characterizing the role of the dna damage response in

158
Characterizing the role of the DNA Damage Response in Class switch recombination. by Shaliny Ramachandran A thesis submitted in conformity with the requirements for the degree of Doctorate of Philosophy Department of Immunology University of Toronto © Copyright by Shaliny Ramachandran 2014

Upload: others

Post on 29-Dec-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Characterizing the role of the DNA Damage Response in

Characterizing the role of the DNA Damage Response in Class switch recombination.

by

Shaliny Ramachandran

A thesis submitted in conformity with the requirements

for the degree of Doctorate of Philosophy

Department of Immunology

University of Toronto

© Copyright by Shaliny Ramachandran 2014

Page 2: Characterizing the role of the DNA Damage Response in

ii

Characterizing the role of the DNA damage response in Class switch recombination.

Shaliny Ramachandran

Doctorate of Philosophy

Department of Immunology

University of Toronto

2014

Abstract

B cells undergo secondary antibody diversification to elicit an effective immune

response. Secondary antibody diversification generates antibodies with high affinity for the

cognate antigen, and antibodies of different classes. Different classes of antibodies mediate

different effector functions, enabling the humoral response to be fine-tuned to the infection at

hand. The different classes are generated by a process called class switch recombination (CSR),

which requires the generation and repair of double-stranded DNA breaks (DSBs) at the antibody

heavy chain locus, and relies on the 53BP1-dependent DNA damage response (DDR). However,

the molecular mechanism and function of DDR in CSR is not fully understood. To characterize

the role of DDR in CSR, we first established an in vitro CH12F3-2 mouse B cell line/ RNA

interference system and used this system to demonstrate that the DDR factors RNF8 and RN168,

play a role in CSR. Using this system, we then developed and conducted a genome-wide screen

to identify novel CSR factors, followed by a secondary screen to identify DDR factors among the

CSR candidates. Based on this study, we selected one candidate, the Drosophila enhancer of

yellow 2 homolog (Eny2), for further characterization. Eny2 is a structural component of the

H2B-deubiquitination module of the SAGA complex. We found that Eny2 likely acts

downstream of the DSBs generated for CSR, and that Eny2 might act with the SAGA complex in

Page 3: Characterizing the role of the DNA Damage Response in

iii

mediating H2B-deubiquitination to facilitate CSR. We also provide evidence that, in addition to

its role in CSR, Eny2 is important for H2B-deubiquitination upon irradiation-induced DNA

damage, and may be required for optimal DNAPK activation.

Page 4: Characterizing the role of the DNA Damage Response in

iv

Acknowledgements

I have many people to thank for the support and encouragement they provided me with, through my

doctoral studies. Alberto, I am grateful that you are my PhD thesis mentor. I have grown tremendously

with your supervision, and I have learned so much from you, especially the need to fly. I am thankful for

all your support, and I truly appreciate your encouragement of more scientific independence towards

the end of my PhD. It has been a pleasure, and I will continue to look to you for guidance as I pursue my

career in research. I wish you all the very best with your career and with your growing family.

My supervisory committee, Dr. Durocher and Dr. Carlyle, I cannot thank you enough for all your

suggestions, scientific ideas and directions. I learned so much from you both, and your input have been

absolutely invaluable to my thesis. I am also very appreciative of your supportive and encouraging

words, especially during the more difficult times in my research. I could not have envisioned a more

perfect supervisory committee, and I thank you for that.

The Immunology department administrative staff, Anna, Ed, Lynne, Sherry and William, thank you so

much for all your help and efforts with the administrative aspects of my PhD. Rejeanne, I got to know

you as the Graduate Assistant, but since then we have become good friends. Thank you for all your

advice and your very encouraging words.

Dear Martin lab, thank you for making my PhD a very enjoyable experience. Rajeev, Richard, Anat,

Angela and Alan, thank you for your scientific contributions to my thesis. Rajeev, you are an incredible

scientist. I learned so much from you while working on the CSR screen, and I thank you for that. Richard,

as an incoming graduate student, I was quite lost in the DNA damage response field, and I thank you for

introducing me to its fundamentals. Anat, you are a very patient, caring and understanding person. I

enjoyed working with you on the screen validations. Angela and Alan, I had the pleasure of training you

during your undergraduate research. You are both very smart individuals and have strong graduate

careers ahead of you. Dania, Mike and Conglei, I am happy to have started collaborating with you.

Dania, you are very smart and I know you’ll carry the Eny2 project well. Thank you for your honesty and

your advice in my research, personal and professional life. Mike, you are a strong graduate student, and

I really wish we had gotten to understand each other and work together sooner. Thank you for all your

help with the microscopy. Conglei, you are an exceptional scientist, and I have learned so much about

science and scientists from you in the short time that you’ve been here, and I thank you for that.

Antoaneta, Blerta, Bryant, Mani, Sachin and Tey, I have enjoyed working on the same team as you.

Antoaneta, you are the most efficient person I know. I still have no idea how you managed to achieve so

much work in so little time. Blerta, your ability to juggle all your extracurricular activities along with your

lab work has always impressed me. Bryant, I have enjoyed our conversations and appreciate all your

advice. I’m glad that I’ll have you as a friend and colleague in the U.K. Mani, you are an incredible

scientist with drive and ambition, and I look forward to seeing all your valuable contributions to science

in the years ahead. Sachin, I got to know you as an undergraduate student in our lab, but you have

become a good friend. You are intelligent, and your scientific questions and ability to think is very

impressive. Tey, you are a smart and wonderful addition to our lab. I know you have a strong PhD and

scientific career ahead of you.

Page 5: Characterizing the role of the DNA Damage Response in

v

Maribel, Ahmad, Darina and Jahan, I am so glad to have shared so many years in the lab with you. You

have become very close friends. Maribel, you are one of the most sensible and logical people I know. I

look to you for advice both in my personal and professional life, and I appreciate and trust your

judgement. Thank you for being a good friend. Ahmad, you are probably the most, well-rounded person

I know. You are smart, sincere and hilarious, and you’ll be an amazing M.D. You have the courage to be

honest when you need to be, even though the truth may be hard to hear, and I’m very thankful to you

for that. Darina, you are one of the most, strongest women I know. I have looked up to you in so many

ways, and consider you as one of my role models. I can’t think of anyone more ideal to be the first PhD

graduate from our lab, and to lead the way for the rest of us. You encouraged me to speak up, and I

thank you for that. Jahan, you are certainly the most considerate person I know. You are incredibly

intelligent, dedicated, hard-working, talented, as an artist, and meticulous, as a scientist. I have no

doubt that you have a promising future ahead of you in science. It is no exaggeration to say that you

have been an extremely supportive friend through some hard times, and I can’t thank you enough for

your kind and reassuring words. You have all truly made my time at the Martin lab fun times.

Dionne, Ann, Samia and Thiru, I am glad to have good friends like you in the department. Dionne, thank

you for all the cell sorts and flow advice, but more importantly, thank you for being a good friend. Ann,

Samia and Thiru, I have known you three from our undergraduate years, and graduate school has only

strengthened our friendship. You are amazing individuals and I’m grateful for your friendships. Thank

you for all your support and encouragement, and for all the coffee outings and bootcamp sessions. I will

miss them tremendously in the U.K.

Finally, I owe so much to the most important people in my life, my family. My beloved Sanjayan, you are

caring, respectful, sincere, and incredibly wonderful. Over this last year, you showed me that I couldn’t

have asked for a more supportive better half, and I promise to be just as encouraging of you as you

embark on your PhD. I know that you’ll do amazing things, and I look forward to supporting you along

the way, just like I know you’ll be there for me. You make me feel lucky and happy, and I hope to do the

same for you. Mama, Mami and Sangeetha, I feel truly blessed to become a part of your lovely family.

You have all been so incredibly welcoming, and I will do my best to honor your family values and

traditions. Serothy, you are not only my sister, you are my best friend. You are an amazing person and I

know you’ll be an inspiring teacher. Even though you are younger, I look up to you in so many ways. I

can’t ever be thankful enough for all the times you’ve been there for me, and especially for all the times

you’ve knocked some sense into me. Last but certainly not least, Appa and Amma, words cannot express

how grateful I am to you. You have struggled and made sacrifices to ensure that Serothy and I can have

all the opportunities for a better future. I learned what strength, perseverance, determination and

dedication is from you. I am who I am because of you, and every single achievement of mine is entirely a

reflection of you. I cannot ever repay you for everything that you have done for me, but as a small token

of my appreciation, I dedicate my doctoral thesis to you, Appa and Amma.

Page 6: Characterizing the role of the DNA Damage Response in

vi

Table of contents

Acknowledgements .....................................................................................................................................................iv

Table of contents .........................................................................................................................................................vi

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

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

List of Abbreviations ................................................................................................................................................ xii

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

1.1 B CELLS AND THE HUMORAL IMMUNE RESPONSE ............................................................................ 1 1.1.1 Humoral immune response ...................................................................................................................... 1 1.1.2 B cell activation and signalling ................................................................................................................ 2 1.1.3 The germinal center reaction ................................................................................................................... 4 1.1.4 Antibody protein structure and isotype function ...................................................................................... 5 1.1.5 The antibody loci ..................................................................................................................................... 7

1.2 ANTIBODY DIVERSIFICATION ................................................................................................................. 9 1.2.1 Primary antibody diversification .............................................................................................................. 9 1.2.2 Overview of Secondary antibody diversification ................................................................................... 11

1.2.2.1 Activation-induced cytidine deaminase (AID) .............................................................................. 13 1.2.2.1.1 AID substrates and targeting: ................................................................................................... 13 1.2.2.1.2 AID Regulation: ....................................................................................................................... 15 1.2.2.1.3 AID off-target effects:.............................................................................................................. 16

1.2.2.2 DNA repair pathways engaged by AID activity ............................................................................ 16 1.2.2.2.1 Base Excision Repair (BER): ................................................................................................... 16 1.2.2.2.2 Mismatch Repair (MMR): ....................................................................................................... 18

1.2.2.3 Consequences of AID activity ....................................................................................................... 19 1.2.2.3.1 Somatic hypermutation ............................................................................................................ 19 1.2.2.3.2 Gene conversion ...................................................................................................................... 20 1.2.2.3.3 Class switch recombination ..................................................................................................... 20

1.3 THE DNA DAMAGE RESPONSE (DDR) .................................................................................................. 23 1.3.1 DNA damage response in class switch recombination .......................................................................... 24 1.3.2 ATM ...................................................................................................................................................... 25 1.3.3 Mre11-Rad50-Nbs1 (MRN)................................................................................................................... 28 1.3.4 H2AX ..................................................................................................................................................... 29 1.3.5 MDC1 .................................................................................................................................................... 31 1.3.6 RNF8 ..................................................................................................................................................... 32 1.3.7 RNF168 ................................................................................................................................................. 33 1.3.8 53BP1 .................................................................................................................................................... 34

1.3.8.1 53BP1 in long-range end joining: ................................................................................................. 37 1.3.8.2 53BP1 and DNA end resection: .................................................................................................... 37 1.3.8.3 53BP1 recruits Rif1 and PTIP ...................................................................................................... 39

1.3.9 DDR factors within the ATM network not investigated in the context of CSR ..................................... 40 1.3.9.1 Deubiquitination in DDR .............................................................................................................. 40

Page 7: Characterizing the role of the DNA Damage Response in

vii

1.3.9.2 H2B Ubiquitylation in DDR .......................................................................................................... 41 1.3.9.3 Sumoylation in DDR ..................................................................................................................... 42 1.3.9.4 RNF4 function ............................................................................................................................... 43

1.3.10 DDR leads to cell cycle arrest ........................................................................................................... 44

1.4 DNA REPAIR ............................................................................................................................................. 45 1.4.1 Non-homologous end joining (NHEJ) ................................................................................................... 45 1.4.2 Microhomology-mediated end joining (MMEJ) .................................................................................... 49 1.4.3 Homologous recombination (HR) .......................................................................................................... 50

1.5 OBJECTIVES ............................................................................................................................................. 51

2 THE RNF8/RNF168 UBIQUITIN LIGASE CASCADE FACILITATES CLASS SWITCH

RECOMBINATION. ................................................................................................................................................. 52

2.1 ABSTRACT ................................................................................................................................................. 53

2.2 INTRODUCTION ....................................................................................................................................... 53

2.3 MATERIALS AND METHODS .................................................................................................................. 56 2.3.1 In Vitro Cell Culture .............................................................................................................................. 56 2.3.2 Flow Cytometric analyses ...................................................................................................................... 57 2.3.3 Irradiation sensitivity assays .................................................................................................................. 57 2.3.4 Western blot analyses ............................................................................................................................ 57 2.3.5 Immunoflourescence .............................................................................................................................. 58 2.3.6 Statistical Analysis ................................................................................................................................. 58

2.4 RESULTS ................................................................................................................................................... 58 2.4.1 shRNA-mediated RNAi can establish functional knockdown in CH12F3-2 cells. ................................ 58 2.4.2 RNF8 is required for CSR in CH12F3-2 cells. ...................................................................................... 62 2.4.3 RNF168 is important for CSR in CH12F3-2 cells. ................................................................................ 64 2.4.4 53BP1 deficiency, but not RNF8 or RNF168-deficiency, protects cells from Nutlin-3-induced

apoptosis. ............................................................................................................................................................ 67

2.5 DISCUSSION ............................................................................................................................................. 67 2.5.1 Role of RNF8 and RNF168 in CSR and its relation to 53BP1 localization. .......................................... 67 2.5.2 53BP1 has a unique function that is independent of RNF8 and RNF168. ............................................. 70

3 IDENTIFYING NOVEL DDR FACTORS FROM A GENOME-WIDE RNA INTERFERENCE

SCREEN FOR CSR. .................................................................................................................................................. 72

3.1 ABSTRACT ................................................................................................................................................. 73

3.2 INTRODUCTION ....................................................................................................................................... 73

3.3 MATERIALS AND METHODS .................................................................................................................. 75 3.3.1 Pooled shRNA lentiviral library ............................................................................................................ 75 3.3.2 In Vitro Cell Culture .............................................................................................................................. 75 3.3.3 Flow cytometry cell sorting ................................................................................................................... 76 3.3.4 Microarray analysis................................................................................................................................ 76 3.3.5 Sequencing analysis ............................................................................................................................... 76 3.3.6 Western blot analysis ............................................................................................................................. 76 3.3.7 Statistical Analysis ................................................................................................................................. 76

3.4 RESULTS ................................................................................................................................................... 77

Page 8: Characterizing the role of the DNA Damage Response in

viii

3.4.1 CH12F3-2/shRNA: A system to screen for potential CSR factors. ....................................................... 77 3.4.2 Screen methodology .............................................................................................................................. 77 3.4.3 CSR Screen results and analysis. ........................................................................................................... 81 3.4.4 Validation analysis ................................................................................................................................. 81 3.4.5 Secondary screen to identify possible DDR and repair factors .............................................................. 85

3.5 DISCUSSION ............................................................................................................................................. 91

4 CLASS SWITCH RECOMBINATION REQUIRES H2B UBIQUITINATION AND

DEUBIQUITINATION. ............................................................................................................................................ 98

4.1 ABSTRACT ................................................................................................................................................. 99

4.2 INTRODUCTION ....................................................................................................................................... 99

4.3 MATERIALS AND METHODS ................................................................................................................ 103 4.3.1 In Vitro Cell Culture ............................................................................................................................ 103 4.3.2 Ex vivo mouse experiments .................................................................................................................. 104 4.3.3 Cell staining and Flow cytometry ........................................................................................................ 104 4.3.4 RNA extraction, RT and semi-quantitative and qPCR analysis ........................................................... 104 4.3.5 Plasmids ............................................................................................................................................... 105 4.3.6 Plasmid Integration assay .................................................................................................................... 105 4.3.7 Irradiation assays ................................................................................................................................. 107 4.3.8 Western blot analyses .......................................................................................................................... 107 4.3.9 Statistical Analysis ............................................................................................................................... 107

4.4 RESULTS ................................................................................................................................................. 108 4.4.1 Eny2 acts downstream of AID in CSR ................................................................................................ 108 4.4.2 Eny2 mediates H2B-K120-deubiquitination upon DNA damage, independent of 53BP1 .................. 113 4.4.3 H2B-K120-ubiquitination/deubiquitination is important for CSR. ...................................................... 118 4.4.4 Accumulation of H2B-K120-ubiquitin may inhibit DNAPK activation .............................................. 120

4.5 DISCUSSION ........................................................................................................................................... 122

5 DISCUSSION AND FUTURE DIRECTIONS ............................................................................................. 126

5.1 RNF8 and RNF168 play a role in CSR. ................................................................................................... 126

5.2 DDR does not form a linear signalling cascade in CSR. ......................................................................... 126

5.3 H2B-ubiquitination in DDR and repair. .................................................................................................. 128

5.4 Generating an in vivo mouse model to investigate H2B-ubiquitination. .................................................. 129

Page 9: Characterizing the role of the DNA Damage Response in

ix

List of Tables

Table 1 .......................................................................................................................................... 83

Table 2 .......................................................................................................................................... 90

Table 3 ........................................................................................................................................ 106

Page 10: Characterizing the role of the DNA Damage Response in

x

List of Figures

Figure 1 ........................................................................................................................................... 6

Figure 2 ........................................................................................................................................... 8

Figure 3 ......................................................................................................................................... 12

Figure 4 ......................................................................................................................................... 17

Figure 5 ......................................................................................................................................... 21

Figure 6 ......................................................................................................................................... 26

Figure 7 ......................................................................................................................................... 46

Figure 8 ......................................................................................................................................... 60

Figure 9 ......................................................................................................................................... 61

Figure 10 ....................................................................................................................................... 63

Figure 11 ....................................................................................................................................... 65

Figure 12 ....................................................................................................................................... 66

Figure 13 ....................................................................................................................................... 68

Figure 14 ....................................................................................................................................... 78

Figure 15 ....................................................................................................................................... 79

Figure 16 ....................................................................................................................................... 82

Figure 17 ....................................................................................................................................... 84

Figure 18 ....................................................................................................................................... 86

Figure 19 ....................................................................................................................................... 87

Figure 20 ....................................................................................................................................... 88

Figure 21 ....................................................................................................................................... 89

Figure 22 ....................................................................................................................................... 92

Figure 23 ....................................................................................................................................... 94

Figure 24 ..................................................................................................................................... 109

Page 11: Characterizing the role of the DNA Damage Response in

xi

Figure 25 ..................................................................................................................................... 112

Figure 26 ..................................................................................................................................... 114

Figure 27 ..................................................................................................................................... 115

Figure 28 ..................................................................................................................................... 117

Figure 29 ..................................................................................................................................... 119

Figure 30 ..................................................................................................................................... 121

Page 12: Characterizing the role of the DNA Damage Response in

xii

List of Abbreviations

3'RR 3' Regulatory Region

53BP1 p53 Binding Protein 1

dA Adenine

AID Activation-Induced cytidine Deaminase

APOBEC Apolipoprotein B mRNA-editing catalytic

APRIL A proliferation-inducing ligand, TNFSF 13a

A-T Ataxia-telangiectasia

ATM Ataxia-telangiectasia Mutated

ATR Ataxia-telangiectasia and Rad3 related

ATRIP ATR interacting protein

BAFF B cell activation factor of the TNF family, TNFSF 13b

Bcl2 B cell leukemia/lymphoma 2

Bcl6 B cell leukemia/lymphoma 6

BER Base Excision Repair

BRCA1 breast cancer 1

C Constant

Chk1 Checkpoint kinase 1

Chk2 Checkpoint kinase 2

c-myc myelocytomatosis oncogene

CSR Class Switch Recombination

CtIP CtBP-interacting protein; retinoblastoma binding protein 8

D Diversity

dA Deoxyadenosine

dC Deoxycytosine

DDR DNA Damage Response

dG Deoxyguanosine

DNAPK DNA-dependent protein kinase

DNAPKcs DNA-dependent protein kinase catalytic subunit

DSB Double-Strand Break

dT Deoxythymine

dU Deoxyuridine

Eµ Heavy chain µ intronic enhancer

Eny2 Enhancer of Yellow 2 homolog (Drosophila)

Exo1 Exonuclease 1

Fc

Fucci

Fragment crystallisable

Fluorescent Ubiquitination based cell cycle indicator

GCV Gene Conversion

HR Homologous Recombination

Hsp90 Heat shock protein 90

Page 13: Characterizing the role of the DNA Damage Response in

xiii

I Intronic promoter

IFNγ Interferon γ

Igα Immunoglobulin α chain

Igβ Immunoglobulin β chain

IL4 Interleukin 4

IL5 Interleukin 5

IRIF Irradiation-induced nuclear foci

ITAM Immunoreceptor tyrosine-based activation motif

J Joining

K Lysine

LPS Lipopolysaccaride

mAG1 Azami-Green1

MDC1 Mediator of DNA damage checkpoint 1

MDM2 Transformed mouse 3T3 cell double minute 2

MHC Major Histocompatibility Complex

Mlh1 mutL homolog 1

MMEJ Microhomology-mediated end joining

MMR Mismatch repair

Mre11 meiotic recombination 11

MRN Mre11-Rad50-Nbs1

Msh2 mutS homolog 2

Msh6 mutS homolog 6

mTOR mechanistic target of rapamycin (serine/threonine kinase)

Nbs1 nijmegen breakage syndrome 1 protein

NHEJ Non-homologous end joining

p53 tumor protein p53

PI3-Kinase phosphatidylinositol 3-kinase

PI3-KK phosphatidylinositol 3-kinase-like kinase

pIgR polymeric Ig receptor

PKA Protein Kinase A

Pms2 postmeiotic segregation increased 2 (S. cerevisiae)

PTIP Pax Transactivation-Domain Interacting Protein

R Purine base pairs

Rad50 Rad50 homolog (S. cerevisiae)

RAG1 Recombination activating gene 1

RAG2 Recombination activating gene 2

Rif1 Rap1 interacting factor 1 homolog (yeast)

RNF168 Ring finger protein 168

RNF20 Ring finger protein 20

RNF40 Ring finger protein 40

RNF8 Ring finger protein 8

Page 14: Characterizing the role of the DNA Damage Response in

xiv

RPA Replication Protein A

S Switch

SAGA Spt-Ada-Gcn5 acetyltransferase complex

SCID Severe Combined Immunodeficiency

Ser Serine

SHM Somatic Hypermutation

SIM SUMO-Interacting Motif

SMG1 SMG1 homolog, phosphatidylinositol 3-kinase-related kinase (C. elegans)

S/TQ Serine/Threonine-Glutamine

SUMO Small Ubiquitin-like Modifier

TGFβ Transforming growth factor, beta

Thr Threonine

TNF Tumor necrosis factor

TRRAP transformation/transcription domain-associated protein

Ubc13 ubiquitin conjugating enzyme 13

UNG uracil-DNA glycosylase

V Variable

W Adenine or Thymine

XRCC2 X-ray repair complementing defective repair in Chinese hamster cells 2

XRCC4 X-ray repair complementing defective repair in Chinese hamster cells 6

Page 15: Characterizing the role of the DNA Damage Response in

1

1 INTRODUCTION

1.1 B CELLS AND THE HUMORAL IMMUNE RESPONSE

1.1.1 Humoral immune response

Humoral immunity is responsible for protecting an organism by eliminating extracellular

pathogens or toxins, and by preventing the spread of intracellular infections. The humoral

immune response is carried out by B lymphocytes. There are two types of mature B

lymphocytes, called B-1 and B-2 cells. B-1 cells are considered innate-like, mainly express CD5,

undergo self-renewal, mainly reside within the peritoneal and pleural cavities and produce

natural serum antibodies (Baumgarth 2011). B-2 cells are the conventional B cells which arise in

the bone marrow, function within the lymphoid tissue and produce antibodies in response to an

infection. Upon activation with its cognate antigen, B cells can either differentiate into memory

cells or plasma cells. Memory cells are long-lived quiescent B cells that get activated upon

subsequent exposure to the same pathogen that they previously encountered. Upon activation,

they typically produce high-affinity antibodies. On the other hand, plasma cells are short-lived

cells that secrete large amounts of high affinity antibodies to the pathogen that serve to eliminate

the infection.

Antibodies can eliminate an infection by three different mechanisms: neutralization,

opsonization, and complement activation. Neutralization is achieved, for example, when

antibodies bind to receptors on a pathogen that are used by the pathogen to infect cells, thus

disabling the pathogen from infecting cells (Smith et al. 1996; Edelson et al. 2001; Haberstroh et

al. 2008). Opsonization occurs when antibodies coat the surface of a pathogen, but in this case,

the bound antibody promotes phagocytosis of the antibody-coated pathogen, by phagocytic

Page 16: Characterizing the role of the DNA Damage Response in

2

immune cells such as macrophages (Joller et al. 2011). Upon phagocytosis, the pathogen is

enveloped within the phagosome, which fuses with the lysosome, in turn causing degradation of

the pathogen (Joller et al. 2011). Complement deposition is achieved by the activation of the

complement cascade by pathogen-bound antibody (Mold et al. 2002). Complement activation

can either augment opsonization or create pores in the pathogen membrane, directly inducing

lysis of the pathogen (Dunkelberger et al. 2010). All three of these functions of antibodies play a

key role in controlling extracellular bacterial infections. Antibodies can also cause antibody-

dependent cellular cytotoxicity, in which case the antibodies bind to an infected cell and this is

typically recognized by natural killer kills that induce lysis of the infected cell (Bradford et al.

2001; Sulica et al. 2001). In addition to natural killer cells, macrophages, myeloid cells and

neutrophils can also mediate antibody-dependent cellular cytotoxicity (Siders et al. 2010; Braster

et al. 2013).

1.1.2 B cell activation and signalling

Naïve B cells are activated when they first encounter foreign antigens, leading to antibody

secretion and the generation of a germinal center response. B cell activation involves cross-

linking of the B cell receptor expressed on the surface of the B cell (Gold et al. 1992). The B cell

receptor complex contains the B cell receptor in association with the Igα and Igβ polypeptides

(DeFranco et al. 1995). The B cell receptor itself is incapable of transmitting a signal and relies

instead on the associated Igα and Igβ polypeptides for signal transmission to the cell (Taddie et

al. 1994; DeFranco et al. 1995). Igα and Igβ each contain an immunoreceptor tyrosine-based

activation motif (ITAM) on their cytosolic end that is phosphorylated by Src-family kinases

upon B cell receptor engagement to the antigen (DeFranco et al. 1995). B cell receptor signalling

Page 17: Characterizing the role of the DNA Damage Response in

3

also relies on the B cell co-receptor complex, which includes CD21, CD19, CD81 and Leu-13

(Sato et al. 1997; Fearon et al. 2000; Mongini et al. 2001; Mongini et al. 2002). CD21, also

known as complement receptor 2, binds to complement fragments on the antigen, clustering the

co-receptor complex with the B cell receptor, which leads to phosphorylation of CD19 by B cell

receptor-associated kinases, thus initiating PI3-kinase signalling cascade (Fearon et al. 2000).

B cell activation requires a second signal in addition to B cell receptor cross-linking. This

second signal depends on the type of antigen, and can either be T cell-dependent or T cell-

independent. The B cell receptor binds the pathogen and internalizes it by endocytosis, then

fragments it into peptides, and some of these peptides are then presented on the B cell surface in

the context of MHC class II (Sproul et al. 2000; Clark et al. 2004). An antigen presented in the

context of MHC class II on B cells can be recognized by the cognate T cell receptor of a helper T

cell, which in turn provides a second signal for B cell activation through CD154-CD40 ligation

and cytokine stimulation (Gowthaman et al. 2010). T cell-dependent activation in most cases

requires that the helper T cell and activated B cell recognize the same antigen. This leads to the

germinal center reaction and the generation of high-affinity antibodies, which bind to and

eliminate the pathogen by the mechanisms described above. Alternatively, some antigens lead to

T cell-independent B cell activation. Certain microbial antigens can activate an innate immune

receptor in addition to the B cell receptor, providing the second signal needed for B cell

activation (Ruprecht et al. 2006). A well-known example of this is LPS mediated activation of

Toll-like receptor 4 (Hoshino et al. 1999). A microbial antigen with a highly repetitive epitope

can also lead to B cell activation, simply by causing enhanced cross-linking of the B cell receptor

(Vos et al. 2000). T cell independent B cell activation cannot generate a germinal center

Page 18: Characterizing the role of the DNA Damage Response in

4

response but can lead to long-lived plasma cells (Taillardet et al. 2009; Defrance et al. 2011;

Bortnick et al. 2013).

1.1.3 The germinal center reaction

Upon activation, B cells either enter or produce one de novo germinal center, where they

undergo high levels of proliferation and secondary antibody diversification. B cells expressing

antibodies with a high affinity for antigen are then selected in the germinal center to differentiate

either into antibody-secreting plasma cells or memory cells, while B cells expressing low affinity

antibodies undergo apoptosis (Victora et al. 2012). The germinal center is organized into a dark

zone and light zone. The dark zone is concentrated with hyper-proliferating B cells, while the

light zone is characterized by a network of follicular dendritic cells that capture antigen and

provides the basis for selecting B cells that produce high-affinity antibodies to the antigen in

question (Victora et al. 2012).

Traditionally, the germinal center response was viewed as a simple model in which

activated B cells enter the dark zone, where they become centroblasts that undergo hyper-

proliferation and somatic hypermutation (Allen et al. 2007). Somatic hypermutation is a process

by which antibodies accumulate mutations in the antibody encoding genes (see below), changing

the affinity of the antibody for the antigen. The cells then enter the light zone and become

centrocytes that compete with each other to bind antigen-immune complexes presented by the

follicular dendritic cells (Allen et al. 2007). As antigen engagement provides survival signals to

the B cells, the B cells with high affinity for antigen are selected for, while B cells with low

affinity for antigen do not receive survival signals and undergo apoptosis (Victora et al. 2012). B

cells can receive survival signals from TNF ligand family members, including BAFF and APRIL

Page 19: Characterizing the role of the DNA Damage Response in

5

(Mackay et al. 2003). This process of selection is called affinity maturation. B cells with high

affinity are selected to survive, proliferate and interact with the cognate helper T cell, initiating

differentiation into plasma cells or memory cells. However, two-photon microscopy technology

developed recently for real-time imaging of B cells in the germinal center has suggested that B

cell proliferation occurs in both dark and light zones, and B cells migrate in both directions

between the two compartments suggesting that the traditional view of the germinal center

reaction may not be completely accurate (Allen et al. 2007).

1.1.4 Antibody protein structure and isotype function

An antibody is composed of two heavy chains and two light chains held together by

disulfide bridges, to form a Y-shaped structure (Figure 1). The two arms of the Y-shaped

antibody join at a flexible region called the hinge region. At the N-terminal ends of the two

antibody arms lies the antigen binding sites, which is constructed by the variable (V) regions of

both the heavy and light chains. The C-terminal end of the antibody is referred to as the constant

region, because there is minimal sequence diversity in this region compared to the V-region. The

constant region dictates the effector function of the antibody. There are five different constant

regions. These five classes or isotypes are IgM, IgD, IgG, IgE and IgA. IgM and IgD are initially

expressed by mature B cells, but B cells can then switch to express IgG, IgE or IgA. IgG is the

most abundant antibody, IgE is important in allergic responses and IgA is critical for mucosal

immunity. There are four subclasses of IgG in humans – IgG1, IgG2, IgG3 and IgG4; and five

subclasses in mice - IgG1, IgG2a, IgG2b, IgG2c, and IgG3. Inbred mouse strains including

C57BL/6, C57BL/10, SJL and NOD contain the Igh1-b gene produce IgG2c, and contain a

Page 20: Characterizing the role of the DNA Damage Response in

6

Figure 1

Figure 1: Protein structure of IgM, IgD, IgG, IgE and IgA antibodies. Each antibody

consists of 2 heavy chains and 2 light chains, which together form a Y-shaped structure. The

antigen binding variable sites are found at the N-terminal ends of the two antibody arms. The

constant region is at the C-terminal end of the antibody. There are five different constant

regions: IgM, IgD, IgG, IgE and IgA. For IgD, IgG and IgA, the two arms and the leg of the Y-

shaped antibody join at a flexible region called the hinge region. IgM and IgE have a third

domain replacing the hinge region. IgM can form a pentamer held together by the J chain, while

IgA can form a dimer held by the J chain.

Page 21: Characterizing the role of the DNA Damage Response in

7

deletion of the IgG2a gene (Martin et al. 1998). While most antibodies exist as monomers, IgA

can form dimers and IgM can form pentamers/hexamers.

The antibody constant region can be bound by Fc receptors and immunoglobulin receptors

found on cells. Different receptors have different affinities for each of the constant regions, thus

dictating how each class of antibody functions. Fc receptors can lead to a stimulatory or

inhibitory response. For instance, IgG1 binds to FcγRIII and is important for NK cell activation

and antibody-dependent cellular cytotoxicity (Sulica 2001). IgM bound by FcμRI, and IgG1 and

IgG3 bound by IgG receptors can activate the classical complement pathway, while IgA bound

by FcαRI can activate the alternative and lectin complement pathways (Daha 2011). Polymeric

Ig receptors (pIgR) bind to and transport IgA dimers and IgM pentamers across epithelial tissues

(Raghavan et al. 1996). IgE binds to FcεRI on mast cells, eosinophils and basophils, and leads to

the allergic response (MacGlashan 2012).

1.1.5 The antibody loci

The immunoglobulin loci consist of one µ heavy chain locus and two light chain loci, κ

and λ. In humans, the µ heavy chain is on chromosome 14, κ on chromosome 2 and λ on

chromosome 22. In mice, the µ heavy chain is on chromosome 12, κ on chromosome 6 and λ on

chromosome 16 (Honjo 1983). Beginning from the 5’ end, the heavy chain is organized into the

variable gene segments, intronic enhancer Eμ, constant gene segments and finally 3’Regulatory

region (3’RR), which contains the distal enhancers (Figure 2). The variable gene segments can

be subdivided into variable (V), diversity (D) and joining (J) regions on the heavy chain, and V

and J regions on the light chains, which together encode the corresponding variable domain of

the antibody. During maturation, a gene rearrangement process termed V(D)J recombination,

Page 22: Characterizing the role of the DNA Damage Response in

8

Figure 2

Figure 2: Locus structure of immunoglobulin heavy chain in human and mouse. The heavy

chain is organized starting from the 5' end, into the variable gene segments (VDJ), intronic

enhancer (Eμ), constant gene segments and finally 3’Regulatory region (3’RR), which contains

the distal enhancers. Each constant gene segment (C) is associated with a promoter and switch

region (S). The human heavy chain consists of µ, δ, γ3, γ1, α1, γ2, γ4, ε, and α2 constant regions.

The mouse heavy chain consists of µ, δ, γ3, γ1, γ2b, γ2a, ε, and α constant regions. However,

C57BL/6, C57BL/10, SJL and NOD mice contain a deletion of the γ2a gene, while expressing

the Igh1-b gene that produces antibodies with γ2c constant region.

Page 23: Characterizing the role of the DNA Damage Response in

9

which is initiated by the Rag1/2 protein complex, recombines one V, one D and one J gene

segment together on the µ heavy chain locus, and one V and one J gene segment together on the

light chain locus, to form a single coding exon called the variable domain (described in further

detail below). Within the variable domain sequence, there are three distinct hypervariable or

complementarity-determining regions that come together on the folded protein to make up the

antigen-binding hypervariable surface.

1.2 ANTIBODY DIVERSIFICATION

1.2.1 Primary antibody diversification

Humans and mice produce millions of B cells each day. Each B cell produces antibodies

that recognize one epitope, but different B cells recognize different epitopes. Combining all the

millions of B cells produced every day, the human antibody repertoire is estimated to be 1011

This antibody repertoire is generated through primary antibody diversification, a mechanism

based on somatic gene rearrangement referred to as V(D)J recombination (Hozumi et al. 1976).

V(D)J recombination occurs within developing B cells in the bone marrow and involves genetic

rearrangement at the immunoglobulin heavy and light chain loci as described above (Bernard et

al. 1978; Brack et al. 1978).

Primary antibody diversification uses combinatorial diversity and junctional diversity

mechanisms to derive the antibody repertoire. As described above, the light chain is encoded by

two gene segments, the V and J segments, while the µ heavy chain locus contains three gene

segments, the V, D and J segments. Recombination occurs between one of the multiple V regions

and one of the multiple J regions, to assemble one functional light chain. At the heavy chain

locus, recombination occurs between one of the multiple D and J regions, followed by

Page 24: Characterizing the role of the DNA Damage Response in

10

recombination between one of the V regions to the DJ regions to construct a functional heavy

chain (Early et al. 1980; Alt et al. 1984). All the different combinations of V to J in the light

chain, and all the different combinations of V to D to J in the heavy chain, provide combinatorial

diversity. All the possible combinations of µ heavy chain to either the κ or λ light chains, adds a

further layer of combinatorial diversity to the primary antibody diversification. Junctional

diversity, on the other hand, is generated by nucleotide processing of the two DNA ends

involved in generating the joint between V and J in light chain, and the joints between V and D

and between D and J in the heavy chain (Boubnov et al. 1993). The addition of nucleotides at

these joints during the recombination process also contributes to junctional diversity (Lieber et

al. 1988). Together, combinatorial and junctional diversity contribute to the estimated 1011

antibody repertoire.

V(D)J recombination requires the lymphoid-specific RAG1 and RAG2 proteins to

recognize the recombination signal sequences located directly beside the V, D and J segments

(Schatz et al. 1989; Oettinger et al. 1990). In the light chain, the RAG complex binds to the

recombination signal sequences and brings the V and J segments into close proximity. RAG1/2

then mediates cleavage of the recombination signal sequences such that the intervening DNA is

removed, and the two DNA ends of the V and J segments are combined to form a hairpin

structure (Schatz et al. 2011). The hairpin is cleaved by Artemis to generate palindromic or P-

nucleotides, and terminal deoxynucleotidyl transferase adds nucleotides to the DNA ends in a

non-templated fashion generating N-nucleotides (Komori et al. 1993; Rooney et al. 2003). P and

N nucleotides along with nuclease activity that leads to the removal of nucleotides from the

DNA ends contribute to the junctional diversity described above. Finally, the two DNA ends are

ligated together to form a coding joint through non-homologous end joining, which will be

Page 25: Characterizing the role of the DNA Damage Response in

11

described below in more detail, to form a complete chain (Pergola et al. 1993; Taccioli et al.

1993). A similar mechanism of recombining D to J and then V to DJ will generate the complete

variable portion of the heavy chain. During coding joint formation, nucleolytic processing can

cause frame-shifts leading to non-productive rearrangements. Developing B cells bearing these

non-functional rearrangements undergo apoptosis.

1.2.2 Overview of Secondary antibody diversification

While primary antibody diversification produces a wide variety of antibodies against the

array of antigens that challenge the host, the majority of these antibodies possess low affinity for

the target. Antibodies with high affinity for the pathogen are most desirable to mount a strong

immune response, while antibodies of different classes are most desirable to fine-tune the type of

immune response in order to be most effective against the type of pathogen that has infected the

host. Increases in antibody affinity and changes in antibody class are achieved through secondary

antibody diversification processes that are triggered by B cell activation. In the initial phase of

the adaptive response, B cells produce low affinity antibodies of class µ and δ. Upon interacting

with the pathogen, B cells enter the germinal center and undergo secondary antibody

diversification processes that generate both high affinity antibodies and antibodies with different

classes. While the processes of somatic hypermutation (SHM) and gene conversion (GCV)

generate high affinity antibodies, class switch recombination (CSR) generates the different

classes (Figure 3). SHM and CSR occur in humans and mice, while GCV occurs in certain

animals including chicken and sheep. All three secondary diversification processes absolutely

require the B cell specific enzyme, activation-induced cytidine deaminase (AID) (Muramatsu et

al. 2000; Arakawa et al. 2002).

Page 26: Characterizing the role of the DNA Damage Response in

12

Figure 3

Figure 3: Secondary Antibody diversification. Secondary antibody diversification processes

include somatic hypermutation, gene conversion and class switch recombination. Somatic

hypermutation introduces point mutations along the recombined variable region of the heavy and

light chains. Gene conversion is the non-reciprocal insertion of gene segments from the upstream

pseudo V genes into the rearranged variable region. Class switch recombination is the

replacement of the heavy chain constant region by a downstream constant region.

Page 27: Characterizing the role of the DNA Damage Response in

13

1.2.2.1 Activation-induced cytidine deaminase (AID)

AID is a B cell-specific member of the apolipoprotein B mRNA-editing catalytic

component (APOBEC) family (Muramatsu et al. 1999). It contains a cytidine deaminase domain,

a bipartite nuclear localization signal and a nuclear export signal. Abnormalities in AID

expression and function have been linked to hyper IgM syndrome, autoimmunity, lymphoid

hyperplasia and B cell lymphomas (Minegishi et al. 2000; Revy et al. 2000; Greeve et al. 2003;

Xiao et al. 2007).

1.2.2.1.1 AID substrates and targeting:

AID is generally believed to act on single-stranded DNA to catalyze deamination of

deoxycytidines to deoxyuridines (Bransteitter et al. 2003; Dickerson et al. 2003; Larijani et al.

2007). In vitro studies have demonstrated that AID acts on deoxycytidine by preferentially

targeting a WRC motif (W=A/T, R=A/G) for deamination (Pham et al. 2003; Larijani et al.

2005). Transcription of the immunoglobulin locus upon B cell activation produces the single-

stranded DNA substrates required for AID (Chaudhuri et al. 2003; Ramiro et al. 2003). AID

mutates both template and non-template strands equally, as there is no observed bias in mutation

frequencies towards either strand in vivo (Xue et al. 2006). Recent work has suggested that the

molecular substrates of AID are short single-stranded DNA patches that are present equally on

the template and nontemplate strands and are linked to transcription (Parsa et al. 2012).

Specifically, these single-stranded DNA patches are thought to be formed by transcription-

induced negative supercoiling (Parsa et al. 2012).

Although AID is capable of inducing genome-wide mutations in mammalian cells, it

nevertheless preferentially mutates the immunoglobulin locus by orders of magnitude over other

Page 28: Characterizing the role of the DNA Damage Response in

14

loci, indicating that AID is targeted to the immunoglobulin locus (Wagner et al. 1996; Parsa et

al. 2007; Liu et al. 2008; Staszewski et al. 2011). Initial work in the field was focused on

identifying AID-targeting cis elements, with a particular interest in sequences within the

immunoglobulin enhancer regions. However, these enhancers are required for efficient

transcription, making it difficult to delineate whether the enhancers were required for AID

activity through transcription and/or AID targeting (Pavri et al. 2011). However, recent data

shows that the 3’ regulatory region of the µ heavy chain locus is necessary for SHM and CSR,

but only partially required to induce transcription of the immunoglobulin gene, suggesting that

the 3' regulatory region has AID-targeting cis elements (Rouaud et al. 2013). In addition, a

region of 4kb found downstream of the distal 3’ enhancer in chicken λ light chain was suggested

to possess AID targeting abilities (Kothapalli et al. 2008).

Trans acting factors have also been implicated in AID targeting. RNA polymerase II

stalling was linked to AID targeting (Rajagopal et al. 2009; Pavri et al. 2010). Spt5 (Suppressor

of Ty5 homolog), which associates with stalled RNA polymerase II, was found to bind to AID

and recruit it to regions of stalled transcription (Pavri et al. 2010). This is followed by the

recruitment of the RNA exosome, an RNA processing complex, to the AID-Spt5-Polymerase II

complex at the stalled sites (Basu et al. 2011; Pavri et al. 2011). Stalled RNA polymerase II can

often backtrack, uncovering the free 3’ end of the nascent mRNA, thereby allowing the RNA

exosome to load and degrade the nascent RNA, in turn exposing the bottom strand in addition to

the free top strand, to AID deamination activity (Basu et al. 2011; Pavri et al. 2011). This model

also provides a means for targeting AID to the immunoglobulin locus to both the template and

coding DNA strands.

Page 29: Characterizing the role of the DNA Damage Response in

15

Interestingly, N-terminal AID mutations cause reduced SHM but maintain CSR, while C-

terminal AID mutations cause reduced CSR and retain SHM, suggesting SHM and CSR may

require differential cofactors, possibly for AID targeting (Ta et al. 2003; Shinkura et al. 2004).

1.2.2.1.2 AID Regulation:

Owing to the ability of AID to mutate the genome, it is not surprising that AID has been

found to be highly regulated at multiple levels. AID expression is largely restricted to the

centroblast stage of germinal center B cells, the stage at which the B cells undergo SHM and

CSR (Muramatsu et al. 1999). AID transcript levels are regulated by two miRNAs, miR-155 and

miR-181b (de Yebenes et al. 2008; Dorsett et al. 2008; Teng et al. 2008). Moreover, AID

expression and activity is largely restricted to the G1 phase of cell cycle (Schrader et al. 2007).

AID protein is restrained to the cytoplasm. Nuclear AID level is limited by two

mechanisms. First, AID is actively exported from the nucleus through CRM1-dependent nuclear

export (McBride et al. 2004). Second, nuclear AID is less stable than cytoplasmic AID: nuclear

AID has a half-life of ~2.5 hours while cytoplasmic AID has a half-life of ~8 hours (Aoufouchi

et al. 2008). The decreased stability of nuclear AID is likely due to polyubiquitination and

proteasomal degradation within the nucleus (Aoufouchi et al. 2008), while cytoplasmic AID is

protected from proteasomal degradation by hsp90 (Orthwein et al. 2010).

AID can be phosphorylated at multiple Ser/Thr residues. In particular Ser38

phosphorylation by PKA plays an important role in SHM and CSR (Cheng et al. 2009). It is

believed that PKA and AID are independently recruited to S regions, where PKA phosphorylates

AID at Ser38, which allows AID to interact with RPA, a single-stranded DNA binding protein

Page 30: Characterizing the role of the DNA Damage Response in

16

(Vuong et al. 2009). Both PKA-mediated S38 phosphorylation and RPA binding increase AID

binding and activity (Chaudhuri et al. 2004; Basu et al. 2005).

Intriguingly, ectopic expression of AID using AID transgenic mice showed frequent

development of T cell lymphomas instigated by AID-induced chromosomal translocations

(Okazaki et al. 2003). The observation that ectopic AID lead to T cell lymphomas instead of B

cell lymphomas suggests that there are other B cell specific regulators that control AID function

to prevent AID off-target effects and oncogenesis (Okazaki et al. 2007).

1.2.2.1.3 AID off-target effects:

AID-induced off-target effects can lead to oncogenic mutations, chromosomal

translocations, genome instability and promote B cell lymphomas (Perez-Duran et al. 2007;

Staszewski et al. 2011; Gazumyan et al. 2012). Chromosomal translocations involving the

immunoglobulin locus and the c-myc, Bcl2 and Bcl6 have been associated with Burkitt’s

lymphoma, follicular B cell lymphoma and diffuse large B-cell lymphoma respectively (Ramiro

et al. 2007). These translocation events have been ascribed to AID (Lu et al. 2013; Robbiani et

al. 2013), and this is consistent with the ability of AID to mutate the genome at many sites, albeit

at significantly lower levels than immunoglobulin loci (Liu et al. 2008).

1.2.2.2 DNA repair pathways engaged by AID activity

1.2.2.2.1 Base Excision Repair (BER):

As shown in Figure 4, BER generates nicks at the AID-induced dU:dG mismatches

(Petersen-Mahrt et al. 2002). Uracil N-Glycosylase (UNG) excises the deoxyuridine formed by

AID activity, producing an abasic site (Di Noia et al. 2002; Rada et al. 2002; Kavli et al. 2005).

Page 31: Characterizing the role of the DNA Damage Response in

17

Figure 4

Figure 4: AID-induced mutations. AID-induced cytidine deamination can be engaged by

DNA replication, base-excision repair (BER) and mismatch repair (MMR) pathways. DNA

replication through the dU:dG mismatch creates transition mutations. In BER, UNG/APE

activity generates an abasic site, and replication through the abasic site produces transition and

transversion mutations. Finally, engagement of the MMR pathway, orchestrated by Msh2 and

Exo1, in concert with UNG/APE-mediated BER, leads to the recruitment of translesional

replicative Polymerase η, producing transversion mutations and mutations at A-T basepairs.

Page 32: Characterizing the role of the DNA Damage Response in

18

The apurinic/apyrimidinic endonuclease (APE) is then recruited to cleave the abasic site to

produce single-stranded DNA breaks (Peled et al. 2008). Translesional replication through the

abasic site can generate transversion and transition mutations. On the other hand, if the

AID/UNG induced nicks occur in very close proximity on both strands of DNA, these lesions

can be converted to staggered double-stranded DNA breaks (DSBs) (Schrader et al. 2005). The

DSBs produced during CSR are thought to proceed via the above described mechanism.

1.2.2.2.2 Mismatch Repair (MMR):

In the context of secondary antibody diversification, MMR paradoxically mutates dA:dT

base-pairs as depicted in Figure 4, contrary to its classic role as a DNA repair pathway (Cascalho

et al. 1998; Roa et al. 2010). AID-induced dU:dG mismatches can be recognized by the

Msh2/Msh6 heterodimer (Martomo et al. 2004; Wilson et al. 2005). Subsequent recruitment of

the Mlh1/Pms2 heterodimer leads to the recruitment of the Exo1 exonuclease to a nearby nick,

allowing Exo1 to excise a stretch of single-stranded DNA, including the mismatch (Saribasak et

al. 2009). This is followed by recruitment of DNA polymerases to fill in the excised strand.

However, in secondary antibody diversification, partly due to the concerted action of UNG (in

BER pathway) and MMR, the error-prone polymerase Polymerase η can be recruited to repair

the excised patch, which fills in the excised strand, while producing mutations at dA:dT

basepairs (Zeng et al. 2001). Alternatively, Exo1-mediated single-stranded DNA track excision

can cause single-stranded DNA breaks to become double-stranded DNA breaks if a nick is

present on the strand opposite to the one being excised (Schrader et al. 2005; Stavnezer et al.

2005; Schrader et al. 2007). As will be discussed below, the MMR system plays a minor but

detectable role in producing the pre-requisite DSBs during CSR.

Page 33: Characterizing the role of the DNA Damage Response in

19

1.2.2.3 Consequences of AID activity

1.2.2.3.1 Somatic hypermutation

SHM is the process by which AID induces point mutations along the rearranged variable

region of the immunoglobulin heavy and light chains, creating antibodies with varying affinities

to the antigen. SHM of the V-region occurs at a mutation rate of 10-3 mutations/base-

pair/generation (Peled et al. 2008), and both strands of DNA are subject to hypermutation

(Milstein et al. 1998). AID-mediated deamination of cytidine to uridine creates dU:dG

mismatches in the DNA, which are processed through DNA replication, BER, or error-prone

MMR to generate point mutations along the immunoglobulin locus (Figure 4). Replication

through the dU:dG mismatch causes transition mutations, engagement by the BER pathway

generates transition and transversion mutations, and involvement of the MMR pathway leads to

mutations at dA:dT basepairs (Di Noia et al. 2007). More specifically, the joint activity of both

BER and MMR results in A/T mutations (Frieder et al. 2009). The UNG-produced abasic site

causes MMR to recruit error-prone polymerase η, which causes A/T mutations (Frieder et al.

2009). However, there are other unknown manners by which these A/T mutations arise, since

A/T mutations still occur in UNG-/- mice. While mutations can occur at any of the four base-

pairs, a general bias towards transition mutations over transversion mutations is observed (Peled

et al. 2008). The mutations are primarily confined to the region starting ~150 basepairs

downstream of the transcription start site to ~2 kilo-basepairs downstream of the transcription

start site, although the mutation rate peaks at ~500 basepairs and then drops as the distance

increases from the promoter (Lebecque et al. 1990; Rada et al. 2001). Transcription is critical to

SHM, given that the promoter is required, and mutation rates have been found to correlate with

transcription rates (Fukita et al. 1998; Bachl et al. 2001). These mutations can be neutral,

Page 34: Characterizing the role of the DNA Damage Response in

20

negative or positive with respect to antigen affinity. B cells harbouring mutations that confer the

antibody with higher affinity to the antigen will be selected during affinity maturation of the

immune response.

1.2.2.3.2 Gene conversion

In some vertebrates including birds, rabbits, cows, pigs, sheep and horses, primary

antibody diversification generates an extremely limited antibody repertoire (Tang et al. 2007).

This is compensated through the secondary antibody process of GCV, which in chickens, occurs

in a highly specialized compartment called the Bursa of Fabricius. Upon completing V(D)J

recombination, the B cells colonize the bursa to undergo GCV. In this case, the immunoglobulin

heavy and light chain loci contain pseudo gene V regions upstream of the rearranged V(D)J

segment. The pseudo V genes differ from the rearranged V-region by 10-20% at the nucleotide

level (Tang et al. 2007). During GCV, parts of the V gene sequence in the rearranged V(D)J

gene are replaced with sequences from the upstream pseudo V regions, in a non-reciprocal

fashion, to create the antibody repertoire. This process likely proceeds through a two-ended DSB

intermediate and requires AID (Tang et al. 2006). AID-induced dU:dG mismatches are

processed by BER and MMR, as described above, to produce DSBs that are then repaired

through a homologous recombination-dependent DNA repair mechanism, to complete GCV.

Homologous recombination is discussed below in more detail.

1.2.2.3.3 Class switch recombination

CSR is the process by which the µ constant region of the antibody gene is replaced by a

downstream heavy chain constant region segment (γ, ε, α), providing the antibody with a

different effector function (Figure 5). CSR occurs at the immunoglobulin locus, through a

Page 35: Characterizing the role of the DNA Damage Response in

21

Figure 5

Figure 5. Class switch recombination. CSR can be broadly divided into two phases: 1) generation

of DSBs and 2) repair of DSBs. AID-induced cytidine deamination is engaged by BER/MMR,

leading to the formation of DSBs at the donor and acceptor switch regions. The DSBs are repaired in

the manner that replaces the donor constant region with the acceptor constant region, while deleting

the intervening DNA.

Page 36: Characterizing the role of the DNA Damage Response in

22

process in which the downstream acceptor constant region is joined with the upstream rearranged

V(D)J segment while the intervening DNA is deleted. This proceeds via a requisite DSB

intermediate, and can be broadly categorized into 2 phases: AID-induced DSB generation and

DSB repair (Stavnezer et al. 2008).

Upon activation, AID is expressed in B cells, and transcription is initiated at the donor and

acceptor switch regions, providing single-stranded DNA substrates for AID-mediated

deamination, as described above. The switch regions are located adjacent and upstream of each

constant region, with the exception of δ. However, δ constant region contains a vestigial

recombination site allowing for rare switching events to IgD (Kluin et al. 1995). Transcription is

initiated from the intronic (I) promoter that is located just upstream of the I exon, which precedes

each switch region (Kinoshita et al. 2001). Transcription proceeds through the I exon into the

associated switch region and ends downstream of the associated constant region, producing a

sterile transcript (also referred to as germline transcripts). Switch regions are highly repetitive

sequences that are G-rich on the nontemplate or coding strand and can range from 1-12 kilo-

basepairs in length (Chaudhuri et al. 2004). While Sμ, Sε and Sα are made of pentameric repeat

sequences, the Sγ subclasses Sγ1, Sγ2a, Sγ2b and Sγ3 are made of repeats of 49-52 basepairs

(Chaudhuri et al. 2004). Depending on the type of B cell activation, different acceptor switch

regions are transcribed. For example, in ex vivo B cells, LPS stimulation leads to Sγ2b and Sγ3

transcription, LPS with IL4 leads to Sγ1 and Sε transcription, LPS with IFNγ leads to Sγ2a

transcription, and LPS with TGFβ and IL5 leads to Sα transcription (Chaudhuri et al. 2004).

Hence, the cytokine milieu of the activated B cell dictates the constant region that will be

selected for CSR.

Page 37: Characterizing the role of the DNA Damage Response in

23

As described above, AID induces dU:dG mismatches at the G1 phase of cell cycle and

these mutations are subsequently processed by MMR or BER (Schrader et al. 2005; Stavnezer et

al. 2005; Schrader et al. 2007). MMR/BER processing of the AID-induced mutations cause

single-stranded DNA breaks, and if these breaks occur on opposite strands and are close enough

to one another, staggered DSBs are generated at the switch regions (Schrader et al. 2005;

Stavnezer et al. 2005). Consistent with this notion, although single-deficient UNG-/- or Msh2-/-

mice have reduced switching, switching is abolished in double-deficient Msh2-/- UNG-/- mice

(Xue et al. 2006). These DSBs elicit the DNA damage response (DDR), which can initiate DNA

repair, leading to long-range end joining between the donor and acceptor switch regions. DNA

repair is primarily mediated by non-homologous end joining (NHEJ) or alternatively by

microhomology-mediated end joining (MMEJ), thus completing CSR in the G1 phase of the cell

cycle (Boboila et al. 2012). However, AID-induced breaks have been reported to escape cell

cycle checkpoints and progress from G1 into S phase of cell cycle, where they are detected and

repaired by homologous recombination (Hasham et al. 2012). DDR and other DNA repair

pathways will be discussed in further detail below.

1.3 THE DNA DAMAGE RESPONSE (DDR)

DSBs can be generated by various means, including environmental cues like irradiation or

ultraviolet light, external chemical hazards, oxidative stress, errors occurring during DNA

replication, and during the process of antibody diversification. These DSBs ubiquitously elicit

the DNA Damage Response (DDR) pathway. DDR is an elaborate pathway in which factors are

recruited to the DSB site within minutes of DSB induction, in a highly regulated and sequential

fashion. This signalling network involves the initial “sensors” that detect the DSB and pass the

Page 38: Characterizing the role of the DNA Damage Response in

24

signal onto “transducers”, which then activate the “effectors” leading to cell cycle arrest and

DNA repair. The DDR pathway activated by DSBs can be broadly divided into three distinct yet

inter-connecting cascades, each initiated by one of three different sensor complexes, Mre11-

Rad50-Nbs1 (MRN)/ATM, the Ku70/Ku80-DNAPKcs, and ATR-ATRIP (Yang et al. 2003). In

each of these 3 pathways, the initial DSB recognition involves a PI3-kinase-like serine/threonine

kinase, ATM, DNAPK and ATR respectively (Yang et al. 2003). This recognition triggers a

cascade of signalling events, with highly regulated recruitment and/or post-translational

modifications of DDR factors. Accumulation of these DDR factors at the ionizing radiation (IR)-

induced DSB can be observed as irradiation-induced nuclear foci (IRIF) by confocal microscopy.

This pathway culminates in cell cycle arrest and DNA repair, or apoptosis in the case of aberrant

repair.

1.3.1 DNA damage response in class switch recombination

During CSR, the AID-induced DSBs activate ATM signalling (Lumsden et al. 2004;

Reina-San-Martin et al. 2004). The MRN/ATM complex is recruited to the DNA break (Lee et

al. 2004; Lee et al. 2005), where ATM phosphorylates histone H2AX (Stiff et al. 2004). γ-

H2AX (phosphorylated H2AX) signals the recruitment of MDC1 (Goldberg et al. 2003; Stewart

et al. 2003; Stucki et al. 2005), which recruits even more MRN/ATM, thereby leading to an

amplification of the γH2AX signal (Lukas et al. 2004; Lou et al. 2006). ATM also

phosphorylates MDC1, allowing for the recruitment of an E3 ubiquitin ligase, RNF8, which acts

with the E2-conjugating enzyme Ubc13, to mediate mono-ubiquitination of H2A-type histones

(Huen et al. 2007; Kolas et al. 2007; Mailand et al. 2007). Mono-ubiquitinated H2A-type

histones signal the recruitment of another E3 ubiquitin ligase, RNF168, which also acts with

Page 39: Characterizing the role of the DNA Damage Response in

25

Ubc13 to protect and amplify the RNF8-mediated ubiquitination (Doil et al. 2009; Pinato et al.

2009; Stewart et al. 2009). Specifically, RNF168 mediates H2A-K15 mono-ubiquitination at the

break site (Mattiroli et al. 2012). This H2A post-translational modification is recognized by the

ubiquitination-dependent recruitment motif of 53BP1 leading to the recruitment of 53BP1

(Fradet-Turcotte et al. 2013). 53BP1 is in turn phosphorylated by the master kinase, ATM, which

leads to the recruitment of Rif1 (Silverman et al. 2004; Di Virgilio et al. 2013; Escribano-Diaz et

al. 2013; Feng et al. 2013). 53BP1-Rif1 control the DNA repair pathway choice in the G1 phase

of the cell cycle by preferentially initiating the NHEJ pathway (Chapman et al. 2013; Escribano-

Diaz et al. 2013; Feng et al. 2013; Zimmermann et al. 2013). As discussed further below, the

DDR factors, ATM, NBS1, γ-H2AX, MDC1, RNF8, RNF168, 53BP1and Rif1 belong to the

53BP1-dependent DDR network, and have all been demonstrated to play a role in CSR (Figure

6). Identifying that RNF8 and RNF168 play a role in CSR is part of my PhD thesis, and will be

discussed further in Chapter 2. This 53BP1 dependent pathway is thought to mediate CSR by

facilitating long-range DNA end joining (Difilippantonio et al. 2008; Dimitrova et al. 2008), and

by protecting DNA ends from DNA resection, thus promoting NHEJ (Bothmer et al. 2010;

Bunting et al. 2010; Bothmer et al. 2011). However, the molecular mechanism by which 53BP1

and associated DDR factors mediate CSR remains to be elucidated.

1.3.2 ATM

Ataxia-telangiectasia (A-T) is an autosomal, recessive disorder in humans and mice that is

caused by a deficiency in ATM, and is characterized by radiosensitivity, neurodegeneration,

genome instability, cancer predisposition and immune deficiency (Savitsky et al. 1995; Ziv et al.

1997). ATM is a PI3-Kinase-like kinase that phosphorylates multiple targets, preferentially at a

Page 40: Characterizing the role of the DNA Damage Response in

26

Figure 6

Figure 6. The DNA damage response in class switch recombination. DDR factors are

recruited to the break in a sequential fashion, displayed above from left to right. The break is

recognized by the MRN/ATM complex, leading to the phosphorylation of H2AX (γH2AX).

γH2AX signals the recruitment of MDC1, which is also phosphorylated by ATM. This recruits

RNF8, which mediates histone H2A and H2AX ubiquitination, which in turn recruits RNF168.

RNF168 adds to the ubiquitination at the DNA break, and specifically leads to H2A-K15-

ubiquitin, leading in turn to the recruitment of 53BP1. 53BP1 is also phosphorylated by ATM

leading to the recruitment of Rif1. 53BP1-Rif1 block DNA resection and promote NHEJ.

Page 41: Characterizing the role of the DNA Damage Response in

27

S/TQ consensus motif (Kim et al. 1999). It belongs to the family of PI3-Kinase-like kinases,

which includes ATR, DNAPKcs, SMG1, mTOR and TRRAP (Shiloh et al. 2013). ATM kinase

activity is important for cell cycle regulation, DNA repair and apoptosis (Lavin et al. 2007).

Deficiency in ATM causes aberrant G1/S phase cell cycle checkpoint (Xie et al. 1998). In

humans, ATM is maintained in an inactive state as homodimers within the nucleus, but upon

DSB induction, ATM dissociates into active monomers, which undergo autophosphorylation at

Ser1981 and relocate to the DNA break site (Bakkenist et al. 2003). It has been suggested that

ATM Ser1981 autophosphorylation is not required for the recruitment of ATM, but rather for its

retention at the DNA break (So et al. 2009). In mice however, ATM autophosphorylation does

not affect activation (Pellegrini et al. 2006). Upon activation, ATM phosphorylates a multitude

of substrates including H2AX, NBS1, MDC1, 53BP1, Chk2, p53 and MDM2(Shiloh et al.

2013).

A-T patients and ATM-/- mice develop lymphomas due to translocations resulting from

abnormal V(D)J recombination, implicating ATM in primary antibody diversification (Taylor et

al. 1996; Xu et al. 1996; Liyanage et al. 2000). Indeed, ATM-/- mice have reduced lymphocyte

development in support of this notion (Xu et al. 1996). This result prompted investigators to test

whether ATM also acts in CSR. Interestingly, ATM-deficient mice exhibit approximately 70%

reduction in CSR without affecting intrinsic B cell growth, germline switch region transcription,

SHM or intra-switch recombination (Lumsden et al. 2004; Reina-San-Martin et al. 2004). The

CSR defect observed in ATM-/- mice was attributed to defective long-range DNA end joining

(Reina-San-Martin et al. 2004).

Page 42: Characterizing the role of the DNA Damage Response in

28

1.3.3 Mre11-Rad50-Nbs1 (MRN)

ATM alone cannot act as a DNA damage sensor. Rather, to obtain complete activation,

ATM must interact with the MRN complex (Uziel et al. 2003; Lee et al. 2004; Lee et al. 2005),

as deficiency in either Mre11 or Nbs1 can lead to reduced ATM recruitment to DSBs, reduced

ATM autophosphorylation at S1981 and reduced phosphorylation of ATM substrates. Patients

with Nbs1 mutations present with Nijmegan breakage syndrome and patients with mutations in

Mre11 present with ataxia-telangiectasia-like disorder, both of which display similar clinical

symptoms to A-T patients, which include radiosensitivity, neurodegeneration, increased cancer

predisposition and immune deficiency, further indicating an epistasis between ATM and MRN

(Varon et al. 1998; Stewart et al. 1999). Deletion of any of the three MRN components leads to

embryonic lethality in mice (Xiao et al. 1997; Luo et al. 1999; Zhu et al. 2001).

Mre11 forms the central element of the complex, as it binds to Rad50, DNA and Nbs1

(Williams et al. 2007). Nbs1 in turn interacts with ATM, and together it forms the MRN-ATM

sensor complex (You et al. 2005). Mre11 can mediate DNA end processing due to its 3’-5’

exonuclease, single-stranded DNA endonuclease and DNA unwinding capabilities. Rad50, on

the other hand, forms a homodimer and holds the two DNA ends together. The Rad50 protein

contains ATP-Binding cassette (ABC) ATPase, Zn hook, and coiled coil domains. Nbs1 contains

BRCA1 carboxy-terminal (BRCT) and Forkhead-associated (FHA) domains, and S/TQ sites; and

is linked to cell cycle checkpoint regulation (Kobayashi et al. 2004). BRCT and FHA domains

are very common to DDR factors. While BRCT is a protein-protein interaction motif, FHA is a

phospho-peptide binding motif. Together, the MRN complex is important for 1) sensing DSBs

and forming a complex with ATM to initiate the DDR, 2) aiding in DNA repair via DNA end

Page 43: Characterizing the role of the DNA Damage Response in

29

processing, and 3) providing stability by tethering the DNA DSB ends together and preventing

chromosomal detachment (Williams et al. 2007).

Mouse models have been used to investigate the role of MRN in CSR. Conditional Nbs1

knock-out mice, in which Nbs1 is specifically deleted in the B cell compartment, have decreased

B cell proliferation and increased genome instability (Reina-San-Martin et al. 2005).

Importantly, CSR is reduced by about 50% in Nbs1-/- mice (Reina-San-Martin et al. 2005).

Previous work had demonstrated that Nbs1 foci can form at immunoglobulin heavy chain locus

in an AID-dependent manner during CSR, which further implicates Nbs1 in CSR (Petersen et al.

2001). Mre11 was also found to co-localize with immunoglobulin heavy chain locus at a much

higher frequency in the presence of AID as compared to AID-/- cells, suggesting that like Nbs1,

Mre11 forms AID-dependent foci during CSR (Reina-San-Martin et al. 2005). Mouse B cells

that were deficient for the whole MRN complex, generated by deletion of Mre11, showed 80%

decrease in CSR ex vivo, while Mre11 nuclease-deficient mouse B cells, which maintained

normal levels of MRN, showed a 50% decrease in CSR (Dinkelmann et al. 2009). Persistence of

chromosomal breaks at the immunoglobulin locus of MRN-deficient cells demonstrated that it

likely plays a role in DNA repair during CSR (Dinkelmann et al. 2009).

1.3.4 H2AX

H2AX is one of three H2A core histone variants, and corresponds to 2-25% of the H2A

found within mammalian cells (Yin et al. 2009). Deficiency of H2AX leads to radiation

sensitivity, genome instability and G2/M checkpoint defect in mammalian cells (Bassing et al.

2002). H2AX-/- mice display radiation sensitivity, genome instability, retarded growth, reduced

spermatogenesis, and immune deficiency (Celeste et al. 2002). Upon DNA damage, H2AX is

Page 44: Characterizing the role of the DNA Damage Response in

30

immediately phosphorylated, typically within minutes, on its C-terminus at Ser139 by PI3K-like

kinases, ATM, ATR and DNAPK (Burma et al. 2001; Hammond et al. 2003; Stiff et al. 2004).

Accepted as a classical marker of DNA damage, phosphorylated histone variant H2AX (γH2AX)

is part of the initial phase of the DDR. In response to DNA damage, amplified ATM kinase

activity can lead to γH2AX formation extending up to 2 mega-basepairs surrounding the DNA

break site. This forms a docking site for accumulation of downstream DDR factors necessary for

cell cycle checkpoint arrest and DNA repair.

The observation that RAG induces γH2AX formation on the T cell receptor α locus and

immunoglobulin κ light chain locus suggested that γH2AX may play a role in V(D)J

recombination (Chen et al. 2000). However, H2AX-/- mice do not exhibit reduced lymphocyte

development, suggesting that H2AX does not have a direct role in the repair of RAG-induced

DSBs. Instead, H2AX-/- mice displayed increased chromosomal translocations originating from

RAG-induced DSBs (Bassing et al. 2003). Hence, γH2AX may prevent DSBs from progressing

to chromosomal translocations, possibly facilitating the tethering of DNA strands together, so as

to prevent their detachment and translocation (Franco et al. 2006; Yin et al. 2009). γH2AX foci

were also observed in the constant region of switching cells (Petersen et al. 2001). These γH2AX

foci co-localized with Nbs1 foci and were dependent on AID. H2AX deficiency leads to

approximately 70-80% reduction of CSR, without affecting germline transcription, B cell

proliferation, SHM and intra-switch recombination (Reina-San-Martin et al. 2003). Akin to

ATM, γH2AX is thought to be important for the long-range DNA end joining aspect of CSR

(Reina-San-Martin et al. 2003).

Page 45: Characterizing the role of the DNA Damage Response in

31

1.3.5 MDC1

MDC1 acts in conjunction with H2AX to promote cell cycle arrest and recruit DNA repair

factors to the DNA break site (Stewart et al. 2003). MDC1-deficient cells display defects in

intra-S- and G2/M-phase cell cycle arrest upon IR (Stewart et al. 2003). This protein contains

two tandem BRCT domains at its C-terminus, and a FHA domain at its N-terminus (Stewart et

al. 2003). As well, MDC1 has a cluster of S/TQ sites near its N-terminus, targeted for

phosphorylation by PI3K-like kinases including ATM and ATR (Stewart et al. 2003). MDC1 co-

localizes with γH2AX within minutes of IR induced DNA damage, and requires γH2AX to form

IR-induced foci, suggesting that MDC1 recruitment follows γH2AX formation (Stewart et al.

2003). MDC1 requires its tandem BRCT domain to bind to the C-terminus of γH2AX (Stucki et

al. 2005). This γH2AX-MDC1 interaction is required for effective recruitment and foci

formation of factors in the DDR network, including 53BP1, BRCA, Nbs1 and phosphorylated

ATM, and for achieving radio-resistance in cells, but is dispensable for regulating intra-S phase

checkpoint (Stewart et al. 2003; Stucki et al. 2005). MDC1 may not be required for the initial

γH2AX signal, but is required for its maintenance, either by blocking de-phosphorylation of

γH2AX by phosphatases or by mediating retention of activated ATM at the break site (Stucki et

al. 2005). MDC1 likely facilitates the accumulation of activated ATM at the break, by physically

interacting with Nbs1, which is a component of the MRN complex that associates with ATM

(Lukas et al. 2004; Chapman et al. 2008). Through phosphorylation-dependent MDC1-Nbs1

interactions, recruitment of MDC1 to the break site leads to further deposition and retention of

MRN-ATM complex surrounding the break, which in turn can amplify the ATM-dependent

γH2AX signal at the break (Lou et al. 2006). This MDC1-dependent accumulation of active

ATM is also important for the ATM-dependent phosphorylation of downstream factors like

Page 46: Characterizing the role of the DNA Damage Response in

32

Chk1 and Chk2 (Lou et al. 2006). MDC1-/- mice, like ATM-/- and H2AX-/- mice, show radiation

sensitivity, growth retardation, male infertility, genome instability and immune deficiency (Lou

et al. 2006). Interestingly, MDC1-/- mice only display a 25-50% decrease in the ability to

undergo CSR (Lou et al. 2006).

1.3.6 RNF8

Upon IR, phosphorylation of MDC1 in an ATM-dependent manner leads to the recruitment

of RNF8 (Kolas et al. 2007). In mice, RNF8 deficiency causes defective spermatogenesis,

growth retardation, radiosensitivity, genome instability, cancer predisposition and immune

deficiency (Li et al. 2010; Santos et al. 2010). RNF8 is an E3 ubiquitin ligase containing a C-

terminal RING finger domain and an N-terminal FHA domain (Huen et al. 2007; Mailand et al.

2007). Upon IR, RNF8 co-localizes with γH2AX (Huen et al. 2007; Mailand et al. 2007). This

co-localization depends on H2AX, Ser139 residue on H2AX, and MDC1; and requires the RNF8

FHA domain to bind to phosphorylated MDC1 (Huen et al. 2007; Kolas et al. 2007; Mailand et

al. 2007). The recruitment of DDR factors downstream of RNF8 requires the RNF8 RING

domain to mediate ubiquitination at the DNA break (Huen et al. 2007; Kolas et al. 2007;

Mailand et al. 2007). RNF8 acts with the E2 ubiquitin conjugating enzyme, Ubc13, to mediate

mono-ubiquitination of histones H2A and H2AX (Huen et al. 2007; Kolas et al. 2007; Mailand

et al. 2007). This ubiquitination is a prerequisite for the recruitment of downstream factors,

BRCA1 and 53BP1 (Huen et al. 2007; Kolas et al. 2007; Mailand et al. 2007). RNF8-/- cells

displayed partial 53BP1 IR-induced foci formation, suggesting the possibility of an RNF8-

independent mechanism of 53BP1 recruitment (Li et al. 2010). RNF8 deficiency leads to

increased IR sensitivity and disruption of G2/M cell cycle checkpoint (Huen et al. 2007; Kolas et

Page 47: Characterizing the role of the DNA Damage Response in

33

al. 2007). As part of my PhD thesis, I collaborated with Richard Chahwan to demonstrate that

RNF8 also plays a role in CSR (Chapter 2 below). Subsequent studies on RNF8-/- mice showed a

50-70% decrease in CSR (Li et al. 2010; Santos et al. 2010).

1.3.7 RNF168

Following RNF8 recruitment and activity, another E3 ubiquitin ligase, RNF168, binds to

the DNA break (Doil et al. 2009). RNF8-mediated ubiquitination of histones H2A at the DNA

break site can be bound by the RNF168 Motif Interacting with Ubiquitin (MIU) domain (Stewart

et al. 2009). RNF168 contains two MIU domains, an N-terminal and C-terminal, in which the C-

terminal MIU is predominantly responsible for binding to ubiquitin at the DNA break (Stewart et

al. 2009). RNF168 also contains a RING domain that acts with the E2-conjugated enzyme,

Ubc13, in protecting and amplifying the RNF8-mediated ubiquitination, leading to the formation

of K63-linked poly-ubiquitin chains at the DNA break site (Doil et al. 2009; Panier et al. 2009;

Stewart et al. 2009). Further studies showed that RNF168 mediates mono-ubiquitination of

histone H2A on residues K13-15 (Mattiroli et al. 2012). RNF168 is crucial for instigating both

arms of DDR, leading to the recruitment of Brca1 on one side and 53BP1 on the other (Doil et

al. 2009; Stewart et al. 2009). The Brca1 complex, made up of Bard1, Abraxas and Rap80, binds

to the RNF8/168-mediated ubiquitination at the DNA break through the Rap80 ubiquitin-binding

motif (Doil et al. 2009). The Brca1 pathway leads to DNA end resection, promoting HR, and is

beyond the scope of this thesis and so will not be further discussed. RNF168 is also crucial for

the recruitment of the highly characterized DDR factor, 53BP1, which blocks DNA resection,

promoting NHEJ, and will be discussed further below.

Page 48: Characterizing the role of the DNA Damage Response in

34

Along with RNF8, Richard Chahwan and I also published that RNF168 plays a role in

CSR (Chapter 2 below). Subsequent studies on RNF168-/- mice corroborated our work,

demonstrating that RNF168-/- mice have 40-50% reduction in CSR (Bohgaki et al. 2011). These

mice, akin to all DDR factors, are radiosensitive, have impaired spermatogenesis and are

immunodeficient due to deficiencies in both CSR and V(D)J recombination (Bohgaki et al.

2011). Interestingly, RNF168 deficiency is linked to RIDDLE syndrome that has been identified

in one patient who displays with radiosensitivity, immunodeficiency, dysmorphic features and

learning difficulties, similar to the clinical features of A-T, ATLD and NBS patients mentioned

above (Stewart et al. 2007; Stewart 2009; Stewart et al. 2009; Bohgaki et al. 2011).

1.3.8 53BP1

53BP1 was initially identified as a p53-interacting factor in yeast 2-hybrid studies

(Iwabuchi et al. 1994). In comparison to p53-/- mice, p53-/- 53BP1-/- double-deficient mice

develop increased incidence of tumors, particularly T cell lymphomas and to a lesser extent, B

cell lymphomas, sarcomas and teratomas (Morales et al. 2006). While p53-/- mice present with

aneuploidy, the double-deficient mice present with both aneuploidy and chromosomal

translocations, suggesting that 53BP1 acts synergistically with p53 to maintain genomic stability,

possibly through its functions in cell cycle checkpoint regulation and DNA repair (Morales et al.

2006).

53BP1 shares homology with the prototypical cell cycle checkpoint factors Rad9 in

budding yeast and Crb2 in fission yeast (Schultz et al. 2000). Human cells deficient in 53BP1 are

sensitive to IR, and have deregulated G2/M and intra-S phase checkpoints (Wang et al. 2002).

Upon IR, 53BP1 is itself phosphorylated in an ATM-dependent manner (Anderson et al. 2001;

Page 49: Characterizing the role of the DNA Damage Response in

35

Rappold et al. 2001), and 53BP1 contributes to optimal ATM-mediated phosphorylation of

Checkpoint kinase 2 (Chk2), leading to cell cycle arrest and G2/M checkpoint control

(Fernandez-Capetillo et al. 2002; Wang et al. 2002; Peng et al. 2003; Wilson et al. 2008).

53BP1 protein contains a glycine-arginine stretch, a tandem BRCT domain that to date has

not been identified with a specific function, a tandem tudor domain, and several phosphorylation

sites. 53BP1 forms foci that co-localize with γH2AX, Mre11 and Nbs1 upon DNA damage

(Schultz et al. 2000). 53BP1 relocates to DSBs upon DNA damage, as its tudor domain tethers to

the DNA break site by binding to dimethylated-H4K20 residues that surround the DSB (Botuyan

et al. 2006). Dimethylated-H4K20 is a constitutively present histone mark, initially causing a

puzzle as to how 53BP1 recruitment to a constitutive histone mark was restricted to DNA

damage. Initially, it was believed that dimethylated-H4K20 may be buried in the absence of

DNA damage, but DSBs cause chromatin relaxation/modification exposing this constitutive

histone mark for 53BP1 access. Additionally, the methyltransferase MMSET was found to

increase H2K20 methylation specifically at DSBs, providing additional substrates for 53BP1

binding (Pei et al.). However, further investigations suggested alternative, RNF8/168-dependent

mechanisms of exposing this histone mark for 53BP1 binding. One study suggested that

KDM4A, also called JMJD2A, constitutively binds to dimethylated-H4K20, preventing 53BP1

binding (Mallette et al.). However, DSB-induced RNF8/168 activity leads to the degradation of

KDM4A/JMJD2A at the break site, exposing dimethylated-H4K20 for 53BP1 recruitment

(Mallette et al.). Another study suggested that the VCP/p97-UFD1-NPL4 complex is recruited to

the DNA break in a RNF8-dependent fashion, and is important for the recruitment of 53BP1 and

BRCA1 (Meerang et al.). Acs et al. also showed that VCP/p97 is recruited to DSBs in an

RNF8/168-dependent fashion, but they suggest that it is important for the ubiquitination and

Page 50: Characterizing the role of the DNA Damage Response in

36

removal of a Polycomb protein L3MBTL1 from dimethylated-H4K20, thus exposing this histone

mark for 53BP1 (Acs et al. 2011). However, the ground-breaking finding that 53BP1 contains an

ubiquitination-dependent recruitment motif, located adjacent to its tandem tudor domain, that

binds to RNF168-mediated H2A-K15 mono-ubiquitin, finally solved the puzzle of DNA

damage-induced 53BP1 recruitment (Fradet-Turcotte et al. 2013). 53BP1 accumulation upon

DNA damage is dependent on RNF168 function, as 53BP1 localizes to the DNA break site by

binding to dimethylated-H4K20 via its tandem tudor domain and H2A-K15-mono-ubiquitin via

its ubiquitination-dependent recruitment motif (Fradet-Turcotte et al. 2013). However, low levels

of 53BP1 can transiently bind at the DNA break in a manner independent of MDC1, RNF8 and

RNF168, and this is likely reflective of low levels of 53BP1 binding to dimethylated-H4K20me2

independent of H2A-K15-mono-ubiquitin (Fradet-Turcotte et al. 2013).

Most interestingly, CSR is almost completely ablated in 53BP1-deficient mice (Manis et

al. 2004; Ward et al. 2004). The 53BP1 tudor domain, oligomerization domain and N-terminal

ATM phosphorylation motifs are all important for CSR (Bothmer et al. 2011). These mice have

no abnormality in AID expression, germline switch region transcription, or SHM, suggesting that

while the function of AID in CSR is not affected, 53BP1 may be involved in the ligation of DNA

ends (Manis et al. 2004; Ward et al. 2004). More specifically, 53BP1 is thought to facilitate the

switch-synapse formation by enhancing long-range DNA end joining and by protecting the DNA

ends from DNA resection (Details provided below). However, the precise mechanism of 53BP1

function and the downstream events following 53BP1 recruitment, at the molecular level, in CSR

remains unknown. Intriguingly, of the DDR factors characterized to date, 53BP1 deficiency

leads to the most severe CSR defect.

Page 51: Characterizing the role of the DNA Damage Response in

37

1.3.8.1 53BP1 in long-range end joining:

53BP1 has been suggested to mediate long-range end joining in CSR. In support of this

notion, 53BP1-deficient B cells show increased levels of intra-switch recombination (Reina-San-

Martin et al. 2007). Furthermore, 53BP1 facilitates long-range NHEJ during V(D)J

recombination at the TCR locus (Difilippantonio et al. 2008), and 53BP1 promotes NHEJ of

dysfunctional telomeres by increasing chromatin mobility (Dimitrova et al. 2008).

During the DDR to DSBs in heterochromatin, 53BP1 recruitment forms part of an

amplification loop involving Mre11 and Nbs1, and this leads to the concentration of active ATM

at the DNA break (Noon et al. 2010). This amplification of active ATM, in addition to the first

MDC1-Nbs1 dependent amplification described above, causes high accumulation of active ATM

leading to ATM-mediated phosphorylation of Kap1 on residue S824 at the break site, which can

be observed as phosphorylated Kap1 foci (Noon et al. 2010). Kap1 phosphorylation is

particularly important for heterochromatic DNA repair, as phosphorylation of Kap1 leads to

release of chromatin from a repressed state, possibly allowing for increased chromatin mobility

(Ziv et al. 2006). Although the 53BP1-dependent, ATM-mediated Kap1 foci are crucial to

heterochromatic DNA repair, these findings may be extended to CSR in which relaxed chromatin

may promote increased DNA end mobility facilitating long-range end joining, although this

notion has yet to be investigated (Noon et al. 2010).

1.3.8.2 53BP1 and DNA end resection:

Emerging evidence indicates that 53BP1 can block DNA ends from resection, thus

promoting NHEJ-mediated DNA repair over the homology-based DNA repair mechanisms,

including HR and MMEJ (Bothmer et al. 2010; Bunting et al. 2010). 53BP1-null cells displayed

Page 52: Characterizing the role of the DNA Damage Response in

38

increased resection of DNA ends leading to MMEJ repair (Bothmer et al. 2010). This DNA

resection, which is antagonized by the presence of 53BP1 at the DNA ends, requires ATM

activation (Bothmer et al. 2010). The DNA resection and processing are mediated by CtIP, Exo1

and RecQ DNA helicase (Bothmer et al. 2013). The 53BP1 chromatin-binding tudor domain,

oligomerization domain and the cluster of ATM phosphorylation sites in the N-terminus are all

required for the ability of 53BP1 to protect DNA ends from processing (Bothmer et al. 2011). In

agreement with these findings, 53BP1-mediated blockade of DNA end resection inhibited

homologous recombination in BRCA1-null cells, leading to error-prone NHEJ-mediated repair

in S phase cells, in turn causing genome instability (Bunting et al. 2010). Deleting 53BP1

rescued genomic instability and the translocation phenotype of BRCA1-deficient cells, by

alleviating the 53BP1-mediated block on homologous recombination (Bunting et al. 2010).

53BP1 possibly inhibits DNA resection by binding to and blocking DNA ends from access to

DNA end processing enzymes. While 53BP1-mediated DNA end protection is advantageous in

G1 cells, promoting NHEJ and CSR, it is deleterious in cells that are in S phase, since it would

block HR and promote mutagenic repair by NHEJ (Bunting et al. 2010). Parallel studies in the

context of V(D)J recombination showed that γH2AX and MDC1, both of which act upstream of

53BP1, inhibited CtIP-mediated DNA resection of RAG-induced DSBs (Helmink et al. 2011).

Intriguingly, γH2AX formation and CtIP activation are both dependent on ATM activity

(Bothmer et al. 2010; Bothmer et al. 2013). Hence ATM initiates both the pathway that

suppresses DNA resection and the pathway that promotes DNA resection (Bothmer et al. 2010).

CtIP and DNA resection are largely restricted to S/G2/M cell cycles, where HR dominates, and

are blocked in G1 cell cycle to promote NHEJ (Escribano-Diaz et al. 2013). In cells at G1, CtIP

Page 53: Characterizing the role of the DNA Damage Response in

39

activation would promote error-prone MMEJ, causing chromosomal deletions and genome

instability (Helmink et al. 2011).

1.3.8.3 53BP1 recruits Rif1 and PTIP

While the ATM phosphorylation motifs clustered towards the N-terminus of 53BP1 are

important for end joining and CSR, they are dispensable for 53BP1 foci formation, suggesting

that ATM phosphorylation of 53BP1 is not required for its recruitment, but rather for possibly

recruiting downstream factors that promote NHEJ and inhibit HR (Escribano-Diaz et al. 2013).

Very recent work has revealed that a key downstream factor is Rif1, which acts in concert with

53BP1 to protect DNA ends from resection (Chapman et al. 2013; Di Virgilio et al. 2013;

Escribano-Diaz et al. 2013; Feng et al. 2013; Zimmermann et al. 2013). Rif1 depletion in

BRCA1-deficient cells phenocopies 53BP1 depletion in BRCA1-deficient cells, confirming

53BP1-Rif1 epistasis (Chapman et al. 2013; Escribano-Diaz et al. 2013; Feng et al. 2013).

ATM phosphorylation of 53BP1 is absolutely required for the 53BP1-Rif1 interaction,

leading to the recruitment of Rif1 to DNA breaks (Silverman et al. 2004; Di Virgilio et al. 2013;

Escribano-Diaz et al. 2013; Feng et al. 2013). Rif1 functions to inhibit DNA end processing by

CtIP, Exo1 and BLM helicase (Zimmermann et al. 2013). Rif1 and BRCA1 control DNA repair

pathway choice as the cell progresses through the cell cycle (Chapman et al. 2013; Escribano-

Diaz et al. 2013; Feng et al. 2013). First, 53BP1-Rif-dependent inhibition of DNA resection in

G1 antagonizes the BRCA1-CtIP dependent HR pathway and MMEJ, while promoting NHEJ

(Chapman et al. 2013; Escribano-Diaz et al. 2013; Feng et al. 2013). Second, the BRCA1-CtIP

pathway antagonizes Rif1 in S/G2, promoting DNA end resection and HR, whilst blocking NHEJ

(Chapman et al. 2013; Escribano-Diaz et al. 2013; Feng et al. 2013). Importantly, Rif1 depletion

Page 54: Characterizing the role of the DNA Damage Response in

40

leads to 80-90% reduction in CSR (Chapman et al. 2013; Di Virgilio et al. 2013; Escribano-Diaz

et al. 2013). Hence, Rif1 is an important CSR factor acting downstream of 53BP1.

Further investigation revealed that a 53BP1 mutant, with alanine substitution mutations in

the first 8 N-terminal phosphorylation sites, maintained Rif1 recruitment and CSR (Callen et al.

2013). However, this mutant was unable to rescue abnormal repair and genome instability in

BRCA1-deficient cells, and this is attributed to the inability of the 53BP1 mutant to bind to PTIP

(Callen et al. 2013). 53BP1 facilitates deleterious mutagenic repair in BRCA1-deficient cells by

promoting NHEJ and blocking HR in S phase cells (Bunting et al. 2010). While both the

functions of blocking DNA end resection and HR require 53BP1 N-terminal phosphorylation by

ATM, blocking DNA resection relies on 53BP1 interacting with Rif1, while blocking HR and

promoting mutagenic NHEJ during S phase rely on 53BP1 recruitment of PTIP (Callen et al.

2013). While CSR requires that 53BP1 interacts with Rif1, the interaction of 53BP1 with PTIP

appears to be dispensable for CSR (Callen et al. 2013).

1.3.9 DDR factors within the ATM network not investigated in the context of CSR

1.3.9.1 Deubiquitination in DDR

Multiple deubiquitylases (DUB) have been implicated in suppressing the DDR, including

USP3, USP16, BRCC36, POH1, Usp44 and OTUB1 (Jackson et al. 2013). These enzymes

antagonize RNF8-RNF168 ubiquitylation (Jackson et al. 2013). USP3 is involved in H2A/H2B-

deubiquitination, and has been linked to DDR (Nicassio et al. 2007). USP16 removes RNF8 and

RNF168-mediated ubiquitylation (Shanbhag et al. 2010). BRCC36 and the proteasome-

associated POH1 have a propensity towards removing K63-linked ubiquitin (Jackson et al.

2013). BRCC36 is part of the Rap80-BRCA1 complex that gets recruited to DSBs downstream

Page 55: Characterizing the role of the DNA Damage Response in

41

of RNF168. Usp44 was recently demonstrated to oppose the RNF8-RNF168-mediated H2A

ubiquitylation (Mosbech et al. 2013). Intriguingly, OTUB1 opposes RNF168 ubiquitylation

independent of its isopeptidase enzymatic activity, which displays specificity towards K48-

linked chains (Nakada et al. 2010). Instead, OTUB1 binds to and inhibits the RNF168 E2

ubiquitin conjugating enzyme, Ubc13 (Nakada et al. 2010; Blackford et al. 2011). OTUB1

simultaneously binds Ubc13 conjugated to ubiquitin and a free ubiquitin molecule in a manner

that places the two ubiquitins in a configuration similar to K48-linked ubiquitin, the inherent

product of OTUB1 enzyme function (Juang et al. 2012; Wiener et al. 2012). The authors suggest

that by co-opting the recognition of K48-linked chains, OTUB1 inhibits Ubc13 from facilitating

RNF168 K63-linked ubiquitylation (Juang et al. 2012).

1.3.9.2 H2B Ubiquitylation in DDR

In addition to H2A ubiquitylation, H2B ubiquitylation also plays a role in DDR (Moyal et

al. 2011). The RNF20-RNF40 complex mediates histone H2B mono-ubiquitylation at K120 in

humans and K123 in yeast, and has been linked to transcription (Kao et al. 2004; Zhu et al.

2005; Jung et al. 2012). DNA damage leads to the recruitment of a RNF20-RNF40 heterodimer

to DSBs, where it locally mono-ubiquitylates histone H2B (Moyal et al. 2011). While ATM is

dispensable for the recruitment of RNF20-RNF40, ATM-mediated phosphorylation of RNF20 is

required for its retention at DSBs (Moyal et al. 2011). However, this H2B mono-ubiquitylation

represents an arm of the ATM network that is independent of the MDC1-RNF8 pathway (Moyal

et al. 2011). H2B mono-ubiquitylation at DSBs is thought to facilitate chromatin decondensing,

allowing for the recruitment of DNA repair factors (Fierz et al. 2011).

Page 56: Characterizing the role of the DNA Damage Response in

42

Indeed, the RNF20-RNF40 complex was shown to be required for efficient repair by both

the NHEJ and HR repair pathways (Moyal et al. 2011). H2B-ubiquitination is prerequisite to

histone H3 methylation at K4 and K79 (Ng et al. 2002; Sun et al. 2002). RNF20 and H2B

mono-ubiquitination enables H3K4 methylation and recruitment of chromatin remodeling factor

SNF2h to DSBs (Nakamura et al. 2011). RNF20 and SNF2h recruitment to DSBs is critical for

CtIP-dependent DNA end resection, and BRCA1 and Rad51 loading for HR (Nakamura et al.

2011). RNF20 acts in HR, in part to promote chromatin remodelling, as the DNA intercalater

chloroquine, which induces chromatin relaxation by modifying chromosome structure, can

partially bypass the requirement for RNF20 during HR (Nakamura et al. 2011). Similarly,

RNF40 is important for cell cycle checkpoint activation, H3K56 acetylation and recruitment of

the Facilitates Chromatin Transcription (FACT) complex to chromatin upon DNA damage (Kari

et al. 2011). RNF40 and the FACT complex component, SUPT16H, are required for efficient

DNA resection and DNA repair, likely by promoting chromatin remodelling (Kari et al. 2011).

1.3.9.3 Sumoylation in DDR

Recent work has identified sumoylation as another post-translation modification relevant to

DDR. The PIAS4 Small Ubiquitin-like Modifier (Sumo) E3 ligase was shown to be an important

component of DDR, required for RNF168, K63-linked ubiquitin chain, 53BP1 and BRCA1 foci

formation, while the PIAS1 Sumo E3 ligase was specifically required for BRCA1-Rap80

recruitment (Galanty et al. 2009). PIAS1 and PIAS4 are themselves recruited to DNA breaks and

are responsible for Sumo1/2/3 accumulation at the breaks (Galanty et al. 2009). PIAS1 targets

BRCA1 for sumoylation, while PIAS4 can target RNF168, BRCA1 and 53BP1 for sumoylation

(Galanty et al. 2009). Both Sumo E3 ligases are important for HR and NHEJ and depleting both

Page 57: Characterizing the role of the DNA Damage Response in

43

ligases confers the cells with radiation sensitivity (Galanty et al. 2009). Sumoylation is another

post-translational modification, in addition to phosphorylation and ubiquitination, utilized by the

DDR pathway.

1.3.9.4 RNF4 function

The discovery of RNF4 extended the complexities of the inter-crossing ubiquitination and

sumoylation networks in DDR. RNF4 is an E3 ubiquitin ligase and contains both a RING

domain and a SUMO-interacting motif (SIM). RNF4 accumulates at DSBs by virtue of its

SUMO-interaction motif that binds to the sumoylation added by PIAS1 and PIAS4 activity

(Galanty et al. 2012). Upon recruitment to DSBs, RNF4 functions in the protein turnover of

MDC1 and RPA (Galanty et al. 2012; Luo et al. 2012; Vyas et al. 2012). By ubiquitinating these

factors, RNF4 targets MDC1 and RPA for proteasomal degradation (Galanty et al. 2012; Luo et

al. 2012; Vyas et al. 2012). MDC1 turnover may facilitate DNA repair and timely recovery from

cell cycle arrest, possibly by providing access to allow for the recruitment of downstream DDR

factors, including Rap80-BRCA1 (Galanty et al. 2012; Guzzo et al. 2012; Luo et al. 2012). RPA

turnover is necessary for its replacement by BRCA2 and Rad51 on resected DNA in order to

complete homologous recombination (Galanty et al. 2012; Luo et al. 2012; Vyas et al. 2012).

However, an independent study suggested that RNF4 was required for RPA recruitment rather

than turnover (Yin et al. 2012). This was attributed to reduced recruitment of CtIP and DNA

resection, and hence fewer single-stranded DNA for RPA loading, in RNF4-depleted cells (Yin

et al. 2012). RNF4 was also suggested to regulate BRCA2 protein turnover (Vyas et al. 2012).

RNF4 depletion results in the persistence of γH2AX, RNF8, RNF168 and 53BP1 at DSBs and

inefficient DNA repair (Galanty et al. 2012; Luo et al. 2012; Yin et al. 2012). RNF4 functions in

Page 58: Characterizing the role of the DNA Damage Response in

44

both NHEJ and HR (Galanty et al. 2012; Luo et al. 2012; Vyas et al. 2012; Yin et al. 2012).

RNF4-deficient cells display increased sensitivity to DNA damage, and RNF-4 deficient mice

display increased sensitivity to DNA damage and defective spermatogenesis (Vyas et al. 2012).

1.3.10 DDR leads to cell cycle arrest

Cells can either be at a resting state, referred to as G0, or undergo cycling through the G1,

S, G2 and M phases of cell cycle. S phase, during which DNA replication occurs, and M phase,

during which mitosis occurs, are separated by the G1 and G2 phases of cell cycle. Cell cycle

progression is controlled by cyclin and cyclin-dependent protein kinases. G1/S transition, intra-S-

phase and G2/M transition checkpoints are in place to ensure that the cell cycle progression is

regulated. DNA damage-induced DDR will activate cell cycle checkpoints and cause an arrest in

cell cycle until the damage is repaired, preventing the cell from progressing to the next phase

without repairing the damage (Langerak et al. 2011). This is accomplished by the

phosphorylation and activation of the effector kinases, Chk1 and Chk2, by components of DDR,

including ATM and ATR (Latif et al. 2004; Kiyokawa et al. 2008; Reinhardt et al. 2009). While

replication-induced DNA damage activates the ATR pathway, leading to Chk1 activation, IR-

induced DNA damage activates the ATM pathway, leading to Chk2 activation (Matsuoka et al.

1998; Chaturvedi et al. 1999; Heffernan et al. 2002). Both effector kinases lead to the

phosphorylation and inactivation of Cdc25 phosphatase family members that regulate

cyclin/cyclin-dependent kinases, including Cdc2/Cdk1, involved in cell cycle checkpoint

regulation, thus causing cell cycle arrest allowing for DNA repair (Sancar et al. 2004; Kiyokawa

et al. 2008; Reinhardt et al. 2009).

Page 59: Characterizing the role of the DNA Damage Response in

45

1.4 DNA REPAIR

1.4.1 Non-homologous end joining (NHEJ)

Antibody diversification primarily relies on NHEJ for DNA repair (Soulas-Sprauel et al.

2007). Both V(D)J recombination and CSR mechanisms utilize NHEJ (Soulas-Sprauel et al.

2007). NHEJ is active during the entire cell cycle, but is primarily used during the G0 and G1-

phases of cell cycle (Lieber 2010). Since AID is expressed and prompts DSBs in the G1-phase of

the cell cycle (Schrader et al. 2007), these breaks are engaged by NHEJ to complete CSR.

NHEJ proceeds through a blunt-end ligation type of mechanism, and thus does not rely on

regions of homology to dictate repair (Lieber 2010). Hence, NHEJ joints are characterized by

very little or no homology (Stavnezer et al. 2010). NHEJ suppresses DNA resection and

chromosomal translocations, but can lead to error-prone DNA repair, due to short nucleotide

insertions or deletions at the joint (Boboila et al. 2012). As illustrated in Figure 7, the break is

recognized by the Ku70/Ku80 complex, which then recruits the catalytic domain, DNAPKcs, to

form the DNAPK holoenzyme (Gottlieb et al. 1993; Smider et al. 1998). This is followed by the

recruitment of XRCC4 and DNA Ligase IV to mediate blunt end joining (Grawunder et al. 1997;

Chen et al. 2000). Artemis and XLF/Cernunnos have also been implicated in NHEJ (Ahnesorg et

al. 2006; Buck et al. 2006). However, Ku70, Ku80, XRCC4 and Ligase IV are evolutionarily

conserved and make up the core components of NHEJ, while Artemis and DNAPKcs are

auxiliary components found in vertebrates (Boboila et al. 2012).

The first phase of NHEJ involves sensing of the break and formation of the break synapsis.

Ku70/Ku80 heterodimer recognizes and binds to the break, protecting the ends from resection,

although Ku displays lyase activity towards terminal abasic sites at DSBs (Roberts et al. 2010).

The Ku heterodimer then recruits DNAPKcs to form DNAPK. DNAPKcs can autophosphorylate

Page 60: Characterizing the role of the DNA Damage Response in

46

Figure 7

Figure 7. Non-homologous end joining. The DSB recruits Ku70/Ku80, which is followed by

the recruitment of DNAPKcs. Together, this forms the DNAPK holoenzyme. Subsequent

recruitment of the XRCC4-DNA Ligase IV completes repair. XRCC4-DNA Ligase IV function

is enhanced by XLF/Cernunnos.

Page 61: Characterizing the role of the DNA Damage Response in

47

itself and/or phosphorylate other proteins involved in NHEJ (Meek et al. 2008). Indeed,

DNAPKcs and ATM have incomplete redundancy, as deficiency of both kinases leads to a more

severe effect on end joining compared to single deficiency (Zha et al. 2011). Further, the

DNAPKcs autophosphorylation status can affect the protein’s affinity for DNA, and thus

regulate its dynamics at the break site (Uematsu et al. 2007). Autophosphorylation leads to

release of DNAPKcs from the break, providing the downstream NHEJ factors access to the break

(Cui et al. 2005).

The second phase of NHEJ is the ligation of the break, which requires the XRCC4/Ligase

IV complex, and the ancillary factor XLF. XRCC4/Ligase IV are recruited to the break, where

XRCC4 provides a structural role in stabilizing Ligase IV, which catalyzes the ligation of the

break. XLF can directly bind to and enhance XRCC4/Ligase IV-mediated blunt-end ligation

(Ahnesorg 2006). Mutations in XLF can cause radiosensitivity and defective V(D)J

recombination in humans (Buck et al. 2006).

Artemis and DNAPKcs may be relevant to NHEJ of breaks that require processing before

repair, especially if the two ends are incompatible. Artemis endonuclease is activated by

DNAPKcs-mediated phosphorylation (Ma et al. 2002). This endonuclease plays a key role in

V(D)J recombination, owing to its ability to cleave the coding joint hairpin, a requirement for

V(D)J recombination (Ma et al. 2002). Artemis nuclease activity may play a role in DNA end

resection in CSR (Franco et al. 2008). Incompatible ends may also require gap-filling by DNA

polymerases Polλ and Polμ (Ogiwara et al. 2011). Such nucleotide processing at the DNA breaks

can lead to error-prone DNA repair.

In vivo mouse models have been generated to test the role of each factor in CSR. Mice with

NHEJ core component deficiencies cannot complete B cell development, due to impaired V(D)J

Page 62: Characterizing the role of the DNA Damage Response in

48

recombination. The B cell development defect can be bypassed by engineering transgenic mice

with prearranged heavy and light chains (Ramiro et al. 2007). Using this technology,

Ku70/Ku80-deficient mice were shown to exhibit reduced CSR (Casellas et al. 1998; Manis et

al. 1998). This defect was shown to be independent of proliferative defects and was shown to be

B cell intrinsic (Casellas et al. 1998; Manis et al. 1998). Ku-deficient cells were able to generate

DSBs, but were unable to complete the repair, showing that Ku heterodimer is involved in the

DNA repair phase of CSR (Casellas et al. 1998; Manis et al. 1998). DNAPKcs, on the other

hand, showed inconsistent results. Using the same technology described above for Ku-deficient

mice, DNAPKcs-null mice with rescued B cell development exhibited reduced CSR to all

isotypes except IgG1 (Manis et al. 2002). However, severe combined immunodeficient (SCID)

mice transgenic for immunoglobulin heavy and light chains, but with a truncated, enzymatically

inactive DNAPKcs, maintained similar levels of CSR as control mice, suggesting that DNAPKcs

kinase activity may be dispensable to CSR (Bosma et al. 2002). The discrepancy between these

two studies could be due to the possibility that the SCID mice expressed a truncated or

enzymatically dead protein that plays a structural role in CSR (Ramiro et al. 2007).

Furthermore, DNAPKcs-deficient cells displayed impaired repair of a subset of AID-induced

breaks at the immunoglobulin heavy chain locus by FISH experiments, implicating DNAPKcs in

CSR (Franco et al. 2008). Artemis was considered dispensable for CSR, as Artemis-deficient

mice exhibited normal CSR, but fluorescent in situ hybridization assays also demonstrated that a

subset of AID-induced breaks relied on Artemis for repair (Franco et al. 2008). XRCC4 and

Ligase IV deletion in mice causes embryonic lethality due to neuronal apoptosis. However,

embryonic lethality can be rescued by deletion of p53 (Yan et al. 2007). Combining the p53-null

mutation with either a deletion of XRCC4 or Ligase IV also exhibited defective but not a

Page 63: Characterizing the role of the DNA Damage Response in

49

complete abrogation of CSR (Yan et al. 2007). A mouse model with a B-cell specific conditional

deletion of XRCC4 has also shown that in the absence of XRCC4 and NHEJ, there is a two-fold

reduction in CSR (Soulas-Sprauel et al. 2007).

1.4.2 Microhomology-mediated end joining (MMEJ)

MMEJ is an alternate pathway to NHEJ that has emerged recently. Mouse B cells deficient

in the core NHEJ factors, Ku, XRCC4, and Ligase IV, displayed ~50% reduced CSR compared

to controls (Soulas-Sprauel et al. 2007; Yan et al. 2007; Han et al. 2008; Boboila et al. 2010;

Boboila et al. 2010). Switch joints from these mice demonstrated short regions of homology,

suggestive of an alternative end joining mechanism to NHEJ. Indeed, patients with Ligase IV

deficiencies also showed marked increase in the use of microhomology at switch junctions (Pan-

Hammarstrom et al. 2005). This pathway, which is secondary to NHEJ and likely acts in the

absence of NHEJ, is referred to as MMEJ. MMEJ mediates ligation of the two DNA ends but

relies on sequence microhomology consisting of only a few nucleotides. MMEJ is more error-

prone than NHEJ, more often causing deletions and chromosomal translocations, leading to

genome instability (Yan et al. 2007; Boboila et al. 2010). However, the enzymes and factors

involved, as well as the mechanism of the MMEJ pathway, largely remain elusive. Msh2, Mlh1,

Exo1, Parp1, XRCC1, DNA Ligases I and III, CtIP and Mre11, have all been directly implicated

in MMEJ (Zhong et al. 2002; Liang et al. 2008; Robert et al. 2009; Rahal et al. 2010; Eccleston

et al. 2011; Saribasak et al. 2011; Zhang et al. 2011; Jia et al. 2013). However, XRCC1 and

DNA Ligase III were also shown to be dispensable for MMEJ in CSR (Boboila et al. 2012). CtIP

likely mediates DNA end resection, providing regions of microhomology required for MMEJ

(Zhang et al. 2011). Recent findings suggest that there may be more than one alternative end-

Page 64: Characterizing the role of the DNA Damage Response in

50

joining pathway to NHEJ (Boboila et al. 2010). Importantly, MMEJ is an alternate pathway

distinct from the core NHEJ factors, Ku70, Ku80, XRCC4 and Ligase IV (Yan et al. 2007;

Boboila et al. 2010).

1.4.3 Homologous recombination (HR)

HR relies on the genetic information present on the undamaged duplicate sister chromatid

or homologous chromosome to repair DNA damage, and so it is active during the S and G2

phases of cell cycle, after DNA replication has occurred. HR requires large regions of homology,

typically greater than 100 bp, to direct the repair (Dudas et al. 2004). Similar to MMEJ, HR

begins with DNA end resection at the 5’ ends, forming two 3’ single-stranded DNA ends

(Truong et al. 2013). One 3’ end invades the homologous chromosome producing a D loop

structure, where the complimentary strand of the intact chromosome is displaced and gets bound

by the other 3’single-stranded end (Dudas et al. 2004). DNA synthesis begins at the two 3’ ends

using the undamaged chromosomal strands as templates (Dudas et al. 2004). Subsequent ligation

forms Holliday junctions that can be resolved to produce a crossover or non-crossover end

product (Dudas et al. 2004). Rad51, Rad52, Rad55, Rad57, Brca1, Brca2, XRCC2, RPA and

MRN participate in HR, but the molecular functions have not been completely defined (Sancar et

al. 2004).

HR is critical for gene conversion, and was largely considered dispensable for CSR.

However, recent evidence suggests that AID-induced DSBs, and in particular off-target DSBs,

are engaged by HR (Hasham et al. 2010; Hasham et al. 2012; Lamont et al. 2013). The XRCC2-

dependent HR pathway plays a key role in protecting B cells from AID-induced off-target

effects, thus maintaining genome stability (Hasham et al. 2010). Although AID induces the off-

Page 65: Characterizing the role of the DNA Damage Response in

51

target DSBs in G1-phase of cell cycle, these breaks can bypass the G1/S cell cycle checkpoint,

progressing into S phase, where they are repaired by HR (Hasham et al. 2012). HR not only

protects the B cells from potential AID-induced genome instability, but can also successfully

complete CSR in S-phase (Hasham et al. 2012).

1.5 OBJECTIVES

It is well established that the 53BP1-dependent DDR pathway is required to mediate

CSR. However, the exact molecular function of this pathway, and the factors acting downstream

53BP1 leading up to DNA repair in CSR, are not fully understood. My objective during my

thesis work was to better understand how the DDR pathway functions at the molecular level in

CSR. To this end, I first established an in vitro system to study the role of DDR factors in CSR.

Using this system, I then screened for novel CSR factors, and from these CSR candidates,

identified potential DDR factors for further characterization. By characterizing the molecular

functions of novel DDR factors in CSR, we can better understand how the DDR pathway

facilitates CSR.

Page 66: Characterizing the role of the DNA Damage Response in

52

2 THE RNF8/RNF168 UBIQUITIN LIGASE CASCADE FACILITATES

CLASS SWITCH RECOMBINATION.

Published Work:

Ramachandran S1, Chahwan R1, Nepal RM, Frieder D, Panier S, Roa S, Zaheen A, Durocher

D, Scharff MD, Martin A. 2010. The RNF8/RNF168 ubiquitin ligase cascade facilitates class

switch recombination. PNAS USA. 107(2): 809-14

1 S.R. and R.C. contributed equally to this work.

Author contributions:

S. Ramachandran, R.C., D.D., M.D.S., and A.M. designed research;

S. Ramachandran, R.C., R.N., D.F., S.P., S. Roa, and A.Z. performed research;

S.P. and D.D. contributed new reagents/analytic tools;

S. Ramachandran, R.C., S. Roa, D.D., M.D.S., and A.M. analyzed data;

S. Ramachandran, R.C., D.D., M.D.S., and A.M. wrote the paper.

Page 67: Characterizing the role of the DNA Damage Response in

53

2.1 ABSTRACT

An effective immune response requires B cells to produce several classes of antibodies through

the process of class switch recombination (CSR). Activation-induced cytidine deaminase (AID)

initiates CSR by deaminating cytidines at switch regions within the immunoglobulin locus. This

activity leads to double-stranded DNA break formation at the donor and recipient switch regions

that are subsequently synapsed and ligated in a 53BP1-dependent process that remains poorly

understood. The DNA Damage Response E3 ubiquitin ligases, RNF8 and RNF168, were

recently shown to facilitate recruitment of 53BP1 to sites of DNA damage. Here, we show that

the ubiquitination pathway mediated by RNF8 and RNF168 plays an integral part in CSR. Using

the CH12F3-2 mouse B cell line that undergoes CSR to IgA at high rates, we demonstrate that

knockdown of RNF8, RNF168, and 53BP1 leads to a significant decrease in CSR. We also show

that 53BP1-deficient CH12F3-2 cells are protected from apoptosis mediated by the MDM2

inhibitor, Nutlin-3. In contrast, deficiency in either E3 ubiquitin ligase does not protect cells

from Nutlin-3-mediated apoptosis, indicating that RNF8 and RNF168 do not regulate all

functions of 53BP1.

2.2 INTRODUCTION

Part of an effective immune response requires the production of antibodies of different

classes, each of which mediates a different effector function. This process is initiated by

activation-induced cytidine deaminase (AID), which induces class switch recombination (CSR)

by deaminating cytidines within immunoglobulin switch regions (Di Noia et al. 2007; Stavnezer

et al. 2008). The recognition and subsequent processing of the mutated residues by the base

excision repair and/or mismatch repair machineries generates double-stranded DNA breaks

Page 68: Characterizing the role of the DNA Damage Response in

54

(DSBs) at switch regions (Saribasak et al. 2009). In an attempt to mend the ensuing breaks, B

cells mount a damage response similar to that signalled by irradiated cells (Stavnezer et al.

2008). Ultimately, AID-induced double-strand DNA breaks are repaired predominantly by non-

homologous end joining (Pan-Hammarstrom et al. 2005; Franco et al. 2008) and, to a lesser

extent, by an alternative end joining pathway(s) (Yan et al. 2007).

Generally, DSBs are readily recognized by sensor-kinase protein complexes. These include

the Mre11-Rad50-Nbs1 (MRN)/ATM, the Ku70/Ku80-DNAPKcs, and the ATR-ATRIP

complexes (Wang et al. 2004). Upon binding to a DSB site, these factors are activated to trigger

a cascade of events culminating in cell cycle delay and/or DNA repair (Wang et al. 2004).

During this process, sequential chromatin modifications around the break sites appear crucial for

the adequate resolution of the DNA breaks. Active chromatin modification is initiated by ATM-

dependent phosphorylation of the histone variant H2AX, to the γH2AX form (Rogakou et al.

1998; Burma et al. 2001; Celeste et al. 2002). γH2AX then preferentially binds the tandem-

BRCT motifs of Mediator of DNA Damage Checkpoint 1 (MDC1) (Stucki et al. 2005; Lou et al.

2006), which in turn recruits more MRN/ATM, thereby amplifying the γ-H2AX signal (Stucki et

al. 2005; Lou et al. 2006). Recently, it was shown that ATM also phosphorylates MDC1 at a

TQxF consensus, allowing for the recruitment of a novel E3 ubiquitin ligase, called RNF8, via its

phospho-binding FHA module (Huen et al. 2007; Kolas et al. 2007; Mailand et al. 2007). Once

at the break, RNF8 mediates the mono-ubiquitination of H2A-type histones, among other

potential proteins. Mono-ubiquitinated H2A-type histones subsequently signal the recruitment of

another E3 ubiquitin ligase, RNF168 (Doil et al. 2009; Stewart et al. 2009), which was recently

found to be mutated in an immunodeficient and radiosensitive Riddle Syndrome patient (Stewart

et al. 2007; Stewart et al. 2009). RNF168 protects and amplifies the RNF8-mediated

Page 69: Characterizing the role of the DNA Damage Response in

55

ubiquitination by catalyzing the addition of K63-linked poly-ubiquitin chains at DNA damage

sites (Doil et al. 2009; Stewart et al. 2009). Whilst low levels of 53BP1 can transiently localize

to DNA breaks via direct binding to di-methylated H4K20 (Botuyan et al. 2006), ubiquitination

by RNF8 and RNF168 seem crucial for the accumulation and stabilization of 53BP1 at break

sites.

Most of the factors required for DNA Damage Response (DDR) also appear to contribute

to the process of CSR (Stavnezer et al. 2008). However, one important distinction between

general DDR and CSR is the major reliance of the latter on 53BP1 (Ward et al. 2003; Manis et

al. 2004; Ward et al. 2004), since 53BP1-deficiency results in a drastic decrease in CSR. It is

still unclear whether this marked defect is due to a repair function that 53BP1 fulfills and/or

because it is required to bring distal switch regions into close proximity and assist in switch

synapse formation (Stavnezer et al. 2008). Recent data seem to focus on the latter aspect,

especially since 53BP1-deficient B cells show increased levels of intra-switch region

recombination (Reina-San-Martin et al. 2007), which is consistent with 53BP1 promoting and/or

stabilizing the synapsis of distant switch regions during CSR. Similarly, 53BP1 facilitates long-

range end joining during V(D)J recombination at the TCR locus (Difilippantonio et al. 2008),

and also promotes end joining of dysfunctional telomeres by increasing chromatin mobility

(Dimitrova et al. 2008). Historically, 53BP1 was first described as one of two proteins to interact

with p53 (Iwabuchi et al. 1994). This aspect of 53BP1 function is generally thought to mediate

apoptosis and is seen as distinct from its ability to localize to DSB.

We and others have previously shown that the ubiquitination of PCNA is important for

both somatic hypermutation and CSR (Langerak et al. 2007; Roa et al. 2008). In a continued

effort to delineate the importance of the ubiquitin system in antibody diversification, we examine

Page 70: Characterizing the role of the DNA Damage Response in

56

here whether the RNF8 and RNF168 members of the E3 ubiquitin ligase cascade, that are

upstream of 53BP1 in the DDR pathway, are also required to mediate CSR.

2.3 MATERIALS AND METHODS

2.3.1 In Vitro Cell Culture

CH12F3-2 cells were maintained and CSR assays were performed as previously described

(Nakamura et al. 1996). Briefly, cells were stimulated with 1 ng/mL recombinant human TGFβ1

(R&D systems), 10 ng/mL recombinant mouse IL-4 (R&D systems) and 2 µg/mL functional

grade purified anti-mouse CD40 (eBiosciences), and analyzed by flow cytometry, as described

below. Lentiviral shRNA constructs and protocols for 53BP1 (V2LMM 83391), RNF8

(AAF59E8 and AAF59E12), RNF168 (V2MM 16141), and non-silencing negative control

(RHS4346 and SHC002) shRNA were purchased from Open Biosystems. CH12F3-2 cells were

transduced with lentivirus by centrifugation at 800xg for 90 min at room temperature in the

presence of 5 μg/mL polybrene. Cells were then incubated for 3 days and positively transduced

clones were subsequently obtained by limiting dilution cloning and puromycin selection. For

growth curve analysis, CH12F3-2 cells were diluted to a concentration of 1x105 cells/ml and

aliquoted in duplicate on a 96-well plate. At various time points, the numbers of live-trypan blue

excluded cells were counted using a haemocytometer. NIH3T3 cells were treated with an siRNA

SMARTpool (from ThermoFisher) targeting murine RNF168. siRNA transfections were

performed using Dharmafect 1 (ThermoFisher) in forward transfection mode. CH12F3-2 cells

were treated with 25 μM Nutlin-3 (Sigma) or DMSO as a negative control.

Page 71: Characterizing the role of the DNA Damage Response in

57

2.3.2 Flow Cytometric analyses

CH12F3-2 cells were analysed by intracellular staining with PE conjugated anti-mouse IgA

clone 11-44-2 (eBiosciences), requiring the use of Cytofix/Cytoperm and Perm/wash buffers

(BD Biosciences). CH12F3-2 cells were surface stained for Annexin V using Annexin V-APC

Apoptosis Detection Kit (eBiosciences). Stained cells were analyzed by FACSCalibur (BD Bio)

and FlowJo software (Truestar Inc.).

2.3.3 Irradiation sensitivity assays

CH12F3-2 cell lines were irradiated with various doses of X-rays and subsequently plated at

various dilutions in duplicate 96 well plates. Survival was determined by counting the number of

expanding clones normalized to plating efficiency.

2.3.4 Western blot analyses

53BP1 (Alexis Biochemicals and Novus), RNF8 (Abcam), H2AX (Upstate), PCNA (Santa Cruz

Biotechnology), β-Actin (Abcam), Msh2 (BD Pharmingen) antibodies were used as specified by

the manufacturers protocols. The RNF168 polyclonal antibody was raised against a murine GST-

RNF168381-567 fusion protein and affinity purified using a murine HIS6-RNF168381-567 Sepharose

column.

Page 72: Characterizing the role of the DNA Damage Response in

58

2.3.5 Immunoflourescence

Cells were pelleted at 1000 rpm for 3 min in Eppendorf tubes. Cells were then fixed with 4%

paraformaldehyde (Sigma), permeabilized with 0.3% Triton X (Roche), and blocked with 10%

Fetal Calf Serum (Hyclone) and 0.01% saponin (Sigma) in PBS. The following antibodies were

used: mouse FITC-anti-γH2AX (Millipore), rabbit anti-53BP1 (Novus), and anti-rabbit Alexa

Fluor 568 (Invitrogen). Stains were done overnight at 4C, and washes were done with blocking

buffer. Cells were pelleted on slides using a Shandon cytospin machine at 400 rpm for 2 min.

DAPI stain in mounting dye (Vectashield) was used to detect DNA.

2.3.6 Statistical Analysis

Analyses were performed on GraphPad Prism. For Student’s t tests, two-way analysis of

variance (ANOVA), and Mann-Whitney tests, p values of 0.05 or less were considered

significant.

2.4 RESULTS

2.4.1 shRNA-mediated RNAi can establish functional knockdown in CH12F3-2 cells.

We first ascertained that the CH12F3-2 cell line is a faithful model of CSR, with respect to the

role of DDR factors. To do this, we assessed whether RNA interference (RNAi) by shRNA for

53BP1, which has been shown to be an essential CSR factor (Manis et al. 2004; Ward et al.

2004), can achieve functional knockdowns in CH12F3-2 cells. Cells were transduced with a

lentiviral shRNA construct specific for 53BP1 (sh53BP1), and individual clones were isolated by

puromycin selection. Compared to clones transduced with a scrambled, non-silencing shRNA

Page 73: Characterizing the role of the DNA Damage Response in

59

construct (control), 53BP1 knockdown cells showed a ~5 fold reduction in 53BP1 transcript

levels, as determined by reverse transcriptase-polymerase chain reaction (RT-PCR) analysis, as

well as by western blot analysis (Figure 8A). These cells have a slightly reduced rate of growth

compared to control cells (Figure 8B) and demonstrate no major differences in apoptosis, before

and after stimulation (Figure 8C). Although 53BP1-deficient cells have previously been shown

to have a mild sensitivity to irradiation (Nakamura et al. 2006), 53BP1-depleted CH12F3-2 cells

did not display an increased sensitivity to irradiation in comparison to control cells (Figure 8D).

We next tested whether 53BP1 knockdown cells were deficient in CSR. While CSR was

not affected in control clones relative to the parental cells (Figure 8E), eight individual 53BP1

knockdown clones showed a complete abrogation of CSR at 48 hours after stimulation (Figure

8E), even though both groups of cells expressed similar levels of AID (Figure 9). This defect in

CSR persisted even at 4 days post-stimulation (Figure 8F). Although the 53BP1-depleted cells

showed a mild proliferation defect, previous findings have established a role for 53BP1 in CSR

(Manis et al. 2004; Ward et al. 2004), suggesting that the mild proliferative defect observed in

these cells is not responsible for the dramatic lack of CSR. These results confirm the importance

of 53BP1 for CSR in the CH12F3-2 cells used in these studies. Furthermore, we show that

shRNA-mediated RNAi can achieve a functional knockdown in CH12F3-2 cells, thereby setting

the stage for testing whether RNF8 and RNF168 are involved in CSR.

Page 74: Characterizing the role of the DNA Damage Response in

60

Figure 8

Figure 8. shRNA-mediated RNAi knockdown of 53BP1 in CH12F3-2 cells abrogates CSR. (A) Western blot analysis of control and sh53BP1 CH12F3-2 cells quantifying 53BP1 and Msh2

(loading control). (B) Two clones of each indicated CH12F3-2 cell population were analyzed for

growth in duplicate and statistical significance was tested by 2-way ANOVA test. **** =

P<0.0001 (C) Percentage of CH12F3-2 populations undergoing apoptosis (surface Annexin V+)

with and without stimulation for 2 days. Assay was carried out with at least 2 individual clones

for each population. (D) Clonogenic survival assays with asynchronous CH12F3-2 populations

exposed to varying doses of X-rays. Assay was carried out with at least 2 individual clones for

each population. (E) Six individual control and eight sh53BP1 CH12F3-2 clones were stimulated

for 2 days and analyzed for IgA expression. Values were normalized by dividing the % IgA-

positive cells in the experimental group to the % IgA-positive cells in the stimulated parental

CH12F3-2 cells. CSR assays were performed on each clone in duplicate and statistical

significance was tested by one-tailed t-test. (F) Three control and three sh53BP1 CH12F3-2

clones were stimulated for 2, 3 and 4 days and IgA expression was analyzed. CSR assays were

performed in duplicate for each clone. Statistical significance tested by 2-way ANOVA. **** =

P<0.0001.

Page 75: Characterizing the role of the DNA Damage Response in

61

Figure 9

Figure 9. AID expression is not affected by RNAi of CH12F3-2 cells. Control, shRNF168,

and sh53BP1 CH12F3-2 cells were stimulated (+) and analyzed for AID expression by western

blot analysis, with β-actin as the loading control. Unstimulated (-) and stimulated (+) parental

CH12F3-2 cells (CH12F3) represent the negative and positive controls, respectively. Molecular

weigh markers have been indicated on the right of the figure.

29 kDa

22 kDa

42 kDa

Page 76: Characterizing the role of the DNA Damage Response in

62

2.4.2 RNF8 is required for CSR in CH12F3-2 cells.

To achieve knockdown of RNF8, we tested five different shRNA lentiviral constructs

specific for RNF8. Two of those constructs (shRNA #1 and 2) achieved strong depletion of

RNF8 protein following puromycin selection in CH12F3-2 cells, as verified by Western blot

analysis (Figure 10A). We then assessed whether RNF8 depletion in those lines affected the

response to DNA double-strand breaks. As expected, activated cells had a considerably higher

γH2AX signal in comparison to non-activated samples (Figure 10B). However, there was no

significant difference in γH2AX protein levels in control and RNF8-depleted CH12F3-2 cells,

indicating that both groups of cells responded similarly to AID-induced DNA double-strand

breaks.

To test the effects of RNF8 depletion on CSR, we selected eight individual clones from

each transduction for subsequent analysis. Although different switching efficiencies were

achieved within each group of cells, there was significantly reduced CSR in RNF8-depleted

clones (Figure 10C). The extent of the switching defect was greater in the clones (shRNA #2)

that were more depleted for RNF8 (Figure 10A and C). In addition, we followed CSR at different

time intervals in a representative sample of the clones, and found that CSR in RNF8-depleted

cells remained low relative to control cells (Figure 10D). RNF8-depleted cells did not show a

defect in proliferation (Figure 8B) or apoptosis, before or after stimulation (Figure 8C). In

addition, RNF8-depleted cells and control cells were equally sensitive to irradiation (Figure 8D).

These data indicate that the reduced CSR observed in RNF8-depleted cells is not due to

proliferation defects, and that RNF8 function is required for efficient CSR.

To gain insight into the role of RNF8 in CSR, we studied the effects of RNF8 knockdown

by immunofluorescence. When unstimulated CH12F3-2 cells were fixed and stained with anti-

Page 77: Characterizing the role of the DNA Damage Response in

63

Figure 10

Figure 10. CSR is impaired in RNF8-deficient CH12F3-2 cells. (A) Western blot analysis

quantifying RNF8, 53BP1, and PCNA (loading control) in CH12F3-2 cells with control shRNA,

and two different lentiviral vectors targeting RNF8 (1) and (2) (B) Western blot analysis of

stimulated and unstimulated control or shRNF8 treated CH12F3-2 cells, quantifying RNF8 and

γH2AX protein levels with PCNA as the loading control. (C) Eight individual clones for each of

control, shRNF8 (1), shRNF8 (2) CH12F3-2 cells were stimulated for 3 days and analyzed for IgA

expression. Values were normalized by dividing the % IgA-positive cells in the experimental group

to the % IgA-positive cells in the stimulated parental CH12F3-2 cells. Statistical significance was

tested by two-tailed t-test. * = P<0.05 and **** = P<0.0001. (D) Two control, two shRNF8 (1), and

two shRNF8 (2) treated CH12F3-2 clones were stimulated and IgA expression was analyzed at 2, 3

and 4 days post-stimulation. Statistical significance tested by two-way ANOVA. *** = P<0.001 and

**** = P<0.0001. (E) Stimulated control and shRNF8 CH12F3-2 cells were collected at time 0 h

and 48 h, and stained with the denoted antibodies and DAPI. Cells, from two separate experiments

(each with n=1000), having single H2AX foci with 53BP1 (H2AX + 53BP1 foci) or without

53BP1 (H2AX + 53BP1 foci) co-localization were counted and analyzed as shown. Greater than

90% of the cells did not exhibit H2AX foci after cytokine stimulation. Statistical significance was

tested by two-tailed t-test between control 48hr and shRNF8(2) 48hr. (* = P=0.014).

Page 78: Characterizing the role of the DNA Damage Response in

64

H2AX and anti-53BP1, almost all cells expectedly manifested a diffuse 53BP1 nuclear pattern

and a lack of H2AX signal (Figure 11 - top panel). Upon activation, control cells showed a

marked increase in single H2AX foci that co-localise with single 53BP1 foci, indicative of

induced DSBs, and hence active switching events (Figure 10E, Figure 11 bottom panel).

Importantly, compared to control cells, shRNF8-transduced cells had a significant increase in

H2AX foci with a diffuse 53BP1 signal (Figure 10E, Figure 11). 53BP1 depletion has no effect

on germline transcription ((Manis et al. 2004) or AID expression (Figure 9). In addition, we

could show that shRNF8 cells generated similar levels of AID-induced H2AX, as assayed by

immunoblotting and immunofluorescence, as control cells. This suggests that the cascade

upstream of AID-induced breaks, which is dependent on germline transcription and AID, is most

likely intact. Taken together, this suggests that the inability of shRNF8-treated cells to undergo

efficient CSR is due to their inability to signal the accumulation and/or stabilization of 53BP1 at

AID-induced breaks.

2.4.3 RNF168 is important for CSR in CH12F3-2 cells.

We next examined the contribution of RNF168 to CSR. To achieve knockdown of

RNF168, we tested three different shRNA lentiviral constructs specific for RNF168, one of

which led to a ~2-3 fold reduction in RNF168 protein levels (Figure 12A). Akin to the shRNF8

transduced cells, shRNF168-transduced cells and control cells displayed similar growth rates

(Figure 8B), apoptosis levels before and after stimulation (Figure 8C), and sensitivity to

irradiation (Figure 8D). On the other hand, we observed that RNF168-knockdown cells had 46%

lower CSR levels compared to control cells (Figure 12B and C). Like the RNF8-depleted cells,

this ~2-fold reduction of CSR correlates with the level of knockdown achieved in the CH12F3-2

Page 79: Characterizing the role of the DNA Damage Response in

65

Figure 11

Figure 11. 53BP1 foci formation is reduced in RNF8-knockdown clones. Control and

shRNF8 CH12F3-2 cells were fixed and stained with the denoted antibodies and DAPI, and

visualized by confocal microscopy. Stimulated control and shRNF8 CH12F3-2 cells were

collected at time 0 h and 48 h. A representative sample of cells is shown in both panels. Apparent

DAPI foci are most likely due to the high levels of heterochromatin in B cells.

Page 80: Characterizing the role of the DNA Damage Response in

66

Figure 12

Figure 12. CSR is impaired in RNF168-deficient CH12F3-2 cells. (A) Western blot analysis

of control or shRNF168 CH12F3-2 cells quantifying RNF168, and β-actin as the loading control.

NIH3T3 cells treated with control siRNA and siRNF168 were used as negative and positive

controls for RNF168 protein expression, respectively. (B) Six individual control and seven

individual shRNF168 CH12F3-2 clones were stimulated for 2 days and analyzed for IgA

expression. Values were normalized by dividing the % IgA-positive cells in the experimental

group to the % IgA-positive cells in the stimulated parental CH12F3-2 cells. CSR assays were

performed on each clone in triplicate and statistical significance was tested by two-tailed t-test.

** = P=0.0023 (C) Three control and three shRNF168 CH12F3-2 clones were stimulated for 2, 3

and 4 days and IgA expression was analyzed. CSR assays were performed in duplicate for each

clone. Statistical significance tested by two-way ANOVA. (** = P=0.0034).

Page 81: Characterizing the role of the DNA Damage Response in

67

cells. Thus, the CSR defect observed in the absence of RNF8 can be recapitulated by depleting

RNF168. Taken together, these findings indicate that the E3 ubiquitin ligase cascade composed

of RNF8 and RNF168 plays a significant role in CSR.

2.4.4 53BP1 deficiency, but not RNF8 or RNF168-deficiency, protects cells from Nutlin-3-

induced apoptosis.

Nutlin-3 is a small molecule inhibitor of MDM2 that induces activation of p53 and

apoptosis of cancer cells in a process that is dependent on 53BP1 (Brummelkamp et al. 2006).

That is, 53BP1-deficient cells are resistant to Nutlin-3-mediated apoptosis. To test whether this

activity of 53BP1 is influenced by RNF8 or RNF168, we examined whether deficiency in any of

these proteins protects CH12F3-2 cells from the effects of Nutlin-3. In contrast to control cells

incubated with Nutlin-3, which exhibited dramatically increased levels of apoptosis, Nutlin-3 did

not induce high levels of apoptosis in 53BP1-depleted cells (Figure 13A). Strikingly, RNF8-

depleted cells and RNF168-depleted cells remained sensitive to Nutlin-3-mediated apoptosis

(Figure 13A). These data indicate that 53BP1 has functions that are independent of RNF8 and

RNF168 (Figure 4B).

2.5 DISCUSSION

2.5.1 Role of RNF8 and RNF168 in CSR and its relation to 53BP1 localization.

Although the necessity of 53BP1 for CSR has been shown before (Wang et al. 2002;

Manis et al. 2004; Ward et al. 2004), the exact function of the 53BP1-dependent pathway in

CSR remains elusive. Recent findings demonstrate that RNF8 and RNF168 play essential roles

Page 82: Characterizing the role of the DNA Damage Response in

68

Figure 13

Figure 13. 53BP1 knockdown cells, but not RNF8/168 knockdown cells, are resistant to

Nutlin-3-mediated apoptosis. (A) Percentage of control, sh53BP1, shRNF8 (2), and shRNF168

CH12F3-2 cells undergoing apoptosis (i.e. surface Annexin V+) after treatment with Nutlin-3 or

DMSO for 1 day. Each assay was carried out with 2 different clones of each population.

Statistical significance tested by 2-way ANOVA. **** = P<0.0001. (B) Schematic illustrating

two different roles for 53BP1. AID indirectly induces a DSB by the actions of the mismatch

repair (MMR) and base excision repair (BER) pathways. 53BP1 is recruited to the double-

stranded DNA break either by directly binding di-methylated H4K20, or through the γH2AX-

RNF8-RNF168 amplification pathway, the latter of which leads to the accumulation and

stabilization of 53BP1 at the break site. These pathways ultimately lead to end processing and

resolution of the DNA break. 53BP1 stabilizes p53 (independent of RNF8/168) that has been

dissociated from the inhibitor MDM2 to induce apoptosis.

Page 83: Characterizing the role of the DNA Damage Response in

69

in recruiting and stabilizing 53BP1 at sites of DSBs during the irradiation-induced DDR (Huen et

al. 2007; Kolas et al. 2007; Mailand et al. 2007; Doil et al. 2009; Stewart et al. 2009). In support

of a role for 53BP1 localization to sites of AID-induced DSBs, our data show that knockdown of

either RNF8 or RNF168 causes a significant decrease in CSR. These findings support the notion

RNF8/RNF168 and is required for the efficient bridging of the two distal DNA ends

(Difilippantonio et al. 2008; Dimitrova et al. 2008; Stavnezer et al. 2008).

Since the recruitment of RNF8 and subsequent recruitment of RNF168 to DSB sites

requires ATM/MRN, H2AX and MDC1, it would be reasonable to assume that all these factors

have similar effects on CSR (Figure 4B). Yet, in reality the loss of these DDR factors seem to

have different effects on CSR. While ATM-, MRN-, and H2AX-deficiency leads to a ~70%

reduction of CSR (Reina-San-Martin et al. 2003; Lahdesmaki et al. 2004; Reina-San-Martin et

al. 2004; Reina-San-Martin et al. 2005; Dinkelmann et al. 2009), MDC1-deficiency has a milder

effect (25%-50% reduction (Lou et al. 2006)), and 53BP1-deficiency leads to the most profound

effect (~90% CSR reduction; this study and (Manis et al. 2004; Ward et al. 2004)). A lesser

dependence of CSR on MDC1 may be either due to: 1) the residual levels of MDC1 protein in

the MDC1 gene-trap knock-out mice that was sufficient to amplify the ATM/MRN-mediated

H2AX signal (Lou et al. 2006); or 2) a redundant MDC1-independent mechanism that allows

for recruitment of low levels of RNF8, and subsequently 53BP1, allowing for intermediate levels

of CSR. On the other hand, the difference between the low levels of CSR observed in ATM-,

MRN-, and γH2AX-deficiency and the near complete lack of CSR in 53BP1-deficient cells

might be explained by the potential transient recruitment of 53BP1 to switch regions via direct

binding to either di-methylated H4K20 that are exposed at the DSB site (Botuyan et al. 2006) or

through other yet to be determined factors. As a result, this transient recruitment of 53BP1 could

Page 84: Characterizing the role of the DNA Damage Response in

70

be sufficient for the resolution of a subset of DNA breaks mediating low levels of CSR in the

absence of the upstream proteins required for 53BP1 accumulation and stabilization (Figure 4B).

It remains to be seen whether complete knock-out of RNF8 or RNF168 affect CSR in a manner

akin to the loss of 53BP1 or H2AX. Our data suggest that RNF8/RNF168 have an incomplete

effect on CSR. However our shRNA constructs cannot completely deplete these factors. The

finding that serum IgG is reduced by ~10 fold in a Riddle syndrome patient deficient in RNF168

(Stewart et al. 2007) suggests that complete ablation of RNF8 or RNF168 would produce a

pronounced phenotype.

2.5.2 53BP1 has a unique function that is independent of RNF8 and RNF168.

It was previously shown that Nutlin-3, which inhibits the p53-Mdm2 interaction, induces

p53-mediated apoptosis in a 53BP1-dependent manner (Brummelkamp et al. 2006). Consistent

with this finding, our results show that 53BP1-deficient CH12F3-2 cells are resistant to Nutlin-3-

induced apoptosis. However, we found that RNF8- and RNF168-deficient CH12F3-2 cells were

still sensitive to the apoptotic effects of Nutlin-3. 53BP1 was initially discovered as a p53-

binding protein (Iwabuchi et al. 1994), interacting with p53 through its C-terminal tandem-

BRCT motifs in a phosphorylation-independent manner. This interaction is thought to stabilize

p53 and stimulate p53-mediated transcriptional activation (Iwabuchi et al. 1998). It is possible

that p53, upon release from MDM2-mediated inhibition, is degraded in 53BP1-deficient cells

and as such, apoptosis is not induced, or that 53BP1 functions as a co-factor of p53 to induce

apoptosis. Whether this function of 53BP1 is related to its role in localization at sites of DSB

remains unclear. However, since the p53-53BP1 interaction is phosphorylation-independent and

does not seem to require ATM activation (Iwabuchi et al. 1998), it is plausible that this might be

Page 85: Characterizing the role of the DNA Damage Response in

71

a function of 53BP1 distinct from foci formation at DSBs. The finding that Nutlin-3-mediated

apoptosis is not affected in RNF8- or RNF168-deficient CH12F3-2 cells supports this notion and

suggests that 53BP1 has distinct roles in response to genotoxic stress (Figure 13B).

Page 86: Characterizing the role of the DNA Damage Response in

72

3 IDENTIFYING NOVEL DDR FACTORS FROM A GENOME-WIDE

RNA INTERFERENCE SCREEN FOR CSR.

Contributions:

Cell culture, cell sorting and genomic DNA extractions: Shaliny Ramachandran, Rajeev Nepal

and Dionne White.

Microarray and sequencing: Troy Ketela, Bohdana Fedyshyn, Dahlia Kasimer and Kaajal Nagar

Microarray and Sequencing analysis: Shaliny Ramachandran and Troy Ketela

Screen Validation analysis: Shaliny Ramachandran, Maribel Berru, Anat Kapelnikov, Alan

Jiao, Michael Le

AID expression analysis: Shaliny Ramachandran and Angela Zhang

Page 87: Characterizing the role of the DNA Damage Response in

73

3.1 ABSTRACT

B cells undergo CSR to generate different classes of antibody, which direct diverse

functions, resulting in a humoral immune response that can be fine-tuned to the infection

at hand. CSR requires AID-generated DSBs at the antibody locus, and subsequent repair

of this lesion by DDR. Using the CH12F3-2 B cell line, we demonstrated that shRNA-

mediated knockdown of the DDR factors, RNF8, RNF168 and 53BP1, led to a significant

decrease in CSR (Chapter 2). However, the molecular events following 53BP1 and

leading up to DNA repair in CSR remain elusive. To identify novel DDR factors in CSR,

we conducted a pooled, loss-of-function, genome-wide shRNA screen for CSR factors,

using the CH12F3-2/RNA interference system established above in Chapter 2. Using

microarray and sequencing-based readouts, hairpins that inhibit CSR were identified with

validation frequencies of 63.2% and 61.5%, respectively. The CSR candidates were then

screened for AID expression to narrow down the CSR defect relative to AID.

Surprisingly, nearly 80% of the 33 tested hairpins showed a significant effect on AID

expression. In some instances, the effect on AID expression might be attributed to off-

target effects that have an indirect effect on AID protein levels. Given that CSR is tightly

linked to AID expression, future RNA interference studies on CSR in this system must

ensure that any effects on CSR is specifically due to the gene of interest and not due to

off-target effects on AID protein levels.

3.2 INTRODUCTION

B cells undergo CSR to shift from expressing the IgM antibody to IgG, IgE and IgA

classes. Each class of antibody is attributed with a different effector function, allowing for a fine-

Page 88: Characterizing the role of the DNA Damage Response in

74

tuned response that depends on the nature of the infection. During CSR, AID is absolutely

required to catalyze the deamination of deoxycytidines within the switch regions (Kracker et al.

2011). This deamination forms dU:dG mismatches, which are subsequently engaged by the

MMR and BER pathways leading to the production of nicks and/or stretches of excised DNA

strands within the switch regions (Stavnezer et al. 2008). Owing to the highly repetitive nature of

the switch regions, BER- and MMR-generated nicks or strand excision tracts formed within

close proximity on opposite strands can become staggered DSBs (Schrader et al. 2005; Stavnezer

et al. 2005). AID activity produces staggered DSBs at both donor and acceptor switch regions,

which can then be repaired in a manner that places the downstream acceptor region immediately

adjacent to the rearranged V(D)J region, while deleting the intervening DNA containing the

donor switch region (Chaudhuri et al. 2004). This replaces the donor constant region with the

acceptor constant region on the antibody, completing CSR.

CSR requires an intricate network of factors that function in a highly regulated manner that

lead to expression of AID and its requisite co-factors, switch region transcription, and factors

involved in the DNA repair pathways required to process the AID-induced mutations to generate

the staggered DSBs, and to repair them (Xu et al. 2007; Stavnezer et al. 2008; Pavri et al. 2011;

Stavnezer 2011). AID-induced DSBs activate DDR, leading to cell cycle arrest and repair

primarily via NHEJ (Chaudhuri et al. 2007). While much of the CSR process is known, the exact

molecular events within the DDR and the repair aspects of CSR have not been completely

characterized. In this report, we carried out a genome-wide CSR screen designed to identify

genes involved in the all the steps leading up to CSR. We used the mouse B cell line CH12F3-2

that undergoes CSR specifically from IgM to IgA at high levels upon stimulation (Nakamura et

al. 1996). We established a CH12F3-2/RNA interference system to enrich for factors involved in

Page 89: Characterizing the role of the DNA Damage Response in

75

the CSR process. Using this RNA interference approach, we previously demonstrated that RNF8,

RNF168 and 53BP1 play a role in CSR, in Chapter 2 above (Ramachandran et al. 2010). We

conducted a pooled, loss-of-function screen testing approximately 80,000 lentiviral shRNA

hairpins in total and analyzed the results using microarray-based and next-generation sequencing

readout methods, preformed validation analyses to determine the confirmation rates, and finally

screened a subset of CSR candidates for AID expression to identify possible DDR and DNA

repair factors.

3.3 MATERIALS AND METHODS

3.3.1 Pooled shRNA lentiviral library

Lentiviral library targeting the mouse genome was generated by The RNAi Consortium.

Containing a total of ~78,000 shRNA haipirns targeting the mouse genome, the virus used for

transduction was a pool of 10 sub-pools, each containing ~8000 shRNA hairpins. Lentivirus sub-

pools were generated as previously described (Ketela et al. 2011).

3.3.2 In Vitro Cell Culture

CH12F3-2 cells were maintained as previously described in Chapter 2 above (Ramachandran et

al. 2010). Lentiviral shRNA constructs and protocols were performed as previously described in

Chapter 2 above (Ramachandran et al. 2010). Briefly, for lentiviral transductions, CH12F3-2

cells were incubated with virus at 37ºC for 24 hours at an MOI of 0.2-0.3. Positively transduced

cells were selected for puromycin resistance, with 1 μg/mL puormycin for 3 days, followed by

CSR stimulation. Cells were stimulated with 1 ng/mL TGFβ, 10 ng/mL IL-4 and 2 μg/mL anti-

CD40 for 3 days.

Page 90: Characterizing the role of the DNA Damage Response in

76

3.3.3 Flow cytometry cell sorting

Stimulated CH12F3-2 cells were subjected to extracellular staining with PE conjugated anti-

mouse IgA clone (Southern Biotech). Stained cells were sorted by FACS Aria (BD Bio).

3.3.4 Microarray analysis

Genomic DNA was extracted from each of the sorted population using the Blood Maxi prep kit

(Qiagen). shRNA hairpin barcodes were PCR-amplified and used as probe in mircroarray

experiments as described previously (Ketela et al. 2011).

3.3.5 Sequencing analysis

Genomic DNA was extracted and subjected to PCR-amplification as described above. PCR

product was then subjected to a second PCR reaction in order to be tagged with an Illumina

adapter sequence, as previously described (Ketela et al. 2011). Illumina sequencing and analysis

was performed as previously described (Ketela et al. 2011).

3.3.6 Western blot analysis

AID (Cell Signalling) and β-Actin (Sigma) antibodies were used as specified by the

manufacturers’ protocols.

3.3.7 Statistical Analysis

Analyses were performed on GraphPad Prism. For Student’s t-tests and linear regression

analysis, p values of 0.05 or less were considered significant.

Page 91: Characterizing the role of the DNA Damage Response in

77

3.4 RESULTS

3.4.1 CH12F3-2/shRNA: A system to screen for potential CSR factors.

The CH12F3-2 mouse B cell line undergoes CSR from IgM specifically to IgA at very high

levels upon stimulation with anti-CD40, IL4 and TGFβ (CIT) (Figure 14). After 72 hours of

stimulation, the system reaches saturation with over 60% of cells typically expressing IgA.

CH12F3-2 is a subclone variant isolated from the IgM-expressing CH12.LX mouse B line, a

transformed cell line originating from a lymphoma that developed in a B10-2a4b mouse

immunized with sheep red blood cells (Kunimoto et al. 1988; Nakamura et al. 1996). We

previously established a protocol to individually test for potential CSR factors using an RNA

interference approach in CH12F3-2 cells in Chapter 2 above (Ramachandran et al. 2010). Using

shRNA, we achieved knock-down of target proteins in CH12F3-2, and ascertained whether the

knock-down affected CSR, thus determining whether the target protein is important for CSR.

Taking advantage of this system, we developed a screen to identify novel CSR factors. For this

screen, we utilized the pLKO.1 vector-based library of shRNA targeting the mouse genome

(from The RNAi Consortium), with a total of 77,690 hairpins targeting approximately 16,000

mouse genes, with 4-5 different hairpins targeting each gene (Moffat et al. 2006). These shRNA

have been designed to specifically target the gene of interest, and have at least 3 nucleotide

mismatches to other related genes so as to prevent off-target effects (Moffat et al. 2006).

3.4.2 Screen methodology

We designed a pooled, loss-of function shRNA screen (Figure 15). CH12F3-2 cells were

transduced with the library of 77,690 shRNA at a low MOI of 0.2 - 0.3 to minimize the

Page 92: Characterizing the role of the DNA Damage Response in

78

Figure 14

Figure 14. CH12F3-2 in vitro class switch recombination assay. CH12F3-2 cells were

stimulated at a concentration of 100,000 cells/mL with 1ng/µL TGFβ, 10 ng/µL IL4 and 2

µg/mL α-CD40 antibody, and incubated at 37 °C, 6% CO2. At 2, 3 and 4 days post-stimulation,

cells were stained analyzed for IgA expression by flow cytometry. Unstimulated cells were

assayed in parallel as non-switched negative controls.

Page 93: Characterizing the role of the DNA Damage Response in

79

Figure 15

Figure 15. CSR Screen methodology. CH12F3-2 cells were transduced with the lentiviral

shRNA library at a MOI ~ 0.2- 0.3. Positively transduced cells were selected for puromycin

resistance and stimulated to undergo CSR (in 3 biological replicates). Switched IgA+ cells were

sorted from non-switched IgA- cells. Genomic DNA was extracted from each population for

PCR-mediated amplification of the shRNA barcode, which was then subjected to microarray

analysis and sequencing analysis, to determine the level of each hairpin in each population. For

each hairpin, CSR screen score is determined as fold difference in hairpin representation in IgA-

population over IgA+ population.

Page 94: Characterizing the role of the DNA Damage Response in

80

probability of integrating more than one shRNA hairpin into any one individual cell.

Transductions were scaled-up so that each hairpin could have at least 1000 independent genomic

representations. Positively transduced cells were selected for puromycin resistance, and

subsequently stimulated with the cocktail of TGFβ, IL4 and anti-CD40 in three biological

replicates. After 3 days of stimulation, switched IgA+ cells were sorted from non-switched IgA-

cells. Genomic DNA was extracted from each population for PCR-mediated amplification of the

shRNA sequence, which acts as the unique barcode used to identify the hairpin and the target

gene. The PCR product was subjected to microarray-based Gene Modulation Array platform

analysis and next-generation sequencing analysis, to determine the representation of each

individual hairpin in eachof the sorted IgA- and IgA+ populations. Gene Modulation Array

platform analysis was performed in 2 technical replicates, while next-generation sequencing

analysis was performed in 1 technical replicate. For each hairpin, a CSR screen score was

determined as fold difference in hairpin representation in IgA- population over the IgA+

population. Hairpins that are equally-represented in both IgA- and IgA+ populations likely target

neutral genes that have no effect on CSR. Hairpins that are under-represented in both IgA- and

IgA+ populations, as observed relative to the entire cohort of hairpins analyzed, likely target

genes involved in cell survival and proliferation. Hairpins under-represented in the non-switched

IgA- population as compared to switched IgA+ population may target genes that inhibit CSR.

Hairpins over-represented in the non-switched IgA- population as compared to switched IgA+

population are considered to target potential CSR factors.

Page 95: Characterizing the role of the DNA Damage Response in

81

3.4.3 CSR Screen results and analysis.

A total of 77,690 shRNA hairpins were screened for the ability to affect CSR. The raw

results are displayed in Figure 16. ShRNA hairpins with a high coefficient of variation between

technical replicates (CV>0.3 for microarray, no technical replicates for sequencing); a low signal

(Signal<10 for microarray, Signal<1 for sequencing); and a high coefficient of variation between

biological replicates (CV>0.4 for microarray, CV>0.6 for sequencing) were eliminated (Table 1).

The analyzed results are displayed in Figure 17. Hairpins with a screen score >= 1.75, that is fold

representation of a given hairpin in IgA- population divided by that in the IgA+ population, were

considered to be candidate CSR factors. 1070 shRNA that target 1035 genes and 3044 shRNA

targeting 2743 genes were identified as hits by the microarray and sequencing analyses,

respectively. A comparison between the 1070 hairpins identified by the microarray analysis and

3044 hairpins identified by the sequencing analysis revealed 271 common hairpins identified as

strong hits from both analyses. This represents 268 potential CSR factors.

3.4.4 Validation analysis

34 hairpins out of the 1070 hairpins identified as a CSR candidate (screen score >= 1.75) in

the microarray analysis and 243 of the 57,057 non-candidate hairpins (screen score <1.75) were

selected for validation analysis. Similarly, 38 of the 3044 candidates in the sequencing analysis

and 239 of the 39,983 non-candidates were selected for validation analysis. Importantly, all of

these non-candidate hairpins target the same genes as the candidates selected for validation, but

these non-candidates likely could not generate a functional knock-down of the targeted gene.

Typically, one of the five hairpins desgined to target a gene appeared as a CSR candidate on the

screen. The ability of each hairpin to affect CSR was individually tested and normalized to a

Page 96: Characterizing the role of the DNA Damage Response in

82

Figure 16

Figure 16. Raw CSR Screen Results. Raw results of the screen are displayed for microarray

(A) and sequencing (B) analysis. Each hairpin is represented by a single point. The

representation of a hairpin in the IgA- population is graphed on the x-axis, while the

representation of the hairpin in the IgA+ population is on the y-axis.

Page 97: Characterizing the role of the DNA Damage Response in

83

Table 1

Microarray analysis Sequencing analysis

Criteria Number of Hairpins

Criteria Number of Hairpins

Total - 77690 - 77690

Technical replicates CV < 0.3 67659 No technical replicates 77690

Background signal IgA- OR IgA+ Signal > 10 67482 IgA- OR IgA+ Signal > 1 58665

Biological Replicates CV < 0.4 58127 CV < 0.6 43027

Hits – Hairpins Screen score =>1.75 1070 Screen score =>1.75 3044

Hits – Genes Screen score =>1.75 1035 Screen score =>1.75 2743

CSR Screen analysis. A total of 77,690 shRNA hairpins were screened for the ability to affect

CSR. ShRNA hairpins with a high coefficient of variation between technical replicates (CV>0.3

for microarray, no technical replicates for sequencing), a low signal (Signal<10 for microarray,

Signal<1 for sequencing), and a high coefficient of variation between biological replicates

(CV>0.4 for microarray, CV>0.6 for sequencing) were eliminated. Hits were set as hairpins with

a screen score => 1.75. Screen score is fold representation of hairpin in IgA- population divided

by IgA+ population.

Page 98: Characterizing the role of the DNA Damage Response in

84

Figure 17

Figure 17. Analyzed CSR Screen Results. Screen results are displayed for microarray (A) and

sequencing (B) analysis. Each hairpin is represented by a single point. The representation of a

hairpin in the IgA- population is graphed on the x-axis, while the representation of the hairpin in

the IgA+ population is on the y-axis.

Page 99: Characterizing the role of the DNA Damage Response in

85

control shRNA hairpin targeting GFP set in parallel. Validation analysis revealed an 8.7% false-

negative frequency, and a 37.8% false-positive frequency, for the microarray analysis; and

13.5% false-negative and 38.5% false-positive frequencies respectively, for the sequencing

analysis (Figure 18). This translates into a confirmation rate of 63.2% and 61.5% for microarray

and sequencing readouts, respectively. Furthermore, an inverse correlation between the observed

CSR, when individually tested in the validation analysis, and the CSR screen score was

observed, with r2 = 0.4103 (P<0.0001) for microarray and r2 = 0.1854 (P=0.002) for sequencing

(Figure 19).

Z score and volcano plot analyses was performed on both the microarray and sequencing

results (Figure 20 and 21). The microarray analysis revealed 2 independent hairpins that target

AID (TRC clone ID: TRCN0000112031, P=0.0026; and TRCN0000112033, P=0.0365) as top

hits of the screen, while sequencing analysis revealed 3 independent hairpins that target AID

(TRC clone ID: TRCN0000112031, P=0.0450; TRCN0000112032, P=0.0184; and

TRCN0000112033, P=0.0598) as top hits of the screen (Table 2). Indeed, hairpins

TRCN0000112031 and TRCN0000112033 have been previously shown to reduce AID protein

levels by 5.8- and 2.4-fold, and CSR levels by 5- and 2.5-fold, respectively (Parsa et al. 2012).

These results further validated the screen and revealed that CSR in CH12F3-2 cells is highly

sensitive to any effects on AID expression.

3.4.5 Secondary screen to identify possible DDR and repair factors

In order to decipher the DDR and repair factors among the CSR candidates identified, a

secondary screen was performed on a subset of genes identified to be interest. Genes that have

been previously associated with DNA damage, genes with multiple targeting hairpins identified

Page 100: Characterizing the role of the DNA Damage Response in

86

Figure 18

Figure 18. Validation analysis of CSR screen. Validations of microarray (A) and sequencing

(B) analyses are displayed. 34 hairpins identified as a hit (score >= 1.75) and 243 non-hit

hairpins (score <1.75) were selected for validation analysis of the microarray results. Similarly,

38 hits and 239 non-hits were selected for validation analysis of the sequencing results. The

ability of each hairpin to affect CSR was individually tested and normalized to a control shRNA

targeting GFP set in parallel. Validation analysis reveals a 8.7% false-negative rate and a 37.8%

false-positive rate for the microarray analysis, and 13.5 % false-negative and 38.5 % false-

positive rates for sequencing analysis.

Page 101: Characterizing the role of the DNA Damage Response in

87

Figure 19

Figure 19. Validation results correlate with screen scores. Correlations are displayed for

microarray (A) and sequencing (B) analysis. For both the microarray and sequencing analysis,

the normalized CSR levels of each hairpin (obtained from the validation analysis) is displayed on

the y-axis and the CSR screen score is displayed on the x-axis.

Page 102: Characterizing the role of the DNA Damage Response in

88

Figure 20

Figure 20. Z score analysis on the microarray (A) and sequencing (B) results. Z score values

are shown on the y-axis. Each dot represents a hairpin, displayed in order of highest Z score to

lowest on the x-axis. A total of 58,127 hairpins and 43,027 hairpins are included for microarray

and sequencing analysis, respectively.

Page 103: Characterizing the role of the DNA Damage Response in

89

Figure 21

Figure 21. Volcano plot analysis on the microarray (A) and sequencing (B) results. For each

hairpin, the negative log (to the base 10) of the P-value was calculated and plotted on the y-axis.

The P-value was determined an using paired, two-tailed student’s t-test between the hairpin

levels in the IgA- population versus the IgA+ population. For each hairpin, the log (to the base 2)

of the CSR screen score (fold difference in the hairpin level in IgA- population relative to the

IgA+ population) was calculated and plotted on the x-axis. A total of 58,127 hairpins and 43,027

hairpins are included for microarray and sequencing analysis, respectively.

Page 104: Characterizing the role of the DNA Damage Response in

90

Table 2

TRC Clone ID

Microarray screen score

Sequencing screen score

TRCN0000112031 2.313781422 3.527886661

TRCN0000112032 < 1.75 4.411919732

TRCN0000112033 2.391796705 7.042824572

CSR screen scores for AID-targeting shRNA. AID-shRNA with CSR screen scores above 1.75

for microarray and sequencing analysis are displayed.

Page 105: Characterizing the role of the DNA Damage Response in

91

as screen hits, or genes that appeared in both microarray and sequencing analyses were selected

for the secondary screen to identify hairpins that did not affect AID expression. The hairpins that

reduce CSR without affecting AID likely target genes acting downstream of AID, and possibly

act in the repair phase of CSR. Surprisingly, a significant proportion of hairpins identified to

target potential CSR factors by the screen also affected AID expression in CH12F3-2 cells

(Figure 22). Of the 33 hairpins tested by the screen, 26 hairpins reduced AID protein levels. The

remaining 7 hairpins had no effect on AID protein levels and hence likely target CSR factors that

act downstream of AID expression and may play in role in DDR and DNA repair. These may be

strong candidates for detailed characterization (Chapter 4).

3.5 DISCUSSION

We developed and conducted a pooled, loss-of-function, genome-wide screen for CSR

using CH12F3-2 cells in conjunction with lentiviral-based RNA interference technology. A

previous study used the RNAi approach to identify novel factors involved in CSR, but used an

entirely different methodology (Pavri et al. 2010). Pavri et al conducted a limited screen using

8797 shRNA hairpins targeting 1745 genes, that were primarily selected based on expression in

CH12F3-2 cells and germinal centre B cells, and factors that could be associated with AID or

DNA repair based on molecular functions and literature review (Pavri et al. 2010). They

individually tested the effect of each hairpin on CSR, and identified 181 hits, of which 28 have

been previously linked to CSR (Pavri et al. 2010). The authors selected Spt5 for in-depth

characterization (Pavri et al. 2010). We, however, adopted a different approach by developing a

genome-wide screen using a pooled format (Blakely et al. 2011). While our approach tests

almost the entire set of mouse genes, owing to our large-scale approach, it is likely that our

Page 106: Characterizing the role of the DNA Damage Response in

92

Figure 22

Figure 22. A significant number of CSR candidates identified by the screen affect the

expression of AID in CH12F3-2 cells. Of the 33 hairpins identified to target potential CSR

factors by the screen (either microarray, sequencing or combined analysis), 26 hairpins affect

AID expression. 7 hairpins likely target CSR factors that act downstream of AID expression and

may play in role in DNA repair.

Page 107: Characterizing the role of the DNA Damage Response in

93

approach would fail to identify all the factors involved in the CSR process. Indeed, Spt5 did not

appear as a CSR candidate in our screen. In the microarray analysis, all 5 hairpins targeting Spt5

had a CSR screen score of ~1, and in the sequencing analysis, all 5 hairpins were eliminated

either due to low signal or a high coefficient of variation between replicates. However, our

unbiased approach enabled us to uncover novel and possibly unexpected genes that might

function in CSR. Since many genes are not differentially expressed in stimulated and

unstimulated cells, or are expressed at low levels, the approach conducted by Pavri et al might

preclude genes that fall in this category but still function in CSR.

Two distinct screen readout methods, relying on microarray-based technology and

sequencing technology, were used in parallel to screen for the strongest CSR candidates. Direct

comparison of the screen score obtained for each hairpin revealed a low correlation, with r2 of

0.2985 (P<0.0001) between the two readouts (Figure 23). Previous studies have also shown a

modest correlation between microarray and sequencing methods of analysis (Ketela et al. 2011).

This difference may be partially explained by a difference in the relationship of readout signal to

barcode PCR concentration between the two methods (Ketela et al. 2011). The absolute fold-

change observed for each hairpin varies between microarray and sequencing analyses (Ketela et

al. 2011). While sequencing provides a signal that is linearly proportional to the levels of a

shRNAs in the two populations, microarray readout demonstrates a nonlinear relationship with

shRNA levels, within the reliable signal range (Ketela et al. 2011). There is a similar trend

observed between the two readout methods, but the absolute fold-changes observed show a low

degree of equivalence. Having used only one technical replicate for the sequencing analysis and

possibly having a poor quality of genomic DNA in sequencing may have added to the deviation

between the sequencing and microarray results. The validation analysis of the screen revealed a

Page 108: Characterizing the role of the DNA Damage Response in

94

Figure 23

Figure 23. Correlation between CSR results of microarray readout and sequencing

readout. Each dot represents a hairpin. Microarray screen scores are displayed on the y-axis and

sequencing screen scores are displayed on the x-axis.

Page 109: Characterizing the role of the DNA Damage Response in

95

false-negative rate of 8.7% for microarray and 13.5% for sequencing. Given the nature of this

high-throughput, pooled genomic screen, these false-negative rates are quite minimal, suggesting

that most CSR factors may have been identified as hits by the screen (Blakely et al. 2011).

Nevertheless, cross-hybridization and signal saturation can confound the true effects of a

shRNA, leading to false-negative readout on microarray. Similarly, secondary structures formed

by shRNA sequence that affect its binding efficiency to the flow cell of the sequencer can cause

false-negative readout in sequencing analysis.

A point of surprise and concern from the screen results was the lack of known CSR factors

appearing as hits, and in particular DDR factors involved in CSR like 53BP1. This can be

attributed to multiple reasons, including those described above in explaining the causes of false-

negative readings. With respect to DDR factors specifically, certain genes may be not appear as

hits if the shRNA hairpins targeting these factors are ineffective. With respect to 53BP1, while

the pGIPZ vector based 53BP1 shRNA used in Chapter 2 above is highly effective, the pLKO.1

vector 53BP1 shRNA sequence may not be as effective, although this remains to be tested.

Alternatively, they could be effective but lead to a reduction in survival and growth or an

enhancement in apoptosis, and in both cases these targets will be lost in the pooled format, and

not be detected by the screen.

This loss-of-function approach is likely useful in identifying factors that function in CSR

at different stages. Included within this group of genes are genes important for the first phase of

CSR, which involves AID expression, stability, targeting and other factors important in DSB

formation, as well as genes involved in the second phase of CSR requiring DSB-induced DDR

and repair. As we are interested in the second phase of CSR involving DDR and repair, we

utilized a secondary screen on a subset of genes to find potential DDR and repair factors. By

Page 110: Characterizing the role of the DNA Damage Response in

96

screening for hairpins that maintained AID levels, we ruled out any genes involved in AID

expression and protein stability, and specifically focused on genes that are likely to function

downstream of AID, possibly in DDR and repair. Intriguingly, nearly 80% of the hairpins tested

affected AID expression. This result emphasizes that AID expression in CH12F3-2 cells either

relies on a whole host of factors (involved in cytokine signalling and in AID protein stability)

and/or may be highly susceptible to off-target effects. We have previously demonstrated that

CSR is highly sensitive to small changes in AID expression in CH12F3-2 cells (Parsa et al.

2012). Even the slightest effect on a single factor, whether on-target or off-target, upstream of

AID could affect AID expression. Hence, these results highlight that in the CH12F3-2 system,

AID expression is highly sensitive to alterations in the expression of many other factors. And

since CSR is highly sensitive to AID expression, future work examining CSR using the CH12F3-

2/ RNA interference system should be highly cautious in interpreting results, ensuring that any

effects on CSR are due to the gene in question and not due to off-target effects on AID. These

results further stress the pertinence to test off-target effects, by AID reconstitution experiments,

when using RNA interference-based systems.

Alternative to our secondary screen approach, another approach that utilizes the in vitro

system developed by Tasuku Honjo can be useful to focus the efforts on CSR events downstream

of AID expression (Kobayashi et al. 2009). In this system, AID is fused to the hormone-binding

domain of the estrogen receptor and can be induced by estrogen analogue, 4-hydroxytamoxifen

(Kobayashi et al. 2009). For this secondary screen, an AID-deficient CH12F3-2 cell line with an

integrated AID-estrogen receptor fusion construct described above would be generated. Only

those hairpins that were identified to target CSR factors from the primary screen would be

selected for the secondary screen. Hence, using this modified system can specifically identify

Page 111: Characterizing the role of the DNA Damage Response in

97

those hairpins that target factors acting downstream of AID expression, likely in the DDR and

repair phase of CSR.

Although this genome-wide screen has its caveats, the very high confirmation rates of the

screens prove advantageous in discovering and characterizing novel factors within the pathway,

and can be translated for other genome-wide studies as well. In particular, variations of the

established CH12F3-2/shRNA system can be used to answer novel questions. For instance,

MMEJ, the alternative end-joining pathway to NHEJ in CSR, remains completely elusive. While

a few factors may be implicated in MMEJ, very little is known about this pathway. A parallel

CSR screen approach using an NHEJ-deficient CH12F3-2 cell line instead, in which the cells

will solely rely on MMEJ to complete CSR, would identify MMEJ candidates. In this context,

genome-wide, loss-of-function, shRNA-mediated knock-down approach can be used to identify

the genes that completely abrogate CSR in the NHEJ-deficient cell line, as these are potential

MMEJ candidates.

Page 112: Characterizing the role of the DNA Damage Response in

98

4 CLASS SWITCH RECOMBINATION REQUIRES H2B

UBIQUITINATION AND DEUBIQUITINATION.

All experiments and data analysis (with the exception of ex vivo mouse CSR assay) were

performed by Shaliny Ramachandran.

Ex vivo mouse experiment was performed by Conglei Li.

Page 113: Characterizing the role of the DNA Damage Response in

99

4.1 ABSTRACT

B cells produce several classes of antibodies to mediate an effective humoral response.

This process, termed class switch recombination (CSR), requires the generation and repair of

DSBs at the antibody locus, and relies on the 53BP1-dependent DNA Damage Response (DDR).

However, the molecular mechanism and function of DDR in CSR is not fully understood. To

identify novel DDR factors in the 53BP1-signalling network of CSR, we developed and

conducted a genome-wide screen for CSR factors. In this study, we have characterized one

promising CSR candidate identified by the screen, the Drosophila enhancer of yellow 2 homolog

(Eny2). Eny2 is a structural component of the deubiquitination module of the SAGA complex,

which primarily deubiquitinates histone H2B at K120. We demonstrate that Eny2 likely acts with

the SAGA complex to facilitate CSR downstream of AID-induced DSBs, and may be required

for optimal DNAPK activation.

4.2 INTRODUCTION

Class switch recombination is the process by which B cells switch from expressing the

IgM antibody to IgG, IgE or IgA. Each class of antibody mediates a different effector function.

Critical for eliciting an effective immune response, CSR requires the enzyme activation-induced

cytidine deaminase (AID), to catalyze the deamination of deoxycytidines at the immunoglobulin

switch regions, specifically in the G1 phase of cell cycle (Schrader et al. 2007; Stavnezer et al.

2008). These AID-induced mutations are processed by the mismatch repair and base excision

repair pathways, leading to the generation of staggered DSBs at the donor and acceptor switch

regions (Schrader et al. 2009). These AID-induced DNA breaks instigate the 53BP1-dependent

DDR, a signalling network akin to that elicited by irradiation-induced DNA damage (Manis et al.

Page 114: Characterizing the role of the DNA Damage Response in

100

2004; Ward et al. 2004). DDR in turn can initiate the DNA repair program. Repair of the DSBs

in a manner that places the acceptor switch region immediately adjacent to the recombined

V(D)J segment, while deleting the intervening DNA, marks successful completion of CSR

(Chaudhuri et al. 2004). This repair process is primarily completed during the G1 phase of cell

cycle and largely depends on NHEJ, but can also rely on alternative end joining mechanisms

(Yan et al. 2007).

DSBs are ubiquitously recognized by DDR factors. This type of DNA break can be

recognized by DDR sensor complexes, including the Mre11-Rad50-Nbs1

(MRN)/ataxiatelangiectasia-mutated (ATM) complex ; the Ku70/Ku80-DNA-PKcs complex; the

ataxiatelangiectasia and Rad3-related (ATR)-ATR-interacting protein complex (Yang et al.

2003); and the more recently characterized Parp1 (Ali et al. 2012). Each sensor can initiate a

cascade of signalling events leading to cell cycle arrest and DNA repair, and in the absence of

successful repair, apoptosis. Specifically, the MRN/ATM complex has been identified to play a

role in CSR (Lumsden et al. 2004; Reina-San-Martin et al. 2004; Reina-San-Martin et al. 2005).

Upon recognition of the AID-induced DSB, ATM kinase catalyzes the phosphorylation of the

S/TQ motif in an array of substrates, key among these substrates being histone variant, H2AX

(Stiff et al. 2004). ATM-mediated phosphorylation of H2AX at S139 produces γH2AX, forming

a docking site for MDC1 (Stewart et al. 2003; Stucki et al. 2005). MDC1 in turn binds to and

further recruits Nbs1, concentrating the MRN/ATM complex at the break (Lukas et al. 2004;

Lou et al. 2006). This leads to an amplification of the γH2AX signal, which can be visualized as

distinct nuclear foci by fluorescence microscopy. ATM can also phosphorylate MDC1 on the

S/TQ consensus site, leading to the recruitment of the E3 ubiquitin ligase, RNF8 (Huen et al.

2007; Kolas et al. 2007; Mailand et al. 2007). RNF8 mediates mono-ubiquitination of histone

Page 115: Characterizing the role of the DNA Damage Response in

101

H2A and H2AX leading to the requirement of another E3 ubiquitin ligase, RNF168, which

further augments the ubiquitin levels at the break (Doil et al. 2009; Pinato et al. 2009; Stewart et

al. 2009). RNF168-mediated H2A-K15 mono-ubiquitination is recognized by the ubiquitination-

dependent recruitment motif of 53BP1, leading to the recruitment of 53BP1 (Fradet-Turcotte et

al. 2013). 53BP1 localizes to the DNA break site by simultaneously binding the RNF168-

mediated H2A-K15 mono-ubiquitin and the constitutively expressed histone H4 dimethylated-

K20 (Fradet-Turcotte et al. 2013). 53BP1 is in turn phosphorylated by the master kinase, ATM,

which leads to the recruitment of Rif1 (Silverman et al. 2004; Di Virgilio et al. 2013; Escribano-

Diaz et al. 2013; Feng et al. 2013). 53BP1-Rif1 control the DNA repair pathway choice in G1

phase of cell cycle by preferentially promoting NHEJ (Chapman et al. 2013; Escribano-Diaz et

al. 2013; Feng et al. 2013; Zimmermann et al. 2013). ATM, NBS1, Mre11, γ-H2AX, MDC1,

RNF8, RNF168, 53BP1 and Rif1 belong to the 53BP1-dependent DDR pathway, and also play a

role in CSR (Reina-San-Martin et al. 2003; Lumsden et al. 2004; Manis et al. 2004; Reina-San-

Martin et al. 2004; Ward et al. 2004; Reina-San-Martin et al. 2005; Lou et al. 2006; Li et al.

2010; Ramachandran et al. 2010; Santos et al. 2010; Bohgaki et al. 2011; Chapman et al. 2013;

Di Virgilio et al. 2013; Escribano-Diaz et al. 2013). 53BP1 mediates CSR by facilitating long-

range DNA end joining (Difilippantonio et al. 2008; Dimitrova et al. 2008), and by blocking the

DNA ends from resection, and promoting NHEJ (Bothmer et al. 2010; Bunting et al. 2010).

The established DDR pathway involves modifications of histones, but has been largely

focused to histone H2A and variant H2AX. Recently, H2B has also been suggested to participate

in the DDR (Moyal et al. 2011). Upon DNA damage, the RNF20/40 E3 ubiquitin ligase

heterodimer is recruited to DNA breaks, where it gets retained in an ATM-dependent manner

and catalyzes H2B-ubiquitination (Moyal et al. 2011). H2B-ubiquitin is thought to initiate

Page 116: Characterizing the role of the DNA Damage Response in

102

chromatin disassembly, providing access to DNA repair factors in NHEJ and HR to bind and

complete repair (Fierz et al. 2011; Moyal et al. 2011). The RNF20/40 complex primarily

mediates H2B-K120-mono-ubiquitination in mammals (and K123 in yeast) (Shiloh et al. 2011).

H2B-K120/123-ubiquitination can then be engaged by the SAGA complex, which catalyzes the

deubiquitination of H2B-K120/123 (Henry et al. 2003; Daniel et al. 2004). The SAGA complex

has two enzymatic activities, deubiquitination and acetylation (Zhang 2003). Acetylation is

mediated by the Gcn5-containing module, while deubiquitination is mediated by the Usp22

deubiquitinase, which requires Eny2 and Atnx7 as structural components in order to function

(Zhao et al. 2008). RNF20/40-mediated H2B-K120 mono-ubiquitination and Eny2-dependent,

SAGA-mediated H2B-K120 deubiquitination are known to play a role in transcription (Weake et

al. 2008).

Intriguingly, a component of the SAGA complex, Eny2, was identified as a strong

candidate in the microarray and in the sequencing-based analysis of the CSR screen above

(Chapter 3). Furthermore, Eny2-shRNA (hairpin 3 in Figure 21) reduced CSR without affecting

AID expression, suggesting that Eny2 may play a role in DDR or DNA repair. In this

investigation, we demonstrated that Eny2 plays a role in CSR. Knockdown of Eny2 led to

reduced CSR in CH12F3-2 and ex vivo mouse splenic B cells. We found that Eny2 acts

downstream of AID, and a lack of Eny2 leads to G1 cell cycle arrest and apoptosis specifically in

cells stimulated to undergo CSR. As expected, Eny2 is necessary for H2B-K120-

deubiquitination in the absence of DNA damage, and especially upon irradiation-induced DNA

damage. However, Eny2 does not appear to protect cells from irradiation-induced DNA damage,

unlike its role in CSR. Eny2 likely functions within the context of H2B-deubiquitination to

facilitate CSR, as we found that other factors in the pathway, including RNF20, RNF40 and

Page 117: Characterizing the role of the DNA Damage Response in

103

Atxn7, also affect CSR. Eny2 and H2B-deubiquitination may be required for optimal DNAPK

activation and NHEJ.

4.3 MATERIALS AND METHODS

4.3.1 In Vitro Cell Culture

CH12F3-2 cells were maintained and CSR assays were performed as described above in Chapter

2 (Ramachandran et al. 2010). Ligase IV-/- CH12F3-2 cells were provided by Keifei Yu (Han et

al. 2008). Briefly, cells were stimulated with 1 ng/mL recombinant human TGFβ1 (R&D

systems), 10 ng/mL recombinant mouse IL-4 (R&D systems) and 2 µg/mL functional grade

purified anti-mouse CD40 (eBiosciences), and analyzed by flow cytometry, as described below.

Lentiviral shRNA constructs targeting 53BP1 (V2LMM 83391) and non-silencing negative

control (RHS4346) were obtained from Open Biosystems and used as described above in

Chapter 2 (Ramachandran et al. 2010). Lentiviral shRNA constructs and protocols for shRNA

from TRC were provided by Jason Moffat (Moffat et al. 2006). shRNA constructs targeting the

negative control of GFP (TRCN0000072181), Eny2 (TRCN0000086039 and

TRCN0000086041), RNF40 (TRCN0000041048) and Atxn7 (TRCN0000104973) were used

from TRC. CH12F3-2 cells were transduced with lentivirus for 24 hours, on 24 well plates for

bulk transductions or diluted and plated on 96 well plates to obtain clones. Positively transduced

cells were selected with 1 µg/mL puromycin for 3 days. For growth curve analysis, CH12F3-2

cells were diluted to a concentration of 1x105 cells/ml and aliquoted in duplicate on a 96-well

plate. At various time points, the numbers of live-trypan blue excluded cells were counted using

a haemocytometer. ATM inhibitor (KU 55933) and DNAPK inhibitor (NU 7441) were obtained

from Tocris Bioscience.

Page 118: Characterizing the role of the DNA Damage Response in

104

4.3.2 Ex vivo mouse experiments

Splenic B cells were purified from wild-type C57BL/6 mouse, using a negative selection mouse

B cell enrichment kit (Stemcell technology). The cells were cultured in complete RPMI media

with 25 ng/mL LPS for 24 hours, followed by lentiviral transduction in the presence of 8 ng/mL

polybrene and LPS at a final concentration of 25 ng/mL for 24 hours. Positively transduced cells

were then selected with 0.6 μg/mL puromycin and stimulated with 25 ng/mL IL4 for 4 days, and

then analyzed by flow cytometery as described below.

4.3.3 Cell staining and Flow cytometry

For CSR analysis, CH12F3-2 cells were stained with PE conjugated anti-mouse IgA (Southern

Biotech), and ex vivo mouse B cells were stained with PE conjugated anti-mouse IgG1 antibody

clone A85-1 (BD Bioscience). For cell cycle analysis, cells were incubated with 10 µM BrdU for

1 hour followed by ethanol fixation overnight. Cells were then washed and incubated in 2 M HCl

for 30 minutes, washed and stained with FITC conjugated anti-BrdU antibody clone PRB-1

(eBiosciences), and then washed and incubated in 20 µg/mL PI and 10 µg/mL RNAse A for 30

minutes. For apoptosis analysis, CH12F3-2 cells were surface stained for Annexin V using

Annexin V-APC Apoptosis Detection Kit (eBiosciences). Stained cells were analyzed by

FACSCalibur (BD Bio) and FlowJo software (Truestar Inc.).

4.3.4 RNA extraction, RT and semi-quantitative and qPCR analysis

RNA extraction was performed using Trizol (Invitrogen), followed by DNase treatment

(Fermentas) and reverse transcription reactions using a Superscript III kit (Invitrogen) or

Maxima kit (Thermo Scientific) to prepare cDNA. cDNA samples were diluted and subject to

Page 119: Characterizing the role of the DNA Damage Response in

105

PCR reactions for semi-quantitative RT-PCR or qPCR reactions. For semi-quantitative PCR, Iµ-

Cµ and Iα-Cα amplification primers are same as previously described (Muramatsu et al. 2000),

Gapdh pimers (Gapdh RTPCR For and Rev) can be found in Table 3. For qPCR analysis,

primers for Gapdh (Gapdh qPCR For and Rev), Eny2 (Eny2 qPCR For and Rev), RNF20 (Rnf20

qPCR For and Rev), RNF40 (Rnf40 qPCR For and Rev) and Atnx7 (Atxn7 qPCR For and Rev)

can be found on Table 3.

4.3.5 Plasmids

pFucci-S/G2/M Green vector (AM-V9016) was obtained from Amalgam. Eny2 was PCR-

amplified from CH12F3-2 cDNA using Eny2 Clone For and Rev primers. PCR product was

digested with HindIII and NotI restriction enzymes (NEB), and ligated into pCDNA3.1+ vector

using T4 DNA ligase (NEB). shRNA-resistant Eny2 vector was generated by site-directed

mutagenesis of the Eny2-pCDNA3.1+ plasmid, by PCR with Eny2-SDM For and Rev primers

that introduced 7 silent mutations in the shRNA-complimentary sequence of Eny2 cDNA,

followed by DpnI (NEB) digestion, ethanol precipitation and bacterial transformation. Primers

used are listed in Table 3. All plasmids sequences were verified by the TCAG sequencing

facility.

4.3.6 Plasmid Integration assay

Plasmid integration was performed by plasmid linearization and electroporation at 550V and 50

µF. For Eny2 reconstitution assays, shRNA-resistant Eny2 expression vector and control vector

were linearized by BglII (NEB), electroporated, diluted and plated on 96 well plates to obtain

clones. For blunt-end plasmid integration assay, pCDNA3.1+ plasmid was linearized with SspI

Page 120: Characterizing the role of the DNA Damage Response in

106

Table 3

Primers

Primer Purpose Forward Sequence Reverse Sequence

Gapdh Semi-quantitative

PCR

AACTTTGGCATTGTGGAAGG GGAGACAACCTGGTCCTCAG

Gapdh qPCR AACTTTGGCATTGTGGAAGG GGAGACAACCTGGTCCTCAG

Eny2 qPCR GGTAATTAAAGAAAAAGGA

CTAGAACACGTT

GAGCTCCTTCTTTACACTGT

CAGGTA

RNF40 qPCR ATGCAGATGCTGACGAGATC

CT

CATCCTTCTTGCGGGTGTTA

C

Atxn7 qPCR TCTGAATGCACAGCTATGGA

AGA

AGGAACACGGGCAGAAACA

G

Eny2 Clone Cloning GAACAAGCTTCACCATGGTG

GTTAGCAAGATGAA

TGTCGCGGCCGCTCAAGCGT

AATCTGGAACATCGTATGGG

TATGATCCAAGGCTGGCATG

CTGAGCAAGG

Eny2 SDM Site-directed

mutagenesis

Aactggagaaagagaacgcctcaaagaact

gttgcgggcgaagttaattgagtgcggctgga

aggatcag

Ctgatccttccagccgcactcaattaacttcgc

ccgcaacagttctttgaggcgttctctttctcca

gtt

Page 121: Characterizing the role of the DNA Damage Response in

107

(NEB), electroporated plated on 96 well plates at various dilutions. Plasmid integration

efficiency was determined by counting the number of expanding clones dividing by the total

number of cells plated.

4.3.7 Irradiation assays

CH12F3-2 cell lines were irradiated with various doses of γradiation using a gammacell 1000

irradiator. For clonal survival assays, cells were subsequently plated at various dilutions in

duplicate on 96 well plates, and survival was determined by counting the number of expanding

clones normalized to plating efficiency. For apoptosis assays, cells were plated on 24 well plates

and analyzed by Annexin V staining as described above. For western blot analysis, cells were

harvested to prepare lysates at various time points.

4.3.8 Western blot analyses

H2AX (Upstate), H2B-K120-ubiquitin (Millipore), Kap1-S824-phosphorylation (Cell

signalling), p53-S15-phosphorylation, βactin (Sigma) antibodies were used as specified by the

manufacturers’ protocols.

4.3.9 Statistical Analysis

Analyses were performed on GraphPad Prism. For Mann Whitney tests and linear regression

analysis, p values of 0.05 or less were considered significant.

Page 122: Characterizing the role of the DNA Damage Response in

108

4.4 RESULTS

4.4.1 Eny2 acts downstream of AID in CSR

To understand how DDR facilitates CSR at a molecular level, we developed and conducted

a screen for novel CSR factors from which we identified potential DDR and repair factors

(Chapter 3). Briefly, we performed a loss-of-function, genome-wide screen to identify CSR

factors, and subsequently screened a subset of strong CSR candidates for normal AID

expression, to identify possible DDR and repair factors. Eny2 was identified as a strong CSR

candidate in the microarray and in the sequencing based screen analysis (Chapter 3), and knock-

down of Eny2 did not affect AID expression (Hairpin 3 in Figure 22). We first determined

whether Eny2 is indeed a CSR factor, using the CH12F3-2 mouse B cell line. We ascertained

whether shRNA-mediated knock-down of Eny2 led to a decrease in CSR. Of the five Eny2-

targeting shRNA lentiviral vectors tested, shEny2-39 and shEny2-41 mediated knock-down of

Eny2 transcript by 50% and 80%, respectively, as compared to a control hairpin that targets GFP

(shGFP-81) (Figure 24A). The shEny2-39 and shEny2-41 hairpins led to a 30% and 70%

decrease in CSR to IgA in CD40/Il-4/TGFβ (CIT)-stimulated CH12F3-2 cells, respectively

(Figure 24B). Given that shEny2-39-mediated knock-down of Eny2 is minimal, we carried out

subsequent analysis using the shEny2-41 lentiviral vector. ShEny2-41 maintains knock-down of

CSR in CH12F3-2 cells over a 3-day period (Figure 24C). The effect of this hairpin on CSR is

independent of off-target effects, as reconstitution of a shEny2-41-knock-down clone with a

shRNA-resistant Eny2-cDNA expression vector, rescued both Eny2 transcript levels and CSR

(Figures 24D and 24E). On the other hand, reconstitution with an empty vector did not rescue

Eny2 transcript levels or CSR (Figure 24D). Furthermore, ex vivo spleen mouse B cells

Page 123: Characterizing the role of the DNA Damage Response in

109

Figure 24

Figure 24. ShRNA-mediated knock-down of Eny2 reduces CSR. (A) Eny2 transcript levels

relative to Gapdh were determined by qPCR analysis from CH12F3-2 cells with Eny2-shRNA

hairpins, shEny2-39 and shEny2-41, and compared against GFP-targeting control shRNA,

shGFP-81. (B) CSR analysis on CH12F3-2 cells with two Eny2-shRNA, shEny2-39 and shEny2-

41, compared to control shGFP-81. (C) CSR time course analysis of CH12F3-2 cells with Eny2-

shRNA (shEny2-41) compared to control-shRNA (shGFP-81). Statistical significance was tested

by two-tailed, Mann-Whitney test at each time point. (D) CSR analysis of clones derived from a

CH12F3-2 shEny2-41-knock-down clone reconstituted with pCDNA3.1 vector control

(pCDNA3.1) that maintained low Eny2 transcript levels, or clones derived from the CH12F3-2

shEny2-41-knock-down clone reconstituted with shRNA-resistant Eny2 expression vector

(Eny2-pCDNA3.1) that rescued Eny2 transcript levels. CSR was compared to a parental

CH12F3-2 clone (clone I). Statistical significance was tested by two-tailed Mann Whitney test.

(E) Correlation of CSR and Eny2 transcript levels on clones derived from CH12F3-2 shEny2-41

clone reconstituted with either pCDNA3.1 control vector or shRNA-resistant Eny2 expression

vector. Grey circles are clones reconstituted with pCDNA3.1, white circles are clones

Page 124: Characterizing the role of the DNA Damage Response in

110

reconstituted with Eny2-pCDNA3.1, and the black circle is parental CH12F3-2 clone I.

Statistical significance was tested by linear regression analysis. (F) CSR analysis of ex vivo

mouse spleen B cells with Eny2 shRNA (shEny2-41) compared to control-shRNA (shGFP-81).

Page 125: Characterizing the role of the DNA Damage Response in

111

transduced with shEny2-41 also demonstrated a reduced ability to undergo CSR, further

validating Eny2 as a CSR factor (Figure 24F).

To confirm that Eny2 acts downstream of AID (based on observations from secondary

screen in Chapter 3), we determined whether AID protein expression or germline transcription,

both of which are critical for CSR, were affected by shEny2-41. As shown in Figures 25A and

25B, upon stimulation with the CIT cocktail, there was no significant difference in AID

expression or germline transcripts in CH12F3-2 cells containing the shEny-41 or shGFP-81

vectors. Interestingly, we found that while unstimulated shEny2-41 cells grew at a rate similar to

control cells, upon stimulation with the CIT cocktail, there was a significant decrease in the

ability of shEny2-41 cells to multiply in comparison to the controls (Figures 25C and 25D). To

assess whether knock-down of Eny2 affects cell cycle progression, we analysed the cell cycle

profile of cells stimulated for 24 hours using BrdU/PI staining, and found that compared to

control cells, shEny2-41 cells maintained a higher level of G1 cells and a lower level of S-phase

cells, suggesting that cells with reduced Eny2 undergo G1-cell cycle arrest upon stimulation

(Figure 25E). Furthermore, we found that stimulation also led to a significant increase in

apoptosis of shEny2-41 cells when compared to control cells at 48 hours post-stimulation (Figure

25F). Taken together, these data suggest that upon stimulation, AID-induced DSBs form in G1,

but in Eny2-deficient cells this leads to G1 cell cycle arrest and apoptosis most likely due to an

inability to repair the DSBs. Collectively, this implicates Eny2 as a factor acting downstream of

AID, likely in DDR and/or the DNA repair phase of CSR.

Page 126: Characterizing the role of the DNA Damage Response in

112

Figure 25

Figure 25. Eny2 likely acts downstream of AID in CSR. (A) Western blot analysis on AID

expression relative to βactin, in CH12F3-2 cells with Eny2-shRNA (shEny2-41) compared to

control-shRNA (shGFP-81). Cells were stimulated with the CIT-cocktail for 24 hours. Statistical

significance tested by two-tailed, Mann Whitney test. (B) Semi-quantitative RT-PCR analysis on

germline transcription through the µ switch region (Iμ-Cμ) and α switch region (Iα-Cα) with

Gapdh loading control, on CH12F3-2 cells with Eny2-shRNA (shEny2-41) compared to control-

shRNA (shGFP-81). Cells were stimulated with CIT for 24 hours. The blot is representative of

two replicates tested at both, 24 hours and 48 hours post-stimulation. Growth of unstimulated (C)

and CIT-stimulated (D) CH12F3-2 cells with Eny2-shRNA (shEny2-41) compared to control-

shRNA (shGFP-81) over time. (E) Cell cycle profile of CH12F3-2 cells with Eny2-shRNA

(shEny2-41) compared to control-shRNA (shGFP-81) determined by BrdU/PI staining at 24

hours post-CIT-stimulation. Statistical significance was tested by two-tailed, Mann Whitney test.

(F) Apoptosis levels of CH12F3-2 cells with Eny2-shRNA (shEny2-41) compared to control-

shRNA (shGFP-81) as determined by Annexin V staining at 48 hours post-CIT-stimulation.

Statistical significance tested by two-tailed, Mann Whitney test.

Page 127: Characterizing the role of the DNA Damage Response in

113

4.4.2 Eny2 mediates H2B-K120-deubiquitination upon DNA damage, independent of 53BP1

Previous work demonstrated that Eny2 provides a structural role to the SAGA complex,

which mediates H2B-K120 deubiquitination during transcription. We first ascertained whether

Eny2 carries out a similar function in CH12F3-2 cells. As such, we tested the steady-state levels

of H2B-K120-ubiquitin in CH12F3-2 cells. Indeed, shEny2-41 led to a 5-fold increase in steady-

state H2B-K120-ubiquitin (Figures 26A and 26B). Moreover, we tested the steady-state H2B-

K120-Ub levels in clones expressing different levels of Eny2 transcript, and found that there is a

strong inverse correlation between Eny2 transcript levels and steady-state H2B-K120-ubiquitin

levels (Figure 26C), suggesting that Eny2 is required for H2B-K120-deubiquitination in

CH12F3-2 cells.

To determine whether Eny2 plays a role in DDR, we tested whether Eny2-deficiency

causes radiation sensitivity. Irradiation of asynchronous cells did not affect clonal survival or

apoptosis of shEny2-41 cells when compared to control cells (Figures 27A and 27B), suggesting

that Eny2-knock-down does not affect radiation sensitivity. Given that Eny2-knock-down led to

increased G1 cell cycle arrest and apoptosis upon CSR stimulation (Figures 27E and 27F), and

given that AID-induced mutations occur in G1 cell cycle, we postulated that Eny2 may confer

radiation resistance to cells specifically during the G1 cell cycle. To test this notion, we utilized

the mAG1-hGEM S/G2/M-labeling fluorescent ubiquitination-based cell cycle indicator (Fucci),

which expresses an Azami-Green1 (mAG1) fluorescent protein fused to part of the human

Geminin protein, which accumulates in S/G2/M-phases of cell cycle. By transfected cells with

this expression vector and sorting for mAG1-negative cells (i.e. G1 cells), we enriched for G1

cells by up to ~65% G1, which is considerable given only ~30% of unsorted CH12F3-2 cells are

in G1 (data not shown). However, irradiation of mAG1-negative CH12F3-2 shEny2-41 cells that

Page 128: Characterizing the role of the DNA Damage Response in

114

Figure 26

Figure 26. Eny2 is important for H2B-K120-deubiquitination in CH12F3-2 cells.

Representative (A) and compiled (B) western blot analysis on steady-state levels of H2B-K120-

ubiquitin relative to βactin, in CH12F3-2 cells with Eny2-shRNA (shEny2-41) compared to

control-shRNA (shGFP-81). Statistical significance tested by two-tailed Mann Whitney test. (C)

Correlating H2B-K120-ubiquitin levels to Eny2 transcript levels on clones derived from a

CH12F3-2 shEny2-41 clone reconstituted with either pCDNA3.1 control vector or shRNA-

resistant Eny2 expression vector from Figure 1D and 1E above. Grey circles are clones

reconstituted with pCDNA3.1, white circles are clones reconstituted with Eny2-pCDNA3.1, and

the black circle is parental CH12F3-2 clone I. Statistical significance was tested by linear

regression analysis.

Page 129: Characterizing the role of the DNA Damage Response in

115

Figure 27

Figure 27. Eny2 mediates H2B-K120-deubiquitination in DDR, but is independent of the

53BP1-pathway. Clonal survival assays on asynchronous (A), or S/G2/M-Fucci excluded (C),

CH12F3-2 cells, with Eny2-shRNA (shEny2-41) or control-shRNA (shGFP-81), exposed to

varying doses of γ-radiation. Apoptosis assays, as determined by Annexin V staining, on

asynchronous (B) or S/G2/M-Fucci excluded (D) CH12F3-2 cells, with Eny2-shRNA (shEny2-

41) or control-shRNA (shGFP-81), exposed to varying doses of γ-radiation. CH12F3-2 cells with

Eny2-shRNA (shEny2-41) or control-shRNA (shGFP-81) were exposed to 8 Grays of γ-

radiation, and harvested at various time points for western blot analysis of H2B-K120-ubiquitin

levels relative to βactin. Representative blot (E) and compiled data (F) are displayed. A CH12F3-

2 clone with 53BP1-shRNA (sh53BP1) and a control clone with non-silencing shRNA (Non-

silencing) were exposed to 8 Grays of γ-radiation, and harvested at various time points for

western blot analysis of H2B-K120-ubiquitin levels relative to βactin.

Page 130: Characterizing the role of the DNA Damage Response in

116

were primarily in the G1 phase of the cell cycle, also did not affect clonal survival and apoptosis

as compared to mAG1-negative control cells (Figure 27C and 27D). This suggests that Eny2

may not have a significant role in the cellular response to irradiation-induced DNA damage.

Interestingly, shEny2-41 cells acquire very high accumulation of H2B-K120-ubiquitin starting

about 2 hours post-irradiation when compared to control cells (Figures 27E and 27F). Together,

these data suggest that Eny2 is required for deubiquitination of H2B-K120-ubiquitin upon γ-

irradiation, but unlike its role in CSR, Eny2 may not be as important for repair of irradiation-

induced DNA damage.

To further characterize H2B-K120-ubiquitin in DDR, we tested the kinetic profile of H2B-

K120-ubiquitin in parallel with Kap1-phosphorylation and γH2AX formation. While γH2AX and

Kap1-phosphorylation formed immediately upon irradiation (Figures 28A and 28B), H2B-K120-

ubiquitin formation lagged behind, with the signal becoming visible at about 1.5 hours post-

irradiation and peaking at 2-3 hours (Figure 28C). This suggests that H2B-K120-ubiquitin is part

of a later phase of DDR, unlike Kap1-phosphorylation and γH2AX, and may either be dependent

on the 53BP1 pathway or be an independent arm of DDR. To test whether H2B-K120-ubiquitin

requires the 53BP1 pathway, we determined whether knock-down of 53BP1 affected H2B-K120-

ubiquitin accumulation upon irradiation. Although a ~5-fold reduction in 53BP1 protein

completely abrogated CSR (Figure 8A, 8E and 8F), there was no difference in H2B-K120-

ubiquitin upon irradiation of sh53BP1 cells in comparison to control cells, suggesting that H2B-

K120-ubiquitin may be independent of the 53BP1 arm, or upstream of 53BP1 in DDR (Figure

27G).

Page 131: Characterizing the role of the DNA Damage Response in

117

Figure 28

Figure 28. Time-course analysis of irradiation-induced post-translational modifications

involved in DDR. CH12F3-2 control (shGFP-81) cells were exposed to 8 Grays of γ-radiation,

and harvested at various time points for western blot analysis of γH2AX (A), Kap1-Ser824-

phosphorylation (B) and H2B-K120-ubiquitin (C), all relative to βactin.

Page 132: Characterizing the role of the DNA Damage Response in

118

4.4.3 H2B-K120-ubiquitination/deubiquitination is important for CSR.

We hypothesized that Eny2 might also function in H2B-K120-deubiquitination in CSR. To

examine this, we first tested whether other factors in the H2B-K120-

ubiquitination/dequibitionation pathway are required for CSR. First, we identified a shRNA

hairpin targeting Atxn7 that decreased Atxn7 transcript levels (Figure 29A). Interestingly, akin to

Eny2 knock-down, Atxn7-shRNA lead to an increased accumulation of irradiation-induced H2B-

K120-ubiquitin, and most importantly, also decreased CSR (Figure 29B and 29E). This suggests

that Eny2 likely functions with the Atxn7-containing, SAGA complex in DDR and CSR. To

compare the extent of H2B-K120-ubiquitin in cells with knock-down of Eny2 relative to control

cells, western blot analysis was performed on Eny2-knock-down cell lysates loaded in 10-fold

seriel dilutions along with GFP-shRNA control cells (Figure 29E). These blots were exposed for

short (60 second exposure) in which H2B-K120-ubiquitin is apparent in cells with shEny2-41

but not control cells, and a long (400 second exposure) in which trace amounts of H2B-K120-

ubiquitin is visible in the control cells, highlighting that while H2B-K120-ubiquitination occurs

in irradiated control cells, the accumulation of H2B-K120-ubiquitin is significantly higher in

Eny2-knock-down cells.

The SAGA complex acts downstream of RNF20/40 in the H2B-ubiquitination pathway,

and so we tested whether RNF20/40 also plays a role in CSR. We identified a shRNA targeting

RNF40 that decreased RNF40 transcript levels (Figure 29C). This RNF40-shRNA also lead to

decreased irradiation-induced H2B-K120-ubiquitin accumulation, and most importantly, also

reduced CSR, thus implicating RNF20/40 in CSR (Figure 29D and 29F). Together, these data

suggest that RNF20/40 and the SAGA complex, which are factors involved in the H2B-

ubiquitination/deubiquitination pathway, play a role in CSR.

Page 133: Characterizing the role of the DNA Damage Response in

119

Figure 29

Figure 29. The H2B-K120-ubiquitination/deubiquitination pathway facilitates CSR. (A)

qPCR analysis of CH12F3-2 cells with Atxn7-shRNA (shAtxn7-73) compared to control-shRNA

(shGFP-81). (B) CH12F3-2 cells with Atxn7-shRNA (shAtxn7-73), Eny2-shRNA (shEny2-41)

as positive control, and negative control-shRNA (shGFP-81), were exposed to 8 Grays of γ-

radiation, and harvested at 3 hours for western blot analysis of H2B-K120-ubiquitin levels

relative to βactin. (C) qPCR analysis of CH12F3-2 cells with RNF40-shRNA (shRNF40-48)

compared to control-shRNA (shGFP-81). (D) CH12F3-2 cells with RNF40-shRNA (shRNF40-

48) and control-shRNA (shGFP-81) were exposed to 8 Grays of γ-radiation, and harvested at

various time points for western blot analysis of H2B-K120-ubiquitin levels relative to βactin. (E)

Comparing H2B-K120-ubiquitin levels between CH12F3-2 cells with Eny2-shRNA (shEny2-41)

and control-shRNA (shGFP-81) at 3 hours after 8 Gy IR. shEny-41 cell lysates were loaded in

10-fold serial dilutions, and blots were exposed for short (60 seconds) and long (400 second)

exposure. (F) CSR analysis on CH12F3-2 cells with Atxn7-shRNA (shAtxn7-73), RNF40-

shRNA (shRNF40-48) compared to control-shRNA (shGFP-81). (G) CSR analysis on CH12F3-2

clones with either normal steady-state levels of H2B-K120-ubitquitin or high steady-state levels

of H2B-K120-ubitquitin. Grey circles are clones derived from Eny2-knock-down clone

reconstituted with pCDNA3.1, white circles are clones reconstituted with Eny2-pCDNA3.1, and

the black circle is parental CH12F3-2 clone I from Figure 1D and 1E above. Statistical

significance was tested by two-tailed, Mann Whitney test.

Page 134: Characterizing the role of the DNA Damage Response in

120

To further examine whether H2B-K120-ubiquitin is linked to CSR, we determined

whether there was a correlation between H2B-K120-ubiquitin levels and CSR. CH12F3-2 clones

with increasing Eny2 transcript levels had an inversely correlating decrease in H2B-K120-

ubiquitin steady-state levels (Figure 26C). Taking advantage of these clones, we compared the

clones containing wild-type or low steady-state H2B-K120-ubiquitin levels, against clones that

contained high steady-state H2B-K120-ubiquitin (at least ~ 2.5-fold above parental levels) for

their ability to switch. Indeed clones with wild-type or low steady-state H2B-K120-ubiquitin

maintained wild-type levels of CSR, while clones with high steady-state H2B-K120-ubiquitin

demonstrated significantly reduced levels of CSR (Figure 29G). Indeed, this inverse correlation

between H2B-K120-ubiquitin and CSR suggests that H2B-deubiquitination may be important to

CSR.

4.4.4 Accumulation of H2B-K120-ubiquitin may inhibit DNAPK activation

Next, we tested whether Eny2 and H2B-K120-ubiquitin may directly affect the DDR, by

examining various DDR signalling events in Eny2-knock-down cells. Specifically, we tested

whether shEny2-41 affected the induction of Kap1-Ser824-phosphorylation, p53-Ser15

phosphorylation and γH2AX formation upon irradiation. Interestingly, shEny2-41 did not affect

Kap1-Ser824 or p53-Ser15 phosphorylation, both of which are post-translational modifications

that rely solely on ATM (Figures 30A and 30B). However, shEny2-41 modestly reduced the

formation of γH2AX, a phosphorylation event that can rely on either ATM or DNAPK in

irradiation-induced DDR (Figure 30C). This suggests that knock-down of Eny2 may interfere

with optimal DNAPK activity. To delineate the Eny2 and DNAPK relationship further, we tested

γH2AX induction upon irradiation in cells pre-incubated with either ATM inhibitor (ATMi) or

Page 135: Characterizing the role of the DNA Damage Response in

121

Figure 30

Figure 30. Accumulation of H2B-K120-ubiquitin may inhibit DNAPK. CH12F3-2 Eny2-

shRNA (shEny2-41) cells and control-shRNA (shGFP-81) cells were exposed to 8 Grays of γ-

radiation, and harvested at various time points for western blot analysis of p53-Ser15

phosphorylation (A), Kap1-Ser824-phosphorylation (B) and γH2AX formation (C), all relative to

βactin. CH12F3-2 Eny2-shRNA (shEny2-41) cells and control (shGFP-81) cells were exposed to

8 Grays of γ-radiation, and harvested at various time points for western blot analysis of γH2AX

formation relative to βactin, in the presence of either DNAPKi (D) and (E), or ATMi (F) and (G).

(D) and (F) are representative blots and (E) and (G) are the compiled data. (H) Blunt-ended

plasmid integration efficiencies of CH12F3-2 and Ligase IV-deficient (Ligase IV-/-) cells

containing either Eny2-shRNA (shEny2-41) or control-shRNA (shGFP-81).

Page 136: Characterizing the role of the DNA Damage Response in

122

DNAPK inhibitor (DNAPKi). Cells pre-incubated with ATMi will largely depend on DNAPK

activity to produce γH2AX upon irradiation. Similarly, cells pre-incubated with DNAPKi will

largely rely on ATM to produce the γH2AX signal upon irradiation. Intriguingly, in cells with

ATMi, shEny2-41 led to reduced γH2AX signal when compared to control cells (Figures 30D

and 30E), while DNAPKi showed no difference between the shEny2-41 and control cells after

irradiation (Figures 30F and 30G). Together, this suggests that knock-down of Eny2 may be

affecting DNAPK recruitment and/or activation, hence affecting NHEJ. Furthermore, we find

that while the difference between blunt-ended plasmid integration efficiency in Eny2-knock-

down and control cells is not statistically significant, there was a trend in the Eny2-knock-down

cells to display a decreased ability to integrate blunt-ended plasmid. Interestingly, there is no

difference between shEny2-41 and control cells in Ligase IV-deficient CH12F3-2 cells,

suggesting a possible epistasis between Eny2 and NHEJ (Figure 30H).

4.5 DISCUSSION

H2B ubiquitination has been suggested to primarily facilitate chromatin disassembly (Fierz

et al. 2011), providing access to downstream DNA repair factors to repair the DSBs (Shiloh et

al. 2011). While RNF20/40 functions downstream of ATM, this H2B ubiquitination pathway is

independent from the ATM-MDC1-RNF8 pathway, suggesting that H2B-ubiquitination and

MDC1-RNF8-mediated DDR are two distinct arms of the ATM-signalling cascade (Moyal et al.

2011). In a similar manner, we showed that H2B-ubiquitination is independent of 53BP1, which

acts downstream RNF8. Our findings further demonstrate that ATM orchestrates the DDR by

initiating multiple branches of signalling pathways. While the MDC1-RNF8-53BP1 pathway

promotes NHEJ, the RNF20/40 and Eny2-SAGA complex-mediated H2B-

Page 137: Characterizing the role of the DNA Damage Response in

123

ubiquitination/deubiquitination pathway functions distinctly, possibly in disassembling

chromatin and providing DSB access to the DNA repair factors recruited by the 53BP1-

dependent pathway.

A surprising finding was that while Eny2-knock-down caused increased G1 cell cycle

arrest and apoptosis upon CSR stimulation, irradiation of both asynchronous and ~65% G1-

sorted cells neither affected clonal survival nor apoptosis. This finding should first be confirmed

using a more homogenous population of G1 cells, which may be achieved using the mAG1-

hGEM-S/G2/M Fucci marker in conjunction with the mKO2-Cdt1-G1 Fucci marker.

Alternatively, it is possible the Eny2 does not confer protection from irradiation-induced

apoptosis specifically in CH12F3-2 mouse B cells. Hence, the role of Eny2 in irradiation-

induced clonal survival and apoptosis must be tested in other cell lines, including U2OS. If Eny2

is not important for radiation resistance, this suggests that Eny2 may exhibit differential roles in

the ubiquitous DDR pathway, versus the DDR pathway in CSR. We speculate that AID

expression may cause the CH12F3-2 B cells to be more prone to apoptosis in comparison to the

cells cycling in culture, which are used for irradiation. Indeed, it has been previously

demonstrated that AID-/- mice show significantly reduced apoptosis of germinal centre B cells

(Zaheen et al. 2009). Given the increased tendency to undergo apoptosis, Eny2-knock-down

cells that are unable to repair the AID-induced damage may immediately resort to death.

However, cells cycling in culture may be able recover from DNA damage, even in Eny2-knock-

down.

Previous studies that focused on RNF20/40 demonstrated that H2B-ubiquitination upon

DNA damage is important for both the HR and NHEJ repair pathways (Moyal et al. 2011). We

observed that Eny2 and H2B-deubiquitination displays epistasis with NHEJ, although we have

Page 138: Characterizing the role of the DNA Damage Response in

124

not tested whether Eny2-mediated H2B-deubiquitination also plays a role HR. This can be tested

using in vitro NHEJ and HR substrates.

Intriguingly, we observed that Eny2 knock-down may potentially affect DNAPK

activation. It is possible that H2B-deubiquitination is required for DNAPK recruitment.

Alternatively, DNAPK enzymatic activity may be affected in the absence of H2B-

deubiquitination. DNAPKcs contains two autophosphorylation clusters, an ABCDE cluster

associated with increased DNA end processing and a PQR cluster associated with decreased end

processing (Meek et al. 2008). DNAPKcs autophosphorylation is required for efficient DNA

repair, and in the absence of autophosphorylation, the DNAPKcs is retained at the unrepaired

break, suggesting that autophosphorylation can cause DNAPKcs dissociation from the break

(Uematsu et al. 2007). It is thought that autophosphorylation mediates DNAPKcs association and

dissociation from the break and that DNAPKcs dynamics may be important for NHEJ (Cui et al.

2005). In particular, S2056 autophosphorylation within the PQR cluster occurs upon irradiation-

induced DNA damage, is important for NHEJ, and is attenuated in S-phase cells (Chen et al.

2005). Investigating S2056 autophosphorylation in Eny2-deficient cells may provide further

insights on the link between Eny2 and DNAPKcs. Observing an effect on DNAPKcs S2056

autophosphorylation in Eny2-knock-down cells would suggest that Eny2 may be important for

the recruitment and/or activation of DNAPKcs, which in turn can can affect NHEJ. Consistent

with this notion, we observed a possible epistasis between Eny2 and NHEJ using a plasmid

integration assay, implicating that Eny2 is involved in NHEJ and not MMEJ. However, this can

be tested using an in vitro MMEJ substrate, or by determining the extent of microhomology

usage in switch junctions of Eny2-knock-down cells (Stavnezer et al. 2010).

Page 139: Characterizing the role of the DNA Damage Response in

125

H2B-ubiquitination is also considered a prerequisite for histone H3 trimethylation at K4

(H3K4me3) (Chandrasekharan et al. 2010). H3K3me3 has been linked to DDR (Faucher et al.

2010), and to CSR (Daniel et al. 2010; Stanlie et al. 2010). In the context of CSR, H3K4me3 is

associated with germline transcripts and is suggested to be important for AID targeting and

switch region DSB formation (Daniel et al. 2010; Stanlie et al. 2010; Begum et al. 2012).

However, we observed no difference in germline transcripts, an increase in G1 cell cycle arrest

and an increase in apoptosis upon stimulation of Eny2-knock-down cells, suggesting that AID-

induced DSB is likely unaffected by Eny2 deficiency. H3K4me3 has also been linked to PTIP

(Cho et al. 2007), and PTIP has previously been linked to CSR (Daniel et al. 2010; Schwab et al.

2011). This study found that PTIP can act downstream of DSBs, and may be involved in

chromatin moification and accessibility (Daniel et al. 2010). Furthermore, PTIP is known to

interact with 53BP1 (Jowsey et al. 2004; Wu et al. 2009; Callen et al. 2013), suggesting that the

53BP1-PTIP interaction may be linked to H2B-ubiquitination and H3K4me3. However, recent

work suggested that the 53BP1-PTIP interaction is dispensable for CSR (Callen et al. 2013),

making it unlikely that H3K4me3-PTIP-53BP1 association is important to CSR. Further

investigation is needed to understand whether H2B-ubiquitination directly affects H3K4me3 at

AID-induced DSBs and whether this is linked to CSR, either upstream or upstream of AID.

Page 140: Characterizing the role of the DNA Damage Response in

126

5 DISCUSSION AND FUTURE DIRECTIONS

5.1 RNF8 and RNF168 play a role in CSR.

Using the in vitro CH12F3-2/ shRNA system that we established, we demonstrated that the

E3 ubiquitin ligasae cascade in DDR, composed of RNF8 and RNF168, facilitate CSR. By using

a knock-down approach, we showed that reducing RNF8 and RNF168 protein levels caused a

correlating reduction in CSR. Subsequent in vivo mouse gene knock-out models confirmed our

findings. RNF8 knock-out mice displayed a 50-70% decrease CSR, and RNF168 knock-out mice

displaye a 40-50% decrease CSR (Li et al. 2010; Santos et al. 2010; Bohgaki et al. 2011), further

validating our findings from an in vitro system, using an in vivo system.

5.2 DDR does not form a linear signalling cascade in CSR.

While factors in the 53BP1-dependent DDR pathway, including ATM, Mre11, NBS1, γ-

H2AX, MDC1, RNF8, RNF168, 53BP1and Rif1, have all been demonstrated to play a role in

CSR, in all likelihood, this cascade forms part of a network rather than a linear succession of

signalling events. Within this network, the significance and functional contribution of each

protein to CSR differs, possibly due to functional redundancy and the ability of other factors

within the network to compensate for the lack of certain factors. In vivo studies using gene-

deletion mouse models, have revealed that each of the DDR factors involved do not equally

contribute to CSR. ATM-, MRN-, and H2AX-deficiency leads to a ~70% reduction in CSR

(Reina-San-Martin et al. 2003; Lumsden et al. 2004; Reina-San-Martin et al. 2004; Reina-San-

Martin et al. 2005), MDC1-deficiency leads to 25%-50% reduction (Lou et al. 2006), RNF8-

deficiency causes 50-70% decrease (Li et al. 2010; Santos et al. 2010), RNF168-deficiency has

Page 141: Characterizing the role of the DNA Damage Response in

127

40-50% (Bohgaki et al. 2011), 53BP1-deficiency leads to the most profound effect, reducing

CSR by ~90% reduction (Manis et al. 2004; Ward et al. 2004), and Rif1 causes 80-90%

reduction (Chapman et al. 2013; Di Virgilio et al. 2013). It is possible that in certain studies, the

use of gene-trap knock-out mice allowed for residual levels of protein expression, leading to a

less pronounced effect on CSR (Ramachandran et al. 2010). Alternatively, functional

redundancy between certain proteins can also explain the reduced effects. For instance, given

that 53BP1 and Rif1 exhibit the most drastic effects on CSR, it is likely that these factors are

absolutely crucial to the process. As the upstream factors, ATM, MRN, H2AX, MDC1, RNF8

and RNF168 do not exhibit the same degree of effect on CSR, it is plausible that there are

alternative methods of 53BP1-Rif1 recruitment and activation leading to CSR. For instance, low

levels of 53BP1 can bind to H4K20me2 in a MDC1-RNF8-RNF168-independent fashion

(Fradet-Turcotte et al. 2013), and this recruitment may account for the reduced contribution of

MDC1-RNF8-RNF168 to CSR. 53BP1-Rif may be critical, either because these proteins have

multiple molecular functions in CSR and/or they do not exhibit functional redundancies with any

other proteins.

Recent work has shown that 53BP1-Rif1 promote NHEJ, while blocking DNA-resection-

dependent pathways (Chapman et al. 2013; Di Virgilio et al. 2013; Escribano-Diaz et al. 2013;

Feng et al. 2013; Zimmermann et al. 2013), and so 53BP1 and Rif1 may be considered as a bona

fide NHEJ factors. However, while 53BP1-deletion leads to 90% reduction in CSR (Manis et al.

2004; Ward et al. 2004), Ligase IV- or XRCC4- deletion only leads to 50-70% reduction (Yan et

al. 2007), suggesting that 53BP1 may also have NHEJ-independent functions. Moreover, switch

junctions of 53BP1-deficient B cells do not display the increased use of microhomology that is

observed with Ligase IV-deficiency (Manis et al. 2004). While this suggests that 53BP1 could

Page 142: Characterizing the role of the DNA Damage Response in

128

potentially assist alternative-end joining mechanisms like MMEJ, the observation that 53BP1

blocks DNA resection, and thus microhomology-based repair pathways, provides opposing

evidence (Bothmer et al. 2010; Bunting et al. 2010), creating a subject of controversy.

Delineating the exact molecular link between 53BP1-Rif and NHEJ and a possible relation with

MMEJ would provide a better understanding of the interconnections between these pathways.

Specifically, creating a NHEJ- and 53BP1- double-deficient system, and investigating the extent

of microhomology usage in the absence of both NHEJ and 53BP1, may provide further insights

into the role of 53BP1 in NHEJ.

5.3 H2B-ubiquitination in DDR and repair.

In addition to investigating aspects of the 53BP1-dependent DDR, we have also

characterized a 53BP1-independent aspect of DDR, H2B-ubiquitination. While we and others

have demonstrated that H2B-ubiquitination occurs independently of MDC1, RNF8 and 53BP1

(Moyal et al. 2011), it is, however, possible that H2B-ubiquitination/deubiquitination may affect

aspects of the MDC-RNF8/168-53BP1 pathway. While the recruitment of MDC1, RNF8 and

53BP1 was not affected by depletion of the H2B-ubiquitinase, RNF20, the disappearance of

γH2AX and 53BP1 foci was at a slower rate (Moyal et al. 2011). This persistence of γH2AX and

53BP1 is likely due to lack of DNA repair. However, there might be molecular links between

H2B-deubiquitination and the 53BP1-dependent DDR pathway. Future work examining 53BP1

recruitment to, and dissociation from DNA breaks, by foci formation assays in the context of

persisting H2B-ubiquitination upon irradiation of Eny2-depleted cells, can start underpinning

any possible links between these two arms of ATM signalling.

Page 143: Characterizing the role of the DNA Damage Response in

129

Interestingly, our work has suggested that H2B-deubiquitination may be linked to

DNAPKcs recruitment and/or activation. It is possible that either H2B-ubiquitination actively

blocks DNAPKcs or H2B-deubiquitination promotes DNAPKcs activation. Firstly, whether

H2B-ubiquitin specifically affects DNAPKcs recruitment and/or activation must be investigated

by foci-formation analysis on DNAPKcs, and the activated form of DNAPKcs, which is

autophosphorylated at S2056 in human cells (Chen et al. 2005). Subsequent studies can focus on

the molecular mechanism linking H2B-ubiquitination to DNAPKcs activation, possibly by

observing foci formation kinetics of H2B-ubiquitin and phosphorylated DNAPKcs, and by in

vitro biochemical studies.

5.4 Generating an in vivo mouse model to investigate H2B-ubiquitination.

The findings from our in vitro studies should be further corroborated using in vivo systems.

Generating a mouse model in which Eny2 is deleted at the genetic level would be ideal.

However, given the potential role of Eny2 in transcription, it is likely that a complete deletion

would cause embryonic lethality. Indeed, Atanassov et al. attempted to create a Gcn5 knock-out

mouse (Atanassov et al. 2009). Gcn5 is a component of the histone acetyltransferase module of

the SAGA complex (Zhang 2003). Gcn5 deletion leads to embryonic lethality at embryonic day

8.5, due to increased apoptosis (Atanassov et al. 2009). However, embryos with a catalytically-

dead Gcn5 not only survive longer, but also do not display the increased apoptosis seen in Gcn5-

deletion, suggesting that the embryonic lethality at day 8.5 is not due to histone acetyltransferase

function of Gcn5 (Atanassov et al. 2009). It has been shown that in order to maintain the

deubiquitination function, both the acetylation and deubiquitination modules need to associate

and form a complete SAGA complex (Atanassov et al. 2009). Given that the enzymatic function

Page 144: Characterizing the role of the DNA Damage Response in

130

requires all the structural components of the complex, it is very likely that akin to Gcn5, deletion

of Eny2 will cause embryonic lethality. However, generating conditional knock-out mice using

the Cre-LoxP system, in which Eny2 is specifically deleted in either the B cell compartment or

germinal center B cells would be a valuable tool. In addition to Eny2-deletion, which would

eliminate H2B-deubiquitination, creating a conditional RNF20 knock-out mouse, which would

eliminate H2B-ubiquitination, would provide further insights. Briefly, if the mice with complete

knock-out of Eny2 are viable, it would be most interesting to test radiation sensitivity, genome

stability, cancer predisposition, V(D)J recombination and CSR in these mice. However, if Eny2

is specifically deleted in the B cell compartment then the role of Eny2 in V(D)J recombination

and CSR would be investigated.

Page 145: Characterizing the role of the DNA Damage Response in

131

References Acs, K., et al. (2011). "The AAA-ATPase VCP/p97 promotes 53BP1 recruitment by removing L3MBTL1

from DNA double-strand breaks." Nat Struct Mol Biol 18(12): 1345-50. Ahnesorg, P., et al. (2006). "XLF interacts with the XRCC4-DNA ligase IV complex to promote DNA

nonhomologous end-joining." Cell 124(2): 301-13. Ali, A. A., et al. (2012). "The zinc-finger domains of PARP1 cooperate to recognize DNA strand breaks."

Nat Struct Mol Biol 19(7): 685-92. Allen, C. D., et al. (2007). "Germinal-center organization and cellular dynamics." Immunity 27(2): 190-

202. Alt, F. W., et al. (1984). "Ordered rearrangement of immunoglobulin heavy chain variable region

segments." Embo J 3(6): 1209-19. Anderson, L., et al. (2001). "Phosphorylation and rapid relocalization of 53BP1 to nuclear foci upon DNA

damage." Mol Cell Biol 21(5): 1719-29. Aoufouchi, S., et al. (2008). "Proteasomal degradation restricts the nuclear lifespan of AID." J Exp Med

205(6): 1357-68. Arakawa, H., et al. (2002). "Requirement of the Activation-Induced Deaminase (AID) Gene for

Immunoglobulin Gene Conversion." Science 295(5558): 1301-6. Atanassov, B. S., et al. (2009). "Gcn5 and SAGA regulate shelterin protein turnover and telomere

maintenance." Mol Cell 35(3): 352-64. Bachl, J., et al. (2001). "Increased transcription levels induce higher mutation rates in a hypermutating

cell line." J Immunol 166(8): 5051-7. Bakkenist, C. J., et al. (2003). "DNA damage activates ATM through intermolecular autophosphorylation

and dimer dissociation." Nature 421(6922): 499-506. Bassing, C. H., et al. (2002). "Increased ionizing radiation sensitivity and genomic instability in the

absence of histone H2AX." Proc Natl Acad Sci U S A 99(12): 8173-8. Bassing, C. H., et al. (2003). "Histone H2AX: a dosage-dependent suppressor of oncogenic translocations

and tumors." Cell 114(3): 359-70. Basu, U., et al. (2005). "The AID antibody diversification enzyme is regulated by protein kinase A

phosphorylation." Nature 438(7067): 508-11. Basu, U., et al. (2011). "The RNA exosome targets the AID cytidine deaminase to both strands of

transcribed duplex DNA substrates." Cell 144(3): 353-63. Baumgarth, N. (2011). "The double life of a B-1 cell: self-reactivity selects for protective effector

functions." Nat Rev Immunol 11(1): 34-46. Begum, N. A., et al. (2012). "The histone chaperone Spt6 is required for activation-induced cytidine

deaminase target determination through H3K4me3 regulation." J Biol Chem 287(39): 32415-29. Bernard, O., et al. (1978). "Sequences of mouse immunoglobulin light chain genes before and after

somatic changes." Cell 15(4): 1133-1144. Blackford, A. N., et al. (2011). "When cleavage is not attractive: non-catalytic inhibition of ubiquitin

chains at DNA double-strand breaks by OTUB1." DNA Repair (Amst) 10(2): 245-9. Blakely, K., et al. (2011). "Pooled lentiviral shRNA screening for functional genomics in mammalian cells."

Methods Mol Biol 781: 161-82. Boboila, C., et al. (2012). "Classical and alternative end-joining pathways for repair of lymphocyte-

specific and general DNA double-strand breaks." Adv Immunol 116: 1-49. Boboila, C., et al. (2010). "Alternative end-joining catalyzes robust IgH locus deletions and translocations

in the combined absence of ligase 4 and Ku70." Proc Natl Acad Sci U S A 107(7): 3034-9. Boboila, C., et al. (2012). "Robust chromosomal DNA repair via alternative end-joining in the absence of

X-ray repair cross-complementing protein 1 (XRCC1)." Proc Natl Acad Sci U S A 109(7): 2473-8.

Page 146: Characterizing the role of the DNA Damage Response in

132

Boboila, C., et al. (2010). "Alternative end-joining catalyzes class switch recombination in the absence of both Ku70 and DNA ligase 4." J Exp Med 207(2): 417-27.

Bohgaki, T., et al. (2011). "Genomic instability, defective spermatogenesis, immunodeficiency, and cancer in a mouse model of the RIDDLE syndrome." PLoS Genet 7(4): e1001381.

Bortnick, A., et al. (2013). "What is and what should always have been: long-lived plasma cells induced by T cell-independent antigens." J Immunol 190(12): 5913-8.

Bosma, G. C., et al. (2002). "DNA-dependent protein kinase activity is not required for immunoglobulin class switching." J Exp Med 196(11): 1483-95.

Bothmer, A., et al. (2011). "Regulation of DNA end joining, resection, and immunoglobulin class switch recombination by 53BP1." Mol Cell 42(3): 319-29.

Bothmer, A., et al. (2010). "53BP1 regulates DNA resection and the choice between classical and alternative end joining during class switch recombination." J Exp Med 207(4): 855-65.

Bothmer, A., et al. (2013). "Mechanism of DNA resection during intrachromosomal recombination and immunoglobulin class switching." J Exp Med 210(1): 115-23.

Botuyan, M. V., et al. (2006). "Structural basis for the methylation state-specific recognition of histone H4-K20 by 53BP1 and Crb2 in DNA repair." Cell 127(7): 1361-73.

Boubnov, N. V., et al. (1993). "V(D)J recombination coding junction formation without DNA homology: processing of coding termini." Mol Cell Biol 13(11): 6957-68.

Brack, C., et al. (1978). "A complete immunoglobulin gene is created by somatic recombination." Cell 15(1): 1-14.

Bradford, H. E., et al. (2001). "Antibody-dependent killing of virus-infected targets by NK-like cells in bovine blood." J Vet Med B Infect Dis Vet Public Health 48(8): 637-40.

Bransteitter, R., et al. (2003). "Activation-induced cytidine deaminase deaminates deoxycytidine on single-stranded DNA but requires the action of RNase." Proc Natl Acad Sci U S A 100(7): 4102-7.

Braster, R., et al. (2013). "Myeloid cells as effector cells for monoclonal antibody therapy of cancer." Methods.

Brummelkamp, T. R., et al. (2006). "An shRNA barcode screen provides insight into cancer cell vulnerability to MDM2 inhibitors." Nat Chem Biol 2(4): 202-6.

Buck, D., et al. (2006). "Cernunnos, a novel nonhomologous end-joining factor, is mutated in human immunodeficiency with microcephaly." Cell 124(2): 287-99.

Bunting, S. F., et al. (2010). "53BP1 inhibits homologous recombination in Brca1-deficient cells by blocking resection of DNA breaks." Cell 141(2): 243-54.

Burma, S., et al. (2001). "ATM phosphorylates histone H2AX in response to DNA double-strand breaks." J Biol Chem 276(45): 42462-7.

Callen, E., et al. (2013). "53BP1 mediates productive and mutagenic DNA repair through distinct phosphoprotein interactions." Cell 153(6): 1266-80.

Cascalho, M., et al. (1998). "Mismatch repair co-opted by hypermutation." Science 279: 1207-10. Casellas, R., et al. (1998). "Ku80 is required for immunoglobulin isotype switching." Embo J 17(8): 2404-

11. Celeste, A., et al. (2002). "Genomic instability in mice lacking histone H2AX." Science 296(5569): 922-7. Chandrasekharan, M. B., et al. (2010). "Histone H2B ubiquitination and beyond: Regulation of

nucleosome stability, chromatin dynamics and the trans-histone H3 methylation." Epigenetics 5(6): 460-8.

Chapman, J. R., et al. (2013). "RIF1 is essential for 53BP1-dependent nonhomologous end joining and suppression of DNA double-strand break resection." Mol Cell 49(5): 858-71.

Chapman, J. R., et al. (2008). "Phospho-dependent interactions between NBS1 and MDC1 mediate chromatin retention of the MRN complex at sites of DNA damage." EMBO Rep 9(8): 795-801.

Page 147: Characterizing the role of the DNA Damage Response in

133

Chaturvedi, P., et al. (1999). "Mammalian Chk2 is a downstream effector of the ATM-dependent DNA damage checkpoint pathway." Oncogene 18(28): 4047-54.

Chaudhuri, J., et al. (2004). "Class-switch recombination: interplay of transcription, DNA deamination and DNA repair." Nat Rev Immunol 4(7): 541-52.

Chaudhuri, J., et al. (2007). "Evolution of the immunoglobulin heavy chain class switch recombination mechanism." Adv Immunol 94: 157-214.

Chaudhuri, J., et al. (2004). "Replication protein A interacts with AID to promote deamination of somatic hypermutation targets." Nature 430(7003): 992-8.

Chaudhuri, J., et al. (2003). "Transcription-targeted DNA deamination by the AID antibody diversification enzyme." Nature 422(6933): 726-30.

Chen, B. P., et al. (2005). "Cell cycle dependence of DNA-dependent protein kinase phosphorylation in response to DNA double strand breaks." J Biol Chem 280(15): 14709-15.

Chen, H. T., et al. (2000). "Response to RAG-mediated VDJ cleavage by NBS1 and gamma-H2AX." Science 290(5498): 1962-5.

Chen, L., et al. (2000). "Interactions of the DNA ligase IV-XRCC4 complex with DNA ends and the DNA-dependent protein kinase." J Biol Chem 275(34): 26196-205.

Cheng, H. L., et al. (2009). "Integrity of the AID serine-38 phosphorylation site is critical for class switch recombination and somatic hypermutation in mice." Proc Natl Acad Sci U S A.

Cho, Y. W., et al. (2007). "PTIP associates with MLL3- and MLL4-containing histone H3 lysine 4 methyltransferase complex." J Biol Chem 282(28): 20395-406.

Clark, M. R., et al. (2004). "B-cell antigen receptor signaling requirements for targeting antigen to the MHC class II presentation pathway." Curr Opin Immunol 16(3): 382-7.

Cui, X., et al. (2005). "Autophosphorylation of DNA-dependent protein kinase regulates DNA end processing and may also alter double-strand break repair pathway choice." Mol Cell Biol 25(24): 10842-52.

Daniel, J. A., et al. (2010). "PTIP promotes chromatin changes critical for immunoglobulin class switch recombination." Science 329(5994): 917-23.

Daniel, J. A., et al. (2004). "Deubiquitination of histone H2B by a yeast acetyltransferase complex regulates transcription." J Biol Chem 279(3): 1867-71.

de Yebenes, V. G., et al. (2008). "miR-181b negatively regulates activation-induced cytidine deaminase in B cells." J Exp Med 205(10): 2199-206.

Defrance, T., et al. (2011). "T cell-independent B cell memory." Curr Opin Immunol 23(3): 330-6. DeFranco, A. L., et al. (1995). "Signal transduction by the B-cell antigen receptor." Ann N Y Acad Sci 766:

195-201. Di Noia, J., et al. (2002). "Altering the pathway of immunoglobulin hypermutation by inhibiting uracil-

DNA glycosylase." Nature 419(6902): 43-8. Di Noia, J. M., et al. (2007). "Molecular mechanisms of antibody somatic hypermutation." Annu Rev

Biochem 76: 1-22. Di Virgilio, M., et al. (2013). "Rif1 prevents resection of DNA breaks and promotes immunoglobulin class

switching." Science 339(6120): 711-5. Dickerson, S. K., et al. (2003). "AID Mediates Hypermutation by Deaminating Single Stranded DNA." J Exp

Med 197(10): 1291-6. Difilippantonio, S., et al. (2008). "53BP1 facilitates long-range DNA end-joining during V(D)J

recombination." Nature 456(7221): 529-33. Dimitrova, N., et al. (2008). "53BP1 promotes non-homologous end joining of telomeres by increasing

chromatin mobility." Nature 456(7221): 524-8. Dinkelmann, M., et al. (2009). "Multiple functions of MRN in end-joining pathways during isotype class

switching." Nat Struct Mol Biol 16(8): 808-13.

Page 148: Characterizing the role of the DNA Damage Response in

134

Doil, C., et al. (2009). "RNF168 binds and amplifies ubiquitin conjugates on damaged chromosomes to allow accumulation of repair proteins." Cell 136(3): 435-46.

Dorsett, Y., et al. (2008). "MicroRNA-155 suppresses activation-induced cytidine deaminase-mediated Myc-Igh translocation." Immunity 28(5): 630-8.

Dudas, A., et al. (2004). "DNA double-strand break repair by homologous recombination." Mutat Res 566(2): 131-67.

Dunkelberger, J. R., et al. (2010). "Complement and its role in innate and adaptive immune responses." Cell Res 20(1): 34-50.

Early, P., et al. (1980). "An immunoglobulin heavy chain variable region gene is generated from three segments of DNA: VH, D and JH." Cell 19(4): 981-92.

Eccleston, J., et al. (2011). "Mismatch repair proteins MSH2, MLH1, and EXO1 are important for class-switch recombination events occurring in B cells that lack nonhomologous end joining." J Immunol 186(4): 2336-43.

Edelson, B. T., et al. (2001). "Intracellular antibody neutralizes Listeria growth." Immunity 14(5): 503-12. Escribano-Diaz, C., et al. (2013). "A cell cycle-dependent regulatory circuit composed of 53BP1-RIF1 and

BRCA1-CtIP controls DNA repair pathway choice." Mol Cell 49(5): 872-83. Faucher, D., et al. (2010). "Methylated H3K4, a transcription-associated histone modification, is involved

in the DNA damage response pathway." PLoS Genet 6(8). Fearon, D. T., et al. (2000). "Regulation of B lymphocyte responses to foreign and self-antigens by the

CD19/CD21 complex." Annu Rev Immunol 18: 393-422. Feng, L., et al. (2013). "RIF1 counteracts BRCA1-mediated end resection during DNA repair." J Biol Chem

288(16): 11135-43. Fernandez-Capetillo, O., et al. (2002). "DNA damage-induced G2-M checkpoint activation by histone

H2AX and 53BP1." Nat Cell Biol 4(12): 993-7. Fierz, B., et al. (2011). "Histone H2B ubiquitylation disrupts local and higher-order chromatin

compaction." Nat Chem Biol 7(2): 113-9. Fradet-Turcotte, A., et al. (2013). "53BP1 is a reader of the DNA-damage-induced H2A Lys 15 ubiquitin

mark." Nature 499(7456): 50-4. Franco, S., et al. (2006). "H2AX prevents DNA breaks from progressing to chromosome breaks and

translocations." Mol Cell 21(2): 201-14. Franco, S., et al. (2008). "DNA-PKcs and Artemis function in the end-joining phase of immunoglobulin

heavy chain class switch recombination." J Exp Med 205(3): 557-64. Frieder, D., et al. (2009). "The concerted action of Msh2 and UNG stimulates somatic hypermutation at

A . T base pairs." Mol Cell Biol 29(18): 5148-57. Fukita, Y., et al. (1998). "Somatic hypermutation in the heavy chain locus correlates with transcription."

Immunity 9: 105-114. Galanty, Y., et al. (2012). "RNF4, a SUMO-targeted ubiquitin E3 ligase, promotes DNA double-strand

break repair." Genes Dev 26(11): 1179-95. Galanty, Y., et al. (2009). "Mammalian SUMO E3-ligases PIAS1 and PIAS4 promote responses to DNA

double-strand breaks." Nature 462(7275): 935-9. Gazumyan, A., et al. (2012). "Activation-induced cytidine deaminase in antibody diversification and

chromosome translocation." Adv Cancer Res 113: 167-90. Gold, M. R., et al. (1992). "Membrane Ig cross-linking regulates phosphatidylinositol 3-kinase in B

lymphocytes." J Immunol 148(7): 2012-22. Goldberg, M., et al. (2003). "MDC1 is required for the intra-S-phase DNA damage checkpoint." Nature

421(6926): 952-6. Gottlieb, T. M., et al. (1993). "The DNA-dependent protein kinase: requirement for DNA ends and

association with Ku antigen." Cell 72(1): 131-42.

Page 149: Characterizing the role of the DNA Damage Response in

135

Gowthaman, U., et al. (2010). "T cell help to B cells in germinal centers: putting the jigsaw together." Int Rev Immunol 29(4): 403-20.

Grawunder, U., et al. (1997). "Activity of DNA ligase IV stimulated by complex formation with XRCC4 protein in mammalian cells." Nature 388(6641): 492-5.

Greeve, J., et al. (2003). "Expression of activation-induced cytidine deaminase in human B-cell non-Hodgkin lymphomas." Blood 101(9): 3574-80.

Guzzo, C. M., et al. (2012). "RNF4-dependent hybrid SUMO-ubiquitin chains are signals for RAP80 and thereby mediate the recruitment of BRCA1 to sites of DNA damage." Sci Signal 5(253): ra88.

Haberstroh, A., et al. (2008). "Neutralizing host responses in hepatitis C virus infection target viral entry at postbinding steps and membrane fusion." Gastroenterology 135(5): 1719-1728 e1.

Hammond, E. M., et al. (2003). "ATR/ATM targets are phosphorylated by ATR in response to hypoxia and ATM in response to reoxygenation." J Biol Chem 278(14): 12207-13.

Han, L., et al. (2008). "Altered kinetics of nonhomologous end joining and class switch recombination in ligase IV--deficient B cells." J Exp Med 205(12): 2745-53.

Han, L., et al. (2008). "Altered kinetics of nonhomologous end joining and class switch recombination in ligase IV-deficient B cells." J Exp Med 205(12): 2745-53.

Hasham, M. G., et al. (2010). "Widespread genomic breaks generated by activation-induced cytidine deaminase are prevented by homologous recombination." Nat Immunol 11(9): 820-6.

Hasham, M. G., et al. (2012). "Activation-induced cytidine deaminase-initiated off-target DNA breaks are detected and resolved during S phase." J Immunol 189(5): 2374-82.

Heffernan, T. P., et al. (2002). "An ATR- and Chk1-dependent S checkpoint inhibits replicon initiation following UVC-induced DNA damage." Mol Cell Biol 22(24): 8552-61.

Helmink, B. A., et al. (2011). "H2AX prevents CtIP-mediated DNA end resection and aberrant repair in G1-phase lymphocytes." Nature 469(7329): 245-9.

Henry, K. W., et al. (2003). "Transcriptional activation via sequential histone H2B ubiquitylation and deubiquitylation, mediated by SAGA-associated Ubp8." Genes Dev 17(21): 2648-63.

Honjo, T. (1983). "Immunoglobulin genes." Annu Rev Immunol 1: 499-528. Hoshino, K., et al. (1999). "Cutting edge: Toll-like receptor 4 (TLR4)-deficient mice are hyporesponsive to

lipopolysaccharide: evidence for TLR4 as the Lps gene product." J Immunol 162(7): 3749-52. Hozumi, N., et al. (1976). "Evidence for somatic rearrangement of immunoglobulin genes coding for

variable and constant regions." Proc Natl Acad Sci U S A 73(10): 3628-32. Huen, M. S., et al. (2007). "RNF8 transduces the DNA-damage signal via histone ubiquitylation and

checkpoint protein assembly." Cell 131(5): 901-14. Iwabuchi, K., et al. (1994). "Two cellular proteins that bind to wild-type but not mutant p53." Proc Natl

Acad Sci U S A 91(13): 6098-102. Iwabuchi, K., et al. (1998). "Stimulation of p53-mediated transcriptional activation by the p53-binding

proteins, 53BP1 and 53BP2." J Biol Chem 273(40): 26061-8. Jackson, S. P., et al. (2013). "Regulation of DNA damage responses by ubiquitin and SUMO." Mol Cell

49(5): 795-807. Jia, Q., et al. (2013). "Poly(ADP-ribose)polymerases are involved in microhomology mediated back-up

non-homologous end joining in Arabidopsis thaliana." Plant Mol Biol 82(4-5): 339-51. Joller, N., et al. (2011). "Antibody-Fc receptor interactions in protection against intracellular pathogens."

Eur J Immunol 41(4): 889-97. Jowsey, P. A., et al. (2004). "Human PTIP facilitates ATM-mediated activation of p53 and promotes

cellular resistance to ionizing radiation." J Biol Chem 279(53): 55562-9. Juang, Y. C., et al. (2012). "OTUB1 co-opts Lys48-linked ubiquitin recognition to suppress E2 enzyme

function." Mol Cell 45(3): 384-97.

Page 150: Characterizing the role of the DNA Damage Response in

136

Jung, I., et al. (2012). "H2B monoubiquitylation is a 5'-enriched active transcription mark and correlates with exon-intron structure in human cells." Genome Res 22(6): 1026-35.

Kao, C. F., et al. (2004). "Rad6 plays a role in transcriptional activation through ubiquitylation of histone H2B." Genes Dev 18(2): 184-95.

Kari, V., et al. (2011). "The H2B ubiquitin ligase RNF40 cooperates with SUPT16H to induce dynamic changes in chromatin structure during DNA double-strand break repair." Cell Cycle 10(20): 3495-504.

Kavli, B., et al. (2005). "B cells from hyper-IgM patients carrying UNG mutations lack ability to remove uracil from ssDNA and have elevated genomic uracil." J Exp Med 201(12): 2011-21.

Ketela, T., et al. (2011). "A comprehensive platform for highly multiplexed mammalian functional genetic screens." BMC Genomics 12: 213.

Kim, S. T., et al. (1999). "Substrate specificities and identification of putative substrates of ATM kinase family members." J Biol Chem 274(53): 37538-43.

Kinoshita, K., et al. (2001). "A hallmark of active class switch recombination: transcripts directed by I promoters on looped-out circular DNAs." Proc Natl Acad Sci U S A 98(22): 12620-3.

Kiyokawa, H., et al. (2008). "In vivo roles of CDC25 phosphatases: biological insight into the anti-cancer therapeutic targets." Anticancer Agents Med Chem 8(8): 832-6.

Kluin, P. M., et al. (1995). "IgD class switching: identification of a novel recombination site in neoplastic and normal B cells." Eur J Immunol 25(12): 3504-8.

Kobayashi, J., et al. (2004). "NBS1 and its functional role in the DNA damage response." DNA Repair (Amst) 3(8-9): 855-61.

Kobayashi, M., et al. (2009). "AID-induced decrease in topoisomerase 1 induces DNA structural alteration and DNA cleavage for class switch recombination." Proc Natl Acad Sci U S A 106(52): 22375-80.

Kolas, N. K., et al. (2007). "Orchestration of the DNA-damage response by the RNF8 ubiquitin ligase." Science 318(5856): 1637-40.

Komori, T., et al. (1993). "Lack of N regions in antigen receptor variable region genes of TdT-deficient lymphocytes." Science 261(5125): 1171-5.

Kothapalli, N., et al. (2008). "Cutting edge: a cis-acting DNA element targets AID-mediated sequence diversification to the chicken Ig light chain gene locus." J Immunol 180(4): 2019-23.

Kracker, S., et al. (2011). "Insights into the B cell specific process of immunoglobulin class switch recombination." Immunol Lett 138(2): 97-103.

Kunimoto, D. Y., et al. (1988). "Regulation of IgA differentiation in CH12LX B cells by lymphokines. IL-4 induces membrane IgM-positive CH12LX cells to express membrane IgA and IL-5 induces membrane IgA-positive CH12LX cells to secrete IgA." J Immunol 141(3): 713-20.

Lahdesmaki, A., et al. (2004). "Delineation of the role of the Mre11 complex in class switch recombination." J Biol Chem 279(16): 16479-87.

Lamont, K. R., et al. (2013). "Attenuating homologous recombination stimulates an AID-induced antileukemic effect." J Exp Med 210(5): 1021-33.

Langerak, P., et al. (2007). "A/T mutagenesis in hypermutated immunoglobulin genes strongly depends on PCNAK164 modification." J Exp Med 204(8): 1989-98.

Langerak, P., et al. (2011). "Regulatory networks integrating cell cycle control with DNA damage checkpoints and double-strand break repair." Philos Trans R Soc Lond B Biol Sci 366(1584): 3562-71.

Larijani, M., et al. (2005). "The mutation spectrum of purified AID is similar to the mutability index in Ramos cells and in ung(-/-)msh2(-/-) mice." Immunogenetics 56(11): 840-5.

Larijani, M., et al. (2007). "AID associates with single-stranded DNA with high affinity and a long complex half-life in a sequence-independent manner." Mol Cell Biol 27(1): 20-30.

Page 151: Characterizing the role of the DNA Damage Response in

137

Latif, C., et al. (2004). "DNA damage checkpoint maintenance through sustained Chk1 activity." J Cell Sci 117(Pt 16): 3489-98.

Lavin, M. F., et al. (2007). "ATM activation and DNA damage response." Cell Cycle 6(8): 931-42. Lebecque, S. G., et al. (1990). "Boundaries of somatic mutation in rearranged immunoglobulin genes: 5'

boundary is near the promoter, and 3' boundary is approximately 1 kb from V(D)J gene." Journal of Experimental Medicine 172(6): 1717-27.

Lee, J. H., et al. (2004). "Direct activation of the ATM protein kinase by the Mre11/Rad50/Nbs1 complex." Science 304(5667): 93-6.

Lee, J. H., et al. (2005). "ATM activation by DNA double-strand breaks through the Mre11-Rad50-Nbs1 complex." Science 308(5721): 551-4.

Li, L., et al. (2010). "Rnf8 deficiency impairs class switch recombination, spermatogenesis, and genomic integrity and predisposes for cancer." J Exp Med 207(5): 983-97.

Liang, L., et al. (2008). "Human DNA ligases I and III, but not ligase IV, are required for microhomology-mediated end joining of DNA double-strand breaks." Nucleic Acids Res 36(10): 3297-310.

Lieber, M. R. (2010). "The mechanism of double-strand DNA break repair by the nonhomologous DNA end-joining pathway." Annu Rev Biochem 79: 181-211.

Lieber, M. R., et al. (1988). "Lymphoid V(D)J recombination: nucleotide insertion at signal joints as well as coding joints." Proc Natl Acad Sci U S A 85(22): 8588-92.

Liu, M., et al. (2008). "Two levels of protection for the B cell genome during somatic hypermutation." Nature 451(7180): 841-5.

Liyanage, M., et al. (2000). "Abnormal rearrangement within the alpha/delta T-cell receptor locus in lymphomas from Atm-deficient mice." Blood 96(5): 1940-6.

Lou, Z., et al. (2006). "MDC1 maintains genomic stability by participating in the amplification of ATM-dependent DNA damage signals." Mol Cell 21(2): 187-200.

Lu, Z., et al. (2013). "BCL6 breaks occur at different AID sequence motifs in Ig-BCL6 and non-Ig-BCL6 rearrangements." Blood 121(22): 4551-4.

Lukas, C., et al. (2004). "Mdc1 couples DNA double-strand break recognition by Nbs1 with its H2AX-dependent chromatin retention." Embo J 23(13): 2674-83.

Lumsden, J. M., et al. (2004). "Immunoglobulin class switch recombination is impaired in Atm-deficient mice." J Exp Med 200(9): 1111-21.

Luo, G., et al. (1999). "Disruption of mRad50 causes embryonic stem cell lethality, abnormal embryonic development, and sensitivity to ionizing radiation." Proc Natl Acad Sci U S A 96(13): 7376-81.

Luo, K., et al. (2012). "Sumoylation of MDC1 is important for proper DNA damage response." Embo J 31(13): 3008-19.

Ma, Y., et al. (2002). "Hairpin opening and overhang processing by an Artemis/DNA-dependent protein kinase complex in nonhomologous end joining and V(D)J recombination." Cell 108(6): 781-94.

Mackay, F., et al. (2003). "BAFF AND APRIL: a tutorial on B cell survival." Annu Rev Immunol 21: 231-64. Mailand, N., et al. (2007). "RNF8 ubiquitylates histones at DNA double-strand breaks and promotes

assembly of repair proteins." Cell 131(5): 887-900. Mallette, F. A., et al. "RNF8- and RNF168-dependent degradation of KDM4A/JMJD2A triggers 53BP1

recruitment to DNA damage sites." Embo J 31(8): 1865-78. Manis, J. P., et al. (2002). "IgH Class Switch Recombination to IgG1 in DNA-PKcs-Deficient B Cells."

Immunity 16(4): 607-617. Manis, J. P., et al. (1998). "Ku70 is required for late B cell development and immunoglobulin heavy chain

class switching." J Exp Med 187(12): 2081-9. Manis, J. P., et al. (2004). "53BP1 links DNA damage-response pathways to immunoglobulin heavy chain

class-switch recombination." Nat Immunol 5(5): 481-7.

Page 152: Characterizing the role of the DNA Damage Response in

138

Martin, R. M., et al. (1998). "The need for IgG2c specific antiserum when isotyping antibodies from C57BL/6 and NOD mice." J Immunol Methods 212(2): 187-92.

Martomo, S. A., et al. (2004). "A role for Msh6 but not Msh3 in somatic hypermutation and class switch recombination." J Exp Med 200(1): 61-8.

Matsuoka, S., et al. (1998). "Linkage of ATM to cell cycle regulation by the Chk2 protein kinase." Science 282(5395): 1893-7.

Mattiroli, F., et al. (2012). "RNF168 ubiquitinates K13-15 on H2A/H2AX to drive DNA damage signaling." Cell 150(6): 1182-95.

McBride, K. M., et al. (2004). "Somatic hypermutation is limited by CRM1-dependent nuclear export of activation-induced deaminase." J Exp Med 199(9): 1235-44.

Meek, K., et al. (2008). "DNA-PK: the means to justify the ends?" Adv Immunol 99: 33-58. Meerang, M., et al. "The ubiquitin-selective segregase VCP/p97 orchestrates the response to DNA

double-strand breaks." Nat Cell Biol 13(11): 1376-82. Milstein, C., et al. (1998). "Both DNA strands of antibody genes are hypermutation targets." Proc Natl

Acad Sci U S A 95: 8791-4. Minegishi, Y., et al. (2000). "Mutations in activation-induced cytidine deaminase in patients with hyper

IgM syndrome." Clin Immunol 97(3): 203-10. Moffat, J., et al. (2006). "A lentiviral RNAi library for human and mouse genes applied to an arrayed viral

high-content screen." Cell 124(6): 1283-98. Mold, C., et al. (2002). "Protection from Streptococcus pneumoniae infection by C-reactive protein and

natural antibody requires complement but not Fc gamma receptors." J Immunol 168(12): 6375-81.

Mongini, P. K., et al. (2001). "Cytokine dependency of human B cell cycle progression elicited by ligands which coengage BCR and the CD21/CD19/CD81 costimulatory complex." Cell Immunol 207(2): 127-40.

Mongini, P. K., et al. (2002). "Antigen receptor triggered upregulation of CD86 and CD80 in human B cells: augmenting role of the CD21/CD19 co-stimulatory complex and IL-4." Cell Immunol 216(1-2): 50-64.

Morales, J. C., et al. (2006). "53BP1 and p53 synergize to suppress genomic instability and lymphomagenesis." Proc Natl Acad Sci U S A 103(9): 3310-5.

Mosbech, A., et al. (2013). "The deubiquitylating enzyme USP44 counteracts the DNA double-strand break response mediated by the RNF8 and RNF168 ubiquitin ligases." J Biol Chem 288(23): 16579-87.

Moyal, L., et al. (2011). "Requirement of ATM-dependent monoubiquitylation of histone H2B for timely repair of DNA double-strand breaks." Mol Cell 41(5): 529-42.

Muramatsu, M., et al. (2000). "Class switch recombination and hypermutation require activation-induced cytidine deaminase (AID), a potential RNA editing enzyme." Cell 102(5): 553-63.

Muramatsu, M., et al. (1999). "Specific expression of activation-induced cytidine deaminase (AID), a novel member of the RNA-editing deaminase family in germinal center B cells." J Biol Chem 274(26): 18470-6.

Nakada, S., et al. (2010). "Non-canonical inhibition of DNA damage-dependent ubiquitination by OTUB1." Nature 466(7309): 941-6.

Nakamura, K., et al. (2011). "Regulation of homologous recombination by RNF20-dependent H2B ubiquitination." Mol Cell 41(5): 515-28.

Nakamura, K., et al. (2006). "Genetic dissection of vertebrate 53BP1: a major role in non-homologous end joining of DNA double strand breaks." DNA Repair (Amst) 5(6): 741-9.

Nakamura, M., et al. (1996). "High frequency class switching of an IgM+ B lymphoma clone CH12F3 to IgA+ cells." Int Immunol 8(2): 193-201.

Page 153: Characterizing the role of the DNA Damage Response in

139

Ng, H. H., et al. (2002). "Ubiquitination of histone H2B by Rad6 is required for efficient Dot1-mediated methylation of histone H3 lysine 79." J Biol Chem 277(38): 34655-7.

Nicassio, F., et al. (2007). "Human USP3 is a chromatin modifier required for S phase progression and genome stability." Curr Biol 17(22): 1972-7.

Noon, A. T., et al. (2010). "53BP1-dependent robust localized KAP-1 phosphorylation is essential for heterochromatic DNA double-strand break repair." Nat Cell Biol 12(2): 177-84.

Oettinger, M. A., et al. (1990). "RAG-1 and RAG-2, adjacent genes that synergistically activate V(D)J recombination." Science 248(4962): 1517-23.

Ogiwara, H., et al. (2011). "Essential factors for incompatible DNA end joining at chromosomal DNA double strand breaks in vivo." PLoS One 6(12): e28756.

Okazaki, I. M., et al. (2003). "Constitutive expression of AID leads to tumorigenesis." J Exp Med 197(9): 1173-81.

Okazaki, I. M., et al. (2007). "Role of AID in tumorigenesis." Adv Immunol 94: 245-73. Orthwein, A., et al. (2010). "Regulation of activation-induced deaminase stability and antibody gene

diversification by Hsp90." J Exp Med 207(12): 2751-65. Pan-Hammarstrom, Q., et al. (2005). "Impact of DNA ligase IV on nonhomologous end joining pathways

during class switch recombination in human cells." J Exp Med 201(2): 189-94. Panier, S., et al. (2009). "Regulatory ubiquitylation in response to DNA double-strand breaks." DNA

Repair (Amst) 8(4): 436-43. Parsa, J. Y., et al. (2007). "AID mutates a non-immunoglobulin transgene independent of chromosomal

position." Mol Immunol 44(4): 567-75. Parsa, J. Y., et al. (2012). "Negative supercoiling creates single-stranded patches of DNA that are

substrates for AID-mediated mutagenesis." PLoS Genet 8(2): e1002518. Pavri, R., et al. (2010). "Activation-induced cytidine deaminase targets DNA at sites of RNA polymerase II

stalling by interaction with Spt5." Cell 143(1): 122-33. Pavri, R., et al. (2011). "AID targeting in antibody diversity." Adv Immunol 110: 1-26. Pei, H., et al. "MMSET regulates histone H4K20 methylation and 53BP1 accumulation at DNA damage

sites." Nature 470(7332): 124-8. Peled, J. U., et al. (2008). "The biochemistry of somatic hypermutation." Annu Rev Immunol 26: 481-511. Pellegrini, M., et al. (2006). "Autophosphorylation at serine 1987 is dispensable for murine Atm

activation in vivo." Nature 443(7108): 222-5. Peng, A., et al. (2003). "NFBD1, like 53BP1, is an early and redundant transducer mediating Chk2

phosphorylation in response to DNA damage." J Biol Chem 278(11): 8873-6. Perez-Duran, P., et al. (2007). "Oncogenic events triggered by AID, the adverse effect of antibody

diversification." Carcinogenesis 28(12): 2427-33. Pergola, F., et al. (1993). "V(D)J recombination in mammalian cell mutants defective in DNA double-

strand break repair." Mol Cell Biol 13(6): 3464-71. Petersen-Mahrt, S. K., et al. (2002). "AID mutates E. coli suggesting a DNA deamination mechanism for

antibody diversification." Nature 418(6893): 99-104. Petersen, S., et al. (2001). "AID is required to initiate Nbs1/gamma-H2AX focus formation and mutations

at sites of class switching." Nature 414(6864): 660-5. Pham, P., et al. (2003). "Processive AID-catalysed cytosine deamination on single-stranded DNA

simulates somatic hypermutation." Nature 424(6944): 103-7. Pinato, S., et al. (2009). "RNF168, a new RING finger, MIU-containing protein that modifies chromatin by

ubiquitination of histones H2A and H2AX." BMC Mol Biol 10: 55. Rada, C., et al. (2001). "The intrinsic hypermutability of antibody heavy and light chain genes decays

exponentially." Embo J 20(16): 4570-6.

Page 154: Characterizing the role of the DNA Damage Response in

140

Rada, C., et al. (2002). "Immunoglobulin Isotype Switching Is Inhibited and Somatic Hypermutation Perturbed in UNG-Deficient Mice." Curr Biol 12(20): 1748-55.

Raghavan, M., et al. (1996). "Fc receptors and their interactions with immunoglobulins." Annu Rev Cell Dev Biol 12: 181-220.

Rahal, E. A., et al. (2010). "ATM regulates Mre11-dependent DNA end-degradation and microhomology-mediated end joining." Cell Cycle 9(14): 2866-77.

Rajagopal, D., et al. (2009). "Immunoglobulin switch mu sequence causes RNA polymerase II accumulation and reduces dA hypermutation." J Exp Med 206(6): 1237-44.

Ramachandran, S., et al. (2010). "The RNF8/RNF168 ubiquitin ligase cascade facilitates class switch recombination." Proc Natl Acad Sci U S A 107(2): 809-14.

Ramiro, A., et al. (2007). "The role of activation-induced deaminase in antibody diversification and chromosome translocations." Adv Immunol 94: 75-107.

Ramiro, A. R., et al. (2003). "Transcription enhances AID-mediated cytidine deamination by exposing single-stranded DNA on the nontemplate strand." Nat Immunol 4(5): 452-6.

Rappold, I., et al. (2001). "Tumor suppressor p53 binding protein 1 (53BP1) is involved in DNA damage-signaling pathways." J Cell Biol 153(3): 613-20.

Reina-San-Martin, B., et al. (2004). "ATM is required for efficient recombination between immunoglobulin switch regions." J Exp Med 200(9): 1103-10.

Reina-San-Martin, B., et al. (2007). "Enhanced intra-switch region recombination during immunoglobulin class switch recombination in 53BP1-/- B cells." Eur J Immunol 37(1): 235-9.

Reina-San-Martin, B., et al. (2003). "H2AX is required for recombination between immunoglobulin switch regions but not for intra-switch region recombination or somatic hypermutation." J Exp Med 197(12): 1767-78.

Reina-San-Martin, B., et al. (2005). "Genomic instability, endoreduplication, and diminished Ig class-switch recombination in B cells lacking Nbs1." Proc Natl Acad Sci U S A 102(5): 1590-5.

Reinhardt, H. C., et al. (2009). "Kinases that control the cell cycle in response to DNA damage: Chk1, Chk2, and MK2." Curr Opin Cell Biol 21(2): 245-55.

Revy, P., et al. (2000). "Activation-induced cytidine deaminase (AID) deficiency causes the autosomal recessive form of the Hyper-IgM syndrome (HIGM2)." Cell 102(5): 565-75.

Roa, S., et al. (2008). "Ubiquitylated PCNA plays a role in somatic hypermutation and class-switch recombination and is required for meiotic progression." Proc Natl Acad Sci U S A 105(42): 16248-53.

Roa, S., et al. (2010). "MSH2/MSH6 complex promotes error-free repair of AID-induced dU:G mispairs as well as error-prone hypermutation of A:T sites." PLoS One 5(6): e11182.

Robbiani, D. F., et al. (2013). "Chromosome translocation, B cell lymphoma, and activation-induced cytidine deaminase." Annu Rev Pathol 8: 79-103.

Robert, I., et al. (2009). "Parp1 facilitates alternative NHEJ, whereas Parp2 suppresses IgH/c-myc translocations during immunoglobulin class switch recombination." J Exp Med 206(5): 1047-56.

Roberts, S. A., et al. (2010). "Ku is a 5'-dRP/AP lyase that excises nucleotide damage near broken ends." Nature 464(7292): 1214-7.

Rogakou, E. P., et al. (1998). "DNA double-stranded breaks induce histone H2AX phosphorylation on serine 139." J Biol Chem 273(10): 5858-68.

Rooney, S., et al. (2003). "Defective DNA repair and increased genomic instability in Artemis-deficient murine cells." J Exp Med 197(5): 553-65.

Rouaud, P., et al. (2013). "The IgH 3' regulatory region controls somatic hypermutation in germinal center B cells." J Exp Med 210(8): 1501-7.

Ruprecht, C. R., et al. (2006). "Toll-like receptor stimulation as a third signal required for activation of human naive B cells." Eur J Immunol 36(4): 810-6.

Page 155: Characterizing the role of the DNA Damage Response in

141

Sancar, A., et al. (2004). "Molecular mechanisms of mammalian DNA repair and the DNA damage checkpoints." Annu Rev Biochem 73: 39-85.

Santos, M. A., et al. (2010). "Class switching and meiotic defects in mice lacking the E3 ubiquitin ligase RNF8." J Exp Med 207(5): 973-81.

Saribasak, H., et al. (2011). "XRCC1 suppresses somatic hypermutation and promotes alternative nonhomologous end joining in Igh genes." J Exp Med 208(11): 2209-16.

Saribasak, H., et al. (2009). "Hijacked DNA repair proteins and unchained DNA polymerases." Philos Trans R Soc Lond B Biol Sci 364(1517): 605-11.

Sato, S., et al. (1997). "Regulation of B lymphocyte development and activation by the CD19/CD21/CD81/Leu 13 complex requires the cytoplasmic domain of CD19." J Immunol 159(7): 3278-87.

Savitsky, K., et al. (1995). "A single ataxia telangiectasia gene with a product similar to PI-3 kinase." Science 268(5218): 1749-53.

Schatz, D. G., et al. (2011). "Recombination centres and the orchestration of V(D)J recombination." Nat Rev Immunol 11(4): 251-63.

Schatz, D. G., et al. (1989). "The V(D)J recombination activating gene, RAG-1." Cell 59(6): 1035-48. Schrader, C. E., et al. (2007). "Activation-induced cytidine deaminase-dependent DNA breaks in class

switch recombination occur during G1 phase of the cell cycle and depend upon mismatch repair." J Immunol 179(9): 6064-71.

Schrader, C. E., et al. (2009). "The roles of APE1, APE2, DNA polymerase beta and mismatch repair in creating S region DNA breaks during antibody class switch." Philos Trans R Soc Lond B Biol Sci 364(1517): 645-52.

Schrader, C. E., et al. (2005). "Inducible DNA breaks in Ig S regions are dependent on AID and UNG." J Exp Med 202(4): 561-8.

Schultz, L. B., et al. (2000). "p53 binding protein 1 (53BP1) is an early participant in the cellular response to DNA double-strand breaks." J Cell Biol 151(7): 1381-90.

Schwab, K. R., et al. (2011). "Role of PTIP in class switch recombination and long-range chromatin interactions at the immunoglobulin heavy chain locus." Mol Cell Biol 31(7): 1503-11.

Shanbhag, N. M., et al. (2010). "ATM-dependent chromatin changes silence transcription in cis to DNA double-strand breaks." Cell 141(6): 970-81.

Shiloh, Y., et al. (2011). "RNF20-RNF40: A ubiquitin-driven link between gene expression and the DNA damage response." FEBS Lett 585(18): 2795-802.

Shiloh, Y., et al. (2013). "The ATM protein kinase: regulating the cellular response to genotoxic stress, and more." Nat Rev Mol Cell Biol 14(4): 197-210.

Shinkura, R., et al. (2004). "Separate domains of AID are required for somatic hypermutation and class-switch recombination." Nat Immunol 5(7): 707-12.

Siders, W. M., et al. (2010). "Involvement of neutrophils and natural killer cells in the anti-tumor activity of alemtuzumab in xenograft tumor models." Leuk Lymphoma 51(7): 1293-304.

Silverman, J., et al. (2004). "Human Rif1, ortholog of a yeast telomeric protein, is regulated by ATM and 53BP1 and functions in the S-phase checkpoint." Genes Dev 18(17): 2108-19.

Smider, V., et al. (1998). "Failure of hairpin-ended and nicked DNA To activate DNA-dependent protein kinase: implications for V(D)J recombination." Mol Cell Biol 18(11): 6853-8.

Smith, T. J., et al. (1996). "Neutralizing antibody to human rhinovirus 14 penetrates the receptor-binding canyon." Nature 383(6598): 350-4.

So, S., et al. (2009). "Autophosphorylation at serine 1981 stabilizes ATM at DNA damage sites." J Cell Biol 187(7): 977-90.

Soulas-Sprauel, P., et al. (2007). "Role for DNA repair factor XRCC4 in immunoglobulin class switch recombination." J Exp Med 204(7): 1717-27.

Page 156: Characterizing the role of the DNA Damage Response in

142

Soulas-Sprauel, P., et al. (2007). "V(D)J and immunoglobulin class switch recombinations: a paradigm to study the regulation of DNA end-joining." Oncogene 26(56): 7780-91.

Sproul, T. W., et al. (2000). "A role for MHC class II antigen processing in B cell development." Int Rev Immunol 19(2-3): 139-55.

Stanlie, A., et al. (2010). "Histone3 lysine4 trimethylation regulated by the facilitates chromatin transcription complex is critical for DNA cleavage in class switch recombination." Proc Natl Acad Sci U S A 107(51): 22190-5.

Staszewski, O., et al. (2011). "Activation-induced cytidine deaminase induces reproducible DNA breaks at many non-Ig Loci in activated B cells." Mol Cell 41(2): 232-42.

Stavnezer, J. (2011). "Complex regulation and function of activation-induced cytidine deaminase." Trends Immunol 32(5): 194-201.

Stavnezer, J., et al. (2010). "Mapping of switch recombination junctions, a tool for studying DNA repair pathways during immunoglobulin class switching." Adv Immunol 108: 45-109.

Stavnezer, J., et al. (2008). "Mechanism and regulation of class switch recombination." Annu Rev Immunol 26: 261-92.

Stavnezer, J., et al. (2005). "Mismatch repair converts AID-instigated nicks to double-strand breaks for antibody class-switch recombination." Trends Genet.

Stewart, G. S. (2009). "Solving the RIDDLE of 53BP1 recruitment to sites of damage." Cell Cycle 8(10). Stewart, G. S., et al. (1999). "The DNA double-strand break repair gene hMRE11 is mutated in individuals

with an ataxia-telangiectasia-like disorder." Cell 99(6): 577-87. Stewart, G. S., et al. (2009). "The RIDDLE syndrome protein mediates a ubiquitin-dependent signaling

cascade at sites of DNA damage." Cell 136(3): 420-34. Stewart, G. S., et al. (2007). "RIDDLE immunodeficiency syndrome is linked to defects in 53BP1-mediated

DNA damage signaling." Proc Natl Acad Sci U S A 104(43): 16910-5. Stewart, G. S., et al. (2003). "MDC1 is a mediator of the mammalian DNA damage checkpoint." Nature

421(6926): 961-6. Stiff, T., et al. (2004). "ATM and DNA-PK function redundantly to phosphorylate H2AX after exposure to

ionizing radiation." Cancer Res 64(7): 2390-6. Stucki, M., et al. (2005). "MDC1 directly binds phosphorylated histone H2AX to regulate cellular

responses to DNA double-strand breaks." Cell 123(7): 1213-26. Sulica, A., et al. (2001). "Ig-binding receptors on human NK cells as effector and regulatory surface

molecules." Int Rev Immunol 20(3-4): 371-414. Sun, Z. W., et al. (2002). "Ubiquitination of histone H2B regulates H3 methylation and gene silencing in

yeast." Nature 418(6893): 104-8. Ta, V. T., et al. (2003). "AID mutant analyses indicate requirement for class-switch-specific cofactors."

Nat Immunol 4(9): 843-8. Taccioli, G. E., et al. (1993). "Impairment of V(D)J recombination in double-strand break repair mutants."

Science 260(5105): 207-10. Taddie, J. A., et al. (1994). "Activation of B- and T-cells by the cytoplasmic domains of the B-cell antigen

receptor proteins Ig-alpha and Ig-beta." J Biol Chem 269(18): 13529-35. Taillardet, M., et al. (2009). "The thymus-independent immunity conferred by a pneumococcal

polysaccharide is mediated by long-lived plasma cells." Blood 114(20): 4432-40. Tang, E. S., et al. (2006). "NHEJ-deficient DT40 cells have increased levels of immunoglobulin gene

conversion: evidence for a double strand break intermediate." Nucleic Acids Res 34(21): 6345-51.

Tang, E. S., et al. (2007). "Immunoglobulin gene conversion: synthesizing antibody diversification and DNA repair." DNA Repair (Amst) 6(11): 1557-71.

Taylor, A. M., et al. (1996). "Leukemia and lymphoma in ataxia telangiectasia." Blood 87(2): 423-38.

Page 157: Characterizing the role of the DNA Damage Response in

143

Teng, G., et al. (2008). "MicroRNA-155 is a negative regulator of activation-induced cytidine deaminase." Immunity 28(5): 621-9.

Truong, L. N., et al. (2013). "Microhomology-mediated End Joining and Homologous Recombination share the initial end resection step to repair DNA double-strand breaks in mammalian cells." Proc Natl Acad Sci U S A 110(19): 7720-5.

Uematsu, N., et al. (2007). "Autophosphorylation of DNA-PKCS regulates its dynamics at DNA double-strand breaks." J Cell Biol 177(2): 219-29.

Uziel, T., et al. (2003). "Requirement of the MRN complex for ATM activation by DNA damage." Embo J 22(20): 5612-21.

Varon, R., et al. (1998). "Nibrin, a novel DNA double-strand break repair protein, is mutated in Nijmegen breakage syndrome." Cell 93(3): 467-76.

Victora, G. D., et al. (2012). "Germinal centers." Annu Rev Immunol 30: 429-57. Vos, Q., et al. (2000). "B-cell activation by T-cell-independent type 2 antigens as an integral part of the

humoral immune response to pathogenic microorganisms." Immunol Rev 176: 154-70. Vuong, B. Q., et al. (2009). "Specific recruitment of protein kinase A to the immunoglobulin locus

regulates class-switch recombination." Nat Immunol 10(4): 420-6. Vyas, R., et al. (2012). "RNF4 is required for DNA double-strand break repair in vivo." Cell Death Differ

20(3): 490-502. Wagner, S. D., et al. (1996). "Somatic hypermutation of immunoglobulin genes." Annu Rev Immunol 14:

441-57. Wang, B., et al. (2002). "53BP1, a mediator of the DNA damage checkpoint." Science 298(5597): 1435-8. Wang, C. L., et al. (2004). "Genome-wide somatic hypermutation." Proc Natl Acad Sci U S A 101(19):

7352-6. Ward, I. M., et al. (2003). "p53 Binding protein 53BP1 is required for DNA damage responses and tumor

suppression in mice." Mol Cell Biol 23(7): 2556-63. Ward, I. M., et al. (2004). "53BP1 is required for class switch recombination." J Cell Biol 165(4): 459-64. Weake, V. M., et al. (2008). "Histone ubiquitination: triggering gene activity." Mol Cell 29(6): 653-63. Wiener, R., et al. (2012). "The mechanism of OTUB1-mediated inhibition of ubiquitination." Nature

483(7391): 618-22. Williams, R. S., et al. (2007). "Mre11-Rad50-Nbs1 is a keystone complex connecting DNA repair

machinery, double-strand break signaling, and the chromatin template." Biochem Cell Biol 85(4): 509-20.

Wilson, K. A., et al. (2008). "NFBD1/MDC1, 53BP1 and BRCA1 have both redundant and unique roles in the ATM pathway." Cell Cycle 7(22): 3584-94.

Wilson, T. M., et al. (2005). "MSH2-MSH6 stimulates DNA polymerase eta, suggesting a role for A:T mutations in antibody genes." J Exp Med 201(4): 637-45.

Wu, J., et al. (2009). "PTIP regulates 53BP1 and SMC1 at the DNA damage sites." J Biol Chem 284(27): 18078-84.

Xiao, Y., et al. (1997). "Conditional gene targeted deletion by Cre recombinase demonstrates the requirement for the double-strand break repair Mre11 protein in murine embryonic stem cells." Nucleic Acids Res 25(15): 2985-91.

Xiao, Z., et al. (2007). "Known components of the immunoglobulin A:T mutational machinery are intact in Burkitt lymphoma cell lines with G:C bias." Mol Immunol 44(10): 2659-66.

Xie, G., et al. (1998). "Requirements for p53 and the ATM gene product in the regulation of G1/S and S phase checkpoints." Oncogene 16(6): 721-36.

Xu, Y., et al. (1996). "Targeted disruption of ATM leads to growth retardation, chromosomal fragmentation during meiosis, immune defects, and thymic lymphoma." Genes Dev 10(19): 2411-22.

Page 158: Characterizing the role of the DNA Damage Response in

144

Xu, Z., et al. (2007). "Regulation of aicda expression and AID activity: relevance to somatic hypermutation and class switch DNA recombination." Crit Rev Immunol 27(4): 367-97.

Xue, K., et al. (2006). "The in vivo pattern of AID targeting to immunoglobulin switch regions deduced from mutation spectra in msh2-/- ung-/- mice." J Exp Med 203(9): 2085-94.

Yan, C. T., et al. (2007). "IgH class switching and translocations use a robust non-classical end-joining pathway." Nature 449(7161): 478-82.

Yang, J., et al. (2003). "ATM, ATR and DNA-PK: initiators of the cellular genotoxic stress responses." Carcinogenesis 24(10): 1571-80.

Yin, B., et al. (2009). "Histone H2AX stabilizes broken DNA strands to suppress chromosome breaks and translocations during V(D)J recombination." J Exp Med 206(12): 2625-39.

Yin, Y., et al. (2012). "SUMO-targeted ubiquitin E3 ligase RNF4 is required for the response of human cells to DNA damage." Genes Dev 26(11): 1196-208.

You, Z., et al. (2005). "ATM activation and its recruitment to damaged DNA require binding to the C terminus of Nbs1." Mol Cell Biol 25(13): 5363-79.

Zaheen, A., et al. (2009). "AID constrains germinal center size by rendering B cells susceptible to apoptosis." Blood 114(3): 547-54.

Zeng, X., et al. (2001). "DNA polymerase eta is an A-T mutator in somatic hypermutation of immunoglobulin variable genes." Nat Immunol 2(6): 537-41.

Zha, S., et al. (2011). "Ataxia telangiectasia-mutated protein and DNA-dependent protein kinase have complementary V(D)J recombination functions." Proc Natl Acad Sci U S A 108(5): 2028-33.

Zhang, Y. (2003). "Transcriptional regulation by histone ubiquitination and deubiquitination." Genes Dev 17(22): 2733-40.

Zhang, Y., et al. (2011). "An essential role for CtIP in chromosomal translocation formation through an alternative end-joining pathway." Nat Struct Mol Biol 18(1): 80-4.

Zhao, Y., et al. (2008). "A TFTC/STAGA module mediates histone H2A and H2B deubiquitination, coactivates nuclear receptors, and counteracts heterochromatin silencing." Mol Cell 29(1): 92-101.

Zhong, Q., et al. (2002). "BRCA1 facilitates microhomology-mediated end joining of DNA double strand breaks." J Biol Chem 277(32): 28641-7.

Zhu, B., et al. (2005). "Monoubiquitination of human histone H2B: the factors involved and their roles in HOX gene regulation." Mol Cell 20(4): 601-11.

Zhu, J., et al. (2001). "Targeted disruption of the Nijmegen breakage syndrome gene NBS1 leads to early embryonic lethality in mice." Curr Biol 11(2): 105-9.

Zimmermann, M., et al. (2013). "53BP1 regulates DSB repair using Rif1 to control 5' end resection." Science 339(6120): 700-4.

Ziv, Y., et al. (1997). "Recombinant ATM protein complements the cellular A-T phenotype." Oncogene 15(2): 159-67.

Ziv, Y., et al. (2006). "Chromatin relaxation in response to DNA double-strand breaks is modulated by a novel ATM- and KAP-1 dependent pathway." Nat Cell Biol 8(8): 870-6.