Chromatin Regulators and Transcriptional Control of Drosophila Development

Qi Dai

Stockholm University

The cover shows transgenic Drosophila embryos hybridized with a lacZ RNA probe. The lacZ reporter gene is driven by a modified rho neuroectoderm en- hancer (NEE) with synthetic Snail binding sites. The endogenous Snail tran- scription factor represses reporter gene expression in the wild-type (wt) embryo, but fails to repress in an embryo devoid of the co-repressor Ebi (ebi).

Qi Dai, Stockholm 2007

ISBN (978-91-7155-547-2)

Printed in Sweden by US-AB, Stockholm 2007 Distributor: Wenner-Gren Institute, Division of Developmental Biology

To my beloved family

Abstract

The development of a multicellular organism is programmed by complex temporal and spatial patterns of gene expression. In eukaryotic cells, DNA is packaged by histone into chromatin. Chromatin regulators affect gene expression by influencing access of proteins to DNA and often function as transcription co-factors. The interplay between transcriptional activators and repressors is modulated by co-factors at cis-regulatory DNA modules (CRMs) to generate precise gene expression patterns. In this study, we have investigated the in vivo function of four co-factors, dAda2b, Reptin, Ebi and Brakeless during development of the fruit fly Dro- sophila melanogaster. dAda2b and Reptin belong to histone acetyl trans- ferase (HAT) complexes, a SAGA-like complex and the Tip60 complex, respectively. We generated dAda2b mutants and found that lack of dAda2b strongly affects global histone acetylation and viability. dAda2b mutants are hyper-sensitive to irradiation-induced DNA damage. Similarly, yeast SAGA mutants are sensitive to the genetoxic agent MMS. We propose that Ada2 and its associated complexes may be involved in DNA repair. Reptin belongs to several chromatin regulatory complexes including the Tip60 HAT complex which is actively involved in transcription, DNA repair and other cellular processes. Our studies revealed new roles of Reptin and other Tip60 complex components in Polycomb Group mediated repression and heterochromatin formation, thereby promoting generation of silent chromatin. During embryogenesis, transcriptional repressors establish localized and tissue-specific patterns of gene expression. Their activities often utilize ubiquitously distributed co-repressors. In this thesis, we identified two novel co-repressors in the early embryo, Ebi and Brakeless. Ebi genetically and physically interacts with the Snail repressor. The Ebi-interaction motif in the Snail is essential for Snail function in vivo and is evolutionarily con- served in insects. We further demonstrated that Ebi associates with histone deacetylase 3 (HDAC3) and that histone deacetylation is part of the mecha- nism by which Snail mediates transcriptional repression. We isolated Brakeless in a genetic screen for novel regulators of gene ex- pression during embryogenesis. We found that mutation of brakeless impairs function of the Tailless repressor. Brakeless and Tailless bind to each other and interact genetically. Brakeless also associates with Atrophin, another Tailless corepressor, and both are recruited to a Tailless-regulated gene in embryos where they function together in Tailless-mediated repression. In summary, transcription co-factors, including chromatin regulators, are selectively required in distinct processes during development.

List of papers

I Dai Qi, Jan Larsson, and Mattias Mannervik (2004) Drosophila Ada2b is required for viability and normal histone H3 acetylation. Mol Cell Biol. 24(18):8080-9.

II Dai Qi, Haining Jin, Tobias Lilja, and Mattias Mannervik (2006) Drosophila Reptin and other TIP60 complex components pro- mote generation of silent chromatin. Genetics. 174(1):241-51.

III Dai Qi, Mattias Bergman, Hitoshi Aihara, Yutaka Nibu, and Mattias Mannervik (2007) Drosophila Ebi mediates Snail- dependent transcriptional repression through HDAC3-induced histone deacetylation. (Submitted).

IV Achim Haecker, Dai Qi, Tobias Lilja, Bernard Moussian, Luiz Paulo Andrioli, Stefan Luschnig, and Mattias Mannervik (2007) Drosophila Brakeless interacts with Atrophin and is required for Tailless-mediated transcriptional repression in early embryos. PLoS Biol. 5(6):1298-1308.

Articles are reprinted with permission from publishers; American Society for Microbiology, Genetics Society of America and Public Library of Science

Abbreviations

Ada2 Adaptor protein 2 AP axis Anterior-posterior axis ATM Ataxia-telangiectasia mutated ATR ATM and Rad-3 related Atro Atrophin Bks Brakeless CRM Cis regulatory module CtBP C-terminal binding protein DV axis Dorsal-ventral axis Gro Groucho GTF General H3K9 Histone H3 lysine 9 HAT Histone acetyl transferase Hb Hunchback HDAC Histone deacetylase HMT Histone methyl transferase HP1 Heterochromatin protein 1 Knirps Kni Kruppel Kr NCoR Nuclear corepressor PcG Polycomb group PEV Position effect variegation PRC1 Polycomb repressive complex 1 Rho Rhomboid Sim Single-minded SMRT Silencing mediator for retinoic acid and thyroid hormone receptors Sog Short-gastrulation TF Transcription factor Tip60 Tat interaction protein 60kD Tll Tailless TrxG Trithorax group

Contents

Introduction...... 11 Transcription machinery in eukaryotes...... 11 Chromatin and its function ...... 12

Drosophila as a model system ...... 13 The life cycle of Drosophila...... 13 Early embryonic development of Drosophila...... 14 Dorsal-ventral patterning...... 15 Snail in gastrulation and mesoderm formation ...... 16 Anterior-posterior patterning...... 17

Histone modifications ...... 20 Functions of histone modifications ...... 21 Histone acetylation and its function ...... 22 Histone acetylation in gene regulation...... 22 Histone acetylation in DNA repair...... 23 Histone acetyl-transferase complexes...... 25 GNAT complexes ...... 26 Ada2 proteins...... 27 Tip60 HAT complexes ...... 28 Reptin...... 29 Histone deacetylases and their associated complexes...... 31 NCoR/SMRT/HDAC3 complex...... 32 Epigenetic control ...... 33 Heterochromatin and position effect variegation...... 33 Polycomb group and Trithorax group...... 34

Co-regulators...... 36 CtBP...... 38 Groucho ...... 38 Atrophin and Brakeless...... 39

Aim of this thesis ...... 41

Results and discussion of project...... 42 Paper I ...... 42 Paper II ...... 43

Paper III ...... 44 Paper IV ...... 45

Conclusion and Perspectives ...... 47

Acknowledgements ...... 48

References ...... 50

Introduction

A human being contains around 200 different cell types, all of which arise from a single cell, the fertilized egg. This is probably the major issue that fascinates developmental biologists, how a single-cell egg can develop into a mature adult organism. The developmental process is directed by genetic information that is stored in the fertilized egg. In the cell, DNA molecules carry the genetic information units, called genes. Genes are first transcribed into RNAs which subsequently deliver (translate) the genetic code to pro- teins, the fundamental performers of most functions in the cell. This is called the central dogma of molecular biology and is valid in most organisms except RNA viruses. By controlling where and when proteins are synthe- sized, genes control development. In other words, the genetically identical cells come to differ from one another by expressing distinct sets of genes during development.

Transcription machinery in eukaryotes

TF TF Chromatin regulators Mediator

B RNA pol II A D F TATA E H

Figure 1. Typical transcription machinery of eukaryotic genes. RNA Pol II binds to the together with GTFs (TFIIA, B, D, E, F and H). TATA box is the recognition site by a subunit of TFIID, TATA box binding protein (TBP). Tran- scription factors (TFs) bind CRMs in DNA. The mediator complexes bridge GTFs and TFs. Chromatin regulators change the status of chromatin.

Gene expression can be regulated at several steps, such as transcriptional initiation, transcriptional elongation, RNA splicing and translation. Most of eukaryotic genes are transcribed by RNA polymerase II (Pol II). RNA poly-

11

merase II recognizes transcriptional start sites with help from general tran- scription factors (GTFs) consisting of TFIID, TFIIA, TFIIB, TFIIE, TFIIF and TFIIH. Additionally, the typical transcription machinery includes media- tor complexes, accessory complexes including chromatin regulators and transcription factors (TFs) (Figure 1) (reviewed in Hahn, 2004). TFs bind to DNA and function as either activators that help transcription initiation or repressors that hinder transcription initiation.

Chromatin and its function The genome of a eukaryote is wrapped in proteins called histones to form nucleosomes. One nucleosome contains a histone octamer (an H3/H4 tetramer and two H2A/H2B dimers) wrapped around 147 base pairs of DNA (Luger et al., 1997). Numerous nucleosomes build up chromatin. Chromatin is packaged into chromosomes. So the transcription machinery is present with a partially concealed substrate. Changing the accessibility of DNA for the transcription machinery is an important mean for regulating gene expres- sion. Histones contribute to the function of chromatin through at least three principles: First, chromatin-remodeling factors, such as the SWI/SNF com- plexes alter the position or properties of the nucleosomes by affecting the stability of histone-DNA interactions (Peterson, 2002). Second, the modifi- cations of histones regulate accessibility of chromatin for the cellular ma- chinery (Strahl and Allis, 2000). Finally, the exchange of core histones with specialized histone variants, such as H2AX and H3.3, affect the composition and function of nucleosomes (Smith, 2002).

12

Drosophila as a model system

The basic principle of development and many developmental genes are con- served in different organisms. We used the fruitfly Drosophila melanogaster as a model system to investigate gene function in development. The fruitfly has lots of basic advantages, such as inexpensive culturing condition, rapid life cycle, large amount of progenies, and ease of manipulation. It also offers many sophisticated tools of classical and molecular genetics. In Drosophila, it is possible to genetically investigate the function of endogenously and exogenously introduced genes in a way that is impossible in humans and impractical in mice. For example, large scale screens using chemical- induced mutagenesis can be used to analyze gene functions involved in spe- cific developmental and cellular processes. Transposon-based technologies are used to introduce transgenes, cause lethal mutations and screen for gene expression patterns. It is also very efficient to create mosaic clones in so- matic tissue as well as in the germ line in Drosophila. Furthermore, the genes of interest can be inactivated and misexpressed in a tissue of choice. Taken together, Drosophila provides us a very useful system to study the in vivo function of regulators in gene expression.

The life cycle of Drosophila The Drosophila life cycle starts from a very rapid embryonic stage (0-24 hours after fertilization (AF)), followed by three larval stages: the first instar L1 (24-48 hours AF), the second instar L2 (48-72 hours) and the third instar L3 (72-120 hours AF), and then the pupa stage (metamorphosis) (day 6-9 AF). After pupariation, the animal develops into an adult fly. The early Dro- sophila embryo is one of the best-characterized developmental systems and provides a particularly useful framework for the analysis of gene function (see below). During larval development, one of the notable events is the growth of imaginal discs. They are initially small and composed of fewer than 100 cells in the mid-stage embryo but contain tens of thousands of cells in the mature larva. There is a pair of discs for every set of appendages (for example, labia, antennae, eyes, wings, three pairs of legs and genitalia). Imaginal discs differentiate into their appropriate adult structures during metamorphosis. Another useful system are the polytene chromosomes in the

13

salivary glands in the third instar larva. The polytene chromosomes result from successive rounds of replication without nuclear division or cytokinesis. They are particularly useful in studying chromosome structure and localiza- tion of chromatin binding proteins.

Early embryonic development of Drosophila

amnioserosa B A

zen Dorsal

Dorsal ectoderm Acron

s Telson

o

g

o Dorsal gradient h r Posterior Ventral ectoderm Anterior Sna Ventral Head Tail Mesoderm Thorax Abdomen Figure 2. The dorsal-ventral axis and anterior-posterior axis of Drosophila. A.The schematic picture shows a section view of a pre-cellularization stage embryo. The nuclear gradient of the Dorsal protein specifies four regions along DV axis from ven- tral to dorsal side: the mesoderm, ventral ectoderm, dorsal ectoderm and amnioserosa. sna, rho, sog and zen are representative genes regulated by Dorsal in these regions. B. Picture of cuticle preparation of a first instar larva. From anterior to posterior, the main body of the animal is divided into the head, thorax and abdomen. The terminal structures include acron and telson.

Before fertilization, many important genes are transcribed to mRNAs and/or translated to proteins by the mother and then these gene products are depos- ited into the egg. In this way, the mother fly has stored enough material in the egg to initiate the early embryonic development (reviewed in St Johnston and Nusslein-Volhard, 1992). After fertilization, the diploid, zygotic nucleus undergoes a series of nearly synchronous divisions which results in a syncytium, meaning a single cell with multiple nuclei. Then most of the resulting nuclei begin to migrate to the periphery. The end result of the mi- gration and more divisions is the formation of a monolayer of approximately 6,000 nuclei surrounding the central yolk. During the following one hour, from 2 to 3 hours after fertilization, cell membranes form between adjacent nuclei. This process is called cellularization. During this short period, each nucleus is rapidly specified to follow a particular pathway of differentiation. The dorsal-ventral (DV) axis and anterior-posterior (AP) axis are both de- termined by this time. Along the DV axis, the embryo becomes divided up into four regions: from ventral to dorsal these are the mesoderm, ventral ectoderm (neurogenic ectoderm), dorsal ectoderm and the amnioserosa (Fig-

14

ure 2A). Along the AP axis, the embryo is divided into three broad regions which will become the head, thorax and abdomen (Figure 2B).

Dorsal-ventral patterning Dorsal-ventral patterning is implemented by the graded distribution of a maternal transcription factor, Dorsal, and culminates in the differentiation of specialized tissues (reviewed in St Johnston and Nusslein-Volhard, 1992). Approximately 3-5 hours after fertilization, each of these tissues generates multiple cell types. The key player in this process is the Dorsal protein. It is deposited by the mother and initially distributed throughout the cytoplasm of the unfertilized egg. After fertilization, the Dorsal protein forms a nuclear gradient along dorsal-ventral axis with the highest concentration in the ventral most region and provides positional information for the determination of different cell fates. The generation of the nuclear gradient of Dorsal originates from oogenesis and involves at least 17 maternal factors (reviewed in Moussian and Roth, 2005). The procedure can be generally divided into two steps. First, a gradient of the activated Spaetzle ligand is established from ventral to dorsal in the extracellular matrix (Morisato, 2001; Roth, 2003). Second, Spaetzle binds to the cell surface receptor Toll and thus triggers another sig- naling cascade inside the cell, which finally induces the translocation of Dorsal from the cytoplasm into nuclei (Bergmann et al., 1996; Towb et al., 1998; Yang and Steward, 1997). Once this nuclear gradient is achieved, Dorsal functions as a transcrip- tional regulator, leading to the differential expression of nearly 50 genes across the dorsal-ventral axis (Stathopoulos et al., 2002). Roughly half the genes encode transcriptional factors, whereas the other half encodes signal- ing components. They play fundamental roles in embryogenesis. The Dorsal gradient specifies multiple thresholds of gene activation. High levels of Dor- sal activate the transcription of twist and snail in the ventral-most cells, which invaginate into the embryo to form mesoderm (Rusch and Levine, 1996). In the lateral region, intermediate levels of Dorsal activate genes (e.g. rhomboid (rho)) that are required for the formation of ventral neuroectoderm. In the more lateral region, low levels of Dorsal are sufficient to activate the short-gastrulation (sog) gene in both the ventral and dorsal neurogenic ecto- derm. The neuroectoderm gives rise to the ventral nerve cord and the ventral epidermis. In principle, rho and sog can also be activated by high levels of Dorsal. However, their expression is precluded from the ventral most cells due to repression by Snail. How Dorsal produces so many different tran- scription outputs is extensively studied (Stathopoulos and Levine, 2002). It was suggested that the nature of the Dorsal binding sites (affinity, number)

15

determines the binding efficiency and thus determines how much nuclear Dorsal is sufficient. For example, the twist and snail cis regulatory modules (CRM) only contain low-affinity binding sites, so they need highest level of Dorsal. However, the rho CRM contains one high-affinity site and a few low-affinity sites, which allow both high and intermediate levels of Dorsal to activate. Finally, the sog CRM contains four evenly-spaced optimal sites, which allows lowest levels of Dorsal to activate transcription. The combina- tory role with other factors is also involved in regulating Dorsal target genes. Twist protein regulates nearly half of the known Dorsal target genes syner- gistically once it is synthesized in the presumptive mesoderm. Snail also regulates a large number of Dorsal target genes, but functions as a repressor (Leptin, 1991). In this way, the broadly distributed activator and the local- ized repressor set up restricted expression of their target genes. Dorsal protein can repress transcription as well (Jiang et al., 1993; Pan and Courey, 1992). Some genes, like zerknullt (zen) and decapentaplegic (dpp), are expressed in the dorsal portion of the embryo and are required for the establishment of the amnioserosa and the dorsal epidermis. They are repressed by Dorsal in the ventrolateral and ventral cells. The CRMs of these genes contain not only high affinity Dorsal binding sites but also AT-rich sites which are bound by the Cut and Dead-ringer proteins (Chen et al., 1998). These proteins cooperatively recruit the Groucho co-repressor, medi- ating transcriptional repression (Valentine et al., 1998). Another factor, Capicua, was also shown to be involved in the repression of zen (Jimenez et al., 2000).

Snail in gastrulation and mesoderm formation Gastrulation is the first morphogenetic process in embryonic development, during which cells of the blastula undergo rearrangement and move to the correct positions to set up the three germ layers (ectoderm, mesoderm and endoderm). This requires the coordination of cell-fate determination, cell- cycle control, individual cell-shape changes, and movement of groups of cells (Ip and Gridley, 2002; Leptin, 1999; Leptin et al., 1992). During cellu- larization, the expression of Snail is sharply restricted to the 18 cell-width ventral domain mainly due to the cooperative function between Dorsal and Twist. These Snail-expressing cells correspond precisely to the presumptive mesoderm. Either expanding or reducing the expression of Snail correlates with changes in ventral cell invagination and the formation of future meso- derm (Leptin et al., 1992). Genetic rescue by Snail in a twist mutant can promote ventral-cell invagination but not vice versa, indicating Snail has a more direct role in regulating downstream events leading to gastrulation (Ip et al., 1994). It was proposed that Snail can regulate two separate sets of target genes, based on the fact that some snail hypomorphic mutants show

16

phenotypes on derepression of neuroectodermal genes and mesoderm differ- entiation, but not on ventral cell invagination. Snail represses neuroectoder- mal genes such as single-minded (sim), rho, brinker (brk), lethal of scute and so on by direct binding to their enhancers. And it also promotes the expres- sion of mesodermal genes such as dGATAb and zfh-1 to help the establish- ment of the mesodermal cell fate. The second set of genes (e.g. folded gas- trulation) control the cell shape changes and cell movements during gas- trulation (reviewed in Ip and Gridley, 2002). However, deregulation of indi- vidual Snail target genes does not interfere with mesoderm cell fate or invagination, suggesting that the cumulative effect of regulating all these genes is essential for gastrulation (Hemavathy et al., 1997). The molecular function of the Snail repressor depends on two functional domains, a zinc-finger DNA-binding domain in the C-terminus and a repres- sion domain in the N-terminus (Boulay et al., 1987; Hemavathy et al., 1997). The repression domain contains two conserved motifs that allow interaction with the co-repressor dCtBP (Drosophila homolog of C-terminal binding protein) (Nibu et al., 1998a). Our studies demonstrated that the full repres- sion activity of Snail requires another co-repressor, Ebi, which interacts with Snail through a conserved motif in the N-terminus of Snail protein (Paper III).

Anterior-posterior patterning The Drosophila larva has a very obvious feature, the regular segmentaion of the larval cuticle along the AP axis, each segment carrying cuticular struc- tures that define it as, for example, thorax or abdomen (Figure 2B). Similar to DV patterning, the determination of AP axis is initiated by maternal fac- tors and then carried out by a segmentation hierarchy of zygotic transcription factors (reviewed in Niessing et al., 1997). The whole procedure requires three classes of maternal determinants in the egg that control anterior pat- terning, posterior patterning and terminal patterning of the embryo respec- tively (reviewed in St Johnston and Nusslein-Volhard, 1992). The hierarchy produces broad bands of gap gene expression, and the striped patterns of pair-rule genes that subsequently control the segmental organization deter- mined by segment polarity genes (Figure 3). By this time, the cell fates can be determined at single-cell resolution along the AP axis. Bicoid is the maternal determinant that patterns the head and the thorax of the embryo. Bicoid protein forms gradient from anterior to posterior. An- other maternal factor, Caudal, patterns the posterior part of the embryo. A Caudal protein gradient is established from posterior to anterior. Bicoid and Caudal initiate the activation of zygotic gap genes including orthodenticle, hunchback (hb), Kruppel (Kr), knirps (kni) and giant, all of which code for transcription factors (Ip et al., 1992; Lebrecht et al., 2005; Ochoa-Espinosa et al., 2005; Rivera-Pomar et al., 1995). A steep gradient of Hb protein is

17

established from anterior to posterior. The Hb gradient, together with Bicoid, activates or represses other gap genes. Interactions between gap genes also help to sharpen and establish their border of expression. For example, the anterior border of kni is set by Hb, and its posterior border is limited by Gi- ant and Tailless (Tll) (Rivera-Pomar 1995). The anterior boundary of Kr expression is repressed by Hb, but its posterior boundary repressed by Kni, Giant and Tll (Hoch et al., 1992). The expression of tll is not repressed by other gap genes. Instead, it is a downstream gene in response to Torso signaling.

Bcd Cad Maternal effects

Gap genes gt, Kr, kni, tll etc

Pair-rule genes eve, ftz, odd, hairy, etc

Segment polarity genes en, wg, hh, etc

Figure 3. Segmentation is achieved by a hierarchy of maternal and zygotic tran- scription factors. Bicoid (Bcd) and Caudal (Cad) form morphogen gradients with highest level at anterior and posterior respectively. They activate the first set of zygotic genes, the gap genes, which are expressed in rather broad regions. Gap gene products define the periodical expression patterns of pair-rule genes, whose products subsequently determine the expression of segment polarity genes. Pair-rule genes are expressed in seven stripes, wheras segment polarity genes are expressed in fourteen stripes.

The terminal regions are specified by a third group of maternal genes in which a key factor is called Torso. Torso is a receptor tyrosine kinase uni- formly distributed throughout the fertilized egg plasma membrane. But it is only activated at the two ends of the embryo through binding to its ligand. This signal induces the activation of zygotic genes, such as tll and huckebein, at both poles, thus defining the two extremities of the embryo (reviewed in Niessing et al., 1997). Different concentrations and combinations of the gap genes set up the lo- calized striped pattern of pair-rule genes (Jackle et al., 1992; Niessing et al.,

18

1997), including even-skipped (eve) and hairy. Pair-rule genes are expressed in seven stripes, with a periodicity corresponding to alternate parasegments. For example, eve defines odd-numbered parasegments, whereas ftz defines even-numbered parasgments. The striped expression of primary pair-rule genes requires the regulation by separate CRMs, working independently in the regulatory region (Small et al., 1992). Each CRM contains multiple bind- ing sites for both activators and repressors. A famous example is the regula- tion of eve whose regulatory region contains 5 separate CRMs, together pro- ducing the seven stripes. This enhancer autonomy is partially due to short- range repression, meaning that a repressor bound to one enhancer does not interfere with another (more discussion later). Pair-rule genes also regulate the expression of each other, thereby establishing a refinement of the stripe pattern (reviewed in Niessing et al., 1997). The segment-polarity genes are activated by products of pair-rule genes and are involved in stablizing the parasegment boundary and setting up a signaling center at the boundary that patterns the segments (Nasiadka et al., 2000; Nasiadka and Krause, 1999). Well-studied segment polarity genes include (en), wingless and hedgehog. En is a transcription factor, whereas Wingless and Hedgehog are signaling molecules. The segment identity is specified by homeotic selector genes (Homeotic genes). The two Homeotic gene clusters in Drosophila are called the Bitho- rax and Antennapedia gene complexes (Regulski et al., 1985; Wedeen et al., 1986). The Bithorax complex is responsible for diversification of the poste- rior segments 5-12 and the Antennapedia complex for that of the anterior segments 1-5. These genes are activated by gap genes and pair-rule genes at an early embryonic stage but their expression pattern are maintained by Polycomb Group genes (PcG) and Trithorax Group (trxG) genes later in development (Breen and Harte, 1991; Harding and Levine, 1988; LaJeunesse and Shearn, 1995; Simon et al., 1992).

19

Histone modifications

Histone modifications play important roles in different biological processes including transcription regulation, DNA repair, replication and heterochro- matin formation (reviewed in Kouzarides, 2007). Their functions depend on the type of chemical modifications, the location in the histone octamer and even the combinatorial effects of multiple modifications. The classic modifi- cations include lysine acetylation, lysine and arginine methylation, serine phosphorylation and lysine ubiquitylation. Recent studies have discovered more modifications that include lysine sumoylation, ADP ribosylation, ar- ginine deimination and proline isomeration (reviewed in Kouzarides, 2007). The possible sites for modification in histones are numerous (Figure 4), and the forms of modification are also various such as mono-, di-, or trimethyl for lysines and mono- or di- (asymmetric or symmetric) for arginines. Most modification sites were identified in the flexible N-terminal tails of histones, such as histone H3 Lysine 4 (H3K4), H3K9, H3K14, H4K5, H4K8, and so on (Figure 4). Interestingly, recent studies suggested that residues within core histones can also be modified (e.g., H3K56 and H3K79) (Xu 2005). Furthermore, most of these modifications, such as acetylation, phosphoryla- tion and methylation, are all reversible. This fact of complexity gives histone modifications enormous potential for functional responses.

P Ac Ac Ac Ac Ac U H2A S K K K K K K 1 5 9 13 15 36 119 M Ac Ac P Ac Ac M Ac U H2B K K S K K K K K 5 12 14 15 20 23 24 120 M M M P M AcAc P Ac Ac M Ac Ac M M P Is M M Is Ac H3 R T K R K S T K R K K R K S P K K P K 2 3 4 8 9 10 11 14 17 18 23 26 2728 30 36 37 38 56 M M P M Ac Ac Ac Ac Ac H4 S R K K K K K 1 3 5 8 12 16 20 Figure 4. Most of histone modifications, including lysine methylation, arginine me- thylation, lysine acetylation, serine phosphorylation, lysine ubiquitylation and proline isomeration on histone H2A, H2B, H3 and H4, are depicted.

20

Functions of histone modifications There are two mechanisms by which histone modifications contribute to chromatin function. First, some modifications (e.g. acetylation) reduce the positive charge of histones, in principle resulting in the alteration of the chromatin structure directly. Second, the modifications can generate binding surfaces for non-histone proteins. Specifically, methylated lysines can be recognized and bound by chromo-like domains (chromo, tudor, MBT) and PHD domains, acetylated lysines can be recognized by bromodomains, and phosphorylation is recognized by a domain within 14-3-3 proteins (Figure 5). These nonhistone proteins often possess enzymatic activities which can fur- ther modify chromatin. For example, many histone acetyltransferases (HATs) contain a bromodomain, facilitating the maintenance of the acetylated chro- matin by further modifying regions that are already acetylated. Similarly, histone methylases (HMTs) often contain a chromodomain, which is impor- tant for the maintenance of methylated nucleosomes. Other proteins contain- ing these domains may not be histone modifying enzymes but instead deliver enzymes to histones. For example, the Polycomb protein binds to H3K27me through its chromodomain and is associated with ubiquitin ligase (Ring1A) activity specific for H2A (de Napoles et al., 2004; Fang et al., 2004). Het- erochromatin protein1 (HP1) binds H3K9me and is associated with Su(var)3-9, a HMT for H3K9. It was also suggested that histone modifica- tions can prevent the targeting of nonhistone protein onto chromatin (Margueron et al., 2005). Moreover, one modification, can affect another modification, positively or negatively. The phosphorylation of H3S10 dis- rupts the binding of HP1 to methylated H3K9 (Fischle et al., 2005) but fa- cilitates the binding of GCN5 acetyltransferase to acetylated H3K9 (Clements et al., 2003). The effect of histone modifications seems to be con- text dependent. Methylation of H3K36 has a positive effect when it is pre- sent in the coding region and a negative effect when in the promoter (Landry et al., 2003; Li et al., 2007a; Strahl et al., 2002).

HP1 Eaf3 Pc Rsc4 14-3-3 CHD1

Bdf1 Brd2 Taf1 Bromo Chromo Chromo JMJD2A Chromo Chromo CRB2 Ac P Me Me Me T H3 PH B Bromo Me D M H4 Bromo Bromo Tudor K K S K K K Ac 27 14 10 9 L(3)MBTL Ac Ac Me 36 BPTF 4 TAF3 K K K ING2 8 12 16 K 20 Figure 5. The known modification sites for protein domains in non-histone proteins. Proteins containing the bromodomain bind to acetylated lysine. Proteins containing a chromodomain, Tudor domain, PHD domain or MBT domain bind to lysine me- thylation. 14-3-3 protein binds to phosphorylated serine.

21

The number, variety and interdependence of histone modifications led to the histone code hypothesis, multiple histone modifications, acting in a combi- natorial or sequential fashion on one or multiple histone tails, specify unique downstream functions (Strahl and Allis, 2000). It has provided a framework for the interpretation of many discoveries regarding the mechanisms by which chromatin functions. However, the fact that the function of histone modifications is highly context dependent suggests the code cannot be uni- versal.

Histone acetylation and its function Histone acetylation is one of the best-characterized histone modifications, an event with a big impact on chromatin structure and function. It is possibly involved in all chromatin-based processes, such as gene transcription, gene silencing, DNA repair, DNA replication and chromosome condensation. Acetylation can occur on all core histones (Figure 4) and is carried out by HATs, which catalyze the transfer of an acetyl group from acetyl-CoA to the lysine -amino groups on the N-terminal tails of histones. This process is reversible. Histone deacetylases (HDACs) remove the acetyl group from acetylated lysines.

Histone acetylation in gene regulation The correlation between histone acetylation and transcription activation has been established for a long time based on several facts. First, acetylation has the most potential to unfold chromatin by neutralizing the basic charge of the lysine, which makes DNA more accessible for targeting. Second, many HATs were identified as co-activators, such as GCN5 and CBP/P300 (Bantignies et al., 2000; Brownell et al., 1996). Third, the actively tran- scribed regions tend to be hyperacetylated, whereas transcriptionally silent regions tend to be hypoacetylated. The most-characterized sites are within the N-terminal tail of the histones, such as H3K9, H3K14, H4K5, H4K8, H4K12, and H4K16 (Figure 4). Some other sites were recently identified, such as H3K18, whose pattern tightly correlates with gene activation in pros- tate cancer tissue (Seligson et al., 2005). Histone deacetylation makes chromatin more compacted and generally correlates with transcription repression. Consistently, many HDACs are part of corepressor complexes such as Rpd3 (reduced potassium dependency 3) in the Sin3-Rpd3 complex and HDAC3 in the NCoR/SMRT complex. How- ever, recent evidence, that hypoacetylation of H4K16 and H2BK11/16 is linked to transcription activation, challenges this over-simplified view (Kurdistani et al., 2004; Wiren et al., 2005).

22

Histone acetylation/deacetylation also regulates gene expression through gene silencing which involves heterochromatin establishment and DNA me- thylation. Histone deacetylation could be involved in both cases. In budding yeast, HDACs are recruited to unacetylated histones in silenced regions, thus reinforcing and maintaining the deacetylated state in heterochromatin (Braunstein et al., 1996). How heterochromatin is established is not fully understood. However, it has been shown that the involvement of H3K9 me- thylation and heterochromatin protein 1 (HP1) is important. Additionally, it may involve the changes of several other histone marks (reviewed in Ebert et al., 2006), such as H3K4 demethylation (Rudolph et al., 2007), H3S10 dephosphorylation (Zhang et al., 2006c), H3K9 deacetylation, H4K12 acety- lation (Swaminathan et al., 2005). In mammalian cells, HDACs can be re- cruited to methyl groups attached to the cytosine residues within CpG is- lands via proteins such as methyl-CpG binding proteins and methyl-CpG- binding-domain-containing proteins, or via the DNA methylases, resulting in tightly-packed chromatin and gene silencing.

Histone acetylation in DNA repair DNA double strand breaks (DSBs) cause genetic instabilities and gene muta- tions if they are not accurately repaired. The response of eukaryotic cells to DSBs is highly conserved and involves both checkpoint functions that arrest the cell cycle, allowing time for repair to occur, and repair functions that directly fix the break (reviewed in Khanna and Jackson, 2001). Defects in DNA repair result in the accumulation of DNA damage and consequently lead to apoptosis or enhanced carcinogenesis. DSBs are repaired through two major pathways, homologous recombination (HR) and non-homologous end joining (NHEJ) (reviewed in Bernstein et al., 2002; Khanna and Jackson, 2001). The DNA damage response pathway includes a series of events: de- tection by sensor proteins, delivery by transducer proteins and downstream actions by effecter proteins (Figure 6A). Ataxia-telangiectasia mutated (ATM) and ATM and Rad-3 related (ATR) kinases are transducer proteins and belong to the phosphoinositide 3-kinase related kinase family (PIKK). Upon DNA damage, ATM and ATR get auto-phosphorylated and activate the second group of kinases, Chk2 and Chk1, which subsequently phos- phorylate effecter proteins. ATM is actively involved in DNA repair, check- point response and apoptosis through modifying some key factors in these processes (e.g. BRCA1, Chk2, ) (Banin et al., 1998; Cortez et al., 1999; Matsuoka et al., 1998). Chk2 can also phosphorylate the signal effector, p53, to induce apoptosis and G1/S checkpoint (Hirao et al., 2000). p53 is a tumor suppressor gene which is mutated in more than half of all human cancers (Greenblatt et al., 1994). The phosphorylation stabilizes and stimulates p53 activity. Once activated, it functions as a transcription factor and regulates

23

expression of genes which are involved in apoptosis and in DNA repair (reviewed in Oren, 2003). A p53 homolog is absent in yeast. The Drosophila homolog of p53 is only involved in apoptosis but not in cycle arrest upon DNA damage (Figure 6B) (Brodsky et al., 2004; Lee et al., 2003). Another early response to DSBs is the accumulation and the spreading of phosphory- lated histone variant H2AX (fly H2Av) and yeast H2A around the break site. The phosphorylation is possibly mediated by ATM as well (Rogakou et al., 1998; van Attikum and Gasser, 2005).

IR B A DSB DSB P H2Av

Sensors Rad17, Rad9, Rad1, Hus1, Mei-41(ATR) P MRN complex P dATM

Chk2 P Mus304 P Transducers ATM, ATR, DNA-PK

P p53 Grp(Chk1) P

CDC25 Effectors P53 reaper, grim, hid, skl

Cell cycle arrest DNA repair Apoptosis Apoptosis G2 arrest Figure 6. A. The general organization of the DNA-damage response pathway. It is composed of sensor proteins, transducer proteins and effector proteins. The effec- tors subsequently activate signaling pathways for DNA repair, cell cycle arrest and apoptosis. B. IR induced G2 arrest and apoptosis in Drosophila. Cell cycle arrest and apoptosis pathways are rather independent. IR induced apoptosis is initiated by dATM. dATM gets auto-phosphorylated, which subsequently activates Chk2. dATM and Chk2 activate p53 by phosphorylation. p53 activates its target genes, such as reaper, grim, hid (head involution defect) and skl (sickle), whose activities are responsible for apoptosis. In parallel, G2 arrest is prominently implemented by Mei-41 (dATR), Mus 304 (dATRIP) and Grp (Chk1). dATM and Chk2 may also play roles in the initiation of G2 arrest.

Chromatin dynamics need to be adjusted to facilitate the recruitment of the repair machinery to the break sites. Chromatin remodeling and histone modi- fications can make the lesion more accessible for damage response machin- ery (reviewed in Peterson and Cote, 2004; van Attikum and Gasser, 2005; Wurtele and Verreault, 2006). It was suggested that chromatin remodeling complexes including the RSC (remodels the structure of chromatin) complex, the SWI/SNF complex and the INO80 complex play roles at multiple steps during the detection and repair of DSBs (reviewed in Downs et al., 2007). In addition to the phosphorylation of H2AX (H2Av in fly, H2A in yeast), sev- eral other histone modifications are involved in DNA repair (Peterson and

24

Cote, 2004; van Attikum and Gasser, 2005; Wurtele and Verreault, 2006). The acetylation of conserved lysine residues in the N-terminal tails of H3 and H4 are important for DNA-damage responses (Bird et al., 2002; Downs et al., 2007; Tamburini and Tyler, 2005). In yeast, histone H3 acetylation by the Hat1(Qin and Parthun, 2006) complex and Gcn5 (Teng et al., 2002) and H4 by the NuA4 (nucleosome acetyltransferase of H4) complex is impotant for DNA repair (Bird et al., 2002). In mammalian cells, the Tip60 HAT complex is recruited to sites of DSBs, where they induce H4 acetylation and HR (Murr et al., 2006). The Drosophila Tip60 complex specifically acety- lates phosphorylated H2Av, which is a prerequisite for the exchange with unmodified H2A during DNA repair (Kusch et al., 2004). Not only was the initial recruitment of HATs observed on DSBs, but also the subsequent asso- ciation of several HDACs with chromatin near DSBs (Tamburini and Tyler, 2005). Possibly HDACs are required for the restoration of chromatin higher- order structure following DSB repair.

Histone acetyl-transferase complexes The histone acetyltransferases are divided into five families (Figure 7), in- cluding the Gcn5-related N-acetyltransferases (GNATs); the MYST (for MOZ, Ybf2/Sas3, Sas2 and Tip60)-related HATs; p300/CBP HATs; the general transcription factor HATs, and the nuclear hormone-related HATs SRC1 and ACTR (SRC3) (reviewed in Roth et al., 2001).

HAT HDAC

Five families: GCN5 Three types: MYST Type I: HDAC1, 2, 3, 8 CBP Type II: HDAC4, 5, 7, 9 SRC-1 HDAC6, 10 TAF1 Type III: Sir2 family

Figure 7. Histone acetyl transferases and histone deacetylases. Histone acetyl trans- ferases are classified into five families. Histone deacetylases are classified into three families. Histone acetylation generally correlates with more opened chromatin which facilitates transcription. By contrast, deacetylation state of histone results in more closed chromatin which inhibits transcription.

25

GNAT complexes All the GNAT complexes contain HATs that show sequence and structural similarity to Gcn5 (reviewed in Roth et al., 2001). They commonly share two functional conserved domains: a HAT domain that encompasses three regions of conserved amino acid sequence spanning approximately eighty residues, and a C-terminal bromodomain that is believed to interact with acetylated lysines. The metazoan members of the family share an additional domain which is not found in yeast, the N-terminal PCAF region (reviewed in Carrozza et al., 2003). Gcn5 relatives have been identified in virtually all eukaryotes. Besides Gcn5 homologs, this family also includes the chromatin- assembly-related Hat1, the elongator complex subunit Elp3, the mediator complex subunits Nut1 and Hpa2. Recombinant Gcn5 mainly acetylates free histone H3, but exhibits little HAT activity when using nucleosomal histones as substrates, suggesting that incorporation into complexes is required for Gcn5 to obtain the capacity to acetylate nucleosomes. The most characterized GNAT complexes are yeast SAGA (for Spt-Ada-Gcn5-acetyltransferase) and the ADA complex. Gcn5 and the accessory proteins Ada2 and Ada3 form the catalytic core of these complexes and enable them to acetylate nucleosomes. Previous data also indicated that incorporation of Gcn5 into complexes alters lysine specificity. Gcn5 alone acetylates mainly H3K14 on free histones, but SAGA acetylates lysines 9, 14, 18 and 23, and ADA acetylates 9, 14 and 18 (Grant et al., 1999). GNAT complexes were identified from different organisms and they have similar composition and functions, but the complexes from higher metazoans are more diverse than those in yeast. One Drosophila SAGA-like complex has been identified. Three known mammalian SAGA-like complexes are called PCAF, STAGA and TFTC (Ogryzko, 2000). The SAGA-like com- plexes basically contain: (i) Proteins for HAT activity (Ada2, Ada3, and Ada4/Gcn5) and complex assembly (Ada1 and Ada5/Spt20), which were isolated in a genetic screen as proteins interacting with the activators yeast and the herpes simplex virus activation domain VP16; (ii) Proteins for TBP binding, the Spt proteins, (Spt3, Spt7, and Spt8), which were initially identified as suppressors of transcription initiation defects caused by pro- moter insertions of the Ty transposable element. (iii) A subset of TBP- associated factors (TAFs) (TAF5, TAF6, TAF9, TAF10 and TAF12). (iv) Activator interaction protein Tra1, an ATM/PI-3 kinase-related homologue of the human TRRAP protein and (v) proteins for deubiquitylation (Ubp8 and Sgf11), mediating the deubiquitylation of H2B (Shukla et al., 2006). Several other SAGA-associated proteins were identified and characterized recently. Sgf73p facilitates formation of the preinitiation complex assembly at promoters (Shukla et al., 2006). The mRNA export factor Sus1 is involved in SAGA-mediated H2B deubiquitinylation through its interaction with

26

Ubp8 and Sgf11 (Kohler et al., 2006). Both normal and polyglutamine- ex- panded ataxin-7 are components of TFTC-type GCN5 histone acetyltrans- ferase- containing complexes (Helmlinger et al., 2006). SAGA-related com- plexes can function as co-activators through the interactions with TBP and acidic activators, such as GCN4, Pho4 and Gal4 (Barbaric et al., 2003; Kuo et al., 2000; Larschan and Winston, 2001). Recent studies by Govind et al. suggested that SAGA also occupies coding sequences and contributes to H3 acetylation, H3K4 methylation and even nucleosome eviction (Govind et al., 2007). Intriguingly, the SAGA complex has been implicated in transcrip- tional repression by the Arg/Mcm1 repressor complex (Ricci et al., 2002). Mammalian TFTC complex exhibits increased binding to UV-damaged DNA in vitro, implicating the function of GNAT complexes in DNA repair (Brand et al., 2001). GCN5 related proteins play critical roles during development. GCN5-null mice are embryonic lethal with the lack of somite formation. The deletion of PCAF, however, does not cause lethality, presumably due to compensation by GCN5. Animals carrying deletions in both genes die earlier than GCN5- null animals, indicating that PCAF does have distinct functions from GCN5 in early development (Xu et al., 2000; Yamauchi et al., 2000). dGCN5 null mutants in Drosophila have defects in both oogenesis and metamorphosis, while hypomorphic alleles of dGCN5 affect the formation of adult append- ages and cuticle (Carre et al., 2005). Mutations in Arabidopsis GCN5 also result in developmental defects (Vlachonasios et al., 2003). GCN5 possesses both HAT-dependent and independent functions, as indicated by the fact that GCN5 null mice die much earlier with increased apoptosis than GCN5 HAT mutant mice with defects in neural tube closure but with no abnormal apop- tosis (Bu et al., 2007). GCN5 also targets non-histone proteins as substrates which are involved in diverse cellular processes. For example, PCAF acety- lates high mobility group protein (HMG)-A1 that positively regulates tran- scription of the interferon gene (Munshi et al., 2001), the oncoprotein c- , and Ku70 that is involved in DNA repair through the non-homologous end-joining (NHEJ) pathway (reviewed in Glozak et al., 2005).

Ada2 proteins The Ada2 protein was first identified from yeast, in which its inactivation could relieve the toxicity caused by overexpression of the strong transcrip- tional activation domain in the viral VP16 protein (Berger et al., 1992). Sub- sequently, Ada2 proteins were shown to exist in all known Gcn5 containing complexes and to be required for the assembly of these complexes. In con- trast with the single Ada2 gene present in Saccharomyces cerevisiae, the Arabidopsis thaliana, fly and human genome all contain two genes, encod- ing two Ada2 proteins, Ada2a and Ada2b (Barlev et al., 2003; Kusch et al., 2003; Muratoglu et al., 2003; Stockinger et al., 2001). However, Ada2a and Ada2b are present in different complexes which possess different functional

27

activities. For example, in Drosophila, the SAGA-like complex contains dAda2b and acetylates histone H3K9 and H3K14 (Pankotai et al., 2005; Qi et al., 2004). However, the ATAC (Ada Two A Containing) complex can acetylate histone H4K5 and H4K12 (Guelman et al., 2006; Pankotai et al., 2005). Ada2 homologues share several conserved regions including a ZZ domain, a SANT domain and three ADA boxes (Kusch et al., 2003). The ZZ domain is a potential zinc-binding motif and is required for the interaction with Gcn5 (Candau and Berger, 1996). The SANT (SWI3, Ada2, N-CoR and TFIIIB) domain of Ada2 is able to associate with chromatin and is essential for the catalytic activity of Gcn5 (Boyer et al., 2002).

Tip60 HAT complexes Tip60 (Tat-interactive protein 60kDa) is one of the best characterised MYST proteins whose family members function in a broad range of biological proc- esses, such as gene regulation, dosage compensation, DNA damage repair and tumourigenesis (reviewed in Utley and Cote, 2003). Tip60 homologues have been identified in various organisms from yeast to human. They share at least two functional domains, an N-terminal chromodomain and a C- terminal MYST catalytic domain. Recombinant Tip60 acetylates core his- tones H2A, H3 and H4 in vitro (Kimura and Horikoshi, 1998; Yamamoto and Horikoshi, 1997). When assembled in multiprotein complex, Tip60 can modify nucleosomal H2A and H4 (Ikura et al., 2000). Tip60, like GCN5 and CBP, can also acetylate non-histone proteins, including , upstream binding transcription factor, c-Myc, forkhead family protein FOXP3 and so on (Li et al., 2007a; Sapountzi et al., 2006).

Tip60Tip60 E(Pc)E(Pc) ING3ING3 TRRAPTRRAP DominoDomino DMAP1Reptin Pontin BAF53BAF53DMAP1Reptin Pontin Eaf6 AActinctin Eaf6 Brd8Brd8 GAS41GAS41 MRG15MRG15 Figure 8. The composition of Tip60 complex. Tip60, E(Pc) and ING3 constitutes a core enzymatic sub-complex.

The composition of the Tip60 complex is conserved between Drosophila and human. It was suggested that the metazoan Tip60 complex is a hybrid of the yeast NuA4 and SWR1 complexes (Doyon et al., 2004). Three subunits, Tip60 (Esa1), EPC1 (yeast homolog EPL1, Drosophila homolog E(Pc) (En- hancer of Polycomb)) and ING3 (yeast Yng2, Drosophila dIng3) form a core

28

enzymatic subcomplex and exist as a distinct entity responsible for the global non-targeted acetylation of chromatin in vivo (Boudreault et al., 2003). Drosophila E(Pc) is a suppressor of position-effect variegation. Dom- ino/p400 (Swr1), an ATPase possessing chromatin remodeling activity (Mizuguchi et al., 2004), is able to exchange phosphorylated H2Av for un- modified H2Av after DNA damage (Kusch et al., 2004). Mortality factor 4 related gene 15 (MRG15) (Eaf3) contains a chromodomain and can recruit HDAC1 to chromatin in the Rpd3S complex (Carrozza et al., 2005). The TRRAP (transformation/transcription domain-associated protein, yeast Tra1) protein is a common subunit in SAGA-related complexes and interacts with transcription factors. Pontin (Tip49a, RuvBL1) and Reptin (Tip49b, RuvBL2) are putative helicases, related to the bacterial repair protein RuvB (reviewed in Gallant, 2007). Previous studies have indicated two major aspects of roles of the Tip60 complexes, in transcription and in DNA DSBs response. In a lot of cases, Tip60 was associated with transcription activation. Examples of Tip60- coactivated transcription factors are androgen receptor (AR) (Gaughan et al., 2002), NF- B (Baek et al., 2002; Kim et al., 2005), c-myc (Patel et al., 2004), (Taubert et al., 2004), p53 (Doyon et al., 2004) and so on. However, Tip60 was also shown to have transcriptional repression activity. For exam- ple, Tip60 corepresses CREB (cAMP response element binding protein) by direct binding to CREB and repressing CREB phosphorylation by protein kinase A (Gavaravarapu and Kamine, 2000). Tip60 is believed to recruit HDAC7 to STAT3 (signal transducer and activator of transcription 3) regu- lated promoters and thus represses transcription (Xiao et al., 2003). Tip60 promotes the acetylation of FOXP3 protein and interacts with HDAC7, which are both required for the repressor function (Li et al., 2007b). The role of Tip60 complexes in DNA repair has also been well studied. In yeast cells, after DNA damage NuA4 complex is recruited to the vicinity of DNA le- sions, which promotes the recruitment of the chromatin remodeling com- plexes INO80 and SWR1 (Downs et al., 2004). In Drosophila, Tip60 acety- lates phosphorylated H2Av and Domino exchanges the acetylated and phos- phorylated H2Av with unmodified H2Av, which is required for DSB repair (Kusch et al., 2004). In mammalian cells, the Tip60 complex is involved in the DNA damage-induced H4 hyperacetylation and homologous recombina- tion (HR) (Ikura et al., 2000; Murr et al., 2006). In a word, this subfamily of MYST complexes play multiple roles in a diverse variety of cellular path- ways, involving not only normal cellular conditions but also abnormal proc- esses including viral infection, neurodegernerative disease and cancer (reviewed in Sapountzi et al., 2006).

Reptin Reptin and the related Pontin protein are members of the AAA+ family of helicases (ATPases associated with diverse cellular activities) (reviewed in

29

Ogura and Wilkinson, 2001). Reptin and Pontin can form a dodecamer con- sisting of two differently shaped, juxtaposed hexamers (Puri et al., 2007). Their ATPase activities are interdependent as either protein alone is almost inactive (Makino et al., 1999). Reptin and Pontin were found to be integral subunits of different chromatin-modifying complexes, including the Tip60 complex, the INO80 complex and the Swr1 (SRCAP in metazoan) complex (Kobor et al., 2004; Krogan et al., 2003; Mizuguchi et al., 2004; Shen et al., 2000). Drosophila Reptin was co-purified with Polycomb repressive com- plex 1 (PRC1) which maintains the repressive state of HOX genes during development (Saurin et al., 2001). Integration in these complexes indicates potential roles of Reptin and/or Pontin in diverse cellular processes, includ- ing the response to DNA double-strand breaks and transcription regulation (reviewed in Gallant, 2007). Reptin and Pontin were also identified by their physical interaction with the transcription-associated protein -catenin (Bauer et al., 2000), with the transcription factors TBP (Bauer et al., 1998; Ohdate et al., 2003), Myc (Wood et al., 2000), E2F1 (Dugan et al., 2002) and ATF2 (Activating transcription factor 2) (only Reptin) (Cho et al., 2001). Along with components of the Tip60 complex, Reptin and/or Pontin were shown to bind to the promoters of transcriptional targets of c-myc, E2F1 and NF- B (Frank et al., 2003; Kim et al., 2005; Taubert et al., 2004). Interest- ingly, Reptin and Pontin may play similar or antagonistic roles in a context- dependent manner. For example, both Reptin and Pontin enhance the repres- sion function of c-myc in the control of cell proliferation in Drosophila and Xenopus (Bellosta et al., 2005; Etard et al., 2005). However, in the cases of -catenin and NF- B, they have opposite activities. Drosophila and zebraf- ish Reptin act as transcriptional repressors , whereas Pontin are activators in -Catenin/Wnt signaling (Bauer et al., 2000; Rottbauer et al., 2002). Similar activities were observed in a human cell line in the regulation of a NF- B p50 target, KAI1(Kim et al., 2005). KAI1 is a tetraspanin protein inhibiting tumor metastasis. At the regulatory region of KAI1, p50 recruits the Tip60/Pontin coactivator for transcriptional activation, whereas it recruits the -catenin/Reptin corepressor for repression. In metastatic cells, KAI1 is downregulated due to increased -catenin and decreased Tip60. Reptin/ - catenin mediated repression partially depends on HDAC1. Reptin is sumoy- lated on lysine 456. It was suggested that sumoylation of Reptin stimulates the repressive potential of Reptin by promoting the nuclear localization of Reptin and by increasing its interaction with HDAC1 (Kim et al., 2006). These studies indicate an additional function of Reptin independent of the Tip60 complex.

30

Histone deacetylases and their associated complexes The histone deacetylases (HDACs) are conserved proteins in different organ- isms. They are classified into three subfamilies (Figure 7): class I, class II and Sir2 related HDACs, based on their protein sequence similarity to yeast Rpd3, HDA1 and Sir2 respectively. Class I and class II proteins are evolu- tionarily related and share a common enzymatic mechanism, the zinc- catalyzed hydrolysis of the acetyl-lysine amide bond. Class III proteins are functionally different and catalyze the transfer of the acetyl group onto the sugar moiety of NAD (reviewed in Blander and Guarente, 2004). HDACs have diverse functions through multiple mechanisms: the incorporation into different complexes, the modulation of deacetylase activity by protein- protein interactions or by post-translational modifications, tissue specific pattern, variable subcellular localizations and splice variants (reviewed in de Ruijter et al., 2003). The substrates of HDACs are not only histones, but also non-histone proteins, including p53, , -tubulin, MyoD and (Gregoire et al., 2007; Hubbert et al., 2002; Juan et al., 2000). Class I HDACs include HDAC1 (Rpd3), 2, 3 and 8. HDAC1 and HDAC2 are closely related and present in common co-repressor complexes, such as Sin3, NuRD (nucleosome remodeling and deacetylating) and Co-REST complexes (Kasten et al., 1997; You et al., 2001; Zhang et al., 1999). These complexes are widely involved in transcription regulation and developmen- tal control. HDAC1 and HDAC2 are also regulated by phosphorylation, which promotes both deacetylase activity and complex formation (Galasinski et al., 2002). HDAC3, however, is stably associated with NCoR/SMRT ( corepressor/ silencing mediator for retinoic acid and thyroid hormone receptors) corepressor complexes. HDAC8 was recently identified and its function remains to be discovered (Buggy et al., 2000). Class II HDACs include HDAC4, 5, 6, 7, 9 and 10. The subgroup con- taining HDAC4, 5, 7 and 9 is able to shuttle in and out of the nucleus in re- sponse to certain cellular signals (reviewed in Verdin et al., 2003). Upon phosphorylation by several kinases, they are confined to the cytosol by inter- action with 14-3-3 proteins. In the nucleus they associate with transcription factors, for example the MEF and Runx families (Vega et al., 2004; Zhang et al., 2002a). HDAC6 is able to deacetylate Tubulin and Hsp90, suggesting its special role in cell motility and protein folding (Hubbert et al., 2002; Kawa- guchi et al., 2003; Matsuyama et al., 2002). HDAC10 is the most recently discovered member of the class II HDACs. It was suggested that it associates with many other HDACs so it might function as a recruiter rather than as a deacetylase although it does show deacetylase activity in vitro (de Ruijter et al., 2003). HDAC11 is more closely related to the class I HDACs, but no clear classification has been made due to the low similarity of the overall sequence (Gao et al., 2002).

31

Class III HDACs are homologous to yeast Sir2 (Silent information regu- lator) protein that is known to be required for silencing of the mating type loci and of telomeres. Recent evidence suggests that Sir2 and its related pro- teins are involved in various biological pathways including longevity, apop- tosis, and skeletal myogenesis (Fulco et al., 2003; Guarente, 2001; Luo et al., 2001).

NCoR/SMRT/HDAC3 complex NCoR and the related protein SMRT (SMRTER in Drosophila) are involved in repression by unliganded thyroid hormone and retinoic acid receptors (TR and RAR) (Chen and Evans, 1995; Horlein et al., 1995). In addition, they are recruited by several unrelated transcription repressors such as Notch binding protein CBF-1 (Kao et al., 1998) and homeodomain proteins including Rpx2, Pit, and Pbx (Asahara et al., 1999; Xu et al., 1998) in mammals. NCoR/SMRT complexes additionally include GPS2 (G protein pathway suppressor 2), HDAC3, TBL1 (Transducin like-1) and TBLR1 (TBL re- lated-1) proteins (Guenther et al., 2000; Li et al., 2000; Zhang et al., 2002a). NCoR and SMRT are highly related but distinct proteins, sharing similar domain structure and function. They both contain a conserved deacetylase- activating domain (DAD) that is essential for HDAC3 activation. Addition- ally, NCoR/SMRT can transiently recruit other HDACs (e.g. HDAC1-2) for full repression activity (Guenther et al., 2000). SMRTER, the Drosophila ortholog of mammalian SMRT, was identified through its interaction with ecdyson receptor (Tsai et al., 1999). TBL1 is an F box/WD-40-containing protein, originally cloned in relationship to an X-linked human disorder called Ocular Albinism with late-onset Sensorineural Deafness (OASD) (Bassi et al., 1999). WD-repeat domains are also found in two other well- studied corepressors, Drosophila Groucho (Fisher and Caudy, 1998) and yeast Tup1 (Wahi et al., 1998). In mammalian cells, TBL1 and its related TBLR1 protein share very high homology, and they have overlapping func- tions but distinct specificity (Perissi et al., 2004; Yoon et al., 2005). TBL1 and TBLR1 can directly interact with hypoacetylated histones H2B and H4 (Yoon et al., 2005), which is important for the stable association of NCoR/SMRT complexes with chromatin (Yoon et al., 2003; Yoon et al., 2005). Interestingly, the presence of the F-box domains in TBL1, TBLR1 and their homologs implies a second function of this protein family, the abil- ity to recruit the ubiquitin machinery. TBL1 associates with Siah-1 E3 ligase and SIP (Siah-interacting protein), promoting -catenin degradation linked to p53 responses (Matsuzawa and Reed, 2001). It was suggested that TBL1 and TBLR1 recruiting ubiquitin/19S proteaseome complexes to nuclear receptor- regulated targets is required for a corepressor/coactivator exchange upon ligand binding (Perissi et al., 2004). Ebi, the fly homolog of TBL1, was first identified as a downstream regulator of epidermal growth factor receptor

32

(EGFR) signaling during eye development (Dong et al., 1999). The Ebi/SMRTER complex mediates cross-talk between EGFR and Notch sig- naling by regulating Delta expression in the eye (Tsuda et al., 2002)., This is achieved by Ebi/SMRTER complex regulating expression of the charlatan gene together with the transcription factor Suppressor of Hairless (Su(H)) (Tsuda et al., 2006). Additionally, Ebi collaborates with Sina (Seven in ab- sentia) E3 ligase and the adaptor Phyllopod, inducing poly-ubiquitylation and degradation of Tramtrack during eye development (Boulton et al., 2000; Dong et al., 1999; Li et al., 2002).

Epigenetic control In contrast to the term genetic , meaning information coded by DNA se- quence, a definition of epigenetic is nuclear inheritance which is not based on differences in DNA sequence (Holliday, 1994). When gene expression is under epigenetic control, the expression pattern can be kept in a stable state during cell division, meaning heritable . There are a number of phenomena that involve epigenetic control, including position effect variegation (PEV) in Drosophila, imprinting, X chromosome inactivation in female mammals, heterochromatin formation, nuclear reprogramming, and gene silencing. Some histone modifications and DNA methylation are related to epigenetic phenomena, so they are collectively termed as epigenetic marks (reviewed in Berger, 2007). DNA methylation is found in several organisms and it occurs frequently in CpG sites in the DNA sequence. Methylated DNA can interfere with transcription factor binding and can recruit repressor proteins which specifically bind to methylated CpG. DNA methylation plays a critical role in imprinting and X chromosome inactivation in mammals. Several histone methylation sites are linked to heterochromatin formation, such as methyla- tion on H3K9, H3K27 and H4K20. Methylation of H3K27 is involved in gene silencing controlled by the Polycomb groups of proteins in Drosophila. However, it is still unclear how epigenetic persistence of chromatin states is achieved and which modifications are truly heritable.

Heterochromatin and position effect variegation Heterochromatin was named for its appearance under the light microscope: it appears dense compared to other chromatin, the euchromatin. Heterochro- matin is usually located in the centromeres and the telomeres, and is required for chromosome integrity. These regions contain many repetitious sequences but few genes. The histones in heterochromatin are often hypoacetylated and the residue H3K9 is methylated (Stewart et al., 2005). When boundary ele- ments between heterochromatin and euchromatin are removed, heterochro- matin packaging is able to spread into euchromatin (reviewed in Grewal and

33

Jia, 2007). This was firstly noticed in a phenomenon, position effect variega- tion (PEV), in Drosophila. There when a reporter gene (e.g. the white gene) is juxtaposed with heterochromatin through chromosome rearrangement or transposition, the expression of the reporter gene will show a variegated pattern. PEV has been a critical tool for identifying factors involved in het- erochromatin formation, particularly the investigation of the suppressors of PEV (Su(var) genes). The identified genes typically encode either hetero- chromatin components or enzymes that control changes in the histone modi- fications. For example, Su(var)3-9, a H3K9 methyl transferase, work to- gether with HP1, a methyl binder, and Su(var)3-7 (Jaquet et al., 2002) in a complex to maintain stable gene silencing in heterochromatin (reviewed in Ebert et al., 2006). It was suggested that the establishment of heterochro- matin in gene silencing requires a sequential set of reactions: histone de- methylase (LSD1) removes the mark of active genes, methylated H3K4, HDACs (HDAC1) deacetylates H3K9, Su(var) 3-9 methylates H3K9, HP1 binds to methylated H3K9 and finally recruit more histone methyl trans- ferase to make more heterochromatin marks (Grewal and Jia, 2007). Studies from fission yeast discovered the involvement of the RNAi machinery in the nucleation of heterochromatin (Hall et al., 2002). siRNA generated by the RNAi machinery is able to target histone-modifying proteins (e.g. clr 4, ho- molog of Su(var)3-9) to chromatin which will subsequently methylate H3K9 (Verdel and Moazed, 2005). The methylation of H3K9 not only recruits Swi6 (homolog of HP1) and other factors but also stabilize the association of the RNAi complexes with chromatin, facilitating the spreading of hetero- chromatin. Similarly, recent investigations also implicate the RNAi system in heterochromatic silencing in Drosophila (Pal-Bhadra et al., 2004).

Polycomb group and Trithorax group Polycomb Group (PcG) proteins and Trithorax Group (TrxG) proteins are conserved transcriptional regulators essential for maintaining repressed and active expression patterns of critical genes respectively during development. For example, PcG genes are necessary for silencing the Ubx gene in wing discs and the Scr gene in second and third thoracic discs in Drosophila (Beuchle et al., 2001; Ringrose and Paro, 2004). In addition, they are in- volved in cell proliferation, stem cell identity, cancer, genomic imprinting in plants and mammals, and X inactivation (reviewed in Schuettengruber et al., 2007). The DNA elements that recruit PcG and TrxG factors to chromatin are called PcG and trxG response elements (PREs and TREs) respectively. Three PcG complexes have been identified, PRC1, PRC2 and PhoRC. In Drosophila, the core components of PRC2 include E(z) (Enhancer of zeste), Esc (Extra sex comb), Su(z)12 (Suppressor of zeste 12) and Nurf55. E(z) is a HMT and able to tri-methylate H3K27 (Czermin et al., 2002), a mark recog- nized and bound by Pc (Polycomb) that is a core subunit in the PRC1 com-

34

plex. PRC1 additionally contains Ph (Polyhomeotic), Psc (Posterior sex combs) and dRing. dRing was shown to be a E3 ubiquitin ligase that monoubiquitylates histone H2A (Wang et al., 2004). The PhoRC complex was newly identified and contains the DNA-binding protein Pho (Pleio- homeotic) and dSfmbt (Scm-related gene containing four malignant brain tumor domains) protein which binds to mono-and dimethylated H3K9 and H4K20 (Klymenko et al., 2006). TrxG genes are not always acting in complexes, but there are a few trxG complexes classified. The Drosophila SWI/SNF chromatin remodeling complex, also called the BRM complex, contains the proteins BRM, MOIRA, OSA and SNR1. Trx (Trithorax), a H3K4 HMT, constitutes a TAC1 com- plex together with CBP (Petruk et al., 2001). Ash1, another H3K4 HMT, genetically and physically interacts with CBP as well (Bantignies et al., 2000; Byrd and Shearn, 2003; Petruk et al., 2001). The trxG protein Ash2 is also present in a complex but its partners haven ¯t been i dentifi ed(Papoulas et al., 1998). It is notable that the functions of PcG and trxG proteins involve sev- eral histone modifications, such as H3K27me3, H3K4me3, H2BK123ub1 and H2A119ub1 (Schuettengruber et al., 2007). Recently, it was shown that noncoding RNAs also contribute to PcG protein mediated gene silencing (reviewed in Costa, 2005).

35

Co-regulators

Transcription co-regulators are recruited to DNA by transcription factors and thus they cooperate to integrate complex regulatory information. In general, there are three means by which coregulators modulate transcription. First, they change the accessibility of chromatin for RNA polymerase II and other general transcription factors (GTFs). This class of proteins includes histone modification enzymes and chromatin remodeling factors. Second, they inter- act directly with GTFs, such as the TBP associated factors (TAFs) and the mediator complexes. Finally, they post-translationally modify transcription factors or GTFs. Examples are HATs, HDACs, and protein kinases. These modifications can either promote or disrupt the function of transcription factors. Some coregulators are integrated into complexes which constitute multiple functional units. This fulfills the requirements for complex devel- opmental programs. Notably, one coregulator complex can be used by dif- ferent transcription factors and one transcription factor can recruit distinct classes of coregulator complexes in a context dependent manner (Rosenfeld et al., 2006). Co-activators are recruited by activators, aiding transcription. A number of co-activators have been extensively studied (e.g. CBP and SAGA, reviewed in Carrozza et al., 2003; Daniel and Grant, 2007; Mannervik et al., 1999). By contrast, co-repressors interact with repressors, inhibiting tran- scription. Besides NCoR/SMRT corepressor complexes, well-known co- repressors include yeast Tup1 protein (reviewed in Malave and Dent, 2006), Groucho (reviewed in Buscarlet and Stifani, 2007; Chen and Courey, 2000), CtBP (reviewed in Chinnadurai, 2002; Mannervik et al., 1999) and Sin3/Rpd3 complexes (reviewed in Ayer, 1999; Schreiber-Agus and De- Pinho, 1998; Silverstein and Ekwall, 2005). Based on transcriptional studies in transgenic embryos, the Drosophila repressors have been grouped into two different types: the short-range rep- ressors work over short distances (less than 100bp) from the activator sites or the core promoter and the long range repressors function over long dis- tances (more than 1000 bp) (reviewed in Courey and Jia, 2001; Mannervik et al., 1999). For example, there are three binding clusters for Kr in the eve stripe 2 CRM. Two of them directly overlap the sites of Bicoid activator, and may function through competition. On the third site, Kr can recruit the dCtBP co-repressor that helps to inhibit the action of Bicoid bound nearby. On the other hand, the long-range repressors, such as Hairy and Dorsal, are able to suppress the activity of multiple enhancers. The mechanism of long-

36

range and short-range repression has not been fully understood. Three mechanisms were proposed to interpret how repressors work: competing for binding sites with activatiors; antagonizing the activity of the activator- coactivator complex (quenching); inhibiting the activity of a transcription complex at the core promoter (direct repression). Quenching and direct re- pression mechanisms involve the recruitment of co-factors. All of the identi- fied short-range repressors appear to mediate their repression dependent on dCtBP (Nibu and Levine, 2001; Nibu et al., 1998a; Zhang and Levine, 1999). In contrast, two long-range repressors require Groucho (Barolo and Levine, 1997). Table 1 summarizes repressors and co-repressors in pre-cellular em- bryos.

Table 1. The known co-repressors of repressors in pre-cellular Drosophila embryos

Co- Repressor Family Range Reference repressor dCtBP Kr short (Nibu et al., 1998a)

Kni Nuclear receptor short (Nibu et al., 1998a)

Giant bZIP short (Nibu and Levine, 2001; Strunk et al., 2001)

Snail Zinc finger short (Nibu et al., 1998a)

Ebi Snail Zinc finger short (paper III)

Groucho Hairy bHLH Long (Paroush et al., 1994)

Dorsal Rel Long (Dubnicoff et al., 1997)

Huckebein Zinc finger ? (Goldstein et al., 1999)

Capicua HMG-box ? (Jimenez et al., 2000)

Odd-skipped Zinc finger ? (Goldstein et al., 2005)

Eve Homeodomain ? (Kobayashi et al., 2001)

Runt Runt ? (Aronson et al., 1997)

Atrophin Tailless Nuclear receptor ? (Wang et al., 2006)

Brakeless Tailless Nuclear receptor ? (Paper IV)

37

CtBP The C-terminal-binding protein (CtBP) was originally identified as a cellular protein that interacts with a C-terminal motif (PXDLSXK) of adenovirus E1A oncoproteins (Boyd et al., 1993). CtBP family proteins are highly con- served among invertebrates and vertebrates. They share a high degree of amino acid homology with NAD-dependent 2-hydroxy acid dehydrogenases, so it is conceivable that they may mediate repression through their enzymatic activity (Kumar et al., 2002). The biochemical and genetic studies of Droso- phila CtBP (dCtBP) have revealed that dCtBP functions as a transcriptional corepressor (Nibu et al., 1998b; Poortinga et al., 1998). Kr, Knirps and Snail interact with dCtBP through conserved CtBP-interaction motifs (Nibu et al., 1998b). However, a similar interaction motif has not been found in the Giant protein sequence (Nibu and Levine, 2001). Interestingly, these repressors harbor both dCtBP-dependent and dCtBP-independent activities (Keller et al., 2000; La Rosee-Borggreve et al., 1999; Strunk et al., 2001). Hairy uses Groucho for its repression acitivity but it contains a divergent CtBP- interacting motif (PLSLV) which may antagonize the activity of Groucho (Zhang and Levine, 1999). In addition, dCtBP is involved in signaling path- ways, including the Notch and Wnt pathways. Specifically, dCtBP can be recruited by Hairless to Supressor of Hairless (Su(H) targeted promoters, repressing Notch target genes (Barolo et al., 2002; Morel et al., 2001). dCtBP was shown to associate with Wnt response elements and was able to repress Wnt target genes (Fang et al., 2006). The precise mechanism by which CtBP proteins mediate transcriptional repression remains to be eluci- dated. Human CtBP has been reported to associate with HDACs (Sundqvist et al., 1998). However, CtBP-mediated repression of certain promoters seems to be sensitive to HDAC inhibitor trichostatin A (TSA), while others are not (Chinnadurai, 2002). In Drosophila, the function of dCtBP is not impaired in Rpd3 mutant embryos (Mannervik and Levine, 1999). Further- more, dCtBP-mediated repression in cell culture is insensitive to TSA (Ryu and Arnosti, 2003). So the involvement of HDACs in dCtBP function is not clear.

Groucho The Drosophila Groucho (Gro) protein was identified as a corepressor of Hairy, a bHLH (basic helix-loop-helix) repressor that is essential for seg- mentation and neurogenesis (Paroush et al., 1994). Since then, more repres- sors have been identified to interact with Gro. These factors include Dorsal (Dubnicoff et al., 1997; Flores-Saaib et al., 2001), Runt (Aronson et al., 1997), Eve (Kobayashi et al., 2001), Odd-skipped (Goldstein et al., 2005), En (Jimenez et al., 1997), Huckebein (Goldstein et al., 1999), Brinker

38

(Hasson et al., 2001) and Capicua (Jimenez et al., 2000). It was shown that the recruitment of Gro by certain transcription factors is through the WRPW- like motifs (e.g. WRPW in Hairy, WRPY in Runt) or the engrailed homol- ogy domain 1 (EH-1) motifs (e.g. FSISNIL in Engrailed) (reviewed in Bus- carlet and Stifani, 2007; Chen and Courey, 2000). The human homologs of Gro have been named Transducin-like enhancer of split (TLE) proteins (Stifani et al., 1992). The Gro/TLE proteins contain two conserved domains, a WD-repeat domain and a glutamine-rich (Q) domain, both of which are required for the repression function. Specifically, the WD-repeat domain can mediate protein-protein interactions (Chen and Courey, 2000; Jimenez et al., 1997) and the Q domain mediates homotetramerization and higher-order oligomerization (Chen et al., 1998; Song et al., 2004). It was suggested that the oligomerization of Gro and the ability to interact with histones and Rpd3 may contribute to the long-range repression activity of Gro (Chen et al., 1999; Courey and Jia, 2001; Song et al., 2004). Studies in Drosophila indi- cated that Gro may function to convert an activator into a repressor. For ex- ample, Dorsal activates ventral-and-lateral- expressed genes (e.g. twist, rho, and sog) but it represses dorsal-expressing genes (e.g. dpp, zen and tld). Its repression activity requires the association with Cut and Dead Ringer (Dri), which will together facilitate the recruitment of Gro onto the target promoter (Valentine et al., 1998). Another example is dTcf (Drosophila homolog of T cell factor) in Wingless signaling. In the absence of Wingless, dTcf is con- verted to a repressor, which involves the recruitment of Gro (Cavallo et al., 1998; Levanon et al., 1998).

Atrophin and Brakeless Atrophin proteins represent a new class of nuclear receptor corepressors (Wang et al., 2006; Zhang et al., 2006a). Human Atrophin1 (ATN1) and Atrophin2 (ATN2, RERE) are polyglutamine (poly-Q) proteins and expan- sion of the poly-Q domain of ATN1 causes the neurodegenerative disease Dentatorubral-pallidoluysian atrophy (DRPLA) (Kanazawa, 1999). Mouse and zebrafish Atrophin2 recruit HDAC1/2 and are involved in Fgf8 signal- ing during embryogenesis (Asai et al., 2006; Plaster et al., 2007; Zoltewicz et al., 2004). Mouse Atrophin1 is not required for viability (Shen et al., 2007) but interacts with TLX, which may contribute to prevent retinal malforma- tion and degeneration during development (Zhang et al., 2006a). The TLX repressor is the homolog of Drosophila Tll repressor and both belong to the NR2 subfamily of orphan nuclear receptors (reviewed in Moran and Jimenez, 2006). Mutations in Drosophila atrophin cause a variety of patterning de- fects (Erkner et al., 2002; Fanto et al., 2003; Kankel et al., 2004; Zhang et al., 2002b). It has been suggested that Drosophila Atrophin is recruited to DNA by Tll and represses transcription by interacting with HDAC1 and HDAC2

39

(Wang et al., 2006). Tll specifies the terminal structures of embryo by re- pressing transcription of genes such as Kr and kni. Atrophin was also shown to interact with Eve and Hkb (Zhang et al., 2002b). Additionally, Atrophin has been classed as a member of the Trithorax Group of genes (Kankel et al., 2004). Atrophin has been shown to associate with the atypical cadherin Fat and is required for establishment of planar cell polarity in the eye and wing (Fanto et al., 2003). Recently, Atrophin was shown to genetically interact with Groucho and Brakeless (Wehn and Campbell, 2006). Brakeless (Bks) is also known as Scribbler, and as Master of thickveins in Drosophila. Previous studies have discovered three kinds of mutant pheno- types: 1) a behavioral locomotor phenotype in larvae (Yang et al., 2000); 2) deregulation of expression of the thickveins (tkv) gene in the wing imaginal disc (Funakoshi et al., 2001); 3) defects in axonal guidance in the eye due to the derepression of the runt gene in R2 and R5 photo receptors (Kaminker et al., 2002; Rao et al., 2000; Senti et al., 2000). Bks encodes a nuclear protein. The facts that Bks genetically interacts with Atrophin and Groucho and that bks mutants have similar phenotypes to atrophin mutants, imply its function in transcriptional repression (Wehn and Campbell, 2006). The bks gene lo- cus encodes two isoforms of proteins, BksA and BksB. They share the N- terminal 929 amino acids, but BksB is longer and has a 1373-residue C- terminal extension (Rao et al., 2000; Senti et al., 2000). There is a zinc finger domain as well as a glutamine-rich domain in the unique part of BksB. BksA can rescue the behavioral phenotype but only partially rescue the axon guid- ance phenotype. However, BksB fully rescued the defect in axon guidance (Senti et al., 2000; Yang et al., 2000) This result indicates that these two isoforms have non-overlapping function. The human homologs of Bks are two putative zinc finger proteins, ZFP608 and ZFP609. Recently, it was suggested that ZFP608 may play an inhibitory role in rag1 (recombination- activating gene1) and rag2 expression (Zhang et al., 2006b).

40

Aim of this thesis

The purpose of this thesis has been to understand the in vivo function of transcriptional co-regulators, using Drosophila melanogaster as a model. These co-regulators, associated with DNA binding transcription factors, regulate gene expression and thus affect the developmental process. dAda2b and Reptin belong to HAT complexes, whereas Ebi associates with HDAC3. This class of proteins may affect transcription by changing the accessibility of chromatin. Brakeless is a nuclear protein with previously unknown function that we isolated from a genetic screen. Our aim was to investigate the function of these factors and their partners in Drosophila development.

41

Results and discussion of project

Paper I The SAGA complex is evolutionarily conserved and contains a trimeric sub- complex which consists of Gcn5, Ada2 and Ada3. There are two gene loci encoding two Ada2 proteins in the Drosophila genome. Although dAda2a and dAda2b are both associated with Gcn5, they exist in different complexes. dAda2b is present in a SAGA-like complex (Kusch et al., 2003; Muratoglu et al., 2003). We generated dAda2b mutant flies from a P-element transpo- son line by using imprecise excision. The homozygous mutant animals die during the early pupa stage without any gross morphological abnormalities. We showed that dAda2b is able to interact with Gcn5 in vitro and that dAda2b and Gcn5 are expressed in many of the same tissues. Since yeast Ada2 is required for maintaining Gcn5 in the SAGA complex and stimulates the catalytic activity of Gcn5, we performed immunostainings of Drosophila embryos and polytene chromosomes in salivary glands to investigate histone acetylation level in dAda2b mutants by using antibodies against different acetylated forms of histone H3 and histone H4. Consistent with previous data from yeast which showed that Gcn5 preferentially acetylates histone H3, we found that the acetylation of histone H3 on lysine 9 and lysine 14 is dra- matically reduced in homozygous dAda2b mutants but that of histone H4 on tested lysines is normal. In addition, chromatin acetylation is globally af- fected by the dAda2b mutation, rather than affecting a subset of gene loci. We conclude that dAda2b is required for viability and normal H3 acetylation levels either through affecting the catalytic activity of Gcn5 or proper target- ing of Gcn5 to chromatin. It has been shown that the activation domain in p53 can interact with both mammalian and Drosophila SAGA complexes (Candau et al., 1997; Kusch et al., 2003; Utley et al., 1998). In order to investigate whether dAda2b con- tributes to p53 function in vivo, we analyzed the p53-dependent induction of apoptosis and reaper (rpr) gene expression in response to X-rays. Contradic- tory to the in vitro results, p53-induced apoptosis and reaper gene expres- sion in response to DNA damage are not impaired in dAda2b mutants, indi- cating that the dAda2b/Gcn5 subcomplex of SAGA is not necessary for p53 function in vivo. Unexpectedly, increased apoptosis was observed after

42

treatment with gamma radiation in dAda2 mutants. However, when we in- duced reaper expression by using a reaper transgene instead of by DNA damage, we did not detect any excessive apoptosis in dAda2b mutants. Fur- ther, we generated a dAda2b and p53 double mutant to examine whether the increased apoptosis is dependent on p53. The double mutants did not show an increase in apoptosis following irradiation, suggesting that the excessive apoptosis in dAda2b single mutants is p53 dependent. We propose that in- creased apoptosis in dAda2b mutants could result from inefficient DNA re- pair, forcing more cells into apoptosis. If this is true, the function may be evolutionarily conserved. When we tested yeast strains lacking Ada2p, Ada3p or Gcn5p proteins on plates containing the DNA damaging agent MMS, they all showed sensitivity to MMS. Taken together, our results show that Ada2 and its associated proteins are sensitive to genotoxic agents in both yeast and flies, indicating that an Ada2-Gcn5 complex may facilitate efficient DNA repair or the generation of a DNA damage signal.

Paper II Reptin was shown to co-purify with the Polycomb complex PRC1 in Droso- phila (Saurin et al., 2001), prompting us to test if the biochemical interaction with PRC1 was accompanied by a genetic interaction. We examined three components of the PRC1 complex, Pc, Psc and Ph, and two components of the E(z)-Esc complex (also called PRC2 complex), Esc and Pcl. As shown in paper II, reptin genetically interacts with members from the PRC1 complex to regulate expression of the Hox gene Scr (Sex comb reduced). Additionally, reptin interacts with Pcl from the PRC2 complex. A lacZ transgene driven by the IDE enhancer and a PRE from another Hox gene, Ubx, is derepressed in reptin mutant clones in the wing imaginal disc. However, the endogenous Ubx gene is still silent in the reptin mutant clones in the wing discs, suggest- ing that Reptin is not essential for the expression of Ubx and that repression of Ubx gene is not as sensitive to the loss of Reptin as the Ubx PRE. reptin mutants suppress PEV, which is another feature different from most PcG genes. So we do not consider reptin a bona fide PcG gene. Since Reptin is present in a Tip60 complex in mammals and recently was shown to be a component of a Drosophila Tip60 complex (Doyon et al., 2004; Kusch et al., 2004), it is possible that the genetic interactions with PcG genes are through the presence of Reptin in the Tip60 complex. In fact, genes encoding three other components in the Tip60 complex, E(Pc), domino and dMRG15, share with reptin the ability to genetically interact with PcG genes and suppress PEV. The Tip60 complex possesses at least two enzymatic activities, a HAT activity associated with Tip60 and an ATPase activity associated with Dom- ino. Domino protein is similar to P400 and to SRCAP in mammals and to Swr1 in yeast (Eissenberg et al., 2005). This class of proteins has histone

43

exchange activities (Kobor et al., 2004; Krogan et al., 2003; Mizuguchi et al., 2004). For example, Swr1 exchange H2A for the variant histone H2A.Z in nucleosomes (Mizuguchi et al., 2004). The fly Domino protein exchanges phosphorylated and acetylated H2Av for unmodified H2Av after DNA dam- age (Kusch et al., 2004). What can be the basis for the genetic interaction between Tip60 compo- nents and PcG genes and that for the function in silent chromatin formation? One possibility is that the Tip60 complex regulates PcG expression. How- ever, we did not observe reduced Pc expression in reptin mutant embryos. Another possibility is that the enzymatic activities of the Tip60 complex cooperate with PcG genes to mediate transcriptional silencing. Since binding of Pc to polytene chromosomes is abolished in H2Av mutant animals (Swaminathan et al., 2005), histone variant exchange might cause the genetic interaction with PRC1. However, we found that binding of PcG proteins to polytene chromosomes is unaffected in domino mutant larvae. Htz1 is in- volved in the spreading of silenced chromatin (Babiarz et al., 2006; Me- neghini et al., 2003). However, no change was found in binding of H2Av to polytene chromosomes from domino mutant larvae. It is possible that PRC1- mediated H2A ubiquitylation helps to recruit the Tip60 complex, whose histone acetylation or histone exchange activity assists in transcriptional repression. Alternatively, histone acetylation or exchange facilitates binding of the PRC1 complex to the PREs. Crosstalk among different histone modi- fications may also contribute to the function of the Tip60 complex in hetero- chromatin.

Paper III The Snail-family of repressors play central roles in morphogenesis by induc- ing mesoderm formation in several organisms, including flies and mammals (reviewed in Nieto, 2002).We identified Ebi as an essential co-repressor for the Drosophila Snail protein. In ebi mutant embryos, Snail-target genes are de-repressed in the presumptive mesoderm. Ebi and Snail interact both ge- netically and physically. We further identified an evolutionarily conserved interaction motif in Snail that is necessary for Ebi binding and for the repres- sion activity of Snail. Ebi functions independently of another Snail co- repressor, dCtBP, based on several lines of evidence. First, dCtBP protein level is not affected in ebi mutant embryos. Second, the ebi mutation did not affect the function of the Kr and Kni repressors, which also requires dCtBP. Third, the Ebi-interaction domain (Snail 1-40) does not bind to dCtBP but still has repression activity in S2 cells and in transgenic embryos. Fourth, removing the Ebi-interacting motif from the Snail protein abolishes its re- pression activity. Our experiments show that repression of several Snail- target genes requires both dCtBP and Ebi, and that Snail recruits both CtBP

44

and Ebi to the same rho enhancer in the presumptive mesoderm. Since the ebi mutation prevents repression of rho, sog and sim but not brk, the ebi mutant behaves as a snail hypomorph. Moreover, the Ebi-dependent repres- sion domain, Snail 1-40, has weaker activity than the full-length repression domain (Snail 1-245) in transgenic embryos. Taken together, these results suggest that Snail requires both Ebi and dCtBP for full repressor activity.

HDAC3

Ebi

Y--CPLKKRP a.a. 245 Zinc fingers a.a. 390 Sna P-DLS-K P-DLS-R DNA-binding CtBP CtBP

Figure 9. The interactions of Snail with Ebi and CtBP co-repressors. Schematic drawing of Snail protein shows the C-terminal zinc-finger DNA binding domain and N-terminal repression domain. Ebi interacts with Snail through the motif Y-- CPLKKRP. CtBP interacts with Snail through two motifs, P-DLS-K and P-DLS- R. Ebi and HDAC3 are present in the same complex and HDAC activity is part of mechanism by which Snail mediates repression.

In mammals, the Ebi homolog TBL1 is part of the N-CoR/SMRT/HDAC3 co-repressor complex. Histone deacetylation is functionally linked to the activity of this complex. Our studies found that Drosophila HDAC3 and Ebi associate and that both are required for Snail repressor function in S2 cells, as determined by RNAi and inhibition of HDAC activity. In the early em- bryo, Ebi is recruited to a Snail target gene in a Snail-dependent manner, which coincides with histone hypoacetylation. These results indicate that histone deacetylation is part of the mechanism by which Snail mediates re- pression and is a repression mechanism in early Drosophila development (Figure 9).

Paper IV From a screen for maternal factors that control patterning of Drosophila embryos, we selected mutants which specifically affect gene expression in the early embryo. We found that embryo segmentation is severely affected in bks germline clone mutants, likely resulting from the derepression of two gap genes, kni and Kr. However, mutations in bks do not affect the expres- sion of other gap genes that are known to control the transcription of kni and Kr, indicating that Bks may be required for function of one or more gap pro-

45

teins. We examined the function of Kni, Kr, Gt, Hb and Tll in bks mutants. All of them except Tll function normally, suggesting that Bks is specifically required for Tll-mediated repression. Additionally, we found that Bks ge- netically and physically interacts with Tll. Bks is recruited to the CRMs of Tll target genes, and represses transcription when tethered to DNA. Tll was recently shown to utilize Atrophin as a co-repressor (Wang et al., 2006). We show that both Bks and Atrophin are recruited to the kni CRM. Additionally, we demonstrate that Atrophin and Bks interact in vitro and that they can be co-immunoprecipitated from S2 cells. The interaction is conserved between fly and human as the human ZFP608 (homolog of Bks) protein interacts with human Atrophin1. We propose that Bks and Atrophin function together as a co-repressor complex and the complex may function throughout develop- ment as bks mutants share similar phenotypes with atrophin mutants in sev- eral developmental processes (Wehn and Campbell, 2006). Tll may simulta- neously or separately interact with Bks and Atrophin. In either case, both Bks and Atrophin are required for full Tll activity. However, overexpressing Tll from a heat-shock promoter can repress the posterior kni stripe in both wild-type and bks mutant embryos. This result indicates that Bks activity is dispensable at high enough level of Tll. We believe that Bks and Atrophin are cooperating as Tll co-repressors, so that Tll function is only partially impaired by the absence of either one. Interaction with HDACs may be part of the mechanism by which Atrophin mediates repression. We found that Bks-mediated repression in S2 cells is insensitive to the HDAC inhibitor TSA (Trichostatin A). So it is possible that Bks could repress transcription through a separate mechanism.

46

Conclusion and Perspectives

Gene regulation in multi-cellular organisms is a key process to implement the genetic program that guides development. In the Drosophila embryo, zygotically expressed transcription factors require maternal co-regulators for their activity. In this thesis, we have studied the in vivo function of four co- factors, dAda2b, Reptin, Bks and Ebi in Drosophila development. Although these proteins are widely expressed, it appears that they take part in specific cellular and developmental processes. dAda2b mutation strongly affects global histone acetylation in the central nervous system (CNS) during the embryonic stage and that on polytene chromosomes at the third instar larva stage. dAda2b mutants are hyper-sensitive to irradiation-induced DNA dam- age and yeast Ada2 and other SAGA complex components are sensitive to the DNA-damage agent MMS. This conserved feature indicates a role of Ada2 and its associated SAGA complex in DNA repair. It would be interest- ing to find out the precise mechanism. Our studies revealed a new role of Reptin and other TIP60 HAT complex components in PcG-mediated repres- sion and in formation of silent chromatin. How does a HAT complex influ- ence gene silencing? Again the molecular basis needs to be explored. To establish localized and tissue-specific patterns of gene expression, transcrip- tional repressors require co-regulators for their full activities. Here we identi- fied Ebi and Bks as novel co-repressors of Snail and Tailless respectively. We further found that histone deacetylation is necessary for Snail-mediated repression. This is the first evidence that histone deacetylation may be in- volved in cell fate specification during Drosophila embryonic development. Chromatin regulators are actively involved in transcription. However, in- vestigation of their roles in development is far from complete. Our studies give insight into certain aspects of the function of these proteins in Droso- phila development.

47

Acknowledgements

During these years in Sweden, I have received tremendous help and support from my family, friends and colleagues. I would like to acknowledge

My supervisor Mattias Mannervik, for giving me your invaluable scientific guidance throughout my studies, your continuous support and encourage- ment as well as your great patience and kindness.

Professor Christos Samakovlis, for giving me good advice on my work and my career and always being there when we need help.

My former supervisor, Xiushan Wu, for bringing me to Sweden, being sup- portive to my decisions and giving me many good suggestions.

My former colleagues, Tobias, for taking a lot of general responsibilities in the lab and translating Swedish documents for me; Mattias Bergman, for helping with the Ebi project and many funny talks; Achim, for the pleasant discussions about projects and movies; Haojiang and Haining, for all useful information and help; my present colleagues, Filip and Fang, for being in the lab with me during my last stay.

Monika, for taking care of our flies, making embryo injections and sharing your Chinese mushrooms; Kostas, for making fly food and taking a lot of boring tasks in the department.

Magdalena, for keeping track of us all and giving me a lot of personal help.

All the rest of the members at Devbio, for making it such a nice place to work, especially, Stefan, for providing yeast SAGA strains and suggestions on my projects; other yeast guys Emad, Sid and Jiang, for the nice occa- sional communications and the great pub time; Dana, for celebrating our common birthday and the tasty Slovakia wine; Nafiseh, for helping with cell culture work; Kirsten, for the protocols and suggestions; Katarina, for orga- nizing the nice boat trip; Satish, for being a good party-mood maker; Vasil- ios, for showing us how devoted a scientist should be; Nicole, Naumi, Olivi- ano, Chie and Ryo, for creating a good working atmosphere.

48

The former PH.D students, Peggy, Andreas, Johanna, Nikos and Marco, for always being so kind and helpful to the juniors and exemplifying a good student.

My friends in Sweden, Bing Tang, Wei Zhao, Liqun He, Ying Sun, Wei Jiao, Zhi Wang, Guiling Zhao and Dongbo Zhao, for helping us with mov- ing, parties and communications; all my friends in China, especially, Yongqun Zhang, for sending me good wishes and my favorite food from 8000 kms away.

Last but not least, I am deeply grateful to my family for the support: my parents, thanks for your loving and spoiling; my brother Tao, thanks for following me to Sweden and making my life here easier and happier; Qiu ¯s family, thanks for believing in me; my dear husband Qiu, for sharing all the happiness and difficulties with me and giving me endless encouragement and love.

49

References

Aronson, B. D., Fisher, A. L., Blechman, K., Caudy, M. and Gergen, J. P. (1997). Groucho-dependent and -independent repression activities of Runt domain proteins. Mol Cell Biol 17, 5581-7. Asahara, H., Dutta, S., Kao, H. Y., Evans, R. M. and Montminy, M. (1999). Pbx-Hox heterodimers recruit coactivator-corepressor complexes in an isoform-specific manner. Mol Cell Biol 19, 8219-25. Asai, Y., Chan, D. K., Starr, C. J., Kappler, J. A., Kollmar, R. and Hudspeth, A. J. (2006). Mutation of the atrophin2 gene in the zebrafish disrupts signaling by fibroblast growth factor during development of the inner ear. Proc Natl Acad Sci U S A 103, 9069-74. Ayer, D. E. (1999). Histone deacetylases: transcriptional repression with SINers and NuRDs. Trends Cell Biol 9, 193-8. Babiarz, J. E., Halley, J. E. and Rine, J. (2006). Telomeric heterochro- matin boundaries require NuA4-dependent acetylation of histone variant H2A.Z in Saccharomyces cerevisiae. Genes Dev 20, 700-10. Baek, S. H., Ohgi, K. A., Rose, D. W., Koo, E. H., Glass, C. K. and Rosenfeld, M. G. (2002). Exchange of N-CoR corepressor and Tip60 coac- tivator complexes links gene expression by NF-kappaB and beta-amyloid precursor protein. Cell 110, 55-67. Banin, S., Moyal, L., Shieh, S., Taya, Y., Anderson, C. W., Chessa, L., Smorodinsky, N. I., Prives, C., Reiss, Y., Shiloh, Y. et al. (1998). En- hanced phosphorylation of p53 by ATM in response to DNA damage. Sci- ence 281, 1674-7. Bantignies, F., Goodman, R. H. and Smolik, S. M. (2000). Functional interaction between the coactivator Drosophila CREB-binding protein and ASH1, a member of the trithorax group of chromatin modifiers. Mol Cell Biol 20, 9317-30. Barbaric, S., Reinke, H. and Horz, W. (2003). Multiple mechanistically distinct functions of SAGA at the PHO5 promoter. Mol Cell Biol 23, 3468- 76. Barlev, N. A., Emelyanov, A. V., Castagnino, P., Zegerman, P., Bannister, A. J., Sepulveda, M. A., Robert, F., Tora, L., Kouzarides, T., Birshtein, B. K. et al. (2003). A novel human Ada2 homologue functions with Gcn5 or Brg1 to coactivate transcription. Mol Cell Biol 23, 6944-57. Barolo, S. and Levine, M. (1997). hairy mediates dominant repression in the Drosophila embryo. Embo J 16, 2883-91. Barolo, S., Stone, T., Bang, A. G. and Posakony, J. W. (2002). Default repression and Notch signaling: Hairless acts as an adaptor to recruit the

50

corepressors Groucho and dCtBP to Suppressor of Hairless. Genes Dev 16, 1964-76. Bassi, M. T., Ramesar, R. S., Caciotti, B., Winship, I. M., De Grandi, A., Riboni, M., Townes, P. L., Beighton, P., Ballabio, A. and Borsani, G. (1999). X-linked late-onset sensorineural deafness caused by a deletion in- volving OA1 and a novel gene containing WD-40 repeats. Am J Hum Genet 64, 1604-16. Bauer, A., Chauvet, S., Huber, O., Usseglio, F., Rothbacher, U., Aragnol, D., Kemler, R. and Pradel, J. (2000). Pontin52 and reptin52 function as antagonistic regulators of beta-catenin signalling activity. Embo J 19, 6121- 30. Bauer, A., Huber, O. and Kemler, R. (1998). Pontin52, an interaction partner of beta-catenin, binds to the TATA box binding protein. Proc Natl Acad Sci U S A 95, 14787-92. Bellosta, P., Hulf, T., Balla Diop, S., Usseglio, F., Pradel, J., Aragnol, D. and Gallant, P. (2005). Myc interacts genetically with Tip48/Reptin and Tip49/Pontin to control growth and proliferation during Drosophila devel- opment. Proc Natl Acad Sci U S A 102, 11799-804. Berger, S. L. (2007). The complex language of chromatin regulation during transcription. Nature 447, 407-12. Berger, S. L., Pina, B., Silverman, N., Marcus, G. A., Agapite, J., Regier, J. L., Triezenberg, S. J. and Guarente, L. (1992). Genetic isolation of ADA2: a potential transcriptional adaptor required for function of certain acidic activation domains. Cell 70, 251-65. Bergmann, A., Stein, D., Geisler, R., Hagenmaier, S., Schmid, B., Fer- nandez, N., Schnell, B. and Nusslein-Volhard, C. (1996). A gradient of cytoplasmic Cactus degradation establishes the nuclear localization gradient of the dorsal morphogen in Drosophila. Mech Dev 60, 109-23. Bernstein, C., Bernstein, H., Payne, C. M. and Garewal, H. (2002). DNA repair/pro-apoptotic dual-role proteins in five major DNA repair pathways: fail-safe protection against carcinogenesis. Mutat Res 511, 145-78. Beuchle, D., Struhl, G. and Muller, J. (2001). Polycomb group proteins and heritable silencing of Drosophila Hox genes. Development 128, 993- 1004. Bird, A. W., Yu, D. Y., Pray-Grant, M. G., Qiu, Q., Harmon, K. E., Megee, P. C., Grant, P. A., Smith, M. M. and Christman, M. F. (2002). Acetylation of histone H4 by Esa1 is required for DNA double-strand break repair. Nature 419, 411-5. Blander, G. and Guarente, L. (2004). The Sir2 family of protein deacety- lases. Annu Rev Biochem 73, 417-35. Boudreault, A. A., Cronier, D., Selleck, W., Lacoste, N., Utley, R. T., Allard, S., Savard, J., Lane, W. S., Tan, S. and Cote, J. (2003). Yeast enhancer of polycomb defines global Esa1-dependent acetylation of chroma- tin. Genes Dev 17, 1415-28. Boulay, J. L., Dennefeld, C. and Alberga, A. (1987). The Drosophila de- velopmental gene snail encodes a protein with nucleic acid binding fingers. Nature 330, 395-8.

51

Boulton, S. J., Brook, A., Staehling-Hampton, K., Heitzler, P. and Dyson, N. (2000). A role for Ebi in neuronal cell cycle control. Embo J 19, 5376-86. Boyd, J. M., Subramanian, T., Schaeper, U., La Regina, M., Bayley, S. and Chinnadurai, G. (1993). A region in the C-terminus of adenovirus 2/5 E1a protein is required for association with a cellular phosphoprotein and important for the negative modulation of T24-ras mediated transformation, tumorigenesis and metastasis. Embo J 12, 469-78. Boyer, L. A., Langer, M. R., Crowley, K. A., Tan, S., Denu, J. M. and Peterson, C. L. (2002). Essential role for the SANT domain in the function- ing of multiple chromatin remodeling enzymes. Mol Cell 10, 935-42. Brand, M., Moggs, J. G., Oulad-Abdelghani, M., Lejeune, F., Dilworth, F. J., Stevenin, J., Almouzni, G. and Tora, L. (2001). UV-damaged DNA- binding protein in the TFTC complex links DNA damage recognition to nucleosome acetylation. Embo J 20, 3187-96. Braunstein, M., Sobel, R. E., Allis, C. D., Turner, B. M. and Broach, J. R. (1996). Efficient transcriptional silencing in Saccharomyces cerevisiae requires a heterochromatin histone acetylation pattern. Mol Cell Biol 16, 4349-56. Breen, T. R. and Harte, P. J. (1991). Molecular characterization of the trithorax gene, a positive regulator of homeotic gene expression in Droso- phila. Mech Dev 35, 113-27. Brodsky, M. H., Weinert, B. T., Tsang, G., Rong, Y. S., McGinnis, N. M., Golic, K. G., Rio, D. C. and Rubin, G. M. (2004). Drosophila melanogaster MNK/Chk2 and p53 regulate multiple DNA repair and apop- totic pathways following DNA damage. Mol Cell Biol 24, 1219-31. Brownell, J. E., Zhou, J., Ranalli, T., Kobayashi, R., Edmondson, D. G., Roth, S. Y. and Allis, C. D. (1996). Tetrahymena histone acetyltransferase A: a homolog to yeast Gcn5p linking histone acetylation to gene activation. Cell 84, 843-51. Bu, P., Evrard, Y. A., Lozano, G. and Dent, S. Y. (2007). Loss of Gcn5 acetyltransferase activity leads to neural tube closure defects and exen- cephaly in mouse embryos. Mol Cell Biol 27, 3405-16. Buggy, J. J., Sideris, M. L., Mak, P., Lorimer, D. D., McIntosh, B. and Clark, J. M. (2000). Cloning and characterization of a novel human histone deacetylase, HDAC8. Biochem J 350 Pt 1, 199-205. Buscarlet, M. and Stifani, S. (2007). The 'Marx' of Groucho on develop- ment and disease. Trends Cell Biol 17, 353-61. Byrd, K. N. and Shearn, A. (2003). ASH1, a Drosophila trithorax group protein, is required for methylation of lysine 4 residues on histone H3. Proc Natl Acad Sci U S A 100, 11535-40. Candau, R. and Berger, S. L. (1996). Structural and functional analysis of yeast putative adaptors. Evidence for an adaptor complex in vivo. J Biol Chem 271, 5237-45. Candau, R., Scolnick, D. M., Darpino, P., Ying, C. Y., Halazonetis, T. D. and Berger, S. L. (1997). Two tandem and independent sub-activation do- mains in the amino terminus of p53 require the adaptor complex for activity. Oncogene 15, 807-16.

52

Carre, C., Szymczak, D., Pidoux, J. and Antoniewski, C. (2005). The histone H3 acetylase dGcn5 is a key player in Drosophila melanogaster metamorphosis. Mol Cell Biol 25, 8228-38. Carrozza, M. J., Li, B., Florens, L., Suganuma, T., Swanson, S. K., Lee, K. K., Shia, W. J., Anderson, S., Yates, J., Washburn, M. P. et al. (2005). Histone H3 methylation by Set2 directs deacetylation of coding regions by Rpd3S to suppress spurious intragenic transcription. Cell 123, 581-92. Carrozza, M. J., Utley, R. T., Workman, J. L. and Cote, J. (2003). The diverse functions of histone acetyltransferase complexes. Trends Genet 19, 321-9. Cavallo, R. A., Cox, R. T., Moline, M. M., Roose, J., Polevoy, G. A., Clevers, H., Peifer, M. and Bejsovec, A. (1998). Drosophila Tcf and Groucho interact to repress Wingless signalling activity. Nature 395, 604-8. Chen, G. and Courey, A. J. (2000). Groucho/TLE family proteins and tran- scriptional repression. Gene 249, 1-16. Chen, G., Fernandez, J., Mische, S. and Courey, A. J. (1999). A func- tional interaction between the histone deacetylase Rpd3 and the corepressor groucho in Drosophila development. Genes Dev 13, 2218-30. Chen, G., Nguyen, P. H. and Courey, A. J. (1998). A role for Groucho tetramerization in transcriptional repression. Mol Cell Biol 18, 7259-68. Chen, J. D. and Evans, R. M. (1995). A transcriptional co-repressor that interacts with nuclear hormone receptors. Nature 377, 454-7. Chinnadurai, G. (2002). CtBP, an unconventional transcriptional corepres- sor in development and oncogenesis. Mol Cell 9, 213-24. Cho, S. G., Bhoumik, A., Broday, L., Ivanov, V., Rosenstein, B. and Ronai, Z. (2001). TIP49b, a regulator of activating transcription factor 2 response to stress and DNA damage. Mol Cell Biol 21, 8398-413. Clements, A., Poux, A. N., Lo, W. S., Pillus, L., Berger, S. L. and Mar- morstein, R. (2003). Structural basis for histone and phosphohistone bind- ing by the GCN5 histone acetyltransferase. Mol Cell 12, 461-73. Cortez, D., Wang, Y., Qin, J. and Elledge, S. J. (1999). Requirement of ATM-dependent phosphorylation of brca1 in the DNA damage response to double-strand breaks. Science 286, 1162-6. Costa, F. F. (2005). Non-coding RNAs: new players in eukaryotic biology. Gene 357, 83-94. Courey, A. J. and Jia, S. (2001). Transcriptional repression: the long and the short of it. Genes Dev 15, 2786-96. Czermin, B., Melfi, R., McCabe, D., Seitz, V., Imhof, A. and Pirrotta, V. (2002). Drosophila enhancer of Zeste/ESC complexes have a histone H3 methyltransferase activity that marks chromosomal Polycomb sites. Cell 111, 185-96. Daniel, J. A. and Grant, P. A. (2007). Multi-tasking on chromatin with the SAGA coactivator complexes. Mutat Res 618, 135-48. de Napoles, M., Mermoud, J. E., Wakao, R., Tang, Y. A., Endoh, M., Appanah, R., Nesterova, T. B., Silva, J., Otte, A. P., Vidal, M. et al. (2004). Polycomb group proteins Ring1A/B link ubiquitylation of histone H2A to heritable gene silencing and X inactivation. Dev Cell 7, 663-76.

53

de Ruijter, A. J., van Gennip, A. H., Caron, H. N., Kemp, S. and van Kuilenburg, A. B. (2003). Histone deacetylases (HDACs): characterization of the classical HDAC family. Biochem J 370, 737-49. Dong, X., Tsuda, L., Zavitz, K. H., Lin, M., Li, S., Carthew, R. W. and Zipursky, S. L. (1999). ebi regulates epidermal growth factor receptor sig- naling pathways in Drosophila. Genes Dev 13, 954-65. Downs, J. A., Allard, S., Jobin-Robitaille, O., Javaheri, A., Auger, A., Bouchard, N., Kron, S. J., Jackson, S. P. and Cote, J. (2004). Binding of chromatin-modifying activities to phosphorylated histone H2A at DNA damage sites. Mol Cell 16, 979-90. Downs, J. A., Nussenzweig, M. C. and Nussenzweig, A. (2007). Chroma- tin dynamics and the preservation of genetic information. Nature 447, 951-8. Doyon, Y., Selleck, W., Lane, W. S., Tan, S. and Cote, J. (2004). Struc- tural and functional conservation of the NuA4 histone acetyltransferase complex from yeast to humans. Mol Cell Biol 24, 1884-96. Dubnicoff, T., Valentine, S. A., Chen, G., Shi, T., Lengyel, J. A., Paroush, Z. and Courey, A. J. (1997). Conversion of dorsal from an activator to a repressor by the global corepressor Groucho. Genes Dev 11, 2952-7. Dugan, K. A., Wood, M. A. and Cole, M. D. (2002). TIP49, but not TRRAP, modulates c-Myc and E2F1 dependent apoptosis. Oncogene 21, 5835-43. Ebert, A., Lein, S., Schotta, G. and Reuter, G. (2006). Histone modifica- tion and the control of heterochromatic gene silencing in Drosophila. Chro- mosome Res 14, 377-92. Eissenberg, J. C., Wong, M. and Chrivia, J. C. (2005). Human SRCAP and Drosophila melanogaster DOM are homologs that function in the notch signaling pathway. Mol Cell Biol 25, 6559-69. Erkner, A., Roure, A., Charroux, B., Delaage, M., Holway, N., Core, N., Vola, C., Angelats, C., Pages, F., Fasano, L. et al. (2002). Grunge, related to human Atrophin-like proteins, has multiple functions in Drosophila de- velopment. Development 129, 1119-29. Etard, C., Gradl, D., Kunz, M., Eilers, M. and Wedlich, D. (2005). Pontin and Reptin regulate cell proliferation in early Xenopus embryos in collabora- tion with c-Myc and Miz-1. Mech Dev 122, 545-56. Fang, J., Chen, T., Chadwick, B., Li, E. and Zhang, Y. (2004). Ring1b- mediated H2A ubiquitination associates with inactive X chromosomes and is involved in initiation of X inactivation. J Biol Chem 279, 52812-5. Fang, M., Li, J., Blauwkamp, T., Bhambhani, C., Campbell, N. and Cadigan, K. M. (2006). C-terminal-binding protein directly activates and represses Wnt transcriptional targets in Drosophila. Embo J 25, 2735-45. Fanto, M., Clayton, L., Meredith, J., Hardiman, K., Charroux, B., Ker- ridge, S. and McNeill, H. (2003). The tumor-suppressor and cell adhesion molecule Fat controls planar polarity via physical interactions with Atrophin, a transcriptional co-repressor. Development 130, 763-74. Fischle, W., Tseng, B. S., Dormann, H. L., Ueberheide, B. M., Garcia, B. A., Shabanowitz, J., Hunt, D. F., Funabiki, H. and Allis, C. D. (2005).

54

Regulation of HP1-chromatin binding by histone H3 methylation and phos- phorylation. Nature 438, 1116-22. Fisher, A. L. and Caudy, M. (1998). Groucho proteins: transcriptional corepressors for specific subsets of DNA-binding transcription factors in vertebrates and invertebrates. Genes Dev 12, 1931-40. Flores-Saaib, R. D., Jia, S. and Courey, A. J. (2001). Activation and re- pression by the C-terminal domain of Dorsal. Development 128, 1869-79. Frank, S. R., Parisi, T., Taubert, S., Fernandez, P., Fuchs, M., Chan, H. M., Livingston, D. M. and Amati, B. (2003). MYC recruits the TIP60 his- tone acetyltransferase complex to chromatin. EMBO Rep 4, 575-80. Fulco, M., Schiltz, R. L., Iezzi, S., King, M. T., Zhao, P., Kashiwaya, Y., Hoffman, E., Veech, R. L. and Sartorelli, V. (2003). Sir2 regulates skeletal muscle differentiation as a potential sensor of the redox state. Mol Cell 12, 51-62. Funakoshi, Y., Minami, M. and Tabata, T. (2001). mtv shapes the activity gradient of the Dpp morphogen through regulation of thickveins. Develop- ment 128, 67-74. Galasinski, S. C., Resing, K. A., Goodrich, J. A. and Ahn, N. G. (2002). Phosphatase inhibition leads to histone deacetylases 1 and 2 phosphorylation and disruption of corepressor interactions. J Biol Chem 277, 19618-26. Gallant, P. (2007). Control of transcription by Pontin and Reptin. Trends Cell Biol 17, 187-92. Gao, L., Cueto, M. A., Asselbergs, F. and Atadja, P. (2002). Cloning and functional characterization of HDAC11, a novel member of the human his- tone deacetylase family. J Biol Chem 277, 25748-55. Gaughan, L., Logan, I. R., Cook, S., Neal, D. E. and Robson, C. N. (2002). Tip60 and histone deacetylase 1 regulate androgen receptor activity through changes to the acetylation status of the receptor. J Biol Chem 277, 25904-13. Gavaravarapu, S. and Kamine, J. (2000). Tip60 inhibits activation of CREB protein by protein kinase A. Biochem Biophys Res Commun 269, 758- 66. Glozak, M. A., Sengupta, N., Zhang, X. and Seto, E. (2005). Acetylation and deacetylation of non-histone proteins. Gene 363, 15-23. Goldstein, R. E., Cook, O., Dinur, T., Pisante, A., Karandikar, U. C., Bidwai, A. and Paroush, Z. (2005). An eh1-like motif in odd-skipped me- diates recruitment of Groucho and repression in vivo. Mol Cell Biol 25, 10711-20. Goldstein, R. E., Jimenez, G., Cook, O., Gur, D. and Paroush, Z. (1999). Huckebein repressor activity in Drosophila terminal patterning is mediated by Groucho. Development 126, 3747-55. Govind, C. K., Zhang, F., Qiu, H., Hofmeyer, K. and Hinnebusch, A. G. (2007). Gcn5 promotes acetylation, eviction, and methylation of nu- cleosomes in transcribed coding regions. Mol Cell 25, 31-42. Grant, P. A., Berger, S. L. and Workman, J. L. (1999). Identification and analysis of native nucleosomal histone acetyltransferase complexes. Methods Mol Biol 119, 311-7.

55

Greenblatt, M. S., Bennett, W. P., Hollstein, M. and Harris, C. C. (1994). Mutations in the p53 tumor suppressor gene: clues to cancer etiology and molecular pathogenesis. Cancer Res 54, 4855-78. Gregoire, S., Xiao, L., Nie, J., Zhang, X., Xu, M., Li, J., Wong, J., Seto, E. and Yang, X. J. (2007). Histone deacetylase 3 interacts with and deacety- lates myocyte enhancer factor 2. Mol Cell Biol 27, 1280-95. Grewal, S. I. and Jia, S. (2007). Heterochromatin revisited. Nat Rev Genet 8, 35-46. Guarente, L. (2001). SIR2 and aging--the exception that proves the rule. Trends Genet 17, 391-2. Guelman, S., Suganuma, T., Florens, L., Swanson, S. K., Kiesecker, C. L., Kusch, T., Anderson, S., Yates, J. R., 3rd, Washburn, M. P., Abmayr, S. M. et al. (2006). Host cell factor and an uncharacterized SANT domain protein are stable components of ATAC, a novel dAda2A/dGcn5-containing histone acetyltransferase complex in Drosophila. Mol Cell Biol 26, 871-82. Guenther, M. G., Lane, W. S., Fischle, W., Verdin, E., Lazar, M. A. and Shiekhattar, R. (2000). A core SMRT corepressor complex containing HDAC3 and TBL1, a WD40-repeat protein linked to deafness. Genes Dev 14, 1048-57. Hahn, S. (2004). Structure and mechanism of the RNA polymerase II tran- scription machinery. Nat Struct Mol Biol 11, 394-403. Hall, I. M., Shankaranarayana, G. D., Noma, K., Ayoub, N., Cohen, A. and Grewal, S. I. (2002). Establishment and maintenance of a heterochro- matin domain. Science 297, 2232-7. Harding, K. and Levine, M. (1988). Gap genes define the limits of anten- napedia and bithorax gene expression during early development in Droso- phila. Embo J 7, 205-14. Hasson, P., Muller, B., Basler, K. and Paroush, Z. (2001). Brinker re- quires two corepressors for maximal and versatile repression in Dpp signal- ling. Embo J 20, 5725-36. Helmlinger, D., Hardy, S., Eberlin, A., Devys, D. and Tora, L. (2006). Both normal and polyglutamine- expanded ataxin-7 are components of TFTC-type GCN5 histone acetyltransferase- containing complexes. Biochem Soc Symp, 155-63. Hemavathy, K., Meng, X. and Ip, Y. T. (1997). Differential regulation of gastrulation and neuroectodermal gene expression by Snail in the Drosophila embryo. Development 124, 3683-91. Hirao, A., Kong, Y. Y., Matsuoka, S., Wakeham, A., Ruland, J., Yoshida, H., Liu, D., Elledge, S. J. and Mak, T. W. (2000). DNA damage-induced activation of p53 by the checkpoint kinase Chk2. Science 287, 1824-7. Hoch, M., Gerwin, N., Taubert, H. and Jackle, H. (1992). Competition for overlapping sites in the regulatory region of the Drosophila gene Kruppel. Science 256, 94-7. Holliday, R. (1994). Epigenetics: an overview. Dev Genet 15, 453-7. Horlein, A. J., Naar, A. M., Heinzel, T., Torchia, J., Gloss, B., Kurokawa, R., Ryan, A., Kamei, Y., Soderstrom, M., Glass, C. K. et al. (1995).

56

Ligand-independent repression by the thyroid mediated by a nuclear receptor co-repressor. Nature 377, 397-404. Hubbert, C., Guardiola, A., Shao, R., Kawaguchi, Y., Ito, A., Nixon, A., Yoshida, M., Wang, X. F. and Yao, T. P. (2002). HDAC6 is a micro- tubule-associated deacetylase. Nature 417, 455-8. Ikura, T., Ogryzko, V. V., Grigoriev, M., Groisman, R., Wang, J., Horikoshi, M., Scully, R., Qin, J. and Nakatani, Y. (2000). Involvement of the TIP60 histone acetylase complex in DNA repair and apoptosis. Cell 102, 463-73. Ip, Y. T. and Gridley, T. (2002). Cell movements during gastrulation: snail dependent and independent pathways. Curr Opin Genet Dev 12, 423-9. Ip, Y. T., Levine, M. and Small, S. J. (1992). The bicoid and dorsal morphogens use a similar strategy to make stripes in the Drosophila embryo. J Cell Sci Suppl 16, 33-8. Ip, Y. T., Maggert, K. and Levine, M. (1994). Uncoupling gastrulation and mesoderm differentiation in the Drosophila embryo. Embo J 13, 5826-34. Jackle, H., Hoch, M., Pankratz, M. J., Gerwin, N., Sauer, F. and Bron- ner, G. (1992). Transcriptional control by Drosophila gap genes. J Cell Sci Suppl 16, 39-51. Jaquet, Y., Delattre, M., Spierer, A. and Spierer, P. (2002). Functional dissection of the Drosophila modifier of variegation Su(var)3-7. Develop- ment 129, 3975-82. Jiang, J., Cai, H., Zhou, Q. and Levine, M. (1993). Conversion of a dor- sal-dependent silencer into an enhancer: evidence for dorsal corepressors. Embo J 12, 3201-9. Jimenez, G., Guichet, A., Ephrussi, A. and Casanova, J. (2000). Relief of gene repression by torso RTK signaling: role of capicua in Drosophila ter- minal and dorsoventral patterning. Genes Dev 14, 224-31. Jimenez, G., Paroush, Z. and Ish-Horowicz, D. (1997). Groucho acts as a corepressor for a subset of negative regulators, including Hairy and En- grailed. Genes Dev 11, 3072-82. Juan, L. J., Shia, W. J., Chen, M. H., Yang, W. M., Seto, E., Lin, Y. S. and Wu, C. W. (2000). Histone deacetylases specifically down-regulate p53-dependent gene activation. J Biol Chem 275, 20436-43. Kaminker, J. S., Canon, J., Salecker, I. and Banerjee, U. (2002). Control of photoreceptor axon target choice by transcriptional repression of Runt. Nat Neurosci 5, 746-50. Kanazawa, I. (1999). Molecular pathology of dentatorubral-pallidoluysian atrophy. Philos Trans R Soc Lond B Biol Sci 354, 1069-74. Kankel, M. W., Duncan, D. M. and Duncan, I. (2004). A screen for genes that interact with the Drosophila pair-rule segmentation gene fushi tarazu. Genetics 168, 161-80. Kao, H. Y., Ordentlich, P., Koyano-Nakagawa, N., Tang, Z., Downes, M., Kintner, C. R., Evans, R. M. and Kadesch, T. (1998). A histone deacety- lase corepressor complex regulates the Notch signal transduction pathway. Genes Dev 12, 2269-77.

57

Kasten, M. M., Dorland, S. and Stillman, D. J. (1997). A large protein complex containing the yeast Sin3p and Rpd3p transcriptional regulators. Mol Cell Biol 17, 4852-8. Kawaguchi, Y., Kovacs, J. J., McLaurin, A., Vance, J. M., Ito, A. and Yao, T. P. (2003). The deacetylase HDAC6 regulates aggresome formation and cell viability in response to misfolded protein stress. Cell 115, 727-38. Keller, S. A., Mao, Y., Struffi, P., Margulies, C., Yurk, C. E., Anderson, A. R., Amey, R. L., Moore, S., Ebels, J. M., Foley, K. et al. (2000). dCtBP-dependent and -independent repression activities of the Drosophila Knirps protein. Mol Cell Biol 20, 7247-58. Khanna, K. K. and Jackson, S. P. (2001). DNA double-strand breaks: sig- naling, repair and the cancer connection. Nat Genet 27, 247-54. Kim, J. H., Choi, H. J., Kim, B., Kim, M. H., Lee, J. M., Kim, I. S., Lee, M. H., Choi, S. J., Kim, K. I., Kim, S. I. et al. (2006). Roles of sumoyla- tion of a reptin chromatin-remodelling complex in cancer metastasis. Nat Cell Biol 8, 631-9. Kim, J. H., Kim, B., Cai, L., Choi, H. J., Ohgi, K. A., Tran, C., Chen, C., Chung, C. H., Huber, O., Rose, D. W. et al. (2005). Transcriptional regula- tion of a metastasis suppressor gene by Tip60 and beta-catenin complexes. Nature 434, 921-6. Kimura, A. and Horikoshi, M. (1998). Tip60 acetylates six lysines of a specific class in core histones in vitro. Genes Cells 3, 789-800. Klymenko, T., Papp, B., Fischle, W., Kocher, T., Schelder, M., Fritsch, C., Wild, B., Wilm, M. and Muller, J. (2006). A Polycomb group protein complex with sequence-specific DNA-binding and selective methyl-lysine- binding activities. Genes Dev 20, 1110-22. Kobayashi, M., Goldstein, R. E., Fujioka, M., Paroush, Z. and Jaynes, J. B. (2001). Groucho augments the repression of multiple Even skipped target genes in establishing parasegment boundaries. Development 128, 1805-15. Kobor, M. S., Venkatasubrahmanyam, S., Meneghini, M. D., Gin, J. W., Jennings, J. L., Link, A. J., Madhani, H. D. and Rine, J. (2004). A pro- tein complex containing the conserved Swi2/Snf2-related ATPase Swr1p deposits histone variant H2A.Z into euchromatin. PLoS Biol 2, E131. Kohler, A., Pascual-Garcia, P., Llopis, A., Zapater, M., Posas, F., Hurt, E. and Rodriguez-Navarro, S. (2006). The mRNA export factor Sus1 is involved in Spt/Ada/Gcn5 acetyltransferase-mediated H2B deubiquitinyla- tion through its interaction with Ubp8 and Sgf11. Mol Biol Cell 17, 4228-36. Kouzarides, T. (2007). Chromatin modifications and their function. Cell 128, 693-705. Krogan, N. J., Keogh, M. C., Datta, N., Sawa, C., Ryan, O. W., Ding, H., Haw, R. A., Pootoolal, J., Tong, A., Canadien, V. et al. (2003). A Snf2 family ATPase complex required for recruitment of the histone H2A variant Htz1. Mol Cell 12, 1565-76. Kumar, V., Carlson, J. E., Ohgi, K. A., Edwards, T. A., Rose, D. W., Escalante, C. R., Rosenfeld, M. G. and Aggarwal, A. K. (2002). Tran- scription corepressor CtBP is an NAD(+)-regulated dehydrogenase. Mol Cell 10, 857-69.

58

Kuo, M. H., vom Baur, E., Struhl, K. and Allis, C. D. (2000). Gcn4 acti- vator targets Gcn5 histone acetyltransferase to specific promoters independ- ently of transcription. Mol Cell 6, 1309-20. Kurdistani, S. K., Tavazoie, S. and Grunstein, M. (2004). Mapping global histone acetylation patterns to gene expression. Cell 117, 721-33. Kusch, T., Florens, L., Macdonald, W. H., Swanson, S. K., Glaser, R. L., Yates, J. R., 3rd, Abmayr, S. M., Washburn, M. P. and Workman, J. L. (2004). Acetylation by Tip60 is required for selective histone variant ex- change at DNA lesions. Science 306, 2084-7. Kusch, T., Guelman, S., Abmayr, S. M. and Workman, J. L. (2003). Two Drosophila Ada2 homologues function in different multiprotein complexes. Mol Cell Biol 23, 3305-19. La Rosee-Borggreve, A., Hader, T., Wainwright, D., Sauer, F. and Jackle, H. (1999). hairy stripe 7 element mediates activation and repression in response to different domains and levels of Kruppel in the Drosophila embryo. Mech Dev 89, 133-40. LaJeunesse, D. and Shearn, A. (1995). Trans-regulation of thoracic ho- meotic selector genes of the Antennapedia and bithorax complexes by the trithorax group genes: absent, small, and homeotic discs 1 and 2. Mech Dev 53, 123-39. Landry, J., Sutton, A., Hesman, T., Min, J., Xu, R. M., Johnston, M. and Sternglanz, R. (2003). Set2-catalyzed methylation of histone H3 represses basal expression of GAL4 in Saccharomyces cerevisiae. Mol Cell Biol 23, 5972-8. Larschan, E. and Winston, F. (2001). The S. cerevisiae SAGA complex functions in vivo as a coactivator for transcriptional activation by Gal4. Genes Dev 15, 1946-56. Lebrecht, D., Foehr, M., Smith, E., Lopes, F. J., Vanario-Alonso, C. E., Reinitz, J., Burz, D. S. and Hanes, S. D. (2005). Bicoid cooperative DNA binding is critical for embryonic patterning in Drosophila. Proc Natl Acad Sci U S A 102, 13176-81. Lee, J. H., Lee, E., Park, J., Kim, E., Kim, J. and Chung, J. (2003). In vivo p53 function is indispensable for DNA damage-induced apoptotic sig- naling in Drosophila. FEBS Lett 550, 5-10. Leptin, M. (1991). twist and snail as positive and negative regulators during Drosophila mesoderm development. Genes Dev 5, 1568-76. Leptin, M. (1999). Gastrulation in Drosophila: the logic and the cellular mechanisms. Embo J 18, 3187-92. Leptin, M., Casal, J., Grunewald, B. and Reuter, R. (1992). Mechanisms of early Drosophila mesoderm formation. Dev Suppl, 23-31. Levanon, D., Goldstein, R. E., Bernstein, Y., Tang, H., Goldenberg, D., Stifani, S., Paroush, Z. and Groner, Y. (1998). Transcriptional repression by AML1 and LEF-1 is mediated by the TLE/Groucho corepressors. Proc Natl Acad Sci U S A 95, 11590-5. Li, B., Carey, M. and Workman, J. L. (2007a). The role of chromatin dur- ing transcription. Cell 128, 707-19.

59

Li, B., Samanta, A., Song, X., Iacono, K. T., Bembas, K., Tao, R., Basu, S., Riley, J. L., Hancock, W. W., Shen, Y. et al. (2007b). FOXP3 interac- tions with histone acetyltransferase and class II histone deacetylases are re- quired for repression. Proc Natl Acad Sci U S A 104, 4571-6. Li, J., Wang, J., Wang, J., Nawaz, Z., Liu, J. M., Qin, J. and Wong, J. (2000). Both corepressor proteins SMRT and N-CoR exist in large protein complexes containing HDAC3. Embo J 19, 4342-50. Li, S., Xu, C. and Carthew, R. W. (2002). Phyllopod acts as an adaptor protein to link the sina ubiquitin ligase to the substrate protein tramtrack. Mol Cell Biol 22, 6854-65. Luger, K., Mader, A. W., Richmond, R. K., Sargent, D. F. and Rich- mond, T. J. (1997). Crystal structure of the nucleosome core particle at 2.8 A resolution. Nature 389, 251-60. Luo, J., Nikolaev, A. Y., Imai, S., Chen, D., Su, F., Shiloh, A., Guarente, L. and Gu, W. (2001). Negative control of p53 by Sir2alpha promotes cell survival under stress. Cell 107, 137-48. Makino, Y., Kanemaki, M., Kurokawa, Y., Koji, T. and Tamura, T. (1999). A rat RuvB-like protein, TIP49a, is a germ cell-enriched novel DNA helicase. J Biol Chem 274, 15329-35. Malave, T. M. and Dent, S. Y. (2006). Transcriptional repression by Tup1- Ssn6. Biochem Cell Biol 84, 437-43. Mannervik, M. and Levine, M. (1999). The Rpd3 histone deacetylase is required for segmentation of the Drosophila embryo. Proc Natl Acad Sci U S A 96, 6797-801. Mannervik, M., Nibu, Y., Zhang, H. and Levine, M. (1999). Transcrip- tional coregulators in development. Science 284, 606-9. Margueron, R., Trojer, P. and Reinberg, D. (2005). The key to develop- ment: interpreting the histone code? Curr Opin Genet Dev 15, 163-76. Matsuoka, S., Huang, M. and Elledge, S. J. (1998). Linkage of ATM to cell cycle regulation by the Chk2 protein kinase. Science 282, 1893-7. Matsuyama, A., Shimazu, T., Sumida, Y., Saito, A., Yoshimatsu, Y., Seigneurin-Berny, D., Osada, H., Komatsu, Y., Nishino, N., Khochbin, S. et al. (2002). In vivo destabilization of dynamic microtubules by HDAC6- mediated deacetylation. Embo J 21, 6820-31. Matsuzawa, S. I. and Reed, J. C. (2001). Siah-1, SIP, and Ebi collaborate in a novel pathway for beta-catenin degradation linked to p53 responses. Mol Cell 7, 915-26. Meneghini, M. D., Wu, M. and Madhani, H. D. (2003). Conserved histone variant H2A.Z protects euchromatin from the ectopic spread of silent hetero- chromatin. Cell 112, 725-36. Mizuguchi, G., Shen, X., Landry, J., Wu, W. H., Sen, S. and Wu, C. (2004). ATP-driven exchange of histone H2AZ variant catalyzed by SWR1 chromatin remodeling complex. Science 303, 343-8. Moran, E. and Jimenez, G. (2006). The tailless nuclear receptor acts as a dedicated repressor in the early Drosophila embryo. Mol Cell Biol 26, 3446- 54.

60

Morel, V., Lecourtois, M., Massiani, O., Maier, D., Preiss, A. and Schweisguth, F. (2001). Transcriptional repression by suppressor of hairless involves the binding of a hairless-dCtBP complex in Drosophila. Curr Biol 11, 789-92. Morisato, D. (2001). Spatzle regulates the shape of the Dorsal gradient in the Drosophila embryo. Development 128, 2309-19. Moussian, B. and Roth, S. (2005). Dorsoventral axis formation in the Dro- sophila embryo--shaping and transducing a morphogen gradient. Curr Biol 15, R887-99. Munshi, N., Agalioti, T., Lomvardas, S., Merika, M., Chen, G. and Thanos, D. (2001). Coordination of a transcriptional switch by HMGI(Y) acetylation. Science 293, 1133-6. Muratoglu, S., Georgieva, S., Papai, G., Scheer, E., Enunlu, I., Komonyi, O., Cserpan, I., Lebedeva, L., Nabirochkina, E., Udvardy, A. et al. (2003). Two different Drosophila ADA2 homologues are present in distinct GCN5 histone acetyltransferase-containing complexes. Mol Cell Biol 23, 306-21. Murr, R., Loizou, J. I., Yang, Y. G., Cuenin, C., Li, H., Wang, Z. Q. and Herceg, Z. (2006). Histone acetylation by Trrap-Tip60 modulates loading of repair proteins and repair of DNA double-strand breaks. Nat Cell Biol 8, 91- 9. Nasiadka, A., Grill, A. and Krause, H. M. (2000). Mechanisms regulating target gene selection by the homeodomain-containing protein Fushi tarazu. Development 127, 2965-76. Nasiadka, A. and Krause, H. M. (1999). Kinetic analysis of segmentation gene interactions in Drosophila embryos. Development 126, 1515-26. Nibu, Y. and Levine, M. S. (2001). CtBP-dependent activities of the short- range Giant repressor in the Drosophila embryo. Proc Natl Acad Sci U S A 98, 6204-8. Nibu, Y., Zhang, H., Bajor, E., Barolo, S., Small, S. and Levine, M. (1998a). dCtBP mediates transcriptional repression by Knirps, Kruppel and Snail in the Drosophila embryo. Embo J 17, 7009-20. Nibu, Y., Zhang, H. and Levine, M. (1998b). Interaction of short-range repressors with Drosophila CtBP in the embryo. Science 280, 101-4. Niessing, D., Rivera-Pomar, R., La Rosee, A., Hader, T., Schock, F., Purnell, B. A. and Jackle, H. (1997). A cascade of transcriptional control leading to axis determination in Drosophila. J Cell Physiol 173, 162-7. Nieto, M. A. (2002). The snail superfamily of zinc-finger transcription fac- tors. Nat Rev Mol Cell Biol 3, 155-66. Ochoa-Espinosa, A., Yucel, G., Kaplan, L., Pare, A., Pura, N., Oberstein, A., Papatsenko, D. and Small, S. (2005). The role of binding site cluster strength in Bicoid-dependent patterning in Drosophila. Proc Natl Acad Sci U S A 102, 4960-5. Ogryzko, V. (2000). Mammalian histone acetyltransferase complexes. Medicina (B Aires) 60 Suppl 2, 21-6. Ogura, T. and Wilkinson, A. J. (2001). AAA+ superfamily ATPases: common structure--diverse function. Genes Cells 6, 575-97.

61

Ohdate, H., Lim, C. R., Kokubo, T., Matsubara, K., Kimata, Y. and Kohno, K. (2003). Impairment of the DNA binding activity of the TATA- binding protein renders the transcriptional function of Rvb2p/Tih2p, the yeast RuvB-like protein, essential for cell growth. J Biol Chem 278, 14647- 56. Oren, M. (2003). Decision making by p53: life, death and cancer. Cell Death Differ 10, 431-42. Pal-Bhadra, M., Leibovitch, B. A., Gandhi, S. G., Rao, M., Bhadra, U., Birchler, J. A. and Elgin, S. C. (2004). Heterochromatic silencing and HP1 localization in Drosophila are dependent on the RNAi machinery. Science 303, 669-72. Pan, D. and Courey, A. J. (1992). The same dorsal binding site mediates both activation and repression in a context-dependent manner. Embo J 11, 1837-42. Pankotai, T., Komonyi, O., Bodai, L., Ujfaludi, Z., Muratoglu, S., Ci- urciu, A., Tora, L., Szabad, J. and Boros, I. (2005). The homologous Dro- sophila transcriptional adaptors ADA2a and ADA2b are both required for normal development but have different functions. Mol Cell Biol 25, 8215-27. Papoulas, O., Beek, S. J., Moseley, S. L., McCallum, C. M., Sarte, M., Shearn, A. and Tamkun, J. W. (1998). The Drosophila trithorax group proteins BRM, ASH1 and ASH2 are subunits of distinct protein complexes. Development 125, 3955-66. Paroush, Z., Finley, R. L., Jr., Kidd, T., Wainwright, S. M., Ingham, P. W., Brent, R. and Ish-Horowicz, D. (1994). Groucho is required for Dro- sophila neurogenesis, segmentation, and sex determination and interacts directly with hairy-related bHLH proteins. Cell 79, 805-15. Patel, J. H., Du, Y., Ard, P. G., Phillips, C., Carella, B., Chen, C. J., Ra- kowski, C., Chatterjee, C., Lieberman, P. M., Lane, W. S. et al. (2004). The c-MYC oncoprotein is a substrate of the acetyltransferases hGCN5/PCAF and TIP60. Mol Cell Biol 24, 10826-34. Perissi, V., Aggarwal, A., Glass, C. K., Rose, D. W. and Rosenfeld, M. G. (2004). A corepressor/coactivator exchange complex required for transcrip- tional activation by nuclear receptors and other regulated transcription fac- tors. Cell 116, 511-26. Peterson, C. L. (2002). Chromatin remodeling: nucleosomes bulging at the seams. Curr Biol 12, R245-7. Peterson, C. L. and Cote, J. (2004). Cellular machineries for chromosomal DNA repair. Genes Dev 18, 602-16. Petruk, S., Sedkov, Y., Smith, S., Tillib, S., Kraevski, V., Nakamura, T., Canaani, E., Croce, C. M. and Mazo, A. (2001). Trithorax and dCBP act- ing in a complex to maintain expression of a homeotic gene. Science 294, 1331-4. Plaster, N., Sonntag, C., Schilling, T. F. and Hammerschmidt, M. (2007). REREa/Atrophin-2 interacts with histone deacetylase and Fgf8 signaling to regulate multiple processes of zebrafish development. Dev Dyn 236, 1891- 904.

62

Poortinga, G., Watanabe, M. and Parkhurst, S. M. (1998). Drosophila CtBP: a Hairy-interacting protein required for embryonic segmentation and hairy-mediated transcriptional repression. Embo J 17, 2067-78. Puri, T., Wendler, P., Sigala, B., Saibil, H. and Tsaneva, I. R. (2007). Dodecameric structure and ATPase activity of the human TIP48/TIP49 complex. J Mol Biol 366, 179-92. Qi, D., Larsson, J. and Mannervik, M. (2004). Drosophila Ada2b is re- quired for viability and normal histone H3 acetylation. Mol Cell Biol 24, 8080-9. Qin, S. and Parthun, M. R. (2006). Recruitment of the type B histone ace- tyltransferase Hat1p to chromatin is linked to DNA double-strand breaks. Mol Cell Biol 26, 3649-58. Rao, Y., Pang, P., Ruan, W., Gunning, D. and Zipursky, S. L. (2000). brakeless is required for photoreceptor growth-cone targeting in Drosophila. Proc Natl Acad Sci U S A 97, 5966-71. Regulski, M., Harding, K., Kostriken, R., Karch, F., Levine, M. and McGinnis, W. (1985). Homeo box genes of the Antennapedia and bithorax complexes of Drosophila. Cell 43, 71-80. Ricci, A. R., Genereaux, J. and Brandl, C. J. (2002). Components of the SAGA histone acetyltransferase complex are required for repressed tran- scription of ARG1 in rich medium. Mol Cell Biol 22, 4033-42. Ringrose, L. and Paro, R. (2004). Epigenetic regulation of cellular memory by the Polycomb and Trithorax group proteins. Annu Rev Genet 38, 413-43. Rivera-Pomar, R., Lu, X., Perrimon, N., Taubert, H. and Jackle, H. (1995). Activation of posterior gap gene expression in the Drosophila blas- toderm. Nature 376, 253-6. Rogakou, E. P., Pilch, D. R., Orr, A. H., Ivanova, V. S. and Bonner, W. M. (1998). DNA double-stranded breaks induce histone H2AX phosphoryla- tion on serine 139. J Biol Chem 273, 5858-68. Rosenfeld, M. G., Lunyak, V. V. and Glass, C. K. (2006). Sensors and signals: a coactivator/corepressor/epigenetic code for integrating signal- dependent programs of transcriptional response. Genes Dev 20, 1405-28. Roth, S. (2003). The origin of dorsoventral polarity in Drosophila. Philos Trans R Soc Lond B Biol Sci 358, 1317-29; discussion 1329. Roth, S. Y., Denu, J. M. and Allis, C. D. (2001). Histone acetyltransferases. Annu Rev Biochem 70, 81-120. Rottbauer, W., Saurin, A. J., Lickert, H., Shen, X., Burns, C. G., Wo, Z. G., Kemler, R., Kingston, R., Wu, C. and Fishman, M. (2002). Reptin and pontin antagonistically regulate heart growth in zebrafish embryos. Cell 111, 661-72. Rudolph, T., Yonezawa, M., Lein, S., Heidrich, K., Kubicek, S., Schafer, C., Phalke, S., Walther, M., Schmidt, A., Jenuwein, T. et al. (2007). Het- erochromatin formation in Drosophila is initiated through active removal of H3K4 methylation by the LSD1 homolog SU(VAR)3-3. Mol Cell 26, 103-15. Rusch, J. and Levine, M. (1996). Threshold responses to the dorsal regula- tory gradient and the subdivision of primary tissue territories in the Droso- phila embryo. Curr Opin Genet Dev 6, 416-23.

63

Ryu, J. R. and Arnosti, D. N. (2003). Functional similarity of Knirps CtBP- dependent and CtBP-independent transcriptional repressor activities. Nucleic Acids Res 31, 4654-62. Sapountzi, V., Logan, I. R. and Robson, C. N. (2006). Cellular functions of TIP60. Int J Biochem Cell Biol 38, 1496-509. Saurin, A. J., Shao, Z., Erdjument-Bromage, H., Tempst, P. and King- ston, R. E. (2001). A Drosophila Polycomb group complex includes Zeste and dTAFII proteins. Nature 412, 655-60. Schreiber-Agus, N. and DePinho, R. A. (1998). Repression by the Mad(Mxi1)-Sin3 complex. Bioessays 20, 808-18. Schuettengruber, B., Chourrout, D., Vervoort, M., Leblanc, B. and Cavalli, G. (2007). Genome regulation by polycomb and trithorax proteins. Cell 128, 735-45. Seligson, D. B., Horvath, S., Shi, T., Yu, H., Tze, S., Grunstein, M. and Kurdistani, S. K. (2005). Global histone modification patterns predict risk of prostate cancer recurrence. Nature 435, 1262-6. Senti, K., Keleman, K., Eisenhaber, F. and Dickson, B. J. (2000). brake- less is required for lamina targeting of R1-R6 axons in the Drosophila visual system. Development 127, 2291-301. Shen, X., Mizuguchi, G., Hamiche, A. and Wu, C. (2000). A chromatin remodelling complex involved in transcription and DNA processing. Nature 406, 541-4. Shen, Y., Lee, G., Choe, Y., Zoltewicz, J. S. and Peterson, A. S. (2007). Functional architecture of atrophins. J Biol Chem 282, 5037-44. Shukla, A., Bajwa, P. and Bhaumik, S. R. (2006). SAGA-associated Sgf73p facilitates formation of the preinitiation complex assembly at the promoters either in a HAT-dependent or independent manner in vivo. Nu- cleic Acids Res 34, 6225-32. Silverstein, R. A. and Ekwall, K. (2005). Sin3: a flexible regulator of global gene expression and genome stability. Curr Genet 47, 1-17. Simon, J., Chiang, A. and Bender, W. (1992). Ten different Polycomb group genes are required for spatial control of the abdA and AbdB homeotic products. Development 114, 493-505. Small, S., Blair, A. and Levine, M. (1992). Regulation of even-skipped stripe 2 in the Drosophila embryo. Embo J 11, 4047-57. Smith, M. M. (2002). Histone variants and nucleosome deposition pathways. Mol Cell 9, 1158-60. Song, H., Hasson, P., Paroush, Z. and Courey, A. J. (2004). Groucho oligomerization is required for repression in vivo. Mol Cell Biol 24, 4341-50. St Johnston, D. and Nusslein-Volhard, C. (1992). The origin of pattern and polarity in the Drosophila embryo. Cell 68, 201-19. Stathopoulos, A. and Levine, M. (2002). Dorsal gradient networks in the Drosophila embryo. Dev Biol 246, 57-67. Stathopoulos, A., Van Drenth, M., Erives, A., Markstein, M. and Levine, M. (2002). Whole-genome analysis of dorsal-ventral patterning in the Dro- sophila embryo. Cell 111, 687-701.

64

Stewart, M. D., Li, J. and Wong, J. (2005). Relationship between histone H3 lysine 9 methylation, transcription repression, and heterochromatin pro- tein 1 recruitment. Mol Cell Biol 25, 2525-38. Stifani, S., Blaumueller, C. M., Redhead, N. J., Hill, R. E. and Arta- vanis-Tsakonas, S. (1992). Human homologs of a Drosophila Enhancer of split gene product define a novel family of nuclear proteins. Nat Genet 2, 343. Stockinger, E. J., Mao, Y., Regier, M. K., Triezenberg, S. J. and Thomashow, M. F. (2001). Transcriptional adaptor and histone acetyltrans- ferase proteins in Arabidopsis and their interactions with CBF1, a transcrip- tional activator involved in cold-regulated gene expression. Nucleic Acids Res 29, 1524-33. Strahl, B. D. and Allis, C. D. (2000). The language of covalent histone modifications. Nature 403, 41-5. Strahl, B. D., Grant, P. A., Briggs, S. D., Sun, Z. W., Bone, J. R., Cald- well, J. A., Mollah, S., Cook, R. G., Shabanowitz, J., Hunt, D. F. et al. (2002). Set2 is a nucleosomal histone H3-selective methyltransferase that mediates transcriptional repression. Mol Cell Biol 22, 1298-306. Strunk, B., Struffi, P., Wright, K., Pabst, B., Thomas, J., Qin, L. and Arnosti, D. N. (2001). Role of CtBP in transcriptional repression by the Drosophila giant protein. Dev Biol 239, 229-40. Sundqvist, A., Sollerbrant, K. and Svensson, C. (1998). The carboxy- terminal region of adenovirus E1A activates transcription through targeting of a C-terminal binding protein-histone deacetylase complex. FEBS Lett 429, 183-8. Swaminathan, J., Baxter, E. M. and Corces, V. G. (2005). The role of histone H2Av variant replacement and histone H4 acetylation in the estab- lishment of Drosophila heterochromatin. Genes Dev 19, 65-76. Tamburini, B. A. and Tyler, J. K. (2005). Localized histone acetylation and deacetylation triggered by the homologous recombination pathway of double-strand DNA repair. Mol Cell Biol 25, 4903-13. Taubert, S., Gorrini, C., Frank, S. R., Parisi, T., Fuchs, M., Chan, H. M., Livingston, D. M. and Amati, B. (2004). E2F-dependent histone acetyla- tion and recruitment of the Tip60 acetyltransferase complex to chromatin in late G1. Mol Cell Biol 24, 4546-56. Teng, Y., Yu, Y. and Waters, R. (2002). The Saccharomyces cerevisiae histone acetyltransferase Gcn5 has a role in the photoreactivation and nu- cleotide excision repair of UV-induced cyclobutane pyrimidine dimers in the MFA2 gene. J Mol Biol 316, 489-99. Towb, P., Galindo, R. L. and Wasserman, S. A. (1998). Recruitment of Tube and Pelle to signaling sites at the surface of the Drosophila embryo. Development 125, 2443-50. Tsai, C. C., Kao, H. Y., Yao, T. P., McKeown, M. and Evans, R. M. (1999). SMRTER, a Drosophila nuclear receptor coregulator, reveals that EcR-mediated repression is critical for development. Mol Cell 4, 175-86.

65

Tsuda, L., Kaido, M., Lim, Y. M., Kato, K., Aigaki, T. and Hayashi, S. (2006). An NRSF/REST-like repressor downstream of Ebi/SMRTER/Su(H) regulates eye development in Drosophila. Embo J 25, 3191-202. Tsuda, L., Nagaraj, R., Zipursky, S. L. and Banerjee, U. (2002). An EGFR/Ebi/Sno pathway promotes delta expression by inactivating Su(H)/SMRTER repression during inductive notch signaling. Cell 110, 625- 37. Utley, R. T. and Cote, J. (2003). The MYST family of histone acetyltrans- ferases. Curr Top Microbiol Immunol 274, 203-36. Utley, R. T., Ikeda, K., Grant, P. A., Cote, J., Steger, D. J., Eberharter, A., John, S. and Workman, J. L. (1998). Transcriptional activators direct histone acetyltransferase complexes to nucleosomes. Nature 394, 498-502. Valentine, S. A., Chen, G., Shandala, T., Fernandez, J., Mische, S., Saint, R. and Courey, A. J. (1998). Dorsal-mediated repression requires the for- mation of a multiprotein repression complex at the ventral silencer. Mol Cell Biol 18, 6584-94. van Attikum, H. and Gasser, S. M. (2005). ATP-dependent chromatin remodeling and DNA double-strand break repair. Cell Cycle 4, 1011-4. Vega, R. B., Matsuda, K., Oh, J., Barbosa, A. C., Yang, X., Meadows, E., McAnally, J., Pomajzl, C., Shelton, J. M., Richardson, J. A. et al. (2004). Histone deacetylase 4 controls chondrocyte hypertrophy during skeletogene- sis. Cell 119, 555-66. Verdel, A. and Moazed, D. (2005). RNAi-directed assembly of hetero- chromatin in fission yeast. FEBS Lett 579, 5872-8. Verdin, E., Dequiedt, F. and Kasler, H. G. (2003). Class II histone deace- tylases: versatile regulators. Trends Genet 19, 286-93. Vlachonasios, K. E., Thomashow, M. F. and Triezenberg, S. J. (2003). Disruption mutations of ADA2b and GCN5 transcriptional adaptor genes dramatically affect Arabidopsis growth, development, and gene expression. Plant Cell 15, 626-38. Wahi, M., Komachi, K. and Johnson, A. D. (1998). Gene regulation by the yeast Ssn6-Tup1 corepressor. Cold Spring Harb Symp Quant Biol 63, 447- 57. Wang, H., Wang, L., Erdjument-Bromage, H., Vidal, M., Tempst, P., Jones, R. S. and Zhang, Y. (2004). Role of histone H2A ubiquitination in Polycomb silencing. Nature 431, 873-8. Wang, L., Rajan, H., Pitman, J. L., McKeown, M. and Tsai, C. C. (2006). Histone deacetylase-associating Atrophin proteins are nuclear receptor corepressors. Genes Dev 20, 525-30. Wedeen, C., Harding, K. and Levine, M. (1986). Spatial regulation of Antennapedia and bithorax gene expression by the Polycomb locus in Dro- sophila. Cell 44, 739-48. Wehn, A. and Campbell, G. (2006). Genetic interactions among scribbler, Atrophin and groucho in Drosophila uncover links in transcriptional repres- sion. Genetics 173, 849-61. Wiren, M., Silverstein, R. A., Sinha, I., Walfridsson, J., Lee, H. M., Laurenson, P., Pillus, L., Robyr, D., Grunstein, M. and Ekwall, K.

66

(2005). Genomewide analysis of nucleosome density histone acetylation and HDAC function in fission yeast. Embo J 24, 2906-18. Wood, M. A., McMahon, S. B. and Cole, M. D. (2000). An AT- Pase/helicase complex is an essential cofactor for oncogenic transformation by c-Myc. Mol Cell 5, 321-30. Wurtele, H. and Verreault, A. (2006). Histone post-translational modifica- tions and the response to DNA double-strand breaks. Curr Opin Cell Biol 18, 137-44. Xiao, H., Chung, J., Kao, H. Y. and Yang, Y. C. (2003). Tip60 is a co- repressor for STAT3. J Biol Chem 278, 11197-204. Xu, L., Lavinsky, R. M., Dasen, J. S., Flynn, S. E., McInerney, E. M., Mullen, T. M., Heinzel, T., Szeto, D., Korzus, E., Kurokawa, R. et al. (1998). Signal-specific co-activator domain requirements for Pit-1 activation. Nature 395, 301-6. Xu, W., Edmondson, D. G., Evrard, Y. A., Wakamiya, M., Behringer, R. R. and Roth, S. Y. (2000). Loss of Gcn5l2 leads to increased apoptosis and mesodermal defects during mouse development. Nat Genet 26, 229-32. Yamamoto, T. and Horikoshi, M. (1997). Novel substrate specificity of the histone acetyltransferase activity of HIV-1-Tat interactive protein Tip60. J Biol Chem 272, 30595-8. Yamauchi, T., Yamauchi, J., Kuwata, T., Tamura, T., Yamashita, T., Bae, N., Westphal, H., Ozato, K. and Nakatani, Y. (2000). Distinct but overlapping roles of histone acetylase PCAF and of the closely related PCAF-B/GCN5 in mouse embryogenesis. Proc Natl Acad Sci U S A 97, 11303-6. Yang, J. and Steward, R. (1997). A multimeric complex and the nuclear targeting of the Drosophila Rel protein Dorsal. Proc Natl Acad Sci U S A 94, 14524-9. Yang, P., Shaver, S. A., Hilliker, A. J. and Sokolowski, M. B. (2000). Abnormal turning behavior in Drosophila larvae. Identification and molecu- lar analysis of scribbler (sbb). Genetics 155, 1161-74. Yoon, H. G., Chan, D. W., Huang, Z. Q., Li, J., Fondell, J. D., Qin, J. and Wong, J. (2003). Purification and functional characterization of the human N-CoR complex: the roles of HDAC3, TBL1 and TBLR1. Embo J 22, 1336-46. Yoon, H. G., Choi, Y., Cole, P. A. and Wong, J. (2005). Reading and func- tion of a histone code involved in targeting corepressor complexes for re- pression. Mol Cell Biol 25, 324-35. You, A., Tong, J. K., Grozinger, C. M. and Schreiber, S. L. (2001). CoREST is an integral component of the CoREST- human histone deacety- lase complex. Proc Natl Acad Sci U S A 98, 1454-8. Zhang, C. L., McKinsey, T. A., Chang, S., Antos, C. L., Hill, J. A. and Olson, E. N. (2002a). Class II histone deacetylases act as signal-responsive repressors of cardiac hypertrophy. Cell 110, 479-88. Zhang, C. L., Zou, Y., Yu, R. T., Gage, F. H. and Evans, R. M. (2006a). Nuclear receptor TLX prevents retinal dystrophy and recruits the corepressor atrophin1. Genes Dev 20, 1308-20.

67

Zhang, F., Thomas, L. R., Oltz, E. M. and Aune, T. M. (2006b). Control of thymocyte development and recombination-activating gene expression by the zinc finger protein Zfp608. Nat Immunol 7, 1309-16. Zhang, H. and Levine, M. (1999). Groucho and dCtBP mediate separate pathways of transcriptional repression in the Drosophila embryo. Proc Natl Acad Sci U S A 96, 535-40. Zhang, S., Xu, L., Lee, J. and Xu, T. (2002b). Drosophila atrophin ho- molog functions as a transcriptional corepressor in multiple developmental processes. Cell 108, 45-56. Zhang, W., Deng, H., Bao, X., Lerach, S., Girton, J., Johansen, J. and Johansen, K. M. (2006c). The JIL-1 histone H3S10 kinase regulates di- methyl H3K9 modifications and heterochromatic spreading in Drosophila. Development 133, 229-35. Zhang, Y., Ng, H. H., Erdjument-Bromage, H., Tempst, P., Bird, A. and Reinberg, D. (1999). Analysis of the NuRD subunits reveals a histone deacetylase core complex and a connection with DNA methylation. Genes Dev 13, 1924-35. Zoltewicz, J. S., Stewart, N. J., Leung, R. and Peterson, A. S. (2004). Atrophin 2 recruits histone deacetylase and is required for the function of multiple signaling centers during mouse embryogenesis. Development 131, 3-14.

68 This document was created with Win2PDF available at http://www.win2pdf.com. The unregistered version of Win2PDF is for evaluation or non-commercial use only. This page will not be added after purchasing Win2PDF.