A Dissertation

entitled

Bromodomain Containing in Melanocyte Differentiation and Melanoma

by

Archit Rakeshkumar Trivedi

Submitted to the Graduate Faculty as partial fulfillment of the requirements for the

Doctor of Philosophy Degree in Biomedical Sciences

______Dr. Ivana de la Serna, Committee Chair

______Dr. William Maltese, Committee Member

______Dr. Kam Yeung, Committee Member

______Dr. Robert Trumbly, Committee Member

______Dr. James Willey, Committee Member

______Dr. Patricia R. Komuniecki, Dean, College of Graduate Studies

The University of Toledo

August 2015

Copyright 2015, Archit Rakeshkumar Trivedi

This document is copyrighted material. Under copyright law, no parts of this document may be reproduced without the expressed permission of the author.

An Abstract of

Bromodomain Containing Proteins in Melanocyte Differentiation and Melanoma

by

Archit Rakeshkumar Trivedi

Submitted to the Graduate Faculty as partial fulfillment of the requirements for the

Doctor of Philosophy Degree in Biomedical Sciences

The University of Toledo

August 2015

Bromodomains belong to a family of evolutionarily conserved modules originally found in proteins associated with and in nearly all nuclear acetyltransferases. modulate enzyme activities, protein assembly, and protein-protein interactions via with broad implications in a wide variety of cellular events, such as transcriptional activation and .

They are the sole protein domains which function as acetyl-lysine binding residues and are epigenetic readers which bind to the acetylated residues regulating chromatin structure and expression. We assessed two groups of bromodomain containing proteins with their relevance to melanocyte differentiation and melanoma.

In the first study, the function of bromodomain (BRD) and extra-C terminal domain (BET) protein family in modulating pigmentation and melanocyte differentiation was deduced. BET protein family members (BRD2, BRD3, BRD4, and BRDT) are epigenome readers that bind to acetyl-lysine residues on the tails of H3 and H4, and exert key roles at the interface between chromatin remodeling and transcriptional iii regulation. BRD4 is a BET protein which acts as a transcriptional and was initially known as mitotic associated protein (MCAP). BRD4 associates with acetylated chromatin throughout the cell cycle and regulates at targeted loci. It is implicated in cellular growth control, cell cycle progression and cancer development. However, the role of BRD4 in melanocyte differentiation has not been investigated.

In this study, we utilized an in vitro murine model of immortalized murine melanoblasts (melb-a cells) which can be differentiated by treatment with alpha-MSH.

Alpha-MSH upon binding to melanocortin 1 activates the microphthalmia- associated (MITF) which is the master regulator of melanocyte differentiation and melanoma . MITF in turn activates required for melanin production, proliferation, and survival. JQ1 is a cell permeable small molecule inhibitor of BET proteins which binds to BRD4 with the highest affinity and specificity, inhibiting its binding to the acetylated lysine residues. JQ1 has an excellent shape complementarity with the acetyl-lysine binding cavity of bromodomains of BET family.

Treating melb-a cells with JQ1 compromised MITF target , melanin synthesis, and reduced proliferation of melb-a cells as well as neonatal human epidermal melanocytes (NHEM) and melanoma cells. Furthermore, we determined that BRD4 occupied the promoters of MITF target genes that regulate pigmentation. Treatment with

JQ1 inhibited the binding of both BRD4 and MITF to these promoters. Moreover, treating the melb-a cells as well as NHEMs disrupts histone H3 lysine 4 trimethylation

(H3K4me3) which is an active mark of transcription. The effects on gene expression were recapitulated by siRNA mediated silencing of BRD4. We showed a physical iv interaction between the two proteins. Thus, we hypothesize that BRD4 plays a crucial role in melanocyte differentiation by interacting with MITF and can be a potential novel target for treatment of hyperpigmentation disorders and in melanoma.

In the second project, the role of bromodomain containing proteins of an ATP- dependent chromatin remodeling complex SWI/SNF were assessed. SWI/SNF complexes can be categorized as BAF (BRG1/BRM-associated factor) complex bearing either one of

ARID1A/ARID1B or PBAF (Polybromo-associated BAF) complex harboring ARID2,

BAF180 and BRD7 subunits based on purified complexes from mammalian cells. In particular, we investigated the role of PBAF SWI/SNF complex in the context of melanoma tumorigenicity.

Ultraviolet (UV)-radiation causes DNA damage and is implicated in the etiology of cutaneous melanoma, an aggressive malignancy that is notoriously chemoresistant.

SWI/SNF chromatin remodeling enzymes have critical functions in the regulation of gene expression. Importantly, PBAF specific SWI/SNF subunits are disrupted in a number of human cancers including melanoma. We were interested in assessing a potential tumor suppressive role of PBAF complexes in the context of melanocytes and melanoma cells.

We hypothesized that the PBAF specific SWI/SNF complex mediates tumor suppression by activating target genes in response to UV-radiation. ARID2 has been identified as a gene that is frequently mutated in melanoma. We surveyed a panel of melanoma cell lines and found that a subset of melanoma cells is deficient in the expression of the other

PBAF specific subunits, BAF180 and BRD7. Our data indicate that BRG1, BAF180,

BRD7, and ARID2 promote expression of multiple p53 target genes as well as cell cycle arrest in UV-irradiated melanocytes and melanoma cells. BRG1, the central ATPase in v

PBAF SWI/SNF complex promoted cell cycle arrest in melanoma cells. Thus, our work suggests that a specific configuration of the SWI/SNF complex may have a tumor suppressive role in melanocytes and that aberrant expression or mutations of SWI/SNF components may contribute to melanoma tumorigenicity.

vi

I would like to dedicate my dissertation research to my parents Mr. Rakeshkumar

Trivedi (Bapu) and Mrs. Hemangini Trivedi (Ma) for their unconditional love and blessings they have showered on me. I am where I am today in life is because of their upbringing and their sacrifices. I am indebted to them for believing in me and supporting my endeavors of pursuing doctoral research in the United States of America. I would also like to take this opportunity to thank my late uncle Mr. Bhavnesh Jani (Mama) and my cousin Mr. Vishrut Trivedi (Mota Bhai) for always being a source of inspiration. One of the reasons I pursued PhD is because they wanted to call me a Doctor.

Acknowledgements

First of all, I would like to thank my advisor Dr. Ivana de la Serna for providing me with the opportunity to pursue doctoral research in her lab. She is a true leader who inspires immensely with her work ethics and has been instrumental in steering my dissertation research in a proper direction. Her passion and sincerity towards science has been a constant source of motivation. I thank my advisory committee members for their guidance and words of wisdom which has helped immensely in shaping up my dissertation research. I acknowledge my past and present lab members for maintaining a cordial and conducive lab environment to thrive in research. Special thanks to Vinod and

Himangi for their friendships and efforts in making me familiarize with the techniques during my early days in the lab. I am indebted to Dr. Mithun Khattar and Dr. Gurpanna

Saggu for their unconditional support throughout my doctoral research. Special thanks to

Prabhat for being a great friend. The ride of graduate school would not have been so smooth without the most efficient administrative staff of our department. Lastly, I would like to thank my parents for giving me the independence to pursue a career of my choice and always being supportive of my ventures. I am indebted to them for everything they have done for me. I would never forget the University of Toledo where I found a friend who became my colleague and eventually my life partner. I am at lose of words to describe how lucky I am to have Dr. Shweta Aras in my life. I would also like to acknowledge my in-laws for their support and encouragement. This journey would have been incomplete without their love and affection.

Finally, Glory goes to God for everything I have in life.

viii

Table of Contents

Abstract ...... iii

Acknowledgements ...... viii

Table of Contents ...... ix

List of Tables ...... xi

List of Figures ...... xii

List of Abbreviations ...... xiii

List of Symbols ...... xiv

1 Introduction and Literature Review ...... 1

1.1 Melanocytes: Proliferation and Differentiation ...... 1

1.2 Chromatin Remodeling ...... 7

1.3 Bromodomains ...... 17

1.4 JQ1: A selective and potent inhibitor of BET family bromodomains ...... 30

1.5 SWI/SNF Enzymes and Melanoma ...... 33

2 Role of BRD4 in melanocyte differentiation ...... 37

2.1 Introduction ...... 37

2.2 Materials and Methods ...... 39

2.3 Results ...... 45

2.4 Discussion ...... 51

2.5 Figure Legends ...... 54

2.6 Figures ...... 63

ix

3 The PBAF complex promotes expression of p53 target genes in reponse to Ultraviolet radiation in melanoma ...... 71

3.1 Introduction ...... 71

3.2 Materials and Methods ...... 72

3.3 Results ...... 76

3.4 Discussion ...... 80

3.5 Figure Legends ...... 83

3.6 Figures ...... 89

4 Overall Discussion ...... 95

References ...... 106

x

List of Tables

1.1 BET Bromodomain inhibitors in cancer ...... 31

1.2 Bromodomain inhibitors in clinical development ...... 33

1.3 SWI/SNF mutations in cancer ...... 35

xi

List of Figures

1-1 Regulators of MITF ...... 5

1-2 MITF Targets ...... 6

1-3 DNA Packs into a Mitotic Chromosome...... 8

1-4 ATPases of SWI2/SNF2-related chromatin remodeling complexes ...... 9

1-5 SWI/SNF chromatin remodeling complexes ...... 13

1-6 Epigenetic Regulators ...... 14

1-7 Epigenetic Reader Domains ...... 17

1-8 Structure of Bromodomain ...... 19

1-9 BET Family Proteins ...... 29

1-10 Structure and selectivity of JQ1 ...... 31

xii

List of Abbreviations

Alpha-MSH...... Alpha- melanocortin stimulating hormone ARID...... AT-rich interactive domain

BAF………………………………………………………..BRG1/BRM associated factors BET……………………………………………..Bromodomain and extra terminal domain BRG1……………………………………………………………… Brahma related gene 1 BRM………………………………………………………………………………..Brahma

CHD……………………………………………………………………….. CREB/CBP/CREBBP…………………….Cyclic-AMP response element-binding protein CTM………………………………………………………………………C-terminal motif

DCT……………………………………………………………...Dopachrome tautomerase

HAT……………………………………………………………...Histone acetyltransferase HDAC………………………………………………………………... Histone deacetylase HMT…………………………………………………………….Histone methyltransferase

MCAP…………………………………………….Mitotic chromosome-associated protein MC1R…………………………………………………………….Melanocortin -1 receptor MITF……………………………………...Microphthalmia-associated transcription factor

PBAF………………………………………………………….Polybromo-associated BAF P-TEFb…………………………………………….Positive transcription elongation factor

SWI/SNF………………………………….Mating type switching/sucrose non fermenting

TRP1………………………………………………………….Tyrosinase related protein 1 TYR……………………………………………………………………………..Tyrosinase

UVR…………………………………………………………………..Ultra violet radiation

xiii

List of Symbols

α- Alpha β- Beta κ- Kappa ε- Epsilon γ- Gamma

xiv

Chapter 1

Introduction and Literature Review

1.1 Melanocytes: Proliferation and Differentiation

Melanocytes are neural-crest derived cells located in the basal layer of epidermis that produce the pigment melanin [1, 2]. Undifferentiated precursors of melanocytes are called melanoblasts, which are derived from cells in embryonic skin [2, 3].

Melanoblasts are multipotent and are believed to be differentiated into melanocytes only when it exhibits pigmentation and shows structural characteristics such as dendrite formation and small granular structures known as melanosomes which contain melanin

[4, 5]. Melanin is responsible for providing photo-protection and by absorbing UV radiation and transforming it to harmless heat [4]. Melanocytes produce two types of melanin: brownish-black eumelanin and reddish-yellow pheomelanin.

Eumelanin-containing melanosomes (eumelanosomes) are elliptical, with longitudinal pigment deposits in intraluminal fibrils whereas pheomelanin-containing melanosomes

(pheomelanosomes) are spherical, with granular pigment deposits in multivesicular bodies. Thus, differences in melanosome morphology depict the differences in the type of melanin synthesized [2]. Once melanin is produced, it is stored in the melanosomes, which are transferred from melanocytes to keratinocytes, thus giving skin its 1

characteristic color. Melanoblast/melanocyte proliferation and differentiation are regulated by tissue microenvironment, especially by keratinocytes, which synthesize endothelins, steel factor, hepatocyte growth factor, leukemia inhibitory factor and granulocyte-macrophage colony-stimulating factor. Alpha-melanocyte stimulating hormone also stimulates melanocyte differentiation [2]. Three major enzymes regulate the melanin synthesis in melanocytes. The melanocyte-specific enzyme, tyrosinase

(TYR), is required for the rate-limiting step of melanin synthesis while two other members of the tyrosinase family, tyrosinase related protein 1 (TRP1) and dopachrome tautomerase (DCT), regulate the type of melanin synthesized [6, 7]. The expression of these enzymes is highly specific to melanocytes and hence their expression is often utilized as indicators of melanocyte differentiation. Thus, melanocyte proliferation and differentiation are not only regulated by genes encoding typical growth factors and their receptors but also by genes classically known for their role in pigment formation.

Disruption of normal melanocyte function can lead to skin cancer and other congenital pigmentary disorders such as Waardenburg syndrome and Tietz syndrome that affect the level of pigmentation [8-10]. Thus, understanding the mechanisms which regulate melonocyte differentiation and pigmentation could yield interesting cues regulating these processes and intervention of such players could be useful for treating pigment-related disorders.

Microphthalmia-associated transcription factor

The microphthalmia-associated transcription factor (MITF) is a basic helix-loop- helix transcription factor that binds to DNA recognition sequences called

E-box elements in the promoter region of the target genes with the consensus sequence 2

CA[T/C]GTG. Related family members of this group includes transcription factor E3

(TFE3), transcription factor EB (TFEB), and transcription factor EC (TFEC) [11-13].

MITF plays a crucial role in the development, differentiation, function and survival of melanocytes and hence, is a master regulator of melanocytes [14-16]. MITF binds to the regulatory regions of many genes which regulate pigmentation, as well as genes important for melanocyte survival, and activates their expression. In regards to activating pigmentation gene expression, our lab was the first to establish a novel requirement of SWI/SNF enzymes for MITF activity [17, 18]. It was reported that MITF physically interacts with BRG1, which is the core ATPase of the SWI/SNF complex and also with other subunits of the complex. Thus, pharmacological suppression of MITF expression can decrease pigmentation. Modulation of MITF expression or activity is therefore a useful strategy for treating hyperpigmentation disorders.

In addition to melanocytes, MITF is expressed in retinal pigment epithelium cells, as well as in certain non-pigmented cell lineages including osteoclast, mast cells [19], and cardiomyocytes. MITF is transcribed from several alternative promoters, giving rise to various splice isoforms. There are nine different isoforms of MITF known till date, which exhibit tissue-specific expression patterns. In melanocytes, the MITF-M isoform is selectively expressed [20].

An autosomal dominant condition known as Waardenburg syndrome type IIA occurs in humans expressing heterozygous mutation of MITF [21]. Such individuals exhibit a congenital white forelock and sensorineural deafness. In mice, complete lack of

MITF results in characteristics such as white fur, deafness, and small eyes

3

(microphthalmia) [22, 23]. Loss of pigmentation in such mice is due to the absence of melanocytes, and not the defect of melanin synthesis in viable melanocytes.

MITF has been shown to be a lineage-specific survival oncogene in melanoma, which is amplified in 5-20% of total human melanoma cases. In melanoma patients,

MITF amplification was associated with a decreased five-year survival [24]. MITF has also been shown to play an important role in controlling the proliferative, invasive and metastatic properties of melanoma cells [24-26].

MITF is regulated both transcriptionally and post-translationally. At the transcriptional level, MITF-M is regulated by transcription factors which include MITF itself, paired box gene 3 (PAX3), cAMP-responsive element binding protein (CREB), sex-determining region Y-box 10 (SOX10), lymphoid enhancer-binding factor 1

(LEF1/TCF), one cut domain 2 (ONECUT-2), and the mitogen-activated protein kinase

(MAPK) pathway [27].

At the post-translational level, MITF is known to be phosporylated [28, 29], sumoylated [30, 31], and ubiquitinated [32]. Phosphorylation of MITF occurs mainly by

MAPK, ribosomal S6 kinase (RSK), glycogen synthase kinase-3β (GSK3β) and p38. In melanocytes, c-kit activation causes MITF phosphorylation by ERK2 and RSK. Once phosphorylated, MITF induces recruitment of the transcriptional coactivator p300, and targets itself for ubiquitination and proteolysis. Protein inhibitor of activated STAT3

(PIAS3) mediates MITF sumoylation repressing its activity.

4

Figure 1-1 Regulators of MITF. Reprinted from Archives of Biochemistry and Biophysics, Vol 563: p. 28-34, Copyright (2014) with permission from Elsevier.

MITF is thought to play paradoxical roles in melanoma proliferation, survival, and invasion. Whereas lower levels of MITF correspond to increased motility and invasive capacity in melanoma, higher levels are thought to impart increased proliferative capacity [33]. High levels of POU3F2 (BRN2) transcription factor, a direct repressor of

MITF transcription is observed in melanoma cells that express low levels of MITF [25,

34].

MITF targets a variety of genes with wide implications in biological processes.

Potentially important functional transcriptional targets of MITF include regulators of cell cycle (CDK2, p21CIP1, p16INK4A), differentiation, survival and metastasis (BCL2,

BCL2A1, c-Met), cAMP levels (PDE4D3), metabolism, miRNA processing (DICER),

DNA repair, invasiveness, and pigmentation [27]. 5

Figure 1-2 MITF Targets. Reprinted from Archives of Biochemistry and Biophysics, Vol 563: p. 28-34, Copyright (2014) with permission from Elsevier.

DIAPH1 is a gene that encodes the diaphanous-related formin Dia1, which is involved in actin polymerization and coordinating actin cytoskeleton. MITF has been shown to regulate the expression of DIAPH1 to control invasiveness and proliferation.

Knock down of MITF leads to downregulation of Dia1, a cyclin-dependent kinase inhibitor 1B (p27Kip1)-dependent G1 arrest, reorganization of the actin cytoskeleton, and increased invasiveness. Proliferation occurs when moderate MITF levels lead to actin polymerization and suppressed p27Kip1 [26]. Senescence can be triggered in melanoma cells by prolonged MITF suppression [35].

MITF can interact with chromatin remodeling enzymes and thus promote changes in chromatin structure. “Epigenetic programming” through altered chromatin structure and histone modifications are believed to dictate how cells such as melanocytes specifically respond to certain environmental cues.

6

1.2 Chromatin Remodeling

Eukaryotic genomes are packed into condensed chromatin fibers to fit over a meter of DNA within a microscopic volume of the nucleus. Histones are the unique proteins which govern this compaction [36]. Pierre Chambon‟s laboratory was the first to name the nucleosome [37, 38], which is the basic building block of chromatin. It consists of 147 base pairs of DNA wrapped in a left-handed superhelix 1.7 times around a core histone octamer (two copies each of histones H2A, H2B, H3, and H4). Histone H1 serves as a linker histone between the adjacent nucleosomes and directs the path of DNA that makes up the chromatin fiber. This compacted structure is rendered inherently repressive to a major of cellular processes that require access to the DNA sequence. Each core histone comprises of two functional domains: “Histone-fold domain” which suffices histone-histone and histone-DNA contacts within the nucleosome and NH2-terminal and

COOH-terminal “tail” domain comprising of sites essential for posttranslational modifications such as acetylation, , phosphorylation, and ubiquitination [36].

The higher order chromatin structure can be divided into transcriptionally “active” euchromatin or “inactive” or “repressive” [39]. Thus, the chromatin structure needs to be remodeled for active gene transcription to take effect. Chromatin remodeling is the regulated alteration of the chromatin structure which is accomplished by covalent modification of the histones or by the action of ATP-dependent remodeling complexes [40].

7

FIGURE 1-3: DNA Packs into a Mitotic Chromosome. Reprinted from Life: The Science of Biology, Seventh Edition with permission from Sinaur Associates and W.H. Freeman & Co.

Chromatin Remodeling Enzymes

Chromatin-modifying complexes can be broadly classified into two major groups based on their modes of action as follows.

1) ATP-dependent complexes, which uses the energy from ATP hydrolysis to locally

disrupt or alter the association of histones with DNA.

2) Histone acetlytransferase (HAT) and histone deacetylase (HDAC) complexes,

which regulate the transcriptional activity of genes by determining the acetylation

levels of the amino-terminal domains of nucleosomal histones associated with

them.

8

Based on a unique subunit composition and the presence of a distinct ATPase, four

different classes of ATP-dependent chromatin remodeling enzymes can be recognized

namely SWI/SNF, ISWI, CHD, and Ino80 [41, 42].

ATPases of the four main families of ATP-dependent Swi2/Snf2-related chromatin remodeling complexes. All ATPase- domains comprise an amino-terminal DEXDc and a carboxy-terminal HELICc subdomain, shown in dark blue.

Figure 1-4 ATPases of SWI2/SNF2-related chromatin remodeling complexes. Reprinted from Biochimica et Biophysica Acta Vol 1681, Pg:59-73 Copyright (2005) with permission from Elsevier

The ISWI Group

These complexes contain the ISWI (Imitation Switch) protein as the ATPase

subunit and were purified from extracts using biochemical methods based on

their ability to disrupt and /or generate regularly spaced nucleosomal arrays [43]. The

ISWI-based complexes are much smaller and have fewer subunits as compared to the

SWI/SNF complexes containing between two and six peptides [40]. The presence of two

unique domains namely SANT and SLIDE is specific to this group of complexes. These

domains constitute a nucleosomal recognition unit which allows it to bind to DNA and

unmodified H4 tails. Acetylation of H4K16 has been shown to reduce the ISWI activity 9

and is targeted away from transcriptionally active chromatin. ACF (ATP-utilizing chromatin assembly and remodeling factor), NURF (Nucleosome-remodeling factor), and

CHRAC (Chromatin accessibility complex) are the most extensively studied members of this group. These complexes have been implicated in regulation of several functions such as heterochromatin formation, DNA replication, and transcriptional repression [43, 44].

The CHD Group

The Chromodomain-Helicase-DNA binding (CHD) family comprises of 9 members which have been further divided into 3 sub-families namely CHD1, Mi-2 and

CHD7. This family of chromatin remodeling complexes possesses two chromo domains at the amino terminal. Apart from chromatin remodeling, this complexes also possess deacetylase activites presumably by means of HDAC subunits [43]. The CHD members can function as monomers or associate with other factors to form complexes, of which the best characterized is the Nucleosome remodeling and deacetylase (NURD) complex.

It has been suggested that Mi-2 could be recruited to specific genes by repressors [45, 46] implicating this complex in transcriptional repression and as a regulator of genomic stability [41, 43].

The Ino80 Group

The ATPase domain of the Ino80 (Inositol requiring 80) family of chromatin- modifying complexes is referred to as the split ATPase domain due to a split in the middle due to the presence of a spacer region [47]. This complex exhibits DNA helicase activity through Pontin (Rvb1) and Reptin (Rvb2) and absence of the activity of either of them leads to abrogation of the activity of Ino80 complexes [43]. This complex also has an affinity for H2A variants of H2AX and H2AZ. The complex mediates nucleosome 10

eviction at double strand breaks and promotes homologous recombination repair implicating its importance in DNA repair. The complex has been shown to be critical for regulating telomere length and maintaining genomic stability [43, 47].

SWI/SNF Chromatin Remodeling Enzymes:

SWI/SNF enzymes are evolutionarily conserved, multisubunit complexes of 1-2

MDa that utilize the energy derived from ATP hydrolysis to alter the nucleosome structure by disrupting the histone-DNA contacts [48]. Each complex consists of a central

ATPase catalytic domain which can either be Brahma (BRM) or Brahma related gene 1

(BRG1) and 9-12 accessory proteins known as BRG1/BRM associated factors (BAFs).

BRG1 was characterized in yeast and was the first chromatin remodeling protein to be identified [49, 50]. BRM was first identified in Drosophila as a suppressor of body segment defects observed in Polycomb . The SWI/SNF complex does not possess any sequence specificity and is recruited to target regions primarily via interactions between BAFs and key transcriptional modulators [17]. SWI/SNF activity has been shown to be required for a variety of physiological processes including embryonic pluripotency [51, 52], development of neural progenitors [53], cardiac development [54], liver development [55], erythropoiesis [56] and differentiation of a plethora of cell types including muscle differentiation [57, 58], hematopoietic [59], neural differentiation [60], adipogenesis [61] and very recently melanocyte differentiation[17, 18]. Two variants of the complex have been identified in mammalian cells – BAF (also known as SWI/SNF-

A) and PBAF (also known as SWI/SNF-B) based on the purified complexes from mammalian cells. These two complexes have highly related but distinct subunit composition [62]. Specific components of chromatin remodelers confer specialized 11

activities and selective gene targeting to distinct complexes. The “core” SWI/SNF complex subunits includes BAFs 47 (SNF5/INI1), 53, 57, 60, 155, 170 and actin.

Additionally “specificity” subunits like ARID2 (AT-rich interactive domain 2, BAF200) and Polybromo (BAF180) distinguish Polybromo-associated BAF (PBAF) from BAF complexes, which are specified by ARID1 (BAF250) [63]. Notably, although the complex normally consists of 9-12 subunits, only four of these – BRG1 or BRM

(mutually exclusive ATPase subunits that are human homologues of the yeast Swi2/Snf2

ATPase subunit), SNF5 (INI1/BAF47), BAF155 and BAF170 – are required in vitro to remodel nucleosomes and additional BAFs are required for interactions with transcriptional activators and repressors in vivo [62]. SWI/SNF enzymes can play complex roles in the regulation of gene expression, and participate in activation as well as repression of a specific gene, owing to its heterogeneity.

Many of the BAFs have been established as bonafide tumor suppressors which get inactivated in multiple malignancies as well as in cancer cell lines [62, 64-66]. BRG1 and

BRM share about 74% sequence identity and exert similar biochemical activities in vitro

[67]. However, the function of these ATPases is very diverse in vivo as is indicated by their disruption/inactivation studies in mice. Whereas BRG1 inactivation is embryonic lethal in homozygotes, and heterozygotes show accelerated development of tumors, BRM inactivation exhibits mild proliferation defects and showed upregulation of BRG1 in certain tissues, suggestive of the fact that BRG1 can compensate for BRM loss in several cases [68, 69]. Whereas BRG1 was found to interact with proteins, BRM was found to be preferentially binding with ankyrin repeat proteins [70].

12

Figure 1-5 SWI/SNF chromatin remodeling complexes. Reprinted by permission from Macmillan Publishers Ltd: [NATURE REVIEWS CANCER] Vol 11, Pg 481-492 Copyright (2011).

Histone modifying enzymes

Covalent modifications of DNA or histones coordinate the maintenance of both local and global chromatin architecture [71, 72]. These modifications regulate DNA- templated processes, including transcription, repair, and replication [73]. Chromatin modifications determine accessibility to specific DNA loci by physically enhancing or weakening the non-covalent interactions between histones or between histones and DNA, and providing a platform for recruiting epigenetic regulators. Examples of histone modifying enzymes are Histone acetyltransferases (HATs), Histone deacetylases

(HDACs), kinases, histone methyltransferases (HMT) and demethylases. These enzymes modify histone tails by means of addition or removal of methyl, acetyl or phosphate groups post-translationally [9]. Epigenetic regulators can be broadly categorized as epigenetic writers, which are responsible for addition of covalent modifications on histones and DNA, epigenetic erasers, which are responsible for removal of covalent modifications on histones and DNA, and the epigenetic readers, proteins with

13

specialized binding domains that recognize and bind to covalent modifications of histones and DNA [72].

Figure 1-6 Epigenetic Regulators. Adapted from Dawson, M.A., T. Kouzarides, and B.J. Huntly,. N Engl J Med, 2012. 367(7): p. 647-57.

Histone Acetylation: The Nε-acetylation of lysine residues is the histone modification which is associated with “open” chromatin conformation and transcriptional activation.

Lysine acetylation is one of the most abundant post-translational modifications in the human proteome. Acetylation marks have been associated with processes like metabolism, DNA repair, protein-protein interactions and protein stability [74].

Acetylation is a highly dynamic process and is mainly regulated by competing activities of two enzymatic families, the histone lysine acetyltransferases (HATs) and the histone deacytylases (HDACs).There are two major types of HATs: (1) type-B, which are predominantly cytoplasmic and modify free histones, and (2) type-A, which are primarily nuclear and consists of CBP/p300 families. Global histone acetylation patterns are perturbed in cancers [75, 76]. 14

Histone Deacetylation: HDACs are enzymes that reverse lysine acetylation and 18 such enzymes are subdivided into 4 major classes depending on the . Class

I (HDAC 1-3 and HDAC8), class II (HDAC 4-7 and HDAC 9-10) and class IV

(HDAC11) share a related catalytic mechanism requiring zinc metal ion but does not involve the use of a cofactor. In contrast, class III HDACs (sirtuins 1-7) employs a distinct catalytic mechanism which is NAD+-dependent. HDACs target both histone and non-histone proteins and substrate specificity for these enzymes are largely mediated by components of multi-subunit complexes such as Mi2/NuRD, comprising of HDACs [75,

77].

Histone Methylation: Methylation in histones occurs on side chains of arginine, lysine, and histidine residues. The best-characterized sites of histone methylation are the ones on lysine residues. Some of these (H3K4, H3K36, and H3K79) are associated with active genes in euchromatin, whereas others (H3K9, H3K27, and H4K20) are associated with heterochromatin regions of the genome. Different methylation states on the same residue can also localize differently. For example, H3K4me2/3 spans the transcriptional start site

(TSS) of active genes, whereas H3K4me1 is associated with active enhancers. Similarly, whereas monomethylation of H3K9 is seen at active genes, trimethylation of H3K9 is associated with gene repression [75, 78].

Histone Demethylation: Two classes of histone demethylases have been studied so far.

LSD1 (KDM1A), demethylates lysine via an amine oxidation reaction with FAD as a cofactor, and belongs to first class. The Jumonji demethylases belong to second class of demethylases and have been studied more extensively. They have a conserved JmjC

15

domain, which functions via an oxidative mechanism and radical attack (involving Fe(II) and α-ketoglutarate) [75].

Histone Phosphorylation: Phosphorylation influences the overall structure and function of the local chromatin environment and is integral to essential cellular processes such as mitosis, , DNA repair, replication and transcription. Histone phosphorylation sites on core histones can be divided into two broad categories: (1) those involved in transcription regulation, and (2) those involved in chromatin condensation.

Kinases: They are the main players of pathways which convey extracellular cues within the cell. Coding mutations, altered expressions, and recurrent translocations involving signaling kinases are among the most frequent oncogenic phenomena described in cancer [79].

Histone Acetylation and Histone Methylation Readers: The primary readers of Nε- acetylation of lysine residues are bromodomains, which contain an evolutionarily conserved binding motif. These consists a diverse group of proteins which function as chromatin remodelers, histone acetyltransferases, histone methyltransferases, and transcriptional co-activators [80]. Bromodomains have been discussed in detail in separate sections.

Lysine methylation leads to considerable physicochemical diversity of lysine; these modification states are read and interpreted by proteins containing different specialized recognition motifs. Methylation readers are broadly classified into two major families, the Royal Family (Tudor domains, Chromo domains, and malignant brain tumor [MBT] domains) and PHD fingers [81].

16

Figure 1-7 Epigenetic Reader Domains. Adapted from Dawson, M.A., T. Kouzarides, and B.J. Huntly,. N Engl J Med, 2012. 367(7): p. 647-57.

1.3 Bromodomains

Bromodomains are evolutionarily conserved 110- protein modules that recognize and bind to acetylated lysine residues (KAc) on histone proteins and regulate epigenetic information by turning genes on and off [82]. Bromodomains were first reported in the Drosophila protein brahma [83, 84], and are present in many chromatin- associated proteins and nearly all known nuclear histone acetlytrasferases (HATs) [85]. It was long known from yeast genetic studies that bromodomains play a critical role in chromatin remodeling but specific biological functions began to emerge only after it was discovered that they function as acetyl-lysine binding domains [86]. Lysine acetylation is dynamic, as this modification directs both structural changes to chromatin and gene transcription. It is believed that the bromodomain is the sole known to recognize acetyl-lysine residues on proteins [74, 87].

Bromodomains can contribute to highly specific histone acetylation by tethering transcriptional HATs to specific chromosomal sites, and to the assembly and activity of 17

multi-protein complexes of chromatin remodeling enzymes such as SAGA and NuA4

[88, 89]. Bromodomain-containing proteins have been implicated in cellular differentiation and transcriptional regulation[90]. Detailed functions of bromodomains have been subsequently discussed in a separate section.

Structure of the Bromodomain

The three-dimensional structure of a bromodomain was first determined by performing nuclear magnetic resonance (NMR) spectroscopy of the bromodomain from the human transcriptional coactivator PCAF (p300/CBP-associated factor). A number of three-dimensional structures of bromodomains have been reported thereafter which include bromodomains either in the free form or bound to an acetyl-lysine containing peptide [91]. These include bromodomains from human transcriptional protein TAFII250

[91, 92], human and Saccharomyces cerevisiae GCN5, human PCAF [91, 93], human coactivator CBP [91, 94], human BET (bromodomains and extra terminal domain) family protein BRD2 [91, 95], human transcription factor BPTF (bromodomain and PHD domain transcription factor) [91, 96], and human SWI/SNF chromatin-remodeling complex protein Brg1 (Brahma-related gene 1) [91, 97]. It was evident from the structural analysis of all these bromodomains that most of them, if not all, adopt an atypical left handed bundle of four helices (αZ, αA, αB, αC) arranged in an antiparellel fashion [98]. The evolutionarily conserved structural fold is termed the „BrD fold‟ [99].

The structural arrangement places the ZA and BC loops at one end of the molecule and the AB loop and the N- and C-terminal ends on the opposite ends of the molecule. There is an additional secondary structural element called the ZA helix (a short single-turn helix) within the ZA loop due to the left handed twist topology of the bundle, which is 18

less common as compared to the right handed twist [98]. The inter-helical ZA and BC loops are of variable length and sequence, and together form a hydrophobic pocket, which serves to stabilize the structure and interact with acetyl-lysine modifications [86,

91, 99]. It was observed from aligning the bromodomain sequences that the highest degree of conservation occurs within the ZA loop region. Significantly, several of these residues were accessible for protein-interactions. Interestingly, it also became known that

ZA loop has three positions which are a common site of amino acid insertions or deletions imparting importance to this region for substrate specificity. Henceforth, the ZA loop is thought to play an important role in mediating bromodomain function in which conserved residues impart conserved substrate interactions and non-conserved residues mediate protein-specific bromodomain function [98].

Figure 1-8 Structure of Bromodomain. Reprinted from Gene, Vol 272(1-2):Pg:1-9 Copyright (2001) with permission from Elsevier

19

Acetyl-lysine recognition of bromodomain

The unique ability of NMR spectroscopy to measure changes in the local chemical environment and/or conformation of a protein induced upon binding to a ligand led to the discovery of acetyl-lysine recognition by bromodomains. NMR has the ability to detect weak but highly specific interactions between a protein and a ligand [86]. The overall sequence similarity between members of the bromodomain family is not high and there are significant variations in the sequences of the ZA and BC loops despite the conserved BrD fold [85, 99]. Tyr1125, Tyr1167 and Asn1168 in CREBBP (or CBP) are amongst the most conserved residues in the large BrD family which are engaged in acetyl-lysine recognition [94, 99-101]. The alternative modes of acetyl-lysine recognition by the BrD fold came to light when it was found out that TRIM28 BrD does not bind to lysine-acetylated residues, whereas the BrD6 of Polybromo does [99, 102]. A network of water-mediated hydrogen bonds with protein backbone carbonyl groups also contributes to acetyl-lysine binding. Site-directed mutagenesis played a critical role in cementing the function of bromodomains as acetyl-lysine binding domains [86, 103].The BET family of

BrD-containing proteins (BRDT, BRD3, BRD4) have the ability to recognize two acetylation marks with a single binding pocket [99, 104-106].

Human bromodomain family

There are 61 human bromodomains contained within 46 different bromodomain containing proteins, with some proteins containing multiple bromodomains [74, 107]. Till date, depending on the structure-based classification of human bromodomains, these 61 human bromodomains have been categorized into 8 families that share similar sequence length and at least 35% sequence identity [74, 82, 87, 108]. In contrast to this, the yeast 20

genome encodes only 9 bromodomain-containing proteins with a total of 14 BrDs [87,

109].

The first subfamily consists of a functionally diverse group of proteins that includes the HAT p300/CBP-associated factor (PCAF). In fact, it was the BRD structure of PCAF that demonstrated the ability of BRDs to function as acetyl-lysine binding domains [82, 103]. Additional proteins which belong to this group include the general control of amino-acid synthesis 5-like 2 (GCN5L), transcription factor Fetal Alzheimer antigen (FALZ) and the chromatin remodeling factor cat eye syndrome chromosome region 2 (CECR2), all of which localize in the nucleus.

The Bromodomain and Extra Terminal (BET) proteins are grouped in the second subfamily. They share a common architecture consisting of two N-terminal BRDs that exhibit high levels of sequence conservation as well as an extra-terminal (ET) domain and a more diverse C-terminal recruitment domain. This subfamily comprises of 4 proteins namely BRD2, BRD3, BRD4 and BRDT. It has been shown that these BET proteins are recruited to transcriptional start sites during mitosis [110-112] and the BET protein BRD4 tethers the positive transcription elongation factor (P-TEFb) to these sites via its unique C-terminus [74]. Highly specific small molecule inhibitors of this family have recently emerged as promising therapeutic agents in inflammation and cancer [82,

113].

The third subfamily of BRDs contains the transcriptional regulator bromodomain containing 8B (BRD8B), the HAT enzymes CREB binding protein (CREBBP) and E1A binding protein p300 (EP300), the C-terminal domain of the chromatin remodeling factors WD repeat domain 9 (WDR9 domain 2), the bromodomain adjacent to zinc finger 21

domain 1B (BAZ1B), the C-terminal domain of the JAK/STAT pathway related bromodomain-containing protein disrupted in leukemia (BRWD3 domain 2) and the C- terminal domain of the insulin signaling related pleckstrin homology domain interacting protein (PHIP domain 2).

The fourth subfamily contains the transcriptional regulators bromodomain containing 7 (BRD7) [63, 74], bromodomain containing protein 1 (BRD1), bromodomain and PHD finger-containing protein 1 (BRPF1), two AAA domain containing protein (ATAD2) as well as the KIAA1240 protein (KIAA1240), the bromodomain containing 9 (BRD9) and the bromodomain and PHD finger-containing protein 3 (BRPF3) [74].

Subfamily V of the human BRDs comprises of the transcriptional repressor tripartite motif-containing 66 (TRIM66), the tripartite motif-containing 33 (TRIM33), the transcriptional regulator transcriptional intermediary factor 1 (TIF1α), the transcriptional regulators nuclear auto-antigen Sp-100 (SP100), nuclear auto-antigen Sp-

110 (SP110) and SP140 nuclear body protein (SP140) as well as the SP140-like protein

(LOC93349), the transcriptional repressor bromodomain adjacent to zinc finger domain

2A (BAZ2A) and the bromodomain adjacent to zinc finger domain 2B (BAZ2B).

Subfamily VI of BRDs includes the histone methyl-transferase myeloid/lymphoid or mixed-lineage leukemia (MLL) and the transcriptional co-regulator tri-partite motif- containing 28 (TRIM28).

Subfamily VII contains the transcriptional repressor zinc finger MYND domain containing 11 protein (ZMYND11), the transcriptional initiators TAF1 RNA polymerase

II TATA box binding protein (TBP)-associated factor (TAF1) and TAF-1 like RNA 22

polymerase II TATA box binding protein (TBP)-associated factor (TAF1L) and the N- terminal BRDs of the chromatin remodeling factors WD repeat domain 9 (WDR9 domain

1), the JAK/STAT pathway related bromodomain-containing protein disrupted in leukemia (BRWD3 domain 1) and the insulin signaling related pleckstrin homology domain interacting protein (PHIP domain 1).

The last subfamily of the human bromodomains (VIII) contains the methyl- transferase ash1 (absent, small, or homeotic)-like (ASH1L), the chromatin remodeling factors SWI/SNF-related matrix-associated actin-dependent regulator of chromatin a2

(SMARCA2) and the SWI/SNF-related matrix-associated actin-dependent regulator of chromatin a4 (SMARCA4) as well as the Polybromo 1 (PB1) [74].

Function of Bromodomains

Targeting of chromatin-modifying enzymes to specific sites is often carried out by bromodomains. In Drosophila, a methyltransferase ASH1L activates ultrabithorax expression and its mammalian homologue has been associated with actively transcribed genes. Mixed lineage leukemia (MLL), a multidomain methyltransferase, regulates self- renewal of haematopoietic stem cells by controlling HOX () gene expression and displays haplo-insufficiency [87] meaning a single functional copy of the gene is insufficient for bringing about a wild-type condition, leading to the diseased state.

HATs CREBBP and EP300 act as transcriptional coactivators controlling a variety of biological processes, including cell growth, genomic stability, development, neuronal plasticity and memory formation, as well as energy homeostasis [114].

Homozygous knockout of CREBBP leads to severe defects in blood vessel formation in the central nervous system, developmental retardation and mice die in utero [115]. 23

Homozygous deletion of EP300 leads to death of mice between days 9 and 11.5 of gestation with defects in neurulation, cell proliferation and heart development [116].

Gcn5-knockout mice die during embryogenesis [117] whereas homozygous deletion of a closely related gene, PCAF leads to short-term memory defects and conditioned fear associated with increased plasma corticosterone levels [118].

ATAD2 has been shown to regulate expression of genes required for cell cycle progression. There is a suggestive role of ATAD2 in prostate cancer by mediating specific functions involved in cancer cell survival and proliferation

[119].

TAF1 modulates the rate of transcription initiation by interacting with transcriptional regulators and binding to the core promoter sequence encompassing transcriptional start sites [120]. It acts as a general transcriptional activator and regulates various essential biological processes like myogenesis, DNA-damage response, cell cycle and apoptosis [87].

The WD repeat proteins of BRWD1 and BRWD3 are involved in a variety of cellular processes, including cell cycle progression, signal transduction, apoptosis and gene regulation [121, 122]. BRWD1 associates with SWI/SNF complex component and plays a vital role in chromatin remodeling [123]. In mice, BRWD1 is required for normal spermiogenesis and a mutation in BRWD1 leads to phenotypically normal, but infertile mice [87].

TRIM24 mediates ligand-dependent activation of the androgen receptor and the . It also interacts with nuclear receptors such as thyroid, vitamin D3 and oestrogen receptors [124]. TRIM28 plays a key role in early embryogenesis by acting 24

as a for Kruppel-associated box-domain-containing zinc finger proteins

[125].

TRIM33 plays an important role in development which was advocated by the death of mice in utero, which are deficient of TRIM33 [126]. TRIM66 functions as a transcriptional silencer by associating with heterochromatin-associated factors and is mainly expressed in testis. [127]

SWI/SNF complexes remodel chromatin structure, leading to either transcriptional activation or repression of target genes depending on the composition of various complexes [49].Mammalian SWI/SNF complexes have a vital role in cell differentiation and proliferation and are essential components of the embryonic stem cell core pluripotency transcriptional network [128]. PB1 is a key regulator of senescence and has roles in cell cycle regulation [129].

ISWI complexes are key regulators of transcription, heterochromatin replication and chromatin structure [130]. The DNA-replication machinery requires the activity of

BAZ1A to penetrate condensed chromatin structure [131]. BAZ2A is a key subunit of

NoRC (nucleolar remodeling complex), which mediates transcriptional silencing of ribosomal RNA [132].

The bromodomain extra-C terminal domain (BET) family of proteins includes

BRDT, BRD2, BRD3 and BRD4 [133]. These proteins have been implicated in the regulation of transcription and proliferation [98]. BRD2 functions as a transcriptional coactivator or corepressor in a promoter-specific or tissue-specific manner. BRD2 is essential for embryonic development in mice [134, 135] and an association between

BRD2 and juvenile myoclonic epilepsy in humans has been reported [136]. BRDT is the 25

testis-specific BET family member which is essential for [137]. Deletion of BRDT in mice led to sterility and an altered histone modifications (enrichment of acetylated H3K9 and H3K27me3) in the exon region of BRDT gene of subfertile patients was observed [138]. Genome-wide association studies (GWAS) linked polymorphism in

BRDT to sterility in European male [139]. BET proteins are epigenome readers that play a key role in gene expression by acting at the interface between chromatin remodeling and transcriptional regulation.

BRD4

BRD4, originally named MCAP (Mitotic chromosome-associated protein), binds preferentially to acetylated lysine residues found in histones (H3K14ac and H4 dual K5 and K12 or K8 or K16) and non-histone proteins (RelA subunit of nuclear factor (NF)κB) [140, 141]. BRD4 is associated with acetylated chromatin throughout the cell cycle, and regulates transcription at targeted loci [140]. This association of BRD4 with acetylated chromatin is critical for the rapid expression of early G1 genes upon exiting mitosis [142]. BRD4 has been proposed to transmit epigenetic memory across cell division by bookmarking transcriptionally active genes by an unknown mechanism [143].

BRD4 is a key of transcriptional elongation by recruiting the positive transcription elongation factor complex (P-TEFb) [144]. This association of BRD4 with active P-TEFb is suggestive of its role in transcription by RNA polymerase II [145].

Functional transition of BRD4 between chromatin targeting and transcriptional regulation leads to speculation of diverse biological roles of BRD4 [140, 143]. Interestingly, there seems to be multiple functionally distinct pools of BRD4. One remains associated with chromatin during mitosis responsible for accelerating post mitotic gene activation [146]. 26

Another BRD4 pool recruits P-TEFb to the sites of active transcription [142] and a third pool which is recruited by the transcriptional Mediator complex independent of P-TEFb

[145]. BRD4 stimulates G1 gene transcription and promotes cell cycle progression to S- phase [87, 147]. BRD4 activation also predicts the survival of patients with breast cancer

[87, 148]. Activation of BRD4 manipulates the response of tumors to their microenvironments in vivo, resulting in reduction of tumor growth and metastasis in mice

[148]. BRD4 has been implicated in the differentiation of NUT-midline carcinoma [149], as well as skeletal muscle differentiation in association with MyoD and [150].

BRD4 controls the transcription of a number of viral genes. It regulates HIV transcription by inducing the phosphorylation of CDK9 (cyclin-dependent kinase 9) in the HIV transcription initiation complex and inhibiting the kinase activity [151]. BRD4 also inhibits the proteasomal degradation of the papillomavirus E2 protein [152]. It associates with Kaposi‟s sarcoma-associated herpesvirus-encoded LANA-1 (latency- associated nuclear antigen) and murine γ-herpesvirus protein which is required for establishing viral latency in vivo [153, 154].

BRD4 heterozygotes display pre- and post-natal growth defects associated with a reduced proliferation rate. These mice also exhibit a variety of anatomical abnormalities such as head malformations, absence of subcutaneous fat, cataracts, and abnormal liver cells. In primary cell cultures, heterozygous cells also display reduced proliferation rates.

Embryos nullizygous for BRD4 die shortly after implantation and are compromised in their ability to maintain inner cell mass in vitro, indicating a role in fundamental cellular processes [155].

27

Recent studies with small molecule inhibitors of the BET proteins have revealed a critical role of BRD4 in the development of several hematopoietic and somatic cancers, such as Burkitt‟s lymphoma, [156], melanoma [157], colon [158] and breast cancer [159].

The chemical inhibition of BRD4 alters gene expression and results in reduced proliferation of cancerous cells. As a result, BET bromodomain proteins are being targeted for use in male contraception and in the treatment of NUT-Midline carcinoma, leukemia and other cancers, HIV, cardiac failure, inflammation, and viral infection.

Domain organization of BRD4 family proteins

Bromodomain-containing protein 4 (BRD4) is a member of BET family that contains two tandem bromodomains (BDI and BDII) and an extraterminal (ET) domain.

In humans, four BET proteins (BRD2, BRD3, BRD4 and BRDT) exhibit similar domain organizations, gene arrangements, and some functional properties [145]. Except for

BRDT, which is expressed specifically in testis and ovary, BRD2, BRD3, and BRD4 are widely distributed. Although BDI, BDII, and the ET domains are characteristics of the

BET family proteins, other domains, such as motifs B and SEED (Ser/Glu/Asp-rich region), are also highly conserved. The C-terminal motif (CTM) is only present in BRD4

[160]. The ET domain mediates protein-protein interactions, which are conserved for the human BET proteins [161].

28

Figure 1-9 BET Family Proteins. Adapted from The Journal of Biological Chemistry Vol 282, No. 18, pp 13141-13145, Copyright (2007)

Therapeutic potential for inhibition of BRD4

The methyl ester analogue (MS417) of JQ1 selectively inhibits NF-κB transcriptional activation of proinflammatory genes in kidney cells treated with TNFα or infected by HIV. Genetic knockdown of BRD4 or inhibition with JQ1 causes a decrease in c- mRNA levels in (AML) cells and a profound anti- leukemic effect in vitro and in vivo. I-BET151 demonstrated significant activity in mixed-lineage leukemia (MLL)-fusion leukemia through inhibition of transcription at

BCL2, CDK6, and c-MYC. JQ1, I-BET762, and I-BET151 downregulates MYCN transcriptional program in neuroblastoma cell lines. JQ1 down-regulated c-MYC and induced profound changes in gene expression via both c-MYC-dependent and independent mechanisms in glioblastoma samples. Treating patient-derived samples with

JQ1 led to growth arrest of the malignant cells in NUT-midline carcinoma [133].

29

1.4 JQ1: A selective and potent inhibitor of BET family bromodomains

Epigenetic proteins have been an intently pursued target in ligand discovery for therapeutic gain which is evident by the FDA approval of DNA methyltransferase inhibitors and HDAC-I, but with the advancement of medicinal chemistry, chromatin- associated proteins such as bromodomains are gaining importance, as modulating bromodomain/acetyl-lysine interactions with small molecules provide new avenues for tailoring therapies to specific diseases. Based on the observation by Mitsubishi

Pharmaceuticals that simple thienodiazepines possessed binding activity of BRD4, a prototype ligand, known as JQ1, was developed which specifically inhibits the BET family of bromodomain proteins from interacting with chromatin [162]. The BET antagonist JQ1 is a cell-permeable small molecule that binds competitively to acetyl- lysine recognition motifs, such as bromodomains. JQ1 is highly potent and specific to a subset of human bromodomains, having excellent shape complementarity with the acetyl- lysine binding cavity. A general binding assay based on differential scanning fluorimetry

(DSF) [163] was conducted to observe the specificity of JQ1 as specific substrates of most bromodomains are unknown. Binding of JQ1 significantly increases the thermal stability of all the bromodomains of the BET bromodomain and extra-terminal family, highest being that of BRD4, while no significant stability shifts are detectable outside the

BET family, indicating that JQ1 is highly selective. In contrast, the stereoisomer (-)-JQ1 showed no significant interaction with any bromodomain present in the panel. To assess the competitive binding of (+)-JQ1 with acetyl-lysine, a luminescence proximity homogenous assay (alpha-screen) [164] was adapted to the BET bromodomains. (+)-JQ1 proved to be a potent, highly specific and Kac-competitive inhibitor for the BET family of 30

bromodomains [162]. Evaluations of JQ1 have indicated that BET bromodomain proteins can be targeted for use in male contraception and in the treatment of human diseases including cancer, inflammation, and viral infection.

Figure 1-10 Structure and selectivity of JQ1. Reprinted by permission from Macmillan Publishers Ltd: [NATURE ] Vol 468(7327): p. 1067-73 Copyright (2010).

BET Bromodomain inhibitors in cancer

Table 1-1 BET Bromodomain inhibitors in cancer. Adapted and modified from Oncotarget. Mar 20;6(8):5501-16 Copyright (2015)

31

32

Table 1-2 Bromodomain inhibitors in clinical development. Reprinted by permission from Macmillan Publishers Ltd: [NATURE REVIEWS DRUG DISCOVERY] Vol 13, Pg 337-356 Copyright (2014)

1.5 SWI/SNF Enzymes in Cancer:

SWI/SNF complex lacks intrinsic DNA sequence specificity and are typically recruited to genomic sites through physical interactions with sequence-specific transcription factors. SWI/SNF chromatin remodeling enzyme functions as an epigenetic regulator assisting in gene expression by turning genes on and off. In this context, the

SWI/SNF complex serves as an oncogene or a tumor suppressor, depending on the interacting protein partners [165]. Meta-analyses of cancer genome-sequencing data have revealed that mutations that inactivate SWI/SNF subunits are found in nearly 20% of human cancers [166]. These mutations are generally loss of function, implicating

SWI/SNF as a major tumor suppressor in diverse cancers.

It has been shown that Microphthalmia transcription factor (MITF) which is a master regulator of melanocyte differentiation interacts with SWI/SNF enzymes [17].

Our lab has shown that BRG1 promotes melanoma differentiation and increases melanoma cell survival on exposure to a chemotherapeutic agent or to UV radiation. The 33

BRM subunit of SWI/SNF chromatin remodelers has been implicated in melanocyte senescence [167]. BRG1 expression is increased in human cutaneous melanoma and it is a novel binding partner of tumor suppressor p16INK4a which is an important melanoma susceptibility gene and is also associated with increased melanoma invasiveness [168-

170]. Our lab has shown that down-regulation of BRM in BRG1-deficient melanoma cells significantly inhibits the expression of proproliferative and survival genes CDK2,

TBX2, and BCL2, thus compromising tumorigenicity. However, the requirement for

BRG1 is not compensated by BRM in several melanoma cell lines [171].

BAF180 has been identified to be mutated in 41% of renal cell carcinomas and a small minority of breast cancers [172]. It is thought to function as a tumor suppressor by interacting with the p53 protein to promote cell cycle arrest and senescence [129, 173].

ARID2, a PBAF SWI/SNF specific subunit is required for stability of BAF180 [174].

ARID2 mutations have been detected in hepatitis C virus-associated hepatocellular carcinomas, non-small-cell lung cancer, pancreatic cancer, and melanoma. Analysis of large-scale melanoma exome data has identified ARID2 as the most frequently mutated

SWI/SNF component [175, 176]. Another PBAF SWI/SNF specific subunit, BRD7, is frequently deleted in a subset of breast tumors that express wild-type, but not , p53. It has been implicated as a cofactor for p53 that activates oncogene-induced senescence. BRD7 along with its PBAF SWI/SNF component BAF180 are critical regulators of p53 required for replicative senescence [129]. BRD7 inhibits the G1-S progression of nasopharyngeal carcinoma (NPC) cell lines and is thought to be a [177].

34

The ARID1A subunit of SWI/SNF complexes has been found to specifically mutated in primary human cancers. It is mutated in 50% of ovarian clear cell carcinoma and in 30% of endometrial carcinoma. It is occasionally mutated in medulloblastomas, primary breast cancer and lung adenocarcinoma cell line indicating its tumor suppressor activities in various lineages. ARID1B mutations have been reported to be highly frequent in breast cancer [178]. Both, ARID1A and ARID1B mutations have been reported in hepatocellular carcinoma, neuroblastoma, as well as in melanoma, at a low frequency [175].

BAF47, a core component of SWI/SNF complex acts as a classic tumor suppressor with inactivating mutations occuring in nearly all malignant rhabdoid tumors.

These lethal tumors are poorly responsive to current therapies and arise in the brain, kidney and other soft tissues [178]. Down-regulation of BAF47 occurs in significant number of primary and metastic melanomas and correlates with poor prognosis [175].

Table 1-1 SWI/SNF mutations in cancer. Reprinted by permission from Macmillan Publishers Ltd: [NATURE REVIEWS CANCER] Vol 11, Pg 481-492

35

There are multiple studies suggestive of the important role played by SWI/SNF enzymes in melanoma. A number of SWI/SNF subunits were on a microarray that identified

MAPK dependent genes in melanoma cells [179].

36

Chapter 2

Role of BRD4 in melanocyte differentiation

2.1 Introduction

Bromodomains belong to a family of evolutionarily conserved protein modules originally found in proteins associated with chromatin and in nearly all nuclear histone acetyltransferases. Bromodomains modulate enzyme activities, protein assembly, and protein-protein interactions via lysine acetylation with broad implications in a wide variety of cellular events, such as transcriptional activation and chromatin remodeling.

They are the sole protein domains which function as acetyl-lysine binding residues and are epigenetic readers which bind to the acetylated residues regulating chromatin structure and gene expression.

BET protein family members (BRD2, BRD3, BRD4, and BRDT) are epigenome readers that bind to acetyl-lysine residues on the tails of histones H3 and H4, and exert key roles at the interface between chromatin remodeling and transcriptional regulation.

BRD4 is a BET protein which acts as a transcriptional coactivator and was initially known as mitotic chromosome associated protein (MCAP). BRD4 associates with acetylated chromatin throughout the cell cycle and regulates transcription at targeted loci.

37

It is implicated in cellular growth control, cell cycle progression and cancer development.

However, the role of BRD4 in melanocyte differentiation has not been investigated.

In this study, we utilized an in vitro murine model of immortalized murine melanoblasts (melb-a cells) which can be differentiated by treatment with alpha-MSH.

Alpha-MSH upon binding to melanocortin 1 receptor activates the microphthalmia- associated transcription factor (MITF) which is the master regulator of melanocyte differentiation and melanoma oncogene. MITF in turn activates genes required for melanin production, proliferation, and survival. JQ1 is a cell permeable small molecule inhibitor of BET proteins which binds to BRD4 with the highest affinity and specificity, inhibiting its binding to the acetylated lysine residues. JQ1 has an excellent shape complementarity with the acetyl-lysine binding cavity of bromodomains of BET family.

Treating melb-a cells with JQ1 compromised MITF target gene expression, melanin synthesis, and reduced proliferation of melb-a cells as well as neonatal human epidermal melanocytes (NHEM) and melanoma cells. Furthermore, we determined that BRD4 occupied the promoters of MITF target genes that regulate pigmentation. Treatment with

JQ1 inhibited the binding of both BRD4 and MITF to these promoters. Moreover, treating the melb-a cells as well as NHEMs disrupts histone H3 lysine 4 trimethylation

(H3K4me3) which is an active mark of transcription. The effects on gene expression were recapitulated by siRNA mediated silencing of BRD4. We showed a physical interaction between the two proteins. Thus, we hypothesize that BRD4 plays a crucial role in melanocyte differentiation by interacting with MITF and can be a potential novel target for treatment of hyperpigmentation disorders and in melanoma.

38

2.2 Materials and methods

Cell Culture

Human Embryonic Kidney (HEK) 293T cells were obtained from the ATCC.

They were cultured in DMEM Media (Life Technologies, Grand Island, N.Y.) containing

10% fetal bovine serum.Mouse melanoblast cells (Melb-a) were obtained from Dr.

Dorothy Bennett (The Welcome Trust, UK). The growth media consisted of RPMI1640

(Life Technologies, Grand Island, N.Y.), 10% fetal calf serum (Life Technologies, Grand

Island, N.Y.), 20ng/ml stem cell factor, and 20 picoM fibroblast growth factor 2. When cells grew to cover 70% of the dish, they were shifted to differentiation media. The differentiation media consisted of DMEM (Life Technologies, Grand Island, N.Y.), 10% fetal calf serum, 200 nMphorbal 12-myristate 13-acetate (Sigma-Aldrich, St. Louis, MO) and 2 nM Nle4, D-Phe7-alpha-MSH (Sigma-Aldrich, St. Louis, MO). Neonatal human epidermal melanocytes (NHEMs) were isolated as described in [180] and cultured in

Media 254 (Invitrogen, Carlsbad, CA, USA).

RNA isolation and quantitative real time PCR

Total RNA was isolated using Trizol (Invitrogen) and cDNA was prepared using the QiagenQuantitect Reverse Transcription kit. Quantitative PCR (qPCR) was performed in SYBR Green master mix (Qiagen) with an Applied Biosystems 7500 PCR and analyzed with the SDS software as described [18]. Mouse Bcl2, p21, Tyrosinase,

TRP1, TRPM1 and BRD4 mRNA levels were normalized to mouse RPL7. Human TRP1 mRNA levels were normalized to human RPL19. All the primers were obtained from

Integrated DNA Technologies (IDT).

Primer sequences for mouse melanoblasts are as follows: 39

Bcl2: 5‟- AGT TCG GTG GGG TCA TGT GTG -3‟ and 5‟- CCA GGT ATG CAC CCA

GAG TG -3‟ p21: 5‟- ACA CAC AGA GAG AGG GCT AAG G -3‟ and 5‟- AGA TCC ACA GCG

ATA TCC AGA C -3‟

Tyrosinase: 5‟- TTC AAA GGG GTG GAT GAC CG-3‟ and 5‟- GAC ACA TAG TAA

TGC ATC C -3‟

TRP1: 5‟- GCC CCA ACT CTG TCT TTT CTC AAT-3‟ and 5‟- GAT CGG CGT TAT

ACC TCC TTA GC-3‟

TRPM1: 5‟- CCT ACG ACA CCA AGC CAG AT -3‟ and 5‟- GAC GAC ACC AGT

GCT CAC AC -3‟

BRD4: 5‟- TAA AAA CTC CAA CCC CGA TG -3‟ and 5‟- TGC TCT CCG ACT CAG

AGG AT -3‟

RPL7: 5‟- GGAGGAAGCTCATCTATGAGAAGG-3‟ and 5‟- AAG ATC TGT GGA

AGA GGA AGG AGC -3‟

Primer sequences for human melanocytes are as follows:

TRP1: 5‟- TGG GAT CCA GAA GCA ACT TT -3‟ and 5‟- TGT GGT TCA GGA AGA

CGT TG -3‟

RPL19: 5‟- AAA CAA GCG GAT TCT CAT GG -3‟ and 5‟- TTG GTC TCT TCC TCC

TTG GAT -3‟

siRNA knockdown

A SMARTPool of siRNAs targeting mouse BRD4, and a non-targeting siRNA

[26] were obtained from Dharmacon. Transfection was performed according to 40

manufacturer‟s instructions. Media was replaced with the differentiation media consisting of DMEM (Life Technologies, Grand Island, N.Y.), 10% fetal bovine serum, 200 nMphorbal 12-myristate 13-acetate (Sigma-Aldrich, St. Louis, MO) and 2 nM Nle4, D-

Phe7-alpha-MSH (Sigma-Aldrich, St. Louis, MO)72 hours post transfection and cells were then harvested 48 hours later and subjected to Western blotting, Quantitative real time PCR, Chromatin immunoprecipitation and FAIRE analysis.

Cell extracts and immunoblot analysis

Cell extracts were prepared and Western Blotting was performed as described[18].The MITF, BRD4 and β-Actin were from Abcam (Cambridge,

MA, USA).The Tyrosinase and TRP1 antibodies were from Santa Cruz Biotechnology

(Santa Cruz, CA). The Tubulin was from Cell Signaling (Beverly, MA, USA).

Co-immunoprecipitations

HEK 293T cells obtained from ATCC were transfected with FLAG tagged MITF construct using Lipofectamine LTX (Invitrogen, Carlsbad, CA). Cells were harvested 48 hours post-transfection.Co-immunoprecipitations were performed as previously described[18].Species matched IgG (Santa Cruz Technologies) was used as a control.

Differentiated Melb-a cells were incubated with BRD4 antibody and western blotting was performed using MITF and BRD4 antibodies.

Chromatin immunoprecipitations (ChIPs)

ChIPs were performed with a control rabbit IgG, control mouse IgG, BRD4 antibody,

MITF antibody, Histone H3 antibody, Histone 3 Lysine 4 trimethylation (H3K4me3)

41

antibody, Histone H4 antibody, Histone 4 Lysine 16 acetylation () antibody as described[18].Primers used were as follows:

Mouse Tyrosinase Promoter: 5‟-AGT CAT GTG CTT TGC AGA AGA T-3‟ and 5‟-

CAG CCA AGA ACA TTT TCT CCT T-3‟

Mouse TRP1 Promoter: 5‟-GCA AAA TCT CTT CAG CGT CTC-3‟ and 5‟-AGC CAG

ATT CCT CAC ACT GG-3‟

Mouse IgH Enhancer: 5‟-GCCGATCAGAACCAGAAC ACC-3‟ and 5‟-

TGGTGGGGCTGGACAGAGTGTTTC-3‟.

Human TRP1 Promoter: 5‟-AGT GCT TCT GGC CTT TTC TT-3‟ and 5‟-TGT GGA

TTG CTG CCT GAT AA-3‟

Human CD25 Promoter: 5‟-AAA TAG CAA TTT CGC CGT TG-3‟ and 5‟-TAC GTT

AGT GCT GGG GCT TT-3‟

Propidium iodide (PI) staining and fluorescence-activated cell sorting (FACS)

Approximately 1 x 106 cells were fixed with 100% ethanol for 1 h, stained with

PI-RNAse solution for 30 min and loaded on a FACS-Calibur (BD Biosciences, San Jose,

CA) at the University of Toledo Flow Cytometry Core Facility. Data was analyzed using

Cell Quest Pro (BD Biosciences).

Apoptosis assay

Cells were stained with Annexin V as described in [171]and analyzed on a FACS-

Calibur (BD Biosciences) at the University of Toledo Flow Cytometry Core Facility.

42

Melanin assay

Melb-a cells were differentiated in the presence of JQ1 or vehicle control and harvested at the time points indicated. Cells were analyzed for melanin content as describedin [181].Briefly, cells were trypsinized and counted using Scepter 2.0 handheld automated cell counter (Millipore) and equal number of cells from each plate were lysed in 0.1M NaOH and vortexed for 20 minutes. Melanin content was calculated based on the absorbance at 475nm as compared to the standard curve obtained using synthetic melanin

(Sigma).

Cell Images

An equal number of mouse melanoblasts which are responsive to α-MSH were seeded onto polystyrene dishes and cultured in growth media at 37°Cand 5% CO2. The growth media consisted of RPMI1640 (Life Technologies, Grand Island, N.Y.), 10% fetal calf serum (Life Technologies, Grand Island, N.Y.), 20ng/ml stem cell factor, and 20 picoM fibroblast growth factor 2. When cells grew to cover 70% of the dish, they were shifted to differentiation media. The differentiation media consisted of DMEM (Life

Technologies, Grand Island, N.Y.), 10% fetal calf serum, 200 nMphorbal 12-myristate

13-acetate (Sigma-Aldrich, St. Louis, MO) and 2 nM Nle4, D-Phe7-alpha-MSH (Sigma-

Aldrich, St. Louis, MO). 500 nMactive enantiomer of JQ1 (+JQ1) or 500nM inactive enantiomer of JQ1 (-JQ1) was added to the cells that were maintained in growth media and to the cells that were shifted to differentiation media. Cells were then photographed using EVOSfl microscope (AMG, Bothell, WA)at 10Xmagnification at the indicated time points.

Chromatin accessibility (FAIRE analysis) 43

Melb-a cells were differentiated for 48 hours in the presence of active and inactive isomer of JQ1 and harvested for FAIRE analysis. NHEMs were cultured in the presence of an inactive isomer of JQ1 (-JQ1) or the active isomer of JQ1 (+JQ1) for 96 hours and harvested for FAIRE analysis.

FAIRE analysis was performed as described in [182]. Briefly, cells were cross- linked with 1% formaldehyde for 6 minutes and neutralized with glycine for 5 mins.

Chromatin was sheared by sonication for 4 minutes. 10% chromatin was removed to be used as Input or unFAIRE material and remaining chromatin material was processed for

FAIRE by phenol-chloroform extraction. Samples were subjected to qPCR and normalized to inputs and IgH Enhancer.

Cell counts

Cells were trypsinized at the indicated times. Cell counts were taken using the

Scepter 2.0 Hand-Held Automated Cell Counter (Millipore, Billerica, MA).

Statistical analysis

Statistical significance was calculated by the student‟s t-test.

44

2.3 Results

JQ1 treatment inhibits proliferation by inducing cell cycle arrest

In order to determine if BET proteins have a role in melanocyte differentiation, immortalized mouse melanoblasts, melb-a cells were differentiated in the presence and absence of the BET protein inhibitor JQ1. An equal number of melb-a cells were seeded onto polystyrene dishes and cultured in growth media. When cells grew to cover 70% of the dish, they were shifted to fresh growth media or to differentiation media. 500nM active enantiomer of JQ1 (+JQ1) or 500nM inactive enantiomer of JQ1 (-JQ1) was added to the cells that were maintained in growth media and to the cells that were shifted to differentiation media. Cells were then photographed using EVOSfl microscope (AMG,

Bothell, WA) at 10X magnification at the indicated time points. Cells that were differentiated in the presence of the inactive enantiomer of JQ1 accumulated melanin granules while cells differentiated with active JQ1 accumulated fewer melanin granules, indicating that JQ1 inhibited melanin synthesis (Fig. 1A).

The images in Fig. 1A indicated that JQ1 also caused a reduction in cell proliferation. In order to assess the effects of JQ1 on proliferation, cell counts were obtained after melanoblasts that were differentiated in the presence or absence of JQ1 for the indicated time periods. JQ1 significantly suppressed proliferation (Fig. 1B).

We then conducted an Annexin Assay in order to gain insights into the anti- proliferative effect of JQ1 on melb-a cells. Melb-a cells were cultured in presence of active and inactive JQ1 and harvested at the time points indicated and subjected to

Annexin Assay. Percent Annexin V positive cells indicate the percent of cells undergoing apoptosis. As observed, the exposure of cells to JQ1 resulted in minimal cell death and 45

the extent of cell death which was measured by percent annexin V positive cells was similar to the cells treated with inactive JQ1 (Fig. 1C). Therefore, treatment with JQ1 did not result in a significant increase in apoptotic cells.

To determine if the effects of JQ1 on proliferation were due to cell cycle changes, cells were stained with propidium iodide and subjected to flow cytometry. JQ1 promoted cell cycle arrest in G1 (Fig. 1D).

In order to further assess the effects of active isomer of JQ1 on melanin synthesis, melb-a cells were maintained in growth media and differentiation media in presence of a vehicle control (DMSO) or active JQ1 or were left untreated. Cells were harvested at the time points indicated in eppendorf tubes and the pellets were photographed. As was visible, untreated cells and the cells treated with vehicle control showed a similar extent of pigmentation upon differentiation whereas JQ1 treatment led to reduced pigmentation

(Fig. 2A).

The melanin content vehicle treated and JQ1 treated cells was measured and normalized to cell number. JQ1 treated cells had a significantly lower melanin content compared to vehicle treated cells (Fig. 2B).

Similarly, neonatal human epidermal melanocytes (NHEMs) were treated with an active or inactive enantiomer of JQ1. Treatment of NHEMs with active JQ1 reduced the number of melanin granules and altered cell morphology (Fig. 3A).

Similarly to mouse melanoblasts, JQ1 did not cause a significant increase in apoptosis of NHEMs (Fig. 3B) but did promote the cell cycle arrest at G1 (Fig. 3C).

To further characterize the effects of JQ1 treatment on pigmentation of NHEMs, cells were treated with vehicle control or JQ1 and harvested at the indicated time points. 46

Cell pellets were collected and photographed. Cell pellets from JQ1treated cultures were substantially lighter than control pellets, indicating that JQ1 reduced melanin synthesis

(Fig. 3D).

JQ1 treatment abrogates the expression of a subset of MITF target genes

The data in Fig. 1A, Fig. 2 and Fig. 3D indicated that melanin synthesis was abrogated by treatment of both melb-a cells and NHEMs with JQ1. Therefore, we assessed the transcriptional effects of JQ1 on the Microphthalmia-Associated

Transcription Factor (MITF) and on the expression of MITF target genes that regulate melanin synthesis. To assay the effects of JQ1 on gene expression at the protein level, total cell extracts were prepared from melb-a cells and NHEMs after treatment with JQ1 at the indicated time points and then subjected to Western Blotting. JQ1 treatment had no significant effect on the levels of MITF whereas the protein levels of MITF target genes tyrosinase and tyrosinase related protein 1 (TRP1) were decreased upon JQ1 treatment

(Fig. 4A & Fig. 4C).

To assay the effects of JQ1 on gene expression at the mRNA level, RNA was isolated from melb-a cells and NHEMs at the time points indicated, reverse-transcribed, and subjected to quantitative PCR. JQ1 inhibited expression of tyrosinase and TRP1 at the mRNA level (Fig. 4B & Fig. 4D).Transient receptor cation channel protein 1

(TRPM1) is an MITF target gene that is associated with melanocyte differentiation [183].

We found that JQ1 also abrogated TRPM1 expression (Fig. 4B).

The cyclin dependent kinase inhibitor, p21 and the pro-survival gene, BCL2 are also MITF targets. Because treatment with JQ1 inhibited proliferation and cell cycle progression, we investigated the effects of JQ1 on the expression of p21 and BCL2. We 47

detected a significant increase in p21 mRNA levels when melb-a cells were differentiated with JQ1 (Fig. 4E). Interestingly, JQ1treatment did not have a significant effect on expression of BCL2 (Fig. 4F). Thus, JQ1 has differential effects on different classes of

MITF target genes.

BRD4 occupies the promoter of MITF target genes and JQ1 treatment disrupts the enrichment of BRD4 and MITF on promoters of MITF target genes

Previous studies have shown that JQ1 binds to BRD4 with higher affinity than to other BET proteins and disrupts the acetyl-lysine binding activity of BRD4. Since JQ1 treatment inhibits the expression of MITF target genes (Fig. 4), we assessed the occupancy of BRD4 on promoters of MITF target genes. By mining publicly available data sets, we analyzed a ChIP-Seq data set which was performed in the SK-MEL-5 melanoma cell line (GSM1121099). We detected peaks of BRD4 binding on a number of loci that regulate pigmentation (Fig. 5A). We then performed chromatin immunoprecipitation with antibodies to BRD4 (Fig. 5B & Fig. 5D) and MITF (Fig. 5C &

Fig. 5E) on differentiated melb-a cells and NHEMs that were cultured in the presence of an inactive isomer of JQ1 (-JQ1) or the active isomer of JQ1 (+JQ1). BRD4 and MITF enrichment was quantified by qPCR using primers to the promoters of pigmentation genes (TRP1 and Tyrosinase in melb-a cells, TRP1 in NHEMs). It was observed that the occupancy of BRD4 and MITF on melanocyte specific promoters was disrupted when the cells were cultured in JQ1 (Fig. 5B-Fig. 5E).

JQ1 disrupts the histone mark of H3K4me3 and reduces chromatin accessibility

H3K4me3 is a histone modification mark known to be associated with transcriptionally active genes. Chromatin immunoprecipitations were performed with an 48

antibody to histone H3 trimethylated on lysine 4 (H3K4me3) on melb-a cells (Fig. 6A) and NHEMs (Fig. 6B) that were cultured in the presence of an inactive isomer of JQ1 (-

JQ1) or the active isomer of JQ1 (+JQ1). H3K4me3 enrichment was quantified by qPCR using primers to the indicated promoters. The data indicates that JQ1 disrupts this epigenetic mark (Fig. 6A & Fig. 6B).

As JQ1 disrupted the active mark of transcription, we evaluated the effect of JQ1 on chromatin accessibility by performing the formaldehyde assisted isolation of regulatory elements (FAIRE) assay. Melb-a cells (Fig. 6C) and NHEMs (Fig. 6D) that were cultured in the presence of an inactive isomer of JQ1 (-JQ1) or the active isomer of

JQ1 (+JQ1) were subjected to FAIRE analysis. The fold accessibility of chromatin was quantified by qPCR using primers to the indicated promoters. It was observed that JQ1 led to reduced chromatin accessibility at the melanocyte specific promoters (Fig. 6C &

Fig. 6D).

Knocking down BRD4 inhibits expression of genes that regulate melanin synthesis

To determine if the effect of JQ1 is specific to BRD4, an RNA interference approach was utilized. To assay the effects of knocking down BRD4 on gene expression at the protein level, total cell extracts were prepared from differentiated mouse melanoblasts and were subjected to Western Blotting. Knocking down BRD4 had no significant effect on the expression level of MITF whereas MITF target genes of

Tyrosinase and TRP1 were inhibited (Fig. 7A).

To assay the effects of knocking down BRD4 on gene expression at the mRNA level, RNA was isolated from differentiated mouse melanoblasts, reverse-transcribed, and

49

subjected to quantitative PCR. BRD4 knockdown inhibited expression of tyrosinase and

TRP1 at the mRNA and protein levels (Fig. 7B & Fig. 7C).

BRD4 knock down disrupts the active mark of transcription and reduces chromatin accessibility

Chromatin immunoprecipitations were performed with an antibody to H3K4me3 on melb-a cells that were cultured in the presence of a control siRNA or siBRD4.

H3K4me3 enrichment was quantified by qPCR using primers to the indicated promoters.

The data indicates that knocking down BRD4 disrupts this epigenetic mark (Fig. 7D).

We also checked the effect of knocking down BRD4 on chromatin accessibility. Melb-a cells that were cultured in the presence of siC or siBRD4 were subjected to FAIRE analysis. The fold accessibility of chromatin was quantified by qPCR using primers to the indicated promoters. It was observed that BRD4 knockdown led to reduced chromatin accessibility at the melanocyte specific promoters (Fig. 7E).

BRD4 physically interacts with MITF

Our data indicated that BRD4 occupies the promoters of MITF target genes and that disruption of BRD4 activity with the BET inhibitor JQ1 led to down regulation of

MITF target genes. Therefore, we investigated if BRD4 physically interacts with MITF.

To assay for physical interactions between BRD4 and MITF, 293T cells were transfected with epitope-tagged (FLAG) MITF using Lipofectamine LTX (Life Technologies, Grand

Island, N.Y.). Total cell extracts were immunoprecipitated with an antibody to BRD4 and were subjected to Western Blotting. The Western Blot was probed with an antibody to

BRD4 (Abcam, Cambridge, MA) and an antibody to the FLAG epitope (Sigma-Aldrich,

50

St. Louis, MO). It was observed that BRD4 co-immunoprecipitated with FLAG-MITF suggesting a physical interaction of BRD4 with MITF (Fig. 8A).

Next, we probed for physical interactions between BRD4 and MITF in melb-a cells. The BRD4 antibody pulled down MITF in differentiated melb-a cells, suggesting a physical interaction between the two proteins (Fig. 8B).

2.4 Discussion

Melanocytes are cells located in the basal layer of the epidermis that produce the pigment melanin. Once melanin is produced, it is stored in melanosomes, which are transferred from melanocytes to keratinocytes, thus giving skin its characteristic color.

Disruption of normal melanocyte function can lead to skin cancer and other skin diseases that affect the level of pigmentation. Our studies identified BRD4, a BET bromodomain protein, as a possible therapeutic target for treatment of hyperpigmentation disorders. We propose a novel mechanism where BRD4 is essential for carrying out the process of melanocyte differentiation and pigmentation by regulating the transcription of genes essential for melanin synthesis. We utilized mouse melanoblasts (melb-a) cells and physiologically relevant neonatal human epidermal melanocytes (NHEMs) to test our hypothesis. JQ1, a well known BET bromodomain inhibitor was used in our studies to block the activity of BRD4. We observed that JQ1 treatment led to a reduced proliferation in our cell lines. We performed annexin assay and cell cycle assay using flow cytometry analysis to check for the cause of reduced proliferation in these cell lines and detected that JQ1 did not affect apoptosis, but caused cell cycle arrest at G1 phase.

Dysfunction of melanin can result in a number of hyperpigmentation disorders. Upon differentiation, melb-a cells accumulated melanin granules and increased pigmentation 51

which was evident from our cell images. We observed a reduction in the number of melanin granules accumulated after treating the cells with JQ1. Inhibiting the BET bromodomain activity with JQ1 led to down regulation of melanin pigment.

Microphthalmia-associated transcription factor (MITF) is a master regulator of melanocyte differentiation which is activated by α-MSH binding to the MC1R receptor.

MITF regulates transcription of a wide variety of genes with implications in cell survival, melanogenesis and cell cycle arrest. The melanocyte-specific enzyme, tyrosinase (TYR), is required for the rate-limiting step of melanin synthesis while two other member of the tyrosinase family, tyrosinase related protein 1 (TRP1) and dopachrome tautomerase

(DCT), regulate the type of melanin synthesized. Treating the cells with JQ1 led to down regulation of pigment specific genes. However, although the effects of JQ1 on gene expression of TRP1 was profound and consistent in Melb-a as well as NHEM cell lines, tyrosinase gene expression seems to be differentially regulated in these cell lines, which was evident from the fact that tyrosinase was affected at a much later time point in case of NHEMs. We assessed the binding sites of BRD4 by analyzing a ChIP-seq data and found out that BRD4 occupied the promoters of MITF target genes. JQ1 treatment disrupts the binding of BRD4 and MITF to the promoters of pigment-specific genes.

H3K4me3 is an active mark of transcription which was disrupted by treating the cells with BET inhibitor, JQ1. Lysine acetylation is a predominant post translational modification which is characterized by an open chromatin structure. BRD4, with its bromodomain modules recognizes the acetyl-lysine residues. We detected that JQ1 treatment led to reduced chromatin accessibility. Since JQ1 inhibits BET bromodomain family members, we wanted to assess if the effects on gene expression were recapitulated 52

by siRNA mediated silencing of BRD4. Indeed, knocking down BRD4 led to similar effects as observed upon treating the cells with JQ1. Next, we wanted to assess the interaction between BRD4 and MITF since JQ1 treatment, as well as knocking down

BRD4, led to down-regulation of MITF target genes and disrupted the occupancy of

MITF on the promoters of pigment-specific genes. We carried out a co- immunoprecipitation assay in 293T‟s and differentiated melb-a cells and detected a physical interaction between the two proteins. Thus, we hypothesized that BRD4 plays a crucial role in melanogenesis by interacting with MITF and can be a potential novel target for treatment of hyperpigmentation disorders.

53

2.5 Figure Legends

Fig.1. JQ1 treatment inhibits proliferation by inducing cell cycle arrest in mouse melanoblasts.

(A) Representative mouse melanoblasts (melb-a) cell images taken with EVOSfl

microscope at 10X magnification at indicated time points in presence of inactive

and active isomer of JQ1.

(B) In order to assess the effect of JQ1 on proliferation, treated Melb-a cells were

harvested at the time points indicated and counted with a Scepter 2.0 Hand-Held

automated cell counter. Cell counts plotted are the average of 3 independent

experiments and statistical significance was calculated. * p<0.05, ** p<0.01.

(C) Melb-a cells were cultured in the presence of inactive and active isomer of JQ1

for the indicated time points. Cells were then harvested by trypsinization and

stained with Annexin V and subjected to flow cytometry. Percent Annexin V

positive cells indicate the percentage of cells undergoing apoptosis. The

percentages plotted are the average of 3 independent experiments. Standard error

bars are shown.

(D) To determine if the effects of JQ1 on proliferation were due to cell cycle changes,

Melb-a cells were harvested at indicated time points and stained with propidium

iodide and subjected to flow cytometry on a FACS-Calibur (BD Biosciences, San

Jose, CA, USA at the University of Toledo Flow Cytometry Core Facility). Data

was analyzed using Cell Quest Pro (BD Biosciences). Percentage of cells in G1, S

54

and G2 phases were plotted and statistical significance was calculated.

***p<0.001

Fig.2. JQ1 inhibits melanin synthesis.

(A) Melb-a cells were left untreated, treated with vehicle control (DMSO) and active

isomer of JQ1 and harvested at the indicated time points. Cells were collected in

eppendorf tubes and pellets were photographed.

(B) Harvested cells in Fig. 2A which were cultured with vehicle control and active

isomer of JQ1 were subjected to melanin assay and melanin content per cell

(µg/µl) was plotted.

Fig.3. JQ1 inhibits proliferation, pigmentation and induces cell cycle arrest in neonatal human epidermal melanocytes.

(A) Representative neonatal human epidermal melanocyte (NHEM) cell images taken

with EVOSfl microscope at 10X magnification at indicated time points in

presence of inactive and active isomer of JQ1.

(B) NHEM cells were cultured in the presence of inactive and active isomer of JQ1

for the indicated time points. Cells were then harvested by trypsinization and

stained with Annexin V and subjected to flow cytometry. Percent Annexin V

positive cells indicate the percentage of cells undergoing apoptosis. The

percentages plotted are the average of 3 independent experiments. Standard error

bars are shown.

55

(C) To determine if the effects of JQ1 on proliferation were due to cell cycle changes,

NHEM cells were harvested at indicated time points and stained with propidium

iodide and subjected to flow cytometry on a FACS-Calibur (BD Biosciences, San

Jose, CA, USA at the University of Toledo Flow Cytometry Core Facility). Data

was analyzed using Cell Quest Pro (BD Biosciences). Percentage of cells in G1, S

and G2 phases were plotted and statistical significance was calculated.

(D) NHEM cells were treated with vehicle control (DMSO) and active isomer of JQ1

and harvested at the indicated time points. Cells were collected in eppendorf tubes

and pellets were photographed.

Fig.4. JQ1 abrogates the expression of pigmentation genes in Melb-a and NHEM cells

(A) Total cell extracts were prepared from Melb-a cells harvested at the time points

indicated and were subjected to Western Blotting. Antibodies to detect Tyrosinase

and TRP1 were from Santa Cruz Biotechnology (Santa Cruz, CA). Antibody to

detect MITF was from Abcam (Cambridge, MA) and the loading control, Tubulin

was from Cell Signaling (Beverly, MA, USA). A representative blot is shown.

(B) RNA was isolated from Melb-a cells at time points indicated, reverse-transcribed,

and subjected to quantitative PCR. Primers for mouse Tyrosinase, TRP1 and

TRPM1 were obtained from IDT. Relative mRNA expression levels were

normalized to RPL7 as a control and are shown relative to undifferentiated sample

treated with inactive isomer of JQ1. Standard error bars and statistical

significance are shown. * p<0.05, ** p<0.01, ***p<0.001 56

(C) Total cell extracts were prepared from NHEM cells harvested at the time points

indicated and were subjected to Western Blotting. Antibodies to detect Tyrosinase

and TRP1 were from Santa Cruz Biotechnology (Santa Cruz, CA). Antibody to

detect MITF and the loading control β-Actin was from Abcam (Cambridge, MA).

A representative blot is shown.

(D) RNA was isolated from NHEM cells at time points indicated, reverse-transcribed,

and subjected to quantitative PCR. Primers for human TRP1 were obtained from

IDT. Relative mRNA expression levels were normalized to RPL19 as a control

and are shown relative to sample untreated with MSH. Standard error bars and

statistical significance are shown. ***p<0.001

(E) Total RNA was isolated as in Fig. 1D, reverse-transcribed, and subjected to qPCR

with primers to p21 and RPL7. The data shown are the average of 3 or more

independent experiments that were subjected to qPCR. Standard error bars are

shown and statistical significance was calculated. * p<0.05

(F) Total RNA from Melb-a cells treated with active and inactive isomer of JQ1 were

harvested at the indicated time points, reverse-transcribed, and subjected to qPCR

with primers to BCl2 and RPL7. The data shown are the average of 3 or more

independent experiments that were subjected to qPCR. Standard error bars are

shown.

57

Fig.5. BRD4 occupies the promoter of MITF target genes and JQ1 treatment disrupts the enrichment of BRD4 and MITF on promoters of MITF target genes

(A) ChIP-seq (GSM1121099) shows BRD4 peaks on several loci important for

melanogenesis. This ChIP-seq data was available in GEO profiles performed on

SK-MEL5 melanoma cells. TSS: Transcriptional start site

(B) Chromatin immunoprecipitation (ChIPs) were performed with an antibody to

BRD4 and a control antibody IgG in Melb-a cells differentiated for 48 hours in

presence of active and inactive isomer of JQ1. Fold enrichment of BRD4 on the

TRP1 and Tyrosinase Promoter is shown relative to a control IgH Enhancer

region. The data are representative of three independent experiments. Standard

error bars and statistical significance are shown. ** p<0.01

(C) Chromatin immunoprecipitation (ChIPs) were performed with an antibody to

MITF and a control antibody IgG in Melb-a cells differentiated for 48 hours in

presence of active and inactive isomer of JQ1. Fold enrichment of MITF on the

TRP1 and Tyrosinase Promoter is shown relative to a control IgH Enhancer

region. The data are representative of three independent experiments. Standard

error bars and statistical significance are shown. * p<0.05, ***p<0.001

(D) Chromatin immunoprecipitation (ChIPs) were performed with an antibody to

BRD4 and a control antibody IgG in NHEM cells cultured in presence of active

and inactive isomer of JQ1 for 96 hours. Fold enrichment of BRD4 on the TRP1

Promoter is shown relative to a control CD25 region. The data are representative

58

of three independent experiments. Standard error bars and statistical significance

are shown. ***p<0.001

(E) Chromatin immunoprecipitation (ChIPs) were performed with an antibody to

MITF and a control antibody IgG in NHEM cells cultured in presence of active

and inactive isomer of JQ1 for 96 hours. Fold enrichment of MITF on the TRP1

Promoter is shown relative to a control CD25 region. The data are representative

of three independent experiments. Standard error bars and statistical significance

are shown. ***p<0.001

Fig.6. JQ1 disrupts the H3K4me3 enrichment and reduces chromatin accessibility

(A) Chromatin immunoprecipitation (ChIPs) were performed with an antibody to

H3K4me3 and a control antibody histone H3 in Melb-a cells differentiated for 48

hours in presence of active and inactive isomer of JQ1. Fold enrichment of

H3K4me3 on the TRP1 and Tyrosinase Promoter is shown relative to a control

IgH Enhancer region. The data are representative of three independent

experiments. Standard error bars and statistical significance are shown. **

p<0.01,***p<0.001

(B) Chromatin immunoprecipitation (ChIPs) were performed with an antibody to

H3K4me3 and a control antibody histone H3 in NHEM cells cultured in presence

of active and inactive isomer of JQ1 for 96 hours. Fold enrichment of H3K4me3

on the TRP1 Promoter is shown relative to a control CD25 region. The data are

representative of three independent experiments. Standard error bars and

statistical significance are shown. ***p<0.001 59

(C) Chromatin accessibility assay was performed using FAIRE technique in Melb-a

cells differentiated for 48 hours in presence of active and inactive isomer of JQ1.

Fold accessibility on the TRP1 and Tyrosinase Promoter is shown relative to a

control IgH Enhancer region. The data are representative of three independent

experiments. Standard error bars and statistical significance are shown.

***p<0.001

(D) Chromatin accessibility assay was performed using FAIRE technique in NHEM

cells cultured in presence of active and inactive isomer of JQ1 for 96 hours. Fold

accessibility on the TRP1 Promoter is shown relative to a control CD25 region.

The data are representative of three independent experiments. Standard error bars

and statistical significance are shown. ** p<0.01

Fig.7. Knocking down BRD4 inhibits expression of pigmentation genes, disrupts the active mark of transcription and reduces chromatin accessibility

(A) Total cell extracts were prepared from Melb-a cells cultured in presence of a

control scrambled siRNA or specific siRNA against BRD4. Melb-a cells were

subjected to siRNA containing growth media for 72 hours and then the media was

changed to differentiation media. Melb-a cells were harvested 48 hours post

differentiation and were subjected to Western Blotting. Antibody to detect BRD4

and MITF was from Abcam (Cambridge, MA). Antibodies to detect Tyrosinase

and TRP1 were from Santa Cruz Biotechnology (Santa Cruz, CA). Antibody for

the loading control Tubulin was from Cell Signaling (Beverly, MA, USA). A

representative blot is shown. 60

(B) RNA was isolated from Melb-a cells cultured in presence of a control scrambled

siRNA or specific siRNA against BRD4. Melb-a cells were subjected to siRNA

containing growth media for 72 hours and then the media was changed to

differentiation media. Melb-a cells were harvested 48 hours post differentiation

and were subjected to trizol extraction. RNA was reverse-transcribed, and

subjected to quantitative PCR. Primers for mouse BRD4 was obtained from IDT.

Relative mRNA expression levels were normalized to RPL7 as a control and are

shown relative to sample treated with control siRNA. Standard error bars and

statistical significance are shown. ***p<0.001

(C) RNA was isolated as in Fig. 7B, reverse-transcribed, and subjected to quantitative

PCR. Primers for mouse Tyrosinase and TRP1 was obtained from IDT. Relative

mRNA expression levels were normalized as in Fig. 7B. Standard error bars and

statistical significance are shown. * p<0.05, **p<0.01

(D) Melb-a cells were cultured in presence of a control scrambled siRNA or specific

siRNA against BRD4. Melb-a cells were subjected to siRNA containing growth

media for 72 hours and then the media was changed to differentiation media.

Melb-a cells were harvested 48 hours post differentiation and were subjected to

chromatin immunoprecipitation(ChIPs). ChIP were performed with an antibody to

H3K4me3 and a control antibody histone H3. Fold enrichment of H3K4me3 on

the TRP1 and Tyrosinase Promoter is shown relative to a control IgH Enhancer

region. The data are representative of three independent experiments. Standard

error bars and statistical significance are shown. * p<0.05

61

(E) Melb-a cells were harvested as described in Fig 7D. FAIRE chromatin

accessibility assay was performed. Fold accessibility on the TRP1 and Tyrosinase

Promoter is shown relative to a control IgH Enhancer region. The data are

representative of three independent experiments. Standard error bars and

statistical significance are shown. ***p<0.001

Fig.8. BRD4 physically interacts with MITF

(A) 293T cells were transfected with epitope-tagged (FLAG) MITF using

Lipofectamine LTX (Life Technologies, Grand Island, NY).Total cell extracts

were immunoprecipitated with an antibody to BRD4 and were subjected to

Western Blotting. The Western Blot was probed with an antibody to BRD4

(Abcam, Cambridge, MA) and an antibody to the FLAG epitope (Sigma Aldrich,

St. Louis, MO).

(B) Melb-a cells differentiated for 48 hours were harvested and whole cell extracts

were used for immunoprecipitation with an antibody to BRD4 or a control IgG.

The immunoprecipitated material was subjected to Western Blotting and

immunoblotted for MITF and BRD4. 10% of the input was loaded on the gel.

62

2.6 FIGURES

Figure 1

63

Figure 2

UN-Untreated VC-Vehicle Control Figure 3

64

Figure 4

65

66

Figure 5

A.

GeneSymbol chrome peak start peak end peak length TSS-peak dist score DCT chr13 95130023 95130348 325 -1913 0.301134 MITF chr3 69975084 69976434 1350 -10666 0.011149 PMEL chr12 56359132 56360057 925 -714 0.422908 SOX10 chr22 38380378 38380953 575 -161 0.576604 TRPM1 chr15 31393932 31394607 675 3 0.5293 TYR chr11 88901226 88903726 2500 -9813 0.019738 TYRP1 chr9 12693174 12694499 1325 -211 0.506313

67

Figure 6

68

Figure 7

69

Figure 8

70

Chapter 3

The PBAF complex promotes expression of p53 target genes in response to ultraviolet radiation in melanoma

3.1 Introduction

Ultraviolet (UV)-radiation causes DNA damage and is implicated in the etiology of cutaneous melanoma, an aggressive malignancy that is notoriously chemoresistant.

SWI/SNF chromatin remodeling enzymes are multi-protein complexes composed of one of two related ATPases, BRG1 or BRM, and 9-12 associated factors (BAFs), that have critical functions in the regulation of gene expression. SWI/SNF complexes have been categorized as BAF and PBAF based on purified complexes from mammalian cells. The

PBAF complex contains only BRG1 as the catalytic subunit and includes specific subunits namely BAF180 (Polybromo), BRD7, and ARID2. Importantly, PBAF specific

SWI/SNF subunits are disrupted in a number of human cancers including melanoma. We were interested in assessing a potential tumor suppressive role of PBAF complexes in the context of melanocytes and melanoma cells. We hypothesized that the PBAF specific

SWI/SNF complex mediates tumor suppression by activating p53 target genes in

71

response to UV-radiation. ARID2 has been identified as a gene that is frequently mutated in melanoma. We surveyed a panel of melanoma cell lines and found that a subset of melanoma cells is deficient in the expression of the other PBAF specific subunits,

BAF180 and BRD7. Our data indicate that BRG1, BAF180, BRD7, and ARID2 promote expression of multiple p53 target genes as well as cell cycle arrest in UV-irradiated melanocytes and melanoma cells. BRG1, the central ATPase in PBAF SWI/SNF complex promoted cell cycle arrest in melanoma cells. Thus, our work suggests that a specific configuration of the SWI/SNF complex may have a tumor suppressive role in melanocytes and that aberrant expression or mutations of SWI/SNF components may contribute to melanoma tumorigenicity.

3.2 Materials and methods

Cell lines

Mouse melanoblast cells (Melb-a) were obtained from Dr. Dorothy Bennett (The

Welcome Trust, UK). The growth media consisted of RPMI1640 (Life Technologies,

Grand Island, N.Y.), 10% fetal calf serum (Life Technologies, Grand Island, N.Y.),

20ng/ml stem cell factor, and 20 picoM fibroblast growth factor 2. When cells grew to cover 70% of the dish, they were shifted to differentiation media. The differentiation media consisted of DMEM (Life Technologies, Grand Island, N.Y.), 10% fetal calf serum, 200 nMphorbal 12-myristate 13-acetate (Sigma-Aldrich, St. Louis, MO) and 2 nM

Nle4, D-Phe7-alpha-MSH (Sigma-Aldrich, St. Louis, MO). Neonatal human epidermal melanocytes (NHEMs) were isolated as described in [180] and cultured in Media 254

72

(Invitrogen, Carlsbad, CA, USA). SK-MEL5, WM-266-4, A375, SK-MEL147, MEL224 and MEL505 melanoma cells were obtained from ATCC. YUMAC melanoma cell lines were from Yale Cell Culture Core Facility (New Haven, CT, USA). SK-MEL5 cells expressing an empty vector or BRG1 were described in [18]. Melanoma cell lines were cultured as described in [18].

UV-irradiation

Melanoma cells were irradiated with a lamp that emits 75% in the UVB range and 25% in the UVA range (National Biologics, Twinsburg, OH, USA) at dose of 50mJ/cm2. UVC was blocked with a Kodacel sheet (Eastman Kodak, Rochester, NY, USA).

RNA isolation and quantitative real time PCR

Total RNA was isolated using Trizol (Invitrogen) and cDNA was prepared using the

Qiagen Quantitect Reverse Transcription kit. Quantitative PCR (qPCR) was performed in

SYBR Green master mix (Qiagen) with an Applied Biosystems 7500 PCR and analyzed with the SDS software as described [18]. Mouse p21 mRNA levels were normalized to mouse RPL7. Human p21, GADD45A, p53, BAF180 and BRD7 mRNA levels were normalized to human GAPDH. Primers for human GADD45A and BRD7 were obtained from SABiosciences (Qiagen). All the other human primers were obtained from

Integrated DNA Technologies (IDT).

73

Primer sequences for mouse melanoblasts are as follows: p21: 5‟- ACA CAC AGA GAG AGG GCT AAG G -3‟ and 5‟- AGA TCC ACA GCG

ATA TCC AGA C -3‟

RPL7: 5‟- GGAGGAAGCTCATCTATGAGAAGG-3‟ and 5‟- AAG ATC TGT GGA

AGA GGA AGG AGC -3‟

Primer sequences for human melanocytes and melanoma cell lines are as follows: p21: 5‟- GGA TGT CCG TCA GAA CCC A-3‟ and 5‟- CAG GTC CAC ATG GTC TTC

C-3‟ p53: 5‟- AGA ACC AAC ACG ACG TGG CTG C-3‟ and 5‟- GCC AGA CTG AGG

CAA GTG AGG G-3‟

BAF180: 5‟- TGC TGC AGG CCT CTC AGG CT-3‟ and 5‟- TCA CCA GCT GAA

TCC TCC CAC AGT-3‟

GAPDH: 5‟- TGG TCA CCA GGG CTG CTT TT-3‟ and 5‟- GGT GAA GAC GCC

AGT GGA CT-3‟ siRNA knockdown siRNA targeting mouse ARID2 and a non-targeting siRNA were obtained from

74

Dharmacon. Transfection was performed according to manufacturer‟s instructions. Media was replaced with the differentiation media consisting of DMEM (Life Technologies,

Grand Island, N.Y.), 10% fetal bovine serum, 200 nMphorbal 12-myristate 13-acetate

(Sigma-Aldrich, St. Louis, MO) and 2 nM Nle4, D-Phe7-alpha-MSH (Sigma-Aldrich, St.

Louis, MO) 72 hours post transfection and cells were then harvested 48 hours later and subjected to Western blotting and Quantitative real time PCR analysis. siRNAs targeting human p53, BAF180, BRD7 and ARID2 and a non-targeting siRNA were obtained from

Dharmacon (Lafayette, CO, USA) and transfected according to the manufacturer‟s instructions. Cells were UV-irradiated 72 after siRNA transfection and assayed by

Western blotting and Quantitative real time PCR, 48 hr after irradiation.

Cell extracts and immunoblot analysis

Cell extracts were prepared and Western Blotting was performed as described [18].The

BRG1 antiserum was previously described in [17]. The BAF47 antibody was from BD

Transduction Laboratories (BD Biosciences, San Jose, CA). The BAF180 antibody was from Bethyl Labs (Montgomery, TX, USA). The p21, p53 and ARID2 antibodies were from Santa Cruz Biotechnology (Santa Cruz, CA). The BRD7 antibody was from

Proteintech Group (Chicago, IL, USA). The Tubulin antibody was from Cell Signaling

(Beverly, MA, USA).

Co-immunoprecipitations

SK-MEL5 cells expressing BRG1 were transfected with FLAG tagged BRG1 construct

75

using Lipofectamine LTX (Invitrogen, Carlsbad, CA). Cells were harvested 48 hours post-transfection.Co-immunoprecipitations were performed as previously described

[18].Species matched IgG (Santa Cruz Technologies) was used as a control.

Propidium iodide (PI) staining and fluorescence-activated cell sorting (FACS)

Approximately 1 x 106 cells were fixed with 100% ethanol for 1 h, stained with PI-

RNAse solution for 30 min and loaded on a FACS-Calibur (BD Biosciences, San Jose,

CA) at the University of Toledo Flow Cytometry Core Facility. Data was analyzed using

Cell Quest Pro (BD Biosciences).

Statistical analysis

Statistical significance was calculated by the student‟s t-test.

3.3 Results

Expression levels of PBAF specific SWI/SNF components

We assessed the expression levels of PBAF specific SWI/SNF components in a panel of established melanoma cell lines and compared them with the neonatal human epidermal melanocytes (NHEM). We found that the levels of BRD7 were down regulated to a certain extent in all the melanoma cell lines as compared to NHEMs. BRG1 expression levels were almost similar in all the cell lines with the exception of SK-MEL-5 cells expressing empty vector. ARID2 and BAF180 protein expression levels were varying in different cell lines. However, it was observed that SK-MEL-5 cells expressing BRG1 had

76

increased expression of all the PBAF specific SWI/SNF components suggestive of stabilizing PBAF SWI/SNF complex in this cell line (Fig 1).

BRG1 promotes cell cycle arrest in response to UV-irradiation

To determine whether BRG1 also protects against UV-induced DNA damage, SK-MEL-

5 cells expressing either empty vector or BRG1 were subjected to sham or UV- irradiation. Cells were harvested 48 hours post irradiation and subjected to cell cycle analysis using flow cytometry. It was observed that SK-MEL-5 cells expressing BRG1 led to cell cycle arrest at G1 phase post UV-irradiation (Fig 2).

BRG1 promotes expression of a subset of p53 target genes

The p53 pathway activates upon induction of DNA damage. We irradiated SK-MEL-5 cells expressing either empty vector or BRG1 with a UV lamp. We performed a time course of protein expression and found that p53 accumulated equally in control and

BRG1 expressing cells. We have previously determined that re-expression of BRG1 in

SK-MEL-5 cells promotes expression of cyclin-dependent kinase inhibitor and cell cycle arrest gene, p21 [18]. We observed that p21 accumulated at higher levels in BRG1 expressing cells at all time points following UV-irradiation (Fig 3). At the RNA level,

BRG1 promoted 3-4 fold increase in p21 levels compared to control cells that were sham- irradiated. Although p21 expression was induced by UV-irradiation in both control as well as BRG1 expressing SK-MEL-5 cells, the levels were higher in UV-irradiated

BRG1 expressing cells (Fig 4). GADD45A is a p53 target gene which has been

77

implicated in growth arrest and DNA-damage. BRG1 significantly increased the expression of GADD45A at the mRNA level in both sham and UV-irradiated cells with the greatest effects observed in UV-irradiated BRG1 expressing SK-MEL-5 cells (Fig 4).

Knocking down p53 abrogates BRG1 mediated activation of a subset of p53 target genes

To determine if the observed BRG1 dependent increase in the expression of p21 and

GADD45A in response to UV-irradiation was dependent on p53, we introduced a short- hairpin RNA targeting p53 in sham and UV-irradiated SK-MEL-5 cell stably expressing

BRG1. We observed an increase in p53 and subsequent increase in p21 expression levels in the UV-irradiated cells treated with control siRNA. We detected a decrease in p21 expression at the protein level in both sham and UV-irradiated cells (Fig 5A).

Furthermore, expression of p21 and GADD45A was significantly decreased at the mRNA level in UV-irradiated SK-MEL-5 cells expressing BRG1 (Fig 5B).

BAF180 interacts with BRG1 and stabilizes PBAF SWI/SNF complex in melanoma cells

BAF180 has been implicated as a tumor suppressor that is mutated in breast and renal cell carcinomas and as a SWI/SNF subunit that promotes p53 transcriptional activity

[129, 172, 173].To determine whether BAF180 is involved in BRG1 mediated activation of p53 target gene expression in melanoma cells, we compared expression of BAF180 on sham or UV-irradiated control SK-MEL-5 cells and SK-MEL-5 cells that express BRG1.

78

We observed that BAF180 protein expression was increased in BRG1 expressing cells but was not dependent on UV-irradiation, whereas protein expression of the BAF47 subunit of the SWI/SNF complex was unaffected (Fig 6A). Protein-protein interactions among SWI/SNF subunits have been shown to regulate the stability of the SWI/SNF components [184]. BAF180 was co-immunoprecipitated with BRG1, suggesting that

BRG1 stabilized BAF180 protein levels by restoring the PBAF complex in these melanoma cells (Fig 6B).

Requirement of BAF180 in BRG1-mediated activation of a subset of p53 target genes

In order to investigate whether BAF180 is required for the observed BRG1-dependent increase in the expression of p21 and other p53 target gene, we introduced a short-hairpin

RNA that targets BAF180 in SK-MEL-5 cells expressing BRG1. Down-regulation of

BAF180 resulted in a small reduction in p53 protein levels in sham-irradiated cells but not in UV-irradiated cells, whereas p21 protein levels were affected on knocking down

BAF180 in sham as well as UV-irradiated cells (Fig 7A). Moreover, knocking down

BAF180 compromised the expression of p21 and GADD45A at the mRNA level in SK-

MEL5- cells expressing BRG1 (Fig 7B).

PBAF specific SWI/SNF components are required for activation of a subset of p53 target genes

To assess if BAF180 is required for p21 expression in other cells, we utilized WM-266-4

79

cells which have high levels of PBAF specific SWI/SNF subunits. Upon knocking down

BAF180, BRD7 (Fig 8A) and ARID2 (Fig 8C) individually in WM-266-4 cell line, we detected significant down regulation of the p53 target genes, p21, and GADD45A (Fig

8B and Fig 8D).

Knocking down ARID2 compromised the activation of p53 target genes in human melanocytes and mouse melanoblasts

Finally, to investigate the effects of knocking down the ARID2 component of PBAF

SWI/SNF complex in NHEMs and Melb-a cells, we introduced short hairpin RNA that targets ARID2 in NHEMs (Fig 9A) and Melb-a cells (Fig 9C). Down regulation of

ARID2 significantly compromised the expression of p21 (Fig 9C and Fig 9D).

3.4 Discussion

UV-radiation induces DNA damage and is the major environmental risk factor for all skin cancers, including melanoma [185]. The DNA damage response is a multi-step process that involves activation of cell cycle checkpoints and DNA repair. The response to DNA damage involves activation of the p53 pathway, which promotes expression of p21 and cell cycle arrest at the G1 phase. p21 is a p53 trans-activation target that mediates p53 driven G1 arrest in response to DNA damage [186]. In spite of extensive evidence from animal models in support of p53 as a tumor suppressor in melanoma initiation and progression, the majority of human melanomas retain wild-type p53.

Components of the PBAF SWI/SNF complex have been implicated in the regulation of

80

DNA damage response by their role in transcriptional activation of p53 target genes.

BRG1, the ATPase in PBAF SWI/SNF complex, has been shown to promote melanoma differentiation as well as to increase survival of melanoma cells that were exposed to a chemotherapeutic agent or to UV radiation [18, 171]. To determine if BRG1 protects against UV-induced DNA damage, we carried out a cell cycle analysis wherein SK-

MEL-5 cells expressing empty vector or BRG1 were subjected to sham and UV- irradiation and harvested 48 hours post treatment. Harvested cells were processed for flow cytometry analysis and it was observed that BRG1 containing SK-MEL-5 cells exhibited a significant cell cycle arrest at G1 phase. It suggested that BRG1 promotes an anti-proliferative effect in response to UV-irradiation. p21 is one of the critical regulators of DNA damage response and a cyclin-dependent kinase inhibitor. We observed that p21 accumulated at higher level in BRG1 expressing cells at all the time points following

UV-irradiation. Thus, in melanoma cells, BRG1 modulates the response to UV-radiation by promoting expression of cell cycle regulators involved in check point control and

DNA damage. This function was highly dependent on the BAF180 subunit of the

SWI/SNF complex. BAF180 is a bromodomain containing subunit of the PBAF complex which exclusively contains BRG1 as the catalytic subunit. Recent studies have implicated

BAF180 as a tumor suppressor in several human cancers [129, 172, 173]. We detected an increase in the expression levels of BAF180 in BRG1 expressing cells independent of

UV treatment. BAF180 was shown to be interacting with BRG1 suggesting the stabilization of PBAF SWI/SNF complex in these cells. We could assess the direct interactions between BAF180 and BRG1 using yeast-two hybrid assay or GST pull-down

81

assay. Our data indicate that depletion of BAF180 compromises p53 target gene expression in melanoma indicating that it may play a tumor suppressive role in melanoma. BRD7 and ARID2 are also PBAF specific SWI/SNF subunits [166]. We elucidated the requirement of these components in regulating the expression of a subset of p53 target genes in melanoma cell lines. Our data indicates that BRD7 is down regulated in melanoma cell lines as compared to melanocytes suggestive of its tumor suppressive role in melanoma. Two genome-wide sequencing studies reported that mutations in ARID2 occur in melanoma at a significant frequency and may be driver mutations [176, 187]. There is a lack of knowledge into the biological functions of

ARID2 in melanocytes and how ARID2 might act as a tumor suppressor. Our data with neonatal human epidermal melanocytes and mouse melanoblast indicates that knocking down ARID2 compromised p21 gene expression. Thus, our work suggests that a specific configuration of the SWI/SNF complex may have a tumor suppressive role in melanocytes and that aberrant expression or mutations of SWI/SNF components may contribute to melanoma tumorigenicity.

82

3.5 Figure Legends

Fig.1. Expression levels of PBAF specific SWI/SNF components in a panel of melanoma cell lines compared to neonatal human epidermal melanocytes

Total cell extracts were prepared from NHEM cells and melanoma cell lines indicated and were subjected to Western Blotting. Antiserum to detect BRG1 was described [17].

The BAF180 antibody was from Bethyl Labs (Montgomery, TX, USA). The ARID2 antibody was from Santa Cruz Biotechnology (Santa Cruz, CA). The BRD7 antibody was from Proteintech Group (Chicago, IL, USA). The Tubulin antibody was from Cell

Signaling (Beverly, MA, USA). A representative blot is shown.

Fig. 2. BRG1 promotes cell cycle arrest in response to UV-irradiation.

SK-MEL5 cells stably expressing either empty vector (EV) or BRG1 were sham- irradiated or UV-irradiated for 48 hours. Cells were then harvested, stained with propidium iodide and subjected to flow cytometry on a FACS-Calibur (BD Biosciences,

San Jose, CA, USA at the University of Toledo Flow Cytometry Core Facility). Data was analyzed using Cell Quest Pro (BD Biosciences). Percentage of cells in G1, S and G2 phases were plotted and statistical significance was calculated.

Fig. 3. DNA damage response upon UV-irradiation.

SK-MEL5 cells stably expressing either empty vector (EV) or BRG1 were sham- irradiated or UV-irradiated. Cell extracts were prepared 0, 2, 12, 24, and 48 hrs after

83

irradiation and subjected to Western Blotting with antibodies to BRG1, p53 and p21.

Tubulin is a loading control. A representative blot is shown.

Fig. 4. BRG1 promotes expression of a subset of p53 target genes

SK-MEL5 cells stably expressing either empty vector (EV) or BRG1 were sham- irradiated or UV-irradiated. Total RNA was harvested, reverse-transcribed, and subjected to qPCR with primers to p21, GADD45A and GAPDH. The data shown are the average of 3 or more independent experiments that were subjected to qPCR. Standard error bars are shown. **p <0.01

Fig. 5. Down-regulation of p53 abrogates BRG1 mediated activation of p53 target genes.

(A) Total cell extracts were prepared from SK-MEL5 cells stably expressing BRG1

which were cultured in presence of a control siRNA or specific siRNA against

p53. Cells were either sham-irradiated or UV-irradiated 72 hr after siRNA

transfection and assayed by Western blotting 48 hr after irradiation for antibodies

p53 and p21. Tubulin is a loading control. A representative blot is shown.

(B) Total RNA was harvested, reverse-transcribed, and subjected to qPCR with

primers to p53, p21, GADD45A and GAPDH for samples harvested as in Fig 5A.

The data shown are the average of 3 or more independent experiments that were

subjected to qPCR. Standard error bars are shown. **p <0.01

84

Fig. 6. Stabilization and interaction of BAF180 with BRG1.

(A) SK-MEL5 cells stably expressing either empty vector (EV) or BRG1 were sham-

irradiated or UV-irradiated. Cell extracts were prepared 0, 2, 12, 24, and 48 hrs

after irradiation and subjected to Western Blotting with antibodies to BRG1,

BAF180 and BAF47. Tubulin is a loading control. A representative blot is shown.

(B) SK-MEL5 cells stably expressing BRG1 were transfected with epitope-tagged

(FLAG) BRG1 using Lipofectamine LTX (Life Technologies, Grand Island,

NY).Total cell extracts were immunoprecipitated with an antibody to BRG1 and

were subjected to Western Blotting. The Western Blot was probed with an

antibody to BRG1, BAF47 and BAF180. 10% of the input was loaded on the gel.

Fig. 7. BAF180 is indispensable in BRG1-mediated activation of subset of p53 target genes.

(A) Total cell extracts were prepared from SK-MEL5 cells stably expressing BRG1

which were cultured in presence of a control siRNA or specific siRNA against

BAF180. Cells were either sham-irradiated or UV-irradiated 72 hr after siRNA

transfection and assayed by Western blotting 48 hr after irradiation for antibodies

BAF180, p53 and p21. Tubulin is a loading control. A representative blot is

shown.

(B) Total RNA was harvested, reverse-transcribed, and subjected to qPCR with

primers to BAF180, p21, GADD45A and GAPDH for samples harvested as in Fig

85

7A. The data shown are the average of 3 or more independent experiments that

were subjected to qPCR. Standard error bars are shown. *p <0.05, **p <0.01

Fig. 8. PBAF specific SWI/SNF components are required in mediating activation of p53 target genes.

(A) Total RNA was harvested, reverse-transcribed, and subjected to qPCR with

primers to BAF180, BRD7 and GAPDH for samples which were prepared from

WM-266-4 cells which were cultured in presence of a control siRNA or specific

siRNA against BAF180 or BRD7. The data shown are the average of 3 or more

independent experiments that were subjected to qPCR. Standard error bars are

shown. *p <0.05, **p <0.01

(B) Total RNA was harvested, reverse-transcribed, and subjected to qPCR with

primers to p21, GADD45A, and GAPDH for samples which were prepared from

WM-266-4 cells which were cultured in presence of a control siRNA or specific

siRNA against BAF180 or BRD7. The data shown are the average of 3 or more

independent experiments that were subjected to qPCR. Standard error bars are

shown. *p <0.05, **p <0.01

(C) Total cell extracts were prepared from WM-266-4 cells cultured in presence of a

control siRNA or specific siRNA against ARID2. Cells were harvested 72 hr after

siRNA transfection and assayed by Western blotting for antibody against ARID2.

Tubulin is a loading control. A representative blot is shown.

86

(D) Total RNA was harvested, reverse-transcribed, and subjected to qPCR with

primers to p21, GADD45A, and GAPDH for samples which were prepared from

WM-266-4 cells which were cultured in presence of a control siRNA or specific

siRNA against ARID2. The data shown are the average of 3 or more independent

experiments that were subjected to qPCR. Standard error bars are shown. **p

<0.01, ***p < 0.001

Fig.9.ARID2 is required for activation of a subset of p53 target genes in human melanocytes and mouse melanoblasts.

(A) Total cell extracts were prepared from NHEM cells cultured in presence of a

control siRNA or specific siRNA against ARID2. Cells were harvested 72 hr after

siRNA transfection and assayed by Western blotting for antibody against ARID2.

Tubulin is a loading control. A representative blot is shown.

(B) Total RNA was harvested, reverse-transcribed, and subjected to qPCR with

primers to p21 and GAPDH for samples which were prepared from NHEM cells

which were cultured in presence of a control siRNA or specific siRNA against

ARID2. The data shown are the average of 3 or more independent experiments

that were subjected to qPCR. Standard error bars are shown. *p <0.05

(C) Total cell extracts were prepared from Melb-a cells cultured in presence of a

control siRNA or specific siRNA against ARID2. Cells were harvested 72 hr after

siRNA transfection and assayed by Western blotting for antibody against ARID2.

87

Tubulin is a loading control. A representative blot is shown.

(D) Total RNA was harvested, reverse-transcribed, and subjected to qPCR with

primers to p21 and GAPDH for samples which were prepared from Melb-a cells

which were cultured in presence of a control siRNA or specific siRNA against

ARID2. The data shown are the average of 3 or more independent experiments

that were subjected to qPCR. Standard error bars are shown. *p <0.05

88

3.6 FIGURES

Figure 1

Figure 2

89

Figure 3

Figure 4

90

Figure 5

Figure 6

91

Figure 7

Figure 8

92

Figure 9

93

94

Chapter 4 Chapter 2

Overall Discussion Role of BRD4 in melanocyte differentiation

Bromodomains are an extensive family of evolutionarily conserved protein modules that function as acetyl-lysine binding domains. This recognition serves as a pivotal mechanism for regulating protein-protein interactions in numerous cellular processes including chromatin remodeling and transcriptional activation. Two groups of bromodomain containing proteins will be discussed in this section with their relevance to melanocyte differentiation and melanoma.

In the first project, the function of BET group of bromodomains in modulating pigmentation and melanocyte differentiation was deduced. BET protein family members are the epigenetic readers that bind to acetyl-lysine residues on the tails of histones H3 and H4, and exert key roles at the interface between chromatin remodeling and transcriptional regulation. There is increasing evidence of their role in human diseases, and recently a number of small-molecule inhibitors have been reported to have remarkable pre-clinical efficacy in various malignancies. These findings have provided

95

for a thrust of exploring BET proteins as novel targets in cancer therapy. We dissected the role of BRD4 in melanocyte differentiation and the potential of inhibiting BRD4 by a small molecule inhibitor JQ1 in treating hyperpigmentation disorders.

Ultraviolet (UV) radiation from sunlight is an environmental risk factor involved in the etiology of skin cancer, the most dangerous being melanoma. Melanoma is a highly aggressive form of skin cancer with an extremely poor prognosis after advancing to metastatic stage. Moreover, exposure to UV radiation can result in development of hyperpigmented lesions called solar lentigines. Lentigines are pigmented macules characterized by a localized proliferation of melanocytes, which appear to replace keratinocytes in the basal layer of the epidermis. Melanocytes are the pigment producing cells which synthesize melanin and distribute the pigment to keratinocytes, thus giving skin its characteristic color. The regulation of skin pigmentation is complex and involves not only melanin synthesis by melanocytes but also proper distribution of melanin to surrounding keratinocytes. Disruption of normal melanocyte function leads to over- stimulation of melanin which results in hyperpigmentation disorders. Current treatment options for hyperpigmentation include chemical peels, intense pulse light therapy, and bleaching creams such as hydroquinone or arbutin. The topical application of hydroquinone is the most effective treatment strategy available but it has adverse side effects that include increased sensitivity to UV radiation and increased risk for developing melanoma and other skin cancers. Thus, there is a need for the development of improved therapeutic options for hyperpigmentary conditions.

96

We utilized a novel approach of targeting an epigenetic regulator to gain insights in the modulation of melanogenesis and melanocyte function. We found that treatment of cultured human neonatal epidermal melanocytes (NHEMs) with the BET inhibitor, JQ1 resulted in a loss of pigmentation over time. We also assessed the effects of JQ1 in an in vitro mouse model of melanocyte pigmentation. Mouse melanoblasts (melb-a) cells are unpigmented cells that can be induced to differentiate into pigmented melanocytes. We observed that treatment of melb-a cells with JQ1 suppressed pigmentation by suppressing melanin synthesis. Furthermore, we determined that JQ1 suppressed proliferation of these cells by inducing cell cycle arrest and not causing cell death. However, treatment with

JQ1 may be toxic to keratinocytes. Keratinocytes are the principal cell type of the epidermis protected from UV radiation by melanin. It needs to be studied how JQ1 affects keratinocyte growth and response to UV radiation. We will perform a dose dependent response experiment and monitor cell death and melanin synthesis to determine a dose that alters melanogenesis but does not cause keratinocyte death.

Alternatively, we will test other BET inhibitors, such as PFI-1, which has similar effects on pigmentation, to identify one that causes minimal cell death.

The microphthalmia-associated transcription factor (MITF) is an important regulator of pigmentation-specific genes and melanocyte proliferation. MITF is a master regulator of melanocyte differentiation that binds to the regulatory regions of many genes which regulate pigmentation, as well as genes important for melanocyte survival, and activates their expression. Thus, modulation of MITF expression or activity could be a useful strategy for treating hyperpigmentation disorders. The melanocyte specific enzyme,

97

tyrosinase (TYR), is required for the rate limiting steps of melanin synthesis while two other melanocyte specific enzymes, tyrosinase related protein 1 (TRP1) and dopachrome tautomerase (DCT) regulate the type of melanin synthesized. We found that treatment of melb-a cells and NHEMs with JQ1 inhibited expression of genes that these enzymes. However, although the effects of JQ1 on gene expression of TRP1 was profound and consistent in Melb-a as well as NHEM cell lines, tyrosinase gene expression seems to be differentially regulated in these cell lines, which was evident from the fact that tyrosinase was affected at a much later time point in case of NHEMs.

Transient receptor cation channel protein 1 (TRPM1) is an MITF target gene that is associated with melanocyte differentiation. We found that JQ1 also abrogated TRPM1 expression. The cyclin dependent kinase inhibitor, p21 and the pro-survival gene, BCL2 are also MITF targets. Because treatment with JQ1 inhibited proliferation and cell cycle progression, we investigated the effects of JQ1 on the expression of p21 and BCL2. We detected a significant increase in p21 mRNA levels when melb-a cells were differentiated with JQ1. Interestingly, JQ1 treatment did not have a significant effect on expression of

BCL2. Thus, JQ1 has differential effects on different classes of MITF target genes. It would also be interesting to assess the effects of how quickly is melanogenesis restored when JQ1 is removed from the culture media.

Previous studies showed that JQ1 binds to BRD4 with higher affinity than to other BET proteins and is the BET protein most likely to mediate the suppressive effects of JQ1.

Although BRD4 is the BET protein most sensitive to JQ1, it also affects other BET proteins to a lesser extent. Thus, there is a small possibility that inhibition of other BET

98

proteins underlie the suppressive effects of JQ1 on the expression of MITF target genes, or that activities of other BET proteins can compensate for BRD4. We will down-regulate other BET proteins by siRNA approach, alone and in combination with BRD4 knockdown to deduce their effects on pigmentation. A CRISPR/Cas9 approach to mutate or flox multiple genes in vitro or in vivo can also be utilized.

Next, we wanted to identify BRD4 binding sites in melanocytes and investigate the changes in BRD4 binding that occur as a result of BET inhibition. By mining publicly available data sets, we came across a ChIP-seq data that was performed in a melanoma cell line SK-MEL-5. We detected peaks of BRD4 binding on a number of loci that regulate pigmentation. By performing chromatin immunoprecipitation in melb-a cells, we indicate that BRD4 is enriched on the promoters of pigment-specific genes of TRP1 and tyrosinase and JQ1 displaces BRD4 from these sites. BET inhibition also disrupts MITF binding onto the promoters of its target genes. Addition of a parallel gene-expression study, e.g. RNA-seq would indicate the functions of BRD4-binding sites. Performing a genome-wide study to understand the role of BRD4 in modulating chromatin architecture and gene expression simultaneously could yield a global picture of how these genes are functioning in the context of melanocytes. Furthermore, JQ1 disrupts histone H3 lysine 4 trimethylation (H3K4me3), a mark of active transcription. This suggests that BRD4 directly regulates the transcription of these genes involved in melanin synthesis. It would be useful to deduce the effect of BET inhibition on inhibitory marks of transcription like histone H3 lysine 27 trimethylation (H3K27me3) or histone H3 lysine 9 trimethylation

99

(H3K9me3).

BRD4 has a role in regulating higher order chromatin structure by mediating communication between promoters and enhancers. A chromosome conformation capture assay can be performed in the presence of active and inactive enantiomer of JQ1 to analyze the changes in chromosome conformation and evaluate the requirement of BRD4 at the promoters of pigment specific genes. We also investigated the effect of BET inhibition on chromatin accessibility. It was observed by formaldehyde-assisted isolation of regulatory elements (FAIRE) assay that JQ1 treatment as well as knocking down

BRD4 led to reduced chromatin accessibility at the melanocyte specific promoters.

Because BRD4 is an epigenetic reader and not a chromatin remodeling enzyme, the

BRD4 elicited changes in chromatin structure are thought to be due to BRD4 interactions with other proteins. Our data indicated that BRD4 occupied the promoters of MITF target gens and that disruption of BRD4 activity with the BET inhibitor led to down regulation of MITF target genes, we set to investigate the interaction of BRD4 with MITF. Using co-immunoprecipitation assays, we deduced the physical interactions between these two proteins. To further analyze if the interaction between BRD4 and MITF is direct, GST- pull down assays or yeast-two hybrid assays need to be performed.

We have performed an extensive study to deduce the role of a BET protein in melanocyte differentiation and pigmentation in vitro. It needs to be assessed if JQ1 responds in a similar way on intact skin. Studying the effects of BET inhibition on human skin equivalents or animals (hk14-Kitl mice) which possess epidermal melanocytes would

100

strengthen the results of our study. However, the investigation of the role of BRD4 in regulating melanin synthesis in melanocytes is novel. We have combined BET inhibiton by JQ1 with RNAi strategies to interrogate the function of BRD4 in melanocyte differentiation and pigmentation. The outcome of our studies has a potential to translate into novel therapeutic options for treating skin disorders and for developing products that protect against damage from solar radiation by targeting epigenetic regulators. In total, our results demonstrate that BRD4 plays an important role in melanocyte differentiation by interacting with MITF, and can be targeted by BET protein inhibitors for treatment of hyperpigmentation disorders.

In the second project, the role of bromodomain containing proteins of an ATP-dependent chromatin remodeling complex, SWI/SNF, were assessed. In particular, we investigated the role of PBAF SWI/SNF complex in the context of melanoma tumorigenicity.

SWI/SNF is an evolutionarily conserved ATP-dependent multi-subunit chromatin- remodeling complex that performs fundamental roles in gene regulation, cell lineage specification, and organ development. Genomic studies have revealed a tumor- suppressive role for SWI/SNF complex, with inactivating mutations in nearly 20% of human cancers. Mammalian SWI/SNF complexes can be classified broadly as BAF

(BRG1/BRM-associated factors) complex bearing either one of ARID1A/ARID1B or

PBAF (Polybromo-associated BAF) complex harboring ARID2, BAF180 and BRD7 subunits. Loss-of-function mutation of genes encoding BRG1 and BAF180 are particularly common in lung and renal cell cancers, respectively. Specific mutations in

BRG1 have been identified in pancreatic, breast, lung and prostate cancer cell lines.

101

Mutations in PBAF specific SWI/SNF components have been reported widely in various cancer types.

Melanoma is an extremely aggressive malignancy originating in pigment-producing melanocytes in the basal layer of epidermis. The alarming increase in its rate of incidence

(5-7% annually) and lack of effective treatment underscores the necessity to investigate the molecular mechanisms by which melanoma develops. In that regards, studying the regulatory pathways involved in melanoma development and progression is of utmost importance. UV-radiation causes DNA damage and is the major environmental risk factor in the development of melanoma.

There is an increasing evidence of some SWI/SNF subunits being down-regulated or mutated in melanoma. Our preliminary data shows BRD7 protein expression levels to be down-regulated in a panel of melanoma cell lines as compared to neonatal human epidermal melanocytes. The studies of melanoma identified ARID2 as the SWI/SNF component that is most frequently mutated. BRG1, the ATPase in PBAF SWI/SNF complex, has been shown to promote melanoma differentiation as well as to increase survival of melanoma cells that were exposed to a chemotherapeutic agent or to UV radiation.

The tumor suppressor, p53 is activated is response to DNA damage and serves as a critical regulator of cell cycle checkpoints. The response to DNA damage involves activation of the p53 pathway, which promotes expression of p21 and cell cycle arrest at the G1 phase. p21 is a p53 trans-activation target that mediates p53 driven G1 arrest in

102

response to DNA damage. In spite of extensive evidence from animal models in support of p53 as a tumor suppressor in melanoma initiation and progression, the majority of human melanomas retain wild-type p53. Some studies suggest that the loss of p53 in melanoma may be unnecessary due to the existence of compensatory mechanisms that involve over-expression of p53 inhibitory proteins such as MDM2, silencing of downstream effectors, such as APAF1, and by activation of anti-apoptotic factors and inhibitors of apoptosis.

BRG1, the central ATPase in PBAF SWI/SNF complex, has been reported to be required for p53-driven transcriptional activation and cell-cycle regulation. Components of the

PBAF SWI/SNF complex (BRD7 and BAF180) have been implicated in the regulation of

DNA damage response by their role in transcriptional activation of p53 target genes.

We wanted to assess the tumor suppressive role of PBAF SWI/SNF complex in context of melanoma by activating p53 pathway in response to UV-radiation. To determine if

BRG1 protected against UV-induced DNA damage, we carried out a cell cycle analysis wherein SK-MEL-5 cells expressing empty vector or BRG1 were subjected to sham and

UV-irradiation. It was observed that BRG1 containing SK-MEL-5 cells exhibited a significant cell cycle arrest at G1 phase. It suggested that BRG1 exhibited anti- proliferative effect on melanoma cells under UV-irradiation. The requirement of PBAF specific SWI/SNF subunits could possibly be assessed by performing similar cell cycle analysis by knocking down these components individually or in combinations. p21 is a cyclin-dependent kinase inhibitor and one of the critical regulators of DNA

103

damage response. We detected that BRG1 expressing cells accumulated p21 at higher levels, following UV treatment, suggestive of its role in the regulation of this gene.

BAF180 is a PBAF SWI/SNF specific subunit which is implicated as a tumor suppressor in several human cancers like renal cell carcinoma and breast cancer. We detected an increase in the expression levels of BAF180 in BRG1 expressing cells independent of

UV treatment. BAF180 was shown to be interacting with BRG1 suggesting the stabilization of PBAF SWI/SNF complex in these cells. We could assess the direct interactions between BAF180 and BRG1 using yeast-two hybrid assay or GST pull-down assay. Our studies suggest BAF180 to be indispensable for BRG1-mediated activation of a subset of p53 target genes. Our data indicate that depletion of BAF180 compromises p53 target gene expression in melanoma indicating that it may play a tumor suppressive role in melanoma.

SWI/SNF mutations can be mutually exclusive with other tumor suppressor mutations

(e.g., PTEN and p53) in certain tumor types, suggesting that SWI/SNF performs tumor protective functions that overlap with known pathways. It would be interesting to check for the p53 independent activity of SWI/SNF complex in context of melanoma progression. Co-immunoprecipitation of p53 with PBAF specific SWI/SNF subunits along with chromatin immunoprecipitation assays on the promoters of a subset of p53 target genes could shed light on the requirement of specific sub-units for activating these p53 target genes.

We elucidated the requirement of BRD7 and ARID2 components of PBAF SWI/SNF

104

complex in regulating the expression of a subset of p53 target genes in melanoma cell lines. Similar to BAF180, our data indicated the requirement of BRD7 and ARID2 to mediating activation of p21 and GADD45A. There is a lack of knowledge into the biological functions of ARID2 in melanocytes and how ARID2 might act as a tumor suppressor. Our data with neonatal human epidermal melanocytes and mouse melanoblast indicates that knocking down ARID2 compromised p21 gene expression.

Thus, our work suggests that a specific configuration of the SWI/SNF complex may have a tumor suppressive role in melanocytes and that aberrant expression or mutations of

SWI/SNF components may contribute to melanoma tumorigenicity.

In our zeal for unraveling the tumor suppressive functions of a multi-subunit PBAF

SWI/SNF complex in the context of melanoma, understanding the mechanisms by which tumor progresses, is of utmost importance. It might be due to an inactivating mutation of a specific subunit, dysfunction of the aberrant residual complex or the disintegration of the entire complex that could lead to abnormal cell growth. Understanding these aspects will give us a better outlook on the activity of PBAF SWI/SNF complex and p53 in the context of melanoma.

105

References

1. Ernfors, P., Cellular origin and developmental mechanisms during the formation

of skin melanocytes. Exp Cell Res, 2010. 316(8): p. 1397-407.

2. Hirobe, T., How are proliferation and differentiation of melanocytes regulated?

Pigment Cell Melanoma Res, 2011. 24(3): p. 462-78.

3. Mayer, T.C., The migratory pathway of neural crest cells into the skin of mouse

embryos. Dev Biol, 1973. 34(1): p. 39-46.

4. Lin, J.Y. and D.E. Fisher, Melanocyte biology and skin pigmentation. Nature,

2007. 445(7130): p. 843-50.

5. Opdecamp, K., et al., Melanocyte development in vivo and in neural crest cell

cultures: crucial dependence on the Mitf basic-helix-loop-helix-zipper

transcription factor. Development, 1997. 124(12): p. 2377-86.

6. Ito, S. and K. Wakamatsu, Diversity of human hair pigmentation as studied by

chemical analysis of eumelanin and pheomelanin. J Eur Acad Dermatol Venereol,

2011. 25(12): p. 1369-80.

7. Ito, S. and K. Wakamatsu, Human hair melanins: what we have learned and have

not learned from mouse coat color pigmentation. Pigment Cell Melanoma Res,

2011. 24(1): p. 63-74.

106

8. Bondurand, N., et al., Deletions at the SOX10 gene cause Waardenburg

syndrome types 2 and 4. Am J Hum Genet, 2007. 81(6): p. 1169-85.

9. Mollaaghababa, R. and W.J. Pavan, The importance of having your SOX on: role

of SOX10 in the development of neural crest-derived melanocytes and glia.

Oncogene, 2003. 22(20): p. 3024-34.

10. Wegner, M., Secrets to a healthy Sox life: lessons for melanocytes. Pigment Cell

Res, 2005. 18(2): p. 74-85.

11. Mitra, D. and D.E. Fisher, Transcriptional regulation in melanoma. Hematol

Oncol Clin North Am, 2009. 23(3): p. 447-65, viii.

12. Vance, K.W. and C.R. Goding, The transcription network regulating melanocyte

development and melanoma. Pigment Cell Res, 2004. 17(4): p. 318-25.

13. Steingrimsson, E., N.G. Copeland, and N.A. Jenkins, Melanocytes and the

microphthalmia transcription factor network. Annu Rev Genet, 2004. 38: p. 365-

411.

14. Bejar, J., Y. Hong, and M. Schartl, Mitf expression is sufficient to direct

differentiation of medaka blastula derived stem cells to melanocytes.

Development, 2003. 130(26): p. 6545-53.

15. Planque, N., et al., Expression of the microphthalmia-associated basic helix-loop-

helix leucine zipper transcription factor Mi in avian neuroretina cells induces a

pigmented . Cell Growth Differ, 1999. 10(7): p. 525-36.

16. Tachibana, M., et al., Ectopic expression of MITF, a gene for Waardenburg

syndrome type 2, converts fibroblasts to cells with melanocyte characteristics. Nat

107

Genet, 1996. 14(1): p. 50-4.

17. de la Serna, I.L., et al., The microphthalmia-associated transcription factor

requires SWI/SNF enzymes to activate melanocyte-specific genes. J Biol Chem,

2006. 281(29): p. 20233-41.

18. Keenen, B., et al., Heterogeneous SWI/SNF chromatin remodeling complexes

promote expression of microphthalmia-associated transcription factor target

genes in melanoma. Oncogene, 2010. 29(1): p. 81-92.

19. Hershey, C.L. and D.E. Fisher, Genomic analysis of the Microphthalmia locus

and identification of the MITF-J/Mitf-J isoform. Gene, 2005. 347(1): p. 73-82.

20. Fuse, N., et al., Identification of a melanocyte-type promoter of the

microphthalmia-associated transcription factor gene. Biochem Biophys Res

Commun, 1996. 219(3): p. 702-7.

21. Tassabehji, M., V.E. Newton, and A.P. Read, Waardenburg syndrome type 2

caused by mutations in the human microphthalmia (MITF) gene. Nat Genet, 1994.

8(3): p. 251-5.

22. Herbarth, B., et al., Mutation of the Sry-related Sox10 gene in Dominant

megacolon, a mouse model for human Hirschsprung disease. Proc Natl Acad Sci

U S A, 1998. 95(9): p. 5161-5.

23. Hemesath, T.J., et al., microphthalmia, a critical factor in melanocyte

development, defines a discrete transcription factor family. Genes Dev, 1994.

8(22): p. 2770-80.

24. Garraway, L.A., et al., Integrative genomic analyses identify MITF as a lineage

108

survival oncogene amplified in malignant melanoma. Nature, 2005. 436(7047): p.

117-22.

25. Goodall, J., et al., Brn-2 represses microphthalmia-associated transcription factor

expression and marks a distinct subpopulation of microphthalmia-associated

transcription factor-negative melanoma cells. Cancer Res, 2008. 68(19): p. 7788-

94.

26. Carreira, S., et al., Mitf regulation of Dia1 controls melanoma proliferation and

invasiveness. Genes Dev, 2006. 20(24): p. 3426-39.

27. Hsiao, J.J. and D.E. Fisher, The roles of microphthalmia-associated transcription

factor and pigmentation in melanoma. Arch Biochem Biophys, 2014. 563: p. 28-

34.

28. Hemesath, T.J., et al., MAP kinase links the transcription factor Microphthalmia

to c-Kit signalling in melanocytes. Nature, 1998. 391(6664): p. 298-301.

29. Mansky, K.C., et al., Microphthalmia transcription factor is a target of the p38

MAPK pathway in response to receptor activator of NF-kappa B ligand signaling.

J Biol Chem, 2002. 277(13): p. 11077-83.

30. Miller, A.J., et al., Sumoylation of MITF and its related family members TFE3

and TFEB. J Biol Chem, 2005. 280(1): p. 146-55.

31. Murakami, H. and H. Arnheiter, Sumoylation modulates transcriptional activity of

MITF in a promoter-specific manner. Pigment Cell Res, 2005. 18(4): p. 265-77.

32. Wu, M., et al., c-Kit triggers dual phosphorylations, which couple activation and

degradation of the essential melanocyte factor Mi. Genes Dev, 2000. 14(3): p.

109

301-12.

33. Goding, C.R., Commentary. A picture of Mitf in melanoma immortality.

Oncogene, 2011. 30(20): p. 2304-6.

34. Kobi, D., et al., Genome-wide analysis of POU3F2/BRN2 promoter occupancy in

human melanoma cells reveals Kitl as a novel regulated target gene. Pigment

Cell Melanoma Res, 2010. 23(3): p. 404-18.

35. Giuliano, S., et al., Microphthalmia-associated transcription factor controls the

DNA damage response and a lineage-specific senescence program in melanomas.

Cancer Res, 2010. 70(9): p. 3813-22.

36. Horn, P.J. and C.L. Peterson, Molecular biology. Chromatin higher order folding-

-wrapping up transcription. Science, 2002. 297(5588): p. 1824-7.

37. Wolffe, A.P., Chromatin remodeling: why it is important in cancer. Oncogene,

2001. 20(24): p. 2988-90.

38. Germond, J.E., et al., Folding of the DNA double helix in chromatin-like

structures from simian virus 40. Proc Natl Acad Sci U S A, 1975. 72(5): p. 1843-

7.

39. de la Serna, I.L. and A.N. Imbalzano, Unfolding heterochromatin for replication.

Nat Genet, 2002. 32(4): p. 560-2.

40. Aalfs, J.D. and R.E. Kingston, What does 'chromatin remodeling' mean? Trends

Biochem Sci, 2000. 25(11): p. 548-55.

41. Vignali, M., et al., ATP-dependent chromatin-remodeling complexes. Mol Cell

Biol, 2000. 20(6): p. 1899-910.

110

42. Mohrmann, L. and C.P. Verrijzer, Composition and functional specificity of

SWI2/SNF2 class chromatin remodeling complexes. Biochim Biophys Acta, 2005.

1681(2-3): p. 59-73.

43. Hargreaves, D.C. and G.R. Crabtree, ATP-dependent chromatin remodeling:

genetics, genomics and mechanisms. Cell Res, 2011. 21(3): p. 396-420.

44. Bozhenok, L., P.A. Wade, and P. Varga-Weisz, WSTF-ISWI chromatin

remodeling complex targets heterochromatic replication foci. EMBO J, 2002.

21(9): p. 2231-41.

45. Kehle, J., et al., dMi-2, a hunchback-interacting protein that functions in

polycomb repression. Science, 1998. 282(5395): p. 1897-900.

46. Kim, J., et al., Ikaros DNA-binding proteins direct formation of chromatin

remodeling complexes in lymphocytes. Immunity, 1999. 10(3): p. 345-55.

47. Morrison, A.J. and X. Shen, Chromatin remodelling beyond transcription: the

INO80 and SWR1 complexes. Nat Rev Mol Cell Biol, 2009. 10(6): p. 373-84.

48. de la Serna, I.L., Y. Ohkawa, and A.N. Imbalzano, Chromatin remodelling in

mammalian differentiation: lessons from ATP-dependent remodellers. Nat Rev

Genet, 2006. 7(6): p. 461-73.

49. Neigeborn, L. and M. Carlson, Genes affecting the regulation of SUC2 gene

expression by glucose repression in Saccharomyces cerevisiae. Genetics, 1984.

108(4): p. 845-58.

50. Peterson, C.L. and I. Herskowitz, Characterization of the yeast SWI1, SWI2, and

SWI3 genes, which encode a global activator of transcription. Cell, 1992. 68(3):

111

p. 573-83.

51. Ho, L., et al., An embryonic stem cell chromatin remodeling complex, esBAF, is

an essential component of the core pluripotency transcriptional network. Proc

Natl Acad Sci U S A, 2009. 106(13): p. 5187-91.

52. Ho, L., et al., An embryonic stem cell chromatin remodeling complex, esBAF, is

essential for embryonic stem cell self-renewal and pluripotency. Proc Natl Acad

Sci U S A, 2009. 106(13): p. 5181-6.

53. Lessard, J., et al., An essential switch in subunit composition of a chromatin

remodeling complex during neural development. Neuron, 2007. 55(2): p. 201-15.

54. Lickert, H., et al., Baf60c is essential for function of BAF chromatin remodelling

complexes in heart development. Nature, 2004. 432(7013): p. 107-12.

55. Gresh, L., et al., The SWI/SNF chromatin-remodeling complex subunit SNF5 is

essential for hepatocyte differentiation. EMBO J, 2005. 24(18): p. 3313-24.

56. Xu, Z., et al., Recruitment of the SWI/SNF protein Brg1 by a multiprotein complex

effects transcriptional repression in murine erythroid progenitors. Biochem J,

2006. 399(2): p. 297-304.

57. de la Serna, I.L., et al., MyoD targets chromatin remodeling complexes to the

myogenin locus prior to forming a stable DNA-bound complex. Mol Cell Biol,

2005. 25(10): p. 3997-4009.

58. Ohkawa, Y., et al., Myogenin and the SWI/SNF ATPase Brg1 maintain myogenic

gene expression at different stages of skeletal myogenesis. J Biol Chem, 2007.

282(9): p. 6564-70.

112

59. Gebuhr, T.C., et al., The role of Brg1, a catalytic subunit of mammalian

chromatin-remodeling complexes, in T cell development. J Exp Med, 2003.

198(12): p. 1937-49.

60. Yoo, A.S., et al., MicroRNA-mediated switching of chromatin-remodelling

complexes in neural development. Nature, 2009. 460(7255): p. 642-6.

61. Pedersen, T.A., et al., Cooperation between C/EBPalpha TBP/TFIIB and

SWI/SNF recruiting domains is required for adipocyte differentiation. Genes Dev,

2001. 15(23): p. 3208-16.

62. Roberts, C.W. and S.H. Orkin, The SWI/SNF complex--chromatin and cancer.

Nat Rev Cancer, 2004. 4(2): p. 133-42.

63. Kaeser, M.D., et al., BRD7, a novel PBAF-specific SWI/SNF subunit, is required

for target gene activation and repression in embryonic stem cells. J Biol Chem,

2008. 283(47): p. 32254-63.

64. Decristofaro, M.F., et al., Characterization of SWI/SNF protein expression in

human breast cancer cell lines and other malignancies. J Cell Physiol, 2001.

186(1): p. 136-45.

65. DeCristofaro, M.F., et al., Alteration of hSNF5/INI1/BAF47 detected in rhabdoid

cell lines and primary rhabdomyosarcomas but not Wilms' tumors. Oncogene,

1999. 18(52): p. 7559-65.

66. Reisman, D.N., et al., Loss of BRG1/BRM in human lung cancer cell lines and

primary lung cancers: correlation with poor prognosis. Cancer Res, 2003. 63(3):

p. 560-6.

113

67. Phelan, M.L., et al., Reconstitution of a core chromatin remodeling complex from

SWI/SNF subunits. Mol Cell, 1999. 3(2): p. 247-53.

68. Reyes, J.C., et al., Altered control of cellular proliferation in the absence of

mammalian brahma (SNF2alpha). EMBO J, 1998. 17(23): p. 6979-91.

69. Bultman, S., et al., A Brg1 null mutation in the mouse reveals functional

differences among mammalian SWI/SNF complexes. Mol Cell, 2000. 6(6): p.

1287-95.

70. Kadam, S. and B.M. Emerson, Transcriptional specificity of human SWI/SNF

BRG1 and BRM chromatin remodeling complexes. Mol Cell, 2003. 11(2): p. 377-

89.

71. Segal, E. and J. Widom, What controls nucleosome positions? Trends Genet,

2009. 25(8): p. 335-43.

72. Dawson, M.A., T. Kouzarides, and B.J. Huntly, Targeting epigenetic readers in

cancer. N Engl J Med, 2012. 367(7): p. 647-57.

73. Kouzarides, T., Chromatin modifications and their function. Cell, 2007. 128(4): p.

693-705.

74. Filippakopoulos, P. and S. Knapp, The bromodomain interaction module. FEBS

Lett, 2012. 586(17): p. 2692-704.

75. Dawson, M.A. and T. Kouzarides, : from mechanism to

therapy. Cell, 2012. 150(1): p. 12-27.

76. Fraga, M.F., et al., Loss of acetylation at Lys16 and trimethylation at Lys20 of

histone H4 is a common hallmark of human cancer. Nat Genet, 2005. 37(4): p.

114

391-400.

77. Bantscheff, M., et al., Chemoproteomics profiling of HDAC inhibitors reveals

selective targeting of HDAC complexes. Nat Biotechnol, 2011. 29(3): p. 255-65.

78. Barski, A., et al., High-resolution profiling of histone in the human

genome. Cell, 2007. 129(4): p. 823-37.

79. Hanahan, D. and R.A. Weinberg, Hallmarks of cancer: the next generation. Cell,

2011. 144(5): p. 646-74.

80. Chung, C.W. and J. Witherington, Progress in the discovery of small-molecule

inhibitors of bromodomain--histone interactions. J Biomol Screen, 2011. 16(10):

p. 1170-85.

81. Taverna, S.D., et al., Long-distance combinatorial linkage between methylation

and acetylation on histone H3 N termini. Proc Natl Acad Sci U S A, 2007. 104(7):

p. 2086-91.

82. Barbieri, I., E. Cannizzaro, and M.A. Dawson, Bromodomains as therapeutic

targets in cancer. Brief Funct Genomics, 2013. 12(3): p. 219-30.

83. Tamkun, J.W., et al., brahma: a regulator of Drosophila homeotic genes

structurally related to the yeast transcriptional activator SNF2/SWI2. Cell, 1992.

68(3): p. 561-72.

84. Haynes, S.R., et al., The bromodomain: a found in human,

Drosophila and yeast proteins. Nucleic Acids Res, 1992. 20(10): p. 2603.

85. Jeanmougin, F., et al., The bromodomain revisited. Trends Biochem Sci, 1997.

22(5): p. 151-3.

115

86. Zeng, L. and M.M. Zhou, Bromodomain: an acetyl-lysine binding domain. FEBS

Lett, 2002. 513(1): p. 124-8.

87. Sanchez, R. and M.M. Zhou, The role of human bromodomains in chromatin

biology and gene transcription. Curr Opin Drug Discov Devel, 2009. 12(5): p.

659-65.

88. Brown, C.E., et al., Recruitment of HAT complexes by direct activator

interactions with the ATM-related Tra1 subunit. Science, 2001. 292(5525): p.

2333-7.

89. Sterner, D.E., et al., Functional organization of the yeast SAGA complex: distinct

components involved in structural integrity, nucleosome acetylation, and TATA-

binding protein interaction. Mol Cell Biol, 1999. 19(1): p. 86-98.

90. Belkina, A.C. and G.V. Denis, BET domain co-regulators in obesity,

inflammation and cancer. Nat Rev Cancer, 2012. 12(7): p. 465-77.

91. Mujtaba, S., L. Zeng, and M.M. Zhou, Structure and acetyl-lysine recognition of

the bromodomain. Oncogene, 2007. 26(37): p. 5521-7.

92. Jacobson, R.H., et al., Structure and function of a human TAFII250 double

bromodomain module. Science, 2000. 288(5470): p. 1422-5.

93. Mujtaba, S., et al., Structural basis of lysine-acetylated HIV-1 Tat recognition by

PCAF bromodomain. Mol Cell, 2002. 9(3): p. 575-86.

94. Mujtaba, S., et al., Structural mechanism of the bromodomain of the coactivator

CBP in p53 transcriptional activation. Mol Cell, 2004. 13(2): p. 251-63.

95. Nakamura, Y., et al., Crystal structure of the human BRD2 bromodomain:

116

insights into dimerization and recognition of acetylated histone H4. J Biol Chem,

2007. 282(6): p. 4193-201.

96. Li, H., et al., Molecular basis for site-specific read-out of histone H3K4me3 by

the BPTF PHD finger of NURF. Nature, 2006. 442(7098): p. 91-5.

97. Shen, W., et al., Solution structure of human Brg1 bromodomain and its specific

binding to acetylated histone tails. Biochemistry, 2007. 46(8): p. 2100-10.

98. Marmorstein, R. and S.L. Berger, Structure and function of bromodomains in

chromatin-regulating complexes. Gene, 2001. 272(1-2): p. 1-9.

99. Sanchez, R., J. Meslamani, and M.M. Zhou, The bromodomain: from epigenome

reader to druggable target. Biochim Biophys Acta, 2014. 1839(8): p. 676-85.

100. Plotnikov, A.N., et al., Structural insights into acetylated-histone H4 recognition

by the bromodomain-PHD finger module of human transcriptional coactivator

CBP. Structure, 2014. 22(2): p. 353-60.

101. Zeng, L., et al., Structural basis of site-specific histone recognition by the

bromodomains of human coactivators PCAF and CBP/p300. Structure, 2008.

16(4): p. 643-52.

102. Zeng, L., et al., Structural insights into human KAP1 PHD finger-bromodomain

and its role in gene silencing. Nat Struct Mol Biol, 2008. 15(6): p. 626-33.

103. Dhalluin, C., et al., Structure and ligand of a histone acetyltransferase

bromodomain. Nature, 1999. 399(6735): p. 491-6.

104. Filippakopoulos, P., et al., Histone recognition and large-scale structural analysis

of the human bromodomain family. Cell, 2012. 149(1): p. 214-31.

117

105. Moriniere, J., et al., Cooperative binding of two acetylation marks on a histone

tail by a single bromodomain. Nature, 2009. 461(7264): p. 664-8.

106. Gamsjaeger, R., et al., Structural basis and specificity of acetylated transcription

factor GATA1 recognition by BET family bromodomain protein Brd3. Mol Cell

Biol, 2011. 31(13): p. 2632-40.

107. Fu, L.L., et al., Inhibition of BET bromodomains as a therapeutic strategy for

cancer drug discovery. Oncotarget, 2015. 6(8): p. 5501-16.

108. Sanchez, R., et al., Protein structure modeling for structural genomics. Nat Struct

Biol, 2000. 7 Suppl: p. 986-90.

109. Schultz, J., et al., SMART: a web-based tool for the study of genetically mobile

domains. Nucleic Acids Res, 2000. 28(1): p. 231-4.

110. Kanno, T., et al., Selective recognition of acetylated histones by bromodomain

proteins visualized in living cells. Mol Cell, 2004. 13(1): p. 33-43.

111. Dey, A., et al., The double bromodomain protein Brd4 binds to acetylated

chromatin during interphase and mitosis. Proc Natl Acad Sci U S A, 2003.

100(15): p. 8758-63.

112. Dey, A., et al., A bromodomain protein, MCAP, associates with mitotic

and affects G(2)-to-M transition. Mol Cell Biol, 2000. 20(17): p.

6537-49.

113. Prinjha, R.K., J. Witherington, and K. Lee, Place your BETs: the therapeutic

potential of bromodomains. Trends Pharmacol Sci, 2012. 33(3): p. 146-53.

114. Kalkhoven, E., CBP and p300: HATs for different occasions. Biochem

118

Pharmacol, 2004. 68(6): p. 1145-55.

115. Tanaka, Y., et al., Extensive brain hemorrhage and embryonic lethality in a

mouse null mutant of CREB-binding protein. Mech Dev, 2000. 95(1-2): p. 133-45.

116. Yao, T.P., et al., Gene dosage-dependent embryonic development and

proliferation defects in mice lacking the transcriptional integrator p300. Cell,

1998. 93(3): p. 361-72.

117. Bu, P., et al., Loss of Gcn5 acetyltransferase activity leads to neural tube closure

defects and exencephaly in mouse embryos. Mol Cell Biol, 2007. 27(9): p. 3405-

16.

118. Maurice, T., et al., Altered memory capacities and response to stress in

p300/CBP-associated factor (PCAF) histone acetylase knockout mice.

Neuropsychopharmacology, 2008. 33(7): p. 1584-602.

119. Zou, J.X., et al., Androgen-induced coactivator ANCCA mediates specific

androgen receptor signaling in prostate cancer. Cancer Res, 2009. 69(8): p.

3339-46.

120. Wassarman, D.A. and F. Sauer, TAF(II)250: a transcription toolbox. J Cell Sci,

2001. 114(Pt 16): p. 2895-902.

121. Kalla, C., et al., Translocation t(X;11)(q13;q23) in B-cell chronic lymphocytic

leukemia disrupts two novel genes. Genes Chromosomes Cancer, 2005. 42(2): p.

128-43.

122. Ramos, V.C., et al., Characterisation and expression analysis of the WDR9 gene,

located in the Down critical region-2 of the human chromosome 21. Biochim

119

Biophys Acta, 2002. 1577(3): p. 377-83.

123. Huang, H., et al., Expression of the Wdr9 gene and protein products during mouse

development. Dev Dyn, 2003. 227(4): p. 608-14.

124. Le Douarin, B., et al., A possible involvement of TIF1 alpha and TIF1 beta in the

epigenetic control of transcription by nuclear receptors. EMBO J, 1996. 15(23):

p. 6701-15.

125. Friedman, J.R., et al., KAP-1, a novel corepressor for the highly conserved KRAB

repression domain. Genes Dev, 1996. 10(16): p. 2067-78.

126. Kim, J. and V. Kaartinen, Generation of mice with a conditional allele for

Trim33. Genesis, 2008. 46(6): p. 329-33.

127. Khetchoumian, K., et al., TIF1delta, a novel HP1-interacting member of the

transcriptional intermediary factor 1 (TIF1) family expressed by elongating

. J Biol Chem, 2004. 279(46): p. 48329-41.

128. Singhal, N., et al., Chromatin-Remodeling Components of the BAF Complex

Facilitate Reprogramming. Cell, 2010. 141(6): p. 943-55.

129. Burrows, A.E., A. Smogorzewska, and S.J. Elledge, Polybromo-associated

BRG1-associated factor components BRD7 and BAF180 are critical regulators of

p53 required for induction of replicative senescence. Proc Natl Acad Sci U S A,

2010. 107(32): p. 14280-5.

130. Muller, S., P. Filippakopoulos, and S. Knapp, Bromodomains as therapeutic

targets. Expert Rev Mol Med, 2011. 13: p. e29.

131. Collins, N., et al., An ACF1-ISWI chromatin-remodeling complex is required for

120

DNA replication through heterochromatin. Nat Genet, 2002. 32(4): p. 627-32.

132. Strohner, R., et al., NoRC--a novel member of mammalian ISWI-containing

chromatin remodeling machines. EMBO J, 2001. 20(17): p. 4892-900.

133. Garnier, J.M., P.P. Sharp, and C.J. Burns, BET bromodomain inhibitors: a patent

review. Expert Opin Ther Pat, 2014. 24(2): p. 185-99.

134. Shang, E., et al., Double bromodomain-containing gene Brd2 is essential for

embryonic development in mouse. Dev Dyn, 2009. 238(4): p. 908-17.

135. Gyuris, A., et al., The chromatin-targeting protein Brd2 is required for neural

tube closure and embryogenesis. Biochim Biophys Acta, 2009. 1789(5): p. 413-

21.

136. Pal, D.K., et al., BRD2 (RING3) is a probable major susceptibility gene for

common juvenile myoclonic epilepsy. Am J Hum Genet, 2003. 73(2): p. 261-70.

137. Shang, E., et al., The first bromodomain of Brdt, a testis-specific member of the

BET sub-family of double-bromodomain-containing proteins, is essential for male

germ cell differentiation. Development, 2007. 134(19): p. 3507-15.

138. Steilmann, C., et al., The interaction of modified histones with the bromodomain

testis-specific (BRDT) gene and its mRNA level in sperm of fertile donors and

subfertile men. Reproduction, 2010. 140(3): p. 435-43.

139. Aston, K.I., et al., Evaluation of 172 candidate polymorphisms for association

with oligozoospermia or in a large cohort of men of European

descent. Hum Reprod, 2010. 25(6): p. 1383-97.

140. Chiang, C.M., Brd4 engagement from chromatin targeting to transcriptional

121

regulation: selective contact with acetylated histone H3 and H4. F1000 Biol Rep,

2009. 1: p. 98.

141. Huang, B., et al., Brd4 coactivates transcriptional activation of NF-kappaB via

specific binding to acetylated RelA. Mol Cell Biol, 2009. 29(5): p. 1375-87.

142. Zhao, R., et al., Gene bookmarking accelerates the kinetics of post-mitotic

transcriptional re-activation. Nat Cell Biol, 2011. 13(11): p. 1295-304.

143. Devaiah, B.N. and D.S. Singer, Two faces of : mitotic bookmark and

transcriptional lynchpin. Transcription, 2013. 4(1): p. 13-7.

144. Ai, N., et al., Signal-induced Brd4 release from chromatin is essential for its role

transition from chromatin targeting to transcriptional regulation. Nucleic Acids

Res, 2011. 39(22): p. 9592-604.

145. Wu, S.Y. and C.M. Chiang, The double bromodomain-containing chromatin

adaptor Brd4 and transcriptional regulation. J Biol Chem, 2007. 282(18): p.

13141-5.

146. Dey, A., et al., Brd4 marks select genes on mitotic chromatin and directs

postmitotic transcription. Mol Biol Cell, 2009. 20(23): p. 4899-909.

147. Mochizuki, K., et al., The bromodomain protein Brd4 stimulates G1 gene

transcription and promotes progression to S phase. J Biol Chem, 2008. 283(14):

p. 9040-8.

148. Crawford, N.P., et al., Bromodomain 4 activation predicts breast cancer survival.

Proc Natl Acad Sci U S A, 2008. 105(17): p. 6380-5.

149. Grayson, A.R., et al., MYC, a downstream target of BRD-NUT, is necessary and

122

sufficient for the blockade of differentiation in NUT midline carcinoma.

Oncogene, 2014. 33(13): p. 1736-42.

150. Du, C., et al., Effects of myogenin on expression of late muscle genes through

MyoD-dependent chromatin remodeling ability of myogenin. Mol Cells, 2012.

34(2): p. 133-42.

151. Zhou, M., et al., Bromodomain protein Brd4 regulates human immunodeficiency

virus transcription through phosphorylation of CDK9 at threonine 29. J Virol,

2009. 83(2): p. 1036-44.

152. Gagnon, D., et al., Proteasomal degradation of the papillomavirus E2 protein is

inhibited by overexpression of bromodomain-containing protein 4. J Virol, 2009.

83(9): p. 4127-39.

153. Ottinger, M., et al., Kaposi's sarcoma-associated herpesvirus LANA-1 interacts

with the short variant of BRD4 and releases cells from a BRD4- and

BRD2/RING3-induced G1 cell cycle arrest. J Virol, 2006. 80(21): p. 10772-86.

154. Ottinger, M., et al., The interaction of the gammaherpesvirus 68 orf73 protein

with cellular BET proteins affects the activation of cell cycle promoters. J Virol,

2009. 83(9): p. 4423-34.

155. Houzelstein, D., et al., Growth and early postimplantation defects in mice

deficient for the bromodomain-containing protein Brd4. Mol Cell Biol, 2002.

22(11): p. 3794-802.

156. Zuber, J., et al., RNAi screen identifies Brd4 as a therapeutic target in acute

myeloid leukaemia. Nature, 2011. 478(7370): p. 524-8.

123

157. Segura, M.F., et al., BRD4 sustains melanoma proliferation and represents a new

target for epigenetic therapy. Cancer Res, 2013. 73(20): p. 6264-76.

158. Rodriguez, R.M., et al., Aberrant epigenetic regulation of bromodomain BRD4 in

human colon cancer. J Mol Med (Berl), 2012. 90(5): p. 587-95.

159. Alsarraj, J. and K.W. Hunter, Bromodomain-Containing Protein 4: A Dynamic

Regulator of Breast Cancer Metastasis through Modulation of the Extracellular

Matrix. Int J Breast Cancer, 2012. 2012: p. 670632.

160. Paillisson, A., et al., Bromodomain testis-specific protein is expressed in mouse

oocyte and evolves faster than its ubiquitously expressed paralogs BRD2, -3, and

-4. Genomics, 2007. 89(2): p. 215-23.

161. Rahman, S., et al., The Brd4 extraterminal domain confers transcription

activation independent of pTEFb by recruiting multiple proteins, including NSD3.

Mol Cell Biol, 2011. 31(13): p. 2641-52.

162. Filippakopoulos, P., et al., Selective inhibition of BET bromodomains. Nature,

2010. 468(7327): p. 1067-73.

163. Niesen, F.H., H. Berglund, and M. Vedadi, The use of differential scanning

fluorimetry to detect ligand interactions that promote protein stability. Nat

Protoc, 2007. 2(9): p. 2212-21.

164. Quinn, A.M., et al., A homogeneous method for investigation of methylation-

dependent protein-protein interactions in epigenetics. Nucleic Acids Res, 2010.

38(2): p. e11.

165. Marquez, S.B., et al., Beyond Mutations: Additional Mechanisms and

124

Implications of SWI/SNF Complex Inactivation. Front Oncol, 2014. 4: p. 372.

166. Hohmann, A.F. and C.R. Vakoc, A rationale to target the SWI/SNF complex for

cancer therapy. Trends Genet, 2014. 30(8): p. 356-63.

167. Bandyopadhyay, D., et al., Dynamic assembly of chromatin complexes during

cellular senescence: implications for the growth arrest of human melanocytic

nevi. Aging Cell, 2007. 6(4): p. 577-91.

168. Lin, H., et al., BRG1 expression is increased in human cutaneous melanoma. Br J

Dermatol, 2010. 163(3): p. 502-10.

169. Becker, T.M., et al., The chromatin remodelling factor BRG1 is a novel binding

partner of the tumor suppressor p16INK4a. Mol Cancer, 2009. 8: p. 4.

170. Saladi, S.V., et al., Modulation of extracellular matrix/adhesion molecule

expression by BRG1 is associated with increased melanoma invasiveness. Mol

Cancer, 2010. 9: p. 280.

171. Saladi, S.V., et al., BRG1 promotes survival of UV-irradiated melanoma cells by

cooperating with MITF to activate the melanoma inhibitor of apoptosis gene.

Pigment Cell Melanoma Res, 2013. 26(3): p. 377-91.

172. Varela, I., et al., Exome sequencing identifies frequent mutation of the SWI/SNF

complex gene PBRM1 in renal carcinoma. Nature, 2011. 469(7331): p. 539-42.

173. Xia, W., et al., BAF180 is a critical regulator of p21 induction and a tumor

suppressor mutated in breast cancer. Cancer Res, 2008. 68(6): p. 1667-74.

174. Yan, Z., et al., PBAF chromatin-remodeling complex requires a novel specificity

subunit, BAF200, to regulate expression of selective interferon-responsive genes.

125

Genes Dev, 2005. 19(14): p. 1662-7.

175. Mehrotra, A., et al., SWI/SNF chromatin remodeling enzymes in melanocyte

differentiation and melanoma. Crit Rev Eukaryot Gene Expr, 2014. 24(2): p. 151-

61.

176. Hodis, E., et al., A landscape of driver mutations in melanoma. Cell, 2012.

150(2): p. 251-63.

177. Zhou, J., et al., BRD7, a novel bromodomain gene, inhibits G1-S progression by

transcriptionally regulating some important molecules involved in ras/MEK/ERK

and Rb/ pathways. J Cell Physiol, 2004. 200(1): p. 89-98.

178. Wilson, B.G. and C.W. Roberts, SWI/SNF nucleosome remodellers and cancer.

Nat Rev Cancer, 2011. 11(7): p. 481-92.

179. Shields, J.M., et al., Lack of extracellular signal-regulated kinase mitogen-

activated protein kinase signaling shows a new type of melanoma. Cancer Res,

2007. 67(4): p. 1502-12.

180. Liu, T.F., G. Kandala, and V. Setaluri, PDZ domain protein GIPC interacts with

the cytoplasmic tail of melanosomal membrane protein gp75 (tyrosinase-related

protein-1). J Biol Chem, 2001. 276(38): p. 35768-77.

181. Eller, M.S., K. Ostrom, and B.A. Gilchrest, DNA damage enhances

melanogenesis. Proc Natl Acad Sci U S A, 1996. 93(3): p. 1087-92.

182. Giresi, P.G., et al., FAIRE (Formaldehyde-Assisted Isolation of Regulatory

Elements) isolates active regulatory elements from human chromatin. Genome

Res, 2007. 17(6): p. 877-85.

126

183. Guo, H., J.A. Carlson, and A. Slominski, Role of TRPM in melanocytes and

melanoma. Exp Dermatol, 2012. 21(9): p. 650-4.

184. Chen, J. and T.K. Archer, Regulating SWI/SNF subunit levels via protein-protein

interactions and proteasomal degradation: BAF155 and BAF170 limit expression

of BAF57. Mol Cell Biol, 2005. 25(20): p. 9016-27.

185. Lo, J.A. and D.E. Fisher, The melanoma revolution: from UV carcinogenesis to a

new era in therapeutics. Science, 2014. 346(6212): p. 945-9.

186. Harms, K., S. Nozell, and X. Chen, The common and distinct target genes of the

p53 family transcription factors. Cell Mol Life Sci, 2004. 61(7-8): p. 822-42.

187. Krauthammer, M., et al., Exome sequencing identifies recurrent somatic RAC1

mutations in melanoma. Nat Genet, 2012. 44(9): p. 1006-14.

127