ABSTRACT

Bellingham-Johnstun, Kimberly Suzanne. Knockdown of the Upstream Stimulatory Factors 1 and 2 Lead to Long-term Changes in the Expression Response to DNA Damage. (Under the direction of Dr. Michael Sikes).

Despite dramatic advances in treatment, cancer remains one of the world’s most significant public health challenges. In the United States alone, one in three women and one in two men will develop some form of cancer. Although the types of cancers and their root causes are varied, all cancers develop from the accumulation of critical genetic mutations and the loss of genomic stability. My research interests have focused on the molecular programs that cells employ to preserve genomic stability in the face of environmentally-induced DNA damage. Upon genotoxic insult, cells rapidly activate -dependent programs that pause the cell cycle and direct either DNA repair or apoptosis. Beyond P53, the long-term transcriptional programs that direct either cell recovery or senescence, depending on the success of DNA repair, are unclear. In this study, we used RNAi to investigate the potential of Upstream Stimulatory Factor 1 (USF1) and Upstream Stimulatory Factor 2 (USF2) stress-response transcription factors as directors of long-term DNA damage responses in the P53-deficient mouse B lymphocyte cell line, M12.

Microarray analysis revealed 765 differentially expressed (≥1.50-fold change, n=3) in

USF-depleted cells when compared with cells expressing a scrambled shRNA. In contrast, 7 days after cells were exposed to ionizing radiation (IR), 2866 genes now showed altered expression in the USF-depleted cells. Microarray findings were recapitulated in separate biological replicates, and further confirmed for a panel of constitutively altered and IR-induced genes by quantitative reverse transcription PCR (RT-qPCR). USF-depletion led to the up- regulation of a number of genes critical for immune function and strongly linked to cancer development, including genes for the DNA mutator Activation Induced Cytidine Deaminase (AID), the CD300a lipid , the Ig kinases B Lymphocyte Kinase (BLK) and B Cell Linker

(BLNK), and multiple members of the Nuclear Factor Kappa-Light-Chain Enhancer (NFB) family including Transcription Factor p65 (REL-A), REL-B, c-REL, p50,

IB and IB. The dramatic up-regulation of the genes for AID (Aicda, 7.99 ± 1.11), REL-A

(Rela 15.53 ± 0.39), p50 (Nfkb1 9.38 ± 3.07), and IB (Nfkbia 11.87 ± 0.87) were particularly surprising as neither USF has previously been implicated in the regulation of these well- studied genes that are central to lymphocyte function. Usf1-/- thymic RNA and RNA harvested from M12 cells selectively depleted of either USF1 or USF2 exhibited significantly lower levels of NFB and Aicda expression than observed in M12 cells depleted of both USF1 and USF2.

Despite increased Aicda transcription, sequencing analysis of the functional Ig Vba9*/J1 rearrangement in 66 independent clones revealed no somatic hypermutation in long-term cultures of USF-depleted M12 cells, supporting previous studies that somatic hypermutation in germinal center B cells is a multifaceted response to antigenic and cytokine stimulation, and involves increased expression, activation and nuclear translocation of the AID protein. Taken together, the findings of my Master’s research provide the first evidence that USF1 and USF2 collaborate to play a novel role in regulating long-term response to DNA damage. These findings also illustrate the significance of investigating USF’s regulatory role using models that overcome the functional overlap shared by USF1 and USF2. Finally, by using our USF RNAi model cell line, I have provided the first evidence that loss of USF1 and USF2 tumor suppressor results in the overexpression of NFB and AID, suggesting a significant role for USF in regulating B cell activation and tumorigenesis.

© Copyright 2017 by Kimberly Suzanne Bellingham-Johnstun

All Rights Reserved Knockdown of the Upstream Stimulatory Factors 1 and 2 Lead to Long-term Changes in the Gene Expression Response to DNA Damage

by Kimberly Suzanne Bellingham-Johnstun

A thesis submitted to the Graduate Faculty of North Carolina State University in partial fulfillment of the requirements for the degree of Master of Science

Microbiology

Raleigh, North Carolina

2017

APPROVED BY:

______Dr. Michael L. Sikes Dr. Scott M. Laster Committee Chair

______Dr. Frank Scholle Dr. Jeffrey Yoder ii

DEDICATION

This thesis is dedicated to my husband, Joel, whose unending support made this possible, and to my mother and father, who gave me the necessary foundation and love to reach my goals.

iii

BIOGRAPHY

Kimberly Suzanne Bellingham-Johnstun was born to Debra Christine Bellingham and

Robert Vincent Bellingham in Burbank, California on July 15, 1992. Kimberly graduated from McMinnville High School in McMinnville, Oregon and attended Oregon State

University in Corvallis, Oregon, where she graduated magna cum laude with a Bachelor of

Science in Microbiology. Kimberly came to North Carolina State University to pursue her

Master of Science degree in Microbiology in 2014.

iv

ACKNOWLEDGMENTS

I would like to thank my friends and family that have supported me through this period in my life, especially my husband, Joel, who traveled this rocky road with me more closely than anyone else. I would also like to thank my mother and father, whose encouragement through all my schooling led me to greater heights. I would like to thank my lab-mates, both past and present, especially Jenn Stone and John King for not only helping me with the project, but for being great coworkers and friends. Of course, I cannot over-emphasize how much is owed to

Dr. Michael Sikes. He was an outstanding mentor, both forgiving of mistakes and encouraging quality work. He really encouraged me to become a well-rounded scientist, and

I am grateful that I got the opportunity to work in his lab.

v

TABLE OF CONTENTS

LIST OF TABLES ...... vii LIST OF FIGURES ...... viii CHAPTER 1: Literature Review ...... 1 1.1: DNA stress during B cell development ...... 1 1.2: Upstream Stimulatory Factors ...... 3 1.3: USF as a tumor suppressor ...... 6 1.4 USF as a stress response factor ...... 9 1.5 USF regulation by phosphorylation ...... 10 1.6: Activation-induced cytidine deaminase ...... 12 1.7: Regulation of AID ...... 14 1.8: Pathologies associated with AID ...... 17 1.9: Introduction and Objectives of the Thesis Project ...... 19 CHAPTER 2: Materials & Methods ...... 24 2.1 Cell culture ...... 24 2.2 RNA interference and transfection ...... 24 2.3 Western blotting ...... 26 2.4 Ionizing radiation treatment ...... 27 2.5 RNA extraction ...... 28 2.6 Affymetrix Microarray Hybridization ...... 28 2.7 Bioinformatic Analysis of Gene Expression ...... 29 2.8 Quantitative and reverse transcriptase PCR (qPCR and RT-PCR) ...... 29 CHAPTER 3: Results ...... 32 3.1: USF depletion ...... 32 3.2: Transcriptomic analysis of USF depletion...... 34 3.3: clustering analysis ...... 43 3.4: RT-QPCR validation of transcriptomic data ...... 46 3.5: Aicda activity in USF-depleted cells ...... 51 CHAPTER 4: Discussion ...... 53 REFERENCES ...... 62 APPENDICES ...... 82

vi

Appendix A ...... 83

vii

LIST OF TABLES

Chapter 3: Results

Table 1. List of Oligonucleotides…………………………………………………………….31

Table 2: Selected GeneChip mouse whole transcript microarray determinations of fold increase in gene expression in untreated USFKD M12 cells………………………………...39

Table 3: Selected GeneChip mouse whole transcript microarray determinations of fold decrease in gene expression in untreated USFKD M12 cells………………………………..40

Table 4: Selected GeneChip mouse whole transcript microarray determinations of fold change in gene expression in USFKD M12 cells 7 days post-IR……………………………41

Table 5: Selected GeneChip mouse whole transcript microarray determinations of fold change in gene expression in sshRNA M12 cells 7 days post-IR……………………………42

Appendix A

Table 1: Functional clustering of GO terms………………………………………………...83

viii

LIST OF FIGURES

Figure 1. Functional domains and kinase sites of mouse USF1 and USF2…………………...5

Figure 2. Schematic of AID regulation………………………………………………………16

Figure 3. USF Knockdown in M12 cells…………………………………………………….33

Figure 4. Differential gene expression analysis of USF-depleted and irradiated M12 cells...35

Figure 5. Gene expression response to USF depletion and IR………………………………36

Figure 6. Heat map of the clustering among the enriched Gene Ontology biological process terms annotated to the DEGs……………………………………...... 44

Figure 7. Loading control validation in USF-depleted and irradiated cells………………….48

Figure 8. Quantitative PCR (qPCR) validation of microarray findings……………………...50

Figure 9. Representative sequence alignment of Vk region………………………………….52

Figure 10. Schematic depiction of the mouse Aicda gene…………………………………...59

1

CHAPTER 1: Literature Review

1.1: DNA stress during B cell development Humoral immunity is mediated by antibodies, the secreted form of the B cell antigen receptor or Immunoglobulin (Ig). Antibodies are multi-subunit protein complexes composed of two heavy (IgH) and two light (IgL) chains. Each Ig contains repeated domains (so-called

Ig domains) of anti-parallel beta sheets linked by alpha helical loops. The enormous diversity of antigens that can be recognized by antibodies is generated early in B cell development through a program of somatic DNA rearrangements within the 5’ portion of each Ig gene.

These rearrangements assemble the exon that encodes each Ig’s antigen binding domain from pools of subexonic variable (V), diversity (D) and joining (J) gene segments, and produce the capacity to recognize as many as 5x1013 different antigenic molecules1.

V(D)J recombination is initiated by the recombination activating gene 1 and 2 (or

RAG1 and RAG2) , which are only expressed in developing lymphocytes2. The RAG proteins drive a stepwise assembly of the antigen binding exon, with IgH genes assembled by

D-to-J, and then V-to-DJ joining, followed by IgL V-to-J joining (IgL genes lack D gene segments). For each pair of compatible gene segments, recombination begins when RAG1/2 complexes bind conserved recombination signal sequences (RSS) that flank each V, D and J coding segment, and nick the DNA at the RSS:coding junction, catalyzing a transesterification reaction that breaks the DNA into a hairpinned coding end and blunt signal end3. The double-stranded breaks (DSBs) caused by RAG1/2 are repaired through the low- fidelity process of nonhomologous end joining (NHEJ) before the cell is allowed to proceed to the next stage of development. NHEJ relies on Nibrin (NBS1), ATM Serine Threonine

2

Kinase (ATM), and Tumor Suppressor p53-binding Protein 1 (53BP1) to aggregate in the chromatin area surrounding DSB and create a “DNA damage repair focus”4. KU70/80 heterodimers, DNA-dependent Protein Kinase (DNA-PKcs), ARTEMIS, XRCC4-like Factor

(XLF), and DNA Ligase 4 (LIG4) accumulate at this focus, stabilizing free DNA ends, recruiting additional DNA damage response (DDR) factors, and ligating the two blunt signal ends together (most typically as an extrachromosomal signal joint). Restoration of the by ligating the compatible coding junctions requires additional processing that dramatically increases antigen binding diversity5. This “junctional diversity” is generated when RAG proteins at the coding ends, together with and DNA-PKcs again nick the DNA, opening the hairpinned coding ends, which may then be subjected to limited exonuclease activity, and/or templated and non-templated polymerization5. Deficiency of either NBS1, y-

H2AX, ATM, or 53BP1 leads to abnormalities in antigen receptor V(D)J recombination, while loss of KU80 and LIG4 causes a dramatic increase in DNA breaks and chromosomal translocations3,6,7.

After V(D)J recombination, B cells that complete development in the bone marrow are released into circulation as mature naïve B cells that express surface-bound Ig as their B cell receptor (BCR). Should the B cell bind its cognate antigen and become activated while trafficking through secondary lymphoid tissue, the Ig genes are again targeted for diversification through the processes of somatic hypermutation (SHM) and class-switch recombination (CSR)8. During SHM, point mutations are introduced in the IgH and IgL

V(D)J exons9 at a frequency of about one mutation per kilobase (kb) per cell per

3

generation10. Mutations are introduced by Activation-Induced Cytidine Deaminase (or AID), resulting in a C-to-T substitution. In mammals, CSR produces antibodies with different effector activities through DNA rearrangements that replace the 5’-most constant region exon of the IgH gene, Cμ and C, with either ,  or  exons 11,12. Since the effector activities and tissue localization activities of antibodies are encoded by the constant region exons, CSR alters antibody functionality without altering its antigen specificity.9. In CSR, DSBs occur between palindromic switch (S) regions after a S-S synaptosome is formed. The S-S synaptosome is created by AID driving C-to-A substitutions in the donor Sμ and an acceptor

S, S or S region13. The combined DNA targeting activities of V(D)J recombination, somatic hypermutation and class-switch recombination are absolutely essential for humoral immunity, and loss of either results in severe compromise of humoral immune efficacy. Each however also offers the potential for unintended harm to the host through tumorigenic mutations. Indeed, the vast majority of B cell lymphomas are linked to either AID- or RAG- mediated translocations.

1.2: Upstream Stimulatory Factors Upstream Stimulatory Factor (USF) was first identified by the Sharp laboratory in

1985 as Major Late Transcription Factor (MLTF), a single 46 KDa polypeptide responsible for RNA Polymerase II activation of the adenovirus major late promoter14,15 16. In a later paper from the Sawadago lab, it was discovered that two paralogs of the protein, renamed

USF existed. Both USF1 and USF2, consist of 10 exons with basic-Helix-Loop-Helix and

Leucine Zipper domains (Figure 1)17,18. Alternative splicing results in two USF2 isoforms,

4

USF2a and USF2b, with USF2b having the potential to act as a dominant negative regulator of USF1/219, though the significance and tissue distribution of USF2b are unclear.

Both USF1 and USF2 are ubiquitously expressed, though concentrations vary between tissues20. USF1 and USF2 act as transcription factors by binding to canonical

CANNTG targets shared with other so-called E proteins (e.g. E47, E2A, , MAX; all sharing a conserved bHLH “E protein” motif). As with other E proteins, USF binds DNA as a dimer, preferentially formed between USF1 and USF2, though both USF1 and USF2 homodimers are also active19. CANNTG E boxes are found in a wide array of genes distributed throughout the genome. Chromatin immunoprecipitation (ChIP)-seq experiments suggest that USF proteins bind as many as 2500 different human genes21. Both contain a highly-conserved USF-specific region (USR) in N-terminal region responsible for regulation of transcriptional activity18,22,23. However, USF1 and USF2 are not fully interchangeable, as only USF2 can bind the pyrimidine-rich Inr elements in the core promoters of some targets24.

Likewise, studies in available model systems lacking either USF1 or USF2 (but not both) have linked USF2 specifically to prostate cancer, while USF1 appears to provide UV- response independent of USF225. The two USF proteins show significant preference for a canonical CACGTG E box, though they will also bind E boxes with variations at the third and fourth nucleotides, or even noncanonical heptameric E boxes that preserve the palindromic nature of the CAC-GTG half-sites26. MYC:MAX heterodimers also prefer

CACGTG targets, and separate studies have demonstrated that USF and MYC

5

Figure 1. Functional domains and kinase sites of mouse USF1 and USF2.

6

complexes bind competitively at select promoters. For example, expression of the CAD pyrimidine biosynthesis protein in dividing cells requires the binding of MYC:MAX heterodimers at the Cad gene promoter. In non-dividing cells, MYC:MAX is replaced by

USF1:USF2 heterodimers, and expression of the Cad gene is repressed27. The widespread nature of USF and MYC expression, together with their affinity for CACGTG binding sites suggests that competition between USF and MYC may occur at many of the USF target genes28.

1.3: USF as a tumor suppressor The genetic lesion most associated with prostate cancer aggressiveness is localized to human Chromosome 1929. This Prostate Cancer Aggressiveness Locus (19q12-q13.1) is most tightly linked to genes encoding Kallikrein 1 (KLK1) and Kallikrein 3 (KLK3) (or PSA), members of the kallikrein family of serine proteases, positioned at 19q13.3330. The Usf2 gene sits at 19q13.11 and is also frequently lost in prostate cancers bearing 19q deletions31. PC3 human prostate cancer cell lines show loss of USF2 activity despite persistence of the Usf2 gene32, suggesting that linkage of Usf2 deletion and prostate cancer may underestimate

USF2’s prostate cancer association. Indeed, PC3 cells engineered to overexpress a recombinant USF2 transgene showed reduced tumorigenicity in xenograft transplants20.

Although the scope of USF2’s gene regulatory involvement in prostate biology has not been studied, USF2 assembles with (AR) to regulate transcription of AR target genes in prostate epithelial cells33. USF2 knock-out (KO) mice also support a role for USF2 in suppressing prostate cancer. Specifically, USF2 KO mice display marked prostate hyperplasia at a young age20. USF2 (but not USF1) was also found to inhibit tumorigenic

7

transformation in rat embryonic fibroblasts (REFs) induced by the adenovirus onco-protein

E1A22.

Though not as clearly demonstrated as USF2, USF1 also exhibits tumor suppressor activity. Both USF proteins can prevent RAS- and MYC-dependent transformation of REFs

22, and loss of both USF1 and USF2 transcriptional activities was shown in a panel of human breast cancer cell lines (as noted in PC3 cells, this loss of USF transactivating activity was seen despite persistent expression of both USF proteins)34. USF1’s role in tumor suppression is perhaps best understood in skin cells that have been exposed to UV radiation. In mouse skin cells treated with UV light, USF1 appears to bind and stabilize P5335, regulate G1/S and

G2/M transitions through transcriptional regulation of multiple cyclin and cyclin dependent kinase genes 36, and induce expression of pigmentation and nucleotide excision repair (NER) factors37.

In addition to the evidence that links USF1 or USF2 to skin or prostate cancer, respectively, both USF proteins have been linked to a number of other cancer types. An analysis of human breast cancer cell lines found that while USF1 and USF2 protein levels were similar to those seen in untransformed and primary breast tissues, 75% of the breast cancer lines lacked detectable USF2 transactivating activity, and half also lacked USF1 activity34. Similar studies have linked changes in USF expression or activity to a number of cancers. In a screen for human ovarian cancer biomarkers, USF1 exhibited the highest level of enrichment among the 19 ovarian cancer biomarker candidates identified from cancer serum and tumor cell extracts38,39.As with USF’s disregulation in breast cancer lines,

8

subcellular staining in ovarian tumors found that USF1 was mislocalized to a perinuclear compartment that appeared Golgi-related. The authors speculate that mislocalization of USF1 to the Golgi could account for the protein’s appearance in cancer serum samples39.

Though no attempt has been made to systematically define USF’s impact on genome- wide transcription, a number of genes critical to growth regulation have been defined as USF target genes. For example, the Tp53, Brca2 and Apc tumor suppressor genes are all direct targets of USF1/240. Whereas USF activates expression of tumor suppressor genes like Apc and Brca241, they repress expression of human telomerase reverse transcriptase (hTert)42.

Deletion of USF leads to elevated hTERT in oral cancer cell models43, mimicking the MYC- dependent upregulation of hTERT that is seen in over 85% of tumor cells44. USF2 also represses expression of the cell cycle kinase CDK436. As with CAD and perhaps hTERT,

USF’s repression of the Cdk4 gene opposes activation by c-MYC36. Together, the repressive role USF plays in regulating growth promoting genes like hTert and Cdk4, and the activating role it plays in regulating tumor suppressor genes like Brca2 result in a largely antiproliferative function.

Paradoxically, whereas loss of USF activity is associated with prostate, breast and ovarian cancers, one study suggested that USF2 overexpression induces hyperproliferation in human dysplastic lung epithelium and non-small cell lung cancer cell lines45, though no model for this activity was proposed. However, the study’s authors note that expression of

USF2 dominant negative transgenes had no apparent impact on proliferation, calling the pro- proliferative role of USF2 in these models into question.

9

1.4 USF as a stress response factor Stress-induced activation of USF is not limited to the UV radiation response. In cell models of inflammation, USF1 has been shown to counteract pro-inflammatory NFB signals through induction of the gene for tumor necrosis factor alpha-induced protein-3

46 (TNFAPI3 or A20) . Whereas USF1 normally suppresses expression of plasminogen activator inhibitor 1 (PAI-1), it stimulates PAI-1 expression when activated at wound sites47.

USF1 is activated in response to fasting, and in turn regulates the genes for insulin48 and fatty acid synthetase49, suggesting a potential linkage with Type II diabetes50.

The ability of USF to transduce environmental signals extends beyond stressor like

DNA damage, tissue lesion or fasting. USF is essential for early embryonic, neuronal, reproductive and hematopoietic development programs and regulates a wide array of genes involved in the immune response. Attempts to generate mouse genetic models of USF deletion first confirmed its role in neuronal development. USF2-deficient mice in particular showed dramatically shortened male lifespan, reduced viability, growth and fertility.

Moreover, both USF1-/- and USF2-/- mice were prone to spontaneous epileptic seizures51. In contrast, homozygous USF1-/-/USF2-/- mice died during early embryogenesis, as did mice engineered to overexpress A-USF, a USF1 dominant negative transgenic protein that readily dimerizes with either USF1 or USF2, but lacks USF1’s DNA binding and transactivation domains52.

10

USF1 has been shown to promote transcription of the IgH and Igλ2 genes53, the C4 complement protein and 2 microglobulin54–56. USF1/2 heterodimers regulate expression of the J chain and polymeric Ig receptor in plasma cells57. As with Ig expression in B cells, USF regulates expression of multiple T cell receptor genes in developing T cells. Whereas it drives transcriptional activation of the D2 gene segment in Tcrd, our laboratory found that it represses germline transcription of the Tcrb D2 gene segment58. During Tcrb assembly,

DJβ1 joints are preferentially formed over more distal DJβ2 joints. We found that USF occupies the D2 upstream promoter prior to recombination, leaving the D2 RSSs untranscribed and therefore less accessible to V(D)J recombinase at the outset of gene assembly. However, as DJ1 recombination progresses, transient DNA breaks trigger activation of DNA-PKcs, which phosphorylates USF1, leading to its displacement from the

5’Dβ2 promoter. In the absence of USF, the promoter is activated and D2 rearrangements begin to accumulate26. This temporal delay in DJ2 assembly ensures that V gene segments initially target the upstream DJ1 joint, allowing DJ2 to serve as a target for secondary V- to-DJ rearrangements should V-to-DJ1 joints be non-functional due to frameshift.

1.5 USF regulation by phosphorylation Given their ubiquitous and largely stable expression, the ability of USF1 and USF2 to respond to stressors suggests that function of the USF proteins is regulated post- translationally. Separate studies have shown that both USF1 and USF2 interact with the oncogenic transcription factor FRA159, and that USF2 is polyubiquitinylated and degraded in

11

hypoxic cells60. However, the significance of such regulatory schemes has not been extended.

More intriguingly, two independent studies have shown that nuclear localization of USF2 declines in activated mast cells and in hepatocytes of infants suffering from biliary atresia61,62, mirroring the loss of nuclear USF1 seen in ovarian cancer63 and the lack of detectable USF activity in breast cancer cell models34. The mechanisms that regulated altered

USF nuclear translocation in such stress conditions are unknown.

The diverse pathways by which USF1 and USF2 are phosphorylated, and the functional consequence of their phosphorylation are better understood. For example, DNA-

PK phosphorylates USF1 at ser262 in the b-HLH-LZ domain, triggering subsequent lys237 acetylation and USF1 activation64. We have shown in lymphocyte cell lines exposed to IR, that induction of DNA-PK results in USF1 displacement from the Tcrb 5’D2 promoter and loss of promoter repression26. Conversely, DNA-PK induced in fasting adipocytes resulted in enhanced USF1 binding at the (FAS) promoter and promoter activation64.

Whereas IR activates DNA-PK to alter USF1, UV light activates P38-MAP kinase, which phosphorylates USF1 at thr153. Similar to IR-mediated activation, phosphorylation at thr153 leads to lysine acetylation (in this case, lys199) and altered DNA binding affinity65. In trigeminal ganglion neurons, USF1 transactivation is regulated by (MEK)/ERK MAP kinase- dependent phosphorylation at thr15366.

In cardiac muscle, USF1 is phosphorylated by PKC, which increases binding at the cardiac α-myosin heavy chain promoter67, by CDK2, which regulates multiple stages of the cell cycle68. Interestingly, whereas CDK2-phosphorylation of USF1 increased its DNA

12

binding activity, CDK2-phosphorylated c-MYC showed reduced DNA binding69,70. USF1 is also phosphorylated by messenger-independent CK2 in pancreatic insulinoma cell lines, which regulates USF1’s heterodimerization and transactivation69. Both USF1 and USF2 contain putative PKA phosphorylation sites (two for USF1, three for USF2)71 that are linked to USF’s transactivation potential72 73. Likewise, both proteins are phosphorylated by

Glycogen synthase kinase-3 (GSK3), which regulates PI3K/Akt and Wnt/-catenin pathways, and is implicated in a number of pathologies including type-2 diabetes and cancer74. Phosphorylation of USF1 (again, at thr153) was critical for transactivation of genes that promote apoptosis and cell cycle arrest75. In silico modeling of GSK3-phosphorylated

USF2 suggests adoption of a more open conformation that would facilitate DNA binding and transactivation potential. Functional studies using USF2 variants that mimic GSK3 phosphorylated forms found no impact on cell proliferation, but instead found increased levels of cell migration31, suggesting that GSK3 regulation of USF proteins may influence their role in tumorigenesis.

1.5: Activation-induced cytidine deaminase Activation-induced cytidine deaminase (AID) is responsible for initiating both SHM and CSR9. In both cases, AID deaminates cytosines found on single-stranded DNA (ssDNA) during transcription8, and is facilitated by AID’s interaction with RNA Pol II (RNAP II)76,77.

During SHM, AID preferentially targets the hotspot sequence RGYW/WRCY78, with a focus on WRC (A/T, A/G, C) motifs in IgH and IgL genes79. SHM typically starts approximately

150 bp downstream of the transcription start site (TSS) of V region promoters78, with

13

mutations occurring up to 1-2 kb downstream of these TSSs80. Mutations that occur within the V region promoter can lead to a decrease in SHM by preventing transcription initiation80.

Mismatch repair (MMR) and base excision repair (BER) factors, like MSH2/MSH6 and

UNG, help repair U*G mismatches that result from deaminating cytidine to uridine, either restoring the C*G basepair, or producing a new T*A basepair. Deficiencies in these genes are evidenced by the presence of unrepaired uracil bases within the Ig genes, or the accumulation of AID-mediated C*G to T*A transition mutations in genes other than IgH and IgL81. Such

AID off-target genes include Bcl6, h2afx and c-myc, all three of which are induced during germinal center activation78.

Although MMR and BER are typically error-free DNA repair mechanisms, their recruitment to AID-mediated DNA lesions, together with DNA Exonuclease 1 or AP endonuclease 1, respectively, result in excision of a single-stranded or double-stranded patch of DNA containing the mismatch. Single-stranded ends are repaired by an error-prone DNA polymerase like Pol or Pol, while DSBs are first bound by Rad52/51 complexes, resected, and then repaired by Pol82. The error-prone nature of SHM lesion repair is essential for hypermutation of the Ig V sequences.

During CSR, synaptosomes between donor Sμ and a downstream acceptor Switch region are formed in part due to AID deaminating cytidines in the targeted S, S or Sε region83. Switch regions are highly enriched in AID’s preferred WRC target sequences79.

AID-driven mutagenesis introduces transient double stranded DNA breaks within the synapsed S and targeted S region, facilitating microhomology-based NHEJ repair that

14

deletes the sequence between S and target S breaks. Simultaneous targeting of AID to sites outside of IgH can also lead to translocations like those between S and cMyc that result in

Burkitt’s Lymphoma78. CSR is essential for immune competence. The rarity of Burkitt’s-like tumorigenic translocations is again facilitated by the affinity of AID for RNAP II and by its preference for the G-rich R-loops that the highly repetitive Switch regions form during their transcription83,84, causing RNAP II stalling and increased activating histone modifications and chromatin accessibility85,86.

AID is expressed at low levels in most cells including embryonic cells, and participates in pathways unrelated to SHM and CSR. Specifically, AID was found to promote

DNA demethylation through the deamination of meC to thymidine87 and repair of the subsequent mismatch though MMR and BER repair pathways88–90. Some have speculated that this pathway may play a role in pluripotent tissues87. AID expression in transitional B cells may also mediate tolerance91,92. However, the significance of AID activity in processes beyond SHM and CSR, and the mechanisms by which they are mediated, remain highly speculative.

1.6: Regulation of AID To safeguard against inappropriate alteration of the genome, AID is regulated on many levels, including expression, subcellular localization and activation. AID transcription is induced by a variety of cytokines including: T cell-derived cytokines like IL-4 and CD40L,

T-cell independent cyokines like BAFF and APRIL, female sex hormones, and TLR signaling, all of which promote transcription factors to bind to regulatory elements within the

15

Aicda promoter and enhancer12,38,93–95. Transcription factors implicated in Aicda transcription include NF-κB, C/EBP, Smad3/4, Sp1, Oct/Hox and STAT696. The first 2 kb downstream of the Aicda transcription start site contains repressor elements critical to limiting expression in the absence of germinal center activation signaling. Transcription factors implicated in Aicda repression include: PAX5, c-Myb, , NFB and E47/E2-297. Stability of the processed

Aicda mRNA is further repressed by miR-155 and miR-181b 98–101; in acute B cell lymphoma lacking miR-155 increased AID protein levels and nuclear translocations lead to aggressive CSR85,98–100.

The AID protein is stabilized through several mechanisms. In the cytoplasm, where

90% of AID is found before GC induction102, AID is stabilized by heat shock protein 90

(HSP90)103, while eEF1A prevents AID from entering the nucleus104. Recruitment of intranuclear AID to its ssDNA target is further regulated by a series of co-factors including mechanism for the regulation of AID activity is through the recruitment of ssDNA binding adaptor 14-3-3 protein105, RNA-processing and/or splicing factors, polypyrimidine-tract binding protein (PTBP2)106, RNAPII stalling cofactor (Spt5)107, replication protein A

(RPA)108 and RNA exosomes109. RPA stabilizes AID’s interaction with ssDNA108 at immunoglobin variable genes specifically10, while ar1-α provides a similar function at sites of switch recombination77,110.

16

Casellas et al, 2016 Figure 2. Schematic of AID regulation at the transcriptional, post-transcriptional, and post-translational levels.

17

1.7: Pathologies associated with AID Deregulated expression of AID in B cells is associated with chromatid breaks and translocations in nearly every chromosome111, and with 95% of lymphomas having a B cell origin112, a role for AID in B cell pathogenesis is of particular interest. Translocations occur when 1) there are two pairs of DSBs, 2) there is close proximity between the broken ends, and 3) there is joining of the heterologous DNA ends5. They usually occur at fragile sites, which are large genomic regions prone to replicative stress113; fragile sites containing CpG nucleotides may be more susceptible to AID-independent replication fork collapse114, although it has also been proposed that AID expression in early B cells targets CpG sites, resulting in the generation of substrates for RAG- and ARTEMIS-dependent nicking115.

There are many translocations associated with AID from BCR:ATM116 to Myc:IgH. AID overexpression is strongly correlated with oncogenesis, irrespective of cell type, and AID- mediated translocations have been identified in prostate cancer cells117. Indeed, AID disregulation has been implicated in diverse pathologies including blast crisis progression in chronic myeloid leukemia101, prostate malignancies117, and gastric tumors118.

The most common oncogenic lesion associated with AID is the Myc:IgH translocation found in Burkitt’s lymphoma. In this disease, a reciprocal translocation occurs between the coding region of the c-Myc proto-oncogene and the Igh V(D)J coding region119, placing Myc expression under control of the strongly activated IgH 3’ enhancer120. The Myc and Igh genes frequently colocalize in the nuclei of activated B cells, but not in the nuclei of resting B cells or other cell types121,85; this may imply that they share transcriptional regulation, as loci that are jointly regulated are often spatially proximal122. IgH S regions

18

seem to be particularly susceptible to AID-dependent translocation due to the R-loops and other unstable structures, like G-quadruplexes, that form during transcription of the highly repetitive switch regions123. Multiple studies have confirmed that these translocations are

AID-dependent. First, overexpression of AID has been shown to drive DNA damage at the exon1-intron1 junction of Myc124. Second, translocations in B cell lines do not occur if UNG is absent, indicating that the U:G mismatches created by AID are necessary125. Third, Balb/c mice injected with IL-6 generated Myc-IgH rearrangements only when AID was present119,126,127.

Translocations are not the only aberration found due to AID expression. In fact, many cancers caused by improper AID expression are characterized by point mutations in both oncogenes and tumor suppressor genes128. Some genes associated with B cell tumorigenesis, like Myc, Pim1, Pax5, Ocab (Pou2af1), H2afx, Rhoh are able to escape AID-induced mutations by mismatch and base excision repair78. When these repair mechanisms are present, the mutation rate in off-target genes like c-Myc often doesn’t exceed background levels78. When tumor suppressor genes like Tp53 are absent, AID can induce rapid onset mature B cell lymphoma111; p53 functionality can be impaired through normal processes like replication errors, or through the induction of AID expression, which causes mutations in

Tp53 in human gastric cells118 and which is strongly implicated in Tp53 mutation in human breast cancer cells129. Bcl6 is the most frequently mutated non-Ig gene in B cells, and is a gene implicated in diffuse large B cell lymphoma78. Its mutation rate is ascribed to AID, as an analysis of 31 genes from AID-/- cells showed a significantly lower Bcl6 mutation

19

frequency than that in AID-expressing cells78. Further, AID accounted for mutations in 25% of 188 genes randomly surveyed in a germinal center B cell genome, though such off-target genes showed mutation rated of 100-fold lower than the Ig loci78.

1.8: Introduction and Objectives of the Thesis Project Cancer arises from genome instability that arises after damaged DNA is either not repaired or improperly repaired. To avoid genome instability, cells have elaborate programs that are activated in response to damage130. The substitution error rate of human DNA polymerase ε is around 10-8 errors per , the vast majority of which are then repaired by ubiquitous DNA repair enzymes. DNA can similarly be damaged during the process of recombination. But again, that damage is aggressively repaired, and DNA recombination is limited to meiosis in gametes and antigen receptor gene assembly in developing lymphocytes131. More commonly, cancer-causing mutations arise after DNA is damaged by exposure to environmental agents including ultraviolet or ionizing radiation, chemicals or infectious agents. In response to DNA damage, cells undergo a multi-faceted response that includes cell cycle arrest, DNA repair, metabolic regulation, and if necessary, senescence or apoptosis if damage is irreparable. Several of these responses are immediate, driven by the post-translational activation of existing regulators. Ultimately, recovery from DNA damage also involves transcription responses over the following days and weeks. Unfortunately, beyond the cell cycle arrest and proapoptotic programs driven by P53 very few of the transcriptional processes triggered by DNA damage have been investigated.

20

As noted earlier, the ubiquitously expressed USF1 and USF2 stress response proteins were among the first mammalian transcription factors identified based on their involvement in regulating activity of the adenovirus major late promoter 16. They have long been appreciated as tumor suppressors, though evidence supporting such a role is limited. Chiefly,

USF’s tumor suppressor linkages include: (a) their activity being lost in a variety of cancers, including breast23, liver 132, colorectal40, prostate20, and oral cancers43, (b) the ability of USF2 overexpression to reduce in vivo tumorigenicity of mouse prostate cancer xenografts20, (c) their binding activity at the promoters of other tumor suppressor genes including Apc, Brca2 and Tp5340, (d) their ability to repress expression of TERT133, and (e) their antagonism of the transforming function of MYC134.

Much of the deciphering the molecular pathways that link USF to cancer has been conducted in the skin of USF1-deficient mice. These studies demonstrate that upon UV exposure USF1 is activated in keratinocytes and melanocytes by p38 MAPK, and drives expression of genes involved in nucleotide excision repair (NER)135 and melanin deposition as part of the tanning response136. USF1 also appears to stabilize P53 protein levels through nuclear protein:protein interaction35. Unfortunately, all of the studies examining USF1’s role in protecting skin cells from cancer were limited to short-term (ie, ≤ 24 hrs) responses, which have also been the primary focus of most DNA damage response studies. As a consequence, virtually nothing is known about USF’s role in the longer-term response to DNA damage and the slow process of tumorigenesis. We previously found that USF1 was also involved in regulating germline transcription of Tcrb gene segments in response to dsDNA breaks

21

generated either during V(D)J recombination or by experimental exposure of cultured lymphocyte cell lines to sublethal doses of ionizing radiation26. In our lymphocyte cell models however, we found that exposure to IR or prolonged V(D)J recombinase expression resulted in USF activation that persisted long after nonhomologous end-joining (NHEJ) repair programs had attempted to resolve DNA lesions26, suggesting that USF might be an ideal candidate for directing the long-term transcriptional changes that guide cell recovery and limit genomic instability after DNA stress. To test its candidacy however, we would need to improve on existing genetic models of USF deficiency.

The two USF proteins are basic helix-loop-helix (bHLH) transcription factors and function primarily as USF1/USF2 heterodimers, binding E boxes at over 2500 target genes21.

As seen in the example of UV irradiation, stress responsiveness of USF is controlled through differential phosphorylation of USF1 by protein kinases including p38 (as well as ERK

MAPK), DNA-PK, GSK3, CK2 and PKA64 (see Figure 1). Phosphorylation by these protein kinases regulates USF1/2 heterodimerization, DNA binding, and transactivation potential of USF proteins137. Attempts to define the role of USF in DDR and cancer development have been limited by difficulty in generating appropriate genetic model systems. Although USF1 knock-out mice are prone to seizures51, and USF2-deficient mice are underweight and have reduced male sterility138, they each retain a significant degree of

USF functionality through the ability for form functional homodimers of the remaining USF protein139,140. Because mice with homozygous null mutations of both USF1 and USF2 die

22

early in embryogenesis51, no animal model deficient for USF activity has been developed.

Consequently, no profile of USF’s genomic stress-response yet exists.

Germinal center B cells are activated in response to antigen presentation, and each undergo two separate events that induced dsDNA breaks in their Ig heavy chain genes. These events, somatic hypermutation and class-switch recombination, are both essential for B cell function, but are also strongly linked to lymphomagenesis, illustrating the need for B cells to discriminate between DNA damage that is ultimately beneficial or dangerous. Building on our finding that USF activity is responsive to DNA breaks in cultured lymphocytes, we modeled USF deficiency using RNAi to simultaneously deplete both USF1 and USF2 from the M12 mouse germinal center B cell line. Cells were transduced with either a scrambled shRNA control lentivirus (sshRNA) or a mixture of USF1 and USF2 shRNA viruses, stably transduced clones were isolated based on their resistance to virally-encoded puromycin, and the transcriptomic response to USF depletion and ionizing radiation was measured using

Affymetrix mouse whole transcript arrays.

Microarray analysis revealed 765 differentially expressed genes (≥1.50-fold change, 0.05 false discovery rate threshold, n=3) in untreated USF-depleted cells when compared with sshRNA controls. Seven days after cells were exposed to ionizing radiation, 2866 genes showed altered expression in the USF-depleted cells. In stark contrast, comparison of untreated USF-KD cells and USF-KD cells only 1 day after IR showed no differentially expressed genes that met both the fold-change and FDR thresholds, supporting a role for

USF in long-term rather than short-term DDR. Microarray findings were confirmed by RT-

23

qPCR for a number of immune- and cancer-related genes including Aicda, Blk, Blnk, and multiple NFB family members, none of which had previously been identified as USF target genes. The value of using a genetic system that simultaneously targets both USF proteins was further supported by expression analyses in Usf1-/- thymus and USF single-gene KD cells, both of which showed trends in differential gene expression similar to but of significantly less magnitude than that seen in the USF1/USF2 double KD cells. These findings provide the first evidence that USF1 and USF2 collaborate to play a novel role in regulating long-term response to DNA damage and lymphocyte activity, and also illustrate the significance of investigating USF’s regulatory role using models that overcome the functional overlap shared by USF1 and USF2.

24

CHAPTER 2: Materials & Methods

2.1 Cell culture The murine M12 germinal center B cell line has been previously described141. All

o M12 cell lines used in this study were cultured at 37 C/5% CO2 in RPMI 1640 media supplemented with 10% fetal calf serum, 2 mM L-glutamine, 0.01% penicillin/streptomycin, and 50 mM -mercaptoethanol.

2.2 RNA interference and transfection For depletion of USF2 in M12 cells, a validated Mission USF2 shRNA

(TRCN0000071575) in pLKO.1 backbone was purchased and prepared as described by the provider (Sigma). Two independent USF1 targeting shRNAs were designed using the

BROAD Institute Genetic Perturbation Platform. Duplexed shRNA USF1shR1 (sense: 5’-

CCGGCAGAAGTTAAGATGCGTACCTCTCGAGAGGTACGCATCTTAACTTCTGTTT

TTG -3’, antisense: 5’-

AATTCAAAAACAGAAGTTAAGATGCGTACCTCTCGAGAGGTACGCATCTTAACT

TCTG-3’) and USF1 shR2 (sense: 5’-

CCGGGAGGGCTCAACATAACGAAGTCTCGAGACTTCGTTATGTTGAGCCCTCTTT

TTG-3’, antisense: 5’-

AATTCAAAAAGAGGGCTCAACATAACGAAGTCTCGAGACTTCGTTATGTTGAGC

CCTC-3’) oligonucleotides were cloned into the AseI site of pLKO.1 (Addgene), and insertion was verified by sequence analysis. The plasmid psshRNA, in which a scrambled shRNA sequence was cloned into the AseI site of pLKO.1 was purchased from Addgene.

25

USF1shR1, USF2shR2 and sshRNA plasmids were independently transfected into

293T cells, together with third generation lentiviral packaging plasmids (kindly provided by

Dr. Frank Scholle) using FugeneHD (Roche). Briefly, for shRNA plasmid transfections, 1 x

107 293T cells were seeded per 10 ml DMEM complete medium (supplemented with 10% fetal calf serum, 2 mM L-glutamine, 0.01% penicillin/streptomycin, and 50 mM - mercaptoethanol) and allowed to attach for 4 hours. USF1shR1, USF1shR2 and USF2shR

(640 ng each) or sshRNA (2.56 g) was mixed together with pFG12 (640 ng), pMDLG (1.6

g), pREV (1.6 g) and pCMV-G (1.0 g) in 870 l serum free DMEM and 56 l

FugeneHD, vortexed briefly, and allowed to complex at room temperature for 10 minutes.

DNA mixture was added dropwise to plated cells. The plate was swirled gently to mix, and returned to the incubator overnight. After overnight transfection, the culture medium was replaced with complete DMEM supplemented with 1% BSA. After 24 hours, the culture supernatant was harvested and stored at 4°C, while the plate was replenished with BSA- supplemented medium and allowed to grow an additional 8 hours, at which point the culture supernatant was harvested, combined with the initial harvest and filtered (0.45 m).

The day before transduction, M12 cells were split to ensure growth. For lentiviral transductions, 105 M12 cells were seeded in 500 l complete RPMI in each well of a 24-well tissue culture plate. Filtered viral supernatant (500 l) and polybrene (8 g/ml final concentration) was added, and the plated cells were centrifuged (1 hr @ 1200 x g @ RT) and allowed to recover. GFP reporter fluorescence was confirmed in transduced wells one day after transduction, and cells were selected with the addition of puromycin (3 g/ml final

26

concentration). Individual clones were manually isolated and screened for USF1 and USF2 expression by Western blotting.

2.3 Western blotting For protein analysis, freshly growing M12 clones (1-5 x 106 cells) were washed in ice-cold PBS, and nuclear or whole cell protein lysates were generated. For nuclear fraction isolates, cells were lysed in NP-40 Buffer (500 l of 10 mM HEPES pH 7.9, 1.5 mM MgCl2,

10 mM KCl, 0.5 mM DTT, 0.1% NP-40) supplemented with PMSF and HALT protease and phosphatase inhibitor cocktail (Thermo Scientific) for 2 min on ice. After lysis, cells were pelleted in a microcentrifuge (2 min @ 12,000 RPM @ 4oC), the nuclear pellet was resuspended in Nuclear Lysis Buffer (100 l 10 mM HEPES pH 7.9, 25% glycerol, 0.42 M

NaCl, 1.5 mM MgCl2, 0.2 mM EDTA, 0.5 mM DTT) supplemented with PMSF and HALT protease and phosphatase inhibitor cocktail, and incubated 30 min on ice with frequent manual agitation. Nuclear lysate was recovered by microcentrifugation (20 min @ 12,000

RPM @ 4oC) and stored at -80oC. Whole cell lysates were generated by lysing pelleted cells in RIPA (1 ml of 150 mM NaCl, 1% NP-40, 0.5% NaDeoxycholate, 0.1% SDS, 50 mM Tris-

HCl 8.0) at 4oC for 30 min with frequent manual agitation, and microcentrifuged at 12,000

RMP for 20 min at 4oC. Protein concentration in the recovered supernatant was assessed by

Coomassie Protein Assay (Thermo Scientific) according to the manufacturer’s instructions.

For Western analysis, nuclear or whole cell lysates (15 g each) were electrophoresed on AnyKD Mini-Protean TGX acrylamide gels (Bio-Rad), and transferred to PVDF using a

27

TransBlot Turbo device (Bio-Rad). The membrane was treated with 1X TBS 3% Casein

Blocking Buffer (Bio-Rad) one hour at room temperature, and then incubated with rabbit anti-mouse polyclonal antibodies (Santa Cruz Biotechnology) against USF1 (sc-229), USF2

(sc-862), Sp1 (sc-59) or -actin (sc-130656) at a 1:5000 dilution in Blocking Buffer overnight at 4°C. After washing 5X in TBST, the membrane was incubated with goat anti- rabbit HRP secondary antibody (Pierce, 32260) for 1hr at RT, rewashed (5X in TBST) before

ECL Prime chemiluminescent detection (GE Heathcare) and autoradiographic exposure.

2.4 Ionizing radiation treatment Triplicate dishes were prepared for parental M12 and USF1/2-depleted M12 clones.

Twenty-four hours prior to irradiation, total cellular RNA and protein whole cell lysates were extracted from 3 x 106 cells per clone, and each clone was reseeded at a density of 3 x 105/ml culture medium in 10 cm dishes. The next day, clones were harvested into independent 15 ml conical centrifuge tubes, transported in a sealed cooler and exposed to a single sublethal dose

(5 Gy) of ionizing radiation using a Gammacell 220 cobalt-60 irradiator (MDS Nordion). After irradiation, cells were pelleted, resuspended in fresh RPMI complete medium, and allowed to recover overnight. Total cellular RNA and protein whole cell lysates were extracted from 3 x

106 cells/plate each at one day and seven days post IR treatment.

28

2.5 RNA extraction Total cellular RNA was extracted from 1-5 x 106 cells using RNeasy Plus (Qiagen) according to the manufacturer’s protocol. RNA levels were quantified using a NanoDrop

1000 spectrophotometer (Thermo Scientific), and stored at -80oC.

2.6 Affymetrix Microarray Hybridization RNA quality was assessed using an AATI Fragment Analyzer (Advanced Analytical

Technologies, Ankeny, IA). RNA samples (3-5 ul) are heated at 70oC for two minutes, placed on ice, and 2 ul of the denatured RNA is added to 22 ul of Normal Sensitivity RNA

Marker-Diluent (Advanced Analytical Technologies, Ankeny, IA) in a 96 well plate. The 96 well plate is loaded on to the instrument and the RNA quality is assessed using the

PROSizeTM analytical software (Advanced Analytical Technologies, Ankeny, IA). Fragment

Analyzer provides a RNA Quality Number (RQN) to assesses the quality or integrity of total

RNA on a scale of 1–10 (one is poor quality, 10 is excellent quality). The average RQN for the samples used in this experiment was 9.4.

Gene expression analysis was conducted using Affymetrix Mouse Transcriptome 1.0 arrays (Affymetrix, Santa Clara, CA). One hundred nanograms (100 ng) of total RNA were amplified and labeled as directed using the Affymetrix WT Plus Reagent Kit (WT Plus Kit).

Five and one half micrograms (5.5 μg) of amplified biotin-cDNAs were fragmented and hybridized to each array for 16 hours at 45°C in a rotating hybridization oven. Array slides were stained with streptavidin/phycoerythrin utilizing a double-antibody staining procedure and then washed for antibody amplification according to the GeneChip Hybridization, Wash

29

and Stain Kit and user manual following protocol FS450-0001. Arrays were scanned in an

Affymetrix Scanner 3000 and data was obtained using the GeneChip® Command Console software.

2.7 Bioinformatic Analysis of Gene Expression The CEL files were preprocessed by computing the robust multiarray average (RMA) of background adjusted, quantile normalized and log2 transformed perfect match (PM) pixel intensity values (PMID: 12582260, PMID: 12925520). Microarray analysis was performed using a three-way (cell line, treatment and time) analysis of variance (ANOVA) model with a three-way interaction term in Partek Genomics Suite v6.6 software. For each cell line, differentially expressed genes (DEGs) were detected by contrasting treatment group means versus the mean of the untreated controls and using a false discovery rate (FDR) < 0.05142 with an absolute fold change > 1.5. The DEGs were enriched for biological pathways

(FDR < 0.05) using gene ontology (GO) biological processes and the Kyoto Encyclopedia of

Genes and Genomes (KEGG) knowledge base (PMID: 10592173) in the Database for

Annotation, Visualization, and Integrated Discovery (DAVID) v6.7 Bioinformatics resource

(PMID: 19131956, PMID: 19033363)143,144.

2.8 Quantitative and reverse transcriptase PCR (qPCR and RT-PCR) For RT-qPCR analyses, each RNA sample (1 g) was reverse transcribed using

RevertAid MMLV reverse transcriptase (Thermo Scientific) and oligo d(T) primers. The resultant cDNAs were amplified in triplicate qPCR reactions with 1 mol each of the

30

indicated primer pairs (Table 1). PCR reaction mixtures (20 l SensiMix Plus; Bioline) were amplified (94oC, 20 sec; 60oC, 30 sec; 72oC, 30 sec) for 40 cycles, followed by melt-curve analysis using an iQ5 iCycler (Bio-Rad). Quantitation of target gene levels was achieved by

2-CT analysis145 normalizing gene expression levels in experimental samples to matched levels in untreated sshRNA M12 cDNA. Variations in sample loading were controlled by normalizing target gene expression levels to those obtained for  actin.

31

Table 1. List of Oligonucleotides Name Sequence Tm Aicda F ΑCACCTCCTGCTCACTGGACT 58.9 Aicda R GGTCCAGGTCCCAGTCTGA 58.6 cdkn1b F TTGGACCAAATGCCTGACTC 55.0 cdkn1b R GGGAACCGTCTGAAACATTTTC 54.4 NFKBID F TCTTTCCCATTCTCTGCTTCTG 54.7 NFKBID R AGGGAAGGCTCAGGATACAG 55.9 BLK F TCTGTTTGACTATGCCGCTG 55.2 BLK R CATAACCTTCTCTTCCTGTGACG 55.1 CD300A F CAGGACCAACACTAGAGACAC 54.7 CD300A R CAGGAGAGCTAACACAGACAA 55.2 BLNK F GAGGATGAGGCTGATTATGTGG 55.1 BLNK R GTGCTTTGAGGAACTGTTTGG 54.5 CTSE F TCGCAGTCCGACACATACAC 57.1 CTSE R CATCCACAGTCAACCCTTCC 55.6 ICOSI F AGCCACAGAGTTAGTCAAGATC 54.1 ICOSI R CATGCAGGTGTAGGTACGTTC 55.5 NFKBIA F AGGAGTACGAGCAAATGGTG 54.7 NFKBIA R CGGCTTCTCTTCGTGGATG 55.8 FOXJ1 F CCATCTACAAGTGGATCACGG 55.1 FOXJ1 R TGTTCAAGGACAGGTTGTGG 55.2 ICAM2 F TGGAGAACAGGAATGGAAGC 54.8 ICAM R TCGGTTGTGGAGATTGGTG 54.9 IRAK1 F AGACTTTGCTGGCTACTGTG 54.9 IRAK1 R AAGAATGTCCAGTCGTTGAGG 54.7 BTK F TGGAGAACAGGAATGGAAGC 54.8 BTK R TCGGTTGTGGAGATTGGTG 54.9 VKBB1*F TGACCCAGTCTCCATCTA 53 JK1R CGTTTCACCTCCACCTTGGT 57.5 CK.2 GTTAACATCTGGAGGTGCCTCAGTCG 60.8 CKR GAAGCTCTTGACAATGGGTGA 55

32

CHAPTER 3: Results

3.1: USF depletion USF1 phosphorylation by P38 MAPK facilitates protective transcriptional responses to DNA damage elicited when keratinocytes and melanocytes are exposed to UV radiation137,146. We have similarly shown that DSBs elicited by IR treatment of lymphocyte cell lines trigger DNA-PK-dependent phosphorylation of USF1 that also regulates its trans- activation capacity26. An appropriate mouse model of USF deficiency does not yet exist.

While single-gene Usf1 or Usf2 knockout mice have been developed, the ability of the remaining USF protein to partially compensate for deletion of the other has prevented the mapping of USF target genes and pathways. At the same time, mice deficient for both Usf genes or constitutively expressing a dominant negative USF transgene fail to develop beyond early embryogenesis. In light of USF’s ubiquitous expression, its involvement in responses to a wide array of stressor including DNA damage, and its function as a tumor suppressor20,35,39,147, a better understanding of how USF regulates stress response and cancer prevention is urgently needed.

Our previous findings suggested that, unlike the rapid response of P53 to DNA damage, dsDNA breaks induce a more slowly developing, but longterm change in USF activity. To visualize the potential role of USF in mediating a protracted transcriptional response to DNA damage, we sought to target USF activity in the mouse M12 B cell line.

Importantly, M12 cells contain point mutations in the DNA binding domain of P53 that inhibit its ability to activate gene transcription148. We used lentiviral-mediated RNA interference to simultaneously deplete USF1 and USF2 proteins from M12 cells. No gross

33

differences in cell viability or growth were observed among the populations of transduced cells. Puromycin-resistant transduced clones expressing short hairpinned RNAs (shRNAs) targeting either Usf1 or Usf2 alone, or in combination, or expressing a scrambled shRNA

(sshRNA) control were isolated and screened for USF depletion by Western blotting of nuclear protein extracts (Fig. 3). Previous studies in USF2-/- mice showed that loss of USF2 led to reduced USF1 expression51. Consistent with this finding, we observed significantly diminished USF1 protein levels in cells transduced with shRNAs targeting either Usf1 or

Usf2 (lanes 3 and 4), and still further USF1 reduction in cells transduced with shRNAs targeting both Usf genes (lane 5). In contrast, depletion of USF2 protein expression was less extensive and was only observed in the presence of Usf2 shRNAs.

USF1

USF2

SP1

Figure 3. USF Knockdown in M12 cells. Western blot analysis of USF1 (upper panel) and USF2 nuclear protein levels (middle panel) in stable USF-depleted cells relative to parental M12 and sshRNA-transduced control cells. SP1 protein levels (lower panel) control for variations in protein loading.

34

3.2: Transcriptomic analysis of USF depletion To test the functional significance of USF depletion on the longterm cellular response to DNA damage, we exposed USFKD and sshRNA control lines (n = 3) to a single sublethal

(5 Gy) dose of ionizing radiation. Total RNA from untreated cells and cells allowed to recover from IR treatment for one or seven days was screened on Affymetrix GeneChip WT arrays to investigate the impact of genotoxic stress on the transcriptome (Fig. 4). We identified 9289 differentially expressed gene probes showing >1.5-fold change (<0.05 FDR) between knockdown and/or treated samples and untreated sshRNA controls (Fig. 4A). ChIP- chip studies suggest that USF proteins bind a wide array of sites across the genome, primarily at the promoter regions of transcriptionally active genes21. However, USF depletion had minimal impact on the pattern of gene expression in unirradiated cells. Untreated sshRNA and USFKD transcription patterns co-segregated in hierarchical Spearman and Ward clustering analyses (Fig. 4B). Indeed, around 10% of the total DEG probes identified showed greater than 1.5-fold change in gene expression in the USF knockdown cells, corresponding to 31 genes with increased expression and 141 genes with decreased expression when compared with untreated sshRNA controls (Fig. 5A).

Figure 4. Differential gene expression analysis of USF-depleted and irradiated M12 cells. (A) Clustering and heatmap representing the 9289 differentially expressed gene(DEG) probes. The data represents the log2 ratio of RMA-normalized intensity data for each sample relative to the mean of the genotype-matched, untreated (no IR) sshRNA controls (0hrs). Red color represents increased expression, blue represents decreased expression and grey denotes no change. Spearman rank as the similarity measure and Ward grouping were used for clustering of the gene probes in the rows (left) and the samples in the columns (top). (B) Number of DEG probes that are statistically significant in the comparisons of the mean expression of the samples in each of the indicated pairwise contrasts.

35

36

Figure 5. Gene expression response to USF depletion and IR. (A and B) Venn diagrams of genes that exhibit ≥2-fold increase (A) or decrease (B) in expression in USFKD cells relative to sshRNA controls before and 7 days after exposure to IR. (C and D) Venn diagrams of genes that exhibit ≥2-fold increase (C) or decrease (D) in expression in USFKD cells or sshRNA cells 7 days post-IR, relative to untreated sshRNA controls.

37

During the first day of recovery, the majority of transcriptional responses to IR are P53- dependent149. As expected, the absence of P53 activity in M12 resulted in only 360 DEG probes meeting the 1.5-fold change and <0.05 FDR thresholds of significance, and did not include genes previously linked to P53 induction (Fig. 4A). Interestingly, no DEG probes met the thresholds of significance in USFKD cells one day post-IR when compared with untreated USFKD cells, suggesting that USF is dispensable for the near-term transcriptional response to DNA damage. In contrast, USF-dependent responses were much more evident seven days post-IR. When normalized to untreated sshRNA controls, both sshRNA and

USFKD cells yielded very similar numbers of DEG probes with >1.5-fold change in expression (Fig. 4A, 5054 and 5035, respectively). However, analysis of the genes identified by these probes showed a markedly different pattern of gene expression seven days post-IR in USF knockdown cells than in sshRNA cells (Fig. 5B). Only half the 240 genes upregulated in the control cells seven days post-IR, were also significantly upregulated in

USF-depleted cells, along with an additional 181 genes. Likewise, of the 446 genes down- regulated in sshRNA cells seven days after IR, only 185 were similarly down-regulated in the knockdown cells, along with an additional 174 genes. The significance of USF’s involvement in such longer-term changes in gene expression is further underscored when the gene expression pattern in USFKD cells after seven days of IR recovery is compared with that of untreated USFKD cells (Fig. 4A). Unlike the relatively modest changes in gene expression resulting from USF knockdown alone, 88% and 76% of the upregulated and downregulated

DEGs, respectively, are restricted to the IR-treated cells. Representative DEGs observed in

38

the USF-depleted cells before (Tables 1 and 2) and 7 days after IR treatment (Table 3), and in sshRNA cells after treatment (Table 4) are shown.

39

40

41

42

43

3.3: Gene Ontology clustering analysis Clustering analyses identified 152 distinct Gene Ontology terms associated with

DEGs in this study, grouped into 12 clusters of related biological function (Fig. 6) that were differentially enriched among experimental samples (Table 5). Irrespective of treatment,

DEGs associated with MHC II-dependent antigen processing and presentation (e.g.:

GO:0002478, GO:0002495, GO: 0002504, GO: 0019886), antigen receptor production (e.g.:

GO:0002377, GO:0002460, GO:0016445), lymphocyte differentiation (GO:0045580,

GO:0045619, GO:0045621) and the activation of an immune response (e.g.: GO:0002250,

GO:0002757, GO:0001817, GO:0006955, GO:0019724, GO:0050863, GO:0051251,

GO:0002253, GO:0050864, GO:0050865, GO:0051249) were enriched in USF-depleted cells. In contrast, DEGs associated with proliferation of immune cells (GO:0030888,

GO:0032944, GO:0050670, GO:0070663) and protein tyrosine phosphorylation

(GO:0018108, GO:0018212) were only enriched in untreated USFKD cells, while DEGs associated with protein folding (GO:0006457, GO:0006458, GO:0051085) and the assembly of large multiprotein or macromolecular complexes (e.g.: GO:0034621, GO:0034622,

GO:0043623, GO:0065003) were only enriched in USFKD samples 7 days after IR treatment.

44

Figure 6. Heat map of the clustering among the enriched Gene Ontology biological process terms annotated to the DEGs. Heat map representation of the clustering of 152 Gene Ontology (GO) biological process (BP) terms enriched (FDR < 0.05) by the statistically significant DEG probes in each of the contrasts listed in Figure 4B. The data represents the –log10 p-value of each BP term from the functional enrichment Fisher Exact Test. Missing values were imputed as 0 (−log base 10(1), i.e., p-value = 1). Euclidean distance as the dissimilarity metric and average linkage grouping were used for clustering the BPs in the rows. The 12 clusters are each labeled with the BP that best represent the cluster as biological theme. The more red the heat map color, the more significant the enrichment.

45

Gene Ontology terms associated with response to ionizing radiation and DNA damage (e.g.: GO:0009314, GO:0009628, GO:0010212, GO:0006302, GO:0008630,

GO:0042770), DSB repair (GO:0006302), chromatin regulation (e.g.: GO:0000070,

GO:0006325, GO:0030261) and apoptosis (GO:0010941, GO:0042981, GO:0043067), as well as nucleic acid processing and transport and regulation of gene expression (e.g.:

GO:0040029, GO:0006350, GO:0006397, GO:0032259, GO:0051169, GO:0051052) were enriched in 7 day post-IR sshRNA controls, but were absent from IR-treated USFKD cells.

However, both sshRNA and USFKD cells were enriched 7 days after IR for GO terms associated with rRNA processing (GO:0006364, GO:0016072), intracellular transport

(GO:0006886, GO:0034613, GO:0046907), ribosome biogenesis (GO:0042254), cell cycle progression (GO:0000278, GO:0007049, GO:0022402, GO:0022403), mitosis (GO:0000087,

GO:0000278, GO:0000280, GO:0007067, GO:0051301) and DNA replication (GO:0006259,

GO:0006260, GO:0006261, GO:0006270), suggesting that USF may be critical for the long- term genomic repair and stabilization, but not for regulating cell growth and division. A complete list of GO terms in each cluster is provided (Appendix 1).

As noted above, USF depletion impacted a number of biological processes associated with immune function. DEGs in USF-depleted cells included those for immune receptors

(e.g.: Il2rg, Il12rb2, Tlr3, Tlr4, Tlr6, Tlr9, Cd19, Cd37, Cd55, Cd74, Fas, and Icosl), intracellular signaling molecules (e.g.: numerous H2-Dma and H2-Dmb genes, Btk, Fyn, Blk,

Btla, Myd88, Plcg2, Ube2n, Calr and Prkdc), and transcription factors (e.g.: Bcl3, Foxj1,

Stat5a, Nfkb1, Nfkbia, Nfkbid and Crebbp), as well as genes for complement C1 proteins.

46

Among the highest observed increases in gene expression upon USF depletion was seen for

Aicda, with or without IR treatment. DEGs that were enriched 7 days after IR in both

USFKD and sshRNA cells included those associated with ribosomal function (e.g.: downregulation of Imp4, Nsa2 and Bms1), DNA replication (e.g.: downregulation of Mcm2,

Mcm4, Mcm5, Mcm7, Pole2 and Polq, but upregulation of Primpol and Polg2), and cell cycling (e.g.: downregulation of Ccna2, Cdc20, Cdc45, Nde1, Smc1a, Smc3). Conversely, those DEGs that were enriched only in the 7 days post-IR control cells included genes critical for response to DNA damage (e.g: downregulation of Trp53, Trp53bp, Rbl1, Brca2, Brcc3,

Chek1 and Chek2 and upregulation of Bbc3, Gadd45b, Atg4a, Btg1 and Btg2). Because M12 cells are P53-deficient, repair of the DNA damage generated during IR exposure depends primarily on ATM/ATR signaling through the Checkpoint kinases encoded by Chek1 and

Chek2150. The failure to appropriately retard expression of DDR genes like Trp53, Rbl1,

Brca2 and the Checkpoint kinase genes, and to enhance expression of the antiproliferative

Gadd45b and B cell translocation genes (Btg1 and 2) genes in irradiated USFKD cells that lack a P53-dependent DNA damage response suggests that USF activity is involved in limiting the cellular response to genotoxic trauma.

3.4: RT-QPCR validation of transcriptomic data To confirm the microarray findings, we submitted two biological replicates of untreated sshRNA, untreated USFKD, and IR-treated USFKD cDNAs to qPCR across several DEGs identified by microarray. Prior to our analyses, we first measured the

47

expression of a series of genes routinely used as PCR loading controls to assess potential impacts of USF depletion and IR exposure. In each case, PCR signals in treated and untreated USFKD samples were normalized to untreated sshRNA cells (Fig. 7A). None of the genes tested were altered by USF depletion. Whereas expression of Histone H1 and

Gapdh genes was significantly affected by IR treatment, expression of 18s rRNA (Rn18s), - actin (Actb), Ig (Cd79a), Hprt and Rplp1 remained relatively constant between treatment conditions. We next amplified Aicda, which was strongly induced in untreated (+3.47-fold) and 7 days post-IR USFKD microarray screens (+4.38-fold), and found that normalizing

Aicda qPCR signals and the qPCRS signals of the genes contained in Figure 8a to -actin and

Ig most directly recapitulated the microarray data (Fig. 7B).

48

Figure 7. Loading control validation in USF-depleted and irradiated cells. (A) The indicated genes were each assessed by qPCR in untreated and irradiated

USFKD cells using ΔΔCT relative to values obtained for untreated sshRNA clones. (B) Aicda gene expression was assessed by qPCR in untreated and irradiated USFKD cells using ΔΔCT relative to values obtained for untreated sshRNA clones, and normalized to the indicated genes to control for variations in loading. Bars represent average fold change in expression (+/- SD) (n=2) of the indicated genes in USFKD cells relative to levels in unirradiated sshRNA cells. Findings are representative of results obtained from two independent biological replicates.

49

We next amplified a panel of representative DEGs identified by microarray, normalizing expression in each sample to untreated sshRNA, and using validated -actin as a loading control. In each case, we were able to recapitulate the directionality of altered gene expression in untreated (Fig. 8A) and 7 days post-IR USFKD cells (Fig. 8B). As expected, the increased sensitivity of qPCR revealed a significantly greater magnitude of increased expression in the USF-depleted cells than was indicated by microarray. In contrast, where qPCR signals diminished near threshold for genes that showed reduced expression in USF- depleted cells, microarray and qPCR magnitudes were equivalent.

50

Figure 8. Quantitative PCR (qPCR) validation of microarray findings in untreated (A), and 7 days post-IR RNAs (B). Bars for qPCR data represent average fold change in expression (n=3 technical replicates) of the indicated genes in USFKD cells relative to levels in unirradiated sshRNA cells (+/- SD). Expression levels were normalized to β- actin to control for sample loading. qPCR values are representative of results obtained from two independent biological replicates.

51

3.5: Aicda activity in USF-depleted cells Activation-induced cytidine deaminase is responsible for initiating both class-switch recombination and somatic hypermutation in germinal center B cells. Overexpression of the

Aicda gene in vitro has been shown to drive IgH:Myc translocations observed in Burkitt’s

Lymphoma119, as well as enhanced somatic hypermutation of Igk antigen binding exons10.

However, AICDA activity is regulated at multiple levels38, and somatic hypermutation in

M12 cells has to date only been demonstrated in a single study141. To determine whether elevated Aicda expression in US-depleted M12 cells is sufficient to drive AICDA activity, we compared sequence variation within the in-frame Igk Vba9/J5 rearrangement141 and downstream C region that had accumulated in untreated USFKD and sshRNA cells. PCR products of the rearranged and control exons were subcloned and sequenced. Despite culture over multiple passages, analysis of the USFKD Vba9/J5 exon in 66 clones revealed no alterations from the published M12 Vba9/J5 sequence (Fig. 9), matching the sequence conservation seen in 33 C controls (data not shown). As such, our data are consistent with previous analyses showing that somatic hypermutations in M12 cells requires exogenous stimulation141. It remains unknown however whether elevated Aicda RNA levels in the USF- depleted cells increase the sensitivity to such activating stimuli.

52

Figure 9. Representative sequence alignment of a portion of the cloned Vκba9*/Jκ5 rearranged antigen binding exons in USFKD using Unipro Ugene v1.26. The nucleotide position of V-to-J recombination is indicated.

53

CHAPTER 4: Discussion

The current study establishes a role for USF transcription factors in regulating the longterm transcriptional response to DNA damage. The ability to properly respond to DNA damage and safeguard the genome is essential to our survival. If DNA lesions, particularly

DSBs, are not resolved or are resolved incorrectly, mutations can accumulate that destabilize the genome and leave the cell susceptible to cancerous transformation. Some participants in the DDR pathway and their contributions are well understood, while others have gone greatly understudied. USF1 and USF2 were among the first mammalian transcription factors identified, are known to bind both tumor suppressor gene and protooncogene promoters40,134,151, and participate in the cellular response to UV-induced DNA damage. USF activity is diminished or lost from several cancers, including breast cancer34, prostate cancer20, oral epithelial cancer43, and liver cancer152. Despite these many linkages to cancer, a critical investigation of how the USF proteins contribute to the DNA damage response and cancer prevention has not been undertaken. This critical gap in knowledge is due in part to the absence of model systems in which both USF activities are targeted, and in part to the focus in DDR studies on activities initiated within the first few hours after genotoxic insult.

Although USF is induced by UV-dependent phosphorylation in skin cells, its role in the immediate DDR appears limited35. However, our findings in lymphocytes show that DSBs can induce longterm changes in USF DNA binding activity26, suggesting that USF’s role in

DDR and cancer prevention may occur not during the P53-dominated immediate response to

54

DNA damage, but during the poorly-studied window in which cells must complete the repair response and either return to homeostasis or senesce and die.

To investigate USF’s role in DDR, we examined the transcriptional response of P53- deficient M12 germinal B cell lymphoma cell lines depleted of both USF1 and USF2 to sublethal ionizing radiation. The resulting data reveal many interesting elements of the role of USF. Despite USF’s widespread DNA binding profile21, only 940 DEG probes were enriched in untreated USFKD cells relative to control cells transduced with a scrambled shRNA (Fig. 4), suggesting that either the USF proteins play a limited role in directing gene expression in the absence of stress, or that sufficient levels of USF1 or USF2 proteins remained in USFKD cells to mediate their regulatory functions. Annotation of DEGs identified in untreated USFKD cells to Gene Ontology biological processes reveals a strong association with immune-related functions (Fig. 6), including lymphocyte proliferation

(NFB family genes, Aicda, the common gamma chain gene Il2rg), antigen recognition

(TLR and MyD88 genes), and immune cell activation (Btk, Fyn, Blk, Blnk). Strikingly, of these genes, only Il2rg has previously been shown to require USF activity153. Given that USF depletion altered the expression of a number of additional transcription factors implicated in immune activity (e.g. Rela and Nfkb1 which the P65 and P50 subunits of NFB, the

IB gene Nfkbia, Nfatc1 and Stat5a, among others), the impact of USF depletion on expression of immune-related genes would be expected to include a mixture of genes like

Il2rg that are directly regulated by USF and genes that are indirectly regulated by USF via its control of their respective transcription factors. In that regard, this study provides a rich

55

source of material for future studies to map both USF’s direct and indirect regulatory cascades. Irrespective of association with promoters of the genes listed above, our study clearly demonstrates that the expression of immune-related genes is particularly sensitive to reductions in USF activity. Since a variety of stimuli from DNA damage to metabolic stress have been shown to at least transiently alter USF activity through phosphorylation154, our data suggest that USF may be a critical unappreciated transcriptional regulator of the immune system.

As noted above, the immediate transcriptional response to DNA damage is dominated by P53, and is focused on pausing the cell cycle and mobilizing DNA repair machinery to the site(s) of DNA lesion155. In P53-deficient cells like M12148, survival after DNA damage is mediated by P38MAPK/MK2156. As expected, transcriptional response to treatment with 5

Gy IR was strongly impacted by P53-deficiency, and argues against a widespread impact of

USF depletion on the immediate DDR. However, unlike the transcription profiles of untreated USFKD samples which clustered together with sshRNA control samples, the transcription profiles of individual USFKD samples one day after IR were distinct from the controls (Fig. 4). This suggests that USF does in fact play a role, even if limited, in the early response to DNA damage. However, a more sensitive system (e.g. a system in which USF1 and USF2 are fully deleted) may be required in order to identify which genes are most impacted by loss of USF immediately following genotoxic stress.

Our current cell line showed a much stronger depletion of USF1 than USF2 (Fig. 3), despite screening a variety of clones. It is possible that loss of USF2 dramatically reduces

56

cell viability, in which case our selection strategy for stable knockdown clones would bias against cells lacking USF2. Indeed, USF2-deficient mice show marked decreases in lifespan51, and erythroid stem cells expressing a USF dominant negative transgene fail to progress through erythropoiesis52. We do not however expect that USF2 deficiency would prevent growth of the transformed M12 B cell line. Indeed, no lymphocyte defect was reported for USF2-deficient mice. A more likely explanation for the persistence of USF2 in our knockdown clones is the apparent positive impact USF1 deficiency has on USF2 expression51, which would be expected to limit the effectiveness of our USF2 RNAi targeting, and again supporting the need for a gene deletion approach to removing USF activity.

By far, the most striking impact of USF depletion was on the transcriptional profile of cells one week after they had been irradiated, well after most analyses of the DNA damage response. Although microarray identified over 5000 DEG probes in both USFKD and sshRNA RNAs seven days after IR when contrasted with untreated controls, the Gene

Ontology profiles of the two cell types were markedly different. Control cells up- or down- regulated genes associated with DDR, apoptosis and chromatin regulation, whereas USF- depleted cells failed to alter expression of these genes, and instead up- or down-regulated expression of genes associated with immune development and function well beyond the number of genes seen in untreated USFKD cells.

As noted above, contrasts between cells 1 day after irradiation and genotype-matched untreated cells revealed few DEGs that met 1.5-fold difference in expression and <0.05 FDR.

57

However, contrasting the 1 day post-IR and 7 day post-IR sshRNA controls does reveal two critical groups of biological processes that differentially expressed during the DDR response and that are not seen in matched USF-depleted cells (Fig. 6). Specifically, genes associated with the response to ionizing radiation and chromatin regulation are altered between 1 and 7 days post-IR in sshRNA cells, but not between 7 days post-IR and untreated sshRNA, indicating that these genes are altered during the initial response to DNA damage, but then return to untreated levels within 7 days. In contrast, genes more broadly associated with the

DNA damage response and with ribosome development and function are differentially expressed when comparing 7 day post-IR sshRNA profiles with untreated controls, but not when comparing 7 day post-IR with 1 day post-IR sshRNA profiles, indicating that these gene expression responses are activated early, but persist a week after irradiation. DEGs associated with both of these short-term and long-term transcriptional responses to IR are absent when USF is depleted, accounting for the paucity of DEGs in 1 day post-IR USFKD cells when contrasted with untreated knockdown cells and part of the divergence seen between the two cell types at 7 days post-IR. Differential expression of genes associated with cell cycle regulation, DNA replication and ribosomal function seven days after irradiation is much less sensitive to USF depletion. Taken together, the altered expression of genes involved in immune development and function, and the failure to alter expression of genes involved in the response to genotoxic damage clearly support a critical role for USF proteins in regulating the transcriptional response to DNA trauma.

58

Beyond the evidence that USF regulates long-term transcriptional responses to DNA stress, our findings particularly support a critical and heretofore unappreciated role for USF in regulating the expression of two sets of genes central to lymphocyte function: Aicda and the NFκB family of transcription factors. The AID protein is responsible for initiating SHM and CSR. If aberrantly expressed, AID is known to have off-target effects, ranging from point mutations in genes like Bcl6 and Myc141 or translocations that fuse Igh and cMyc (the cause of Burkitt’s Lymphoma) due to CSR-dependent DSBs119.

We see that AID is overexpressed in USF-depleted M12 cells relative to sshRNA controls, irrespective of radiation treatment. Past overexpression studies have shown that induction of AID is sufficient to cause an increase in both off-target SHM and CSR119. Yet, our experiments showed no evidence of SHM in the Vk region of USF-depleted cells. The lack of SHM may be because AID requires the presence of an inducer activity, provided in

SHM model systems by treatment with LPS, IL-4, or CD40 ligand93,157. M12 cells have been shown to support SHM in one study141. However, even in that case, SHM was only observed when cells were cultured in medium supplemented with specific batches of fetal calf serum, suggesting that SHM required inducing activity provided by the FBS. Work is underway in our laboratory to determine if elevated AID expression leads to SHM following cytokine treatment. Likewise, we are working to determine if CSR is upregulated in cytokine- stimulated USF-depleted cells. It is intriguing that, outside of germinal center B cells, elevated AID expression is strongly associated with cancer. Its inappropriate upregulation in precursor B cells has been linked to lymphoma114,115,158, while non-lymphoid cancers like

59

breast cancer frequently upregulate AID129. Indeed, mutations in key cancer genes including

Tp53, Apc, and Src among many others bear the hallmark of AID-mediated mutation159, suggesting that upregulation of AID may be a key step in tumorigenesis. ENCODE ChIP-seq analyses160 of USF2 binding in two separate mouse germinal center B cell lines, CH12X and

MEL, indicate two peaks of USF2 binding within the repressor region of Aicda, flanking exon 3 (Fig. 10). These peaks match consensus and near-consensus USF DNA binding sites.

Work is underway in the laboratory to determine a) whether these putative USF binding sites are associated with USF homo- or heterodimers in vivo, and b) whether they contribute to regulation of Aicda expression.

Figure 10. Schematic depiction of the mouse Aicda gene. Exons (filled boxes) and

previously mapped transcriptional enhancer, promoter and repressor elements are indicated, along with transcription factors implicated in Aicda repression. The approximate

positioning of two conserved USF binding sites and their sequences are shown in red. Nucleotide numbering is shown relative to the transcription start site.

60

No family of transcription factors is more closely tied to immune regulation than the

NFκB proteins and their inhibitors. Antigen receptor and TLR4 signaling in particular direct the proteolytic degradation of Iκbα and nuclear translocation of p65 RelA, where it forms the canonical p65p50 NFκB heterodimer161. Among the first targets of NFκB transcriptional regulation is the IκBα gene Nfkbia, whose expression leads to resequestration of p65 in the cytoplasm. It is striking then that USFKD cells show high levels of expression of both Nfkbia and Rela genes. ENCODE ChIP-seq analyses160 of USF2 binding show robust enrichment in the promoter regions of Rela, Nfkbia and Nfkb1. It is not yet known if USF1 is also enriched at the promoter regions of these genes, but the USF2 binding peaks overlap those of cMYC and MAX, suggesting that either MYC/MAX or USF dimers associate with NFκB gene promoters. As such, upregulation of the NFκB genes in USF-depleted cells would be consistent with the competition between activating MYC and repressive USF binding seen at other genes22. Beyond regulating NFκB at the level of protein translocation, the sensitivity of

USF to stress signaling through kinases including PKA, PKC and p38MAPK154 would provide a compelling mechanism to ensure NFκB protein levels in activated cells. Work in the laboratory is underway to define the role of USF in regulating these and other NFκB family members at the level of gene expression.

In conclusion, the findings of my thesis study support our hypothesis that USF mediates a longterm transcriptional response to DNA damage. In cells depleted of both USF1 and USF2, genes associated with leukocyte development and function including such critical genes as Aicda, Rela, Nfkbia,Il2rg and the Tlr genes are disregulated, even in the absence of

61

DNA stress. However, one week after DNA stress, USF-depleted cells show widespread failure to properly regulate genes associated with DNA repair, chromatin regulation and apoptosis, suggesting that USF normally functions in the pathway that safeguards the genome as cells recover from DNA damage. My studies have identified a number of key questions that must be solved as we improve our understanding of cancer, and lay the foundation that is now guiding mechanistic investigations in the laboratory.

62

REFERENCES 1. Pieper, K., Grimbacher, B. & Eibel, H. B-cell biology and development. Journal of

Allergy and Clinical Immunology 131, 959–971 (2013).

2. Schatz, D. G. & Ji, Y. Recombination centres and the orchestration of V(D)J

recombination. Nat. Rev. Immunol. 11, 251–63 (2011).

3. Chen, H. T. et al. Response to RAG-mediated VDJ cleavage by NBS1 and gamma-

H2AX. Science 290, 1962–1965 (2000).

4. Celeste, A. et al. Histone H2AX phosphorylation is dispensable for the initial

recognition of DNA breaks. Nat. Cell Biol. 5, 675–679 (2003).

5. Nussenzweig, A. & Nussenzweig, M. C. Origin of Chromosomal Translocations in

Lymphoid Cancer. Cell 141, 27–38 (2010).

6. Buis, J. et al. Mre11 Nuclease Activity Has Essential Roles in DNA Repair and

Genomic Stability Distinct from ATM Activation. Cell 135, 85–96 (2008).

7. Karanjawala, Z. E., Grawunder, U., Hsieh, C. L. & Lieber, M. R. The nonhomologous

DNA end joining pathway is important for chromosome stability in primary

fibroblasts. Curr. Biol. 9, 1501–1504 (1999).

8. Stavnezer, J., Guikema, J. E. J. & Schrader, C. E. Mechanism and regulation of class

switch recombination. Annu. Rev. Immunol. 26, 261–92 (2008).

9. Kenter, A. L., Kumar, S., Wuerffel, R. & Grigera, F. AID hits the jackpot when

missing the target. Current Opinion in Immunology 39, 96–102 (2016).

10. Yamane, A. et al. Deep-sequencing identification of the genomic targets of the

cytidine deaminase AID and its cofactor RPA in B lymphocytes. Nat. Immunol. 12,

63

62–9 (2011).

11. Storb, U. & Stavnezer, J. Immunoglobulin genes: Generating diversity with AID and

UNG. Current Biology 12, (2002).

12. Revy, P. et al. Activation-Induced Cytidine Deaminase (AID) Deficiency Causes the

Autosomal Recessive Form of the Hyper-IgM Syndrome (HIGM2). Cell 102, 565–575

(2000).

13. Tong, P. & Wesemann, D. R. Molecular Mechanisms of IgE Class Switch

Recombination. Current topics in microbiology and immunology 388, 21–37 (2015).

14. Carthew, R. W., Chodosh, L. A. & Sharp, P. A. An RNA polymerase II transcription

factor binds to an upstream element in the adenovirus major late promoter. Cell 43,

439–448 (1985).

15. Miyamoto, N. G., Moncollin, V., Egly, J. M. & Chambon, P. Specific interaction

between a transcription factor and the upstream element of the adenovirus-2 major late

promoter. EMBO J. 4, (1985).

16. Sawadogo, M. & Roeder, R. G. Interaction of a gene-specific transcription factor with

the adenovirus major late promoter upstream of the TATA box region. Cell 43, 165–

175 (1985).

17. Sawadogo, M., Van Dyke, M. W., Gregor, P. D. & Roeder, R. G. Multiple forms of

the human gene-specific transcription factor USF. I. Complete purification and

identification of USF from HeLa cell nuclei. J. Biol. Chem. 263, 11985–11993 (1988).

18. Groenen, P. M. et al. Structure, sequence, and localization of human

64

USF2 and its rearrangement in a patient with multicystic renal dysplasia. Genomics

38, 141–148 (1996).

19. Viollet, B. et al. Immunochemical characterization and transacting properties of

upstream stimulatory factor isoforms. J. Biol. Chem. 271, 1405–1415 (1996).

20. Chen, N. et al. Tumor-suppression function of transcription factor USF2 in prostate

carcinogenesis. Oncogene 25, 579–87 (2006).

21. Rada-Iglesias, A. et al. Whole-genome maps of USF1 and USF2 binding and histone

H3 acetylation reveal new aspects of promoter structure and candidate genes for

common human disorders. Genome Res. 18, 380–392 (2008).

22. Luo, X. & Sawadogo, M. Antiproliferative properties of the USF family of helix-loop-

helix transcription factors. Proc. Natl. Acad. Sci. U. S. A. 93, 1308–13 (1996).

23. Qyang, Y. et al. Cell-type-dependent activity of the ubiquitous transcription factor

USF in cellular proliferation and transcriptional activation. Mol.Cell Biol. 19, 1508–

1517 (1999).

24. Roy, a L. et al. Cloning of an Inr- and E-box-binding protein, TFII-I, that interacts

physically and functionally with USF1. EMBO J. 16, 7091–104 (1997).

25. Corre, S. et al. UV-induced expression of key component of the tanning process, the

POMC and MC1R genes, is dependent onp the p-38-activated upstream stimulating

factor-1 (USF-1). J. Biol. Chem. 279, 51226–51233 (2004).

26. Stone, J. L. et al. DNA double-strand breaks relieve USF-mediated repression of Dβ2

germline transcription in developing thymocytes. J. Immunol. 188, 2266–75 (2012).

65

27. Boyd, K. & Farnham, P. Myc versus USF: discrimination at the cad gene is

determined by core promoter elements. Mol. Cell. Biol. 17, 2529–37 (1997).

28. Murre, C., McCaw, P. S. & Baltimore, D. A new DNA binding and dimerization motif

in immunoglobulin enhancer binding, daughterless, MyoD, and myc proteins. Cell 56,

777–783 (1989).

29. Osman, I. et al. Chromosome 16 in primary prostate cancer: a microsatellite analysis.

Int J Cancer 71, 580–584 (1997).

30. Lose, F. et al. Genetic Association of the KLK4 Locus with Risk of Prostate Cancer.

PLoS One 7, 1–14 (2012).

31. Horbach, T. et al. GSK3b-dependent phosphorylation alters DNA binding,

transactivity and half-life of the transcription factor USF2. PLoS One 9, (2014).

32. Chen, N. et al. Tumor-suppression function of transcription factor USF2 in prostate

carcinogenesis. Oncogene 25, 579–87 (2006).

33. Kivinen, A., Patrikainen, L., Kurkela, R., Porvari, K. & Vihko, P. USF2 Is Connected

to GAAAATATGATA Element and Associates with Androgen Receptor-Dependent

Transcriptional Regulation in Prostate. Prostate 59, 190–202 (2004).

34. Ismail, P. M., Lu, T. & Sawadogo, M. Loss of USF transcriptional activity in breast

cancer cell lines. Oncogene 18, 5582–91 (1999).

35. Bouafia, A. et al. p53 requires the stress sensor USF1 to direct appropriate cell fate

decision. PLoS Genet. 10, e1004309 (2014).

36. Pawar, S. A., Szentirmay, M. N., Hermeking, H. & Sawadogo, M. Evidence for a

66

cancer-specific switch at the CDK4 promoter with loss of control by both USF and c-

Myc. Oncogene 23, 6125–6135 (2004).

37. Vaulont, S. et al. USF-1 Is Critical for Maintaining Genome Integrity in Response to

UV-Induced DNA Photolesions. PLoS Genetics 8, e1002470 (2012).

38. Crouch, E. E. et al. Regulation of AID expression in the immune response. J. Exp.

Med. 204, 1145–56 (2007).

39. Ramirez, A. B. et al. Use of a single-chain antibody library for ovarian cancer

biomarker discovery. Mol. Cell. Proteomics 9, 1449–60 (2010).

40. Jaiswal, A. S. & Narayan, S. Upstream stimulating factor-1 (USF1) and USF2 bind to

and activate the promoter of the adenomatous polyposis coil (APC) tumor suppressor

gene. J. Cell. Biochem. 81, 262–277 (2001).

41. Wu, K., Jiang, S. W. & Couch, F. J. p53 mediates repression of the BRCA2 promoter

and down-regulation of BRCA2 mRNA and protein levels in response to DNA

damage. J. Biol. Chem. 278, 15652–15660 (2003).

42. Goueli, B. S. & Janknecht, R. Regulation of telomerase reverse transcriptase gene

activity by upstream stimulatory factor. Oncogene 22, 8042–8047 (2003).

43. Chang, J. T.-C., Yang, H.-T., Wang, T.-C. V & Cheng, A.-J. Upstream stimulatory

factor (USF) as a transcriptional suppressor of human telomerase reverse transcriptase

(hTERT) in oral cancer cells. Mol. Carcinog. 44, 183–92 (2005).

44. Wu, a, Ichihashi, M. & Ueda, M. Correlation of the expression of human telomerase

subunits with telomerase activity in normal skin and skin tumors. Cancer 86, 2038–44

67

(1999).

45. Ocejo-Garcia, M. et al. Roles for USF-2 in lung cancer proliferation and bronchial

carcinogenesis. J. Pathol. 206, 151–9 (2005).

46. Tiruppathi, C. et al. The transcription factor DREAM represses the deubiquitinase

A20 and mediates inflammation. Nat. Immunol. 15, 239–47 (2014).

47. Providence, K. M. Epithelial monolayer wounding stimulates binding of USF-1 to an

E-box motif in the plasminogen activator inhibitor type 1 gene. J. Cell Sci. 115, 3767–

3777 (2002).

48. Read, M. L., Clark, a R. & Docherty, K. The helix-loop-helix transcription factor USF

(upstream stimulating factor) binds to a regulatory sequence of the human gene

enhancer. Biochem. J. 295 ( Pt 1, 233–237 (1993).

49. Yuyama, M. & Fujimori, K. Suppression of adipogenesis by valproic acid through

repression of USF1-activated fatty acid synthesis in adipocytes. Biochem. J. 459, 489–

503 (2014).

50. Elbein, S. C., Hoffman, M. D., Teng, K., Leppert, M. F. & Hasstedt, S. J. A genome-

wide search for type 2 diabetes susceptibility genes in Utah Caucasians. Diabetes 48,

1175–1182 (1999).

51. Sirito, Lin, Q., Deng, J. M., Behringer, R. R. & Sawadogo, M. Overlapping roles and

asymmetrical cross-regulation of the USF proteins in mice. Proc. Natl. Acad. Sci. U. S.

A. 95, 3758–3763 (1998).

52. Liang, S. Y. et al. Defective erythropoiesis in transgenic mice expressing dominant-

68

negative upstream stimulatory factor. Mol. Cell. Biol. 29, 5900–5910 (2009).

53. Chang, L. A., Smith, T., Pognonec, P., Roeder, R. G. & Murialdo, H. Identification of

USF as the ubiquitous murine factor that binds to and stimulates transcription from the

immunoglobulin ??2 chain promoter. Nucleic Acids Res. 20, 287–293 (1992).

54. Galibert, M. D., Boucontet, L., Goding, C. R. & Meo, T. Recognition of the E-C4

element from the C4 complement gene promoter by the upstream stimulatory factor-1

transcription factor. J. Immunol. 159, 6176–83 (1997).

55. Gobin, S. J., van Zutphen, M., Woltman, A. M. & van den Elsen, P. J. Transactivation

of classical and nonclassical HLA class I genes through the IFN-stimulated response

element. J. Immunol. 163, 1428–34 (1999).

56. Howcroft, T. K. et al. Upstream stimulatory factor regulates major histocompatibility

complex class I gene expression: the U2DeltaE4 splice variant abrogates E-box

activity. Mol. Cell. Biol. 19, 4788–97 (1999).

57. Wallin, J. J. et al. B cell-specific activator protein prevents two activator factors from

binding to the immunoglobulin J chain promoter until the antigen-driven stages of B

cell development. J. Biol. Chem. 274, 15959–15965 (1999).

58. McMillan, R. E. & Sikes, M. L. Differential activation of dual promoters alters Dbeta2

germline transcription during thymocyte development. J. Immunol. 180, 3218–28

(2008).

59. Pognonec, P. et al. Cross-family interaction between the bHLHZip USF and bZip Fra1

proteins results in down-regulation of AP1 activity. Oncogene 14, 2091–8 (1997).

69

60. Pawlus, M. R., Wang, L., Ware, K. & Hu, C.-J. Upstream Stimulatory Factor 2 and

Hypoxia-Inducible Factor 2 (HIF2 ) Cooperatively Activate HIF2 Target Genes

during Hypoxia. Mol. Cell. Biol. 32, 4595–4610 (2012).

61. Frenkel, S., Kay, G., Nechushtan, H. & Razin, E. Nuclear translocation of upstream

stimulating factor 2 (USF2) in activated mast cells: a possible role in their survival. J

Immunol 161, 2881–2887 (1998).

62. Huang, Y. H. et al. Upstream stimulatory factor 2 is implicated in the progression of

biliary atresia by regulation of hepcidin expression. J. Pediatr. Surg. 43, 2016–2023

(2008).

63. Tsui, H. W., Hasselblatt, K., Martin, A., Mok, S. C. & Tsui, F. W. L. Molecular

mechanisms underlying SHP-1 gene expression. Eur. J. Biochem. 269, 3057–3064

(2002).

64. Wong, R. H. F. et al. A Role of DNA-PK for the Metabolic Gene Regulation in

Response to Insulin. Cell 136, 1056–1072 (2009).

65. Corre, S. & Galibert, M.-D. Upstream stimulating factors: highly versatile stress-

responsive transcription factors. Pigment Cell Res. 18, 337–48 (2005).

66. Park, B. J., Park, J. I., Byun, D. S., Park, J. H. & Chi, S. G. Mitogenic conversion of

transforming growth factor-beta1 effect by oncogenic Ha-Ras-induced activation of

the mitogen-activated protein kinase signaling pathway in human prostate cancer.

Cancer Res. 60, 3031–8 (2000).

67. Xiao, Q., Kenessey, A. & Ojamaa, K. Role of USF1 phosphorylation on cardiac alpha-

70

myosin heavy chain promoter activity. Am. J. Physiol. Heart Circ. Physiol. 283,

H213-9 (2002).

68. Satyanarayana, A. & Kaldis, P. Mammalian cell-cycle regulation: several Cdks,

numerous cyclins and diverse compensatory mechanisms. Oncogene 28, 2925–2939

(2009).

69. Cheung, E., Mayr, P., Coda-Zabetta, F., Woodman, P. G. & Boam, D. S. DNA-

binding activity of the transcription factor upstream stimulatory factor 1 (USF-1) is

regulated by cyclin-dependent phosphorylation. Biochem. J. 344 Pt 1, 145–152

(1999).

70. Hydbring, P. & Larsson, L. G. Tipping the balance: Cdk2 enables Myc to suppress

senescence. Cancer Research 70, 6687–6691 (2010).

71. Sayasith, K., Lussier, J. G. & Sirois, J. Role of upstream stimulatory factor

phosphorylation in the regulation of the prostaglandin G/H synthase-2 promoter in

granulosa cells. J. Biol. Chem. 280, 28885–28893 (2005).

72. Maekawa, T., Sudo, T., Kurimoto, M. & Ishii, S. USF-related transcription factor,

HIV-TF1, stimulates transcription of human immunodeficiency virus-1. Nucleic Acids

Res. 19, 4689–4694 (1991).

73. Berger, A., Cultaro, C. M., Segal, S. & Spiegel, S. The potent lipid mitogen

sphingosylphosphocholine activates the DNA binding activity of upstream stimulating

factor (USF), a basic helix-loop-helix-zipper protein. Biochim. Biophys. Acta - Lipids

Lipid Metab. 1390, 225–236 (1998).

71

74. McCubrey, J. A. et al. GSK-3 as potential target for therapeutic intervention in cancer.

Oncotarget 5, 2881–2911 (2014).

75. Corre, S. et al. Target gene specificity of USF-1 is directed via p38-mediated

phosphorylation-dependent acetylation. J. Biol. Chem. 284, 18851–18862 (2009).

76. Wang, X., Fan, M., Kalis, S., Wei, L. & Scharff, M. D. A source of the single-stranded

DNA substrate for activation-induced deaminase during somatic hypermutation. Nat.

Commun. 5, 4137 (2014).

77. Nambu, Y. et al. Transcription-coupled events associating with immunoglobulin

switch region chromatin. Sci. (New York, NY) 302, 2137–2140 (2003).

78. Liu, M. M. . et al. Two levels of protection for the B cell genome during somatic

hypermutation. Nature 451, 841–845 (2008).

79. Maul, R. W. & Gearhart, P. J. AID and somatic hypermutation. Adv. Immunol. 105,

159–191 (2010).

80. Peters, A. & Storb, U. Somatic hypermutation of immunoglobulin genes is linked to

transcription initiation. Immunity 4, 57–65 (1996).

81. Xue, K., Rada, C. & Neuberger, M. S. The in vivo pattern of AID targeting to

immunoglobulin switch regions deduced from mutation spectra in msh2-/- ung-/-

mice. J. Exp. Med. 203, 2085–94 (2006).

82. Wood, R. D. & Doubli??, S. DNA polymerase ?? (POLQ), double-strand break repair,

and cancer. DNA Repair 44, 22–32 (2016).

83. Matthews, A. J., Zheng, S., DiMenna, L. J. & Chaudhuri, J. Regulation of

72

immunoglobulin class-switch recombination: Choreography of noncoding

transcription, targeted DNA deamination, and long-range DNA repair. Adv. Immunol.

122, 1–57 (2014).

84. Alt, F. W., Zhang, Y., Meng, F. L., Guo, C. & Schwer, B. Mechanisms of

programmed DNA lesions and genomic instability in the immune system. Cell 152,

417–429 (2013).

85. Wang, L., Wuerffel, R., Feldman, S., Khamlichi, A. A. & Kenter, A. L. S region

sequence, RNA polymerase II, and histone modifications create chromatin

accessibility during class switch recombination. J. Exp. Med. 206, 1817–30 (2009).

86. Rajagopal, D. et al. Immunoglobulin switch mu sequence causes RNA polymerase II

accumulation and reduces dA hypermutation. J. Exp. Med. 206, 1237–1244 (2009).

87. Morgan, H. D., Dean, W., Coker, H. A., Reik, W. & Petersen-Mahrt, S. K. Activation-

induced cytidine deaminase deaminates 5-methylcytosine in DNA and is expressed in

pluripotent tissues: Implications for epigenetic reprogramming. J. Biol. Chem. 279,

52353–52360 (2004).

88. Bhutani, N. et al. Reprogramming towards pluripotency requires AID-dependent DNA

demethylation. Nature 463, 1042–7 (2010).

89. Popp, C. et al. Genome-wide erasure of DNA methylation in mouse primordial germ

cells is affected by AID deficiency. Nature 463, 1101–5 (2010).

90. Rai, K. et al. DNA Demethylation in Zebrafish Involves the Coupling of a Deaminase,

a Glycosylase, and Gadd45. Cell 135, 1201–1212 (2008).

73

91. Meyers, G. et al. Activation-induced cytidine deaminase (AID)is required for B-cell

tolerance in humans. Proc. Natl. Acad. Sci. U. S. A. 108, 11554–11559 (2011).

92. Kuraoka, M. et al. Activation-induced cytidine deaminase mediates central tolerance

in B cells. Proc. Natl. Acad. Sci. U. S. A. 108, 11560–5 (2011).

93. Muramatsu, M. et al. Class switch recombination and hypermutation require

activation-induced cytidine deaminase (AID), a potential RNA editing enzyme. Cell

102, 553–63 (2000).

94. Kieffer-Kwon, K. R. et al. Interactome maps of mouse gene regulatory domains reveal

basic principles of transcriptional regulation. Cell 155, 1507–1520 (2013).

95. Castigli, E. et al. TACI and BAFF-R mediate isotype switching in B cells. J. Exp.

Med. 201, 35–39 (2005).

96. Tran, T. H. et al. B cell-specific and stimulation-responsive enhancers derepress Aicda

by overcoming the effects of silencers. Nat. Immunol. 11, 148–154 (2010).

97. Huong, L. T. et al. In Vivo Analysis of Aicda Gene Regulation: A Critical Balance

between Upstream Enhancers and Intronic Silencers Governs Appropriate Expression.

PLoS One 8, (2013).

98. de Yébenes, V. G. et al. miR-181b negatively regulates activation-induced cytidine

deaminase in B cells. J. Exp. Med. 205, 2199–206 (2008).

99. Dorsett, Y. et al. MicroRNA-155 Suppresses Activation-Induced Cytidine Deaminase-

Mediated Myc-Igh Translocation. Immunity 28, 630–638 (2008).

100. Teng, G. et al. MicroRNA-155 Is a Negative Regulator of Activation-Induced

74

Cytidine Deaminase. Immunity 28, 621–629 (2008).

101. Klemm, L. et al. The B Cell Mutator AID Promotes B Lymphoid Blast Crisis and

Drug Resistance in Chronic Myeloid Leukemia. Cancer Cell 16, 232–245 (2009).

102. Aoufouchi, S. et al. Proteasomal degradation restricts the nuclear lifespan of AID. J.

Exp. Med. 205, 1357–1368 (2008).

103. Orthwein, A. et al. Regulation of activation-induced deaminase stability and antibody

gene diversification by Hsp90. J. Exp. Med. 207, 2751–2765 (2010).

104. Hasler, J., Rada, C. & Neuberger, M. S. Cytoplasmic activation-induced cytidine

deaminase (AID) exists in stoichiometric complex with translation elongation factor

1 (eEF1A). Proc. Natl. Acad. Sci. 108, 18366–18371 (2011).

105. Xu, Z. et al. 14-3-3 adaptor proteins recruit AID to 5’-AGCT-3’-rich switch regions

for class switch recombination. Nat. Struct. Mol. Biol. 17, 1124–35 (2010).

106. Nowak, U., Matthews, A. J., Zheng, S. & Chaudhuri, J. The splicing regulator PTBP2

interacts with the cytidine deaminase AID and promotes binding of AID to switch-

region DNA. Nat. Immunol. 12, 160–6 (2011).

107. Pavri, R. et al. NIH Public Access. October 143, 122–133 (2011).

108. Chaudhuri, J., Khuong, C. & Alt, F. W. Replication protein A interacts with AID to

promote deamination of somatic hypermutation targets. Nature 430, 992–998 (2004).

109. Basu, U. et al. The RNA exosome targets the AID cytidine deaminase to both strands

of transcribed duplex DNA substrates. Cell 144, 353–363 (2011).

110. Vuong, B. Q. et al. Specific recruitment of protein kinase A to the immunoglobulin

75

locus regulates class-switch recombination. Nat. Immunol. 10, 420–6 (2009).

111. Robbiani, D. F. et al. AID Produces DNA Double-Strand Breaks in Non-Ig Genes and

Mature B Cell Lymphomas with Reciprocal Chromosome Translocations. Mol. Cell

36, 631–641 (2009).

112. Küppers, R. Mechanisms of B-cell lymphoma pathogenesis. Nat. Rev. Cancer 5, 251–

62 (2005).

113. Chaudhuri, J. & Alt, F. W. Class-switch recombination: interplay of transcription,

DNA deamination and DNA repair. Nat. Rev. Immunol. 4, 541–552 (2004).

114. Barlow, J. H. et al. Identification of early replicating fragile sites that contribute to

genome instability. Cell 152, 620–632 (2013).

115. Tsai, A. G. & Lieber, M. R. Mechanisms of chromosomal rearrangement in the human

genome. BMC Genomics 11 Suppl 1, S1 (2010).

116. Hu, J., Tepsuporn, S., Meyers, R. M., Gostissa, M. & Alt, F. W. Developmental

propagation of V(D)J recombination-associated DNA breaks and translocations in

mature B cells via dicentric . Proc. Natl. Acad. Sci. U. S. A. 111, 10269–

74 (2014).

117. Lin, C. et al. -Induced Chromosomal Proximity and DNA Breaks

Underlie Specific Translocations in Cancer. Cell 139, 1069–1083 (2009).

118. Matsumoto, Y. et al. Helicobacter pylori infection triggers aberrant expression of

activation-induced cytidine deaminase in gastric epithelium. Nat. Med. 13, 470–476

(2007).

76

119. Ramiro, A. R. et al. AID is required for c-myc/IgH chromosome translocations in

vivo. Cell 118, 431–438 (2004).

120. Gostissa, M. et al. Long-range oncogenic activation of Igh-c-myc translocations by the

Igh 3’ regulatory region. Nature 462, 803–807 (2009).

121. Osborne, C. S. et al. Myc dynamically and preferentially relocates to a transcription

factory occupied by Igh. PLoS Biol. 5, 1763–1772 (2007).

122. Spilianakis, C. G., Lalioti, M. D., Town, T., Lee, G. R. & Flavell, R. a.

Interchromosomal associations between alternatively expressed loci. Nature 435, 637–

645 (2005).

123. Chiarle, R. et al. Genome-wide translocation sequencing reveals mechanisms of

chromosome breaks and rearrangements in B cells. Cell 147, 107–119 (2011).

124. Robbiani, D. F. et al. AID Is Required for the Chromosomal Breaks in c-myc that

Lead to c-myc/IgH Translocations. Cell 135, 1028–1038 (2008).

125. Ramiro, A. R. et al. Role of genomic instability and p53 in AID-induced c-myc-Igh

translocations. Nature 440, 105–9 (2006).

126. Kovalchuk, A. L. et al. AID-deficient Bcl-xL transgenic mice develop delayed

atypical plasma cell tumors with unusual Ig/Myc chromosomal rearrangements. J.

Exp. Med. 204, 2989–3001 (2007).

127. Kovalchuk, A. L. et al. Mouse model of endemic Burkitt translocations reveals the

long-range boundaries of Ig-mediated oncogene deregulation. Proc. Natl. Acad. Sci.

U. S. A. 109, 10972–7 (2012).

77

128. Okazaki, I. -m. et al. Constitutive Expression of AID Leads to Tumorigenesis. J. Exp.

Med. 197, 1173–1181 (2003).

129. Muñoz, D. P. et al. Activation-induced cytidine deaminase (AID) is necessary for the

epithelial-mesenchymal transition in mammary epithelial cells. Proc. Natl. Acad. Sci.

U. S. A. 110, E2977-86 (2013).

130. Ciccia, A. & Elledge, S. J. The DNA Damage Response: Making It Safe to Play with

Knives. Molecular Cell 40, 179–204 (2010).

131. Janeway, C. A. J., Travers, P. & Walport, M. The generation of diversity in

immunoglobulins Virtually. Immunobiology: The Immune System in Health and

Disease. 5th edition. (2001).

132. Chen, B., Chen, X.-P., Wu, M.-S., Cui, W. & Zhong, M. Expressions of heparanase

and upstream stimulatory factor in hepatocellular carcinoma. Eur. J. Med. Res. 19, 45

(2014).

133. McMurray, H. R. & McCance, D. J. Human papillomavirus type 16 E6 activates

TERT gene transcription through induction of c-Myc and release of USF-mediated

repression. J Virol 77, 9852–9861 (2003).

134. Walhout, A. J. M., Gubbels, J. M., Bernards, R., Van Der Vliet, P. C. & Timmers, H.

T. M. C-Myc/Max heterodimers bind cooperatively to the E-box sequences located in

the first intron of the rat ornithine decarboxylase (ODC) gene. Nucleic Acids Res. 25,

1493–1501 (1997).

135. Baron, Y. et al. USF-1 is critical for maintaining genome integrity in response to UV-

78

induced DNA photolesions. PLoS Genet. 8, e1002470 (2012).

136. Corre, S. & Galibert, M. D. Upstream stimulating factors: Highly versatile stress-

responsive transcription factors. Pigment Cell Research 18, 337–348 (2005).

137. Corre, S. et al. Target gene specificity of USF-1 is directed via p38-mediated

phosphorylation-dependent acetylation. J. Biol. Chem. 284, 18851–62 (2009).

138. Chen, N., Pawar, S. A., Szentirmay, M. N. & Sawadogo, M. Transcription factor

USF2 suppresses the malignant phenotype of prostate cancer cells. 46, (2005).

139. Vallet, V. S. et al. Differential roles of upstream stimulatory factors 1 and 2 in the

transcriptional response of liver genes to glucose. J. Biol. Chem. 273, 20175–20179

(1998).

140. Park, S., Liu, X., Davis, D. R. & Sigmund, C. D. Gene trapping uncovers sex-specific

mechanisms for upstream stimulatory factors 1 and 2 in angiotensinogen expression.

Hypertension 59, 1212–1219 (2012).

141. Bhattacharya, P. et al. Identification of murine B cell lines that undergo somatic

hypermutation focused to A:T and G:C residues. Eur. J. Immunol. 38, 227–239

(2008).

142. Benjamini, Y. & Hochberg, Y. Controlling the false discovery rate: a practical and

powerful approach to multiple testing. J. R. Stat. Soc. Ser. B … 57, 289–300 (1995).

143. Huang, D. W., Sherman, B. T. & Lempicki, R. A. Bioinformatics enrichment tools:

Paths toward the comprehensive functional analysis of large gene lists. Nucleic Acids

Res. 37, 1–13 (2009).

79

144. Huang, D. W., Sherman, B. T. & Lempicki, R. A. Systematic and integrative analysis

of large gene lists using DAVID bioinformatics resources. Nat. Protoc. 4, 44–57

(2008).

145. Livak, K. J. & Schmittgen, T. D. Analysis of Relative Gene Expression Data Using

Real-Time Quantitative PCR and the 2−ΔΔCT Method. Methods 25, 402–408 (2001).

146. Galibert, M. D., Carreira, S. & Goding, C. R. The Usf-1 transcription factor is a novel

target for the stress-responsive p38 kinase and mediates UV-induced Tyrosinase

expression. EMBO J. 20, 5022–31 (2001).

147. Chang, J. T.-C., Yang, H.-T., Wang, T.-C. V & Cheng, A.-J. Upstream stimulatory

factor (USF) as a transcriptional suppressor of human telomerase reverse transcriptase

(hTERT) in oral cancer cells. Mol. Carcinog. 44, 183–92 (2005).

148. HOLLMANN, A., GONG, Q. & OWENS, T. CD40-Mediated Apoptosis in Murine B-

Lymphoma Lines Containing Mutated p53. Exp. Cell Res. 280, 201–211 (2002).

149. Solozobova, V., Rolletschek, A. & Blattner, C. Nuclear accumulation and activation of

p53 in embryonic stem cells after DNA damage. BMC Cell Biol. 10, 46 (2009).

150. Feijoo, C. et al. Activation of mammalian Chk1 during DNA replication arrest: A role

for Chk1 in the intra-S phase checkpoint monitoring replication origin firing. J. Cell

Biol. 154, 913–923 (2001).

151. Cable, P. L. A. et al. Novel consensus DNA-binding sequence for BRCA1 protein

complexes. Mol. Carcinog. 38, 85–96 (2003).

152. Wang, Y., Wang, B., Tong, J., Chang, B. & Wang, B. USF-1 genetic polymorphisms

80

confer a high risk of nonalcoholic fatty liver disease in Chinese population. 8, 2545–

2553 (2015).

153. Domashenko, A. D. et al. proliferation and engraftment potential of human

hematopoietic progenitor cells TAT-mediated transduction of NF-Ya peptide induces

the ex vivo proliferation and engraftment potential of human hematopoietic progenitor

cells. 116, 2676–2683 (2014).

154. Horbach, T., Götz, C., Kietzmann, T. & Dimova, E. Y. Protein kinases as switches for

the function of upstream stimulatory factors: Implications for tissue injury and cancer.

Front. Pharmacol. 6, 1–11 (2015).

155. Lakin, N. D. & Jackson, S. P. Regulation of p53 in response to DNA damage.

Oncogene 18, 7644–7655 (1999).

156. Eom, H.-J. & Choi, J. p38 MAPK Activation, DNA Damage, Cell Cycle Arrest and

Apoptosis As Mechanisms of Toxicity of Silver Nanoparticles in Jurkat T Cells.

Environ. Sci. Technol. 44, 8337–8342 (2010).

157. Park, S.-R. et al. HoxC4 binds to the promoter of the cytidine deaminase AID gene to

induce AID expression, class-switch DNA recombination and somatic hypermutation.

Nat. Immunol. 10, 540–550 (2009).

158. Duquette, M. L., Huber, M. D. & Maizels, N. G-rich proto-oncogenes are targeted for

genomic instability in B-cell lymphomas. Cancer Res. 67, 2586–2594 (2007).

159. Takaishi, S. & Wang, T. C. Providing AID to p53 mutagenesis. Nat. Med. 13, 404–

406 (2007).

81

160. de Souza, N. The ENCODE project. Nat. Methods 9, 1046–1046 (2012).

161. Wi, S. M., Park, J., Shim, J.-H., Chun, E. & Lee, K.-Y. Ubiquitination of ECSIT is

crucial for the activation of p65/p50 NF-κBs in Toll-like receptor 4 signaling. Mol.

Biol. Cell 26, 151–160 (2014).

82

APPENDICES

83

Appendix A