Characterization of the Schizosaccharomyces Pombe Hat1 Complex: the Role of H4 in Telomeric Silencing

Author: Kevin Tong

Persistent link: http://hdl.handle.net/2345/2222

This work is posted on eScholarship@BC, Boston College University Libraries.

Boston College Electronic Thesis or Dissertation, 2009

Copyright is held by the author, with all rights reserved, unless otherwise noted.

Boston College

The Graduate School of Arts and Sciences

Department of Biology

CHARACTERIZATION OF THE SCHIZOSACCHAROMYCES POMBE HAT1

COMPLEX: THE ROLE OF ACETYLATION IN TELOMERIC SILENCING

a dissertation by

KEVIN TONG

Submitted in partial fulfillment of the requirements

for the degree of

Doctor of Philosophy

August, 2009

 Copyright by KEVIN TONG

2009

Abstract

CHARACTERIZATION OF THE SCHIZOSACCHAROMYCES POMBE HAT1

COMPLEX: THE ROLE OF HISTONE H4 ACETYLATION IN TELOMERIC SILENCING

Author: Kevin Tong

Dissertation Advisor: Anthony T. Annunziato

The Hat1 complex was characterized in S. pombe . Through tandem affinity

purification and mass spectrometry, it was determined that Hat1 is associated with

Mis16 (an orthologue of HAT2). Unlike HAT2 in S. cerevisiae , we confirm mis16 to be an essential in S. pombe . As expected, the S. pombe Hat1 complex was found to acetylate lysines 5 and 12 of histone H4. In contrast to budding yeast, deletion of hat1

alone resulted in the loss of telomeric silencing without concomitant mutations of the H3

N-terminal domain. Deletion of hat1 caused an increase of H4 acetylation at telomeres.

Additionally, the hyperacetylation of also results in the loss of telomeric

silencing. Loss of Hat1 did not affect silencing at the inner most repeat (imr) or outer

repeat (otr) regions of the centromere, but did appear to increase silencing at the central

core region (cnt) of the centromere. The experiments described herein demonstrate

Hat1 to be essential for the establishment of proper telomeric silencing in fission yeast,

and suggest that the timely acetylation of H4 during chromatin assembly is a unique

factor in generating the correct epigenetic state at telomeres in S. pombe . Additionally, Hat1 and its acetylation of new H4 may have entirely different roles during telomeric

silencing than during silencing at the centromeric central core.

Our studies in HeLa cells demonstrated that transcription is involved in the exchange of H2A/H2B in acetylated chromatin regions. The finding that cytosolic H2A can be acetylated at lysine 5 is the first demonstration that cytosolic H2A can be specifically modified in vivo . Our results support a model in which H2A/H2B exchange during transcription is mediated by the NAP1 chaperone.

TABLE OF CONTENTS PAGE

Acknowledgements iii

List of Tables iv

List of Figures v

Abbreviations vii

Introduction 1

Chapter One: Characterization of the Hat1 Complex in S. pombe

Summary 25 Materials and Methods 26 Results 35

Chapter Two: The role of Hat1 and acetylation in S. pombe telomeric silencing

Summary 53 Materials and Methods 54 Results 62

Chapter Three: The deposition of H2A and H2B in human cells

Summary 85 Materials and Methods 86 Results 93

Discussion 109

i References 120

ii ACKNOWLEDGEMENTS

I would first like to thank my thesis advisor, Dr. Anthony T. Annunziato, for his support, encouragement, patience, and mentorship during my doctoral work. His advice and guidance has truly been invaluable during my time at Boston College.

I would also like to thank the members of my thesis committee. First, Dr. Charles

Hoffman for his advice, guidance, and generosity with reagents and equipment.

Second, Dr. Thomas C. Chiles for his generosity with reagents, equipment, and advice.

Lastly, Dr. Junona Moroianu and Dr. John Wing for their constant support and encouragement.

Additionally, I am grateful for the generous gifts of strains from Dr. Robin Allshire and Dr. Mitsuhiro Yanagida. Many of my experiments would not be possible without the

S. pombe strains that they so generously provided.

I would also like to thank past and present members of the Chiles Lab and the

Hoffman Lab; especially; Dr. Derek Blair, Fay Dufort, Dr. F. Douglas Ivey, and Dr. Lili

Wang for their advice and technical expertise.

Additionally, I would like to thank the members of the Annunziato Lab, past and present; especially Laura Benson, Yong Li Gu and Inna Grishkan for their tutelage, support, and friendship.

Lastly, I would like to thank Cheryl, my parents, and my sister for their love and support. Without them, I would not be the person I am today.

iii LIST OF TABLES

Table Page

I S. pombe strain list 40

iv LIST OF FIGURES

Chapter One

Figure Page

1 Generation of a S. pombe Hat1-TAP strain 41

2 Hat2/Mis16 copurifies with Hat1 during tandem affinity purification 43

3 Hat1-TAP and Mis16-myc are associated in a complex 45

4 Mis16 is essential 47

5 Acetylation of H4 N-terminal peptides by S. pombe Hat1 in vitro 49

6 Acetylation of recombinant H4 with S. pombe Hat1 in vitro 51

Chapter Two

7 hat1 deletion results in a loss of telomeric silencing in S. pombe 69

8 Hat1 expressed from a plasmid restores telomeric silencing 71

9 hat1 deletion does not result in the loss of centromeric silencing in 73 S. pombe

10 The increase in relative enrichment of acH4 K5/K12 at the 75 subtelomeric ura DNA sequence in the hat1 deletion strain over the wildtype strain

11 The increase in relative enrichment of acH4 K8/K16 at the 77 subtelomeric ura DNA sequence in the hat1 deletion strain over the wildtype strain

v

12 Hyperacetylation of histones results in the loss of telomeric silencing 79

13 Treatment of synchronized S. pombe cells with sodium butyrate 81 results in a cell cycle delay upon release

14 The accumulation of S. pombe cells with greater than 2C DNA 83 content following extended exposure to histone deacetylase inhibitors

Chapter Three

15 Not all newly synthesized H2A/H2B is deposited together with 97 new H3/H4

16 Deposition of new H2A and H2B onto acetylated chromatin in 99 the absence of DNA replication

17 Thymidine incorporation in the presence of hydroxyurea 101

18 Deposition of H2A and H2B onto non-replicating chromatin is 103 inhibited by 5,6-Dichloro-1-β-D-ribofuranosylbenzimidazole (DRB), an inhibitor RNA synthesis

19 Uridine incorporation in the presence of DRB 105

20 Acetylation of cytosolic H2A 107

vi ABBREVIATIONS

5-FOA, 5-fluoroorotic acid; ac, acetylated; ASF1, anti-silencing factor 1; bp, base pairs; CAF-1, chromatin assembly factor 1; ChIP, chromatin immunoprecipitation; cnt, central core; EMM, Edinburgh minimal medium; EtBr, ethidium bromide; GNAT, GCN5- related N-acetyltransferase; HAT, histone acetyltransferase; Hbo1, histone acetyltransferase binding to ORC; HDAC, histone deacetylase; HU, hydroxyurea; MMS, methyl methane sulfonate; imr, innermost repeat; MYST, MOZ-Ybf2/Sas3-Sas2-Tip60;

ORC, origin recognition complex; otr, outer repeat; PBS, phosphate buffered saline;

PCNA, proliferating cell nuclear antigen; Pot1, protection of telomeres 1; Rif1, Rap1 interacting factor; Rtt106, regulator of Ty1 transposition protein; SPB, spindle pole body;

TAP, tandem affinity purification; Tip60, TAT-interactive protein 60; TRF, telomere repeat factor; TSA, trichostatin A; YEA, yeast extract agar; YEL, yeast extract liquid;

vii

Introduction

Chromatin: History and structure

In eukaryotic cells, genetic information in the form of DNA is contained within the nucleus. The contains approximately 3 x 10 9 base pairs (bp), which must be compacted into a nucleus measuring 10 -5 meters in diameter. To accomplish this compaction, genetic information is organized into a highly conserved structural polymer called chromatin. Chromatin, a supercoiled nucleoprotein complex, consists of a fundamental repeating unit known as the nucleosome. The first evidence for the nucleosomal organization of chromatin came through the use of nuclease digestion studies by Williamson in 1970 and Hewish and Burgoyne in 1973. In these studies, it was observed that chromosomal DNA degraded into fragments of discrete size separated by multiples of 180-200 bp. Shortly thereafter, two groups (Woodcock (1973) and Olins and Olins (1974)) made the first visualizations of nucleosomes through the use of electron microscopy (van Holde 1988). In short, both groups noted that an extended chromatin fiber resembled "beads on a string." Chemical cross-linking experiments by Kornberg in 1974 suggested that core histones are present in a precise stoichiometry (Kornberg and Thomas 1974; Wolffe 1995). This observation coupled with those observed in previous studies led Kornberg to present a model for the structure of the nucleosome in 1974. In a seminal paper, Kornberg proposed that the repeating unit of chromatin or nucleosome consisted of 200 bp of DNA wrapped around a tetramer of

(H3) 2(H4) 2 and two each of H2A and H2B (Kornberg 1974; van Holde 1988). This model

1

was later refined. Experiments performed using micrococcal nuclease digestion produced particles with 200 bp of DNA. Further digestion, however, resulted in a stable

"core" particle consisting of the histone octamer and approximately 140 bp of DNA that is resistant to micrococcal nuclease digestion (Sollner-Webb and Felsenfeld 1975). The current model today presents the nucleosome core as a structure consisting of a highly conserved histone octamer containing two molecules each of H2A, H2B, H3, and H4 around the exterior of which is wrapped 146 bp, or two turns, of DNA (Wolffe 1995;

Verreault et al. 1996). Between each core particle is a variable length of DNA referred to as "linker" DNA. This linker DNA is trimmed during micrococcal nuclease digestion resulting in the core particle (Wolffe 1995). In 1977, Noll and Kornberg demonstrated that the digestion of linker DNA resulted in the concomitant loss of linker histone from the nucleosome (Noll and Kornberg 1977). The identity of this linker histone was identified to be histone H1 in a mobility shift assay (Varshavsky et al. 1976). A later study by Simpson in 1978 demonstrated that histone H1 significantly increased the stability of DNA in the nucleosome core particle (Simpson 1978). These experiments suggested that histone H1 not only played a role in the organization of linker DNA, but also contributed to the organization of the DNA around the core particle (Wolffe 1995).

Although chromatin research since the identification of the nucleosome in 1973 has demonstrated that the dynamic higher order structuring of nucleosomes is required for discrete levels of chromatin organization, the mechanism for how nucleosomal arrays containing histone H1 repeatedly twist and fold into a progressively more compact

2

filament and higher order chromatin structures remains largely unknown (Jenuwein and

Allis 2001). What is apparent, however, is that in addition to their role in the formation of higher order chromatin structures, nucleosomes and their presence on DNA contribute

to the regulation of basal promoter activity and gene transcription (Ushinsky et al. 1997).

Nucleosome core particle

The nucleosome core particle consists of four histones: H2A (14 kDa), H2B (14 kDa), H3 (15 kDa), and H4 (11 kDa). These core histones are low molecular weight basic proteins. Although the core histones share little , they share a common tertiary structure known as the "histone fold" (Arents and Moudrianakis 1995).

This histone fold domain, around which nucleosomal DNA is wrapped, consists of approximately 75% of the core histone protein mass (Arents et al. 1991; Arents and

Moudrianakis 1993; Luger et al. 1997). The histone fold is a helix-strand-helix motif consisting of an 11-residue α-helix, followed by a short loop and β-strand, a 27-residue

α-helix, another short loop and β-strand, and ending with an 11-residue α-helix (Arents

and Moudrianakis 1995). The association of histone fold domains in a head-to-tail

conformation forms the "handshake" motif, which interlock to assemble heterodimers of

H2A/H2B and H3/H4 (Arents and Moudrianakis 1993). These heterodimers comprise

the nucleosome core. The joining of two dimers of H2A/H2B and a tetramer of

(H3) 2(H4) 2 results in a tripartite protein complex around the exterior of which is wrapped

the nucleosome core DNA (Eickbush and Moudrianakis 1978).

3

The remaining mass of core histones (25-30%) is comprised of evolutionarily

conserved "tail" domains that are, for the most part, structurally undefined (Zheng and

Hayes 2003). These domains, which are defined by their sensitivity to proteases, are

found at the N-terminus of each core histone protein and the C-terminus of histone H2A

(van Holde 1988). X-ray crystallography studies of the nucleosome core particle have

demonstrated that the tail domains track minor grooves of DNA to the nucleosome

exterior (Luger et al. 1997; Harp et al. 2000; Suto et al. 2000; White et al. 2001; Davey et al. 2002). This conformation causes the histone tails to be susceptible to enzymes such as methylases, acetyltransferases, kinases, and to post-translational modification.

Within condensed chromatin structures, these tail domains are also capable of mediumting interactions between nucleosomes (Zheng and Hayes 2003). However, the conformations and interactions of the histone tails are largely undefined in these structures due to the ability of the tail domains to assume various conformations in the crystals of nucleosome cores (Luger et al. 1997; Harp et al. 2000; Suto et al. 2000;

White et al. 2001; Davey et al. 2002).

Histone N-terminal tail

The core histone N-termini are highly conserved amongst organisms. With the

exception of a single conservative substitution, the amino acid sequence of both H3 and

H4 N-termini in humans and yeast are indistinguishable (Smith and Andresson 1983).

Unlike the C-terminal domains of core histones, which have similar conformations and

are essential for nucleosome assembly, the N-termini are less structured and do not play

4

a significant role in maintaining the integrity of individual nucleosomes (Wolffe 1999).

For example, removal of the N-terminal histone tails using trypsin does not render the nucleosomes unstable (Whitlock and Simpson 1977; Ausio et al. 1989). Nevertheless, their highly conserved nature suggests that these N-termini possess a complex biological purpose and do not function merely by nonspecific interactions with the DNA component of nucleosomes (Hansen et al. 1998). One proposed function of histone N- terminal tails involves the modulation of chromatin architecture (Cheung et al. 2000).

Evidence suggests that these tails are the targets of ATP-dependent chromatin remodeling factors such as SWI/SNF and NURF (Georgel et al. 1997; Lee et al. 1999;

Krebs et al. 2000). Studies have also demonstrated the N-terminal tails to be susceptible to a number of post-translational modifications, which can modulate histone-

DNA contacts and regulate a variety of DNA related processes due to their reversible nature (Cheung et al. 2000).

Post-translational modifications

Although core histones can have specialized forms known as histone variants, further variation can be imparted through covalent modifications of the histone tail domains. Because the enzymes responsible for these modifications have specific substrate specificities, the possibility of a "histone code" exists which may expand the information capacity of the genetic code (Strahl and Allis 2000; Turner 2000; Jenuwein and Allis 2001). The "histone code" hypothesis predicts that combinations of these post- translational modifications, which include acetylation, methylation, and ubiquitination of

5

lysine residues, phosphorylation of serine and threonine residues, and methylation of

arginine residues, establish global and local patterns of modifications that can be read

by other proteins and translated into biological functions (Fischle et al. 2003). Because histone modifications can be either transient or more permanent in nature, their effects on genomic function can also be short-term or long-term, respectively. Additionally, it is possible that these patterns of histone modifications also provide a means for epigenetic inheritance (heritable variations that do not involve a change in the DNA sequence)

(Turner 2000).

Histone acetylation

Of the histone post-translational modifications mentioned thus far, acetylation is

the most frequent to occur and most widely studied. Originally discovered in 1964 by

Allfrey and colleagues, the reversible acetylation of histone proteins occurs on the ε-

amino groups of lysine residues located in the N-terminal tails of core histones (Allfrey et

al. 1964; Kuo and Allis 1998). The introduction of an acetyl group by histone

acetyltransferases neutralizes the positive charge on the lysine residue and increases

the hydrophobicity of histones. During the reverse reaction, histone deacetylases

remove the acetyl group and re-establish the positive charge (Kuo and Allis 1998).

As DNA is packaged into nucleosomes, the basal level of transcription within a

cell is diminished (Han and Grunstein 1988). To counteract the transcriptional

repression resulting from DNA packaging and allow changes in gene transcription,

nucleosome properties must be altered by processes such as histone acetylation

6

(Shahbazian and Grunstein 2007). One proposal for how histone acetylation may

promote transcription is that the open chromatin structure that results from the neutralization of the positive charge on histone tails facilitates the binding of transcription factors to DNA (Lee et al. 1993; Vettese-Dadey et al. 1996; Nightingale et al. 1998). As

evidence, a study has shown that the acetylation of lysine 56 of histone H3 at the DNA

entry-exit point of the nucleosome may regulate transcription by disrupting histone-DNA

electrostatic interactions, which allows for the recruitment of SWI/SNF remodeling

complexes and facilitates transcription factor binding (Xu et al. 2005). A study has also

demonstrated that histone acetylation assists RNA polymerase progression during

transcription (Marushige 1976). Together, these studies suggest that histone acetylation

promotes transcription by opening chromatin structure to permit transcription factor

access and facilitating RNA polymerase progression (Shahbazian and Grunstein 2007).

Prior to their deposition onto DNA, newly synthesized histones are acetylated.

This acetylation, however, is transient. To ensure proper chromatin maturation following

histone deposition, the acetylation pattern is quickly removed (within 30-60 minutes of

histone deposition) (Ruiz-Carrillo et al. 1975; Jackson et al. 1976; Annunziato and Seale

1983). Presumably, this early acetylation pattern functions specifically during histone

deposition. The deletion of this acetylation pattern following deposition suggests that chromatin functions that take place after deposition would be deleteriously affected by the persistent acetylation of new histones (Shahbazian and Grunstein 2007). This deposition-related acetylation pattern of newly synthesized histone H4 is highly

7

conserved. In many organisms, newly synthesized H4 is acetylated at lysines 5 and 12.

In contrast, the acetylation pattern of newly synthesized H3 is more variable from one

species to the next: in Tetrahymena , K9 and K14 are acetylated; in Drosophila

melanogaster , K14 and K23 are acetylated; and in humans there appears to be no

specific deposition-related acetylation pattern for H3 (Sobel et al. 1995). The Hat1

complex appears to be responsible for the acetylation of new H4. Although no HAT has

currently been identified that acetylates new H3, the HatB3.1 complex, which consists of

Gcn5 and Ada3, is a possible candidate due to its cytoplasmic localization and in vitro

ability to acetylate free, but not nucleosomal H3 (Sklenar and Parthun 2004).

Histone variants

Although histones are slow to evolve and are therefore highly conserved, there

are non-allelic histone variants with considerable differences in primary sequence

(Kamakaka and Biggins 2005). Unlike canonical histones whose expression is restricted

to S phase of the cell cycle, histone variants are expressed at a low constitutive level

throughout the cell cycle (Smith 2002). The incorporation of these histone variants,

many of which have unique biological characteristics, into nucleosomes can lead to

alternative chromatin states. In terms of structure, histone variants are classified into

two families based on the degree of dissimilarity in amino acid sequence from the

canonical histone isoforms. The homomorphous family, which include H2A.1, H2A.2,

H3.1, H3.2 and H3.3, have little amino acid variation from canonical histones. By

comparison, the heteromorphous family, which includes H2A.X, H2A.Z, macroH2A

8

(mH2A), H2A Barr body-deficient (H2A.Bbd) and centromeric protein A (CENP-A),

contain significant differences from canonical histones (West and Bonner 1980; Ausio et

al. 2001). Some histone variants are known as replacement histones/variants due to

their ability to exchange with pre-existing histones (Kamakaka and Biggins 2005). H3.3 is of particular interest in relation to the nucleosome assembly pathway as it functions as the replacement variant for canonical H3 (also called H3.1) during replication-

independent nucleosome assembly (Tagami et al. 2004).

Replication-coupled nucleosome assembly

Replication-coupled nucleosome assembly occurs during S phase in the cell

cycle and is closely associated with the movement of the replication fork (Lucchini and

Sogo 1995). In order to propagate epigenetic information to daughter cells and maintain

genome integrity, two coordinated processes must occur: the segregation of parental

histones to new DNA and the de novo assembly of nucleosomes (Annunziato and

Hansen 2000). Nucleosomes assembled de novo contain a mix of parental and newly

synthesized histones. While nearly all new H3 and H4 are deposited onto newly

replicated DNA, nascent H2A, H2B and H1 are assembled onto non-replicating

chromatin in a process of histone exchange (Jackson and Chalkley 1981; Jackson and

Chalkley 1981; Jackson et al. 1981; Annunziato et al. 1982). This process of de novo

nucleosome assembly occurs in a stepwise manner. First, newly synthesized H3 and

H4 (acetylated in a deposition related pattern) are deposited onto nascent DNA (Smith

and Stillman 1991). Next, a pair of H2A/H2B dimers is deposited (Jackson 1990).

9

Finally, H1 is deposited (Worcel et al. 1978). Following histone deposition, H3 and H4 are deacetylated to complete the process of chromatin maturation (Annunziato 1990;

Annunziato 1995; Wolffe 1999).

The process of replication-coupled nucleosome assembly is essential for the inheritance of epigenetic information during DNA replication and repair (Henikoff et al.

2004; Groth et al. 2007). A protein that is important for the de novo deposition of new histones at the replication fork is chromatin assembly factor 1 (CAF-1) (Smith and

Stillman 1989; Verreault 2000). In most species, CAF-1 consists of three subunits of varying size. In humans, these subunits are: p150, p60, and p48 (Smith and Stillman

1989; Smith and Stillman 1991). This complex was originally identified in human cells based on its ability to assemble nucleosomes de novo during in vitro SV40 DNA replication (Stillman 1986; Smith and Stillman 1989). During replication-coupled nucleosome assembly, CAF-1 binds with newly synthesized H3 (H3.1) and H4 and deposits them onto replicating DNA (Kaufman et al. 1995). This ability of CAF-1 is dependent on its interaction with the proliferating cell nuclear antigen (PCNA), a DNA polymerase sliding clamp (Shibahara and Stillman 1999; Zhang et al. 2000; Krawitz et al. 2002). In humans, the p150 subunit binds directly to PCNA, while the p48 subunit binds directly to histones (Shibahara and Stillman 1999; Moggs et al. 2000; Loyola and

Almouzni 2004). The association of CAF-1 with PCNA allows for the recruitment of

CAF-1 to sites of DNA replication and repair. Studies have also demonstrated an interaction between CAF-1 and two other H3/H4 histone chaperones: anti-silencing

10

factor 1 (Asf1) and regulator of Ty1 transposition protein 106 (Rtt106) (Tyler et al. 2001;

Huang et al. 2005). Both of these histone chaperones assist CAF-1 mediumted histone deposition, but the manner in which CAF-1, Asf1, and Rtt106 are coordinated to result in the rapid deposition of histones onto nascent DNA has yet to be elucidated (Li et al.

2008).

Replication-independent chromatin assembly

Besides the main replication-dependent varieties of histone H3 in human cells, numerous organisms also express a H3 variant known as H3.3. In contrast to H3.1 and

H3.2, the synthesis of new H3.3 occurs throughout the cell cycle and is not limited to S phase (Ahmad and Henikoff 2002). Despite the high degree of similarity in amino acid sequence between H3.3 and H3.1/H3.2, there are differences in the type of histone chaperones that these H3 variants interact with in proliferating cells. While H3.1 and

H3.2 associate predominantly with CAF-1, H3.3 interacts with HIRA (Tagami et al.

2004). The predominant difference between H3.1 and H3.3 reside in three amino acids of the histone fold helix 2. This amino acid sequence (SAVM) in H3.1 targets H3.1 to the replication-coupled assembly pathway. In contrast, a mutated form of this sequence

(AAIG) in H3.3 results in the deposition of this histone variant during replication- independent nucleosome assembly (Ahmad and Henikoff 2002; Rocha and Verreault

2008). In addition to binding H3.1/H4 dimers, Asf1 also binds to H3.3/H4 dimers, consistent with its role in replication-dependent and replication-independent nucleosome assembly (Tagami et al. 2004). During replication-independent nucleosome assembly,

11

H3.3 is primarily deposited into actively transcribed (Mito et al. 2005; Schwartz

and Ahmad 2005; Wirbelauer et al. 2005). It is possible that as Asf1 moves with the

transcriptional machinery, it assists the process of assembly by closing openings left in

the nucleosome array as RNA polymerase II passes (Schwabish and Struhl 2006).

Histone acetyltransferases: Type A vs. Type B

Histone acetyltransferases (HATs) are enzymes capable of removing acetyl

groups from acetyl coenzyme A and adding them to conserved lysine residues on

histones (Roth et al. 2001). HATs are generally grouped into two different subtypes

based on their supposed subcellular location and function. The first type, B-type HATs,

is characterized by their cytoplasmic localization. They are believed to be involved in

acetylation events associated with the transfer of newly synthesized histones from the

cytoplasm to the nucleus for the purpose of replication-coupled nucleosome assembly

(Roth et al. 2001). Specifically, B-type HATs, such as histone acetyltransferase 1

(Hat1), are believed to acetylate free histones prior to their deposition onto newly replicated DNA (Tse et al. 1998). By comparison, the second type of HATs, A-type

HATs, is characterized by its nuclear localization. These HATs are believed to be

involved in acetylation events related to transcription (Roth et al. 2001). Specifically,

they are thought to be participants in the acetylation of core histone tails following

nucleosome and chromatin assembly (Tse et al. 1998).

12

HAT families and functional motifs

In general, HAT enzymes are grouped together based on their conserved sequence domains. Several large families exist. The GNAT (GCN5-related N- acetyltransferase) superfamily was originally described following the recognition that

Hat1, the only known B-type HAT to date, and GCN5 A-type HATs exhibit similarities in acetylation-related structural motifs (Roth et al. 2001). Members of the GNAT superfamily contain up to four conserved structural motifs. The most highly conserved motif, motif A, contains an R/Q-X-X-G-X-G/A sequence, which is essential for the recognition and binding of acetyl-CoA (Dutnall et al. 1998; Wolf et al. 1998).

The largest and most divergent HAT family in mammals is the MYST superfamily. Like the GNAT superfamily, motif A is prevalent in members of the MYST

(Moz-Ybf2/Sas3-Sas2-Tip60) superfamily. The conservation of this motif between the two superfamilies suggests that motif A (known as the MYST domain when referring to the MYST superfamily) is commonly used to bind acetyl-CoA (Roth et al. 2001). Studies have shown MYST family members to be involved in various cellular functions such as apoptosis, transcriptional regulation, cell cycle progression, and DNA replication and repair (Thomas and Voss 2007; Pillus 2008). For example, Hbo1 (histone acetyltransferase binding to ORC) bound to ORC (origin recognition complex) interacts with essential proteins (ORC7 and MCM2) from the pre-replication complex and contributes to the regulation of the assembly of this complex and the initiation of DNA replication (Iizuka and Stillman 1999; Burke et al. 2001; Aggarwal and Calvi 2004;

13

Doyon et al. 2006; Iizuka et al. 2006). The first human MYST member to be identified as

having HAT activity was Tip60 (TAT-interactive protein 60) (Kamine et al. 1996;

Yamamoto and Horikoshi 1997; Kimura and Horikoshi 1998). Similar to the role of the

histone acetyltransferase Esa1 in yeast, Tip60 appears to be involved in the regulation

of cell cycle progression in higher eukaryotes indicating that Tip60 is a functional

homologue of Esa1 (Clarke et al. 1999; Doyon et al. 2004; Latrasse et al. 2008).

A third major family of HATs is the p300/CBP family. CBP and p300 were

originally identified as proteins associated with phosphorylated CREB and adenovirus

E1A protein, respectively (Chrivia et al. 1993; Eckner et al. 1994). Both of these proteins

are highly related and share a similar biological function (Arany et al. 1994; Arany et al.

1995; Lundblad et al. 1995; Shikama et al. 1997). Several distinct domains are found in

p300/CBP, including a bromodomain, a HAT domain, and three putative zinc finger

domains (cys, ZZ, and TAZ domains). Two additional regions have been shown to

associate with a variety of transcription factors (Janknecht and Hunter 1996; Shikama et al. 1997). p300/CBP do not interact directly with DNA. Instead, they are recruited to

promoter regions through their associations with various transcription factors bound to

DNA. Two such transcription factors are phosphorylated CREB and E1A protein (Roth

et al. 2001). Once localized to promoter regions, p300/CBP participate in transcriptional

regulation by serving as transcriptional co-activators (Kwok et al. 1994; Janknecht and

Hunter 1996). Their function as transcriptional co-activators, however, is dependent on

their HAT activities. When mutations are introduced into the HAT active site, p300/CBP

14

lose the ability to activate transcription (Bannister and Kouzarides 1996; Ogryzko et al.

1996; Martinez-Balbas et al. 2000).

Histone substrates of HATs

Studies over the years have shown the acetylation of histones to be a non-random

process. Specific physiological conditions allow for the acetylation of specific amino

acids within specific histones (Chicoine et al. 1986). Two examples of this specificity are

the diacetylation of H4 at K5/K12 during the synthesis and deposition of histones and the

correlation between dosage compensation on the X of Drosophila males

and the acetylation of H4 at K16 (Turner et al. 1992; Bone et al. 1994; Sobel et al. 1995).

Observations such as these led to the hypothesis that distinct patterns of histone

acetylation are the result of the substrate specificity of HATs involved in specific cellular

processes (Brownell and Allis 1996). Studies of the substrate specificities of different

HATs have supported this concept. One example is Hat1, which acetylates histone H4

at K5 and/or K12, a pattern of acetylation consistent with that observed during histone

deposition (Chicoine et al. 1986; Verreault et al. 1996). Other HATs exhibiting specific patterns of acetylation include Gcn5 and PCAF, which preferentially acetylate H3 at K14

(Kuo et al. 1996; Xu et al. 1998; Schiltz et al. 1999). Interestingly, Gcn5 containing

complexes such as SAGA and ADA in yeast have more expanded substrate specificity

and are capable of acetylating multiple sites in H3 and H2B (Grant et al. 1999). Two

HATs capable of generating both site-specific and multi-site patterns of acetylation are

15

p300 and CBP. All four nucleosomal histones are acetylated equally well by these two enzymes (Schiltz et al. 1999).

Histone acetyltransferase 1: Identification

The first HAT to be identified was the B-type HAT, Hat1. Two independent approaches were used to identify the hat1 gene as coding for a histone acetyltransferase. In the first approach, temperature sensitive yeast mutants were screened to identify a strain exhibiting decreased HAT activity from cell extracts.

Although the decreased HAT activity and the temperature sensitive phenotype did not segregate together, the screen did allow the gene for Hat1 to be mapped and cloned

(Kleff et al. 1995). In the second approach, the predominant cytoplasmic HAT activity from yeast was isolated. The gene responsible for the HAT activity was subsequently identified using protein sequence data from the catalytic subunit (Parthun et al. 1996).

Histone acetyltransferase 1: Biochemistry

Two observations led to the original classification of Hat1 as a B-type HAT. First, yeast cytosolic fractions exhibited elevated levels of Hat1 associated acetyltransferase activity, which suggested that Hat1 resides in the cytoplasmic compartment. Second, recombinant and native Hat1 enzymes acetylate free histones, but not histones assembled in nucleosomes (Parthun et al. 1996).

Isolation of Hat1 from budding yeast cytosolic extracts also results in the co- purification of Hat2, a homologue of Rbap46/48 in mammals (Parthun et al. 1996;

Parthun 2007). When associated with Hat1 in a complex, Hat2 has been found to

16

increase the enzyme's catalytic activity ten-fold, presumably by facilitating interactions

between Hat1 and H4 (Parthun et al. 1996). One factor that may contribute to the

inability of Hat1/Hat2 complex to acetylate nucleosomal substrates is the concealment of

the H4 α-helix proximal to the N-terminal tail during nucleosome assembly, which inhibits the binding of Rbap46/48 (Verreault et al. 1998; Vermaak et al. 1999; Furuyama et al.

2006; Parthun 2007). Purification studies performed on a number of different organisms suggest the H4 specific B-type HAT complex to be evolutionarily conserved. B-type

HAT complexes isolated from Xenopus laevis , humans, and maize all contain subunits similar to Hat1 and Hat2 (Eberharter et al. 1996; Chang et al. 1997; Verreault et al.

1998; Imhof and Wolffe 1999; Lusser et al. 1999). Additionally, the substrate specificity of human and Xenopus laevis HAT1 are similar to yeast Hat1 with each enzyme able to acetylate free H4 at K5 and K12 (Chang et al. 1997; Verreault et al. 1997; Imhof and

Wolffe 1999).

The crystal structure of yeast Hat1 in a complex with acetyl-CoA shows Hat1 assuming a curved conformation with the C-terminus and N-terminus at opposite ends.

The active site of Hat1 is positioned near the middle of the concave face of this curved structure. Acetyl-CoA binds in a cleft in the active site (Dutnall et al. 1998; Parthun

2007). The structure of Hat1 also contributes to the substrate specificity of Hat1

(Parthun 2007). It is believed that the specificity of yeast Hat1 for specific lysine residues is achieved through the association with a recognition motif consisting of

GXGKXG (X can be any amino acid) (Parthun et al. 1996). In the current model for the

17

binding of the H4 tail to Hat1, the H4 tail is believed to reside in a shallow channel on the

concave surface of Hat1. The alignment of the K12 next to acetyl-CoA positions K8 and

K16 near the acidic patches located on the surface of Hat1 (Dutnall et al. 1998; Dutnall

et al. 1998). It is believed that the electrostatic interactions between these positive

charged lysine residues and negatively charged patches stabilizes association between

Hat1 and the H4 tail during the acetylation of K12 (Dutnall et al. 1998). Studies from our

own laboratory have provided support for this model. Synthetic peptides previously

acetylated at K8 and K16 prevented both human and yeast Hat1 from acetylating the

peptides (Makowski et al. 2001). Additionally, the substitution of the lysine residues at

these positions for arginine residues did not prevent the acetylation of the peptides by

Hat1, suggesting that the binding of Hat1 to substrates is not dependent on the presence

of individual lysine residues, but rather the positive charge at these positions (Benson et

al. 2007).

Evidence for the in vivo substrate specificity of Hat1 comes from several studies

in S. cerevisiae . A deletion of hat1 in S. cerevisiae does not result in a detectable

phenotype (Kleff et al. 1995; Parthun et al. 1996). However, when this deletion is

combined with mutations of lysine residues in the H3 N-terminal tail, defects in DNA

double-strand break repair and telomeric silencing are observed (Kelly et al. 2000; Qin

and Parthun 2002). In terms of telomeric silencing, combining these same mutations of

the H3 N-terminal tail with a H4 lysine 12 to arginine mutation phenocopies the hat1 mutation. No defect in telomeric silencing is observed with transmutations of lysines 5,

18

8, or 16, which suggests that Hat1 specifically acetylates lysine 12 of H4 (Kelly et al.

2000). In another study, a general increase in acetylation is observed at the four lysine

residues of H4 following the introduction of a double-stranded break at the MAT locus by endonuclease HO. The observed increase of acetylation at lysine 12 requires Hat1, which is recruited to the site of the double-stranded break (Qin and Parthun 2006).

Thus, in relation to DNA repair and telomeric silencing in budding yeast, Hat1 appears to work by way of the acetylation of H4 at K12 (Parthun 2007).

Evidence for the Hat1 dependent acetylation of newly synthesized histones comes from a study by Mosammaparast and colleagues. Immunoprecipitation of protein

A-tagged histone H4 from S. cerevisiae cytoplasmic extracts results in the co-purification

of Kap123, Hat1, and Hat2. The interaction of H4 molecules with Kap123 (a nuclear

import factor) suggests the H4 molecules have been targeted for nuclear import

(Mosammaparast et al. 2002). The association of the Hat1/Hat2 complex with pre-

nuclear H4 molecules is in agreement with the enzyme acetylating newly synthesized

histones (Parthun 2007).

The observation that newly synthesized H3 and H4 are acetylated at the start of

chromatin assembly and rapidly deacetylated at the conclusion of chromatin assembly

suggests that this process of acetylation and deacetylation is involved in de novo histone deposition (Annunziato and Hansen 2000). This question, however, has been difficult to answer due, in large part, to the difficulty in characterizing the in vivo function of Hat1.

Part of the difficulty stems from the finding that neither the conserved Hat1/Hat2 complex

19

nor the conserved acetylation pattern on new H4 is essential for yeast viability (Megee et

al. 1990; Kleff et al. 1995; Parthun et al. 1996; Ma et al. 1998). S. pombe and chicken

DT40 cells also do not require Hat1 for cell viability (Barman et al. 2006; Benson et al.

2007). It appears that Hat1 is not essential for all nucleosome assembly in vivo . The lack of a detectable phenotype relating to chromatin assembly or viability upon the deletion of Hat1 suggests that redundancies may exist in the chromatin assembly process. These redundancies may include multiple chromatin assembly pathways, different acetyltransferases modifying newly synthesized H4 in the absence of Hat1, or the overlap in function between acetylated new H3 and new H4 (Parthun 2007).

Although Hat1 is not essential for viability, there is evidence that Hat1 is involved in chromatin assembly. The observed defects in telomeric silencing and DNA double- stranded break repair following the deletion of hat1 may be a reflection of a deficiency in chromatin assembly (Parthun 2007). Perhaps the most compelling evidence in support of Hat1 having a role in chromatin assembly comes from the isolation of the nuclear

Hat1/Hat2 complex in S. cerevisiae . Isolation of the nuclear Hat1/Hat2 complex results in the co-purification of Hif1 (Hat1 interacting factor 1), a histone H3/H4 specific chaperone in budding yeast with in vitro chromatin assembly activity. The fact that the

NuB4 (nuclear B-type HAT specific for H4) complex interacts with H3 and H4 in vivo and

Hat1/Hat2 are required for the association of Hif1 with histones suggests that Hat1 has a role in chromatin assembly in vivo . (Ai and Parthun 2004; Poveda et al. 2004)

20

One current model of Hat1 activity in the cell proposes that the Hat1/Hat2 complex associates with newly synthesized H4 in the cytoplasm. The Hat1/Hat2 complex subsequently acetylates new H4 at K5 and K12. Newly synthesized H3 binds to H4. With the aid of karyopherins ( i.e. Kap123), the entire complex is imported into the nucleus. Following its import into the nucleus, the Hat1/Hat2/H3/H4 complex interacts with Hif1 creating the aforementioned NuB4 complex, which deposits H3/H4 onto DNA.

It remains unclear, however, whether Hat1 and Hat2 separate from Hif1 prior to histone deposition or whether the NuB4 complex remains intact during this process. This model does suggest that Hat1 plays a larger role than simply acetylating new H4. It has been suggested Hat1 may also act as a chaperone for H4 from the time of its synthesis to the time it is deposited onto DNA (Parthun 2007).

Telomeres

Telomeres were originally described by Muller in 1938 and McClintock in 1941.

Telomeres are specialized heterochromatic structures composed of both DNA and assorted associated proteins. The DNA component contains G-rich repeat sequences that can vary from organism to organism. For example, the repeat sequence in humans

is TTAGGG, in S. pombe is TTACAG 2-5, in S. cerevisiae is TG 2-3(TG) 1-6, and in many plants is TTTAGGG (Shampay et al. 1984; Moyzis et al. 1988; McKnight et al. 1997). A single stranded 3'-overhang of this G-rich sequence, called the G-tail, resides near the distal end of the telomere and extends for roughly 75-200 nucleotides in humans

21

(Makarov et al. 1997; McElligott and Wellinger 1997; Wright et al. 1997; Kanoh and

Ishikawa 2003).

Double-stranded DNA breaks and protein-free DNA ends are highly unstable

structures that are vulnerable to degradation by nucleases and DNA end fusion by DNA

repair mechanisms. Chromosome ends, by comparison, are relatively stable structures

due to the protection provided by telomeres. This imparted stability results from the

formation of a protective cap by telomere associated proteins, which effectively insulate

the chromosome ends against nucleases and the activation of DNA damage checkpoints

(Broccoli 2004). In addition to this protective function, telomeres also have a role during

meiosis during which they cluster toward the spindle pole body (SPB) or near the

centrosomes. In S. pombe , this clustering is essential for the proper pairing of homologous (Chikashige et al. 1994; Scherthan et al. 1994; Trelles-

Sticken et al. 1999; Niwa et al. 2000). Moreover, genes located in a subtelomeric region are transcriptionally silenced by the telomere (referred to as telomeric silencing or telomere position effect) (Nimmo et al. 1994).

S. pombe telomeres

The organization and dynamics of S. pombe chromosomes are analogous to those in humans. Like humans, fission yeast have large, complex, heterochromatic centromeres and an RNAi pathway (Cooper and Hiraoka 2006). The organization of telomeric proteins is also similar to that of humans. To date, Taz1 is the only known double-stranded DNA-binding protein that is specific to the S. pombe telomere and the

22

only known orthologue of mammalian TRF1 (telomere repeat factor 1) and TRF2

(Cooper et al. 1997; Li et al. 2000; Fairall et al. 2001). Taz1 is a 663 amino acid protein

with a helix-loop-helix motif (Myb motif) at the carboxy terminus and a TRFH (TRF

homology) domain in the central region (Kanoh and Ishikawa 2003; Cooper and Hiraoka

2006). Taz1 is believed to have three main functions relating to the telomere: regulating

the length of telomeric DNA, telomeric silencing, and clustering of telomeres toward the

spindle pole body (SPB) during meiosis. Similar to the interaction of Rap1 with TRF2 in

humans, Rap 1 is recruited to fission yeast telomeres by Taz1. Mutations of rap1 in S. pombe have demonstrated Rap1 to be important for telomeric clustering toward the

SPB. The localization of S. pombe Rif1 (Rap1 interacting factor) to telomeres is also dependent on its interaction with Taz1 (Kanoh and Ishikawa 2003). This localization of

Rif1, however, is independent of Rap1. Studies suggest that Rif1 may act as a competitor to Rap1 for Taz1 binding and may be involved in regulating telomere length.

Deletion of rif1 from S. pombe , however, does not result in other defects in telomere function (Kanoh and Ishikawa 2003; Cooper and Hiraoka 2006). Interacting with DNA located at the most distal end of the telomere is Pot1 (protection of telomeres 1)

(Baumann and Cech 2001). It is the only protein known to associate with single- stranded telomeric DNA (Cooper and Hiraoka 2006). Deletion of pot1 results in the

abrupt loss of telomeric DNA, suggesting that Pot1 is essential for the protection of

chromosome ends (Kanoh and Ishikawa 2003). Ku70 and Ku80 in S. pombe form a

heterodimer that localizes to telomeres and double-stranded breaks. Similar to their

23

homologues in S. cerevisiae , Ku70 and Ku80 participate in telomere maintenance.

Deletion of the gene for either protein results in shorter telomeric repeat DNA and results in the recombination of telomeric repeat sequences (Kanoh and Ishikawa 2003).

Similar to humans and S. cerevisiae , telomere-proximal genes in S. pombe are

transcriptionally silenced (Nimmo et al. 1994). This silencing requires a number of

different proteins including: Taz1, Rap1, and a variety of different heterochromatin

proteins consisting of Swi6, Csp4, Rik1, Clr1, Clr2, Clr3, and Clr4 (Allshire et al. 1995;

Cooper et al. 1997; Chikashige and Hiraoka 2001; Kanoh and Ishikawa 2001).

Chromatin immunoprecipitation studies have shown Taz1 to be present up to 10 kb

away from the telomere and Swi6, a HP1 homolog, to associate with the final 50 kb of

the chromosome (Cooper and Hiraoka 2006). Regions of chromatin associated with

Swi6 are normally transcriptionally silent. The binding of Swi6 requires Clr4 (cryptic loci

regulator 4) and Rik1 (Ekwall et al. 1996; Nakayama et al. 2000). The recruitment of

Clr4 to chromatin by Rik1 results in the methylation of H3 at lysine 9 by Clr4 (a H3 K9

methyltransferase), prompting the subsequent binding of Swi6 (Nakayama et al. 2001).

Once it is recruited to a specific region, Swi6 is capable of self-assembling and

spreading to adjacent regions by interacting with histone modifiers that generate binding

sites on nearby nucleosomes (Kanoh et al. 2005). This spreading of Swi6 leads to the

spreading of heterochromatin.

24

Chapter One

Characterization of the Hat1 complex in S. pombe

Summary

A Hat1-TAP strain was created in the fission yeast Schizosaccharomyces pombe . Through tandem affinity purification (TAP) and mass spectrometry, we have demonstrated the in vivo association of Hat1 and Mis16. Additionally, mis16 was confirmed to be an essential gene in S. pombe . Lastly, in vitro assays examining the

activity of Hat1 show that this enzyme is capable of acetylating histone H4 at lysines 5

and 12.

25

Materials and Methods

S. pombe strain construction

S. pombe were cultured and maintained in YEA medium. The genotypes

of strains are listed in Table 1. The construction of KTP1 was performed as follows.

The S. pombe Hat1 protein (SPAC139.06; Uniprot accession number Q9UTM7) was

TAP tagged using a modified PCR based method originally described by Bahler et al.

(Bahler et al. 1998). PCR primers hat1-tagfor (5'-

CTACCCAAGCTTAAGGAAGATTCGCCT-

CGAAAACGCCAAAAACTTGCTCAATCTTCTTCCCGGATCCCCGGGTTAATTAA-3')

and hat1-tagrev (5'-

AAGCTTTCAAAAGCAAATTATATAAAAAGTAATTGCGTCCAATAG-

TGTAATTTAGTCGATGAATTCGAGCTCGTTTAAAC-3') were used to amplify the TAP

tag from plasmid pFA6A-CBP 4.5X protein A (TEV)-kanMX6. The reaction conditions

were as follows: 1 cycle at 95 0C for 2 min; 30 cycles at 95 0C for 30 sec, 50 0C for 30 sec, 72 0C for 5 min; and 1 cycle at 72 0C for 15 min. This amplicon, containing sequence homology to the COOH terminus of hat1 , was integrated into a wildtype strain

(975) by transformation as previously described (Keeney and Boeke 1994). Briefly, a colony was grown overnight at 30 0C, pelleted, and washed twice with sterile water before being washed with LiAc/TE (0.1 M lithium acetate, 10 mM Tris-HCl, pH 7.6, 1 mM EDTA). An equal volume of LiAc/TE was added to the cell pellet. In an Eppendorf tube, 100 µl of this re-suspension was combined with 2 µl boiled salmon testes DNA (10

26

mg/ml stock, Sigma Aldrich) and 10 µl of transforming DNA. The suspension was

incubated at room temperature for 10 min before adding 260 µl of 40% PEG/LiAc/TE

and incubating at 30 0C for 60 min. DMSO (43 µl) was added and the suspension was heat shocked for 5 min at 42 0C. The cells were pelleted, washed, and re-suspended with sterile water before plating onto YEA and G418 (100 mg/ml) plates. Candidate cells were selected for their resistance to G418.

The TAP tagging of Hat1 was confirmed by yeast colony PCR using 10 pmol each of primers hat1-f2 (5'-TTGACGTTCATTC-GGCAAATAG-3') and kanrev22 (5'-

GCT-TATTTAGAAGTGGCGCG-3') under the following conditions: 1 cycle at 98 0C for

10 min; 1 cycle at 94 0C for 2 min; 35 cycles at 94 0C for 1 min, 55 0C for 1 min, 72 0C for

2 min 30 sec; and 1 cycle at 72 0C for 7 min. PCR reactions were run on 1% agarose gels containing 0.5 ng/µl ethidium bromide (EtBr) and visualized under UV light. 1kb plus DNA ladder (Invitrogen) was included on the gels as a marker for size.

To confirm the gel electrophoresis results, the PCR reactions were purified using the Wizard SV Gel and PCR Clean-Up System (Promega) and sequenced on a CEQ

8000 Genetic Analysis System (Beckman Coulter). Briefly, PCR reactions using primers hat1-f2 and kanrev22 were combined with five volumes of Membrane Binding Solution and added to a SV Minicolumn assembly. Samples were centrifuged, washed twice with

Membrane Wash Solution, and eluted with water. O.D. 260/280 readings were taken on a SmartSpec Plus spectrophotometer (Bio-Rad). PCR sequencing reactions were set- up using the CEQ DTCS quick start kit (Beckman Coulter): 100 fmol of DNA was

27

vacuumed dried, re-suspended in 5 µl water, and combined with 4 µl DTCS Quick Start

Master Mix and 1 µl of hat1-f2 (5 pmol/µl). The reaction conditions were as follows: 36

cycles at 96 0C for 20 sec, 50 0C for 20 sec, 60 0C for 4 min. After thermal cycling, 0.5 µl

of glycogen solution and 2 µl of stop solution (1.5 M NaOAc, 50 µM EDTA) were added.

DNA pellets were ethanol precipitated, re-suspended in 35 µl sample loading solution,

and sent for sequencing (Boston College).

A Hat1-TAP Mis16-myc tagged strain (KTP40) was constructed as

follows. KTP1 was mated with mis16-myc on sporulation medium (MEA) to introduce a

ura4- marker into KTP1. This new strain, designated KTP22, was subsequently mated

with mis16-myc. Tetrad dissections on YEA plates were replica plated onto EMM-ura, 5-

fluoroorotic acid (5-FOA), and G418 (100mg/ml) plates to select for Ura+, G418 resistant

cells.

The construction of a S. pombe diploid strain heterozygous for deletion of

mis16 (SPCC1672.10, Uniprot accession number O94244) (KTP7) was performed as follows: PCR primers hat2delfor2 (5'-

ATGTCAGAGGAAGTAGTCCAGGATGCACCTCTCGAGA-

ATAATGAACTCAATGCCGAGATACGGATCCCCGGGTTAATTAA-3') and hat2rev2 (5'-

TGGTGTTATAGAAATGTAGTCTGATTTATAACAGTAGTTTTGATGTATTTACAAGGCG

ACGAATTCGAGCTCGTTTAAAC-3') were used to amplify the kanMX6 selectable marker from plasmid PFA6a-3HA-kanMX6. The reaction conditions were as follows: 1 cycle at 94 0C for 2 min; 30 cycles at 94 0C for 30 sec, 66 0C for 30 sec, 68 0C for 2 min

28

30 sec; and 1 cycle at 68 0C for 10 min. The amplicon, which contains sequence homology to regions flanking mis16 , was integrated into diploid strains by transformation. Deletion of mis16 was confirmed by colony PCR using primers hat2test2

(5'-TTCA-GACTTAAGAGTGCGCTAG-3') and hat2test (5'-TAGTACGGAGAGAGC-

CCTGG-3'). The reaction conditions were as follows: 1 cycle at 98 0C for 10 min; 1 cycle at 94 0C for 2 min; 35 cycles at 94 0C for 1 min, 53 0C for 1 min, 72 0C for 2 min 45 sec; and 1 cycle at 72 0C for 7 min. Agarose gel electrophoresis was used to confirm amplicon size. To determine if mis16 is essential, tetrad dissections of a heterozygous diploid strain was performed on YEA plates, grown at 30 0C, and replica plated to G418

(400 mg/ml) and YEA plates.

Tandem affinity purification

The tandem affinity purification method originally described by Rigaut et

al . was performed as modified (Puig et al. 2001) (Rigaut et al. 1999). Two liters of yeast

were grown to ~ 1 x 10 7 cells/ml. After harvest, cells were filtered through a glass fiber filter circle (Fisher brand). Cells and filter were frozen with liquid nitrogen and ground

into a powder before being re-suspended in 10 ml of NP-40 buffer (6 mM NaH 2PO 4, 1%

NP-40, 150 mM NaCl, 2mM EDTA, 50 mM NaF, 0.1 mM Na 3VO 4, 1.3 mM Benzamidine,

1mM PMSF, Complete tablet, EDTA free (Roche)). The lysate was centrifuged for 5 min at 5000 rpm at 4 oC and the supernatant collected. The pellet was washed with 5 ml of

NP-40 buffer and centrifuged for 5 min at 5000 rpm at 4 oC. The supernatant was clarified by centrifugation for 1 hr at 38000 rpm at 4 oC. The clarified supernatant was

29

incubated on a rotating platform for 2 hrs at 4 oC following the addition of 800 µl of IgG

Sepharose beads (GE Healthcare) in NP-40 buffer without proteinase inhibitors (1:1).

The lysate and beads were added to a Bio-Rad Poly Prep chromatography column and allowed to pack by gravity. The beads were washed with 30 ml of IPP 150 buffer (10 mM Tris-HCL pH 8.0, 150 mM NaCl, 0.1% NP-40), followed by 10 ml of TEV cleavage buffer (10 mM Tris-HCl pH 8.0, 150 mM NaCl, 0.1% NP-40). The beads were re- suspended in 1 ml TEV cleavage buffer and 500 units AcTEV Protease (Invitrogen) and incubated for 2 hrs at 16 0C with gentle rocking. The eluate was drained into a new

column and the old column was washed with 1 ml TEV cleavage buffer. Three volumes

of calmodulin binding buffer (CBB) (10 mM Tris-Hcl pH 8.0, 150 mM NaCl, 1 mM Mg 2+

Acetate, 1 mM Imidazole, 2 mM CaCl 2, 10 mM β-mercaptoethanol), 3 µl of 1M CaCl 2 per

ml of IgG eluate, and 300 µl calmodulin affinity resin (Stratagene) in CBB (1:1) were

added to the eluate. The eluate was incubated for 1 hr at 4 oC with gentle rocking. The

beads were washed twice with 10 ml of CBB containing 0.1% NP-40 followed by one

wash with 10 ml CBB containing 0.02% NP-40. The protein complexes were eluted with

1 ml calmodulin elution buffer containing 0.02% NP-40 and precipitated with 25%

trichloroacetic acid (TCA). Resulting pellets were washed once with acidified acetone

(0.05N HCl), once with acetone alone, and vacuum dried.

Mass spectrometry

Protein samples were run through approximately 2 cm of a 10% SDS-PAGE gel.

Each lane was subsequently cut out above the dye front and placed in separate 1.5 ml

30

Eppendorf tubes. The gel fragments were fixed for 30 minutes with a solution containing

50% methanol and 5% acetic acid. After fixation, the gel fragments were washed 2 x 30 min with distilled water. The samples were sent to the Tapalin Biological Mass

Spectrometry Facility at Harvard University where they were analyzed using an LCQ

DECA ion-trap mass spectrometer.

Methyl methanesulfonate (MMS) assays

The analysis of the sensitivity of wildtype and mutant S. pombe strains was performed as previously described (Qin and Parthun 2002). Briefly, wildtype and mutant strains were grown overnight in 3 ml of YEL. Cells were counted on a hemacytometer and re-suspended at 2 x 10 6 cells/1000 µl. Five fold serial dilutions were made and 10 µl of each dilution was spotted on Edinburgh minimal medium (EMM) (Moreno et al. 1991) plates containing 0.01% MMS (Sigma Aldrich). The plates were incubated 3 to 4 days at

30 0C and photographed.

Gel electrophoresis and immunoblotting

To separate the subunits of the HAT1 complex, TAP extracts were subjected to

SDS-PAGE in 10% to 12.5% polyacrylamide gels (Laemmli 1970). The resolved proteins were transferred from SDS gels to Immobilon-P transfer membranes (Millipore) at 30V overnight using transfer buffer (10% methanol, 25 mM Tris, and 192 mM glycine).

To visualize the proteins after transfer, the membranes were stained with Ponceau-S solution (Sigma Aldrich) and de-stained with 10% acetic acid. The membranes were washed three times with TBS (0.3 M NaCl, 20 mM Tris, pH 7.4) prior to immunoblotting.

31

Immunoblotting was performed according to the Western-Star system (Applied

Biosystems). Briefly, the membrane was blocked for 1 hour at room temperature using

blocking buffer (1x TBS, 0.2% I-Block reagent, 0.1% Tween-20 detergent). Primary

antibodies were diluted in blocking buffer: anti-calmodulin binding protein epitope tag

(#05-932, Millipore) was diluted 1:5000; anti-protein A was diluted 1:5000; anti-c-myc

(sc-40, Santa Cruz Biotechnology) was diluted 1:5000. Membranes were incubated with

primary antibody for 2 hours at room temperature and subsequently washed 2 x 5 min

with blocking buffer. Secondary antibodies conjugated to alkaline phosphatase were

diluted 1:5000 and applied to the membranes for 1 hour at room temperature. The

membranes were washed 3 x 5 min with blocking buffer followed by 2 x 2 min washes

with 1X assay buffer. Following a 5 min incubation with CDP-Star substrate, antibody reactions were visualized by exposing the membranes to HyBlot CL autoradiography film

(Denville) and developing.

In vitro HAT filter binding assay

In vitro HAT filter binding assays were essentially performed as described

(Benson and Annunziato 2004). To set up 80 µl reactions, 40 ng of yHat1p or 50 µl of

tandem affinity purified Hat1p was added to a 1.5 ml eppendorf tube. To this tube, the

following was added: 100 mM sodium butyrate, pH 7.2 to 5 mM concentration, 10 mg/ml

acetylated BSA to 1 mg/ml concentration, 4 µg of substrate peptide, and 0.1 mCi/ml

[3H]acetyl-CoA (Perkin Elmer Life Sciences) to 4-8 µci/ml. Buffer (10 mM Tris pH 8.0,

150 mM NaCl) was added to bring the final volume to 80 µl. Reactions were mixed and

32

incubated at 37 0C for 30 min. To terminate the reactions, duplicate 35 µl aliquots were

spotted onto Whatman p81 filter circles and submerged in 50 mM NaHCO 3, pH 9.2 for

30 min at room temperature. The filters were washed 4 x 10 min in 50 mM NaHCO 3, pH

9.2 with agitation. After drying overnight, the filters were placed in scintillation vials

along with 5 ml of Ecoscint A (National Diagnostics) and counted on a Tri-Carb

2100/2900 TR liquid scintillation counter (Perkin Elmer). To account for background, two

blank filter circles were submerged and washed with NaHCO 3,Si pH 9.2. The average

background counts per minute (CPM) was subtracted from the average experimental

CPM to give the final experimental CPM.

Histone acetyltransferase assay

Hat1p was tandem affinity purified from KTP1 and collected following elution. To

set-up 100 µl reactions, the following were combined in a 1.5 ml Eppendorf tube: 100

mM sodium butyrate, pH 7.2 to 5 mM concentration, 10 mg/ml acetylated BSA (GE

Healthcare) to 1 mg/ml concentration, 0.5 µg recombinant H4 (#14-697, Millipore), 1mM

unlabeled acetyl-CoA (Sigma Aldrich) to 10 µM concentration, and 79 µl of purified

Hat1p. The reactions were mixed, incubated at 30 0C for 1 hour, and cooled on ice.

Recombinant H4 was precipitated with 25% TCA, washed with acidified acetone (0.05 N

HCl) and then acetone (Annunziato and Seale 1983). Protein pellets were vacuum dried, re-suspended in SDS sample buffer (0.17 M Tris, pH 6.8, 2.3% SDS, 2.3 M Urea,

23% glycerol, 4.7 mM EDTA, 2.3% 2-mercaptoethanol, 0.002% bromophenol blue), and subjected to SDS-PAGE in a 13% polyacrylamide gel. The resolved proteins were

33

transferred to Immobilon-P transfer membranes and immunoblotted as previously described above. The acetylated (ac) antibodies used and their dilutions are as follows: anti-acH4 K5/K12 (7481) was diluted 1:5000; anti-acH4 K5 (#07-290, Millipore) was diluted 1:600; anti-acH4 K12 (#07-595, Millipore) was diluted 1:800; anti-acH4 K8 (#07-

378, Millipore) was diluted 1:200; anti-total H4 (#05-858, Millipore) was diluted 1:30000.

34

Results

Composition of the cytosolic Hat1 complex in S. pombe

In order to purify Hat1 and its partners from S. pombe , a TAP tag was incorporated on the C-terminus of Hat1. To confirm the presence of the TAP tag in the candidate strain (KTP1), extracted proteins were subjected to SDS-PAGE and probed for the presence of protein A (Fig. 1A). Western blot analysis showed the presence of a band (64 kDa) in KTP1 corresponding to the approximate size of Hat1 (44 kDa) and the

TAP tag (20 kDa). A band was also visible in the TAP marker lane which contained extracted proteins from a S. pombe strain expressing a TAP tagged catalytic domain of adenylate cyclase. No bands were visible in the wildtype (975) control lane indicating that protein A and therefore the TAP tag were not inherently present in S. pombe . DNA sequencing analysis confirmed the incorporation of the TAP tag on the C-terminus of

Hat1 in KTP1 (data not shown).

Previous studies in our own lab have shown that a deletion of hat1 in S. pombe results in the heightened sensitivity to the DNA damaging agent MMS (Benson et al.

2006). To determine whether TAP-tagged Hat1 is functional, MMS assays were performed using the following strains: KTP1 (Hat1-TAP), LBP6 ( hat1 ∆), and 975 (WT)

(FIg. 1B). When grown on medium containing 0.01% MMS, LBP6 displayed a growth defect as compared to 975. KTP1 by comparison did not display a growth defect and grew similarly to 975. The growth of the three strains on medium lacking MMS was

35

similar indicating that the effects seen on MMS were not due to inconsistencies in cell

number.

To determine the components of the Hat1 complex in S. pombe , Hat1 and its

partners were purified from KTP1 by tandem affinity purification and analyzed by mass

spectrometry. Along with Hat1, four other proteins were identified by mass

spectrometry. Of these proteins, only Hat2 (an orthologue of Mis16 in S. pombe ) was unique to the TAP purification of KTP1 (Fig. 2). The other proteins were most likely contaminants inherent to the TAP purification as they were eluted from a wildtype strain deficient for TAP tagged Hat1. These data suggest an in vivo association between Hat1 and Mis16 in S. pombe .

To confirm the interaction between Hat1 and Mis16 observed in the mass spectrometry data, a Hat1-TAP, Mis16-myc tagged strain (KTP40) was generated in S. pombe . Following tandem affinity purification, western blots performed using an anti- myc antibody reveal a band corresponding to myc tagged Mis16 (~ 80 kDa) (Fig. 3A, left panels). Stripping and re-probing the blot with an antibody against calmodulin binding protein (CBP), the remnant of the TAP tag after purification, also revealed the presence of CBP-Hat1. A band corresponding to Mis16-myc was not visible when HAT1-TAP was purified from KTP1 due to the lack of myc tagged Mis16 (Fig. 3A, middle panels). As expected, S. pombe does not inherently contain myc or CBP epitopes. Bands were not observed when wildtype protein extracts were probed with anti-myc and anti-CBP antibodies (Fig. 3A, right panels).

36

As a control for non-specific protein interaction with the TAP tag, myc-tagged

Mis16 was not evident following a control TAP purification using protein extracts from a

myc-tagged Mis16 strain (Fig. 3B). Hat1-CBP was also not evident, as Hat1 is not TAP

tagged in this strain. In contrast, a band corresponding to myc-tagged Mis16 was clearly

visible when the purification was performed with the Hat1-TAP Mis16-myc strain,

suggesting that Mis16-myc is not pulled down in a tandem affinity purification without

Hat1-TAP. Together with the mass spectrometry data, these results suggest an in vivo interaction between Hat1 and Mis16.

Previous studies have shown decreased cell viability due to unequal chromosome segregation with a G1 arrested temperature sensitive mis16 mutant upon its introduction into complete medium at the restrictive temperature (37 0C) (Hayashi et al. 2004). To confirm the essential nature of mis16 , a S. pombe diploid strain heterozygous for deletion of mis16 (KTP7) was constructed. Tetrad dissections on YEA plates following sporulation resulted in two viable spores per tetrad after three days growth at 30 0C (Fig. 4, panel A). To determine if these haploid yeast colonies were resistant to G418 and therefore mis16 ∆ mutants, the colonies were replica plated to

G418 and YEA plates. Substantial colony growth was not observed on medium containing G418 after incubating for 2 days at 30 0C suggesting that these colonies were wildtype for mis16 (Fig. 4, panel B). The replica colonies grew on control YEA plates, which excludes the possibility of insufficient yeast transfer during replica plating (Fig. 4, panel C).

37

In vitro activity of the S. pombe Hat1 complex

Human Hat-B has been shown to acetylate lysines 5 and 12 of histone H4 in vivo

(Chang et al. 1997; Verreault et al. 1997). To determine if the Hat1 complex in S. pombe is capable of generating this deposition related K5/K12 pattern, in vitro analysis of histone acetyltransferase activity was performed.

The purified Hat1 complex was capable of acetylating an unacetylated H4 N- terminal peptide (aa 1-18), but not a peptide already acetylated at lysines 5 and 12 (Fig.

5). The level of acetylation with the K5/K12 peptide was the same as a "no peptide"

(NP) control reaction using Hat1 and [ 3H]acetyl coenzyme A. These results suggest that the S. pombe Hat1 complex has in vitro activity capable of acetylating histone H4 at K5 and K12.

To confirm the results from the peptide assays, in vitro Hat assays were performed using recombinant H4 and non-radioactive acetyl-CoA. The results were monitored by western blotting using anti-total H4 and anti-H4 acetylated at K5 and/or

K12 antibodies. The relative amount of total H4 was similar between the control recombinant H4 lane and the in vitro reaction lane, ruling out a possible error in protein load (Fig. 6, panel A). From the data collected, it was apparent that the S. pombe Hat1 complex was able to acetylate recombinant H4 at K5 and K12 (Fig. 6, panel B). These observations were in agreement with the results from the in vitro assays using H4 tail peptides.

38

To further elucidate the acetylation pattern generated by the Hat1 complex, experiments were performed using antibodies that recognize H4 monoacetylated at K5 or K12. The results indicate that the Hat1 complex was capable of acetylating histone

H4 at K5 and K12 (Fig. 6, panels C and D). Western blot analysis using an antibody that recognized H4 acetylated at K8 showed this complex to be capable of acetylating H4 at

K8 to a limited degree (Fig. 6, panel E).

39

Tables, Figures and Legends

Table I: S. pombe strain list

Strain Genotype

975 h + wildtype

972 h - wildtype

LBP6 h + hat1 ∆::kan

FY941 h + ade6-210 leu 1-32 ura4-DS/E tRNAPhe-otr1L (Xho1-BamHI fragment) Hpa1::ura4

FY1872 h 90 ade6-210 leu1-32 ura4-DS/E otrRSph1::ade6 TEL2L-ura4

FY336 h - ade6-210 leu1-32 ura4-DS/E TM1::ura4

FY496 h + ade6-210 leu 1-32 ura4-DS/E imr1L (dg-glu) NcoI::ura4 oril

FY648 h + ade6-210 leu 1-32 ura4-DS/E otr1R (dg-glu BamHI-Spe1 fragment) Sph1::ura4

mis16-myc h - leu1 ura4 mis16-myc [ura4+]

FWP93 h - ura4::fbp1-lacz leu1-32 ade6-210

KTP1 h + hat1-4xPACTAP

KTP7 h - / h + mis16 +/mis16 ∆::kan ade6-M216/ade6-M210 leu1 +/leu1-32

KTP22 h + ura4 hat1-4xPACTAP

KTP24 h - hat1 ∆::kan ade6-210 leu1-32 ura4::fbp1-lacz

KTP25 h + hat1 ∆::kan ade6-210 leu1-32 ura4::fbp1-lacz

KTP29 h - hat1 ∆::kan ade6-210 leu1-32 ura4-DS/E imr 1L (dg-glu) NcoI :: ura4 oriI

KTP30 h - hat1 ∆::kan ade6-210 leu1-32 ura4-DS/E otr1R (dg-glu BamHI - SpeI fragment) sphI::ura4

KTP33 h - hat1 ∆::kan ade6-210 leu1-32 ura4-DS/E TM1 :: ura4

KTP35 h + hat1 ∆::kan ade6-210 leu1-32 ura4-DS/E tRNA Phe-otr 1L (XhoI-BamHI fragment) HpaI :: ura4

KTP36 h 90? hat1 ∆::kan ade6-210 leu1-32 ura4-DS/E otr1 Rsph1 :: ade6 TEL2L-ura4 12C

KTP40 h + hat1-4xPACTAP mis16-myc

40

Figure 1: Generation of a S. pombe Hat1-TAP strain . ( A) Cell extracts were prepared from Hat1-TAP ( KTP1 ) and wildtype ( 975 ) S. pombe cells. Extracted proteins were resolved by SDS-PAGE and analyzed by western blot analysis using antibodies that recognize protein A (1:5000 dilution). ( B) Wildtype and experimental S. pombe strains were cultured on EMM plates in the presence or absence of 0.01% MMS. Cells were grown three (EMM) or four (EMM + 0.01% MMS) days at 30 0C. KTP1: Hat1-TAP ;

LBP6: hat1 . Note: When grown on medium containing 0.01% MMS, LBP6 displayed a

growth defect compared to 975. KTP1 by comparison did not display a growth defect

and grew similarly to 975.

41

Figure 1

42

Figure 2: Hat2/Mis16 co-purifies with Hat1 during tandem affinity purification. Cell extracts were prepared from Hat1-TAP (KTP1) S. pombe cells and tandem affinity purified. Proteins were subjected to SDS-PAGE and analyzed using a LCQ DECA ion- trap mass spectrometer. Note: Hat2 (also known as Mis16, an orthologue of Rbap46) is part of the Hat1 complex.

43

Figure 2

44

Figure 3: Hat1-TAP and Mis16-myc are associated in a complex . ( A) Cell extracts were prepared from wildtype and experimental S. pombe strains prior to tandem affinity

purification. Purified proteins were resolved by SDS-PAGE and analyzed by western

blotting with antibodies against calmodulin binding protein epitope tag (1:5000 dilution)

and c-myc (1:5000 dilution). KTP40: Hat1-TAP Mis16-myc ; KTP1: Hat1-Tap . Note: for

this experiment, membranes were stripped and re-probed. ( B) Cell extracts prepared

from Mis16-myc and Hat1-TAP Mis16-myc S. pombe strains were tandem affinity

purified. Total protein (Input) and bound proteins (Bound) were subjected to SDS-

PAGE, and western blot analysis. The antibodies used were against calmodulin binding

protein epitope tag (1:5000 dilution) and c-myc (1:5000 dilution). Note: Mis16-myc is not

pulled down in a tandem affinity purification without Hat1-TAP.

45

Figure 3

46

Figure 4: Mis16 is essential . A S. pombe diploid strain (KTP7) heterozygous for deletion

of mis16 was sporulated on MEA plates. ( A) Tetrads were dissected on YEA plates.

Spores were allowed to grow for 3 days at 30 0C prior to replica plating onto ( B) G418

(400 mg/ml) and ( C) YEA plates. Replica plated colonies were allowed to grow for 2 days. Note: Substantial colony growth was not observed on medium containing G418 after incubating for 2 days at 30 0C suggesting that these colonies were wildtype for mis16.

47

Figure 4

48

Figure 5: Acetylation of H4 N-terminal peptides by S. pombe Hat1 in vitro. Unacetylated

(UN ) and K5/K12 diacetylated ( 5/12 ) H4 N-terminal peptides (4 µg each) were incubated

in vitro for 30 min at 37° with affinity-purified yHat1 and [ 3H]acetyl-CoA. Reactions were also performed without added peptide ( NP ), and with the UN peptide minus Hat1 ( UN-

Hat1 ). Reactions were spotted onto P-81 filters and prepared for scintillation counting.

Results are expressed as a percentage of radioactivity incorporated into the

Unacetylated peptide.

49

Figure 5

50

Figure 6: Acetylation of recombinant H4 with S. pombe Hat1 in vitro. Recombinant H4

(0.4 µg) was incubated in vitro for 60 min at 30° with affinity-purified yHat1p an d unlabeled acetyl-CoA. Proteins from the reaction were resolved by electrophoresis, and analyzed by western blotting using antibodies that recognize total H4 ( A); or H4 acetylated at K5-and/or-K12 ( B), K5 ( C), K12 ( D), or K8 (E) .

51

Figure 6

52

Chapter Two

The role of Hat1 and acetylation in S. pombe telomeric silencing

Summary

Studies in S. cerevisiae have shown that the deletion of hat1 results in the loss of telomeric silencing, but only when combined with mutations of acetylatable lysines of the histone H3 N-terminal tail. In contrast, silencing assays performed in S. pombe revealed that the loss of hat1 alone results in the loss of telomeric silencing, indicating that the

redundant functions between Hat1 and H3 acetylation during telomeric silencing in S.

cerevisiae are absent from S. pombe . Deleting hat1 caused an increase of acetylation of H4 at telomeres. This finding, coupled with the observation that histone hyperacetylation results in the loss of telomeric silencing in hat1 + S. pombe, suggests

that the establishment of the correct epigenetic state at telomeres requires the timely

acetylation of H4 during chromatin assembly.

53

Materials and Methods

Strain construction

Strains KTP24 and KTP25 were generated by crossing LBP6 with FWP93.

Tetrad analysis was performed. Colonies exhibiting leucine, adenine, and uracil auxotrophies as well as G418 resistance were selected.

Centromeric and telomeric marker strains deleted for hat1 were constructed as

follows: KTP29 ( ura4-centromeric inner most repeat marker, hat1 ) was generated by

crossing FY496 with KTP24. KTP30 ( ura4-centromeric outer repeat marker, hat1 ) was

generated by crossing FY648 with KTP25. KTP33 ( ura4-centromeric central core marker, hat1 ) was generated by crossing FY336 with KTP25. KTP35 ( ura4-

euchromatic marker, hat1 ∆) was generated by crossing FY941 with KTP24. KTP36

(ura4-telomeric marker, hat1 ∆) was generated by crossing FY1872 with KTP24. Tetrad

analysis was performed. Colonies exhibiting a uracil prototrophy, G418 resistance, and

an inability to produce ß-galactosidase (white colonies in ß-galactosidase filter lift assay)

were selected.

ß-galactosidase filter lift assay

Dissected tetrads were replica plated directly to a BioTrace NT Pure

Nitrocellulose Transfer Membrane (0.2 µM, 82 mm; Pall Life Sciences). The membrane

was submerged in liquid nitrogen for ~ 1 min, removed, and allowed to thaw. A circular

piece of blotting paper (~ 90 mm diameter) was placed in a petri dish containing 2.5 ml Z

buffer (60 mM Na 2HPO 4·7H 2O, 40 mM NaH 2PO 4·H 2O, 1mM MgSO 4·7H 2O, 10 mM KCl,

54

0.26% 2-mercaptoethanol) and 50 µl X-gal (40 mg/ml). The thawed membrane was

positioned on the blotting paper with the cells facing upward. The petri dish containing

the blotting paper and membrane were incubated at 30 0C for 30 min before scoring the

results.

Telomeric and centromeric silencing assays

Telomeric (KTP36) and centromeric marker (KTP29, KTP33, and KTP35) strains

along with indicated controls were used to inoculate overnight cultures in YEA. Cultures

were re-suspended at 2 x 10 6 cells/1000 µl. 5-fold serial dilutions of the suspensions were made and 10 µl of each dilution were spotted on YEA, YEA containing 0.04% 5-

FOA, EMMG, and EMMG-ura plates (Moreno et al. 1991). The plates were incubated at

30 0C for 2-3 days.

The effect of Trichostatin A (TSA) on telomeric silencing was tested as follows: pre-cultures (3 ml) of FY1872 and 972 were grown overnight in YEL at 30 0C. New 3 ml

YEL cultures were seeded with 6.2 x 10 5 cells from one of the pre-cultures. Each culture was treated as follows: FY1872 was treated with TSA (50 µg/ml, Wako) or Optima methanol (volume equal to TSA, Fisher Scientific). 972 was treated with Optima methanol (volume equal to TSA) (Ekwall et al. 1997). Cells were grown ~ 44 hrs (~ 3 doublings in TSA, methanol treated cells were re-seeded if overgrown). Prior to counting on a hemacytometer, cells were sonicated on a Sonifier 250 (Branson) at 90% duty cycle; output setting 1 for 5 sec. Cells were re-suspended at 2 x 10 6 cells/1000 µl.

2.5- to 5-fold serial dilutions of the suspensions were made and 5 µl of each dilution

55

were spotted on EMMG (minimal glutamate), EMMG-ura, and 5-FOA (1g/L) plates. The plates were incubated 2 to 4 days at 30 0C and photographed.

Chromatin Immunoprecipitation

Cells were grown to 1 x 10 7 cells/ml in 50 ml of YEL. The culture was incubated in 3% paraformaldehyde for 30 min at 30 0C with gentle shaking. Glycine was added to 0.125 M concentration. After centrifugation for 5 min at 3000 rpm, the cells were washed 3 times with 1 ml ice-cold PBS containing 50 mM sodium butyrate before being centrifuged at 5000 rpm for 5 min. The pelleted cells were re-suspended in 400 µl of ice-cold lysis buffer (50 mM Hepes-KOH pH 7.5, 140 mM NaCl, 1 mM EDTA, 1%

Triton X-100, 0.1% DOC, 1 mM PMSF, Complete tablet, EDTA free) containing 23.5 µM depudecin, 10 µg/ml TSA, and 50 mM sodium butyrate. The cells were combined with

0.6 g of ice-cold glass beads and lysed 3 x 5 min in a mini bead beater (Biospec

Products). The base of the tube was punctured, placed on top of a 5 ml round bottom glass tube, and centrifuged for 1 min at 1500 rpm. Lysis buffer was added to bring the volume of the lysate up to 750 µl. The lysate was sonicated 12 times on ice using a

Sonifier 250 for 5 sec each at 90% duty cycle to shear the chromatin to ~ 600 bp. After centrifuging 3 x 15000 rpm for 5 min at 4 oC, the lysate was pre-cleared for 1 hr at 25 oC with inversion and centrifuged 2 x 2000 rpm for 1 min at 4 oC. An aliquot (70 µl, input fraction) was removed from the lysate and adjusted to 50 mM Tris pH 8.0, 10 mM EDTA, and 1% SDS final concentration. The input fraction was incubated overnight at 65 0C, allowed to cool to room temperature, ethanol precipitated, and vacuumed dried. The

56

pellet was re-suspended in 100 µl of a solution containing 50 mM Tris pH. 8.0 and 1%

SDS. 8 µl of 20 mg/ml Proteinase K was added to the sample prior to a 2 hr incubation

at 37 oC. The input fraction was extracted twice with phenol/chloroform, ethanol

precipitated and re-suspended in 20 µl TE.

For ChIPs performed using α-acH4 K12 (Millipore) or purified rabbit IgG

(Bethyl Labs), 5 µl of anti-serum were added to equal aliquots of remaining lysate and

incubated at 37 oC for 2 hrs with rotation. 30 µl of hydrated Protein A Sepharose beads

(GE Healthcare) were added and the beads were incubated at 37 0C for 1 hr with

rotation. For the ChIPs performed using α-H4 K5/12ac and α-H4 K8/16ac, equal aliquots of lysates were added to 30 µl of Protein A Sepharose beads containing immobilized anti-serum or rabbit non-immune serum. The samples were incubated at 37 oC for 2 hrs with rotation. Immunocomplexes were centrifuged 2 x 5000 rpm for 1 min and washed with the following: 1 ml lysis buffer containing 0.5 M NaCl, 50 mM sodium butyrate, and

10 µg/ml TSA; 1 ml lysis buffer containing 0.5 M NaCl and 50 mM sodium butyrate; 1 ml wash buffer (10 mM Tris-HCL pH 8.0, 0.25 M LiCl, 0.5% NP-40, 0.5% DOC, 1 mM

EDTA, 50 mM sodium butyrate); 1 ml TE. 65 µl of TES (50 mM Tris-Hcl pH 8.0, 10 mM

EDTA, 1% SDS) were added to the beads. The beads were incubated at 65 o C for 15 min and centrifuged 2 x 5000 rpm for 1 min at 4 oC. The supernatant was saved and incubated overnight at 65 oC. The sample was cooled to room temperature before adding 8 µl of 20 mg/ml Proteinase K and incubating the sample at 37 oC for 2 hrs. The

57

ChIP DNA (bound fraction) was extracted twice with phenol/chloroform, ethanol

precipitated and re-suspended in 20 µl TE.

Real-time polymerase chain reaction (PCR)

To amplify immunoprecipitated ura4 DNA sequence, real-time PCR was performed as follows: 0.5 µM of PCR primers Ura4Chip-F (5'-CAAGGCCTCAAAGAAG-

TTGG-3') and Ura4Chip-R (5'-GATGATATCGCTACCG-CAG-3') were combined with 9

µl of 2.5x Real Master Mix SYBR ROX (5 Prime), and 1 µl of ChIP DNA (diluted 1:5 water) in a PCR tube. The final volume was brought up to 20 µl with PCR grade water.

Real-time PCR reactions were run on a Mastercycler ep realplex 2 (Eppendorf) under the following conditions: 1 cycle at 95 0C for 2 min; 40 cycles at 95 0C for 30 sec, 57 0C for

30 sec, and 72 0C for 1min; 1 cycle at 95 0C for 15 sec; 1 cycle at 60 0C for 15 sec; ramp for 20 min; 1 cycle at 95 0C for 15 sec.

To amplify immunoprecipitated fus1 DNA sequence, real-time PCR was performed as follows: 0.5 µM of PCR primers fus1F (5'-AGAG-CACAACCCCGTCC-3') and fus1R (5'-TTTGCTATTGGTAGTACCGTAG-CC-3') were combined with 2.5x Real

Master Mix SYBR ROX, and ChIP DNA as described above. Real-time PCR conditions were as follows: 1 cycle at 95 0C for 2 min; 40 cycles at 95 0C for 30 sec, 65 0C for 30 sec, and 72 0C for 1min; 1 cycle at 95 0C for 15 sec; 1 cycle at 60 0C for 15 sec; ramp for

20 min; 1 cycle at 95 0C for 15 sec.

Analysis of real-time PCR data was performed as follows: triplicate C t values for input and bound fractions were averaged from ChIPs using anti-serum or non-immune

58

serum. Average net C t for ura4 and fus1 primers were calculated by subtracting the

average input C t from the average bound C t for the two primers. For each individual

primer, the average net C t for FY1872 was subtracted from the average net Ct of KTP36

to give the average net C t difference between the wildtype strain and hat1 ∆. The

negative of this value was used as an exponent for the base 1.9 to calculate the relative

level of immunoprecipitated acetylated histones at the ura4 or fus1 DNA sequence.

These values were graphed on Excel software (Microsoft). A summary of the

calculations is given by the equations below:

Avg. net C t ura4 = Avg. bound C t ura4 - Avg. input C t

Avg. net C t fus1 = Avg. bound C t fus1 - Avg. input C t

Avg. net C t ura4 difference = Avg. net KTP36 C t ura4 - Avg. net FY1872 C t ura4

Avg. net C t fus1 difference = Avg. net KTP36 C t fus1 - Avg. net FY1872 C t fus1

Relative level of IP histones at ura4 in hat1 ∆ over WT = 1.9 - Avg. net Ct ura4 difference

Relative level of IP histones at fus1 in hat1 ∆ over WT = 1.9 - Avg. net Ct fus1 difference

F-tests were performed to determine if the variances from the ChIPs using anti-serum are significantly different (< 0.05) from the ChIPs using non-immune serum. T-tests

(two-sample unequal variances or two-sample equal variances) were performed to

59

determine if difference between the ChIPs using anti-serum are significantly different (<

0.05) from the ChIPs using non-immune serum. Both F-tests and T-tests were

performed using Microsoft Excel.

S. pombe cell cycle assays

A 10 ml pre-culture of 972 was grown at 30 0C to ~ 10 7 cells/ml. 8 x 10 6 cells

were pelleted and washed with 1 ml EMM-Nitrogen (MP Biomedicals). Cells were

synchronized by re-suspending in 10 ml EMM-Nitrogen and incubating at 30 0C for 38

hrs. Two aliquots (10 7 cells each) were removed, washed with 1 ml YEL, and pelleted.

Each pellet was re-suspended in 10 ml YEL containing either 55 mM sodium butyrate or

55 mM sodium chloride. The cultures were incubated at 30 0C. Alternatively, asynchronous cells were also exposed to 55 mM NaButyrate and 50 µg/ml TSA for 48 hours at 30 0C. A 300 µl (~ 10 6 to 10 7 cells) aliquot was removed every half hour for 8

hrs and fixed with 1 ml cold 70% ethanol prior to analysis by flow cytometry.

Flow cytometry

S. pombe cells to be analyzed by flow cytometry were pelleted and fixed with

cold 70% ethanol (1 ml). 0.3 ml of fixed cells (~10 5 to 10 7 cells) were removed, placed in

a 15 ml centrifuge tube (Corning), and washed with 3 ml of 50 mM sodium citrate. The

cells were pelleted, re-suspended in 0.5 ml of 50 mM sodium citrate containing 0.1

mg/ml RNase A (Sigma Aldrich), and incubated in a 37 0C water bath for 2 hrs. To

disrupt cell clusters, the cell suspension was sonicated at 100% duty cycle; output

setting 1 for 45 sec. A 250 µl aliquot was removed as an unstained control. 0.25 ml of

60

50 mM sodium citrate containing 8 µg/ml propidium iodide (Sigma Aldrich) was added to each sample. The samples were incubated on ice for ~ 10 min and analyzed on a BD

FACS Canto flow cytometer with BD FACS Diva software (BD Biosciences).

61

Results

Effect of hat1 ∆ on S. pombe telomeric silencing

The deletion of hat1 in S. cerevisiae results in an increased sensitivity to MMS and loss of telomeric silencing when combined with mutations in certain acetylatable lysines of histone H3 (Kelly et al. 2000; Barman et al. 2006). S. pombe hat1 ∆ mutants,

however, exhibit a similar sensitivity to MMS without additional mutations of H3 (Benson

et al. 2006). To determine if a deletion of hat1 alone results in a loss of telomeric

silencing, a hat1 ∆ S. pombe mutant (KTP36) was generated, containing ura4 integrated into a sub-telomeric region of the chromosome that is normally silenced to repress transcription (Manolis et al. 2001). The loss of silencing was monitored by the sensitivity of these strains on medium containing 5-FOA and also by the growth of these strains on medium lacking uracil.

To determine if the loss of hat1 affects transcription in euchromatic regions,

silencing assays were performed using a hat1 ∆ S. pombe mutant (KTP35) with ura4

integrated into a euchromatic region of the chromosome that is not silenced to repress

transcription and a control strain (FY941) containing the euchromatic marker alone

(Allshire et al. 1995). FY941 did not grow on medium containing 5-FOA (Fig 7A, panel

A). As seen with KTP35, the deletion of hat1 did not restore growth on 5-FOA

suggesting that hat1 is not involved in the repression of transcription in euchromatic

regions.

62

In contrast to the above data, the hat1 ∆ S. pombe mutant KTP36 containing a

ura4 subtelomeric marker was more sensitive to 5-FOA than FY1872, a parental strain

containing the subtelomeric marker alone (Fig. 7A, panel B). Wildtype and LBP6

negative control strains did not grow on 5-FOA due to endogenously expressed Ura4.

As a positive control for growth, the other parental strain (KTP24) with a deletion of hat1 and a disruption in ura4 was also tested and displayed the least sensitivity to 5-FOA. On medium lacking uracil, the growth of KTP36 was comparable to LBP6 and the wildtype strain, both of which contain endogenous Ura4 (Fig. 7B). The growth of all strains on complete medium was similar indicating that the effects seen with 5-FOA medium and medium lacking uracil were not due to inconsistencies in cell number. Interestingly, the loss of silencing phenotype occurred upon the deletion of hat1 alone and did not require the concurrent mutation of the H3 N-terminal domain as seen in S. cerevisiae (Kelly et

al. 2000).

It is formally possible that the loss of silencing phenotype observed with KTP36

on 5-FOA may not have been caused by the deletion of hat1 , but rather a mutation that

occurred during the construction of this strain by homologous recombination. In order to

exclude this possibility, KTP36 was transformed with a plasmid containing hat1

controlled by an nmt1 promoter. Transformants were selected based on their ability to

grow on EMM-leu. Spot assays on 5-FOA revealed that the expression of exogenous

Hat1 restored telomeric silencing in KTP36 (Fig. 8). A control transformation with empty

vector alone also allowed for some growth on medium containing 5-FOA, but not to the

63

same degree as the transformation with the hat1 plasmid. These results combined with those showing a loss of telomeric silencing upon deletion of hat1 suggest Hat1 to be essential for the establishment of proper telomeric silencing in S. pombe .

Effect of hat1 ∆ on S. pombe centromeric silencing

To determine the effect of the loss of hat1 on centromeric silencing, three hat1 ∆

S. pombe mutants were generated. Each strain contained ura4 integrated into one of the following centromeric regions: the central core domain (KTP33), inner most repeats

(KTP29), and outer repeats (KTP30). These centromeric regions are normally silenced to repress transcription (Allshire et al. 1995). The effect of hat1 deletion on S. pombe centromeric silencing was monitored by the sensitivity of these strains on 5-FOA medium.

The deletion of hat1 did not alter the sensitivity of KTP29 and KTP30 on 5-FOA medium (Fig. 9, panels B and C). Both strains grew similarly to their respective parent strains (FY496 and FY648) containing wildtype hat1 , suggesting that Hat1 does not have a significant role in silencing at the inner or outer repeat regions of the centromere. In contrast, when hat1 was deleted from the centromeric central core ura4 marker strain, the growth of this strain (KTP33) on medium containing 5-FOA was more robust than the parental strain (FY336) containing wildtype hat1 (Fig.9, panel A).

The relative enrichment of acetylated histones at subtelomeric ura4 DNA sequences

The results presented thus far suggest that the loss of hat1 results in the loss of telomeric silencing. Previous studies have shown native Hat1 isolated from S.

64

cerevisiae to be capable of acetylating H4 lysine 12 and the recombinant enzyme capable of acetylating lysines 5 and 12 (Kleff et al. 1995; Parthun et al. 1996). As such, one might predict a lower level of histone acetylation, especially at lysines 5 and 12 of histone H4, in the subtelomeric region of a hat1 ∆ mutant relative to a strain wildtype for hat1 . To test this prediction, chromatin immunoprecipitation assays were performed on

KTP36 and FY1872 using various acetylated histone H4 antibodies and primers specific

for ura4 and fus1 . Interestingly, ChIPs performed on the hat1 ∆ mutant using antibodies

against H4 acetylated at K5 and/or K12 displayed an increased enrichment of this

acetylated H4 at the subtelomeric ura4 DNA sequence relative to the non-deletion strain

(Fig. 10A). This relative enrichment at the ura4 DNA sequence was 3- to 6-fold greater

than that observed with control ChIPs using rabbit normal immune serum. In contrast to

the results at the subtelomeric ura4 DNA sequence, ChIPs performed on the hat1 ∆

mutant did not reveal a significant increase in the relative enrichment of acetylated H4 at

K5/K12 at the non-subtelomeric fus1 DNA sequence. The level of relative enrichment

observed was similar to the relative enrichment at the ura4 and fus1 sequences

following control ChIPs using rabbit normal immune serum.

Similar results were obtained with ChIPs using antibodies against

monoacetylated H4 at K12. The relative enrichment of this monoacetylated H4 at the

ura4 DNA sequence was 6- to 15-fold greater than that observed with control ChIPs

using rabbit normal immune serum (Fig. 10B). Similar to the results from the ChIPs

using diacetylated H4 antibodies, the relative enrichment of monoacetylated H4 K12 at

65

the fus1 DNA sequence was not significantly different from control ChIPs using rabbit

normal immune serum.

To determine if the deletion of hat1 affects the relative enrichment of histone H4

acetylated at non-substrate lysine residues, ChIPs were performed using antibodies

against H4 acetylated at K8 and/or K16. Surprisingly, an increased relative enrichment

of H4 acetylated at K8/K16 was observed at the subtelomeric ura4 DNA sequence (Fig.

11). This relative enrichment was 2- to 4-fold greater than that observed with control

ChIPs using rabbit normal immune serum. The relative enrichment of H4 acetylated at

K8/K16 at the fus1 DNA sequence was not significantly different from control ChIPs using rabbit normal immune serum. The combined results from the ChIP experiments suggest that the deletion of hat1 leads to an increase in the acetylation of histone H4 at genes located in subtelomeric regions at multiple acetylatable sites in the H4 tail domain.

Effect of histone hyperacetylation on telomeric silencing

Trichostatin A (TSA) is a known inhibitor of histone deacetylases (HDACs) in

humans and other eukaryotes. Studies have shown that S. pombe cells treated with

TSA for 5 and 10 doublings display increased levels of histone acetylation (Ekwall et al.

1997).

To determine the effect of global histone hyperacetylation on telomeric silencing

in S. pombe , 2.5 fold serial dilutions of FY1872 cells treated with TSA were spotted on

EMMG, EMMG-ura, and 5-FOA plates. Similar to the hat1 ∆ mutant, treatment of the telomeric ura4 marker strain with TSA for approximately 4-5 generations resulted in a

66

growth defect on 5-FOA plates relative to the untreated control (Fig. 12A). The degree

of 5-FOA sensitivity was similar to a wildtype strain expressing native Ura4. On EMMG-

ura plates, cells treated with TSA exhibited a more robust growth relative to untreated

cells (Fig. 12B). However, this growth was less than the growth observed with a

wildtype strain. The growth of all strains on EMMG was comparable indicating that the

observed effects were not due to inconsistencies in cell number.

Effect of histone deacetylase inhibitors on S. pombe cell cycle

Studies have shown TSA and sodium butyrate to arrest HeLa cells in S phase

and early G1 phase respectively (Xue and Rao 1981; Littlefield et al. 1982; Toth et al.

2004). To determine the effects, if any, of sodium butyrate on the cell cycle of S. pombe , wildtype cells were starved to allow for their accumulation in G1 phase. The release of these cells in the presence of sodium butyrate resulted in a delay in cell cycle release of approximately 30 min relative to untreated cells (Fig. 13). However, sodium butyrate did not arrest S. pombe cells at a specific cell cycle phase during the 8-hour time course.

To determine the effect on fission yeast of prolonged exposure to HDAC inhibitors, asynchronous S. pombe cells were exposed to either 55 mM sodium butyrate

or 50 µg/ml TSA for 48 hours. After 48 hours, control cells treated with either NaCl or

MeOH exhibited a 2C peak similar to that of untreated control cells (Fig. 14, panel A). In

contrast, cells treated with sodium butyrate or TSA displayed a broader peak indicative

of a cell population having a greater than 2C DNA content (Fig. 14, panel B). S. pombe

typically have a very short period during the cell cycle during which cells have more than

67

a 2C DNA content (one cell with two late S-phase nuclei) (Gomez and Forsburg 2004).

This period, however, is not generally detectable by flow cytometry. The appearance of a peak representing a greater than 2C DNA content suggests a possible block in late S- phase resulting from the treatment with sodium butyrate and TSA.

68

Figure 7: hat1 deletion results in a loss of telomeric silencing in S. pombe .

Wild type and experimental yeast strains were cultured on ( A) YEA plates in the presence (5-FOA) or absence (YEA) of added 5-FOA and (B) EMMG plates in the presence (EMMG) or absence (EMMG-ura) of added uracil. Spot cultures represent 5- fold dilutions. Cells were grown for two (YEA and EMMG) or three (5-FOA and EMMG- ura) days at 30°. FY941 : ura4-euchromatic marker , KTP35 : ura4-euc, hat1 ∆; FY1872 : ura4-telomeric marker ; KTP36 : ura4-tel, hat1 ∆; KTP24 : ura4-disrupted, hat1 ∆; LBP6 , hat1 ∆. Note: the ura4 + gene is silent in FY1872, but expressed in KTP36. Also, deletion of hat1 + has no effect on silencing of the endogenous ura4 + gene ( LBP6 ).

69

Figure 7

70

Figure 8: Hat1 expressed from a plasmid restores telomeric silencing. The ura4 - telomeric marker, hat1 strain ( KTP36 ) was transformed with an empty vector ( EV ) or a plasmid containing hat1 + controlled by an nmt1 promoter ( phat1 ). Transformed cells were cultured on EMMG plates in the presence or absence of 5-FOA. Spot cultures represent 5-fold serial dilutions. Cells were grown for two (EMMG) or three days (5-

FOA) at 30 0C. FY1872 : ura4-telomeric marker ; KTP36 : ura4-tel, hat1 ∆; KTP24 : ura4- disrupted, hat1 ∆. Note: The expression of exogenous Hat1 restored telomeric silencing in KTP36.

71

Figure 8

72

Figure 9: hat1 deletion does not result in the loss of centromeric silencing in S. pombe.

Wild type and experimental yeast strains were cultured on YEA plates in the presence

(5-FOA) or absence (YEA) of added 5-FOA. Spot cultures represent 5-fold dilutions.

Cells were grown for two (YEA and EMMG) or three (5-FOA and EMMG-ura) days at

30°. FY336: ura4-centromeric central core marker ; KTP33: ura4-centromeric central core marker, hat1 ; KTP25: h+ ura4-disrupted, hat1 ; FY496: ura4-centromeric inner

most repeat marker ; KTP29: ura4-centromeric inner most repeat marker, hat1 ; KTP24:

h- ura4-disrupted, hat1 ; FY648: ura4-centromeric outer repeat marker ; KTP30: ura4-

centromeric outer repeat marker, hat1 ; LBP6: hat1 . Note: The growth of KTP33 on

medium containing 5-FOA was more robust than the parental strain (FY336) containing

wildtype hat1 +.

73

Figure 9

74

Figure 10: The increase in relative enrichment of acH4 K5/K12 at the subtelomeric ura

DNA sequence in the hat1 deletion strain over the wildtype strain. Chromatin

immunoprecipitation assays were performed on ura4 -telomeric marker ( FY1872 ) and

ura4 -telomeric marker , hat1 ∆ (KTP36 ) S. pombe strains using antibodies against ( A) H4

acetylated at K5/K12 (acH4 K5/K12) and ( B) H4 acetylated at K12 (acH4 K12). RT-PCR

was performed using primers specific for ura4 and fus1. The average net C t difference

between the hat1 ∆ and wildtype strain was used to calculate the relative enrichment of immunoprecipitated acetylated histones at the ura4 or fus1 DNA sequence. Note : The

loss of hat1 resulted in the increase of acH4 K5/K12 at the subtelomeric ura4 DNA

sequence.

75

Figure 10

76

Figure 11: The increase in relative enrichment of acH4 K8/K16 at the subtelomeric ura

DNA sequence in the hat1 deletion strain over the wildtype strain. Chromatin immunoprecipitation assays were performed on ura4 -telomeric marker ( FY1872 ) and ura4 -telomeric marker , hat1 ∆ (KTP36 ) S. pombe strains using antibodies against H4 acetylated at K8/K16 (acH4 K8/K16). RT-PCR was performed using primers specific for

ura4 and fus1. The average net C t difference between the hat1 ∆ and wildtype strain was used to calculate the relative enrichment of immunoprecipitated acetylated histones at the ura4 or fus1 DNA sequence. Note : The loss of hat1 resulted in the increase of acH4

K8/K16 at the subtelomeric ura4 DNA sequence.

77

Figure 11

78

Figure 12: Hyperacetylation of histones results in the loss of telomeric silencing. Strains of S. pombe (FY1872 and 972 ) were cultured in medium containing (+ TSA) or lacking (-

TSA) trichostatin A and allowed to grow for approximately three generations at 30 0C.

Spot cultures represent 2.5 fold serial dilutions. Cells were grown for two to four days at

30 0C on EMMG (two to three days), 5-FOA (two days), and EMMG-ura (four days).

FY1872 : ura4-telomeric marker ; 972 : wildtype

79

Figure 12

80

Figure 13: Treatment of synchronized S. pombe cells with sodium butyrate results in a cell cycle delay upon release. Wildtype ( 972 ) S. pombe cells were nitrogen starved in

EMM-nitrogen medium for 38 hours at 30 0C to allow for their accumulation in G1 phase.

Cells were subsequently released in the presence or absence of sodium butyrate (55 mM final concentration). An aliquot was removed every half hour for 8 hrs and analyzed by flow cytometry. Note: The presence of sodium butyrate resulted in a delay in cell

cycle release of approximately 30 min relative to untreated cells. However, sodium

butyrate did not arrest S. pombe cells at a specific cell cycle phase during the 8 hr time

course.

81

Figure 13

82

Figure 14: The accumulation of S. pombe cells with greater than 2C DNA content following extended exposure to histone deacetylase inhibitors. Asynchronous S. pombe

(972 ) cells were cultured in the presence or absence of histone deacetylase inhibitors

(55 mM sodium butyrate or 50 µg/ml TSA) for 48 hours and analyzed by flow cytometry.

Note: The treatment of asynchronous S. pombe cells with sodium butyrate or TSA for 48 hours results in the accumulation of cells with greater than 2C DNA content.

83

Figure 14

84

Chapter Three

The deposition of H2A and H2B in human cells

Summary

During histone nucleosome assembly, virtually all of new H3 and H4 are

deposited at the replication fork. New H2A and H2B are also deposited into

nucleosomes, but there is evidence that considerable deposition takes place on non-

replicating chromatin. In the presence of hydroxyurea, a DNA synthesis inhibitor, only

new H2A and H2B are deposited into chromatin. One postulated mechanism for this is

H2A/H2B exchange. To test this hypothesis histones were radiolabeled in the presence

or absence of either hydroxyurea (HU), or 5,6-Dichloro-1-β-D- ribofuranosylbenzimidazole (DRB), an inhibitor of transcription. Chromatin immunoprecipitations (ChIPs) were performed using antibodies that recognize H4 acetylated at either lysines 5 and 12 (which are acetylated in newly synthesized H4), or at lysines 8 and 16 (which are acetylated in transcriptionally active chromatin, but not in new H4). ChIPs of chromatin labeled in the presence of hydroxyurea or DRB indicate that transcription is involved in the exchange of H2A and H2B in acetylated chromatin regions. However, H2A and H2B are deposited at a low level even when transcription is inhibited. Additional experiments showed that cytosolic H2A is acetylated. This cytosolic acetylated H2A is not associated with cytosolic acetylated H4. We postulate that this H2A has been exchanged out of transcriptionally active chromatin regions.

85

Materials and Methods

HeLa cell culture, inhibitor treatments, and radiolabeling

Spinner culture of HeLa cells were maintained at 37 0C in minimal essential medium (MEM; Joklik modification) supplemented with 10% calf serum (Lonza), 1% L- glutamine (20mM, Gibco), and 0.5% penicillin-streptomycin (Gibco). HeLa cells were harvested and re-suspended in 10 to 20 ml of lysine-free MEM containing 5% calf serum. RNA synthesis was inhibited with the addition DRB to 40 µM final concentration followed by an incubation at 37 0C for 12 min. DNA synthesis was inhibited with the addition of HU to 10 mM final concentration followed by an incubation at 37 0C for 10 min. After incubation, either TSA (1 µM final concentration) or sodium butyrate (50 mM final concentration) was added to inhibit histone deacetylation. Cells were labeled with

50 µCi/ml [ 3H]lysine (1 µCi/µl, Perkin Elmer) at 37 0C for 10 min.

HeLa Nuclear isolation

To isolate HeLa nuclei, 200 to 300 ml of harvested cells were washed twice with

Buffer A (10 mM Tris, 2 mM 2-mercaptoethanol, 2 mM MgCl 2, 10 mM sodium butyrate,

pH 7.6), re-suspended in 40 ml Buffer A, and incubated on ice for 15 min. Cells were

lysed with a Dounce homogenizer. Nuclei were pelleted and washed with Buffer A.

Micrococcal nuclease digestion and immunoprecipitation of non-fixed chromatin

Optical density readings of isolated nuclei were taken on a DU-50

spectrophotometer (Beckman) in 1% SDS. Isolated nuclei were pelleted and re-

suspended in PIPES buffer (10 mM PIPES, 80 mM NaCl, 20 mM sodium butyrate, pH

86

7.0) at 40 OD/ml. Calcium chloride (CaCl 2) and micrococcal nuclease (Mnase) were added to 0.5 mM and 5 units/ml final concentrations, respectively. Adding 0.1 M EGTA to 5 mM final concentration and incubating the reaction on ice for 30 min terminated

Mnase digestion. The soluble chromatin fraction (S1) was collected after centrifugation and optical density readings were taken on the sample. The S1 was adjusted as follows:

10% Triton X-100 to 0.25% concentration, 100 mM DTT to 0.5 mM concentration, 100 mM PMSF to 1 mM final concentration, 100 mM EGTA to 1 mM final concentration, pH

8.0.

Protein A Sepharose beads were re-hydrated in high salt TSE (0.1% SDS, 1%

Triton X-100, 2 mM EDTA, 500 mM NaCl, 20 mM Tris, pH 8.1), washed twice with 1 ml high salt TSE, and washed twice with 1 ml low salt TSE (0.1% SDS, 1% Triton X-100, 2 mM EDTA, 150 mM NaCl, 20 mM Tris, pH 8.1). Beads were re-suspended 1:1 (vol:vol) with low salt TSE. Approximately 100 µl of beads were collected and incubated with

~100 µl rabbit non-immune serum or anti-serum (2240, anti-acH4 K5/K12; 6815, anti- acH4K8/K16) at 37 0C for 90 min. Beads with immobilized anti-serum were collected and washed 3 x 5 min with 0.5 ml low salt TSE containing 10 µl of 100 mM PMSF.

0.2 to 0.4 OD of S1 was added to beads with immobilized anti-serum. A separate aliquot was removed as the input fraction. Samples were incubated overnight at 4 0C with rotation followed by an additional 2 hr incubation at room temperature.

Samples were centrifuged at 2000 rpm for 30 sec. The input and unbound fraction

(supernatant) were adjusted to 10 mM MgCl 2 and ethanol precipitated. The bound

87

fraction (beads) was washed 5 x 10 min with 0.5 ml high salt TSE and pelleted. Input,

unbound, and bound fractions were vacuumed dried and re-suspended in SDS sample

buffer.

Preparation of HeLa cytosolic extract (S100)

After harvest, HeLa cells were pelleted and washed twice with Buffer A, twice

with HB buffer (20 mM HEPES free acid, 5 mM KCl, 1.5 mM MgCl 2, 10 mM sodium butyrate, pH 7.5), and re-suspended in 0.5 ml HB buffer containing 0.5 mM DTT. After a

15 min incubation on ice, 1.5 ml of HB buffer was added. Cells were lysed in a Dounce homogenizer and incubated 30 min on ice. The lysate was centrifuged at 12000 rpm for

10 min at 4 0C. The supernatant was removed and centrifuged at 12000 rpm for 5 min at

4 0C. The supernatant was saved and centrifuged at 43000 rpm for 90 min at 4 0C in a

Beckman TL-100 Ultracentrifuge using a Beckman TLA 45 rotor. The resulting

supernatant (S100) was adjusted to 10 mM MgCl 2, incubated on ice for 30 min, and centrifuged at 10000 rpm to remove residual chromatin.

Immunoprecipitation of S100

As described previously, Protein A Sepharose beads were hydrated and

immobilized with the following anti-serum: anti-acH4 K5/K12 (7481) and rabbit IgG

(Bethyl Laboratories). S100 was adjusted as follows: 100 mM PMSF to 1 mM

concentration, 100 mM DTT to 0.5 mM concentration, 100 mM EGTA to 1 mM

concentration, 10% Triton X-100 to 0.25% concentration, 0.1 mg/ml leupeptin to 0.6

µg/ml concentration, 0.1 mg/ml pepstatin to 0.8 µg/ml concentration, 0.1 mg/ml

88

mycrocystein to 0.5 µg/ml concentration. Equal aliquots of adjusted sample (~ 500 µl)

were added to the beads with immobilized anti-serum and non-immune serum. Samples

were incubated overnight with rotation at 4 0C followed by a 2 hr incubation at room temperature. Samples were centrifuged at 2000 rpm for 30 sec. The unbound fraction

(supernatant) was adjusted with 18 M sulfuric acid to 0.2 M final concentration and incubated overnight at 4 0C. Samples were centrifuged at 14000 rpm for 10 min at 4 0C.

The supernatant was adjusted to 25% TCA, incubated on ice for 30 min, and centrifuged at 14000 rpm for 10 min at 4 0C. Pellets were washed with acidified acetone (0.05 N

HCl) and then acetone. Protein pellets were vacuum dried and re-suspended in SDS sample buffer.

The bound fraction (beads) was washed 5 x 10 min at room temperature with high salt TSE followed by one wash with 10 mM Tris, pH 8.1 for 5 min at room temperature. The supernatant was removed after centrifugation at 2000 rpm for 30 sec.

The beads were dried by vacuum and re-suspended in SDS sample buffer.

Gel electrophoresis, fluorography, and immunoblotting

To separate acetylated histone isoforms, proteins were subjected to SDS-PAGE

on 18% polyacrylamide gels (Thomas and Kornberg 1975) as previously described

(Perry and Annunziato 1989). Gels were stained overnight with Coomassie blue and

destained with Coomassie destain (20% methanol, 10% acetic acid). To prepare for

fluorography, gels were washed 3 x 30 min in 600 ml DMSO followed by a 90 min wash

89

in 600 ml DMSO/PPO. Gels were washed 30 min in distilled water, dried, and exposed to Kodak Biomax XAR film (Bonner and Laskey 1974) ((Laskey and Mills 1975).

For immunoblotting, resolved proteins were transferred to Immobilon-P as described previously. Membranes were stained with Ponceau-S, destained, equilibrated with TBS, and sliced to separate the various histone regions. Membranes were blocked for 1 hour at room temperature using blocking buffer containing 10% goat serum and 5% BSA. Anti-acH4 K5/K12 (7481) was diluted in dilution buffer (5% goat serum, 3% BSA) at 1:5000. Membranes were incubated with primary antibody for 2 hrs at room temperature and subsequently washed 3 x 15 min with TBS. Alkaline phosphatase conjugated secondary antibody was diluted 1:5000 in dilution buffer and applied to the membranes for 1 hr at room temperature. The membranes were washed

3 x 15 min with TBS. Antibody reactions were visualized by a color reaction following the addition of a solution containing BCIP/NBT tablets (Sigma Aldrich).

Uridine incorporation assays

HeLa cells were harvested, spilt into two aliquots, and re-suspended in 5 ml pre- warmed MEM containing 0.1% calf serum and either 40 µM DRB or ethanol (100% stock, same volume as DRB). Cells were pre-incubated 20 min at 37 0C, pelleted, and re-suspended again in the same medium. After 5 min, 12.5 µCi of [ 3H]uridine (1 µCi/µl,

Perkin Elmer) was added to the cells. A 100 µl aliquot was removed from the non-DRB treated cells and suspended in 1 ml Buffer A containing 10% sodium azide at the following time intervals: 2, 4, 6, 8, 10, 12, 15, 20, 25, 30, and 35 min. A 100 µl aliquot

90

was removed from the DRB treated cells and suspended in 1 ml Buffer A containing

10% sodium azide at the following time intervals: 2.5, 4.5, 6.5, 8.5, 10.5, 12.5, 15.5,

20.5, 25.5, 30.5, and 35.5 min. Samples were centrifuged 2 min at 1500 rpm and re- suspended in 1 ml distilled water. TCA (100%) was added to 10% concentration.

Samples were incubated 30 min on ice and filtered onto glass filters. The filters were dried and placed in glass scintillation vials. Ecoscint A (5 ml) was added and the samples were counted on a liquid scintillation counter.

Thymidine incorporation assays

HeLa cells were harvested, re-suspended in 20 ml pre-warmed MEM containing

10% calf serum and incubated at 37 0C. After 5 min, 100 µCi of [ 3H]thymidine (1 µCi/µl,

Perkin Elmer) were added to the cells. A 100 µl aliquot of cells was removed every minute for 6 minutes and suspended in 1 ml Buffer A containing 0.1% sodium azide. The cells were split into two equal aliquots. To one aliquot, 10 mM HU was added to 10 mM final concentration. To the other aliquot, pre-warmed MEM containing 10% calf serum was added (same volume as HU). A 100 µl aliquot was removed from the control cells and suspended in 1 ml Buffer A containing 10% sodium azide at the following time intervals: 8, 9, 10, 12, 15, and 20 min. A 100 µl aliquot was removed from the HU treated cells and suspended in 1 ml Buffer A containing 10% sodium azide at the following time intervals: 8.5, 9.5, 10.5, 12.5, 15.5, and 20.5 min. Samples were centrifuged 2 min at 1500 rpm and re-suspended in 1 ml distilled water. TCA (100%) was added to 10% concentration. Samples were incubated 30 min on ice, filtered onto

91

glass filters, dried, placed in glass scintillation vials with 5 ml Ecoscint A, and counted on a liquid scintillation counter.

92

Results

Deposition of newly synthesized histones

Previous studies have shown that during chromatin replication and assembly,

newly synthesized histone H4 is acetylated at lysines 5 and 12 prior to its deposition

onto nascent DNA (Chicoine et al. 1986; Sobel et al. 1995). To elucidate the deposition

pattern of newly replicated histones, HeLa histones radiolabeled with [ 3H]lysine were immunoprecipitated with antibodies specific for H4 acetylated at either K5/K12 or

K8/K16. Rabbit normal immune serum was used as a negative control. Almost all new

H4 was found in the bound fraction when immunoprecipitated with antibodies against H4 acetylated at K5/K12 (Fig 15). However, not all new H2A and H2B were immunoprecipitated with this same antibody. Some H2A and H2B remained in the unbound fraction. In contrast, immunoprecipitations performed using antibodies against

H4 acetylated at K8/K16 resulted in no new H4 in the bound fraction. However, some

H2A and H2B were immunoprecipitated by this antibody. The results from these immunoprecipitations indicate that not all new H2A/H2B is deposited together with new

H3/H4.

Role of replication and transcription in H2A/H2B exchange

To confirm the inhibition of DNA replication in S. pombe by HU, thymidine

incorporation assays were performed in the presence or absence of HU. Untreated cells

exhibited an increase in thymidine incorporation, while thymidine incorporation ceased in

93

the HU treated cells (Fig. 16). Uridine incorporation assays performed under the same

conditions demonstrate that HU does not prevent RNA synthesis (data not presented).

To inhibit DNA replication, HeLa cells were incubated in MEM containing 10 mM

HU prior to radiolabeling with [ 3H]lysine. The addition of HU reduced the amount of new

H3/H4 in chromatin (Input) and in the bound fraction of the immunoprecipitation using anti-H4 acetylated at K5/K12 (Fig. 17). H2A and H2B deposition, however, was relatively unaffected by the presence of HU. Immunoprecipitations using anti-H4 acetylated at K8/K16 exhibited a band in the bound fraction suggesting that considerable

H2A and H2B deposition takes place on non-replicating chromatin that is acetylated.

To confirm the inhibition of RNA synthesis by DRB, uridine incorporation assays were performed in the presence or absence of DRB. DRB essentially eliminated uridine incorporation by S. pombe (Fig. 18). DRB reduced DNA replication approximately 2-fold after 50 minutes of treatment after which time thymidine incorporation also was strongly inhibited (data not presented).

To determine the effect of DRB on histone deposition, HeLa cells were incubated in MEM containing 40 µM DRB prior to radiolabeling with [ 3H]lysine. Results using untreated control cells were similar to those observed previously (Fig. 19A).

Immunoprecipitation of DRB-treated cells using antibodies against H4 acetylated at

K5/K12 did not yield a significant reduction of H3 and H4 in the bound fraction suggesting that the deposition of new H3 and H4 was unaffected by the inhibition of

RNA synthesis (Fig. 19B). This is consistent with the deposition of new H3/H4 onto

94

newly replicated DNA. In contrast, immunoprecipitation using antibodies against H4 acetylated at K8/K16 revealed a reduction of H2A and H2B in the bound fraction. DRB appeared to inhibit the deposition of H2A and H2B, suggesting that transcription may be a factor involved in H2A/H2B exchange.

A previous study by the Annunziato laboratory has shown the association of free

H2A with the nucleosome assembly factor NAP-1 in HeLa cytosolic extracts (Chang et al. 1997). The report from that study suggested that NAP-1 is a factor involved in the transcription-mediumted exchange of H2A/H2B dimers and in replication-coupled chromatin assembly. A specific deposition related pattern of histone modification has not been observed for newly synthesized H2A (Jackson et al. 1976; Sealy and Chalkley

1979; Cousens and Alberts 1982). This finding, however, does not exclude the possibility that H2A/H2B may be post-translationally modified if H2A/H2B dimers are exchanged from transcriptionally active chromatin. To determine if cytosolic H2A can be acetylated, H2A in a cytosolic S100 fraction was immunoblotted using anti-H2A acetylated at K5 antibodies. Western blot analysis revealed that cytosolic H2A exhibits acetylation at K5 (Fig. 20A). To confirm that this cytosolic H2A is free and not associated in nucleosomes, H4 was immunoprecipitated from HeLa cytosolic extracts

(S100) using antibodies against H4 acetylated at K5/K12 or control rabbit normal immune serum. A similar immunoprecipitation was performed with isolated chromatin as a control. A previous study has shown the association of new H2A, H2B, and H3 with

H4 following the immunoprecipitation of intact nucleosomes using anti-acetylated H4

95

antibodies (Perry et al. 1993). Control immunoprecipitations of isolated chromatin demonstrated the association of acetylated nuclear H2A with acetylated nuclear H4 (Fig.

20B). In contrast, cytosolic acetylated H2A from the S100 remained in the unbound fraction and did not precipitate with cytosolic acetylated H4 (Fig. 20A). This finding confirms that acetylated H2A in the cytosol is not complexed in nucleosomes. Work from the Annunziato lab has shown that most cytosolic H2A is not acetylated (Barrows, unpublished results). This is consistent with the finding that newly synthesized H2A, which is also found in the S100, is not modified (Smith and Stillman 1991; Chang et al.

1997; Benson et al. 2006). This suggests that cytosolic acetylated H2A might originate from non-replicating acetylated chromatin, possibly through a process of transcription- mediumted H2A exchange.

96

Figure 15: Not all newly synthesized H2A/H2B is deposited together with new H3/H4.

HeLa cells were radiolabeled with [ 3H]lysine for 10 min. Radiolabeled histones were immunoprecipitated with antibodies specific for H4 acetylated at either K5/K12 or

K8/K16. Rabbit normal immune serum (RNIS) was used as a negative control. Almost all new H3 and H4 are found in the bound fraction when immunoprecipitated with an anti-H4 diacetylated on K5 and K12 antibody ( α-acH4 K5/K12). By comparison, not all

H2A and H2B are immunoprecipitated with this same antibody. Notably, a fraction of

H2A and H2B is immunoprecipitated with an anti-H4 diacetylated on K8 and K16

antibody ( α-acH4 K8/K16), independently of new H3/H4.

97

Figure 15

98

Figure 16: Thymidine incorporation in the presence of hydroxyurea. HeLa cells were incubated in vitro with 100 µCi of [ 3H]thymidine for 20 minutes. After 6 minutes of incubation, the cells were split into two aliquots. Hydroxyurea was added to one aliquot to 10 mM final concentration. At the indicated time intervals, aliquots were removed, filtered through glass filter circles, and prepared for scintillation counting. Note:

Untreated cells exhibited an increase in thymidine incorporation while the level of thymidine incorporation remained unchanged in the hydroxyurea treated cells.

99

Figure 16

100

Figure 17: Deposition of new H2A and H2B onto acetylated chromatin in the absence of

DNA replication. HeLa cells were incubated in MEM containing 10 mM hydroxyurea for

25 min prior to radiolabeling for 10 min with [ 3H]lysine. The addition of hydroxyurea

reduced the amount of new H3/H4 in chromatin (Input) and in the bound fraction of the

immunoprecipitation using α-H4 K5/K12ac. H2A and H2B deposition, however, was not

as affected by the presence of hydroxyurea. H2A and H2B were immunoprecipitated

with α-H4 K8/K16ac. This result suggests that considerable H2A and H2B deposition

takes place on non-replicating chromatin that is acetylated.

101

Figure 17

102

Figure 18: Uridine incorporation in the presence of DRB. HeLa cells were pre-incubated for 20 min at 37 0C in the presence or absence of added DRB (40 M final concentration). Following pre-incubation, 12.5 µCi of [ 3H]uridine was added to each culture. At the indicated time intervals, aliquots were removed, filtered through glass filter circles, and prepared for scintillation counting. Note: Untreated cells exhibited a greater increase in uridine incorporation than cells treated with DRB.

103

Figure 18

104

Figure 19: Deposition of H2A and H2B onto non-replicating chromatin is inhibited by

5,6-Dichloro-1-β-D-ribofuranosylbenzimidazole (DRB), an inhibitor RNA synthesis. HeLa

cells were incubated for approximately 25 minutes in minimum essential medium (MEM)

in the presence or absence of 40 M of DRB prior to radiolabeling for 10 min with

[3H]lysine. Unlike the case with new H3/H4, new H2A and H2B were depleted in the bound fractions of the immunoprecipitation with α-acH4 K8/K16 in cells treated with

DRB. This suggests that transcription is a factor involved in H2A/H2B exchange.

105

Figure 19

106

Figure 20: Acetylation of cytosolic H2A. A HeLa cytosolic extract ( A) or soluble HeLa

nucleosomes ( B) was immunoprecipitated (IP) using antibodies that recognize H4 acetylated at K5/K12 or using control rabbit serum (RNIS). Input, unbound (U), and bound (B) fractions were resolved by SDS-PAGE and analyzed by Western blotting using antibodies that recognize H2A acetylated at K5 (acH2A; 1:1000 dilution) or H4 acetylated at K5/K12 (acH4; 1:5000 dilution). Note: In contrast to acetylated

nucleosomal H2A, which associated with acetylated H4, cytosolic acetylated H2A from

the S100 remained in the unbound fraction and did not precipitate with cytosolic

acetylated H4.

107

Figure 20

108

Discussion

Part I: Characterization of the S. pombe Hat1 complex

Previous studies by Lopez-Rodas et al. have described a type-B HAT in S. cerevisiae with an enzymatic activity specific for free histone H4 (Lopez-Rodas et al.

1985; Lopez-Rodas et al. 1991). Biochemical characterization of this HAT-B enzyme from S. cerevisiae cytosolic extracts identified a complex consisting of two subunits:

Hat1 (catalytic subunit) and Hat2 (a homolog mammalian Rbap46 and Rbap48). This complex was determined to specifically acetylate free H4 at lysine 12. In contrast, recombinant Hat1 was found to acetylate free H4 at lysines 5 and 12, with a slight activity towards H2A (Kleff et al. 1995; Parthun et al. 1996). More recently, a study by

Poveda et al . using the yeast two-hybrid system has demonstrated Hif1 to be part of a heterotrimeric nuclear Hat-B complex, which includes Hat1 and Hat2 (Poveda et al.

2004).

An initial goal of my study was to characterize the in vivo Hat1 complex in S. pombe. The rationale for examining this complex in fission yeast is that S. pombe and

S. cerevisiae diverged evolutionarily over a billion years ago and therefore have considerable biological differences. As such, it is possible that the components and activity of the Hat1 complex differ between the two yeast species. To determine the members of this complex in S. pombe , the Hat1 protein was TAP tagged. Previous studies in our own lab by Benson et al. have demonstrated that the loss of Hat1 activity in S. pombe cells leads to an increased sensitivity to the DNA damaging agent MMS

109

without concurrent mutations of histone H3 (Benson et al. 2007). MMS assays using the

Hat1-TAP strain did not reveal a similar increased sensitivity to MMS. This result

establishes Hat1-TAP as a functional enzyme and demonstrates that the C-terminal TAP

tag does not significantly interfere with Hat1 activity.

In accordance with the results from S. cerevisiae , mass spectrometry of the tandem affinity purified S. pombe Hat1 complex revealed the identity of two subunits:

Hat1 and Hat2 (Parthun et al. 1996). Based on homology, Hat2 is referred to as Mis16 in S. pombe . Western blot analysis following the tandem affinity purification of Hat1 confirmed the association between Hat1 and Mis16 in S. pombe . This is the first evidence for the in vivo association of these two proteins in fission yeast. The addition of S. po mbe to the growing list of organisms in which proteins similar to Hat1 and Hat2 interact suggests that the Hat1/Hat2 complex may be the conserved core of the HAT-B complex.

The data presented herein confirm mis16 as an essential gene in S. pombe

(Ruiz-Garcia et al. 1998; Hayashi et al. 2004). The essential nature of mis16 in S. pombe may be explained by its role in the loading of CENP-A (the centromere specific

H3 variant) and the maintenance of a deacetylated state for histones located in the

central domains of centromeres (Hayashi et al. 2004). To date, no evidence has been

provided demonstrating similar functions for Hat2 in S. cerevisiae . This is not to say that

Hat2 is not involved in these processes in S. cerevisiae . It is possible that redundant

110

mechanisms were developed during the course of evolution that have rendered Hat2 non-essential in budding yeast.

Previous studies in S. cerevisiae have shown native Hat1 to acetylate H4 at K12, while recombinant Hat1 acetylates H4 at K5 and K12 (Kleff et al. 1995; Parthun et al.

1996). Results from HAT assays using tandem affinity purified Hat1 from S. pombe and in vitro filter binding assays suggest this enzyme to have a similar activity in fission yeast. Purified Hat1 appears is able to acetylate recombinant H4 at K5 and K12.

Additionally, results from the HAT assays appear to show some slight activity for K8.

The filter binding assays also do not support a major activity for the acetylation of K8.

The percent acetylation with the no peptide control and the peptide previously acetylated on K5 and K12 ("5/12" peptide) are essentially the same. If K8 of H4 is a substrate of

Hat1, one would expect to see an increased acetylation with the "5/12" peptide over the no peptide control since K8 is not blocked by a previous acetylation event. However, it is possible that the H4 tail peptides may not be as good a substrate for Hat1 as recombinant H4 proteins in vitro .

Our experiments focused on HAT1 complexes released into standard "whole cell" extracts under relatively physiological ionic strength conditions. In future experiments, it will be interesting to specifically characterize the nuclear Hat-B complex in S. pombe . As stated earlier, Poveda et al. have shown that the nuclear Hat-B complex contains an additional subunit, Hif1, which is not a member of the cytosolic Hat-

B complex (Poveda et al. 2004). Although there is no obvious orthologue of Hif1 in S.

111

pombe , Hif1 does share 42% identity (20% similarity) with Xenopus N1, a H3/H4 chaperone, which in turn is related to human NASP protein (Ai and Parthun 2004). Like

N1, NASP is capable of binding histones and may have an in vivo role as a chaperone

for linker histone (Richardson et al. 2000; Alekseev et al. 2003; Alekseev et al. 2005;

Richardson et al. 2006; Finn et al. 2008). Additionally, isolation of H3/H4 pre-deposition

complexes from HeLa cell nuclei has resulted in the co-purification of Hat1 and NASP.

Blast searches have identified a possible NASP orthologue in S. pombe called Sim3, a histone H3 and CENP-A specific chaperone (Dunleavy et al. 2007).

Part II: The role of Hat1 in S. pombe telomeric and centromeric silencing

As stated earlier, the deletion of HAT1 + in S. cerevisiae combined with mutations in certain acetylatable lysines of histone H3 results in an increased sensitivity to MMS

(Qin and Parthun 2002). Similarly, Kelly et al. have shown that a deletion of HAT1 + or

HAT2 + combined with mutations of the histone H3 N-terminal tail results in the loss of telomeric silencing in S. cerevisiae . Mutational analysis of the H4 tail demonstrated that the function of Hat1 in telomeric silencing was mediumted by lysine 12. In contrast to other histone acetyltransferases, Hat1 activity appeared to be involved in transcriptional repression rather than gene activation (Kelly et al. 2000). Previous studies in our own lab have demonstrated that the deletion of hat1 in S. pombe does not require concomitant mutations of the H3 tail for cells to exhibit an increased sensitivity to MMS

(Benson et al. 2007). Therefore, it remained possible that if Hat1 is involved in S. pombe telomeric silencing, the loss of telomeric silencing upon the deletion of hat1

112

would also not require concurrent mutations of the H3 tail. Spot assays supported this

hypothesis, revealing that the deletion of hat1 from a strain of S. pombe carrying a

telomeric ura4 marker results in the loss of telomeric silencing without concomitant H3 mutations. Thus, the redundant functions between Hat1 and H3 acetylation during telomeric silencing in S. cerevisiae are absent from S. pombe . The fact that Hat1 is involved in transcriptional repression at telomeres in both S. cerevisiae and S. pombe

suggests that the role of type-B histone acetyltransferases during telomeric silencing is

evolutionarily conserved. However, it remains unclear why H3 acetylation has a

redundant function to Hat1 in S. cerevisiae , but not in S. pombe . One possibility is that the two species do not have the same necessity for the deposition-related acetylation of

H3. Another possibility is simply that S. cerevisiae has more compensatory chromatin assembly pathways than S. pombe.

The deletion of hat1 also resulted in an increase of H4 acetylation at telomeres.

This result is interesting because one might have predicted a decrease in the acetylation of H4 at telomeres, particularly at lysines 5 and 12, following the deletion of hat1 , the gene coding for the HAT believed to be responsible for H4 K5/K12 acetylation. The fact that increased acetylation of H4 occurs at all the acetylatable lysine residues following the deletion of hat1 suggests that other HATs must be acetylating histone H4 in the subtelomeric region. However, the activity of these other HATs does not appear to be a compensatory reaction to the loss of hat1 , with respect to silencing. With Hat1 present, acetylation is actually lower in the wildtype strain than the hat1 strain. The finding that

113

the hyperacetylation of histones through TSA treatment mimics the loss of telomeric

silencing phenotype observed with a hat1 mutant further suggests that histone

acetylation state plays a key regulatory role during telomeric silencing. As for how this

increased acetylation and loss of telomeric silencing occur following the deletion of hat1 , one possible explanation centers on the processes of chromatin assembly and chromatin maturation. The results presented herein demonstrate that the loss of Hat1 or the inhibition of histone deacetylation results in the loss of telomeric silencing. As mentioned previously, studies have shown that during chromatin replication and synthesis, the timely acetylation and subsequent deacetylation of new H4 are required for the proper assembly and maturation of chromatin (Jackson et al. 1976; Annunziato and Seale 1983; Chicoine et al. 1986; Sobel et al. 1995). Therefore, it is possible that the proper establishment of telomeric silencing requires the preservation of this cycle of acetylation and deacetylation. By deleting hat1 , this cycle is presumably disrupted in hat1 mutants. Gomez et al . have demonstrated that the S. pombe HAT, Mst2, negatively regulates telomeric silencing (Gomez et al. 2005). When silencing does not occur, the subtelomeric region may become erroneously accessible to other HATs, such as Mst2, resulting in excess acetylation, which would result in a euchromatic region.

Alternatively, the region may become transcriptionally active first and become hyperacetylated during the course of transcription. In either case, it is apparent that the proper establishment of S. pombe telomeric silencing requires Hat1 and that the timely

114

acetylation of H4 during chromatin assembly may be required to produce the correct

epigenetic state at S. pombe telomeres.

Interestingly, the inhibition of histone deacetylation by TSA and sodium butyrate prolonged the arrest of synchronized S. pombe cells and resulted in the accumulation of cells with greater than 2C DNA content. Typically, S. pombe have a very short period

during their cell cycle where cells will have more than 2C DNA content, an indicator of

cells completing S-phase. Cells from this short period are generally undetectable by

flow cytometry (Gomez and Forsburg 2004). The appearance of cells with more than 2C

content after HDAC inhibitor treatment suggests a delay in cells exiting S-phase.

Experiments in human cells have also shown HDAC inhibitors to have an effect on

human cell cycle progression. As an example, HeLa cells treated with sodium butyrate

for 24 hours exhibit an arrest at G1 phase (Littlefield et al. 1982). The reasons for why

the S. pombe cell cycle should be affected by the inhibition of histone deacetylation remain unclear. Nevertheless, the results presented here demonstrate that proper S. pombe cell cycle progression requires the maintenance of correct histone acetylation / deacetylation regulation.

The effect of hat1 deletion on silencing at the S. pombe centromeric central core region was the opposite of that observed on telomeric silencing. The results demonstrate that hat1 mutants experience increased centromeric silencing at the

central core following the loss of Hat1. The different effects of hat1 deletion on telomeric and centromeric silencing may be due to the different silencing mechanisms employed

115

at the two regions. As an example of these differences, studies have shown that swi6 mutants in S. pombe have defects in telomeric silencing, but not in centromeric silencing at the central core region (Allshire et al. 1995). Conversely, mutations of CENP-A, the centromere specific S. pombe H3 variant, alleviate silencing within the central core of centromeres, but do not affect other silent chromatin (Pidoux et al. 2003). Because of these differences, Hat1 (and the acetylation of new H4) may function differently during the establishment of telomeric silencing than it does during silencing at the centromeric central core. Nevertheless, the results described herein demonstrate Hat1 to be important for the establishment of "normal" silencing at the telomere and the centromeric central core.

Part III: The deposition of H2A and H2B

A previous study by Nadeau et al. has demonstrated that a decreased level of histone synthesis persists for several hours following the inhibition of DNA replication

(Nadeau et al. 1978). With DNA replication inhibited, newly synthesized H3 and H4 accumulate in free pools, while new H2A and H2B continue to be deposited onto chromatin (Louters and Chalkley 1985; Bonner et al. 1988; Jackson 1990). These findings suggest that while new H3 and H4 deposition occurs at the replication fork, considerable H2A and H2B deposition occurs on non-replicating chromatin. The results presented herein support these findings. While the addition of hydroxyurea reduced the amount of new H3/H4 in chromatin and in the bound fraction of the immunoprecipitation using α-H4 K5/K12ac, the deposition of H2A and H2B remained largely unaffected.

116

Additionally, H2A and H2B were immunoprecipitated with both α-acH4 K5/K12 and α- acH4 K8/K16 suggesting that considerable H2A and H2B exchange occurs in acetylated non-replicating chromatin.

A study by Jackson has previously shown that the deposition of H2A and H2B is at least partially dependent on transcription. A decrease in deposition was observed when both replication and transcription were inhibited. The same reliance on transcription was observed in density labeling experiments in the presence of the RNA synthesis inhibitor actinomycin D (Jackson 1990). Our finding that the RNA synthesis inhibitor DRB decreases the level of new H2A and H2B deposition onto non-replicating chromatin is in agreement with those of Jackson and provides further evidence that transcription is involved in the exchange of H2A and H2B. Moreover, our results show that this transcription-related exchange occurs on acetylated chromatin regions. The question regarding how transcription facilitates this exchange has not been definitively answered. According to Perry et al. , the observation that new H2A and H2B are exchanged into chromatin may indicate that a pool of old H2A/H2B is present in HeLa cells (Perry et al. 1993). One possibility suggested by Jackson is that the dissolution of nucleosomes during transcription and replication is important in generating this histone pool (Jackson 1990). As further evidence, an in vitro study by Kireeva et al. has shown

that transcription by RNA polymerase II through nucleosome cores results in the

displacement of one H2A/H2B dimer from the octamer upon polymerase passage

(Kireeva et al. 2002). Additionally, ChIP experiments performed in Physarum

117

polycelphalum have shown H2A/H2B dimer exchange to be much more prevalent on transcribed genes than on silent loci. Furthermore, the finding that this exchange was limited to the structural gene within transcribed loci indicates that the active process of elongation is responsible for the exchange of H2A/H2B with free histone pools and is not merely the result of increased passive exchange into acetylated, transcriptionally active chromatin regions (Thiriet and Hayes 2005). It is apparent, however, that histone displacement must occur in a fashion such that the old displaced histones are not immediumtely reassembled into the same positions. One possibility is that old histones are essentially diluted by being introduced into a free histone pool containing predominantly nascent histones. During S-phase, nascent H3/H4 tetramers have been shown to associate with bold old and new H2A/H2B dimers on newly replicated DNA

(Jackson et al. 1981). Therefore, one can conclude that a portion of old H2A/H2B dimers is reassembled from a histone pool.

As stated previously, newly synthesized H2A is not uniquely modified in a deposition-related pattern. If histones are exchanged from acetylated transcriptionally active regions, as studies have suggested, then displaced H2A/H2B may carry post- translational modifications. The results presented herein, offer indirect evidence for this hypothesis by providing the first demonstration that cytosolic H2A can be specifically modified by acetylation in vivo . Notably, acetylated cytosolic H2A is not associated with acetylated cytosolic H4. As cytosolic H2A/H2B is associated with the histone chaperone

118

NAP1 (Chang et al. 1997), our results support a model in which H2A/H2B exchange during transcription is mediumted by the NAP1 chaperone.

119

References

Aggarwal, B. D. and B. R. Calvi (2004). "Chromatin regulates origin activity in Drosophila follicle cells." Nature 430 (6997): 372-376. Ahmad, K. and S. Henikoff (2002). "The histone variant H3.3 marks active chromatin by replication-independent nucleosome assembly." Mol Cell 9(6): 1191-200. Ai, X. and M. R. Parthun (2004). "The nuclear Hat1p/Hat2p complex: A molecular link between type B histone acetyltransferases and chromatin assembly." Mol Cell 14 (2): 195-205. Alekseev, O. M., D. C. Bencic, et al. (2003). "Overexpression of the Linker histone- binding protein tNASP affects progression through the cell cycle." J Biol Chem 278 (10): 8846-52. Alekseev, O. M., E. E. Widgren, et al. (2005). "Association of NASP with HSP90 in mouse spermatogenic cells - Stimulation of ATPase activity and transport of linker histones into nuclei." J Biol Chem 280 (4): 2904-2911. Allfrey, V. G., R. Faulkner, et al. (1964). "Acetylation and Methylation of Histones and Their Possible Role in the Regulation of RNA Synthesis." Proc Natl Acad Sci U.S.A. 51 : 786-794. Allshire, R. C., E. R. Nimmo, et al. (1995). "Mutations derepressing silent centromeric domains in fission yeast disrupt chromosome segregation." Genes Dev 9(2): 218- 33. Annunziato, A. T. (1990). Chromatin Replication and Nucleosome Assembly. The Eukaryotic Nucleus: Molecular Biochemistry and Macromolecular Assemblies . P. R. Strauss and S. H. Wilson. Caldwell, NJ, The Telford Press. 2: 687-712. Annunziato, A. T. (1995). Histone Acetylation During Chromatin Replication and Nucleosome Assembly. The Nucleus . A. P. Wolffe. Greenwich, CT, JAI Press. 1: 31-55. Annunziato, A. T. and J. C. Hansen (2000). "Role of histone acetylation in the assembly and modulation of chromatin structures." Gene Expr 9: 37-61. Annunziato, A. T., R. K. Schindler, et al. (1982). "Association of Newly Synthesized Histones with Replicating and Nonreplicating Regions of Chromatin." J. Biol. Chem. 257 : 8507-8515.

120

Annunziato, A. T. and R. L. Seale (1983). "Chromatin replication, reconstitution and assembly." Mol Cell Biochem 55 (2): 99-112. Annunziato, A. T. and R. L. Seale (1983). "Histone Deacetylation is Required for the Maturation of Newly Replicated Chromatin." J. Biol. Chem. 258 : 12675-12684. Arany, Z., D. Newsome, et al. (1995). "A family of transcriptional adaptor proteins targeted by the E1A oncoprotein." Nature 374 (6517): 81-4. Arany, Z., W. R. Sellers, et al. (1994). "E1A-associated p300 and CREB-associated CBP belong to a conserved family of coactivators." Cell 77 (6): 799-800. Arents, G., R. W. Burlingame, et al. (1991). "The Nucleosomal Core Histone Octamer at 3.1-A Resolution - A Tripartite Protein Assembly and a Left-Handed Superhelix." Proc Natl Acad Sci USA 88 (22): 10148-10152. Arents, G. and E. N. Moudrianakis (1993). "Topography of the Histone Octamer Surface - Repeating Structural Motifs Utilized in the Docking of Nucleosomal DNA." Proc Natl Acad Sci USA 90 (22): 10489-10493. Arents, G. and E. N. Moudrianakis (1995). "The histone fold: A ubiquitous architectural motif utilized in DNA compaction and protein dimerization." Proc Natl Acad Sci USA 92 (24): 11170-11174. Ausio, J., D. W. Abbott, et al. (2001). "Histone variants and histone modifications: A structural perspective." Biochem Cell Biol 79 (6): 693-708. Ausio, J., F. Dong, et al. (1989). "Use of selectively trypsinized nucleosome core particles to analyze the role of the histone "tails" in the stabilization of the nucleosome." J Mol Biol 206 (3): 451-63. Bahler, J., J. Q. Wu, et al. (1998). "Heterologous modules for efficient and versatile PCR-based gene targeting in Schizosaccharomyces pombe." Yeast 14 (10): 943- 51. Bannister, A. J. and T. Kouzarides (1996). "The CBP co-activator is a histone acetyltransferase." Nature 384 (6610): 641-643. Barman, H. K., Y. Takami, et al. (2006). "Histone acetyltransferase 1 is dispensable for replication-coupled chromatin assembly but contributes to recover DNA damages created following replication blockage in vertebrate cells." Biochem Biophys Res Commun 345 (4): 1547-57. Baumann, P. and T. R. Cech (2001). "Pot1, the putative telomere end-binding protein in fission yeast and humans." Science 292 (5519): 1171-5.

121

Benson, L. J. and A. T. Annunziato (2004). "In vitro analysis of histone acetyltransferase activity." Methods 33 (1): 45-52. Benson, L. J., Y. L. Gu, et al. (2006). "Modifications of H3 and H4 during chromatin replication, nucleosome assembly, and histone exchange." J Biol Chem 281 (14): 9287-9296. Benson, L. J., J. A. Phillips, et al. (2006). "Properties of the type B histone acetyltransferase HAT1: H4 tail interaction, site preference, and involvement in DNA repair." J Biol Chem . Benson, L. J., J. A. Phillips, et al. (2007). "Properties of the type B histone acetyltransferase Hat1 - H4 tail Interaction, site preference, and involvement in DNA repair." J Biol Chem 282 (2): 836-842. Bone, J. R., J. Lavender, et al. (1994). "Acetylated histone H4 on the male X chromosome is associated with dosage compensation in Drosophila." Genes Dev 8(1): 96-104. Bonner, W. M. and R. A. Laskey (1974). "A film detection method for tritium-labeled proteins and nucleic acids in polyacrylamide gels." Eur J Biochem 46 : 83-88. Bonner, W. M., R. S. Wu, et al. (1988). "Kinetics of Accumulation and Depletion of Soluble Newly Synthesized Histone in the Reciprocal Regulation of Histone and DNA Synthesis." Biochemistry 27 : 6542-6550. Broccoli, D. (2004). "Function, replication and structure of the mammalian telomere." Cytotechnology 45 (1-2): 3-12. Brownell, J. E. and C. D. Allis (1996). "Special HATs for special occasions: linking histone acetylation to chromatin assembly and gene activation." Curr Opin Genet Dev 6: 176-184. Burke, T. W., J. G. Cook, et al. (2001). "Replication factors MCM2 and ORC1 interact with the histone acetyltransferase HBO1." J Biol Chem 276 (18): 15397-15408. Chang, L., S. S. Loranger, et al. (1997). "Histones in transit: Cytosolic histone complexes and diacetylation of H4 during nucleosome assembly in human cells." Biochemistry 36 (3): 469-480. Cheung, P., C. D. Allis, et al. (2000). "Signaling to chromatin through histone modifications." Cell 103 (2): 263-271. Chicoine, L. G., I. G. Schulman, et al. (1986). "Nonrandom utilization of acetylation sites in histones isolated from Tetrahymena : Evidence for functionally distinct H4 acetylation sites." J Biol Chem 261 : 1071-1076.

122

Chikashige, Y., D. Q. Ding, et al. (1994). "Telomere-led premeiotic chromosome movement in fission yeast." Science 264 (5156): 270-3. Chikashige, Y. and Y. Hiraoka (2001). "Telomere binding of the Rap1 protein is required for meiosis in fission yeast." Curr Biol 11 (20): 1618-23. Chrivia, J. C., R. P. Kwok, et al. (1993). "Phosphorylated CREB binds specifically to the nuclear protein CBP." Nature 365 (6449): 855-9. Clarke, A. S., J. E. Lowell, et al. (1999). "Esa1p is an essential histone acetyltransferase required for cell cycle progression." Mol Cell Biol 19 (4): 2515-2526. Cooper, J. P. and Y. Hiraoka (2006). Fission Yeast Telomeres . Cold Spring Harbor, Cold Spring Harbor Laboratory Press. Cooper, J. P., E. R. Nimmo, et al. (1997). "Regulation of telomere length and function by a Myb-domain protein in fission yeast." Nature 385 (6618): 744-747. Cousens, L. S. and B. M. Alberts (1982). "Accessibility of Newly Synthesized Chromatin to Histone Acetylase." J. Biol. Chem. 257 : 3945-3949. Davey, C. A., D. F. Sargent, et al. (2002). "Solvent mediated interactions in the structure of the nucleosome core particle at 1.9 angstrom resolution." J Mol Biol 319 (5): 1097-1113. Doyon, Y., C. Cayrou, et al. (2006). "ING tumor suppressor proteins are critical regulators of chromatin acetylation required for genome expression and perpetuation." Mol Cell 21 (1): 51-64. Doyon, Y., W. Selleck, et al. (2004). "Structural and functional conservation of the NuA4 histone acetyltransferase complex from yeast to humans." Mol Cell Biol 24 (5): 1884-1896. Dunleavy, E. M., A. L. Pidoux, et al. (2007). "A NASP (N1/N2)-related protein, Sim3, binds CENP-A and is required for its deposition at fission yeast centromeres." Mol Cell 28 (6): 1029-44. Dutnall, R. N., S. T. Tafrov, et al. (1998). "Structure of the histone acetyltransferase Hat1: A paradigm for the GCN5-related N-acetyltransferase superfamily." Cell 94 (4): 427-438. Dutnall, R. N., S. T. Tafrov, et al. (1998). "Structure of the yeast histone acetyltransferase Hat1: Insights into substrate specificity and implications for the Gcn5-related N-acetyltransferase superfamily." Cold Spring Harbor Symp Quant Biol 63 : 501-507.

123

Eberharter, A., T. Lechner, et al. (1996). "Purification and characterization of the cytoplasmic histone acetyltransferase B of maize embryos." FEBS Lett 386 (1): 75-81. Eckner, R., M. E. Ewen, et al. (1994). "Molecular cloning and functional analysis of the adenovirus E1A-associated 300-kD protein (p300) reveals a protein with properties of a transcriptional adaptor." Genes Dev 8(8): 869-84. Eickbush, T. H. and E. N. Moudrianakis (1978). "The Histone Core Complex: An Octamer Assembled by Two Sets of Protein-Protein Interactions." Biochemistry 17 : 4956-4964. Ekwall, K., E. R. Nimmo, et al. (1996). "Mutations in the fission yeast silencing factors clr4+ and rik1+ disrupt the localisation of the chromo domain protein Swi6p and impair centromere function." J Cell Sci 109 ( Pt 11) : 2637-48. Ekwall, K., T. Olsson, et al. (1997). "Transient inhibition of histone deacetylation alters the structural and functional imprint at fission yeast centromeres." Cell 91 (7): 1021-1032. Fairall, L., L. Chapman, et al. (2001). "Structure of the TRFH dimerization domain of the human telomeric proteins TRF1 and TRF2." Mol Cell 8(2): 351-61. Finn, R. M., K. Browne, et al. (2008). "sNASP, a histone H1-specific eukaryotic chaperone dimer that facilitates chromatin assembly." Biophys J 95 (3): 1314-25. Fischle, W., Y. M. Wang, et al. (2003). "Histone and chromatin cross-talk." Curr Opin Cell Biol 15 (2): 172-183. Furuyama, T., Y. Dalal, et al. (2006). "Chaperone-mediated assembly of centromeric chromatin in vitro." Proc Nat Acad Sci Usa 103 (16): 6172-6177. Georgel, P. T., T. Tsukiyama, et al. (1997). "Role of histone tails in nucleosome remodeling by Drosophila NURF." EMBO J 16 (15): 4717-4726. Gomez, E. B., J. M. Espinosa, et al. (2005). "Schizosaccharomyces pombe mst2(+) encodes a MYST family histone acetyltransferase that negatively regulates telomere silencing." Mol Cell Biol 25 (20): 8887-8903. Gomez, E. B. and S. L. Forsburg (2004). "Analysis of the fission yeast Schizosaccharomyces pombe cell cycle." Methods Mol Biol 241 : 93-111. Grant, P. A., A. Eberharter, et al. (1999). "Expanded lysine acetylation specificity of Gcn5 in native complexes." J Biol Chem 274 (9): 5895-5900. Groth, A., W. Rocha, et al. (2007). "Chromatin challenges during DNA replication and repair." Cell 128 (4): 721-33.

124

Han, M. and M. Grunstein (1988). "Nucleosome loss activates yeast downstream promoters in vivo." Cell 55 (6): 1137-45. Hansen, J. C., C. Tse, et al. (1998). "Structure and function of the core histone N-termini: more than meets the eye." Biochemistry 37 (51): 17637-17641. Harp, J. M., B. L. Hanson, et al. (2000). "Asymmetries in the nucleosome core particle at 2.5 angstrom resolution." Acta Crystallogr D Biol Cryst 56 : 1513-1534. Hayashi, T., Y. Fujita, et al. (2004). "Mis16 and Mis18 are required for CENP-A loading and histone deacetylation at centromeres." Cell 118 (6): 715-729. Henikoff, S., T. Furuyama, et al. (2004). "Histone variants, nucleosome assembly and epigenetic inheritance." Trends Genet 20 (7): 320-326. Huang, S., H. Zhou, et al. (2005). "Rtt106p is a histone chaperone involved in heterochromatin-mediated silencing." Proc Natl Acad Sci U S A 102 (38): 13410- 5. Iizuka, M., T. Matsui, et al. (2006). "Regulation of replication licensing by acetyltransferase Hbo1." Mol Cell Biol 26 (3): 1098-1108. Iizuka, M. and B. Stillman (1999). "Histone acetyltransferase HBO1 interacts with the ORC1 subunit of the human initiator protein." J Biol Chem 274 (33): 23027-23034. Imhof, A. and A. P. Wolffe (1999). "Purification and properties of the Xenopus Hat1 acetyltransferase: Association with the 14-3-3 proteins in the oocyte nucleus." Biochemistry 38 (40): 13085-13093. Jackson, V. (1990). "In Vivo Studies on the Dynamics of Histone-DNA Interaction: Evidence for Nucleosome Dissolution during Replication and Transcription and a Low Level of Dissolution Independent of Both." Biochemistry 29 : 719-731. Jackson, V. and R. Chalkley (1981). "A New Method for the Isolation of Replicative Chromatin: Selective Deposition of Histone on Both New and Old DNA." Cell 23 : 121-134. Jackson, V. and R. Chalkley (1981). "A Reevaluation of New Histone Deposition on Replicating Chromatin." J. Biol. Chem. 256 : 5095-5103. Jackson, V., S. Marshall, et al. (1981). "The Sites of Deposition of Newly Synthesized Histone." Nucleic Acids Res. 9(18): 4563-4581. Jackson, V., A. Shires, et al. (1976). "Modifications to histones immediately after synthesis." J Mol Biol 104 : 471-483. Janknecht, R. and T. Hunter (1996). "Transcription. A growing coactivator network." Nature 383 (6595): 22-3.

125

Jenuwein, T. and C. D. Allis (2001). "Translating the histone code." Science 293 (5532): 1074-1080. Kamakaka, R. T. and S. Biggins (2005). "Histone variants: deviants?" Gene Develop 19 (3): 295-310. Kamine, J., B. Elangovan, et al. (1996). "Identification of a cellular protein that specifically interacts with the essential cysteine region of the HIV-1 Tat transactivator." Virology 216 (2): 357-66. Kanoh, J. and F. Ishikawa (2001). "spRap1 and spRif1, recruited to telomeres by Taz1, are essential for telomere function in fission yeast." Curr Biol 11 (20): 1624-1630. Kanoh, J. and F. Ishikawa (2003). "Composition and conservation of the telomeric complex." Cell Mol Life Sci 60 (11): 2295-302. Kanoh, J., M. Sadaie, et al. (2005). "Telomere binding protein Taz1 establishes Swi6 heterochromatin independently of RNAi at telomeres." Curr Biol 15 (20): 1808-19. Kaufman, P. D., R. Kobayashi, et al. (1995). "The p150 and p60 subunits of chromatin assembly factor I: A molecular link between newly synthesized histones and DNA replication." Cell 81 (7): 1105-1114. Keeney, J. B. and J. D. Boeke (1994). "Efficient targeted integration at leu1-32 and ura4- 294 in Schizosaccharomyces pombe." Genetics 136 (3): 849-56. Kelly, T. J., S. Qin, et al. (2000). "Type B histone acetyltransferase Hat1p participates in telomeric silencing." Mol Cell Biol 20 (19): 7051-7058. Kimura, A. and M. Horikoshi (1998). "Tip60 acetylates six lysines of a specific class in core histones in vitro." Genes Cells 3(12): 789-800. Kireeva, M. L., W. Walter, et al. (2002). "Nucleosome remodeling induced by RNA polymerase II: Loss of the H2A/H2B dimer during transcription." Mol Cell 9(3): 541-552. Kleff, S., E. D. Andrulis, et al. (1995). "Identification of a gene encoding a yeast histone H4 acetyltransferase." J Biol Chem 270 (42): 24674-24677. Kornberg, R. D. (1974). "Chromatin structure: a repeating unit of histones and DNA." Science 184 (139): 868-71. Kornberg, R. D. and J. O. Thomas (1974). "Chromatin Structure: Oligomers of Histones." Science 184 : 865-868. Krawitz, D. C., T. Kama, et al. (2002). "Chromatin assembly factor I mutants defective for PCNA binding require Asf1/Hir proteins for silencing." Mol Cell Biol 22 (2): 614- 625.

126

Krebs, J. E., C. J. Fry, et al. (2000). "Global role for chromatin remodeling enzymes in mitotic gene expression." Cell 102 (5): 587-598. Kuo, M. H. and C. D. Allis (1998). "Roles of histone acetyltransferases and deacetylases in gene regulation." Bioessays 20 (8): 615-626. Kuo, M. H., J. E. Brownell, et al. (1996). "Transcription-linked acetylation by Gcn5p of histones H3 and H4 at specific lysines." Nature 383(6597): 269-272. Kwok, R. P., J. R. Lundblad, et al. (1994). "Nuclear protein CBP is a coactivator for the transcription factor CREB." Nature 370 (6486): 223-6. Laemmli, U. K. (1970). "Cleavage of structural proteins during the assembly of the head of bacteriophage T4." Nature 227 (5259): 680-5. Laskey, R. A. and A. D. Mills (1975). "Quantitative film detection method of 3H and 14C in polyacrylamide gels by fluorography." Eur J Biochem 56 : 335-341. Latrasse, D., M. Benhamed, et al. (2008). "The MYST histone acetyltransferases are essential for gametophyte development in Arabidopsis." BMC Plant Biol 8: 121. Lee, D. Y., J. J. Hayes, et al. (1993). "A Positive Role for Histone Acetylation in Transcription Factor Access to Nucleosomal DNA." Cell 72 (1): 73-84. Lee, K. M., S. Sif, et al. (1999). "hSWI/SNF disrupts interactions between the H2A N- terminal tail and nucleosomal DNA." Biochemistry Usa 38 (26): 8423-8429. Li, B., S. Oestreich, et al. (2000). "Identification of human Rap1: implications for telomere evolution." Cell 101 (5): 471-83. Li, Q., H. Zhou, et al. (2008). "Acetylation of histone H3 lysine 56 regulates replication- coupled nucleosome assembly." Cell 134 (2): 244-55. Littlefield, B. A., N. B. Cidlowski, et al. (1982). "Lack of effect of butyrate on S phase DNA synthesis despite presence of histone hyperacetylation." Exp Cell Res 141 (2): 283-91. Lopez-Rodas, G., J. E. Perez-Ortin, et al. (1985). "Partial purification and properties of two histone acetyltransferases from the yeast, Saccharomyces cerevisiae." Arch Biochem Biophys 239 (1): 184-90. Lopez-Rodas, G., V. Tordera, et al. (1991). "Subcellular localization and nucleosome specificity of yeast histone acetyltransferases." Biochemistry 30 (15): 3728-32. Louters, L. and R. Chalkley (1985). "Exchange of Histones H1, H2A, and H2B in Vivo." Biochemistry 24 : 3080-3085. Loyola, A. and G. Almouzni (2004). "Histone chaperones, a supporting role in the limelight." Bba Gene Struct Express 1677 (1-3): 3-11.

127

Lucchini, R. and J. M. Sogo (1995). "Replication of transcriptionally active chromatin." Nature 374 (6519): 276-280. Luger, K., A. W. Mader, et al. (1997). "Crystal structure of the nucleosome core particle at 2.8 angstrom resolution." Nature 389 (6648): 251-260. Lundblad, J. R., R. P. Kwok, et al. (1995). "Adenoviral E1A-associated protein p300 as a functional homologue of the transcriptional co-activator CBP." Nature 374 (6517): 85-8. Lusser, A., A. Eberharter, et al. (1999). "Analysis of the histone acetyltransferase B complex of maize embryos." Nucl Acid Res 27 (22): 4427-4435. Ma, X. J., J. S. Wu, et al. (1998). "Deposition-related sites K5/K12 in histone H4 are not required for nucleosome deposition in yeast." Proc Natl Acad Sci USA 95 (12): 6693-6698. Makarov, V. L., Y. Hirose, et al. (1997). "Long G tails at both ends of human chromosomes suggest a C strand degradation mechanism for telomere shortening." Cell 88 (5): 657-666. Makowski, A. M., R. N. Dutnall, et al. (2001). "Effects of acetylation of histone H4 at lysines 8 and 16 on activity of the Hat1 histone acetyltransferase." J Biol Chem 276 (47): 43499-43502. Manolis, K. G., E. R. Nimmo, et al. (2001). "Novel functional requirements for non- homologous DNA end joining in Schizosaccharomyces pombe." Embo J 20 (1-2): 210-21. Martinez-Balbas, M. A., U. M. Bauer, et al. (2000). "Regulation of E2F1 activity by acetylation." Embo J 19 (4): 662-71. Marushige, K. (1976). "Activation of chromatin by acetylation of histone side chains." Proc Natl Acad Sci U S A 73 (11): 3937-41. McElligott, R. and R. J. Wellinger (1997). "The terminal DNA structure of mammalian chromosomes." Embo J 16 (12): 3705-14. McKnight, T. D., M. S. Fitzgerald, et al. (1997). "Plant telomeres and telomerases. A review." Biochemistry (Mosc) 62 (11): 1224-31. Megee, P. C., B. A. Morgan, et al. (1990). "Genetic Analysis of Histone H4: Essential Role of Lysines Subject to Reversible Acetylation." Science 247 : 841-845. Mito, Y., J. G. Henikoff, et al. (2005). "Genome-scale profiling of histone H3.3 replacement patterns." Nat Genet 37 (10): 1090-1097.

128

Moggs, J. G., P. Grandi, et al. (2000). "A CAF-1-PCNA-mediated chromatin assembly pathway triggered by sensing DNA damage." Mol Cell Biol 20 (4): 1206-18. Moreno, S., A. Klar, et al. (1991). "Molecular genetic analysis of fission yeast Schizosaccharomyces pombe." Methods Enzymol 194 : 795-823. Mosammaparast, N., Y. R. Guo, et al. (2002). "Pathways mediating the nuclear import of histones H3 and H4 in yeast." J Biol Chem 277 (1): 862-868. Moyzis, R. K., J. M. Buckingham, et al. (1988). "A highly conserved repetitive DNA sequence, (TTAGGG)n, present at the telomeres of human chromosomes." Proc Natl Acad Sci U S A 85 (18): 6622-6. Nadeau, P., D. R. Oliver, et al. (1978). "Effect of Inhibition of DNA Synthesis on Histone Synthesis and Deposition." Biochemistry 17 : 4885-4893. Nakayama, J., A. J. Klar, et al. (2000). "A chromodomain protein, Swi6, performs imprinting functions in fission yeast during mitosis and meiosis." Cell 101 (3): 307- 17. Nakayama, J., J. C. Rice, et al. (2001). "Role of histone H3 lysine 9 methylation in epigenetic control of heterochromatin assembly." Science 292 (5514): 110-3. Nightingale, K. P., R. E. Wellinger, et al. (1998). "Histone acetylation facilitates RNA polymerase II transcription of the Drosophila hsp26 gene in chromatin." EMBO J 17 (10): 2865-2876. Nimmo, E. R., G. Cranston, et al. (1994). "Telomere-associated chromosome breakage in fission yeast results in variegated expression of adjacent genes." EMBO J 13 (16): 3801-3811. Niwa, O., M. Shimanuki, et al. (2000). "Telomere-led bouquet formation facilitates homologous chromosome pairing and restricts ectopic interaction in fission yeast meiosis." Embo J 19 (14): 3831-40. Noll, M. and R. D. Kornberg (1977). "Action of micrococcal nuclease on chromatin and the location of histone H1." J Mol Biol 109 (3): 393-404. Ogryzko, V. V., R. L. Schiltz, et al. (1996). "The transcriptional coactivators p300 and CBP are histone acetyltransferases." Cell 87 (5): 953-959. Parthun, M. R. (2007). "Hat1: the emerging cellular roles of a type B histone acetyltransferase." Oncogene 26 (37): 5319-28. Parthun, M. R., J. Widom, et al. (1996). "The major cytoplasmic histone acetyltransferase in yeast: Links to chromatin replication and histone metabolism." Cell 87 (1): 85-94.

129

Perry, C. A. and A. T. Annunziato (1989). "Influence of Histone Acetylation on the Solubility, H1 content and DNase I Sensitivity of Newly Assembled Chromatin." Nucleic Acids Res 17 (11): 4275-4291. Perry, C. A., C. A. Dadd, et al. (1993). "Analysis of Nucleosome Assembly and Histone Exchange Using Antibodies Specific for Acetylated H4." Biochemistry 32 (49): 13605-13614. Pidoux, A. L., W. Richardson, et al. (2003). "Sim4: a novel fission yeast kinetochore protein required for centromeric silencing and chromosome segregation." J Cell Biol 161 (2): 295-307. Pillus, L. (2008). "MYSTs mark chromatin for chromosomal functions." Curr Opin Cell Biol 20 (3): 326-33. Poveda, A., M. Pamblanco, et al. (2004). "Hif1 is a component of yeast histone acetyltransferase B, a complex mainly localized in the nucleus." J Biol Chem 279 (16): 16033-16043. Puig, O., F. Caspary, et al. (2001). "The tandem affinity purification (TAP) method: a general procedure of protein complex purification." Methods 24 (3): 218-29. Qin, S. and M. R. Parthun (2002). "Histone H3 and the histone acetyltransferase Hat1p contribute to DNA double-strand break repair." Mol Cell Biol 22 (23): 8353-8365. Qin, S. and M. R. Parthun (2006). "Recruitment of the type B histone acetyltransferase Hat1p to chromatin is linked to DNA double-strand breaks." Mol Cell Biol 26 (9): 3649-3658. Richardson, R. T., O. M. Alekseev, et al. (2006). "Nuclear autoantigenic sperm protein (NASP), a linker histone chaperone that is required for cell proliferation." J Biol Chem 281 (30): 21526-21534. Richardson, R. T., I. N. Batova, et al. (2000). "Characterization of the histone H1-binding protein, NASP, as a cell cycle-regulated somatic protein." J Biol Chem 275 (39): 30378-30386. Rigaut, G., A. Shevchenko, et al. (1999). "A generic protein purification method for protein complex characterization and proteome exploration." Nat Biotechnol 17 (10): 1030-2. Rocha, W. and A. Verreault (2008). "Clothing up DNA for all seasons: Histone chaperones and nucleosome assembly pathways." FEBS Lett 582 (14): 1938-49. Roth, S. Y., J. M. Denu, et al. (2001). "Histone acetyltransferases." Annu Rev Biochem 70 : 81-120.

130

Ruiz-Carrillo, A., L. J. Wangh, et al. (1975). "Processing of newly synthesized histone molecules." Science 190 : 117-128. Ruiz-Garcia, A. B., R. Sendra, et al. (1998). "HAT1 and HAT2 proteins are components of a yeast nuclear histone acetyltransferase enzyme specific for free histone H4." J Biol Chem 273 (20): 12599-12605. Scherthan, H., J. Bahler, et al. (1994). "Dynamics of chromosome organization and pairing during meiotic prophase in fission yeast." J Cell Biol 127 (2): 273-285. Schiltz, R. L., C. A. Mizzen, et al. (1999). "Overlapping but distinct patterns of histone acetylation by the human coactivators p300 and PCAF within nucleosomal substrates." J Biol Chem 274 (3): 1189-1192. Schwabish, M. A. and K. Struhl (2006). "Asf1 mediates histone eviction and deposition during elongation by RNA polymerase II." Mol Cell 22 (3): 415-422. Schwartz, B. E. and K. Ahmad (2005). "Transcriptional activation triggers deposition and removal of the histone variant H3.3." Gene Develop 19 (7): 804-814. Sealy, L. and R. Chalkley (1979). "Modification of histones immediately following synthesis." Arch Biochem Biophys 197 : 78-82. Shahbazian, M. D. and M. Grunstein (2007). "Functions of site-specific histone acetylation and deacetylation." Annu Rev Biochem 76 : 75-100. Shampay, J., J. W. Szostak, et al. (1984). "DNA sequences of telomeres maintained in yeast." Nature 310 (5973): 154-7. Shibahara, K. and B. Stillman (1999). "Replication-dependent marking of DNA by PCNA facilitates CAF-1-coupled inheritance of chromatin." Cell 96 (4): 575-585. Shikama, N., J. Lyon, et al. (1997). "The p300/CBP family: Integrating signals with transcription factors and chromatin." Tr Cell Biol 7(6): 230-236. Simpson, R. T. (1978). "Structure of the chromatosome, a chromatin particle containing 160 base pairs of DNA and all the histones." Biochemistry 17 (25): 5524-31. Sklenar, A. R. and M. R. Parthun (2004). "Characterization of yeast histone H3-specific type B histone acetyltransferases identifies an ADA2-independent Gcn5p activity." BMC Biochem 5: 11. Smith, M. M. (2002). "Histone variants and nucleosome deposition pathways." Mol Cell 9(6): 1158-60. Smith, M. M. and O. S. Andresson (1983). "DNA sequences of yeast H3 and H4 histone genes from two non-allelic gene sets encode identical H3 and H4 proteins." J Mol Biol 169 (3): 663-90.

131

Smith, S. and B. Stillman (1989). "Purification and Characterization of CAF-I, a Human Cell Factor Required for Chromatin Assembly during DNA Replication in vitro ." Cell 58 : 15-25. Smith, S. and B. Stillman (1991). "Stepwise Assembly of Chromatin During DNA Replication in vitro." EMBO J 10 (4): 971-980. Sobel, R. E., R. G. Cook, et al. (1995). "Conservation of deposition-related acetylation sites in newly synthesized histones H3 and H4." Proc Natl Acad Sci USA 92 (4): 1237-1241. Sollner-Webb, B. and G. Felsenfeld (1975). "A comparison of the digestion of nuclei and chromatin by staphylococcal nuclease." Biochemistry 14 (13): 2915-20. Stillman, B. (1986). "Chromatin Assembly during DNA Replication in vitro." Cell 45 : 555- 565. Strahl, B. D. and C. D. Allis (2000). "The language of covalent histone modifications." Nature 403 (6765): 41-45. Suto, R. K., M. J. Clarkson, et al. (2000). "Crystal structure of a nucleosome core particle containing the variant histone H2A.Z." Nature Struct Biology 7(12): 1121-1124. Tagami, H., D. Ray-Gallet, et al. (2004). "Histone H3.1 and H3.3 complexes mediate nucleosome assembly pathways dependent or independent of DNA synthesis." Cell 116 (1): 51-61. Thiriet, C. and J. J. Hayes (2005). "Replication-independent core histone dynamics at transcriptionally active loci in vivo." Gene Develop 19 (6): 677-682. Thomas, J. O. and R. D. Kornberg (1975). "An Octamer of Histones in Chromatin and Free in Solution." Proc Natl Acad Sci USA 72 : 2626-2630. Thomas, T. and A. K. Voss (2007). "The diverse biological roles of MYST histone acetyltransferase family proteins." Cell Cycle 6(6): 696-704. Toth, K. F., T. A. Knoch, et al. (2004). "Trichostatin A-induced histone acetylation causes decondensation of interphase chromatin." J Cell Sci 117 (18): 4277-4287. Trelles-Sticken, E., J. Loidl, et al. (1999). "Bouquet formation in budding yeast: initiation of recombination is not required for meiotic telomere clustering." J Cell Sci 112 ( Pt 5): 651-8. Tse, C., E. I. Georgieva, et al. (1998). "Gcn5p, a transcription-related histone acetyltransferase, acetylates nucleosomes and folded nucleosomal arrays in the absence of other protein subunits." J BIOL CHEM 273 (49): 32388-32392.

132

Turner, B. M. (2000). "Histone acetylation and an epigenetic code." Bioessays 22 (9): 836-845. Turner, B. M., A. J. Birley, et al. (1992). "Histone H4 Isoforms Acetylated at Specific Lysine Residues Define Individual Chromosomes and Chromatin Domains in Drosophila Polytene Nuclei." Cell 69 : 375-384. Tyler, J. K., K. A. Collins, et al. (2001). "Interaction between the Drosophila CAF-1 and ASF1 chromatin assembly factors." Mol Cell Biol 21 (19): 6574-6584. Ushinsky, S. C., H. Bussey, et al. (1997). "Histone H1 in Saccharomyces cerevisiae." Yeast 13 (2): 151-161. van Holde, K. E. (1988). Chromatin . New York, Springer-Verlag. Varshavsky, A. J., V. V. Bakayev, et al. (1976). "Heterogeneity of chromatin subunits in vitro and location of histone H1." Nucleic Acids Res 3(2): 477-92. Vermaak, D., P. A. Wade, et al. (1999). "Functional analysis of the SIN3-Histone deacetylase RPD3-RbAp48-histone H4 connection in the Xenopus oocyte." Mol Cell Biol 19 (9): 5847-5860. Verreault, A. (2000). "De novo nucleosome assembly: new pieces in an old puzzle." Gene Develop 14 (12): 1430-1438. Verreault, A., P. D. Kaufman, et al. (1996). "Nucleosome assembly by a complex of CAF-1 and acetylated histones H3/H4." Cell 87 (1): 95-104. Verreault, A., P. D. Kaufman, et al. (1997). "Nucleosomal DNA regulates the core- histone-binding subunit of the human Hat1 acetyltransferase." Curr Biol 8(2): 96- 108. Verreault, A., P. D. Kaufman, et al. (1998). "Nucleosomal DNA regulates the core- histone-binding subunit of the human Hat1 acetyltransferase." Curr Biol 8(2): 96- 108. Vettese-Dadey, M., P. A. Grant, et al. (1996). "Acetylation of histone H4 plays a primary role in enhancing transcription factor binding to nucleosomal DNA in vitro." Embo J 15 (10): 2508-18. West, H. H. P. and W. M. Bonner (1980). "Histone 2A, a Heteromorphous Family of Eight Protein Species." Biochemistry 19 : 3238-3245. White, C. L., R. K. Suto, et al. (2001). "Structure of the yeast nucleosome core particle reveals fundamental changes in internucleosome interactions." Embo J 20 (18): 5207-5218.

133

Whitlock, J. P., Jr. and R. T. Simpson (1977). "Localization of the sites along nucleosome DNA which interact with NH2- terminal histone regions." J Biol Chem 252 (18): 6516-20. Wirbelauer, C., O. Bell, et al. (2005). "Variant histone H3.3 is deposited at sites of nucleosomal displacement throughout transcribed genes while active histone modifications show a promoter-proximal bias." Gene Develop 19 (15): 1761-1766. Wolf, E., A. Vassilev, et al. (1998). "Crystal structure of a GCN5-related N- acetyltransferase: Serratia marcescens aminoglycoside 3-N-acetyltransferase." Cell 94 (4): 439-449. Wolffe, A. P. (1995). Chromatin: Structure and Function . San Diego, Academic Press, Inc. Wolffe, A. P. (1999). Chromatin: Structure and Function . San Diego, CA, Academic Press, Inc,. Worcel, A., S. Han, et al. (1978). "Assembly of newly replicated chromatin." Cell 15 : 969- 977. Wright, W. E., V. M. Tesmer, et al. (1997). "Normal human chromosomes have long G- rich telomeric overhangs at one end." Genes Dev 11 (21): 2801-9. Xu, F., K. Zhang, et al. (2005). "Acetylation in histone H3 globular domain regulates gene expression in yeast." Cell 121 (3): 375-85. Xu, W. T., D. G. Edmondson, et al. (1998). "Mammalian GCN5 and P/CAF acetyltransferases have homologous amino-terminal domains important for recognition of nucleosomal substrates." Mol Cell Biol 18 (10): 5659-5669. Xue, S. and P. N. Rao (1981). "Sodium butyrate blocks HeLa cells preferentially in early G1 phase of the cell cycle." J Cell Sci 51 : 163-71. Yamamoto, T. and M. Horikoshi (1997). "Novel substrate specificity of the histone acetyltransferase activity of HIV-1-Tat interactive protein Tip60." J Biol Chem 272 (49): 30595-30598. Zhang, Z. G., K. Shibahara, et al. (2000). "PCNA connects DNA replication to epigenetic inheritance in yeast." Nature 408 (6809): 221-225. Zheng, C. Y. and J. J. Hayes (2003). "Structures and interactions of the core histone tail domains." Biopolymers 68 (4): 539-546.

134