complementarity of symmetry tests at the energy and

154
Complementarity of Symmetry Tests at the Energy and Intensity Frontiers by Tao Peng A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy (Physics) at the University of Wisconsin–Madison 2017 Date of final oral examination: 5/4/2017 The dissertation is approved by the following members of the Final Oral Committee: Akif Baha Balantekin, Professor, Physics Michael Ramsey-Musolf, Professor, Physics Akikazu Hashimoto, Professor, Physics Matthew Herndon, Professor, Physics Wesley Smith, Professor, Physics Pupa Gilbert, Professor, Geoscience, Chemistry, Materials Science, Physics

Upload: others

Post on 10-Jun-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Complementarity of Symmetry Tests at the Energy and

Complementarity of Symmetry Tests at

the Energy and Intensity Frontiers

by

Tao Peng

A dissertation submitted in partial fulfillment of

the requirements for the degree of

Doctor of Philosophy

(Physics)

at the

University of Wisconsin–Madison

2017

Date of final oral examination: 5/4/2017

The dissertation is approved by the following members of the Final Oral Committee:

Akif Baha Balantekin, Professor, Physics

Michael Ramsey-Musolf, Professor, Physics

Akikazu Hashimoto, Professor, Physics

Matthew Herndon, Professor, Physics

Wesley Smith, Professor, Physics

Pupa Gilbert, Professor, Geoscience, Chemistry, Materials Science, Physics

Page 2: Complementarity of Symmetry Tests at the Energy and

i

Complementarity of Symmetry Tests at

the Energy and Intensity Frontiers

Tao Peng

Under the supervision of Professors Michael Ramsey-Musolf and Akif Baha

Balantekin

At the University of Wisconsin–Madison

Abstract

We studied several symmetries and interactions beyond the Standard Model and

their phenomenology in both high energy colliders and low energy experiments. The

lepton number conservation is not a fundamental symmetry in Standard Model (SM).

The nature of the neutrino depends on whether or not lepton number is violated. Lep-

togenesis also requires lepton number violation (LNV). So we want to know whether

lepton number is a good symmetry or not, and we want to compare the sensitivity of

high energy collider and low energy neutrinoless double-β decay (0νββ) experiments.

To do this, We included the QCD running effects, the background analysis, and the

long-distance contributions to nuclear matrix elements. Our result shows that the

reach of future tonne-scale 0νββ decay experiments generally exceeds the reach of the

14 TeV LHC for a class of simplified models. For a range of heavy particle masses at

the TeV scale, the high luminosity 14 TeV LHC and tonne-scale 0νββ decay experi-

ments may provide complementary probles. The 100 TeV collider with a luminosity of

Page 3: Complementarity of Symmetry Tests at the Energy and

ii

30 ab−1 exceeds the reach of the tonne-scale 0νββ experiments for most of the range

of the heavy particle masses at the TeV scale.

We considered a non-Abelian kinetic mixing between the Standard Model gauge

bosons and a U(1)′ gauge group dark photon, with the existence of an SU(2)L scalar

triplet. The coupling constant between the dark photon and the SM gauge bosons ǫ

is determined by the triplet vacuum expectation value (vev), the scale of the effective

theory Λ, and the effective operator Wiloson coeffcient. The triplet vev is constrained

to <∼ 4 GeV. By taking the effective operator Wiloson coeffcient to be O(1) and Λ

> 1 TeV, we will have a small value of ǫ which is consistent with the experimetal

constraint. We outlined the possible LHC signatures and recasted the current ATLAS

dark photon experimental results into our non-Abelian mixing scenario.

We analyzed the QCD corrections to dark matter (DM) interactions with SM

quarks and gluons. Because we like to know the new physics at high scale and the

effect of the direct detection of DM at low scale, we studied the QCD running for a list

of dark matter effective operators. These corrections are important in precision DM

physics. Currently little is known about the short-distance physics of DM. We find

that the short-distance QCD corrections generate a finite matching correction when

integrating out the electroweak gauge bosons.

The high precision measurements of electroweak precision observables can pro-

vide crucial input in the search for supersymmetry (SUSY) and play an important

role in testing the universality of the SM charged current interaction. We studied the

SUSY corrections to such observables ∆CKM and ∆e/µ, with the experimental con-

straints on the parameter space. Their corrections are generally of order O(10−4).

Future experiments need to reach this precision to search for SUSY using these ob-

Page 4: Complementarity of Symmetry Tests at the Energy and

iii

servables.

Page 5: Complementarity of Symmetry Tests at the Energy and

iv

To my family

“It is a miracle that curiosity survives formal education.”

Albert Einstein

Page 6: Complementarity of Symmetry Tests at the Energy and

v

Acknowledgements

First and foremost, I would like to thank my advisors Professor Michael Ramsey-

Musolf and Professor Baha Balantekin for guiding me in the physics study and re-

search, and in the all the administrative paperwork. Professor Ramsey-Musolf taught

me courses in Particle Physics and Collider Physics Phenomenology, in which I learned

a lot of important theories and skills in theoretical physics and phenomenology. In

research, Professor Ramsey-Musolf also provided very good and interesting research

project topics. He guided me, worked in parallel with me, and offered much help in

research. Without his help, I could never be able to finish the research. Professor

Ramsey-Musolf also provided me very good opportunities to visit UMass and attend

workshops and conferences to communicate with and learn from colleagues.

I thank Professor Baha Balantekin very much for willing to be my advisor in

UW. He helped me a lot in my paperwork, which made it convenient for me to work

remotely with Professor Ramsey-Musolf.

In the research projects, I thank my collaborators Michael Ramsey-Musolf, Peter

Winslow, Grigory Ovanesyan, Wei Chao, Carlos Argüelles, Xiao-Gang He, and Haolin

Li. Without their work and help, I could not finish the projects efficiently. I especially

would like to thank Michael Ramsey-Musolf, Peter Winslow, Grigory Ovanesyan and

Wei Chao, because we did most of the research work in parallel but independently, so

that we can cross check and make sure that our results are all correct and reliable.

I am also grateful to all my graduation committee members: Professors Michael

Page 7: Complementarity of Symmetry Tests at the Energy and

vi

Ramsey-Musolf, Baha Balantekin, Akikazu Hashimoto, Matthew Herndon, Wesley

Smith. For my preliminary examination, I thank Professors Michael Ramsey-Musolf,

Baha Balantekin, Akikazu Hashimoto, and Sau Lan Wu for being on my committee

and for their feedbacks.

I would also like to thank Professor Akikazu Hashimoto for teaching me Ad-

vanced Quantum Mechanics for two semesters, and I would like to thank Professor

Ludwig Bruch for teaching me Theoretical Physics Dynamics.

I sincerely thank everyone in the High Energy Phenomenology and Theory group

in the University of Wisconsin-Madison for creating a friendly and nice environment.

I also thank everyone in my research group for their insights and ideas in group

meetings: Michael Ramsey-Musolf, Peter Winslow, Grigory Ovanesyan, Wei Chao,

Huaike Guo, Chien Yeah Seng, Haolin Li, Jiang-Hao Yu, Hiren Patel, Kaori Fuyuto,

Simon Shen, Satoru Inoue, Yong Du, Mario Pitschmann, Martin Gonzalez-Alonso,

and Sky Bauman.

Finally, I would like to thank my family for their support, understanding, and

encouragement.

Page 8: Complementarity of Symmetry Tests at the Energy and

vii

Contents

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i

1 Introduction 1

1.1 Particle physics standard model and beyond . . . . . . . . . . . . . . . 2

1.2 Lepton number and its conservation . . . . . . . . . . . . . . . . . . . . 3

1.3 Lepton number violation . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.4 Neutrinoless double beta decay . . . . . . . . . . . . . . . . . . . . . . 5

1.5 Experimental tests and searches for LNV . . . . . . . . . . . . . . . . . 6

1.6 Future colliders at 100 TeV . . . . . . . . . . . . . . . . . . . . . . . . . 10

1.7 The triplet scalar model and LHC studies . . . . . . . . . . . . . . . . 11

1.8 The motivation of non-Abelian kinetic mixing . . . . . . . . . . . . . . 13

1.9 Dark matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

1.10 Motivations for studying dark matter operator mixing and running . . 16

1.11 Supersymmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

1.12 Charged current universality and the weak charge . . . . . . . . . . . . 19

2 Lepton number violation collider study at 14 TeV and comparison

with 0νββ decay 23

2.1 The models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

Page 9: Complementarity of Symmetry Tests at the Energy and

viii

2.2 The effect of LNV operator running . . . . . . . . . . . . . . . . . . . . 27

2.3 Constraint from neutrinoless double beta decay . . . . . . . . . . . . . 33

2.4 Collider studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

3 Lepton number violation collider study at 100 TeV 52

3.1 LHC signal and backgrounds . . . . . . . . . . . . . . . . . . . . . . . . 53

3.2 Cut analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

3.3 LHC results and comparison to 0νββ decay experiments . . . . . . . . 56

3.4 Comparison to machine learning results . . . . . . . . . . . . . . . . . . 60

4 LHC Signatures of Non-Abelian Kinetic Mixing 67

4.1 The model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

4.2 Collider phenomenology . . . . . . . . . . . . . . . . . . . . . . . . . . 73

4.3 Triplet-like scalar decay branching ratios . . . . . . . . . . . . . . . . . 79

4.4 ATLAS recast . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

4.5 More on the UV completion . . . . . . . . . . . . . . . . . . . . . . . . 85

5 QCD corrections for dark matter effective interactions 89

5.1 Purpose of this study . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

5.2 The effective operators . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

5.3 Loop corrections and the anomalous dimension matrix . . . . . . . . . 94

5.4 Phenomenological effects of QCD corrections . . . . . . . . . . . . . . . 98

5.5 Box graph corrections and factorizability . . . . . . . . . . . . . . . . . 103

5.5.1 Fermion dark matter . . . . . . . . . . . . . . . . . . . . . . . . 105

5.5.1.1 Dirac dark matter . . . . . . . . . . . . . . . . . . . . 107

Page 10: Complementarity of Symmetry Tests at the Energy and

ix

5.5.1.2 Majorana dark matter . . . . . . . . . . . . . . . . . . 113

5.5.1.3 Inelastic dark matter . . . . . . . . . . . . . . . . . . . 114

5.5.2 Scalar dark matter . . . . . . . . . . . . . . . . . . . . . . . . . 116

6 SUSY radiative corrections 118

6.1 Parameter scans in MSSM . . . . . . . . . . . . . . . . . . . . . . . . . 119

6.2 Correction results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

7 Summary 127

Bibliography 130

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

Page 11: Complementarity of Symmetry Tests at the Energy and

x

List of Tables

2.1 Cut-flow part 1/2. Designed for optimizing signal relative to back-

grounds. The backgrounds include diboson and charge-flip. For the cut

flow of jet-fake backgrounds, see Table ??. . . . . . . . . . . . . . . . . 45

2.2 Cut-flow part 2/2. Designed for optimizing signal relative to back-

grounds. The backgrounds include jet-fake. For the cut flow of diboson

and charge-flip backgrounds, see Table ??. . . . . . . . . . . . . . . . . 47

3.1 Cut-flow part 1/2 for 100 TeV. Designed for optimizing signal relative

to backgrounds. The backgrounds include diboson and charge-flip. For

the cut flow of jet-fake backgrounds, see Table ??. . . . . . . . . . . . . 56

3.2 Cut-flow part 2/2 for 100 TeV. Designed for optimizing signal relative

to backgrounds. The backgrounds include jet-fake. For the cut-flow of

diboson and charge-flip backgrounds, see Table ??. . . . . . . . . . . . 59

5.1 Operator basis and their corresponding dimensions. . . . . . . . . . . . 92

Page 12: Complementarity of Symmetry Tests at the Energy and

xi

List of Figures

1.1 Graphic illustration of 0νββ decay. . . . . . . . . . . . . . . . . . . . . 7

1.2 The light Majorana neutrino model that makes 0νββ decay . . . . . . . 7

1.3 The left-right symmetric model that makes 0νββ decay . . . . . . . . . 7

1.4 The R Parity Violation Supersymmetry model that makes 0νββ decay 8

2.1 The lepton number violating process extracted from neutrinoless double

beta decay. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2.2 The lepton number violating process, with fictitious intermediate par-

ticles S+ and F 0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2.3 Cross section of pp → e−e− + jets vs. mass of S+ and F 0, taking the

couplings C1 = C2 = 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

2.4 The one-loop correction to the dd→ uuee LNV process. There are two

more diagrams symmetric to the two diagrams on the lower part. . . . 29

2.5 The running of the Wilson coefficients. Assuming that at the TeV scale,

C1 = 1, C2 = C3 = C4 = 0. . . . . . . . . . . . . . . . . . . . . . . . . . 35

2.6 The process of the two pion-two electron operator in Eq. ??. . . . . . . 35

2.7 The charge-flip Z/γ∗ → e+e− process. . . . . . . . . . . . . . . . . . . . 40

2.8 The charge-flip tt process. The b’s are not tagged. . . . . . . . . . . . . 40

Page 13: Complementarity of Symmetry Tests at the Energy and

xii

2.9 The HT distribution for signal and backgrounds at 14 TeV. . . . . . . . 46

2.10 The ml1l2 distribution for signal and backgrounds at 14 TeV. . . . . . . 46

2.11 The MET distribution for signal and backgrounds at 14 TeV. . . . . . . 48

2.12 Significance of the e−e− + dijet signal as a function of integrated lumi-

nosity, assuming that the C1/Λ5 is consistent with the GERDA 0νββ

half-life limit. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

2.13 Current and future exclusion reach of 0νββ decay and LHC searches

for the TeV LNV interaction as a function of the coupling geff and mass

scale Λ. The coupling is defined as geff = C1/41 , where the C1 is the

coupling in Eq. ??. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

2.14 Current and future discovery reach of 0νββ decay and LHC searches

for the TeV LNV interaction as a function of the coupling geff and mass

scale Λ. The coupling is defined as geff = C1/41 , where the C1 is the

coupling in Eq. ??. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

3.1 The HT distribution for signal and backgrounds at 100 TeV. . . . . . . 57

3.2 The ml1l2 distribution for signal and backgrounds at 100 TeV. . . . . . 57

3.3 The MET distribution for signal and backgrounds at 100 TeV. . . . . . 58

3.4 The leading lepton pT distribution for signal and backgrounds at 100

TeV. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

Page 14: Complementarity of Symmetry Tests at the Energy and

xiii

3.5 Current and future exclusion reach of 0νββ decay and 100 TeV LHC

searches for LNV interaction as function of the coupling geff and mass

scale Λ. The requirement for exclusion is S/√S + B ≥ 2. The cou-

pling is defined as geff = C1/41 , where the C1 is the coupling in Eq. ??.

The blue shaded areas are for the uncertainty of M0, whose borders

correspond to M0 = −1.0 and M0 = −1.99. . . . . . . . . . . . . . . . . 61

3.6 Current and future discovery reach of 0νββ decay and 100 TeV LHC

searches for LNV interaction as function of the coupling geff and mass

scale Λ. The requirement for discovery is S/√S + B ≥ 5. The cou-

pling is defined as geff = C1/41 , where the C1 is the coupling in Eq. ??.

The blue shaded areas are for the uncertainty of M0, whose borders

correspond to M0 = −1.0 and M0 = −1.99. . . . . . . . . . . . . . . . . 62

3.7 Comparison of the discovery reaches using cut-based analysis (left) and

machine learning analysis (right). The random forest method is used in

the machine learning analysis. The machine learning analysis was done

by Peter Winslow. In the right figure, the notation M means Λ and yeff

means geff . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

4.1 Feynman diagrams that may generate the non-abelian mixing effective

operator O(5)WX . The intermediate particles in the loops are (a) fermions,

(b) scalars, or (c) other sources from non-perturbative dynamics. . . . . 76

Page 15: Complementarity of Symmetry Tests at the Energy and

xiv

4.2 The Feynman diagrams for LHC production and the subsequent decay

of the particles in the non-Abelian mixing model with the triplet scalars.

Diagrams (a) and (b) are the scalar pair productions, followed by the

scalar decays mediated by the non-Abelian mixing operator O(5)WX . Di-

agrams (c) and (d) are production and decays of H and X. In all four

diagrams, the incoming vector bosons are all virtual. . . . . . . . . . . 76

4.3 LHC production cross sections for pp → V → φφ and pp → V → Xφ

at√s = 8 TeV, where φ = H+, H2. The mφ = 130 GeV, and mX = 0.4

GeV. For the processes with final states of a single charged scalar and

one neutral boson, we summed the cross sections for both charges, for

example: σ(H+H2) + σ(H−H2). . . . . . . . . . . . . . . . . . . . . . . 77

4.4 LHC production cross sections for pp → V → φφ and pp → V → Xφ

at√s = 8 TeV, where φ = H+, H2. The mφ = 300 GeV, and mX = 0.4

GeV. For the processes with final states of a single charged scalar and

one neutral boson, we summed the cross sections for both charges, for

example: σ(H+H2) + σ(H−H2). . . . . . . . . . . . . . . . . . . . . . . 78

4.5 Branching ratios forH+ decays as a function of ǫ (upper horizontal axis)

and β/Λ (bottom horizontal axis), for the triplet vev vΣ = 1 GeV. The

dark photon mass is chosen as mX = 0.4 GeV. The top plot corresponds

to mH+ = 130 GeV, and the bottom plot corresponds to mH+ = 300

GeV. The solid black line is the branching ratio for H+ → W+X.

Branching ratios for other final states are as indicated by other colors. . 80

Page 16: Complementarity of Symmetry Tests at the Energy and

xv

4.6 Branching ratios for H+ decays as a function of ǫ (upper horizontal

axis) and β/Λ (bottom horizontal axis), for the triplet vev vΣ = 10−3

GeV. The dark photon mass is chosen as mX = 0.4 GeV. The top

plot corresponds to mH+ = 130 GeV, and the bottom plot corresponds

to mH+ = 300 GeV. The solid black line is the branching ratio for

H+ → W+X. Branching ratios for other final states are as indicated

by other colors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

4.7 Constraints on the cτ of X from the ATLAS exclusion. The ATLAS

exclusion in the (cτ , σ × BR) plane [99], where the region above the

parabola is excluded. The diagonal curves are the dependence of σ×BR

on cτ for different values of vΣ. This figure was made by our collaborator

G. Ovanesyan. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

4.8 Constraints on the non-Abelian kinetic mixing model parameters, re-

casted from the ATLAS results in Ref. [99]. The curves give the exclu-

sion regions in the(vΣ, Λ/β) parameter plane for mX = 0.4 GeV (the

red region) and mX = 1.5 GeV (the yellow region). This figure was

made by our collaborator G. Ovanesyan. . . . . . . . . . . . . . . . . . 86

5.1 The diagrams for the one loop QCD corrections to the operators in

Table. ??. The grey dot represents insertion of the operators in Table. ??. 95

5.2 The Feynman rule for the χχgg vertex. . . . . . . . . . . . . . . . . . . 99

5.3 The dependence of the functions X(r), Y (r), S(r) on the new physics

scale Λ, with r = αs(µ)/αs(Λ), in the RG equation solutions in Eq. ??

and Eq. ??. Here we used the low energy scale µ = 1 GeV. . . . . . . . 99

Page 17: Complementarity of Symmetry Tests at the Energy and

xvi

5.4 Ratio of the NLO order to LO DM-nucleon cross sections from the

operators O3 and O4. There is no quark mass factor in operators O3

and O4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

5.5 The QCD running effect on the relic abundance curve for the operators

O3 and O4. There is no quark mass factor in operators O3 and O4. . . . 104

5.6 Box graphs which describe the χ-quark interactions at the one loop level.106

5.7 Feynman rules for the Dirac DM and gauge boson interaction. . . . . . 108

6.1 Upper: ∆CKM vs M2, the parameter values are ml2= 120 GeV and µ =

M1 = 80 GeV. Lower: ∆CKM, the parameter values are ml2= 120 GeV

and M1 =M2 = 80 GeV. The resulting charginos are sufficiently heavy

as to obey the LEP limits. . . . . . . . . . . . . . . . . . . . . . . . . . 121

6.2 ∆CKM vs M1, the parameter values are ml2= 120 GeV and µ = M2 =

80 GeV. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

6.3 ∆CKM and ∆e/µ vs µ and M2, where µ = M2. The difference between

the two figures is that the upper figure has M1 = 500 GeV, while the

lower figure has M1 = 1 TeV. . . . . . . . . . . . . . . . . . . . . . . . 124

6.4 The ∆CKM and ∆e/µ scatter plots for parameters constrained by weak

charge and LHC results. The upper one shows only the constrained

plot. The lower one shows the comparison between the constrained

plot and the plot with completely random parameters. . . . . . . . . . 125

Page 18: Complementarity of Symmetry Tests at the Energy and

1

Chapter 1

Introduction

Page 19: Complementarity of Symmetry Tests at the Energy and

2

1.1 Particle physics standard model and beyond

The Standard Model (SM) is a great success in fundamental particles and in-

teractions. In this model, there are three types of particles: quarks, leptons, and

gauge bosons. The quarks include u, d, s, c, b, t quarks, the leptons include

e, νe, µ, νν , τ, ντ . The interactions include electromagnetic, weak, and strong interac-

tions, where the electomagnetic and weak interactions are unified in a SU(2)L⊗U(1)Y

electroweak theory, and the strong interaction is described by the SU(3)c quantum

chromodynamics (QCD) theory.

Although the Standard Model has huge and continuing success in experiments,

it is still often believed to be not sufficient to explain all phenomena and not a com-

plete theory of matter and interactions. For example, it does not incorporate general

relativity for the gravitation interaction. It does not have massive neutrinos and thus

does not explain the neutrino oscillation experiments. It also does not explain the dark

matter which is observed in cosmology. It has the hierachy problem, which is the large

difference between the magnitudes of weak force and gravity, and the large correction

to the Higgs mass. It also lacks naturalness, which means that many parameters differ

by many orders of magnitude. The baryon asymmetry, which means the imbalance in

baryonic and antibaryonic matter in the universe, is also a problem of SM.

To solve the above and other problems. Many theories or models beyond the

Standard Model have been proposed, such as the Grand Unified Theories, the seesaw

mechanisms for neutrino mass, string theory, extra dimensions, supersymmetry, and

other theories beyond the Standard Model.

Page 20: Complementarity of Symmetry Tests at the Energy and

3

In this paper, we study the test and search of lepton number violation, dark

matter, and supersymmetry.

1.2 Lepton number and its conservation

The lepton number in particle physics is defined as the number of leptons minus

the number of antileptons in a reaction process. Written in equation, it is Ltotal =

Ll −Ll, where Ll is the number of letpons, Ll is the number of antileptons, and Ltotal

is the total lepton number. Each lepton and antilepton has a lepton number value of

+1 and −1 respectively. Besides total lepton number of a system, each leptonic family

has its own lepton number, like Le, Lµ, and Lτ .

Lepton number conservation is a law which says that in particle reaction pro-

cesses, the total lepton number must remain the same. The lepton number for each

type of lepton must also remain the same. An example is β decay n → p + e− + νe,

in which the left hand side has lepton number 0, and the right hand side has lepton

number 0 + 1− 1 = 0. Another example is muon decay µ− → e− + νe + νµ, in which

the lepton number of each lepton type is conserved, respectively. Lepton number con-

servation is useful in determining whether a particle reaction process is possible to

happen.

1.3 Lepton number violation

While lepton number conservation is an important conservation law in the Stan-

dard Model, and is supported by many experiments, there are several reasons why we

care whether the lepton number is a good symmetry of nature.

First of all, the Standard Model requires lepton number conservation because

Page 21: Complementarity of Symmetry Tests at the Energy and

4

its Lagrangian is invariant under the global U(1)e ⊗ U(1)µ ⊗ U(1)τ rotations of the

lepton fields, assuming that there are no neutrino mass terms. This results in the

conservation of the total lepton number and the conservation of the lepton number of

each of the lepton types. However, this is just an “accidental" consequence of the fact

that there are no possible renormalizable Lagrangian terms that violate the lepton

numbers. In other words, there is no corresponding fundamental symmetry behind

the lepton number conservation. So in a new model beyond the Standard Model,

it is possible that the lepton number is violated. Moreover, lepton number is not

conserved in the SM at the level of quantum corrections. The B+L anaomaly implies

lepton number is not conserved.

The second reason is that there is a general obstacle to treating the lepton

number and baryon number as fundamental symmetries of nature, since they are

violated by non-perturbative electroweak effects. In Ref [1], it is shown that the

lepton number conservation law is violated by Bell-Jackiw anomalies, in models of

fermions coupled to gauge fields.

The neutrinoless double beta decay (0νββ decay), which will be explained in

the next subsection, is a typical lepton number violating process. If 0νββ decay is

observed, then the Schechter-Valle Theorem [2] indicates that neutrinos are Majorana

particles. This is important for neutrino property.

Another important factor is that leptogenesis requires lepton number violation.

In Ref [3], the following lepton number violating Lagrangian is presented:

L = LW.S. + N tR∂N

tR +MtN

tcRNR + h.c.+ hijN

iRl

jLφ

† + h.c., (1.1)

where N tR is a right-handed Majorana neutrino. In this model, the decays of NR:

Page 22: Complementarity of Symmetry Tests at the Energy and

5

NR → lL+ φ and NR → lL+φ have different day rates through the one-loop radiative

correction by a Higgs particle if CP is violated. This gives the net lepton number

production.

Finally, the lowest dimension non-renormalizable operator violates the lepton

number. For example, the dimension 5 operator

L5 =1

MLiLjHkH lǫikǫjl, (1.2)

which gives a neutrino mass

mν =λαβM

v2

2, (1.3)

violates lepton number.

1.4 Neutrinoless double beta decay

Beta decay n→ pe−νe is a process of radioactive decay in nuclear physics. In this

process, a neutron becomes a proton with the emission of an electron with some missing

energy. Double beta decay nn → ppe−e−νeνe is a process of radioactive decay in

nuclear physics. In this process, two neutrons are simultaneously transformed into two

protons inside a nucleus. Neutrinoless double beta decay (0νββ decay) nn→ ppe−e−

is a process like a double beta decay, but with no neutrinos in the final state. A

graphic illustration of 0νββ decay is as shown in Fig 1.1.

0νββ decay can happen in many models [63]. For example, the Light Majorana

neutrino model [6], in which the neutrino is a Majorana particle and at least one type

of neutrino has non-zero mass, the two neutrinos annihilate each other without going

out to the final state. This scenario is shown in Fig 1.2. Another model that makes

0νββ decay possible is the left-right symmetric model [7, 24, 9], in which a heavy

Page 23: Complementarity of Symmetry Tests at the Energy and

6

right-handed neutrino is involved. This scenario is shown in Fig 1.3. The R-Parity

Violating Supersymmetry (RPV SUSY) can also give rise to a 0νββ decay [10, 11, 12].

R-Parity is defined as PR = (−1)3(B−L)+2s, where B is bayron number, L is lepton

number, and s is spin. R-Parity was introduced as a new symmetry to eliminate the

possibility of B and L violating terms in the renormalizable superpotential in SUSY

models. So in the SUSY models which violates R-Parity (PRV SUSY), lepton number

violation is allowed. In particular, in RPV SUSY, two selectrons become a neutralino,

and emits two electrons without neutrinos. This gives the 0νββ decay. This scenario

is shown in Fig 1.4.

All those are possible models that can make 0νββ decay happen. There might

be more such models. Here in this paper we are trying to do a model independent

study, which applies to all these models for 0νββ decay.

1.5 Experimental tests and searches for LNV

There are several ways to test the lepton number violation. For examples: the

0νββ decay nn → ppe−e− mentioned before, which has a half-life longer than 1025

years [13, 14, 15, 16]. The conversion of muon type lepton to electron type lepton

µ− + (Z, A) → e+ + (Z − 2, A), with experimental branching ratio smaller than

10−12 [17]. The kaon decay K+ → µ+µ+π−, with experimental branching ratio smaller

than 3 × 10−9 [18]. Of all the potential lepton number violating processes, the 0νββ

decay is by far the most sensitive test of lepton number violation. This is why people

are most interested in 0νββ decay, and why the low energy experiments of the search

for lepton number violation are all about 0νββ decay.

There are several experimental searches for 0νββ decay. In 2001, the Heidelberg-

Page 24: Complementarity of Symmetry Tests at the Energy and

7

u

d

d

d

d

u

u

d

u

e−

e−

u

d

u

n

n

p

p

Fig. 1.1.— Graphic illustration of 0νββ decay.

d

d

u

e−

e−

u

W−

W−

νe

Fig. 1.2.— The light Majorana neutrino model that makes 0νββ decay

d

d

u

e−

e−

u

WR

WR

N

Fig. 1.3.— The left-right symmetric model that makes 0νββ decay

Page 25: Complementarity of Symmetry Tests at the Energy and

8

Moscow experiment gave a bound to the half-life of 76Ge, which is T1/2(76Ge) > 1.9×

1025 yr at 90% CL [13]. In 2006, a subset of the Heidelberg-Moscow experiment

announced a limit of T1/2(76Ge) = 2.23 × 1025 yr [14], but this result needs to be

confirmed. The EXO-200 experiments on 136Xe set a half-life limit of T1/2 > 1.1×1025

yr at 90% CL [15]. The GERDA Phase I experiment on 76Ge set the lower limit

T1/2(76Ge) > 2.1 × 1025 years at 90% CL, and the combination with the previous

experimental resutls about 76Ge sets T1/2(76Ge) > 3.0 × 1025 yr [16]. The latest

result of the KamLAND-Zen experiment on 136Xe gives the half-life T1/2 > 1.07×1026

yr at 90% CL [20]. The next generation of tonne scale experiments aim for a half-

life sensitiviy of ∼ 1027 years [19]. A comparison of several experiments, including

GERDA and EXO, are shown in Fig. 2 in Ref. [16].

The 0νββ decay experiments are on even-even nuclei, such as 76Ge and 136Xe,

because for the nuclei which has one atomic number higher have smaller binding

energy, preventing single beta decay. However, the nuclei which has two atomic number

higher have larger binding energy, making double beta decay allowed.

The above experiments are all classic searches at the intensity frontier. It is

d

d

u

e−

e−

u

e−

e−

χ0

Fig. 1.4.— The R Parity Violation Supersymmetry model that makes 0νββ decay

Page 26: Complementarity of Symmetry Tests at the Energy and

9

also possible to search for 0νββ decay at the energy frontier using LHC, to make

complementary studies.

One reason to do the energy frontier searches is that the 0νββ decay lifetime

measurement does not provide the means for determining the underlying mechanism.

To see this, let’s consider the case when 0νββ decay is generated by light massive

Majorana neutrino exchange [4]. We have the half life

1

T 0ν1/2

= G0ν |M0ν |2| 〈mββ〉 |2, (1.4)

where the mass 〈mββ〉 = |Σi|Uei|2mieiαi |, or else if the 0νββ decay is generated by

heavy particle exchange, assuming the dynamics is at scale Λ, then [24]

AH

AL

∼ M4W 〈k2〉

Λ5 〈mββ〉,⟨

k2⟩

∼ (100 MeV)2. (1.5)

For 〈mββ〉 ∼ 0.1 − 0.5 eV and Λ ∼ 1 TeV, we have AH/AL ∼ O(1). Therefore we

do not know whether it is heavy or light particle exchange. And it does not provide

the way to know the underlying mechanism. One the other hand, the underlying

mechanism can be better studied by going to the higher energy scale.

The best way to search for lepton number violation in high energy colliders is

to look for the same-sign dilepton signals, for example pp → e−e− + jets in which

the lepton number is violated by two. The ATLAS Collaboration [25] searched for

the same-sign dilepton signals at LHC for the Type II seesaw model, which is defined

as having an addition of Higgs triplet (H++, H+, H0) whose coupling to the (l, ν)L

doublets gives a small neutrino mass. Assuming pair production, couplings to left-

handed fermions, and a branching ratio of 100% for each final state, masses below

409 GeV, 398 GeV, and 375 GeV are excluded at 95% credence level for e±e±, µ±µ±,

Page 27: Complementarity of Symmetry Tests at the Energy and

10

and e±µ± final states, respectively. The ATLAS Collaboration also searched for the

same sign and opposite sign lepton pairs in LHC for supersymmetry models [26] pp→

gg + SS/OS lepton pairs, in which they set a limit of 550 GeV for gluino mass at

95% credence level. In 2015, ATLAS Collaboration did their latest search for the new

physics with same-sign dilepton signatures with an integrated luminosity of 20.3 fb−1 at

8 TeV. Exclusion limits are derived for a specific model of doubly charged Higgs boson

production [21]. CMS searched for same-sign dilepton and jets with an integrated

luminosity of 19.5 fb−1 at 8 TeV [22]. Constraints are set on several RPV SUSY models.

CMS also searched for same-sign leptons with two or more jets and missing transverse

momentum with an integrated luminosity of 2.3 fb−1 at 13 TeV [23]. Constraints

are set on various SUSY models, with gluinos and bottom squarks masses regions

excluded.

In Chapter. 2, we study the lepton number vilation signatures in LHC at an

energy of 14 TeV and comapre its discovery and exlcusion reach with the low energy

0νββ decay experiments.

1.6 Future colliders at 100 TeV

The second operational run of LHC is designed to be between 2015 and 2018. The

operating energy will be 13 TeV for 2015 to 2017. In 2016, it focused on increasing the

integrated luminosity and the collision rate. It achieved a luminosity of 1034cm−2s−1.

Other than LHC, people are thinking about building more powerful colliders in

the future. The FCC-pp is a proposed future energy-frontier hadron collider [27, 28],

which could make protons or heavy-ions collide at a center-of-mass energy of 100

TeV. It is designed to deliver an integrated luminosity of more than several 100fb−1

Page 28: Complementarity of Symmetry Tests at the Energy and

11

per year. An ultimate goal of an integrated luminosity of 30 ab−1 is proposed. The

SppC is also designed for a center-of-mass energy above 50 TeV and a luminosity of

1.2× 1035 cm−2s−1. These high energy and high luminosity can provide a much more

powerful way to search for the lepton number violation.

In Chapter 3, we will study searching for lepton number violation signals at a

100 TeV collider, wth high integrated luminosity of 3 − 30 ab−1, which is within the

design of the FCC-pp and SppC colliders.

1.7 The triplet scalar model and LHC studies

The next important symmetry we consider is the dark U(1)′ gauge symmetry

and its corresponding collider phenomenology. We look at the electroweak symmetry

breaking and the scalar sector of the Standard Model and see the possible extension

to it. Ref. [31] studied the predictions of a possible extension of the Standard Model

where the Higgs sector consists of a real triplet and an SU(2)L doublet. Because the

non-Abelian mixing model we studied in Chapter 4 is the extension of the model of

Ref. [31], we explain this model here.

Although the Standard Model has achieved great success, the scalar sector of

the theory which is responsible for the electroweak symmetry breaking (EWSB) has

to be confirmed experimentally, which is one of the primary goals of the LHC. In

Ref. [31], they considered the possibility that a light real triplet Σ = (Σ+,Σ0,Σ−)

that transforms as (1, 3, 0) under SU(3)C ⊗SU(2)L⊗U(1)Y is added to the Standard

Model scalar sector. This is a simple and natural extension to the SM scalar sector

because it only added a scalar triplet to the doublet. This model can also provide

a light charged scalar which can be a dark matter candidate. In this model, the

Page 29: Complementarity of Symmetry Tests at the Energy and

12

Lagrangian of the scalar sector is

Lscalar = (DµH)†(DµH) + Tr(DµΣ)†(DµΣ)− V (H,Σ) (1.6)

where H is the Standard Model Higgs doublet:

H =

(

φ+

φ0

)

, (1.7)

and Σ is the real triplet, which can be written as its components:

Σ =1

2

(

Σ0√2Σ+

√2Σ− −Σ0

)

(1.8)

A compact form of the most general renormalizable scalar potential is

V (H,Σ) = −µ2 H†H + λ0(

H†H)2 −1

2M2

ΣF +b44F 2 + a1 H

†ΣH+a22H†HF , (1.9)

where

F ≡(

Σ0)2

+ 2Σ+Σ−. (1.10)

Both H and Σ can be written in terms of their vacuum expectation value v0 and

x0 respectively as the following:

H =

(

φ+

(v0 + h0 + iξ0)/√2

)

(1.11)

and

Σ =1

2

(

x0 + σ0√2Σ+

√2Σ− −x0 − σ0

)

. (1.12)

The relations between the parameters can be obtained by minimizing the tree-level

potential.

After electroweak symmetry breaking, the mass term of the neutral scalars can

be written as:

V = ...+1

2

(

h0 σ0)

M20

(

h0

σ0

)

+ ..., (1.13)

Page 30: Complementarity of Symmetry Tests at the Energy and

13

where the mass matrix is

M20 =

(

2λ0v20 −a1v0/2 + a2v0x0

−a1v0/2 + a2v0x0 2b4x20 +

a1v204x0

)

, (1.14)

and similarly, we have the mass matrix for the charged scalars:

M2± =

(

a1x0 a1v0/2

a1v0/2a1v204x0

)

. (1.15)

The masses of the eigenstates of the neutral and charged scalars are given by:

(

H1

H2

)

=

(

cos θ0 sin θ0− sin θ0 cos θ0

)(

h0

σ0

)

, (1.16)

(

)

=

(

− sin θ± cos θ±cos θ± sin θ±

)(

φ±

Σ±

)

, (1.17)

where the θ0 and θ± are the mixing angles.

Ref. [31] find that in this model, the decay of the Standard Model like Higgs boson

into two photons can be different substantially from that of the Standard Model. If

the neutral tripletlike Higgs has a vanishing vev, the charged scalars can be long-lived,

which can have a distinctive single or double charged track plus MET signals at the

LHC. If the vev is non-vanishing, the γγ decays of the triplet-like neutral scalar can

have a large rate for the γγτν and γγbb states.

We then consider the scenario of the mixing between the SM SU(2)L and a dark

sector U(1)′ gauge group via non-Abelian kinetic mixing, with the presence of a scalar

SU(2)L triplet.

1.8 The motivation of non-Abelian kinetic mixing

The search for weakly coupled light vector bosons (dark photons) has been one of

the interests in recent years. This interest is because dark photons are possible cause

Page 31: Complementarity of Symmetry Tests at the Energy and

14

of the (g − 2) anomaly [37]. It is helpful to achieve the Sommerfeld enhancement for

dark matter annihilation, which is often needed in many DM scenarios to obtain the

right relic density. And by interacting with SM photons, dark photon can also allow

the existence of a dark sector.

The searches were in several different ways, like low energy and high energy

colliders, meson decays, and beam dump experiments [38, 39]. Previous theoretical

studies mostly considered that the interactions of the new vector bosons with the SM

fields are mediated by Abelian kinetic mixing between the SM hypercharge and the

dark U(1)’ gauge groups, or via the mass terms in the Lagrangian [40, 41, 42, 43, 44].

For both the Abelian and mass term mixing, the effects arise from the renormalizable

operators. The coupling between the vector boson and the SM fields is described by

a parameter ǫ which is constrained by experiments to be less than O(10−3). However,

this small value of ǫ is not natural and needs to be explained. In Chapter 4, we show

how the non-Abelian kinetic mixing between the U(1)′ and the SM SU(2)L gauge

groups can provide a natural explanation of the small ǫ. There we also discuss the

possibilities for future LHC tests of this scenario.

The idea of non-Abelian kinetic mixing is not original to us. Ref. [32] considered

a U(1)Y ⊗ SU(2)′ model, in which the SU(2)′ is a dark gauge group [46]. The dark

SU(2) gauge invariance requires an extra scalar triplet. However, for large values of

the dark triplet vev, a small ǫ value requires a small operator coefficient. Applications

in astrophysical anomalies and other constraints were studied in this model in a follow-

up work of them [33]. Ref [34] used this non-Abelian kinetic mixing to explain the 3.55

keV X-ray line. Ref. [35] considered the SU(2)L⊗U(1)′ kinetic mixing in a dimension

Page 32: Complementarity of Symmetry Tests at the Energy and

15

six operator

C

Λ2H†T aHW a

µνXµν , (1.18)

where H is the usual Standard Model Higgs doublet. The analysis gives the coupling

ǫ ∼ C(v/Λ)2. If this dimension six operator arises at one-loop with a mediator with

mass mϕ, one can show that Λ ∼ 4πmϕ. For Λ >∼ 10 TeV, which is mϕ>∼ 1 TeV,

the experimental constraints on ǫ can be satisfied for C ∼ O(1). The authors also

considered an explicit model, which has a scalar mediator ϕ ∼ (1, 3, 0, qD) and a dark

Higgs hD ∼ (1, 1, 0, qD) that generates the dark photon mass. They also analyzed the

collider signatures of the dark boson.

1.9 Dark matter

Dark matter is another important topic of physics beyond the Standard Model.

Dark matter is a type of matter that does not have electromagnetic interaction, which

means that it does not emit, absorb, or reflect electromagnetic waves, and is therefore

invisible via the electromagnetic spectrum. Dark matter has not been observed directly

due to the lack of electromagnetic interaction, but its existence can be inferred from

its interaction with visible matter via gravitational effects. Although not observed,

dark matter is very important. Dark matter constitutes more than 80% of the total

mass of the universe, and dark mass plus dark energy constitute more than 95% of

the energy density of the present universe. Thus, dark matter can influence the large-

scale structure of the universe, the galaxies, and can affect the cosmic microwave

background of the universe. It can also cause gravitational lensing.

In experiments, dark matter particles may be produced at high energy colliders

like LHC, and because of the lack of electromagnetic interaction, they will not be ob-

Page 33: Complementarity of Symmetry Tests at the Energy and

16

served by the detectors. But we can infer their existence by calculating the energy and

momentum carried by them, which is one part of the missing energy and momentum.

In theory, the current most popular hypothesis for dark matter is that it is

weakly interacting massive particles (WIMPs), which interact with other particles

through only gravitational force and weak force. The WIMPs can pass through ordi-

nary matter without being noticed because they are weakly interacting, but they are

massive because they take part in the gravitational interaction. The neutrino in the

Standard Model is an example of a WIMP particle. However, the neutrino mass is

too small, if any, to contribute to the large dark matter mass in the universe. Some

physics theories beyond the Standard Model can provide such dark matter particles

as the neutralinos in supersymmetry, particles in extra dimension theories, and axions

which were originally proposed to explain the neutron’s lack of electrical dipole mo-

ment (EDM), which needs explanation because the θQCD term is naturally non-zero in

QCD and it can induce an EDM for the neutron [45]. Experiments have not detected

these particles yet.

1.10 Motivations for studying dark matter operator mixing

and running

As shown in the above section, although we have evidence for the existence of

dark matter in the universe and the important role it plays in the universe, very

little is known about its features and its non-gravitational interactions. In the past

decade, several experiments have searched for dark matter interactions with atomic

nuclei. Some of these experiments claimed to have observed the signal [36, 47], while

Page 34: Complementarity of Symmetry Tests at the Energy and

17

others reported no signal [48, 49, 50, 51]. The purpose of the next generation of direct

detection experiments is to clarify and improve this situation.

Since little is known about the short-distance physics of the dark matter particle

interactions, it is useful to first study them in effective theory, which is in a model

independent way [52, 53, 54, 55, 56, 57]. For the low-energy DM interactions, the

heavy intermediate particles have been integrated out. So DM effective theory is

most appropriate for such interactions. In this way, we consider a set of operators

which can generate interactions between dark matter particles and Standard Model

particles. Basically, the effective theory has a set of non-renormalizable operators

which contains both the dark matter and Standard Model fields. Refs. [52, 53, 54, 55]

studied scenarios with spin-zero and spin-1/2 dark matter operators, and Refs. [56, 57]

studied spin one dark operators. One advantage of this approach is that we can use

the constraints from the LHC direct detection on dark matter in a model independent

way. For example, as shown in Refs. [58, 59]. But Refs. [60, 61, 62] pointed out

that this approach has limitations, and when the mediator is not heavy, more work is

needed to obtain the collider constraints on dark matter searches [62]. However, in the

case when the energy scale is low compared to the mediator mass, the effective theory

is still a useful method for studying the result of the direct detection experiments.

In this context of dark matter effective theory, the effects from beyond the leading

order can play an important role in some cases [56, 63, 64, 65, 66, 67, 68, 69, 70, 71,

72, 73, 74, 75]. The electroweak loop correction effects [64, 71] can generate mixing

between the spin-independent and spin-dependent dark matter operators for direct

detections. The meson-exchange currents which can be thought of as the long-range

QCD effects, were shown in Refs. [63, 66, 69] to play an important role in the theoretical

Page 35: Complementarity of Symmetry Tests at the Energy and

18

precision in the calculations of WIMP-DM cross sections. As shown in Ref. [69], in the

case of the isospin-violating dark matter model [76], these long-distance QCD effects

can lead to significantly different phenomenology [69]. Ref. [67] shows that the loop

effects can change the LHC monojet bounds on dark matter couplings by several orders

of magnitude. In Chapter. 5, we study the effect of the loop corrections and mixings

of the dark matter operators in a model independent way using effective theory.

1.11 Supersymmetry

Supersymmetry is one of the most well-motivated new physics beyond the Stan-

dard Model. It proposes a new symmetry which relates the bosons and fermions. In

this symmetry, each boson has a superpartner fermion and each fermion has a super-

partner boson. The spin of the superpartner differs from itself by a half-integer. For

example, the electron has a superpartner selectron with spin 0. If supersymmetry is a

perfect symmetry, each pair of superpartners should have the same mass. But since no

superpartners of the Standard Model particles have been observed in experiments, the

superpartners must have different mass from the SM particles. This difference can be

generated from a spontaneously broken symmetry. The simplest form of the sponta-

neously broken symmetry is the Minimal Supersymmetric Standard Model (MSSM),

which is the best candidate for supersymmetric theories. The benefit of supersymme-

try is that provides a potential solution to the hierarchy problem, and some natural

dark matter candidate, and a way for the gauge grand unification.

The searches for supersymmetric models have been ongoing for many years, in

both the measurement of low-energy observables, the dark matter density measure-

ment and collider experiments including the LHC. The first run of the LHC found

Page 36: Complementarity of Symmetry Tests at the Energy and

19

no direct evidence for supersymmetry, and thus many scenarios and parameter spaces

are constrained. The low energy precision measurement experiments can be a com-

plementary search for the LHC experiments.

1.12 Charged current universality and the weak charge

The universality of the charged current weak interaction is a feature of the Stan-

dard Model and it has been tested with high precision, thus, the experiments place

stringent constraints on the Beyond Standard Models with non-universality, which

includes the MSSM. So testing the universality is an important way to discover or

exclude MSSM models.

Two useful quantities to test the charged current universality are ∆CKM and

Re/µ [77]. The ∆CKM describes the deviation of the square sum of the first-row

Cabibbo-Kobayashi-Maskawa (CKM) matrix from unity:

∆CKM =(

|Vud|2 + |Vus|2 + |Vub|2)

− 1. (1.19)

The correction to the largest entry Vud can be related to the Fermi constant GβV

in the following way [77]:

GβV = GµVud

[

1 + ∆r(V )β −∆rµ

]

gV (0), (1.20)

where ∆r(V )β and ∆rµ are the corrections to the amplitudes of the β-decay and muon

decay respectively. Both of these corrections can be from the Standard Model and

Beyond Standard Model, like the MSSM.

In the difference (∆r(V )β −∆rµ), the SM W-boson propagator modifications cancel

due to universality, leaving only the non-universal corrections. The shift in the value

Page 37: Complementarity of Symmetry Tests at the Energy and

20

of ∆CKM due to beyond SM physics is then in the form

δ∆CKM = −2|Vud|2[

∆r(V )β −∆rµ

]

(BSM). (1.21)

The MSSM corrections to this difference, which are non-universal, were calculated at

one-loop in Ref [77] and studied numerically in Ref [78].

The experimental value of ∆CKM is currently [80]

∆CKM = −0.0001± 0.0006, (1.22)

and the largest theoretical numerical value from MSSM in Ref [78] can reach order

10−3 with some chosen parameter values.

The Re/µ is the ratio of pion decay branching ratios:

Re/µ =Γ [π+ → e+ν(γ)]

Γ [π+ → µ+ν(γ)]. (1.23)

The advantage of calculating Re/µ is that many hadronic uncertainties cancel

from this ratio. Recent work gives the deviation of SM prediction from experiments [81]

∆e/µ ≡ ∆Re/µ

Re/µ

≡Rexp

e/µ −RSMe/µ

RSMe/µ

= −0.0034± 0.0030± 0.0001, (1.24)

where the first error is experimental and the second is theoretical. The MSSM con-

tribution to this ∆e/µ was calculated in Ref [79] and numerically studied in Ref [78],

which shows that the numerical value can reach order 10−3.

The weak charge is another quantity to search for SUSY and new physics beyond

the SM [82, 83]. The parity-violating electron scattering experiments can provide

ways to measure it. The weak charge of the fermion is defined in the effective A× V

Lagrangian [83]:

Leff = − Gµ

2√2Qf

W eγµγ5efγµf , (1.25)

Page 38: Complementarity of Symmetry Tests at the Energy and

21

where Gµ above is the Fermi constant. At tree level in SM, the weak charges of

electrons and protons are QeW = 1 − 4 sin2 θW ≈ 0.1, and the one loop electroweak

corrections reduces the values to QeW = −0.0449 [82, 84] and Qp

W = 0.0716 [82]. This

significant suppression of their values in SM makes them more transparent to possible

effects of new physics beyond SM.

At tree level in SM, the fermion weak charge is

QfW = 2If − 4Qf sin

2 θW . (1.26)

With higher order corrections, the fermion weak charge can be written in the

following form [83]:

QfW = ρPV

[

2I3f − 4QfκPV sin2 θW]

+ λf (1.27)

By comparing with Eq. (1.26), we see that at tree level ρPV = κPV = 1 and λf = 0.

At one loop level we can write [83]:

ρPV = 1 + δρSM + δρSUSY (1.28)

κPV = 1 + δκSM + δκSUSY (1.29)

λf = λSMf + λSUSYf (1.30)

The variables ρPV and κPV can be written in terms of oblique parameters S, T

Page 39: Complementarity of Symmetry Tests at the Energy and

22

and U , which are defined as the following [85]

S =4s2c2

αM2Z

Re

{

ΠZZ(0)− ΠZZ(M2Z) +

c2 − s2

cs

[

ΠZγ(M2Z)− ΠZγ(0)

]

+ Πγγ(M2Z)

}New

,

T =1

αM2W

{

c2(

ΠZZ(0) +2s

cΠZγ(0)

)

− ΠWW (0)

}New

,

U =4s2

α

{

ΠWW (0)− ΠWW (M2W )

M2W

+ c2ΠZZ(M

2Z)− ΠZZ(0)

M2Z

+ 2csΠZγ(M

2Z)− ΠZγ(0)

M2Z

+ s2Πγγ(M

2Z)

M2Z

}New

, (1.31)

where "New" means that we only include new physics contributions to the self-energies.

Then the variables ρPV and κPV can be written as the following [83]:

δρSUSY = αT − δµV B (1.32)

δκSUSY =

(

c2

c2 − s2

)(

α

4s2c2S − αT + δµV B

)

+c

s

[ΠZγ(q2)

q2− ΠZγ(M

2Z)

M2Z

]SUSY

+( c2

c2 − s2

)[

−Πγγ(M2Z)

M2Z

+∆α

α

]SUSY

+ 4c2F eA(q

2)SUSY (1.33)

Any parameters combination leading to values of S, T and U lying outside the present

95% confidence limit will be ruled out.

Page 40: Complementarity of Symmetry Tests at the Energy and

23

Chapter 2

Lepton number violation collider study at 14

TeV and comparison with 0νββ decay

Page 41: Complementarity of Symmetry Tests at the Energy and

24

2.1 The models

In this chapter, we study the lepton number violation at 14 TeV LHC and the

low energy 0νββ decay experiments and compare their sensitivity. This discussion

is based on the work published in Ref [86]. The purpose of this section is to study

the lepton number violation at LHC, by searching for the same sign dilepton signals.

We can extract the effective LNV operators from the neutrinoless double beta decay,

which has the corresponding process shown in Fig. 2.1.

Here we used effective operators. The reason is that we try to be model inde-

pendent. It is also possible that new physics is too heavy to be produced on shell at

the LHC. Some new physics may be light enough to be produced on shell, but we still

would like to study the effect of new particle mass.

We have implemented the model in FeynRules, and generated events with Mad-

Graph and MadEvent for pp collisions at 14 TeV, carrying out showering, jet matching,

and hadronization with Pythia and detector simulation with PGS. We implemented

the model with the following simplest operator:

L = ǫij(QαLdRα)(Q

βiL dRβ)(L

jLL

cR) + h.c.+ LSM , (2.1)

but MadGraph was not able to generate the corresponding events, because MadGraph

could not generate events for lepton number violating operators with dimension higher

than 6. Therefore, we generated the events in an alternative way, using some heavy

fictitious intermediate particles S+ and F 0 [87], as shown in Fig. 2.2.

The particle S+ is a complex sclar field with electric charge +1, and F 0 is a

Page 42: Complementarity of Symmetry Tests at the Energy and

25

d

d

u

u

e−

e−

Fig. 2.1.— The lepton number violating process extracted from neutrinoless double

beta decay.

u

d

u

d

e− e−

S+ S+F 0

Fig. 2.2.— The lepton number violating process, with fictitious intermediate particles

S+ and F 0.

Page 43: Complementarity of Symmetry Tests at the Energy and

26

neutral Majorina fermion. Their masses were both initially set at 1 TeV. We later

varied their masses.

The Lagrangian which can lead to the process in Fig. 2.2 is

L = C1udS + C2eFS∗ + h.c.+ LSM , (2.2)

and written in Gauge invariant form, it is

Leff = C1QαLdRαD + C2ǫ

ijLiLFD

∗j, (2.3)

where D is the SU(2) doublet scalar field for the S+ particle, with hypercharge

1/2:

D =

(

S

S ′

)

(2.4)

and F is the SU(2) singlet fermion field for the F 0 particle, with hypercharge 0.

Gigen the Lagrangian in Eq. 2.2, we have the following formulae for the decay

widths of S+ and F 0, which have been verified numerically by MadGraph:

ΓS =mS

(

Nc|C1|2 + θ(1− xF )|C2|2(1− xF )2)

, (2.5)

ΓF = θ(xF − 1)ΓF1 + ΓF2 , (2.6)

where

ΓF1 =Nc|C2|216π

mF

(

1− 1

xF

)2

, (2.7)

ΓF2 =Nc|C1|2|C2|2

256π3

mF

x2F

·[

2(

(1− xF )2 + xS(2xF − 3)

)

cot−1 (√s) + tan−1

(

xF−1√xS

)

√xS

+xF (4− 3xF ) + (xS + xF (4− xF )− 3) ln

(

1 + xS(1− xF )2 + xS

)

]

, (2.8)

xF ≡ m2F

m2S

, xS ≡ Γ2S

m2S

, (2.9)

Page 44: Complementarity of Symmetry Tests at the Energy and

27

and θ is the Heaviside step function:

θ(x) ≡{

0, x < 0,

1, x ≥ 0,(2.10)

and Nc ≡ 3 is the number of quark colors. We included the decay widths of S+ and

F 0 as parameters when generating events with MadGraph.

The use of particles S+ and F 0 not only enables us to generate events with

MadGraph, but what more importantly, it is a simplified model that can be matched

onto a more UV complete model, as shown in Fig. 2.2.

Fig. 2.3 shows the relation between σ(pp→ e−e−+jets) and the mass of S+ and

F 0, where the cross sections were obtained by running MadGraph. We can see that

the logarithm of the cross section is almost a linear function of mS+ and mF 0 , which

is expected because according to Fig. 2.2, we should have σ ∝ m−8S+m

−2F 0 . The curve

has a change in slope at 6 TeV, above which the particles S+ and F 0 can no longer be

integrated out.

For mS+ = mF 0 = 1 TeV, we have σ = 0.06913 pb. This means that for

an integrated luminosity of 300 fb−1, the number of events is roughly 20 thousand.

Therefore it is appropriate to generate 1 million events for the collider study.

2.2 The effect of LNV operator running

We want to constrain the coupling constant of the lepton number violating op-

erators using the low energy neutrinoless double beta decay experimental results, and

use the constrained coupling constant to do the collider studies. Because the neutrino-

less double beta decay and the collider experiments are at very different energy scales,

we would like to know the effect of the operator running with energy scale changes.

Page 45: Complementarity of Symmetry Tests at the Energy and

28

In particular, we would like to study the effect of the running of the lepton number

violating operators from new physics scales to EW scales, and then to hadronic scale.

To compute the running, we consider the following operators extracted from

neutrinoless double beta decay as the basis operators, which are the effective operator

from the operators in Eq. 2.2 when the particles S+ and F 0 are integrated out

O = ǫij(QLΓdR)(QiLΓdR)(L

jLL

cR), (2.11)

where

Γ = 1, σµν , ta, taσµν . (2.12)

In the above equation, σµν ≡ i2[γµ, γν ], and ta is the generator of the SU(3)c group.

We found that the following set of operators is a closed set of operators in the

mixing and running, and it also contains the operator (uLdR)(uLdR)(eLecR), which is

the simplest one, and the one we are interested in.

O1 = (uLdR)(uLdR)(eLecR), (2.13)

O2 = (uLσµνdR)(uLσµνdR)(eLe

cR), (2.14)

O3 = (uLtadR)(uLt

adR)(eLecR), (2.15)

O4 = (uLtaσµνdR)(uLt

aσµνdR)(eLecR). (2.16)

To study the mixing and running, we need to compute the one-loop correction

diagrams. The independent diagrams are shown in Fig. 2.4.

After calculations, we can show that when Γ = 1, the infinite terms of the

diagrams in Fig. 2.4 are

g2

16π2ǫ

1

((M)2)2−d/2

(

−O3 +1

4O4

)

, (2.17)

Page 46: Complementarity of Symmetry Tests at the Energy and

29

2 4 6 8 1010-15

10-12

10-9

10-6

0.001

1

mS+ =mF0 [TeV]

σ(p

p→

e-e-

jj)[p

b]

Fig. 2.3.— Cross section of pp → e−e− + jets vs. mass of S+ and F 0, taking the

couplings C1 = C2 = 1.

dR

dR

uL

uL

eL

eL

g

dR

dR

uL

uL

eL

eL

g

dR

dR

uL

uL

eL

eL

g

dR

dR

uL

uL

eL

eL

g

Fig. 2.4.— The one-loop correction to the dd → uuee LNV process. There are two

more diagrams symmetric to the two diagrams on the lower part.

Page 47: Complementarity of Symmetry Tests at the Energy and

30

where ǫ ≡ 2−d/2, and we worked in the modified minimal substraction renormalization

scheme.

Similarly, when Γ = σαβ, the infinite terms are

g2

16π2ǫ

1

((M)2)2−d/2(12O3 − 3O4) , (2.18)

When Γ = ta, the infinite terms are

g2

16π2ǫ

1

((M)2)2−d/2

(

−2

9O1 +

1

18O2 +

1

3O3 −

1

12O4

)

, (2.19)

when Γ = taσαβ, the infinite terms are

g2

16π2ǫ

1

((M)2)2−d/2

(

8

3O1 −

2

3O2 − 4O3 +O4

)

. (2.20)

Doing the same calculations for the upper-right graph in Fig. 2.4, which has the

same results as above.

For the lower-left graph in Fig. 2.4, the infinite terms are:

g2

16π2ǫ

1

((M)2)2−d/2

(

O3 +1

4O4

)

, if Γ = 1, (2.21)

g2

16π2ǫ

1

((M)2)2−d/2(12O3 + 3O4) , if Γ = σαβ, (2.22)

g2

16π2ǫ

1

((M)2)2−d/2

(

2

9O1 +

1

18O2 +

7

6O3 +

7

24O4

)

, if Γ = ta, (2.23)

g2

16π2ǫ

1

((M)2)2−d/2

(

24

9O1 +

2

3O2 + 14O3 +

7

2O4

)

, if Γ = taσαβ. (2.24)

Page 48: Complementarity of Symmetry Tests at the Energy and

31

For the lower-right graph in Fig. 2.4, the infinite terms are:

g2

16π2ǫ

1

((M)2)2−d/2

(

16

3O1

)

, if Γ = 1, (2.25)

0, if Γ = σαβ, (2.26)

g2

16π2ǫ

1

((M)2)2−d/2

(

−2

3O3

)

, if Γ = ta, (2.27)

0, if Γ = taσαβ. (2.28)

For other graphs, the results are the same as the above results due to symmetry.

Summing up all the graphs, we get the following corrections at the one-loop

level:

O1 → g2

16π2ǫ

(

32

3O1 +O4

)

, (2.29)

O2 → g2

16π2ǫ(48O3) , (2.30)

O3 → g2

16π2ǫ

(

2

9O2 +

5

3O3 +

5

12O4

)

, (2.31)

O4 → g2

16π2ǫ

(

32

3O1 + 20O3 + 9O4

)

, (2.32)

In order to calculate the anomalous dimension matrix, we define the renormal-

ization matrix Z as the following [88]:

OiR = (Z−1)ijOj

0 (2.33)

= (Z−1)ijZnq/2q Z

nl/2l Oj, (2.34)

where OR is the renormalized field, O is the unrenormalized field, and O0 is the

bare field. nq and nl are the number of quark fields and the number of lepton fields

respectively. The Zq and Zl are the wave function renormalization constants for the

Page 49: Complementarity of Symmetry Tests at the Energy and

32

quark and lepton fields, with values [89]:

Zq = 1 + δ2 = 1− g2

12π2ǫ, (2.35)

Zl = 1. (2.36)

The renormalized Lagrangian is

LReff =

j

CjORj , (2.37)

and we require LReff be independent from the scale M :

Md

dMLR

eff = 0, (2.38)

and because the bare field O0 is M independent, the above equation means

Md

dMLR

eff =Md

dM

j

CjORj =M

d

dM

j

Cj(Z−1O0)j = 0, (2.39)

which is(

Md

dMC

)

Z−1O0 + C

(

Md

dMZ−1

)

O0 = 0 (2.40)

Md

dMCj +

i

Ciγij = 0, (2.41)

where

γij =∑

k

(

Md

dMZ−1

ik

)

Zkj, (2.42)

is the anomalous dimension matrix.

Combining the above equations, we can obtain

γ =αs

8 0 0 1

0 −8/3 48 0

0 2/9 −1 5/12

32/3 0 20 19/3

(2.43)

Page 50: Complementarity of Symmetry Tests at the Energy and

33

Eq. 2.41 is the Renormalization Equation. In matrix form, it is

Md

dMC + γTC = 0. (2.44)

The Wilson coefficients then evolve in the above way.

After considering the running of αs, we have solved the Renormalization Equa-

tion analytically. The analytical solution is too long to list here. The numerical

solution to the Renormalization Equation is shown in Fig. 2.5. Under this evolution,

we find, for example, that if only C1(M = Λ) is non-vanishing at the high scale, then

the magnitude of the Wilson coefficients Cj(M = 1 GeV) are: C1 = 0.203C1(Λ),

C2 = −0.007C1(Λ), C3 = 0.266C1(Λ), and C4 = −0.055C1(Λ).

2.3 Constraint from neutrinoless double beta decay

For the low energy neutrinoless double beta decay, we can write the effective

Lagrangian as

LeffLNV =

j

Cj

Λ5Oj + h.c.. (2.45)

At the neutrinoless double beta decay scale, which is the GeV scale, it is no longer

appropriate to use the quark degrees of freedom. So we have to match the operators

Oj onto operators at hadronic degrees of freedom [90]. To do this, we followed Ref. [63]

to find the SU(2)L×SU(2)R chiral and parity transformation properties of the Oj. We

also did the Fierz transformation of O3,4 to quark bilinears which are all color singlets,

which gives an effective O1:

Ceff ≈ C1(1 GeV)− 5

12C3(1 GeV) = 0.092C1(Λ) (2.46)

where we omitted the contributions from C2,4(1 GeV) due to running from high to

low scale.

Page 51: Complementarity of Symmetry Tests at the Energy and

34

We can then write O1 in the notation of Ref. [63] as:

LeffLNV =

Ceff

2Λ5

(

O++2+ −O++

2−)

eLecR + h.c. , (2.47)

where ecR ≡ (eL)C and

Oab2± = qRτ

aqLqRτbqL ± qLτ

aqRqLτbqR (2.48)

with qTL,R = (u, d)L,R. Because O++2− has odd parity, and the 0νββ-decay process of

experimental interest is the 0+ → 0+ transition, we only keep the O++2+ part of (2.47).

At the hadronic level, the O++2+ eLe

cR operator can be matched onto the pion-

electron operator because they have the same SU(2)L×SU(2)R chiral transformation

and parity transformation properties:

Ceff

ΛO++

2+ eLecR + h.c.→ CeffΛ

2HF

2Λ5π−π−eLe

cR + h.c. , (2.49)

where Fπ = 92.2 ± 0.2 MeV is the pion decay constant [91], and ΛH is the mass

scale related to the matrix element of the operator O++2+ . ΛH can be estimated as

ΛH = m2π/(mu +md) ≈ 2.74 GeV for mπ+ = 139 MeV and mu +md = 7 MeV [92].

The corresponding process of the two pion-two electron operator in Eq. 2.49 is

shown in Fig. 2.6.

Following Ref. [63], we can obtain the matrix element from the operator in

Eq. 2.49:

M2π0 =

1

12π

g2AG2FΛ

2H

Λue1γ

2γ0uTe2Oππ0 , (2.50)

where gA = 1.27 is the axial pion-nucleon coupling which is related to the coupling

gπNN by the Goldberger-Treiman relation.

The squared amplitude, after simplification, is then

spin

|Mππ0 |2 = − 1

144π2

g4AG4FΛ

4H

R2Λ2ββ

|M0|2∑

spin

ue1γ2γ0uTe2 · uTe2γ2γ0ue1, (2.51)

Page 52: Complementarity of Symmetry Tests at the Energy and

35

200 400 600 800 10000.0

0.2

0.4

0.6

0.8

1.0

Λnp (GeV)

C1,C

2,C

3,C

4

C1

C3

Fig. 2.5.— The running of the Wilson coefficients. Assuming that at the TeV scale,

C1 = 1, C2 = C3 = C4 = 0.

n

n

p

p

e−

e−

p1

p2

Fig. 2.6.— The process of the two pion-two electron operator in Eq. 2.49.

Page 53: Complementarity of Symmetry Tests at the Energy and

36

Where M0 is the NME and is given in Ref. [63] as

M0 = 〈Ψf |∑

i,j

R

ρij[F1~σi · ~σj + F2Tij ] τ

+i τ

+j |Ψi〉 (2.52)

where Tij = 3~σi · ρij~σj · ρij − ~σi · ~σj, R = r0A1/3, ~ρij is the distance between nucleons i

and j, and the functions F1,2(|~ρij|) are all given in Ref. [63].

After computing the contractions and traces in the squared amplitude, we have

the following expression for the width of the neutrinoless double beta decay:

Γ = − 1

2Mi

d3p1(2π)3

1

2E1

d3p2(2π)3

1

2E2

d3pf(2π)3

1

2Ef

1

144π2

g4AG4FΛ

4HM

2

R2Λ2ββ

·|M0|2 · 4(

m2e + p1 · p2 − 2E1E2

)

(2π)4δ(4)(pi − p1 − p2 − pf ). (2.53)

We can then compute the integral in the width, and use the relation 1/T1/2 =

Γ/ ln 2, we can then obtain the half-life formula for the neutrinoless double beta decay:

1

T1/2=

~c2

288π5 ln 2· g

4AG

4FΛ

4H

R2Λ2ββ

∫ Eββ−me

me

dE1 · F (Z + 2, E1)F (Z + 2, E2)

·(

p1E1p2E2 − p1p2m2e

)

|M0|2. (2.54)

To find the numerical value of the integral in the above equation, we can use the

value of G(A,Z)0ν , which is defined as below and its value is tabulated in Ref. [93]:

G(A,Z)0ν ≡ (GF cos θcgA)

4

(

~c

R

)21

32π5~ ln 2

·∫ Eββ−me

me

dE1F (Z + 2, E1)F (Z + 2, E2)p1E1p2E2, (2.55)

where for 76Ge, it is computed as(

GGe

)−1

= 4.09× 1025 eV2 yrs.

The value for M0 was calculated by Ref. [90], using the quasiparticle random

phase approximation (QRPA). In QRPA, the particle numbers, isospin, and angular

momentum are not good quantum numbers in the basis states but they are conserved

Page 54: Complementarity of Symmetry Tests at the Energy and

37

on average. Some of these symmetries are partially restored after the equations of

motion are solved [4]. For 76Ge, it is MGe0 = −1.99. However, one thing we need to

note is that both ΛH and theM0 are subject to theoretical uncertainties. For the 0νββ-

decay mediated by light Majorana neutrinos, as an example, the NME computations

using the nuclear shell model are often a factor of two smaller than the value from

QRPA. To include the impact of both sources of uncertainty, we will later show the

results for the two different values of the product M0Λ2H which differ by a factor of

two.

Combining Eqs. 2.54 and 2.55, and considering the effect of the running of the

coefficient, we have the simplified expression for the half-life:

1

T1/2= G(A,Z)

(

ΛH

TeV

)4(1

18

)

( v

TeV

)8

×(

1

gA cos θC

)4

|M0|2[

C2eff

(Λ/TeV)10

]

, (2.56)

We can then use the above equation and the experimental values of the half-life

to constrain the effective LNV coupling constant Ceff . Then we can use the constrained

Ceff in collider studies.

2.4 Collider studies

We implemented the lepton number violating model in Eq. 2.3 using FeynRules,

and generated events using MadGraph and MadEvent for pp collisions at 14 TeV, car-

rying out showering, jet matching, and hadronization with Pythia and detector simu-

lation with PGS. For every process in the signal and backgrounds, one million events

were generated using the Titan cluster of the University of Massachusetts Amherst.

Page 55: Complementarity of Symmetry Tests at the Energy and

38

Because we are looking for the dd → uuee process, the signal we look for is the

same sign dilepton plus jets process:

pp→ e−e− + jets (2.57)

There are three categories of backgrounds [94, 26]. The first category of back-

grounds is the diboson, in which the events contains two gauge bosons. It contains

the following three backgrounds:

• WW + jets

• WZ + jets

• ZZ + jets

The diboson backgrounds have very small cross sections at 14 TeV, which is 10−6 ∼

10−5 fb. So they are subdominant backgrounds.

The second category of backgrounds is the charge-flip. Their cross sections can

be hundreds of fb, so they are one of the dominant backgrounds. In these events,

the opposite charged leptons are misreconstructed as same-sign lepton events by the

detector, which looks like the same-sign lepton signals. This often occurs when a

lepton undergoes bremsstrahlung, emitting a photon and converts to lepton and anti-

lepton pairs. The photon passes the majority of its momentum to the lepton with the

opposite charge. The overall effect is that the actual original lepton was identified as

having the opposite charge. There are two such processes:

• Z/γ∗ → e+e−,

• tt semileptonic decays.

Page 56: Complementarity of Symmetry Tests at the Energy and

39

The graph illustration of the charge-flip process Z/γ∗ → e+e− is illustrated in

Fig. 2.7. In this process, the e+ transfers most of its transverse momentum via the

photon to the e−, making the e+ looks like a e−. The graph illustration of the charge-

flip process tt is illustrated in Fig. 2.8. In this process, the e+ transfers most of its

pT to the e− via the photon, looking like a e−. The b’s are not tagged, making this

process look like the signal pp→ e−e− + jets.

The charge-flip background is the largest in the ee channel, and is also present

in the eµ channel. It does not appear in the µµ channel because of the near absence of

photon decaying to two muons. It is also possible to have a charge-flip background even

when the bremsstrahlung photon does not carry the larger share of the momentum; in

this case, the misidentification happens when the lepton ID track is misreconstructed.

This is also considered to be charge-flip.

In experiment, the lepton charge misidentification probability can be measured

and is dependent on the pseudorapidity of the lepton. Here we use the ATLAS lepton

charge misidentification probability which is shown in the left panel of Fig. 23 in

Ref. [96]. The zeros in the regions 1.37 < |η| < 1.52 are indicative of a physical hole

in the sensitivity of the ATLAS detector in the transition region between barrel and

endcap EM calorimeters.

As a cross check, we calculated the global misidentification probability on the

charge-flip events we generated and obtained the following value, which is in agreement

with the value in the left panel of Fig. 23 in Ref. [96]:

Global misidentification probability =∑

bins

NiPi

N≈ 2.41%, (2.58)

where Pi is the misidentification probability in the i-th η bin, Ni is the number of

Page 57: Complementarity of Symmetry Tests at the Energy and

40

g

g

Z

e+

e−

e−

e+

γ

Fig. 2.7.— The charge-flip Z/γ∗ → e+e− process.

gt

t

W

W

b

e+

ν

ν

e−

e−

e+

γb

Fig. 2.8.— The charge-flip tt process. The b’s are not tagged.

Page 58: Complementarity of Symmetry Tests at the Energy and

41

events with the lepton having η within the i-th η bin, and N is the total number of

events.

In our collider simulation analysis, we implemented the charge-flip probabilities

in this way. For every charge-flip event, we bin it according to its lepton pseudorapidity

and the η bins in the left panel of Fig. 23 in Ref. [96]. When making histograms of the

physics variables, if an event has η falling into any of the bins in the left panel of Fig. 23

in Ref. [96], the event weight is then multiplied by the corresponding misidentification

probability in that η bin. When making the cut-flow, we also multiply the cross

sections and the event weight by the misidentification probability in the corresponding

η bin.

The third category of backgrounds is the jet-fake. They are also dominant back-

grounds because the cross sections can be from several fb to several hundred fb. Jet-

fake is when a jet is identified as a lepton. This occurs when a lepton originates from

a jet, and the lepton is observed instead of the jet. There are four jet-fake processes:

• tt semileptonic decays,

• Single t decay,

• W + jets,

• QCD multijet.

Similarly to charge misidentification probability, a jet also has some probability

to be misidentified as a lepton. For electrons, medium criteria are used, instead of

tight criteria [96]. The medium cuts can be seen as a conservative choice. There is

no strong kinematic dependence of the jet-fake probabilities over the η region. In an

Page 59: Complementarity of Symmetry Tests at the Energy and

42

event with multiple jets, it is impossible to determine which jet will fake a lepton. To

account for all possibilities, for each physical variable we perform an event-by-event

average over all jets that could have faked the electron in a given sample, which will

be explained later.

In our analysis, we implemented jet-faking by recalculating the cross sections

and averaging the physics variables when making the distributions. The jet-fake cross

sections were calculated using the following formula:

σJF, before cuts = σJF, MG+Pythia+PGS ×(

1

5000× 1

2

)# of jet fakes

×(

# of jets

# of jet fakes

)

,

(2.59)

where 1/5000 is the medium jet-fake probability [96] and the factor of 1/2 is because

a jet fakes an electron and a positron with equal probability. The combinatorial factor

means the number of ways to choose the jets which will fake leptons among all the

jets.

To average the physical variables, we need to distinguish the ordinary jet-fake

backgrounds (tt, single t, W + jets) and the QCD multijet backgrounds. For the

ordinary jet-fake backgrounds, we do the following because, in these backgrounds, we

only need one jet to fake lepton and we required at least three jets in the signal selection

cuts, so we select the leading three jets as the potential jets which may fake leptons,

and we call their transverse momentum p(1)T , p

(2)T , p

(3)T . For HT , we first calculate it

by summing all the jet pT , and we call it H(raw)T , and calculate the average HT as in

the following way:

H(average)T =

1

3

((

H(rawT − p

(1)T

)

+(

H(raw)T − p

(2)T

)

+(

H(raw)T − p

(3)T

))

, (2.60)

where in the above equation, we subtract the pT of the leading jets from H(rawT because

Page 60: Complementarity of Symmetry Tests at the Energy and

43

these jets are faking leptons, so they should not be included in the HT value. The

factor of 1/3 is because each of the three leading jets have equal probability of faking

leptons. We can do similar things for other variables, like the lepton invariant mass

mll, leading lepton pT , etc. The MET is not affected by jet-faking, so we can calculate

it in the ordinary way.

For the QCD multijet background, the difference is that we need two jets to fake

leptons and we required four jets in the signal selection cuts. So we select the leading

four jets, from these four jets we select pairs of two jets and subtract their pT from

H(raw)T in a similar way as Eq. 2.60, and divide the result by 6, because there are six

different ways to choose two jets from four jets. We can do similar things for other

variables, like mll, and leading lepton pT , etc. Again, the MET is not affected by

jet-faking so we can calculate it in the ordinary way.

In order to make sure every background behaves like the signal, we imposed the

signal selection for the various types of backgrounds and the signal, as shown below:

• For signal: Njet ≥ 2, Ne− ≥ 2, Nb = 0,

• For diboson: Njet ≥ 2, Ne− ≥ 2, Nb = 0,

• For charge-flip: Njet ≥ 2, Ne− ≥ 1, Ne+ ≥ 1, Nb = 0,

• For tt, t+ jets, W + jets jet-fake: Njet ≥ 3, Ne− ≥ 1, Nb = 0,

• For QCD jet-fake: Njet ≥ 4, Nb = 0,

When generating events at MadGraph level, we also required that pTj,b,ℓ±

>

20 GeV, |ηj| < 2.8, |ηℓ± | < 2.5.

Page 61: Complementarity of Symmetry Tests at the Energy and

44

For the events which pass the signal selections, we find that the cuts on HT ,

ml1l2 , and MET are very effective in reducing the background while still maintaining

the signal. These variables are defined and calculated as below:

HT ≡∑

jets

pT (2.61)

ml1l2 : invariant mass of the two leading leptons (2.62)

MET : the missing transverse energy. (2.63)

We then made the distributions of these variables for events which passed the

signal selections at 14 TeV, which are shown in Figs. 2.9, 2.10, and 2.11.

From the HT distributions in Fig. 2.9, we see that there is a very good separation

between the signal and the backgrounds, and a numerical analysis on the relation

between the S/√S + B value and the location of cut shows that a cut at HT > 650

GeV is the optimal cut to suppress the backgrounds while maintaining the signal.

From the ml1l2 distribution in Fig. 2.10, we see that the distributions of processes

which have a Z as the intermediate particle have sharp peaks around the mZ ∼ 91

GeV. We can do a Z veto, which is to select the events which have ml1l2 falling outside

a region near mZ : [mZ −∆mZ , mZ +∆mZ ], but after studying the relation between

the S/√S + B value and the cuts, we find that a cut like ml1l2 > 130 GeV works

better than a Z veto cut.

From the MET distribution in Fig. 2.11, we see that there is not as clear a

separation between the signal and the backgrounds as with the HT distribution, but

the MET is still useful, because the shapes of the MET distribution are divided into

two groups. The signal and a few processes without neutrinos in the final states (like

jjzz, and QCD) have small MET, so their MET distributions are peaked at very small

Page 62: Complementarity of Symmetry Tests at the Energy and

45

σ(fb) Signal Backgrounds

Diboson Charge Flip

W−W−+2j W−Z+2j ZZ+2j Z/γ∗+2j tt

Before Cuts 0.142 0.541 6.682 0.628 903.16 68.2

Signal Selection 0.091 0.358 4.66 0.435 721.7 28.9

HT (jets) > 650 GeV 0.054 0.04 0.187 0.015 5.6 0.266

mℓ1ℓ2 > 130 GeV 0.039 0.029 0.105 0.008 0.163 0.127

MET < 40 GeV 0.036 0.005 0.036 0.007 0.126 0.014

Table 2.1: Cut-flow part 1/2. Designed for optimizing signal relative to backgrounds.

The backgrounds include diboson and charge-flip. For the cut flow of jet-fake back-

grounds, see Table 2.2.

value. In contrast, the other processes, which have neutrinos in the final states, have

more flat MET distributions. So MET can still be very effective in suppressing the

latter group of processes. A similar study on the relation between the S/√S + B value

and the cuts shows that a cut on MET < 40 GeV is an optimal cut.

After determining the cuts from the distributions of the variables, we can make

a cut flow of all the processes. The cut flow is shown in Table 2.1 and Table 2.2. In

this cut flow, the signal is generated for MS = MF = 1 TeV and the coupling in the

model in Eq. 2.2 is C1 = C2 = 0.176, which corresponds to a neutrinoless double beta

decay rate consistent with the present GERDA upper bound.

From the cut flow we see that before the cuts, the dominant backgrounds, which

are the charge-flip Z/γ∗+2j and tt, and the jet-fake W−+3j and QCD 4j, have cross

sections several orders larger than the signal. However, after the cuts, they are com-

parable or smaller than the signal. This means that the cuts based analysis is very

effective in suppressing the backgrounds. After the cuts, the charge-flip Z/γ∗+2j back-

ground still dominates, but is much smaller and acceptable for the current luminosity.

Page 63: Complementarity of Symmetry Tests at the Energy and

46

0 500 1000 1500 20000.00

0.05

0.10

0.15

0.20

HT [GeV]

Nu

mb

er

of

ev

en

ts(n

orm

ali

zed

) Signal

jjww

jjwz

jjzz

CF Zjets

CF tt

JF W+jets

JF t+jets

JF tt

QCD

Fig. 2.9.— The HT distribution for signal and backgrounds at 14 TeV.

0 200 400 600 800 10000.0

0.1

0.2

0.3

0.4

ml1, l2[GeV]

Nu

mb

er

of

ev

en

ts(n

orm

ali

zed

) Signal

jjww

jjwz

jjzz

CF Zjets

CF tt

JF W+jets

JF t+jets

JF tt

QCD

Fig. 2.10.— The ml1l2 distribution for signal and backgrounds at 14 TeV.

Page 64: Complementarity of Symmetry Tests at the Energy and

47

σ(fb) Signal Backgrounds S√S+B

(√

fb)

Jet Fake

tt t+3j W−+3j 4j

Before Cuts 0.142 6.7 0.45 15.09 362.352 0.0038

Signal Selection 0.091 2.37 0.22 11.73 72.03 0.0031

HT (jets) > 650 GeV 0.054 0.025 0.0003 0.102 0.027 0.0213

mℓ1ℓ2 > 130 GeV 0.039 0.024 3× 10−4 0.101 0.027 0.0493

MET < 40 GeV 0.036 0.005 3× 10−5 0.03 0.017 0.0684

Table 2.2: Cut-flow part 2/2. Designed for optimizing signal relative to backgrounds.

The backgrounds include jet-fake. For the cut flow of diboson and charge-flip back-

grounds, see Table 2.1.

Using the cross sections of the signal and backgrounds after cuts, we can make

the curves of the significance of the e−e− + dijet signal as a function of integrated

luminosity, as shown in Fig. 2.12. Here, for the C1/Λ5 constraint, we used the GERDA

0νββ half-life limit, which is T1/2 (76Ge) < 3 × 1025 years. The two dashed curves

correspond to values of the NME of M0 = −1.0 and M0 = −1.99, respectively. From

Fig. 2.12, we see that a nonobservation with luminosity at ∼ 735 fb−1 for M0 = −1.99

and ∼ 70 fb−1 for M0 = −1.0 would imply exclusion at a level consistent with the

present GERDA limit. The requirement for discovery is S/√S + B ≥ 5, which means a

requirement of luminosity of >∼ 435 fb−1 for M0 = −1.99 and >∼ 4.6 ab−1 for M0 = −1.0.

It is striking to see that a difference of a factor of 2 in M0, after transferal to the limit

on C1/Λ5, implies an order of magnitude of difference in the required LHC luminosity

for exclusion and discovery.

We then did a parameter scan for the signal model, for 0 TeV ≤ Λ ≤ 5 TeV and

0 ≤ geff ≤ 1.5, where geff ≡ C1 = C2 and C1, C2 are from Eq. 2.2. In Fig. 2.13 and

Fig. 2.14, we show in the scanned region, the exclusion and discovery reache curves

Page 65: Complementarity of Symmetry Tests at the Energy and

48

0 50 100 150 2000.00

0.05

0.10

0.15

0.20

0.25

0.30

MET [GeV]

Nu

mb

er

of

ev

en

ts(n

orm

ali

zed) Signal

jjww

jjwz

jjzz

CF Zjets

CF tt

JF W+jets

JF t+jets

JF tt

QCD

Fig. 2.11.— The MET distribution for signal and backgrounds at 14 TeV.

0 1 2 3 40.5

1

5

10

ℒ [ab-1]

S/

S+

B

M0=-1

M0=-1.99

Discovery

Exclusion

Fig. 2.12.— Significance of the e−e− + dijet signal as a function of integrated lumi-

nosity, assuming that the C1/Λ5 is consistent with the GERDA 0νββ half-life limit.

Page 66: Complementarity of Symmetry Tests at the Energy and

49

for both the LHC at different luminosities (100 fb−1, 300 fb−1, 1000 fb−1). We also

show the exclusion and discovery reach curves for the low energy 0νββ decay GERDA

experiment and the future 1 Tonne experiment. For the 1 Tonne experiment, we used

a prospective sensitivity of T1/2 (76Ge) = 6×1027 years. For the GERDA and 1 Tonne

experiments, the solid and dotted curves indicate the impact of varying M0 by a factor

of 2.

From Fig. 2.13, we observe that with luminosity ≥ 100 fb−1, the LHC would

begin to extend the present GERDA exclusion limit for Λ between 1 TeV and 3 TeV.

And from Fig. 2.14, we see that the opportunities for discovery with a luminosity of

300 fb−1 appear more limited, for both the smaller and bigger nuclear and hadronic

matrix elements. However, at a higher luminosity of 3 ab−1, it could open the chance

for discovery in a range of Λ which depends on the value of M0.

We see that the reach of the 1 Tonne scale 0νββ decay experiments tend to ex-

ceed the reach of the high-luminosity LHC reaches over almost the entire range of the

parameter space. Therefore, in terms of the reach of the LHC, the above conclusion is

not as optimistic as obtained in Ref. [97] and Ref. [98]. These papers delineated the

simplified models and did a first round of collider analysis. They found that for the

simplified model we use, the 14 TeV LHC with an integrated luminosity of 300 fb−1 is

much more sensitive than the future low energy 0νββ decay experiments. We expect

that our findings regarding the three perspectives, which are the full background stud-

ies, the QCD running, and the long-range NME contributions, to generalize to other

simplified LNV models. However, it is always interesting to compare the prospects

for the high energy LHC and the low energy 0νββ decay experiments, because the

observation of an LNV signal in both these experiments is possible, and may point to

Page 67: Complementarity of Symmetry Tests at the Energy and

50

1 2 3 4 50.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

Λ [TeV]

ge

ff

GERDA(M0

=1) GERDA(M0

=2)

1 Tonne (M0=1)

1 Tonne (M0=2)

100fb

-1

300fb

-1

3000fb

-1

Fig. 2.13.— Current and future exclusion reach of 0νββ decay and LHC searches for

the TeV LNV interaction as a function of the coupling geff and mass scale Λ. The

coupling is defined as geff = C1/41 , where the C1 is the coupling in Eq. 2.45.

1 2 3 4 50.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

Λ [TeV]

ge

ff

GERDA(M0

=1) GERDA(M0

=2)

1 Tonne (M0=1)

1 Tonne (M0=2)

100fb

-1

300fb

-1

3000fb

-1

Fig. 2.14.— Current and future discovery reach of 0νββ decay and LHC searches for

the TeV LNV interaction as a function of the coupling geff and mass scale Λ. The

coupling is defined as geff = C1/41 , where the C1 is the coupling in Eq. 2.45.

Page 68: Complementarity of Symmetry Tests at the Energy and

51

the existence of LNV interactions.

Page 69: Complementarity of Symmetry Tests at the Energy and

52

Chapter 3

Lepton number violation collider study at

100 TeV

Page 70: Complementarity of Symmetry Tests at the Energy and

53

3.1 LHC signal and backgrounds

As shown in the Introduction, the future FCC-pp is designed for a center-of-mass

energy of 100 TeV and an integrated luminosity of 10 − 20 ab−1. Therefore, here we

extend our previous study of 14 TeV lepton number violation to a 100 TeV collider wth

integrated luminosity of 30 ab−1, which is within the design of the FCC-pp collider.

The model will be the same as our previous 14 TeV study, which is shown in

Eq. 2.45. We used this model to generate one million events for the signal and every

process in backgrounds, using the UMass Titan Cluster [95].

For the signal, we found that it is useful to include the e+e+ in the final state

leptons as well, because the protons contain more positively charged u-quarks than d-

quarks, making the e+e+ state in the proton-proton collisions enhanced relative to the

e−e− state. Therefore, it is better to search for the pp→ e±e± + 2j signal, rather than

the pp → e−e− + 2j signal. We ran MadGraph at 100 TeV and found the following

ratio in the cross sections:

σ(pp→ e±e± + 2j)

σ(pp→ e−e− + 2j)≃ 3.5, (3.1)

which is larger than a simple factor of 2. We will also apply the e±e± signal to the 14

TeV LHC study.

The backgrounds are similar to the 14 TeV case. The only difference is that we

now should generate e±e± instead of e−e−. So we have the following backgrounds:

The diboson backgrounds are shown below:

• WW + jets

Page 71: Complementarity of Symmetry Tests at the Energy and

54

• WZ + jets

• ZZ + jets

The charge-flip backgrounds are shown below. We use same charge misidenti-

fication probability as shown in the left panel of Fig. 23 in Ref. [96]. We calculated

the global misidentification probability in the same way as Eq. 2.58, and our results

were consistent with Ref. [96]. So it is reasonbale to continue using the misidentifica-

tion probability there. We also do the same as in the 14 TeV case to implement the

charge-flip misidentification probability in our collider analysis and calculate the cross

sections in the cut-flow.

• Z/γ∗ → e+e−,

• tt semileptonic decays.

The jet-fake backgrounds are shown below. We use the same jet-fake probabil-

ity as in the 14 TeV study, and the same formula to calculate the cross section as

in Eq. 2.59. We also use the same averaging scheme when calculating the physical

variables like HT , ml1l2 , and MET, which are defined in Eq. 2.63.

• tt semileptonic decays,

• Single t decay,

• W + jets,

• QCD multijet.

Page 72: Complementarity of Symmetry Tests at the Energy and

55

3.2 Cut analysis

Similar to the 14 TeV studies, we impose the signal selections for all the signal and

background processes. However, since we are now searching for the e±e± + 2j signal,

the signal selections will be different from the 14 TeV studies, where we searched for

the e−e− + 2j signal. The signal selections are shown below:

• For signal: Njet ≥ 2, Ne− ≥ 2, Ne+ ≥ 2, Nb = 0,

• For diboson: Njet ≥ 2, Ne− ≥ 2, Ne+ ≥ 2, Nb = 0,

• For charge-flip: Njet ≥ 2, Ne− ≥ 1, Ne+ ≥ 1, Nb = 0,

• For tt, t+ jets, W + jets jet-fake: Njet ≥ 3, Ne− ≥ 1, Ne+ ≥ 1, Nb = 0,

• For QCD jet-fake: Njet ≥ 4, Nb = 0,

When generating events at MadGraph level, we also required that pTj,b,ℓ±

>

20 GeV, |ηj| < 2.8, |ηℓ± | < 2.5.

Using the events which pass the signal selections, we can make the distributions

of variables like HT , ml1l2 , and MET, which are defined in Eq. 2.63, and determine

the optimal cuts on them to suppress the backgrounds. We also find that the leading

lepton pT is a good variable in separating the signal from the backgrounds. The

distributions of the variables are shown in Fig. 3.1, Fig. 3.2, Fig. 3.3, and Fig. 3.4.

From these distributions, we again see that the signal jets are more energetic

than the backgrounds, making HT a very good variable to select the signal from the

backgrounds. The ml1,l2 are good at suppressing the processes which have Z in the

backgrounds. The MET can suppress the backgrounds which have no neutrinos in the

Page 73: Complementarity of Symmetry Tests at the Energy and

56

σ(fb) Signal Backgrounds

Diboson Charge Flip

W−W−+2j W−Z+2j ZZ+2j Z/γ∗+2j tt

Before cuts 28.3 49.2 591 26.5 8× 106 3× 106

Signal selections 4.5 6.3 68 5.8 1.3× 103 137

HT > 500GeV 3.1 2.2 13 0.97 72 8.6

pT (lead e) > 150 GeV 2.1 0.8 7.1 0.4 22 3.0

MET < 40GeV 2.0 0.8 5.7 0.37 1.3 2.8

Z veto 1.2 0.09 1.1 0.16 0.7 0.8

Table 3.1: Cut-flow part 1/2 for 100 TeV. Designed for optimizing signal relative to

backgrounds. The backgrounds include diboson and charge-flip. For the cut flow of

jet-fake backgrounds, see Table 3.2.

final state, and the signal has a very different distribution of leading lepton pT than

the backgrounds, making this another useful variable in selecting the signal.

3.3 LHC results and comparison to 0νββ decay experiments

After determining the best cuts from the distributions, we can do the cut flow;

and the result is shown in Table 3.1 and Table 3.2. From the cut flow, we see that

similarly to the 14 TeV case, before the cuts, the dominant backgrounds are the charge-

flip and jet-fake backgrounds, especially the charge-flip Z/γ∗+2j and tt backgrounds,

which are several orders larger than the signal. After the cuts, their cross sections

are smaller than the signal. After the cuts, the cross sections of W−Z+2j, W−+3j,

and the QCD 4j backgrounds have comparable or even larger cross sections than the

signal, but not too much larger than the signal. So the cuts are still very effective in

suppressing the backgrounds and maintaining the signal.

We can then scan the parameters of the signal model and do the cuts on the

models of all the scanned parameter values, and use the cross sections of all the

Page 74: Complementarity of Symmetry Tests at the Energy and

57

0 500 1000 1500 2000 2500 30000.00

0.05

0.10

0.15

0.20

HT [GeV]

Nu

mb

er

of

ev

en

ts(n

orm

ali

zed

) Signal

jjww

jjwz

jjzz

CF Zjets

CF tt

JF W+jets

JF t+jets

JF tt

QCD

Fig. 3.1.— The HT distribution for signal and backgrounds at 100 TeV.

0 500 1000 1500 20000.00

0.05

0.10

0.15

0.20

0.25

0.30

ml1, l2[GeV]

Nu

mb

er

of

ev

en

ts(n

orm

ali

zed

) Signal

jjww

jjwz

jjzz

CF Zjets

CF tt

JF W+jets

JF t+jets

JF tt

QCD

Fig. 3.2.— The ml1l2 distribution for signal and backgrounds at 100 TeV.

Page 75: Complementarity of Symmetry Tests at the Energy and

58

0 50 100 150 200 250 3000.00

0.02

0.04

0.06

0.08

MET [GeV]

Nu

mb

er

of

ev

en

ts(n

orm

ali

zed) Signal

jjww

jjwz

jjzz

CF Zjets

CF tt

JF W+jets

JF t+jets

JF tt

QCD

Fig. 3.3.— The MET distribution for signal and backgrounds at 100 TeV.

0 200 400 600 800 10000.00

0.05

0.10

0.15

Leading lepton PT [GeV]

Nu

mb

er

of

ev

en

ts(n

orm

ali

zed

) Signal

jjww

jjwz

jjzz

CF Zjets

CF tt

JF W+jets

JF t+jets

JF tt

QCD

Fig. 3.4.— The leading lepton pT distribution for signal and backgrounds at 100 TeV.

Page 76: Complementarity of Symmetry Tests at the Energy and

59

σ(fb) Signal Backgrounds

Jet Fake

tt t+3j W−+3j 4j

Before cuts 28.3 573 25.5 585 3× 103

Signal selections 4.5 38.9 1.4 120 264

HT > 500GeV GeV 3.1 2.4 0.03 9.9 4.3

pT (lead e) > 150 GeV GeV 2.1 2.2 0.03 9.7 4.3

MET < 40GeVGeV 2.0 2.2 0.03 9.7 4.3

Z veto 1.2 0.3 3× 10−3 1.8 1.6

Table 3.2: Cut-flow part 2/2 for 100 TeV. Designed for optimizing signal relative

to backgrounds. The backgrounds include jet-fake. For the cut-flow of diboson and

charge-flip backgrounds, see Table 3.1.

backgrounds after the cuts in the cut flow tables, to make the curves of exclusion and

discovery reaches of future 100 TeV colliders. The curves are shown in Fig. 3.5 and

Fig. 3.6.

From Fig. 3.5, we can see that a luminosity of 1fb−1 of the collider would begin

to exceed the present GERDA exclusion reach for the smaller value of M0. For a

luminosity of 100fb−1, it is comparable for the future 1 Tonne experiment for smaller

M0 values in most Λ regions, and is even comparable to the 1 Tonne experiment for

larger M0 values near Λ ∼ 2.7 TeV. For the future designed larger luminosity, which

is 3ab−1 and 30ab−1, they are all beyond the exclusion reach of the future 1 Tonne

experiment for Λ > 3 TeV, making the future high energy collider more effective than

the future low energy 0νββ decay experiment in excluding lepton number violation if

no signals are observed.

From Fig. 3.6, we see that a 10fb−1 of luminosity of the collider is beyond the

discovery reach of the present GERDA experiment for the smaller M0, and is com-

Page 77: Complementarity of Symmetry Tests at the Energy and

60

parable to the GERDA experiment for the larger M0 for Λ in the range 3.5− 4 TeV.

A 100fb−1 of luminosity of the collider is beyond the discovery reach of the GERDA

experiment, but is still less effective than the future 1 Tonne experiment. For the

designed luminosity of 3ab−1 for the future designed collider, the discovery reach is

comparable to the future 1 Tonne experiment for the larger value of M0 at Λ ∼ 3

TeV, but is slightly less sensitive in other Λ regions. For the designed luminosity

of 30ab−1 for the future designed collider, the discovery reach exceeds that of the 1

Tonne experiment for the larger value of M0 for Λ > 3 TeV. Therefore, a future 100

TeV collider (like FCC-pp) with the designed luminosity of 30ab−1 is more effective

for both excluding and discovering the lepton number violation than the future low

energy 0νββ experiments.

3.4 Comparison to machine learning results

We can also compare the discovery reaches from the cut based analysis with that

from the Machine Learning analysis, which is shown in Figure 3.7.

The machine learning analysis was done by Peter Winslow. In this machine

learning analysis, a random forest classifier is trained to classify the events. For the

random forest, we first build a collection of decision trees. A decision tree is very

similar to a cut based analysis. Each node of a decision tree is a feature used to

separate the events. A feature can be a single or a combination of several physical

variables, like Ht, MET, etc. On a node, the events are divided into different groups

according to the value of the physical variable. Each group is a child of the original

node. The leaves of the tree are the result of decisions, i.e., whether an event is a

signal or background.

Page 78: Complementarity of Symmetry Tests at the Energy and

61

0 1 2 3 4 50.0

0.5

1.0

1.5

2.0

Λ [TeV]

ge

ff

GERDA

1 Tonne

1 fb-1

10 fb-1

100 fb-1

3 ab-1

30 ab-1

Fig. 3.5.— Current and future exclusion reach of 0νββ decay and 100 TeV LHC

searches for LNV interaction as function of the coupling geff and mass scale Λ. The

requirement for exclusion is S/√S + B ≥ 2. The coupling is defined as geff = C

1/41 ,

where the C1 is the coupling in Eq. 2.45. The blue shaded areas are for the uncertainty

of M0, whose borders correspond to M0 = −1.0 and M0 = −1.99.

Page 79: Complementarity of Symmetry Tests at the Energy and

62

0 1 2 3 4 50.0

0.5

1.0

1.5

2.0

Λ [TeV]

ge

ff

GERDA

1 Tonne

1 fb-1

10 fb-1

100 fb-1

3 ab-1

30 ab-1

Fig. 3.6.— Current and future discovery reach of 0νββ decay and 100 TeV LHC

searches for LNV interaction as function of the coupling geff and mass scale Λ. The

requirement for discovery is S/√S + B ≥ 5. The coupling is defined as geff = C

1/41 ,

where the C1 is the coupling in Eq. 2.45. The blue shaded areas are for the uncertainty

of M0, whose borders correspond to M0 = −1.0 and M0 = −1.99.

Page 80: Complementarity of Symmetry Tests at the Energy and

63

In this way, we can build many decision trees, each differing in the structure

or how we separate events in the nodes. The collection of all decision trees forms an

ensemble, which is often called a forest. There are several ways to utilize the forest as

a classifier. One way is to randomly pick up several trees from the forest each time,

and either let the trees vote on the decision of a particular event or take the average

of the decision of the picked trees. This is called the random forest method. Another

ensemble method is boosting, in which we make a distribution of the events in each

iteration, and the distribution gives higher weight (or probability) to the events which

we classified wrong, so that in the next iteration we can place more emphasis on the

“harder" problems. The right figure of Figure 3.7 is the result using a random forest

as the classifier.

From the right figure of Figure 3.7, we observe that a luminosity of 1fb−1 of a

100 TeV collider can exceed the discovery reach of the present GERDA experiment

for Λ greater than 2-3 TeV. A luminosity of 10fb−1 of a 100 TeV collider can be close

to the discovery reach of the future 1 Tonne experiment, and a luminosity of 100fb−1

can be comparable to the future 1 Tonne experiment.

By comparing the two figures in Figure 3.7, we see that using the machine

learning methods makes the collider experiments much more effective. This is because

of two reasons. First, the machine learning method uses many more physics variables

to construct the features, so it uses much more information. Many of these variables

are not very useful in the cut based analysis, but by combining them together using the

ensemble algorithms, they can be much more effective. Second, the cut based analysis

is equivalent to using only one decision tree, while the random forest method used in

the machine learning analysis uses a collection of many trees. Each tree can make a

Page 81: Complementarity of Symmetry Tests at the Energy and

64

decision, and the final decision is the average or vote of all the trees. This makes the

final result more reliable and less likely to suffer from over-fitting. However, as it uses

many features and constructs and trains many trees, the machine learning analysis is

much more computationally expensive, especially when we have large volumes of data.

As a summary, both cut based and machine learning analysis show that a future

100 TeV collider with high designed luminosity is more effective in searching for lepton

number violation than the future low energy 0νββ experiments.

Page 82: Complementarity of Symmetry Tests at the Energy and

65

0 1 2 3 4 50.0

0.5

1.0

1.5

2.0

Λ [TeV]

ge

ff

GERDA

1 Tonne

1 fb-1

10 fb-1

100 fb-1

3 ab-1

30 ab-1

Fig. 3.7.— Comparison of the discovery reaches using cut-based analysis (left) and

machine learning analysis (right). The random forest method is used in the machine

learning analysis. The machine learning analysis was done by Peter Winslow. In the

right figure, the notation M means Λ and yeff means geff .

Page 83: Complementarity of Symmetry Tests at the Energy and

66

Acknowledgement

I need to thank Professor Michael Ramsey-Musolf and Michael Graesser for the

helpful discussion on the LNV signal at 100 TeV, especially the discussion on searching

for the e±e± signal.

I also thank Dr. Peter Winslow for his machine learning analysis, for sharing his

result figures, and for generating the events at the UMass Titan Cluster.

Page 84: Complementarity of Symmetry Tests at the Energy and

67

Chapter 4

LHC Signatures of Non-Abelian Kinetic

Mixing

Page 85: Complementarity of Symmetry Tests at the Energy and

68

4.1 The model

In this chapter, we study the non-Abelian kinetic mixing between the Standard

Model SU(2)L and a dark sector U(1)′ gauge group, with the presentence of a scalar

triplet, and the corresponding phenomenology. This discussion is based on work pub-

lished in Ref [30]. We consider the Standard Model fields with the scalar triplet Σ,

and the scalar triplet-doublet potential, which were described in Section 1.8:

Σ =1

2

(

Σ0√2Σ+

√2Σ− −Σ0

)

, DµΣ = ∂µΣ + ig

[

3∑

a=1

W aµT

a,Σ

]

, (4.1)

V (H,Σ) = −µ2H†H + λ0(

H†H)2 − µ2

ΣG+ b4G2 + a1H

†ΣH + a2H†HG, (4.2)

where G ≡ TrΣ†Σ =(Σ0)

2

2+ Σ+Σ−, and T a is the SU(2)L generator.

We focus on the dimension-five operator

O(5)WX = −β

ΛTr (WµνΣ)X

µν , (4.3)

where Xµν is the U(1)′ vector boson and Wµν the SU(2)L vector boson fields. Λ

is the mass scale of the intermediate fields which are integrated out. Σ is defined as

Σ = ΣaT a. After Σ acquires its vacuum expectation value 〈Σ0〉 ≡ vΣ, the U(1)′ boson

Xµ and the neutral SU(2)L gauge boson W 3µ will mix, with the mixing parameter

ǫ = β(vΣΛ

)

sin θW , (4.4)

Page 86: Complementarity of Symmetry Tests at the Energy and

69

with θW being the usual weak mixing angle. The coupling between Xµ and other

fields are the same as the photon and other SM fields, except for a universal rescaling

factor ǫ. As can be seen from Eq. 4.4, the magnitude of ǫ is controlled by the ratio

vΣ/Λ. As we saw in Section 1.8, ǫ is constrained by experiments to be smaller than

O(10−3), which was hard to explain in other models. Here we see that ǫ will satisfy

the experimental bounds for β ∼ O(1) and Λ larger than 1 TeV.

In the model, we add to the Standard Model Lagrangian the dimension four and

five operators, which involve the dark photon and the real triplet fields:

L = LSM +∆L(d=4) +∆L(d=5) + . . . . (4.5)

In the above Lagrangian, the dimension four and five operators are in the form:

∆L(d=4) = −1

4XµνX

µν +ǫ0

2 cWBµνX

µν + Tr[

(DµΣ)†DµΣ

]

− V (Σ, H) + ∆L(d=4),

∆L(d=5) = − 1

ΛTr (WµνΣ) (αB

µν + βXµν) ≡ O(5)WB +O(5)

WX . (4.6)

The dimension four operator ∆L(d=4) contains the abelian kinetic mixing term

(the XB term), and cW is the cosine of the usual weak mixing angle. The terms in

∆L(d=4) which breaks the dark U(1)′ gauge group are not explicitly shown here.

Here we wrote the O(5)WX as an effective theory, leaving the model-dependent

details unspecified and focusing on the corresponding collider phenomenology. This

effective operator O(5)WX can be generated in several ways, as shown in Fig. 4.1. It

can be generated via loops, in which the mediators in the loop can be either fermions

or scalars. Or it can be generated from other degrees of freedom in non-perturbative

theories. These can be considered as possible UV complete theory. After the heavy

intermediate states are integrated out, we have the effective operator O(5)WX . Similar

graphs as shown in Fig. 4.1 may also generate the same effective operator.

Page 87: Complementarity of Symmetry Tests at the Energy and

70

The effective operator O(5)WX can contribute to the S parameter after electroweak

symmetry breaking (EWSB):

αemS = 4cW sWαvΣΛ. (4.7)

This sets a confidence level bound of 90% for αvΣ/Λ ∼< 0.0008. So we set α = 0 and

focus on the phenomenology of the operator O(5)WX .

In Eq. 4.6, after the hypercharge zero field Σ acquires a vev, we have the mixing

term in the Lagrangian:

∆Lmixing = −1

2W 0,3

µν

(

ǫWBB0,µν + ǫWXX

0,µν)

− 1

2ǫBXB

0µνX

0,µν , (4.8)

where ǫWB = αvΣ/Λ, ǫWX = βvΣ/Λ. We can then rewrite the B0, W 0,3 fields in

terms of A0, Z0, and we have

LSM +∆Lmixing ⊃ −1

4(1 + αA)A

0µνA

0,µν − 1

4(1 + αZ)Z

0µνZ

0,µν − 1

4X0

µνX0,µν

−1

2αAZA

0µνZ

0,µν − 1

2αAXA

0µνX

0,µν − 1

2αZXZ

0µνX

0,µν , (4.9)

where the α parameters are defined as

αA = 2cW sW ǫWB, αZ = −2cW sW ǫWB, αAZ = ǫWB

(

c2W − s2W)

,

αAX = ǫBXcW + ǫWXsW , αZX = −ǫBXsW + ǫWXcW . (4.10)

We can then diagonalize the kinetic terms in the Lagrangian into the following

form:

−1

4AµνA

µν − 1

4ZµνZ

µν − 1

4XµνX

µν , (4.11)

Page 88: Complementarity of Symmetry Tests at the Energy and

71

using the following transformation matrix:

A

Z

X

=

√1 + αA αAZ

√1 + αZ αAX

0√

1− α2AZ

−αAX αAZ+αZX√1−α2

AZ

0 0

√(1−α2

AX)(1−α2

AZ)−(αZX−αAX αAZ)2√

1−α2AZ

A0

Z0

X0

.

(4.12)

The above transformation has the inverse form:

A0

Z0

X0

=

1√1+αA

− αAZ√1+αA

√1−α2

AZ

−αAX+αAZ αZX√1+αA

√1−α2

AZ

√(1−α2

AX)(1−α2

AZ)−(αZX−αAX αAZ)2

0 1√1−αAZ

√1+αZ

αAX αAZ−αZX√1+αZ

√1−α2

AZ

√(1−α2

AX)(1−α2

AZ)−(αZX−αAX αAZ)2

0 0

√1−α2

AZ√(1−α2

AX)(1−α2

AZ)−(αZX−αAX αAZ)2

A

Z

X

,

(4.13)

where the α parameters are defined as

αAZ ≡ αAZ√1 + αA

√1 + αZ

, αAX =αAX√1 + αA

, αZX =αZX√1 + αZ

. (4.14)

After spontaneous symmetry breaking, the mass terms for the gauge boson are

∆Lmass =1

2m2Z0

µZ0,µ +

1

2m2

XX0µX

0,µ, (4.15)

and after being transformed to the fields {A, Z, X}, the mass matrix becomes off-

diagonal:

∆Lmass =1

2(M)2ij V

iµV

jµ, (4.16)

where V 1,2 = Z, X and

M2 =

m2Z

(1−α2AZ

)(1+αZ)

−m2Z(αZX−αAX αAZ)

(1−α2AZ

)(1+αZ)√

(1−α2AX

)(1−α2AZ

)−(αZX−αAX αAZ)2

−m2Z(αZX−αAX αAZ)

(1−α2AZ

)(1+αZ)√

(1−α2AX

)(1−α2AZ

)−(αZX−αAX αAZ)2

m2X(1−α2

AZ)+m2Z

(−αAXαAZ+αZX )2

(1−α2AZ

)(1+αZ )

(1−α2AX

)(1−α2AZ

)−(αZX−αAX αAZ)2

.

(4.17)

Now we can concentrate on the non-Abelian mixing part, and we can set the

Abelian mixing parameters α = ǫBX = 0. Then the following two terms in the

Page 89: Complementarity of Symmetry Tests at the Energy and

72

Standard Model Lagrangian becomes

eQfγµA0,µf → eQfγµ

(

Aµ − ǫWX sW√

1− ǫ2WX

)

f, (4.18)

gZ fγµ(I3L−Qs2W )Z0,µf → gZ fγµ(I3L−Qs2W )

(

Zµ − ǫWX sW√

1− ǫ2WX

)

f.

Here we see that with the new fields, the actual value of the electron charge and the

Weinberg angle in the Standard Model are shifted from experimentally known values

of e, sW , cW .

We can then derive the following Feynman rules for the interactions between

the dark bosons fields, SM leptons, gauge bosons, the neutral and charged Higgs

bosons. The new Feynman rules of the non-Abelian kinetic mixing scenario of the form

W±H∓X, ZH1X, ZH2X, AH1X, AH2X dictate novel collider signatures, which are

listed below.

Interaction Feynman rule

Xl+l− ie(

ǫ0 − βv∆sWΛ

)

W±H∓X iβΛ(gµνpp′ − pνp′µ) c∓

ZH1XiβΛ(gµνpp′ − pνp′µ) cW s0

ZH2XiβΛ(gµνpp′ − pνp′µ) cW c0

AH1XiβΛ(gµνpp′ − pνp′µ) sW s0

AH2XiβΛ(gµνpp′ − pνp′µ) sW c0

W+µ (p1)W

−ν (p2)H1Xα(p3)

iβ gΛ

(pµ3gνα − pν3g

µα) s0W+

µ (p1)W−ν (p2)H2Xα(p3)

iβ gΛ

(pµ3gνα − pν3g

µα) c0W±

µ (p1)Zν(p2)H∓Xα(p3) ∓ iβ g

Λ(pµ3g

να − pν3gµα) cW c∓

W±µ (p1)Aν(p2)H

∓Xα(p3) ∓ iβ gΛ

(pµ3gνα − pν3g

µα) sW c∓

In the above Feynman rules, c∓ ≡ cos θ∓ and c0 ≡ cos θ0 are as defined in Ref. [31]. In

the table above, all the momentum of the particles flow into the vertices. The LHC sig-

nature of the SM with an additional scalar field have been studied in Ref. [31], in which

all the Feynman rules needed for the production of scalar particles like H1, H2, H±

at colliders are listed.

Page 90: Complementarity of Symmetry Tests at the Energy and

73

We also comment here that the kinetic mixing of gauge bosons can arise from

non-abelian gauge groups. For instance, in an SU(N)×SU(M) theory in which the

gauge fields are W and Y , we can introduce a scalar field ∆ab, which transfroms as

the adjoint representation under both the SU(N)×SU(M) groups. Here the indices

“a” and “b” correspond to the indices of the SU(N) and SU(M) groups, respectively.

We can then build the d = 5 operator W aµνY bµν∆

ab similar to the operator O(5)WX . The

kinetic mixing between W and Y occurs after the ∆ab acquires a non-zero vacuum

expectation value. There can also be some renormalizable models that generate this

effective operator at the one-loop level.

4.2 Collider phenomenology

Ref. [31] studied the predictions of a simple extension to SM, where the Higgs

sector inlcudes the usual SU(2)L and the scalar triplet mentioned in Section 4.1. This

model predicts a pair of charged scalars and a dark matter candidate for vanishing

triplet vev. This model predicts a significant excess of the two-photon events compared

to that in SM. With the existence of the non-Abelian kinetic mixing operator O(5)WX ,

the collider phenomenology related to the real triplet can be very different from those

studied in Ref. [31]. Here we make the assumption that the doublet-triplet mixing

angle, which is proportional to vΣ, to be some small but non-zero value. From Eq. 1.16,

we see that in this case, the neutral scalar sector has two mass eigenstates H1,2, where

H1 is mostly the Standard Model Higgs boson andH2 is mostly Σ0. And from Eq. 1.17,

we see that the charged scalars H± are not pure triplet states, instead they are the

mixtures of Σ± and the charged component of the doublet scalar, with the other

mixtures being the longitudinal components of weak gauge bosons. We can also see

Page 91: Complementarity of Symmetry Tests at the Energy and

74

that if the doublet-triplet mixing angle is zero, then the neutral component Σ0 does

not couple to the Standard Model fermions and can be a dark matter candidate.

When the triplet gets a vev, as needed for the non-Ablelian mixing mechanism, Σ0

can no longer be a DM candidate. The coupling of the mass eigenstates H± and H2

to the SM fermions through the Yukawa interactions is enabled by the presence of a

non-vanishing doublet-triplet mixing angle.

If we have a zero vΣ value instead, then the triplet states have a common mass

m2Σ = −µ2

Σ + a2v2/2. In this case, the loop effect increases the mass of the charged

components and makes the mass splitting between it and the neutral component to

be ∼ 166 MeV. This makes the decay H+ → H2π+ possible. In our studies, which are

under the assumption that vΣ 6= 0, the choice of parameter values will not alter this

mass splitting substantially.

To do the collider studies, We implemented in FeynRules the triplet given in

Eq. 4.1 and the model given in Eq. 4.5. In the model file, we choose the parame-

ters {vΣ, λ0, b4, a1, a2, v0} in Eq. 4.2 as the fundamental parameters, where v0 is

the Higgs vev. However, in the collider studies, our input is the scalar masses like

{

M2H1, M2

H2, M2

}

. So we solved the fundamental parameters in terms of the input

parameters as following, using the relations in Ref. [31]:

a1 =M2

H+

vΣ (1 + v20/(4x20)),

λ0 =M2

H1+M2

H2±√

p

4v20,

b4 =M2

H1+M2

H2∓√

p− a1v20/(2vΣ)

4x20, (4.19)

Page 92: Complementarity of Symmetry Tests at the Energy and

75

where

p =1

2

(

r ±√

r2 − 4qr)

,

r =(

M2H1

−M2H2

)2,

q = (a1v0 − 2vΣa2v0)2 (4.20)

We can then consider the production and decays of the triplet-like scalars. The

LHC production and decay mechanisms of interest are shown in Fig. 4.2. Diagrams

(a) and (b) are the Drell-Yan pair production process: pp → V ∗ → φφ, where the

symbol φ denotes any of the physical scalars (H1,2, H±), with the subsequent scalar

decays φ→ XV . These graphs have the topology XXV V . As shown in the discussion

above, when the mixing angle is small, the φ states are mostly triplet-like. Diagrams

(c) and (d) are the production pp→ V ∗ → φX which is mediated by the non-Abelian

mixing operator O(5)WX , with a scalar decay φ→ XV . These graphs have the topology

XXV .

We applied the model files in MadGraph to obtain the LHC cross sections for

the production of the neutral and charged scalars. Fig. 4.3 and Fig. 4.4 shows the

LHC production cross sections at√s = 8 for different channels, for mφ = 130 GeV

and mφ = 300 GeV, respectively. For both the mφ values we see that the Drell-Yan

pair production dominates for β/Λ ∼< 1 /TeV, and the O(5)WX-mediated production

dominates for β/Λ ∼> 1 /TeV. For an LHC energy of√s = 14 TeV, we did the similar

study and it shows that the corresponding transition between the Drell-Yan process

and the O(5)WX-mediated production process dominates for almost the same β/Λ values.

Page 93: Complementarity of Symmetry Tests at the Energy and

76

W µaXν

Σa

F

W µaXν

Σa

S

W νa

Σa

(a) (b) (c)

Fig. 4.1.— Feynman diagrams that may generate the non-abelian mixing effective

operator O(5)WX . The intermediate particles in the loops are (a) fermions, (b) scalars,

or (c) other sources from non-perturbative dynamics.

(a) (b) (c) (d)

Fig. 4.2.— The Feynman diagrams for LHC production and the subsequent decay of

the particles in the non-Abelian mixing model with the triplet scalars. Diagrams (a)

and (b) are the scalar pair productions, followed by the scalar decays mediated by the

non-Abelian mixing operator O(5)WX . Diagrams (c) and (d) are production and decays

of H and X. In all four diagrams, the incoming vector bosons are all virtual.

Page 94: Complementarity of Symmetry Tests at the Energy and

77

10- 2 0.1 1 10 100

10- 4

10- 2

1

100

� / � [TeV-1

]

�[pb]

pp � H ± X

pp � H 2 X

pp � H + H -

pp � H ± H 2

Fig. 4.3.— LHC production cross sections for pp → V → φφ and pp → V → Xφ

at√s = 8 TeV, where φ = H+, H2. The mφ = 130 GeV, and mX = 0.4 GeV. For

the processes with final states of a single charged scalar and one neutral boson, we

summed the cross sections for both charges, for example: σ(H+H2) + σ(H−H2).

Page 95: Complementarity of Symmetry Tests at the Energy and

78

10- 2 0.1 1 10 100

10- 4

10- 2

1

100

� / � [TeV-1

]

�[pb]

pp � H ± X

pp � H 2 X

pp � H + H -

pp � H ± H 2

Fig. 4.4.— LHC production cross sections for pp → V → φφ and pp → V → Xφ

at√s = 8 TeV, where φ = H+, H2. The mφ = 300 GeV, and mX = 0.4 GeV. For

the processes with final states of a single charged scalar and one neutral boson, we

summed the cross sections for both charges, for example: σ(H+H2) + σ(H−H2).

Page 96: Complementarity of Symmetry Tests at the Energy and

79

4.3 Triplet-like scalar decay branching ratios

Besides the final states considered in Ref. [31], the triplet-like scalar H± can also

decay to W±X and H2 can decay to Z/γ X. For illustrative purposes, we show the

decay width of the tree level H± → W±X below, which is sufficient for the analysis

we consider below.

Γ(H± → W±X) (4.21)

=

1− 2(m2X+M2

W±)

M2H±

+(m2

X−M2

W±)2

M4H±

16πMH+

[

1

2

(

M2H± −m2

X −M2W±

)2+M2

XM2W±

]

β2

Λ2c2∓ ,

where c∓ is the mixing angle between the charged scalar fields. Combining other H+

decay channels given in Ref. [31], we calculated the branching ratios of all the channels,

which is shown in Fig. 4.5, Fig. 4.6, which are for vΣ = 1 GeV and vΣ = 10−3 GeV,

respectively. Both figures have mX = 0.4 GeV. In Fig. 4.5, the top plot corresponds

to mH+ = 130 GeV, and the bottom plot corresponds to mH+ = 300 GeV, and the

same applies to Fig. 4.6.

From Fig. 4.5, we see that for vΣ = 1GeV, which is near the maximum allowed

by electroweak precision tests, the H+ → W+X channel has almost 100% branching

ratio for ǫ ∼> 10−4. This means a range of β/Λ ∼> 0.1/TeV. For smaller values of

β/Λ, H+ → W+X can have any branching ratio from zero to one. From Fig. 4.6,

we see that for smaller value of vΣ, the branching ratio of H+ → W+X is essentially

100% for all values of ǫ. From these two figures, we have the observation that when

β/Λ ∼> 0.1/TeV, the branching ratio of H+ → W+X is near 100% and is independent

on vΣ, while for lower β/Λ values, the branching ratio can be any value and depends

strongly on vΣ.

Page 97: Complementarity of Symmetry Tests at the Energy and

80

10- 5

10- 4

10- 3

10- 2

10- 1

10- 5

10- 4

10- 3

10- 2

0.1

110

- 810

- 710

- 610

- 5

� / � [TeV-1

]

Br(H

+)

W+X

W+Z

W+H 1

c s

t b

� + �

10- 5

10- 4

10- 3

10- 2

10- 1

10- 6

10- 4

10- 2

110

- 810

- 710

- 610

- 5

� / � [TeV-1

]

Br(H

+)

W+X

W+Z

W+H 1

c s

t b

� + �

Fig. 4.5.— Branching ratios for H+ decays as a function of ǫ (upper horizontal axis)

and β/Λ (bottom horizontal axis), for the triplet vev vΣ = 1 GeV. The dark photon

mass is chosen as mX = 0.4 GeV. The top plot corresponds to mH+ = 130 GeV, and

the bottom plot corresponds to mH+ = 300 GeV. The solid black line is the branching

ratio for H+ → W+X. Branching ratios for other final states are as indicated by other

colors.

Page 98: Complementarity of Symmetry Tests at the Energy and

81

10- 2 0.1 1 10 100

10- 8

10- 5

10- 2

1010

- 810

- 710

- 610

- 5

� / � [TeV-1

]

Br(H

+)

W+X

W+Z

W+H 1

c s

t b

� + �

H 2 � +

10- 2 0.1 1 10 100

10- 7

10- 5

10- 3

0.1

1010

- 810

- 710

- 610

- 5

� / � [TeV-1

]

Br(H

+)

W+X

W+Z

W+H 1

c s

t b

� + �

Fig. 4.6.— Branching ratios for H+ decays as a function of ǫ (upper horizontal axis)

and β/Λ (bottom horizontal axis), for the triplet vev vΣ = 10−3 GeV. The dark photon

mass is chosen as mX = 0.4 GeV. The top plot corresponds to mH+ = 130 GeV, and

the bottom plot corresponds to mH+ = 300 GeV. The solid black line is the branching

ratio for H+ → W+X. Branching ratios for other final states are as indicated by other

colors.

Page 99: Complementarity of Symmetry Tests at the Energy and

82

From the production cross sections in Fig. 4.3, Fig. 4.4 and the decay branching

ratios in Fig. 4.5, and Fig. 4.6, we see that the LHC signatures and thus the detection

strategies vary depending on the value of β/Λ. Therefore, we divide the search regions

into the following three parts, which lead to different phenomenology for the 8 TeV

LHC search.

For β/Λ ∼ 1/TeV. In this region, the Drell-Yan pair production pp → φφ

dominates in the LHC production rates. The branching ratio of φ→ XV can be any

value from zero to one, depending on the value of vΣ.

For β/Λ <∼ 0.1/TeV. In this region, the Drell-Yan pair production pp → φφ

continues to dominate. The decay φ → XV also dominates in all the channels and

has almost 100% branching ratio.

For β/Λ ∼> 1/TeV. In this region, the pp→ Xφ process dominates. This process

is mediated by the non-Abelian kinetic mixing operator O(5)WX , which is our interest.

In addition, the branching ratio of φ → XV is close to one. In this case, the LHC

production final states are those in Fig. 4.2 (c) and (d).

We did the similar study for the production at 14 TeV, which shows that the

transition between the non-Abelian mixing O(5)WX operator mediated production and

the Drell-Yan pair production pp → φφ also happens at β/Λ ∼ 1/TeV. While all the

above three regions are interesting to explore in the future, for illustrative purposes

and interest in the O(5)WX operator mediated production, we focus below on the third

region, which is β/Λ ∼> 1/TeV.

Page 100: Complementarity of Symmetry Tests at the Energy and

83

4.4 ATLAS recast

Now we focus on the third region, which is β/Λ ∼> 1/TeV, and pp → Xφ domi-

nates and Br(φ→ XV ) is close to one. ATLAS has done the dark photon search and

has their constraints [99], which is shown in the left panel of Figure 16 in Ref. [99]. We

can recast the ATLAS result into constraints in our scenario. The ATLAS analysis in

Ref. [99] assumes the SM Higgs boson decays to 2 γd and 4 γd, leading to displaced

vertices and lepton jets in the final states. This is similar to our process, which is

shown in Fig. 4.2 (c), where an intermediate off-shell vector boson becomes two X

bosons and an on-shell vector boson. We need to note that the ATLAS analysis only

used cuts to isolate events with lepton jets and displaced vertices. They did not recon-

struct the Higgs boson invariant mass, nor did they apply cuts on the missing energy.

They did not require the presence of a final state vector boson either. Therefore, their

analysis is inclusive enough to accommodate the scenario in our case, although it as-

sumed different underlying X-boson productions. In future studies, we can improve

the LHC sensitivity to the non-Abelian kinetic mixing operator O(5)WX by including

more criteria and cuts to identify the final state vector boson.

ATLAS used the limits in the left panel of Figure 16 in Ref. [99] to obtain their

constraints in the parameter space {ǫ,mX}. However, there are several distinctions

between their analysis and our scenario. For examples, the left panel of Figure 16 in

Ref. [99] is the 95% exclusion limits for σ(H)×Br(H → 2X+· · · ) and the dark photon

lifetime cτ . In our case, however, the 95% limits is for σ(φX)×Br(φ→ V X) instead.

Moreover, in the ATLAS analysis, the σ(H) and Br(H → 2X + · · · ) are independent

form the ǫ value and their dependence on mX is negligible when mX is small. In our

Page 101: Complementarity of Symmetry Tests at the Energy and

84

scenario, however, the corresponding production cross section and branching ratios do

depend on the parameters that govern ǫ, σ(HX) ∼ (β/Λ)2 ∼ 1/(τv2Σ), where τ is the

lifetime of the dark photon X. Therefore, we need to recast the ATLAS limits in the

left panel of Figure 16 in Ref. [99] to the constraints to the parameters in our case.

To make the constraints of our parameters, we first used the ATLAS 95% CL

limit on the process σ(φX)× Br(φ→ V X) in the left panel of Figure 16 in Ref. [99],

where mX = 0.4GeV, we then plotted in the same figure the lines of several constant

cross sections of our process σ(pp→ φX) summed over all φ for three different values

of the triplet vev: vΣ = 1MeV, vΣ = 1.5MeV and vΣ = 2.5MeV. The plots are shown

in Fig. 4.7. In each of our lines, Br (φ→ V X) ≈ 100%. The region above the ATLAS

curve is excluded. So for each of our constant vΣ curves, the points of intersection with

the ATLAS curve determine the boundaries of the excluded region of cτ , which is the

decay length of X. We see that the ATLAS exclusion then applies to the exclusion of

vΣ in the MeV range, which is well below the ρ-parameter limit. The cτ can be related

to Λ/β and vΣ in the following way. We first calculated the decay width (life-time) of

X:

Γ(X → ff) ≡ 1

τ=g2Xff

Nc

12πmX(1 + 2r2)

√1− 4r2, (4.22)

where r ≡ mf/mX , and the coupling is

gXff = ie

(

ǫ0 −βvΣsW

Λ

)

. (4.23)

And then from the constraint of cτ , we can make the constraints on the {Λ/β, vΣ}

plane, which is shown in Fig. 4.8. We made this exclusion for mX = 0.4 GeV and

mX = 1.5 GeV. From the figure, we see that Λ/β can be excluded up to several hundred

GeV, for proper values of mX and vΣ. In Fig. 4.8, we observe that the exclusion limits

Page 102: Complementarity of Symmetry Tests at the Energy and

85

mX=0.4 GeV

mH+=mH2

=130 GeV

5 10 50 100 500 1000

0.1

1

10

100

cΤ HmmL

Σ*BrHpbL

ATLAS 95% CL limit on Σ*BrΣHvS=1.0 MeV,cΤLΣHvS=1.5 MeV,cΤLΣHvS=2.5 MeV,cΤL

Fig. 4.7.— Constraints on the cτ of X from the ATLAS exclusion. The ATLAS

exclusion in the (cτ , σ × BR) plane [99], where the region above the parabola is

excluded. The diagonal curves are the dependence of σ×BR on cτ for different values

of vΣ. This figure was made by our collaborator G. Ovanesyan.

on Λ/β in the right panel is weaker with smaller values of vΣ. This is because from

Eq. 4.22 we see that for a given value of Λ/β, a smaller value of vΣ means larger value

of cτ , which may fall below the ATLAS exclusion curve in Fig. 4.7.

4.5 More on the UV completion

Our analysis has been as model-independent as possible. But it is also inter-

esting to consider the possible dynamics which can generate the non-Abelian kinetic

mixing operator O(5)WX that we are interested in, and analyze the implications for the

Page 103: Complementarity of Symmetry Tests at the Energy and

86

ATLAS

mH+=mH2

=130 GeV

0.0 0.5 1.0 1.5 2.0 2.5 3.00.0

0.2

0.4

0.6

0.8

vS HMeVL

L�ÈΒÈHTeVL

mX=0.4 GeVmX=1.5 GeV

Fig. 4.8.— Constraints on the non-Abelian kinetic mixing model parameters, recasted

from the ATLAS results in Ref. [99]. The curves give the exclusion regions in the(vΣ,

Λ/β) parameter plane for mX = 0.4 GeV (the red region) and mX = 1.5 GeV (the

yellow region). This figure was made by our collaborator G. Ovanesyan.

current and future LHC searches. We have shown in Fig. 4.1 some possible diagrams

that generate this mixing operator, which involves fermion loops, scalar loops, and

non-perturbative dynamics. For a new vector-like fermion F with mass MF , the di-

mensional analysis estimates that Λ/β ∼ 16π2MF/y, where y is the FFΣ coupling,

and we took the gauge couplings to be O(1). The fermion needs to be vector-like

because it can have gauge invariant mass term without the Higgs which means it

can be arbitrarily heavy. If the fermion F is sufficiently light, it would likely have

been observed because it carries the SU(2) charge. In Ref. [100], the non-observation

Page 104: Complementarity of Symmetry Tests at the Energy and

87

of the pairs of new charged particles like the vector-like leptons at LHC indicates a

lower bound of MF>∼ 200 GeV, which translates to the bound of Λ/β >∼ 3.2 TeV for

y ∼ O(1). To reach this level of sensitivity, we need much larger integrated luminosity

or a search which analyzes the final state gauge boson reconstruction. In the scalar

case, for a new electroweak scalar S with mass MS, the similar dimensional analysis

gives Λ/β ∼ 64π2M2S/aS, where aS is the SSΣ coupling which has the dimension of

mass. To evade the LEP II limits, we assume MS>∼ 100 GeV and take aS ∼MS. This

also gives Λ/β ∼ 6.4 TeV. But aS can be a few times larger than MS. Therefore, we

can anticipate the Λ/β to be the upper end of the exclusion region in Fig. 4.8.

The dynamics in Fig. 4.1 are all loop or non-perturbative processes. The reason

why they are more important than the tree-level processes is the following. It is

possible that the charged triplet state H+ decay to a pair of scalar mediators S at

tree-level. So in order to have the H+ → SS tree level decay not allowed and the

H+ → W+X decay dominate, if there is a mediator S in the loop in Fig. 4.1, the mass

of S needs to be greater than half of the triplet mass.

In our foregoing analysis, we discussed two cases. The first case is mH+ =

130 GeV, which satisfies this constraint of mass and it is also what we used in Fig. 4.1

when recasting the ATLAS results. So in this case, we do not need to worry about

the on shell decays to the S mediators. The second case is mH+ = 300 GeV. It was

used in some of our branching ratio plots. But we did not use it to do the recast in

Fig. 4.1 in deriving our bounds on Λ/β. Without the tree-level process, we can also

see that our assumption of a ∼ 100% branching of H+ → W+X is consistent.

For an internal particle mass near 100 GeV, we may be near the border of the

region of a valid pure effective theory to treat the collider phenomenology. To do

Page 105: Complementarity of Symmetry Tests at the Energy and

88

a more quantitatively realistic study, we need to either study an explicit model that

generates the mixing operator coefficient or include a form factor. This is similar to the

application of the Higgs effective theory when studying the Higgs boson observables.

For example, Ref. [101] discussed in the case of Standard Model di-Higgs production

together with an additional high pT jet. Currently, the lepton jet reconstruction

efficiency of ATLAS peaks at the region close to pXT ∼ 40 GeV, which is larger than

the masses of the intermediate H±/H2 scalars and the W -bosons in the final state.

So in a more realistic model-dependent analysis, we would expect a little degradation

of the signal strength in the collider. However, we still consider our results about

the current 8 TeV LHC reach as more an indication of the sensitivity than a definite

quantitative result.

As a summary, we see that the mixing between the SU(2)L and the dark U(1)′

gauge groups, which is mediated by the non-Abelian kinetic mixing operator O(5)WX ,

can provide a natural small value of the mixing parameter ǫ, whose value is at the

scale of vΣ/Λ if we have the Wilson coefficient β at the order of O(1). This model

has very distinctive collider phenomenology, because the O(5)WX mediated production

of X dominates the production of all final states when Λ/β >∼ 1 TeV at both the

LHC energy of√s = 8 TeV and

√s = 14 TeV. The current ATLAS bounds, which is

based on the inclusive search for the displaced vertices and lepton jets, can be recast

to our case and indicates an exclusion of Λ/β up to about several hundreds of GeV,

depending on the value of the triplet vev vΣ and the dark photon mass mX . The

future collection of additional data of the ATLAS run will extend the reach of the

search. The additional search criteria associated with the vector bosons in the final

state can help to distinguish this non-Abelian mixing from other scenarios.

Page 106: Complementarity of Symmetry Tests at the Energy and

89

Chapter 5

QCD corrections for dark matter effective

interactions

Page 107: Complementarity of Symmetry Tests at the Energy and

90

5.1 Purpose of this study

As shown in the Section. 1.10, it is useful to study the dark matter operators

in a model independent way, using the effective theory, and it is also important to

study the effect of loop corrections and the mixing of dark matter operators. In the

following, we study the short-distance effective operators and the beyond leading order

QCD effects from renormalization group running and mixing of these operators. These

short-distance QCD effects may be useful in extracting the dark matter couplings, to

consider ahead the era of precision dark matter phenomenology after discovery and

identification of the dark matter’s particle nature and interaction features. While new

physics may be generated at a high scale and probed at the collider scale, the direct

detection of dark matter may probe the Wilson coefficients of the effective operators

at a different low scale. Since we have such different scales and we may like to compare

experiments results in these scales, we should consider the effect of the variations of

corresponding interactions due to QCD evolution. The one-loop anomalous dimension

matrix due to QCD corrections of the dark matter effective operators have been studied

in Refs. [56, 57, 72, 102].

In this study, besides doing the full one-loop QCD corrections for all the possible

effective operators, we consider the following two additional issues with short-distance

QCD corrections. First, the QCD running of one set of effective operators, which

was not considered in previous studies. This issue involves dark matter interactions

with quark scalar and pseudoscalar densities. In many of these interactions in the

dark matter models, there are explicit factors of quark Yukawa couplings, which leads

Page 108: Complementarity of Symmetry Tests at the Energy and

91

to quark mass factor in the operators, as shown in Refs. [52, 53, 54, 55, 58]. In the

following, we discuss a scenario in which such factor suppression in the Yukawa does

not appear. We will also show that the differences in the QCD evolution of the two

scenarios can be noticeable.

The second additional issue is more subtle and it is about the fact that one can

factorize the Standard Model and dark matter components of the effective operators

when computing the αs corrections. This issue involves dark matter fields which are

charged in the Standard Model electroweak gauge group [103, 104]. In this scenario,

the Wilson coefficients of the dark matter operators can have substantial contribu-

tions from box graphs with the exchange of two standard model gauge bosons. From

Ref. [105], we see that the QCD corrections need not factorize for low scale semi-

leptonic weak processes. In this case, we have to calculate the αs corrections by

computing the entire box graphs. In this work, we compute these entire box graphs in

scenarios where the dark matter fields are Dirac fermions, Majorana fermions, inelastic

fermions, and scalars. Our results show that there are non-factorizable contributions,

which corresponds to a matching correction when integrating out the standard model

gauge bosons. Although the matching correction is not in general universal, it still

involve a universal factor of the form (1−αs/π). This correction does not apply when

the Wilson coefficient is generated by the tree-level interactions.

5.2 The effective operators

Here we study the set of operators as our basis shown in Table 5.1. They are

from Ref. [53], with the addition of operators involving scalar and pseudoscalar quark

operators having the factor of quark mass mq in the coefficient.

Page 109: Complementarity of Symmetry Tests at the Energy and

92

i OΓi ni

1 mqχΓχ qq 7

2 mqχΓχ qγ5q 7

3 χΓχ qq 6

4 χΓχ qγ5q 6

5 χΓαχ qγαq 6

6 χΓαχ qγαγ5q 6

7 χΓαβχ qσαβq 6

8 αsχΓχ (Gaαβ)

2 7

9 iαs χΓχGaαβG

aαβ 7

Table 5.1: Operator basis and their corresponding dimensions.

The effective Lagrangian, written in terms of the operator basis, is

Leff =∑

i

CΓi (µ)

Λni−4OΓ

i (µ) + h.c., (5.1)

where Λ is the scale at which the heavy intermediate particles are integrated out, which

is taken to be of the order of electroweak scale or higher. µ is the renormalization

scale and CΓi (µ) is the Wilson coefficient at scale µ. Γ denotes the scalar, pseudoscalar,

vector, axial vector, and tensor bilinears. For fermion dark matters, they are the Dirac

matrix form χΓχW for Γ = 1, γ5, σµν , γµ, γµγ5, respectively. For scalar dark matters,

these bilinears become χ†χ form, and only appear for i = 1− 4, 8, 9 in Table. 5.1.

In Table. 5.1, for the scalar and pseudo-scalar operators (i = 1 − 4), we have

included both the operators with and without the quark mass operator. To our knowl-

edge, the literature does not typically consider the scalar DM-quark effective operators

without quark mass factor [52, 53, 54, 55, 58]. To motivate the operators without quark

mass factor (O3 and O4 in Table. 5.1), we consider an illustrative scenario, which in-

volves the SU(2)L fermion doublets Ψ(1),Ψ(2). To make the Lagrangian U(1) invariant,

they should carry hypercharge ∓1/2. We also consider a gauge singlet fermion field

Page 110: Complementarity of Symmetry Tests at the Energy and

93

χR. We can then write down the Lagrangian, which has the following dimension four

operators that can generate the S and P operators without the quark mass factor:

LDM = ǫij

(

c1 Ψ(1)i dR Qj χR + c2 Qi dR Ψ

(1)j χR

)

+ǫij

(

c3 Ψ(2)i uR Qj χR + c4 Qi uR Ψ

(2)j χR

)

⊃(c14− c2

2

)

Ψ(1)1 χR

(

dd+ dγ5d)

+(c34− c4

2

)

Ψ(2)2 χR (uu+ uγ5u) , (5.2)

where Q is the Standard Model quark SU(2)L doublet. Ψ(1)i and Ψ

(2)i (i = 1, 2) are the

components of the SU(2)L doublets Ψ(1) and Ψ(2). Ψ(1)1 is neutral and Ψ

(1)2 is charged,

while Ψ(2)1 is charged and Ψ

(2)2 is neutral. The neutral components of Ψ(1), Ψ(2), and

χR, can mix into the neutral mass eigenstates:

χ0j = Nj1Ψ

(1)1 +Nj2Ψ

(2)2 +Nj3χR (5.3)

where j = 1, 2, 3. The parameters leading to the Nj1,2,3 must be chosen to make

the DM-Z boson couplings small enough to evade the direct detection limits on the

Z-exchange cross section. The lowest mass eigenstate χ01 can be the dark matter

candidate, since its stability is guaranteed by using a Z2 symmetry. We can then invert

Eq. 5.3 and substitute it into Eq. 5.2 to obtain the interaction operator between DM

and quarks. Note that in the operators obtained, there is no quark mass factor. These

are exactly the form of operators O3 and O4 in Table. 5.1. Therefore, we include O3

and O4 in our set of operators.

We next consider the effective Lagrangian in Eq. 5.2 at low energies, which ap-

pears due to the renormalization group running associated with next-to-leading order

QCD corrections, and the possible phenomenological consequences of the corrections.

Page 111: Complementarity of Symmetry Tests at the Energy and

94

5.3 Loop corrections and the anomalous dimension matrix

To calculate the anomalous dimension matrix, we first compute the one loop

QCD corrections to the operators in Table. 5.1. There are seven possible diagrams

for the one loop correction, as shown in Fig. 5.1. The diagrams with two gluons in

the final state also include the crossed graphs and the symmetry factors. We then

extracted the ultraviolet poles of these diagrams using dimensional regularization.

From the operators in Table. 5.1, we have the following Feynman Rule for the

χχgg vertex is shown in Fig. 5.2.

Extracting the divergent term from the diagrams, we have

iMdivergent =g2

16π2ǫ

1

(µ2)ǫ· 13(χΓ1χ) (qγ

µγνΓ2γνγµq) , (5.4)

where ǫ ≡ 2− d/2, and µ is the energy scale.

By inserting Γi = 1, γ5, σµν , γµ, γµγ5 in the above result, we have

O1,3 → g2

16π2ǫ

1

(µ2)ǫ

(

16

3O1,3

)

,

O2,4 → g2

16π2ǫ

1

(µ2)ǫ

(

16

3O2,4

)

,

O5 → g2

16π2ǫ

1

(µ2)ǫ

(

4

3O5

)

,

O6 → g2

16π2ǫ

1

(µ2)ǫ

(

4

3O6

)

,

O7 → 0. (5.5)

We did similar calculation to other diagrams and other operators.

The way to calculate the anomalous dimension matrix is similar to Section. 2.2,

so we do not repeat the procedure here. The result of the anomalous dimension matrix

Page 112: Complementarity of Symmetry Tests at the Energy and

95

χ

χ

q

q

g

χ

χ

g

g

χ

χ

q

q

χ

χ

g

g

χ

χ

g

g

χ

χ

g

g

χ

χ g

g

Fig. 5.1.— The diagrams for the one loop QCD corrections to the operators in Ta-

ble. 5.1. The grey dot represents insertion of the operators in Table. 5.1.

Page 113: Complementarity of Symmetry Tests at the Energy and

96

is

γO =αs

4πγ(0), where

γ(0)33 = γ

(0)44 = −6CF , γ

(0)77 = 2CF ,

γ(0)81 = γ

(0)92 = 24αsCF , (5.6)

where for QCD CF = 4/3. The other elements of the anomalous dimension matrix

are zero. All of the anomalous dimension entries except for entries for the O3 and O4

have been listed in Ref. [72], and for those listed, we have exact agreement except for

the tensor operator anomalous dimension. This is because the tensor operator in [72]

is defined differently from us: imq qσµνγ5q.

Once we have the anomalous dimension matrix, we can know the running of the

Wilson coefficients by solving the equation

dC

d lnµ= γTO · C (5.7)

where C is the vector of the Wilson coefficients and µ is the renormalization scale.

From Eq. 5.6, we see that most of the entries of the anomalous dimension matrix are

zero, so it is relatively straightforward to solve the renormalization group equation

analytically. When solving the equations, we recall that there are additional quark

mass factors in operators O1 and O2, and the entire mq qq and mq qγ5q do not run.

We also need to consider the scale dependence of the αs coupling in operators O8 and

O9. The effect of this scale dependence is that there is no diagonal running of αsGG

Page 114: Complementarity of Symmetry Tests at the Energy and

97

and αsGG. The full solution to the RG equations is

c1,2(µ) = c1,2(Λ) + c8,9(Λ)Y (r),

c3,4(µ) = c3,4(Λ)X(r),

c5,6,8,9(µ) = c5,6,8,9(Λ),

c7(µ) = c7(Λ)S(r) . (5.8)

where r is the ratio as r = αs(µ)/αs(Λ), and the functions X, Y, S are defined as

X(r) = r3CFβ0 ,

Y (r) = −12CF

β0(αs(µ)− αs(Λ)) ,

S(r) = r−CF

β0 . (5.9)

From the above solution, we see that the operators with quark vector current O5 and

axial vector current O6 do not run, and the operators O8 and O9 with the square

of gluon fields αsGG do not run either. We also see that the operators with quark

scalar and pseudoscalar part and explicit of quark mass factor O1 and O2 mix with

the gluonic operators O8 and O9 when running, while the scalar and pseudoscalar

operators without the quark mass factor O3 and O4 and the tensor operators O7 run

diagonally and do not mix with other operators.

To get more ideas about the effect of the running of the Wilson coefficients, we

show in Fig. 5.3 the dependence of the functions X(r), Y (r), S(r) on the scale Λ from

the new physics scale Λ ∼ 100GeV to the lowest perturbative scale ∼ 1GeV, where

we fix the low energy scale µ = 1 GeV. From the S(Λ) curve, we see that the running

of the tensor operator O7 is mild. And from the X(Λ) curve, we see that the diagonal

running of the scalar and pseudoscalar operators with quark mass factor O3 and O4

Page 115: Complementarity of Symmetry Tests at the Energy and

98

can be significant when Λ runs from low to above 100 GeV. A similar significant effect

also happens to the mixing of the scalar and pseudoscalar operators without quark

mass factor O1 and O2 into the gluonic operators.

5.4 Phenomenological effects of QCD corrections

We can then use the result of the evolution to do the phenomenological analyses.

Here we focus on the direct detection experiment results. The running and mixing

can also affect the collider phenomenology, but a robust collider study needs to con-

sider more of other higher-order corrections, like the corrections in soft and collinear

resummations. So we will not do the collider study here.

For the direct detection experiments to search for dark matter, the relevant

scale is µ ∼ q < mπ, which is the scale of the momentum transfer in the elastic

scattering between a dark matter particle and a nucleus. We do not evolve the Wilson

coefficients at this scale for two reasons. First, in this scale, the running of the strong

couplings is no longer perturbative. Second, to evaluate the cross sections, we match

our operators onto nucleon level operators at the scale 1 GeV, but at lower scales, the

nucleons, instead of the quark operators, become the effective degrees of freedom. To

have an idea of the theoretical uncertainty of the result, we vary the low scale in a

range µ = 1− 2 GeV. As we see in Fig. 5.3, the Wilson coefficients can change by an

order of 50% depending on the heavy scale Λ. To illustrate the running and mixing

effects, we consider the effect of running on the direct detection cross section σ from

the operators O8,9. At LO, there is no direct effect from O8,9 since they do not run

diagonally. But at NLO, they contribute to the cross section directly as well as by

mixing into O1,2.

Page 116: Complementarity of Symmetry Tests at the Energy and

99

χ

χ

p

q

a, µ

b, ν

= 4iδab(

gaνgβµ − gµνgαβ)

pαpβχχǫaµǫ

Fig. 5.2.— The Feynman rule for the χχgg vertex.

20 40 60 80 100-1.0

-0.5

0.0

0.5

1.0

1.5

2.0

Λ [GeV]

X,Y

,S

X

Y

S

Fig. 5.3.— The dependence of the functions X(r), Y (r), S(r) on the new physics scale

Λ, with r = αs(µ)/αs(Λ), in the RG equation solutions in Eq. 5.8 and Eq. 5.9. Here

we used the low energy scale µ = 1 GeV.

Page 117: Complementarity of Symmetry Tests at the Energy and

100

To be concrete, we take the fermionic dark matter and choose Γ = 1 as an

example. To evaluate the cross sections, we first consider the operators O1,2,8,9 and

match them into the χ-nucleon interactions:

LNeff = χΓχ

(

CN1 NN + CN

2 Niγ5N)

, (5.10)

where the Wilson coefficients are given in Ref. [106]:

CN1 =

q=u,d,s

cq1mN f(N)Tq

+2

27fTG

·(

q=c,b,t

cq1mN − 12πcg8mN

)

,

CN2 =

q=u,d,s

mN [(−i)(cq2 − Ctot) + 4πicg9m] ∆(N)q ,

(5.11)

where m = (1/mu+1/md+1/ms)−1, and Ctot =

q cq2 m/mq. The numerical values of

the constants f(N)Tq

, f(N)TG

,∆(N)q are also given in Ref. [106]. There are some differences

between Eq. 5.11 and the corresponding equations Eqs. (45a) and (45b) in Ref. [106].

This is because we included the quark mass factor in our definition of O1,2, and our

overall normalization factor in operators O8,9 are also different from the factors in

Ref. [106]. Moreover, Ref. [106] does not have the prefactor of 1/2 in the definition of

the dual tensor Gµνa = (1/2)ǫµνρσGa

ρσ.

From the above χ-nucleon Lagrangian and the solution to the RG equation, we

can calculate the ratio of the direct detection cross sections σNLO/σLO. Here σLO is

the leading order DM-nucleus cross section generated by O8,9, and σNLO is the next-

to-leading order cross section, which includes the mixing of O8,9 into O1,2. We assume

that at the new physics scale Λ, there are only O8,9. At low scale µ, the O8,9 mix into

O1,2.

The results for OΓ=18 and OΓ=1

9 are that for these operators, the overall effect is

Page 118: Complementarity of Symmetry Tests at the Energy and

101

modest: roughly a 10-15% increase (decrease) in the OΓ=18 (OΓ=1

9 ) with mild depen-

dence on the DM mass at it varies in the range 20 GeV − 1TeV . We have used for

Λ the lower bound on the new physics scale derived from LHC monojet data Λ(mDM)

from Ref. [], hence the mild dependence of σNLO/σLO on mDM, with the effect becom-

ing smaller as mDM approaches 1TeV. The range of the σNLO/σLO corresponds to

varying the low scale µ in 1− 2GeV. So we have c1(Λ) = 0 and

c1,2(µ) = c1,2(Λ) + c8,9(Λ)Y (r) = c8,9(Λ)Y (r)

c8,9(µ) = c8,9(Λ). (5.12)

And the Wilson coefficients in the χ-neucleon Lagrangians at different scales are

CN1 (Λ) = − 2

27fTG

· 12πc8(Λ)mN

CN1 (µ) =

q=u,d,s

c1(µ)mN f(N)Tq

+2

27fTG

(

q=c,b,t

c1(µ)mN − 12πc8(µ)mN

)

(5.13)

And then we have the ratio

σNLO

σLO=

CN1 (µ)

CN1 (Λ)

2

. (5.14)

By plugging in CN1 (µ) and CN

1 (Λ) from Eq. 5.13, we can calculate the ratio σNLO/σLO

numerically as a function of the high scale Λ for operators O8,9. We then use the

lower bound on the new physics scale from the LHC monojet result from Ref. [58] as

a function of Λ(mDM), to derive the numerical relation between the ratio σNLO/σLO

and the dark matter mass mDM. Our result shows a modest dependence of the cross

section ratio on the dark matter mass: a variation of mDM in the range 20 GeV−1TeV

causes approximately a 10-15% increase (decrease) in operator O8 (O9), with the effect

becoming even milder as mDM approaches 1TeV. In this numerical study, we varied

the low scale µ in 1− 2GeV.

Page 119: Complementarity of Symmetry Tests at the Energy and

102

We then consider the operators O3,4 as another illustration. From Eq. 5.8 and

Fig. 5.3, we see that there is a significant diagonal running effect on the cross section,

which is described by the factor X(r)2. To our knowledge, currently there is no LHC

bound for these operators. However, one would have a stronger bound on O3,4 than

O1,2, for which the scale Λ is very loosely bounded from LHC for Λ >∼ 20−30GeV [58].

Here we take Λ in 10 − 100 GeV as an example. We make a numerical plot of the

ratio σNLO/σLO as a function of the new physics scale Λ, which is shown in Fig. 5.4.

For O3,4, the NLO cross section can be nearly three times larger than the LO cross

section. The nucleon matrix elements cancel in the ratio of the cross sections. So the

entire ratio in Fig. 5.4 is given by only X(µ,Λ)2, as can be seen from the RG running

Eq. 5.8.

We can also study the effect of the operator running on the dark matter relic

density constraints. Here we again consider the operators O3,4 because their diagonal

running effects are the most significant. We take the low energy scale µ to be the order

of the dark matter mass, since the energy released to the dark matter annihilation

products is governed by this energy scale. More specifically, we take µ = mDM and

µ = mDM/2 as two examples. We can write the LO and the NLO annihilation cross

section in the following form

〈σv〉LO = f(mDM)/Λ4,

〈σv〉NLO = f(mDM)/Λ4 ×X[r(mDM,Λ)]2, (5.15)

where the v is the relative velocity of the dark matter particles, and the function

f(mDM) depends on the type of the operator (OS,P3 , OS,P

4 ). It is given in Appendix A

of Ref. [59]. Because we did not include the quark mass factor in operators O3,4, we

Page 120: Complementarity of Symmetry Tests at the Energy and

103

need to remove a factor of m2q/Λ

2 from the equations of f(mDM) there.

From Eq. 5.15, we have

ΛLO =

(

f(mDM)

〈σv〉LO

) 14

, (5.16)

and we can solve ΛNLO numerically in the same way.

We then solve the scale of new physics ΛLO and ΛNLO in terms of µ and mDM

by setting the LO or NLO order annihilation cross sections 〈σv〉LO and 〈σv〉NLO to be

equal to the cross section corresponding to the relic abundance of the dark matter,

which is 3× 10−26cm/s. The ratio of ΛLO and ΛNLO for the operators OS3 and OP

3 is

shown in Fig. 5.5. The curves for operators type OP3 and OP

4 , which are also identical to

each other, are approximately larger by a factor of 1.025 from the curves in Figure 5.5.

5.5 Box graph corrections and factorizability

Next we study the QCD corrections in the case when dark matter is charged

under SU(2)L ⊗U(1)Y . The effective operators in Section 5.3 may be generated from

the box graphs with electroweak gauge bosons as the intermediate fields. These graphs

shown in Fig. 5.6. This makes it necessary to consider the QCD corrections at the

scale of mW . These corrections will affect the evolution of the operators from the

electroweak scale to the hadronic scale, so they will be different from the corrections

in Section 5.3.

To be concrete, we consider the Standard Model with an additional electroweak

n-tuplet field. The neutral component of the n-tuplet is the lowest mass state, and can

thus be a candidate of the dark matter particle [103, 104]. Here we consider both the

scalar and fermion dark matter fields. For the fermion dark matter, we also consider

Page 121: Complementarity of Symmetry Tests at the Energy and

104

20 40 60 80 1000.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

Λ [GeV]

σN

LO/σ

LO

μ = 1 GeV

μ = 2 GeV

Fig. 5.4.— Ratio of the NLO order to LO DM-nucleon cross sections from the opera-

tors O3 and O4. There is no quark mass factor in operators O3 and O4.

10 15 20 25 301.00

1.05

1.10

1.15

1.20

1.25

mDM [GeV]

ΛN

LO

/ΛL

O

μ = mDM/2

μ = mDM

Fig. 5.5.— The QCD running effect on the relic abundance curve for the operators

O3 and O4. There is no quark mass factor in operators O3 and O4.

Page 122: Complementarity of Symmetry Tests at the Energy and

105

the Dirac, Majorana, and the inelastic dark matter. We do the box graph calculation

for each of these cases, respectively.

For the QCD corrections for the loop mediated χ-quark interactions, we focus

on the box graphs with gauge bosons as intermediate fields. Ref. [105] studied the

graphs of the semileptonic weak interactions. These graphs have a non-factorizable

pattern. This suggests that one should explicitly check whether the graphs in our case

have a non-factorizable pattern or not. The non-factorizability is what we are going to

study. The loop interactions mediated by other fields such as Higgs bosons have been

studied and the corresponding QCD corrections for the spin-independent interactions

are given in Ref. [75]. Next we consider the box graphs for each of the dark matter

types.

5.5.1 Fermion dark matter

The loop level interactions of dark matter with quarks are more phenomeno-

logically favorable than the tree level interactions because of the strong constraints

from the direct detection experiments of dark matter. In this part we consider the

case when the dark matter is a fermion field which is charged under the SM gauge

group. It can be achieved by taking the hypercharge Y = 0 or allowing Y 6= 0 for

both Majorana fermion or inelastic fermion dark matter. Some examples are given in

Ref. [65]. Fig. 5.6 shows the W and Z box graphs which generate the leading effective

χ-quark Lagrangian. The leading effective Lagrangian, separating the spin-dependent

Page 123: Complementarity of Symmetry Tests at the Energy and

106

χ χχ

qq

q

W,Z W,Z

χ χχ

qq

q

W,Z W,Z

Fig. 5.6.— Box graphs which describe the χ-quark interactions at the one loop level.

Page 124: Complementarity of Symmetry Tests at the Energy and

107

and independent terms, is in the following form:

LW/Zeff =

q=u,d,s

dq χγλγ5χ qγλγ5q + fqmq χχ qq + hq χγλχ qγ

λq

+g(1)q

χi∂µγνχOqµν +

g(2)q

m2χ

χ(i∂µ)(i∂ν)χOqµν , (5.17)

where the effective coefficients dq, fq, hq, g(1)q , g

(2)q are from the loop integration, and

the DM field χ can be either Dirac or Majorana fermions. The Oqµν is the twist-2

operator defined in Ref. 65:

Oqµν ≡ 1

2qi

(

Dµγν +Dνγµ −1

2gµν /D

)

q. (5.18)

5.5.1.1 Dirac dark matter

In the Dirac case, we consider an n-tuplet field and the dark matter is the neutral

component. So it has T 3 = −Y . The corresponding Lagrangian is

LDM = χ(

i /D +M)

. (5.19)

Then we can derive the Feynman rules shown in Fig. 5.7. Using those Feynman

rules, we can calculate the box graphs in Fig. 5.6. The left graph in Fig. 5.6 with W

as the intermediate particle as an example, can be written down as

iM =

d4k

(2π)4ig√2χγµT− i(/p2 + /k +M)

(p2 + k)2 −M2

ig√2γνT+χ

· ig√2uLγ

µ′ i(/p4 − /k)

(p4 − k)2ig√2γν

uL · −igµµ′

k2 −m2W

−igνν′(p1 − p2 − k)2 −m2

W

. (5.20)

And then do the loop integrals and the matrix contractions. We can do the same

thing for other graphs and the Z bosons. The result is the effective Lagrangian shown

in Eq. 5.17, and the coefficients are

Page 125: Complementarity of Symmetry Tests at the Energy and

108

χ χ

W+µ

=ig√2γµT−

χ χ

W−µ

=ig√2γµT+

χ χ

= ieQγµ

χ χ

=ig

cos θW

(

T 3 − sin2 θWQ

)

γµ

Fig. 5.7.— Feynman rules for the Dirac DM and gauge boson interaction.

Page 126: Complementarity of Symmetry Tests at the Energy and

109

dLOq =

n2 − (4Y 2 + 1)

4

α22

m2W

gAV(w)

+8[

(aVq )2 + (aAq )

2]

Y 2

cos4 θW

α22

m2Z

gAV(z), (5.21)

fLOq =

4[

(aVq )2 − (aAq )

2]

Y 2

cos4 θW

α22

m3Z

gS(z), (5.22)

hLOq = ± Y α2

2

4m2W

gV(w), (5.23)

g(1),LOq =

n2 − (4Y 2 − 1)

4

α22

m3W

gT1(w)

+8(

(aVq )2 + (aAq )

2)

Y 2

cos4 θW

α22

m3Z

gT1(z),

g(2),LOq =

n2 − (4Y 2 − 1)

4

α22

m3W

gT2(w)

+8(

(aVq )2 + (aAq )

2)

Y 2

cos4 θW

α22

m3Z

gT2(z), (5.24)

where the ratios are defined as w = m2W/m

2χ, z = m2

Z/m2χ; and the couplings aVq =

12T3q − Qq sin2 θW , a

Aq = −1

2T3q are the vector and axial-vector couplings of quarks

with Z boson, respectively; the + and − signs correspond to the up and down type

quarks, respectively.

Our result of the functions gAV(x) and gS(x) agree exactly with Ref. [65]. To our

knowledge, the function gV(x) is new, which is

gV(x) =

√x[2 + (2− x)x]

2 bxarctan

(

2bx√x

)

+2− 2x+ x2 ln x

2, (5.25)

where bx =√

1− x/4. For completeness, we also list the results for the following

Page 127: Complementarity of Symmetry Tests at the Energy and

110

functions

gAV(x) =

√x(8− x− x2)

24bxarctan

(

2bx√x

)

− x [2− (3 + x) ln x]

24,

gS(x) =(4− 2x+ x2)

4bxarctan

(

2bx√x

)

+

√x (2− x ln x)

4.

gT1(x) =1

3bx(2 + x2) arctan

(

2bx√x

)

(5.26)

+1

12

√x (1− 2x− x(2− x) ln(x)) ,

gT2(x) =1

4bxx(2− 4x+ x2) arctan

(

2bx√x

)

,

−1

4

√x (1− 2x− x(2− x) ln(x)) .

We then proceed to find the next-to-leading QCD corrections, which is the order

O(αs), to the leading results of the coefficients. We use the method explained in

Ref. [105], which is to express the amplitudes of the box graphs in terms of the time

ordered products of two currents. The terms, which can be written in the form of

the Ward identity receive the O(αs) correction. For illustration purpose, we show

the procedure of computing the O(αs) correction for the W exchange box graph. For

the other graphs and other DM scenarios, we give our final answer for the correction

directly.

Written in terms of the currents, the W exchange box graph in Fig. 5.6 can be

written as

M(±)box

W =g4

4N∓

d4l

(2π)41

(l2 −m2W + iε)2

· uχγµ1

/l + /k −mχ + iεγν uχ T

±µνW (l),

(5.27)

where k is the momentum of the incoming dark matter particle. The superscripts ±

means the charges of the dark matter component in the dark matter multiplet. If the

quark is up-type, then M(+)box

W is the amplitude of the upper graph in Fig. 5.6, and

Page 128: Complementarity of Symmetry Tests at the Energy and

111

M(−)box

W is the lower graph. If the quark is down-type, then M(+)box

W is the amplitude

of the lower graph in Fig. 5.6, and M(−)box

W is the upper graph. The factor N± is

defined as

N± ≡(

T± · T∓)

χ0χ0=n2 − (1± 2Y )2

4, (5.28)

where T± is the raising and lowering generator that acts on the dark matter multiplet.

The TW (l) is defined as

T±λρW (l) =

d4x eilx〈p′T(

J±λW (x)J∓ρ

W (0))

〉p, (5.29)

and the current J±µW (x) is

J±µW (x) = QT±γµPLQ, (5.30)

where Q is the quark SU(2)L doublet. We then use the identity

γµγνγρ = γµgνρ + γρgµν − γνgµρ − iǫσµνργσγ5 ,

to rewrite the γµΓγν terms in terms of single γ matrices, and use the equal time

commutator identity

[

J±0 (x), J

∓ν (0)

]

x0=0= ±2J3

ν (x) δ(3)(~x) , (5.31)

and the result of rewriting Eq. 5.27 becomes

M(±)box

W =g4

4N∓

d4l

(2π)41

[l2 −m2W + iε]2

1

(l + k)2 −m2χ + iε

·{

uχγλuχ

[

± 4i〈p′J3λ(0)〉p+

(

kνgµλ + kµgνλ − (l + k)λgµν

)

T±µνW (l)

]

+[

mχuχγµγνuχ + iǫσµνα(l + k)αuχγσγ5uχ

]

T µνW (l)

}

.

(5.32)

Page 129: Complementarity of Symmetry Tests at the Energy and

112

By summing up the two graphs in Fig. 5.6, we have

M(+)box

W +M(−)box

W =4Y 2 + 1− n2

2

α22

m2W

gAV(w) uχγλγ5uχ uγλPLu

± α22 Y

2m2W

uχγλuχ uγλPLu

[

g1V (w) + g2V (w)]

+3α2

2

2m3W

n2 − 1− 4Y 2

4uχuχ u /p PLu [g1T(w) + g2T(w)] ,

(5.33)

where p is the momentum of the external quarks, and in the ± α22 Y

2m2W

term, the + sign

means the quark is up-type and the − means the quark is down-type. The functions

g1V,T, g2V,T are calculated as

g2V(x) =(2− x)

√x arctan 2bx√

x

bx+ x ln x ,

g2T(x) =

√x

6(2− x ln x) +

x(x− 2) arctan 2bx√x

6bx,

g1i(x) = gi(x)− g2i(x). (5.34)

Now we can figure out the O(αs) corrections. According to Ref. [105], the terms

in Eq. 5.32 which are proportional to the structure kνTµν(k), kµT

µν(k) do not receive

the O(αs) correction. These terms are identified as the g2V and g1T terms. The other

terms receive a factor of 1 − αs/π correction at next-to-leading order. They are the

gAV, g1V and g2T terms.

We then do the same calculation to the Z exchange box graphs in Fig. 5.6, and

Page 130: Complementarity of Symmetry Tests at the Energy and

113

the coefficients with the next-to-leading order correction are

dNLOq = dLO

q

(

1− αs

π

)

,

fNLOq =

4α22

[

(aVq )2 − (aAq )

2]

Y 2

m3Z cos4 θW

·[(

1− αs

π

)

g1S(z) + g2S(z)]

,

hNLOq = ± Y α2

2

4m2W

[(

1− αs

π

)

g1V(w) + g2V(w)]

,

g(1),NLOq = g(1),LO

q , (5.35)

g(2),NLOq = g(2),LO

q

(

1− αs

π

)

.

where the extra loop functions other than shown in Eq. 5.34 are defined as

g2S(x) = −g2T(x) ,

g1S(x) = gS(x)− g2S(x) . (5.36)

We can see the non-factorizable terms in Eq. 5.36, which are independent from the

(1− αs/π) factor.

5.5.1.2 Majorana dark matter

We then consider the case when we have a Majorana multiplet λ of the SM

gauge group. The dark matter particle χ is one of the Majorana components of this

multiplet. The corresponding Lagrangian is

LMajoranaDM = λiγµD

µλ/2. (5.37)

There is no tree-level coupling between the Z boson and the dark matter field, so

all the Z exchange box graphs in Fig. 5.6 vanish. Also, note a factor of 1/2 difference

between the Majorana Lagrangian Eq. 5.37 and the Dirac Lagrangian Eq. 5.19. This

is a difference in the standard normalization factors between Dirac and Majorana

Page 131: Complementarity of Symmetry Tests at the Energy and

114

effective Lagrangians. This factor brings an overall factor of 2 difference in the effective

coefficients compared to the Dirac dark matter. Considering these two effects, the

effective coefficients in this case are

dLOq =

n2 − (4Y 2 + 1)

4

α22

2m2W

gAV(w),

fLOq = hLO

q = 0,

g(1),LOq =

n2 − (4Y 2 + 1)

4

α22

m3W

gT1(w),

g(2),LOq =

n2 − (4Y 2 + 1)

4

α22

m3W

gT2(w). (5.38)

Following the same procedure for the O(αs) correction in Section 5.5.1.1, we

can compute the O(αs) correction for the Majorana case. What is different in the

Majorana case is that the Z box graphs do not contribute, and for the Majorana fields

the term χγµχ in the effective Lagrangian vanishes. So the results are

dNLOq = dLO

q

(

1− αs

π

)

,

fNLOq = hNLO

q = 0, (5.39)

g(1),NLOq = g(1),LO

q ,

g(2),NLOq = g(2),LO

q

(

1− αs

π

)

.

5.5.1.3 Inelastic dark matter

We also consider the case when there is a Dirac multiplet and the neutral com-

ponent of the multiplet is split into two Majorana states:

ψ0 =1√2(χ0 + iη0), (5.40)

where χ0 is the lightest mass state and can be the dark matter candidate. Because

the Majorana feature χγµχ = 0, the χχZ vertex does not exist. But the vertex χηZ

Page 132: Complementarity of Symmetry Tests at the Energy and

115

is allowed at tree level. Here we assume that the splitting among χ, η is large enough,

so that the Z exchange tree level inelastic scattering vanishes. The Z exchange box

graph with the χηZ vertex, however, is non-zero. In this case, the effective coefficients

are

dLOq =

n2 − (4Y 2 + 1)

4

α22

2m2W

gAV(w)

+8[

(aVq )2 + (aAq )

2]

Y 2

cos4 θW

α22

2m2Z

gAV(z), (5.41)

fLOq =

4[

(aVq )2 − (aAq )

2]

Y 2

cos4 θW

α22

2m3Z

gS(z), (5.42)

hLOq = 0,

g(1),LOq =

n2 − (4Y 2 − 1)

8

α22

m3W

gT1(w)

+4(

(aVq )2 + (aAq )

2)

Y 2

cos4 θW

α22

m3Z

gT1(z),

g(2),LOq =

n2 − (4Y 2 − 1)

8

α22

m3W

gT2(w)

+4(

(aVq )2 + (aAq )

2)

Y 2

cos4 θW

α22

m3Z

gT2(z), (5.43)

The above results agree exactly with Ref. [75]. We then use the same method

described in Section 5.5.1.1 to compute the O(αs) correction. The results are

dNLOq = dLO

q

(

1− αs

π

)

,

fNLOq =

4[

(aVq )2 − (aAq )

2]

Y 2

cos4 θW

α22

2m3Z

[(

1− αs

π

)

g1S(z) + g2S(z)]

,

hNLOq = 0 , (5.44)

g(1),NLOq = g(1),LO

q ,

g(2),NLOq = g(2),LO

q

(

1− αs

π

)

.

Page 133: Complementarity of Symmetry Tests at the Energy and

116

5.5.2 Scalar dark matter

We now consider a complex scalar dark matter. The corresponding Lagrangian

is

LDM = |Dµχ|2 −m2χ|χ|2. (5.45)

The W and Z exchange box graphs generate the effective Lagrangian in the

following form

LW/Zeff =

q=u,d,s

hq χ†i∂µχ qγ

µq + fqχ†χmq qq ++

g(2)q

m2χ

χ(i∂µ)(i∂ν)χOqµν ,

(5.46)

To our knowledge, our following results of the LO and NLO effective coefficients

are not computed by other people and are new:

hLOq = ∓ α2

2 Y

4m2W

gscalarV (w),

fLOq =

α22 Y

2[

(aVq )2 − (aAq )

2]

m2Z cos4 θW

gscalarS (z),

g(2),LOq =

α22

8m2W

n2 − (4Y 2 + 1)

4gscalarT (w)

+α22 Y

2[

(aVq )2 + (aAq )

2]

m2Z cos4 θW

gscalarT (z)] , (5.47)

where in hLOq , the − sign is for the up type quark and the + is for the down type

quark. The functions gscalarV , gscalar

S , gscalarT are

gscalarV (x) = −

(4− x)(1− x)√x arctan 2bx√

x

2bx− (3− x)x ln x+ 2x+ 4

2,

gscalarS (x) = x ln x+ 4 +

(4− x)(2 + x) arctan 2bx√x

bx√x

,

gscalarT (x) = x ln x− 2 +

(4− x)(2 + x) arctan 2bx√x

bx√x

. (5.48)

Page 134: Complementarity of Symmetry Tests at the Energy and

117

Similarly to the Dirac case, we compute the NLO QCD corrections to the coef-

ficients. The results are

hNLOq = ∓ α2

2 Y

4m2W

[(

1− αs

π

)

gscalar1V (w) + gscalar

2V (w)]

,

fNLOq =

α22Y

2[

(aVq )2 − (aAq )

2]

cos4 θW m2Z

·[(

1− αs

π

)

gscalar1S (z) + gscalar

2S (z)]

,

g(2),NLOq = g(2),LO

q

(

1− αs

π

)

,

(5.49)

where the loop functions are

gscalar2V = −

√x

2bx[8 + (x− 7)x] arctan

2bx√x+x[−2 + (x− 5) ln x]

2,

gscalar2S = x ln x− 4 +

√x(2− x) arctan 2bx√

x

bx, (5.50)

g1i ≡ gi − g2i, where i = V, S.

As a summary for this section, we considered fermion and scalar DM. For

fermion, we considered Dirac, Majorana, and inelastic cases. For each case, we calcu-

lated the LO effective Lagrangian from the W and Z box graphs and the NLO O(αs)

corrections using the method in Ref. [105]. In several cases we found a non-factorizable

O(αs) correction, which is well known in electron-quark interactions. This correction

occurs in the effective coefficients fq for Dirac, inelastic and scalar DM cases; in hq for

Dirac and scalar DM; and in g(2)q for all fermion DM and scalar DM. For fermion DM

and scalar DM, the coefficient dq has the correction in the form of an overall factor of

(1− αs/π).

Page 135: Complementarity of Symmetry Tests at the Energy and

118

Chapter 6

SUSY radiative corrections

Page 136: Complementarity of Symmetry Tests at the Energy and

119

6.1 Parameter scans in MSSM

From Section 1.12, we see that ∆CKM, ∆e/µ and the weak charge are very useful

in searching for the MSSM. So below we extend the numerical studies in Ref [78]. We

make broader constraints in the MSSM parameter space from values of ∆CKM and

∆e/µ, with additional constraints from the weak charge values and LHC constraints

on the sparticle masses. When doing the numerical calculations, we used the codes

by the authors of Ref. [78] and Ref. [79]. From Eq. 1.22 and Eq. 1.24, we see that

the experimental precision on ∆CKM and ∆e/µ can be O(10−4) - O(10−3). So if we

can find some region in the parameter space in MSSM which can make the correction

to ∆CKM and ∆e/µ to be of order O(10−4) - O(10−3), then it will be convenient for

experiments to exclude these parameter regions.

To evaluate the magnitude of the MSSM corrections to ∆CKM and ∆e/µ, we

scanned several variables in the MSSM parameter space, while we took random values

for other parameters. We scanned or randomized the following parameters: the super-

symmetric Higgs-Higgsino mass parameter µ, as defined in the following superpotential

for the MSSM [107]:

W = uyuQHu − dydQHd − eyeLHd + µHuHd, (6.1)

where the µ term is the supersymmetric version of the mass of the Standard Model

Higgs boson. The the gaugino mass parameters M1, M2, M3, M1 and M2 are the bino

and wino mass parameters, which are defined in the following soft supersymmetry

Page 137: Complementarity of Symmetry Tests at the Energy and

120

breaking interaction for gauginos in MSSM [107]:

Lsoft = −1

2

(

M3gg +M2WW +M1BB + c.c.)

+ ..., (6.2)

and if mZ ≪ |µ±M1| , |µ±M2|, then the neutralinos are nearly a bino-like, a wino-

like, and two higgsino-like mass eigenstates. We also scanned or randomized the ratio

between the up and down type Higgs v.e.v. tan β, the left and right handed slepton

mass m2L and m2

R, and the squark masses m2Q, m

2U , m

2D.

6.2 Correction results

We studied the dependence of ∆CKM on MSSM parameters M1, M2 and µ,

which are defined in the previous section. The dependence of ∆CKM and ∆e/µ on

other parameters, like the sfermion masses, have been studied by Ref. [78].

We varied M1, M2, and µ values, and studied how ∆CKM changes with them.

From Ref. [78], we see that to have large corrections, many parameters, like the second

generation slepton mass, should be small. So we took ml2= 120 GeV and M1 =M2 =

80 GeV when making the plots. The plots of ∆CKM versus M2 and µ are in Fig. 6.1.

From these figures we can see that ∆CKM are mostly O(10−4) and can reach the order

O(10−3) when M2 and µ are smaller than roughly 100 GeV. In Fig. 6.2, we plotted the

dependence of ∆CKM on M1, and we see that ∆CKM almost has a very mild dependence

on M1.

Fig. 6.3 shows the dependence of ∆CKM and ∆e/µ vs µ and M2 when µ = M2.

We see that both ∆CKM and ∆e/µ have similar behavior in terms of their magnitudes.

And again, for most regions, ∆CKM and ∆e/µ are mostly O(10−4) and can reach the

order O(10−3) when M2 and µ are smaller than roughly 100 GeV. We also studied

Page 138: Complementarity of Symmetry Tests at the Energy and

121

0 100 200 300 400 500

- 0.0014

- 0.0012

- 0.0010

- 0.0008

- 0.0006

- 0.0004

- 0.0002

0.0000

M 2 [GeV ]

�CKM

0 100 200 300 400 500

- 0.0014

- 0.0012

- 0.0010

- 0.0008

- 0.0006

- 0.0004

- 0.0002

0.0000

� [GeV ]

�CKM

Fig. 6.1.— Upper: ∆CKM vs M2, the parameter values are ml2= 120 GeV and

µ = M1 = 80 GeV. Lower: ∆CKM, the parameter values are ml2= 120 GeV and

M1 = M2 = 80 GeV. The resulting charginos are sufficiently heavy as to obey the

LEP limits.

Page 139: Complementarity of Symmetry Tests at the Energy and

122

the case when µ 6= M2, and the behaviors of ∆CKM and ∆e/µ are very similar to this

figure.

We then studied the effect of adding the constraints on the weak charge and the

collider constraints on the slepton masses. For the weak charge, we selected the data

sets where δ(QeW )SUSY/(Q

eW )S > 2.4% and the S, T , U parameters ranging in 2σ.

For the squark masses, we use the constraints from [109], which searched for

the squark and gluino final states characterized by high-pT jets, missing energy and

absence of electrons and muons at the energy of√s = 8 TeV and a luminosity of

20.3 fb−1 in LHC. For the chargino and slepton masses, we use the constraints from

Ref. [108], which searched for the neutralinos, charginos, and sleptons final states

characterized by the presence of two leptons and missing energy at the same energy

and luminosity in LHC.

After imposing the constraints from the weak charge and the sparticle masses,

the ∆CKM and ∆e/µ values are shown in the scatter plots in Fig. 6.4. For comparison,

we also show the points without constraints from the weak charge and the sparticle

masses. For those points, the sparticle masses are chosen to be random. From the

figure, we see that without the weak charge and sparticle mass constraints, ∆CKM and

∆e/µ can reach near O(10−3), while with these constraints, ∆CKM and ∆e/µ can only

reach O(10−4).

From the figures in this section, we can make the conclusion that the SUSY

correction to ∆CKM and ∆e/µ are or order O(10−4) for most parameter regions. Future

experimental measurements of them also need to reach the precision of O(10−4) in

order to have a reliable exclusion of the MSSM parameter space.

Page 140: Complementarity of Symmetry Tests at the Energy and

123

0 100 200 300 400 500

- 0.0014

- 0.0012

- 0.0010

- 0.0008

- 0.0006

- 0.0004

- 0.0002

0.0000

M 1 [GeV ]

�CKM

Fig. 6.2.— ∆CKM vs M1, the parameter values are ml2= 120 GeV and µ = M2 =

80 GeV.

Page 141: Complementarity of Symmetry Tests at the Energy and

124

0 100 200 300 400 500- 0.002

- 0.001

0.000

0.001

0.002

� =M 2 [GeV ]

� CKM

� e / �

M 1 = 500 GeV

ml

2

= 120 GeV

0 100 200 300 400 500- 0.002

- 0.001

0.000

0.001

0.002

� =M 2 [GeV ]

� CKM

� e / �

M 1 = 1 TeV

ml

2

= 120 GeV

Fig. 6.3.— ∆CKM and ∆e/µ vs µ and M2, where µ = M2. The difference between the

two figures is that the upper figure has M1 = 500 GeV, while the lower figure has

M1 = 1 TeV.

Page 142: Complementarity of Symmetry Tests at the Energy and

125

-0.0002 -0.0001 0.0000 0.0001 0.0002-0.0002

-0.0001

0.0000

0.0001

0.0002

Δe/μ

ΔCKM

-0.0010 -0.0005 0.0000 0.0005 0.0010-0.0010

-0.0005

0.0000

0.0005

0.0010

Δe/μ

ΔCKM

With constraints

No constraint

Fig. 6.4.— The ∆CKM and ∆e/µ scatter plots for parameters constrained by weak

charge and LHC results. The upper one shows only the constrained plot. The lower

one shows the comparison between the constrained plot and the plot with completely

random parameters.

Page 143: Complementarity of Symmetry Tests at the Energy and

126

Acknowledgement

We thank S. Bauman for his codes in calculating the ∆CKM and ∆e/µ values. We

also thank W. Chao in imposing the weak charge and LHC constraints, and thank S.

Su for her codes in imposing these constraints.

Page 144: Complementarity of Symmetry Tests at the Energy and

127

Chapter 7

Summary

As a summary, we considered several symmetries and interactions beyond the Standard

Model, and studied their phenomenology in both high energy colliders and low energy

experiments.

In particular, we studied the lepton number violation. We did this study because

the lepton number conservation is not a fundamental symmetry in Standard Model

(SM). The nature of the neutrino depends on whether or not lepton number is violated.

Leptogenesis also requires lepton number violation (LNV). To test LNV, we compared

the sensitivity of high energy collider and low energy neutrinoless double-β decay

(0νββ) experiments. We included the QCD running effects to obtain the constraints of

the effective coefficient in the operator at the LHC scale. We included the long-distance

contributions to nuclear matrix elements and matched the quark level operators to the

hadronic operators using the SU(2)L × SU(2))R chiral transformation properties of

the operators and obtained the 0νββ decay half life time. We also did the background

analysis for the collider search at both 14 TeV and 100 TeV. We used simplified models,

which allows for the case when one or more heavy particle goes on shell. Our result

Page 145: Complementarity of Symmetry Tests at the Energy and

128

shows that the reach of future tonne-scale 0νββ decay experiments generally exceeds

the reach of the 14 TeV LHC for a class of simplified models. For a range of heavy

particle masses at the TeV scale, the high luminosity 14 TeV LHC and tonne-scale

0νββ decay experiments may provide complementary probles. The 100 TeV collider

with a luminosity of 30 ab−1 exceeds the reach of the tonne-scale 0νββ experiments

for most of the range of the heavy particle masses at the TeV scale. These results can

show how the future colliders and low energy experiments can improve sensitivity and

do the complementary tests for LNV.

We then studied the non-Abelian kinetic mixing and its LHC signatures. We

introduced a U(1)′ vector boson to the Standard Model with an extra scalar triplet.

The U(1)′ vector boson can mix with the SU(2)L gauge boson via non-Abelian kinetic

mixing. The benefit of this scenario is that it can explain the smallness value of the

coupling ǫ naturally. We then studied the collider phenomenology of the non-Abelian

kinetic mixing operator. In particular, we studied the production rates of the charged

and neutral Higgs scalars and the dark photon. We see that in some regions, the Drell-

Yan pair production pp→ φφ dominates in the LHC production rates, while in other

regions the production of pp→ Xφ dominates. We also studied the decay width of the

charged scalar, and we see for some regions the branching ratio of H+ → W+X can be

nearly 100%. We then outlined the possible LHC signatures and recasted the current

ATLAS dark photon experimental results into our non-Abelian mixing scenario. Our

result shows that the exclusion can reach Λ/β to several hundreds of GeV depending

on the values of the dark photon mass and the triplet vev. The future experiments

may improve sensitivity and may distinguish this scenario from the Abelian mixing by

searching for the lepton jets and displaced vertices and the vector boson signatures.

Page 146: Complementarity of Symmetry Tests at the Energy and

129

We then studied the QCD corrections for dark matter effective interactions. we

studied the QCD running for a list of dark matter effective operators. This will be

useful to connect the new physics at high scale and the effect of the direct detection

of DM at low scale. We then studied the phenomenological effects of QCD corrections

by analyzing the effects on direct detection experiments and the dark matter relic

density constraints. These results are important in precision DM physics. Currently

little is known about the short-distance physics of DM. We calculated the box graphs

that can generate the dark matter effective operators for several different types of

dark matter fields. We then calculated the next to leading order QCD corrections,

and studied the non-factorizability. The results shows that the short-distance QCD

corrections generate a finite matching correction when integrating out the electroweak

gauge bosons.

Finally, we studied the supersymmetry. The high precision measurements of

electroweak precision observables can provide crucial input in the search for super-

symmetry (SUSY) and play an important role in testing the universality of the SM

charged current interaction. We calculated the impact of the radiative corrections to

the charged current universality in MSSM, which involves the observables ∆CKM and

∆e/µ, with the experimental constraints from the weak charge and the LHC constraints

on the sparticle masses. The order of magnitude of the corrections implies that future

experiments on charged current universality need to improve the precision to an order

of 10−4 in order to make reliable exclusions in the MSSM parameter space.

Page 147: Complementarity of Symmetry Tests at the Energy and

130

References

[1] G. ’t Hooft, Phys. Rev. Lett. 37, 8 (1976).

[2] J. Schechter and L. Valle, Phys. Rev. D 25, 2951 (1982).

[3] M. Fukugita and T. Yanagida, Phys. Lett. B 174, 45 (1986).

[4] A. de Gouvea and P. Vogel, arXiv: 1303.4097 (2013).

[5] G. Prezeau, M. Ramsey-Musolf, P. Vogel, Phys. Rev. D 68, 034016 (2003).

[6] S. Pascoli and S.T. Petcov, Phys. Lett. B 544, 239 (2002).

[7] J.D. Vergados, Phys. Rep. 361, 1 (2002).

[8] R.N. Mohapatra and J.C. Pati, Phys. Rev. D 11, 2558 (1975).

[9] G. Senjanovic and R.N. Mohapatra, Phys. Rev. D 12, 150 (1975).

[10] R.N. Mohapatra, Phys. Rev. D 34, 3457 (1986).

[11] J.D. Vergados, Phys. Lett. B 184, 55 (1987).

[12] M. Hirsch, H.V. Klapdor-Kleingrothaus, and S.G. Kovalenko, Phys. Rev. D 53,

1329 (1996).

[13] H. Klapdor-Kleingrothaus, A. Dietz, L. Baudis, et al. Eur Phys J A 12, 147

(2001). doi:10.1007/s100500170022.

[14] H. Klapdor-Kleingrothaus, and I. Krivosheina, Mod. Phys. Lett. A 21 1547-1566

(2006).

Page 148: Complementarity of Symmetry Tests at the Energy and

131

[15] J. B. Albert et al. (EXO-200 Collaboration), Nature (London) 510, 229 (2014).

[16] M. Agostini et al. (GERDA Collaboration), Phys. Rev. Lett. 111, 122503 (2013).

[17] U. Bellgardt et al. (SINDRUM Collaboration), Nucl. Phys. B 299, 1 (1988).

[18] E. Goudzovski, Moriond EW 2011 conference, arXiv:1105.5957.

[19] MAJORANA Collaboration, arXiv:1109.4154 [nucl-ex].

[20] A. Gando et al. (KamLAND-Zen Collaboration), Phys. Rev. Lett. 117, 082503

(2016).

[21] ATLAS Collaboration, JHEP 03, 041 (2015).

[22] CMS Collaboration, JHEP 01, 163 (2014).

[23] CMS Collaboration, Eur. Phys. J. C 76, 439 (2016).

[24] R.N. Mohapatra, Nucl. Phys. Proc. Suppl. 77, 376 (1999).

[25] ATLAS Collaboration, Eur. Phys. J. C 72, 2244 (2012).

[26] ATLAS Collaboration, Phys. Rev. Lett. 108 24, 241802 (2012).

[27] V. Shiltsev, arXiv:1509.08369 [physics.acc-ph].

[28] M. Benedikt, B. Goddard, D. Schulte, F. Zimmermann, M.J. Syphers, FCC-hh

Hadron Collider - Parameter Scenarios and Staging Options, Proceedings of

IPAC2015.

[29] CEPC-SPPC Preliminary Conceptual Design Report, The CEPC-SPPC Study

Group, IHEPCEPC-DR-2015-01 (2015).

Page 149: Complementarity of Symmetry Tests at the Energy and

132

[30] C.A. Argüelles, X.-G. He, G. Ovanesyan, T. Peng, M.J. Ramsey-Musolf, Phys.

Lett. B770, 101.

[31] P. F. Perez, H. H. Patel, M. J. Ramsey-Musolf, and K. Wang, Phys. Rev. D 79,

055024 (2009).

[32] F. Chen, J. M. Cline, and A. R. Frey, Phys.Rev. D79, 063530 (2009), 0901.4327.

[33] F. Chen, J. M. Cline, and A. R. Frey, Phys.Rev. D80, 083516 (2009), 0907.4746.

[34] J. M. Cline and A. R. Frey, (2014), 1408.0233.

[35] G. Barello, S. Chang, and C. A. Newby, (2015), 1511.02865.

[36] R. Bernabei et al., DAMA Collaboration, Eur.Phys.J., C56:333-355 (2008),

arXiv:0804.2741.

[37] G.W. Bennett et al., Phys. Rev. D73, 072003 (2006).

[38] R. Essig et al., Working Group Report: New Light Weakly Coupled Particles,

in Community Summer Study 2013: Snowmass on the Mississippi (CSS2013)

Minneapolis, MN, USA, July 29-August 6, 1311.0029 (2013).

[39] D. Curtin, R. Essig, S. Gori, and J. Shelton, JHEP 02, 157 (2015), 1412.0018.

[40] B. Holdom, Phys.Lett. B166, 196 (1986).

[41] R. Foot and X.-G. He, Phys.Lett. B267, 509 (1991).

[42] P. Fayet, Phys.Rev. D70, 023514 (2004), hep-ph/0403226.

Page 150: Complementarity of Symmetry Tests at the Energy and

133

[43] H. Davoudiasl, H.-S. Lee, and W. J. Marciano, Phys.Rev. D85, 115019 (2012),

1203.2947.

[44] H. Davoudiasl, H.-S. Lee, and W. J. Marciano, Phys.Rev. D89, 095006 (2014),

1402.3620.

[45] R.D. Peccei, CP Violation, C. Jarlskog, ed. (World Scientific, 1989).

[46] S. Baek, P. Ko, and W.-I. Park, JCAP 1410, 067 (2014), 1311.1035.

[47] R. Agnese et al., CDMS Collaboration, Phys.Rev.Lett. (2013), arXiv:1304.4279.

[48] D.S. Akerib et al., CDMS Collaboration. Phys.Rev., D82:122004 (2010),

arXiv:1010.4290.

[49] J. Angle et al., XENON10 Collaboration. Phys.Rev.Lett., 107:051301, (2011),

arXiv:1104.3088.

[50] E. Aprile et al., XENON100 Collaboration. Phys.Rev.Lett., 109:181301, (2012),

arXiv:1207.5988.

[51] D.S. Akerib et al., LUX Collaboration, Phys. Rev. Lett. (2013), arXiv:1310.8214.

[52] J. Goodman, M. Ibe, A. Rajaraman, W. Shepherd, T.M.P. Tait, et al, Phys.Lett.,

B695:185-188, (2011), arXiv:1005.1286.

[53] J. Goodman, M. Ibe, A. Rajaraman, W. Shepherd, T.M.P. Tait, et al, Phys.Rev.,

D82:116010, (2010), arXiv:1008.1783.

[54] J. Goodman, M. Ibe, A. Rajaraman, W. Shepherd, T.M.P. Tait, et al, Nucl.Phys.,

B844:55-68, (2011), arXiv:1009.0008.

Page 151: Complementarity of Symmetry Tests at the Energy and

134

[55] A. Rajaraman, W. Shepherd, T.M.P. Tait, A.M. Wijangco, Phys.Rev., D84:095013

(2011), arXiv:1108.1196.

[56] R.J. Hill and M.P. Solon, Phys.Lett., B707:539-545 (2012), arXiv:1111.0016.

[57] L. Vecchi, arXiv:1312.5695.

[58] N. Zhou, D. Berge, and D. Whiteson, Phys.Rev.D87(9):095013 (2013),

arXiv:1302.3619.

[59] M.R. Buckley, Phys.Rev.D88:055028 (2013), arXiv:1308.4146.

[60] M. Papucci, A. Vichi, and K.M. Zurek, arXiv:1402.2285.

[61] J. Abdallah et al, arXiv:1409.2893.

[62] S. Baek, P. Ko, M. Park, W.-I. Park, and C. Yu, arXiv:1506.06556.

[63] G. Prezeau, A. Kurylov, M. Kamionkowski, and P. Vogel, Phys.Rev.Lett.,

91:231301 (2003), arXiv:astro-ph/0309115.

[64] M. Freytsis and Z. Ligeti, Phys.Rev., D83:115009 (2011), arXiv:1012.5317.

[65] J. Hisano, K. Ishiwata, N. Nagata, and T. Takesako, JHEP, 1107:005 (2011),

arXiv:1104.0228.

[66] V. Cirigliano, M.L. Graesser, and G. Ovanesyan, JHEP, 1210:025 (2012),

arXiv:1205.2695.

[67] U. Haisch, F. Kahlhoefer, and J. Unwin, JHEP, 1307:125 (2013), arXiv:1208.4605.

[68] R.J. Hill and M.P. Solon, arXiv:1309.4092.

Page 152: Complementarity of Symmetry Tests at the Energy and

135

[69] V. Cirigliano, M.L. Graesser, G. Ovanesyan, and I.M. Shoemaker,

arXiv:1311.5886.

[70] R.J. Hill and M.P. Solon, arXiv:1401.3339.

[71] A. Crivellin, F. D’Eramo, and M. Procura, arXiv:1402.1173.

[72] R. J. Hill and M.P. Solon, Phys. Rev. D91:043505 (2015), arXiv:1409.8290.

[73] G. Ovanesyan and L. Vecchi, JHEP 07:128 (2015), arXiv:1410.0601.

[74] F. D’Eramo and M. Procura, JHEP 04:054 (2015), arXiv:1411.3342.

[75] J. Hisano, K. Ishiwata, and N. Nagata, JHEP, 06:097 (2015), arXiv:1504.00915.

[76] J.L. Feng, J. Kumar, D. Marfatia, and D. Sanford, Phys.Lett. B703:124-127

(2011), arXiv:1102.4331.

[77] A. Kurylov and M. J. Ramsey-Musolf, Phys. Rev. Lett. 88, 071804 (2002).

[78] S. Bauman, J. Erler, and M. J. Ramsey-Musolf, Phys. Rev. D 87, 035012 (2013).

[79] M. Ramsey-Musolf, S. Su, and S. Tulin, Phys. Rev. D 76, 095017 (2007).

[80] I. S. Towner and J. C. Hardy, Rept. Prog. Phys. 73, 046301 (2010).

[81] V. Cirigliano and I. Rosell, JHEP 0710, 005 (2007) [arXiv: 0707.4464 [hep-ph]].

[82] J. Erler, A. Kurylov, and M.J. Ramsey-Musolf, Phys. Rev. D68, 016006 (2003).

[83] A. Kurylov, M. J. Ramsey-Musolf and S. Su, Phys. Rev. D 68, 035008 (2003)

[hep-ph/0303026].

Page 153: Complementarity of Symmetry Tests at the Energy and

136

[84] A. Czarnecki and W. Marciano, Phys. Rev. D53, 1066 (1996).

[85] G. Degrassi, B. A. Kniehl, and A. Sirlin, Phys. Rev. D48, 3963 (1993).

[86] T. Peng, M.J. Ramsey-Musolf, and P. Winslow, Phys. Rev. D93, 093002 (2016).

[87] M. Hirsch, talk BLV-2013, Heidelberg, 2013.

[88] N. Bell, V. Cirigliano, M.J. Ramsey-Musolf, P. Vogel, and M. Wise, PRL 95

151802 (2005).

[89] M. E. Peskin and D. V. Schroeder, An Introduction to Quantum Field Theory,

Addison-Wesley, Reading, 1995.

[90] A. Faessler, S. Kovalenko, and F. Simkovic, Phys. Rev. D 58, 115004 (1998),

hep-ph/9803253.

[91] J. Donoghue, E. Golowich, and B. Holstein, Dynamics of the Standard Model

(Second Edition) Cambridge Univ. Press (2014).

[92] Particle Data Group, K. A. Olive et al., Chin. Phys. C 38, 090001 (2014).

[93] M. Doi, T. Kotani, and E. Takasugi, Prog. Theor. Phys. Suppl. 83, 1 (1985).

[94] ATLAS Collaboration, JHEP 1110, 107 (2011) [arXiv:1108.0366].

[95] UMass Cluster: https://twiki.cern.ch/twiki/bin/view/Main/UMassCluster

[96] ATLAS Collaboration, Eur. Phys. J. C72, 1909 (2012) [arXiv:1110.3174].

[97] J. Helo, M. Hirsch, S. Kovalenko, and H. Pas, Phys. Rev. D 88, 011901 (2013),

1303.0899.

Page 154: Complementarity of Symmetry Tests at the Energy and

137

[98] J. Helo, M. Hirsch, H. Pas, and S. Kovalenko, Phys. Rev. D 88, 073011 (2013),

1307.4849.

[99] ATLAS, G. Aad et al., JHEP 11, 088 (2014), 1409.0746.

[100] N. Kumar and S. P. Martin, Phys. Rev. D92, 115018 (2015), 1510.03456.

[101] M. J. Dolan, C. Englert, and M. Spannowsky, JHEP 10, 112 (2012), 1206.5001.

[102] B. Grinstein and L. Randall, Phys.Lett. B217:335 (1989).

[103] M. Cirelli, N. Fornengo, and A. Strumia, Nucl.Phys. B753:178-194 (2006),

arXiv:hep-ph/0512090.

[104] M. Cirelli, A. Strumia, and M. Tamburini, Nucl.Phys. B787:152-175 (2007),

arXiv:0706.4071.

[105] A. Sirlin, Rev.Mod.Phys. 50:573 (1978).

[106] M. Cirelli, E.D. Nobile, and P. Panci, JCAP 1310:019 (2013), arXiv:1307.5955.

[107] S.P. Martin, A Supersymmetry Primer, arXiv:hep-ph/9709356.

[108] G. Aad et al. [ATLAS Collaboration], JHEP 1405, 071 (2014) [arXiv:1403.5294

[hep-ex]].

[109] G. Aad et al. [ATLAS Collaboration], arXiv:1405.7875 [hep-ex].