Study of Double Parton Scattering in Photon + 3 Jets Final State
Study of Double Parton Scattering in Photon + 3 Jets Final State
You-Hao Chang
Study of Double
Parton Scattering
in Photon + 3 Jets
Final State
In Proton–Proton Collisions at √s = 7 TeV
with the CMS Experiment at the LHC
Springer Theses
The series “Springer Theses” brings together a selection of the very best Ph.D.
theses from around the world and across the physical sciences. Nominated and
endorsed by two recognized specialists, each published volume has been selected
for its scientific excellence and the high impact of its contents for the pertinent field
of research. For greater accessibility to non-specialists, the published versions
include an extended introduction, as well as a foreword by the student’s supervisor
explaining the special relevance of the work for the field. As a whole, the series will
provide a valuable resource both for newcomers to the research fields described,
and for other scientists seeking detailed background information on special
questions. Finally, it provides an accredited documentation of the valuable
contributions made by today’s younger generation of scientists.
123
Author Supervisor
Dr. You-Hao Chang Prof. Paolo Bartalini
College of Science CERN Meyrin
National Taiwan University Geneva
Taipei Switzerland
Taiwan
This work does focus on the study of photon þ 3 jet final states in proton–proton
pffiffi
collisions at s ¼ 7 TeV, looking for patterns of two (or more) distinct hard
scatterings in the same collision, i.e. the so-called Double Parton Scattering (DPS).
The analysis is performed in the light of the state-of-the-art Monte Carlo generators
providing higher order corrections to the description of the Single Parton Scattering
(SPS) background.
The conclusions on the DPS observation on this channel are prudent as the
current theoretical uncertainties related to the SPS background are found to be
large. At the same time, a rich set of DPS-sensitive measurements is reported for
possible further interpretation. The analysis constitutes an important milestone in
the progress of Multiple Parton Interactions (MPI) studies at the Large Hadron
Collider. It had a big impact on all the DPS papers published by the CMS col-
laboration for reasons of its open-minded approach inviting to deeply review the
methodologies adopted in the early DPS measurements reported at Tevatron, in
particular those focusing on the photon þ 3 jet benchmark channel.
vii
Abstract
The probability of having more than one interaction per collision is nonnegligible at
the LHC. These additional interactions might reach a hard scale comparable to the
primary scattering and become experimentally distinguishable at high energies.
Distributions sensitive to double parton scattering are investigated in the photon + 3
jets final state in proton–proton collisions at a center-of-mass energy of 7 TeV. The
data were collected by the CMS experiment at the LHC with an integrated lumi-
nosity of 36 pb−1 in 2010. The cross section r for a final state with a photon and a
jet of transverse momentum pT [ 75 GeV together with 2 jets of pT [ 20 GeV,
where the photon and jets are within the fiducial volume of the CMS detector, is
measured to be 124.9 8.9 (stat.) 22.6 (syst.) pb. The differential cross sections
are measured as a function of the difference in azimuthal angles and the transverse
momentum balance between the photon-jet pair and the di-jet pair. Further it is
investigated whether additional contributions from double parton scattering can
improve the agreement between the measured data and the Monte Carlo predictions.
ix
Acknowledgements
—Paolo Bartalini
I would like to show my deepest gratitude to whom I meet with the completion of
this dissertation. I do believe those inspiration accumulated in my life makes who I
am.
First I must thank my supervisors, Prof. Min-Zu Wang and Prof. Paolo Bartalini,
for guiding me into a free learning field of experimental high energy physics. They
are great advisors and mentors who accompanied my course from being guided to
be able to independently complete my research. In addition, I appreciate my
co-supervisors Prof. Shin-Shan Yu and Dr. Yuan Chao. They not only provided me
crucial technical support in academic researches but also wise suggestions and aids
to my life. I also thank the faculties in Lab of NTU High Energy Physics. I have
developed new attitudes concerning cooperation, and staying competitive scien-
tifically or socially because of Prof. Wei-Shu Hou, Prof. Yee Hsiung, Prof. Pao-Ti
Chang, Prof. Kai-Feng Chen, Prof. Stathes Paganis and Dr. Jing-Ge Shiu. I am
especially indebted to Profs. Wei-Shu Hou and Min-Zu Wang because of the
general advices and supports they give, so I learn better to reflect from my mistakes.
Many thanks to Yeon-Jyi Lei, Yeng-Ming Tzeng, Yu-Wei Chang, Kai-Yi Kao,
Kuan-Hsin Chen, Yun-Ju Lu, Jr-Kan Hsieh and Lei Hong for providing me spiritual
and substantial supports on my journey in the field of experimental high energy
physics. With your company beside, I do not feel alone anymore while doing
research abroad. I also appreciate the experiences and advices shared by Mark
Chiang, Jui-Te Wei, Bean Huang, Be-Bo Ho and Po-Yuan 56 which inspired me to
cope with my obstacles in a more mellow way. Aside from scientific works, my
interaction with Jelov Peng, Jia-Jia Chang, Kacaw Kaokoy Cikatopay, Small-Pig
Shiu, Dong-Di 56, Jian-Wei 56, Zu-Ta 56, Yun-Tsung 56, Kai-Ming 56, Po-Hsun
56, Shankar Tien, Yu-Tan Chen, This Lin, Dallas Python Wang, Tzu-An Sheng,
Chia-Hao Tu, Mars Chen, Chou Dong, Jenny Huang, Phoebe Liao and Link Liu
broaden my horizon and enrich my life.
xi
xii Acknowledgements
Last but not least, I am grateful beyond words towards my family and all friends.
With their consideration, I am fearless to walk on this winding long road to do my
Ph.D. With their assistance, I achieve a small goal of my life.
Being a scientist to explore our world has been my dream since I was a child.
This is a dream I would never forget. This is a dream I would ever bear in my mind.
My sincere appreciation to all the people I meet on the road of science.
Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2 Experimental Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.1 Overview of the LHC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 The CMS Detector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.1 Magnet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2.2 Tracking Detectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2.3 Electromagnetic Calorimeter (ECAL) . . . . . . . . . . . . . . . . 14
2.2.4 Hadronic Calorimeter (HCAL) . . . . . . . . . . . . . . . . . . . . . 16
2.2.5 Jet Reconstruction at CMS . . . . . . . . . . . . . . . . . . . . . . . . 18
2.2.6 Muon System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2.7 Global Event Reconstruction . . . . . . . . . . . . . . . . . . . . . . 21
2.3 Trigger System. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3.1 Level-1 Trigger (L1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3.2 High Level Trigger (HLT) . . . . . . . . . . . . . . . . . . . . . . . . 23
2.4 Luminosity Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.4.1 The Primary Luminosity . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.4.2 The Absolute Luminosity . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.5 Computing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3 Data and Monte Carlo Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4 Event Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.1 Trigger Requirement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.2 Pileup Reweighting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.3 Photon Identification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.4 Jet Identification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5 Discriminating Observables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
xiii
xiv Contents
Quantum Chromodynamics (QCD) is the underlying theory to describe hard and soft
high-energy hadron collisions. Much of the complexity involved in describing proton-
proton scatters derives from the composite nature of hadrons. In order to compare
theory predictions with experimental data, the set of final-state partons is transformed
into a set of hadrons in the fragmentation step of initial-state (ISR) and final-state
radiation (FSR), hardest parton interaction (HI), multiple parton interaction (MPI)
and beam remnants. The underlying event (UE) then is everything in a single proton-
proton interaction except for the hard scattering component, or, associated with all
particles produced by the hadron remnants “after” the hard scattering. Figure 1.1
shows a typical proton-proton collision.
The basic idea of the multiple interaction models is assuming that QCD fac-
torization theorem is working not only for the hard scattering process but also for
additional scatters. Hence, the hadronic cross section for the additional interaction
(the second, the third…etc.) and the modeling of multiple interactions when simu-
lating proton-proton collisions make use of several adjustable parameters. The set
of parameters which describes a particular final state is referred to as an underlying
event tune. Most underlying event observation only probe the sum of these effects.
But multiple parton interactions models are particularly successful in accounting for
various features of the underlying event at hadron colliders, to exactly quantify their
contribution to the underlying event is a nontrivial task. Indeed, including multiple
parton-parton scatters in the underlying event model has been shown to significantly
improve the predictions of different Monte Carlo (MC) [1–3]. Observation of an
excess of balanced pairs of jets where pairs are uncorrelated with respect to each
other would support the hypothesis of multiple interactions taking place. If these
interactions reach relatively large values of the exchanged transverse momentum,
x, the observation of multiple parton interaction mostly relies on the double parton
scattering (DPS), which is one of the direct evidence for multiple parton interactions.
It is definitely important of understand double parton scattering since such additional
interactions result in higher rate of multi-jet production than the normal predictions,
© Springer Nature Singapore Pte Ltd. 2017 1
Y.-H. Chang, Study of Double Parton Scattering in Photon + 3 Jets Final State,
Springer Theses, DOI 10.1007/978-981-10-3824-2_1
2 1 Introduction
Fig. 1.1 Proton-proton collision at high energies: hard process (in red) with initial-state radiation
(in pink), final-state radiation (in brown), multiple parton interaction (in green) and beam remnants
(in black)
which don’t take the possibilities of additional interactions into account. Thus, it will
produce relevant and background in searching new physic phenomena.
By taking the QCD factorization into account for double parton scattering, the
corresponding cross section σA,BDPS
can be expressed as [4–9]:
m dx D(x , x , b)D(x , x , b)σ̂ (x , x )σ̂ (x , x ).
σDPS = dbdxA dxB dxA B A A B B 1 A B 2 A B
2 p⊥min
(1.1)
The partonic cross sections σ̂1 (xA xB ) and σ̂2 (xA , xB ) where indices 1, 2 reflect
possibly different subprocesses) are convoluted with two-parton density functions
D(xA , xA , b) and D(xB , xB , b) evaluated at the parton momentum fractions xA , xA ,
xB , xB . The m factor has the value m = 2 when two subprocesses are distinguishable
scattering, and m = 1 when they are indistinguishable [10]. A transverse distance
scale b represents the amount of spatial overlap between the hadrons’ wave functions:
The two-parton distribution functions refer to conditional probabilities D(x, x , b) to
find one parton with momentum fraction x and one parton with momentum fraction
x separated by a distance b in transverse space in the same proton. The cross section
integral is evaluated for scatters above a minimal transverse momentum scale thresh-
old p ⊥ min, i.e. for scatters producing hard enough jets. In the simplest approach,
the two-parton densities D are assumed to factorize, which is equivalent to assuming
uncorrelated two-parton distributions:
Fig. 1.2 Published measurements of effective cross section, σeff as a function of center of mass
energy of collision
with f eff being an effective parton density function and F(b) a matter overlap density.
If this assumption holds, the cross section for double parton scattering is simply
proportional to the product of the respective cross sections for single scatters:
m m σ1 σ2
σDPS = σ1 σ2 F 2 (b)d2 b = , (1.3)
2 2 σeff
where the dependence on the transverse impact parameter b is absorbed in the effec-
tive cross section eff. The current published values of eff as a function of center of
mass energy of collision are shown in Fig. 1.2.
The production of four high transverse momentum ( pT ) jets is thus the most
prominent process to study the impact of multiple interactions directly: Two inde-
pendent scatters in the same proton-proton collision each produce two jets. But
searches for double parton scattering in four-jet events at hadron colliders face sig-
nificant backgrounds from other sources of jet production, in particular from QCD
Bremsstrahlung. Typical thresholds employed in jet triggers bias the event sample
toward hard scatterings. However, a high- pT jet parton is more likely to radiate addi-
tional partons, thus producing further jets. Thus, the relative fraction of jets from
final-state showers above a given threshold is enlarged in jet trigger streams which
is an unwanted bias. On the other hand, looking for four jets in a minimum-bias
stream will yield very small statistics. Hence, studying the double parton scattering
in one photon and three jets final state is a novel approach. Analyses which try to
identify two hard scatters in multi-jet events typically rely on methods to overcome
combinatorics as there are three possible ways to group four objects into two pairs:
Combinations are commonly selected pairwisely balanced in azimuth and energy.
4 1 Introduction
Fig. 1.3 Dominant Feynman diagrams for the direct photon subprocesses: qg → qγ (left) and
qq → gγ (center) and for the photon + 3 jets production from double parton scattering (right)
In this inclusive analysis, we focus on the study of double parton scattering in one
photon and three jets final state.
At the Large Hadron Collider (LHC), the probability of having more than one
parton-parton interaction per proton-proton collision is sizeable [8]. These addi-
tional interactions become experimentally distinguishable in high energy hadronic
collisions [11]. Up to the present, double parton scattering measurements in hadron-
hadron collisions have been suggested in double Drell-Yan, four-jet, same-sign WW,
W plus 2 jets and single photon plus 3 jets production [4, 7, 12–15].
√ Experimentally,
DPS has been studied in the photon + 3 jets final √ state at s = 1.8 TeV [16] and
1.96 TeV [17–19], in the four-jet final state at s = 63 GeV [20], 630 GeV [21],
1.8 TeV [22], and 7 TeV [23], in √ the 2 b-jets + 2 jets final state at 1.96 TeV [19], and
in the W + 2 jets production at s = 7 TeV [24, 25].
At leading order in perturbative quantum chromodynamics (pQCD), several
processes contribute to the production of single prompt photons with large trans-
verse momentum: photons directly produced in the Compton-like gluon scattering
process qg → qγ (Fig. 1.3 (left)) and the quark-antiquark annihilation process qq →
gγ (Fig. 1.3 (center)) [26]. An additional contribution to prompt photon production
arises from the fragmentation of coloured partons and is referred to as fragmentation
photon. Multiple photons from electromagnetic decays of neutral light mesons (such
as π 0 and η) can mimic prompt photon production, when the opening angle between
the multiple photons from the meson decay is too small to effectively discriminate
them through electromagnetic shower properties. DPS signal events in the photon
+ 3 jets final state include pairwise balanced photon-jet and di-jet combinations
where the photon-jet pair and the di-jet pair are produced in the first and second hard
interaction of the same collision, respectively (Fig. 1.3 (right)). The photon source
of DPS signal events can be direct photons and fragmentation photons. DPS signal
events and single parton scattering (SPS) events are characterized by different kine-
matics, where the SPS events produce one photon and one jet in one interaction,
with the evolution of the jet resulting in additional jets at lower momenta (2 → 4
partonic process). Events containing photon + 3 jets produced from SPS or misiden-
1 Introduction 5
The LHC accelerator is situated at the national boundary between Switzerland and
French in the existing Large Electron-Positron (LEP) tunnel with a length of 26.7 Km
as deep as 175 m. The schematics of the LHC with the beam injection sequence
is presented in Fig. 2.1. Protons are generated from the ionization of H2 plasma
enhanced by an electron beam, and injected into the Linear accelerator (Linac2) to
get an energy of 50 MeV. The Proton Synchrotron Booster (PSB), which is the first
and the smallest circular proton accelerator, speeds up the 50 MeV protons to 1.4 GeV.
Finally, the protons with an energy of 26 GeV coming from the Proton Synchrotron
(PS) are further accelerated by the Super Proton Synchrotron (SPS) to 450 GeV, and
then injected into the LHC, where they will be quickened to 7 TeV in bunches with
a nominal number of 1.15 · 1011 particles per bunch. The cylindrical bunches with a
radius of 16.6 µm and a length of 7.55 cm group the protons within a bunch spacing
of 25 ns, corresponding to a bunch crossing frequency of 40 MHz. In the LHC, two
kinds of superconducting magnets operating at a temperature of 1.9 K with liquid
© Springer Nature Singapore Pte Ltd. 2017 7
Y.-H. Chang, Study of Double Parton Scattering in Photon + 3 Jets Final State,
Springer Theses, DOI 10.1007/978-981-10-3824-2_2
8 2 Experimental Apparatus
Fig. 2.1 The schematics of the LHC with the beam injection sequence
helium are used to control the beam. During the acceleration, 1232 dipole magnets
providing a magnetic field of 8.33 T guide the particles along the design orbit, and
392 quadrupole magnets confine the particles in the vicinity of the design orbit, from
which most particles will unavoidably deviate. The Radio-Frequency (RF) system,
proving a RF voltage of 8–16 MV/beam, accelerates the particles and compensates
an energy loss of 7 KeV per turn due to the synchrotron radiation.
The six experiments operating at the LHC are shown in Fig. 2.2. Two experiments,
including the CMS [30] and ATLAS (A Toroidal LHC Apparatus) [31] detectors, are
optimized towards high- pT physics for general purposes, involving the verification of
the SM, the Higgs boson hunting, new physics beyond the SM, and heavy ion physics.
The motivation of the ALICE (A Large Ion Collider Experiment) [32] detector is to
study heavy ion collisions with an excellent ability of the low- pT measurement. The
purpose of the LHCb (the Large Hadron Collider beauty experiment) [33] detector
2.1 Overview of the LHC 9
is to study the B physics in order to solve the CP-violation (CP standing for charge
parity). For another two special-purpose experiments, the TOTEM (TOTal Elastic and
diffractive cross section Measurement) experiment [34] in front of the CMS detector
is dedicated to measure the total cross section, elastic scattering, and diffractive
processes, and the LHCf (Large Hadron Collider forward) experiment [35] located
close to the ATLAS detector studies cosmic rays in laboratory conditions by using
the forward particles generated inside the LHC.
A LHC√ commissioning run with proton–proton collisions at center-of-mass ener-
gies of s = 7 TeV started on 30th March 2010. During the 2010 run, the recorded
efficiency of the CMS is about 92, and 84% of recorded data, (36.1 ± 1.4) pb−1 ,1
are certificated for physics analyses. This dissertation uses the 2010 CMS data in
proton–proton collisions at 7 TeV.
1 ‘p’indicates a number of an order of magnitude, −12, and a barn (symbol ‘b’) stands for a cross
section unit, 10−24 cm2 .
10 2 Experimental Apparatus
The CMS detector is situated on the interaction point 5 (P5) of the LHC at a depth of
100 m. In the LHC environment, the CMS detector should experimentally challenge
the physics benchmark channels with a small cross section against the overwhelm-
ing
√ background from QCD jet production, whose proton–proton cross section at
s = 14 TeV is roughly 100 mb with an event rate of approximately 109 inelas-
tic event/s. For reaching the physics goals, rejecting background with the optimal
efficiency in the interesting rare channels has to be performed with a high power.
Additionally, the design of the CMS is optimized to well identify muons, charged
particles, electrons, and jets with an excellent momentum and corresponding two-
particles mass resolution. Good E Tmiss 2 resolution has also been achieved according
to the related objects with an excellent momentum resolution. Moreover, secondary
vertices and impact parameters have to be precisely measured for an efficient iden-
tification of heavy flavor and τ -lepton decays.
An excess of inelastic (hard-core scattering) collisions happen in the interesting
events. That is, additional proton–proton interactions within the same beam crossing
(pileup). A reduction of the pileup effect can be performed using high-granularity
detectors with a good time resolution, resulting in low occupancy. In addition, the
sub-detectors and front-end electronics have to be required radiation-hard because
of a high radiation arising from the large flux of particles near the interaction region.
The CMS coordinate system adopts a right-handed style such that the origin
is centered at the nominal collision point, the x-axis directs south to the center
of the LHC ring, the y-axis points vertically upward, and the z-axis is along the
beam direction to the west. The azimuthal angle φ is measured from the x-axis in
the x y plane, and the polar angle θ is measured from the z-axis. Pseudorapidity,
defined as η = −ln tan(θ/2), is introduced to replace the polar angle θ because of
a property of Lorentz invariant for boosting along the z-axis. Besides, η is a good
approximation of the rapidity ( 21 ln E+ pz c
E− pz c
) in a massless-particles case. The transverse
momentum (denoted pT ) relative to the beam direction is computed from the x- and
y-components, while the transverse energy E T is obtained from Esinθ .
The overview of the CMS apparatus is presented in Fig. 2.3. Briefly speaking, the
CMS detector is composed of a silicon tracking system, an electromagnetic calorime-
ter (ECAL), a hadronic calorimeter (HCAL), and a muon system. A superconducting
solenoid magnet at the heart of the CMS detector provides a magnetic field of 3.8 T.
The CMS detector is 21.6 m long with a diameter of 14.6 m, and has overall weight
of 12500 tons. The detailed descriptions are presented in the following sub-sections.
2 The imbalance of energy is measured in the transverse plane, and also denoted by MET.
2.2 The CMS Detector 11
2.2.1 Magnet
3 The momentum resolution is proportional to p/(B R 2 ), where B is the magnetic field in Tesla, and
R indicates the radius of the charged-particle curvature in meter.
12 2 Experimental Apparatus
Fig. 2.4 The CMS tracker layout. The outer radius is around 110 cm, and the total length is approx-
imately 540 cm
In terms of various radii, the tracker shown in Fig. 2.4 within a range of |η| < 2.5
can be divided into three regions coping with the different charged-particle flux. In
the innermost region (r ≈ 10 cm), since the particle flux closest to the interaction
vertex is the highest as 107 /s, pixel detectors with the fine-granularity size of ≈
100 × 150 µm2 /pixel are installed. In the intermediate region (20 < r < 55 cm),
the particle flux becomes lower to allow use of silicon microstrip detectors with a
minimum cell size of 10 cm × 80 µm. In the outermost region (r > 55 cm), the
particle flux has sharply decreased, and then use of larger-pitch silicon microstrips
with a maximum cell size of 25 cm × 180 µm is enough. With the design of the
inner tracker system, a measurement of the charged-particle trajectory and vertices
reconstructions can be performed. The Fig. 2.5 displays the material budget of the
CMS tracker in units of radiation length (X 0 ) [36].4 Taking muons with pT = 1 GeV/c
as an example, the pT resolution (δpT / pT ) shown in Fig. 2.6 has a good performance
of ∼1% [36]. The (δd0 ) is around 100 µm, and the longitudinal part (δz0 ) is between
100 and 1000 µm.
Pixel Tracker
The pixel detector comprise three barrel layers and two endcap disks on each side
shown in Fig. 2.7. The three barrel layers are located at mean radii of 4.4, 7.3, and
10.2 cm with a length of 53 cm. The two endcap disks, covering from 6 to 15 cm in
radius, are placed on each side at |z| = 34.5 and 46.5 cm in a turbine-like geometry
with blades rotated by 20◦ .5
Fig. 2.5 Material budget of the CMS tracker in units of radiation length X 0 as a function of |η| for
different detector types
Fig. 2.6 Resolution of track pT (left), d0 (center), and z0 (right) with muon pT of 1, 10, 100 GeV/c
Strip Tracker
The strip tracker in the barrel region is composed of two parts including a TIB
(Tracker Inner Barrel) and a TOB (Tracker Outer Barrel). The TIB with a half-length
of |z| < 65 cm is made of four layers, whose first two layers consist of ‘stereo’
modules in order to provide a measurement in both r-φ and r-z coordinates. The
detailed parameters for different detector types are presented in Table 2.2. The TOB
with a half-length of |z| < 110 cm is made of six layers, whose first two layers also
provide a ‘stereo’ measurement.
14 2 Experimental Apparatus
In the endcap region, the strip tracker can be separated into the TEC (Tracker
End Cap) and TID (Tracker Inner Disks). Each TEC, extending to the region 120 <
|z| < 280 cm, comprises nine disks whose first two and fifth rings are made of ‘stereo’
modules. Each TID, filling the gap between the TIB and the TEC, consists of three
small disks whose first two rings have ‘stereo’ modules.
6 The Moliere radius is a characteristic constant of a material, indicating the radius of a cylinder
containing on average 90% of the shower’s energy deposition.
2.2 The CMS Detector 15
within 25 ns) and high radiation resistance (up to 10 Mrad).7 This sub-detector is
hermetic and homogeneous with a large coverage in pseudorapidity (|η| < 3), and
divided into barrel and endcap parts as illustrated in Fig. 2.8. The ECAL barrel (EB)
in the region of |η| < 1.479 consists of 61200 crystals with a crystal front face of
22 × 22 mm2 , and the ECAL endcap (EE) in the rest region comprises 7324 crystals,
whose front face is 28.6 × 28.6 mm2 . Thus, a design of a fine granularity and a com-
pact calorimeter is ensured. A barrel crystal is 23 cm long corresponding to 25.8 X 0 ,
while the length of a endcap crystal is 22 cm equal to 24.7 X 0 . The scintillation light
emitting from the crystals is received by avalanche photodiodes (APD)8 with an
active area of 5 × 5 mm2 in the barrel, and by vacuum phototriodes (VPT)9 with an
active area of 280 mm2 in the EE. The readout has a strong temperature dependence,
and this effect induces a challenge to the ECAL energy calibration.
A special-purpose sub-detector, prEShower (ES) system, placed in front of the
EE crystal is designed to have a good spatial resolution to measure the position of
incoming particles, and aid particle identification in the EE. Furthermore, the ES
can distinguish a single photon from two closely-separated photons so that rejecting
neutral pions (π 0 ) can be performed. In each endcap, there are two orthogonal layers
of 1.9 mm-pitch silicon sensors, each preceded by thin layers of lead (3 X 0 ) that
initiate electromagnetic showers.
The ECAL energy resolution can be parametrized by the following expression:
σ 2 2 2
S N
= √ + + C2 , (2.1)
E E E
than the APD since the radiation levels are highest in the EE.
16 2 Experimental Apparatus
Fig. 2.9 ECAL energy resolution as a function of the energy, which is measured in an electron test
beam. The central value of the stochastic (S), noise (N), and constant (C) terms are shown in the
legend
where S is the stochastic term, N stands for the noise term, and C represents the
constant term. In an electron test beam, the parameters of ECAL energy resolution
1
were measured to be S = 0.028 GeV 2 , N = 0.12 GeV, and C = 0.003 shown in
Fig. 2.9 [36].
The energy measurement in the calorimeter system is completed using the ECAL and
the hadronic calorimeter (HCAL), which is designed to measure energies coming
from hadrons, jets, or neutrinos. The HCAL is a sampling calorimeter which com-
prises alternating layers of massive absorbing brass plates providing a short inter-
action length (λl ),10 and plastic scintillator tiles, which embed wavelength-shifting
fibers.11 The photodetection readout is based on multi-channel hybrid photodiodes
(HPDs). The first scintillators in front of the first absorber plate is used to sample
10 It indicates the mean free path, meaning the average distance between collisions.
11 It shifts the blueviolet light emitted by the scintillator to green light which is sent to HPDs.
2.2 The CMS Detector 17
Table 2.3 Tower segmentation in φ and η for the hadronic barrel, endcap, and forward calorimeter
HB/HO HE |η| ≤ 2.5 HE |η| > 2.5 HF |η| ≤ 4.7 HF |η| > 4.7
δφ × δη 0.087 × 0.087 0.087 × 0.087 0.175 × 0.175 0.175 × 0.175 0.175 × 0.35
showers producing in the material, and the last scintillators installed in back of the last
absorber plate ensures if the late producing showers leak out. The HCAL is divided
into three components, including the barrel and outer hadronic calorimeter (HB and
HO) in the barrel region of |η| < 1.3, the endcap part (HE) at 1.3 < |η| < 3, and the
forward sub-detector (HF) extending up to |η| < 5.2 as presented in Fig. 2.10.
In the barrel region, hadron showers can not be completely absorbed by the EB plus
HB. Therefore, the HB is supplemented by the HO placed between the solenoid and
the muon detectors, while the solenoid is regarded as additional absorbing material
providing sufficient containment with a thickness of 11.8 λl . The granularity of the
calorimeters in δη × δφ is summarized in Table 2.3.
In the forward region, the HF placed at 11.2 m starting from the interaction point is
designed to measure the energetic forward jets12 using steel absorber plates, which
is used to detect the Cherenkov light.13 Thus, signals coming from electrons and
photons can be separated from signals generated by hadrons by considering a longi-
tudinal segmentation.
12
√For example, there is an energy of 760 GeV deposited on average in proton–proton collisions at
(s) = 14 TeV.
13 The Cherenkov light is emitted by a particle when its speed is fast more than the speed of light in
the material.
18 2 Experimental Apparatus
Jets are reconstructed offline from the energy deposits in the calorimeter towers,
clustered by the anti-kt algorithm [27, 28] with a size parameter of 0.5. In this
process, the contribution from each calorimeter tower is assigned a momentum,
the absolute value and the direction of which are given by the energy measured in
the tower, and the coordinates of the tower. The raw jet energy is obtained from
the sum of the tower energies, and the raw jet momentum by the vectorial sum of
the tower momenta, which results in a nonzero jet mass. The raw jet energies are
then corrected to establish a relative uniform response of the calorimeter in η and a
calibrated absolute response in transverse momentum pT .
The muon system, situated at the outermost part of the CMS detector, consists of
three kinds of gaseous detector chambers in order to identify muon and measure
momentum. The coverage of the muon system reaches up to |η| = 2.4. In the barrel
region, drift tube (DT) chambers are adopted since the muon rate and the neutron
induced background are both small under a lower magnetic field. On the other hand,
the muon and the background flux is much higher in the endcap region, and then
cathode strip chambers (CSCs) are used because of a faster response, a higher gran-
ularity, and a better resistance against radiation. Besides, resistive plate chambers
(RPCs) construct a redundant trigger system. In total, the muon system possesses
250 DT chambers, 540 CSCs, and 610 RPCs, and the overview is shown in Fig. 2.11.
For the performance of the muon system, the momentum resolution (Δp/ p) only
based on the muon system shown in Fig. 2.12 can be improved by considering the
inner tracker because the effects from multiple scattering and energy loss are not
included [36].
DT Chamber
Fig. 2.12 The muon momentum resolution (Δp/ p) as function of the muon momentum for the
muon system only, the inner tracker only, or both (‘full system’) with a region of |η| < 0.2 (left)
and 1.8 < |η| < 2.0 (right)
CSC
The CSCs are multiwire proportional chambers with a trapezoidal shape. This type
provides a two-dimensional position measurement, where the r and φ coordinates
are determined by the copper strips and the anode wires, respectively.
2.2 The CMS Detector 21
RPC
The RPCs are composed of two high resistive plastic plates with a bias, and separated
by a gas volume. This type providing a fast response with a time resolution of 1 ns is
used in the muon trigger system. The RPCs are sandwiched between DT chambers
or CSCs.
The global event reconstruction (also called particle-flow event reconstruction [39,
40]) consists in reconstructing and identifying each single particle with an optimized
combination of all subdetector information. In this process, the identification of
the particle type (photon, electron, muon, charged hadron, neutral hadron) plays an
important rôle in the determination of the particle direction and energy. Photons (e.g.
coming from π 0 decays or from electron bremsstrahlung) are identified as ECAL
energy clusters not linked to the extrapolation of any charged particle trajectory to
the ECAL. Electrons (e.g. coming from photon conversions in the tracker material or
from b-hadron semileptonic decays) are identified as a primary charged particle track
and potentially many ECAL energy clusters corresponding to this track extrapolation
to the ECAL and to possible bremsstrahlung photons emitted along the way through
the tracker material. Muons (e.g. from b-hadron semileptonic decays) are identified
as a track in the central tracker consistent with either a track or several hits in the
muon system, associated with an energy deficit in the calorimeters. Charged hadrons
are identified as charged particle tracks neither identified as electrons, nor as muons.
Finally, neutral hadrons are identified as HCAL energy clusters not linked to any
charged hadron trajectory, or as ECAL and HCAL energy excesses with respect to
the expected charged hadron energy deposit.
The energy of photons is directly obtained from the ECAL measurement, cor-
rected for zero-suppression effects. The energy of electrons is determined from a
combination of the track momentum at the main interaction vertex, the correspond-
ing ECAL cluster energy, and the energy sum of all bremsstrahlung photons attached
to the track. The energy of muons is obtained from the corresponding track momen-
tum. The energy of charged hadrons is determined from a combination of the track
momentum and the corresponding ECAL and HCAL energy, corrected for zero-
suppression effects and for the response function of the calorimeters to hadronic
showers. Finally, the energy of neutral hadrons is obtained from the corresponding
corrected ECAL and HCAL energy.
In the barrel section of the ECAL, energy resolutions as good as 1% are achieved
for unconverted or late-converting photons in the tens of GeV energy range. The
remaining barrel photons have a resolution of about 1.3% up to a pseudorapidity of
|η| = 1, rising to about 2.5% at |η| = 1.4. In the endcaps, the resolution of uncon-
verted or late-converting photons is about 2.5%, while the remaining endcap photons
have a resolution between 3 and 4% [41].
22 2 Experimental Apparatus
The CMS trigger system is designed to deal with the unprecedented high interaction
rate ∼109 Hz, caused by the bunch crossing rate of 40 MHz at nominal luminosity.
The data recording rate should be reduced to the order of a few 100 Hz, which allows
for permanent storage of an event. The trigger system is divided into two steps,
including the Level-1 trigger (L1) providing a fast decision with custom electronics,
and the High Level Trigger (HLT) system relying on commercial processors based
on software algorithms. The flow chart is displayed in Fig. 2.13 [42].
The L1 trigger latency, allocated for the transit and for archiving a decision from a
bunch crossing, is 3.2 µs. It turns out that the L1 trigger calculation must be done
in less than 1 µs. The L1 trigger is decided using information from the calorimeters
plus the muon system, and the presence of trigger primitive objects such as photons,
missing
electrons, muons, jets, E T , and E T . Once an event is triggered by L1, the
full detector information (∼1 MB) is transferred to the processor farms (also called
‘event filter farms’) via the data acquisition (DAQ) system at a rate of up to 100 kHz.
Events, passing through the L1 trigger, are proceeding with the HLT. A data recording
rate is limited up to few 100 Hz, while only data accepted by the HLT is recorded for
offline physics analysis. Traditionally, this is performed in two stages, involving a
hardware/software-based Level-2 trigger quickly providing a large rejection factor,
and a Level-3 processor farm based on more sophisticated algorithms.
Two methods for extracting a relative instantaneous luminosity with the online HF
system have been studied. The first method based on zero counting infers the mean
number of interactions per bunch crossing according to the average fraction of empty
towers. In the second method, the linear relationship between the average transverse
energy per tower and the luminosity is exploited. However, one unexpected result
first observed in 2011 running is that the HF scale is non-linear for single-bunch
luminosities due to the pile-up effect. Thus, the zero-counting method is mainly
applied for the primary luminosity measurement.
The separation scan method for determining the absolute luminosity was first pio-
neered by S. Van Der Meer at the Intersecting Storage Rings (ISR) [44]. The beam
size can be measured using the relative interaction rate as a function of the transverse
beam separations. Once the beam profile in x and y is assumed to be given by the
function F(x, y) = f x (x) f y (y), the instantaneous luminosity can be re-written as,
N1 N2 νor bit n b F(0, 0)
L0 = , (2.3)
f x (Δx) dΔx f y (Δy) dΔy
where Δx and Δy are the beam separations in the x (horizontal) and y (verti-
cal) planes, respectively. Besides, double-Gaussian distributions are adopted for the
functions f x and f y . It results in the instantaneous luminosity with beam profiles
expressed as,
N1 N2 νor bit n b
L0 ≡ ef f ef f
, (2.4)
2π σx σy
ef f
where σ j means the effective beam size for each scan plane j.
2.5 Computing
All data has to be stored for the lifetime of the experiment. In particular, data analy-
sis of all running periods has to be supported at all times. Consequently, transfer,
storage and analysis of large data sets is foreseen by the CMS Computing Model.
Real-time (raw) detector data passes pattern recognition, filtering and data reduction
steps. Physics analysis activities are carried out regardless of the physical loca-
tion of the physicist at several computing centers distributed throughout the world.
2.5 Computing 25
RAW: This type of event content consists of the entire detector information, includ-
ing trigger decisions and meta data, for instance on the run conditions. According to
the trigger signature, RAW data is classified into distinct primary datasets.
RECO: Reconstructed events are obtained after RAW data has been passed through
pattern recognition and compression algorithm steps. Thus, in this event content,
high-level physics objects are contained, including all basic detector input from
which they were reconstructed.
AOD: The Analysis Object Data event content consists of high-level physics objects
together with information needed to refit the kinematics. AOD data is obtained from
RECO data by means of filtering.
In addition, event reconstruction relies on information about non-event data, e.g.
construction data, equipment management data, configuration data, and conditions
data. Conditions data includes information on alignment and calibration constants
as well as information on the detector status.
The CMS computing needs exceed all presently available resources of a single
site. Computing resources provided by collaborating institutes worldwide are thus
integrated into a common hierarchical architecture of Tiered centers: One Tier-0
center at CERN, six Tier-1 centers at large computing facilities and approximately
25 Tier-2 centers at medium to large computing facilities. The data ow between tiers
is sketched in Fig. 2.14 and described in the following [36].
26 2 Experimental Apparatus
The CERN Tier-0 center records experimental data at a mass storage system,
promptly reconstructs RAW data into first-pass RECO datasets, and exports a copy of
the RAW data to Tier-1 centers for re-reconstruction. The Tier-1 centers are operated
around the clock and provide large computing farms, dedicated mass storage, and
fast network links to each other and to the Tier-0. RAW data are stored permanently
to offer reconstruction whenever necessary, for instance if improved alignment and
calibration constants are available. In addition, skimming and data-intensive analysis
applications run on the Tier-1 with the output being transferred to associated Tier-2
centers. These centers support the bulk of user analysis and thus have to grant access
to large computing farms to all CMS users with a valid grid certificate. Off-line
calibration and alignment studies as well as detector analyses are typically performed
at the Tier-2 centers as well. Thirdly, Tier-2s produce large amounts of simulated
data which are transferred to the associated Tier-1 center for storage. In order to
provide fast feedback on key analyses, the CERN Analysis Facility (CAF) provides
large CPU resources together with fast access to the full data.
The Worldwide LHC Computing Grid provides an infrastructure for job submis-
sions and remote data access. It does so, while ensuring robust security and account-
ing. Nonetheless, some CMS-specific software is needed on sites. The Dataset Book-
keeping System catalogues existing event data. Existing event data is mapped to its
physical location or locations (if more than one copy of a dataset exists) with the
help of the Data Location Service. Each site employs a Local File Catalogue which
translates a logical file name into the actual path in the respective file system. CMS-
specific tools for large scale data transfers schedule, monitor and verify the data
transfer. Very large-scale data processing is performed with specialized tools, com-
prised of automated job managers. Since remote analysis introduces additional layers
of complexity to a physics analysis, dedicated tools have been developed to facilitate
job submission, monitoring, and retrieval [45].
Chapter 3
Data and Monte Carlo Simulation
This analysis is based on the full 2010 data sample, collected with the CMS
detector and reprocessed with the CMS reconstruction software. The full data sam-
ple processed for this analysis amounts to 36.1 ± 1.4 pb-1 . A detailed list of the
reprocessed data sets can be found in Table 3.1. Only certified good runs are used
in the official list with respect to specific un-prescaled single photon triggers, which
are summarized in Chap. 4. The treatment of pile-up effect in data is also described
in Chap. 4.
Simulated event samples for single photon production associated with 3 jets
are produced with different MC event generators: pythia 8 (version 8.165) [46],
pythia 6 (version 6.426) [47], and MadGraph 5 (version 5.1.1.0) [48–50] in com-
bination with pythia interface. Multiple parton interactions (MPI) are simulated in
both pythia 6 and pythia 8. The modeling of MPI simulation makes use of several
adjustable parameters in the underlying event tune. MPI-related parameters of base
tunes for LHC experiments are studied in Reference [51]. All simulated event sam-
ples are processed and reconstructed in the same manner as done for the collision
data. The full detector response is simulated with geant4 [52, 53]. MC samples
from different generators are used to compare predictions at both stable-particle and
detector level, which can be also used for unfolding process of data.
pythia 6 and pythia 8 are leading order (LO) MC event generators with 2 →
2 matrix element (ME) calculations. The pythia 6 generator with the tune Z2 [54]
uses LO CTEQ6L1 [55] parton distribution functions (PDF) and applies a new MPI
model [56]. The pythia 8 generator with the tune 4C [57] also uses LO CTEQ6L1
PDFs and implements a similar MPI model with respect to pythia 6. It tunes the
free parameters to the underlying event data obtained at the LHC and considers color
reconnection and rescattering between the partons [58]. The pythia 8 generator can
also be used to study the DPS signal distribution by generating two hard scatterings in
the same proton-proton collision. The MadGraph 5 generator with LO CTEQ6L1
PDFs contains LO ME calculations up to 2 → 5 processes. Events generated by
Table 3.1 Summary of analyzed data sets. The full data sets correspond to 36.1 pb-1
Run ranges Dataset
136035–144114 /EG/Run2010A-Apr21ReReco-v1/AOD
146428–149294 /Photon/Run2010B-Apr21ReReco-v1/AOD
Table 3.2 Summary of pythia-8 MC samples used in the analysis at detector level. All MC samples
listed here are produced using pythia-8 generator with the 4 C tune. The cross section values are
all LO predictions
pythia-8, γ +jet 4 C tune
MC set σ , pb # events L dt, pb-1 pˆT range, GeV/c
PhotonJet_Pt30to50_7TeV 1.80×104 2.5×105 13.9 30–50
PhotonJet_Pt50to80_7TeV 2.90×103 1.5×105 51.7 50–80
PhotonJet_Pt80to120_7TeV 470 1.0×105 213 80–120
PhotonJet_Pt120to170_7TeV 89.0 1.0×105 1.12×103 120–170
PhotonJet_Pt170to300_7TeV 22.6 5.0×104 2.21×103 170–300
PhotonJet_Pt300to470_7TeV 1.49 5.0×104 3.35×104 300–470
pythia-8, QCD 4C tune
MC set σ , pb # events L dt, pb-1 pˆT range, GeV/c
QCD_Pt30to50_7TeV 6.10×107 1.1×106 1.81×10−2 30–50
QCD_Pt50to80_7TeV 7.22×106 1.1×106 1.53×10−1 50–80
QCD_Pt80to120_7TeV 8.83×105 1.1×106 1.23 80–120
QCD_Pt120to170_7TeV 1.29×105 8.5×105 6.6 120–170
QCD_Pt170to300_7TeV 2.70×104 8.5×105 31.7 170–300
QCD_Pt300to470_7TeV 1.29×103 5.3×105 412 300–470
means of the MadGraph generator are interfaced with pythia 6, tune Z2 and tune
Z2* [56, 59], for hadronization, parton showering, and MPI simulation.
The data are also compared to the multijet-improved MC event generator, sherpa
2.1.0 [60–64] with next-to-leading order (NLO) CT10 [65] PDFs and tune CT10 [60]
is considered. sherpa contains particle production at tree level with merging between
ME and parton shower (PS) (2 → 2 + N, where N = 1 in this analysis). The pre-
diction of MadGraph, interfaced with pythia 8 tune 4C, is also included for the
comparisons at stable-particle level. We also preform a deep investigation of LO
and NLO cross section of different MCs in Appendix A. The lists of fully simulated
samples are presented in Tables 3.2, 3.3, 3.4, 3.5, 3.6 and 3.7.
For computation reason, the productions of all MC samples are organized by
splitting the different processes in several bins. For pythia and sherpa productions, a
generation in slices according to the transverse momentum exchanged in the partonic
3 Data and Monte Carlo Simulation 29
Table 3.3 Summary of pythia-6 MC samples used in the analysis at detector level. All MC samples
listed here are produced using pythia-6 generator with the Z2 tune. The cross section values are
all LO predictions
pythia-6, γ +jet Z2 tune
MC set σ , pb # events L dt, pb-1 pˆT range, GeV/c
PhotonJet_Pt30to50_7TeV 1.67×104 2.19×106 131 30–50
PhotonJet_Pt50to80_7TeV 2.72×103 2.04×106 748 50–80
PhotonJet_Pt80to120_7TeV 447 2.05×106 4.58×103 80–120
PhotonJet_Pt120to170_7TeV 84.2 2.09×106 2.48×104 120–170
PhotonJet_Pt170to300_7TeV 22.6 2.07×106 9.14×104 170–300
PhotonJet_Pt300to470_7TeV 1.49 2.08×106 1.39×106 300–470
pythia-6, QCD Z2 tune
MC set σ , pb # events L dt, pb-1 pˆT range, GeV/c
QCD_Pt30to50_7TeV 5.31×107 6.58×106 1.24×10−1 30–50
QCD_Pt50to80_7TeV 6.36×106 6.60×106 1.04 50–80
QCD_Pt80to120_7TeV 7.84×105 6.59×106 8.4 80–120
QCD_Pt120to170_7TeV 1.15×105 6.13×106 53.2 120–170
QCD_Pt170to300_7TeV 2.43×104 6.22×106 256 170–300
QCD_Pt300to470_7TeV 1.17×103 6.43×106 5.51×103 300–470
Table 3.4 Summary of MadGraph MC samples used in the analysis at detector level. All MC
samples listed here are generated by MadGraph and passed to pythia-6 Z2 tune hadronizer for
hadronization and parton showering. The cross section values are all LO predictions
MadGraph, γ +jet, with pythia-6 Z2 tune Hadronizer
MC set σ , pb # events L dt, pb-1 HT range, GeV/c
PhotonJet_Ht40to100_7TeV 2.57×104 1.27×107 496 40–100
PhotonJet_Ht100to200_7TeV 5.21×103 1.54×106 295 100–200
PhotonJet_Ht200toinf_7TeV 798.3 9.38×106 1.17×104 200–∞
MadGraph, QCD, with pythia-6 Z2 tune Hadronizer
MC set σ , pb # events L dt, pb-1 HT range, GeV/c
QCD_Ht100to250_7TeV 4.19×106 2.26×107 5.38 100–250
QCD_Ht250to500_7TeV 1.99×105 2.07×107 104 250–500
QCD_Ht500to1000_7TeV 5.86×103 1.44×107 2.47×103 500–1000
QCD_Ht1000toinf_7TeV 122.6 6.29×106 5.13×104 1000–∞
30 3 Data and Monte Carlo Simulation
Table 3.5 Summary of MadGraph MC samples used in the analysis at stable-particle level. All
MC samples listed here were generated by MadGraph and passed to pythia-6 Z2* tune hadronizer
for hadronization, parton showering and multiple parton interaction
MadGraph, γ +jet, with pythia-6 Z2* tune Hadronizer with MPI
MC set σ , pb # events L dt, pb-1 HT range, GeV/c
PhotonJet_Ht40to100_7TeV 2.55×104 1.17×107 460.4 40–100
PhotonJet_Ht100to200_7TeV 5.30×10 3 1.54×106 290.0 100–200
PhotonJet_Ht200toinf_7TeV 779.4 9.79×106 1.26×104 200–∞
MadGraph, QCD, with pythia-6 Z2* tune Hadronizer with MPI
MC set σ , pb # events L dt, pb-1 HT range, GeV/c
QCD_Ht100to250_7TeV 4.17×106 1.54×107 3.68 100–250
QCD_Ht250to500_7TeV 1.99×105 2.14×107 108.0 250–500
QCD_Ht500to1000_7TeV 5.82×103 1.49×107 2.56×103 500–1000
QCD_Ht1000toinf_7TeV 122.5 7.84×106 6.10×104 1000–∞
Table 3.6 Summary of MadGraph MC samples used in the analysis at stable-particle level. All
MC samples listed here were generated by MadGraph and passed to pythia-8 4 C tune hadronizer
for hadronization, parton showering and multiple parton interaction
MadGraph, γ +jet, with pythia-8 4 C tune Hadronizer with MPI
MC set σ , pb # events L dt, pb-1 HT range, GeV/c
PhotonJet_Ht40to100_7TeV 2.33×104 1.07×107 460.3 40–100
PhotonJet_Ht100to200_7TeV 4.67×10 3 1.35×10 6 290.0 100–200
PhotonJet_Ht200toinf_7TeV 624.9 7.85×106 1.26×104 200–∞
MadGraph, QCD, with pythia-8 4 C tune Hadronizer with MPI
MC set σ , pb # events L dt, pb-1 HT range, GeV/c
QCD_Ht100to250_7TeV 3.27×106 1.21×107 3.68 100–250
QCD_Ht250to500_7TeV 1.51×105 1.63×107 108.0 250–500
QCD_Ht500to1000_7TeV 4.60×103 1.18×107 2.56×103 500–1000
QCD_Ht1000toinf_7TeV 98.3 6.00×106 6.10×104 1000–∞
Table 3.7 Summary of sherpa tune CT10 MC samples used in the analysis at stable-particle level
sherpa, γ +jet
MC set σ , pb # events L dt, pb-1 pˆT range, GeV/c
PhotonJet_Pt45_7TeV 5.42×103 5.00×106 922.5 45–∞
sherpa, QCD
MC set σ , pb # events L dt, pb-1 pˆT range, GeV/c
QCD_Pt55_7TeV 6.31×106 1.20×109 190.5 55–∞
3 Data and Monte Carlo Simulation 31
interaction ( pˆT ) is processed. And, for MadGraph production, generation slices are
based on the scalar summation of the transverse momentum of all the jets (HT ).
In this analysis, the contributions of different bins are combined together. Accord-
ing to the corresponding cross section and number of generated events, each bin is
weighted to the effective integrated luminosity of 2010 data sets, 36.1 pb-1 . The data
is then compared with different combined MCs of photon+jet and QCD samples. It
must be noted that for the QCD processes, the size of the MC samples corresponds
to an integrated luminosity much lower than the one analyzed in data, especially,
for pythia-8 samples. Although after applying all the selection designed for this
analysis, which would be introduced in the following section, there is no any con-
tribution from pˆT range between 30 to 50 GeV/c for both pythia-8 and pythia-6.
The sample size of MadGraph production is relatively sufficient and is mainly used
for detector level comparisons in this analysis and also the unfolding procedure for
correcting data. The related results pythia-8 and pythia-6 is taken as comparison
between different theoretical models. All the data sets and MC samples are analyzed
using the cmssw framework with recommended global setting.
Chapter 4
Event Selection
Table 4.1 Summary of the single photon triggers, corresponding run ranges and effective integrated
luminosities
HLT path name Run range L dt, pb−1
HLT_Photon20_Cleaned_L1R 138564–143962 2.46
HLT_Photon30_Cleaned_L1R 144010–147116 5.81
HLT_Photon50_Cleaned_L1R_v1 147196–148058 9.47
HLT_Photon70_Cleaned_L1R_v1 148822–149294 18.4
4.1 Trigger Requirement 35
order to correct for possible inefficiencies in the different regions of the phase space.
According to the selection of HLT in this analysis, The HLT efficiency of data has
+0.5%
been measured, 99.1%−1.2% , in the considered kinematic region.
The phenomenon when there is more than one interaction appeared inside the same
bunch crossing would be defined as pileup condition. Generally, the pileup effect
sometime would seriously affect or bias the results of analysis. This effect must be
well-understood. In order to study the pileup contribution for each data sample, all
the MC samples have been generated with distributions for the number of pileup
interaction.
The reweighting procedure is based on an iterative process: the distribution of
reconstructed vertex extracted from full data and MCs in real cross section scale. The
MCs are reweighted according to the ratio as a function of pileup interaction s for
each event. Then the weighted distribution of vertex reconstructed from MCs is used
to study the difference with respect to the distribution of vertex reconstructed from
data by calculating their ratio. MCs are reweighted repeatedly until the difference is
negligible. After at least 4 times of iteration, there is a good agreement between the
data and weighted MCs. The related results of pileup reweighting for different MC
samples are shown in Fig. 4.1.
106 3
7TeV DATA(2010)
Pythia8 MC(GJet+QCD), PU-reweight
Pythia6 MC(GJet+QCD)
Madgraph MC(GJet+QCD), PU-reweight
104 Madgraph MC(GJet+QCD), no PU-reweight
2 Madgraph MC(GJet+QCD)
DATA / MC
3
10
1.5
102
1
10
0.5
1
0
0 5 10 15 20 25 30 0 2 4 6 8 10 12 14
number of vertex number of vertex
Fig. 4.1 Reconstructed vertex distribution of uncorrected data and MC samples. a Comparison
between uncorrected data, MCs without pileup reweighting and MCs with pileup reweighting. b
Ratio plot between uncorrected data and reweighted MCs
36 4 Event Selection
number of events
103 Pythia6 Z2
102
10
1
0 5 10 15 20 25 30 35 40
Ideally, most of analyses focus on studying the physic phenomena inside a single
interaction. The LHC circulates protons inside its beam-pipes not in a continuous
stream but in several very closely packed bunches. As a matter of known fact, the
pileup effect is already unavoidable in the later run of 2010 and even becomes much
more serious. To some extent, analyzing particles from different proton-proton colli-
sion cause large uncertainties in final results when the pileup part is high pT interac-
tion. But in order not to over-kill the problem of background rejection from pileup,
we try to reject part of beam-crossings where there are more than one proton-proton
interactions as they may give rise to background from pileup. An special selection
on the number of reconstructed
2 vertex is introduced to reduce influences from pileup
effect. Events with pT i−th track (2nd vertex) < 10 GeV2 are accepted, where pT
nd
i−th track is the transverse momentum of the i-th track associated to the 2 primary
vertex. Figure 4.2 shows the related distribution for events with at least two primary
vertices. The value of selection criterion is a relatively safe setting by taking into
the pT threshold of jets account. This selection criterion rejects additional jets from
pileup interactions because the contribution of tracks associated to the 2nd primary
vertex is low in the jet reconstruction. The uncertainty associated to this selection cri-
terion has been taken into account as a source of systematic uncertainty, as discussed
later.
Figure 4.3 shows the comparison of number of reconstructed vertices between
uncorrected data and different combined MC samples. A good agreement between
data and MCs can be found in the normalization scale. But different MCs have
significant differences in the absolute scale. Such disagreement is quite uniform in
all the bins of vertices distribution. The prediction of MadGraph interfaced with
pythia 6 tune Z2 describes data better than pythia 6 tune Z2 and pythia 8 tune
4 C within the statistics uncertainty. pythia 6 tune Z2 and pythia 8 tune 4 C are
underestimated and with relatively large statistic uncertainty.
4.3 Photon Identification 37
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
105 10
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
104 PYTHIA8 4C PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
jets: | η|<2.4 PYTHIA6 Z2 1 jets: | η|<2.4 PYTHIA6 Z2
number of events
3
10 MADGRAPH+PY6 Z2 MADGRAPH+PY6 Z2
dN/N
102 10-1
10
10-2
10-3
2.5 3.5
MC / DATA
MC / DATA
3
2 number of vertex 2.5 number of vertex
1.5 2
1 1.5
1
0.5
0.5
0 0
0 2 4 6 8 10 0 2 4 6 8 10
number of vertex number of vertex
Fig. 4.3 The detector level comparison of number of reconstructed vertices between uncorrected
data and different combined MC samples in a real cross section scale and b normalization scale
Photon candidates are reconstructed from energy deposits in the ECAL, and are
referred to as super-clusters (SC). The calorimeters and tracker system of CMS are
contained within a 3.8 T magnetic field. The radiation length of the material between
the interaction point and the ECAL varies as a function of pseudorapidity (η), and
is significant enough to cause about half of photons from the hard interaction to
convert before reaching the ECAL. To allow recovery of energy emitted by the
bremsstrahlung from e+ e− pairs via photon conversions in the barrel regions of
CMS, super-clusters are up to 35 crystals wide in φ and 5 crystals wide in η. In the
endcap region of CMS, super-clusters are constructed out of at least one array of
5×5 crystals. Further details can be found in [66].
Reconstructed photon candidates are required to have pT > 75 GeV to assure a
fully efficient trigger, and |η| < 1.4442 and 1.566 < |η| < 2.5 to avoid instrumental
inefficiencies of the calorimeter. Photons are reconstructed and selected as described
in Reference [69, 70].
• ISOTRK : The track isolation is the sum of the pT of tracks located in a cone
ΔR = 0.4 centered around the line connecting the center of the super-cluster to
the primary vertex. To avoid misidentification of converted photons, tracks within
ΔR = 0.04 and Δη = 0.015 are not included in the sum.
• ISOECAL : The ECAL isolation is the sum of the E T of energy deposits in the indi-
vidual crystals located in a cone of ΔR = 0.4, centered around the super-cluster.
In order to remove the contribution due to the photon candidate, an exclusion
region consisting of an inner cone of radius ΔR = 0.06 and a strip of dimension
Δη × Δφ = 0.04 × 0.4 is defined. The sum is performed only for crystals where
|E| > 80 MeV.
38 4 Event Selection
• ISOHCAL : The Hadron Calorimeter (HCAL) isolation is defined as the sum of the
energy deposited in the HCAL towers in a cone ΔR = 0.4 around the photon
direction, excluding a cone of radius ΔR = 0.15.
• H/E: A photon reaching the ECAL without undergoing conversion to an e+ e− pair
deposits most of its energy in a 3 × 3 crystal matrix. Only a very small fraction of
the energy from the resulting shower leaks into the HCAL. The ratio of Hadronic
to Electromagnetic energy is defined by taking the ratio of energy deposited in the
HCAL and ECAL in a cone of radius ΔR = 0.15 around the photon direction.
• σ ηη : A particular shower shape variable which measures the width of the photon
super-cluster in the η direction is defined as:
where the sum runs over all the 5 × 5 crystal matrix and ωi = max(0, 4.7 +
log(E i /E 5×5 )), where E i is the energy of i-th crystal and E 5×5 is the total energy
of the crystal matrix; ηi = η̂i × δη, where η̂i is the η index of i-th crystal and
δη = 0.0174.
To have high purity of selected photons, and to be above the turn on curve for all
trigger paths used for 2010 data, listed in Sect. 4.1. Photon candidates are required
to have pT > 75 GeV. Furthermore, to remove contamination from beam halo, an
additional shower shape variable, σφφ (which have similar definition of σηη but in
the φ plane), is required to have σφφ > 0.009 and measured hit time of the most
energetic crystal, tseed , is required to have |tseed | < 1.5. These two selection criteria
are 100% for MC signal photons, but remove 99% of photon arising from beam halo.
The position of super-cluster was extrapolated to the interaction point and required
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
104 2j: pT jet2, pT jet3 > 20GeV 10 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
jets: | η|<2.4 PYTHIA6 Z2 jets: | η|<2.4 PYTHIA6 Z2
Events / 25 GeV
3
10 1
MADGRAPH+PY6 Z2 MADGRAPH+PY6 Z2
dN/N
2 -1
10 10
10 10-2
10-3
1
3.5 4
MC / DATA
MC / DATA
3 3.5
2.5 3
2.5
2
2
1.5
1.5
1 1
0.5 0.5
0 0
75 100 125 150 175 200 225 250 275 300 325 75 100 125 150 175 200 225 250 275 300 325
p (γ ) (GeV) p (γ ) (GeV)
T T
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
1
dN/N
Events / 0.4 (rad)
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
103 jets: | η|<2.4 PYTHIA6 Z2 jets: | η|<2.4 PYTHIA6 Z2
MADGRAPH+PY6 Z2 MADGRAPH+PY6 Z2
10-1
102
2.5 3
MC / DATA
MC / DATA
2 2.5
2
1.5
1.5
1
1
0.5 0.5
0 0
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3
η (γ ) η (γ )
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
1
dN/N
Events / 0.315 (rad)
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C
103
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
jets: | η|<2.4 PYTHIA6 Z2 jets: | η|<2.4 PYTHIA6 Z2
MADGRAPH+PY6 Z2 MADGRAPH+PY6 Z2
10-1
102
3 3
MC / DATA
MC / DATA
2.5 2.5
φ (γ )
2 2
1.5 1.5
1 1
0.5 0.5
0 0
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3
φ (γ ) φ (γ )
Fig. 4.4 The detector level comparison of leading photon’s transverse momentum pT (top row),
pseudorapidity η (center row) and azimuthal angle φ (bottom row) between uncorrected data and
different combined MC samples in real cross section scale (left) and normalization scale (right)
to have no match with hits in the pixel detector compatible with tracks from the hard
interaction. Such kind of requirement could reduce the possible source of background
from electron misidentification.
The summary of photon identification criteria used in this analysis is listed in
Table 4.2.
The photon energy scale, 0.43%, is determined with electrons from reconstructed
Z -boson decays with an uncertainty estimated to be 0.6% (1.5%) in the barrel (end-
40 4 Event Selection
cap) region [71–73]. Electron samples from Z -boson decays are used to estimate
the efficiency of the isolation criteria, εisolation . The εisolation in the data is found to
be higher (lower) than in simulation in the barrel (endcap) region, therefore a scale
factor of 1.008 ± 0.005 (0.984 ± 0.005) is applied to the simulation. The photon
pixel veto efficiency, εpixel veto , is estimated by analyzing final-state-radiation photons
in Z → μ+ μ− γ events. The εpixel veto in the data is found to be higher (lower) than
in simulation in the barrel (endcap) region, therefore a scale factor of 1.009 ± 0.029
(0.959 ± 0.062) is included in the simulation [70].
Comparison of basic properties of the selected leading photon with highest pT
are shown from Fig. 4.4.
Jets used in this analysis are reconstructed by using the particle flow method [74].
Associated energy correction of jets are applied to correct the pT of jets and to make
jet response flat in the η plane [75]. Jets are reconstructed with the anti-k T clustering
algorithm with a cone size of ΔR = 0.5 [27, 28]. All the jets are requested to pass
loose jet identification criteria [76]. The leading jet is requested to have pT > 75 GeV.
The other jets are requested to have pT > 20 GeV. This kind of setting is helpful
to discriminate the different sub-processes, photon-jet process and di-jet process.
Besides, all the jets are counted only if they are within the acceptance of the CMS
tracker(|η| < 2.4). Jets are rejected if they overlap with the selected leading photon
within a cone of ΔR = 0.5. The leading photon and leading jet are tagged as the
“photon-jet pair”, while the other two jets are the “di-jet pair”.
In order to reduce the probability to select jets from pile-up, jet-vertex association
algorithm is introduced. The basic idea of jet-vertex association algorithm is to select
jets according to the correlation between the target jet and different vertices. By
checking the component tracks, we can know their dependence on each vertex. For
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
104 10
dN/N
Events / 25 GeV
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
jets: | η|<2.4 PYTHIA6 Z2 1 jets: | η|<2.4 PYTHIA6 Z2
103
MADGRAPH+PY6 Z2 MADGRAPH+PY6 Z2
10-1
102
10-2
10
4 4
MC / DATA
MC / DATA
3.5 3.5
3 3
2.5 2.5
2 2
1.5 1.5
1 1
0.5 0.5
0 0
75 100 125 150 175 200 225 250 275 300 325 75 100 125 150 175 200 225 250 275 300 325
p (jet1) (GeV) p (jet1) (GeV)
T T
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
1
dN/N
Events / 0.4 (rad)
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C
103 γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
jets: | η|<2.4 PYTHIA6 Z2 jets: | η|<2.4 PYTHIA6 Z2
MADGRAPH+PY6 Z2 MADGRAPH+PY6 Z2
-1
10
102
10
10-2
3 3
MC / DATA
MC / DATA
2.5 2.5
2 2
1.5 1.5
1 1
0.5 0.5
0 0
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3
η (jet1) η (jet1)
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
103 1
dN/N
Events / 0.315 (rad)
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
jets: | η|<2.4 PYTHIA6 Z2 jets: | η|<2.4 PYTHIA6 Z2
MADGRAPH+PY6 Z2 MADGRAPH+PY6 Z2
102
10-1
3 3
MC / DATA
MC / DATA
2.5 2.5
2 2
1.5 1.5
1 1
0.5 0.5
0 0
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3
φ (jet1) φ (jet1)
Fig. 4.5 The detector level comparison of leading jet’s transverse momentum pT (top row), pseudo-
rapidity η (center row) and azimuthal angle φ (bottom row) between uncorrected data and different
combined MC samples in real cross section scale (left) and normalization scale (right)
42 4 Event Selection
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
104 10
dN/N
Events / 20 GeV
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
103 jets: | η|<2.4 PYTHIA6 Z2 1 jets: | η|<2.4 PYTHIA6 Z2
MADGRAPH+PY6 Z2 MADGRAPH+PY6 Z2
102 10-1
10 10-2
MC / DATA
3.5
2.5 p (jet2) (GeV) 3
p (jet2) (GeV)
2 T T
2.5
1.5 2
1 1.5
1
0.5 0.5
0 0
20 40 60 80 100 120 140 160 180 200 220 240 260 20 40 60 80 100 120 140 160 180 200 220 240 260
p (jet2) (GeV) p (jet2) (GeV)
T T
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
1
dN/N
Events / 0.4 (rad)
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
3
10
jets: | η|<2.4 PYTHIA6 Z2 jets: | η|<2.4 PYTHIA6 Z2
MADGRAPH+PY6 Z2 MADGRAPH+PY6 Z2
10-1
102
10
10-2
3-3 -2 -1 0 1 2 3 3-3 -2 -1 0 1 2 3
MC / DATA
MC / DATA
2.5
η (jet2) 2.5
η (jet2)
2 2
1.5 1.5
1 1
0.5 0.5
0 0
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3
η (jet2) η (jet2)
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
1
dN/N
Events / 0.315 (rad)
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
3
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
10 PYTHIA8 4C PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
jets: | η|<2.4 PYTHIA6 Z2 jets: | η|<2.4 PYTHIA6 Z2
MADGRAPH+PY6 Z2 MADGRAPH+PY6 Z2
10-1
102
3 -3 -2 -1 0 1 2 3 3 -3 -2 -1 0 1 2 3
MC / DATA
MC / DATA
2.5
φ (jet2) 2.5
φ (jet2)
2 2
1.5 1.5
1 1
0.5 0.5
0 0
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3
φ (jet2) φ (jet2)
Fig. 4.6 The detector level comparison of second leading jet’s transverse momentum pT (top row),
pseudorapidity η (center row) and azimuthal angle φ (bottom row) between uncorrected data and
different combined MC samples in real cross section scale (left) and normalization scale (right)
4.4 Jet Identification 43
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
103
dN/N
Events / 33.3 GeV
105 γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 102 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C
104 γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
jets: | η|<2.4 PYTHIA6 Z2 10 jets: | η|<2.4 PYTHIA6 Z2
3
10
MADGRAPH+PY6 Z2 1 MADGRAPH+PY6 Z2
102
10-1
10 10-2
1 10-3
10-4
10-1
40
10-54
20 40 60 80 100 120 140 160 180 200 220 0 20 40 60 80 100 120 140 160 180 200 220
MC / DATA
MC / DATA
3.5 3.5
3
p (jet3) (GeV) 3
p (jet3) (GeV)
T T
2.5 2.5
2 2
1.5 1.5
1 1
0.5 0.5
0 0
0 50 100 150 200 0 50 100 150 200
p (jet3) (GeV) p (jet3) (GeV)
T T
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
103 1
dN/N
Events / 0.4 (rad)
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
jets: | η|<2.4 PYTHIA6 Z2 jets: | η|<2.4 PYTHIA6 Z2
MADGRAPH+PY6 Z2 MADGRAPH+PY6 Z2
102 10-1
10-2
3-3 -2 -1 0 1 2 3 3-3 -2 -1 0 1 2 3
MC / DATA
MC / DATA
2.5
η (jet3) 2.5
η (jet3)
2 2
1.5 1.5
1 1
0.5 0.5
0 0
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3
η (jet3) η (jet3)
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
103 1
dN/N
Events / 0.315 (rad)
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
jets: | η|<2.4 PYTHIA6 Z2 jets: | η|<2.4 PYTHIA6 Z2
MADGRAPH+PY6 Z2 MADGRAPH+PY6 Z2
102
10-1
3 -3 -2 -1 0 1 2 3 3 -3 -2 -1 0 1 2 3
MC / DATA
MC / DATA
2.5
φ (jet3) 2.5
φ (jet3)
2 2
1.5 1.5
1 1
0.5 0.5
0 0
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3
φ (jet3) φ (jet3)
Fig. 4.7 The detector level comparison of third leading jet’s transverse momentum pT (top row),
pseudorapidity η (center row) and azimuthal angle φ (bottom row) between uncorrected data and
different combined MC samples in real cross section scale (left) and normalization scale (right)
44 4 Event Selection
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
105 102
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C 10 PYTHIA8 4C
104 γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
jets: | η|<2.4 PYTHIA6 Z2 jets: | η|<2.4 PYTHIA6 Z2
MADGRAPH+PY6 Z2 1 MADGRAPH+PY6 Z2
103
Events
dN/N
10-1
102
10-2
10
10-3
1
2 2.5
MC / DATA
MC / DATA
1.8
1.6 2
1.4
1.2 1.5
1
0.8 1
0.6
0.4 0.5
0.2
0 0
2 4 6 8 10 12 2 4 6 8 10 12
number of jets (PT > 20 GeV) number of jets (PT > 20 GeV)
Fig. 4.8 The detector level comparison of inclusive number of jets with pT > 20 GeV between
uncorrected data and different combined MC samples in real cross section scale (left) and normal-
ization scale (right)
example, there are two reconstructed vertices (ver tex A and ver tex B) and two
reconstructed jets ( jet 1 and jet 2) in one specific event. If most of component
tracks of jet 1 overlap to those tracks which are used to reconstructed ver tex A,
then we will say jet 1 is highly associated with ver tex A. That is, jet 1 is from
interaction A which is respect to ver tex A. The summary of jet identification
criteria used in this analysis is listed in Table 4.3.
Comparison of basic properties of the selected leading jets and also the exclusive
jet multiplicity ( pT > 20 GeV) are shown from Figs. 4.5, 4.6, 4.7 and 4.8.
Chapter 5
Discriminating Observables
A final state arising from background SPS events tends to have a strongly correlated
configuration in the azimuthal angle and pT -balance distribution between the two
sub-systems, while a DPS signal event has a preferred back-to-back topology and
good pT -balance for the separated systems that are not correlated. To differentiate
between SPS and DPS processes, a set of three discriminating observables are inves-
tigated:
• the difference in azimuthal angle between the jets belonging to the di-jet pair,
DPS and SPS will have different configurations for the photon and 3 jets in the
final state in the phase space of angular variable. Δφ23 computes the difference of
azimuthal angle in the di-jet sub-system. SPS background events lead to a broad
distribution for these quantities while DPS signal events contribute most at values
close to π . The later would perform a back-to-back topology of the di-jet pair.
• the transverse momentum balance of the jets belonging to the di-jet pair,
|pT jet2 + pT jet3 |
Δrel pT,23 = ; (5.2)
|pT jet2 | + |pT jet3 |
Δrel pT,23 is sensitive to the momentum balance associated with the di-jet sub-
system. A back-to-back configuration of the di-jet sub-system contributes at low
value Δrel pT,23 , while correlated pairs of jets result in a broader distribution over
the whole phase space.
Fig. 5.1 A possible orientation of the transverse momenta of the photon and the three jets in photon
+ 3 jets events (left). Illustration of the definition of the ΔS observable, applied to a photon + 3 jets
SPS event (center) and a photon + 3 jets DPS event (right)
• the azimuthal angle between the pT vectors of the photon-jet pair and the di-jet
pair,
where γ , jet1, jet2, and jet3 stand for the leading photon, the leading jet, the second
leading jet, and the third leading jet, respectively. For the di-jet pair of DPS event,
the two jets balance each other and are also independent from the first interaction.
Hence Δφ23 will be close to π and Δrel pT,23 will be small. Figure 5.1 (left) illustrates
the definition of ΔS with a possible orientation of the transverse momentum vectors
of the selected photon and jets. Figure 5.1 (center) shows a possible topology of
a background SPS event with two radiation jets emitted close to the leading jet’s
direction, which is recoiling against the photon’s direction in φ are preferred. The
correlation due to momentum conservation biases ΔS towards 180◦ (π ). For a typical
DPS signal event shown in Fig. 5.1 (right), we have better pairwise balance among the
photon-jet pair and the di-jet pair. The low correlation between the two sub-systems
leads to a flatter ΔS distribution.
Except of the above observables, there are more defined in Appendix C which can
be used to analyze the contributions of the two sub-systems. But we only include the
relevant results of the most powerful observables in the main body of this thesis.
The stable-particle level comparisons between the DPS signal and all possible
background components for different MC predictions are shown in Fig. 5.2 using
area-normalized distributions of these three observables within the phase space as
defined in Table 5.1. The hadronic energy surrounding the photon is required to be
at most 5 GeV within ΔR < 0.4 [69].
All DPS signal distributions are produced by pythia 8 tune 4C, and are taken
as a reference for the comparisons. In such DPS signal production, one hard parton
interaction generates a prompt photon and a jet and the second one generates 2
jets in the same proton–proton collision. pythia 8 tune 4C, MadGraph interfaced
with pythia 8 tune 4C, and sherpa tune CT10 are separately used to produce
different background distributions: direct photon + 3 jets SPS events, fragmentation
5 Discriminating Observables 47
CMS Simulation Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Simulation Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Simulation Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
1 γ -j: pT γ , pT jet1 > 75GeV DPS signal γ -j: p , pT jet1 > 75GeV
Tγ
DPS signal γ -j: p , pT jet1 > 75GeV
Tγ
DPS signal
σ dΔφ [1/rad]
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
direct photon+3jets, SPS direct photon+3jets, SPS direct photon+3jets, SPS
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
σ dΔS [1/rad]
jets: |η|<2.4 frag. photon+3jets, SPS 1 jets: |η|<2.4 frag. photon+3jets, SPS 10 jets: |η|<2.4 frag. photon+3jets, SPS
23
1 dσ
T,23
1 dσ
σ dΔrelp
1 dσ
10-1 10-1
10-1
10-2
10-2
0 0.5 1 1.5 2 2.5 3 0 0.2 0.4 0.6 0.8 1 0 0.5 1 1.5 2 2.5 3
Δφ (rad) ΔrelpT,23 ΔS (rad)
23
CMS Simulation Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Simulation Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Simulation Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
1 γ -j: pT γ , pT jet1 > 75GeV DPS signal γ -j: p , pT jet1 > 75GeV
Tγ
DPS signal γ -j: p , pT jet1 > 75GeV
Tγ
DPS signal
σ dΔφ [1/rad]
2j: pT jet2, pT jet3 > 20GeV direct photon+3jets, SPS 2j: pT jet2, pT jet3 > 20GeV direct photon+3jets, SPS 2j: pT jet2, pT jet3 > 20GeV direct photon+3jets, SPS
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
frag. photon+3jets, SPS frag. photon+3jets, SPS frag. photon+3jets, SPS
σ dΔS [1/rad]
jets: |η|<2.4 1 jets: |η|<2.4 10 jets: |η|<2.4
23
1 dσ
σ dΔrelp
1 dσ
10-1 10-1
10-1
10-2
10-2
0 0.5 1 1.5 2 2.5 3 0 0.2 0.4 0.6 0.8 1 0 0.5 1 1.5 2 2.5 3
1 γ -j: pT γ , pT jet1 > 75GeV DPS signal γ -j: p , pT jet1 > 75GeV
Tγ
DPS signal 10 γ -j: pT γ , pT jet1 > 75GeV DPS signal
σ dΔφ [1/rad]
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
direct photon+3jets, SPS direct photon+3jets, SPS direct photon+3jets, SPS
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
σ dΔS [1/rad]
jets: |η|<2.4 frag. photon+3jets, SPS 1 jets: |η|<2.4 frag. photon+3jets, SPS jets: |η|<2.4 frag. photon+3jets, SPS
23
1 dσ
1 dσ
σ dΔrelp
1 dσ
10-1 10-1
10-1
10-2
10-2
0 0.5 1 1.5 2 2.5 3 0 0.2 0.4 0.6 0.8 1 0 0.5 1 1.5 2 2.5 3
Fig. 5.2 Comparison of MC-based signal and background distributions for different DPS-
discriminating observables: Δφ23 (left), Δrel pT,23 (center), and ΔS (right). The signal distributions
are all produced by pythia 8 tune 4C. The comparison of background distributions are shown for
three MC event generators: MadGraph interfaced with pythia 8 tune 4C (top row), pythia 8 tune
4C (center row), and sherpa tune CT10 (bottom row). Three kinds of background contributions
are included in these comparisons: direct photon + 3 jets SPS events, fragmentation photon (frag.
photon) + 3 jets SPS events and misidentified photon (fake photon) + 3 jets events
Table 5.1 Phase space Leading photon pT > 75 GeV, |η| < 1.4442,
definition for photon + 3 jets 1.566 < |η| < 2.5
production at stable-particle
level Leading jet pT > 75 GeV, |η| < 2.4
Second leading jet pT > 20 GeV, |η| < 2.4
Third leading jet pT > 20 GeV, |η| < 2.4
48 5 Discriminating Observables
photon + 3 jets SPS events and misidentified photon + 3 jets events. The background
distributions of direct photon + 3 jets SPS events are obtained by requiring the final
state photon to come from qg Compton scattering or qq annihilation and by removing
the MPI contribution. The background distributions of fragmentation photon + 3 jets
SPS events are obtained with the same method except for requiring the final state
photon to be from the fragmentation of a parton. The background distributions of
misidentified photon + 3 jets events are extracted by requiring the final state photon
to stem from a π 0 or η meson.
The area-normalized distribution of Δφ23 is shown in Fig. 5.2 (left column). The
jets are strongly (weakly) correlated at high (low) Δφ23 values. A local maximum
is visible around Δφ23 0.5–1.0 because the anti-kT clustering jet algorithm merges
jets with a separation less than the distance parameter of R = 0.5. Figure 5.2 (center
column) shows the transverse momentum balance of the jets belonging to the di-jet
pair, Δrel pT,23 . Better transverse momentum balance of jets tends to smaller values
of Δrel pT,23 . The peak around Δrel pT,23 ∼ π indicates the imbalance in pT of di-jet
pair, which is because the jets are inexactly matched or originate from radiation of
the initial- or final-state of the leading jet. The azimuthal angle between the planes
of the photon-jet pair and the di-jet pair, ΔS, is shown in Fig. 5.2 (right column). At
low ΔS values, the photon-jet pair and the di-jet pair are not correlated.
The DPS signal distribution shows different features with respect to each back-
ground distribution and is distinguishable in the DPS-sensitive region: at high Δφ23
values, low Δrel pT,23 values, and low to medium ΔS values. A sizeable model depen-
dence can be observed from comparisons of background distributions predicted by the
investigated MC event generators indicating e.g. differences in the merging between
matrix elements and parton showers.
For pythia 8 tune 4C (Fig. 5.2 top row), the three kinds of background distribu-
tions present similar behavior to each other in Δφ23 and Δrel pT,23 but give distinct
contributions in the DPS-sensitive region of ΔS distribution. MadGraph, interfaced
with pythia 8 tune 4C (Fig. 5.2 center row), predicts the distributions of fragmenta-
tion photon + 3 jets SPS events with less back-to-back topology of Δφ23 and worse
pT -balance of Δrel pT,23 for the di-jet pair than pythia 8 tune 4 C and sherpa tune
CT10. In the region of ΔS < 2.5, fragmentation photon + 3 jets SPS events give less
contributions with respect to the other background components. sherpa tune CT10
(Fig. 5.2 bottom row) gives quite different predictions of misidentified photon + 3
jets background events with respect to the other two MC generators. Its misidentified
photon + 3 jets background events behave similar to DPS signal events in ΔS distri-
bution and also show a more back-to-back topology of Δφ23 and better pT -balance
of Δrel pT,23 for the di-jet pair than the other two kinds of background distributions.
Chapter 6
Photon + 3 Jets Events at Detector Level
In this chapter, the discriminating observables defined in Chap. 5 are used to measure
the photon + 3 jets event in data and compared with the predictions of various MC
simulations, pythia 6 tune Z2, pythia 8 tune 4C and MadGraph interfaced with
pythia 6 tune Z2, at the detector level. Furthermore, the photon purity, defined
as the fraction of isolated prompt photons in data, is estimated from an extended
maximum likelihood two-component fit with related photon signal and background
fitting templates.
Table 6.1 Summary of number of events passing the applied selection criteria for different 2010
data sets
Data sample Good vertex Trigger Photon Jets selection 2nd vertex
requirement requirement selection selection
EG 4.59 × 107 1.35 × 107 2956 279 209
Photon 2.04 × 107 6.07 × 106 31408 3122 1581
Table 6.2 Summary of number of events passing the applied selection criteria for different MC
simulations. Contributions of different processes are weighted to the effective integrated luminosity
of 2010 data, 36 pb-1 , and combined together
Data sample Good vertex Trigger Photon Jets selection 2nd vertex
requirement requirement selection selection
pythia-8 7.21 × 105 1.16 × 105 19987 1053 648
photon-jet
pythia-8 2.20 × 109 7.26 × 105 9514 1799 657
QCD
pythia-6 6.68 × 105 1.08 × 105 17378 999 508
photon-jet
pythia-6 1.95 × 109 6.60 × 105 9357 1550 771
QCD
MadGraph 1.06 × 106 1.33 × 105 22092 2600 1358
photon-jet
MadGraph 1.46 × 108 2.55 × 105 2732 312 183
QCD
Figure 6.1 shows the detector level comparisons of the data with MC simulations
for Δφ23 , Δrel pT,23 , and ΔS. In real scale, MadGraph interfaced with pythia 6
tune Z2 describes data quite well within statistic uncertainties. pythia 6 tune Z2
and pythia 8 tune 4C simulations predict a slightly smaller total number of events
than data by about 20–30%. But both two kinds of pythia simulations are with
larger statistic uncertainties, especially for pythia 8 tune 4C. There is an overall
agreement in shape of all DPS-discriminating distributions for the data and different
MC simulations except for some DPS-sensitive regions. MadGraph, interfaced with
pythia 6 tune Z2, is consistent with the data. At low and medium ΔS values, pythia 8
tune 4C and pythia 6 tune Z2 underestimate the data. At low Δrel pT,23 values,
pythia 6 tune Z2 shows much worse pT -balance of the di-jet pair than the other two
MC simulations.
6.1 DPS Performance at Detector Level 51
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
103 1
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C
dN/N
10-1
102
2 2
1.8 1.8
MC / DATA
1.6 1.6
MC / DATA
1.4 1.4
1.2 1.2
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
Δφ23 (rad) Δφ23 (rad)
CMS Preliminary, s = 7 TeV, L = 36 pb , pp → γ +3j -1
CMS Preliminary, s = 7 TeV, L = 36 pb , pp → γ +3j -1
104
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
Events / 0.101
10-1
102
10-2
10
2.5 2.5
MC / DATA
2 2
MC / DATA
1.5 1.5
1 1
0.5 0.5
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Δrelp Δrelp
T,23
T,23
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C 10 PYTHIA8 4C
104 γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
Events / 0.315 (rad)
103
10-1
102
10-2
10
2.5 2
1.8
MC / DATA
2 1.6
MC / DATA
1.4
1.5 1.2
1
1 0.8
0.6
0.5 0.4
0.2
0 0
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
ΔS (rad) ΔS (rad)
Fig. 6.1 The detector level comparisons of DPS discriminating observables, Δφ23 (top row),
Δrel pT,23 (center row), and ΔS (bottom row) between the data and various MC simulations in
real scale (left column) and normalization scale (right column), The lower panel shows the ratio
of the MC simulation to the data. The error bars on the ratio histogram represent the statistical
uncertainty of the data and simulated samples added in quadrature
52 6 Photon + 3 Jets Events at Detector Level
The photon purity, defined as the fraction of isolated prompt photons in data, of
reconstructed events satisfying the full selection is estimated with a binned extended
maximum likelihood fit to the σηη distribution with the expected signal and back-
ground components of photons. The shapes of signal and background distributions
are obtained as described in [69]. According to the definition of signal photon, the
photon source of DPS signal events can be direct photons and fragmentation photons.
The background photon is due to the electromagnetic decays of light mesons such
as π 0 and η. Events are selected with almost the same selection criteria except for
the σηη selection criteria in Table 4.2. These are performed by requiring two regions
of barrel, |η| < 1.4442, and endcap, 1.566 < |η| < 2.5, for photons and two bins of
E T,γ : 75 < E T,γ < 95 GeV and E T,γ > 95 GeV, in each region separately.
The signal distribution is extracted from the photon-jet samples of the MC simu-
lation of MadGraph interfaced with pythia 6 tune Z2 by taking the σηη distribution
of direct photons identified using MC truth information. These signal distribution
of direct photons are cross-checked with the signal distribution of fragmentation
photons, which is derived from the di-jet samples of the MC simulation of
1 1
dN/N
dN/N
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
0 0.005 0.01 0.015 0.02 0.025 0.03 0 0.005 0.01 0.015 0.02 0.025 0.03
σiηiη σiηiη
1 1
CMS Preliminary 2010 CMS Preliminary 2010
s=7TeV, L=36 pb-1 direct photon s=7TeV, L=36 pb-1 direct photon
0.9 0.9
pp→γ +3jet pp→γ +3jet
γ : 75<E <95 (GeV), Endcap γ : ET>95 (GeV), Endcap
0.8 T
γ : 1.566<|η|<2.5, Endcap 0.8 γ : 1.566<|η|<2.5, Endcap
fragmentation photon fragmentation photon
0.7 0.7
0.6 0.6
dN/N
dN/N
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05
σiηiη σiηiη
Fig. 6.2 Distributions of σηη for direct photons from the photon-jet MadGraph MC sample and
fragmentation photons from the di-jet MadGraph MC sample in different E T,γ bins and different
η regions. The two types of photons have the same σηη distributions
6.2 Purity of Isolated Prompt Photons 53
0.8 0.8
CMS Preliminary 2010 Data side-band CMS Preliminary 2010 Data side-band
s=7TeV, L=36 pb-1 s=7TeV, L=36 pb-1
0.7 pp→γ +3jet 0.7 pp→γ +3jet
γ : 75<E <95 (GeV), Barrel MC side-band γ : E >95 (GeV), Barrel MC side-band
T T
γ : |η|<1.4442, Barrel γ : |η|<1.4442, Barrel
0.6 MC background 0.6 MC background
0.5 0.5
dN/N
dN/N
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.005 0.01 0.015 0.02 0.025 0.03 0 0.005 0.01 0.015 0.02 0.025 0.03
σiηiη σiηiη
0.6 0.6
CMS Preliminary 2010 Data side-band CMS Preliminary 2010 Data side-band
s=7TeV, L=36 pb-1 s=7TeV, L=36 pb-1
pp→γ +3jet pp→γ +3jet
0.5 γ : 75<ET<95 (GeV), Endcap
MC side-band 0.5 γ : ET>95 (GeV), Endcap
MC side-band
γ : 1.566<|η|<2.5, Endcap γ : 1.566<|η|<2.5, Endcap
MC background MC background
0.4 0.4
dN/N
dN/N
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05
σiηiη σiηiη
Fig. 6.3 Distributions of σηη for true background in MC (orange) and candidates from the ISOTRK
side-band in data (black) and MC (green) for different E T,γ bins and different η regions
MadGraph interfaced with pythia 6 tune Z2. All of the σηη distributions are shown
in Fig. 6.2.
The background σηη distribution is obtained from the data by requiring the iso-
lation criterion of the scalar sum of the charged-particle pT within ΔR = 0.4 for
photons larger than 2 GeV but less than 5 GeV, while keeping all other selection cri-
teria (Table 4.2) unchanged. This can be thought of as a side-band to the signal cut,
ISOTRK < 2 GeV/c. Under this procedure, the σiηiη background component shapes
are obtained for:
• The side-band in data, labelled Data side-band.
• The side-band in MC, labelled MC side-band.
• The signal region in MC, requiring the final state photon to have ISOTRK < 2
GeV/c and come from a hadron or have no match, labelled MC backgr ound.
From Fig. 6.3, we can find the agreement between all of three distributions is good
but the comparison is limited by the statistics in data in endcap region.
For each region, data is fitted with f (σηη ) = NS S(σηη ) + NB B(σηη ), where NS and
NB are the estimated numbers of signal and background events in different E T,γ and
different η region. The fit is performed using a binned extended maximum likelihood
by minimizing:
54 6 Photon + 3 Jets Events at Detector Level
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j+X CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j+X
600 800
Data Data
γ : 75 < E < 95 GeV γ: E > 95 GeV
T,γ 700 T,γ
500 |η| < 1.4442, Barrel Fit result |η| < 1.4442, Barrel Fit result
Background 600 Background
400
500
γ candidates
γ candidates
300 400
300
200
200
100
100
0 0
0 0.005 0.01 0.015 0.02 0.025 0.03 0 0.005 0.01 0.015 0.02 0.025 0.03
σηη σηη
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j+X CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j+X
200
140 Data Data
γ : 75 < E < 95 GeV 180 γ: E > 95 GeV
T,γ T,γ
1.566 < |η| < 2.5, Endcap Fit result 1.566 < |η| < 2.5, Endcap Fit result
120 160
Background Background
140
100
120
γ candidates
γ candidates
80
100
60 80
60
40
40
20
20
00 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05
00 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05
σηη σηη
Fig. 6.4 Measured σηη distributions. The extended maximum likelihood fit result is overlaid in
each plot and the component for background is shown separately
Table 6.3 Estimated signal yield as a function of the photon η. For illustration, the purity in the
region σηη < 0.01 of barrel and the region σηη < 0.03 of endcap is also reported
η region, E T bin Fitted yield Purity
Barrel, 75–95 388.16 ± 11.82 85.9% ± 3.0% (stat.)
Barrel, 95–∞ 530.06 ± 16.02 85.5% ± 3.0%3.0 (stat.)
Endcap, 75–95 243.59 ± 10.72 78.8% ± 3.0%4.4 (stat.)
Barrel, 95–∞ 368.15 ± 11.56 90.3% ± 3.0%3.1 (stat.)
n
L = −lnL = −(NS + NB ) + Ni ln(NS Si + NB Bi ) (6.1)
i=1
where n is the number of bins in σηη , Ni is the observed number of events for the i-th
σηη bin and Si and Bi are the values of the corresponding components in that bin. The
results of fitting for different E T,γ and different η regions are shown in Fig. 6.4. The
raw yield of isolated photons (direct photons and fragmentation photons) is reported
in Table 6.3.
6.2 Purity of Isolated Prompt Photons 55
250 250
200 200
150 150
100 100
50 50
0 0
-10 -8 -6 -4 -2 0 2 4 6 8 10 -10 -8 -6 -4 -2 0 2 4 6 8 10
300 300
250 250
200 200
150 150
100 100
50 50
0 0
-10 -8 -6 -4 -2 0 2 4 6 8 10 -10 -8 -6 -4 -2 0 2 4 6 8 10
Fig. 6.5 Distributions of pull from the fitter test. The input values of shape parameters have been
Gaussian randomized. From the top left to the bottom are for (1) 75 < E T,γ < 95 GeV and |η| <
1.4442, (2) 75 < E T,γ < 95 GeV and 1.566 < |η| < 2.5, (3) E T,γ > 95 GeV and |η| < 1.4442 and
(4) E T,γ > 95 GeV and 1.566 < |η| < 2.5
Table 6.4 The fit bias in η region, E T bin Fitter bias (%)
different E T,γ and different η
regions Barrel, 75–95 5.0
Barrel, 95–∞ 6.7
Endcap, 75–95 2.8
Barrel, 95–∞ 4.4
E T,γ and different η regions and the Gaussian fit results. If the pull mean differs from
Input
zero it means that the fitter model is not ideal and a fit bias (NsFit − Ns /NsFit ) is
introduced into the results of photon purity. Table 6.4 summarizes the fit bias of for
photons in different E T,γ bins and barrel and endcap regions, respectively. The fit
bias is included in the final systematic uncertainty of photon purity study.
The background component shape, the σηη < distribution for data in a ISOTRK
side-band of 2 < ISOTRK < 5 GeV/c, may systematically affect the results. Instead of
using side-band distribution of data, Data side-band, we perform the same fitting
method for data by using the other two kinds of background templates, MC side-
band and MC backgr ound. Then the difference will be introduced as the systematic
uncertainty of background shape. Besides, the background templates for fitting are
56 6 Photon + 3 Jets Events at Detector Level
Table 6.5 The final systematic uncertainty of the background component shape for different E T,γ
and different η regions
η region, E T bin Background component shape (%) Background fluctuation (%)
Barrel, 75–95 3.0 1.4
Barrel, 95–∞ 7.5 1.4
Endcap, 75–95 1.5 2.4
Barrel, 95–∞ 4.3 1.1
Table 6.6 The photon purity of reconstructed events satisfying the full selection in the different
E T,γ and different η regions
E T,γ bin, GeV |η| < 1.4442 1.566 < |η| < 2.5
75–95 85.9% ±3.0% (stat.) ±4.0% (syst.) 78.8% ±4.4% (stat.) ±4.0% (syst.)
95–∞ 85.5% ±3.0% (stat.) ±10.2% (syst.) 90.3% ±3.1% (stat.) ±6.2% (syst.)
fluctuated in order to take into account the effects of insufficient sample size on
producing the background fitting templates. The final systematic uncertainties of the
background component shapes are summarized in Table 6.5.
The final result of photon purity is listed in Table 6.6 for each E T,γ bin in the barrel
and endcap regions. The dominant contribution of systematic uncertainty is caused
by the background distribution [69]. For σηη < 0.01 in the barrel region, the photon
purity in the data after full selection is about 86%. However, for σηη < 0.03 in the
endcap region, the photon purity increases from 79% at E T,γ = 75 GeV to 90% at
E T,γ = 95 GeV.
Chapter 7
Correction and Unfolding
where R(xtrue , xreco ) is the response matrix. The response matrix encapsulates the
detector response for the measurement. Response matrices are obtained from Mad-
Graph interfaced with pythia 6 tune Z2 full simulation sample here. The steps
to unfolding involves building a response matrix that maps the true distribution at
generator-truth level to the measured distribution at detector-reconstruction level.
For a given number of events in a bin of the true distribution, the response matrix
element R(xtrue , xreco ) gives the fraction of events that end up in the measured bin.
The response matrices are constructed for each measured observables, as follows.
• Events are accepted if they can satisfy the generator-truth level selection and
detector-reconstruction level selection with the same kinematic acceptance.
• If the event satisfies the generator-truth level acceptance criteria, but has no coun-
terpart inside the detector acceptance, such event is tagged as ”missing” and is
recorded as lost due to detector inefficiency.
• If the event accepted at detector-reconstruction level, but has no counterpart at
generator-truth level, such event is tagged as “fake” and is also recorded in the
response matrices.
For the migration effects inside the phase space, purity and stability are typical
quantities used in related investigation. The purity is defined as the fraction of the
matched events that remain in the same bin as the detector level selection, while the
stability is the fraction of the matched events that stay in the same bin as the generator
level selection. The detailed formulas of these two quantities are shown below:
Nmatched (Edet ∈ bin i ∧ Egen ∈ bin i)
Pi = (7.4)
Nmatched (Edet ∈ bin i)
Acceptance, background, purity and stability are studied for every measured dis-
criminating observable. The related results are shown in Fig. 7.1. Basically, a good
7.1 Acceptance, Background, Purity and Stability 59
Events / 0.101
Events / 0.315 rad
1
1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
0 0.5 1 1.5 2 2.5 3 0 0.2 0.4 0.6 0.8 1
Δφ(jet2,jet3) (rad) ΔrelPT(Jet2,jet3)
1.4
CMS Preliminary 2010
s=7TeV, L=36 pb-1 Acceptance
Purity
1.2 pp→γ +3jet
Stability
Background
Events / 0.315 (rad)
0.8
0.6
0.4
0.2
0
0 0.5 1 1.5 2 2.5 3
ΔS (photonjet,dijet) (rad)
Fig. 7.1 Acceptance, purity, stability and background for Δφ23 (upper left), Δrel pT,23 (upper right)
and ΔS (lower)
measurement should have high purity, high stability, high acceptance but low back-
ground over the whole phase space. These properties directly reflect the quality of
measurement. All the related results are performed by analyzing MadGraph inter-
faced with pythia 6 tune Z1 full simulation sample.
The acceptance for each measured observable is overall around 0.3–0.4 without
significant fluctuations. The phenomenon of low acceptance had been investigated. It
is due to the photon identification and the selection of 2nd primary vertex. Basically,
we present a comparison between original study and a special study by removing the
photon identification selection, which is described in Sect. 4.3, except for selection
criteria of the pT threshold and η acceptance region. In order to have higher photon
purity, photon isolation criteria are introduced to reduce the background rate, caused
by any possible fake photons or photons come from the neutral light meson decay,
such as π 0 and η. But the cost is that some of signal photons fail the selection cuts.
Moreover, in order to reduce the influence from pileup, a selection on the 2nd vertex
described before will make the analysis only focus on events with only one vertex or
events whose pileup is relatively soft. The acceptance decreases due to the composite
influence of these effects.
60 7 Correction and Unfolding
On the other hand, the background curves for different observables are fluctu-
ating a bit more around 0.1–0.3. As described before, background events are due
to the unmatched events, especially those events which are accepted at the detector-
reconstruction level but are not presented at the generator-truth level because of detec-
tion inefficiencies, migration effects, or pileup events, where one or more selected
particles at detector-reconstruction level come from pileup interactions; such events
can not have a counterpart at generator-truth level because there is no information of
particles from the pileup.
The purity and stability are overall above 0.6 for all the measured observ-
ables except for the low degree region of ΔS. For the angular observables, i.e.
Δφ23 , the curves are around 0.7–0.9, while for the pT balance observables, i.e.
Δrel pT,23 , the purity and stability are relatively lower and around 0.5–0.9 due to the
effect of pT resolution of selected photon and jets. And discriminating observables
involved more jets will also have relatively lower purity and stability. In the region
ΔS < 2.5, the purity and stability curves are fluctuating a bit more around 0.2–0.4.
ΔS is a composite discriminating observable contain both angular and momentum
properties. The lower purity and stability in the low ΔS region indicate the migration
effect would be relatively serious.
Measurements must be unfolded for detector effects. A direct comparison is then per-
mitted of distributions from different experiments without knowledge of the detector
response for each experiment. Data after unfolding can facilitated the automated
tuning of Monte Carlo generators using in multiple measurements, which can be a
great feedback to theorists to improve their models.
In this analysis, the iterative (D’Agostini) unfolding method [77], which is imple-
mented in the RooUnfold package [78], is mainly used to remove the known effects
of measurement resolutions, systematic biases, and detector inefficiency.
This unfolding method is based on Bayes’ theorem in the following form:
P(Ek | Cj )P(Cj )
P(Cj | Ek ) = (7.6)
nC
P(Ek | Cj )P(Cj )
j=1
where P(Cj ) is the probability of the j-th cause (generator-truth level object),
P(Ek | Cj ) is the conditional probability of j-th cause to produce the k-th effect
(detector-reconstruction level object), and P(Cj | Ek ) is the probability that the k-th
effect is due to the j-th cause.
7.2 Unfolding with ‘RooUnfold’ 61
• Step 1. Let p0 ≡ (p1 , p2 , ..., pM ) be an initial set of probabilities (derived from the
input distribution,
or even a constant value) for an event to be found in each bin,
and ntotal = ni be the total number of entries.
i
• Step 2. Define
μ̂0 ≡ ntotal p0 (7.8)
⎛ ⎞
1
N
⎜ Rij pi ⎟
μ̂i = ⎜ ⎟ nj (7.9)
εi ⎝ ⎠
j=1 Rkj pk
k
and we update a new value for μ by using the response matrices Rjk .
• Step 3. Extract a new set of probability p
μ̂k
pk = (7.10)
ntotal
• Step 4. Iterate step 2 and 3 on Monte Carlo simulation sample, until the change
in Λ2 between iterations indicates the corrected distribution has stabilized. A
number of 4–6 iterations has turned to be sufficient to get a stable result for the
matrix inversion.
62 7 Correction and Unfolding
R = U S V T, (7.11)
10 4
CMS Preliminary 2010 CMS Preliminary 2010
Bayesian Unfolding Bayesian Unfolding
s = 7 TeV, L = 36 pb-1, pp → γ +3jets s = 7 TeV, L = 36 pb-1, pp → γ +3jets
γ jet: p > 75 GeV, dijet: pT > 20 GeV SVD Unfolding γ jet: p > 75 GeV, dijet: pT > 20 GeV SVD Unfolding
3
T
γ : |η| < 1.4442, 1.566 < |η| < 2.5 Gen. MC truth 10 T
γ : |η| < 1.4442, 1.566 < |η| < 2.5 Gen. MC truth
10 2 Reco. Measurement Reco. Measurement
jets: |η| < 2.4 jets: |η| < 2.4
2
10
σ [pb]
σ [pb]
10 10
1 10-1
0 0.5 1 1.5 2 2.5 3 0 0.2 0.4 0.6 0.8 1
Δφ(jet2,jet3) (rad) ΔrelPT(jet2,jet3)
3.5
3 4
Reco. Level ratio
3.5
Reco. Level ratio
2.5 3
2 2.5
1.5 2
1.5
1
1
0.5 0.5
0 0
0 0.5 1 1.5 2 2.5 3 0 0.2 0.4 0.6 0.8 1
5
10
CMS Preliminary 2010
-1 Bayesian Unfolding
s = 7 TeV, L = 36 pb , pp → γ +3jets
4
10 γ jet: p > 75 GeV, dijet: p > 20 GeV
T T
SVD Unfolding
γ : |η| < 1.4442, 1.566 < |η| < 2.5 Gen. MC truth
jets: |η| < 2.4 Reco. Measurement
3
10
σ [pb]
2
10
10
-1
10
0 0.5 1 1.5 2 2.5 3
ΔS (photonjet,dijet) (rad)
3.5
Reco. Level ratio
3
2.5
2
1.5
1
0.5
0
0 0.5 1 1.5 2 2.5 3
Fig. 7.2 Cross section measurements obtained from the Set 2 MadGraph sample at detector-
reconstruction and generator-truth level and unfolded distributions obtained with the SVD correction
and the Bayesian method: Δφ23 (upper left), Δrel pT,23 (upper right) and ΔS (lower). The response
matrices are extracted from the Set 1 MadGraph sample
7.2 Unfolding with ‘RooUnfold’ 63
4
10
CMS Preliminary 2010 CMS Preliminary 2010
-1 -1
s = 7 TeV, L = 36 pb , pp → γ +3jets Corrected data, Bayesian Unfolding s = 7 TeV, L = 36 pb , pp → γ +3jets Corrected data, Bayesian Unfolding
γ jet: p > 75 GeV, dijet: pT > 20 GeV Corrected data, SVD Unfolding 3 γ jet: p > 75 GeV, dijet: pT > 20 GeV Corrected data, SVD Unfolding
T
γ : |η| < 1.4442, 1.566 < |η| < 2.5
10 T
γ : |η| < 1.4442, 1.566 < |η| < 2.5
2
10 Uncorrected data Uncorrected data
jets: |η| < 2.4 jets: |η| < 2.4
2
10
σ [pb]
σ [pb]
10
10
1 10
-1
4.5 4.5
4 4
3.5 3.5
3 3
2.5 2.5
2 2
1.5 1.5
1 1
0.5 0.5
0 0
0 0.5 1 1.5 2 2.5 3 0 0.2 0.4 0.6 0.8 1
105
CMS Preliminary 2010
-1
s = 7 TeV, L = 36 pb , pp → γ +3jets Corrected data, Bayesian Unfolding
104 γ jet: p > 75 GeV, dijet: pT > 20 GeV Corrected data, SVD Unfolding
T
γ : |η| < 1.4442, 1.566 < |η| < 2.5
Uncorrected data
jets: |η| < 2.4
103
σ [pb]
2
10
10
10-1
0 0.5 1 1.5 2 2.5 3
ΔS (photonjet,dijet) (rad)
Corrected data/Uncorrected data
0
0 0.5 1 1.5 2 2.5 3
Fig. 7.3 Cross section measurements obtained with the complete 2010 data samples at detector-
reconstruction level and unfolded distributions obtained with the SVD correction and the Bayesian
method: Δφ23 (upper left), Δrel pT,23 (upper right) and ΔS (lower). The response matrices are
extracted from MadGraph sample
Several studies are performed to assess in which way quantities entering the absolute
cross section measurements and normalized shape distributions can be systematically
affecting the results. Analyses using photons and jets have to consider in particular
the impact of their corresponding energy scale and energy resolution. The uncer-
tainty on the correction of photon identification efficiency needs to be taken into
account. Studies on the 2nd primary vertex selection, jet-vertex association algo-
rithm, pileup reweighting, luminosity, trigger efficiency and model dependence of
unfolding are also performed. The sample size of MC simulation are treated as an
additional systematic uncertainty since the sample size affect the output response
matrix of unfolding.
All of the systematic uncertainties are expressed as variations with respect to
the nominal values, Ni /Ni − 1, where Ni is the perturbed bin content and Ni is
the nominal bin content of i th bin in the measured distribution. That is, the average
discrepancies between the perturbed and the nominal distribution are taken bin-by-
bin and referred as the uncertainties. Whenever an effect is found to be smaller than
0.1% it is not reported.
For both of real cross sections and normalized shape distributions, the minimum
and the maximum value of the uncertainties along the whole phase space are listed
together with their average. This average is taken as the reference of systematic
uncertainty.
Table 8.1 Effect of the photon energy scale uncertainty on the event yield for measurements of
different discriminating observables in absolute cross section scale. The maximum, the minimum
and the average value are listed
Measured observable Δφ23 (%) Δrel pT,23 (%) ΔS (%)
Max. 6.8 5.0 3.9
Min. 1.5 1.2 1.8
Avg. 3.9 2.8 2.7
Table 8.2 Effect of the photon energy scale uncertainty on the normalized shape for measurements
of different discriminating observables in normalization scale. The maximum, the minimum and
the average value are listed
Measured observable Δφ23 (%) Δrel pT,23 (%) ΔS (%)
Max. 4.6 2.8 4.2
Min. 0.7 0.4 0.3
Avg. 2.2 1.2 0.8
shown in Tables 8.1 and 8.2. A contribution is around 2–5% for the absolute cross
section and around 1–2% for the normalized shape distributions.
Table 8.3 Effect of the photon energy resolution uncertainty on the event yield for measurements
of different discriminating observables in absolute cross section scale. The maximum, the minimum
and the average value are listed
Measured observable Δφ23 (%) Δrel pT,23 (%) ΔS (%)
Max. 1.1 1.7 6.2
Min. 0.3 0.2 0.4
Avg. 0.6 0.8 2.4
Table 8.4 Effect of the photon energy resolution uncertainty on the normalized shape for measure-
ments of different discriminating observables in normalization scale. The maximum, the minimum
and the average value are listed
Measured observable Δφ23 (%) Δrel pT,23 ΔS (%)
Max. 0.6 2.2% 6.6
Min. 0.1 – 0.2
Avg. 0.3 0.7% 2.2
Besides the high quality requirement is needed for photon, this analysis is also based
on the selection of high jet multiplicity. The detailed understanding of the jet energy
scale is of crucial importance for analyses involved jets. The corresponding uncer-
tainty of jet energy scale is believed to play an important role that affects the total
systematic uncertainty. The transverse momentum of each jet is varied up and down
according to the uncertainty associated to the reconstructed jet pT [82]. The resulting
effects in the event yield and normalized shape are shown in Tables 8.5 and 8.6. A
contribution is around 5–17% for the absolute cross section and around 1–9% for
the normalized shape distributions. Relatively, the effect of jet energy scale makes
smaller uncertainties in those discriminating observables connected to the configura-
tion of the azimuthal angle for the absolute cross section. But the situation is opposite
for the normalized shape distribution.
Table 8.5 Effect of the jet energy scale uncertainty on the event yield for measurements of different
discriminating observables in absolute cross section scale. The maximum, the minimum and the
average value are listed
Measured observable Δφ23 (%) Δrel pT,23 (%) ΔS (%)
Max. 11.1 11.1 29.1
Min. 2.9 7.2 6.8
Avg. 7.5 9.3 17.2
68 8 Systematic Uncertainties
Table 8.6 Effect of the jet energy scale uncertainty on the normalized shape for measurements of
different discriminating observables in normalization scale. The maximum, the minimum and the
average value are listed
Measured observable Δφ23 (%) Δrel pT,23 (%) ΔS (%)
Max. 6.0 2.2 18.4
Min. 0.6 0.5 2.3
Avg. 2.7 1.5 8.9
The jet energy response does not exactly correspond to the true value of the measured
physical quantity. Measurements show that the jet energy resolution in data is worse
than in the simulation. The method of studying the effect of jet energy resolution is
similar to what had
been done for photon energy resolution. Except for the matching
aperture ΔR ≡ (Δη)2 + (Δφ)2 , the value would be 0.3. The MC reconstructed jet
gen gen
energy was smeared in the following way: ( pTreco smear ed) = pT ± A×( pTreco − pT )
where the factor A is the official parameters of the detector resolution for the 2010
gen
data as described in [82] and pT , pTreco were transverse momenta of the two matched
jets at the generator level and detector level. The resulting effects in the event yield
and normalized shape are shown in Tables 8.7 and 8.8. A contribution is around
5–7% for the absolute cross section and around 4–6% for the normalized shape
distributions.
Table 8.7 Effect of the jet energy resolution uncertainty on the event yield for measurements of
different discriminating observables in absolute cross section scale. The maximum, the minimum
and the average value are listed
Measured observable Δφ23 (%) Δrel pT,23 (%) ΔS (%)
Max. 9.7 13.2 10.9
Min. 0.8 0.8 2.4
Avg. 5.2 7.0 7.0
Table 8.8 Effect of the jet energy resolution uncertainty on the normalized shape for measurements
of different discriminating observables in normalization scale. The maximum, the minimum and
the average value are listed
Measured observable Δφ23 (%) Δrel pT,23 (%) ΔS (%)
Max. 8.2 8.1 6.2
Min. 1.1 1.2 1.2
Avg. 4.0 5.5 5.0
8.5 Model Dependence Uncertainty 69
Table 8.9 Effect of the model dependence uncertainty on the event yield for measurements of
different discriminating observables in absolute cross section scale. The maximum, the minimum
and the average value are listed
Measured observable Δφ23 (%) Δrel pT,23 (%) ΔS (%)
Max. 10.6 10.5 18.0
Min. 3.7 1.9 4.0
Avg. 8.0 6.9 11.6
Table 8.10 Effect of the model dependence uncertainty on the normalized shape for measurements
of different discriminating observables in normalization scale. The maximum, the minimum and
the average value are listed
Measured observable Δφ23 (%) Δrel pT,23 (%) ΔS (%)
Max. 3.9 3.9 11.9
Min. 1.0 1.1 3.1
Avg. 2.5 2.3 5.7
The half discrepancy between MadGraph and pythia-6 samples at the detector level
is used to study the effect arising from different physic models. The data are unfolded
with response matrices obtained from the two different MC generators pythia 6 and
MadGraph interfaced with pythia 6 tune Z2. The resulting effects in the event yield
and normalized shape are shown in Tables 8.9 and 8.10. A contribution is around
7–12% for the absolute cross section and around 2–6% for the shape distributions.
Systematic uncertainty for photon identification efficiency is assigned for the differ-
ence between the efficiency results in MC and those measured from data by using
the “Tag and Probe” technique. Both corrections of the isolation and the pixel seed
veto efficiency are taken into account for evaluating their effects on the final results,
at described in Sect. 4.3. The efficiency of the isolation criteria, εisolation , in the data is
found to be higher (lower) than in simulation in the barrel (endcap) region, therefore
a scale factor of 1.008 ± 0.005 (0.984 ± 0.005) is applied to the simulation. The
photon pixel veto efficiency, εpixel veto , in the data is found to be higher (lower) than
in simulation in the barrel (endcap) region, therefore a scale factor of 1.009 ± 0.029
(0.959 ± 0.062) is included in the simulation [70]. Such uncertainties arising from
the correction of photon identification efficiencies are quite uniform. Around 4% was
estimated in the real cross section distributions and 0.1% in the normalized shape
distributions.
70 8 Systematic Uncertainties
number of events
400
300
200
100
0
0 2 4 6 8 10 12 14 16 18 20
number of pile-up
The number of interactions in the data is estimated from the measured luminosity
in each bunch crossing times an average total inelastic cross section. The uncer-
tainty due to pileup reweighting can be estimated by shifting the overall mean of the
“reweighting target” (i.e., the data distributions) number of interactions up or down
by the corresponding amount, then redoing the reweighting. Figure 8.1 shows the
estimated distribution of number of pileup with a fitted Poisson function. By taking
a variation ±3.6% [83] for shifting the mean of the fitted Poisson function is suffi-
cient to cover the uncertainties due to pileup modeling. The resulting effects in the
event yield and normalized shape are shown in Tables 8.11 and 8.12. A contribution
is around 2–3% for the absolute cross section and less than 1% for the normalized
shape distributions.
Table 8.11 Effect of the pileup reweighting uncertainty on the event yield for measurements of
different discriminating observables in absolute cross section scale. The maximum, the minimum
and the average value are listed
Measured observable Δφ23 (%) Δrel pT,23 (%) ΔS (%)
Max. 2.7 4.2 4.8
Min. 2.4 2.3 1.7
Avg. 2.6 2.9 2.9
8.8 Second Primary Vertex Selection Uncertainty 71
Table 8.12 Effect of the pileup reweighting uncertainty on the normalized shape for measurements
of different discriminating observables in normalization scale. The maximum, the minimum and
the average value are listed
Measured observable Δφ23 Δrel pT,23 (%) ΔS (%)
Max. 0.9% 1.9 4.1
Min. – 0.1 0.1
Avg. 0.3% 0.7 1.0
the pT of track associated to 2nd primary vertex is varied up and down according
to its corresponding uncertainty [84]. And the resulting effects in the event yield
and shape are quite uniform. Around 1.4% was estimated in the real cross section
distributions and around 0.5% in normalized shape distributions.
Table 8.13 Statistics uncertainties of Monte Carlo (MadGraph) simulation. The maximum, the
minimum and the average value are listed
Measured observable Δφ23 (%) Δrel pT,23 (%) ΔS (%)
Max. 7.3 10.5 11.2
Min. 2.4 2.0 1.4
Avg. 4.0 6.4 8.0
72 8 Systematic Uncertainties
systematic uncertainty due to the size of MC samples in both absolute cross section
scale and normalization scale. Table 8.13 shows the related uncertainties for different
discriminating observables.
From Table 8.14 and the pT threshold used in this analysis, we can know only Pho-
ton70 trigger has inefficiency problem in the +η endcap region. Furthermore, the
inefficiency only exists for events whose photon’s E T is between 75 and 80 GeV.
From Fig. 8.2, only 16 events with HLT_Photon70 requirement are in the +η
endcap region. Hence, the correction of HLT efficiency in such region should be
+0.5%
negligible. The HLT efficiency of data has been measured, 99.1%−1.2% , in the con-
sidered kinematic region. We simply quote the systematic uncertainty of HLT by
taking every uncertainty of HLT_Photon70 and the fractions of events in different
region into account. The systematic uncertainty of HLT is 1.2% and assigned to both
results in absolute and normalization scales.
Table 8.14 Efficiency of single photon HLT paths as a function of the photon transverse energy in
barrel and endcaps
HLT path name Photon E T ( GeV) −η Endcap Barrel +η Endcap
HLT_Photon70 75 - Inf. 100 +0.
−11.6 100 +0.
−1.4 90.0 +5.5
−7.9
CMS Preliminary 2010, s = 7 TeV, L = 36.14 pb-1, pp → γ +3jets CMS Preliminary 2010, s = 7 TeV, L = 36.14 pb-1, pp → γ +3jets
100 100
γ -jet: pT > 75 GeV γ -jet: pT > 75 GeV
90 2jets: p > 20 GeV 90 2jets: p > 20 GeV
T T
γ : 1.566<η< 2.5, +η endcap γ : 1.566<η< 2.5, +η endcap
80
jets: |η|<2.4 DATA 80
jets: |η|<2.4
DATA, HLT_Photon70
Events / 5 GeV
Events / 5 GeV
70 70
60 60
50 50
40 40
30 30
20 20
10 10
0 0
100 150 200 250 300 100 150 200 250 300
PT (γ ) (GeV) PT (γ ) (GeV)
Fig. 8.2 Photon pT distribution of data in the in the +η endcap region. a all passed events. b events
with HLT_Photon70 requirement
8.12 Total Systematic Uncertainty 73
For the separate contributions, jet energy scale, jet energy resolution, photon energy
scale, model dependence and the sample size of MC simulation dominate the sys-
tematic uncertainties in both absolute scale of real cross section and normalization
scale of shape. Basically, the systematic uncertainties in the normalization scale are
smaller than those in the absolute cross section scale.
0.4 0.3
CMS Preliminary 2010 Total Uncertainty CMS Preliminary 2010 Total Uncertainty
Jet Energy Scale Jet Energy Scale
s=7TeV, L=36 pb-1 Jet Energy Resolution s=7TeV, L=36 pb-1 Jet Energy Resolution
0.35 pp→γ +3jet Photon ID Efficiency pp→γ +3jet Photon ID Efficiency
Photon Energy Scale 0.25
Relative uncertainty
Relative uncertainty
Photon Energy Scale
Photon Energy Resolution
Photon Energy Resolution
Model dependence
0.3 Model dependence
MC Statistics
HLT efficiency MC Statistics
2nd vertex selection 0.2 HLT efficiency
0.25 Pile-up Re-weighting 2nd vertex selection
Luminosity Pile-up Re-weighting
0.2 0.15
0.15
0.1
0.1
0.05
0.05
0 0
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
Relative uncertainty
0.2 0.15
0.15
0.1
0.1
0.05
0.05
0 0
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
ΔrelPT(jet2,jet3) ΔrelPT(jet2,jet3)
0.5 0.5
Total Uncertainty Total Uncertainty
CMS Preliminary 2010 CMS Preliminary 2010
Jet Energy Scale Jet Energy Scale
0.45 s=7TeV, L=36 pb-1 Jet Energy Resolution 0.45 s=7TeV, L=36 pb-1 Jet Energy Resolution
pp→γ +3jet Photon ID Efficiency pp→γ +3jet Photon ID Efficiency
Photon Energy Scale
Photon Energy Scale
0.4 0.4
Relative uncertainty
Relative uncertainty
0.25 0.25
0.2 0.2
0.15 0.15
0.1 0.1
0.05 0.05
0 0
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
Fig. 8.3 Single contributions and total systematic uncertainty of Δφ23 (top row), Δrel pT,23 (center
row) and ΔS (bottom row) in absolute scale of real cross section (left column) and normalization
scale of shape (right column)
74 8 Systematic Uncertainties
Table 8.15 Total systematic uncertainties affecting the absolute cross section distributions for each
measured discriminating observable: the maximum, the minimum and the average value are listed
Measured observable Δφ23 (%) Δrel pT,23 (%) ΔS (%)
Max. 18.1 22.6 35.3
Min. 14.0 12.7 13.2
Avg. 15.8 17.6 25.4
Table 8.16 Total systematic uncertainties affecting the normalized distributions for each measured
discriminating observable: the maximum, the minimum and the average value are listed
Measured observable Δφ23 (%) Δrel pT,23 (%) ΔS (%)
Max. 11.9 13.1 24.7
Min. 4.6 4.4 5.0
Avg. 7.9 9.8 15.3
For 2010 data, the official uncertainty on the luminosity is around 4% and this
value is taken into account for this analysis and included in the final combination
of different systematic uncertainties. The combination has been evaluated by sum-
ming in quadrature the single contributions, assuming absence of correlation among
the different sources. The final results of total systematic uncertainties and sepa-
rate contribution are shown in Fig. 8.3 for different discriminating observables. In
Tables 8.15 and 8.16, the total systematic uncertainties are listed for each measured
discriminating observables in absolute cross section scale and normalization scale.
In absolute scale of real cross section, the value of total systematic uncertainties are
about 15–25%. Jet energy scale, jet energy resolution, photon energy scale, photon
efficiency correction, model dependence and Monte-Carlo statistics have relatively
large systematic uncertainties while the others are small or even almost negligible.
And the largest uncertainties are caused by jet energy scale. In normalization scale of
shape, the total systematic uncertainties are smaller than those in the absolute scale
by 5–10%. The values are around 8–15%. The uncertainties of Monte-Carlo statistics
remain the same and become relatively important in this scale. And the constant 4%
of luminosity uncertainty is not taken into account since its effect will disappear after
normalization. ΔS observable has relatively larger total systematic uncertainties than
the other observables by 8–10% in absolute scale of real cross section and 5–6% in
normalization scale of shape. Especially, large systematic uncertainties exist in the
DPS sensitive region of ΔS.
Chapter 9
Results
After the data unfolding, the measurement of the total integrated cross section and the
normalized differential cross section as a function of different DPS-discriminating
observables, fully corrected for detector effects and selection efficiency, are shown
in Table 9.1, Figs. 9.1 and 9.2. The cross sections for a final state with one photon and
at least three jets span a phase space with a photon and a jet of PT > 75 GeV together
with 2 jets of PT > 20 GeV, where photon is measured within η < 1.4442 or 1.566
<η < 2.5 and jets are within η < 2.5. The hadronic energy surrounding the photon
is required to be at most 5GeVwithin ΔR < 0.4 at the stable-particle level. This
isolation sum is calculated as the sum of the E T ofall charged and neutral particles,
after removing the photon, within a cone of ΔR≡ (η − ηγ )2 + (φ − φ γ )2 , ηγ and
φ γ being the coordinates of the photon. The 5GeVthreshold at the generator level
is chosen to ensure greater than 95% efficiency for direct photons and minimize
dependence of the efficiency on the variation of underlying event models. For each
i-th bin, the absolute and normalized differential cross sections are defined as
dσi Ni
= ; (9.1)
dβ L · Ci · Δβi
1 dσi 1 Ni
= · ; (9.2)
σ dβ Ntotal L · Ci · Δβi
Table 9.1 Total integrated cross sections for measurement of the data and different MC predictions
in photon + 3 jets final state
Sample Tune Total Cross Section σ , pb
Data – 124.9 ± 8.9 (stat.)± 22.6 (syst.)
pythia 6 Z2 79.7 ± 3.2 (stat.)
pythia 8 4C 119.1 ± 0.7 (stat.)
MadGraph + pythia 6 Z2* 104.5 ± 1.8 (stat.)
Z2* (NO MPI) 101.6 ± 1.8 (stat.)
MadGraph + pythia 8 4C 95.2 ± 2.1 (stat.)
4C (NO MPI) 84.9 ± 2.0 (stat.)
sherpa CT10 190.2 ± 1.2 (stat.)
CT10 (NO MPI) 159.7 ± 1.3 (stat.)
pythia interfaces for hadronization, parton showering and also MPI simulation. The
MadGraph, interfaced with pythia 6 tune Z2*, gives a lightly lower value but still
in agreement with the measured cross section. The prediction of the MadGraph,
interfaced with pythia 8 tune 4C, is compatible with the measurement even if it
shows a lower value than the prediction of the MadGraph, interfaced with pythia 6
tune Z2*. sherpa tune CT10 overestimates the cross section value. Switching off
the MPI simulation for multijet-improved MC affect the prediction of cross section.
It reduces the prediction of cross section by 3% for the MadGraph, interfaced with
pythia 6 tune Z2*, and by 10% for the MadGraph, interfaced with pythia 8 tune
4C. The value of the cross section predicted by sherpa tune CT10 without MPI is
decreased by 16%, yet is still higher than the measured result.
The fractions of the different types of events, direct photon + 3 jets, fragmentation
photon + 3 jets and misidentified photon + 3 jets, contributing to each MC prediction
are shown in Table 9.2. Each fraction is calculated as the ratio of the integrated cross
section of the corresponding photon + 3 jets events to the total integrated cross section.
Different types of events are categorized according to the definitions described in
Chap. 5.
The main difference between pythia 6 and pythia 8 is the contribution of frag-
mentation photon + 3 jets events. MadGraph predicts a higher contribution of direct
photon + 3 jets events than pythia and sherpa by 40–60% and gives a lower con-
tribution of fragmentation photon + 3 jets events by 40–60%. sherpa shows similar
compositions of different photon + 3 jets events to pythia 6. The contributions of
misidentified photon + 3 jets events are similar (3–9%) among different MC simula-
tions. The theoretical fragmentation scale used in the calculation of MC simulation
will affect the consideration of a photon as a fragmentation photon or a higher-order
photon. Implementing different fragmentation scales then changes the SPS contri-
bution of fragmentation photon + 3 jets events.
9 Results 77
Table 9.2 Fractions of different types of events based on the MC estimation. Three kinds of
contributions are included in this table: direct photon + 3 jets events, fragmentation photon (frag.
photon) + 3 jets events, and misidentified photon (fake photon) + 3 jets events
Sample Tune Direct photon + 3 Frag. photon (%) Fake photon + 3
jets (%) + 3 jets jets (%)
pythia 6 Z2 42.6 48.2 9.2
pythia 8 4C 28.6 66.1 5.3
MadGraph + Z2* 92.1 3.5 4.4
pythia 6
Z2* (NO MPI) 85.0 9.1 5.9
MadGraph + 4C 83.8 13.3 2.9
pythia 8
4C (NO MPI) 83.6 14.0 2.4
sherpa CT10 52.0 41.9 6.0
CT10 (NO MPI) 49.2 45.5 5.3
The absolute real cross section as a function of Δφ23 is shown in Fig. 9.1 (left
column) and is well described by pythia 8 and MadGraph, with some slight fluctu-
ations. The DPS contribution is visible at Δφ ∼ π , as illustrated by the predictions of
MadGraph and sherpa without MPI. pythia 6 has lower prediction than pythia 8,
MadGraph. And sherpa shows significantly overestimated prediction w.r.t. all the
other MC predictions. In Fig. 9.1 (center column), the absolute real cross section as
a function of Δrel pT,23 is shown to be in agreement with pythia 8 and MadGraph
for Δrel pT,23 < 0.5 but visible differences arise in the high value region. pythia 6
has lower prediction than pythia 8, MadGraph. And sherpa shows significantly
overestimated prediction w.r.t. all the other MC predictions. The predictions of Mad-
Graph and sherpa without MPI exhibit the need of the missing DPS contribution
in this region.
Figure 9.2 (right column) shows the absolute real cross section as a function of
ΔS. pythia 8 and MadGraph give compatible predictions to the measurement.
pythia 6 still gives lower prediction than pythia 8, MadGraph in the whole region.
And sherpa still shows significantly overestimated prediction w.r.t. all the other MC
predictions in the whole region.
The normalized differential cross section as a function of Δφ23 is shown in Fig. 9.2
(left column) and is well described by all MC predictions, with some slight fluctua-
tions. The DPS contribution shows very little sensitivity at Δφ ∼ π , as illustrated by
the predictions of MadGraph and sherpa without MPI. In Fig. 9.2 (center column),
the normalized differential cross section as a function of Δrel pT,23 is shown to be in
agreement with all MC prediction for Δrel pT,23 > 0.4 but visible differences arise
78 9 Results
CMS Preliminary 2010, s = 7 TeV, L = 36.14 pb-1, pp → γ +3jets CMS Preliminary 2010, s = 7 TeV, L = 36.14 pb-1, pp → γ +3jets CMS Preliminary 2010, s = 7 TeV, L = 36.14 pb-1, pp → γ +3jets
σ [pb] 102
σ [pb]
σ [pb]
γ -jet: p > 75 GeV DATA γ -jet: p > 75 GeV DATA γ -jet: p > 75 GeV DATA
T T T
2jets: p > 20 GeV 2jets: p > 20 GeV 2jets: p > 20 GeV
T T T
γ : |η|<1.4442, 1.566<|η|< 2.5 Pythia8 4C γ : |η|<1.4442, 1.566<|η|< 2.5 Pythia8 4C 103 γ : |η|<1.4442, 1.566<|η|< 2.5 Pythia8 4C
jets: |η|<2.4 jets: |η|<2.4 jets: |η|<2.4
Pythia6 Z2 Pythia6 Z2 Pythia6 Z2
102
Stat. + syst. uncertainty Stat. + syst. uncertainty Stat. + syst. uncertainty
102
10
10
10
1
2 2.5 3.5
MC/DATA
MC/DATA
MC/DATA
1.8
1.6
1.4
2 ΔrelPT(jet2,jet3) 3
2.5
1.2 1.5 2
1
0.8 1 1.5
0.6 1
0.4 0.5
0.2 0.5
0 0 0
0 0.5 1 1.5 2 2.5 3 0 0.2 0.4 0.6 0.8 1 0 0.5 1 1.5 2 2.5 3
Δφ(jet2,jet3) (rad) ΔrelPT(jet2,jet3) ΔS (photonjet,dijet) (rad)
3
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j 3
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
10 10
σ dΔφ [1/rad]
σ dΔS [1/rad]
T,23
γ -j: p ,p > 75GeV DATA γ -j: p ,p > 75GeV DATA 104 γ -j: p ,p > 75GeV DATA
Tγ T jet1 Tγ T jet1 Tγ T jet1
MADGRAPH+PY6 Z2* σ dΔrelp MADGRAPH+PY6 Z2* MADGRAPH+PY6 Z2*
1 dσ
2j: p ,p > 20GeV MADGRAPH+PY6 Z2* no MPI 2j: p ,p > 20GeV MADGRAPH+PY6 Z2* no MPI 2j: p ,p > 20GeV MADGRAPH+PY6 Z2* no MPI
T jet2 T jet3 T jet2 T jet3 T jet2 T jet3
γ : |η|<1.4442, 1.566<|η|<2.5 MADGRAPH+PY8 4C γ : |η|<1.4442, 1.566<|η|<2.5 MADGRAPH+PY8 4C γ : |η|<1.4442, 1.566<|η|<2.5 MADGRAPH+PY8 4C
23
1 dσ
MADGRAPH+PY8 4C no MPI MADGRAPH+PY8 4C no MPI 3 MADGRAPH+PY8 4C no MPI
1 dσ
10
10 10
1 1 10-1
2.5 2.5 3
MC / DATA
MC / DATA
MC / DATA
MADGRAPH+PY6 Z2* MADGRAPH+PY6 Z2* 2.5
MADGRAPH+PY6 Z2*
2 2
2
1.5 1.5
1.5
1 1
1
0.5 0.5 0.5
2.5 2.5 3
MC / DATA
MC / DATA
MC / DATA
MADGRAPH+PY8 4C MADGRAPH+PY8 4C
T,23 2.5
MADGRAPH+PY8 4C
2 2
2
1.5 1.5
1.5
1 1
1
0.5 0.5 0.5
2.5 3
MC / DATA
MC / DATA
Fig. 9.1 Absolute real cross sections as a function of Δφ23 (left), Δrel pT,23 (center), and ΔS
(right) compared to different MC predictions. The comparisons are shown for two MC classes:
comparisons between the data and LO MC predictions, pythia 6 and pythia 8, (upper row),
comparisons between the data and multijet-improved MC predictions, MadGraph interfaced with
pythia 6 or pythia 8, and sherpa, (lower row). The lower panels show the ratio of the MC
prediction to the data. For the MC predictions, three kinds of contributions are included: direct
photon + 3 jets events, fragmentation photon + 3 jets events and misidentified photon + 3 jets
events. The band represents the total uncertainty of the data
in the low value region. For Δrel pT,23 < 0.2, pythia 6 has lower prediction than
pythia 8, MadGraph and sherpa. The predictions of MadGraph and sherpa
without MPI exhibit the need of the missing DPS contribution in this region.
Figure 9.2 (right column) shows the normalized differential cross section as a
function of ΔS. In the region of ΔS < 2.5, pythia 6 shows insufficient activity,
while the predictions of pythia 8, MadGraph and sherpa are compatible with the
measured distribution. Different pythia interfaces do not affect the predictions of
MadGraph. sherpa is slightly higher than the other predictions. Switching off the
MPI simulation for MadGraph and sherpa causes about 5–10% differences which
9 Results 79
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
1
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C 10 PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
1
jets: | η|<2.4 jets: | η|<2.4 jets: | η|<2.4
σ dΔφ [1/rad]
σ dΔS [1/rad]
T,23
Total Uncertainty Total Uncertainty Total Uncertainty
1
σ dΔrelp
1 dσ
23
1 dσ
1 dσ
-1
10
10-1
10-1
10-2
20
10-22
0.5 1 1.5 2 2.5 3 0 0.2 0.4 0.6 0.8 1 20 0.5 1 1.5 2 2.5 3
MC / DATA
MC / DATA
MC / DATA
1.8
1.6
Δφ (rad) 1.8
1.6 Δ p
rel 1.8
1.6 ΔS (rad)
23 T,23 1.4
1.4 1.4 1.2
1.2 1.2 1
1 1 0.8
0.8 0.8 0.6
0.4
0.6 0.6 0.2
0.4 0.4 0
0 0.5 1 1.5 2 2.5 3 0 0.2 0.4 0.6 0.8 1 0 0.5 1 1.5 2 2.5 3
Δφ23 (rad) ΔrelpT,23 ΔS (rad)
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j
1
γ -j: p ,p > 75GeV DATA γ -j: p ,p > 75GeV DATA γ -j: p ,p > 75GeV DATA
Tγ T jet1 Tγ T jet1 Tγ T jet1
MADGRAPH+PY6 Z2* MADGRAPH+PY6 Z2* MADGRAPH+PY6 Z2*
2j: p ,p > 20GeV MADGRAPH+PY6 Z2* no MPI 2j: p ,p > 20GeV MADGRAPH+PY6 Z2* no MPI 2j: p ,p > 20GeV MADGRAPH+PY6 Z2* no MPI
T jet2 T jet3 T jet2 T jet3 T jet2 T jet3
MADGRAPH+PY8 4C MADGRAPH+PY8 4C 10 MADGRAPH+PY8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
MADGRAPH+PY8 4C no MPI 1 MADGRAPH+PY8 4C no MPI MADGRAPH+PY8 4C no MPI
jets: |η|<2.4 SHERPA CT10 jets: |η|<2.4 SHERPA CT10 jets: |η|<2.4 SHERPA CT10
σ dΔφ [1/rad]
σ dΔS [1/rad]
T,23
1 dσ
1 dσ
10-1
10-1
10-1
10-2
10-2
2 2 2
MC / DATA
MC / DATA
MC / DATA
1.8 MADGRAPH+PY6 Z2* 1.8 MADGRAPH+PY6 Z2* 1.8 MADGRAPH+PY6 Z2*
1.6 1.6 1.6
1.4 1.4 1.4
1.2 1.2 1.2
1 1 1
0.8 0.8 0.8
0.6 0.6 0.6
0.4 0.4 0.4
20 0.5 1 1.5 2 2.5 3 20 0.2 0.4 0.6 0.8 rel 1 20 0.5 1 1.5 2 2.5 3
Δ p4C
MC / DATA
MC / DATA
MC / DATA
1.8 Δφ (rad)
MADGRAPH+PY8 4C 1.8 MADGRAPH+PY8 1.8 Δ S (rad)
MADGRAPH+PY8 4C
1.6 23 1.6 T,23 1.6
1.4 1.4 1.4
1.2 1.2 1.2
1 1 1
0.8 0.8 0.8
0.6 0.6 0.6
0.4 0.4 0.4
20 0.5 1 1.5 2 2.5 3 20 0.2 0.4 0.6 0.8 rel 1 20 0.5 1 1.5 2 2.5 3
SHERPAΔCT10
MC / DATA
MC / DATA
MC / DATA
1.8 Δφ CT10
SHERPA (rad) 1.8 pT,23 1.8 Δ SCT10
SHERPA (rad)
1.6 23 1.6 1.6
1.4 1.4 1.4
1.2 1.2 1.2
1 1 1
0.8 0.8 0.8
0.6 0.6 0.6
0.4 0.4 0.4
0 0.5 1 1.5 2 2.5 3 0 0.2 0.4 0.6 0.8 1 0 0.5 1 1.5 2 2.5 3
Δφ (rad)
23
Δ relpT,23 Δ S (rad)
Fig. 9.2 Normalized differential cross sections as a function of Δφ23 (left), Δrel pT,23 (center),
and ΔS (right) compared to different MC predictions. The comparisons are shown for two MC
classes: comparisons between the data and LO MC predictions, pythia 6 and pythia 8, (upper
row), comparisons between the data and multijet-improved MC predictions, MadGraph interfaced
with pythia 6 or pythia 8, and sherpa, (lower row). The lower panels show the ratio of the
MC prediction to the data. For the MC predictions, three kinds of contributions are included: direct
photon + 3 jets events, fragmentation photon + 3 jets events and misidentified photon + 3 jets events.
The band represents the total uncertainty of the data
is lower than the results observed in Reference [23, 24]. The reason the MPI causes
only small changes in ΔS is due to various background components as discussed in
Chap. 5. They reduce the possible room for the DPS signal contribution.
Chapter 10
Conclusion
Measurements of observables for the production of one photon and at least three
jets have been performed based on the data collected with the CMS experiment in
2010 with an integrated luminosity of 36 pb−1 . The cross section is measured to be
σ = 124.9 ± 8.9 (stat.) ± 22.6 (syst.) pb for a final state with a photon and a jet of pT
> 75 GeV together with 2 jets of pT > 20 GeV, where the photon is found within η <
1.4442 or 1.566 <η < 2.5 and the jets lie within η < 2.4. The normalized differential
cross sections, as a function of Δφ23 , Δrel pT,23 , and ΔS, are compared with various
theoretical predictions: pythia 6 tune Z2, pythia 8 tune 4C, MadGraph interfaced
with pythia 6 tune Z2*, MadGraph interfaced with pythia 8 tune 4C, and sherpa
tune CT10.
The DPS signal and various background components are studied with the help of
the Δφ23 , Δrel pT,23 , and ΔS observables. The DPS signal distribution shows different
features with respect to each background distribution and is distinguishable in the
DPS-sensitive region, at high Δφ23 values, low Δrel pT,23 values, and low to medium
ΔS values. Model dependence can be observed from comparisons of background
distributions produced by different MC simulations. It indicates that higher order
calculations including merging between matrix elements and parton showers affect
the contributions from SPS events and misidentified photon events. Implementing
different fragmentation scales in the calculation of MC simulation changes the SPS
contribution of fragmentation photon + 3 jets events.
From the comparison of the normalized differential cross sections as a function of
Δφ23 , Δrel pT,23 , and ΔS, the predictions of pythia 8, MadGraph interfaced with
pythia 6 or pythia 8, and sherpa with MPI describe the measured distributions
better than pythia 6 within the uncertainties except for some fluctuations in a few
bins. The MC predictions with and without a simulation of MPI do not differ sig-
nificantly even in the DPS-sensitive region due to various background components
which make the DPS signal contribution less pronounced. Hence, not only must
the predictions including MPI be validated with underlying event measurements, but
also the models used to extrapolate the hardest interaction need to be well understood
before the direct extraction of the underlying DPS fraction.
To sum up, the data are reasonably well described by MC predictions with or
without MPI and any conclusion on a DPS component is not possible within the
given precision of the current data and simulations. And this analysis demonstrates
the importance of understanding the background contribution which is crucial for
the extraction of the DPS fraction in proton-proton collisions at LHC energies.
Appendix A
K-Factor Correction for the LO
Cross-Section of P YTHIA, M AD G RAPH
and S HERPA MCs
In order to get the k-factor of isolated prompt photon production (direct photon +
fragmentation photon) used in this analysis, we compare the LO Pythia, MadGraph
and Sherpa predictions with the NLO pQCD predictions from JETPHOX
1.3.0 [68, 85, 86] using CT10 PDFs [65]. The k-factors of Pythia 6, Pythia 8,
MadGraph and Sherpa are calculated by the following formula with ET and η
dependences:
NLO(ET , η) of JETPHOX
K factor(ET , η) = (A.1)
LO(ET , η) of Monte − Carlo
For the photon-jet samples which contains direct photon production, the selected
photons have to be a stable particle and have at least one mother particle. Its mother
particle should be a photon, which is corresponding to the definition of direct photon
production. Besides, the selected photons are requested to have generator-level iso-
lation less than 5 GeV. The pseudorapidity range of selected photons is from −2.5
to 2.5 rad. For the QCD di-jet samples which contain fragmentation photon produc-
tion, the selected photons have to be a stable particle and have at least one mother
particle. Its mother particle should be a quark or gluon, which is corresponding to
the definition of fragmentation photon production. Besides, the selected photons
are requested to have generator-level isolation less than 5 GeV. The pseudorapidity
range of selected photons is from −2.5 to 2.5 rad. The isolation sum is calculated
as the sum of the ET ofall charged and neutral particles, after removing the photon,
within a cone of ΔR≡ (η − ηγ )2 + (φ − φ γ )2 , ηγ and φ γ being the coordinates of
the photon. The 5 GeV threshold at the generator level was chosen to ensure greater
than 95% efficiency for direct photons and minimise dependence of the efficiency
on the variation of underlying event models.
Figure A.1, Tables A.1 and A.2 show the details of k-factor values for Pythia 6,
Pythia 8 and MadGraph.
In order to take the assumption that the k-factors for photon + 1 jet and photon + 3
jets should be similar based upon Z + 1 jet and Z + 3 jets, we study the BLACKHAT’s
predictions of Z + N-Jet(s) events ratio [87, 88] with MSTW2008 NLO and LO PDF
104
|ηγ |<0.9 JETPHOX 0.9<|ηγ |<1.4442 JETPHOX
104
PYTHIA-6 PYTHIA-6
103
MADGRAPH+PYTHIA-6 MADGRAPH+PYTHIA-6
3
10
PYTHIA-8 PYTHIA-8
102
102
σ (pb)
σ (pb)
10
10
1
1
2 2
JETPHOX
JETPHOX
1.8 1.8
1.6 1.6
MC
MC
1.4 1.4
1.2 1.2
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
50 100 150 200 250 300 350 400 50 100 150 200 250 300 350 400
ET (photon) (GeV) ET (photon) (GeV)
104 104
1.566<|ηγ |<2.1 JETPHOX
2.2<|ηγ |<2.5 JETPHOX
PYTHIA-6 PYTHIA-6
3 103
10
MADGRAPH+PYTHIA-6 MADGRAPH+PYTHIA-6
10
σ (pb)
σ (pb)
10
1
10-1
2 2
JETPHOX
JETPHOX
1.8 1.8
1.6 1.6
MC
MC
1.4 1.4
1.2 1.2
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
50 100 150 200 250 300 350 400 50 100 150 200 250 300 350 400
ET (photon) (GeV) ET (photon) (GeV)
Fig. A.1 K-factors of NLO (JETPHOX) to LO (Monte-Carlo) ratios with respect to different ET
and η regions. a |η| < 0.9 b 0.9 < |η| < 1.4442 c 1.566 < |η| < 2.1 d 2.1 < |η| < 2.5
Table A.1 The k-factors of Pythia 6(PY-6) and Pythia 8(PY-8) are shown with ET and η
dependences
|η| < 0.9 0.9 < |η| < 1.4442 1.566 < |η| < 2.1 2.1 < |η| < 2.5
ET bin PY-6 PY-8 PY-6 PY-8 PY-6 PY-8 PY-6 PY-8
( GeV)
70−80 1.267 0.916 1.233 0.922 1.174 0.888 1.108 0.832
80−100 1.235 0.961 1.225 0.953 1.151 0.881 1.137 0.835
100−120 1.265 1.016 1.229 0.947 1.163 0.947 1.170 0.952
120−200 1.310 1.025 1.261 0.986 1.216 0.967 1.160 0.936
200−300 1.337 1.073 1.298 1.032 1.210 1.014 1.090 0.957
300−400 1.342 1.140 1.281 1.081 1.207 1.010 1.111 0.976
3
10
10
LO of Z+1jets, lo LO of Z+1jets, lo
1 102
-1
10
10-2
-3 1
10
10-4 10-1
-5
10
-6
10-2
10
10-7 -3
10 4-3
4 -2 -1 0 1 2 3
NLO_nlo
NLO_nlo
η (Z) (rad)
LO_lo
LO_lo
3.5 3.5
3 3
2.5 2.5
2 2
1.5 1.5
1 1
0.5 0.5
0 0
0 200 400 600 800 1000 1200 1400 -3 -2 -1 0 1 2 3
PT (Z) (GeV) η (Z) (rad)
3
10
10
LO of Z+1jets, lo LO of Z+1jets, lo
1 102
10-1
Events / 0.1 rad
Events / 20 GeV
NLO_nlo
9 η (jet1) (rad)
LO_lo
LO_lo
3.5
8
7 3
6 2.5
5 2
4 1.5
3 1
2
1 0.5
0 0
0 200 400 600 800 1000 1200 1400 -3 -2 -1 0 1 2 3
PT (jet1) (GeV) η (jet1) (rad)
Fig. A.2 LO and NLO distribution of pT and η of Z boson and leading jet for Z + 1 jet events
sets [89]. Table A.3 shows the selection criteria for Z + N-Jet(s) event. The related
results are shown from Figs. A.2, A.3 and A.4. Basically, an additional correction
factor of 0.73 for photon + 3 jets events is included through studying the BLACK-
HAT’s predictions of Z + N-Jet(s) events ratio. This additional factor is applied to
MadGraph samples due to that they contain 2 → 5 processes (γ + 4 jets).
The associated systematic uncertainty is studied by varying the factorization and
renormalization scale of Z + N-Jet(s) events. The original and recommended scale
is 0.5 Ht. By varying both of the the factorization and renormalization scale to 0.25
Ht and Ht, around 21% uncertainty is estimated and assigned for all MadGraph
samples in this analysis.
86 Appendix A: K-Factor Correction for the LO Cross-Section …
3
1 10
10-1
10-4
-5
10-2
10
-3
10
-6
10
10-4
10-7
-5
4
10 4
NLO_nlo
NLO_nlo
LO_lo
LO_lo
3.5 3.5
3 3
2.5 2.5
2 2
1.5 1.5
1 1
0.5 0.5
0 0
0 200 400 600 800 1000 1200 1400 -3 -2 -1 0 1 2 3
PT (Z) (GeV) η (Z) (rad)
3
10
1 LO of Z+3jets, lo LO of Z+3jets, lo
102
10-1
10
NLO of Z+3jets, nlo NLO of Z+3jets, nlo
10-2
1
-3
10 10-1
10-4 10-2
-5 -3
10 10
-6
10 10-4
-5
10-74 10 4
NLO_nlo
NLO_nlo
LO_lo
LO_lo
3.5 3.5
3 3
2.5 2.5
2 2
1.5 1.5
1 1
0.5 0.5
0 0
0 200 400 600 800 1000 1200 1400 -3 -2 -1 0 1 2 3
PT (jet1) (GeV) η (jet1) (rad)
Fig. A.3 LO and NLO distribution of pT and η of Z boson and first leading jet for Z + 3 jets events
4.5 4.5
3.5 3.5
2.5 2.5
NLO_nlo
NLO_nlo
LO_lo
LO_lo
2 2
1.5 1.5
1 1
0.5 0.5
0 0
1.4 1.4
K-factor Z3j
K-factor Z1j
K-factor Z3j
K-factor Z1j
1.2 1.2
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
0 200 400 600 800 1000 1200 1400 -3 -2 -1 0 1 2 3
PT (Z) (GeV) η (Z) (GeV)
10 4.5
NLO_nlo
LO_lo
LO_lo
6
2.5
5
2
4
1.5
3
1
2
1 0.5
0 0
1.4 1.4
K-factor Z3j
K-factor Z1j
K-factor Z3j
K-factor Z1j
1.2 1.2
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
0 200 400 600 800 1000 1200 1400 -3 -2 -1 0 1 2 3
PT (jet1) (GeV) η (jet1) (GeV)
Fig. A.4 k-factor comparisons between Z + 3 jets event and Z + 1 jet event in a pT spectra of
Z boson, b η distribution of Z boson, c pT spectra of first leading jet and d η distribution of first
leading jet
Appendix B
Study of Low Acceptance of Photon + 3 Jets
Events
In order to figure out why the acceptance of photon + 3 jets events is so low in this
analysis, we try to investigate the effect of different selection cuts on the accep-
tance. A comparison between original study and a special study by removing the
photon identification selection, described in Sect. 4.3, except for cuts of pT and η
region. As already mentioned before, photon isolation cuts are introduced to reduce
the background rate, caused by any possible fake photons or photons come from
the decay of π 0 and η. In this analysis, high photon purity is necessary due to the
multiple background can make the observation of DPS hard. It’s still worth to use
such selection criteria even we will loss much events which maybe are actually DPS
events. On the other hand, a selection cut on the 2nd vertex described before will
make the analysis focus on those events with only one vertex or events whose pile-
up is relatively soft. It’s side effect will make the acceptance of final photon + 3 jets
events becomes quite low. But this is unavoidable if we want to reduce the influence
of SPS events. In this appendix, we only perform the side effect caused by the photon
selection criteria.
It should be noted that we only studied difference of the original event yield
after selection by combining different processes of MCs together. That is, only the
weighting factors for combining MCs would be introduced to scale the distributions.
The results can be found in Fig. B.1. The acceptances would be much higher if we
remove the photon identification selection from our original selection criteria. But in
the same time, the background rate also increased and would be even much higher
than acceptance. This results are just consistent with our understanding described
before.
Events / 0.101
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
0 0.5 1 1.5 2 2.5 3 0 0.2 0.4 0.6 0.8 1
Δφ(jet2,jet3) (rad) ΔrelPT(Jet2,jet3)
0.8
0.6
0.4
0.2
0
0 0.5 1 1.5 2 2.5 3
Δ
S (photonjet,dijet) (rad)
Fig. B.1 Study of low acceptance: acceptance, purity, stability and background for a Δφ(jet2,jet3)
and b Δrel pT(jet2,jet3) and c ΔS
Appendix C
More DPS Discriminating Observables
To further differentiate between the SPS and DPS processes, more discriminating
observables with effective distinguishing power are used to the independence and
pairwise momentum balance of the two hard scatterings in DPS events. Basically,
all of observables included in this thesis are designed based on the study described
in [16, 17]. As already mentioned before, a final state arising from a single chain tends
to have a strongly correlated configuration in the azimuthal angle and pT -balance
distribution between the two sub-systems, while a DPS event has a preferred back-
to-back topology for the separated systems that are not correlated. The additional
four discriminating observables are listed below. Relevant detector-level results are
also included in this appendix.
• The difference in azimuthal angle between the photon and jet belonging to the
photon-jet pair,
Δφ(γ ,jet1) = φγ − φjet1 ; (C.1)
• a combination of the difference in azimuthal angle between the photon and jets of
the whole photon + 3 jets system,
1
Sφ = √ (Δφγ ,jet1 )2 + (Δφjet2,jet3 )2 ; (C.2)
2π 2
Different configurations for the photon and 3 jets in the final state translate in
different region for angular variables. Basically, these variables have good dis-
tinguishing power. Sφ accounts for the difference in azimuthal angle between the
selected photon and jets. Δφ(γ ,jet1) computes the difference of azimuthal angle in
photon-jet pair. SPS events lead to a broad distribution for these quantities, while
DPS events contribute most at values close to π , meaning the correlation between
the photon-jet pair and di-jet pair is very low. It should be noted that the variable
Sφ have been normalized by a factor of √ 1 2 .
2π
• the transverse momentum balance of the photon and jet belonging to the photon-jet
pair,
|pT γ + pT jet1 |
Δrel pT(γ ,jet1) = ; (C.3)
|pT γ | + |pT jet1 |
Δrel pT(γ ,jet1) is sensitive to the level of momentum balance in photon-jet pair.
A back-to-back for the two separate sub-systems contributes at low value of
Δrel pT(γ ,jet1) , while correlated pairs of jets result in a broader distribution over
the whole phase space. The balance level of transverse momentum is not only
analyzed in separate sub-systems but also in a compound scale.
• compound information of transverse momenta from all selected photon and 3 jets,
2 2
1 pT γ + pT jet1 pT jet2 + pT jet3
SpT =√ + ; (C.4)
2 |pT γ | + |pT jet1 | |pT jet2 | + |pT jet3 |
The distribution of SpT is similar to the one expected for the single variables
Δrel pT(γ ,jet1) and Δrel pT(jet2,jet3) for considering the transverse momentum balance
of correlated and uncorrelated systems but it gives a better overview of the process,
by taking into account information of the whole photon + 3 jets final state. The
better pT -balance the system have, the smaller value SpT will be. It should be noted
the variable SpT have been normalized by a factor of √12 .
Figures C.1, C.2, C.3 and C.4 represent the measurement of discriminating observ-
ables Δrel pT(γ ,jet1) and Δrel pT(jet2,jet3) , respectively, for both event distribution and
normalized shape comparison between data and different MC simulation. It can
be found the results are quite consistent to those shown in Sect. 6.1. In real scale,
MadGraph samples predict data quite well within statistic uncertainty. Pythia-
6 and Pythia-8 samples predicts a slightly smaller total number of events about
20–30% for both Pythia-6 and Pythia-8 but with larger statistic uncertainties,
especially for Pythia-8. There is an overall good agreement of shapes of all the
distribution for uncorrected data and different MCs in normalization scale.
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j+X CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j+X
105 γ -j: pT γ , pT jet1 > 75GeV γ -j: pT γ , pT jet1 > 75GeV
DATA DATA
2j: pT jet2, pT jet3 > 20GeV 10 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C PYTHIA8 4C
4 γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
10
Events / 0.315 rad
102
10-1
10
10-2
1
4 5
MC / DATA
3.5 4.5
MC / DATA
3 4
3.5
2.5 3
2 2.5
1.5 2
1 1.5
1
0.5 0.5
0 0
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
Δφ(γ jet1) (rad) Δφ(γ jet1) (rad)
Fig. C.1 The detector level comparison of Δφ(γ ,jet1) between uncorrected data and different com-
bined MC samples. a Event distribution. b Shape comparison
Appendix C: More DPS Discriminating Observables 91
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j+X CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j+X
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C 1 PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
103 jets: | η|<2.4 PYTHIA6 Z2 jets: | η|<2.4 PYTHIA6 Z2
Events / 0.101
MADGRAPH+PY6 Z2 MADGRAPH+PY6 Z2
dN/N
10-1
102
10-2
10
3.5 1 3.5
MC / DATA
3 3
MC / DATA
2.5 2.5
2 2
1.5 1.5
1 1
0.5 0.5
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
ΔPT(γ jet1) ΔPT(γ jet1)
Fig. C.2 The detector level comparison of Δrel pT(γ ,jet1) between uncorrected data and different
combined MC samples. a Event distribution. b Shape comparison
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j+X CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j+X
102
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
105
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C 10 PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
4
10
jets: | η|<2.4 PYTHIA6 Z2 jets: | η|<2.4 PYTHIA6 Z2
Events / 0.101
1
103
MADGRAPH+PY6 Z2 MADGRAPH+PY6 Z2
dN/N
10-1
102
10-2
10
1 10-3
10-1
10-4
3 4
3.5
MC / DATA
2.5
MC / DATA
3
2 2.5
1.5 2
1 1.5
1
0.5 0.5
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Sφ Sφ
Fig. C.3 The detector level comparison of Sφ between uncorrected data and different combined
MC samples. a Event distribution. b Shape comparison
CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j+X CMS Preliminary, s = 7 TeV, L = 36 pb-1, pp → γ +3j+X
104
γ -j: pT γ , pT jet1 > 75GeV DATA γ -j: pT γ , pT jet1 > 75GeV DATA
2j: pT jet2, pT jet3 > 20GeV 2j: pT jet2, pT jet3 > 20GeV
PYTHIA8 4C 1 PYTHIA8 4C
γ : |η|<1.4442, 1.566<|η|<2.5 γ : |η|<1.4442, 1.566<|η|<2.5
PYTHIA6 Z2 PYTHIA6 Z2
Events / 0.101
10-1
102
10-2
10
2 2.5
MC / DATA
MC / DATA
1.8
1.6 2
1.4
1.2 1.5
1
0.8 1
0.6
0.4 0.5
0.2
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
SPT SPT
Fig. C.4 The detector level comparison of SpT between uncorrected data and different combined
MC samples. a Event distribution. b Shape comparison
Curriculum Vitae
Work Experience
June 2015 to date Principal Engineer of Data Analysis at TSMC, Taiwan
Algorithm optimization, modeling and numerical simulation
Nov 2011–Jul 2015 Manager Assistant at Grid Center of NTU, Taipei, Taiwan
Assisted with the management of Unix/Linux server
Assisted with the establishment of cluster system
Monitored network and system operations
Designed and set up the website
Education
July 2015 Ph.D – Experimental High Energy Physics
National Taiwan University, Taipei, Taiwan
GPA: 4.0
Thesis: Study of Double Parton Scattering in Photon + 3 Jets
Final State in Proton-proton Collisions at 7 TeV at the CMS at the LHC.
Publications
Apr 2015 Study of Double Parton Scattering in Photon + 3 Jets
Final State in Proton-proton Collisions at 7TeV
The CMS Collaboration, CMS-PAS-FSQ-12-017
Jan 2014 Double Parton Scattering Study in Photon + 3 Jets Final State
at the CMS Experiment, the Physical Society of the Republic
of China (PSROC) Annual Meeting 2014, Taichung, Taiwan
Languages
Taiwan Mandarin: Mother tongue
English: Fluent
Japanese: Intermediate
French: Basic Knowledge
German: Basic Knowledge
1. CMS Collaboration,√ V. Khachatryan et al., “First Measurement of the Underlying Event Activ-
ity at the LHC with s = 0.9 TeV,” Eur. Phys. J. C 70 (2010) 555–572, arXiv:1006.2083.
2. CMS Collaboration,
√ S. Chatrchyan et al., “Measurement
√ of the Underlying Event Activity at
the LHC with s = 7 TeV and Comparison with s = 0.9 TeV,” JHEP 1109 (2011) 109,
arXiv:1107.0330.
3. CMS Collaboration, S. Chatrchyan et al., “Measurement
√ of the underlying event in the Drell-
Yan process in proton-proton collisions at s = 7 TeV,” Eur. Phys. J. C 72 (2012) 2080,
arXiv:1204.1411.
4. C. Goebel, F. Halzen, and D. Scott, “Double Drell-Yan Annihilations in Hadron Collisions:
Novel Tests of the Constituent Picture,” Phys. Rev. D 22 (1980) 2789.
5. N. Paver and D. Treleani, “Multi - Quark Scattering and Large pT Jet Production in Hadronic
Collisions,” Nuovo Cim. A 70 (1982) 215.
6. B. Humpert, “ARE THERE MULTI - QUARK INTERACTIONS?,” Phys. Lett. B 131 (1983)
461.
7. B. Humpert and R. Odorico, “Multiparton Scattering and QCD Radiation as Sources of Four
Jet Events,” Phys. Lett. B 154 (1985) 211.
8. T. Sjöstrand and M. van Zijl, “A Multiple Interaction Model for the Event Structure in Hadron
Collisions,” Phys. Rev. D 36 (1987) 2019.
9. G. Calucci and D. Treleani, “Double parton scatterings in high-energy hadronic collisions,”
Nucl. Phys. Proc. Suppl. 71 (1999) 392–399, arXiv:hep-ph/9711225.
10. M. Drees and T. Han, “Signals for Double Parton Scattering at the Fermilab Tevatron,” Phys.
Rev. Lett. 77 (Nov, 1996) 4142–4145.
11. P. Bartalini, E. L. Berger, B. Blok, G. Calucci, R. Corke, et al., “Multi-Parton Interactions at
the LHC,” ((2011)), no. ANL-HEP-PR-11-65. CMS-CR-2011-048. DESY 11-185. KA-TP-
32-2011. TTK-11-52, arXiv:1111.0469.
12. M. Mekhfi, “Multiparton Processes: an Application to Double Drell-Yan,” Phys. Rev. D 32
(1985) 2371.
13. L. Ametller, N. Paver, and D. Treleani, “Possible Signature of Multiple Parton Interactions in
Collider Four Jet Events,” Phys. Lett. B 169 (1986) 289.
14. R. Godbole, S. Gupta, and J. Lindfors, “Double Parton Scattering Contribution to W + Jets,”
Z. Phys. C 47 (1990) 69.
15. J. R. Gaunt, C.-H. Kom, A. Kulesza, and W. J. Stirling, “Same-sign W pair production as a
probe of double parton scattering at the LHC,” Eur. Phys. J. C 69 (2010) 53, arXiv:1003.3953.
√
16. CDF Collaboration, F. Abe et al., “Double parton scattering in pp collisions at s = 1.8 TeV,”
Phys. Rev. D 56 (Oct, 1997) 3811.
© Springer Nature Singapore Pte Ltd. 2017 97
Y.-H. Chang, Study of Double Parton Scattering in Photon + 3 Jets Final State,
Springer Theses, DOI 10.1007/978-981-10-3824-2
98 References
17. D0 Collaboration,
√ V. Abazov et al., “Double parton interactions in photon+3 jet events in pp
collisions s = 1.96 TeV,” Phys. Rev. D 81 (2010) 052012, arXiv:0912.5104.
18. D0 Collaboration, V. M. Abazov et al., “Azimuthal decorrelations√ and multiple parton interac-
tions in photon+2 jet and photon+3 jet events in pp̄ collisions at s = 1.96 TeV,” Phys. Rev.
D 83 (2011) 052008, arXiv:1101.1509.
19. D0 Collaboration, V. M. Abazov et al., √ “Double parton interactions in γ + 3 jet and γ +
b/cjet + 2 jet events in pp̄ collisions at s = 1.96 TeV,” Phys.Rev. D89 (2014), no. 7, 072006,
arXiv:1402.1550.
20. Axial Field Spectrometer
√ Collaboration, T. Akesson et al., “Double Parton Scattering in pp
Collisions at s = 63 GeV,” Z. Phys. C 34 (1987) 163.
21. UA2 Collaboration, J. Alitti et al., “A Study of multi-jet events at the CERN pp collider and a
search for double parton scattering,” Phys. Lett. B 268 (1991) 145.
22. CDF Collaboration, F. Abe et√ al., “Study of four jet events and evidence for double parton
interactions in pp̄ collisions at s = 1.8 TeV,” Phys. Rev. D 47 (1993) 4857.
23. CMS Collaboration,
√ S. Chatrchyan et al., “Measurement of four-jet production in proton-proton
collisions at s=7 TeV,” Phys. Rev. D 89 (2014) 092010, arXiv:1312.6440.
24. CMS Collaboration, S. Chatrchyan et√al., “Study of double parton scattering using W + 2-jet
events in proton-proton collisions at s = 7 TeV,” JHEP 03 (2014) 032, arXiv:1312.5729.
25. ATLAS Collaboration, G. Aad √ et al., “Measurement of hard double-parton interactions in
W (→ lν)+ 2 jet events at s=7 TeV with the ATLAS detector,” New J. Phys. 15 (2013)
033038, arXiv:1301.6872.
26. M. Glück, L. Gordon, E. Reya, and W. Vogelsang, “High P(T ) photon production at pp̄ collider,”
Phys. Rev. Lett. 73 (1994) 388.
27. M. Cacciari, G. P. Salam, and G. Soyez, “The Anti-k(t) jet clustering algorithm,” JHEP 04
(2008) 063, arXiv:0802.1189.
28. M. Cacciari, G. P. Salam, and G. Soyez, “FastJet User Manual,” Eur. Phys. J. C 72 (2012)
1896, arXiv:1111.6097.
29. O. S. Bruning, (Ed.) et al., “LHC design report. Vol. I: The LHC main ring,”. CERN-2004-
003-V-1.
30. CMS Collaboration, “CMS, the Compact Muon Solenoid: Technical proposal,”
CERN/LHCC 94-38 (1994).
31. ATLAS Collaboration, W. W. Armstrong et al., “ATLAS: Technical proposal for a general-
purpose pp experiment at the Large Hadron Collider at CERN,”. CERN-LHCC-94-43.
32. ALICE Collaboration, S. Ahmad et al., “ALICE: Technical proposal for a Large Ion collider
Experiment at the CERN LHC,”. CERN-LHCC-95-71.
33. LHCb Collaboration, S. Amato et al., “LHCb technical proposal,”. CERN-LHCC-98-04.
34. TOTEM Collaboration, V. Berardi et al., “TOTEM: Technical design report. Total cross
section, elastic scattering and diffraction dissociation at the Large Hadron Collider at CERN,”.
CERN-LHCC-2004-002.
35. LHCf Collaboration, O. Adriani et al., “Technnical proposal for the CERN LHCf experiment:
Measurement of photons and neutral pions in the very forward region of LHC,”. CERN-LHCC-
2005-032.
36. CMS Collaboration, S. Chatrchyan et al., “The CMS experiment at the CERN LHC,” JINST 3
(2008) S08004.
37. CMS Collaboration, CMS, “Jet Energy Corrections Determination at 7 TeV,” CMS Physics
Analysis Summary CMS-PAS-JME-10-010, CERN, 2010.
38. M. Cacciari and G. P. Salam, “Pileup subtraction using jet areas,” Phys. Lett. B 659 (2008) 119,
arXiv:0707.1378.
39. CMS Collaboration, “Particle-Flow Event Reconstruction in CMS and Performance for Jets,
Taus, and ETmiss ,” CMS Physics Analysis Summary CMS-PAS-PFT-09-001, CERN, 2009.
40. CMS Collaboration, “Commissioning of the Particle-flow Event Reconstruction with the first
LHC collisions recorded in the CMS detector,” CMS Physics Analysis Summary CMS-PAS-
PFT-10-001, CERN, 2010.
References 99
41. CMS Collaboration, V. Khachatryan et al., “Performance of photon √ reconstruction and iden-
tification with the CMS detector in proton-proton collisions at s = 8 TeV.” Submitted to
JINST, 2015.
42. CMS Collaboration, S. Chatrchyan et al., “Commissioning of the CMS High-Level Trigger
with Cosmic Rays,” JINST 5 (2010) T03005, arXiv:0911.4889.
43. CMS Collaboration, “Absolute Calibration of Luminosity Measurement at CMS: Summer
2011 Update,” CMS Physics Analysis Summary CMS-PAS-EWK-11-001 (2011).
44. S. van der Meer, “Calibration of the effective beam height in the ISR,” CERN-ISR-PO/68-31
(1968).
45. CMS Collaboration, D. Spiga, “CMS workload management,” Nucl. Phys. Proc. Suppl 172
(2007) 141–144.
46. T. Sjöstrand, S. Mrenna, and P. Skands, “A Brief Introduction to PYTHIA 8.1,” Comput. Phys.
Comm. 178 (2007) 852, arXiv:hep-ph/0710.3820.
47. T. Sjöstrand, S. Mrenna, and P. Skands, “PYTHIA 6.4 physics and manual,” JHEP 05 (2006)
026, arXiv:hep-ph/0603175.
48. F. Maltoni and T. Stelzer, “MadEvent: Automatic event generation with MadGraph,” JHEP 02
(2003) 027, arXiv:hep-ph/0208156.
49. A. Johan, H. Michel, M. Fabio, M. Olivier, and S. Tim, “MadGraph 5: Going Beyond,” JHEP 06
(2011) 128, arXiv:hep-ph/1106.0522.
50. J. Alwall, R. Frederix, S. Frixione, V. Hirschi, F. Maltoni, O. Mattelaer, H.-S. Shao, T. Stelzer,
P. Torrielli, and M. Zaro, “The automated computation of tree-level and next-to-leading order
differential cross sections, and their matching to parton shower simulations,” JHEP 07 (2014)
079, arXiv:1405.0301.
51. A. Buckley, H. Hoeth, H. Lacker, H. Schulz, and J. E. von Seggern, “Systematic event generator
tuning for the LHC,” Eur. Phys. J. C 65 (2010) 331–357, arXiv:0907.2973.
52. S. Agostinelli, J. Allison, K. Amako, J. Apostolakis, H. Araujo, P. Arce, M. Asai, D. Axen,
S. Banerjee, G. Barrand, F. Behner, L. Bellagamba, J. Boudreau, L. Broglia, A. Brunengo,
H. Burkhardt, S. Chauvie, J. Chuma, R. Chytracek, G. Cooperman, G. Cosmo, P. Degtyarenko,
A. Dell’Acqua, G. Depaola, D. Dietrich, R. Enami, A. Feliciello, C. Ferguson, H. Fesefeldt,
G. Folger, F. Foppiano, A. Forti, S. Garelli, S. Giani, R. Giannitrapani, D. Gibin, J. G. Cadenas,
I. González, G. G. Abril, G. Greeniaus, W. Greiner, V. Grichine, A. Grossheim, S. Guatelli,
P. Gumplinger, R. Hamatsu, K. Hashimoto, H. Hasui, A. Heikkinen, A. Howard, V. Ivanchenko,
A. Johnson, F. Jones, J. Kallenbach, N. Kanaya, M. Kawabata, Y. Kawabata, M. Kawaguti,
S. Kelner, P. Kent, A. Kimura, T. Kodama, R. Kokoulin, M. Kossov, H. Kurashige, E. Lamanna,
T. L. V. Lara, V. Lefebure, F. Lei, M. Liendl, W. Lockman, F. Longo, S. Magni, M. Maire,
E. Medernach, K. Minamimoto, P. M. de Freitas, Y. Morita, K. Murakami, M. Nagamatu,
R. Nartallo, P. Nieminen, T. Nishimura, K. Ohtsubo, M. Okamura, S. O’Neale, Y. Oohata,
K. Paech, J. Perl, A. Pfeiffer, M. Pia, F. Ranjard, A. Rybin, S. Sadilov, E. D. Salvo, G. Santin,
T. Sasaki, N. Savvas, Y. Sawada, S. Scherer, S. Sei, V. Sirotenko, D. Smith, N. Starkov,
H. Stoecker, J. Sulkimo, M. Takahata, S. Tanaka, E. Tcherniaev, E. S. Tehrani, M. Tropeano,
P. Truscott, H. Uno, L. Urban, P. Urban, M. Verderi, A. Walkden, W. Wander, H. Weber,
J. Wellisch, T. Wenaus, D. Williams, D. Wright, T. Yamada, H. Yoshida, and D. Zschiesche,
“Geant4–a simulation toolkit,” NIM A 506 (2003) 250.
53. J. Allison, K. Amako, J. Apostolakis, H. Araujo, P. Dubois, M. Asai, G. Barrand, R. Capra,
S. Chauvie, R. Chytracek, G. A. P. Cirrone, G. Cooperman, G. Cosmo, G. Cuttone, G. G.
Daquino, M. Donszelmann, M. Dressel, G. Folger, F. Foppiano, J. Generowicz, V. Grichine,
S. Guatelli, P. Gumplinger, A. Heikkinen, I. Hrivnacova, A. Howard, S. Incerti, V. Ivanchenko,
T. Johnson, F. Jones, T. Koi, R. Kokoulin, M. Kossov, H. Kurashige, V. Lara, S. Larsson,
F. Lei, O. Link, F. Longo, M. Maire, A. Mantero, B. Mascialino, I. McLaren, P. Lorenzo,
K. Minamimoto, K. Murakami, P. Nieminen, L. Pandola, S. Parlati, L. Peralta, J. Perl, A. Pfeif-
fer, M. Pia, A. Ribon, P. Rodrigues, G. Russo, S. Sadilov, G. Santin, T. Sasaki, D. Smith,
N. Starkov, S. Tanaka, E. Tcherniaev, B. Tome, A. Trindade, P. Truscott, L. Urban, M. Verderi,
A. Walkden, J. P. Wellisch, D. Williams, D. Wright, and H. Yoshida, “Geant4 Developments
and Applications,” IEEE Trans. Nucl. Sci. 53 (2006) 270.
100 References
54. R. Field, “Early LHC Underlying Event Data - Findings and Surprises,” (2010)
arXiv:1010.3558.
55. J. Pumplin, D. Stump, J. Huston, H. Lai, P. M. Nadolsky, et al., “New generation of
parton distributions with uncertainties from global QCD analysis,” JHEP 07 (2002) 012,
arXiv:hep-ph/0201195.
56. P. Z. Skands and D. Wicke, “Non-perturbative QCD effects and the top mass at the Tevatron,”
Eur. Phys. J. C 52 (2007) 133–140, arXiv:hep-ph/0703081.
57. R. Corke and T. Sjöstrand, “Interleaved Parton Showers and Tuning Prospects,” JHEP 03
(2011) 032, arXiv:1011.1759.
58. R. Corke and T. Sjöstrand, “Multiparton Interactions and Rescattering,” JHEP 01 (2010) 035,
arXiv:0911.1909.
59. CMS Collaboration, √ S. Chatrchyan et al., “Study of the underlying event at forward rapidity
in pp collisions at s = 0.9, 2.76, and 7 TeV,” JHEP 04 (2013) 072, arXiv:1302.2394.
60. T. Gleisberg, S. Höche, F. Krauss, M. Schonherr, S. Schumann, et al., “Event generation with
SHERPA 1.1,” JHEP 02 (2009) 007, arXiv:0811.4622.
61. S. Schumann and F. Krauss, “A Parton shower algorithm based on Catani-Seymour dipole
factorisation,” JHEP 03 (2008) 038, arXiv:0709.1027.
62. T. Gleisberg and S. Höche, “Comix, a new matrix element generator,” JHEP 12 (2008) 039,
arXiv:0808.3674.
63. S. Höche, F. Krauss, S. Schumann, and F. Siegert, “QCD matrix elements and truncated show-
ers,” JHEP 05 (2009) 053, arXiv:0903.1219.
64. M. Schönherr and F. Krauss, “Soft Photon Radiation in Particle Decays in SHERPA,” JHEP 12
(2008) 018, arXiv:0810.5071.
65. H.-L. Lai, M. Guzzi, J. Huston, Z. Li, P. M. Nadolsky, et al., “New parton distributions for
collider physics,” Phys. Rev. D 82 (2010) 074024, arXiv:1007.2241.
√
66. CMS Collaboration, “Photon reconstruction and identification at s = 7TeV,” CMS Physics
Analysis Summary CMS-PAS-EGM-10-005, CERN, 2010.
67. CMS Collaboration, “Electromagnetic calorimeter commissioning and first results with 7 TeV
data,” CMS Note 2010/012, CERN, 2010.
68. CMS Collaboration, “Measurement of the Differential Cross Section for Isolated Prompt
Photon Production in pp Collisions at 7 TeV,” Phys. Rev. D 84 (2011) 052011. 45 p,
arXiv:hep-ex/1108.2044.
69. CMS Collaboration, V. Khachatryan et al.,√“Measurement of the Isolated Prompt Photon
Production Cross Section in pp Collisions at s = 7 TeV,” Phys. Rev. Lett. 106 (2011) 082001,
arXiv:1012.0799.
70. CMS Collaboration, S. Chatrchyan et al., “Measurement of the Differential Cross Section for
Isolated Prompt Photon Production in pp Collisions at 7 TeV,” Phys. Rev. D 84 (2011) 052011,
arXiv:1108.2044.
71. CMS Collaboration, “Electromagnetic calorimeter calibration with 7 TeV data,” CMS Physics
Analysis Summary CMS-PAS-EGM-10-003, CERN, Geneva, 2010.
72. CMS Collaboration, S. Chatrchyan√et al., “Measurement of the Inclusive W and Z Production
Cross Sections in pp Collisions at s = 7 TeV,” JHEP 10 (2011) 132, arXiv:1107.4789.
73. CMS Collaboration, “ECAL 2010 performance results,” Tech. Rep. CMS-DP-2011-008,
CERN, Geneva, 2011.
74. CMS Collaboration, “Commissioning of the Particle-Flow reconstruction in Minimum-Bias
and Jet Events from pp Collisions at 7 TeV,” Tech. Rep. CMS-PAS-PFT-10-002, CERN, Geneva,
2010.
75. CMS Collaboration, “Jet Energy Corrections determination at 7 TeV,” Tech. Rep. CMS-PAS-
JME-10-010, CERN, Geneva, 2010.
76. CMS Collaboration,
√ S. Chatrchyan et al., “Measurement of the Inclusive Jet Cross Section in
pp Collisions at s = 7 TeV,” Phys. Rev. Lett. 107 (2011) 132001, arXiv:1106.0208.
77. G. D’Agostini, “A Multidimensional unfolding method based on Bayes’ theorem,” Nucl.
Instrum. Meth. A 362 (1995) 487.
78. T. Adye, “Unfolding algorithms and tests using RooUnfold,” (2011) arXiv:1105.1160.
References 101
79. A. Höcker and V. Kartvelishvili, “SVD approach to data unfolding,” Nucl. Instrum. Meth. A 372
(1996) 469, arXiv:hep-ph/9509307.
80. CMS Collaboration, S. Chatrchyan et al., “Measurement of √ the triple-differential cross section
for photon+jets production in proton-proton collisions at s=7 TeV,” JHEP 1406 (2014) 009,
arXiv:1311.6141.
81. CMS Collaboration, S. Chatrchyan et al., “Energy √ calibration and resolution of the CMS
electromagnetic calorimeter in pp collisions at s = 7 TeV,” JINST 8 (2013) P09009,
arXiv:1306.2016.
82. CMS Collaboration, S. Chatrchyan et al., “Determination of jet energy calibration and trans-
verse momentum resolution in CMS,” JINST 6 (2011) P11002, arXiv:1107.4277.
83. CMS Collaboration, “CMS twiki page: Pile-up Systematic Errors,” tech. rep., CERN, 1900.
Geneva, 2012.
84. CMS Collaboration, S. Chatrchyan et al., “Description and performance of track and primary-
vertex reconstruction with the CMS tracker,” JINST 9 (2014), no. 10, P10009, arXiv:1405.6569.
85. S. Catani, M. Fontannaz, J. Guillet, and E. Pilon, “Cross-section of isolated prompt photons in
hadron hadron collisions,” JHEP 0205 (2002) 028, arXiv:hep-ph/0204023.
86. P. Aurenche, M. Fontannaz, J.-P. Guillet, E. Pilon, and M. Werlen, “A New critical study of pho-
ton production in hadronic collisions,” Phys.Rev. D73 (2006) 094007, arXiv:hep-ph/0602133.
87. C. Berger, Z. Bern, L. J. Dixon, F. Febres Cordero, D. Forde, et al., “Next-to-Leading Order
Jet Physics with BlackHat,” PoS RADCOR2009 (2010) 002, arXiv:0912.4927.
88. Z. Bern, L. Dixon, F. F. Cordero, S. Hoeche, H. Ita, et al., “Ntuples for NLO Events at Hadron
Colliders,” arXiv:1310.7439.
89. A. Martin, W. Stirling, R. Thorne, and G. Watt, “Parton distributions for the LHC,”
Eur.Phys.J. C63 (2009) 189–285, arXiv:0901.0002.