V llrllllllllllrl' IllllIllll 0 I 2 Isolation E, (GeV) O Data 21-2 GOV IIIIIIIIIIIIIIIIIIII llllIlLll IsolationEl. (GeV) " O Data ”-39 GOV IIIIIIIIIIIIIII lIlIllIllll Isolation E, (GeV) 165 I Data 12— 15 GeV 1" + 1' J" '1 lLJlIllll 0 1 Isolation I1, (GeV) lllllllllTITITlll — I I Z O TIIIIIIIIIIII I Data 559-54 GOV IllllIllll 0 1 Isolation E1. (GeV) IDatatS 1O -MCn 1 1| J ”"1 "l1'.1 J 1 .1 Isolation E, (GeV) ITIIIIIIIITWTTTI £- a ’fi- I Data 30-33 GOV I TIIIIITTFIIIIII ILlllIllll 0 ‘l 2 Isolation E, (GeV) I Data 54-60 GOV filllllllllllllllllll JlJIlIllll 0 1 2 Isolation E, (GeV) Figure F3: Compan'son of data (points) and composite Monte Carlo (histogram) isolation ET distributions in the central region. 166 I Data 60- 7O GOV I Data 70-88 GOV IIIIIIIIIITII lLllIllll 0 IsolaflonEl. (GeV) lsolnuonlir (Gev) I Data tOO-‘ISOGOV - MC rilllll '-—-II Figure F3 (cont’d) TFIIIIIIII I Data BB-IOOGOV - MC O Data 25—30 Gev FIIIIIIITTI 0.95 " I Data 39-54 GeV 1— - MC 167 0.95 0. 0.95 *' I Data 13-21 GeV I Data 21—25 Gev - 1— ._ - MC 1— ._ 1— P E E 0.95 — I Data 30—33 GeV I Data 3.3—39 GeV : — MC 1._. 1— y— : a l 135 1 EM Fraction I Data 54—60 GeV 1— I Data 60-75 GeV IIIIIIIIIITTTIIT llllllllllllll .0 a u Figure F4: Comparison of data (points) and composite Monte Carlo (histogram) EM fraction distributions in the forward region. 168 1— 0 Data 75—90 Gev I I Z O W—IIIIIITIII - 4 Data 90-1511Gev H" +11. '1 1. sfl‘lllllllllllllllll l 1 0.95 I .0 1 EM Fraction EM Fracflan Figure F.4 (cont’d) I Data 25-30 GOV I Data 39-54 GOV - MC 169 llllllllrllll 1 l I Data 13-21 GOV - MC lllIl 100 150 Hmatrixx’ I Data ISO-3.3 GOV - MC llllll] I Data 54—60 GOV - MC I Data 21 -25 GOV - MC 50 100 150 Hmatrix);1 I Data 3.3-59 GOV I Data 60-75 GOV — MC Figure F5: Comparison of data (points) and composite Monte Carlo (histogram) Hmatrix x2 distributions in the forward region. 170 I Data 75—90 GOV - MC [111 ll—Illljll] '"llrl rill 1111 1111 — _ — I- 1 Figure F5 (cont’d) I Data 90-150 Gev .. J‘C 171 I Data 13-21 Gev fillllllllllllrl t a 4 Isolation E, (GOV) : I Data 25-30 GOV :- - MC E :— J' 1 .I— l l l l l l l 0 1 2 IaolationE, (GOV) Aerllltllllllllllll I Data 359-54 GOV -MC 1 I . I C N 111 I 1 ll 0 1 2 IsolationE,(GeV) I Data 30-33 GOV jtlllllllllll llllIlLll 1 Isolation E, (GOV) I Data 54-60 GOV 1 llllllltllll \ llllIlllL 0 1 Isolation E, (GOV) I Data 21 -25 GOV ,Illlllllr L'llllllJl 0 1 2 Isolation E, (GOV) llllJllll 1 bolation E, (GOV) I Data 60-75 GOV - MC 11 J \ 1+1 '1', t‘ I lllllllllllllllllll ollllIllLl 1 IsolationE,(GeV) Figure E6: Comparison of data (points) and composite Monte Carlo (histogram) isolation ET distributions in the forward region. 172 I Data 75-90 GOV —- I Data 90-150 GOV _ MC ' - MC H TIIIIIIIIIIIIIII $— IITIITIIIITIII Figure F.6 (cont’d) REFERENCES REFERENCES [1] T. Ferbel and W. Molzon, Rev. Mod. Phys. 56, 181 ( 1984) [2] S. Fahey, Ph.D. Thesis, Michigan State University (unpublished) (1995) [3] S. Abachi, et al., Phys. Rev. Lett. 77, 5011 (1996) [4] Review of Particle Properties, Physical Review D54, Part 1 (July, 1996) [5] S. Glashow, Nucl. Phys. 22, 579 (1961);S. Weinberg, Phys. Rev. Lett. 19, 1264 (1967); A. Salarn, in Elementary Particle Theory, W. Svartholm, ed., Almquist and Wiksell, Stockholm (1968); J .D. Bjorken, E.A. Paschos, Phys.Rev.185 1975 (1969) [6] I. Hinchcliffe and J. Womersley, FERMILAB-Conf-96/432 [7] ED. Bloom, et al., Phys. Rev. Lett. 23, 930 (1969); M. Breidenbach et al., Phys. Rev. Lett. 23, 935 (1969) [8] F. J. Hasert, et al., Phys. Lett. 46B, 138 (1973) [9] J .J . Aubert, et al., Phys. Rev. Lett. 33, 1404 (1974); J .E. Augustin, et al., Phys. Rev. Lett. 33, 1406 (1974); G. Goldhaber, et al., Phys. Rev. Lett. 37, 255 (1976); S.W. Herb, et al., Phys. Rev. Lett. 29, 252 (1977); D. Andrews, et al., Phys. Rev. Lett. 45, 219 (1980) [10] R. Brandelik, et al., Phys. Lett. 86B, 243 (1979); DP. Barber, et al., Phys. Rev. Lett. 43, 830 (1979); C. Berger, et al., Phys. Lett. 86B, 418 (1979); W. Bartel, et al., Phys. Lett. 91B, 142 (1980) [11] G. Amison, et al., Phys. Lett. 1223, 103 (1983); M. Banner, et al., Phys. Lett. 122B 476 (1983) 173 [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] 174 S. Abachi, et al., Phys. Rev. Lett. 74, 2632 (1995); F. Abe, et al., Phys. Rev. D52, 2605 ( 1995) W. Vogelsang and A. Vogt, Nucl. Phys. B453, 334 (1995) M. Gliick, et al, Phys. Rev. Lett. 73, 388 (1994) S. Abachi et al, Nucl. Inst. and Meth. A338, 185 (1994) S. Linn, “The Inclusive Cross Section For Isolated Photons in Run 1A,” DO Internal Note 2745 (unpublished) (1995) M. Abolins, et al., Nucl. Inst. and Meth. A280 36 (1989) A. Spadafora, “Analysis of the Electron Energy Scans from Test Beam Load 1,” DO Internal Note 1378 (unpublished) (1992) I. M. Adam, Ph.D. Thesis, Columbia University (unpublished) (1997) M. N arain, et. al, in Proceedings of the American Physical Society Division of Particles and Fields Conference, F ermilab, 1992, R. Raja and J. Yoh, ed., World Scientific, Singapore (1993) C. Shaffer, “The Photon Plus One and Two Jets Cross Section,” DO Internal Note 3104 (unpublished) (1997) T. Sjastrand, Comp. Phys. Comm. 82, 74 (1994) A. Jonckheere, W. Merritt, “DOGEANT User's Guide,” DO Internal Note 969a ( 1995) [24] F. Carminati et. al., GEANT Users Guide, CERN Program Library (1991) [25] [26] [27] [28] [29] H.Baer, J. Ohnemus, and IF. Owens, Phys Rev. D42, 61 (1990) T. Ferbel and W.R. Molzon, Rev.Mod.Phys. 56, 181 (1984) S. Linn, “The Photon Purity in the Central Rapidity Region,” DO Internal Note 2326 (unpublished) (1994) A. Spadafora, private communication A. Peryshkin & R. Raja, “On Sampling Fractions and Electron Shower Shapes,” DO Internal Note 1215 (unpublished) (1991) 175 [30] A. Kotwal, private communication. [31] H.L. Lai, et al., Phys. Rev. D55, 1280 (1997) [32] CDF Collaboration, F. Abe et al., Phys. Rev. D48, 2998 (1993); CDF Collaboration, F. Abe et al., Phys. Rev. Lett. 73, 2662 (1994) [33] J. Huston et al., Phys. Rev. D51, 6139 (1995) [34] H.Baer and M. H. Reno, Phys. Rev. D54, 2017 (1996) [35] E. Berger et al., hep-ph/9610497 [36] S. Linn, private communication. [37] Nang, F., Ph. D. Thesis, Brown University (unpublished) (1996) [38] McKinley, J. and Linnemann, 1., “Description of the Level 2 Electromagnetic Filter Algorithm and Study of the Effects of Shower Generation Method, Zero Suppression, and Event Generation Method on the Level 2 Electromagnetic Filter for D0,” DO Internal Note 1354 (unpublished) (1992) [39] Bantly, J. et al., “DO Luminosity Monitor Constant for the 1994-1996 Tevatron Run,” Fermilab TM-l995 (1997) [40] Chopra, 8., Heintz U., and Narain, M., “Comparison of Electron ID Efficiencies in Run 1A and Run 1B,” DO Internal Note 2351 (unpublished) (1994) [41] Melanson, H., “Electron Efficiency Studies,” DO Internal Note (in preparation) [42] Womersley, 1., “Random Charged Track Overlap Probability in the Central and Forward Direct Photon Samples,” DO Internal Note 2106 (unpublished) (1994) [43] S. Linn, private communication