<<

Insights into Regulation of Human RAD51 Filament Activity During

Homologous Recombination

Dissertation

Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in the Graduate School of The Ohio State University

By

Ravindra Bandara Amunugama, B.S.

Biophysics Graduate Program

The Ohio State University

2011

Dissertation Committee:

Richard Fishel PhD, Advisor

Jeffrey Parvin MD PhD

Charles Bell PhD

Michael Poirier PhD

Copyright by

Ravindra Bandara Amunugama

2011

ABSTRACT

Homologous recombination (HR) is a mechanistically conserved pathway that occurs during meiosis and following the formation of DNA double strand breaks (DSBs) induced by exogenous stresses such as ionization radiation. HR is also involved in restoring replication when replication forks have stalled or collapsed. Defective recombination machinery leads to chromosomal instability and predisposition to tumorigenesis. However, unregulated HR repair system also leads to similar outcomes.

Fortunately, have evolved elegant HR repair machinery with multiple mediators and regulatory inputs that largely ensures an appropriate outcome.

A fundamental step in HR is the homology search and strand exchange catalyzed by the RAD51 recombinase. This process requires the formation of a nucleoprotein filament (NPF) on single-strand DNA (ssDNA). In Chapter 2 of this dissertation I describe work on identification of two residues of human RAD51 (HsRAD51) subunit interface, F129 in the Walker A box and H294 of the L2 ssDNA binding region that are essential residues for salt-induced recombinase activity. Mutation of F129 or H294 leads to loss or reduced DNA induced ATPase activity and formation of a non-functional NPF that eliminates recombinase activity. DNA binding studies indicate that these residues may be essential for sensing the ATP nucleotide for a functional NPF formation.

ii

An intriguing structural motif in the nucleotide-binding subunit interface of

RAD51 is known as the ATP cap. The RAD51 family of recombinases contains a conserved aspartate that forms a salt bridge with the terminal phosphate of ATP, which is the likely source of a nonphysiological cation requirement for the formation of an active

RAD51 NPF. In Chapter 3, I describe the biochemical analyses of a lysine substitution mutation of this conserved aspartate. The prototypical bacterial recombinase RecA as well as most RAD51 paralogs contains a conserved lysine at the analogous position. The

HsRAD51(D316K) substitution mutant possess a reduced turnover from DNA that results in improved recombinase functions. Structural analyses indicates that

HsRAD51(D316K) and its archaebacterial homolog form extended active NPF without the requirement of salt. These studies suggest that HsRAD51(D316) salt bridge may function as a conformational sensor that enhances RAD51 turnover at the expense of recombinase activity.

In Chapter 4, I examination the role of the RAD51 paralog complex RAD51B-

RAD51C as a potential mediator of RAD51 catalyzed D-loop formation. Modeled structures indicate that RAD51, RAD51B and RAD51C appear to have similar domain orientations within a NPF. Furthermore, RAD51B-RAD51C form stable complexes on ssDNA and partially stabilizes RAD51 NPF from the anti-recombinogenic activity of

BLM. At sub-stoichiometric levels, RAD51B-RAD51C may modestly stimulate RAD51 mediated D-loop formation in presence of RPA.

Collectively, these studies provide additional mechanistic and structural insight into the regulation of the RAD51 NPF during the homology search and strand exchange processes that is critical for efficient HR in human cells.

iii

Dedicated to Parents, my wife and my son

iv

ACKNOWLEDGMENTS

It had always been my passion to study biochemical mechanisms of DNA damage repair and it was no surprise that from the first day of my rotation in Fishel Lab I decided to join and pursue the work that I have done so far. I would like to extend my heartfelt gratitude to my PhD advisor Richard Fishel for his continuous support throughout my graduate education in his lab. I greatly appreciate the freedom he allowed me to have to independently address research problems. Many a times I learned the ropes the hard way. But that gave me confidence to pursue a future career as an independent scientist.

The fact that Rick would often challenge my scientific opinions compelled me to be well versed with the literature. I would also like to thank him for carefully reading and editing the drafts of my manuscripts. Finally, I look forward to having him as a mentor for many years ahead for guidance and support.

I would like to thank the members of my committee Chuck Bell, Jeffrey Parvin and Michael Poirier for their guidance and advise throughout my graduate career. I thank

Steve Kowalczykowski, Lorraine Symington, Wolf-Dietrich Heyer and Steve West for stimulating discussions, advice and opinions on the confusing world of Recombination.

I am extremely grateful to Kang-Sup Shim (KS) for advice and teaching me the techniques initially and for many long hours of thoughtful discussions. Also I would like to thank Naduparambil Jacob who has been a wonderful collaborator and a colleague

v

over the years. I am thankful to Sarah Javaid, Kristine Yoder, Michael Mcilhatton and

Samir Acharya for their friendship, advice and helpful discussions. I also like to acknowledge Tom Clanton former director of the Biophysics graduate program, Ralf

Bundschuh current co-director and Susan Hauser for their assistance.

I am grateful to my parents and my brother for believing in me and all their support during my early education. It was the inspiration that I received from my father, being an academic himself, which made me choose the path of academia. Also, I am thankful to my friends Indi, Sharon, Chamika, Mekhala, Suresh, Harshi, Shekar and

Poorani for their friendship and numerous fun times we had that never made us miss home. I am also grateful to my mother-in-law for taking care of our son for a whole year that allowed my wife and me to continue with our education peacefully.

Finally, no words can express the how grateful I am to my dear wife Nirodhini for all her love, support and most importantly her patience throughout our busy years. I am also thankful to my dear son Rehan, who brought us an enormous amount of joy, happiness and meaning to our lives. I want you two to know that even though I have had many shortfalls in trying to be the ideal husband and father, you two are the most important people in my life and the motivation that drives me forward.

vi

VITA

1977...... Born- Paris, France

2003...... B.S. Honors, Biochemistry and Molecular

Biology, University of Colombo, Sri Lanka

2003-2004………………………………….. Teaching Assistant in Biochemistry,

Molecular Biology and Chemistry,

University of Colombo, Sri Lanka

2004-2005 ...... Graduate Teaching Assistant, Department of

Chemistry, University of Iowa, USA

2005-Present ...... Graduate Research Associate, Department

of Molecular Virology, Immunology and

Medical genetics, The Ohio State

University, USA

vii

PUBLICATIONS

1. Amunugama R. and Fishel R. (2011) Subunit interface residues F129 and H294 of human RAD51 are essential for recombinase activity. Plos One 6(8), e23071.

2. Charbonneau, N., Amunugama, R., Schmutte, C., Yoder, K., and Fishel, R. (2009) Evidence that hMLH3 functions primarily in meiosis and in hMSH2-hMSH3 mismatch repair. Cancer Biol Ther 8(14), 1421-20.

3. Su, X., Jacob, N. K., Amunugama, R., Hsu, P.-H., Fishel, R., and Freitas, M. A. (2009) Enrichment and characterization of by two-dimensional hydroxyapatite/reversed-phase liquid chromatography-mass spectrometry. Analytical Biochemistry 388, 47-55.

4. Su, X., Jacob, N. K., Amunugama, R., Lucas, D. M., Knapp, A. R., Ren, C., Davis, M. E., Marcucci, G., Parthun, M. R., Byrd, J. C., Fishel, R., and Freitas, M. A. (2007) Liquid chromatography mass spectrometry profiling of histones. J Chromatogr B Analyt Technol Biomed Life Sci 850, 440-54.

5. Ikura, T., Tashiro, S., Kakino, A., Shima, H., Jacob, N., Amunugama, R., Yoder, K., Izumi, S., Kuraoka, I., Tanaka, K., Kimura, H., Ikura, M., Nishikubo, S., Ito, T., Muto, A., Miyagawa, K., Takeda, S., Fishel, R., Igarashi, K., and Kamiya, K. (2007) DNA damage-dependent acetylation and ubiquitination of H2AX enhances dynamics. Mol Cell Biol 27, 7028-40.

FIELD OF STUDY

Major Field: Biophysics

viii

TABLE OF CONTENTS

ABSTRACT ...... ii

ACKNOWLEDGMENTS ...... v

VITA...... vii

LIST OF TABLES ...... xii

LIST OF FIGURES ...... xiii

CHAPTER 1: Homologous Recombination in Eukaryotes ...... 1

1.1 ABSTRACT ...... 2

1.2 INTRODUCTION TO HOMOLOGOUS RECOMBINATIONAL REPAIR ...... 3 1.3 MEIOSIS...... 4 1.4 DSB REPAIR IN SOMATIC CELLS ...... 6 1.5 RAD52 EPISTASIS GROUP ...... 12 1.6 RECOMBINATION MEDIATORS ...... 17 1.7 DSB REPAIR IN CHROMATIN ...... 33 1.8 POSTSYNAPTIC REMOVAL OF RAD51 ...... 43 1.9 SECOND-END CAPTURE ...... 44 1.10 DOUBLE DISSOLUTION ...... 44 1.11 HOLLIDAY JUNCTION RESOLUTION ...... 45

ix

1.12 HOMEOLOGOUS RECOMBINATION: The interplay between mismatch repair and HR ...... 49 1.13 CONCLUDING REMARKS ...... 50 1.14 FOOTNOTES ...... 51 1.15 REFERENCES ...... 59

CHAPTER 2: Subunit Interface Residues F129 and H294 of Human RAD51 Are Essential For Recombinase Function ...... 90

2.1 ABSTRACT ...... 91

2.2 INTRODUCTION ...... 92

2.3 MATERIALS AND METHODS ...... 94

2.4 RESULTS ...... 98

2.5 DISCUSSION ...... 103

2.6 ACKOWLEDGEMENTS AND FOOTNOTES ...... 105

2.5 REFERENCES ...... 113

CHAPTER 3: The RAD51 ATP Cap Regulates Nucleoprotein Filament Stability 116

3.1 ABSTRACT ...... 117

2.2 INTRODUCTION ...... 118

3.3 MATERIALS AND METHODS ...... 121

3.4 RESULTS ...... 129

3.5 DISCUSSION ...... 138

x

3.6 ACKOWLEDGEMENTS AND FOOTNOTES ...... 142

3.7 REFERENCES ...... 162

CHAPTER 4: Human RAD51B-RAD51C Paralog Heterodimer Complex Enhances and Stabilizes RAD51 Nucleoprotein Filament for D-loop Formation ...... 165

4.1 ABSTRACT ...... 166

4.2 INTRODUCTION ...... 167

4.3 MATERIALS AND METHODS ...... 170

4.4 RESULTS ...... 174

4.5 DISCUSSION ...... 181

4.6 ACKOWLEDGEMENTS AND FOOTNOTES ...... 187

4.7 REFERENCES ...... 202

CHAPTER 5: CONCLUSION...... 208

REFERENCES ...... 212

COMPLETE BIBIOGRAPHY ...... 213

xi

LIST OF TABLES

1.1 Comparison of DSB Repair Factors in Budding Yeast and Human...... 58 2.1 Summary of ATP hydrolysis and nucleotide binding data of HsRAD51 wild type (WT) and HsRAD51(F129V) and HsRAD51(H294V) mutant ...... 112 3.1 Summary of ATP hydrolysis and nucleotide binding data of HsRAD51 wild type (WT) and HsRAD51(D316K) mutant protein...... 159 3.2 Dissociation rate constants of HsRAD51 wild type (WT) and HsRAD51(D316K) ssDNA and dsDNA interactions ...... 160 3.3 X-ray crystallographic data and structure refinement statistics...... 161

xii

LIST OF FIGURES

1.1. DNA double strand break repair (DSBR) by homologous recombination (HR)...... 52 1.2. Alternative DSB repair mechanisms ...... 53 1.3. Break induced replication (BIR) and telomere restoration ...... 54 1.4. HR restores collapsed or stalled replication forks ...... 55 1.5. BRCA2 and its proposed role in HR...... 56 2.1. Cation-induced conformational rearrangement of conserved residues of RadA ...... 106 2.2. Mutation of HsRAD51(F129) and HsRAD51(H294) residues affect ATP turnover ...... 107 2.3. HsRAD51(F129V) and HsRAD51(H294V) are deficient in D-loop formation and strand exchange ...... 109 2.4. F129 and H294 of HsRAD51 are critical for DNA binding in the presence of ATP ...... 111 3.1 ATP Binding and Hydrolysis by HsRAD51 ATP Cap Substitution Mutation ...... 143 3.2 Purification of HsRAD51 and ADP-ATP exchange analysis ...... 145 3.3 HsRAD51(D316K) Exhibits Salt and RPA Independent Strand Exchange Activity ...... 146 3.4 HsRAD51(D316K) Preferentially Binds to Single Stranded DNA ...... 148 3.5 Competition DNA binding by HsRAD51 and HsRAD51(D316K) ...... 150 3.6 Strand Exchange Activity of HsRAD51 and HsRAD51(D316K) with Different Adenosine Nucleotides ...... 152 3.7 HsRAD51(D316K) Displays a Slow Turnover from ssDNA and dsDNA ...... 153 3.8 HsRAD51(D316K) Catalyzes D-loop Formation in the Presence of ATP and Magnesium ...... 155

xiii

3.9 Structural comparison of MvRAD51 and RAD51 ATP Cap Aspartate→Lysine Substitution Mutation ...... 157 4.1 Homology based modeled structures of RAD51, RAD51B and RAD51C indicate similar domain orientations ...... 188 4.2 Sequence alignment of RAD51 homologs and RAD51 paralogs; RAD51B and RAD51C ...... 189 4.3 RAD51B-RAD51C heterodimer forms stable complexes on DNA ...... 190 4.4 RAD51B-RAD51C enhances RAD51 catalyzed D-loop formation in the presence of RPA ...... 191 4.5 Purified proteins used in the D-loop assays ...... 193 4.6 Order of addition of RAD51 and RAD51B-RAD51C for D-loop formation ...... 194 4.7 RAD51B-RAD51C protects RAD51 nucleoprotein filament against anti-recombinase function of BLM ...... 196 4.8 RAD51B-RAD51C does not suppress the overall ATPase activity of the RAD51/RAD51B-RAD51C nucleoprotein filament ...... 198 4.9 RAD51B-RAD51C RAD51B-RAD51C does not alleviate inhibitory effects of RPA ...... 199 4.10 Proposed model for RAD51B-RAD51C mediated stabilization of the RAD51 nucleoprotein filament ...... 201

xiv

CHAPTER 1

Homologous Recombination in Eukaryotes

1

1.1 ABSTRACT

Homologous recombination (HR) is a mechanistically conserved pathway that occurs during meiosis and that ensures maintenance of genomic integrity. During meiosis, HR results in DNA crossover events between homologous that produce the genetic diversity inherent in germ cells. The physical connection established between homologs during the crossover event is essential to facilitate correct segregation. HR is also involved in maintenance of somatic cell genomic stability by restoring replication after a stalled replication fork has encountered a DNA lesion or strand break as well as following exogenous stresses such as ionization radiation that induce DNA double strand breaks (DSBs). The importance of HR can be gauged by the conservation of HR and functions from bacteria to man. Here we review the players and mechanics of eukaryotic HR.

2

1.2 INTRODUCTION TO HOMOLOGOUS RECOMBINATIONAL REPAIR

DNA double-stranded breaks (DSBs) are generated spontaneously by radiation and chemical damage as well as intentionally as part of the chromosome pairing process during meiosis. instability resulting from DSB recombination repair (RR) defects have been linked to a variety of human cancers including hereditary breast cancer

(BRCA1/2) as well as hematopoietic and other solid tumors (Ataxia telangiectasia mutated, ATM; Nijmegen breakage syndrome, NBS; , FANC; Bloom‟s syndrome, BLM) among others (Miki et al. 1994; Ellis et al. 1995; Savitsky et al. 1995;

Tavtigian et al. 1996a; D'Andrea and Grompe 2003). Unlike many repair pathways, RR engenders a complex cascade of responses that include cellular signaling integrated with the physical processes of DSB repair (Hoeijmakers 2001; Lobrich and Jeggo 2007). The

DSB repair reaction itself involves a complex cascade of enzymatic reactions that must manage the chromatin composite on the broken donor DNA in order to search and pair with the assembled chromatin of a homologous acceptor DNA. Deficiencies in any one of the multitudes of steps will affect the outcome of the RR process and ultimately affect genome stability. Understanding of the biophysical events associated with the DSB repair reaction is important since combinatorial chemical strategies are under development, which rely on targeting redundant or overlapping repair pathways that ultimately results in a synergistic therapeutic response in cancer patients (Farmer et al.

2005).

3

1.3 MEIOSIS

All sexually reproducing organisms undergo meiosis: a process that reduces the cellular diploid content to produce haploid gametes. RR has been co-opted and is essential for the completion of meiosis. Meiosis begins with replication that forms sister chromosomes (chromatids) and is followed by a pairing process that spatially associates chromosome homologs (Zickler and Kleckner 1999). The segregation of chromosome homologs is completed in the first meiotic division (meiosis I) and the segregation of chromatids is completed in the second meiotic division (meiosis II); ultimately producing haploid gametes. The mechanism, regulation, and checkpoint functions of meiosis II chromatid segregation appear similar to the well-defined processes associated with mitosis (Marston and Amon 2004). In contrast, meiosis I requires suppression of the tendency to segregate sister chromatids and instead the homologous chromosomes are separated. More than 50% of all spontaneous miscarriages are due to errors in chromosome segregation (non-disjunction) at the first meiotic division (Hassold et al.

2007). Moreover, 90% of Down syndrome cases can be attributed to errors in maternal meiosis (Hassold et al. 1995). With few exceptions, the critical meiosis genes appear identical in all eukaryotes (Heyting 1996; Keeney 2001).

The pairing of homologous chromosomes in meiosis I is a complex process fraught with many pitfalls that may ultimately result in infertility. Homologous chromosome pairing is initiated in Prophase I by the SPO11 product (Klapholz et al.

1985; Keeney 2001) which actively introduces hundreds of DSBs into the sister chromatids (Kleckner 1996). The homologous recombination (HR) repair of these DSBs

4 by the nearest sister is suppressed by the formation of meiosis-specific lateral elements between the chromatids (Cromie and Smith 2007). This leaves the homologous chromosome as the only DNA sequences available for HR repair to restore the exact integrity of the genome (Figure 1.1). The DSBs are first resected by a 5‟3‟ (Alani et al. 1990; Liu et al. 1995). The resulting 3‟ single-stranded DNA

(ssDNA) end is then used in a classic homologous pairing and strand invasion reaction with the chromosome homolog to form a D-loop. Strand invasion requires RAD51 and/or the meiosis-specific DMC1, which are homologs to the prototypical bacterial recombination-initiation protein RecA (Masson et al. 1999; Masson and West 2001). The ssDNA binding (SSB) protein RPA is an essential cofactor in this process (Kantake et al.

2003; Wang and Haber 2004). Mutation of SPO11 or RAD51 result in a dramatic reduction of homologous chromosome pairing, a high frequency of meiosis I non- disjunction, and gamete inviability. In mice, upward of 400 DSB sites are formed that contain RAD51 and RPA beginning in leptotene (Moens et al. 2002). That the DSBs are almost always faithfully repaired is a testament to the accuracy and dependability of the process in the preservation of the many sexual species on earth.

Approximately 90% of the DSB sites are resolved following repair in a process that conversion of one parental homolog DNA sequence to the other parental homolog sequence with concurrent loss of that parental homolog DNA sequence (gene conversion;

(Housworth and Stahl 2003). These events leave the remaining chromosome of both parents intact (Housworth and Stahl 2003). The remaining 10% (40-50 in human) introduce visible chromosomal crossovers known as chiasmata that exchange entire arms

5 of genetic information reciprocally from one parental chromosome to the other (Kneitz et al. 2000). Ultimately, there are two significant events associated with meiotic DSB repair: 1.) genetic information is exchanged between chromosomes that is the basis of modern genetics (Stahl 1979), and 2.) homologous chromosomes become linked via chiasmata that are essential for proper chromosome segregation.

1.4 DSB REPAIR IN SOMATIC CELLS

Eukaryotes have four main pathways that repair DSBs generated spontaneously in somatic cells: HR, Non-homologous end joining (NHEJ), Alternative NHEJ (Alt-NHEJ, also known as microhomology mediated end joining, MMEJ) and single-strand annealing

(SSA; Figure 1.2; (Ciccia and Elledge 2010). The pathway of choice depends on the nature of DSB, the species, cell type and the stage of the cell-cycle stage where the DSB occurs (Heyer et al. 2010).

As discussed above, HR is a template dependent repair process and was long ago recognized to require the formation of a DNA cross-over structure at the site of homology between chromosomes (termed: Holliday Junction; (Holliday 1964). In the 1980‟s a

DSB repair model that involves the formation of a double Holliday Junction (dHJ) was developed based on transformation studies in budding yeast () where a linear plasmid was faithfully integrated into a homologous region of the host genome (Figure 1.1; (Orr-Weaver and Szostak 1983b; Orr-Weaver and Szostak 1983a).

Even though the DSB repair model also explains meiotic recombination products, during mitotic recombination very few crossover events are observed since the sister chromatid

6 appears to be largely used as the template. The use of a sister chromatid strongly suggests that mitotic recombination mainly occurs during late S and G2 phases of the (Heyer et al. 2010). Importantly, crossover events during mitotic recombination could lead to loss of heterozygosity (LOH), which is a common process during tumorigenesis (Wu and Hickson 2006; Heyer et al. 2010; Moynahan and Jasin 2010).

Synthesis dependent strand annealing (SDSA) model was proposed to account for the low numbers of crossover recombination events (Figure 1.2; (Strathern et al. 1982;

Nassif et al. 1994; Ferguson and Holloman 1996). In SDSA once DNA synthesis occurs on the invading strand of a D-loop, it is unwound and displaced such that it may anneal with the second end to prime DNA synthesis on the latter (Figure 1.2). This pathway leads exclusively to noncrossover products. SDSA aside, dHJ based recombination products have been observed between homologous chromosomes in yeast during mitotic

DSB repair, but as a minor pathway (Bzymek et al. 2010).

During NHEJ the broken DSB ends are prevented from resection by the

Ku70/ heterodimer (Mahaney et al. 2009). The strong affinity of /Ku80 for

DSB ends appears to recruit DNA Ligase IV that is capable of sealing the DSB (Figure

1.2). The alternative NHEJ (Alt-NHEJ) pathway appears to resect a 5-25 nt region where microhomology may be used prior to ligation of the ends (Figure 2; (Hartlerode and

Scully 2009; Ciccia and Elledge 2010).

SSA occurs in regions flanked by direct repeat DNA sequences (Figure 1.2;

(Krogh and Symington 2004). Hence, this pathway is seen in higher eukaryotes where direct repeated sequences are prevalent (Lin et al. 1984; Maryon and Carroll 1991;

7

Fishman-Lobell et al. 1992; Krogh and Symington 2004). In SSA, both 5 flanking the

DSB are resected by nucleases and the resulting 3 overhangs annealed by RPA and

RAD52 (Heyer et al. 2010). Since, strand invasion and exchange are not involved, this process is independent of RAD51. 3 single-stranded overhangs are subsequently resected and ligated by nucleases and ligases, respectively (Krogh and Symington 2004).

NHEJ, Alt-NHEJ and SSA are mutation-prone pathways due to the lack of fidelity and loss of genetic information during the repair process.

Collapsed or Stalled Replication Forks - When a replication fork collapses or if a telomere becomes uncapped, a single ended DSB is formed (Lundblad and Blackburn

1993; Garvik et al. 1995; Hackett and Greider 2003). This end may be processed to produce a 3-overhang that can invade a homologous region of the sister chromatid, the homologous chromosome or a homologous region of another chromosome to initiate

DNA synthesis. This process is called Break-Induced Replication (BIR; Figure 1.3).

Use of any other template other than the sister chromatid leads to LOH during BIR.

During mitotic recombination however BIR is disfavored over SDSA apparently because of its slower kinetics (Malkova et al. 2005).

Stalled replication machinery or lesions on leading or lagging strand may lead to formation of DNA gaps or DSBs that often results in replication fork collapse (Heller and

Marians 2006; Branzei and Foiani 2008). In vertebrate cells replication fork collapse occurs in every cell cycle (Tsuzuki et al. 1996; Sonoda et al. 1998). Translesion synthesis (TLS), template switching, or HR can restore replication. TLS is error prone

8 due to the low fidelity of the polymerase employed (Prakash et al. 2005). However, template switching and HR are error-free. Many replication mutants with defective checkpoint activation are dependent on HR gene products for viability (Andreassen et al.

2006; Branzei and Foiani 2007). Uncontrolled HR however during replication fork collapse can lead to gross genomic instability. These observations suggest that cell cycle checkpoints tightly regulate HR pathway to ensure genomic integrity (Andreassen et al.

2006; Branzei and Foiani 2007).

If the nascent strand encounters a nick during replication, the fork may stall and the incomplete replicated strand may undergo resection which can invade the sister chromatid, once the latter is ligated (Figure 1.4, left). After DNA synthesis a resulting partial Holliday Junction may be resolved to reinitiate DNA synthesis at the fork. If the leading strand stalls due to a lesion on its template, the newly synthesized strands may pair via reverse branch migrate to form a chicken foot structure (a pseudo Holliday

Junction; Figure 1.4, middle; (Higgins et al. 1976). Following a short DNA synthesis to fill-in the chicken-foot ssDNA tail, the replication fork may be reinitiated by forward branch migration. A lesion on the lagging strand template would lead to a template switch mechanism where the blocked strand may invade the nascent complementary strand to bypass the lesion by DNA synthesis (Figure 1.4, right). Resolution of the resulting Holliday Junction may then restore replication.

DSB recognition and end resection- The substrate for HR is a ssDNA region with a 3- end generated by resection of the 5 strand of the duplex (Mimitou and Symington

9

2009a). In the event of a DSB by ionization radiation (IR) or chemical agents the terminal nucleotides are often modified structurally or exist as protein bound entities

(McKee and Kleckner 1997; Lobachev et al. 2004; Mimitou and Symington 2009b).

Such modifications posses an obstacle for downstream repair pathways and are effectively removed by nucleases and proteases or both.

The Mre11--Xrs2 (MRX) complex in yeast and the homologous MRE11-

RAD50-NBS1 (MRN) complex in higher eukaryotes recognize DSB‟s (Martin et al.

1999; Lisby et al. 2004; Dupaigne et al. 2008; Wu et al. 2008a; Lisby and Rothstein

2009). Even though MRE11 possesses an inherent 35 nuclease activity, the initial

53 resection is initiated by the Sae2 (CtIP, in mammals) endonuclease in complex with MRX/MRN. Sae2/CtIP appears to process possible adducts on DSB ends for other nucleases to act upon (Krejci et al. 2003; Clerici et al. 2005; Lengsfeld et al. 2007). The cell cycle mediated of Sae2/CtIP by -dependent (CDKs) appears to determine the choice between HR and NHEJ (Huertas et al. 2008; Huertas and

Jackson 2009). In mammalian cells, CtIP is ubiquitinated by BRCA1 during S and G2 phases of the cell cycle, which appears to facilitate its association with DSB sites (Huen and Chen 2010; Huen et al. 2010).

For extensive resection of the 5-strand the Exo1 and Dna2 nucleases as well as the Sgs1-Top3-Rmi1 (STR) complex (BLM-TOP3-RMI1-RMI2 in mammals; (Khavari et al.) is recruited (Zhu et al. 2008; Shim et al. 2010). The MRX complex is implicated in direct recruitment of Exo1 and Dna2 (Zhu et al. 2008). Deletion of Exo1, Dna2 or Sgs1 leads to reduction of resection and the generation of poor HR substrates (Mimitou and

10

Symington 2008; Zhu et al. 2008). Furthermore, Deletion of EXO1 in mammalian cells caused impaired recruitment of RPA and ATR at the DSB sites (Schaetzlein et al. 2007).

The initial resection complex that includes the MRX along with STR and Dna2 has been reconstituted in vitro (Cejka et al. 2010; Niu et al. 2010). Even though, Top3 and Rmi1 stimulate resection by recruiting Sgs1, they are not required for the 5-strand resection processes (Cejka et al. 2010; Niu et al. 2010). Stimulation of Sgs1 helicase activity by

RPA occurs in a species-specific manner, as the bacterial single strand binding protein

(SSB) is unable to stimulate Sgs1 (Cejka et al. 2010; Niu et al. 2010; Nimonkar et al.

2011). RPA also suppresses the inherent 3-endonuclease function of Dna2 while stimulating the 53 exonuclease required for DSB resection (Cejka et al. 2010; Niu et al. 2010). Two functional human resection complexes have been reconstituted in vitro, one comprising MRN-EXO1-BLM-RPA and the other MRN-DNA2-BLM-RPA

(Nimonkar et al. 2011). BLM exhibits direct protein-protein interactions with both

EXO1 and DNA2 (Nimonkar et al. 2008; Nimonkar et al. 2011). Furthermore, the nuclease activity of EXO1 is stimulated by BLM, RPA and MRN.

In the absence of a recombinase or a homologous sequence, resection could continue over several thousand nucleotides at rate of approximately 4 kb/hr in yeast (Zhu et al. 2008; Symington and Gautier 2011). During meiotic recombination resection tracts average ~850 nt (Chung et al. 2010; Zakharyevich et al. 2010). However, the resection length required for recombination between sister chromatids during mitotic recombination has not been determined (Symington and Gautier 2011).

11

1.5 RAD52 EPISTASIS GROUP

Many of the proteins involved in the RR pathway are genes of the RAD52 epistasis group

(Table 1.1). The name was derived from radiation sensitivity genetic screening analysis of budding yeast (Game and Mortimer 1974; McKee and Lawrence 1980; Symington

2002; San Filippo et al. 2008). Among eukaryotes this group of genes are structurally and functionally conserved.

RAD51- RAD51 is unequivocally the central component in HR pathways (Table 1.1). It preserves a high to the prototypical bacterial recombinase RecA

(West 2003; Krogh and Symington 2004; San Filippo et al. 2008). In eukaryotes, homologous pairing and strand exchange is primarily mediated by RAD51 (West 2003;

San Filippo et al. 2008). RAD51 exists as a heptamer in solution (Shin et al. 2003).

Yeast Rad51 and human RAD51 are 43kDa and 37kDa in size, respectively (Shinohara et al. 1992; Benson et al. 1994). The main catalytic ATPase core region that includes the

Walker A/B regions and ssDNA binding domains are conserved among the

RecA/RAD51 recombinases (Aboussekhra et al. 1992; Basile et al. 1992; Shinohara et al.

1992; Wu et al. 2004). However, RAD51 possesses an N-terminal extension that is absent in RecA, whilst in RecA a C-terminal extension is found that is not found in

RAD51 (Baumann and West 1998). These extensions have been implicated as possible double-stranded DNA (dsDNA) binding sites (Aihara et al. 1997; Aihara et al. 1999).

Formation and maintenance of a stable nucleoprotein filament (NPF) is required for the DNA homology search and strand exchange by RecA/RAD51 recombinases

12

(Kowalczykowski 1991; Bugreev and Mazin 2004; San Filippo et al. 2008). RAD51 nucleates on both ssDNA and dsDNA in vitro, forming an extended NPF (Ogawa et al.

1993; Sung and Robberson 1995). Biochemical studies have shown that incorporation of monovalent salts in the reaction buffer biases RAD51 nucleation on ssDNA (Liu et al.

2004b; Shim et al. 2006). This observation appears to presage the possibility that other recombination mediators might be responsible for efficient nucleation of RAD51 on ssDNA. A single RAD51 molecule binds to 3-4 nt or bp, extending the helical pitch by

~50% compared to canonical B-form DNA (Ogawa et al. 1993; Sung and Robberson

1995). Interestingly, when calcium was substituted for magnesium human RAD51 displayed enhance the strand exchange activity in vitro (Bugreev and Mazin 2004). This effect has been attributed to the suppression of ATP hydrolysis by the calcium that appears to enhance the lifetime of the ATP-bound active NPF (Bugreev and Mazin 2004).

This calcium-mediated stimulation is unique to human RAD51 (Bugreev and Mazin

2004).

Budding yeast containing a deletion are viable (Symington 2002).

However, knockout of RAD51 in vertebrates leads to chromosomal instability and embryonic lethality (Tsuzuki et al. 1996; Sonoda et al. 1998). Although no mutations have been reported in the open reading frame, in many cancers and cancer cell lines the expression of RAD51 is increased, presumably providing a replicative advantage to the rapidly dividing cells via its role in the HR repair of collapsed forks (Richardson et al.

2004; Klein 2008; Schild and Wiese 2010).

13

A catalytically conserved lysine residue in the Walker A box [yeast Rad51(K191) and human RAD51(K133)] is essential for ATP binding and hydrolysis. Mutation of this conserved lysine to alanine [Rad51(K191A)] leads to a null phenotype and is a dominant negative phenotypes in diploid cells (Morgan et al. 2002; Krogh and Symington 2004).

Expression of Rad51(K191R) in rad51 null yeast strains leads to resistance to DSB causing agents, suggesting nucleotide binding is sufficient for HR repair in vivo (Sung and Stratton 1996). Human RAD51(K133R) binds ATP but is unable to hydrolyze ATP, similar to the yeast mutation (Morrison et al. 1999; Chi et al. 2006). Overexpression of human RAD51(K133R) in chicken DT40 RAD51 knock out cells confers partial resistance to IR radiation (Morrison et al. 1999). RAD51(K133R) forms a stable NPF on

DNA and has enhanced recombinase activity in vitro (Chi et al. 2006). However, mouse embryonic stem cells generated that expressed RAD51(K133R) display increased sensitivity to DSBs and reduced efficiency of spontaneous sister-chromatid exchanges

(Stark et al. 2002).

ATP binds at the interface region of two adjacent RAD51 monomers within the

NPF (Wu et al. 2004; Wu et al. 2005). The bottom subunit provides the catalytic Walker

A (P-loop) domain while the top subunit shields the nucleotide with an ATP cap containing a conserved proline residue (Wu et al. 2005). For the homology search and strand exchange, ATP binding but not necessarily ATP hydrolysis is required

(Kowalczykowski 1991). When an ATP or a non-hydrolysable ATP analog binds at the subunit interface the NPF adopts an active extended conformation. Several biochemical studies have shown that incorporation of monovalent cations such as ammonium or

14 potassium as well as divalent cations such as calcium extends the NPF ~30% further (Wu et al. 2005; Qian et al. 2006). The NPF exists in the collapsed configuration with ADP

(Qian et al. 2005). The extended conformation is believed to facilitate homology sampling on duplex DNA. The ATP hydrolysis rate of RAD51 is several folds slower compared to the bacterial RecA (Tombline and Fishel 2002). However, this hydrolysis is still important for the dissociation of recombinase from both newly formed heteroduplex

DNA, as well as fortuitously bound RAD51 on duplex DNA (San Filippo et al. 2008).

The RAD51 NPF is not a static structure. Breast cancer susceptibility gene product 2 (BRCA2) is known to nucleate RAD51 NPF formation on ssDNA at the dsDNA-ssDNA junction (Galkin et al. 2005; Yang et al. 2005; Carreira et al. 2009;

Jensen et al. 2010). Many helicases in yeast and human have been identified both in vivo and in vitro that act as anti-recombinases capable of dissociating RAD51 from ssDNA.

These include Srs2 in yeast, BLM and RECQ5 in human (Krejci et al. 2003; Veaute et al.

2003; Bugreev et al. 2007; Hu et al. 2007; Antony et al. 2009). Biophysical studies also indicated that Srs2 augments the Rad51 ATPase activity within the NPF to facilitate rapid protein turnover (Antony et al. 2009). Moreover, both ensemble and single molecule experiments show a direct correlation of recombinase turnover from DNA and its ATPase activity (Kowalczykowski 1991; Bianco et al. 1998; Bugreev and Mazin 2004; Bugreev et al. 2005; Chi et al. 2006; Hilario et al. 2009).

Homology search during the synapsis phase by RAD51/RecA family is by random collision that involves transient non-specific interactions with dsDNA, presumably bound at the secondary DNA binding site (Kowalczykowski 1991; Bianco et

15 al. 1998). The transient triplex DNA formed during the homology search are paramenic and are not topologically interwound (Kowalczykowski 1991). Biophysical studies using fluorescence resonance energy transfer (FRET) and selective substitution of guanine to inosine on both ssDNA and its identical strand on the duplex DNA, showed that human

RAD51 facilitated homology search by a rapid A:T base flipping mechanism (Gupta et al. 1999). Once a homologous sequence is found, strand exchange occurs to produce a topologically interwould plectonemic heteroduplex DNA product (Kowalczykowski

1991; Bianco et al. 1998).

DMC1- DMC1 (Disrupted Meiotic cDNAs) was first isolated from a budding yeast meiotic cDNA library screen (Bishop et al. 1992). DMC1 is only expressed during meiosis and displays considerable homology to RecA/RAD51 family of recombinases

(Table 1.1; (Bishop et al. 1992). During meiosis both Rad51 and Dmc1 co-localize at

DSBs in yeast (Bishop 1994). Disruption of yeast Dmc1 leads to a number of abnormal meiotic phenotypes, including accumulation of DSBs, reduced reciprocal recombination, abnormal synaptonemal complex formation and defective meiotic prophase arrest

(Bishop et al. 1992). Dmc1 exhibits both overlapping yet non-redundant functions to

Rad51 (Tsubouchi and Roeder 2003). Yet, overexpression of Rad51 enables yeast cells to circumvent the defective meiotic phenotype of Dmc1 mutants (Tsubouchi and Roeder

2003). Human DMC1 exists as an octamer in solution (Passy et al. 1999) and functions similar to RAD51 in assays for recombination and ATPase activity in vitro (Sehorn et al.

2004; Bugreev et al. 2005; Sauvageau et al. 2005).

16

1.6 RECOMBINATION MEDIATORS

RPA- (RPA) is a heterotrimeric (70 kD, 32 kD, 14 kD) SSB that binds to ssDNA with high affinity (Table 1.1; (Wold and Kelly 1988; Erdile et al. 1990).

It was first shown to stimulate strand exchange in vitro with strand exchange protein 1

(SEP1) (Heyer et al. 1990; Alani et al. 1992). A similar stimulatory effect was shown later with yeast Rad51 (Sung 1994). RPA has a dual stimulatory role during RAD51 mediated strand exchange. During the presynaptic phase it binds to ssDNA to prevent secondary structure formation that could potentially lead to inhibitory effects during

RAD51 NPF formation (Sugiyama et al. 1997). During the isoenergetic strand exchange phase RPA ensures unidirectional heteroduplex extension by binding to the displaced ssDNA (Eggler et al. 2002; Van Komen et al. 2002). However, during recombination assays in vitro where only RPA and RAD51 are present, if RPA is added to ssDNA prior to the addition of RAD51, strand exchange is inhibited due to the nanomolar affinity of RPA for ssDNA. This inhibition can be overcome by the addition of recombination mediators such as Rad52 or Rad55/Rad57 in yeast Rad51 recombinase reactions, and by the addition of BRCA2 or the RAD51 paralog heterodimer RAD51B-

RAD51C in human RAD51 recombinase reactions (Sung 1997b; Kanaar and

Hoeijmakers 1998; New et al. 1998; Shinohara and Ogawa 1998; Sigurdsson et al. 2001;

Yang et al. 2005; Jensen et al. 2010).

RAD52- Rad52 plays an essential role in HR and SSA and its deletion leads to severe sensitivity to DSB causing agents and defects during meiosis in budding yeast (Table 1.1;

17

(Symington 2002; San Filippo et al. 2008). Electron microscopic (EM) evidence indicates that both yeast and human RAD52 form oligomeric ring structures (Shinohara et al. 1998;

Stasiak et al. 2000; Ranatunga et al. 2001). The EM structure of human RAD52 indicated that the N-terminus is responsible for the formation of a heptameric ring structure and the C-terminus then self-assembles the heptameric rings into a higher ordered structure (Stasiak et al. 2000; Ranatunga et al. 2001). However, two independent

X-ray crystallographic analyses revealed that the yeast Rad52 N-terminal residues 1-201

(1-209 of human RAD52) formed an undecameric (11-subunit) ring (Kagawa et al. 2002;

Singleton et al. 2002). The overall structure resembles a mushroom top with positively charged residues lining a groove on the outside of the ring (Singleton et al. 2002). Even, though no DNA containing structures of RAD52 have been solved, the dimensions of the groove indicate that it is large enough to bind ssDNA in a sequence independent manner that would position the bases away from the protein surface for possible annealing with complementary bases (Singleton et al. 2002; West 2003). The ssDNA binding property of purified yeast Rad52 was first demonstrated by Rothstein and colleagues in 1996 and found to reside in the N-terminus (Mortensen et al. 1996). A similar ssDNA binding pattern was observed when ssDNA-RAD52 complexes were probed for hydroxyl radical hypersensitivity (Parsons et al. 2000).

During meiotic and mitotic DSB repair RAD51 recruitment is dependent on

RAD52 (Gasior et al. 1998; Sugawara et al. 2003; Wolner et al. 2003; Lisby et al. 2004).

Yeast Rad52 has been shown to stimulate Rad51 mediated strand exchange activity by mediating Rad51 NPF formation on RPA coated ssDNA filaments (Sung 1997a; Kanaar

18 and Hoeijmakers 1998; New et al. 1998; Shinohara and Ogawa 1998). This mediator activity is critical to overcome the inhibitory action of RPA, when RPA is added prior to

Rad51. Several methods including yeast-two hybrids, co-immunoprecipitation of whole cell extracts and direct protein-protein affinity pull-down techniques have shown a direct interaction between Rad52 and Rad51 (Shinohara et al. 1992; Milne and Weaver 1993;

Sung 1997a; Krejci et al. 2002). The interaction with Rad51 is mediated by the C- terminal 409-412 residues of Rad52 (Boundy-Mills and Livingston 1993; Krejci et al.

2002). Overexpression of Rad51 alleviates the defective RR phenotype of a (D409-412) C-terminal deletion mutation (Krejci et al. 2002). In vitro and in vivo evidence suggest that human RAD52 interacts with the cognate RAD51 via residues 291-

330, a region that does not share homology with yeast Rad51 (Shen et al. 1996). These results appear to imply a species-specific interaction between RAD51 and RAD52.

Rad52 has also been implicated in yeast Rad51 independent events such as BIR and SSA

(Symington 2002; Krogh and Symington 2004; McEachern and Haber 2006).

Both yeast and human RAD52 have been shown to interact with RPA (Hays et al.

1998; Shinohara et al. 1998; Jackson et al. 2002). Yeast two-hybrid assays indicated

Rad52 interaction with all three subunits of yeast RPA (Hays et al. 1998). However, human RAD52 was shown to bind directly to large (70 kD) and middle (32 kD) subunits of RPA (Jackson et al. 2002). Interestingly, RAD52 interaction with RPA inhibits higher order self-association of RAD52 (Jackson et al. 2002). The strand annealing activity of both yeast and human RAD52 facilitates the second-end capture during HR repair

(McIlwraith and West 2008; Wu et al. 2008b).

19

Even though Rad52 plays an essential role in HR in budding yeast, in vertebrate cells or in cells with BRCA2 homologs, loss of RAD52 gene leads to few phenotypic defects in RR. RAD52 knockdown in mouse embryonic cells and in chicken B-cell line

DT40 cells lines does not cause an apparent sensitivity to IR or DSB causing chemical agents (Rijkers et al. 1998; Yamaguchi-Iwai et al. 1998). In the corn smut Ustilago maydis (which contains the Brh2 BRCA2 homolog) no defects in HR were found in

Rad52 mutants (Kojic et al. 2008). A recent study of human breast cancer cell line suggests that knock-down of RAD52 acts as a synthetically lethal in the presence of

BRCA2 deficiency (Feng et al. 2011). This finding elevates RAD52 as a target for anti- tumorigenic therapy for breast cancer (Feng et al. 2011; Liu and Heyer 2011). A model has been proposed where RAD52 functions in an alternative pathway to BRCA2 (Liu and

Heyer 2011). Human RAD52 does not appear to possess any RR mediator activity in vitro (San Filippo et al. 2008; Jensen et al. 2010). However, chicken DT40 cells were non-viable and exhibited severe HR defects in a double knockdown of RAD52 and the

RAD51 paralog XRCC3 (Fujimori et al. 2001). In addition, U. maydis Rad52 mutants demonstrated an enhanced UV and IR sensitivity when its sole RAD51 paralog rec2 was mutated (Kojic et al. 2008). Collectively, this implies that in human cells RAD52 might function as a recombination mediator in conjunction with any one or a combination of the

RAD51 paralogs, RAD51B, RAD51C, RAD51D, XRCC2 or XRCC3 (Liu and Heyer

2011).

20

BRCA2 - Germline mutations of BRCA2 gene predispose individuals to highly penetrant, autosomal dominant to breast and ovarian cancers as well as predisposition to other types of cancers (Wooster et al. 1994; Wooster et al. 1995; Tavtigian et al. 1996b; Jasin 2002).

Mutations of BRCA2 in metazoans lead to gross chromosomal rearrangements, accumulation of chromosomal breaks, developmental arrest, meiotic defects and increased hypersensitivity of DSB and inter-strand cross-linking (ICL) causing agents

(Sharan et al. 1997; Yu et al. 2000; Moynahan et al. 2001; Xia et al. 2001; Siaud et al.

2004; Martin et al. 2005). Biallelic BRCA2 mutations that lead to expression of truncated forms of BRCA2 predispose individuals to Fanconi Anemia (FA) and the designation as

FANCD1 (Howlett et al. 2002). The similar radiation sensitivity and developmental defective phenotypes observed in RAD51 and BRCA2 deficient cell types suggest that

BRCA2 is intricately involved in RAD51 mediated RR repair (Table 1.1). These assertions have been solidified by an observed interaction between BRCA2 and RAD51 using the yeast two-hybrid system (Mizuta et al. 1997; Sharan et al. 1997).

BRCA2 possesses two spatially distinct RAD51 binding regions. First region involves repeated sequences of BRC motifs and the second RAD51 interaction motif is located at the C-terminus of the protein (C-terminal RAD51 binding domain, CTRB;

Figure 1.5A; (Mizuta et al. 1997; Sharan et al. 1997; Wong et al. 1997; Chen et al. 1998;

Davies and Pellegrini 2007; Esashi et al. 2007; San Filippo et al. 2008). Each BRC repeat consists of about 35 amino acids and several of the residues within each motif is conserved. This conservation is seen among metazoan BRCA2 orthologues (Bork et al.

1996; Bignell et al. 1997). However, the number of repeats in each organism varies. For

21 instance, humans, mice and chicken have eight BRC repeats, Drosophila has four repeats,

Caenorhabditis elegans and U. maydis have a single BRC repeat and the plant species rice and Arabidopsis thaliana have eight and four repeats, respectively (Lo et al. 2003).

The BRC repeats are not functionally equivalent (Holloman 2011). Mutations within

BRC repeats lead to abrogation of RAD51 binding and thus manifests defective DNA repair (Chen et al. 1999). Structural analysis of the human BRCA2 BRC-4 repeat with the core region of RAD51 revealed an interface on BRC-4 that mimics a binding motif of

RAD51 (Pellegrini et al. 2002). This surface was suggested to function as an oligomerization site for RAD51 to facilitate RAD51 NPF formation (Pellegrini et al.

2002). Studies with U. maydis Brh2 also suggested a similar recruitment mechanism of

RAD51 (Yang et al. 2005). All BRC repeats of human BRCA2 bind to RAD51 with variable affinity but with a binding stoichiometry of 1:1 (Carreira and Kowalczykowski

2011). For example, BRC-1, -2, -3 and -4 bind to free RAD51 with a higher affinity compared to BRC-5, -6, -7 and -8 (Carreira and Kowalczykowski 2011). BRC repeats can also modulate the loading of RAD51 onto DNA (Carreira et al. 2009; Shivji et al.

2009; Carreira and Kowalczykowski 2011). RAD51 filament formation on dsDNA leads to a dead-end complex that are recombination deficient both in vivo and in vitro (Solinger et al. 2002; Shah et al. 2010). BRC-1, -2, -3 and -4 are able to suppress the ATP turnover rate of RAD51 and facilitate nucleation on RPA coated ssDNA, while suppressing

RAD51 binding to dsDNA (Jensen et al. 2010; Carreira and Kowalczykowski 2011)

(Figure 1.5B.). This in turn leads to enhanced recombinase activity (Carreira and

Kowalczykowski 2011). The later group, BRC-5, -6, -7 and -8 however, do not enhance

22 preferential filament formation of RAD51 (Carreira and Kowalczykowski 2011). The collective action of these two BRC groups can facilitate RAD51 nucleation and nascent filament formation on ssDNA prior to dissociation of BRCA2 (Carreira and

Kowalczykowski 2011). Similar attenuation of Rad51 ATPase activity is seen with C. elegans BRCA2 homologue BRC-2 (Petalcorin et al. 2006).

The CTRB region that interacts with RAD51 is seen only in vertebrates and this region binds RAD51 filaments, but not free RAD51, likely to stabilize the NPF (Davies and Pellegrini 2007; Esashi et al. 2007). Cyclin dependent (CDK) phosphorylation of BRCA2(S3291) appears to inhibit its interaction with the RAD51 filament leading to its disassembly (Esashi et al. 2005; Davies and Pellegrini 2007). As might be predicted, during when RR is highly active, there is very low

BRCA2(S3291) phosphorylation that gradually increases with the approach of M phase

(Esashi et al. 2005). These results suggest that the CTRB region functions as a cell cycle regulator of RR. In addition, the C-terminus of BRCA2 is essential for the nuclear transport of RAD51 from the cytoplasm since RAD51 lacks a nuclear localization signal

(NLS; (Davies et al. 2001; Gildemeister et al. 2009). Thus, in human pancreatic cancer cell line CAPAN-1 that expresses a truncated version of BRCA2 [BRCA26174delT)]

(Goggins et al. 1996), BRCA2 transportation into the nucleus is compromised and the levels of nuclear RAD51 is greatly diminished (Davies et al. 2001). Finally, it has been recently shown that CTRB region of BRCA2 is essential for protection of stalled replication forks against the MRE11 nuclease by stabilizing the RAD51 nucleoprotein filament (Hashimoto et al. 2010; Schlacher et al. 2011).

23

BRCA2 has also been shown to interact with DMC1 through a mammalian specific 26- amino acid interaction motif containing BRCA2 residues 2386-2411 (Thorslund et al.

2007). This motif contains three critical amino acids BRCA2 F2406, P2408 and P2409

(PhePP motif) that is essential for DMC1 interaction (Thorslund et al. 2007). The N- terminus of BRCA2 has also been shown to interact with RPA in a DNA independent manner by co-immunoprecipitation (Wong et al. 2003) and the cancer predisposing mutation BRCA2(Y42C) compromises this interaction (Wong et al. 2003).

PALB2 (Partner and localizer of BRCA2), also known as FANCN due to its involvement in Fanconi Anemia (Reid et al. 2007), interacts with the N-terminus of

BRCA2 and was shown to be essential for stable nuclear localization, recombination and checkpoint functions of the latter (Xia et al. 2006). The BRCA2(Y42C) also disrupts the interaction with PALB2 (Xia et al. 2006). In addition, PALB2 has been shown to stimulate RAD51 catalyzed D-loop formation by physically interacting with RAD51 and displays a cooperative effect in the presence of RAD51AP1, another stimulator of

RAD51 recombinase activity (Buisson et al. 2010; Dray et al. 2010).

DSS1 (the deleted in split hand/split foot gene) is a small acidic protein first shown to interact with BRCA2 by yeast and mammalian two-hybrid assays (Marston et al. 1999). Disruption of DSS1 leads to compromised RAD51 foci formation and DSB repair in both fungal and mammalian species (Kojic et al. 2003; Gudmundsdottir et al.

2004). DSS1 binds to DNA binding region of BRCA2 (Yang et al. 2002), and in U. maydis Dss1 prevents dimerization of Brh2 allowing the formation of a monomeric

24 functional protein (Zhou et al. 2007). These seemingly contradictory phenotypes will require some resolution in the coming years.

Other than the role of BRCA2 in DSB repair it is also suggested to be involved in post-replication repair of ssDNA gaps by a template switch mechanism due to damage in the original template (Jensen et al. 2010). In fact, studies with Brh2 have confirmed the template switch mechanism in vitro (Mazloum and Holloman 2009).

Recently three independent groups were able to successfully express and purify the full length BRCA2 that had initially been a challenge as a result of its sheer size

(3418 amino acids; (Jensen et al. 2010; Liu et al. 2010; Thorslund et al. 2010). The full length BRCA2 was shown to bind six RAD51 molecules and enabled their nucleation on

RPA coated ssDNA by binding to 3 tailed structures (Jensen et al. 2010; Liu et al. 2010;

Thorslund et al. 2010). Furthermore, as with BRC fragment analysis the full length

BRCA2 suppressed RAD51 filament formation on dsDNA and stabilized RAD51-ssDNA

NPF by suppressing the ATPase activity of the recombinase (Jensen et al. 2010;

Thorslund et al. 2010). EM evidence of BRCA2 bound forked DNA structures illustrates its involvement in DNA replication coupled repair (Thorslund et al. 2010).

RAD51 PARALOGS

Yeast RAD51 paralogs - RAD51 paralogs are products of gene duplication events of

RecA/RAD51 genes that function as accessory proteins in HR repair (Table 1.1; (Lin et al.

2006). Budding yeasts encodes two Rad51 paralogs, Rad55 and Rad57 (Kans and

25

Mortimer 1991; Lovett 1994). Mutants of Rad55 or Rad57 are cold sensitive to DNA damage (Hays et al. 1995). Overexpression of either Rad51 or Rad52 suppresses the

DNA repair defect of these mutants, consistent with the notion that Rad55 and Rad57 function as accessory proteins in recombination (Hays et al. 1995; Johnson and

Symington 1995). In addition, Rad55 has been shown to interact with Rad57 and Rad51 both in vivo and in vitro (Hays et al. 1995; Johnson and Symington 1995). Furthermore, inclusion of the stable heterodimer Rad55-Rad57 in sub-stoichiometric amounts suppresses the inhibitory effects with of RPA in vitro, suggesting an involvement during presynapsis (Sung 1997b). Even though, Rad55-Rad57 paralog heterodimer complex does not exhibit recombinase activity (Sung 1997b), mutation of a Walker A box conserved lysine to alanine in Rad55, but not in Rad57 results in defective meiotic recombination phenotypes (Johnson and Symington 1995). In response to DSB causing genotoxic stress kinase mediated phosphorylation of the Rad55 S2,

S8, S14 and S378 residues has been shown essential for activating homologous recombination (Bashkirov et al. 2000; Herzberg et al. 2006; Janke et al. 2010). In the fission yeast Schizosaccharomyces pombe, mutants of Rhp55 and Rhp57 exhibited similar mutator phenotypes as budding yeast Rad55 or Rad57 mutants, that indicated structural and functional homology between Rad55 and Rad57 with Rhp55 and Rhp57, respectively (Khasanov et al. 1999; Tsutsui et al. 2000).

Vertebrate RAD51 paralogs - Vertebrates five RAD51 paralogs RAD51B

(RAD51L1/ hRec2/ R51H2), RAD51C (RAD51L2), RAD51D (RAD51L3/ R51H3),

26

XRCC2 and XRCC3 that share 20 to 30% homology to RAD51 as well as to one other

(Table 1.1; (Thompson and Schild 2001; Symington 2002). Similar to yeast Rad51 paralogs, these gene products appear to be the result of gene duplication events of an ancestral RecA/RAD51 gene (Lin et al. 2006). However, these paralogs show a high degree of evolutionary divergence from RAD51 as well as from each other (Thompson and Schild 2001). XRCC2 and XRCC3 (X-ray repair cross complementing) were first identified in their ability to complement the extreme sensitivity of irs1 and irs1SF hamster cell lines (Jones et al. 1995; Tebbs et al. 1995; Thacker et al. 1995). Homology searches further identified RAD51B (Albala et al. 1997; Rice et al. 1997; Cartwright et al. 1998), RAD51C (Dosanjh et al. 1998) and RAD51D (Cartwright et al. 1998; Pittman et al. 1998). Yeast two- and three-hybrid analyses, co-immunoprecipitation techniques and biochemical studies have indicated interaction among the five RAD51 paralogs

(Schild et al. 2000; Symington 2002). For example, RAD51B forms a stable heterodimer with RAD51C (Dosanjh et al. 1998; Miller et al. 2002; Lio et al. 2003), while RAD51 interacts with XRCC3 and weakly with RAD51C (Liu et al. 1998; Schild et al. 2000). The latter interaction is improved in the presence of XRCC3 (Schild et al.

2000). RAD51D forms a stable complex with XRCC2 (Braybrooke et al. 2000). In

HeLa cells two discrete complexes containing XRCC3 and RAD51C and the other containing RAD51B, RAD51C, RAD51D and XRCC2 were found (Masson et al. 2001a;

Masson et al. 2001b). Knockout mutants of the RAD51 paralogs in chicken DT40 are viable yet exhibit increased sensitivity to cross-linking agents and ionizing irradiation as well as reduced RAD51 foci formation upon damage induction (Takata et al. 2000;

27

Takata et al. 2001). These phenotypes can be partially corrected by overexpression of

RAD51 (Takata et al. 2001).

Compared to RAD51, the RAD51 paralogs display weaker DNA stimulated

ATPase activities (Braybrooke et al. 2000; Sigurdsson et al. 2001; Lio et al. 2003; Shim et al. 2004). RAD51B and RAD51C bind to ssDNA, dsDNA and 3-tailed dsDNA (Lio et al. 2003). Moreover, RAD51C has an apparent strand exchange activity perhaps by destabilization of dsDNA (Lio et al. 2003). RAD51D preferentially binds to ssDNA

(Braybrooke et al. 2000). RAD51B-RAD51C heterodimer possesses in vitro recombination mediator activity for RAD51 catalyzed strand exchange by suppressing the inhibitory effect of RPA (Sigurdsson et al. 2001). Furthermore, RAD51B-RAD51C is able to suppress the anti-recombinogenic activity of BLM during RAD51 mediated D- loop formation (R. Amunugama and R. Fishel, unpublished data). XRCC2 has been shown to enhance the ATP processing activity of RAD51 by facilitating ADP to ATP exchange by reducing the affinity for ADP (Shim et al. 2004). Both in vivo complexes

RAD51B, RAD51C, RAD51D and XRCC2 and RAD51C-XRCC3 bind to ssDNA, 3- and 5-tailed dsDNA, forked DNA structures and Holliday Junctions (Masson et al.

2001a; Masson et al. 2001b; Compton et al. 2010).

RAD54 - RAD54 is a highly conserved eukaryotic gene of the RAD52 epistasis group

(Table 1.1; (Tan et al. 2003; Heyer et al. 2006; Mazin et al. 2010). RAD54 homologs have been identified in a number of eukaryotes including yeast, Drosophila, plants, zebrafish, chicken, mice and human (Kanaar et al. 1996; Muris et al. 1996; Bezzubova et

28 al. 1997; Kooistra et al. 1997; Petrini et al. 1997; Thoma et al. 2005; Shaked et al. 2006).

A RAD54 homolog has been identified in the archaebacterium Sulfolobus solfataricus, but not in eubacteria (Seitz et al. 2001; Heyer et al. 2006). RAD54 homologs of budding yeast and human share a 66% similarity and 48% homology (Kanaar et al. 1996; Petrini et al. 1997). Budding yeast Rad54 was first discovered in a genetic screen to isolate mutants sensitive for ionizing radiation (Snow 1967; Game and Mortimer 1974). Similar to Rad51 and Rad52 mutants, Rad54 mutants were hypersensitive to ionizing radiation as well as DNA crosslinking and alkylating agents that eventually cause DSBs (Krogh and

Symington 2004). Rad54 mutants display only minor defects in meiotic recombination in yeast due to the presence of the meiotic homolog Rdh54/Tid1 (Klein 1997). RAD54 knock-down in mice causes hypersensitivity to ionizing radiation at embryonic stages but not in adult stages due to rescue by NHEJ repair (Essers et al. 2000). However, all development stages of RAD54 deficient mice are hypersensitive to DNA crosslinking agents (Essers et al. 2000). Mutational analysis of Walker A box conserved lysine indicated that in both mice and yeast ATP hydrolysis of Rad54 is essential for its function in vivo (Clever et al. 1999; Petukhova et al. 1999; Agarwal et al. 2011).

Rad54 expression levels increases during late G1 phase of the cell cycle (Cole et al. 1989; Johnston and Johnson 1995), presumably for HR repair of DSBs during late S and G2 phases (Takata et al. 1998). Rad54 expression levels are up-regulated during

DSB formation and Rad54 foci formation is dependent on Rad51 (Cole et al. 1987; Lisby et al. 2004). However, Rad51 foci formation is not dependent on Rad54, indicating that

Rad54 acts downstream of Rad51 (Lisby et al. 2004). Similarly, RAD54 co-localizes

29 with RAD51 foci following ionizing radiation in mammalian cells (Tan et al. 1999;

Essers et al. 2002).

RAD54 belongs to the Swi2/Snf2 SF2 (Super Family 2) of proteins (Tan et al.

2003; Heyer et al. 2006; Mazin et al. 2010). The Snf2/Swi2 proteins are commonly known for dsDNA dependent ATPase, ATP dependent , DNA translocase and DNA supercoiling activities (Tan et al. 2003; Heyer et al. 2006; Mazin et al. 2010). Like other members of SF1 and SF2 members RAD54 possesses several signature helicase motifs: I, Ia, II, III, IV, V, and VI that constitute the two tandem RecA- like lobes that utilize the energy of ATP binding and hydrolysis for function (Gorbalenya and Koonin 1993; Singleton et al. 2007). However, unlike helicases the SF2 family of proteins does not unwind but translocates on dsDNA (Heyer et al. 2006; Mazin et al.

2010). Also unlike helicases, the ATPase activity of RAD54 is not stimulated by ssDNA nor does it translocate on ssDNA (Swagemakers et al. 1998; Singleton et al. 2007). Both yeast and human RAD54 are strict dsDNA dependent ATPases, with a catalytic turnover rate ranging from 3000-6000 per minute (Mazin et al. 2010). The binding affinity for branched DNA structures such as PX-junctions (partial Holliday Junctions) is approximately 200 times higher than for ssDNA or dsDNA (Bugreev et al. 2006).

RAD54 function has been implicated in all three stages of recombination: pre- synapsis, synapsis and post-synapsis (Tan et al. 2003; Heyer et al. 2006). These include interaction with RAD51 to stabilize the ssDNA NPF, stimulation of homology search and strand exchange by catalyzed by RAD51, chromatin remodeling during the homology

30 search, disruption of RAD51-dsDNA filaments, branch migration of Holliday Junctions and interaction with specific endonuclease to stimulate resolution of Holliday Junctions.

RAD54 interacts with RAD51 in a species-specific manner through its N-terminal domain (Jiang et al. 1996; Clever et al. 1997; Golub et al. 1997; Haseltine and

Kowalczykowski 2009). This interaction is seen both with free RAD51 as well as with the RAD51 ssDNA NPF (Mazin et al. 2003). The ATPase activity of RAD54 is not required for RAD51 NPF stability. This was shown in vivo and in vitro using a RAD54

ATPase deficient mutation where a Walker A box lysine to arginine substitution allows only ATP binding (Mazin et al. 2003; Wolner and Peterson 2005).

Role of RAD54 in stimulating RAD51 mediated three-strand exchange activity and D-loop formation was first demonstrated with recombinant yeast proteins (Petukhova et al. 1998). This stimulation is seen with many RAD54 orthologs and occurs in a species-specific manner (Alexiadis and Kadonaga 2002; Sigurdsson et al. 2002; Mazina and Mazin 2004; Haseltine and Kowalczykowski 2009). Sub-stoichiometric amounts of

RAD54 are sufficient to greatly stimulate the RAD51 recombinase activity in vitro

(Petukhova et al. 1998; Sigurdsson et al. 2002; Mazina and Mazin 2004), indicating that the protein functions in a catalytic manner. In fact, for RAD51-mediated strand exchange stimulation, the ATPase activity of RAD54 is required (Petukhova et al. 1999).

Conversely, RAD51 improves the ATP hydrolysis and translocation ability of RAD54 on dsDNA (Van Komen et al. 2000; Mazina and Mazin 2004).

RAD54 facilitates dissociation of RAD51 from heteroduplex DNA following strand exchange in an ATP dependent manner (Solinger et al. 2002; Li et al. 2007).

31

Overexpression of Rad51 in Rdh54 deficient background leads to arrest of cell growth in yeast due to accumulation of Rad51 on undamaged chromatin (Shah et al. 2010).

Furthermore it was revealed that Rad54 is specialized for removal of Rad51 from damaged induced foci, while Rdh54 is involved in disassembly of Rad51 from undamaged toxic dead-end dsDNA complexes (Shah et al. 2010).

During the post-synaptic phases Rad54 enhances the heteroduplex extension of

Rad51 (Solinger and Heyer 2001). Human RAD54 has a relatively higher affinity for

Holliday Junctions and PX-junctions compared to dsDNA and it exhibits branch migration activity in a multimeric functional complex (Bugreev et al. 2006; Mazina et al.

2007). Furthermore, budding yeast Rad54 and human RAD54 have been shown to physically interact with the Holliday Junction the resolvase Mus81-Mms4 and MUS81-

EME1, respectively, to stimulate Holliday Junction resolution (Mazina and Mazin 2008;

Matulova et al. 2009). Even though, the branch migration activity of RAD54 was not required for MUS81-EME1 stimulation, ATP was required (Mazina and Mazin 2008).

RAD51AP1 - RAD51AP1 (RAD51 Associated Protein 1) previously known as PIR5, enhances RAD51 mediated joint molecule formation by physically interacting with both

RAD51 as well as joint DNA structures and dsDNA molecules (Kovalenko et al. 1997;

Modesti et al. 2007; Wiese et al. 2007). Knockdown of RAD51AP1 in human cells increases genotoxic stress to DSB inducing agents (Modesti et al. 2007; Wiese et al.

2007). However, the detailed role of RAD51AP in RR remains a mystery.

32

1.7 DSB REPAIR IN CHROMATIN

DSB induced modifications - In higher eukaryotes DNA is compacted into chromatin and the basic unit is a nucleosome, which consists of 146 bp of DNA wrapped approximately 1.7 times in left-handed superhelical turns around a tetramer of histones

H3 and H4 with two H2A-H2B dimers (Luger et al. 1997). There are several levels of compaction of chromatin in vivo (Kornberg 1977; Luger et al. 1997). The basic nucleosome structure is arranged into an array of nucleosomes resembling „beads on a string‟ (Thoma and Koller 1977). This array is further compacted into a 30nm filament solenoid through inter-nucleosomal interactions and interactions with linker histones H1 or H5 (Horn and Peterson 2002).

In order to perform DNA replication, repair and transcription the chromatin must be dynamic. The dynamic character is achieved by post-translational modifications

(PMTs) of amino acid residues on both the solvent exposed histone tails as well as the core regions (Kouzarides 2007). The PMTs include phosphorylation, methylation, acetylation, ubiquitination, SUMOylation, ADP-ribosylation and proline-isomerization

(Strahl and Allis 2000; Kouzarides 2007). Collectively these modifications have been termed the „Histone Code‟ (Strahl and Allis 2000; Jenuwein and Allis 2001; Rice and

Allis 2001). Once the histones are modified their affinity for DNA may change. For example, biochemical studies have shown that nucleosomes containing acetylated histone are more mobile than unmodified nucleosomes (Manohar et al. 2009; Simon et al. 2011).

These nucleosomes can then be evicted or pushed from the region of DNA that needs to be replicated, repaired or transcribed by chromatin remodelers using free energy of ATP

33 binding and/or hydrolysis (Gavin et al. 2001; Saha et al. 2002; Flaus and Owen-Hughes

2003; Kassabov et al. 2003; Eberharter and Becker 2004; Ulyanova and Schnitzler 2005;

Lorch et al. 2006).

The initial chromatin mediated response to a DSB is phosphorylation of the C- terminus of either major H2A variant in budding yeast on S129 (often referred to as

H2AX for simplicity (Morrison et al. 2004) or the H2AX variant in mammals on S139

(Rogakou et al. 1998; Downs et al. 2000; Rogakou et al. 2000). H2AX constitutes approximately 10% of the nucleosomal H2A complement and the Phosphorylated form is referred to as H2AX (Rogakou et al. 1998). The phosphorylation is ATM and MDC1 mediated and in yeast it spans several kilobases from the DSB, whist in mammals

H2AX spreads to megabase distances flanking the DSB (Rogakou et al. 1999; Downs et al. 2000; Harper and Elledge 2007). MDC1 directly interacts with H2AX via its C- terminal BRCT motifs (Stucki et al. 2005). Upon DSB formation MDC1 is phosphorylated by Casein Kinase 2 (CK2) on its N-terminal S-D-T tri-amino acid repeats and these phospho-domains interact with FHA and BRCT repeats of NBS1 in the MRN complex, that facilitates its recruitment to the DSB site (Chapman and Jackson 2008;

Melander et al. 2008; Spycher et al. 2008). Phosphorylated MDC1 once recruited to the

DSB site functions as a positive feedback regulator by binding to H2AX via its BRCT domain and to ATM through its FHA domain, respectively, to facilitate ATM mediated additional phosphorylation of H2AX to amplify the DNA damage signal (Lou et al.

2006). Sustaining H2AX flanking a DSB is critical for the recruitment of downstream repair factors (Celeste et al. 2002).

34

H2AX functions as a molecular beacon to recruit the cohesion complex that is involved in linking sister-chromatids during the post-replicative phase of the cell cycle

(Strom et al. 2004; Unal et al. 2004; Strom et al. 2007). This process prevents loss of heterozygosity (LOH) during mitotic HR by allowing the broken strand to use the sister- chromatid as the donor template. Studies on budding yeast indicate that H2AX recruits

ATP-dependent chromatin remodeling factor INO80 through direct interaction (Morrison et al. 2004; van Attikum et al. 2004). Furthermore, ssDNA formation that functions as the substrate for RAD51 nucleoprotein filament assembly is compromised in arp8 INO80 subunit mutants in yeast (van Attikum et al. 2004). The histone acetyl transferase

(HAT) NuA4 has also shown to interact with H2AX and appears to acetylate histone H4 following DSB formation (Downs et al. 2004).

H2AX(Y142) is also constitutively phosphorylated by the WSTF (Williams-

Beuren syndrome transcription factor) kinase (Xiao et al. 2009). Following that advent of a DSB H2AX(Y142) is dephosphorylated by the EYA1/EYA3 phosphatases (Cook et al. 2009; Krishnan et al. 2009). Interestingly, MDC1 interaction with H2AX depends on a dephosphorylation of H2AX(Y142) and the relative amount of dephosphorylation determines the recruitment of either pro-apoptotic or DNA repair factors to the damaged site (Cook et al. 2009). This has suggested that the H2AX(Y142) phosphorylation- dephosphorylation is a “molecular switch” in response to DSB damage (Stucki 2009).

An essential role for histone ubiquitination during the DSB repair response emerged after the discovery of the E3 RNF8 at DSB foci (Huen et al.

2007; Kolas et al. 2007; Mailand et al. 2007). RNF8 is recruitment to DSB is mediated

35 by the interaction between its FHA domain with the phosphorylated motifs of MDC1

(Huen et al. 2007; Kolas et al. 2007; Mailand et al. 2007). RNF8 was shown to catalyze the ubiquitinylation of histone H2A and H2AX upon DSB formation, and knockdown of

RNF8 or disruption of FHA domains leads to failure to recruit the checkpoint activating proteins BRCA1 and 53BP1 to the DSB (Kolas et al. 2007; Mailand et al. 2007).

Furthermore, depletion of the E2 ubiquitin adapter UBC13 compromises RNF8 function (Huen et al. 2007; Kolas et al. 2007). UBC13 is an essential ubiquitin adapter for HR that is also recruited to DSBs (Ikura et al. 2007; Zhao et al. 2007). However,

RNF8 mediated ubiquitination is not sufficient to sustain the damage signal at DSB foci and another ubiquitin E3 ligase, RNF168, that acts with UBC13 to amplify the ubiquitination signal via K63 linked ubiquitination of H2A and H2AX (Doil et al. 2009;

Stewart et al. 2009). Another non-proteolytic E3 ligase, HERC2 is recruited to damage- induced foci and also forms a complex with RNF8 and RNF168 to extend their retention at the repair site (Bekker-Jensen et al. 2010). The K63 linked polyubiquitinated histones function as substrates for BRCA1 A-complex binding that includes BRCA1/BARD1,

ABRAXAS, RAP80, BRCC36 through the ubiquitin-interaction motif (UIM) of RAP80

(Wang and Elledge 2007). ABRAXAS is thought to recruit RAP80 to the DSB site

(Wang et al. 2007). Just as ubiquitination of H2AX is critical for DSB repair factor recruitment, deubiquitination is also tightly regulated during the damage response. In fact, the deubiquitinating enzymes BRCC36 and OTUB1 are simultaneously recruited damage-induced foci (Sobhian et al. 2007; Nakada et al. 2010). BRCC36 is part of the

36

BRCA1 A-complex and OTUB1 physically binds to UBC13 and inhibits RNF8 and

RNF168 mediated polyubiquitination (Sobhian et al. 2007; Nakada et al. 2010).

SUMOylation also appears to be essential for effective DSB damage response.

Recently it was shown that SUMO1, SUMO2 and SUMO3 are recruited to damage- induced foci along with the E3 ligases PIAS1 and PIAS4 (Galanty et al. 2009; Morris et al. 2009). SUMOylation is critical for productive assembly of BRCA1, 53BP1 and

RNF168 at damage-induced foci and PIAS mediated SUMOylation of BRCA1 leads to increased ubiquitin ligase activity of BRCA1/BARD1 heterodimer in vitro (Galanty et al.

2009; Morris et al. 2009).

Histone H2B(K120) has recently been shown to be ubiquitinylated by the E3 ligase RNF20-RNF40 heterodimer following DSB damage. The H2B(K120ub) modification leads to the recruitment of chromatin remodeling factor SNF2h to the DSB

(Moyal et al. 2011; Nakamura et al. 2011).

Other histone modifications include constitutively methylated histone H3(K79)

[H3(K79me)] that has been shown to provide an interacting domain for 53BP1 upon relaxation of higher order chromatin structure during a DSB (Huyen et al. 2004).

Furthermore, 53BP1 also interacts with the dimethylated form of histone H4(K20)

[H4(K20me2)] via its tandem tudor domains (Botuyan et al. 2006). HAT acetylation of histones is common during DNA replication and repair (Kouzarides 2007; Henikoff

2008). Acetylation is found not only on the histone tails but also on the core regions. For example, histones H3(K9) as well as H3(K56) are acetylated by GCN5/ KAT2A and p300 in response to DSB damage in human cells (Das et al. 2009; Tjeertes et al. 2009). It

37 is believed that H3(K56ac) influences the mobility of nucleosomes by neutralizing the positive charge on the lysine and mobilizing the entry-exit region of the nucleosome (Xu et al. 2005; Henikoff 2008). Finally, TIP60/Esa1 has been found to acetylate H4 and

H2AX during DSB repair (Ikura et al. 2000; Bird et al. 2002; Kusch et al. 2004; Murr et al. 2006).

ATP dependent chromatin remodeling - A yeast system that employed a galactose inducible HO endonuclease-provoked single DSB at a defined position in the “mating- type” (MAT) followed by chromatin immunoprecipitation allowed monitoring of chromatin remodeler and HR repair machinery recruitment at the recipient and donor sites in real-time (Azar 2011; Hicks et al. 2011). Once HO endonuclease cleaves the specific site within the MAT locus, that is otherwise protected by highly positioned nucleosomes, the break can be repaired by either NHEJ or HR (Osley et al. 2007). The later pathway is employed if the donor sequence HMRa or HML is present (Lee et al.

1998; Hicks et al. 2011). These studies have identified several chromatin remodelers recruited to DSB site including INO80, SWI/SNF, RSC, SWR1, RAD54 and TIP60

(Cairns et al. 1996; Ikura et al. 2000; Koyama et al. 2002; Cai et al. 2003; Martens and

Winston 2003; Wolner et al. 2003; Doyon et al. 2004; Kusch et al. 2004; Mizuguchi et al.

2004; Morrison et al. 2004; van Attikum et al. 2004; Chai et al. 2005; Shim et al. 2005;

Tsukuda et al. 2005; Wolner and Peterson 2005; Kruhlak et al. 2006; Papamichos-

Chronakis et al. 2006; Kent et al. 2007; Shim et al. 2007; Sinha et al. 2009)].

38

INO80 (Inositol auxotroph 80) is a multi-subunit chromatin remodeling complex that was first characterized in a budding yeast mutant strain that exhibited defective transcription activation following inositol depletion (Table 1.1; (Ebbert et al. 1999).

Relatively widely studied compared to other ATP-dependent chromatin remodelers, it is composed of several subunits that are shared by yeast and other eukaryotes (Osley et al.

2007). These core subunits include INO80 ATPase, two AAA+ ATPases (Rvb1 and

Rvb2 in yeast, RuvB-like 1 and RuvB-like 2 in human), actin and actin-related proteins

Arp4, Arp5 and Arp8, and INO80 subunits (Ies2 and Ies6) (Conaway and Conaway

2009). In addition, the yeast Ino80 complex contains unique polypeptides Taf14, HMG,

Nhp10 and Ies1, Ies3, Ies4, and Ies5 whereas the human homolog contains YY1, Uch37 and NFRKB (Conaway and Conaway 2009). The Ino80ATPase subunit of INO80 appears to essential for its cellular function and for ATP dependent chromatin remodeling in vitro (Shen et al. 2000; Shen et al. 2003; Jonsson et al. 2004). Ino80 is recruited to the

HO endonuclease DSB within an hour (Morrison et al. 2004; van Attikum et al. 2004;

Tsukuda et al. 2005). In yeast, deletion of Ino80ATPase, Arp5 or Arp8 subunits leads to increased sensitivity to DSB inducing agents (Morrison et al. 2004; van Attikum et al.

2004; Tsukuda et al. 2005). Recruitment of Ino80 to the DSB is dependent on a specific interaction between the Nhp10 (an HMG-like subunit of the Ino80 complex) and -H2AX and the loss of either component leads to compromised DSB repair (Morrison et al.

2004). However, two conflicting observations were reported with respect to recruitment of repair mediators at the DSB. One group observed that resection of 5 strands at the break was compromised in Arp8 and H2A mutants (van Attikum et al. 2004), while the

39 other reported that even though the resection occurred regularly in the Ino80 mutant,

Rad51 recruitment was delayed (Tsukuda et al. 2005).

SWI/SNF chromatin remodeling complex was first identified in two genetic screens in budding yeast (Table 1.1). The first gene regulates the mating type switch

(SWI) and the other regulates sucrose non-fermenting (SNF) phenotypes (Neigeborn and

Carlson 1984; Stern et al. 1984; Breeden and Nasmyth 1987). The multi-subunit complex consists of 9-12 subunits and has shown to possess ATP-dependent chromatin remodeling activity in vitro (Cairns et al. 1994; Cote et al. 1994). Several homologs of

SWI/SWF have been identified in metazoans; for example in Drosophila BRM (Brahma) and BAPs (BRM-Associated Proteins) were characterized as SWI/SNF homologs and in human BRM, BRG1 (Brahma-Related Gene1) and BAFs (BRM- or BRG1- Associated

Factors) are found as SWI/SNF complexes (Khavari et al. 1993; Dingwall et al. 1995;

Tamkun 1995; Reisman et al. 2009). SWI/SNF remodels nucleosomes both by nucleosome sliding and nucleosome ejection (Saha et al. 2006). Its activity is implicated both in transcription and DSB repair (Martens and Winston 2003; Chai et al. 2005; Osley et al. 2007). SWI/SNF involvement in DSB repair is extensively studied in yeast.

Although, Rad51 NPF formation does not require chromatin remodelers for homology searches and capture even on positioned nucleosomal surfaces in vivo or in vitro

(Sugawara et al. 2003; Wolner and Peterson 2005; Papamichos-Chronakis et al. 2006;

Sinha and Peterson 2008), Swi/Snf remodelers are essential for recombinational repair within heterochromatin (Sinha et al. 2009). When nucleosomal donor sequences are constrained by Sir2, Sir3 and Sir4 structural proteins that are found at telomeres and

40 silent mating-type loci, remodeling activity of Swi/Snf was required for efficient joint molecule formation (Sinha et al. 2009). Interestingly, Rad54, Ino80, RSC and Swr1 were incapable of promoting joint molecule formation within heterochromatin (Sinha et al.

2009). Several mammalian SWI/SNF complexes have also been shown to interact with

DSB repair response proteins such as BRCA1, and with Fanconi Anemia pathway proteins (Bochar et al. 2000; Otsuki et al. 2001; Lee et al. 2002).

RSC (Remodel Structure of Chromatin) is another multi-subunit ATP dependent chromatin remodeler that is rapidly recruited to DSB site upon damage (Chai et al. 2005).

It is homologous to the SWI/SNF complex (Cairns et al. 1996). Its rapid recruitment to

DSB that coincides with MRX recruitment suggests that initial nucleosome remodeling at the DSB might facilitate DNA end processing by the MRX complex (Kent et al. 2007;

Osley et al. 2007; Shim et al. 2007). RSC is also required for loading of cohesins during

DSB repair to ensure repair occurs between sister chromatids during mitotic HR repair

(Shim et al. 2005; Shim et al. 2007). However, there is also evidence that suggest RSC might be involved in during the latter stages of HR repair pathway, particularly during

DNA ligation step after DNA synthesis (Chai et al. 2005).

SWR1 is closely related to INO80 that is recruited to DSB site in an H2AX dependent manner (Downs et al. 2004; van Attikum et al. 2007). However, its involvement in DSB repair is different than the role of INO80. SWR1 is known for exchanging histone H2A with the H2AZ variant (Krogan et al. 2003; Mizuguchi et al.

2004). SWR1 is also implicated in regulating H2AX levels. Inactivation of SWR1 ceases H2AZ incorporation at nucleosomes surrounding the DSB, however, this restores

41

H2AX levels and checkpoint adaptation; that functions antagonistically to INO80

(Papamichos-Chronakis et al. 2006).

Other than the acetylation of histones H4 and H2AX by TIP60 following DNA damage (Ikura et al. 2000; Bird et al. 2002; Kusch et al. 2004; Murr et al. 2006), the

Drosophila TIP60 has been shown to exchange acetylated histone H2Av (H2AX homolog) with unmodified H2Av using the domino/p400 ATPase (Kusch et al. 2004; Xu et al. 2010). The conjunctional action of TIP60/p400 that leads to nucleosomal destabilization at the DSB is required for efficient RNF8 mediated ubiquitination of histones and recruitment of BRCA1 and 53BP1 for DSB response (Xu et al. 2010).

Furthermore, the exchange of acetylated H2Av with unmodified H2Av might lay a critical for attenuation of DSB signal propagation (Osley et al. 2007).

RAD54 appears to have the ability to remodel chromatin (Haushalter and

Kadonaga 2003). RAD54 from budding yeast, Drosphila and human has been shown to posses ATP-driven chromatin remodeling activity in vitro using assembled mononucleosomes and nucleosomal arrays (Alexiadis and Kadonaga 2002; Alexeev et al.

2003; Jaskelioff et al. 2003; Zhang et al. 2007). The RAD51 nucleoprotein filament stimulates chromatin remodeling activity as well as the dsDNA-dependent ATPase activity of RAD54 (Tan et al. 2003; Heyer et al. 2006; Mazin et al. 2010). Strand exchange by RAD51 is greatly enhanced by RAD54 in chromatinized substrates

(Alexiadis and Kadonaga 2002; Zhang et al. 2007), even though for the initial homology capture by RAD51 nucleoprotein filament on a chromatin substrate RAD54 is not required (Sinha and Peterson 2008). Interestingly, a specific protein-protein interaction

42 between the amino terminus of histone H3 and RAD54 has also been reported, implying a specific role in RAD54 mediated chromatin remodeling (Kwon et al. 2007).

1.8 POSTSYNAPTIC REMOVAL OF RAD51

Postsynaptic RAD51 turnover plays a critical role in regulating the HR repair pathway

(Table 1.1). The deproteinization step in conventional strand exchange studies in vitro obliterates the possible analysis of RAD51 dissociation from dsDNA (Symington and

Heyer 2006). Furthermore, unlike RecA that dissociates from dsDNA upon ATP hydrolysis (Kowalczykowski 1991), RAD51 remains bound to heteroduplex DNA; probably due to its intrinsic slow ATP hydrolysis. These results suggested that eukaryotes accessory proteins have evolved to facilitate RAD51 removal from the nacent heteroduplex to allow the 3-end to prime DNA synthesis (Symington and Heyer 2006;

San Filippo et al. 2008). RAD54 has been shown to dissociate RAD51 from dsDNA

(Solinger et al. 2002; Heyer et al. 2006). Recent studies on the nematode C. elegans identified two more gene products, a helicase HELQ-1 (homologous to human HEL308) and the single RAD51 paralog RFS-1, that are essential for postsynaptic RAD51 turnover during meiotic homologous recombination (Ward et al. 2010). Both these gene products have been shown to have synthetic lethal interactions (Ward et al. 2010). It is suggested that once RAD54 removes the RAD51 from the 3 end of the invading strand HELQ-1 and RFS-1 might be involved in removing the remaining RAD51 from the heteroduplex

DNA (Ward et al. 2010; Williams and Michael 2010).

43

1.9 SECOND-END CAPTURE

When RAD51 nucleoprotein filament forms a D-loop the displace strand could potentially pair with the second 3 overhang that is produced. This process is called second-end capture. In yeast and mammalian cells this process is mediated by RAD52 through its inherent ability of pair with RPA-coated ssDNA (McIlwraith and West 2008;

Nimonkar et al. 2009; Shi et al. 2009; Sugiyama and Kantake 2009). In U. maydis Brh2 is also capable to catalyzing second-end capture in conditions where annealing by

RAD52 is inhibited (Mazloum and Holloman 2009).

1.10 DOUBLE HOLLIDAY JUNCTION DISSOLUTION

In the classical DSB repair model, after the second end capture, DNA synthesis and ligation results in a dHJ which can be either dissolved or resolved. During mitotic recombinational repair the former is the preferred pathway (Wu and Hickson 2006;

Bernstein et al. 2010; Heyer et al. 2010). dHJ dissolution is mainly mediated by RecQ helicases (Table 1.1; (Bernstein et al. 2010). The budding yeast RecQ homolog Sgs1

(fission yeast Rqh1) interacts strongly with the Top3 topoisomerase (Gangloff et al.

1994; Bennett et al. 2000; Fricke et al. 2001; Ahmad and Stewart 2005). In human cells the homologous RecQ helicase BLM strongly interacts with TOP3 (Johnson et al. 2000;

Wu et al. 2000). This helicase-topoisomerase complex has been shown to interact with yeast Rmi1/ Nce4 (RMI1/ BLAP75 in human cells) DNA binding protein. In budding yeast dHJ dissolution is mediated by Sgs1-Top3-Rmi1 (Chang et al. 2005; Mullen et al.

44

2005). In human cells the dHJ dissolution complex is comprised of BLM-TOP3-RMI1-

RMI2 (Yin et al. 2005; Bernstein et al. 2010).

1.11 HOLLIDAY JUNCTION RESOLUTION

Weisberg and colleagues reported the first biochemical evidence for an enzyme that has

Holliday Junction resolution activity in 1982. They identified the T4-endonuclease activity of bacteriophage T4 that was capable of cutting the branched structures of the phage genome before it was packaged into new phage particles (Mizuuchi et al. 1982;

Liu and West 2004; West 2009; Svendsen and Harper 2010). Soon afterwards, Holliday

Junction resolvases were identified in several organisms including budding yeast mitochondrial Cce1 (from cell-free extracts), fission yeast Ydc2 and endonuclease I of the bacteriophage T7 (de Massy et al. 1984; Symington and Kolodner 1985; West and

Korner 1985; Kleff et al. 1992; Whitby and Dixon 1997; Oram et al. 1998). The first prokaryotic resolvase to be identified was E.coli RuvC (Connolly et al. 1991; Sharples and Lloyd 1991). Biochemical characterization of RuvC revealed that the resolvase bound the Holliday Junction as a dimer and unfolds the antiparallel stacked-X structure

(observed in the presence of divalent cation Mg++) into an open planar structure

(observed in the absence of divalent ions) and nicks strands of the same polarity

(Connolly and West 1990; Connolly et al. 1991; Dunderdale et al. 1991; Iwasaki et al.

1991; Sharples and Lloyd 1991; Takahagi et al. 1991; West 1997). RuvC activity in bacterial cells are closely associated with the RuvA, the tetrameric protein that binds to

Holliday Junctions and unfolds it into an open-planar structure and RuvB, an hexameric

45

ATP hydrolysis driven translocase, which promotes branch migration (Parsons et al.

1995a; Parsons et al. 1995b; Hargreaves et al. 1998; Ariyoshi et al. 2000; Yamada et al.

2002; West 2003).

Holliday Junction resolvase activity in mammalian systems was first observed in extracts prepared from homogenized calf thymus tissues (Elborough and West 1990).

Similar resolvase activity was later observed in extracts prepared from cultured cells

(Hyde et al. 1994; Constantinou et al. 2001). Given the complexity of eukaryotic and the stringency of maintaining its integrity, eukaryotic cells have evolved multiple pathways and resolvases to process Holliday Junctions (Klein and Symington

2009). This in turn made identification of single mutants defective in Holliday Junction resolution challenging in eukaryotic model organisms (Symington and Holloman 2008).

In 2001, Mus81-Eme1 (Mus81-Mms4 in budding yeast) heterodimer was identified in fission yeast as an endonuclease capable of cleaving Holliday Junctions as well as branched DNA structures (Table 1.1; (Boddy et al. 2001; Chen et al. 2001; Hollingsworth and Brill 2004). Mus81 is homologous to the XPF subunit of the ERCC1-XPF nucleotide excision repair endonuclease (Boddy et al. 2001; Chen et al. 2001). Eme1 is a non-catalytic subunit (Boddy et al. 2001; Chen et al. 2001; Svendsen and Harper 2010).

Mus81-Eme1 depletion caused some meiotic defects and stalled replication forks (Boddy et al. 2001; Chen et al. 2001). Meiotic defective cells could be rescued by ectopic expression of bacterial Holliday Junction resolvases (Boddy et al. 2001). In 2003, human

MUS81-EME1 was characterized as a replication fork/flap endonuclease that is essential to maintain the integrity of replication, even though it possessed inefficient Holliday

46

Junctions resolvase in vitro (Ciccia et al. 2003). In the case of budding yeast, Mus81 deletion only exhibited modest decrease in crossover formation during meiosis, implying an that Mus81-Eme1 is not the sole resolvase complex of Holliday Junctions (de los

Santos et al. 2001; de los Santos et al. 2003). Similar minor meiotic defects were observed in MUS81 knockout studies in mice (McPherson et al. 2004; Dendouga et al.

2005). Search for a RuvC type Holliday Junction resolvase in eukaryotic cells had been quite challenging by conventional sequence homology based queries due to the absence of conservation of primary amino acid sequences (West 2009). However, tertiary structure level conservation was seen among most of the identified Holliday Junction resolvases categorizing them into two main superfamilies of integrase and nuclease (West

2009). Identification of a Holliday Junction resolvase in eukaryotes that resembled bacterial RuvC was inadvertently assigned to RAD51C-XRCC3 heterodimer that was isolated from mammalian cells (Liu et al. 2004a; Symington and Holloman 2008).

However, the absence of an apparent nuclease domain and the inability of recombinant

RAD51C-XRCC3 to recapitulate the Holliday Junction resolvase activity led to questions regarding the assignment of the Holliday Junction resolvase to the RAD51C-XRCC3 heterodimer (Sharan and Kuznetsov 2007; Svendsen and Harper 2010). In 2008, after a tedious effort by the West laboratory (the group that suggested RAD51C-XRCC3 was a

Holliday Junction resolvase), the eukaryotic ResA Holliday Junction resolvase was identified (Ip et al. 2008). It appears that ResA co-purifies with RAD51C-XRCC3 isolated from human cells (Ip et al. 2008; Symington and Holloman 2008). Two complementary approaches were followed in identification of the ResA complex. One

47 was from fractionation of HeLa cell extracts from column chromatography followed by separating the final fraction on an SDS-PAGE gel (Ip et al. 2008). Following renaturation, one of the isolated bands still possessed Holliday Junction resolution activity in vitro, after mass spectroscopic (MS) analysis this novel protein was named

GEN1 (Ip et al. 2008). In the other approach, fission yeast library was screened where each gene was TAP (tandem affinity purification) tagged at the C-terminus (Ip et al.

2008). Using MS, Western blotting and in vitro Holliday Junction cleavage assays, along with Mus81-Eme1, the GEN1 ortholog in yeast Yen1 was identified (Ip et al. 2008).

Both GEN1 and Yen1 are members of Rad2/ XPG family of nucleases (West 2009) and they cleave Holliday Junctions symmetrically, resembling bacterial RuvC activity (Ip et al. 2008).

Recently another group of Holliday Junction resolvases, namely SLX1-SLX4 complex were identified (Table 1.1; (Andersen et al. 2009; Fekairi et al. 2009; Munoz et al. 2009; Svendsen et al. 2009). Even though Slx1 was conserved, conventional homology searches did not indicate a conservation of Slx4 outside of the yeast genome

(Fricke and Brill 2003; Klein and Symington 2009; Svendsen and Harper 2010). By more refined in silico analyses, human SLX4 was identified as BTBD12 which was the ortholog Drosophila MUS312 and fungal Slx4 (Fekairi et al. 2009; Munoz et al. 2009;

Reymer et al. 2009; Svendsen et al. 2009). In a separate study BTBD12 was identified as a phosphorylation substrate of ATM/ATR kinases (Matsuoka et al. 2007; Svendsen et al.

2009). SLX1 possesses the endonuclease domain for Holliday Junction cleavage, while

SLX4 acts as a protein interacting scaffold that interacts with multiple nucleases that

48 cleave Holliday Junctions both symmetrically and asymmetrically (Klein and Symington

2009; Svendsen and Harper 2010). In fact, SLX4 has been implicated in multiple genome maintenance pathways including replication and repair (Fekairi et al. 2009; Klein and Symington 2009; Munoz et al. 2009; Reymer et al. 2009; Svendsen et al. 2009).

Because SLX4 interacts with other Holliday Junction resolvases such as MUS81-EME1, when SLX1-SLX4 complex was isolated from human cells symmetrical cleavage of

Holliday Junctions was not observed (Matsuoka et al. 2007; Fekairi et al. 2009; Svendsen et al. 2009). However, bacterially expressed the SLX1-SLX4 heterodimer with a truncated SLX4 region that does not interact with MUS81-EME1, did possess symmetrical cleavage ability of Holliday Junctions (Matsuoka et al. 2007; Fekairi et al.

2009; Klein and Symington 2009; Svendsen et al. 2009). Among the many interacting partners of mammalian SLX4 are proteins of diverse functions. These include the endonucleases SLX1, ERCC4-ERCC1 and MUS81-EME1; mismatch repair heterodimer

MSH2-MSH3; telomere proteins TRF2/RAP1 and polo-like kinase PLK1 (Matsuoka et al. 2007; Fekairi et al. 2009; Klein and Symington 2009; Munoz et al. 2009; Svendsen et al. 2009).

1.12 HOMEOLOGOUS RECOMBINATION: the interplay between mismatch repair and HR

Mismatch repair (MMR) is a conserved process that plays an important role in maintaining genome integrity by correcting DNA mismatches formed during replication and recombination (Kolodner 1996; Kolodner and Marsischky 1999; Harfe and Jinks-

49

Robertson 2000). MMR repair ensures that HR occurs between perfectly homologous sequences and suppresses recombination between sequences that contain partial homology (homeologous recombination) (Harfe and Jinks-Robertson 2000). Genetic studies in budding yeast indicates that even a single mismatch reduces the recombination rates by at least four-folds compared to recombination between substrates of perfect homology (Datta et al. 1996; Datta et al. 1997). In yeast Msh2-Msh6, Msh2-Msh3,

Mlh1-Pms1 MMR heterodimers as well as Rad1-Rad10 and Exo1 nucleases and the helicases Sgs1 and Srs2 have been implicated in suppressing homeologous recombination

(Nicholson et al. 2000; Myung et al. 2001; Spell and Jinks-Robertson 2004; Welz-

Voegele and Jinks-Robertson 2008; Heyer et al. 2010). In mice and human cells a BLM

(Sgs1 homolog) deficiency still suppresses homeologous recombination. However, in combination with an Msh2 deficiency the amount of homeologous recombination events increased (Larocque and Jasin 2010). RecQ helicase WRN helicase has also been implicated in suppressing homeologous recombination by exhibiting strong interaction with the MSH2-MSH6, MSH2-MSH3 and MLH1-PMS2 heterodimers (Saydam et al.

2007). To date a reaction that suppresses homeologous recombination in vitro has not been developed.

1.13 CONCLUDING REMARKS

HR is mechanistically conserved in prokaryotes and eukaryotes. However, because of the genome complexity in eukaryotes, additional mediators are required for the successful repair of DSBs. Defective HR leads to genomic instability and tumorigenesis.

50

Paradoxically, unregulated HR also leads to the same outcome. Therefore, eukaryotic cells have evolved elegant mechanisms to regulate each step of HR to ultimately produce an accurate repair outcome.

1.14 FOOTNOTES

This chapter was submitted for publication as: Homologous Recombination in

Eukaryotes. Amunugama R. and Fishel R.

51

Figure 1.1 DNA double strand break repair (DSBR) by homologous recombination (HR). DSBR is initiated by D-loop formation (strand invasion) by the 3 ssDNA overhang that results from strand resection. The invading DNA strand primes DNA synthesis. During double Holliday junction (dHJ) formation the second end is captured and the strands are ligated after DNA synthesis. Branch migration can either dissolve dHJs that result in non-crossover products or stabilize dHJs to undergo resolution. dHJ resolution can result in either crossover or non-crossover products.

52

Figure 1.2 Alternative DSB repair mechanisms. Non-homologous end joining (NHEJ) or Alternative NHEJ (Alt-NHEJ) occurs by either direct ligation of the broken DNA strand or ligation after minimal processing. Both NHEJ and Alt-NHEJ are error- prone repair mechanisms. DSB repair within direct repeat sequences could occur by single-strand annealing (SSA). SSA causes loss of a repeat sequence due to direct resection, annealing and ligation. During DSB repair through synthesis dependent strand annealing (SDSA) synthesized nascent strand is displaced by D-loop dissociation, anneals with the other 3-ssDNA overhang to complete DNA synthesis. SDSA results in non-crossover products. After the initial D-loop formation and DNA synthesis the D- loop can also be cleaved to produce crossover products.

53

Figure 1.3 Break induced replication (BIR) and telomere restoration. BIR initiates from a single-ended strand invasion. If one arm of the chromatid is lost after the DSB or if a telomere is uncapped, a 3-overhang is formed. DNA synthesis can continue to the end of the chromatid either by migration of the D-loop or after D-loop cleavage.

54

Figure 1.4 HR restores collapsed or stalled replication forks. Nicks of template strands lead to replication fork collapse that can be repaired by D-loop formation and Holliday Junction cleavage to restore replication. A lesion on leading strand may result in replication fork regression and DNA synthesis on the leading strand. Subsequent branch migration restores the replication fork.

55

Figure 1.5 BRCA2 and its proposed role in HR. (A) Schematic representation of the functional and structural domains of human BRCA2. (B) Upon formation of a DSB, the 5-strand is resected to leave a 3-ssDNA. RPA binds and prevents secondary structure formation. BRCA2 binds at the dsDNA-ssDNA junction and initiates RAD51 nucleation on RPA coated ssDNA while limiting nucleation on dsDNA. Continued RAD51 nucleoprotein filament growth results in a functional nucleoprotein filament that performs a homology search and strand exchange.

continued

56

Figure 1.5: continued

57

Table 1.1 Comparison of DSB Repair Factors in Budding Yeast and Human.

HR repair process Budding yeast Human

5 resection Mre11-Rad50-Xrs2 (MRX) MRE11-RAD50-NBS1(MRN) Exo1 EXO1 Dna2 DNA2 Sae2 CtIP Sgs1-Top3-RmiI BLM-TOP3-RMII-RMI2 RPA RPA

Pre-synapsis & Rad522 BRCA2 Synapsis Rad51 RAD51 Rad55-Rad57 RAD51B, RAD51C, RAD51D, XRCC2, XRCC3 Rad54 RAD54, RAD54B RAD51AP1 Dmc1 DMC1

DNA synthesis DNA polymerase δ DNA polymerase δ DNA polymerase η

Strand displacement Srs2 BLM Mph1 RTEL1 FANCM

HJ1 dissolution Sgs1-Top3-Rmi1 BLM-TOP3-RMI1-RMI2

HJ1 resolution Yen1 ResA (GEN1) Mus81-Mms4 MUS81-EME1 Slx1-Slx4 SLX1-SLX4

Chromatin remodeling Ino80 INO80 Swi/Snf SWI/SNF Swr1 SWR1 RSC TIP60

1Holliday Junction 2 Rad52 Epistasis Group in Red

58

1.15 REFERENCES

Aboussekhra A, Chanet R, Adjiri A, Fabre F. 1992. Semidominant suppressors of Srs2 helicase mutations of Saccharomyces cerevisiae map in the RAD51 gene, whose sequence predicts a protein with similarities to procaryotic RecA proteins. Mol. Cell. Biol. 12: 3224-3234. Agarwal S, van Cappellen WA, Guénolé A, Eppink B, Linsen SEV, Meijering E, Houtsmuller A, Kanaar R, Essers J. 2011. ATP-dependent and independent functions of Rad54 in genome maintenance. J. Cell Biol. 192: 735-750. Ahmad F, Stewart E. 2005. The N-terminal region of the Schizosaccharomyces pombe RecQ helicase, Rqh1p, physically interacts with Topoisomerase III and is required for Rqh1p function. Mol Genet Genomics 273: 102-114. Aihara H, Ito Y, Kurumizaka H, Terada T, Yokoyama S, Shibata T. 1997. An interaction between a specified surface of the C-terminal domain of RecA protein and double-stranded DNA for homologous pairing. J. Mol. Biol. 274: 213-221. Aihara H, Ito Y, Kurumizaka H, Yokoyama S, Shibata T. 1999. The N-terminal domain of the human Rad51 protein binds DNA: structure and a DNA binding surface as revealed by NMR. J. Mol. Biol. 290: 495-504. Alani E, Padmore R, Kleckner N. 1990. Analysis of wild-type and rad50 mutants of yeast suggests an intimate relationship between meiotic chromosome synapsis and recombination. Cell 61: 419-436. Alani E, Thresher R, Griffith JD, Kolodner RD. 1992. Characterization of DNA-binding and strand-exchange stimulation properties of y-RPA, a yeast single-strand-DNA- binding protein. J. Mol. Biol. 227: 54-71. Albala JS, Thelen MP, Prange C, Fan W, Christensen M, Thompson LH, Lennon GG. 1997. Identification of a Novel Human RAD51 Homolog,RAD51B. Genomics 46: 476-479. Alexeev A, Mazin A, Kowalczykowski SC. 2003. Rad54 protein possesses chromatin- remodeling activity stimulated by the Rad51-ssDNA nucleoprotein filament. Nat Struct Mol Biol 10: 182-186. Alexiadis V, Kadonaga JT. 2002. Strand pairing by Rad54 and Rad51 is enhanced by chromatin. Genes Dev. 16: 2767-2771. Andersen SL, Bergstralh DT, Kohl KP, LaRocque JR, Moore CB, Sekelsky J. 2009. Drosophila MUS312 and the vertebrate ortholog BTBD12 interact with DNA structure-specific endonucleases in DNA repair and recombination. Mol. Cell 35: 128-135. Andreassen PR, Ho GPH, D'Andrea AD. 2006. DNA damage responses and their many interactions with the replication fork. Carcinogenesis 27: 883-892. Antony E, Tomko EJ, Xiao Q, Krejci L, Lohman TM, Ellenberger T. 2009. Srs2 Disassembles Rad51 Filaments by a Protein-Protein Interaction Triggering ATP Turnover and Dissociation of Rad51 from DNA. Mol. Cell 35: 105-115.

59

Ariyoshi M, Nishino T, Iwasaki H, Shinagawa H, Morikawa K. 2000. Crystal structure of the holliday junction DNA in complex with a single RuvA tetramer. Proc. Natl. Acad. Sci. U. S. A. 97: 8257-8262. Azar B. 2011. QnAs with James E. Haber. Proc. Natl. Acad. Sci. U. S. A. 108: 5479. Bashkirov VI, King JS, Bashkirova EV, Schmuckli-Maurer J, Heyer W-D. 2000. DNA Repair Protein Rad55 Is a Terminal Substrate of the DNA Damage Checkpoints. Mol. Cell. Biol. 20: 4393-4404. Basile G, Aker M, Mortimer RK. 1992. Nucleotide sequence and transcriptional regulation of the yeast recombinational repair gene RAD51. Mol. Cell. Biol. 12: 3235-3246. Baumann P, West SC. 1998. Role of the human RAD51 protein in homologous recombination and double-stranded-break repair. Trends Biochem. Sci. 23: 247- 251. Bekker-Jensen S, Rendtlew Danielsen J, Fugger K, Gromova I, Nerstedt A, Lukas C, Bartek J, Lukas J, Mailand N. 2010. HERC2 coordinates ubiquitin-dependent assembly of DNA repair factors on damaged chromosomes. Nat Cell Biol 12: 80- 86; sup pp 81-12. Bennett RJ, Noirot-Gros MF, Wang JC. 2000. Interaction between yeast sgs1 helicase and DNA topoisomerase III. J. Biol. Chem. 275: 26898-26905. Benson FE, Stasiak A, West SC. 1994. Purification and characterization of the human Rad51 protein, an analogue of E. coli RecA. EMBO J. 13: 5764-5771. Bernstein KA, Gangloff S, Rothstein R. 2010. The RecQ DNA helicases in DNA repair. Annu. Rev. Genet. 44: 393-417. Bezzubova O, Silbergleit A, Yamaguchi-Iwai Y, Takeda S, Buerstedde J-M. 1997. Reduced X-Ray Resistance and Homologous Recombination Frequencies in a RAD54 -/- Mutant of the Chicken DT40 Cell Line. Cell 89: 185-193. Bianco PR, Tracy RB, Kowalczykowski SC. 1998. DNA strand exchange proteins: a biochemical and physical comparison. Front. Biosci. 3: D570-603. Bignell G, Micklem G, Stratton MR, Ashworth A, Wooster R. 1997. The BRC Repeats are Conserved in Mammalian BRCA2 Proteins. Hum. Mol. Genet. 6: 53-58. Bird AW, Yu DY, Pray-Grant MG, Qiu Q, Harmon KE, Megee PC, Grant PA, Smith MM, Christman MF. 2002. Acetylation of histone H4 by Esa1 is required for DNA double-strand break repair. Nature 419: 411-415. Bishop DK. 1994. RecA homologs Dmc1 and Rad51 interact to form multiple nuclear complexes prior to meiotic chromosome synapsis. Cell 79: 1081-1092. Bishop DK, Park D, Xu L, Kleckner N. 1992. DMC1: A meiosis-specific yeast homolog of E. coli recA required for recombination, synaptonemal complex formation, and cell cycle progression. Cell 69: 439-456. Bochar DA, Wang L, Beniya H, Kinev A, Xue Y, Lane WS, Wang W, Kashanchi F, Shiekhattar R. 2000. BRCA1 Is Associated with a Human SWI/SNF-Related Complex: Linking Chromatin Remodeling to Breast Cancer. Cell 102: 257-265. Boddy MN, Gaillard PH, McDonald WH, Shanahan P, Yates JR, 3rd, Russell P. 2001. Mus81-Eme1 are essential components of a Holliday junction resolvase. Cell 107: 537-548.

60

Bork P, Blomberg N, Nilges M. 1996. Internal repeats in the BRCA2 protein sequence. Nat. Genet. 13: 22-23. Botuyan MV, Lee J, Ward IM, Kim JE, Thompson JR, Chen J, Mer G. 2006. Structural basis for the methylation state-specific recognition of histone H4-K20 by 53BP1 and Crb2 in DNA repair. Cell 127: 1361-1373. Boundy-Mills KL, Livingston DM. 1993. A Saccharomyces cerevisiae RAD52 allele expressing a C-terminal truncation protein: activities and intragenic complementation of missense mutations. Genetics 133: 39-49. Branzei D, Foiani M. 2007. Interplay of replication checkpoints and repair proteins at stalled replication forks. DNA Repair (Amst) 6: 994-1003. -. 2008. Regulation of DNA repair throughout the cell cycle. Nat Rev Mol Cell Biol 9: 297-308. Braybrooke JP, Spink KG, Thacker J, Hickson ID. 2000. The RAD51 Family Member, RAD51L3, Is a DNA-stimulated ATPase That Forms a Complex with XRCC2. J. Biol. Chem. 275: 29100-29106. Breeden L, Nasmyth K. 1987. Cell cycle control of the yeast HO gene: Cis- and Trans- acting regulators. Cell 48: 389-397. Bugreev DV, Golub EI, Stasiak AZ, Stasiak A, Mazin AV. 2005. Activation of human meiosis-specific recombinase Dmc1 by Ca2+. J. Biol. Chem. 280: 26886-26895. Bugreev DV, Mazin AV. 2004. Ca2+ activates human homologous recombination protein Rad51 by modulating its ATPase activity. Proc. Natl. Acad. Sci. U. S. A. 101: 9988-9993. Bugreev DV, Mazina OM, Mazin AV. 2006. Rad54 protein promotes branch migration of Holliday junctions. Nature 442: 590-593. Bugreev DV, Yu X, Egelman EH, Mazin AV. 2007. Novel pro- and anti-recombination activities of the Bloom syndrome helicase. Genes Dev. 21: 3085-3094. Buisson R, Dion-Cote A-M, Coulombe Y, Launay H, Cai H, Stasiak AZ, Stasiak A, Xia B, Masson J-Y. 2010. Cooperation of breast cancer proteins PALB2 and piccolo BRCA2 in stimulating homologous recombination. Nat Struct Mol Biol 17: 1247- 1254. Bzymek M, Thayer NH, Oh SD, Kleckner N, Hunter N. 2010. Double Holliday junctions are intermediates of DNA break repair. Nature 464: 937-941. Cai Y, Jin J, Tomomori-Sato C, Sato S, Sorokina I, Parmely TJ, Conaway RC, Conaway JW. 2003. Identification of New Subunits of the Multiprotein Mammalian TRRAP/TIP60-containing Histone Acetyltransferase Complex. J. Biol. Chem. 278: 42733-42736. Cairns BR, Kim YJ, Sayre MH, Laurent BC, Kornberg RD. 1994. A multisubunit complex containing the SWI1/ADR6, SWI2/SNF2, SWI3, SNF5, and SNF6 gene products isolated from yeast. Proc. Natl. Acad. Sci. U. S. A. 91: 1950-1954. Cairns BR, Lorch Y, Li Y, Zhang M, Lacomis L, Erdjument-Bromage H, Tempst P, Du J, Laurent B, Kornberg RD. 1996. RSC, an Essential, Abundant Chromatin- Remodeling Complex. Cell 87: 1249-1260.

61

Carreira A, Hilario J, Amitani I, Baskin RJ, Shivji MK, Venkitaraman AR, Kowalczykowski SC. 2009. The BRC repeats of BRCA2 modulate the DNA- binding selectivity of RAD51. Cell 136: 1032-1043. Carreira A, Kowalczykowski SC. 2011. Two classes of BRC repeats in BRCA2 promote RAD51 nucleoprotein filament function by distinct mechanisms. Proc. Natl. Acad. Sci. U. S. A. 108: 10448-10453. Cartwright R, Dunn AM, Simpson PJ, Tambini CE, Thacker J. 1998. Isolation of novel human and mouse genes of the recA/RAD51 recombination-repair gene family. Nucleic Acids Res 26: 1653-1659. Cejka P, Cannavo E, Polaczek P, Masuda-Sasa T, Pokharel S, Campbell JL, Kowalczykowski SC. 2010. DNA end resection by Dna2-Sgs1-RPA and its stimulation by Top3-Rmi1 and Mre11-Rad50-Xrs2. Nature 467: 112-116. Celeste A, Petersen S, Romanienko PJ, Fernandez-Capetillo O, Chen HT, Sedelnikova OA, Reina-San-Martin B, Coppola V, Meffre E, Difilippantonio MJ et al. 2002. Genomic instability in mice lacking histone H2AX. Science 296: 922-927. Chai B, Huang J, Cairns BR, Laurent BC. 2005. Distinct roles for the RSC and Swi/Snf ATP-dependent chromatin remodelers in DNA double-strand break repair. Genes Dev. 19: 1656-1661. Chang M, Bellaoui M, Zhang C, Desai R, Morozov P, Delgado-Cruzata L, Rothstein R, Freyer GA, Boone C, Brown GW. 2005. RMI1/NCE4, a suppressor of genome instability, encodes a member of the RecQ helicase/Topo III complex. EMBO J. 24: 2024-2033. Chapman JR, Jackson SP. 2008. Phospho-dependent interactions between NBS1 and MDC1 mediate chromatin retention of the MRN complex at sites of DNA damage. EMBO Rep 9: 795-801. Chen C-F, Chen P-L, Zhong Q, Sharp ZD, Lee W-H. 1999. Expression of BRC Repeats in Breast Cancer Cells Disrupts the BRCA2-Rad51 Complex and Leads to Radiation Hypersensitivity and Loss of G2/M Checkpoint Control. J. Biol. Chem. 274: 32931-32935. Chen P-L, Chen C-F, Chen Y, Xiao J, Sharp ZD, Lee W-H. 1998. The BRC repeats in BRCA2 are critical for RAD51 binding and resistance to methyl methanesulfonate treatment. Proc. Natl. Acad. Sci. U. S. A. 95: 5287-5292. Chen XB, Melchionna R, Denis CM, Gaillard PH, Blasina A, Van de Weyer I, Boddy MN, Russell P, Vialard J, McGowan CH. 2001. Human Mus81-associated endonuclease cleaves Holliday junctions in vitro. Mol. Cell 8: 1117-1127. Chi P, Van Komen S, Sehorn MG, Sigurdsson S, Sung P. 2006. Roles of ATP binding and ATP hydrolysis in human Rad51 recombinase function. DNA Repair (Amst) 5: 381-391. Chung W-H, Zhu Z, Papusha A, Malkova A, Ira G. 2010. Defective Resection at DNA Double-Strand Breaks Leads to De Novo Telomere Formation and Enhances Gene Targeting. PLoS Genet 6: e1000948. Ciccia A, Constantinou A, West SC. 2003. Identification and characterization of the human mus 81-eme1 endonuclease. J. Biol. Chem. 278: 25172-25178.

62

Ciccia A, Elledge SJ. 2010. The DNA damage response: making it safe to play with knives. Mol. Cell 40: 179-204. Clerici M, Mantiero D, Lucchini G, Longhese MP. 2005. The Saccharomyces cerevisiae Sae2 protein promotes resection and bridging of double strand break ends. J. Biol. Chem. 280: 38631-38638. Clever B, Interthal H, Schmuckli-Maurer J, King J, Sigrist M, Heyer W-D. 1997. Recombinational repair in yeast: functional interactions between Rad51 and Rad54 proteins. EMBO J. 16: 2535-2544. Clever B, Schmuckli-Maurer J, Sigrist M, Glassner BJ, Heyer W-D. 1999. Specific negative effects resulting from elevated levels of the recombinational repair protein Rad54p in Saccharomyces cerevisiae. Yeast 15: 721-740. Cole GM, Schild D, Lovett ST, Mortimer RK. 1987. Regulation of RAD54- and RAD52- lacZ gene fusions in Saccharomyces cerevisiae in response to DNA damage. Mol. Cell. Biol. 7: 1078-1084. Cole GM, Schild D, Mortimer RK. 1989. Two DNA repair and recombination genes in Saccharomyces cerevisiae, RAD52 and RAD54, are induced during meiosis. Mol. Cell. Biol. 9: 3101-3104. Compton SA, Ӧzgür S, Griffith JD. 2010. Ring-shaped Rad51 Paralog Protein Complexes Bind Holliday Junctions and Replication Forks as Visualized by Electron Microscopy. J. Biol. Chem. 285: 13349-13356. Conaway RC, Conaway JW. 2009. The INO80 chromatin remodeling complex in transcription, replication and repair. Trends Biochem. Sci. 34: 71-77. Connolly B, Parsons CA, Benson FE, Dunderdale HJ, Sharples GJ, Lloyd RG, West SC. 1991. Resolution of Holliday junctions in vitro requires the Escherichia coli ruvC gene product. Proc. Natl. Acad. Sci. U. S. A. 88: 6063-6067. Connolly B, West SC. 1990. Genetic recombination in Escherichia coli: Holliday junctions made by RecA protein are resolved by fractionated cell-free extracts. Proc. Natl. Acad. Sci. U. S. A. 87: 8476-8480. Constantinou A, Davies AA, West SC. 2001. Branch migration and Holliday junction resolution catalyzed by activities from mammalian cells. Cell 104: 259-268. Cook PJ, Ju BG, Telese F, Wang X, Glass CK, Rosenfeld MG. 2009. Tyrosine dephosphorylation of H2AX modulates and survival decisions. Nature 458: 591-596. Cote J, Quinn J, Workman J, Peterson C. 1994. Stimulation of GAL4 derivative binding to nucleosomal DNA by the yeast SWI/SNF complex. Science 265: 53-60. Cromie GA, Smith GR. 2007. Branching out: meiotic recombination and its regulation. Trends Cell Biol. 17: 448-455. D'Andrea AD, Grompe M. 2003. The Fanconi anaemia/BRCA pathway. Nature Reviews. Cancer. 3: 23-34. Das C, Lucia MS, Hansen KC, Tyler JK. 2009. CBP/p300-mediated acetylation of histone H3 on lysine 56. Nature 459: 113-117. Datta A, Adjiri A, New L, Crouse GF, Jinks Robertson S. 1996. Mitotic crossovers between diverged sequences are regulated by mismatch repair proteins in Saccaromyces cerevisiae. Mol. Cell. Biol. 16: 1085-1093.

63

Datta A, Hendrix M, Lipsitch M, Jinks-Robertson S. 1997. Dual roles for DNA sequence identity and the mismatch repair system in the regulation of mitotic crossing-over in yeast. Proc. Natl. Acad. Sci. U. S. A. 94: 9757-9762. Davies AA, Masson J-Y, McIlwraith MJ, Stasiak AZ, Stasiak A, Venkitaraman AR, West SC. 2001. Role of BRCA2 in Control of the RAD51 Recombination and DNA Repair Protein. Mol. Cell 7: 273-282. Davies OR, Pellegrini L. 2007. Interaction with the BRCA2 C terminus protects RAD51- DNA filaments from disassembly by BRC repeats. Nat Struct Mol Biol 14: 475- 483. de los Santos T, Hunter N, Lee C, Larkin B, Loidl J, Hollingsworth NM. 2003. The Mus81/Mms4 endonuclease acts independently of double-Holliday junction resolution to promote a distinct subset of crossovers during meiosis in budding yeast. Genetics 164: 81-94. de los Santos T, Loidl J, Larkin B, Hollingsworth NM. 2001. A role for MMS4 in the processing of recombination intermediates during meiosis in Saccharomyces cerevisiae. Genetics 159: 1511-1525. de Massy B, Studier FW, Dorgai L, Appelbaum E, Weisberg RA. 1984. Enzymes and sites of genetic recombination: studies with gene-3 endonuclease of phage T7 and with site-affinity mutants of phage lambda. Cold Spring Harb. Symp. Quant. Biol. 49: 715-726. Dendouga N, Gao H, Moechars D, Janicot M, Vialard J, McGowan CH. 2005. Disruption of murine Mus81 increases genomic instability and DNA damage sensitivity but does not promote tumorigenesis. Mol. Cell. Biol. 25: 7569-7579. Dingwall AK, Beek SJ, McCallum CM, Tamkun JW, Kalpana GV, Goff SP, Scott MP. 1995. The Drosophila snr1 and brm proteins are related to yeast SWI/SNF proteins and are components of a large protein complex. Mol. Biol. Cell 6: 777- 791. Doil C, Mailand N, Bekker-Jensen S, Menard P, Larsen DH, Pepperkok R, Ellenberg J, Panier S, Durocher D, Bartek J et al. 2009. RNF168 binds and amplifies ubiquitin conjugates on damaged chromosomes to allow accumulation of repair proteins. Cell 136: 435-446. Dosanjh MK, Collins DW, Schild D, Fan W, Lennon GG, Albala JS, Shen Z. 1998. Isolation and characterization of RAD51C, a new human member of the RAD51 family of related genes. Nucleic Acids Res 26: 1179-1184. Downs JA, Allard S, Jobin-Robitaille O, Javaheri A, Auger A, Bouchard N, Kron SJ, Jackson SP, Cote J. 2004. Binding of chromatin-modifying activities to phosphorylated histone H2A at DNA damage sites. Mol. Cell 16: 979-990. Downs JA, Lowndes NF, Jackson SP. 2000. A role for Saccharomyces cerevisiae histone H2A in DNA repair. Nature 408: 1001-1004. Doyon Y, Selleck W, Lane WS, Tan S, Cote J. 2004. Structural and Functional Conservation of the NuA4 Histone Acetyltransferase Complex from Yeast to Humans. Mol. Cell. Biol. 24: 1884-1896.

64

Dray E, Etchin J, Wiese C, Saro D, Williams GJ, Hammel M, Yu X, Galkin VE, Liu D, Tsai M-S et al. 2010. Enhancement of RAD51 recombinase activity by the tumor suppressor PALB2. Nat Struct Mol Biol 17: 1255-1259. Dunderdale HJ, Benson FE, Parsons CA, Sharples GJ, Lloyd RG, West SC. 1991. Formation and resolution of recombination intermediates by E. coli RecA and RuvC proteins. Nature 354: 506-510. Dupaigne P, Le Breton C, Fabre F, Gangloff S, Le Cam E, Veaute X. 2008. The Srs2 Helicase Activity Is Stimulated by Rad51 Filaments on dsDNA: Implications for Crossover Incidence during Mitotic Recombination. Mol. Cell 29: 243-254. Ebbert R, Birkmann A, Schuller HJ. 1999. The product of the SNF2/SWI2 paralogue INO80 of Saccharomyces cerevisiae required for efficient expression of various yeast structural genes is part of a high-molecular-weight protein complex. Mol. Microbiol. 32: 741-751. Eberharter A, Becker PB. 2004. ATP-dependent nucleosome remodelling: factors and functions. J. Cell Sci. 117: 3707-3711. Eggler AL, Inman RB, Cox MM. 2002. The Rad51-dependent pairing of long DNA substrates is stabilized by replication protein A. J. Biol. Chem. 277: 39280-39288. Elborough KM, West SC. 1990. Resolution of synthetic Holliday junctions in DNA by an endonuclease activity from calf thymus. EMBO J. 9: 2931-2936. Ellis NA, Groden J, Ye TZ, Straughen J, Lennon DJ, Ciocci S, Proytcheva M, German J. 1995. The Bloom's syndrome gene product is homologous to RecQ helicases. Cell 83: 655-666. Erdile LF, Wold MS, Kelly TJ. 1990. The primary structure of the 32-kDa subunit of human replication protein A. J. Biol. Chem. 265: 3177-3182. Esashi F, Christ N, Gannon J, Liu Y, Hunt T, Jasin M, West SC. 2005. CDK-dependent phosphorylation of BRCA2 as a regulatory mechanism for recombinational repair. Nature 434: 598-604. Esashi F, Galkin VE, Yu X, Egelman EH, West SC. 2007. Stabilization of RAD51 nucleoprotein filaments by the C-terminal region of BRCA2. Nat Struct Mol Biol 14: 468-474. Essers J, Houtsmuller AB, van Veelen L, Paulusma C, Nigg AL, Pastink A, Vermeulen W, Hoeijmakers JHJ, Kanaar R. 2002. Nuclear dynamics of RAD52 group homologous recombination proteins in response to DNA damage. EMBO J. 21: 2030-2037. Essers J, van Steeg H, de Wit J, Swagemakers SMA, Vermeij M, Hoeijmakers JHJ, Kanaar R. 2000. Homologous and non-homologous recombination differentially affect DNA damage repair in mice. EMBO J. 19: 1703-1710. Farmer H, McCabe N, Lord CJ, Tutt AN, Johnson DA, Richardson TB, Santarosa M, Dillon KJ, Hickson I, Knights C et al. 2005. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 434: 917-921. Fekairi S, Scaglione S, Chahwan C, Taylor ER, Tissier A, Coulon S, Dong MQ, Ruse C, Yates JR, 3rd, Russell P et al. 2009. Human SLX4 is a Holliday junction resolvase subunit that binds multiple DNA repair/recombination endonucleases. Cell 138: 78-89.

65

Feng Z, Scott SP, Bussen W, Sharma GG, Guo G, Pandita TK, Powell SN. 2011. Rad52 inactivation is synthetically lethal with BRCA2 deficiency. Proc. Natl. Acad. Sci. U. S. A. 108: 686-691. Ferguson DO, Holloman WK. 1996. Recombinational repair of gaps in DNA is asymmetric in Ustilago maydis and can be explained by a migrating D-loop model. Proc. Natl. Acad. Sci. U. S. A. 93: 5419-5424. Fishman-Lobell J, Rudin N, Haber JE. 1992. Two alternative pathways of double-strand break repair that are kinetically separable and independently modulated. Mol. Cell. Biol. 12: 1292-1303. Flaus A, Owen-Hughes T. 2003. Dynamic properties of nucleosomes during thermal and ATP-driven mobilization. Mol. Cell. Biol. 23: 7767-7779. Fricke WM, Brill SJ. 2003. Slx1-Slx4 is a second structure-specific endonuclease functionally redundant with Sgs1-Top3. Genes Dev. 17: 1768-1778. Fricke WM, Kaliraman V, Brill SJ. 2001. Mapping the DNA topoisomerase III binding domain of the Sgs1 DNA helicase. J. Biol. Chem. 276: 8848-8855. Fujimori A, Tachiiri S, Sonoda E, Thompson LH, Dhar PK, Hiraoka M, Takeda S, Zhang Y, Reth M, Takata M. 2001. Rad52 partially substitutes for the Rad51 paralog XRCC3 in maintaining chromosomal integrity in vertebrate cells. EMBO J. 20: 5513-5520. Galanty Y, Belotserkovskaya R, Coates J, Polo S, Miller KM, Jackson SP. 2009. Mammalian SUMO E3-ligases PIAS1 and PIAS4 promote responses to DNA double-strand breaks. Nature 462: 935-939. Galkin VE, Esashi F, Yu X, Yang S, West SC, Egelman EH. 2005. BRCA2 BRC motifs bind RAD51-DNA filaments. Proc. Natl. Acad. Sci. U. S. A. 102: 8537-8542. Game JC, Mortimer RK. 1974. A genetic study of X-ray sensitive mutants in yeast. Mutation Research/Fundamental and Molecular Mechanisms of Mutagenesis 24: 281-292. Gangloff S, McDonald JP, Bendixen C, Arthur L, Rothstein R. 1994. The yeast type I topoisomerase Top3 interacts with Sgs1, a DNA helicase homolog: a potential eukaryotic reverse gyrase. Mol. Cell. Biol. 14: 8391-8398. Garvik B, Carson M, Hartwell L. 1995. Single-stranded DNA arising at telomeres in cdc13 mutants may constitute a specific signal for the RAD9 checkpoint [published erratum appears in Mol Cell Biol 1996 Jan;16(1):457]. Mol. Cell. Biol. 15: 6128-6138. Gasior SL, Wong AK, Kora Y, Shinohara A, Bishop DK. 1998. Rad52 associates with RPA and functions with Rad55 and Rad57 to assemble meiotic recombination complexes. Genes Dev. 12: 2208-2221. Gavin I, Horn PJ, Peterson CL. 2001. SWI/SNF chromatin remodeling requires changes in DNA topology. Mol. Cell 7: 97-104. Gildemeister OS, Sage JM, Knight KL. 2009. Cellular Redistribution of Rad51 in Response to DNA Damage. J. Biol. Chem. 284: 31945-31952. Goggins M, Schutte M, Lu J, Moskaluk CA, Weinstein CL, Petersen GM, Yeo CJ, Jackson CE, Lynch HT, Hruban RH et al. 1996. Germline BRCA2 Gene

66

Mutations in Patients with Apparently Sporadic Pancreatic Carcinomas. Cancer Res. 56: 5360-5364. Golub EI, Kovalenko OV, Gupta RC, Ward DC, Radding CM. 1997. Interaction of human recombination proteins Rad51 and Rad54. Nucleic Acids Res 25: 4106- 4110. Gorbalenya AE, Koonin EV. 1993. Helicases: amino acid sequence comparisons and structure-function relationships. Curr. Opin. Struct. Biol. 3: 419-429. Gudmundsdottir K, Lord CJ, Witt E, Tutt ANJ, Ashworth A. 2004. DSS1 is required for RAD51 focus formation and genomic stability in mammalian cells. EMBO Rep 5: 989-993. Gupta RC, Folta-Stogniew E, O'Malley S, Takahashi M, Radding CM. 1999. Rapid Exchange of A:T Base Pairs Is Essential for Recognition of DNA Homology by Human Rad51 Recombination Protein. Mol. Cell 4: 705-714. Hackett JA, Greider CW. 2003. End Resection Initiates Genomic Instability in the Absence of Telomerase. Mol. Cell. Biol. 23: 8450-8461. Harfe BD, Jinks-Robertson S. 2000. DNA mismatch repair and genetic instability. Annu. Rev. Genet. 34: 359-399. Hargreaves D, Rice DW, Sedelnikova SE, Artymiuk PJ, Lloyd RG, Rafferty JB. 1998. Crystal structure of E.coli RuvA with bound DNA Holliday junction at 6 A resolution. Nat. Struct. Biol. 5: 441-446. Harper JW, Elledge SJ. 2007. The DNA damage response: ten years after. Mol. Cell 28: 739-745. Hartlerode AJ, Scully R. 2009. Mechanisms of double-strand break repair in somatic mammalian cells. Biochem. J. 423: 157-168. Haseltine CA, Kowalczykowski SC. 2009. An archaeal Rad54 protein remodels DNA and stimulates DNA strand exchange by RadA. Nucleic Acids Res 37: 2757-2770. Hashimoto Y, Chaudhuri AR, Lopes M, Costanzo V. 2010. Rad51 protects nascent DNA from Mre11-dependent degradation and promotes continuous DNA synthesis. Nat Struct Mol Biol 17: 1305-1311. Hassold T, Hall H, Hunt P. 2007. The origin of human aneuploidy: where we have been, where we are going. Hum. Mol. Genet. 16 Spec No. 2: R203-208. Hassold T, Sherman S, Hunt PA. 1995. The origin of trisomy in humans. Prog. Clin. Biol. Res. 393: 1-12. Haushalter KA, Kadonaga JT. 2003. Chromatin assembly by DNA-translocating motors. Nat Rev Mol Cell Biol 4: 613-620. Hays SL, Firmenich AA, Berg P. 1995. Complex formation in yeast double-strand break repair: participation of Rad51, Rad52, Rad55, and Rad57 proteins. Proc. Natl. Acad. Sci. U. S. A. 92: 6925-6929. Hays SL, Firmenich AA, Massey P, Banerjee R, Berg P. 1998. Studies of the Interaction between Rad52 Protein and the Yeast Single-Stranded DNA Binding Protein RPA. Mol. Cell. Biol. 18: 4400-4406. Heller RC, Marians KJ. 2006. Replication fork reactivation downstream of a blocked nascent leading strand. Nature 439: 557-562.

67

Henikoff S. 2008. Nucleosome destabilization in the epigenetic regulation of . Nat Rev Genet 9: 15-26. Herzberg K, Bashkirov VI, Rolfsmeier M, Haghnazari E, McDonald WH, Anderson S, Bashkirova EV, Yates JR, III, Heyer W-D. 2006. Phosphorylation of Rad55 on Serines 2, 8, and 14 Is Required for Efficient Homologous Recombination in the Recovery of Stalled Replication Forks. Mol. Cell. Biol. 26: 8396-8409. Heyer W-D, Li X, Rolfsmeier M, Zhang X-P. 2006. Rad54: the Swiss Army knife of homologous recombination? Nucleic Acids Res 34: 4115-4125. Heyer WD, Ehmsen KT, Liu J. 2010. Regulation of homologous recombination in eukaryotes. Annu. Rev. Genet. 44: 113-139. Heyer WD, Rao MR, Erdile LF, Kelly TJ, Kolodner RD. 1990. An essential Saccharomyces cerevisiae single-stranded DNA binding protein is homologous to the large subunit of human RP-A. EMBO J. 9: 2321-2329. Heyting C. 1996. Synaptonemal Complexes: structure and function. Curr. Biol. 8: 389- 396. Hicks WM, Yamaguchi M, Haber JE. 2011. Real-time analysis of double-strand DNA break repair by homologous recombination. Proc. Natl. Acad. Sci. U. S. A. 108: 3108-3115. Higgins NP, Kato K, Strauss B. 1976. A model for replication repair in mammalian cells. J. Mol. Biol. 101: 417-425. Hilario J, Amitani I, Baskin RJ, Kowalczykowski SC. 2009. Direct imaging of human Rad51 nucleoprotein dynamics on individual DNA molecules. Proc. Natl. Acad. Sci. U. S. A. 106: 361-368. Hoeijmakers JH. 2001. Genome maintenance mechanisms for preventing cancer. Nature 411: 366-374. Holliday RA. 1964. A mechanism for gene conversion in fungi. Genet. Res. 5: 282-304. Hollingsworth NM, Brill SJ. 2004. The Mus81 solution to resolution: generating meiotic crossovers without Holliday junctions. Genes Dev. 18: 117-125. Holloman WK. 2011. Unraveling the mechanism of BRCA2 in homologous recombination. Nat Struct Mol Biol 18: 748-754. Horn PJ, Peterson CL. 2002. Chromatin Higher Order Folding--Wrapping up Transcription. Science 297: 1824-1827. Housworth EA, Stahl FW. 2003. Crossover interference in humans. Am. J. Hum. Genet. 73: 188-197. Epub 2003 May 2022. Howlett NG, Taniguchi T, Olson S, Cox B, Waisfisz Q, de Die-Smulders C, Persky N, Grompe M, Joenje H, Pals G et al. 2002. Biallelic Inactivation of BRCA2 in Fanconi Anemia. Science 297: 606-609. Hu Y, Raynard S, Sehorn MG, Lu X, Bussen W, Zheng L, Stark JM, Barnes EL, Chi P, Janscak P et al. 2007. RECQL5/Recql5 helicase regulates homologous recombination and suppresses tumor formation via disruption of Rad51 presynaptic filaments. Genes Dev. 21: 3073-3084. Huen MS, Chen J. 2010. Assembly of checkpoint and repair machineries at DNA damage sites. Trends Biochem. Sci. 35: 101-108.

68

Huen MS, Grant R, Manke I, Minn K, Yu X, Yaffe MB, Chen J. 2007. RNF8 transduces the DNA-damage signal via histone ubiquitylation and checkpoint protein assembly. Cell 131: 901-914. Huen MS, Sy SM, Chen J. 2010. BRCA1 and its toolbox for the maintenance of genome integrity. Nat Rev Mol Cell Biol 11: 138-148. Huertas P, Cortes-Ledesma F, Sartori AA, Aguilera A, Jackson SP. 2008. CDK targets Sae2 to control DNA-end resection and homologous recombination. Nature 455: 689-692. Huertas P, Jackson SP. 2009. Human CtIP mediates cell cycle control of DNA end resection and double strand break repair. J. Biol. Chem. 284: 9558-9565. Huyen Y, Zgheib O, Ditullio RA, Jr., Gorgoulis VG, Zacharatos P, Petty TJ, Sheston EA, Mellert HS, Stavridi ES, Halazonetis TD. 2004. Methylated lysine 79 of histone H3 targets 53BP1 to DNA double-strand breaks. Nature 432: 406-411. Hyde H, Davies AA, Benson FE, West SC. 1994. Resolution of recombination intermediates by a mammalian activity functionally analogous to Escherichia coli RuvC resolvase. J. Biol. Chem. 269: 5202-5209. Ikura T, Ogryzko VV, Grigoriev M, Groisman R, Wang J, Horikoshi M, Scully R, Qin J, Nakatani Y. 2000. Involvement of the TIP60 histone acetylase complex in DNA repair and apoptosis. Cell 102: 463-473. Ikura T, Tashiro S, Kakino A, Shima H, Jacob N, Amunugama R, Yoder K, Izumi S, Kuraoka I, Tanaka K et al. 2007. DNA damage-dependent acetylation and ubiquitination of H2AX enhances chromatin dynamics. Mol. Cell. Biol. 27: 7028- 7040. Ip SC, Rass U, Blanco MG, Flynn HR, Skehel JM, West SC. 2008. Identification of Holliday junction resolvases from humans and yeast. Nature 456: 357-361. Iwasaki H, Takahagi M, Shiba T, Nakata A, Shinagawa H. 1991. Escherichia coli RuvC protein is an endonuclease that resolves the Holliday structure. EMBO J. 10: 4381-4389. Jackson D, Dhar K, Wahl JK, Wold MS, Borgstahl GEO. 2002. Analysis of the Human Replication Protein A:Rad52 Complex: Evidence for Crosstalk Between RPA32, RPA70, Rad52 and DNA. J. Mol. Biol. 321: 133-148. Janke R, Herzberg K, Rolfsmeier M, Mar J, Bashkirov VI, Haghnazari E, Cantin G, Yates JR, Heyer W-D. 2010. A truncated DNA-damage-signaling response is activated after DSB formation in the G1 phase of Saccharomyces cerevisiae. Nucleic Acids Res 38: 2302-2313. Jasin M. 2002. Homologous repair of DNA damage and tumorigenesis: the BRCA connection. 21: 8981-8993. Jaskelioff M, Van Komen S, Krebs JE, Sung P, Peterson CL. 2003. Rad54p Is a Chromatin Remodeling Enzyme Required for Heteroduplex DNA Joint Formation with Chromatin. J. Biol. Chem. 278: 9212-9218. Jensen RB, Carreira A, Kowalczykowski SC. 2010. Purified human BRCA2 stimulates RAD51-mediated recombination. Nature 467: 678-683. Jenuwein T, Allis CD. 2001. Translating the histone code. Science 293: 1074-1080.

69

Jiang H, Xie Y, Houston P, Stemke-Hale K, Mortensen UH, Rothstein R, Kodadek T. 1996. Direct association between the yeast Rad51 and Rad54 recombination proteins. J. Biol. Chem. 271: 33181-33186. Johnson FB, Lombard DB, Neff NF, Mastrangelo MA, Dewolf W, Ellis NA, Marciniak RA, Yin Y, Jaenisch R, Guarente L. 2000. Association of the with topoisomerase IIIalpha in somatic and meiotic cells. Cancer Res. 60: 1162-1167. Johnson R, Symington L. 1995. Functional differences and interactions among the putative RecA homologs Rad51, Rad55, and Rad57. Mol. Cell. Biol. 15: 4843- 4850. Johnston LH, Johnson AL. 1995. The DNA repair genes RAD54 and UNG1 are cell cycle regulated in budding yeast but MCB promoter elements have no essential role in the DNA damage response. Nucleic Acids Res 23: 2147-2152. Jones NJ, Zhao Y, Siciliano MJ, Thompson LH. 1995. Assignment of the XRCC2 human DNA repair gene to chromosome 7q36 by complementation analysis. Genomics 26: 619-622. Jonsson ZO, Jha S, Wohlschlegel JA, Dutta A. 2004. Rvb1p/Rvb2p recruit Arp5p and assemble a functional Ino80 chromatin remodeling complex. Mol. Cell 16: 465- 477. Kagawa W, Kurumizaka H, Ishitani R, Fukai S, Nureki O, Shibata T, Yokoyama S. 2002. Crystal Structure of the Homologous-Pairing Domain from the Human Rad52 Recombinase in the Undecameric Form. Mol. Cell 10: 359-371. Kanaar R, Hoeijmakers JHJ. 1998. Genetic recombination: From competition to collaboration. Nature 391: 335-338. Kanaar R, Troelstra C, Swagemakers SMA, Essers J, Smit B, Franssen J-H, Pastink A, Bezzubova OY, Buerstedde J-M, Clever B et al. 1996. Human and mouse homologs of the Saccharomyces cerevisiae RAD54 DNA repair gene: evidence for functional conservation. Curr. Biol. 6: 828-838. Kans JA, Mortimer RK. 1991. Nucleotide sequence of the RAD57 gene of Saccharomyces cerevisiae. Gene 105: 139-140. Kantake N, Sugiyama T, Kolodner RD, Kowalczykowski SC. 2003. The recombination- deficient mutant RPA (rfa1-t11) is displaced slowly from single-stranded DNA by Rad51 protein. J. Biol. Chem. 278: 23410-23417. Kassabov SR, Zhang B, Persinger J, Bartholomew B. 2003. SWI/SNF unwraps, slides, and rewraps the nucleosome. Mol. Cell 11: 391-403. Keeney S. 2001. Mechanism and control of meiotic recombination initiation. Curr. Top. Dev. Biol. 52: 1-53. Kent NA, Chambers AL, Downs JA. 2007. Dual chromatin remodeling roles for RSC during DNA double strand break induction and repair at the yeast MAT locus. J. Biol. Chem. 282: 27693-27701. Khasanov FK, Savchenko GV, Bashkirova EV, Korolev VG, Heyer W-D, Bashkirov VI. 1999. A New Recombinational DNA Repair Gene From Schizosaccharomyces pombe With Homology to Escherichia coli RecA. Genetics 152: 1557-1572.

70

Khavari PA, Peterson CL, Tamkun JW, Mendel DB, Crabtree GR. 1993. BRG1 contains a conserved domain of the SWI2/SNF2 family necessary for normal mitotic growth and transcription. Nature 366: 170-174. Klapholz S, Waddell CS, Esposito RE. 1985. The role of the SPO11 gene in meiotic recombination in yeast. Genetics 110: 187-216. Kleckner N. 1996. Meiosis: how could it work?. [Review] [103 refs]. Proc. Natl. Acad. Sci. U. S. A. 93: 8167-8174. Kleff S, Kemper B, Sternglanz R. 1992. Identification and characterization of yeast mutants and the gene for a cruciform cutting endonuclease. EMBO J. 11: 699- 704. Klein HL. 1997. RDH54, a RAD54 homologue in Saccharomyces cerevisiae, is required for mitotic diploid-specific recombination and repair and for meiosis. Genetics 147: 1533-1543. -. 2008. The consequences of Rad51 overexpression for normal and tumor cells. DNA Repair (Amst) 7: 686-693. Klein HL, Symington LS. 2009. Breaking up just got easier to do. Cell 138: 20-22. Kneitz B, Cohen PE, Avdievich E, Zhu L, Kane MF, Hou H, Jr., Kolodner RD, Kucherlapati R, Pollard JW, Edelmann W. 2000. MutS homolog 4 localization to meiotic chromosomes is required for chromosome pairing during meiosis in male and female mice. Genes Dev. 14: 1085-1097. Kojic M, Mao N, Zhou Q, Lisby M, Holloman WK. 2008. Compensatory role for Rad52 during recombinational repair in Ustilago maydis. Mol. Microbiol. 67: 1156-1168. Kojic M, Yang H, Kostrub CF, Pavletich NP, Holloman WK. 2003. The BRCA2- Interacting Protein DSS1 Is Vital for DNA Repair, Recombination, and Genome Stability in Ustilago maydis. Mol. Cell 12: 1043-1049. Kolas NK, Chapman JR, Nakada S, Ylanko J, Chahwan R, Sweeney FD, Panier S, Mendez M, Wildenhain J, Thomson TM et al. 2007. Orchestration of the DNA- damage response by the RNF8 ubiquitin ligase. Science 318: 1637-1640. Kolodner R. 1996. Biochemistry and genetics of eukaryotic mismatch repair. Genes Dev. 10: 1433-1442. Kolodner RD, Marsischky GT. 1999. Eukaryotic DNA mismatch repair. Curr. Opin. Genet. Dev. 9: 89-96. Kooistra R, Vreeken K, Zonneveld JBM, De Jong A, Eeken JCJ, Osgood CJ, Buerstedde JM, Lohman PHM, Pastink A. 1997. The Drosophila melanogaster RAD54 homolog, DmRAD54, is involved in the repair of radiation damage and recombination. Mol. Cell. Biol. 17: 6097-6104. Kornberg RD. 1977. Structure of chromatin. Annu. Rev. Biochem. 46: 931-954. Kouzarides T. 2007. Chromatin modifications and their function. Cell 128: 693-705. Kovalenko OV, Golub EI, Bray-Ward P, Ward DC, Radding CM. 1997. A Novel Nucleic Acid-Binding Protein That Interacts with Human Rad51 Recombinase. Nucleic Acids Res 25: 4946-4953. Kowalczykowski SC. 1991. Biochemistry of genetic recombination: energetics and mechanism of DNA strand exchange. Annu. Rev. Biophys. Biophys. Chem. 20: 539-575.

71

Koyama H, Itoh M, Miyahara K, Tsuchiya E. 2002. Abundance of the RSC nucleosome- remodeling complex is important for the cells to tolerate DNA damage in Saccharomyces cerevisiae. FEBS Lett. 531: 215-221. Krejci L, Song B, Bussen W, Rothstein R, Mortensen UH, Sung P. 2002. Interaction with Rad51 Is Indispensable for Recombination Mediator Function of Rad52. J. Biol. Chem. 277: 40132-40141. Krejci L, Van Komen S, Li Y, Villemain J, Reddy MS, Klein H, Ellenberger T, Sung P. 2003. DNA helicase Srs2 disrupts the Rad51 presynaptic filament. Nature 423: 305-309. Krishnan N, Jeong DG, Jung SK, Ryu SE, Xiao A, Allis CD, Kim SJ, Tonks NK. 2009. Dephosphorylation of the C-terminal tyrosyl residue of the DNA damage-related histone H2A.X is mediated by the protein phosphatase eyes absent. J. Biol. Chem. 284: 16066-16070. Krogan NJ, Keogh MC, Datta N, Sawa C, Ryan OW, Ding H, Haw RA, Pootoolal J, Tong A, Canadien V et al. 2003. A Snf2 family ATPase complex required for recruitment of the histone H2A variant Htz1. Mol. Cell 12: 1565-1576. Krogh BO, Symington LS. 2004. Recombination proteins in yeast. Annu. Rev. Genet. 38: 233-271. Kruhlak MJ, Celeste A, Dellaire G, Fernandez-Capetillo O, Müller WG, McNally JG, Bazett-Jones DP, Nussenzweig A. 2006. Changes in chromatin structure and mobility in living cells at sites of DNA double-strand breaks. J. Cell Biol. 172: 823-834. Kusch T, Florens L, Macdonald WH, Swanson SK, Glaser RL, Yates JR, 3rd, Abmayr SM, Washburn MP, Workman JL. 2004. Acetylation by Tip60 is required for selective histone variant exchange at DNA lesions. Science 306: 2084-2087. Kwon Y, Chi P, Roh DH, Klein H, Sung P. 2007. Synergistic action of the Saccharomyces cerevisiae homologous recombination factors Rad54 and Rad51 in chromatin remodeling. DNA Repair (Amst) 6: 1496-1506. Larocque JR, Jasin M. 2010. Mechanisms of recombination between diverged sequences in wild-type and BLM-deficient mouse and human cells. Mol. Cell. Biol. 30: 1887-1897. Lee D, Kim JW, Seo T, Hwang SG, Choi E-J, Choe J. 2002. SWI/SNF Complex Interacts with Tumor Suppressor p53 and Is Necessary for the Activation of p53-mediated Transcription. J. Biol. Chem. 277: 22330-22337. Lee SE, Moore JK, Holmes A, Umezu K, Kolodner RD, Haber JE. 1998. Saccharomyces Ku70, mre11/rad50 and RPA proteins regulate adaptation to G2/M arrest after DNA damage. Cell 94: 399-409. Lengsfeld BM, Rattray AJ, Bhaskara V, Ghirlando R, Paull TT. 2007. Sae2 is an endonuclease that processes hairpin DNA cooperatively with the Mre11/Rad50/Xrs2 complex. Mol. Cell 28: 638-651. Li X, Zhang X-P, Solinger JA, Kiianitsa K, Yu X, Egelman EH, Heyer W-D. 2007. Rad51 and Rad54 ATPase activities are both required to modulate Rad51-dsDNA filament dynamics. Nucleic Acids Res 35: 4124-4140.

72

Lin FL, Sperle K, Sternberg N. 1984. Model for homologous recombination during transfer of DNA into mouse L cells: role for DNA ends in the recombination process. Mol. Cell. Biol. 4: 1020-1034. Lin Z, Kong H, Nei M, Ma H. 2006. Origins and evolution of the recA/RAD51 gene family: evidence for ancient gene duplication and endosymbiotic gene transfer. Proc. Natl. Acad. Sci. U. S. A. 103: 10328-10333. Lio YC, Mazin AV, Kowalczykowski SC, Chen DJ. 2003. Complex formation by the human Rad51B and Rad51C DNA repair proteins and their activities in vitro. J. Biol. Chem. 278: 2469-2478. Lisby M, Barlow JH, Burgess RC, Rothstein R. 2004. Choreography of the DNA damage response: spatiotemporal relationships among checkpoint and repair proteins. Cell 118: 699-713. Lisby M, Rothstein R. 2009. Choreography of recombination proteins during the DNA damage response. DNA Repair (Amst) 8: 1068-1076. Liu J, Doty T, Gibson B, Heyer W-D. 2010. Human BRCA2 protein promotes RAD51 filament formation on RPA-covered single-stranded DNA. Nat Struct Mol Biol 17: 1260-1262. Liu J, Heyer W-D. 2011. Who's who in human recombination: BRCA2 and RAD52. Proc. Natl. Acad. Sci. U. S. A. 108: 441-442. Liu J, Wu TC, Lichten M. 1995. The location and structure of double-strand DNA breaks induced during yeast meiosis: evidence for a covalently linked DNA-protein intermediate. EMBO J. 14: 4599-4608. Liu N, Lamerdin JE, Tebbs RS, Schild D, Tucker JD, Shen MR, Brookman KW, Siciliano MJ, Walter CA, Fan W et al. 1998. XRCC2 and XRCC3, new human Rad51-family members, promote chromosome stability and protect against DNA cross-links and other damages. Mol. Cell 1: 783-793. Liu Y, Masson JY, Shah R, O'Regan P, West SC. 2004a. RAD51C is required for Holliday junction processing in mammalian cells. Science 303: 243-246. Liu Y, Stasiak AZ, Masson JY, McIlwraith MJ, Stasiak A, West SC. 2004b. Conformational changes modulate the activity of human RAD51 protein. J. Mol. Biol. 337: 817-827. Liu Y, West SC. 2004. Happy Hollidays: 40th anniversary of the Holliday junction. Nat Rev Mol Cell Biol 5: 937-944. Lo T, Pellegrini L, Venkitaraman AR, Blundell TL. 2003. Sequence fingerprints in BRCA2 and RAD51: implications for DNA repair and cancer. DNA Repair (Amst) 2: 1015-1028. Lobachev K, Vitriol E, Stemple J, Resnick MA, Bloom K. 2004. Chromosome fragmentation after induction of a double-strand break is an active process prevented by the RMX repair complex. Curr. Biol. 14: 2107-2112. Lobrich M, Jeggo PA. 2007. The impact of a negligent G2/M checkpoint on genomic instability and cancer induction. Nat Rev Cancer 7: 861-869. Lorch Y, Maier-Davis B, Kornberg RD. 2006. Chromatin remodeling by nucleosome disassembly in vitro. Proc. Natl. Acad. Sci. U. S. A. 103: 3090-3093.

73

Lou Z, Minter-Dykhouse K, Franco S, Gostissa M, Rivera MA, Celeste A, Manis JP, van Deursen J, Nussenzweig A, Paull TT et al. 2006. MDC1 maintains genomic stability by participating in the amplification of ATM-dependent DNA damage signals. Mol. Cell 21: 187-200. Lovett ST. 1994. Sequence of the RAD55 gene of Saccharomyces cerevisiae: similarity of RAD55 to prokaryotic RecA and other RecA-like proteins. Gene 142: 103-106. Luger K, Mader AW, Richmond RK, Sargent DF, Richmond TJ. 1997. Crystal structure of the nucleosome core particle at 2.8 A resolution. Nature 389: 251-260. Lundblad V, Blackburn EH. 1993. An alternative pathway for yeast telomere maintenance rescues est1‚àí senescence. Cell 73: 347-360. Mahaney BL, Meek K, Lees-miller SP. 2009. Repair of ionizing radiation-induced DNA double-strand breaks by non-homologous end-joining. Biochem. J. 417: 639-650. Mailand N, Bekker-Jensen S, Faustrup H, Melander F, Bartek J, Lukas C, Lukas J. 2007. RNF8 ubiquitylates histones at DNA double-strand breaks and promotes assembly of repair proteins. Cell 131: 887-900. Malkova A, Naylor ML, Yamaguchi M, Ira G, Haber JE. 2005. RAD51-Dependent Break-Induced Replication Differs in Kinetics and Checkpoint Responses from RAD51-Mediated Gene Conversion. Mol. Cell. Biol. 25: 933-944. Manohar M, Mooney AM, North JA, Nakkula RJ, Picking JW, Edon A, Fishel R, Poirier MG, Ottesen JJ. 2009. Acetylation of Histone H3 at the Nucleosome Dyad Alters DNA-Histone Binding. J. Biol. Chem. 284: 23312-23321. Marston AL, Amon A. 2004. Meiosis: cell-cycle controls shuffle and deal. Nat Rev Mol Cell Biol 5: 983-997. Marston NJ, Richards WJ, Hughes D, Bertwistle D, Marshall CJ, Ashworth A. 1999. Interaction between the Product of the Breast Cancer Susceptibility Gene BRCA2 and DSS1, a Protein Functionally Conserved from Yeast to Mammals. Mol. Cell. Biol. 19: 4633-4642. Martens JA, Winston F. 2003. Recent advances in understanding chromatin remodeling by Swi/Snf complexes. Curr. Opin. Genet. Dev. 13: 136-142. Martin JS, Winkelmann N, Petalcorin MIR, McIlwraith MJ, Boulton SJ. 2005. RAD-51- Dependent and -Independent Roles of a Caenorhabditis elegans BRCA2-Related Protein during DNA Double-Strand Break Repair. Mol. Cell. Biol. 25: 3127-3139. Martin SG, Laroche T, Suka N, Grunstein M, Gasser SM. 1999. Relocalization of telomeric Ku and SIR proteins in response to DNA strand breaks in yeast. Cell 97: 621-633. Maryon E, Carroll D. 1991. Involvement of single-stranded tails in homologous recombination of DNA injected into Xenopus laevis oocyte nuclei. Mol. Cell. Biol. 11: 3268-3277. Masson JY, Davies AA, Hajibagheri N, Van Dyck E, Benson FE, Stasiak AZ, Stasiak A, West SC. 1999. The meiosis-specific recombinase hDmc1 forms ring structures and interacts with hRad51. EMBO J. 18: 6552-6560. Masson JY, Stasiak AZ, Stasiak A, Benson FE, West SC. 2001a. Complex formation by the human RAD51C and XRCC3 recombination repair proteins. Proc. Natl. Acad. Sci. U. S. A. 98: 8440-8446.

74

Masson JY, Tarsounas MC, Stasiak AZ, Stasiak A, Shah R, McIlwraith MJ, Benson FE, West SC. 2001b. Identification and purification of two distinct complexes containing the five RAD51 paralogs. Genes Dev. 15: 3296-3307. Masson JY, West SC. 2001. The Rad51 and Dmc1 recombinases: a non-identical twin relationship. Trends Biochem. Sci. 26: 131-136. Matsuoka S, Ballif BA, Smogorzewska A, McDonald ER, 3rd, Hurov KE, Luo J, Bakalarski CE, Zhao Z, Solimini N, Lerenthal Y et al. 2007. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 316: 1160-1166. Matulova P, Marini V, Burgess RC, Sisakova A, Kwon Y, Rothstein R, Sung P, Krejci L. 2009. Cooperativity of Mus81·Mms4 with Rad54 in the Resolution of Recombination and Replication Intermediates. J. Biol. Chem. 284: 7733-7745. Mazin AV, Alexeev AA, Kowalczykowski SC. 2003. A Novel Function of Rad54 Protein. J. Biol. Chem. 278: 14029-14036. Mazin AV, Mazina OM, Bugreev DV, Rossi MJ. 2010. Rad54, the motor of homologous recombination. DNA Repair (Amst) 9: 286-302. Mazina OM, Mazin AV. 2004. Human Rad54 Protein Stimulates DNA Strand Exchange Activity of hRad51 Protein in the Presence of Ca2+. J. Biol. Chem. 279: 52042- 52051. -. 2008. Human Rad54 protein stimulates human Mus81–Eme1 endonuclease. Proc. Natl. Acad. Sci. U. S. A. 105: 18249-18254. Mazina OM, Rossi MJ, Thomaä NH, Mazin AV. 2007. Interactions of Human Rad54 Protein with Branched DNA Molecules. J. Biol. Chem. 282: 21068-21080. Mazloum N, Holloman WK. 2009. Second-end capture in DNA double-strand break repair promoted by Brh2 protein of Ustilago maydis. Mol. Cell 33: 160-170. McEachern MJ, Haber JE. 2006. Break-Induced Replication and Recombinational Telomere Elongation in Yeast. Annu. Rev. Biochem. 75: 111-135. McIlwraith MJ, West SC. 2008. DNA Repair Synthesis Facilitates RAD52-Mediated Second-End Capture during DSB Repair. Mol. Cell 29: 510-516. McKee AH, Kleckner N. 1997. Mutations in Saccharomyces cerevisiae that block meiotic prophase chromosome metabolism and confer cell cycle arrest at pachytene identify two new meiosis-specific genes SAE1 and SAE3. Genetics 146: 817-834. McKee RA, Lawrence CW. 1980. Genetic analysis of [gamma]-mutagenesis in yeast III. Double-mutant strains. Mutation Research/Fundamental and Molecular Mechanisms of Mutagenesis 70: 37-48. McPherson JP, Lemmers B, Chahwan R, Pamidi A, Migon E, Matysiak-Zablocki E, Moynahan ME, Essers J, Hanada K, Poonepalli A et al. 2004. Involvement of mammalian Mus81 in genome integrity and tumor suppression. Science 304: 1822-1826. Melander F, Bekker-Jensen S, Falck J, Bartek J, Mailand N, Lukas J. 2008. Phosphorylation of SDT repeats in the MDC1 N terminus triggers retention of NBS1 at the DNA damage-modified chromatin. J. Cell Biol. 181: 213-226.

75

Miki Y, Swensen J, Shattuck Eidens D, Futreal PA, Harshman K, Tavtigian S, Liu Q, Cochran C, Bennett LM, Ding W et al. 1994. A strong candidate for the breast and ovarian cancer susceptibility gene BRCA1. Science 266: 66-71. Miller KA, Yoshikawa DM, McConnell IR, Clark R, Schild D, Albala JS. 2002. RAD51C interacts with RAD51B and is central to a larger protein complex in vivo exclusive of RAD51. J. Biol. Chem. 277: 8406-8411. Milne GT, Weaver DT. 1993. Dominant negative alleles of RAD52 reveal a DNA repair/recombination complex including Rad51 and Rad52. Genes Dev. 7: 1755- 1765. Mimitou EP, Symington LS. 2008. Sae2, Exo1 and Sgs1 collaborate in DNA double- strand break processing. Nature 455: 770-774. -. 2009a. DNA end resection: many nucleases make light work. DNA Repair (Amst) 8: 983-995. -. 2009b. Nucleases and helicases take center stage in homologous recombination. Trends Biochem. Sci. 34: 264-272. Mizuguchi G, Shen X, Landry J, Wu W-H, Sen S, Wu C. 2004. ATP-Driven Exchange of Histone H2AZ Variant Catalyzed by SWR1 Chromatin Remodeling Complex. Science 303: 343-348. Mizuta R, LaSalle JM, Cheng H-L, Shinohara A, Ogawa H, Copeland N, Jenkins NA, Lalande M, Alt FW. 1997. RAB22 and RAB163/mouse BRCA2: Proteins that specifically interact with the RAD51‚Äâprotein. Proc. Natl. Acad. Sci. U. S. A. 94: 6927-6932. Mizuuchi K, Kemper B, Hays J, Weisberg RA. 1982. T4 endonuclease VII cleaves holliday structures. Cell 29: 357-365. Modesti M, Budzowska M, Baldeyron Cl, Demmers JAA, Ghirlando R, Kanaar R. 2007. RAD51AP1 Is a Structure-Specific DNA Binding Protein that Stimulates Joint Molecule Formation during RAD51-Mediated Homologous Recombination. Mol. Cell 28: 468-481. Moens PB, Kolas NK, Tarsounas M, Marcon E, Cohen PE, Spyropoulos B. 2002. The time course and chromosomal localization of recombination-related proteins at meiosis in the mouse are compatible with models that can resolve the early DNA- DNA interactions without reciprocal recombination. J. Cell Sci. 115: 1611-1622. Morgan EA, Shah N, Symington LS. 2002. The Requirement for ATP Hydrolysis by Saccharomyces cerevisiae Rad51 Is Bypassed by Mating-Type Heterozygosity or RAD54 in High Copy. Mol. Cell. Biol. 22: 6336-6343. Morris JR, Boutell C, Keppler M, Densham R, Weekes D, Alamshah A, Butler L, Galanty Y, Pangon L, Kiuchi T et al. 2009. The SUMO modification pathway is involved in the BRCA1 response to genotoxic stress. Nature 462: 886-890. Morrison AJ, Highland J, Krogan NJ, Arbel-Eden A, Greenblatt JF, Haber JE, Shen X. 2004. INO80 and gamma-H2AX interaction links ATP-dependent chromatin remodeling to DNA damage repair. Cell 119: 767-775. Morrison C, Shinohara A, Sonoda E, Yamaguchi-Iwai Y, Takata M, Weichselbaum RR, Takeda S. 1999. The essential functions of human Rad51 are independent of ATP hydrolysis. Mol. Cell. Biol. 19: 6891-6897.

76

Mortensen UH, Bendixen C, Sunjevaric I, Rothstein R. 1996. DNA strand annealing is promoted by the yeast Rad52 protein. Proc. Natl. Acad. Sci. U. S. A. 93: 10729- 10734. Moyal L, Lerenthal Y, Gana-Weisz M, Mass G, So S, Wang SY, Eppink B, Chung YM, Shalev G, Shema E et al. 2011. Requirement of ATM-dependent monoubiquitylation of histone H2B for timely repair of DNA double-strand breaks. Mol. Cell 41: 529-542. Moynahan ME, Jasin M. 2010. Mitotic homologous recombination maintains genomic stability and suppresses tumorigenesis. Nat Rev Mol Cell Biol 11: 196-207. Moynahan ME, Pierce AJ, Jasin M. 2001. BRCA2 Is Required for Homology-Directed Repair of Chromosomal Breaks. Mol. Cell 7: 263-272. Mullen JR, Nallaseth FS, Lan YQ, Slagle CE, Brill SJ. 2005. Yeast Rmi1/Nce4 controls genome stability as a subunit of the Sgs1-Top3 complex. Mol. Cell. Biol. 25: 4476-4487. Munoz IM, Hain K, Declais AC, Gardiner M, Toh GW, Sanchez-Pulido L, Heuckmann JM, Toth R, Macartney T, Eppink B et al. 2009. Coordination of structure- specific nucleases by human SLX4/BTBD12 is required for DNA repair. Mol. Cell 35: 116-127. Muris DFR, Vreeken K, Carr AM, Murray JM, Smit C, Lohman PHM, Pastink A. 1996. Isolation of the Schizosaccharomyces pombe RAD54 homologue, rhp54+, a gene involved in the repair of radiation damage and replication fidelity. J. Cell Sci. 109: 73-81. Murr R, Loizou JI, Yang YG, Cuenin C, Li H, Wang ZQ, Herceg Z. 2006. Histone acetylation by Trrap-Tip60 modulates loading of repair proteins and repair of DNA double-strand breaks. Nat Cell Biol 8: 91-99. Myung K, Datta A, Chen C, Kolodner RD. 2001. SGS1, the Saccharomyces cerevisiae homologue of BLM and WRN, suppresses genome instability and homeologous recombination. Nat. Genet. 27: 113-116. Nakada S, Tai I, Panier S, Al-Hakim A, Iemura S, Juang YC, O'Donnell L, Kumakubo A, Munro M, Sicheri F et al. 2010. Non-canonical inhibition of DNA damage- dependent ubiquitination by OTUB1. Nature 466: 941-946. Nakamura K, Kato A, Kobayashi J, Yanagihara H, Sakamoto S, Oliveira DV, Shimada M, Tauchi H, Suzuki H, Tashiro S et al. 2011. Regulation of homologous recombination by RNF20-dependent H2B ubiquitination. Mol. Cell 41: 515-528. Nassif N, Penney J, Pal S, Engels WR, Gloor GB. 1994. Efficient copying of nonhomologous sequences from ectopic sites via P-element-induced gap repair. Mol. Cell. Biol. 14: 1613-1625. Neigeborn L, Carlson M. 1984. GENES AFFECTING THE REGULATION OF SUC2 GENE EXPRESSION BY GLUCOSE REPRESSION IN SACCHAROMYCES CEREVISIAE. Genetics 108: 845-858. New JH, Sugiyama T, Zaitseva E, Kowalczykowski SC. 1998. Rad52 protein stimulates DNA strand exchange by Rad51 and replication protein A. Nature 391: 407-410.

77

Nicholson A, Hendrix M, Jinks-Robertson S, Crouse GF. 2000. Regulation of mitotic homeologous recombination in yeast. Functions of mismatch repair and nucleotide excision repair genes. Genetics 154: 133-146. Nimonkar AV, Genschel J, Kinoshita E, Polaczek P, Campbell JL, Wyman C, Modrich P, Kowalczykowski SC. 2011. BLM-DNA2-RPA-MRN and EXO1-BLM-RPA- MRN constitute two DNA end resection machineries for human DNA break repair. Genes Dev. 25: 350-362. Nimonkar AV, Ozsoy AZ, Genschel J, Modrich P, Kowalczykowski SC. 2008. Human exonuclease 1 and BLM helicase interact to resect DNA and initiate DNA repair. Proc. Natl. Acad. Sci. U. S. A. 105: 16906-16911. Nimonkar AV, Sica RA, Kowalczykowski SC. 2009. Rad52 promotes second-end DNA capture in double-stranded break repair to form complement-stabilized joint molecules. Proc. Natl. Acad. Sci. U. S. A. 106: 3077-3082. Niu H, Chung W-H, Zhu Z, Kwon Y, Zhao W, Chi P, Prakash R, Seong C, Liu D, Lu L et al. 2010. Mechanism of the ATP-dependent DNA end-resection machinery from Saccharomyces cerevisiae. Nature 467: 108-111. Ogawa T, Yu X, Shinohara A, Egelman E. 1993. Similarity of the yeast RAD51 filament to the bacterial RecA filament. Science 259: 1896-1899. Oram M, Keeley A, Tsaneva I. 1998. Holliday junction resolvase in Schizosaccharomyces pombe has identical endonuclease activity to the CCE1 homologue YDC2. Nucleic Acids Res 26: 594-601. Orr-Weaver TL, Szostak JW. 1983a. Multiple, tandem plasmid integration in Saccharomyces cerevisiae. Mol. Cell. Biol. 3: 747-749. -. 1983b. Yeast recombination: the association between double-strand gap repair and crossing-over. Proc. Natl. Acad. Sci. U. S. A. 80: 4417-4421. Osley MA, Tsukuda T, Nickoloff JA. 2007. ATP-dependent chromatin remodeling factors and DNA damage repair. Mutat. Res. 618: 65-80. Otsuki T, Furukawa Y, Ikeda K, Endo H, Yamashita T, Shinohara A, Iwamatsu A, Ozawa K, Liu JM. 2001. Fanconi anemia protein, FANCA, associates with BRG1, a component of the human SWI/SNF complex. Hum. Mol. Genet. 10: 2651-2660. Papamichos-Chronakis M, Krebs JE, Peterson CL. 2006. Interplay between Ino80 and Swr1 chromatin remodeling enzymes regulates cell cycle checkpoint adaptation in response to DNA damage. Genes Dev. 20: 2437-2449. Parsons CA, Baumann P, Van Dyck E, West SC. 2000. Precise binding of single-stranded DNA termini by human RAD52 protein. EMBO J. 19: 4175-4181. Parsons CA, Stasiak A, Bennett RJ, West SC. 1995a. Structure of a multisubunit complex that promotes DNA branch migration. Nature 374: 375-378. Parsons CA, Stasiak A, West SC. 1995b. The E.coli RuvAB proteins branch migrate Holliday junctions through heterologous DNA sequences in a reaction facilitated by SSB. EMBO J. 14: 5736-5744. Passy SI, Yu X, Li Z, Radding CM, Masson J-Y, West SC, Egelman EH. 1999. Human Dmc1 protein binds DNA as an octameric ring. Proc. Natl. Acad. Sci. U. S. A. 96: 10684-10688.

78

Pellegrini L, Yu DS, Lo T, Anand S, Lee M, Blundell TL, Venkitaraman AR. 2002. Insights into DNA recombination from the structure of a RAD51-BRCA2 complex. Nature 420: 287-293. Petalcorin MIR, Sandall J, Wigley DB, Boulton SJ. 2006. CeBRC-2 Stimulates D-loop Formation by RAD-51 and Promotes DNA Single-strand Annealing. J. Mol. Biol. 361: 231-242. Petrini JHJ, Bressan DA, Yao MS. 1997. TheRAD52epistasis group in mammalian double strand break repair. Semin. Immunol. 9: 181-188. Petukhova G, Stratton S, Sung P. 1998. Catalysis of homologous DNA pairing by yeast Rad51 and Rad54 proteins. Nature 393: 91-94. Petukhova G, Van Komen S, Vergano S, Klein H, Sung P. 1999. Yeast Rad54 Promotes Rad51-dependent Homologous DNA Pairing via ATP Hydrolysis-driven Change in DNA Double Helix Conformation. J. Biol. Chem. 274: 29453-29462. Pittman DL, Weinberg LR, Schimenti JC. 1998. Identification, Characterization, and Genetic Mapping ofRad51d,a New Mouse and HumanRAD51/RecA-Related Gene. Genomics 49: 103-111. Prakash S, Johnson RE, Prakash L. 2005. EUKARYOTIC TRANSLESION SYNTHESIS DNA POLYMERASES: Specificity of Structure and Function. Annu. Rev. Biochem. 74: 317-353. Qian X, He Y, Ma X, Fodje MN, Grochulski P, Luo Y. 2006. Calcium stiffens archaeal Rad51 recombinase from Methanococcus voltae for homologous recombination. J. Biol. Chem. 281: 39380-39387. Qian X, Wu Y, He Y, Luo Y. 2005. Crystal structure of Methanococcus voltae RadA in complex with ADP: hydrolysis-induced conformational change. Biochemistry (Mosc). 44: 13753-13761. Ranatunga W, Jackson D, Lloyd JA, Forget AL, Knight KL, Borgstahl GEO. 2001. Human RAD52 Exhibits Two Modes of Self-association. J. Biol. Chem. 276: 15876-15880. Reid S, Schindler D, Hanenberg H, Barker K, Hanks S, Kalb R, Neveling K, Kelly P, Seal S, Freund M et al. 2007. Biallelic mutations in PALB2 cause Fanconi anemia subtype FA-N and predispose to childhood cancer. Nat. Genet. 39: 162-164. Reisman D, Glaros S, Thompson EA. 2009. The SWI/SNF complex and cancer. Oncogene 28: 1653-1668. Reymer A, Frykholm K, Morimatsu K, Takahashi M, Norden B. 2009. Structure of human Rad51 protein filament from molecular modeling and site-specific linear dichroism spectroscopy. Proc. Natl. Acad. Sci. U. S. A. 106: 13248-13253. Rice JC, Allis CD. 2001. Code of silence. Nature 414: 258-261. Rice MC, Smith ST, Bullrich F, Havre P, Kmiec EB. 1997. Isolation of human and mouse genes based on homology to REC2, a recombinational repair gene from the fungus Ustilago maydis. Proc. Natl. Acad. Sci. U. S. A. 94: 7417-7422. Richardson C, Stark JM, Ommundsen M, Jasin M. 2004. Rad51 overexpression promotes alternative double-strand break repair pathways and genome instability. Oncogene 23: 546-553.

79

Rijkers T, Van Den Ouweland J, Morolli B, Rolink AG, Baarends WM, Van Sloun PPH, Lohman PHM, Pastink A. 1998. Targeted Inactivation of Mouse RAD52 Reduces Homologous Recombination but Not Resistance to Ionizing Radiation. Mol. Cell. Biol. 18: 6423-6429. Rogakou EP, Boon C, Redon C, Bonner WM. 1999. Megabase chromatin domains involved in DNA double-strand breaks in vivo. J. Cell Biol. 146: 905-916. Rogakou EP, Nieves-Neira W, Boon C, Pommier Y, Bonner WM. 2000. Initiation of DNA fragmentation during apoptosis induces phosphorylation of H2AX histone at serine 139. J. Biol. Chem. 275: 9390-9395. Rogakou EP, Pilch DR, Orr AH, Ivanova VS, Bonner WM. 1998. DNA double-stranded breaks induce histone H2AX phosphorylation on serine 139. J. Biol. Chem. 273: 5858-5868. Saha A, Wittmeyer J, Cairns BR. 2002. Chromatin remodeling by RSC involves ATP- dependent DNA translocation. Genes Dev. 16: 2120-2134. Saha A, Wittmeyer J, Cairns BR. 2006. Chromatin remodelling: the industrial revolution of DNA around histones. Nat Rev Mol Cell Biol 7: 437-447. San Filippo J, Sung P, Klein H. 2008. Mechanism of eukaryotic homologous recombination. Annu. Rev. Biochem. 77: 229-257. Sauvageau S, Stasiak AZ, Banville I, Ploquin M, Stasiak A, Masson J-Y. 2005. Fission Yeast Rad51 and Dmc1, Two Efficient DNA Recombinases Forming Helical Nucleoprotein Filaments. Mol. Cell. Biol. 25: 4377-4387. Savitsky K, Bar Shira A, Gilad S, Rotman G, Ziv Y, Vanagaite L, Tagle DA, Smith S, Uziel T, Sfez S et al. 1995. A single ataxia telangiectasia gene with a product similar to PI-3 kinase [see comments]. Science 268: 1749-1753. Saydam N, Kanagaraj R, Dietschy T, Garcia PL, Pena-Diaz J, Shevelev I, Stagljar I, Janscak P. 2007. Physical and functional interactions between Werner syndrome helicase and mismatch-repair initiation factors. Nucleic Acids Res 35: 5706-5716. Schaetzlein S, Kodandaramireddy NR, Ju Z, Lechel A, Stepczynska A, Lilli DR, Clark AB, Rudolph C, Kuhnel F, Wei K et al. 2007. Exonuclease-1 Deletion Impairs DNA Damage Signaling and Prolongs Lifespan of Telomere-Dysfunctional Mice. Cell 130: 863-877. Schild D, Lio Y-c, Collins DW, Tsomondo T, Chen DJ. 2000. Evidence for Simultaneous Protein Interactions between Human Rad51 Paralogs. J. Biol. Chem. 275: 16443- 16449. Schild D, Wiese C. 2010. Overexpression of RAD51 suppresses recombination defects: a possible mechanism to reverse genomic instability. Nucleic Acids Res 38: 1061- 1070. Schlacher K, Christ N, Siaud N, Egashira A, Wu H, Jasin M. 2011. Double-Strand Break Repair-Independent Role for BRCA2 in Blocking Stalled Replication Fork Degradation by MRE11. Cell 145: 529-542. Sehorn MG, Sigurdsson S, Bussen W, Unger VM, Sung P. 2004. Human meiotic recombinase Dmc1 promotes ATP-dependent homologous DNA strand exchange. Nature 429: 433-437.

80

Seitz EM, Haseltine CA, Kowalczykowski SC. 2001. DNA recombination and repair in the Archaea. in Adv. Appl. Microbiol. (ed. B Paul), pp. 101-169. Academic Press. Shah PP, Zheng X, Epshtein A, Carey JN, Bishop DK, Klein HL. 2010. Swi2/Snf2- Related Translocases Prevent Accumulation of Toxic Rad51 Complexes during Mitotic Growth. Mol. Cell 39: 862-872. Shaked H, Avivi-Ragolsky N, Levy AA. 2006. Involvement of the Arabidopsis SWI2/SNF2 Chromatin Remodeling Gene Family in DNA Damage Response and Recombination. Genetics 173: 985-994. Sharan SK, Kuznetsov SG. 2007. Resolving RAD51C function in late stages of homologous recombination. Cell Div 2: 15. Sharan SK, Morimatsu M, Albrecht U, Lim D-S, Regel E, Dinh C, Sands A, Eichele G, Hasty P, Bradley A. 1997. Embryonic lethality and radiation hypersensitivity mediated by Rad51 in mice lacking Brca2. Nature 386: 804-810. Sharples GJ, Lloyd RG. 1991. Resolution of Holliday junctions in Escherichia coli: identification of the ruvC gene product as a 19-kilodalton protein. J. Bacteriol. 173: 7711-7715. Shen X, Mizuguchi G, Hamiche A, Wu C. 2000. A chromatin remodelling complex involved in transcription and DNA processing. Nature 406: 541-544. Shen X, Ranallo R, Choi E, Wu C. 2003. Involvement of actin-related proteins in ATP- dependent chromatin remodeling. Mol. Cell 12: 147-155. Shen Z, Cloud KG, Chen DJ, Park MS. 1996. Specific Interactions between the Human RAD51 and RAD52 Proteins. J. Biol. Chem. 271: 148-152. Shi I, Hallwyl SC, Seong C, Mortensen U, Rothstein R, Sung P. 2009. Role of the Rad52 amino-terminal DNA binding activity in DNA strand capture in homologous recombination. J. Biol. Chem. 284: 33275-33284. Shim EY, Chung W-H, Nicolette ML, Zhang Y, Davis M, Zhu Z, Paull TT, Ira G, Lee SE. 2010. Saccharomyces cerevisiae Mre11/Rad50/Xrs2 and Ku proteins regulate association of Exo1 and Dna2 with DNA breaks. EMBO J. 29: 3370-3380. Shim EY, Hong SJ, Oum JH, Yanez Y, Zhang Y, Lee SE. 2007. RSC mobilizes nucleosomes to improve accessibility of repair machinery to the damaged chromatin. Mol. Cell. Biol. 27: 1602-1613. Shim EY, Ma JL, Oum JH, Yanez Y, Lee SE. 2005. The yeast chromatin remodeler RSC complex facilitates end joining repair of DNA double-strand breaks. Mol. Cell. Biol. 25: 3934-3944. Shim KS, Schmutte C, Tombline G, Heinen CD, Fishel R. 2004. hXRCC2 enhances ADP/ATP processing and strand exchange by hRAD51. J. Biol. Chem. 279: 30385-30394. Shim KS, Schmutte C, Yoder K, Fishel R. 2006. Defining the salt effect on human RAD51 activities. DNA Repair (Amst) 5: 718-730. Shin DS, Pellegrini L, Daniels DS, Yelent B, Craig L, Bates D, Yu DS, Shivji MK, Hitomi C, Arvai AS et al. 2003. Full-length archaeal Rad51 structure and mutants: mechanisms for RAD51 assembly and control by BRCA2. EMBO J. 22: 4566-4576.

81

Shinohara A, Ogawa H, Ogawa T. 1992. Rad51 protein involved in repair and recombination in S. cerevisiae is a RecA-like protein. Cell 69: 457-470. Shinohara A, Ogawa T. 1998. Stimulation by Rad52 of yeast Rad51- mediated recombination. Nature 391: 404-407. Shinohara A, Shinohara M, Ohta T, Matsuda S, Ogawa T. 1998. Rad52 forms ring structures and co-operates with RPA in single-strand DNA annealing. Genes Cells 3: 145-156. Shivji MKK, Mukund SR, Rajendra E, Chen S, Short JM, Savill J, Klenerman D, Venkitaraman AR. 2009. The BRC repeats of human BRCA2 differentially regulate RAD51 binding on single- versus double-stranded DNA to stimulate strand exchange. Proc. Natl. Acad. Sci. U. S. A. 106: 13254-13259. Siaud N, Dray E, Gy I, Gerard E, Takvorian N, Doutriaux M-P. 2004. Brca2 is involved in meiosis in Arabidopsis thaliana as suggested by its interaction with Dmc1. EMBO J. 23: 1392-1401. Sigurdsson S, Van Komen S, Bussen W, Schild D, Albala JS, Sung P. 2001. Mediator function of the human Rad51B-Rad51C complex in Rad51/RPA-catalyzed DNA strand exchange. Genes Dev. 15: 3308-3318. Sigurdsson S, Van Komen S, Petukhova G, Sung P. 2002. Homologous DNA Pairing by Human Recombination Factors Rad51 and Rad54. J. Biol. Chem. 277: 42790- 42794. Simon M, North JA, Shimko JC, Forties RA, Ferdinand MB, Manohar M, Zhang M, Fishel R, Ottesen JJ, Poirier MG. 2011. Histone fold modifications control nucleosome unwrapping and disassembly. Proc. Natl. Acad. Sci. U. S. A. 108: 12711-12716. Singleton MR, Dillingham MS, Wigley DB. 2007. Structure and Mechanism of Helicases and Nucleic Acid Translocases. Annu. Rev. Biochem. 76: 23-50. Singleton MR, Wentzell LM, Liu Y, West SC, Wigley DB. 2002. Structure of the single- strand annealing domain of human RAD52 protein. Proc. Natl. Acad. Sci. U. S. A. 99: 13492-13497. Sinha M, Peterson CL. 2008. A Rad51 presynaptic filament is sufficient to capture nucleosomal homology during recombinational repair of a DNA double-strand break. Mol. Cell 30: 803-810. Sinha M, Watanabe S, Johnson A, Moazed D, Peterson CL. 2009. Recombinational repair within heterochromatin requires ATP-dependent chromatin remodeling. Cell 138: 1109-1121. Snow R. 1967. Mutants of yeast sensitive to ultraviolet light. J. Bacteriol. 94: 571-575. Sobhian B, Shao G, Lilli DR, Culhane AC, Moreau LA, Xia B, Livingston DM, Greenberg RA. 2007. RAP80 Targets BRCA1 to Specific Ubiquitin Structures at DNA Damage Sites. Science 316: 1198-1202. Solinger JA, Heyer W-D. 2001. Rad54 protein stimulates the postsynaptic phase of Rad51 protein-mediated DNA strand exchange. Proc. Natl. Acad. Sci. U. S. A. 98: 8447-8453. Solinger JA, Kiianitsa K, Heyer WD. 2002. Rad54, a Swi2/Snf2-like recombinational repair protein, disassembles Rad51:dsDNA filaments. Mol. Cell 10: 1175-1188.

82

Sonoda E, Sasaki MS, Buerstedde JM, Bezzubova O, Shinohara A, Ogawa H, Takata M, Yamaguchi-Iwai Y, Takeda S. 1998. Rad51-deficient vertebrate cells accumulate chromosomal breaks prior to cell death. EMBO J. 17: 598-608. Spell RM, Jinks-Robertson S. 2004. Examination of the Roles of Sgs1 and Srs2 Helicases in the Enforcement of Recombination Fidelity in Saccharomyces cerevisiae. Genetics 168: 1855-1865. Spycher C, Miller ES, Townsend K, Pavic L, Morrice NA, Janscak P, Stewart GS, Stucki M. 2008. Constitutive phosphorylation of MDC1 physically links the MRE11- RAD50-NBS1 complex to damaged chromatin. J. Cell Biol. 181: 227-240. Stahl FW. 1979. Genetic Recombination. W.H. Freeman and Co., San Francisco. Stark JM, Hu P, Pierce AJ, Moynahan ME, Ellis N, Jasin M. 2002. ATP Hydrolysis by Mammalian RAD51 Has a Key Role during Homology-directed DNA Repair. J. Biol. Chem. 277: 20185-20194. Stasiak AZ, Larquet E, Stasiak A, Müller S, Engel A, Van Dyck E, West SC, Egelman EH. 2000. The human Rad52 protein exists as a heptameric ring. Curr. Biol. 10: 337-340. Stern M, Jensen R, Herskowitz I. 1984. Five SWI genes are required for expression of the HO gene in yeast. J. Mol. Biol. 178: 853-868. Stewart GS, Panier S, Townsend K, Al-Hakim AK, Kolas NK, Miller ES, Nakada S, Ylanko J, Olivarius S, Mendez M et al. 2009. The RIDDLE syndrome protein mediates a ubiquitin-dependent signaling cascade at sites of DNA damage. Cell 136: 420-434. Strahl BD, Allis CD. 2000. The language of covalent histone modifications. Nature 403: 41-45. Strathern JN, Klar AJS, Hicks JB, Abraham JA, Ivy JM, Nasmyth KA, McGill C. 1982. Homothallic switching of yeast mating type cassettes is initiated by a double- stranded cut in the MAT locus. Cell 31: 183-192. Strom L, Karlsson C, Lindroos HB, Wedahl S, Katou Y, Shirahige K, Sjogren C. 2007. Postreplicative formation of cohesion is required for repair and induced by a single DNA break. Science 317: 242-245. Strom L, Lindroos HB, Shirahige K, Sjogren C. 2004. Postreplicative recruitment of cohesin to double-strand breaks is required for DNA repair. Mol. Cell 16: 1003- 1015. Stucki M. 2009. Histone H2A.X Tyr142 phosphorylation: a novel sWItCH for apoptosis? DNA Repair (Amst) 8: 873-876. Stucki M, Clapperton JA, Mohammad D, Yaffe MB, Smerdon SJ, Jackson SP. 2005. MDC1 directly binds phosphorylated histone H2AX to regulate cellular responses to DNA double-strand breaks. Cell 123: 1213-1226. Sugawara N, Wang X, Haber JE. 2003. In vivo roles of Rad52, Rad54, and Rad55 proteins in Rad51-mediated recombination. Mol. Cell 12: 209-219. Sugiyama T, Kantake N. 2009. Dynamic regulatory interactions of rad51, rad52, and replication protein-a in recombination intermediates. J. Mol. Biol. 390: 45-55.

83

Sugiyama T, Zaitseva EM, Kowalczykowski SC. 1997. A single-stranded DNA-binding protein is needed for efficient presynaptic complex formation by the Saccharomyces cerevisiae Rad51 protein. J. Biol. Chem. 272: 7940-7945. Sung P. 1994. Catalysis of ATP-dependent homologous DNA pairing and strand exchange by yeast RAD51 protein. Science 265: 1241-1243. Sung P. 1997a. Function of Yeast Rad52 Protein as a Mediator between Replication Protein A and the Rad51 Recombinase. J. Biol. Chem. 272: 28194-28197. Sung P. 1997b. Yeast Rad55 and Rad57 proteins form a heterodimer that functions with replication protein A to promote DNA strand exchange by Rad51 recombinase. Genes Dev. 11: 1111-1121. Sung P, Robberson DL. 1995. DNA strand exchange mediated by a RAD51-ssDNA nucleoprotein filament with polarity opposite to that of RecA. Cell 82: 453-461. Sung P, Stratton SA. 1996. Yeast Rad51 Recombinase Mediates Polar DNA Strand Exchange in the Absence of ATP Hydrolysis. J. Biol. Chem. 271: 27983-27986. Svendsen JM, Harper JW. 2010. GEN1/Yen1 and the SLX4 complex: Solutions to the problem of Holliday junction resolution. Genes Dev. 24: 521-536. Svendsen JM, Smogorzewska A, Sowa ME, O'Connell BC, Gygi SP, Elledge SJ, Harper JW. 2009. Mammalian BTBD12/SLX4 assembles a Holliday junction resolvase and is required for DNA repair. Cell 138: 63-77. Swagemakers SMA, Essers J, de Wit J, Hoeijmakers JHJ, Kanaar R. 1998. The Human Rad54 Recombinational DNA Repair Protein Is a Double-stranded DNA- dependent ATPase. J. Biol. Chem. 273: 28292-28297. Symington LS. 2002. Role of RAD52 Epistasis Group Genes in Homologous Recombination and Double-Strand Break Repair. Microbiol. Mol. Biol. Rev. 66: 630-670. Symington LS, Gautier J. 2011. Double-Strand Break End Resection and Repair Pathway Choice. Annu. Rev. Genet. 45: 247-271. Symington LS, Heyer WD. 2006. Some disassembly required: role of DNA translocases in the disruption of recombination intermediates and dead-end complexes. Genes Dev. 20: 2479-2486. Symington LS, Holloman WK. 2008. Resolving resolvases: the final act? Mol. Cell 32: 603-604. Symington LS, Kolodner R. 1985. Partial purification of an enzyme from Saccharomyces cerevisiae that cleaves Holliday junctions. Proc. Natl. Acad. Sci. U. S. A. 82: 7247-7251. Takahagi M, Iwasaki H, Nakata A, Shinagawa H. 1991. Molecular analysis of the Escherichia coli ruvC gene, which encodes a Holliday junction-specific endonuclease. J. Bacteriol. 173: 5747-5753. Takata M, Sasaki MS, Sonoda E, Fukushima T, Morrison C, Albala JS, Swagemakers SMA, Kanaar R, Thompson LH, Takeda S. 2000. The Rad51 Paralog Rad51B Promotes Homologous Recombinational Repair. Mol. Cell. Biol. 20: 6476-6482. Takata M, Sasaki MS, Sonoda E, Morrison C, Hashimoto M, Utsumi H, Yamaguchi-Iwai Y, Shinohara A, Takeda S. 1998. Homologous recombination and non- homologous end-joining pathways of DNA double-strand break repair have

84

overlapping roles in the maintenance of chromosomal integrity in vertebrate cells. EMBO J. 17: 5497-5508. Takata M, Sasaki MS, Tachiiri S, Fukushima T, Sonoda E, Schild D, Thompson LH, Takeda S. 2001. Chromosome Instability and Defective Recombinational Repair in Knockout Mutants of the Five Rad51 Paralogs. Mol. Cell. Biol. 21: 2858-2866. Tamkun JW. 1995. The role of brahma and related proteins in transcription and development. Curr. Opin. Genet. Dev. 5: 473-477. Tan TL, Kanaar R, Wyman C. 2003. Rad54, a Jack of all trades in homologous recombination. DNA Repair (Amst) 2: 787-794. Tan TLR, Essers J, Citterio E, Swagemakers SMA, de Wit J, Benson FE, Hoeijmakers JHJ, Kanaar R. 1999. Mouse Rad54 affects DNA conformation and DNA- damage-induced Rad51 foci formation. Curr. Biol. 9: 325-328. Tavtigian SV, Simard J, Rommens J, Couch F, Shattuck Eidens D, Neuhausen S, Merajver S, Thorlacius S, Offit K, Stoppa Lyonnet D et al. 1996a. The complete BRCA2 gene and mutations in chromosome 13q-linked kindreds [see comments]. Nat. Genet. 12: 333-337. Tavtigian SV, Simard J, Rommens J, Couch F, Shattuck-Eidens D, Neuhausen S, Merajver S, Thorlacius S, Offit K, Stoppa-Lyonnet D et al. 1996b. The complete BRCA2 gene and mutations in chromosome 13q-linked kindreds. Nat. Genet. 12: 333-337. Tebbs RS, Zhao Y, Tucker JD, Scheerer JB, Siciliano MJ, Hwang M, Liu N, Legerski RJ, Thompson LH. 1995. Correction of chromosomal instability and sensitivity to diverse mutagens by a cloned cDNA of the XRCC3 DNA repair gene. Proc. Natl. Acad. Sci. U. S. A. 92: 6354-6358. Thacker J, Tambini CE, Simpson PJ, Tsui LC, Scherer SW. 1995. Localization to chromosome 7q36.1 of the human XRCC2 gene, determining sensitivity to DNA- damaging agents. Hum. Mol. Genet. 4: 113-120. Thoma F, Koller T. 1977. Influence of histone H1 on chromatin structure. Cell 12: 101- 107. Thoma NH, Czyzewski BK, Alexeev AA, Mazin AV, Kowalczykowski SC, Pavletich NP. 2005. Structure of the SWI2/SNF2 chromatin-remodeling domain of eukaryotic Rad54. Nat Struct Mol Biol 12: 350-356. Thompson LH, Schild D. 2001. Homologous recombinational repair of DNA ensures mammalian chromosome stability. Mutation Research/Fundamental and Molecular Mechanisms of Mutagenesis 477: 131-153. Thorslund T, Esashi F, West SC. 2007. Interactions between human BRCA2 protein and the meiosis-specific recombinase DMC1. EMBO J. 26: 2915-2922. Thorslund T, McIlwraith MJ, Compton SA, Lekomtsev S, Petronczki M, Griffith JD, West SC. 2010. The breast cancer tumor suppressor BRCA2 promotes the specific targeting of RAD51 to single-stranded DNA. Nat Struct Mol Biol 17: 1263-1265. Tjeertes JV, Miller KM, Jackson SP. 2009. Screen for DNA-damage-responsive histone modifications identifies H3K9Ac and H3K56Ac in human cells. EMBO J. 28: 1878-1889.

85

Tombline G, Fishel R. 2002. Biochemical characterization of the human RAD51 protein. I. ATP hydrolysis. J. Biol. Chem. 277: 14417-14425. Tsubouchi H, Roeder GS. 2003. The Importance of Genetic Recombination for Fidelity of Chromosome Pairing in Meiosis. Developmental Cell 5: 915-925. Tsukuda T, Fleming AB, Nickoloff JA, Osley MA. 2005. Chromatin remodelling at a DNA double-strand break site in Saccharomyces cerevisiae. Nature 438: 379-383. Tsutsui Y, Morishita T, Iwasaki H, Toh H, Shinagawa H. 2000. A Recombination Repair Gene of Schizosaccharomyces pombe, rhp57, Is a Functional Homolog of the Saccharomyces cerevisiae RAD57 Gene and Is Phylogenetically Related to the Human XRCC3 Gene. Genetics 154: 1451-1461. Tsuzuki T, Fujii Y, Sakumi K, Tominaga Y, Nakao K, Sekiguchi M, Matsushiro A, Yoshimura Y, MoritaT. 1996. Targeted disruption of the Rad51 gene leads to lethality in embryonic mice. Proc. Natl. Acad. Sci. U. S. A. 93: 6236-6240. Ulyanova NP, Schnitzler GR. 2005. Human SWI/SNF generates abundant, structurally altered dinucleosomes on polynucleosomal templates. Mol. Cell. Biol. 25: 11156- 11170. Unal E, Arbel-Eden A, Sattler U, Shroff R, Lichten M, Haber JE, Koshland D. 2004. DNA damage response pathway uses histone modification to assemble a double- strand break-specific cohesin domain. Mol. Cell 16: 991-1002. van Attikum H, Fritsch O, Gasser SM. 2007. Distinct roles for SWR1 and INO80 chromatin remodeling complexes at chromosomal double-strand breaks. EMBO J. 26: 4113-4125. van Attikum H, Fritsch O, Hohn B, Gasser SM. 2004. Recruitment of the INO80 complex by H2A phosphorylation links ATP-dependent chromatin remodeling with DNA double-strand break repair. Cell 119: 777-788. Van Komen S, Petukhova G, Sigurdsson S, Stratton S, Sung P. 2000. Superhelicity- Driven Homologous DNA Pairing by Yeast Recombination Factors Rad51 and Rad54. Mol. Cell 6: 563-572. Van Komen S, Petukhova G, Sigurdsson S, Sung P. 2002. Functional cross-talk among Rad51, Rad54, and replication protein A in heteroduplex DNA joint formation. J. Biol. Chem. 277: 43578-43587. Veaute X, Jeusset J, Soustelle C, Kowalczykowski SC, Le Cam E, Fabre F. 2003. The Srs2 helicase prevents recombination by disrupting Rad51 nucleoprotein filaments. Nature 423: 309-312. Wang B, Elledge SJ. 2007. Ubc13/Rnf8 ubiquitin ligases control foci formation of the Rap80/Abraxas/Brca1/Brcc36 complex in response to DNA damage. Proc. Natl. Acad. Sci. U. S. A. 104: 20759-20763. Wang B, Matsuoka S, Ballif BA, Zhang D, Smogorzewska A, Gygi SP, Elledge SJ. 2007. Abraxas and RAP80 form a BRCA1 protein complex required for the DNA damage response. Science 316: 1194-1198. Wang X, Haber JE. 2004. Role of Saccharomyces single-stranded DNA-binding protein RPA in the strand invasion step of double-strand break repair. PLoS Biol 2: E21. Ward JD, Muzzini DM, Petalcorin MI, Martinez-Perez E, Martin JS, Plevani P, Cassata G, Marini F, Boulton SJ. 2010. Overlapping mechanisms promote postsynaptic

86

RAD-51 filament disassembly during meiotic double-strand break repair. Mol. Cell 37: 259-272. Welz-Voegele C, Jinks-Robertson S. 2008. Sequence divergence impedes crossover more than noncrossover events during mitotic gap repair in yeast. Genetics 179: 1251- 1262. West SC. 1997. Processing of recombination intermediates by the RuvABC proteins. Annu. Rev. Genet. 31: 213-244. -. 2003. Molecular views of recombination proteins and their control. Nat Rev Mol Cell Biol 4: 435-445. -. 2009. The search for a human Holliday junction resolvase. Biochem. Soc. Trans. 37: 519-526. West SC, Korner A. 1985. Cleavage of cruciform DNA structures by an activity from Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. U. S. A. 82: 6445-6449. Whitby MC, Dixon J. 1997. A new Holliday junction resolving enzyme from Schizosaccharomyces pombe that is homologous to CCE1 from Saccharomyces cerevisiae. J. Mol. Biol. 272: 509-522. Wiese C, Dray EØ, Groesser T, San Filippo J, Shi I, Collins DW, Tsai M-S, Williams GJ, Rydberg B, Sung P et al. 2007. Promotion of Homologous Recombination and Genomic Stability by RAD51AP1 via RAD51 Recombinase Enhancement. Mol. Cell 28: 482-490. Williams AB, Michael WM. 2010. Eviction notice: new insights into Rad51 removal from DNA during homologous recombination. Mol. Cell 37: 157-158. Wold MS, Kelly T. 1988. Purification and characterization of replication protein A, a cellular protein required for in vitro replication of simian virus 40 DNA. Proc. Natl. Acad. Sci. U. S. A. 85: 2523-2527. Wolner B, Peterson CL. 2005. ATP-dependent and ATP-independent roles for the Rad54 chromatin remodeling enzyme during recombinational repair of a DNA double strand break. J. Biol. Chem. 280: 10855-10860. Wolner B, van Komen S, Sung P, Peterson CL. 2003. Recruitment of the Recombinational Repair Machinery to a DNA Double-Strand Break in Yeast. Mol. Cell 12: 221-232. Wong AKC, Pero R, Ormonde PA, Tavtigian SV, Bartel PL. 1997. RAD51 Interacts with the Evolutionarily Conserved BRC Motifs in the Human Breast Cancer Susceptibility Gene . J. Biol. Chem. 272: 31941-31944. Wong JMS, Ionescu D, Ingles CJ. 2003. Interaction between BRCA2 and replication protein A is compromised by a cancer-predisposing mutation in BRCA2. Oncogene 22: 28-33. Wooster R, Bignell G, Lancaster J, Swift S, Seal S, Mangion J, Collins N, Gregory S, Gumbs C, Micklem G et al. 1995. Identification of the breast cancer susceptibility gene BRCA2. Nature 378: 789-792. Wooster R, Neuhausen S, Mangion J, Quirk Y, Ford D, Collins N, Nguyen K, Seal S, Tran T, Averill D et al. 1994. Localization of a breast cancer susceptibility gene, BRCA2, to chromosome 13q12-13. Science 265: 2088-2090.

87

Wu D, Topper LM, Wilson TE. 2008a. Recruitment and dissociation of nonhomologous end joining proteins at a DNA double-strand break in Saccharomyces cerevisiae. Genetics 178: 1237-1249. Wu L, Davies SL, North PS, Goulaouic H, Riou JF, Turley H, Gatter KC, Hickson ID. 2000. The Bloom's syndrome gene product interacts with topoisomerase III. J. Biol. Chem. 275: 9636-9644. Wu L, Hickson ID. 2006. DNA helicases required for homologous recombination and repair of damaged replication forks. Annu. Rev. Genet. 40: 279-306. Wu Y, He Y, Moya IA, Qian X, Luo Y. 2004. Crystal structure of archaeal recombinase RADA: a snapshot of its extended conformation. Mol. Cell 15: 423-435. Wu Y, Kantake N, Sugiyama T, Kowalczykowski SC. 2008b. Rad51 protein controls Rad52-mediated DNA annealing. J. Biol. Chem. 283: 14883-14892. Wu Y, Qian X, He Y, Moya IA, Luo Y. 2005. Crystal structure of an ATPase-active form of Rad51 homolog from Methanococcus voltae. Insights into potassium dependence. J. Biol. Chem. 280: 722-728. Xia B, Sheng Q, Nakanishi K, Ohashi A, Wu J, Christ N, Liu X, Jasin M, Couch FJ, Livingston DM. 2006. Control of BRCA2 Cellular and Clinical Functions by a Nuclear Partner, PALB2. Mol. Cell 22: 719-729. Xia F, Taghian DG, DeFrank JS, Zeng Z-C, Willers H, Iliakis G, Powell SN. 2001. Deficiency of human BRCA2 leads to impaired homologous recombination but maintains normal nonhomologous end joining. Proc. Natl. Acad. Sci. U. S. A. 98: 8644-8649. Xiao A, Li H, Shechter D, Ahn SH, Fabrizio LA, Erdjument-Bromage H, Ishibe- Murakami S, Wang B, Tempst P, Hofmann K et al. 2009. WSTF regulates the H2A.X DNA damage response via a novel tyrosine kinase activity. Nature 457: 57-62. Xu F, Zhang K, Grunstein M. 2005. Acetylation in histone H3 globular domain regulates gene expression in yeast. Cell 121: 375-385. Xu Y, Sun Y, Jiang X, Ayrapetov MK, Moskwa P, Yang S, Weinstock DM, Price BD. 2010. The p400 ATPase regulates nucleosome stability and chromatin ubiquitination during DNA repair. J. Cell Biol. 191: 31-43. Yamada K, Miyata T, Tsuchiya D, Oyama T, Fujiwara Y, Ohnishi T, Iwasaki H, Shinagawa H, Ariyoshi M, Mayanagi K et al. 2002. Crystal structure of the RuvA-RuvB complex: a structural basis for the Holliday junction migrating motor machinery. Mol. Cell 10: 671-681. Yamaguchi-Iwai Y, Sonoda E, Buerstedde J-M, Bezzubova O, Morrison C, Takata M, Shinohara A, Takeda S. 1998. Homologous Recombination, but Not DNA Repair, Is Reduced in Vertebrate Cells Deficient in RAD52. Mol. Cell. Biol. 18: 6430- 6435. Yang H, Jeffrey PD, Miller J, Kinnucan E, Sun Y, Thoma NH, Zheng N, Chen PL, Lee WH, Pavletich NP. 2002. BRCA2 function in DNA binding and recombination from a BRCA2-DSS1-ssDNA structure. Science 297: 1837-1848.

88

Yang H, Li Q, Fan J, Holloman WK, Pavletich NP. 2005. The BRCA2 homologue Brh2 nucleates RAD51 filament formation at a dsDNA-ssDNA junction. Nature 433: 653-657. Yin J, Sobeck A, Xu C, Meetei AR, Hoatlin M, Li L, Wang W. 2005. BLAP75, an essential component of Bloom's syndrome protein complexes that maintain genome integrity. EMBO J. 24: 1465-1476. Yu VPCC, Koehler M, Steinlein C, Schmid M, Hanakahi LA, van Gool AJ, West SC, Venkitaraman AR. 2000. Gross chromosomal rearrangements and genetic exchange between nonhomologous chromosomes following BRCA2 inactivation. Genes Dev. 14: 1400-1406. Zakharyevich K, Ma Y, Tang S, Hwang PY-H, Boiteux S, Hunter N. 2010. Temporally and Biochemically Distinct Activities of Exo1 during Meiosis: Double-Strand Break Resection and Resolution of Double Holliday Junctions. Mol. Cell 40: 1001-1015. Zhang Z, Fan H-Y, Goldman JA, Kingston RE. 2007. Homology-driven chromatin remodeling by human RAD54. Nat Struct Mol Biol 14: 397-405. Zhao GY, Sonoda E, Barber LJ, Oka H, Murakawa Y, Yamada K, Ikura T, Wang X, Kobayashi M, Yamamoto K et al. 2007. A critical role for the ubiquitin- conjugating enzyme Ubc13 in initiating homologous recombination. Mol. Cell 25: 663-675. Zhou Q, Kojic M, Cao Z, Lisby M, Mazloum NA, Holloman WK. 2007. Dss1 Interaction with Brh2 as a Regulatory Mechanism for Recombinational Repair. Mol. Cell. Biol. 27: 2512-2526. Zhu Z, Chung W-H, Shim EY, Lee SE, Ira G. 2008. Sgs1 Helicase and Two Nucleases Dna2 and Exo1 Resect DNA Double-Strand Break Ends. Cell 134: 981-994. Zickler D, Kleckner N. 1999. Meiotic chromosomes: integrating structure and function. Annu. Rev. Genet. 33: 603-754.

89

CHAPTER 2

Subunit Interface Residues F129 and H294 of Human RAD51 Are Essential For

Recombinase Function

90

2.1 ABSTRACT

RAD51 mediated homologous recombinational repair (HRR) of DNA double-strand breaks (DSBs) is essential to maintain genomic integrity. RAD51 forms a nucleoprotein filament (NPF) that catalyzes the fundamental homologous pairing and strand exchange reaction (recombinase) required for HRR. Based on structural and functional homology with archaeal and yeast RAD51, we have identified the human RAD51 (HsRAD51) subunit interface residues HsRad51(F129) in the Walker A box and HsRad51(H294) in the L2 ssDNA binding region as potentially important participants in salt-induced conformational transitions essential for recombinase activity. We demonstrate that the

HsRad51(F129V) and HsRad51(H294V) substitution mutations reduce DNA dependent

ATPase activity and are largely defective in the formation of a functional NPF, which ultimately eliminates recombinase catalytic functions. Our data are consistent with the conclusion that the HsRAD51(F129) and HsRAD51(H294) residues are important participants in the cation-induced allosteric activation of HsRAD51.

91

2.2 INTRODUCTION

Failure to repair DNA double strand breaks (DSBs) leads to tumorigenesis and genomic instability. Homologous recombination (HR) is an evolutionary conserved repair pathway utilized to restore DSBs. HR mediated DSB repair is initiated by resection of the 5-end of the break to leave a 3-single-stranded DNA (ssDNA) overhang (Mimitou and Symington 2008; Zhu et al. 2008). In eukaryotes, RAD51 forms a nucleoprotein filament (NPF) on the newly formed ssDNA region aided by other recombination mediators such as RAD52 in yeast and BRCA2 in vertebrates (Yang et al. 2002;

Sugawara et al. 2003; Lisby et al. 2004; Yang et al. 2005; Carreira et al. 2009; Jensen et al. 2010). The key function of the RAD51 NPF is to catalyze the homology search and initiate strand exchange. Deletion of Rad51 in mice results in embryonic lethality, while

RAD51 knock down in chicken DT40 cell lines results in increased chromosomal instability (Tsuzuki et al. 1996; Sonoda et al. 1998). Even though, no mutations of

RAD51 have been found in cancers, its expression is elevated in many cancer cell lines; perhaps to provide an advantage to rapidly dividing cells by repairing DSBs that would lead to replication fork collapse (Richardson et al. 2004; Klein 2008; Schild and Wiese

2010).

Despite the functional conservation with bacterial RecA, human RAD51

(HsRAD51) possesses an essential cation salt requirement for efficient strand exchange

+ in vitro. The most effective cation is ammonium (NH4 ), which is unlikely to be physiologically significant. However, potassium (K+) also enhances HsRAD51 recombinase functions to a lesser extent (Liu et al. 2004; Shim et al. 2006) and has been

92

shown to induce single stranded DNA-HsRAD51 complexes that structurally mimic active RecA NPFs (Liu et al. 2004). The effect of salt has been attributed to an induced preferential single stranded DNA (ssDNA) binding by RAD51 over double stranded

DNA (dsDNA) (Shim et al. 2006). However, crystallographic analysis of the methanococcus voltae RadA (MvRAD51) has revealed important K+-induced amino acid and structural rearrangements within the intersubunit region of the NPF (Figure 2.1A and

1B). For example, the MvRAD51(F107) residue in the highly conserved Walker A box rotates away from the ATP binding interface to accommodate K+ cations (Figure 2.1b;

(Wu et al. 2004). This rotation of MvRAD51(F107) induces ordering of the L2 ssDNA binding domain and a conformational transition of MvRAD51(H280) that results in the formation of a direct hydrogen bond with the -phosphate group of the ATP analogue; which is suggested to be involved in polarizing the water molecule involved in ATP hydrolysis. These two residues are conserved in RAD51 homologs from archaebacteria to human (Figure 2.1C).

Here, we have examined substitution mutations of the analogous HsRAD51 residues to MvRAD51(F107) and MvRAD51(H280), HsRAD51(F129) and

HsRAD51(H294). We find that HsRAD51(F129V) and HsRAD51(H294V) affect ATP hydrolysis (ATPase) activity but not adenosine nucleotide binding or ADPATP exchange. Moreover, they alter DNA binding properties in the presence of ATP and salt

+ + cations (K or NH4 ) that ultimately results in defective recombinase functions. These data further delineate the importance of salt-induced allosteric changes at the subunit interface of HsRAD51 that promotes the formation of a functional NPF.

93

2.3 MATERIALS AND METHODS

HsRAD51 Protein Expression and Purification- The mutant RAD51 genes F129V and

H294V were constructed using wild type HsRAD51 gene in the pET24d vector system

(Novagen) using conventional PCR. All mutations were confirmed by DNA sequencing.

Wild type and mutant hRAD51 proteins were expressed and purified following previously published protocols (Baumann et al. 1997; Tombline and Fishel 2002).

Briefly, hRAD51 was expressed in E. coli BLR strain and precipitated using spermidine-

HCl. Resuspended pellet was purified using Reactive-Blue-4-agarose (Sigma-Aldrich),

Heparin sepharose (GE Healthcare), hydroxyapatite (Bio-Rad) and Mono Q (GE

Healthcare) column chromatography. Purity of the fractions was verified by SDS-PAGE analysis. HsRPA was expressed in BL21(AI) cells using pET11d-tRPA purified as described (Henricksen et al. 1994), except for the resuspension of cells where HI buffer was supplemented with 100mM KCl.

DNA Substrates- X174 single-stranded (ss) virion DNA and replicative form I (RFI) were purchased from NEB. X174 RFIII was obtained by linearizing RFI with ApaLI restriction enzyme and gel purifying with Qiaquick Gel Extraction kit (Qiagen). For surface plasmon resonance (SPR) analysis, a 5′ biotinylated oligo dT50 was used as ssDNA and for dsDNA 5′ biotinylated 50-mer 5-TCG AGA GGG TAA ACC ACA-

ATT ATT GAT ATA AAA TAG TTT TGG GTA GGC GA was annealed with its complement. D-loop assay substrates were prepared as described (Van Komen et al.

2002). 94

ATPase assay- ATP hydrolysis was measured as previously described (Tombline and

Fishel 2002). Reactions were performed in 10 L volumes in Buffer A containing 20mM

HEPES (pH 7.5), 10% glycerol, 100g/mL BSA, 1mM DTT and 1mM MgCl2 and when indicated, 150mM KCl. Each reaction mixture contained RAD51 (1M) and 6M (nt or bp) of X174 ssDNA or dsDNA. Reactions were initiated by the addition of protein and incubation at 37°C. After 1 hr 400L of 10% activated charcoal supplemented with

10mM EDTA was added to terminate the reaction and incubated on ice for another 2 hrs.

After centrifuging for 10 min 50L duplicate aliquots were taken for counting [32P] free phosphate by Cerenkov method. Kinetic parameters were obtained by fitting data into

Michaelis-Menten equation using the software Kaleidagraph (Synergy software).

ATPS/ADP binding assay- Experimental procedure was as previously described

(Tombline et al. 2002). Reactions were performed in Buffer A and 150mM KCl, when indicated. RAD51 (1M) and 6M (nt) X174 ssDNA were used with the indicated amount of [-35S]ATPS or [3H]ADP. After a 30 min incubation at 37°C, reactions were kept on ice until filtered. Reaction was added into 4mL of ice-cold reaction buffer and filtered through HAWP nitrocellulose filters (Millipore) presoaked in the same buffer.

Another 4mL of buffer was used to wash the membrane and the filter was dried for 2 hrs.

Radioactivity was counted after an overnight incubation in liquid scintillation fluid.

95

ADP-ATP exchange assay- Reactions were performed in Buffer A with 150mM KCl.

RAD51(1M) and 6M (nt) X174 ssDNA was incubated with 3 M of [3H]ADP in a

60L reaction volume at 37°C for 10 mins. ADP to ATP exchange was initiated by addition of cold ATP to a 5mM final concentration. 10L aliquots were withdrawn at indicated time points, filtered and analyzed as in ADP binding.

SPR analysis- Biotinylated DNA was immobilized on a streptavidin-coated chip (GE

Healthcare) for the analysis. Protein binding and dissociation were analyzed at 25°C with a 5L/min flow rate on a Biacore 3000 (GE Healthcare). Reactions were performed in Buffer containing 20mM HEPES (pH 7.5), 10% glycerol, 1mM DTT and 1mM MgCl2,

0.005% surfactant P-20 (GE Healthcare), 2.5mM of the indicated nucleotide and 150mM

2+ KCl when indicated. For experiments with Ca , 1mM CaCl2 was used instead of MgCl2.

DNA binding was determined by injecting 100nM, 200nM, 400nM, 800nM and 1.6M of protein simultaneously into ssDNA and dsDNA containing flow channels to ensure saturated binding.

D-loop assay- 2.4 M (nt) labeled 90-mer was incubated with HsRAD51 (0.8M) in buffer A with 1mM ATP supplemented with the indicated amounts of MgCl2 or CaCl2 at

37°C for 10 min. Reaction was initiated by adding 35M (bp) supercoiled pBS-SK(-) plasmid and incubated further for 15 min. Samples were deproteinized by addition of 1%

SDS and 1 mg/mL proteinase K to a final concentration and incubated for 15 min at

37°C, mixed with 1/5 volume of gel loading dye and analyzed on 0.9% agarose gel in

96

TAE buffer, run at 4V/cm at 25°C. D-loops were quantified after drying and exposure to

PhosphoImager (Molecular Dynamics) screens.

DNA Strand Exchange Assay- HsRAD51 (5M) and 30M (nt) X174 circular ssDNA were pre-incubated in buffer containing 20mM HEPES (pH 7.5), 10% Glycerol, 1mM

DTT, 1mM MgCl2 supplemented with 2.5mM of ATP at 37°C for 5 min before addition of 150mM of the indicated salt and 15M (bp) linear X174 dsDNA. After another 5 min incubation 2M of HsRPA was added and the incubation was continued. After 3 hrs samples were deproteinized by addition of 3L of stop buffer containing 2% SDS and 10 mg/mL proteinase K, and analyzed on 0.9% agarose gel in TAE buffer. Electrophoresis was carried out at 4V/cm at 25°C with 0.1g/mL ethidium bromide.

97

2.4 RESULTS

Mutation of HsRAD51 inter-subunit residues F129 and H294 affect ATP turnover

HsRAD51(F129) and HsRAD51(H294) were mutated to Val (V) to minimize initial structural perturbations. The respective proteins were then purified to near homogeneity

(Fig. 2.2A). The RecA/RAD51 family of proteins exhibits DNA dependent ATPase activity (Weinstock et al. 1981a; Weinstock et al. 1981b; Pugh and Cox 1987; Baumann et al. 1996; Tombline and Fishel 2002). HsRAD51 possesses a modestly higher ATP turnover (kcat) in the presence of ssDNA compared to dsDNA, although it is about 150 times slower than the bacterial RecA (Weinstock et al. 1981a; Weinstock et al. 1981b;

Pugh and Cox 1987; Tombline and Fishel 2002). The HsRAD51(F129V) substitution mutation displays an ~2-fold reduction in kcat compared to the wild type HsRAD51, while the HsRAD51(H294V) substitution mutation nearly abolishes the ATPase activity (Fig.

2.2B, Table 1).

Reduced kcat could be due to several factors: an ATP binding deficiency, an ADP release deficiency or a deficiency in the catalytic step. To rule out the first two possibilities, we examined ATP binding, ADP binding and ADP-ATP exchange of the mutant proteins. We found no significant difference in ATP binding (Fig. 2.2C, Table 1),

ADP binding (Fig. 2.2D, Table 1) or ADP-ATP exchange (data not shown). Previous studies have confirmed a small but reproducible salt induced ATPase catalytic rate enhancement in the presence of ssDNA compared to dsDNA (Shim et al. 2006). Only

RAD51(F129V) displayed a similar salt (150 mM KCl) induced catalytic rate enhancement in the presence of ssDNA (Fig. 2.2E). Together these observations are

98

consistent with the conclusion that HsRAD51(H294) residue and to a significantly lesser extent the HsRAD51(F129) residue are required for appropriate ATP catalysis.

HsRAD51(F129V) and HsRAD51(H294V) are deficient in D-loop formation and strand exchange

We examined the recombinase activity of the HsRAD51(F129V) and HsRAD51(H294V) substitution mutations. HsRAD51 catalyzes D-loop formation in vitro between a 32P- labeled 90-mer and a homologous supercoiled plasmid (Fig. 2.3A). Efficient D-loop formation occurs when ATP is present in a slow-hydrolysable state (e.g. in the presence of calcium) or when non-hydrolysable ATP analogues such as AMP-PNP are substituted for ATP (Bugreev and Mazin 2004; Chi et al. 2006). Importantly, the Walker A box mutant HsRAD51(K133R) that binds ATP normally but is defective in ATP hydrolysis, catalyzes efficient D-loop formation with ATP and Mg2+ (Chi et al. 2006). These results have suggested that D-loop catalysis requires HsRAD51 to form and maintain an active

ATP-bound NPF (Bugreev and Mazin 2004; Ristic et al. 2005; Chi et al. 2006; San

Filippo et al. 2008).

In the presence of ATP, HsRAD51 catalyzed D-loop formation with Ca2+ but not with Mg2+ (Fig. 2.3B). Previous work has demonstrated that Ca2+ induces the formation of a stable ATP-bound NPF while Mg2+ allows the hydrolysis of ATP that results in a mixed (ADP/ATP) NPF that is largely inactive (Bugreev and Mazin 2004).

HsRAD51(F129V) and HsRAD51(H294V) did not catalyze D-loop formation in either

Mg2+ or Ca2+ (Fig. 3B). Thus, even though HsRAD51(H294V) is ATPase deficient, and both HsRAD51(F129V) and HsRAD51(H294V) display ATP binding that is comparable

99

to the wild type, these mutant proteins are incapable of catalyzing D-loop formation.

Collectively these results indicate that suppression of ATP hydrolysis alone is not sufficient to confer enhanced D-loop recombinase activity.

Catalysis of strand exchange between a duplex DNA substrate and a homologous single stranded circular DNA substrate is a hallmark of RecA/RAD51 proteins (Cox and

Lehman 1981; Menetski et al. 1990; Kowalczykowski 1991; Kowalczykowski and

Eggleston 1994; Baumann et al. 1996; San Filippo et al. 2008). Unlike D-loop formation

+ + that occurs in low salt, strand exchange requires specific cations (either NH4 or K ) to activate HsRAD51 activity (Sigurdsson et al. 2001; Liu et al. 2004; Shim et al. 2006).

Using X174 virion DNA and ApaL1 linearized X174 dsDNA we compared the strand exchange activity of the HsRAD51(F129V) and HsRAD51(H294V) mutant proteins to wild type HsRAD51. We also examined the stimulatory effect of the human single- stranded binding complex RPA (HsRPA). Neither HsRAD51(F129V) nor

HsRAD51(H294V) were able to form joint molecules in the presence of salt and/or RPA

(Fig. 2.3C). We can conclude that the HsRAD51(F129) and HsRAD51(H294) residues are critical for cation-induced HsRAD51 recombinase functions.

HsRAD51(F129) and HsRAD51(H294) are essential for ATP-dependent DNA binding

We examined real-time HsRAD51 ssDNA and dsDNA binding by surface plasmon resonance (SPR, Biacore). Biotinylated dT50 ssDNA and 50 bp dsDNA were immobilized on a streptavidin-coated flow-cell surface. SPR measures the change in the refractive index that reflects protein binding and/or dissociation from DNA. We analyzed ssDNA and dsDNA binding in salt conditions similar to those used in for strand

100

exchange. We titrated the protein from 100 nM to 1.6 M in the presence of ADP, ATP and in the absence of adenosine nucleotide. For clarity the binding and dissociation curves shown correspond to 800 nM of protein (Fig. 2.4). After protein injection nucleoprotein filament formation was analyzed for 150 sec. For steady-state protein disassembly, protein free buffer was injected into the flow cells for 300 sec. Because the ssDNA and dsDNA were introduced into separate channels we could simultaneously examine protein binding and dissociation to the two DNA substrates. In the absence of a adenosine nucleotide only the wild type HsRAD51 and HsRAD51(H294V) bound to ssDNA (Fig. 2.4A, left panel). In the presence of ADP, the wild type and mutant variants bound to ssDNA, but dissociated rapidly (Fig. 2.4A, middle panel). These results are consistent with previous studies that have suggested protein turnover associated with

ADP-bound RAD51 (Bugreev and Mazin 2004; Hilario et al. 2009). Interestingly, in the presence of ATP, wild type HsRAD51 bound ssDNA while both HsRAD51(F129V) and

HsRAD51(H294V) were largely defective in ssDNA binding (Fig. 2.4A, right panel).

Only wild type HsRAD51 displayed dsDNA binding in the presence of ATP as well as in the absence of a nucleotide (Fig. 2.4B). These results suggest that the HsRAD51(F129) and HsRAD51(H294) residues are important for the ATP induced DNA binding that ultimately results in an active NPF required for recombinase function.

The abnormal DNA binding did not appear to fully account for the altered steady- state ATPase activity of HsRAD51(F129V) and HsRAD51(H294V) (compare Fig. 2.2E,

Table 2.1 with Fig. 2.4). This is particularly true for HsRAD51(H294V) which showed reduced but not absent binding to ssDNA, yet near background ATPase activity. The dissociation of ssDNA binding activity from ATPase activity suggests that some other

101

function(s) of these proteins are compromised. Based on the location of these residues in the crystal structures we postulate that cation induced conformational transitions are defective in HsRAD51(F129V) and HsRAD51(H294V) (see Figs. 2.1A and 2.1B).

These conformational transitions do not influence ADP/ATP binding or ADPATP exchange, but do affect the ability of the mutant proteins to interact appropriately with

DNA and to properly catalyze ATP hydrolysis.

102

2.5 DISCUSSION

Structural studies of the MvRAD51 and S.cerevisiae RAD51 (ScRAD51) have detailed significant conformational transitions associated with the NPF. The ScRAD51 residues analogous to HsRAD51(F129) and HsRAD51(H294), ScRAD51(F187) and

ScRAD51(H352), were found in two alternate conformations within the NPF (Conway et al. 2004; Qian et al. 2005). In one conformation ScRAD51(H352) was positioned above the ATP binding site while in the other ScRAD51(F187) excluded ScRAD51(H352) by moving it out of the active site. Importantly, K+ cations induce similar conformational transitions of MvRAD51 of the analogous residues to HsRAD51(F129) and

HsRAD51(H294), MvRAD51(F107) and MvRAD51(H280). In the presence of K+ the

MvRAD51(F107) rotates away from the ATP catalytic interface and MvRAD51(H280) rotates in to form a hydrogen bond interaction with the -phosphate of ATP, while at the same time ordering the L2 ssDNA binding domain (Wu et al. 2004).

Our studies support the importance of these conformational transitions. While

HsRAD51(F129V) can bind and hydrolyze ATP, it displays significantly reduced ssDNA binding in the presence of ATP. These results suggest an impaired ability to incite the conformational transitions necessary to the form an active NPF on ssDNA. We speculate that this impairment involves an inability of the smaller Val residue to incite ordering of the L2 ssDNA-binding domain that is normally provoked by cation-induced rotation of the Phe residue away from the ATP catalytic interface. In contrast, HsRAD51(H294V) binds but does not hydrolyze ATP, binds ssDNA in the absence of adenosine nucleotide, yet displays significantly reduced ssDNA binding in the presence of ATP. We speculate

103

that these impairments are the result of an inability of the Val residue to form an appropriate hydrogen bond interaction with the -phosphate of ATP in spite of the fact that L2 ssDNA-binding domain ordering has been provoked by the MvRAD51(F107) residue (Wu et al. 2005). Modeling of the HsRAD51 structure strongly support such an allosteric communication between ATPase site and L2 ssDNA binding region (Renodon-

Corniere et al. 2008; Reymer et al. 2009). A recent mutational analysis also revealed that the subunit interface residue ScRAD51(H352) is critical for functional NPF formation and strand exchange activity (Grigorescu et al. 2009). Ultimately, the defects associated with both HsRAD51(F129V) and HsRAD51(H294V) result in defective D-loop and strand exchange functions. Because both proteins bind ADP and ATP similar to the wild type HsRAD51, the functional defects are unlikely to be a result of improper folding and/or aggregation.

For RecA/RAD51 proteins ATP binding but not necessarily hydrolysis, is sufficient to catalyze strand exchange (Menetski et al. 1990; Kowalczykowski 1991;

Rehrauer and Kowalczykowski 1993; Kowalczykowski and Eggleston 1994; Shan et al.

1996; Chi et al. 2006). It has been suggested that RAD51 might utilize binding of ATP and the resulting release of the hydrolysis product as a conformational switch for regulating recombinase function similar to other members of the AAA+ superfamily

(Erzberger and Berger 2006). Our data is consistent with the hypothesis that the

HsRAD51(F129) and HsRAD51(H294) residues along with salt cations may play a significant role in this conformational switch during HRR.

104

2.6 ACKNOWLEDGEMENTS AND FOOTNOTES

This chapter was published as: Subunit Interface Residues F129 and H294 of Human

RAD51 Are Essential For Recombinase Function. Amunugama R. and Fishel R.

Plos One (2011) 6(8):e23071.

105

Figure 2.1 Cation-induced conformational rearrangement of conserved amino acid residues of RadA. (A) Subunit interface of MvRadA (MvRAD51) structure in the absence of potassium cation (PDB code 1T4G). (B) Subunit interface region of

MvRAD51 structure in the presence of potassium cation (PDB code 1XU4). Structural figures were generated using Pymol. (C) Sequence alignment of WalkerA/P-loop and L2 ssDNA binding region of H. sapiens (Hs), S. cerevisiae (Sc) and M. voltae (Mv) recombinases. HsRAD51 residues F129 and H294 are indicated with asterisks.

106

Figure 2.2 Mutation of HsRAD51(F129) and HsRAD51(H294) residues affect ATP turnover. (A) Purification of wild type and HsRAD51 substitution mutant proteins. Protein (1 g) analyzed by 12% SDS-PAGE. (B) Steady-state ATPase activity with ssDNA in the presence of 150 mM KCl. (C) ATPS binding by wild type and HsRAD51 substitution mutant proteins in the presence of ssDNA and 150 mM KCl. (D) ADP binding by wild type and HsRAD51 substitution mutant proteins in the presence of ssDNA and 150 mM KCl. (E) ATP turnover (kcat) with ssDNA and dsDNA in the + presence and absence of KCl (K ). kcat values were calculated by Michaelis-Menten analysis. Error bars indicate standard deviation from at least three independent experiments.

Continued

107

Figure 2.2: continued

108

Figure 2.3. HsRAD51(F129V) and HsRAD51(H294V) are deficient in D-loop formation and strand exchange. (A) In vitro D-loop assay reaction schematic. (B) 0.8M of HsRAD51, HsRAD51(F129V), or HsRAD51(H294V) and [P32]-labeled ssDNA (90mer; 2.4 M nt) were preincubated for 10 min at 37°C in the reaction buffer containing 1 mM ATP and 1mM MgCl2 or CaCl2. Reactions were initiated by the addition of supercoiled pBS SK(-) plasmid DNA (35M bp). After 15 min, reactions were terminated by the addition of proteinase-K and SDS. Joint molecules (JMs) were analyzed on a 0.9% agarose gel. (C) Analysis of salt and RPA requirement for strand exchange. Reaction schematic shown above: HsRAD51 (5M) and X174 circular ssDNA (30M nt) were pre-incubated with 2.5mM ATP and 1mM MgCl2 at 37°C for 5 min prior to the addition of 150mM NaNH4HPO4 (if indicated) and linear X174 dsDNA (15M bp). After 5 min, HsRPA (2M) was added (if indicated) and the incubation was continued for 3 hrs. Samples were deproteinized and analyzed on 0.9% agarose gel with 0.1g/mL ethidium bromide.

Continued

109

Figure 2.3: continued

110

Figure 2.4 F129 and H294 of HsRAD51 are critical for DNA binding in the presence of ATP. (A) ssDNA binding analysis of wild type and HsRAD51 substitution mutant proteins by surface Plasmon resonance (SPR, Biacore) in the absence of an adenine nucleotide, in the presence of ADP and in the presence of ATP. Association and dissociation curve corresponding to 800nM of each protein is shown. (B) dsDNA binding analysis of wild type and HsRAD51 substitution mutant proteins.

111

Table 2.1 Summary of ATP hydrolysis and nucleotide binding data of HsRAD51 wild type (WT) and HsRAD51(F129V) and HsRAD51(H294V) mutant proteins.

Kinetic Parameter WT F129V H294V -1 kcat (min ) ssDNA 0.180 ± 0.004 0.094 ± 0.004 0.010 ± 0.001 ssDNA, KCl 0.190 ± 0.004 0.119 ± 0.009 0.008 ± 0.001

dsDNA 0.167 ± 0.002 0.095 ± 0.009 0.021 ± 0.002 dsDNA, KCl 0.135 ± 0.007 0.101 ± 0.008 0.013 ± 0.002

Km (M) ssDNA 17.48 ± 3.25 6.68 ± 1.39 2.59 ± 0.10 ssDNA, KCl 9.34 ± 1.42 6.40 ± 0.032 6.40 ± 3.31

dsDNA 6.54 ± 0.67 4.23 ± 0.38 5.38 ± 0.50 dsDNA, KCl 5.52 ± 0.26 4.90 ± 0.24 8.82 ± 0.26

KD(M) ATPS ssDNA 2.49 ± 0.46 1.53 ± 0.69 0.817 ± 0.14 ssDNA, KCl 1.18 ± 0.22 1.15 ± 0.55 1.40 ± 0.31

Bmax(M) ATPS ssDNA 1.02 ± 0.02 0.85 ± 0.03 0.63 ± 0.01 ssDNA, KCl 0.83 ± 0.04 0.87 ± 0.04 0.77 ± 0.03

KD (M) ADP ssDNA, KCl 1.24 ± 0.32 1.27 ± 0.11 2.66 ± 0.10 Bmax (M) ADP ssDNA, KCl 0.82 ± 0.05 1.03 ± 0.02 0.95 ± 0.01

.

112

2.7 REFERENCES

Baumann P, Benson FE, Hajibagheri N, West SC. 1997. Purification of human Rad51 protein by selective spermidine precipitation. Mutat. Res. 384: 65-72. Baumann P, Benson FE, West SC. 1996. Human Rad51 protein promotes ATP- dependent homologous pairing and strand transfer reactions in vitro. Cell 87: 757- 766. Bugreev DV, Mazin AV. 2004. Ca2+ activates human homologous recombination protein Rad51 by modulating its ATPase activity. Proc. Natl. Acad. Sci. U. S. A. 101: 9988-9993. Carreira A, Hilario J, Amitani I, Baskin RJ, Shivji MK, Venkitaraman AR, Kowalczykowski SC. 2009. The BRC repeats of BRCA2 modulate the DNA- binding selectivity of RAD51. Cell 136: 1032-1043. Chi P, Van Komen S, Sehorn MG, Sigurdsson S, Sung P. 2006. Roles of ATP binding and ATP hydrolysis in human Rad51 recombinase function. DNA Repair (Amst) 5: 381-391. Conway AB, Lynch TW, Zhang Y, Fortin GS, Fung CW, Symington LS, Rice PA. 2004. Crystal structure of a Rad51 filament. Nat Struct Mol Biol 11: 791-796. Cox MM, Lehman IR. 1981. recA protein of Escherichia coli promotes branch migration, a kinetically distinct phase of DNA strand exchange. Proc. Natl. Acad. Sci. U. S. A. 78: 3433-3437. Erzberger JP, Berger JM. 2006. Evolutionary relationships and structural mechanisms of AAA+ proteins. Annu. Rev. Biophys. Biomol. Struct. 35: 93-114. Grigorescu AA, Vissers JH, Ristic D, Pigli YZ, Lynch TW, Wyman C, Rice PA. 2009. Inter-subunit interactions that coordinate Rad51's activities. Nucleic Acids Res 37: 557-567. Henricksen LA, Umbricht CB, Wold MS. 1994. Recombinant replication protein A: expression, complex formation, and functional characterization. J. Biol. Chem. 269: 11121-11132. Hilario J, Amitani I, Baskin RJ, Kowalczykowski SC. 2009. Direct imaging of human Rad51 nucleoprotein dynamics on individual DNA molecules. Proc. Natl. Acad. Sci. U. S. A. 106: 361-368. Jensen RB, Carreira A, Kowalczykowski SC. 2010. Purified human BRCA2 stimulates RAD51-mediated recombination. Nature 467: 678-683. Klein HL. 2008. The consequences of Rad51 overexpression for normal and tumor cells. DNA Repair (Amst) 7: 686-693. Kowalczykowski SC. 1991. Biochemistry of genetic recombination: energetics and mechanism of DNA strand exchange. Annu. Rev. Biophys. Biophys. Chem. 20: 539-575. Kowalczykowski SC, Eggleston AK. 1994. Homologous pairing and DNA strand- exchange proteins. Annu. Rev. Biochem. 63: 991-1043.

113

Lisby M, Barlow JH, Burgess RC, Rothstein R. 2004. Choreography of the DNA damage response: spatiotemporal relationships among checkpoint and repair proteins. Cell 118: 699-713. Liu Y, Stasiak AZ, Masson JY, McIlwraith MJ, Stasiak A, West SC. 2004. Conformational changes modulate the activity of human RAD51 protein. J. Mol. Biol. 337: 817-827. Menetski JP, Bear DG, Kowalczykowski SC. 1990. Stable DNA heteroduplex formation catalyzed by the Escherichia coli RecA protein in the absence of ATP hydrolysis. Proc. Natl. Acad. Sci. U. S. A. 87: 21-25. Mimitou EP, Symington LS. 2008. Sae2, Exo1 and Sgs1 collaborate in DNA double- strand break processing. Nature 455: 770-774. Pugh BF, Cox MM. 1987. Stable binding of recA protein to duplex DNA. Unraveling a paradox. J. Biol. Chem. 262: 1326-1336. Qian X, Wu Y, He Y, Luo Y. 2005. Crystal structure of Methanococcus voltae RadA in complex with ADP: hydrolysis-induced conformational change. Biochemistry (Mosc). 44: 13753-13761. Rehrauer WM, Kowalczykowski SC. 1993. Alteration of the nucleoside triphosphate (NTP) catalytic domain within Escherichia coli recA protein attenuates NTP hydrolysis but not joint molecule formation. J. Biol. Chem. 268: 1292-1297. Renodon-Corniere A, Takizawa Y, Conilleau S, Tran V, Iwai S, Kurumizaka H, Takahashi M. 2008. Structural analysis of the human Rad51 protein-DNA complex filament by tryptophan fluorescence scanning analysis: transmission of allosteric effects between ATP binding and DNA binding. J. Mol. Biol. 383: 575- 587. Reymer A, Frykholm K, Morimatsu K, Takahashi M, Norden B. 2009. Structure of human Rad51 protein filament from molecular modeling and site-specific linear dichroism spectroscopy. Proc. Natl. Acad. Sci. U. S. A. 106: 13248-13253. Richardson C, Stark JM, Ommundsen M, Jasin M. 2004. Rad51 overexpression promotes alternative double-strand break repair pathways and genome instability. Oncogene 23: 546-553. Ristic D, Modesti M, van der Heijden T, van Noort J, Dekker C, Kanaar R, Wyman C. 2005. Human Rad51 filaments on double- and single-stranded DNA: correlating regular and irregular forms with recombination function. Nucleic Acids Res 33: 3292-3302. San Filippo J, Sung P, Klein H. 2008. Mechanism of eukaryotic homologous recombination. Annu. Rev. Biochem. 77: 229-257. Schild D, Wiese C. 2010. Overexpression of RAD51 suppresses recombination defects: a possible mechanism to reverse genomic instability. Nucleic Acids Res 38: 1061- 1070. Shan Q, Cox MM, Inman RB. 1996. DNA strand exchange promoted by RecA K72R. Two reaction phases with different Mg2+ requirements. J. Biol. Chem. 271: 5712- 5724. Shim KS, Schmutte C, Yoder K, Fishel R. 2006. Defining the salt effect on human RAD51 activities. DNA Repair (Amst) 5: 718-730.

114

Sigurdsson S, Trujillo K, Song B, Stratton S, Sung P. 2001. Basis for avid homologous DNA strand exchange by human Rad51 and RPA. J. Biol. Chem. 276: 8798-8806. Sonoda E, Sasaki MS, Buerstedde JM, Bezzubova O, Shinohara A, Ogawa H, Takata M, Yamaguchi-Iwai Y, Takeda S. 1998. Rad51-deficient vertebrate cells accumulate chromosomal breaks prior to cell death. EMBO J. 17: 598-608. Sugawara N, Wang X, Haber JE. 2003. In vivo roles of Rad52, Rad54, and Rad55 proteins in Rad51-mediated recombination. Mol. Cell 12: 209-219. Tombline G, Fishel R. 2002. Biochemical characterization of the human RAD51 protein. I. ATP hydrolysis. J. Biol. Chem. 277: 14417-14425. Tombline G, Shim KS, Fishel R. 2002. Biochemical characterization of the human RAD51 protein. II. Adenosine nucleotide binding and competition. J. Biol. Chem. 277: 14426-14433. Tsuzuki T, Fujii Y, Sakumi K, Tominaga Y, Nakao K, Sekiguchi M, Matsushiro A, Yoshimura Y, MoritaT. 1996. Targeted disruption of the Rad51 gene leads to lethality in embryonic mice. Proc. Natl. Acad. Sci. U. S. A. 93: 6236-6240. Van Komen S, Petukhova G, Sigurdsson S, Sung P. 2002. Functional cross-talk among Rad51, Rad54, and replication protein A in heteroduplex DNA joint formation. J. Biol. Chem. 277: 43578-43587. Weinstock GM, McEntee K, Lehman IR. 1981a. Hydrolysis of nucleoside triphosphates catalyzed by the recA protein of Escherichia coli. Characterization of ATP hydrolysis. J. Biol. Chem. 256: 8829-8834. -. 1981b. Hydrolysis of nucleoside triphosphates catalyzed by the recA protein of Escherichia coli. Steady state kinetic analysis of ATP hydrolysis. J. Biol. Chem. 256: 8845-8849. Wu Y, He Y, Moya IA, Qian X, Luo Y. 2004. Crystal structure of archaeal recombinase RADA: a snapshot of its extended conformation. Mol. Cell 15: 423-435. Wu Y, Qian X, He Y, Moya IA, Luo Y. 2005. Crystal structure of an ATPase-active form of Rad51 homolog from Methanococcus voltae. Insights into potassium dependence. J. Biol. Chem. 280: 722-728. Yang H, Jeffrey PD, Miller J, Kinnucan E, Sun Y, Thoma NH, Zheng N, Chen PL, Lee WH, Pavletich NP. 2002. BRCA2 function in DNA binding and recombination from a BRCA2-DSS1-ssDNA structure. Science 297: 1837-1848. Yang H, Li Q, Fan J, Holloman WK, Pavletich NP. 2005. The BRCA2 homologue Brh2 nucleates RAD51 filament formation at a dsDNA-ssDNA junction. Nature 433: 653-657. Zhu Z, Chung W-H, Shim EY, Lee SE, Ira G. 2008. Sgs1 Helicase and Two Nucleases Dna2 and Exo1 Resect DNA Double-Strand Break Ends. Cell 134: 981-994.

115

CHAPTER 3

The RAD51 ATP Cap Regulates Nucleoprotein Filament Stability

116

3.1 ABSTRACT

RAD51 mediates homologous recombination (HR) by forming an active DNA nucleoprotein filament (NPF). A conserved aspartate that forms a salt bridge with the

ATP -phosphate is found at the nucleotide-binding interface between RAD51 subunits of the NPF known as the ATP cap. The salt bridge accounts for the nonphysiological cation(s) required to fully activate the RAD51 NPF. In contrast, RecA homologs and most RAD51 paralogs contain a conserved lysine at the analogous structural position.

We demonstrate that substitution of human RAD51(D316) with lysine

[HsRAD51(D316K)] decreases NPF turnover and facilitates considerably improved recombinase functions. Structural analysis shows that archaebacterial Methanococcus voltae RadA(D302K) [MvRAD51(D302K)] and HsRAD51(D316K) form extended active NPFs without salt. These studies suggest that the HsRAD51(D316) salt bridge may function as a conformational sensor that enhances turnover at the expense of recombinase activity. It predicts that at least some of the lysine-containing RAD51 paralogs may function to stabilize the RAD51 NPF.

117

3.2 INTRODUCTION

Failure to repair DNA double-stranded breaks (DSBs) may result in genomic instability and tumorigenesis (Bernstein and Rothstein 2009; Jackson and Bartek 2009; Ciccia and

Elledge). HR is an error free pathway that utilizes a template strand to restore DSB ends

(Krogh and Symington 2004). A key component in HR mediated repair is RAD51 (West

2003). Deletion of RAD51 in mice results in embryonic lethality, while RAD51 knock- down in chicken DT40 cell lines results in increased chromosomal instability (Tsuzuki et al. 1996; Sonoda et al. 1998). RAD51 forms a presynaptic NPF that catalyzes homologous pairing and strand-exchange. In addition, RAD51 associates with RAD52,

RAD54 and BRCA2 during recombinational repair (San Filippo et al. 2008). Eukaryotic

HR is further tuned by the presence of several RAD51 paralogs, where some have been shown to enhance human RAD51 (HsRAD51) functionality in vitro (Schild et al. 2000;

Masson et al. 2001; Shim et al. 2004).

Despite functional conservation with the prototypical bacterial homolog RecA,

RAD51 requires unusual salt conditions for an efficient strand exchange in vitro (Liu et

+ al. 2004; Shim et al. 2006). For example, the ammonium (NH4 ) cation appears to be most efficient at enhancing RAD51 recombinase activity, yet is unlikely to occur at physiologically relevant conditions (Liu et al. 2004; Shim et al. 2006). Other cations such as potassium (K+) confer significantly reduced activity (Liu et al. 2004; Shim et al.

2006). Cations that enhance recombinase activity appear to induce an NPF that mimics the active extended RecA NPF (Liu et al. 2004). Recombinase-enhancing cations may also promote preferential binding of single-stranded DNA (ssDNA) over double-stranded

118

DNA (dsDNA), which influences RAD51 ATPase activity (Tombline and Fishel 2002;

Liu et al. 2004; Shim et al. 2006). Structural analysis of MvRAD51 has revealed cation- induced protein conformational rearrangements at the inter-subunit region that results in an active NPF (Wu et al. 2004; Wu et al. 2005).

An evolutionarily intriguing feature of the RAD51/RecA NPF is the ATP cap located at the adenosine nucleotide-binding interface within the inter-subunit region (Wu et al. 2004; Wu et al. 2005). A proline residue that is conserved in all RecA/RAD51 homologs appears to sandwich the adenine nucleotide at the ATP binding interface (Fig.

3.1A). Five amino acids N-terminal to the conserved proline the RAD51 homologs possess a conserved aspartate residue that forms a salt bridge with the -phosphate of

ATP. In contrast, RecA homologs possess a conserved lysine at the analogous position that forms direct hydrogen bonds with the -phosphate (Fig. 3.1A; star). Importantly, four of the five HsRAD51 paralogs contain lysine at this corresponding position (Fig.

3.1A; star). The biophysical role of the HsRAD51 paralogs during HR is largely unknown.

HsRAD51 paralogs are difficult to purify in significant quantities for structural studies and tend to severely aggregate. Here we examined the structure and function of the conserved HsRAD51-paralog ATP cap lysine residue by constructing a substitution mutation where the conserved HsRAD51(D316) was replaced by lysine

[HsRAD51(D316K)]. We found that the HsRAD51(D316K) substitution enabled the uncomplicated formation and maintenance of an active NPF. As a functional consequence, HsRAD51(D316K) displayed reduced ATP hydrolysis (ATPase), enhanced discrimination of ssDNA versus dsDNA, and significantly enhanced recombinase

119

functions. Crystallographic and EM structural analysis indicate that MvRAD51(D302K) and HsRAD51(D316K) form a stable extended NPF in the absence of salt, that mimics salt-induced conformations of the wild type protein. Our results are consistent with the conclusion that the conserved aspartate in the ATP cap functions as a regulatory switch that enhances HsRAD51 NPF turnover, and predict that analogous lysine-containing

HsRAD51 paralogs may function to increase NPF stability. The enhanced stability and recombinase activity of hRAD51(D316K) in physiologically relevant conditions should provide a useful reagent for biochemical studies of HR.

120

3.3 MATERIALS AND METHODS

HsRAD51 Protein Expression and Purification- The HsRAD51(D316K) was constructed using PCR mutagenesis using primers 5- ATC TGC AAA ATC TAC AAA TCT CCC

TGT CT and its complement for mutagenesis (lysine encoding codon in bold), and for cloning into pET24d expression vector (Novagen) primers 5- TAT ACC ATG GCA

ATG CAG ATG CAG CTT GAA and 5- TTC GGA TCC TTA TCA GTC TTT GGC

ATC TCC CA were used that contain NcoI and BamHI restriction sites, respectively. For untagged native protein expression stop codons were introduced upstream of BamHI restriction site. Mutation was confirmed by DNA sequencing. HsRAD51wild type and

HsRAD51(D316K) proteins were expressed and purified following previously published protocols (Baumann et al. 1997; Tombline and Fishel 2002). Briefly, HsRAD51 was expressed in E. coli BLR strain and precipitated using spermidine-HCl. Resuspended pellet was purified using Reactive-Blue-4-agarose (Sigma), Heparin sepharose (GE

Healthcare), Hydroxyapatite (Bio-Rad) and Mono Q (GE Healthcare) column chromatography. Purity of the fractions was verified by SDS-PAGE analysis. HsRPA was expressed in BL21(AI) cells using pET11d-tRPA purified as described (Henricksen et al. 1994), except for resuspension of cells, HI buffer containing 30mM HEPES (pH

7.5), 1mM DTT, 0.25mM EDTA, 0.25% (w/v) inositol and 0.01% (v/v) NP-40 was supplemented with 100mM KCl.

121

DNA Substrates- X174 single-stranded (ss) virion DNA, replicative form I (RFI) was purchased from NEB. X174 RFIII was obtained by linearizing RFI with ApaLI restriction enzyme and gel purifying with Qiaquick Gel Extraction kit (Qiagen). For surface plasmon resonance (SPR) analysis, a 5′ biotinylated oligo dT50 was used as ssDNA and for dsDNA 5′ biotinylated 50-mer 5′-TCG AGA GGG TAA ACC ACA-

ATT ATT GAT ATA AAA TAG TTT TGG GTA GGC GA was annealed with its complement and purified by HPLC on a Gen-Pak FAX column (Waters). For competition DNA binding experiments an oligo dT50 ssDNA and a 50bp dsDNA were used made by annealing the 50-mer 5'- AGA TCT ATA AAC GCA CCT TTG GAA

GCT TGG AAG TGG GCC GAA TCT CCC CA with its complement followed by

HPLC purification as above. D-loop assay substrates were prepared as described (Van

Komen et al. 2002).

ATPase assay- ATP hydrolysis was measured as previously described (Tombline and

Fishel 2002). 10 L reactions were performed in Buffer A containing 20mM HEPES (pH

7.5), 10% glycerol, 100g/mL BSA, 1mM DTT and 1mM MgCl2 and when indicated,

150mM KCl. Each reaction mixture contained HsRAD51(1M) and 6M (nt or bp) of

X174 ssDNA or dsDNA. Reactions were initiated by the addition of protein and were incubated at 37°C for 1 hr. The reaction was stopped by the addition of 400L 10% activated charcoal supplemented with 10mM EDTA and incubated on ice for 2 hrs.

Following centrifuging for 10 min, 50L duplicate aliquots were taken for counting [32P]

122

free phosphate by Cerenkov method. Kinetic parameters were obtained by fitting data into Michaelis-Menten equation using Kaleidagraph software (Synergy).

ATPS/ADP binding assay- Experimental procedure was as previously described

(Tombline and Fishel 2002). Reactions were performed in Buffer A and 150mM KCl, when indicated. HsRAD51(1M) and X174 ssDNA (6M nt) were used with the indicated amount of [-35S]ATPS or [3H]ADP. Following a 30 min incubation at 37°C, reactions were kept on ice until filtered. The reaction mix was added into 4mL of ice cold reaction buffer and filtered through HAWP nitrocellulose filters (Millipore) presoaked in the same buffer. Another 4mL of buffer was used to wash the membrane and the filter was dried for 2 hrs. Radioactivity was counted after an overnight incubation in liquid scintillation fluid.

ADP-ATP exchange assay- Reactions were performed in Buffer A with 150mM KCl.

HsRAD51(1M) and X174 ssDNA (6M nt) was incubated with 3 M of [3H]ADP in a

60L reaction volume at 37°C for 10 min. ADPATP exchange was initiated by addition of cold ATP to a 5mM final concentration. 10L aliquots were withdrawn at indicated time points, filtered and analyzed as in ADP binding.

SPR analysis- Biotinylated DNA was immobilized on a streptavidin-coated chip (GE

Healthcare) for the analysis. Protein binding and dissociation were analyzed at 25°C with

123

a 5L/min flow rate on a Biacore 3000 (GE Healthcare). Reactions were performed in

Buffer containing 20mM HEPES (pH 7.5), 10% glycerol, 1mM DTT and 1mM MgCl2 supplemented, 0.005% surfactant P-20 (GE Healthcare), 2.5mM of the indicated

2+ nucleotide and 150mM KCl when indicated. For experiments with Ca , 1mM CaCl2 was used instead of MgCl2. DNA binding was determined by injecting 100nM, 200nM,

400nM, 800nM and 1.6M of protein simultaneously into ssDNA and dsDNA containing flow channels to ensure saturated binding.

SPR data was performed at five different concentrations for each mutant under each set of conditions, and fit to a single exponential decay in order to obtain rate constants for dissociation (koff). The decay curve, ƒoff(t), is given by,

-koff t ƒoff(t) = A(C) e

Where A(C) is the amplitude, which depends on the concentration of HsRAD51; C, and t is the time. In some cases, binding is insufficient and/or dissociation occurs very rapidly, and cannot be accurately determined by SPR. For these cases we simply report not determinable (ND) for the value.

Gel mobility shift assay - HsRAD51 wild type or HsRAD51(D316K) at indicated

32 concentrations, was incubated with P- labeled oligo dT50 (81 nM nt) in buffer containing 20mM HEPES (pH 7.5), 10% Glycerol, 1mM DTT, 1mM MgCl2 supplemented with 2.5mM of ATP and 150mM KCl at 37°C for 15min. Samples were then kept on ice until resolved by 5% nondenaturing PAGE at 10V/cm in 0.3x TBE

124

buffer at 4°C. Gels were dried and exposed to PhosphoImager (Molecular Dynamics) screens for imaging.

Competition DNA binding analysis- HsRAD51 wild type or HsRAD51(D316K) (730 nM) was incubated with 32P- labeled oligo dT50 (81 nM nt) and the indicated amounts (in bp) of 50bp cold dsDNA in buffer containing 20mM HEPES (pH 7.5), 10% Glycerol,

1mM DTT, 1mM MgCl2 supplemented with 2.5mM of ATP and when indicated, 150mM

KCl, at 37°C for 15min. Samples were kept on ice until resolved by 5% nondenaturing

PAGE at 10V/cm in TBE buffer at 4°C. Dried gels were exposed to PhosphoImager

(Molecular Dynamics) screens for quantification using ImageQuant software (GE

Healthcare).

D-loop assay- Labeled 90-mer (2.4 M nt) was incubated with HsRAD51(0.8M) in buffer A with 1mM ATP supplemented with the indicated amounts of MgCl2 or CaCl2 at

37°C for 10 min. Reaction was initiated by adding supercoiled pBS-SK(-) DNA (35M bp) and incubated further for 15 min. Samples were deproteinized by addition of 1%

SDS and 1 mg/mL proteinase K to a final concentration and incubated for 15 min at

37°C, mixed with 1/5 volume of gel loading dye and analyzed on 0.9% agarose gel in

TAE buffer, run at 4V/cm at 25°C. D-loops were quantified using ImageQuant software

(Molecular Dynamics) after drying and exposure to PhosphoImager screens.

DNA Strand Exchange Assay- HsRAD51 (5M) and X174 circular ssDNA (30M nt) were pre-incubated in buffer containing 20mM HEPES (pH 7.5), 10% Glycerol, 1mM 125

DTT, 1mM MgCl2 supplemented with 2.5mM of ATP at 37°C for 5 minutes before addition of 150mM of the indicated salt and linear X174 dsDNA (15M bp). Following a additional 5 min incubation HsRPA (2M) was added and the incubation was continued. After 3 hrs samples were deproteinized by addition of 3L of stop buffer containing 2% SDS and 10 mg/mL proteinase K, and analyzed on 0.9% agarose gel in

TAE buffer. Electrophoresis was carried out at 4V/cm at 25°C with 0.1mg/mL ethidium bromide. Gels were analyzed on a gel documentation station (Bio-Rad). For quantification of joint molecules ImageJ software (http://rsbweb.nih.gov/ij) was utilized.

For strand exchange reaction in RecA format, X174 circular ssDNA (5M nt) and linear X174 dsDNA (5M bp) was preincubated with HsRAD51(5M) in buffer containing 20mM HEPES (pH 7.5), 10% Glycerol, 1mM DTT, 1mM MgCl2 supplemented with 2.5mM of ATP and 150mM of the indicated salt (when added) at

37°C for 5 min. After 5 min, HsRPA (2M) was added and analyzed as described above.

Electron microscopy- The typical DNA binding reaction for electron microscopy was performed in 20 μl strand exchange buffer containing either no salt, 100 mM ammonium sulfate or 200 mM potassium chloride, a 5 ng/μl concentration of the substrate DNA

(either ss or ds M13 DNA) and HsRad51 proteins at a 3:1 (nt or bp: protein) molar ratio.

The reactions were carried out at 37C for 15 min. For negative staining, an aliquot of the sample was diluted 1:4 in buffer and immediately adsorbed to glow-charged thin carbon foils for 3 min followed by staining with 2% unbuffered uranyl acetate for 5 min and followed by air drying. For direct mounting, the remaining sample volume was passed

126

over 2-ml columns of 6% agarose beads (ABT inc, Burgos Spain) equilibrated with TE buffer (10 mM Tris-HCl (pH 7.4), 0.1 mM EDTA-NaOH) to eliminate unwanted salts and unbound proteins, and fractions enriched for DNA-protein complexes were collected.

Aliquots of the fractions of interest were mixed with a buffer containing spermidine and adsorbed onto copper grids coated with a thin carbon film glow-charged shortly before sample application. After adsorption of the samples for 2–3 min, the grids were dehydrated through a graded ethanol series and rotary shadowcast with tungsten at 10−7 torr. Samples were examined in an FEI T12 TEM and a Philips CM12 TEM equipped with Gatan 2kx2k SC200 CCD cameras at 40 kV (shadowcast samples) or at 80 kV

(stained samples). Dimensions of particles in the images saved from the CCD cameras were analyzed using Digital Micrograph software (Gatan, Inc.). Adobe Photoshop software was used to arrange images into panels for publication.

MvRAD51 protein preparation and crystallization- The recombinant protein was over- expressed in BL21 Rosetta2 (DE3) cells (Novagen). In addition to the D302K mutation, this recombinant protein also carried a N4G mutation and lacked the first three amino acid residues. The soluble protein was purified as reported for the wild-type MvRAD51 protein (Wu et al. 2004; Qian et al. 2006b). In brief, the purification procedure involved steps of polymin P (Sigma) precipitation, high salt extraction and three chromatography steps using heparin (Amersham Biosciences), hydroxyapatite (BioRad) and Sephacryl S-

300 gel filtration (GE Healthcare) columns. The purified MvRAD51(D302K) protein was concentrated to ~30 mg/ml by ultra-filtration.

127

The hexagonal MvRAD51(D302K) crystals (P61 space group) were grown by the hanging drop vapor diffusion method. The maximum dimension was 0.2 mm x 0.2 mm x

0.3 mm. The sample contained ~1 mM of concentrated protein and 2 mM of AMP-PNP.

The reservoir solutions contained 4-5% polyethylene glycol (PEG) 3,350, 50 mM MgCl2,

50 mM Tris-HCl buffer at pH 7.9, 0.4 M NaCl and 14% (w/v) sucrose. One crystal was transferred into a stabilizing solution composed of the reservoir solution supplemented with 28% (w/v) sucrose, looped out of the solution, and frozen in a nitrogen cryo-stream at 100 K. The diffraction data set was acquired at the Canadian Light Source beamline

08ID-1 and was processed using the XDS program (Kabsch 1993). The statistics of the diffraction data is listed in Supplementary Table 2.

The previously solved wild-type MvRAD51 model (PDB code 1T4G) was used as the starting model. After initial refinement, the electron density of the side chain of

Lys-302 was clearly visible in the resulting 2Fo – Fc map. The model was then iteratively rebuilt using XtalView (McRee 1999) and refined using CNS. The molecular figures were generated using Pymol (DeLano 2002). The coordinates and structure factors have been deposited in the (PDB code 3NTU).

128

3.4 RESULTS

HsRAD51(D316K) substitution affects ATPase but not adenosine nucleotide binding or exchange. Comparison of RecA and RAD51 homologs identified a number of conserved residues (Fig. 3.1A). The HsRAD51(D316) residue was examined to test the effect of lysine in the ATP cap that is conserved in RecA and most HsRAD51 paralogs (Fig. 3.1A, red star). We purified the HsRAD51(D316K) protein (Fig. 3.2A) and examined the steady state ATPase activity (Fig. 3.1D; Table 3.1). We found that the ATPase rate was approximately 3-fold lower than the wild type protein. Our results are also consistent with previous studies that have suggested stimulatory cations may modestly enhance the discrimination between ssDNA- and dsDNA-stimulated ATPase (Fig. 3.1E, compare ssDNA to dsDNA ATPase with and without K+; (Liu et al. 2004; Shim et al. 2006).

However, we observe a reduced cation-induced difference in ssDNA and dsDNA stimulated ATPase with HsRAD51(D316K) compared to the wild type protein (Fig.

3.1E).

The difference in HsRAD51(D316K) ATPase compared to the wild type

HsRAD51 was not the result of altered ATP binding activity (Fig. 3.1F; Table 3.1), and there appeared to be no defect associated with ADPATP exchange, although kinetic differences at time points shorter than 30 sec would not have been detected (Fig. 3.2B).

We observed an approximately 4-fold increase in KD for ADP in the presence of ssDNA compared to the wild type HsRAD51 (Fig. 3.1G, Table 3.1). Importantly, the protein preparations appeared fully active since the ATP and ADP maximum binding (Bmax) as well as 1:1 stoichiometry appeared equivalent under a variety of conditions (Fig. 3.1F

129

and 3.1G; Table 3.1). These observations suggest that defects in adenosine nucleotide binding and exchange are unlikely to fully account for the reduced HsRAD51(D316K)

ATPase activity.

HsRAD51(D316K) exhibits RPA and salt independent strand exchange activity.

Replication protein A (RPA) facilitates RAD51 mediated strand exchange at both presynaptic and postsynaptic stages (Sugiyama et al. 1997; Eggler et al. 2002; Van

+ Komen et al. 2002). Efficient strand exchange is facilitated by the NH4 cation, which is a characteristic of RAD51 even though such conditions are unlikely to be physiologically relevant (Sigurdsson et al. 2001a; Liu et al. 2004; Shim et al. 2006). We compared the strand exchange activity of wild type HsRAD51 and HsRAD51(D316K) in the presence

+ of HsRPA and NH4 cation (Fig. 3.3). Remarkably, HsRAD51(D316K) appeared significantly more active than wild type HsRAD51 and was capable of catalyzing efficient strand exchange in the absence of HsRPA (Fig. 3.3A). Moreover, we still

+ observed strand exchange products in the absence of both HsRPA and NH4 cation (Fig.

3.3A). These studies suggest that HsRAD51(D316K) may circumvent both the cation and ssDNA binding (SSB) protein requirements of wild type HsRAD51.

We examined the cation requirements of HsRA51(D316K) and observed efficient strand exchange activity in the presence of sodium (Na+) and K+ cations (Fig. 3.3B).

Moreover, peak strand exchange activity with HsRAD51(D316K) appears to occur with

+ + approximately half the NH4 or K cation concentration compared to the wild type

HsRAD51 (Fig. 3.3C).

130

The enhanced strand exchange activity and reduced SSB protein requirement of

HsRAD51(D316K) appeared similar to bacterial RecA. The order-of-addition requirements for efficient RecA strand exchange are different from RAD51, the latter of which requires a series of steps that include the pre-formation of an NPF on the ssDNA substrate prior to the addition of salt and then the dsDNA donor. In contrast, all of the reaction components including ssDNA and dsDNA substrates may be included with

RecA, and there is no unusual cation requirement. Remarkably, HsRAD51(D316K) appears to efficiently catalyze strand exchange under RecA order-of-addition in the

+ + presence of K and significantly less efficiently in the presence of NH4 cation (Fig.

3.4A). These studies suggest that HsRAD51(D316K) may catalyze fundamental recombination reactions under more physiologically relevant conditions.

A gel based competition analysis was used to examine the relative binding of

HsRAD51 to ssDNA and dsDNA (Fig. 3.4B and 3.4C). In this system HsRAD51 induces a ssDNA NPF gel shift that can be used in a competition analysis with cold ssDNA or dsDNA competitor (Fig. 3.5A-C). As previously suggested for wild type

HsRAD51, salt (KCl) dramatically decreases the ability of cold dsDNA to compete for labeled ssDNA binding while modestly affecting the ability of cold ssDNA to compete for labeled ssDNA binding (Fig. 3.4B, compare WT and WT-KCl; (Shim et al. 2006). In contrast, the HsRAD51(D316K) ssDNA NPF is largely stable to cold dsDNA competition regardless of salt (KCl) while the ssDNA binding appears further enhanced to cold ssDNA competition in the presence of salt (KCl). These results strongly suggest that the HsRAD51(D316K) naturally displays a substantial discrimination between ssDNA and dsDNA that the wild type HsRAD51 only displays in the presence of salt.

131

We also note that the HsRAD51(D316K) mutant protein is unable to catalyze strand exchange in the presence of ATPS, yet performs modest strand exchange in the presence of ADP (Fig. 3.6). Since ATPS binding by wild type HsRAD51 and

HsRAD51(D316K) appear identical, it is likely that this poorly hydrolysable ATP analog induces a conformation that is incompatible with strand exchange activity. Together, these results are consistent with the conclusion that the HsRAD51(D316K) substitution significantly alters the salt requirement for triggering an appropriate adenosine nucleotide-dependent conformational transition necessary for strand exchange (see Fig.

3.1C).

HsRAD51(D316K) affects protein turnover.

RAD51 DNA interaction(s) during pre- and postsynaptic stages is believed to play an important role during HR (Solinger et al. 2002; Chi et al. 2006; Symington and Heyer

2006). We quantitatively examined the DNA binding and dissociation properties of the

HsRAD51 substitution mutant proteins using surface plasmon resonance (SPR) and biotin-streptavidin immobilized dT50 ssDNA or 50bp dsDNA. These studies were performed in physiologically relevant salt (150 mM KCl) with magnesium and ATP similar to strand exchange conditions (see Fig. 3.3). The protein was titrated between

100 nM to 1.6M to examine saturated binding on both ssDNA and dsDNA. After protein injection and NPF formation the sample was analyzed for steady-state protein disassembly (Fig. 3.7).

Previous studies have suggested a multimeric RAD51 binding that would significantly alter the simple concentration dependence of kon (Hilario et al. 2009). A 132

power-dependence of concentration on kon would make any calculations of KD problematic. In contrast the koff behaves as a simple single exponential decay and appears highly accurate for a variety of conditions.

The saturation and shape of the binding curves varied considerably for the wild type HsRAD51 as predicted by previous studies. (Shim et al. 2006). Two general observations emerged: 1.) HsRAD51 does not bind dsDNA in the presence of ADP, and

2.) the HsRAD51 koff increases 2-4 fold in the presence of adenosine nucleotide with the koff•ADP > koff•ATP (Fig. 3.7 and Table 3.2). These results are consistent with previous studies that suggested increased RecA and RAD51 turnover upon ATP hydrolysis

(Kowalczykowski 1991; Bianco et al. 1998; Bugreev and Mazin 2004; Bugreev et al.

2005; Chi et al. 2006).

In contrast, the HsRAD51(D316K) is completely deficient in both ssDNA and dsDNA binding in the absence of adenosine nucleotide (Fig. 3.7). This observation appears to largely mimic the properties of bacterial RecA and yeast Rad51 (Shibata et al.

1979; Pugh and Cox 1987; De Zutter and Knight 1999; Zaitseva et al. 1999). In contrast,

HsRAD51(D316K) binds strongly to ssDNA and dsDNA in the presence of both ADP or

ATP (Fig. 3.7 and Table 3.2). Moreover, the stability of the NPF is increased 7-10 fold

-1 compared to the wild type HsRAD51 in the presence of ATP (koff•ATP•ssDNA = 0.0005 s ;

-1 koff•ATP•dsDNA = 0.0006 s ; Table 3.2). Prolonged analysis confirmed the stability of the

NPF (data not shown). The decreased koff as well as the ability to bind dsDNA in the presence of ADP suggests that the HsRAD51(D316K) mutant protein displays a significantly reduced adenosine nucleotide-induced DNA turnover compared to wild type

HsRAD51.

133

HsRAD51(D316K) catalyzes efficient D-loop formation independent of divalent cation.

RAD51 catalyzes the formation of D-loop joint molecules in the presence of calcium

(Ca++) or nonhydrolyzable ATP analogues (Bugreev et al. 2005; Chi et al. 2006).

Moreover, the P-loop substitution mutation of HsRAD51(K133R) that binds ATP but renders the protein ATPase deficient was found to catalyze efficient D-loop formation in the presence of magnesium (Mg++; (Chi et al. 2006). Together these results are consistent with the conclusion that the catalysis of D-loop joint molecules by RAD51 requires the formation and maintenance of an ATP-bound NPF, which can be enhanced by substitution of Ca++ for Mg++ or by using a hydrolysis-deficient HsRAD51 mutation. The intrinsic stability of the HsRA51(D316K) NPF suggested it might be more efficient at catalyzing D-loop formation.

As expected, we found that HsRAD51 catalyzes the formation of D-loops in the absence of salt, but requires Ca++ to induce an active ATP-bound NPF (Bugreev and

Mazin 2004). The yield of D-loops by wild type HsRAD51 in the presence of Mg++ divalent cation was 8.1 fmol (3%) , and 81 fmol (30%) in the presence of Ca++ (Fig. 3.8B, lanes 3 and 4; Fig. 3.8C). Importantly, HsRAD51(D316K) catalyzed the formation of D- loops in the absence of any divalent cation or in the presence of either Mg++ or Ca++ (107 fmol (38%) and 94 fmol (35%) respectively; Fig. 3.8B, lanes 5 and 6; Figs. 3.8C and

3.8D). Kinetic analysis revealed an increase in the initial rate of D-loop formation by

HsRAD51(D316K) (data not shown). D-loop catalysis by HsRAD51(D316K) in the presence of the nonhydrolyzable nucleotide analogue adenosine 5′-(β,γ- imido)triphosphate (AMP-PNP) appeared similar to the ATP-driven catalysis regardless

134

of divalent cation cofactor (Fig. 3.8D). Moreover, in the presence of Ca++ the

HsRAD51(D316K) mutant protein appeared capable of efficiently catalyzing D-loop formation with ADP (Fig. 3.8D).

We examined the stability of the wild type HsRAD51 and HsRAD51(D316K) presynaptic NPF on ssDNA in the presence of ATP and either Mg++ or Ca++ (Fig. 3.8E and 3.8F). A significant decrease in koff was observed for wild type HsRAD51 in the presence of Ca++ as well as for HsRAD51(D316K) in the presence of both Ca++ and Mg++

(Fig. 3.8E and 3.8F). These results strongly suggest that HsRAD51(D316K) inherently forms and maintains an active ATP-bound NPF regardless of associated divalent cations.

Collectively these results underline previous conclusions that the ability to form and maintain an active adenosine nucleotide bound NPF conformation drives D-loop formation (Shibata et al. 1979; Cox and Lehman 1981; Bugreev and Mazin 2004).

MvRAD51(D302K) induces and active nucleoprotein filament.

We determined the structure of the MvRAD51(D302K) at 1.9Å resolution (Table 3.3).

The MvRAD51(D302K) protein forms a filament with a helical pitch of 104.3Å, similar to the helical pitch of active wild type MvRadA (104.8 Å; Fig. 3.9A). The L2 single strand binding region (D256 to R285) of MvRAD51(D302K) is largely ordered. The non-hydrolysable ATP analogue, AMP-PNP, is buried between RadA monomers as expected (Fig. 3.9B). The adenosine nucleotide moiety is sandwiched between the ATP cap (residues K302 to D308) and the side chain of R158. The amino group of the K302 side chain forms a total of four hydrogen bonds: one with the terminal phosphate of the

ATP analogue and three with the carbonyls of residues G279, H280 and A282. These

135

four groups stabilized a Ca++ cation in a previously reported calcium-containing active form of MvRAD51 (Qian et al. 2006a). The K302 appears to provide further stabilization of the -phosphate in addition to the well-known P-loop residues and Mg++ cation (Wu et al. 2004). Importantly, the K302 side chain replaces both K+ cations including one that normally bridges the wild type D302 with the -phosphate of ATP and additionally coordinates an ordered helix (residues G275 to A282) analogous to helix G of EcRecA (Story and Steitz 1992) (Fig. 3.9C). The H280 side chain makes a direct hydrogen bond with the -phosphate of the ATP analogue and F107 is flipped away from the nucleotide binding pocket as is required to form and active MvRAD51 NPF.

Interestingly, the K248 of the E. coli RecA active nucleoprotein filament is at the analogous position of K302 (Fig. 3.9D)(Chen et al. 2008). The structure of

MvRAD51(D302K) closely resembles the recurrent active form of M. voltae and M. maripaludis RAD51-AMP-PNP complexes (Qian et al. 2005; Wu et al. 2005; Li et al.

2009), but differs from the recurrent inactive form of MvRAD51 (Wu et al. 2004) which has a largely disordered L2 region (compare Fig. 3.9C with Figs. 3.1B and 3.1C).

HsRAD51(D316K) maintains an active nucleoprotein filament in the absence of salt.

To confirm the structural analysis of MvRAD51(D302K) we examined the wild type

HsRAD51 and HsRAD51(D316K) NPFs by electron microscopy (EM) (Fig. 3.9E and

3.9F). These studies were performed in strand exchange conditions with ATP in the

+ + absence of salt or in the presence of equi-normal K or NH4 cations. The wild type

HsRAD51 protein exhibited a 30% increase in the helical pitch with visible striation of

+ + the filament in the presence of K or NH4 (Fig. 3.9E). In contrast, HsRAD51(D316K) 136

maintains an inherently extended conformation (helical pitch of 100-112Å) similar to the

MvRadA(D302K) filament and the salt-induced wild type HsRAD51 filament under all conditions (Fig. 3.9F). These studies are consistent with the biochemical and structural analyses that suggest HsRAD51(D316K) forms and maintains an active adenosine nucleotide bound NPF under a wide range of conditions.

137

3.4 DISCUSSION

The ATP cap of RecA/RAD51 homologs is a structurally conserved region that overlays the ATP binding site of an adjacent protomer within an NPF. It contains a highly conserved proline residue that functions to sandwich the adenine moiety (Fig. 3.1A).

RAD51 homologs possess a conserved aspartate residue adjacent to the conserved proline in the ATP cap, while RecA homologs and a large majority of RAD51 paralogs have a lysine at the analogous position (Fig. 3.1A). Both the aspartate and lysine residues ultimately contact the -phosphate of ATP (Wu et al. 2005; Chen et al. 2008). The lysine residue does this directly, while the cation of the aspartate salt bridge mediates the - phosphate contact. Conformational transition(s) of the MvRAD51 F107 and H280 residues have been shown to be associated with the development of the aspartate salt bridge, which together appear necessary to form an active NPF (Wu et al. 2005). The available data suggest that the cation which occupies this salt bridge must have a sufficient ionic radius to provoke the formation and maintenance of an active ATP-bound

NPF (Shim et al. 2006). In its absence or under conditions in which ATP hydrolysis may readily occur, the RAD51 NPF appears largely inactive. Taken together, these results suggest that the aspartate salt bridge plays a significant role in the management and regulation of RAD51 recombinase functions.

In contrast, RecA homologs and RAD51 containing a lysine residue in the analogous position appear to more easily form and maintain an active NPF that translates to enhanced recombinase function(s) (Chen et al. 2008). The majority of RAD51 paralogs contain a lysine residue in this analogous position (Fig. 3.1A). The human

138

RAD51 paralogs have been purified in a variety of heteromeric forms (Masson et al.

2001; Sigurdsson et al. 2001b; Shim et al. 2004). However, the quantities of these heteromeric forms and their tendency to aggregate have thus far inhibited a complete examination their biophysical and structural role in HR. The HsRAD51(D316K) substitution mutation has allowed us to examine the effect of a RAD51 protein containing an ATP cap lysine on recombinase functions.

A prediction of this analysis is that RAD51 paralogs containing an ATP cap lysine may function to stabilize the RAD51 filament. Such a Rad51 NPF stabilization has recently been demonstrated for the yeast paralogs Rad55-Rad57 (Liu et al.). It is possible that near stoichiometric quantities of these RAD51 paralogs might be required to induce such stability since they would have to occupy a large fraction of the NPF. Of the human RAD51 paralogs, only HsRAD51B and HsRAD51C appear to be expressed in quantities equivalent or in excess of HsRAD51 (Miller et al. 2002). Interestingly, the

HsRAD51B-HsRAD51C heterodimer appears prevalent in human cells and both contain an ATP cap lysine (Miller et al. 2002). Moreover, both HsRAD51B and HsRAD51C display modest binding to ssDNA and HsRAD51B contains the conserved proline in the

ATP cap region (Lio et al. 2003).

The mechanism of ATP cap regulation. The mechanism associated with ATP cap regulation of RAD51 might be inferred from the MvRAD51 structure (Wu et al. 2005).

In the MvRAD51 K+ cation-induced active NPF structure, the H280 residue formed a hydrogen bond with the -phosphate of the ATP analogue, while the F107 residue was moved out of the ATP binding interface to accommodate two K+ cations (one forming the

139

D302 salt bridge). A similarly active MvRAD51 contains an offset calcium that occupies the space of both K+ cations but still forms a salt bridge with the -phosphate (Qian et al.

2006a). Together these observations suggest that activation of MvRAD51 requires salt occupation of the ATP cap and that cations, which disrupt the conformational changes associated with the ATP hydrolytic cycle further enhance activity. The

MvRAD51(K302) and HSRAD51(K316) residues appears to perform these functions naturally, but without the ability to perform post-synaptic turnover.

RPA independence. It has been suggested that RPA enhances strand exchange activity by disrupting ssDNA secondary structure formation and by stabilizing the newly formed joint molecule by binding to the newly released ssDNA (Sugiyama et al. 1997; Mazin and Kowalczykowski 1998). Since HsRAD51(D316K) demonstrates RPA independence in catalyzing strand exchange, it is possible that this protein possesses an enhanced ability to disrupt secondary structures on ssDNA and/or it binds the ssDNA product more efficiently from the triplex joint molecule. Indeed, HsRAD51(D316K) binds ssDNA

~15-fold better that the wild type HsRAD51 in the presence of ATP and ~3-fold better than wild type HsRAD51 in the presence of ADP.

Evolution of aspartate in the ATP cap. An intriguing evolutionary question is why

RAD51 recombinases have selected an aspartate residue in the ATP cap region instead of a lysine found in bacterial RecA? One possibility is that the preservation of the aspartate residue and associated salt bridge fundamentally provides the turnover required during

HR. Perhaps the ability to efficiently recycle HsRAD51 from an expanding presynaptic

140

filament during the construction of an active NPF or the release of HsRAD51 from a completed recombination product is more important in eukaryotic genomes? In this scenario, the aspartate residue and associated salt bridge would function as a conformational sensor in delineating active and inactive protomers in the NPF before and/or after HR. Finally, the development of the HsRAD51(D316K) should significantly enhance biochemical studies of HR when physiologically relevant conditions are required.

141

3.5 ACKNOWLEDGEMENTS AND FOOTNOTES

Crystallographic work was performed by Yujiong He and Yu Luo at the University of

Saskatchewan, Canada. EM studies were performed by Smaranda Willcox and Jack

Griffith at the University of North Carolina. Model fitting of DNA binding data was performed by Robert Forties and Ralf Bundschuh. This chapter was submitted for publication as: The RAD51 ATP Cap Regulates Nucleoprotein Filament Stability,

Amunugama R., He Y. , Willcox S., Forties R.A., Shim K-S., Bundschuh R., Yu Luo,

Griffith J., and Fishel R.

142

Figure 3.1 ATP Binding and Hydrolysis by HsRAD51 ATP Cap Substitution Mutation. (A) Sequence alignment of the ATP cap region of H. sapiens (Hs), S. cerevisiae (Sc), M. voltae (Mv), and E. coli (Ec) recombinases. Sequence alignment of the ATP cap region of HsRAD51 paralogs is indicated below. The analogous position of HsRAD51(D302) is shown with an asterisk. (B) ATP cap region of MvRadA (MvRAD51) structure in the absence of potassium cation (PDB code 1T4G). (C) ATP cap region of MvRAD51 structure in the presence of potassium cation (PDB code 1XU4). Structural figures were generated using Pymol. (D) Steady-state ATPase activity of HsRAD51 wild type or HsRAD51(D316K) with ssDNA in the presence of 150 mM KCl. (E) ATP turnover values (kcat) with ssDNA and dsDNA in the presence and + absence of KCl (K ). kcat values were calculated by Michaelis-Menten analysis. (F) ATPS binding by HsRAD51 wild type or HsRAD51(D316K) in the presence of ssDNA and 150 mM KCl. (G) ADP binding by HsRAD51 wild type or HsRAD51(D316K) in the presence of ssDNA and 150 mM KCl. Error bars indicate standard deviation from at least three independent experiments.

Continued

143

Figure. 3.1: continued

144

Figure 3.2 Purification of HsRAD51 and ADP-ATP exchange analysis (A) Purification of HsRAD51 WT and HsRAD51(D316K) substitution mutant protein. 1 g of protein analyzed by 12% SDS-PAGE. (B) ADP-ATP exchange HsRAD51 WT and HsRAD51(D316K) mutant protein in the presence of ssDNA and 150 mM KCl.

145

Figure 3.3 HsRAD51(D316K) Exhibits Salt and RPA Independent Strand Exchange Activity. (A) Analysis of salt and RPA requirement for strand exchange. Reaction schematic shown above: HsRAD51 wild type or HsRAD51(D316K) (5M) and X174 circular ssDNA (30M nt) were pre-incubated with 2.5mM ATP and 1mM MgCl2 at

37°C for 5 min prior to the addition of 150mM NaNH4HPO4 (if indicated) and linear X174 dsDNA (15M bp). After 5 min, HsRPA (2M) was added (if indicated) and the incubation was continued for 3 hrs. Samples were deproteinized and analyzed on 0.9% agarose gel with 0.1g/mL ethidium bromide. (B) Strand exchange activity of HsRAD51 and HsRAD51(D316K) as a function of cation normality. HsRAD51 (5M) and X174 circular ssDNA (30M nt) were pre-incubated with 2.5mM ATP and 1mM MgCl2 at 37°C for 5 min before addition of indicated amounts of salt and linear X174 dsDNA (15M bp). After another 5 min incubation HsRPA (2M) was added and the incubation was continued. After 3 hrs samples were deproteinized and analyzed on 0.9% agarose gel with 0.1g/mL ethidium bromide. (C) Quantification of the joint molecules in B.

Continued

146

Figure 3.3: continued

147

Figure 3.4 HsRAD51(D316K) Preferentially Binds to Single Stranded DNA. (A) Strand Exchange in RecA format. X174 circular ssDNA (5M nt) and linear X174 dsDNA (5M bp) was preincubated with HsRAD51 wild type or HsRAD51(D316K) (5μM) in buffer containing 20mM HEPES (pH 7.5), 10% Glycerol, 1mM DTT, 1mM

MgCl2 supplemented with 2.5mM of ATP and 150mM KCl at 37°C for 5 min. After 5 min, HsRPA (2M) was added and the incubation was continued. After 3 hrs samples were deproteinized and analyzed on 0.9% agarose gel with 0.1μg/mL ethidium bromide. (B) Quantification of competition DNA binding analysis (see Supp.Fig. 2) of HsRAD51 wild type and HsRAD51(D316K). HsRAD51 wild type or HsRAD51(D316K) (730 nM) 32 was incubated with P- labeled oligo dT50 (81 nM nt) and the indicated amounts of 50bp cold competitor dsDNA in buffer containing 20mM HEPES (pH 7.5), 10% Glycerol,

1mM DTT, 1mM MgCl2 supplemented with 2.5mM of ATP and when indicated, 150mM KCl, at 37°C for 15min. Samples were kept on ice until resolved by 5% nondenaturing PAGE at 10V/cm in TBE buffer at 4°C. Dried gels were exposed to PhosphoImager screens for quantification. (C) Competition DNA binding with oligo dT50 cold competitor DNA. Reaction conditions are similar to B.

Continued

148

Figure 3.4: continued

149

Figure 3.5 Competition DNA binding by HsRAD51 and HsRAD51(D316K). (A) HsRAD51 wild type or HsRAD51(D316K) at indicated concentrations, was incubated 32 with P- labeled oligo dT50 (81 nM) in buffer containing 20mM HEPES (pH 7.5), 10%

Glycerol, 1mM DTT, 1mM MgCl2 supplemented with 2.5mM of ATP and 150mM KCl at 37°C for 15min. Samples were then kept on ice until resolved by 5% nondenaturing PAGE at 10V/cm in TBE buffer at 4°C. A distinct ssDNA-HsRAD51(D316K) complex formed is indicated with an asterisk. (B) HsRAD51 wild type or HsRAD51(D316K) (730 32 nM) was incubated with P- labeled oligo dT50 (81 nM) and the indicated amounts of 50bp cold competitor dsDNA in buffer containing 20mM HEPES (pH 7.5), 10%

Glycerol, 1mM DTT, 1mM MgCl2 supplemented with 2.5mM of ATP and when indicated, 150mM KCl, at 37°C for 15min. Samples were kept on ice until resolved by 5% nondenaturing PAGE at 10V/cm in TBE buffer at 4°C. Dried gels were exposed to PhosphoImager screens for quantification. (C) Competition DNA binding with oligo dT50 cold competitor DNA. Reaction conditions are similar to B.

Continued

150

Figure 3.5: continued

151

Figure 3.6 Strand Exchange Activity of HsRAD51 and HsRAD51(D316K) with Different Adenosine Nucleotides. HsRAD51 (5μM) and X174 circular ssDNA (30μM nt) were pre-incubated in buffer containing 20mM HEPES (pH 7.5), 10% Glycerol, 1mM

DTT, 1mM MgCl2 with 2.5mM of indicated adenosine nucleotide at 37°C for 5min before addition of 150mM NaNH4HPO4 and linear X174 dsDNA (15μM bp). After another 5 min incubation HsRPA (2μM) was added and the incubation was continued. After 3 hrs samples were deproteinized and analyzed on 0.9% agarose gel with 0.1mg/mL ethidium bromide.

152

Figure 3.7 HsRAD51(D316K) Displays a Slow Turnover from ssDNA and dsDNA. (A) ssDNA binding analysis of wild type and HsRAD51(D316K) by surface Plasmon resonance (SPR, Biacore) in the presence of the indicated nucleotide. Association and dissociation curves correspond to 100nM, 200nM, 400nM, 800nM and 1.6M of protein. Fitted cures are overlaid in black lines. (B) dsDNA binding analysis of wild type and HsRAD51(D316K) in the presence of the indicated nucleotide. Protein concentrations are similar to A.

Continued

153

Figure 3.7: continued

154

Figure 3.8 HsRAD51(D316K) Catalyzes D-loop Formation in the Presence of ATP and Magnesium. (A) Reaction schematic. (B) 0.8M of HsRAD51 wild type or HsRAD51(D316K) and [P32]-labeled ssDNA (90mer; 2.4 M nt) were preincubated for

10 min at 37°C in the reaction buffer containing 1 mM ATP and 1mM MgCl2 or CaCl2. Reactions were initiated by the addition of supercoiled pBS SK(-) plasmid DNA (35M bp). After 15 min reactions were terminated by the addition of deproteinization solution and processes similar to Fig. 2. Joint molecules (JMs) were analyzed on a 0.9% agarose gel. (C) Quantification of JMs in B. Error bars indicate standard deviation from at least three independent experiments. (D) D-loop formation analysis with different nucleotides. 0.8M of HsRAD51 wild type or HsRAD51(D316K) and 2.4 M (nt) of P32- labeled ssDNA (90mer) were preincubated for 10 min at 37°C in the reaction buffer containing

1mM of the indicated nucleotide and 1mM MgCl2 , CaCl2 or without a cofactor . Reactions were initiated and analyzed as described in B. (E) ssDNA binding analysis of wild type HsRAD51 using SPR in the presence of magnesium (Mg2+-) ATP and calcium (Ca2+-) ATP. (F) ssDNA binding analysis of HsRAD51(D316K) in the presence of magnesium (Mg2+-) ATP and calcium (Ca2+-) ATP.

Continued

155

Figure 3.8:continued

156

Figure 3.9 Structural comparison of MvRAD51 and RAD51 ATP Cap AspartateLysine Substitution Mutation. (A) Superimposed filament assembly of active wild type MvRAD51 structure in the presence of potassium (K+) cations (cyan) (PDB code 1XU4) and MvRAD51(D302K) (red) (PDB code 3NTU). Helical pitch is indicated with the respective color. (B) Enlarged view of the intersubunit ATP binding region of active wild type MvRAD51 and MvRAD51(D302K). Relative location of AMP-PNP-Mg2+-2K+ of the wild type MvRAD51 (blue) and the AMP-PNP-Mg2+ of the MvRAD51(D302K) (green) are shown. (C) ATP cap region of MvRadA(D302K). F129 and H280 are in active form conformation. - phosphate group of ATP analog forms a direct interaction with -amino group of K302. (D) Active filament form of E.coli RecA - K248 of the ATP cap forms a direct contact with the ALF4 group of the ATP analog (PDB code 3CMU). Electron microscopy (EM) analysis of filaments of wild type HsRAD51 (E) and HsRAD51(D316K) (F) with ssDNA. Reaction conditions and product visualization are described in Experimental Procedures. Reaction conditions are indicated above each EM image and the corresponding helical pitch is indicated below. Insets indicate enlarged images of the helical configuration. Continued

157

Figure 3.9: continued

158

Table 3.1 Summary of ATP hydrolysis and nucleotide binding data of HsRAD51 wild type (WT) and HsRAD51(D316K) mutant protein.

Kinetic Parameter WT D316K -1 kcat (min ) ssDNA 0.180 ± 0.004 0.070 ± 0.005 ssDNA, KCl 0.190 ± 0.004 0.063 ± 0.005

dsDNA 0.167 ± 0.002 0.060 ± 0.007 dsDNA, KCl 0.135 ± 0.007 0.052 ± 0.005

Km (M) ssDNA 17.48 ± 3.25 3.08 ± 0.1 ssDNA, KCl 9.34 ± 1.42 2.14 ± 0.002

dsDNA 6.54 ± 0.67 2.17 ± 0.15 dsDNA, KCl 5.52 ± 0.26 1.70 ± 0.003

KD (M) ATPS ssDNA 2.49 ± 0.46 0.66 ± 0.21 ssDNA, KCl 1.18 ± 0.22 0.72 ± 0.21

Bmax (M) ATPS ssDNA 1.02 ± 0.02 0.90 ± 0.03 ssDNA, KCl 0.83 ± 0.04 0.87 ± 0.05

KD (M) ADP ssDNA, KCl 1.24 ± 0.32 0.26 ± 0.06 Bmax (M) ADP ssDNA, KCl 0.82 ± 0.05 0.86 ± 0.02

159

Table 3.2 Dissociation rate constants of HsRAD51 wild type (WT) and HsRAD51(D316K) ssDNA and dsDNA interactions*

WT (s-1) D316K (s-1) ssDNA,no NTa 0.0014 NDb ssDNA, ADP 0.0070 0.0027 ssDNA, ATP 0.0053 0.0005 dsDNA,no NT 0.0017 ND dsDNA, ADP ND 0.0004 dsDNA, ATP 0.0041 0.0006

*The statistical error of fitting was less than 0.03 for each measure. However, binding drift suggested accuracy of ±10%. a NT- Nucleotide b ND- Not determinable as a result of low binding

160

Table 3.3 X-ray crystallographic data and Structure refinement statistics

Data collection

Space group P61 Cell dimensions a, b, c (Å) 85.0, 85.0, 104.3  (°) 90, 90, 120 Resolution (Å) 50.0 - 1.9 (1.98 - 1.90)*

Rsym 0.049 (0.278) I / I 27.1 (10.0) Completeness (%) 99.1 (98.6) Unique Reflections 33350 (3828) Redundancy 9.5 (9.3) Refinement Resolution (Å) 50.0 – 1.90 No. reflections 33350

Rwork / Rfree 0.188 / 0.214 No. atoms 2501 Protein 2375 Ligand/ion 34 Water 92 B-factors 37.6 Protein 37.7 Ligand/ion 27.6 Water 37.7 R.m.s. deviations Bond lengths (Å) 0.015 Bond angles (°) 1.73

*Values in parentheses are for highest-resolution shell.

161

3.5 REFERENCES

Baumann P, Benson FE, Hajibagheri N, West SC. 1997. Purification of human Rad51 protein by selective spermidine precipitation. Mutat. Res. 384: 65-72. Bernstein KA, Rothstein R. 2009. At loose ends: resecting a double-strand break. Cell 137: 807-810. Bianco PR, Tracy RB, Kowalczykowski SC. 1998. DNA strand exchange proteins: a biochemical and physical comparison. Front. Biosci. 3: D570-603. Bugreev DV, Golub EI, Stasiak AZ, Stasiak A, Mazin AV. 2005. Activation of human meiosis-specific recombinase Dmc1 by Ca2+. J. Biol. Chem. 280: 26886-26895. Bugreev DV, Mazin AV. 2004. Ca2+ activates human homologous recombination protein Rad51 by modulating its ATPase activity. Proc. Natl. Acad. Sci. U. S. A. 101: 9988-9993. Chen Z, Yang H, Pavletich NP. 2008. Mechanism of homologous recombination from the RecA-ssDNA/dsDNA structures. Nature 453: 489-484. Chi P, Van Komen S, Sehorn MG, Sigurdsson S, Sung P. 2006. Roles of ATP binding and ATP hydrolysis in human Rad51 recombinase function. DNA Repair (Amst) 5: 381-391. Ciccia A, Elledge SJ. 2010. The DNA damage response: making it safe to play with knives. Mol. Cell 40: 179-204. Cox MM, Lehman IR. 1981. recA protein of Escherichia coli promotes branch migration, a kinetically distinct phase of DNA strand exchange. Proc. Natl. Acad. Sci. U. S. A. 78: 3433-3437. De Zutter JK, Knight KL. 1999. The hRad51 and RecA proteins show significant differences in cooperative binding to single-stranded DNA. J. Mol. Biol. 293: 769-780. DeLano WL. 2002. The PyMOL Molecular Graphics System. Palo Alto, CA: DeLano Scientific LLC. Eggler AL, Inman RB, Cox MM. 2002. The Rad51-dependent pairing of long DNA substrates is stabilized by replication protein A. J. Biol. Chem. 277: 39280-39288. Henricksen LA, Umbricht CB, Wold MS. 1994. Recombinant replication protein A: expression, complex formation, and functional characterization. J. Biol. Chem. 269: 11121-11132. Jackson SP, Bartek J. 2009. The DNA-damage response in human biology and disease. Nature 461: 1071-1078. Kabsch W. 1993. Automatic processing of rotation diffraction data from crystals of initially unknown symmetry and cell constants. J Appl Cryst 26: 795-800. Kowalczykowski SC. 1991. Biochemistry of genetic recombination: energetics and mechanism of DNA strand exchange. Annu. Rev. Biophys. Biophys. Chem. 20: 539-575. Krogh BO, Symington LS. 2004. Recombination proteins in yeast. Annu. Rev. Genet. 38: 233-271.

162

Li Y, He Y, Luo Y. 2009. Conservation of a conformational switch in RadA recombinase from Methanococcus maripaludis. Acta Crystallogr. D. Biol. Crystallogr. 65: 602- 610. Lio YC, Mazin AV, Kowalczykowski SC, Chen DJ. 2003. Complex formation by the human Rad51B and Rad51C DNA repair proteins and their activities in vitro. J. Biol. Chem. 278: 2469-2478. Liu Y, Stasiak AZ, Masson JY, McIlwraith MJ, Stasiak A, West SC. 2004. Conformational changes modulate the activity of human RAD51 protein. J. Mol. Biol. 337: 817-827. Masson JY, Stasiak AZ, Stasiak A, Benson FE, West SC. 2001a. Complex formation by the human RAD51C and XRCC3 recombination repair proteins. Proc. Natl. Acad. Sci. U. S. A. 98: 8440-8446. Masson JY, Tarsounas MC, Stasiak AZ, Stasiak A, Shah R, McIlwraith MJ, Benson FE, West SC. 2001b. Identification and purification of two distinct complexes containing the five RAD51 paralogs. Genes Dev. 15: 3296-3307. Mazin AV, Kowalczykowski SC. 1998. The function of the secondary DNA-binding site of RecA protein during DNA strand exchange. EMBO J. 17: 1161-1168. McRee DE. 1999. XtalView/Xfit--A versatile program for manipulating atomic coordinates and electron density. J. Struct. Biol. 125: 156-165. Miller KA, Yoshikawa DM, McConnell IR, Clark R, Schild D, Albala JS. 2002. RAD51C interacts with RAD51B and is central to a larger protein complex in vivo exclusive of RAD51. J. Biol. Chem. 277: 8406-8411. Pugh BF, Cox MM. 1987. Stable binding of recA protein to duplex DNA. Unraveling a paradox. J. Biol. Chem. 262: 1326-1336. Qian X, He Y, Ma X, Fodje MN, Grochulski P, Luo Y. 2006a. Calcium stiffens archaeal Rad51 recombinase from Methanococcus voltae for homologous recombination. J. Biol. Chem. 281: 39380-39387. Qian X, He Y, Wu Y, Luo Y. 2006b. Asp302 determines potassium dependence of a RadA recombinase from Methanococcus voltae. J. Mol. Biol. 360: 537-547. Qian X, Wu Y, He Y, Luo Y. 2005. Crystal structure of Methanococcus voltae RadA in complex with ADP: hydrolysis-induced conformational change. Biochemistry (Mosc). 44: 13753-13761. San Filippo J, Sung P, Klein H. 2008. Mechanism of eukaryotic homologous recombination. Annu. Rev. Biochem. 77: 229-257. Schild D, Lio Y-c, Collins DW, Tsomondo T, Chen DJ. 2000. Evidence for Simultaneous Protein Interactions between Human Rad51 Paralogs. J. Biol. Chem. 275: 16443- 16449. Shibata T, Cunningham RP, DasGupta C, Radding CM. 1979. Homologous pairing in genetic recombination: complexes of recA protein and DNA. Proc. Natl. Acad. Sci. U. S. A. 76: 5100-5104. Shim KS, Schmutte C, Tombline G, Heinen CD, Fishel R. 2004. hXRCC2 enhances ADP/ATP processing and strand exchange by hRAD51. J. Biol. Chem. 279: 30385-30394. Shim KS, Schmutte C, Yoder K, Fishel R. 2006. Defining the salt effect on human RAD51 activities. DNA Repair (Amst) 5: 718-730.

163

Sigurdsson S, Trujillo K, Song B, Stratton S, Sung P. 2001a. Basis for avid homologous DNA strand exchange by human Rad51 and RPA. J. Biol. Chem. 276: 8798-8806. Sigurdsson S, Van Komen S, Bussen W, Schild D, Albala JS, Sung P. 2001b. Mediator function of the human Rad51B-Rad51C complex in Rad51/RPA-catalyzed DNA strand exchange. Genes Dev. 15: 3308-3318. Solinger JA, Kiianitsa K, Heyer WD. 2002. Rad54, a Swi2/Snf2-like recombinational repair protein, disassembles Rad51:dsDNA filaments. Mol. Cell 10: 1175-1188. Sonoda E, Sasaki MS, Buerstedde JM, Bezzubova O, Shinohara A, Ogawa H, Takata M, Yamaguchi-Iwai Y, Takeda S. 1998. Rad51-deficient vertebrate cells accumulate chromosomal breaks prior to cell death. EMBO J. 17: 598-608. Story RM, Steitz TA. 1992. Structure of the recA protein-ADP complex. Nature 355: 374-376. Sugiyama T, Zaitseva EM, Kowalczykowski SC. 1997. A single-stranded DNA-binding protein is needed for efficient presynaptic complex formation by the Saccharomyces cerevisiae Rad51 protein. J. Biol. Chem. 272: 7940-7945. Symington LS, Heyer WD. 2006. Some disassembly required: role of DNA translocases in the disruption of recombination intermediates and dead-end complexes. Genes Dev. 20: 2479-2486. Tombline G, Fishel R. 2002. Biochemical characterization of the human RAD51 protein. I. ATP hydrolysis. J. Biol. Chem. 277: 14417-14425. Tombline G, Shim KS, Fishel R. 2002. Biochemical characterization of the human RAD51 protein. II. Adenosine nucleotide binding and competition. J. Biol. Chem. 277: 14426-14433. Tsuzuki T, Fujii Y, Sakumi K, Tominaga Y, Nakao K, Sekiguchi M, Matsushiro A, Yoshimura Y, MoritaT. 1996. Targeted disruption of the Rad51 gene leads to lethality in embryonic mice. Proc. Natl. Acad. Sci. U. S. A. 93: 6236-6240. Van Komen S, Petukhova G, Sigurdsson S, Sung P. 2002. Functional cross-talk among Rad51, Rad54, and replication protein A in heteroduplex DNA joint formation. J. Biol. Chem. 277: 43578-43587. West SC. 2003. Molecular views of recombination proteins and their control. Nat Rev Mol Cell Biol 4: 435-445. Wu Y, He Y, Moya IA, Qian X, Luo Y. 2004. Crystal structure of archaeal recombinase RADA: a snapshot of its extended conformation. Mol. Cell 15: 423-435. Wu Y, Qian X, He Y, Moya IA, Luo Y. 2005. Crystal structure of an ATPase-active form of Rad51 homolog from Methanococcus voltae. Insights into potassium dependence. J. Biol. Chem. 280: 722-728. Zaitseva EM, Zaitsev EN, Kowalczykowski SC. 1999. The DNA binding properties of Saccharomyces cerevisiae Rad51 protein. J. Biol. Chem. 274: 2907-2915.

164

CHAPTER 4

Human RAD51B-RAD51C Paralog Heterodimer Complex Enhances and Stabilizes

RAD51 Nucleoprotein Filament for D-loop Formation

165

4.1 ABSTRACT

There are five RAD51 related proteins, RAD51B, RAD51C, RAD51D, XRCC2 and

XRCC3, termed RAD51 paralogs, that appear to enhance RAD51 mediated homologous recombinational (HR) repair of DNA double strand breaks (DSBs). Here we model the structures of human RAD51, RAD51B and RAD51C and show similar domain orientations within a hypothetical nucleoprotein filament (NPF). We then demonstrate that RAD51B-RAD51C heterodimer forms stable complex on ssDNA and partially stabilizes the RAD51 NPF against the anti-recombinogenic activity of BLM. RAD51B-

RAD51C also appears to stimulate RAD51 mediated D-loop formation in the presence of

RPA. However, RAD51B-RAD51C does not facilitate RAD51 nucleation on a RPA coated ssDNA. These findings provide further insight into RAD51 paralog involvement during presynaptic and synaptic phases of HR.

166

4.2 INTRODUCTION

Homologous recombination (HR) ensures genomic stability during DNA double strand breaks (DSBs) caused by ionizing radiation (IR) and stalled or collapsed replication machinery (Symington 2002; Krogh and Symington 2004; Ciccia and Elledge 2010). In eukaryotes, RAD51 functions as the main recombinase in HR (Symington 2002; West

2003; Krogh and Symington 2004; San Filippo et al. 2008). However, unlike its robust bacterial counterpart RecA, RAD51 exhibits an inefficient homologous pairing and strand exchange activity in vitro (Liu et al. 2004; Shim et al. 2006; San Filippo et al.

2008). This observation would appear to necessitate the requirement of accessory proteins to facilitate efficient RAD51 mediated DSB repair by HR (San Filippo et al.

2008; Heyer et al. 2010).

In yeast, Rad55 and Rad57 display a limited homology to Rad51 (termed Rad51 paralogs; (Symington 2002; Krogh and Symington 2004) and appear to have evolved from gene duplication events (Lin et al. 2006). Mutations of either Rad55 or Rad57 lead to IR sensitivity that can be overcome by overexpression of Rad51 or expression of

Rad51 gain-of-function mutants that possess a higher affinity for DNA (Hays et al. 1995;

Johnson and Symington 1995; Fortin and Symington 2002). The Rad55 and RAD57 proteins were shown to form a stable heterodimer (Sung 1997). Moreover, the inclusion of Rad55-Rad57 during strand exchange reactions relieves the inhibitory effect RPA when Rad51 and RPA are introduced simultaneously with ssDNA (Sung 1997).

Five RAD51 paralogs, RAD51B, RAD51C, RAD51D, XRCC2 and XRCC3, have been identified in vertebrates that share 20-30% homology with RAD51 (Albala et al.

167

1997; Rice et al. 1997; Cartwright et al. 1998; Dosanjh et al. 1998; Pittman et al. 1998;

Thompson and Schild 1999). All five RAD51 paralogs contain canonical Walker A/B domains and all exhibits DNA depended ATPase activities, although to a lesser extent compared to RAD51 (Braybrooke et al. 2000; Sigurdsson et al. 2001; Lio et al. 2003;

Shim et al. 2004). These paralogs exist in a variety of heterodimer complexes in vivo and in vitro that include RAD51B-RAD51C, RAD51D-XRCC2, RAD51B-RAD51C-

RAD51D-XRCC2 and RAD51C-XRCC3. Both XRCC3 and RAD51C have been shown to interact with RAD51 (Liu et al. 1998; Schild et al. 2000), although the RAD51C-

RAD51 interaction appears to be weak (Liu et al. 1998).

Mutations of most RAD51 paralog genes render cells susceptible to DSB causing agents. For example, Chinese hamster cell lines irs1 and irs1SF defective in XRCC2 and

XRCC3, respectively, are sensitive to IR and DNA crosslinking agents (Tebbs et al. 1995;

Liu et al. 1998). These cells also display chromosomal aberrations and missegregation

(Cui et al. 1999; Griffin et al. 2000). Knock out of XRCC2, RAD51D and RAD51B in mice lead to embryonic lethality (Shu et al. 1999; Deans et al. 2000; Pittman and

Schimenti 2000); a phenotype that is similar to RAD51 disruption (Tsuzuki et al. 1996;

Sonoda et al. 1998). In addition, a uterine leiomyoma contains a translocation in the intronic region of RAD51B that leads to premature termination of its transcript (Ingraham et al. 1999; Schoenmakers et al. 1999). Six monoallelic mutations within RAD51C have been shown to confer susceptibility to breast and ovarian cancer (Meindl et al. 2010), while one biallelic mutation caused abnormalities associated with Fanconi anemia (FA)

(Levy-Lahad 2010; Vaz et al. 2010). The intricate involvement of RAD51C in the FA

168

pathway led to its provisional assignment as FANCO, the 14th FA complementation group (Kee and D'Andrea 2010).

RAD51B-RAD51C, RAD51B-RAD51C-RAD51D-XRCC2 and RAD51C-

XRCC3 complexes have been shown to bind ssDNA and, dsDNA with 3- or 5-tails, dsDNA with gapped regions and branched DNA structures (Masson et al. 2001a; Masson et al. 2001b; Sigurdsson et al. 2001; Compton et al. 2010). Previously it was demonstrated that RAD51B-RAD51C functions as a mediator during RAD51 catalyzed three-strand exchange assays by reducing the competition between RPA and RAD51 for ssDNA binding (Sigurdsson et al. 2001).

Here we demonstrate that modeled structures of human RAD51, RAD51B and

RAD51C appear capable of forming filaments with similar domain orientations.

Furthermore, RAD51B-RAD51C heterodimer partially stabilizes the RAD51 NPF against the anti-recombinogenic activity of BLM. At sub-stoichiometric concentrations,

RAD51B-RAD51C appears to enhance D-loop formation in the presence of RPA.

Interestingly, in contrast to its role in alleviation of inhibitory effect of RPA on three- strand exchange (Sigurdsson et al. 2001), RAD51B-RAD51C does not appear to facilitate RAD51 nucleation on an RPA coated ssDNA during D-loop formation in vitro.

169

4.3 MATERIALS AND METHODS

Structure modeling and sequence alignments- Human RAD51, RAD51B and RAD51C structures were modeled using composite iterative threading assembly refinement software I-TASSER (Roy et al. 2010) without additional restraints and the optimized structures were superimposed on the Methanococcus voltae RadA (MvRAD51) crystal structure coordinates ( (Wu et al. 2005) PDB ID: 1XU4) using Pymol molecular imaging software (DeLano 2002). Protein sequence alignments were done using ClustalW2

(http://www.ebi.ac.uk/Tools/msa/clustalw2/)

Proteins expression and purification- Human RAD51B and RAD51C were cloned into pFast-Bac vector (Invitrogen) from cDNA. RAD51B carried a N-terminal six-histidine tag. 2 L of infected SF9 insect cells (Invitrogen) were harvested after 48 hours and resuspended in Buffer P containing 25 mM Tris-Cl (pH 8.0), 300 mM NaCl, 10mM imidazole, 10% Glycerol and protease inhibitors (0.5 mM phenylmethylsulfonyl fluoride,

1 ug/mL leupeptin, 1 ug/mL pepstatin, 1 ug/mL apropitin). The resuspended pellet was flash frozen in liquid nitrogen and stored at -80 °C until purification.

Cell pellet was thawed on ice, sonicated a passed through a 21-guage needle. The insoluble material was separated by centrifugation at 41000 rpm at 4 °C for 1 hr in a

Beckman-Coulter Ti70 rotor). The soluble material was immediately loaded on NiSO4 charged 3 mL Ni-NTA (Nitrilotriacetic acid) Superflow (Qiagen) equilibrated with buffer

P. Column was washed with 10-column volumes of the same buffer followed by 10- column volumes of Buffer P containing 20 mM imidazole. RAD51B-RAD51C complex

170

was eluted in a 50 mL linear gradient from 20 mM - 250 mM imidazole in Buffer P.

RAD51B-RAD51C containing fractions were dialyzed overnight into Buffer R containing 25 mM Tris-Cl (pH 8.0), 100 mM NaCl, 1 mM DTT, 0.5 mM EDTA and 10%

Glycerol. Dialyzed fraction was loaded onto a 10 mL SP-sepharose column (GE

Healthcare) equilibrated with buffer R and the flow through was collected loaded on a 10 mL Q-sepharose column equilibrated in the same buffer. After a 5-column volume wash

RAD51B-RAD51C was eluted in a linear gradient from 100 mM - 600 mM NaCl in

Buffer R. Pooled fractions were dialyzed back into Buffer R with 100 mM NaCl and loaded on a 10 mL Heparin sepharose (GE Healthcare) equilibrated with Buffer R washed with 5-column volumes of the same buffer. RAD51B-RAD51C was eluted in a linear gradient from 100 mM - 800 mM NaCl. Peak fractions containing RAD51B-

RAD51C were concentrated using an Amicon ultra-15 (Millipore) centrifugal filter unit at 4 °C. Concentrated fraction was dialyzed into storage buffer containing 25 mM Tris-

Cl (pH 7.5), 150 mM NaCl, 1 mM DTT, 0.1 mM EDTA and 10% Glycerol and flash frozen in liquid nitrogen in small aliquots and stored in -80 °C.

Human RAD51 and RPA were expressed and purified as described in Chapter 2

(ref.(Amunugama and Fishel 2011) and Chapter 3). BLM was purified as described

(Bhattacharyya et al. 2009) with an additional chromatographic step on S-sepharose (GE

Healthcare).

DNA Substrates- For surface plasmon resonance (SPR) analysis, a 5′ biotinylated oligo dT50 was used as ssDNA and for dsDNA 5′ biotinylated 50-mer 5-TCG AGA GGG TAA

ACC ACA- ATT ATT GAT ATA AAA TAG TTT TGG GTA GGC GA was annealed

171

with its complement and purified by HPLC on a Gen-Pak FAX column (Waters). D-loop assay substrates were prepared as described (Van Komen et al. 2002).

SPR analysis- Biotinylated DNA was immobilized on a streptavidin-coated chip (GE

Healthcare). Binding and dissociation were analyzed following the injection of the indicated amounts of protein at 25°C with a 5 L/min flow rate on a Biacore 3000 (GE

Healthcare). Reactions were performed in Buffer containing 25 mM Tris-OAc (pH 7.5),

1 mM ATP, 1 mM Mg(OAc)2, 2 mM DTT and 0.005% Tween-20.

D-loop assays- RAD51 (1M) was bound with RAD51B-RAD51C (at the indicated concentrations) on 90mer ssDNA (3 M nt) by incubating at 37 °C for 15 min in reaction buffer containing 25 mM Tris-OAc (pH 7.5), 1 mM ATP, 1 mM Mg(OAc)2, 1 mM

CaCl2, 2 mM DTT and BSA (100g/mL). After 15 min, RPA (200 nM) was added (as indicated) and D-loop formation was initiated by the addition of supercoiled dsDNA (50

M bp) and incubation at 37 °C for 15 min. D-loops were analyzed by electrophoresis on a 0.9% agarose gel after deproteinization. Gel was dried and exposed to a

PhosphoImager (Molecular Dynamics) screen for quantification.

Inhibition of D-loop formation- RAD51 (1 M) filaments were formed with RAD51B-

RAD51C (31.2 nM) if indicated, on 90-mer ssDNA (3 M nt) by incubating at 37 °C for

15 min in reaction buffer containing 25 mM Tris-OAc (pH 7.5), 1 mM ATP, 1 mM

Mg(OAc)2, 2 mM DTT, BSA (100 g/mL), 20 mM phosphocreatine and creatine

172

phosphokinase (30 U/mL). After the 10 min incubation BLM at indicated concentrations and RPA (200 nM) were added and incubated for an additional 10 min at 37 °C. 1 mM

CaCl2 was then added and followed by another 10 min incubation at 37°C. D-loop formation was initiated by the addition of supercoiled dsDNA (50 M bp) and incubation at 37°C for 15 min. D-loops were analyzed by electrophoresis on a 0.9% agarose gel after deproteinization. Gel was dried and exposed to a PhosphoImager (GE Healthcare) screen for quantification.

ATPase assays- RAD51 (1 M) and RAD51B-RAD51C at the indicated concentrations with oligo dT50 (3 M nt), X174 ssDNA (3 M nt) or X174 dsDNA (3 M bp) in reaction buffer containing 25 mM Tris-OAc (pH 7.5), 1 mM Mg(OAc)2, 2 mM DTT,

BSA (100 g/mL) and 200 M ATP supplemented with 200 nCi of [-32P] ATP for 1 hr at 37°C. The reaction was stopped by the addition of 400 μL 10% activated charcoal supplemented with 10 mM EDTA and incubated on ice for 2 hrs. Following centrifuging for 10 min, 50 μL duplicate aliquots were taken for counting [32P] free phosphate by

Cerenkov method (Tombline and Fishel 2002). Hydrolyzed ATP percentage is indicated.

173

4.4 RESULTS

Homology based modeled structures of RAD51, RAD51B and RAD51C indicate similar structural and functional domain orientations.

The crystal structure of human RAD51 or any of its paralogs has not been solved.

Utilizing a composite iterative threading assembly refinement technique (Roy et al.

2010), structures of RAD51, RAD51B and RAD51C were modeled. The predicted three-dimensional structures were based on the archaebacterial RadA structure solved in the presence of K+ (Wu et al. 2005). Structural alignments indicated similar orientations of the ATP cap and the ssDNA binding L2 region (Fig. 4.1). The RAD51B predicted structure contains an extra disordered region within L2 region (Fig. 4.1Band Fig. 4.2).

Interestingly, the RAD51(D316) residue aligns in an orientation that seems to coordinate with the K+ ion that functions as a salt bridge between the ATP cap of RadA and the ATP analog (Fig. 4.1A). In both RAD51B and RAD51C, the analogous residues to

RAD51(D316), RAD51B(K324) and RAD51C(K328), respectively, align at a position that would allow direct interaction with the -phosphate group of the ATP-analog (Fig.

4.1B and Fig. 4.1C). These structural predictions and alignments provide intriguing insights into the potential role of the ATP cap domain of RAD51, RAD51B and RAD51C proteins and how these proteins might collectively form a functional NPF.

RAD51B- RAD51C heterodimer forms stable complexes on ssDNA and dsDNA.

RAD51, RAD51B, RAD51C and the RAD51B-RAD51C heterodimer complex have been shown to bind ssDNA, DNA substrates with tails and dsDNA (Benson et al. 1994;

174

Sigurdsson et al. 2001; Lio et al. 2003). The dsDNA binding of RAD51B and RAD51C appears less avid than RAD51 (Benson et al. 1994; Sigurdsson et al. 2001; Lio et al.

2003). We wanted to investigate the RAD51B-RAD51C heterodimer turnover from

DNA and compare it to the turnover properties of RAD51 to ascertain complex stability.

We employed real-time DNA binding/dissociation analysis using surface plasmon resonance (SPR). These studies were performed in 25mM KCl and 100mM KCl to determine the ionic conditions that allow RAD51B-RAD51C binding to DNA.

RAD51 displayed comparable levels of ssDNA and dsDNA binding in both conditions (Fig. 4.3, time scale 100s-300s). However, RAD51 turnover from ssDNA increased at the higher salt concentration (Fig. 4.3, time scale 300s-600s). The increased turnover can be attributed to the salt induced enhancement of the ATPase activity of the recombinase (Liu et al. 2004; Shim et al. 2006); a common manifestation of the RecA/

RAD51 family of proteins (Kowalczykowski 1991; San Filippo et al. 2008).

Interestingly, RAD51B-RAD51C binding to both ssDNA and dsDNA was significantly inhibited in 100 mM KCl (Fig. 4.3A and Fig.4.3B). The inhibition reduced dsDNA binding to nearly the control level (Fig. 4.3B). However, an intriguing feature emerged from RAD51B-RAD51C dissociation pattern. Even though the binding was comparably less than that of RAD51, the bound RAD51B-RAD51C appeared stably associated with the DNA, indicating little or no turnover. This feature was evident in both 25 mM and

100 mM salt on the ssDNA and 25 mM salt on the dsDNA (Fig. 4.3A and Fig.4.3B).

Collectively, these results indicate that RAD51B-RAD51C may form stable complexes on DNA.

175

RAD51B-RAD51C enhances RAD51 catalyzed D-loop formation in the presence of RPA at sub-stoichiometric amounts.

We then investigated the D-loop stimulatory effect of RAD51B-RAD51C catalyzed by

RAD51 and RPA (Fig. 4.4A). Previous studies of RAD51 paralog stimulated strand exchange reactions catalyzed by RAD51 were observed when the paralogs were used at sub-stoichiometric amounts (Sung 1997; Sigurdsson et al. 2001). In the D-loop formation we incubated RAD51 and varying amounts of RAD51B-RAD51C with ssDNA in D-loop formation buffer containing Ca++ (Fig. 4.4), since this divalent cation is a requirement for human RAD51 catalyzed D-loop formation (See figure 4.6B; (Bugreev and Mazin 2004). RAD51B-RAD51C complex did not possess any D-loop forming ability alone (Fig. 4.4B, lanes 2 and 10). The co-incubation of RAD51B-RAD51C complex with RAD51 was necessary to see optimal D-loop formation compared to addition of the paralog complex prior to RAD51 or after formation of RAD51 nucleoprotein filament (Fig. 4.6A and 4.6B). For reactions in lanes 11-17 of figure 4.4B, purified human RPA (Fig. 4.5) was added Prior to initiation of D-loop formation by addition of supercoiled dsDNA. The stimulation by RAD51B-RAD51C occurred at sub- stoichiometric levels (Fig. 4.4B and 4.4C). In the presence of RPA, we observed an increase in D-loop formation that maximized at a molar ratio of ~ 1:30, RAD51B-

RAD51C : RAD51 (Fig. 4.4B, lanes 11-13 and Fig 4.4C). Increasing the concentration of

RAD51B-RAD51C further lead to decreased product formation. We expected addition of RPA to stimulate D-loop formation as it could potentially bind to the displaced ssDNA. These observations suggest that there may be some functional synergism between a RAD51/ RAD51B-RAD51C NPF and RPA.

176

RAD51B-RAD51C protects RAD51 nucleoprotein filament against the anti-recombinase function of BLM.

Yeast Sgs1 and Human BLM of the RecQ family as well as yeast Srs2 have been shown to induce anti-recombinase activity by dissociating RAD51 from ssDNA thus preventing functional NPF formation for strand exchange (Krejci et al. 2003; Veaute et al. 2003;

Bugreev et al. 2007). We hypothesized that RAD51 NPF could be stabilized by the

RAD51B-RAD51C heterodimer as a result of its stable binding activity to ssDNA, as previously demonstrated by SPR.

The assay for RAD51 NPF stability was based on a previous approach to study the inhibitory effects of BLM on RAD51-mediated strand exchange activity (Bugreev et al. 2007). First the RAD51 NPFs were formed either in the absence or presence of the

RAD51B-RAD51C heterodimer in an ATP regeneration system that is required for recombinase NPF formation as well as for the subsequence BLM mediated step. The amount of RAD51B-RAD51C introduced into the reaction was based on the concentration required for optimal D-loop formation (Fig 4.4B). After the initial incubation, purified BLM (Fig. 4.5) and RPA were added, and the reactions were incubated further (Fig. 4.7A). RPA serves to prevent possible RAD51 rebinding events after it has been dislodged by BLM translocase activity on ssDNA (Bugreev et al. 2007).

D-loop formation was initiated by addition of Ca++ and supercoiled DNA. BLM translocation is greatly attenuated in the presence of Ca++ (Bugreev et al. 2007), therefore it was necessary to add the divalent cation prior to the addition of supercoiled dsDNA.

BLM translocation dissociates RAD51 from the NPF, that would lead to reduced D-loop

177

formation. As expected, we observed a decreased RAD51 mediated D-loop formation with increasing concentration of BLM (Fig. 4.7B and 4.7C). As we hypothesized, a partial stabilization of the NPF was observed in the presence of RAD51B-RAD51C paralog complex, resulting modestly increased D-loop formation (compare lanes 2-8 and lanes 9-15 of Fig. 4.7B and 4.7C). This result suggests that RAD51B-RAD51C may confer a partial stability to the RAD51 NPF.

RAD51-RAD51C does not suppress the overall ATPase activity of the RAD51/ RAD51B-

RAD51C nucleoprotein filament.

Both RAD51 and RAD51B-RAD51C are DNA stimulated ATPases (Benson et al. 1994;

Sigurdsson et al. 2001; Tombline and Fishel 2002). The ATPase activity of RAD51B-

RAD51C, derives from both constituents RAD51B and RAD51C (Lio et al. 2003). Since we observed modestly increased D-loop formation with sub-stoichiometric amounts of

RAD51B-RAD51C in the RAD51 catalyzed reaction, we investigated if the paralog complex inhibited the ATP turnover rate of RAD51, as recombinase activity of RAD51 can be enhanced by suppressing the ATP hydrolysis of the protein (Bugreev and Mazin

2004)..

We examined the ATP hydrolysis activity of the RAD51/ RAD51B-RAD51C mixed filament in varying ratios with an oligo dT50, X174 ssDNA and X174 dsDNA substrates (Fig. 4.8). As expected, RAD51 indicated the highest ATP turnover rate with the oligonucleotide substrate (Fig. 4.8). Collectively, the ATP turnover rates of the mixed filament increased with the concentration of the paralog complex, even at ratios that induced enhanced D-loop formation (compare Fig. 4.8 with Fig. 4.4B and 4.4C).

178

This suggests that the stimulatory effect of RAD51B-RAD51C on RAD51 mediated strand exchange activity is not a result of suppression of the ATP turnover rate of the latter.

RAD51B-RAD51C does not alleviate inhibitory effects of RPA to promote RAD51 mediated D-loop formation.

During DSB repair, RPA binds to ssDNA once the 5 ends are resected, to prevent formation of secondary structures (Sugiyama et al. 1997). RPA is eventually replaced by

RAD51 nucleation mediated by either Rad52 in yeast or BRCA2 in vertebrates (Kanaar and Hoeijmakers 1998; New et al. 1998; Shinohara and Ogawa 1998; Yang et al. 2005;

Jensen et al. 2010). During recombinase assays in vitro, pre-incubation of RPA and ssDNA leads to greatly reduced strand exchange activity , due to its strong affinity for the latter. Previously, it was shown that RAD51B-RAD51C heterodimer functions as a presynaptic mediator in a three-strand exchange assay (Sigurdsson et al. 2001). We examined if RAD51B-RAD51C could alleviate the inhibitory effect of RPA without the requirement of BRCA2 during RAD51 catalyzed D-loop formation.

In this approach we pre-incubated ssDNA with the indicated amounts of RPA and then RAD51 or RAD51 with RAD51B-RAD51C was added (Fig. 4.9). Pre-addition of

RPA lead to significantly reduced product formation compared to RPA addition after

RAD51/ RAD51B-RAD51C NPF had formed (compared lanes 2 and 6, and lanes 7 and

11 in Fig. 4.9A and Fig 4.9B). We did not observe any alleviation of RPA inhibition by

RAD51B-RAD51C (Fig. 4.9A and 4.9B).

179

We also examined if RAD51B-RAD51C directly interacted with RPA. Pull- down experiments were performed with Ni-NTA beads exploiting the His-tag on

RAD51B (Fig. 4.9C). However, we were not able to see any direct physical interaction between RAD51B-RAD51C paralog complex and RPA (Fig. 4.9C). Even a pull down experiment performed at a low-salt concentration (50mM KCl) failed to indicate an interaction between RPA and RAD51B-RAD51C (data not shown).

180

4.5 DISCUSSION

Evolution of RAD51 paralogs in eukaryotes.

Gene duplication allows one copy of the gene to retain the existing function while the other evolves to gain a new function by mutations (Lin et al. 2006; Conant and Wolfe

2008). The process also leads to separation of ancestral function (Conant and Wolfe

2008). Phylogenetic analyses indicate that the RecA/ RAD51 family of genes could be divided into three subfamilies; RAD, RAD and RecA that have all evolved from a common ancestral RecA (Lin et al. 2006). RAD contains genes of highly conserved function that include RAD51 and DMC1. RAD includes genes of divergent functions that include RAD51 paralogs RAD51B, RAD51C, RAD51D, XRCC2 and XRCC3 (Lin et al. 2006). RecA is found in eubacteria (Lin et al. 2006). The budding yeast Rad51 paralogs, Rad55 and Rad57 genes appear to be orthologous to vertebrate XRCC2 and

XRCC3, respectively (Lin et al. 2006). Even though RAD51 paralogs share only 20-30% homology with RAD51, there are well conserved within vertebrate cells (Takata et al.

2000; Takata et al. 2001).

Homologous recombination mechanism is conserved in prokaryotes and eukaryotes. The structural and functional properties of the recombinases; bacterial RecA and eukaryotic RAD51 are similar (Ogawa et al. 1993; Story et al. 1993; Benson et al.

1994). However, compared to RecA, RAD51 possesses a greatly reduced recombinase function (Liu et al. 2004; Shim et al. 2006; San Filippo et al. 2008). This necessitates the requirement of other recombination mediator to perform efficient recombinational repair in more complex genomes. 181

Stabilization of the nucleoprotein filament

We previously reported (Chapter3) that the ATP cap of RAD51 is intricately involved in regulating the turnover of RAD51 thereby, affecting the recombinase function. The Mutation, D316K of the ATP cap of RAD51 attenuates the ATPase activity and leads to formation of a stable NPF that results from slow protein turnover from DNA. This in turn, causes an enhanced recombinase activity. In our modeled structures the ATP cap residues RAD51(D316), RAD51B(K324) and RAD51C(K328) are positioned in a similar orientation. Therefore, it is tempting to speculate that the slow protein turnover observed for RAD51B-RAD51C- DNA interaction (see Fig. 4.3), may be mediated by the ATP cap of the paralog complex. However, RAD51B-RAD51C complex does not possess any recombinase function (Fig. 4.4 and 4.6 and ref.(Sigurdsson et al. 2001)). Also, addition of RAD51B-RAD51C does not reduce the amount of

RAD51 required for strand exchange functions (data not shown; (Sigurdsson et al.

2001)). Increasing the amounts of RAD51B-RAD51C in the reactions in fact, leads to reduced strand exchange activity by RAD51 (Fig. 4.4). In our SPR studies we did not observe any apparent reduction in the protein turnover rate when mixed RAD51/

RAD51B-RAD51C NPFs were formed at ratios that were optimal for enhanced D-loop formation (data not shown). Therefore, the partial stabilization of the NPF we observe against the anti-recombination function of BLM may be as a cause of RAD51B-RAD51C functioning as a mediator of D-loop formation after binding to DNA interspersedly (Fig.

4.7).

182

For human RAD51 catalyzed D-loop formation, ATP must be in an active non- hydrolyzable state, i.e. in the presence of Ca++ (Bugreev and Mazin 2004). In the cytosol

Ca++ ions are tightly regulated as they function as secondary messengers and they do not appear in concentrations appreciable for D-loop formation (Ritchie et al. 2011).

Therefore, we hypothesized that RAD51B-RAD51C may function as auxiliary proteins in stabilizing the NPF. However, our ATPase data do not support the fact that RAD51B-

RAD51C stabilizes the NPF by reducing the overall ATP hydrolysis of the RAD51/

RAD51B-RAD51C mixed NPF (Fig. 4.8). Structurally, this is unlikely as both RAD51B and RAD51C paralogs contain Walker A box domains that hydrolyze ATP (Fig. 4.2).

Therefore, it is likely that the partial stabilization we observe against BLM anti- recombination might be as a result of RAD51B-RAD51C being physically bound to

DNA, thereby slowing down the translocation of BLM. However, we cannot rule out the possibility of post-translational modifications or the involvement of other RAD51 paralogs in vivo that might improve the stabilization effect conferred by RAD51B-

RAD51C complex.

RAD51 foci formation during IR is the initial step of HR repair (Haaf et al. 1995;

Scully et al. 1997). BRCA2 is involved in transporting RAD51 to the nucleus, as the latter lacks a nuclear localization signal (NLS) and initiating recombinase nucleation on ssDNA (Gildemeister et al. 2009). The BRC repeats of BRCA2 stabilize RAD51 by suppressing its ATPase activity and facilitates nucleation on ssDNA while inhibiting

RAD51 filament formation on dsDNA (Carreira et al. 2009; Jensen et al. 2010; Carreira and Kowalczykowski 2011). However, in pancreatic epithelial tumor derived CAPAN-1 cells (Goggins et al. 1996), that encode a C- terminal truncated BRCA2 that lacks the

183

NLS, RAD51 foci formation occurs to a lesser extent during S phase and even upon IR in a BRCA2 independent manner (Tarsounas et al. 2003; Gildemeister et al. 2009).

Surprisingly, BRCA2 independent RAD51 nuclear localization was attributed to

RAD51C, which contains a functional NLS (Gildemeister et al. 2009). Therefore,

RAD51B-RAD51C mediated stabilization of the NPF against anti-recombinases of the

RecQ family as we have shown here might be critical in event of BRCA2 absence.

Even though HR is an error-free repair pathway to repair DSBs, unregulated HR could lead to gross chromosomal rearrangements (Wu and Hickson 2006). Therefore, cells have evolved counter mechanism to suppress HR at untimely events. Both cellular and biochemical data indicate that the yeast Srs2 helicase disrupts Rad51-ssDNA filaments to suppress HR (Krejci et al. 2003; Veaute et al. 2003). Overexpression of

RecQ homolog Sgs1 in Srs2 deleted strains suppresses the hyperrecombinogenic phenotypes (Mankouri et al. 2002; Ira et al. 2003). Human protein BLM and RECQL5 has been reported to suppress RAD51 mediated strand exchange activity by dissociating

RAD51-ssDNA filaments (Bugreev et al. 2007; Hu et al. 2007). Recently, yeast Rad55-

Rad57 heterodimer complex was shown to stabilize Rad51 NPF to counteract the anti- recombinogenic activity of Srs2 helicase by a direct protein-protein mediated interaction

(Liu et al. 2011). Furthermore, a genetic study illustrated that the yeast Shu complex that contains Shu1, Shu2, Csm2 and Psy3 (Shu1 and Psy3 being Rad51 paralogs) were involved in suppressing the anti-recombinase activity of Srs2 by physically interacting with the helicase (Bernstein et al. 2011). A similar direct interaction between human

RAD51B-RAD51C paralog complex with the helicase BLM that might inhibit the anti- recombinase function is yet to be characterized. However, BLM has been shown to

184

physically interact with RAD51D of the RAD51D-XRCC2 complex (Braybrooke et al.

2003).

Enhanced D-loop formation in the presence of RPA

RAD51C possesses an ability to destabilize dsDNA that might facilitate DNA strand exchange activity (Lio et al. 2003). In our D-loop assays the RPA induced stimulation appears more apparent in the presence of RAD51B-RAD51C (Fig. 4.4). We could rationalize this observation as RPA capturing the displaced strand that has been destabilized by RAD51C of the RAD51B-RAD51C complex to facilitate RAD51 catalyzed D-loop formation. In fact, previous studies have shown the RPA facilitating the strand exchange activity by binding to the leaving ssDNA (Eggler et al. 2002; Van

Komen et al. 2002).

We can conclude that the RAD51B-RAD51C mediated stimulation of D-loop formation occurs at the presynaptic phase (Fig. 4.6). From our results it is evident that

RAD51B-RAD51C requires co-incubation with RAD51 for optimal NPF formation to catalyze strand invasion (Fig. 4.6). Our D-loop assay does not indicate that RAD51B-

RAD51C has the ability to facilitate nucleation of RAD51, by binding to ssDNA prior to

RAD51. RAD51B-RAD51C also does not appear to bind later to the RAD51 NPF to facilitate D-loop formation (Fig. 4.6). Thus, our conclusion is consistent with a previous observation of RAD51B-RAD51C mediator activity using three-strand exchange assays

(Sigurdsson et al. 2001).

185

In summary, we could conclude that RAD51 paralog complex RAD51B-RAD51C might enable regulation of the stability of the RAD51 NPF by acting as a pro- recombinase against anti-recombinogenic factors (Fig. 4.10). The dynamic property afforded to the RAD51 NPF by these recombination mediators is required for a highly regulated HR mechanism in the cell.

186

4.6 ACKNOWLEDGMENTS AND FOOTNOTES

Purified BLM protein used in this study was kindly provided by Jeremy Keirsey and

Joanna Groden at The Ohio State University. Work in this chapter is in preparation for publication as: Human RAD51B-RAD51C paralog heterodimer complex enhances and stabilizes RAD51 nucleoprotein filament for D-loop formation, Amunugama R. and

Fishel R.

187

Figure 4.1 Homology based modeled structures of RAD51, RAD51B and RAD51C indicate similar domain orientations. (A) Human RAD51 (HsRAD51) structure was modeled using homology based modeling software I-TASSER (Roy et al. 2010) and the optimized structures were superimposed on the Methanococcus voltae RadA (MvRAD51) crystal structure coordinates ((Wu et al. 2005) PDB ID: 1XU4) using Pymol molecular imaging software (DeLano 2002). The non-hydrolysable ATP analog AMP- PNP is bound at the interface between MvRAD51 top monomer (cyan) and the bottom monomer (wheat). The Walker A subunit of the bottom monomer is indicated in grey. L2 ssDNA binding regions of MvRAD51 and HsRAD51 are highlighted in salmon and brown colors, respectively. The ATP caps of MvRAD51 and HsRAD51 are indicated in red and magenta colors, respectively. D316 of HsRAD51 is indicated. Two potassium (K+) ions at the interface of MvRAD51 monomers are also illustrated. Modeled and superimposed structures of Human RAD51B (HsRAD51B) in Blue (B) and Human RAD51C (HsRAD51C) in Green (C). The domain color assignments are similar to A.

188

Figure 4.2 Sequence alignment of RAD51 homologs and RAD51 paralogs; RAD51B and RAD51C. ClustalW sequence aligment of Methanococcus voltae RadA (MvRAD51), human RAD51 (HsRAD51), human RAD51B (HsRAD51B) and human RAD51C (HsRAD51C). Walker A, L2 ssDNA binding region and the ATP domains are highlighted in grey, wheat and pink colors, respectively.

189

Figure 4.3 RAD51B-RAD51C heterodimer forms stable complexes on DNA. (A) ssDNA binding and dissociation curves obtained from surface plasmon resonance (SPR). The SPR curves correspond to 800nM concentration of either RAD51 or RAD51B- RAD51C heterodimer. DNA binding corresponds to the time scale 100s-300s and dissociation from 300s-600s. Binding response is given in Response Units. DNA binding and dissociation were analyzed in 25mM Tris-OAc (pH 7.5), 1mM ATP, 1mM

Mg(OAc)2, 2mM DTT and 0.005% Tween-20. Control (no protein) curve is red, RAD51B-RAD51C binding at 25mM K+ (KCl) curve is black, RAD51B-RAD51C binding at 100mM K+ curve is pink, RAD51 binding at 25mM K+ curve is blue and RAD51 binding at 100mM K+ curve is green. (B) dsDNA binding and dissociation of RAD51B-RAD51C heterodimer and RAD51 analyzed by SPR. The reaction conditions and the binding curve descriptions are same as in A.

190

Figure 4.4 RAD51B-RAD51C enhances RAD51 catalyzed D-loop formation in the presence of RPA. (A) Schematic diagram of the experimental approach. Asterisk indicates the [32P] label on ssDNA. Utilized proteins and DNA substrates are illustrated with labels. B-C, RAD51B-RAD51C heterodimer. RAD51 (1M) filaments were formed with RAD51B-RAD51C at the indicated concentrations (if added), on 90mer ssDNA (3M nt) by incubating at 37°C for 15min in reaction buffer containing 25mM

Tris-OAc (pH 7.5), 1mM ATP, 1mM Mg(OAc)2, 1mM CaCl2, 2mM DTT and BSA (100g/mL). After 15min, RPA (200nM) was added (if indicated) and D-loop formation was initiated by the addition of supercoiled dsDNA (50M bp) and incubation at 37°C for 15min. D-loops were analyzed by electrophoresis on a 0.9% agarose gel after deproteinization. Gel was dried and exposed to a PhosphoImager screen for quantification. (C) Quantification of D-loops formed in B. Error bars indicate S.D.

continued

191

Figure 4.4:continued

192

Figure 4.5 Purified proteins used in the D-loop assays. RAD51B-RAD51C (1g), RPA (RPA1-RPA2-RPA3) (1g), RAD51 (1g) and BLM (500ng) were analyzed by 15% SDS-PAGE.

193

Figure 4.6 Order of addition of RAD51 and RAD51B-RAD51C for D-loop formation. (A) In scheme I, RAD51 (1M) filaments were formed with RAD51B- RAD51C at the indicated concentrations in B, on 90mer ssDNA (3M nt) by incubating at 37°C for 10min in reaction buffer containing 25mM Tris-OAc (pH 7.5), 1mM ATP,

1mM CaCl2, 2mM DTT and BSA (100g/mL). D-loop formation was initiated by the addition of supercoiled dsDNA (50M bp) and incubation at 37°C for 15min. In scheme II, RAD51 (1M) filaments were formed on 90mer ssDNA (3M nt) by incubating at 37°C for 10min in reaction buffer. RAD51B-RAD51C at the indicated concentrations in B were then added and incubated further for 10min at 37°C. In scheme III, RAD51B- RAD51C at the indicated concentrations in B were mixed with 90mer ssDNA (3M nt) and incubated at 37°C for 10min in reaction buffer. RAD51 (1M) was then added and incubated further for 10min at 37°C. In all schemes D-loop formation was initiated by the addition of supercoiled dsDNA (50M bp) and incubation at 37°C for 15min. D- loops were analyzed by electrophoresis on a 0.9% agarose gel after deproteinization. Gels were dried and exposed to PhosphoImager screens for quantification. (C) Quantification of relative amounts of D-loops formed in B. Quantification of D-loops formed in the presence of Mg++ is also shown. Error bars indicate S.D. Continued

194

Figure 4.6: continued

195

Figure 4.7 RAD51B-RAD51C protects RAD51 nucleoprotein filament against anti- recombinase function of BLM. (A) Experimental scheme. Asterisk denotes the [32P] label on ssDNA. Utilized proteins and DNA substrates are illustrated with labels. B-C, RAD51B-RAD51C heterodimer. (B) RAD51 (1M) filaments were formed with RAD51B-RAD51C (31.2 nM) if indicated, on 90mer ssDNA (3M nt) by incubating at 37°C for 15min in reaction buffer containing 25mM Tris-OAc (pH 7.5), 1mM ATP,

1mM Mg(OAc)2, 2mM DTT, BSA (100g/mL), 20mM phosphocreatine and creatine phosphokinase (30U/mL). After the 10min incubation BLM at indicated concentrations and RPA (200nM) were added and incubated further for 10min at 37°C. 1mM CaCl2 was then added and followed by another 10min incubation at 37°C. D-loop formation was initiated by the addition of supercoiled dsDNA (50M bp) and incubation at 37°C for 15min. D-loops were analyzed by electrophoresis on a 0.9% agarose gel after deproteinization. Gel was dried and exposed to a PhosphoImager screen for quantification. (C) Quantification of relative amounts of D-loops formed in B. Error bars indicate S.D.

continued

196

Figure 4.7: continued

197

Figure 4.8 RAD51B-RAD51C does not suppress the overall ATPase activity of the RAD51/RAD51B-RAD51C nucleoprotein filament. ATPase assays were performed by incubating RAD51 (1M) and RAD51B-RAD51C at the indicated concentrations with oligo dT50 (3M nt), X174 ssDNA (3M nt) or X174 dsDNA (3M bp) in reaction buffer containing 25mM Tris-OAc (pH 7.5), 1mM Mg(OAc)2, 2mM DTT, BSA (100g/mL) and 200M ATP supplemented with 200nCi of [-32P] ATP for 1hr at 37°C. The reaction was stopped by the addition of 400μL 10% activated charcoal supplemented with 10mM EDTA and incubated on ice for 2 hrs. Following centrifuging for 10 min, 50μL duplicate aliquots were taken for counting [32P] free phosphate by Cerenkov method. ATP hydrolyzed percentage is indicated.

198

Figure 4.9 RAD51B-RAD51C does not alleviate inhibitory effects of RPA (A) RPA 200nM, 400nM, 800nM and 1600nM (lanes 2-5 and lanes 7-10) was incubated with 90mer ssDNA (3M nt) at 37°C for 10min in reaction buffer containing 25mM Tris-OAc

(pH 7.5), 1mM ATP, 1mM Mg(OAc)2, 1mM CaCl2, 2mM DTT and BSA (100g/mL) for 5min. After 5min, RAD51 (1M) and RAD51B-RAD51C (32.5 nM) if indicated, was added and incubation was continued for another 15min at 37°C. Reaction in lane 6, RAD51 was incubated with 90mer ssDNA (3M nt) at 37°C for 10min in reaction buffer. Reaction in lane 11, RAD51 (1M) and RAD51B-RAD51C (32.5nM) incubated with 90mer ssDNA (3M nt) at 37°C for 10min in reaction buffer, prior to addition of RPA and further incubation at 37°C for 5min. D-loop formation was initiated by the addition of supercoiled dsDNA (50M bp) and incubation at 37°C for 15min. D-loops were analyzed by electrophoresis on a 0.9% agarose gel after deproteinization. Gels were dried and exposed to PhosphoImager screens for quantification. (B) Quantification of D- loops formed in A. Error bars indicate S.D. (C) RAD51B-RAD51C does not interact with RPA directly. Pull down experiments were performed in three separate reactions; RPA (20nM) alone, RAD51B-RAD51C (20nM) alone or RPA (20nM) and RAD51B- RAD51C (20nM) were incubated in reaction buffer containing 25mM Tris-OAc (pH 7.5), 20mM Imidazole, 150mM KCl and 0.05% Tween-20 in 50L for 1hr at 4°C on a rotisserie shaker. Ni-NTA magnetic beads (Qiagen) were then added at a final concentration of 1% (w/v) and incubation continued for another 1hr. Beads were isolated using a spin magnet and the solution was discarded. Beads were then washed twice with 100L of wash buffer (reaction buffer containing 50mM Imidazole). Bound proteins were eluted in 50L of elution buffer (reaction buffer containing 250mM Imidazole). Eluted fractions was analyzed on 15% SDS-PAGE followed by silver staining (indicated as pull down). As markers purified RPA and RAD51B-RAD51C heterodimer were also run.

199

continued Figure 4.9:continued

200

Figure 4.10 Proposed model for RAD51B-RAD51C mediated stabilization of the RAD51 nucleoprotein filament. Here we represent RAD51 mediated D-loop formation that can be enhanced by RAD51B-RAD51C paralogs. The stable nucleoprotein filament that forms could slow the action of anti-recombinases, a property that might be enhanced by additional cofactors. Proteins are illustrated with labels. B-C denotes RAD51B- RAD51C heterodimer.

201

4.7 REFERENCES

Albala JS, Thelen MP, Prange C, Fan W, Christensen M, Thompson LH, Lennon GG. 1997. Identification of a Novel HumanRAD51Homolog,RAD51B. Genomics 46: 476-479. Amunugama R, Fishel R. 2011. Subunit Interface Residues F129 and H294 of Human RAD51 Are Essential for Recombinase Function. PLoS ONE 6: e23071. Benson FE, Stasiak A, West SC. 1994. Purification and characterization of the human Rad51 protein, an analogue of E. coli RecA. EMBO Journal 13: 5764-5771. Bernstein KA, Reid RJD, Sunjevaric I, Demuth K, Burgess RC, Rothstein R. 2011. The Shu complex, which contains Rad51 paralogues, promotes DNA repair through inhibition of the Srs2 anti-recombinase. Mol. Biol. Cell 22: 1599-1607. Bhattacharyya S, Keirsey J, Russell B, Kavecansky J, Lillard-Wetherell K, Tahmaseb K, Turchi JJ, Groden J. 2009. Telomerase-associated Protein 1, HSP90, and Topoisomerase II Associate Directly with the BLM Helicase in Immortalized Cells Using ALT and Modulate Its Helicase Activity Using Telomeric DNA Substrates. J. Biol. Chem. 284: 14966-14977. Braybrooke JP, Li J-L, Wu L, Caple F, Benson FE, Hickson ID. 2003. Functional Interaction between the Bloom's Syndrome Helicase and the RAD51 Paralog, RAD51L3 (RAD51D). J. Biol. Chem. 278: 48357-48366. Braybrooke JP, Spink KG, Thacker J, Hickson ID. 2000. The RAD51 Family Member, RAD51L3, Is a DNA-stimulated ATPase That Forms a Complex with XRCC2. J. Biol. Chem. 275: 29100-29106. Bugreev DV, Mazin AV. 2004. Ca2+ activates human homologous recombination protein Rad51 by modulating its ATPase activity. Proceedings of the National Academy of Sciences of the United States of America 101: 9988-9993. Bugreev DV, Yu X, Egelman EH, Mazin AV. 2007. Novel pro- and anti-recombination activities of the Bloom's syndrome helicase. Genes Dev. 21: 3085-3094. Carreira A, Hilario J, Amitani I, Baskin RJ, Shivji MK, Venkitaraman AR, Kowalczykowski SC. 2009. The BRC repeats of BRCA2 modulate the DNA- binding selectivity of RAD51. Cell 136: 1032-1043. Carreira A, Kowalczykowski SC. 2011. Two classes of BRC repeats in BRCA2 promote RAD51 nucleoprotein filament function by distinct mechanisms. Proc. Natl. Acad. Sci. U. S. A. 108: 10448-10453. Cartwright R, Dunn AM, Simpson PJ, Tambini CE, Thacker J. 1998. Isolation of novel human and mouse genes of the recA/RAD51 recombination-repair gene family. Nucleic Acids Res 26: 1653-1659. Ciccia A, Elledge SJ. 2010. The DNA damage response: making it safe to play with knives. Molecular Cell 40: 179-204. Compton SA, Ӧzgür S, Griffith JD. 2010. Ring-shaped Rad51 Paralog Protein Complexes Bind Holliday Junctions and Replication Forks as Visualized by Electron Microscopy. Journal of Biological Chemistry 285: 13349-13356.

202

Conant GC, Wolfe KH. 2008. Turning a hobby into a job: How duplicated genes find new functions. Nat Rev Genet 9: 938-950. Cui X, Brenneman M, Meyne J, Oshimura M, Goodwin EH, Chen DJ. 1999. The XRCC2 and XRCC3 repair genes are required for chromosome stability in mammalian cells. Mutat. Res. 434: 75-88. Deans B, Griffin CS, Maconochie M, Thacker J. 2000. Xrcc2 is required for genetic stability, embryonic neurogenesis and viability in mice. EMBO J. 19: 6675-6685. DeLano WL. 2002. The PyMOL Molecular Graphics System. Palo Alto, CA: DeLano Scientific LLC. Dosanjh MK, Collins DW, Schild D, Fan W, Lennon GG, Albala JS, Shen Z. 1998. Isolation and characterization of RAD51C, a new human member of the RAD51 family of related genes. Nucleic Acids Res 26: 1179-1184. Eggler AL, Inman RB, Cox MM. 2002. The Rad51-dependent pairing of long DNA substrates is stabilized by replication protein A. Journal of Biological Chemistry 277: 39280-39288. Fortin GS, Symington LS. 2002. Mutations in yeast Rad51 that partially bypass the requirement for Rad55 and Rad57 in DNA repair by increasing the stability of Rad51-DNA complexes. EMBO J. 21: 3160-3170. Gildemeister OS, Sage JM, Knight KL. 2009. Cellular Redistribution of Rad51 in Response to DNA Damage. J. Biol. Chem. 284: 31945-31952. Goggins M, Schutte M, Lu J, Moskaluk CA, Weinstein CL, Petersen GM, Yeo CJ, Jackson CE, Lynch HT, Hruban RH et al. 1996. Germline BRCA2 Gene Mutations in Patients with Apparently Sporadic Pancreatic Carcinomas. Cancer Research 56: 5360-5364. Griffin CS, Simpson PJ, Wilson CR, Thacker J. 2000. Mammalian recombination-repair genes XRCC2 and XRCC3 promote correct chromosome segregation. Nat Cell Biol 2: 757-761. Haaf T, Golub EI, Reddy G, Radding CM, Ward DC. 1995. Nuclear foci of mammalian Rad51 recombination protein in somatic cells after DNA damage and its localization in synaptonemal complexes. Proc. Natl. Acad. Sci. U. S. A. 92: 2298- 2302. Hays SL, Firmenich AA, Berg P. 1995. Complex formation in yeast double-strand break repair: participation of Rad51, Rad52, Rad55, and Rad57 proteins. Proc. Natl. Acad. Sci. U. S. A. 92: 6925-6929. Heyer WD, Ehmsen KT, Liu J. 2010. Regulation of homologous recombination in eukaryotes. Annual Review of Genetics 44: 113-139. Hu Y, Raynard S, Sehorn MG, Lu X, Bussen W, Zheng L, Stark JM, Barnes EL, Chi P, Janscak P et al. 2007. RECQL5/Recql5 helicase regulates homologous recombination and suppresses tumor formation via disruption of Rad51 presynaptic filaments. Genes and Development 21: 3073-3084. Ingraham SE, Lynch RA, Kathiresan S, Buckler AJ, Menon AG. 1999. hREC2, a RAD51-like Gene, is Disrupted by t(12;14)(q15;q24.1) in a Uterine Leiomyoma. Cancer Genet. Cytogenet. 115: 56-61. Ira G, Malkova A, Liberi G, Foiani M, Haber JE. 2003. Srs2 and Sgs1–Top3 Suppress Crossovers during Double-Strand Break Repair in Yeast. Cell 115: 401-411.

203

Jensen RB, Carreira A, Kowalczykowski SC. 2010. Purified human BRCA2 stimulates RAD51-mediated recombination. Nature 467: 678-683. Johnson R, Symington L. 1995. Functional differences and interactions among the putative RecA homologs Rad51, Rad55, and Rad57. Molecular and Cellular Biology 15: 4843-4850. Kanaar R, Hoeijmakers JHJ. 1998. Genetic recombination: From competition to collaboration. Nature 391: 335-338. Kee Y, D'Andrea AD. 2010. Expanded roles of the Fanconi anemia pathway in preserving genomic stability. Genes Dev. 24: 1680-1694. Kowalczykowski SC. 1991. Biochemistry of genetic recombination: energetics and mechanism of DNA strand exchange. Annual Review of Biophysics and Biophysical Chemistry 20: 539-575. Krejci L, Van Komen S, Li Y, Villemain J, Reddy MS, Klein H, Ellenberger T, Sung P. 2003. DNA helicase Srs2 disrupts the Rad51 presynaptic filament. Nature 423: 305-309. Krogh BO, Symington LS. 2004. Recombination proteins in yeast. Annual Review of Genetics 38: 233-271. Levy-Lahad E. 2010. Fanconi anemia and breast cancer susceptibility meet again. Nat. Genet. 42: 368-369. Lin Z, Kong H, Nei M, Ma H. 2006. Origins and evolution of the recA/RAD51 gene family: evidence for ancient gene duplication and endosymbiotic gene transfer. Proceedings of the National Academy of Sciences of the United States of America 103: 10328-10333. Lio YC, Mazin AV, Kowalczykowski SC, Chen DJ. 2003. Complex formation by the human Rad51B and Rad51C DNA repair proteins and their activities in vitro. Journal of Biological Chemistry 278: 2469-2478. Liu J, Renault L, Veaute X, Fabre F, Stahlberg H, Heyer W-D. 2011. Rad51 paralogues Rad55-Rad57 balance the antirecombinase Srs2 in Rad51 filament formation. Nature advance online publication. Liu N, Lamerdin JE, Tebbs RS, Schild D, Tucker JD, Shen MR, Brookman KW, Siciliano MJ, Walter CA, Fan W et al. 1998. XRCC2 and XRCC3, new human Rad51-family members, promote chromosome stability and protect against DNA cross-links and other damages. Molecular Cell 1: 783-793. Liu Y, Stasiak AZ, Masson JY, McIlwraith MJ, Stasiak A, West SC. 2004. Conformational changes modulate the activity of human RAD51 protein. Journal of Molecular Biology 337: 817-827. Mankouri HW, Craig TJ, Morgan A. 2002. SGS1 is a multicopy suppressor of srs2: functional overlap between DNA helicases. Nucleic Acids Res. 30: 1103-1113. Masson JY, Stasiak AZ, Stasiak A, Benson FE, West SC. 2001a. Complex formation by the human RAD51C and XRCC3 recombination repair proteins. Proceedings of the National Academy of Sciences of the United States of America 98: 8440-8446. Masson JY, Tarsounas MC, Stasiak AZ, Stasiak A, Shah R, McIlwraith MJ, Benson FE, West SC. 2001b. Identification and purification of two distinct complexes containing the five RAD51 paralogs. Genes and Development 15: 3296-3307.

204

Meindl A, Hellebrand H, Wiek C, Erven V, Wappenschmidt B, Niederacher D, Freund M, Lichtner P, Hartmann L, Schaal H et al. 2010. Germline mutations in breast and ovarian cancer pedigrees establish RAD51C as a human cancer susceptibility gene. Nat. Genet. 42: 410-414. New JH, Sugiyama T, Zaitseva E, Kowalczykowski SC. 1998. Rad52 protein stimulates DNA strand exchange by Rad51 and replication protein A. Nature 391: 407-410. Ogawa T, Yu X, Shinohara A, Egelman E. 1993. Similarity of the yeast RAD51 filament to the bacterial RecA filament. Science 259: 1896-1899. Pittman DL, Schimenti JC. 2000. Midgestation lethality in mice deficient for the RecA- related gene, Rad51d/Rad51l3. genesis 26: 167-173. Pittman DL, Weinberg LR, Schimenti JC. 1998. Identification, Characterization, and Genetic Mapping ofRad51d,a New Mouse and HumanRAD51/RecA-Related Gene. Genomics 49: 103-111. Rice MC, Smith ST, Bullrich F, Havre P, Kmiec EB. 1997. Isolation of human and mouse genes based on homology to REC2, a recombinational repair gene from the fungus‚ÄâUstilago‚Äâmaydis. Proc. Natl. Acad. Sci. U. S. A. 94: 7417-7422. Ritchie MF, Zhou Y, Soboloff J. 2011. Transcriptional mechanisms regulating Ca2+ homeostasis. Cell Calcium 49: 314-321. Roy A, Kucukural A, Zhang Y. 2010. I-TASSER: a unified platform for automated and function prediction. Nat. Protocols 5: 725-738. San Filippo J, Sung P, Klein H. 2008. Mechanism of eukaryotic homologous recombination. Annual Review of Biochemistry 77: 229-257. Schild D, Lio YC, Collins DW, Tsomondo T, Chen DJ. 2000. Evidence for simultaneous protein interactions between human Rad51 paralogs. J. Biol. Chem. 275: 16443- 16449. Schoenmakers EFPM, Huysmans C, Van de Ven WJM. 1999. Allelic Knockout of Novel Splice Variants of Human Recombination Repair Gene RAD51B in t(12;14) Uterine Leiomyomas. Cancer Res. 59: 19-23. Scully R, Chen J, Ochs RL, Keegan K, Hoekstra M, Feunteun J, Livingston DM. 1997. Dynamic Changes of BRCA1 Subnuclear Location and Phosphorylation State Are Initiated by DNA Damage. Cell 90: 425-435. Shim KS, Schmutte C, Tombline G, Heinen CD, Fishel R. 2004. hXRCC2 enhances ADP/ATP processing and strand exchange by hRAD51. Journal of Biological Chemistry 279: 30385-30394. Shim KS, Schmutte C, Yoder K, Fishel R. 2006. Defining the salt effect on human RAD51 activities. DNA Repair (Amst) 5: 718-730. Shinohara A, Ogawa T. 1998. Stimulation by Rad52 of yeast Rad51- mediated recombination. Nature 391: 404-407. Shu Z, Smith S, Wang L, Rice MC, Kmiec EB. 1999. Disruption of muREC2/RAD51L1 in Mice Results in Early Embryonic Lethality Which Can Be Partially Rescued in a p53-/- Background. Mol. Cell. Biol. 19: 8686-8693. Sigurdsson S, Van Komen S, Bussen W, Schild D, Albala JS, Sung P. 2001. Mediator function of the human Rad51B-Rad51C complex in Rad51/RPA-catalyzed DNA strand exchange. Genes and Development 15: 3308-3318.

205

Sonoda E, Sasaki MS, Buerstedde JM, Bezzubova O, Shinohara A, Ogawa H, Takata M, Yamaguchi-Iwai Y, Takeda S. 1998. Rad51-deficient vertebrate cells accumulate chromosomal breaks prior to cell death. EMBO Journal 17: 598-608. Story R, Bishop D, Kleckner N, Steitz T. 1993. Structural relationship of bacterial RecA proteins to recombination proteins from bacteriophage T4 and yeast. Science 259: 1892-1896. Sugiyama T, Zaitseva EM, Kowalczykowski SC. 1997. A single-stranded DNA-binding protein is needed for efficient presynaptic complex formation by the Saccharomyces cerevisiae Rad51 protein. Journal of Biological Chemistry 272: 7940-7945. Sung P. 1997. Yeast Rad55 and Rad57 proteins form a heterodimer that functions with replication protein A to promote DNA strand exchange by Rad51 recombinase. Genes and Development 11: 1111-1121. Symington LS. 2002. Role of RAD52 Epistasis Group Genes in Homologous Recombination and Double-Strand Break Repair. Microbiology and Molecular Biology Reviews 66: 630-670. Takata M, Sasaki MS, Sonoda E, Fukushima T, Morrison C, Albala JS, Swagemakers SMA, Kanaar R, Thompson LH, Takeda S. 2000. The Rad51 Paralog Rad51B Promotes Homologous Recombinational Repair. Molecular and Cellular Biology 20: 6476-6482. Takata M, Sasaki MS, Tachiiri S, Fukushima T, Sonoda E, Schild D, Thompson LH, Takeda S. 2001. Chromosome Instability and Defective Recombinational Repair in Knockout Mutants of the Five Rad51 Paralogs. Molecular and Cellular Biology 21: 2858-2866. Tarsounas M, Davies D, West SC. 2003. BRCA2-dependent and independent formation of RAD51 nuclear foci. Oncogene 22: 1115-1123. Tebbs RS, Zhao Y, Tucker JD, Scheerer JB, Siciliano MJ, Hwang M, Liu N, Legerski RJ, Thompson LH. 1995. Correction of chromosomal instability and sensitivity to diverse mutagens by a cloned cDNA of the XRCC3 DNA repair gene. Proc. Natl. Acad. Sci. U. S. A. 92: 6354-6358. Thompson LH, Schild D. 1999. The contribution of homologous recombination in preserving genome integrity in mammalian cells. Biochimie 81: 87-105. Tombline G, Fishel R. 2002. Biochemical characterization of the human RAD51 protein. I. ATP hydrolysis. Journal of Biological Chemistry 277: 14417-14425. Tsuzuki T, Fujii Y, Sakumi K, Tominaga Y, Nakao K, Sekiguchi M, Matsushiro A, Yoshimura Y, MoritaT. 1996. Targeted disruption of the Rad51 gene leads to lethality in embryonic mice. Proceedings of the National Academy of Sciences of the United States of America 93: 6236-6240. Van Komen S, Petukhova G, Sigurdsson S, Sung P. 2002. Functional cross-talk among Rad51, Rad54, and replication protein A in heteroduplex DNA joint formation. Journal of Biological Chemistry 277: 43578-43587. Vaz F, Hanenberg H, Schuster B, Barker K, Wiek C, Erven V, Neveling K, Endt D, Kesterton I, Autore F et al. 2010. Mutation of the RAD51C gene in a Fanconi anemia-like disorder. Nat. Genet. 42: 406-409.

206

Veaute X, Jeusset J, Soustelle C, Kowalczykowski SC, Le Cam E, Fabre F. 2003. The Srs2 helicase prevents recombination by disrupting Rad51 nucleoprotein filaments. Nature 423: 309-312. West SC. 2003. Molecular views of recombination proteins and their control. Nat Rev Mol Cell Biol 4: 435-445. Wu L, Hickson ID. 2006. DNA Helicases Required for Homologous Recombination and Repair of Damaged Replication Forks. Annual Review of Genetics 40: 279-306. Wu Y, Qian X, He Y, Moya IA, Luo Y. 2005. Crystal structure of an ATPase-active form of Rad51 homolog from Methanococcus voltae. Insights into potassium dependence. Journal of Biological Chemistry 280: 722-728. Yang H, Li Q, Fan J, Holloman WK, Pavletich NP. 2005. The BRCA2 homologue Brh2 nucleates RAD51 filament formation at a dsDNA-ssDNA junction. Nature 433: 653-657.

207

CHAPTER 5

Conclusion

208

Formation and maintenance of a stable RAD51 nucleoprotein filament (NPF) is essential for efficient homologous recombination in eukaryotes (Bugreev and Mazin 2004; Chi et al. 2006; San Filippo et al. 2008). In this thesis I have addressed by biochemical and structural analyses how the structural elements of the ATP binding interface may have evolved to regulate the stability of the NPF. I have also tested the hypothesis that the

RAD51 paralog complex RAD51B-RAD51C may stabilize the NPF to enhance the recombinase function.

In Chapter 2, I examined two conserved residues in the ATP binding interface,

F129 of the Walker A box and H294 in the L2 ssDNA binding region, and found they are essential for the recombinase function of RAD51. Substitution mutations of these two residues lead to loss or reduced ATPase activity and complete abrogation of recombinase function. I also found that the F129 and H294 residues are essential RAD51 DNA binding in the presence of ATP, implying allosteric communication between the ATP binding and hydrolysis site and the L2 ssDNA binding region.

In chapter 3, I examined an intriguing structural motif in the ATP binding interface termed the ATP cap that appears to function in the regulation of RAD51 turnover from DNA. An aspartate residue resides at position 316 of the ATP cap in

RAD51. Substitution of lysine for this conserved aspartate confers reduced turnover and enhanced recombinase efficiency. We speculated that in complicated eukaryotic genomes the rapid turnover of RAD51 from postsynaptic structures might facilitate downstream homologous recombination repair events. Based on primary sequence analyses we also predicted that some of the RAD51 paralogs containing a lysine residue in the analogous ATP cap position might function to stabilize the NPF.

209

In chapter 4, I describe several biochemical functions of the RAD51 paralog complex RAD51B-RAD51C that ultimately stimulate RAD51 mediated D-loop formation. Modeled structures indicate that RAD51, RAD51B and RAD51C could potentially bind to DNA with similar domain orientations. Furthermore, I showed that

RAD51B-RAD51C form stable complexes on DNA and can partially stabilize the

RAD51 NPF against anti-recombinase activity of BLM. At sub-stoichiometric amounts,

RAD51B-RAD51C could stimulate RAD51-mediated D-loop formation in the presence of ssDNA binding protein RPA.

Taken as a whole these studies suggest that RAD51 may have evolved faster turnover from postsynaptic DNA structures, at the expense of recombinases efficiency.

RAD51 paralogs, that are available in eukaryotes may compensate for the loss in efficiency by functioning as mediators to form stable NPF and enhance strand exchange activity. Even though, our results with RAD51B-RAD51C indicate partial stabilization, we cannot rule out the possibility that post-translational modifications of RAD51 or the

RAD51 paralogs, or the involvement of other RAD51 paralogs and/or mediators, might augment RAD51 function in vivo. In vertebrates, BRCA2 has been shown to nucleate

RAD51 on RPA coated ssDNA (Carreira et al. 2009; Jensen et al. 2010; Carreira and

Kowalczykowski 2011). During double strand break (DSB) repair by homologous recombination (HR) extensive ssDNA regions are form by exonucleolytic resection that ultimately function as substrates for the homology search and strand exchange reactions

(Symington and Gautier 2011). Our results suggest the possibility that RAD51 paralogs such as RAD51B-RAD51C may interspersedly bind to ssDNA further away from the

210

nucleation site to continue growth and/or stabilization of the RAD51 NPF to promote an efficient homology search and strand exchange process.

211

REFERENCES

Bugreev DV, Mazin AV. 2004. Ca2+ activates human homologous recombination protein Rad51 by modulating its ATPase activity. Proc. Natl. Acad. Sci. U. S. A. 101: 9988-9993. Carreira A, Hilario J, Amitani I, Baskin RJ, Shivji MK, Venkitaraman AR, Kowalczykowski SC. 2009. The BRC repeats of BRCA2 modulate the DNA- binding selectivity of RAD51. Cell 136: 1032-1043. Carreira A, Kowalczykowski SC. 2011. Two classes of BRC repeats in BRCA2 promote RAD51 nucleoprotein filament function by distinct mechanisms. Proc. Natl. Acad. Sci. U. S. A. 108: 10448-10453. Chi P, Van Komen S, Sehorn MG, Sigurdsson S, Sung P. 2006. Roles of ATP binding and ATP hydrolysis in human Rad51 recombinase function. DNA Repair (Amst) 5: 381-391. Jensen RB, Carreira A, Kowalczykowski SC. 2010. Purified human BRCA2 stimulates RAD51-mediated recombination. Nature 467: 678-683. San Filippo J, Sung P, Klein H. 2008. Mechanism of eukaryotic homologous recombination. Annu. Rev. Biochem. 77: 229-257. Symington LS, Gautier J. 2011. Double-Strand Break End Resection and Repair Pathway Choice. Annu. Rev. Genet. 45: 247-271.

212

COMPLETE BIBLIOGRAPHY

Aboussekhra A, Chanet R, Adjiri A, Fabre F. 1992. Semidominant suppressors of Srs2 helicase mutations of Saccharomyces cerevisiae map in the RAD51 gene, whose sequence predicts a protein with similarities to procaryotic RecA proteins. Mol. Cell. Biol. 12: 3224-3234. Agarwal S, van Cappellen WA, Guénolé A, Eppink B, Linsen SEV, Meijering E, Houtsmuller A, Kanaar R, Essers J. 2011. ATP-dependent and independent functions of Rad54 in genome maintenance. J. Cell Biol. 192: 735-750. Ahmad F, Stewart E. 2005. The N-terminal region of the Schizosaccharomyces pombe RecQ helicase, Rqh1p, physically interacts with Topoisomerase III and is required for Rqh1p function. Mol Genet Genomics 273: 102-114. Aihara H, Ito Y, Kurumizaka H, Terada T, Yokoyama S, Shibata T. 1997. An interaction between a specified surface of the C-terminal domain of RecA protein and double-stranded DNA for homologous pairing. J. Mol. Biol. 274: 213-221. Aihara H, Ito Y, Kurumizaka H, Yokoyama S, Shibata T. 1999. The N-terminal domain of the human Rad51 protein binds DNA: structure and a DNA binding surface as revealed by NMR. J. Mol. Biol. 290: 495-504. Alani E, Padmore R, Kleckner N. 1990. Analysis of wild-type and rad50 mutants of yeast suggests an intimate relationship between meiotic chromosome synapsis and recombination. Cell 61: 419-436. Alani E, Thresher R, Griffith JD, Kolodner RD. 1992. Characterization of DNA-binding and strand-exchange stimulation properties of y-RPA, a yeast single-strand-DNA- binding protein. J. Mol. Biol. 227: 54-71. Albala JS, Thelen MP, Prange C, Fan W, Christensen M, Thompson LH, Lennon GG. 1997. Identification of a Novel HumanRAD51Homolog,RAD51B. Genomics 46: 476-479. Alexeev A, Mazin A, Kowalczykowski SC. 2003. Rad54 protein possesses chromatin- remodeling activity stimulated by the Rad51-ssDNA nucleoprotein filament. Nat Struct Mol Biol 10: 182-186. Alexiadis V, Kadonaga JT. 2002. Strand pairing by Rad54 and Rad51 is enhanced by chromatin. Genes Dev. 16: 2767-2771. Amunugama R, Fishel R. 2011. Subunit Interface Residues F129 and H294 of Human RAD51 Are Essential for Recombinase Function. PLoS ONE 6: e23071. Andersen SL, Bergstralh DT, Kohl KP, LaRocque JR, Moore CB, Sekelsky J. 2009. Drosophila MUS312 and the vertebrate ortholog BTBD12 interact with DNA structure-specific endonucleases in DNA repair and recombination. Mol. Cell 35: 128-135.

213

Andreassen PR, Ho GPH, D'Andrea AD. 2006. DNA damage responses and their many interactions with the replication fork. Carcinogenesis 27: 883-892. Antony E, Tomko EJ, Xiao Q, Krejci L, Lohman TM, Ellenberger T. 2009. Srs2 Disassembles Rad51 Filaments by a Protein-Protein Interaction Triggering ATP Turnover and Dissociation of Rad51 from DNA. Mol. Cell 35: 105-115. Ariyoshi M, Nishino T, Iwasaki H, Shinagawa H, Morikawa K. 2000. Crystal structure of the holliday junction DNA in complex with a single RuvA tetramer. Proc. Natl. Acad. Sci. U. S. A. 97: 8257-8262. Azar B. 2011. QnAs with James E. Haber. Proc. Natl. Acad. Sci. U. S. A. 108: 5479. Bashkirov VI, King JS, Bashkirova EV, Schmuckli-Maurer J, Heyer W-D. 2000. DNA Repair Protein Rad55 Is a Terminal Substrate of the DNA Damage Checkpoints. Mol. Cell. Biol. 20: 4393-4404. Basile G, Aker M, Mortimer RK. 1992. Nucleotide sequence and transcriptional regulation of the yeast recombinational repair gene RAD51. Mol. Cell. Biol. 12: 3235-3246. Baumann P, Benson FE, Hajibagheri N, West SC. 1997. Purification of human Rad51 protein by selective spermidine precipitation. Mutat. Res. 384: 65-72. Baumann P, Benson FE, West SC. 1996. Human Rad51 protein promotes ATP- dependent homologous pairing and strand transfer reactions in vitro. Cell 87: 757- 766. Baumann P, West SC. 1998. Role of the human RAD51 protein in homologous recombination and double-stranded-break repair. Trends Biochem. Sci. 23: 247- 251. Bekker-Jensen S, Rendtlew Danielsen J, Fugger K, Gromova I, Nerstedt A, Lukas C, Bartek J, Lukas J, Mailand N. 2010. HERC2 coordinates ubiquitin-dependent assembly of DNA repair factors on damaged chromosomes. Nat Cell Biol 12: 80- 86; sup pp 81-12. Bennett RJ, Noirot-Gros MF, Wang JC. 2000. Interaction between yeast sgs1 helicase and DNA topoisomerase III. J. Biol. Chem. 275: 26898-26905. Benson FE, Stasiak A, West SC. 1994. Purification and characterization of the human Rad51 protein, an analogue of E. coli RecA. EMBO J. 13: 5764-5771. Bernstein KA, Gangloff S, Rothstein R. 2010. The RecQ DNA helicases in DNA repair. Annu. Rev. Genet. 44: 393-417. Bernstein KA, Reid RJD, Sunjevaric I, Demuth K, Burgess RC, Rothstein R. 2011. The Shu complex, which contains Rad51 paralogues, promotes DNA repair through inhibition of the Srs2 anti-recombinase. Mol. Biol. Cell 22: 1599-1607. Bernstein KA, Rothstein R. 2009. At loose ends: resecting a double-strand break. Cell 137: 807-810. Bezzubova O, Silbergleit A, Yamaguchi-Iwai Y, Takeda S, Buerstedde J-M. 1997. Reduced X-Ray Resistance and Homologous Recombination Frequencies in a RAD54-/- Mutant of the Chicken DT40 Cell Line. Cell 89: 185-193. Bhattacharyya S, Keirsey J, Russell B, Kavecansky J, Lillard-Wetherell K, Tahmaseb K, Turchi JJ, Groden J. 2009. Telomerase-associated Protein 1, HSP90, and Topoisomerase II Associate Directly with the BLM Helicase in Immortalized

214

Cells Using ALT and Modulate Its Helicase Activity Using Telomeric DNA Substrates. J. Biol. Chem. 284: 14966-14977. Bianco PR, Tracy RB, Kowalczykowski SC. 1998. DNA strand exchange proteins: a biochemical and physical comparison. Front. Biosci. 3: D570-603. Bignell G, Micklem G, Stratton MR, Ashworth A, Wooster R. 1997. The BRC Repeats are Conserved in Mammalian BRCA2 Proteins. Hum. Mol. Genet. 6: 53-58. Bird AW, Yu DY, Pray-Grant MG, Qiu Q, Harmon KE, Megee PC, Grant PA, Smith MM, Christman MF. 2002. Acetylation of histone H4 by Esa1 is required for DNA double-strand break repair. Nature 419: 411-415. Bishop DK. 1994. RecA homologs Dmc1 and Rad51 interact to form multiple nuclear complexes prior to meiotic chromosome synapsis. Cell 79: 1081-1092. Bishop DK, Park D, Xu L, Kleckner N. 1992. DMC1: A meiosis-specific yeast homolog of E. coli recA required for recombination, synaptonemal complex formation, and cell cycle progression. Cell 69: 439-456. Bochar DA, Wang L, Beniya H, Kinev A, Xue Y, Lane WS, Wang W, Kashanchi F, Shiekhattar R. 2000. BRCA1 Is Associated with a Human SWI/SNF-Related Complex: Linking Chromatin Remodeling to Breast Cancer. Cell 102: 257-265. Boddy MN, Gaillard PH, McDonald WH, Shanahan P, Yates JR, 3rd, Russell P. 2001. Mus81-Eme1 are essential components of a Holliday junction resolvase. Cell 107: 537-548. Bork P, Blomberg N, Nilges M. 1996. Internal repeats in the BRCA2 protein sequence. Nat. Genet. 13: 22-23. Botuyan MV, Lee J, Ward IM, Kim JE, Thompson JR, Chen J, Mer G. 2006. Structural basis for the methylation state-specific recognition of histone H4-K20 by 53BP1 and Crb2 in DNA repair. Cell 127: 1361-1373. Boundy-Mills KL, Livingston DM. 1993. A Saccharomyces cerevisiae RAD52 allele expressing a C-terminal truncation protein: activities and intragenic complementation of missense mutations. Genetics 133: 39-49. Branzei D, Foiani M. 2007. Interplay of replication checkpoints and repair proteins at stalled replication forks. DNA Repair (Amst) 6: 994-1003. -. 2008. Regulation of DNA repair throughout the cell cycle. Nat Rev Mol Cell Biol 9: 297-308. Braybrooke JP, Li J-L, Wu L, Caple F, Benson FE, Hickson ID. 2003. Functional Interaction between the Bloom's Syndrome Helicase and the RAD51 Paralog, RAD51L3 (RAD51D). J. Biol. Chem. 278: 48357-48366. Braybrooke JP, Spink KG, Thacker J, Hickson ID. 2000. The RAD51 Family Member, RAD51L3, Is a DNA-stimulated ATPase That Forms a Complex with XRCC2. J. Biol. Chem. 275: 29100-29106. Breeden L, Nasmyth K. 1987. Cell cycle control of the yeast HO gene: Cis- and Trans- acting regulators. Cell 48: 389-397. Bugreev DV, Golub EI, Stasiak AZ, Stasiak A, Mazin AV. 2005. Activation of human meiosis-specific recombinase Dmc1 by Ca2+. J. Biol. Chem. 280: 26886-26895. Bugreev DV, Mazin AV. 2004. Ca2+ activates human homologous recombination protein Rad51 by modulating its ATPase activity. Proc. Natl. Acad. Sci. U. S. A. 101: 9988-9993.

215

Bugreev DV, Mazina OM, Mazin AV. 2006. Rad54 protein promotes branch migration of Holliday junctions. Nature 442: 590-593. Bugreev DV, Yu X, Egelman EH, Mazin AV. 2007. Novel pro- and anti-recombination activities of the Bloom's syndrome helicase. Genes Dev. 21: 3085-3094. Buisson R, Dion-Cote A-M, Coulombe Y, Launay H, Cai H, Stasiak AZ, Stasiak A, Xia B, Masson J-Y. 2010. Cooperation of breast cancer proteins PALB2 and piccolo BRCA2 in stimulating homologous recombination. Nat Struct Mol Biol 17: 1247- 1254. Bzymek M, Thayer NH, Oh SD, Kleckner N, Hunter N. 2010. Double Holliday junctions are intermediates of DNA break repair. Nature 464: 937-941. Cai Y, Jin J, Tomomori-Sato C, Sato S, Sorokina I, Parmely TJ, Conaway RC, Conaway JW. 2003. Identification of New Subunits of the Multiprotein Mammalian TRRAP/TIP60-containing Histone Acetyltransferase Complex. J. Biol. Chem. 278: 42733-42736. Cairns BR, Kim YJ, Sayre MH, Laurent BC, Kornberg RD. 1994. A multisubunit complex containing the SWI1/ADR6, SWI2/SNF2, SWI3, SNF5, and SNF6 gene products isolated from yeast. Proc. Natl. Acad. Sci. U. S. A. 91: 1950-1954. Cairns BR, Lorch Y, Li Y, Zhang M, Lacomis L, Erdjument-Bromage H, Tempst P, Du J, Laurent B, Kornberg RD. 1996. RSC, an Essential, Abundant Chromatin- Remodeling Complex. Cell 87: 1249-1260. Carreira A, Hilario J, Amitani I, Baskin RJ, Shivji MK, Venkitaraman AR, Kowalczykowski SC. 2009. The BRC repeats of BRCA2 modulate the DNA- binding selectivity of RAD51. Cell 136: 1032-1043. Carreira A, Kowalczykowski SC. 2011. Two classes of BRC repeats in BRCA2 promote RAD51 nucleoprotein filament function by distinct mechanisms. Proc. Natl. Acad. Sci. U. S. A. 108: 10448-10453. Cartwright R, Dunn AM, Simpson PJ, Tambini CE, Thacker J. 1998. Isolation of novel human and mouse genes of the recA/RAD51 recombination-repair gene family. Nucleic Acids Res 26: 1653-1659. Cejka P, Cannavo E, Polaczek P, Masuda-Sasa T, Pokharel S, Campbell JL, Kowalczykowski SC. 2010. DNA end resection by Dna2-Sgs1-RPA and its stimulation by Top3-Rmi1 and Mre11-Rad50-Xrs2. Nature 467: 112-116. Celeste A, Petersen S, Romanienko PJ, Fernandez-Capetillo O, Chen HT, Sedelnikova OA, Reina-San-Martin B, Coppola V, Meffre E, Difilippantonio MJ et al. 2002. Genomic instability in mice lacking histone H2AX. Science 296: 922-927. Chai B, Huang J, Cairns BR, Laurent BC. 2005. Distinct roles for the RSC and Swi/Snf ATP-dependent chromatin remodelers in DNA double-strand break repair. Genes Dev. 19: 1656-1661. Chang M, Bellaoui M, Zhang C, Desai R, Morozov P, Delgado-Cruzata L, Rothstein R, Freyer GA, Boone C, Brown GW. 2005. RMI1/NCE4, a suppressor of genome instability, encodes a member of the RecQ helicase/Topo III complex. EMBO J. 24: 2024-2033. Chapman JR, Jackson SP. 2008. Phospho-dependent interactions between NBS1 and MDC1 mediate chromatin retention of the MRN complex at sites of DNA damage. EMBO Rep 9: 795-801.

216

Chen C-F, Chen P-L, Zhong Q, Sharp ZD, Lee W-H. 1999. Expression of BRC Repeats in Breast Cancer Cells Disrupts the BRCA2-Rad51 Complex and Leads to Radiation Hypersensitivity and Loss of G2/M Checkpoint Control. J. Biol. Chem. 274: 32931-32935. Chen P-L, Chen C-F, Chen Y, Xiao J, Sharp ZD, Lee W-H. 1998. The BRC repeats in BRCA2 are critical for RAD51 binding and resistance to methyl methanesulfonate treatment. Proc. Natl. Acad. Sci. U. S. A. 95: 5287-5292. Chen XB, Melchionna R, Denis CM, Gaillard PH, Blasina A, Van de Weyer I, Boddy MN, Russell P, Vialard J, McGowan CH. 2001. Human Mus81-associated endonuclease cleaves Holliday junctions in vitro. Mol. Cell 8: 1117-1127. Chen Z, Yang H, Pavletich NP. 2008. Mechanism of homologous recombination from the RecA-ssDNA/dsDNA structures. Nature 453: 489-484. Chi P, Van Komen S, Sehorn MG, Sigurdsson S, Sung P. 2006. Roles of ATP binding and ATP hydrolysis in human Rad51 recombinase function. DNA Repair (Amst) 5: 381-391. Chung W-H, Zhu Z, Papusha A, Malkova A, Ira G. 2010. Defective Resection at DNA Double-Strand Breaks Leads to De Novo Telomere Formation and Enhances Gene Targeting. PLoS Genet 6: e1000948. Ciccia A, Constantinou A, West SC. 2003. Identification and characterization of the human mus81-eme1 endonuclease. J. Biol. Chem. 278: 25172-25178. Ciccia A, Elledge SJ. 2010. The DNA damage response: making it safe to play with knives. Mol. Cell 40: 179-204. Clerici M, Mantiero D, Lucchini G, Longhese MP. 2005. The Saccharomyces cerevisiae Sae2 protein promotes resection and bridging of double strand break ends. J. Biol. Chem. 280: 38631-38638. Clever B, Interthal H, Schmuckli-Maurer J, King J, Sigrist M, Heyer W-D. 1997. Recombinational repair in yeast: functional interactions between Rad51 and Rad54 proteins. EMBO J. 16: 2535-2544. Clever B, Schmuckli-Maurer J, Sigrist M, Glassner BJ, Heyer W-D. 1999. Specific negative effects resulting from elevated levels of the recombinational repair protein Rad54p in Saccharomyces cerevisiae. Yeast 15: 721-740. Cole GM, Schild D, Lovett ST, Mortimer RK. 1987. Regulation of RAD54- and RAD52- lacZ gene fusions in Saccharomyces cerevisiae in response to DNA damage. Mol. Cell. Biol. 7: 1078-1084. Cole GM, Schild D, Mortimer RK. 1989. Two DNA repair and recombination genes in Saccharomyces cerevisiae, RAD52 and RAD54, are induced during meiosis. Mol. Cell. Biol. 9: 3101-3104. Compton SA, Ӧzgür S, Griffith JD. 2010b. Ring-shaped Rad51 Paralog Protein Complexes Bind Holliday Junctions and Replication Forks as Visualized by Electron Microscopy. J. Biol. Chem. 285: 13349-13356. Conant GC, Wolfe KH. 2008. Turning a hobby into a job: How duplicated genes find new functions. Nat Rev Genet 9: 938-950. Conaway RC, Conaway JW. 2009. The INO80 chromatin remodeling complex in transcription, replication and repair. Trends Biochem. Sci. 34: 71-77.

217

Connolly B, Parsons CA, Benson FE, Dunderdale HJ, Sharples GJ, Lloyd RG, West SC. 1991. Resolution of Holliday junctions in vitro requires the Escherichia coli ruvC gene product. Proc. Natl. Acad. Sci. U. S. A. 88: 6063-6067. Connolly B, West SC. 1990. Genetic recombination in Escherichia coli: Holliday junctions made by RecA protein are resolved by fractionated cell-free extracts. Proc. Natl. Acad. Sci. U. S. A. 87: 8476-8480. Constantinou A, Davies AA, West SC. 2001. Branch migration and Holliday junction resolution catalyzed by activities from mammalian cells. Cell 104: 259-268. Conway AB, Lynch TW, Zhang Y, Fortin GS, Fung CW, Symington LS, Rice PA. 2004. Crystal structure of a Rad51 filament. Nat Struct Mol Biol 11: 791-796. Cook PJ, Ju BG, Telese F, Wang X, Glass CK, Rosenfeld MG. 2009. Tyrosine dephosphorylation of H2AX modulates apoptosis and survival decisions. Nature 458: 591-596. Cote J, Quinn J, Workman J, Peterson C. 1994. Stimulation of GAL4 derivative binding to nucleosomal DNA by the yeast SWI/SNF complex. Science 265: 53-60. Cox MM, Lehman IR. 1981. recA protein of Escherichia coli promotes branch migration, a kinetically distinct phase of DNA strand exchange. Proc. Natl. Acad. Sci. U. S. A. 78: 3433-3437. Cromie GA, Smith GR. 2007. Branching out: meiotic recombination and its regulation. Trends Cell Biol. 17: 448-455. Cui X, Brenneman M, Meyne J, Oshimura M, Goodwin EH, Chen DJ. 1999. The XRCC2 and XRCC3 repair genes are required for chromosome stability in mammalian cells. Mutat. Res. 434: 75-88. D'Andrea AD, Grompe M. 2003. The Fanconi anaemia/BRCA pathway. Nature Reviews. Cancer. 3: 23-34. Das C, Lucia MS, Hansen KC, Tyler JK. 2009. CBP/p300-mediated acetylation of histone H3 on lysine 56. Nature 459: 113-117. Datta A, Adjiri A, New L, Crouse GF, Jinks Robertson S. 1996. Mitotic crossovers between diverged sequences are regulated by mismatch repair proteins in Saccaromyces cerevisiae. Mol. Cell. Biol. 16: 1085-1093. Datta A, Hendrix M, Lipsitch M, Jinks-Robertson S. 1997. Dual roles for DNA sequence identity and the mismatch repair system in the regulation of mitotic crossing-over in yeast. Proc. Natl. Acad. Sci. U. S. A. 94: 9757-9762. Davies AA, Masson J-Y, McIlwraith MJ, Stasiak AZ, Stasiak A, Venkitaraman AR, West SC. 2001. Role of BRCA2 in Control of the RAD51 Recombination and DNA Repair Protein. Mol. Cell 7: 273-282. Davies OR, Pellegrini L. 2007. Interaction with the BRCA2 C terminus protects RAD51- DNA filaments from disassembly by BRC repeats. Nat Struct Mol Biol 14: 475- 483. de los Santos T, Hunter N, Lee C, Larkin B, Loidl J, Hollingsworth NM. 2003. The Mus81/Mms4 endonuclease acts independently of double-Holliday junction resolution to promote a distinct subset of crossovers during meiosis in budding yeast. Genetics 164: 81-94.

218

de los Santos T, Loidl J, Larkin B, Hollingsworth NM. 2001. A role for MMS4 in the processing of recombination intermediates during meiosis in Saccharomyces cerevisiae. Genetics 159: 1511-1525. de Massy B, Studier FW, Dorgai L, Appelbaum E, Weisberg RA. 1984. Enzymes and sites of genetic recombination: studies with gene-3 endonuclease of phage T7 and with site-affinity mutants of phage lambda. Cold Spring Harb. Symp. Quant. Biol. 49: 715-726. De Zutter JK, Knight KL. 1999. The hRad51 and RecA proteins show significant differences in cooperative binding to single-stranded DNA. J. Mol. Biol. 293: 769-780. Deans B, Griffin CS, Maconochie M, Thacker J. 2000. Xrcc2 is required for genetic stability, embryonic neurogenesis and viability in mice. EMBO J. 19: 6675-6685. DeLano WL. 2002. The PyMOL Molecular Graphics System. Palo Alto, CA: DeLano Scientific LLC. Dendouga N, Gao H, Moechars D, Janicot M, Vialard J, McGowan CH. 2005. Disruption of murine Mus81 increases genomic instability and DNA damage sensitivity but does not promote tumorigenesis. Mol. Cell. Biol. 25: 7569-7579. Dingwall AK, Beek SJ, McCallum CM, Tamkun JW, Kalpana GV, Goff SP, Scott MP. 1995. The Drosophila snr1 and brm proteins are related to yeast SWI/SNF proteins and are components of a large protein complex. Mol. Biol. Cell 6: 777- 791. Doil C, Mailand N, Bekker-Jensen S, Menard P, Larsen DH, Pepperkok R, Ellenberg J, Panier S, Durocher D, Bartek J et al. 2009. RNF168 binds and amplifies ubiquitin conjugates on damaged chromosomes to allow accumulation of repair proteins. Cell 136: 435-446. Dosanjh MK, Collins DW, Schild D, Fan W, Lennon GG, Albala JS, Shen Z. 1998. Isolation and characterization of RAD51C, a new human member of the RAD51 family of related genes. Nucleic Acids Res 26: 1179-1184. Downs JA, Allard S, Jobin-Robitaille O, Javaheri A, Auger A, Bouchard N, Kron SJ, Jackson SP, Cote J. 2004. Binding of chromatin-modifying activities to phosphorylated histone H2A at DNA damage sites. Mol. Cell 16: 979-990. Downs JA, Lowndes NF, Jackson SP. 2000. A role for Saccharomyces cerevisiae histone H2A in DNA repair. Nature 408: 1001-1004. Doyon Y, Selleck W, Lane WS, Tan S, Cote J. 2004. Structural and Functional Conservation of the NuA4 Histone Acetyltransferase Complex from Yeast to Humans. Mol. Cell. Biol. 24: 1884-1896. Dray E, Etchin J, Wiese C, Saro D, Williams GJ, Hammel M, Yu X, Galkin VE, Liu D, Tsai M-S et al. 2010. Enhancement of RAD51 recombinase activity by the tumor suppressor PALB2. Nat Struct Mol Biol 17: 1255-1259. Dunderdale HJ, Benson FE, Parsons CA, Sharples GJ, Lloyd RG, West SC. 1991. Formation and resolution of recombination intermediates by E. coli RecA and RuvC proteins. Nature 354: 506-510. Dupaigne P, Le Breton C, Fabre F, Gangloff S, Le Cam E, Veaute X. 2008. The Srs2 Helicase Activity Is Stimulated by Rad51 Filaments on dsDNA: Implications for Crossover Incidence during Mitotic Recombination. Mol. Cell 29: 243-254.

219

Ebbert R, Birkmann A, Schuller HJ. 1999. The product of the SNF2/SWI2 paralogue INO80 of Saccharomyces cerevisiae required for efficient expression of various yeast structural genes is part of a high-molecular-weight protein complex. Mol. Microbiol. 32: 741-751. Eberharter A, Becker PB. 2004. ATP-dependent nucleosome remodelling: factors and functions. J. Cell Sci. 117: 3707-3711. Eggler AL, Inman RB, Cox MM. 2002. The Rad51-dependent pairing of long DNA substrates is stabilized by replication protein A. J. Biol. Chem. 277: 39280-39288. Elborough KM, West SC. 1990. Resolution of synthetic Holliday junctions in DNA by an endonuclease activity from calf thymus. EMBO J. 9: 2931-2936. Ellis NA, Groden J, Ye TZ, Straughen J, Lennon DJ, Ciocci S, Proytcheva M, German J. 1995. The Bloom's syndrome gene product is homologous to RecQ helicases. Cell 83: 655-666. Erdile LF, Wold MS, Kelly TJ. 1990. The primary structure of the 32-kDa subunit of human replication protein A. J. Biol. Chem. 265: 3177-3182. Erzberger JP, Berger JM. 2006. Evolutionary relationships and structural mechanisms of AAA+ proteins. Annu. Rev. Biophys. Biomol. Struct. 35: 93-114. Esashi F, Christ N, Gannon J, Liu Y, Hunt T, Jasin M, West SC. 2005. CDK-dependent phosphorylation of BRCA2 as a regulatory mechanism for recombinational repair. Nature 434: 598-604. Esashi F, Galkin VE, Yu X, Egelman EH, West SC. 2007. Stabilization of RAD51 nucleoprotein filaments by the C-terminal region of BRCA2. Nat Struct Mol Biol 14: 468-474. Essers J, Houtsmuller AB, van Veelen L, Paulusma C, Nigg AL, Pastink A, Vermeulen W, Hoeijmakers JHJ, Kanaar R. 2002. Nuclear dynamics of RAD52 group homologous recombination proteins in response to DNA damage. EMBO J. 21: 2030-2037. Essers J, van Steeg H, de Wit J, Swagemakers SMA, Vermeij M, Hoeijmakers JHJ, Kanaar R. 2000. Homologous and non-homologous recombination differentially affect DNA damage repair in mice. EMBO J. 19: 1703-1710. Farmer H, McCabe N, Lord CJ, Tutt AN, Johnson DA, Richardson TB, Santarosa M, Dillon KJ, Hickson I, Knights C et al. 2005. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 434: 917-921. Fekairi S, Scaglione S, Chahwan C, Taylor ER, Tissier A, Coulon S, Dong MQ, Ruse C, Yates JR, 3rd, Russell P et al. 2009. Human SLX4 is a Holliday junction resolvase subunit that binds multiple DNA repair/recombination endonucleases. Cell 138: 78-89. Feng Z, Scott SP, Bussen W, Sharma GG, Guo G, Pandita TK, Powell SN. 2011. Rad52 inactivation is synthetically lethal with BRCA2 deficiency. Proc. Natl. Acad. Sci. U. S. A. 108: 686-691. Ferguson DO, Holloman WK. 1996. Recombinational repair of gaps in DNA is asymmetric in Ustilago maydis and can be explained by a migrating D-loop model. Proc. Natl. Acad. Sci. U. S. A. 93: 5419-5424.

220

Fishman-Lobell J, Rudin N, Haber JE. 1992. Two alternative pathways of double-strand break repair that are kinetically separable and independently modulated. Mol. Cell. Biol. 12: 1292-1303. Flaus A, Owen-Hughes T. 2003. Dynamic properties of nucleosomes during thermal and ATP-driven mobilization. Mol. Cell. Biol. 23: 7767-7779. Fortin GS, Symington LS. 2002. Mutations in yeast Rad51 that partially bypass the requirement for Rad55 and Rad57 in DNA repair by increasing the stability of Rad51-DNA complexes. EMBO J. 21: 3160-3170. Fricke WM, Brill SJ. 2003. Slx1-Slx4 is a second structure-specific endonuclease functionally redundant with Sgs1-Top3. Genes Dev. 17: 1768-1778. Fricke WM, Kaliraman V, Brill SJ. 2001. Mapping the DNA topoisomerase III binding domain of the Sgs1 DNA helicase. J. Biol. Chem. 276: 8848-8855. Fujimori A, Tachiiri S, Sonoda E, Thompson LH, Dhar PK, Hiraoka M, Takeda S, Zhang Y, Reth M, Takata M. 2001. Rad52 partially substitutes for the Rad51 paralog XRCC3 in maintaining chromosomal integrity in vertebrate cells. EMBO J. 20: 5513-5520. Galanty Y, Belotserkovskaya R, Coates J, Polo S, Miller KM, Jackson SP. 2009. Mammalian SUMO E3-ligases PIAS1 and PIAS4 promote responses to DNA double-strand breaks. Nature 462: 935-939. Galkin VE, Esashi F, Yu X, Yang S, West SC, Egelman EH. 2005. BRCA2 BRC motifs bind RAD51-DNA filaments. Proc. Natl. Acad. Sci. U. S. A. 102: 8537-8542. Game JC, Mortimer RK. 1974. A genetic study of X-ray sensitive mutants in yeast. Mutation Research/Fundamental and Molecular Mechanisms of Mutagenesis 24: 281-292. Gangloff S, McDonald JP, Bendixen C, Arthur L, Rothstein R. 1994. The yeast type I topoisomerase Top3 interacts with Sgs1, a DNA helicase homolog: a potential eukaryotic reverse gyrase. Mol. Cell. Biol. 14: 8391-8398. Garvik B, Carson M, Hartwell L. 1995. Single-stranded DNA arising at telomeres in cdc13 mutants may constitute a specific signal for the RAD9 checkpoint [published erratum appears in Mol Cell Biol 1996 Jan;16(1):457]. Mol. Cell. Biol. 15: 6128-6138. Gasior SL, Wong AK, Kora Y, Shinohara A, Bishop DK. 1998. Rad52 associates with RPA and functions with Rad55 and Rad57 to assemble meiotic recombination complexes. Genes Dev. 12: 2208-2221. Gavin I, Horn PJ, Peterson CL. 2001. SWI/SNF chromatin remodeling requires changes in DNA topology. Mol. Cell 7: 97-104. Gildemeister OS, Sage JM, Knight KL. 2009. Cellular Redistribution of Rad51 in Response to DNA Damage. J. Biol. Chem. 284: 31945-31952. Goggins M, Schutte M, Lu J, Moskaluk CA, Weinstein CL, Petersen GM, Yeo CJ, Jackson CE, Lynch HT, Hruban RH et al. 1996. Germline BRCA2 Gene Mutations in Patients with Apparently Sporadic Pancreatic Carcinomas. Cancer Res. 56: 5360-5364. Golub EI, Kovalenko OV, Gupta RC, Ward DC, Radding CM. 1997. Interaction of human recombination proteins Rad51 and Rad54. Nucleic Acids Res 25: 4106- 4110.

221

Gorbalenya AE, Koonin EV. 1993. Helicases: amino acid sequence comparisons and structure-function relationships. Curr. Opin. Struct. Biol. 3: 419-429. Griffin CS, Simpson PJ, Wilson CR, Thacker J. 2000. Mammalian recombination-repair genes XRCC2 and XRCC3 promote correct chromosome segregation. Nat Cell Biol 2: 757-761. Grigorescu AA, Vissers JH, Ristic D, Pigli YZ, Lynch TW, Wyman C, Rice PA. 2009. Inter-subunit interactions that coordinate Rad51's activities. Nucleic Acids Res 37: 557-567. Gudmundsdottir K, Lord CJ, Witt E, Tutt ANJ, Ashworth A. 2004. DSS1 is required for RAD51 focus formation and genomic stability in mammalian cells. EMBO Rep 5: 989-993. Gupta RC, Folta-Stogniew E, O'Malley S, Takahashi M, Radding CM. 1999. Rapid Exchange of A:T Base Pairs Is Essential for Recognition of DNA Homology by Human Rad51 Recombination Protein. Mol. Cell 4: 705-714. Haaf T, Golub EI, Reddy G, Radding CM, Ward DC. 1995. Nuclear foci of mammalian Rad51 recombination protein in somatic cells after DNA damage and its localization in synaptonemal complexes. Proc. Natl. Acad. Sci. U. S. A. 92: 2298- 2302. Hackett JA, Greider CW. 2003. End Resection Initiates Genomic Instability in the Absence of Telomerase. Mol. Cell. Biol. 23: 8450-8461. Harfe BD, Jinks-Robertson S. 2000. DNA mismatch repair and genetic instability. Annu. Rev. Genet. 34: 359-399. Hargreaves D, Rice DW, Sedelnikova SE, Artymiuk PJ, Lloyd RG, Rafferty JB. 1998. Crystal structure of E.coli RuvA with bound DNA Holliday junction at 6 A resolution. Nat. Struct. Biol. 5: 441-446. Harper JW, Elledge SJ. 2007. The DNA damage response: ten years after. Mol. Cell 28: 739-745. Hartlerode AJ, Scully R. 2009. Mechanisms of double-strand break repair in somatic mammalian cells. Biochem. J. 423: 157-168. Haseltine CA, Kowalczykowski SC. 2009. An archaeal Rad54 protein remodels DNA and stimulates DNA strand exchange by RadA. Nucleic Acids Res 37: 2757-2770. Hashimoto Y, Chaudhuri AR, Lopes M, Costanzo V. 2010. Rad51 protects nascent DNA from Mre11-dependent degradation and promotes continuous DNA synthesis. Nat Struct Mol Biol 17: 1305-1311. Hassold T, Hall H, Hunt P. 2007. The origin of human aneuploidy: where we have been, where we are going. Hum. Mol. Genet. 16 Spec No. 2: R203-208. Hassold T, Sherman S, Hunt PA. 1995. The origin of trisomy in humans. Prog. Clin. Biol. Res. 393: 1-12. Haushalter KA, Kadonaga JT. 2003. Chromatin assembly by DNA-translocating motors. Nat Rev Mol Cell Biol 4: 613-620. Hays SL, Firmenich AA, Berg P. 1995. Complex formation in yeast double-strand break repair: participation of Rad51, Rad52, Rad55, and Rad57 proteins. Proc. Natl. Acad. Sci. U. S. A. 92: 6925-6929.

222

Hays SL, Firmenich AA, Massey P, Banerjee R, Berg P. 1998. Studies of the Interaction between Rad52 Protein and the Yeast Single-Stranded DNA Binding Protein RPA. Mol. Cell. Biol. 18: 4400-4406. Heller RC, Marians KJ. 2006. Replication fork reactivation downstream of a blocked nascent leading strand. Nature 439: 557-562. Henikoff S. 2008. Nucleosome destabilization in the epigenetic regulation of gene expression. Nat Rev Genet 9: 15-26. Henricksen LA, Umbricht CB, Wold MS. 1994. Recombinant replication protein A: expression, complex formation, and functional characterization. J. Biol. Chem. 269: 11121-11132. Herzberg K, Bashkirov VI, Rolfsmeier M, Haghnazari E, McDonald WH, Anderson S, Bashkirova EV, Yates JR, III, Heyer W-D. 2006. Phosphorylation of Rad55 on Serines 2, 8, and 14 Is Required for Efficient Homologous Recombination in the Recovery of Stalled Replication Forks. Mol. Cell. Biol. 26: 8396-8409. Heyer W-D, Li X, Rolfsmeier M, Zhang X-P. 2006. Rad54: the Swiss Army knife of homologous recombination? Nucleic Acids Res 34: 4115-4125. Heyer WD, Ehmsen KT, Liu J. 2010. Regulation of homologous recombination in eukaryotes. Annu. Rev. Genet. 44: 113-139. Heyer WD, Rao MR, Erdile LF, Kelly TJ, Kolodner RD. 1990. An essential Saccharomyces cerevisiae single-stranded DNA binding protein is homologous to the large subunit of human RP-A. EMBO J. 9: 2321-2329. Heyting C. 1996. Synaptonemal Complexes: structure and function. Curr. Biol. 8: 389- 396. Hicks WM, Yamaguchi M, Haber JE. 2011. Real-time analysis of double-strand DNA break repair by homologous recombination. Proc. Natl. Acad. Sci. U. S. A. 108: 3108-3115. Higgins NP, Kato K, Strauss B. 1976. A model for replication repair in mammalian cells. J. Mol. Biol. 101: 417-425. Hilario J, Amitani I, Baskin RJ, Kowalczykowski SC. 2009. Direct imaging of human Rad51 nucleoprotein dynamics on individual DNA molecules. Proc. Natl. Acad. Sci. U. S. A. 106: 361-368. Hoeijmakers JH. 2001. Genome maintenance mechanisms for preventing cancer. Nature 411: 366-374. Holliday RA. 1964. A mechanism for gene conversion in fungi. Genet. Res. 5: 282-304. Hollingsworth NM, Brill SJ. 2004. The Mus81 solution to resolution: generating meiotic crossovers without Holliday junctions. Genes Dev. 18: 117-125. Holloman WK. 2011. Unraveling the mechanism of BRCA2 in homologous recombination. Nat Struct Mol Biol 18: 748-754. Horn PJ, Peterson CL. 2002. Chromatin Higher Order Folding--Wrapping up Transcription. Science 297: 1824-1827. Housworth EA, Stahl FW. 2003. Crossover interference in humans. Am. J. Hum. Genet. 73: 188-197. Epub 2003 May 2022. Howlett NG, Taniguchi T, Olson S, Cox B, Waisfisz Q, de Die-Smulders C, Persky N, Grompe M, Joenje H, Pals G et al. 2002. Biallelic Inactivation of BRCA2 in Fanconi Anemia. Science 297: 606-609.

223

Hu Y, Raynard S, Sehorn MG, Lu X, Bussen W, Zheng L, Stark JM, Barnes EL, Chi P, Janscak P et al. 2007. RECQL5/Recql5 helicase regulates homologous recombination and suppresses tumor formation via disruption of Rad51 presynaptic filaments. Genes Dev. 21: 3073-3084. Huen MS, Chen J. 2010. Assembly of checkpoint and repair machineries at DNA damage sites. Trends Biochem. Sci. 35: 101-108. Huen MS, Grant R, Manke I, Minn K, Yu X, Yaffe MB, Chen J. 2007. RNF8 transduces the DNA-damage signal via histone ubiquitylation and checkpoint protein assembly. Cell 131: 901-914. Huen MS, Sy SM, Chen J. 2010. BRCA1 and its toolbox for the maintenance of genome integrity. Nat Rev Mol Cell Biol 11: 138-148. Huertas P, Cortes-Ledesma F, Sartori AA, Aguilera A, Jackson SP. 2008. CDK targets Sae2 to control DNA-end resection and homologous recombination. Nature 455: 689-692. Huertas P, Jackson SP. 2009. Human CtIP mediates cell cycle control of DNA end resection and double strand break repair. J. Biol. Chem. 284: 9558-9565. Huyen Y, Zgheib O, Ditullio RA, Jr., Gorgoulis VG, Zacharatos P, Petty TJ, Sheston EA, Mellert HS, Stavridi ES, Halazonetis TD. 2004. Methylated lysine 79 of histone H3 targets 53BP1 to DNA double-strand breaks. Nature 432: 406-411. Hyde H, Davies AA, Benson FE, West SC. 1994. Resolution of recombination intermediates by a mammalian activity functionally analogous to Escherichia coli RuvC resolvase. J. Biol. Chem. 269: 5202-5209. Ikura T, Ogryzko VV, Grigoriev M, Groisman R, Wang J, Horikoshi M, Scully R, Qin J, Nakatani Y. 2000. Involvement of the TIP60 histone acetylase complex in DNA repair and apoptosis. Cell 102: 463-473. Ikura T, Tashiro S, Kakino A, Shima H, Jacob N, Amunugama R, Yoder K, Izumi S, Kuraoka I, Tanaka K et al. 2007. DNA damage-dependent acetylation and ubiquitination of H2AX enhances chromatin dynamics. Mol. Cell. Biol. 27: 7028- 7040. Ingraham SE, Lynch RA, Kathiresan S, Buckler AJ, Menon AG. 1999. hREC2, a RAD51-like Gene, is Disrupted by t(12;14)(q15;q24.1) in a Uterine Leiomyoma. Cancer Genet. Cytogenet. 115: 56-61. Ip SC, Rass U, Blanco MG, Flynn HR, Skehel JM, West SC. 2008. Identification of Holliday junction resolvases from humans and yeast. Nature 456: 357-361. Ira G, Malkova A, Liberi G, Foiani M, Haber JE. 2003. Srs2 and Sgs1–Top3 Suppress Crossovers during Double-Strand Break Repair in Yeast. Cell 115: 401-411. Iwasaki H, Takahagi M, Shiba T, Nakata A, Shinagawa H. 1991. Escherichia coli RuvC protein is an endonuclease that resolves the Holliday structure. EMBO J. 10: 4381-4389. Jackson D, Dhar K, Wahl JK, Wold MS, Borgstahl GEO. 2002. Analysis of the Human Replication Protein A:Rad52 Complex: Evidence for Crosstalk Between RPA32, RPA70, Rad52 and DNA. J. Mol. Biol. 321: 133-148. Jackson SP, Bartek J. 2009. The DNA-damage response in human biology and disease. Nature 461: 1071-1078.

224

Janke R, Herzberg K, Rolfsmeier M, Mar J, Bashkirov VI, Haghnazari E, Cantin G, Yates JR, Heyer W-D. 2010. A truncated DNA-damage-signaling response is activated after DSB formation in the G1 phase of Saccharomyces cerevisiae. Nucleic Acids Res 38: 2302-2313. Jasin M. 2002. Homologous repair of DNA damage and tumorigenesis: the BRCA connection. Oncogene 21: 8981-8993. Jaskelioff M, Van Komen S, Krebs JE, Sung P, Peterson CL. 2003. Rad54p Is a Chromatin Remodeling Enzyme Required for Heteroduplex DNA Joint Formation with Chromatin. J. Biol. Chem. 278: 9212-9218. Jensen RB, Carreira A, Kowalczykowski SC. 2010. Purified human BRCA2 stimulates RAD51-mediated recombination. Nature 467: 678-683. Jenuwein T, Allis CD. 2001. Translating the histone code. Science 293: 1074-1080. Jiang H, Xie Y, Houston P, Stemke-Hale K, Mortensen UH, Rothstein R, Kodadek T. 1996. Direct association between the yeast Rad51 and Rad54 recombination proteins. J. Biol. Chem. 271: 33181-33186. Johnson FB, Lombard DB, Neff NF, Mastrangelo MA, Dewolf W, Ellis NA, Marciniak RA, Yin Y, Jaenisch R, Guarente L. 2000. Association of the Bloom syndrome protein with topoisomerase IIIalpha in somatic and meiotic cells. Cancer Res. 60: 1162-1167. Johnson R, Symington L. 1995. Functional differences and interactions among the putative RecA homologs Rad51, Rad55, and Rad57. Mol. Cell. Biol. 15: 4843- 4850. Johnston LH, Johnson AL. 1995. The DNA repair genes RAD54 and UNG1 are cell cycle regulated in budding yeast but MCB promoter elements have no essential role in the DNA damage response. Nucleic Acids Res 23: 2147-2152. Jones NJ, Zhao Y, Siciliano MJ, Thompson LH. 1995. Assignment of the XRCC2 human DNA repair gene to chromosome 7q36 by complementation analysis. Genomics 26: 619-622. Jonsson ZO, Jha S, Wohlschlegel JA, Dutta A. 2004. Rvb1p/Rvb2p recruit Arp5p and assemble a functional Ino80 chromatin remodeling complex. Mol. Cell 16: 465- 477. Kabsch W. 1993. Automatic processing of rotation diffraction data from crystals of initially unknown symmetry and cell constants. J Appl Cryst 26: 795-800. Kagawa W, Kurumizaka H, Ishitani R, Fukai S, Nureki O, Shibata T, Yokoyama S. 2002. Crystal Structure of the Homologous-Pairing Domain from the Human Rad52 Recombinase in the Undecameric Form. Mol. Cell 10: 359-371. Kanaar R, Hoeijmakers JHJ. 1998. Genetic recombination: From competition to collaboration. Nature 391: 335-338. Kanaar R, Troelstra C, Swagemakers SMA, Essers J, Smit B, Franssen J-H, Pastink A, Bezzubova OY, Buerstedde J-M, Clever B et al. 1996. Human and mouse homologs of the Saccharomyces cerevisiae RAD54 DNA repair gene: evidence for functional conservation. Curr. Biol. 6: 828-838. Kans JA, Mortimer RK. 1991. Nucleotide sequence of the RAD57 gene of Saccharomyces cerevisiae. Gene 105: 139-140.

225

Kantake N, Sugiyama T, Kolodner RD, Kowalczykowski SC. 2003. The recombination- deficient mutant RPA (rfa1-t11) is displaced slowly from single-stranded DNA by Rad51 protein. J. Biol. Chem. 278: 23410-23417. Kassabov SR, Zhang B, Persinger J, Bartholomew B. 2003. SWI/SNF unwraps, slides, and rewraps the nucleosome. Mol. Cell 11: 391-403. Kee Y, D'Andrea AD. 2010. Expanded roles of the Fanconi anemia pathway in preserving genomic stability. Genes Dev. 24: 1680-1694. Keeney S. 2001. Mechanism and control of meiotic recombination initiation. Curr. Top. Dev. Biol. 52: 1-53. Kent NA, Chambers AL, Downs JA. 2007. Dual chromatin remodeling roles for RSC during DNA double strand break induction and repair at the yeast MAT locus. J. Biol. Chem. 282: 27693-27701. Khasanov FK, Savchenko GV, Bashkirova EV, Korolev VG, Heyer W-D, Bashkirov VI. 1999. A New Recombinational DNA Repair Gene From Schizosaccharomyces pombe With Homology to Escherichia coli RecA. Genetics 152: 1557-1572. Khavari PA, Peterson CL, Tamkun JW, Mendel DB, Crabtree GR. 1993. BRG1 contains a conserved domain of the SWI2/SNF2 family necessary for normal mitotic growth and transcription. Nature 366: 170-174. Klapholz S, Waddell CS, Esposito RE. 1985. The role of the SPO11 gene in meiotic recombination in yeast. Genetics 110: 187-216. Kleckner N. 1996. Meiosis: how could it work?. [Review] [103 refs]. Proc. Natl. Acad. Sci. U. S. A. 93: 8167-8174. Kleff S, Kemper B, Sternglanz R. 1992. Identification and characterization of yeast mutants and the gene for a cruciform cutting endonuclease. EMBO J. 11: 699- 704. Klein HL. 1997. RDH54, a RAD54 homologue in Saccharomyces cerevisiae, is required for mitotic diploid-specific recombination and repair and for meiosis. Genetics 147: 1533-1543. -. 2008. The consequences of Rad51 overexpression for normal and tumor cells. DNA Repair (Amst) 7: 686-693. Klein HL, Symington LS. 2009. Breaking up just got easier to do. Cell 138: 20-22. Kneitz B, Cohen PE, Avdievich E, Zhu L, Kane MF, Hou H, Jr., Kolodner RD, Kucherlapati R, Pollard JW, Edelmann W. 2000. MutS homolog 4 localization to meiotic chromosomes is required for chromosome pairing during meiosis in male and female mice. Genes Dev. 14: 1085-1097. Kojic M, Mao N, Zhou Q, Lisby M, Holloman WK. 2008. Compensatory role for Rad52 during recombinational repair in Ustilago maydis. Mol. Microbiol. 67: 1156-1168. Kojic M, Yang H, Kostrub CF, Pavletich NP, Holloman WK. 2003. The BRCA2- Interacting Protein DSS1 Is Vital for DNA Repair, Recombination, and Genome Stability in Ustilago maydis. Mol. Cell 12: 1043-1049. Kolas NK, Chapman JR, Nakada S, Ylanko J, Chahwan R, Sweeney FD, Panier S, Mendez M, Wildenhain J, Thomson TM et al. 2007. Orchestration of the DNA- damage response by the RNF8 ubiquitin ligase. Science 318: 1637-1640. Kolodner R. 1996. Biochemistry and genetics of eukaryotic mismatch repair. Genes Dev. 10: 1433-1442.

226

Kolodner RD, Marsischky GT. 1999. Eukaryotic DNA mismatch repair. Curr. Opin. Genet. Dev. 9: 89-96. Kooistra R, Vreeken K, Zonneveld JBM, De Jong A, Eeken JCJ, Osgood CJ, Buerstedde JM, Lohman PHM, Pastink A. 1997. The Drosophila melanogaster RAD54 homolog, DmRAD54, is involved in the repair of radiation damage and recombination. Mol. Cell. Biol. 17: 6097-6104. Kornberg RD. 1977. Structure of chromatin. Annu. Rev. Biochem. 46: 931-954. Kouzarides T. 2007. Chromatin modifications and their function. Cell 128: 693-705. Kovalenko OV, Golub EI, Bray-Ward P, Ward DC, Radding CM. 1997. A Novel Nucleic Acid-Binding Protein That Interacts with Human Rad51 Recombinase. Nucleic Acids Res 25: 4946-4953. Kowalczykowski SC. 1991. Biochemistry of genetic recombination: energetics and mechanism of DNA strand exchange. Annu. Rev. Biophys. Biophys. Chem. 20: 539-575. Kowalczykowski SC, Eggleston AK. 1994. Homologous pairing and DNA strand- exchange proteins. Annu. Rev. Biochem. 63: 991-1043. Koyama H, Itoh M, Miyahara K, Tsuchiya E. 2002. Abundance of the RSC nucleosome- remodeling complex is important for the cells to tolerate DNA damage in Saccharomyces cerevisiae. FEBS Lett. 531: 215-221. Krejci L, Song B, Bussen W, Rothstein R, Mortensen UH, Sung P. 2002. Interaction with Rad51 Is Indispensable for Recombination Mediator Function of Rad52. J. Biol. Chem. 277: 40132-40141. Krejci L, Van Komen S, Li Y, Villemain J, Reddy MS, Klein H, Ellenberger T, Sung P. 2003. DNA helicase Srs2 disrupts the Rad51 presynaptic filament. Nature 423: 305-309. Krishnan N, Jeong DG, Jung SK, Ryu SE, Xiao A, Allis CD, Kim SJ, Tonks NK. 2009. Dephosphorylation of the C-terminal tyrosyl residue of the DNA damage-related histone H2A.X is mediated by the protein phosphatase eyes absent. J. Biol. Chem. 284: 16066-16070. Krogan NJ, Keogh MC, Datta N, Sawa C, Ryan OW, Ding H, Haw RA, Pootoolal J, Tong A, Canadien V et al. 2003. A Snf2 family ATPase complex required for recruitment of the histone H2A variant Htz1. Mol. Cell 12: 1565-1576. Krogh BO, Symington LS. 2004. Recombination proteins in yeast. Annu. Rev. Genet. 38: 233-271. Kruhlak MJ, Celeste A, Dellaire G, Fernandez-Capetillo O, Müller WG, McNally JG, Bazett-Jones DP, Nussenzweig A. 2006. Changes in chromatin structure and mobility in living cells at sites of DNA double-strand breaks. J. Cell Biol. 172: 823-834. Kusch T, Florens L, Macdonald WH, Swanson SK, Glaser RL, Yates JR, 3rd, Abmayr SM, Washburn MP, Workman JL. 2004. Acetylation by Tip60 is required for selective histone variant exchange at DNA lesions. Science 306: 2084-2087. Kwon Y, Chi P, Roh DH, Klein H, Sung P. 2007. Synergistic action of the Saccharomyces cerevisiae homologous recombination factors Rad54 and Rad51 in chromatin remodeling. DNA Repair (Amst) 6: 1496-1506.

227

Larocque JR, Jasin M. 2010. Mechanisms of recombination between diverged sequences in wild-type and BLM-deficient mouse and human cells. Mol. Cell. Biol. 30: 1887-1897. Lee D, Kim JW, Seo T, Hwang SG, Choi E-J, Choe J. 2002. SWI/SNF Complex Interacts with Tumor Suppressor p53 and Is Necessary for the Activation of p53-mediated Transcription. J. Biol. Chem. 277: 22330-22337. Lee SE, Moore JK, Holmes A, Umezu K, Kolodner RD, Haber JE. 1998. Saccharomyces Ku70, mre11/rad50 and RPA proteins regulate adaptation to G2/M arrest after DNA damage. Cell 94: 399-409. Lengsfeld BM, Rattray AJ, Bhaskara V, Ghirlando R, Paull TT. 2007. Sae2 is an endonuclease that processes hairpin DNA cooperatively with the Mre11/Rad50/Xrs2 complex. Mol. Cell 28: 638-651. Levy-Lahad E. 2010. Fanconi anemia and breast cancer susceptibility meet again. Nat. Genet. 42: 368-369. Li X, Zhang X-P, Solinger JA, Kiianitsa K, Yu X, Egelman EH, Heyer W-D. 2007. Rad51 and Rad54 ATPase activities are both required to modulate Rad51-dsDNA filament dynamics. Nucleic Acids Res 35: 4124-4140. Li Y, He Y, Luo Y. 2009. Conservation of a conformational switch in RadA recombinase from Methanococcus maripaludis. Acta Crystallogr D Biol Crystallogr 65: 602- 610. Lin FL, Sperle K, Sternberg N. 1984. Model for homologous recombination during transfer of DNA into mouse L cells: role for DNA ends in the recombination process. Mol. Cell. Biol. 4: 1020-1034. Lin Z, Kong H, Nei M, Ma H. 2006. Origins and evolution of the recA/RAD51 gene family: evidence for ancient gene duplication and endosymbiotic gene transfer. Proc. Natl. Acad. Sci. U. S. A. 103: 10328-10333. Lio YC, Mazin AV, Kowalczykowski SC, Chen DJ. 2003. Complex formation by the human Rad51B and Rad51C DNA repair proteins and their activities in vitro. J. Biol. Chem. 278: 2469-2478. Lisby M, Barlow JH, Burgess RC, Rothstein R. 2004. Choreography of the DNA damage response: spatiotemporal relationships among checkpoint and repair proteins. Cell 118: 699-713. Lisby M, Rothstein R. 2009. Choreography of recombination proteins during the DNA damage response. DNA Repair (Amst) 8: 1068-1076. Liu J, Doty T, Gibson B, Heyer W-D. 2010. Human BRCA2 protein promotes RAD51 filament formation on RPA-covered single-stranded DNA. Nat Struct Mol Biol 17: 1260-1262. Liu J, Heyer W-D. 2011. Who's who in human recombination: BRCA2 and RAD52. Proc. Natl. Acad. Sci. U. S. A. 108: 441-442. Liu J, Renault L, Veaute X, Fabre F, Stahlberg H, Heyer W-D. 2011. Rad51 paralogues Rad55-Rad57 balance the antirecombinase Srs2 in Rad51 filament formation. Nature advance online publication. Liu J, Renault L, Veaute X, Fabre F, Stahlberg H, Heyer WD. Rad51 paralogues Rad55- Rad57 balance the antirecombinase Srs2 in Rad51 filament formation. Nature.

228

Liu J, Wu TC, Lichten M. 1995. The location and structure of double-strand DNA breaks induced during yeast meiosis: evidence for a covalently linked DNA-protein intermediate. EMBO J. 14: 4599-4608. Liu N, Lamerdin JE, Tebbs RS, Schild D, Tucker JD, Shen MR, Brookman KW, Siciliano MJ, Walter CA, Fan W et al. 1998. XRCC2 and XRCC3, new human Rad51-family members, promote chromosome stability and protect against DNA cross-links and other damages. Mol. Cell 1: 783-793. Liu Y, Masson JY, Shah R, O'Regan P, West SC. 2004a. RAD51C is required for Holliday junction processing in mammalian cells. Science 303: 243-246. Liu Y, Stasiak AZ, Masson JY, McIlwraith MJ, Stasiak A, West SC. 2004b. Conformational changes modulate the activity of human RAD51 protein. J. Mol. Biol. 337: 817-827. Liu Y, West SC. 2004. Happy Hollidays: 40th anniversary of the Holliday junction. Nat Rev Mol Cell Biol 5: 937-944. Lo T, Pellegrini L, Venkitaraman AR, Blundell TL. 2003. Sequence fingerprints in BRCA2 and RAD51: implications for DNA repair and cancer. DNA Repair (Amst) 2: 1015-1028. Lobachev K, Vitriol E, Stemple J, Resnick MA, Bloom K. 2004. Chromosome fragmentation after induction of a double-strand break is an active process prevented by the RMX repair complex. Curr. Biol. 14: 2107-2112. Lobrich M, Jeggo PA. 2007. The impact of a negligent G2/M checkpoint on genomic instability and cancer induction. Nat Rev Cancer 7: 861-869. Lorch Y, Maier-Davis B, Kornberg RD. 2006. Chromatin remodeling by nucleosome disassembly in vitro. Proc. Natl. Acad. Sci. U. S. A. 103: 3090-3093. Lou Z, Minter-Dykhouse K, Franco S, Gostissa M, Rivera MA, Celeste A, Manis JP, van Deursen J, Nussenzweig A, Paull TT et al. 2006. MDC1 maintains genomic stability by participating in the amplification of ATM-dependent DNA damage signals. Mol. Cell 21: 187-200. Lovett ST. 1994. Sequence of the RAD55 gene of Saccharomyces cerevisiae: similarity of RAD55 to prokaryotic RecA and other RecA-like proteins. Gene 142: 103-106. Luger K, Mader AW, Richmond RK, Sargent DF, Richmond TJ. 1997. Crystal structure of the nucleosome core particle at 2.8 A resolution. Nature 389: 251-260. Lundblad V, Blackburn EH. 1993. An alternative pathway for yeast telomere maintenance rescues est1‚àí senescence. Cell 73: 347-360. Mahaney BL, Meek K, Lees-miller SP. 2009. Repair of ionizing radiation-induced DNA double-strand breaks by non-homologous end-joining. Biochem. J. 417: 639-650. Mailand N, Bekker-Jensen S, Faustrup H, Melander F, Bartek J, Lukas C, Lukas J. 2007. RNF8 ubiquitylates histones at DNA double-strand breaks and promotes assembly of repair proteins. Cell 131: 887-900. Malkova A, Naylor ML, Yamaguchi M, Ira G, Haber JE. 2005. RAD51-Dependent Break-Induced Replication Differs in Kinetics and Checkpoint Responses from RAD51-Mediated Gene Conversion. Mol. Cell. Biol. 25: 933-944. Mankouri HW, Craig TJ, Morgan A. 2002. SGS1 is a multicopy suppressor of srs2: functional overlap between DNA helicases. Nucleic Acids Res. 30: 1103-1113.

229

Manohar M, Mooney AM, North JA, Nakkula RJ, Picking JW, Edon A, Fishel R, Poirier MG, Ottesen JJ. 2009. Acetylation of Histone H3 at the Nucleosome Dyad Alters DNA-Histone Binding. J. Biol. Chem. 284: 23312-23321. Marston AL, Amon A. 2004. Meiosis: cell-cycle controls shuffle and deal. Nat Rev Mol Cell Biol 5: 983-997. Marston NJ, Richards WJ, Hughes D, Bertwistle D, Marshall CJ, Ashworth A. 1999. Interaction between the Product of the Breast Cancer Susceptibility Gene BRCA2 and DSS1, a Protein Functionally Conserved from Yeast to Mammals. Mol. Cell. Biol. 19: 4633-4642. Martens JA, Winston F. 2003. Recent advances in understanding chromatin remodeling by Swi/Snf complexes. Curr. Opin. Genet. Dev. 13: 136-142. Martin JS, Winkelmann N, Petalcorin MIR, McIlwraith MJ, Boulton SJ. 2005. RAD-51- Dependent and -Independent Roles of a Caenorhabditis elegans BRCA2-Related Protein during DNA Double-Strand Break Repair. Mol. Cell. Biol. 25: 3127-3139. Martin SG, Laroche T, Suka N, Grunstein M, Gasser SM. 1999. Relocalization of telomeric Ku and SIR proteins in response to DNA strand breaks in yeast. Cell 97: 621-633. Maryon E, Carroll D. 1991. Involvement of single-stranded tails in homologous recombination of DNA injected into Xenopus laevis oocyte nuclei. Mol. Cell. Biol. 11: 3268-3277. Masson JY, Davies AA, Hajibagheri N, Van Dyck E, Benson FE, Stasiak AZ, Stasiak A, West SC. 1999. The meiosis-specific recombinase hDmc1 forms ring structures and interacts with hRad51. EMBO J. 18: 6552-6560. Masson JY, Stasiak AZ, Stasiak A, Benson FE, West SC. 2001a. Complex formation by the human RAD51C and XRCC3 recombination repair proteins. Proc. Natl. Acad. Sci. U. S. A. 98: 8440-8446. Masson JY, Tarsounas MC, Stasiak AZ, Stasiak A, Shah R, McIlwraith MJ, Benson FE, West SC. 2001b. Identification and purification of two distinct complexes containing the five RAD51 paralogs. Genes Dev. 15: 3296-3307. Masson JY, West SC. 2001. The Rad51 and Dmc1 recombinases: a non-identical twin relationship. Trends Biochem. Sci. 26: 131-136. Matsuoka S, Ballif BA, Smogorzewska A, McDonald ER, 3rd, Hurov KE, Luo J, Bakalarski CE, Zhao Z, Solimini N, Lerenthal Y et al. 2007. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 316: 1160-1166. Matulova P, Marini V, Burgess RC, Sisakova A, Kwon Y, Rothstein R, Sung P, Krejci L. 2009. Cooperativity of Mus81·Mms4 with Rad54 in the Resolution of Recombination and Replication Intermediates. J. Biol. Chem. 284: 7733-7745. Mazin AV, Alexeev AA, Kowalczykowski SC. 2003. A Novel Function of Rad54 Protein. J. Biol. Chem. 278: 14029-14036. Mazin AV, Kowalczykowski SC. 1998. The function of the secondary DNA-binding site of RecA protein during DNA strand exchange. EMBO J. 17: 1161-1168. Mazin AV, Mazina OM, Bugreev DV, Rossi MJ. 2010. Rad54, the motor of homologous recombination. DNA Repair (Amst) 9: 286-302.

230

Mazina OM, Mazin AV. 2004. Human Rad54 Protein Stimulates DNA Strand Exchange Activity of hRad51 Protein in the Presence of Ca2+. J. Biol. Chem. 279: 52042- 52051. -. 2008. Human Rad54 protein stimulates human Mus81–Eme1 endonuclease. Proc. Natl. Acad. Sci. U. S. A. 105: 18249-18254. Mazina OM, Rossi MJ, Thomaä NH, Mazin AV. 2007. Interactions of Human Rad54 Protein with Branched DNA Molecules. J. Biol. Chem. 282: 21068-21080. Mazloum N, Holloman WK. 2009. Second-end capture in DNA double-strand break repair promoted by Brh2 protein of Ustilago maydis. Mol. Cell 33: 160-170. McEachern MJ, Haber JE. 2006. Break-Induced Replication and Recombinational Telomere Elongation in Yeast. Annu. Rev. Biochem. 75: 111-135. McIlwraith MJ, West SC. 2008. DNA Repair Synthesis Facilitates RAD52-Mediated Second-End Capture during DSB Repair. Mol. Cell 29: 510-516. McKee AH, Kleckner N. 1997. Mutations in Saccharomyces cerevisiae that block meiotic prophase chromosome metabolism and confer cell cycle arrest at pachytene identify two new meiosis-specific genes SAE1 and SAE3. Genetics 146: 817-834. McKee RA, Lawrence CW. 1980. Genetic analysis of [gamma]-mutagenesis in yeast III. Double-mutant strains. Mutation Research/Fundamental and Molecular Mechanisms of Mutagenesis 70: 37-48. McPherson JP, Lemmers B, Chahwan R, Pamidi A, Migon E, Matysiak-Zablocki E, Moynahan ME, Essers J, Hanada K, Poonepalli A et al. 2004. Involvement of mammalian Mus81 in genome integrity and tumor suppression. Science 304: 1822-1826. McRee DE. 1999. XtalView/Xfit--A versatile program for manipulating atomic coordinates and electron density. J. Struct. Biol. 125: 156-165. Meindl A, Hellebrand H, Wiek C, Erven V, Wappenschmidt B, Niederacher D, Freund M, Lichtner P, Hartmann L, Schaal H et al. 2010. Germline mutations in breast and ovarian cancer pedigrees establish RAD51C as a human cancer susceptibility gene. Nat. Genet. 42: 410-414. Melander F, Bekker-Jensen S, Falck J, Bartek J, Mailand N, Lukas J. 2008. Phosphorylation of SDT repeats in the MDC1 N terminus triggers retention of NBS1 at the DNA damage-modified chromatin. J. Cell Biol. 181: 213-226. Menetski JP, Bear DG, Kowalczykowski SC. 1990. Stable DNA heteroduplex formation catalyzed by the Escherichia coli RecA protein in the absence of ATP hydrolysis. Proc. Natl. Acad. Sci. U. S. A. 87: 21-25. Miki Y, Swensen J, Shattuck Eidens D, Futreal PA, Harshman K, Tavtigian S, Liu Q, Cochran C, Bennett LM, Ding W et al. 1994. A strong candidate for the breast and ovarian cancer susceptibility gene BRCA1. Science 266: 66-71. Miller KA, Yoshikawa DM, McConnell IR, Clark R, Schild D, Albala JS. 2002. RAD51C interacts with RAD51B and is central to a larger protein complex in vivo exclusive of RAD51. J. Biol. Chem. 277: 8406-8411. Milne GT, Weaver DT. 1993. Dominant negative alleles of RAD52 reveal a DNA repair/recombination complex including Rad51 and Rad52. Genes Dev. 7: 1755- 1765.

231

Mimitou EP, Symington LS. 2008. Sae2, Exo1 and Sgs1 collaborate in DNA double- strand break processing. Nature 455: 770-774. -. 2009a. DNA end resection: many nucleases make light work. DNA Repair (Amst) 8: 983-995. -. 2009b. Nucleases and helicases take center stage in homologous recombination. Trends Biochem. Sci. 34: 264-272. Mizuguchi G, Shen X, Landry J, Wu W-H, Sen S, Wu C. 2004. ATP-Driven Exchange of Histone H2AZ Variant Catalyzed by SWR1 Chromatin Remodeling Complex. Science 303: 343-348. Mizuta R, LaSalle JM, Cheng H-L, Shinohara A, Ogawa H, Copeland N, Jenkins NA, Lalande M, Alt FW. 1997. RAB22 and RAB163/mouse BRCA2: Proteins that specifically interact with the RAD51‚Äâprotein. Proc. Natl. Acad. Sci. U. S. A. 94: 6927-6932. Mizuuchi K, Kemper B, Hays J, Weisberg RA. 1982. T4 endonuclease VII cleaves holliday structures. Cell 29: 357-365. Modesti M, Budzowska M, Baldeyron Cl, Demmers JAA, Ghirlando R, Kanaar R. 2007. RAD51AP1 Is a Structure-Specific DNA Binding Protein that Stimulates Joint Molecule Formation during RAD51-Mediated Homologous Recombination. Mol. Cell 28: 468-481. Moens PB, Kolas NK, Tarsounas M, Marcon E, Cohen PE, Spyropoulos B. 2002. The time course and chromosomal localization of recombination-related proteins at meiosis in the mouse are compatible with models that can resolve the early DNA- DNA interactions without reciprocal recombination. J. Cell Sci. 115: 1611-1622. Morgan EA, Shah N, Symington LS. 2002. The Requirement for ATP Hydrolysis by Saccharomyces cerevisiae Rad51 Is Bypassed by Mating-Type Heterozygosity or RAD54 in High Copy. Mol. Cell. Biol. 22: 6336-6343. Morris JR, Boutell C, Keppler M, Densham R, Weekes D, Alamshah A, Butler L, Galanty Y, Pangon L, Kiuchi T et al. 2009. The SUMO modification pathway is involved in the BRCA1 response to genotoxic stress. Nature 462: 886-890. Morrison AJ, Highland J, Krogan NJ, Arbel-Eden A, Greenblatt JF, Haber JE, Shen X. 2004. INO80 and gamma-H2AX interaction links ATP-dependent chromatin remodeling to DNA damage repair. Cell 119: 767-775. Morrison C, Shinohara A, Sonoda E, Yamaguchi-Iwai Y, Takata M, Weichselbaum RR, Takeda S. 1999. The essential functions of human Rad51 are independent of ATP hydrolysis. Mol. Cell. Biol. 19: 6891-6897. Mortensen UH, Bendixen C, Sunjevaric I, Rothstein R. 1996. DNA strand annealing is promoted by the yeast Rad52 protein. Proc. Natl. Acad. Sci. U. S. A. 93: 10729- 10734. Moyal L, Lerenthal Y, Gana-Weisz M, Mass G, So S, Wang SY, Eppink B, Chung YM, Shalev G, Shema E et al. 2011. Requirement of ATM-dependent monoubiquitylation of histone H2B for timely repair of DNA double-strand breaks. Mol. Cell 41: 529-542. Moynahan ME, Jasin M. 2010. Mitotic homologous recombination maintains genomic stability and suppresses tumorigenesis. Nat Rev Mol Cell Biol 11: 196-207.

232

Moynahan ME, Pierce AJ, Jasin M. 2001. BRCA2 Is Required for Homology-Directed Repair of Chromosomal Breaks. Mol. Cell 7: 263-272. Mullen JR, Nallaseth FS, Lan YQ, Slagle CE, Brill SJ. 2005. Yeast Rmi1/Nce4 controls genome stability as a subunit of the Sgs1-Top3 complex. Mol. Cell. Biol. 25: 4476-4487. Munoz IM, Hain K, Declais AC, Gardiner M, Toh GW, Sanchez-Pulido L, Heuckmann JM, Toth R, Macartney T, Eppink B et al. 2009. Coordination of structure- specific nucleases by human SLX4/BTBD12 is required for DNA repair. Mol. Cell 35: 116-127. Muris DFR, Vreeken K, Carr AM, Murray JM, Smit C, Lohman PHM, Pastink A. 1996. Isolation of the Schizosaccharomyces pombe RAD54 homologue, rhp54+, a gene involved in the repair of radiation damage and replication fidelity. J. Cell Sci. 109: 73-81. Murr R, Loizou JI, Yang YG, Cuenin C, Li H, Wang ZQ, Herceg Z. 2006. Histone acetylation by Trrap-Tip60 modulates loading of repair proteins and repair of DNA double-strand breaks. Nat Cell Biol 8: 91-99. Myung K, Datta A, Chen C, Kolodner RD. 2001. SGS1, the Saccharomyces cerevisiae homologue of BLM and WRN, suppresses genome instability and homeologous recombination. Nat. Genet. 27: 113-116. Nakada S, Tai I, Panier S, Al-Hakim A, Iemura S, Juang YC, O'Donnell L, Kumakubo A, Munro M, Sicheri F et al. 2010. Non-canonical inhibition of DNA damage- dependent ubiquitination by OTUB1. Nature 466: 941-946. Nakamura K, Kato A, Kobayashi J, Yanagihara H, Sakamoto S, Oliveira DV, Shimada M, Tauchi H, Suzuki H, Tashiro S et al. 2011. Regulation of homologous recombination by RNF20-dependent H2B ubiquitination. Mol. Cell 41: 515-528. Nassif N, Penney J, Pal S, Engels WR, Gloor GB. 1994. Efficient copying of nonhomologous sequences from ectopic sites via P-element-induced gap repair. Mol. Cell. Biol. 14: 1613-1625. Neigeborn L, Carlson M. 1984. GENES AFFECTING THE REGULATION OF SUC2 GENE EXPRESSION BY GLUCOSE REPRESSION IN SACCHAROMYCES CEREVISIAE. Genetics 108: 845-858. New JH, Sugiyama T, Zaitseva E, Kowalczykowski SC. 1998. Rad52 protein stimulates DNA strand exchange by Rad51 and replication protein A. Nature 391: 407-410. Nicholson A, Hendrix M, Jinks-Robertson S, Crouse GF. 2000. Regulation of mitotic homeologous recombination in yeast. Functions of mismatch repair and nucleotide excision repair genes. Genetics 154: 133-146. Nimonkar AV, Genschel J, Kinoshita E, Polaczek P, Campbell JL, Wyman C, Modrich P, Kowalczykowski SC. 2011. BLM-DNA2-RPA-MRN and EXO1-BLM-RPA- MRN constitute two DNA end resection machineries for human DNA break repair. Genes Dev. 25: 350-362. Nimonkar AV, Ozsoy AZ, Genschel J, Modrich P, Kowalczykowski SC. 2008. Human exonuclease 1 and BLM helicase interact to resect DNA and initiate DNA repair. Proc. Natl. Acad. Sci. U. S. A. 105: 16906-16911.

233

Nimonkar AV, Sica RA, Kowalczykowski SC. 2009. Rad52 promotes second-end DNA capture in double-stranded break repair to form complement-stabilized joint molecules. Proc. Natl. Acad. Sci. U. S. A. 106: 3077-3082. Niu H, Chung W-H, Zhu Z, Kwon Y, Zhao W, Chi P, Prakash R, Seong C, Liu D, Lu L et al. 2010. Mechanism of the ATP-dependent DNA end-resection machinery from Saccharomyces cerevisiae. Nature 467: 108-111. Ogawa T, Yu X, Shinohara A, Egelman E. 1993. Similarity of the yeast RAD51 filament to the bacterial RecA filament. Science 259: 1896-1899. Oram M, Keeley A, Tsaneva I. 1998. Holliday junction resolvase in Schizosaccharomyces pombe has identical endonuclease activity to the CCE1 homologue YDC2. Nucleic Acids Res 26: 594-601. Orr-Weaver TL, Szostak JW. 1983a. Multiple, tandem plasmid integration in Saccharomyces cerevisiae. Mol. Cell. Biol. 3: 747-749. -. 1983b. Yeast recombination: the association between double-strand gap repair and crossing-over. Proc. Natl. Acad. Sci. U. S. A. 80: 4417-4421. Osley MA, Tsukuda T, Nickoloff JA. 2007. ATP-dependent chromatin remodeling factors and DNA damage repair. Mutat. Res. 618: 65-80. Otsuki T, Furukawa Y, Ikeda K, Endo H, Yamashita T, Shinohara A, Iwamatsu A, Ozawa K, Liu JM. 2001. Fanconi anemia protein, FANCA, associates with BRG1, a component of the human SWI/SNF complex. Hum. Mol. Genet. 10: 2651-2660. Papamichos-Chronakis M, Krebs JE, Peterson CL. 2006. Interplay between Ino80 and Swr1 chromatin remodeling enzymes regulates cell cycle checkpoint adaptation in response to DNA damage. Genes Dev. 20: 2437-2449. Parsons CA, Baumann P, Van Dyck E, West SC. 2000. Precise binding of single-stranded DNA termini by human RAD52 protein. EMBO J. 19: 4175-4181. Parsons CA, Stasiak A, Bennett RJ, West SC. 1995a. Structure of a multisubunit complex that promotes DNA branch migration. Nature 374: 375-378. Parsons CA, Stasiak A, West SC. 1995b. The E.coli RuvAB proteins branch migrate Holliday junctions through heterologous DNA sequences in a reaction facilitated by SSB. EMBO J. 14: 5736-5744. Passy SI, Yu X, Li Z, Radding CM, Masson J-Y, West SC, Egelman EH. 1999. Human Dmc1 protein binds DNA as an octameric ring. Proc. Natl. Acad. Sci. U. S. A. 96: 10684-10688. Pellegrini L, Yu DS, Lo T, Anand S, Lee M, Blundell TL, Venkitaraman AR. 2002. Insights into DNA recombination from the structure of a RAD51-BRCA2 complex. Nature 420: 287-293. Petalcorin MIR, Sandall J, Wigley DB, Boulton SJ. 2006. CeBRC-2 Stimulates D-loop Formation by RAD-51 and Promotes DNA Single-strand Annealing. J. Mol. Biol. 361: 231-242. Petrini JHJ, Bressan DA, Yao MS. 1997. TheRAD52epistasis group in mammalian double strand break repair. Semin. Immunol. 9: 181-188. Petukhova G, Stratton S, Sung P. 1998. Catalysis of homologous DNA pairing by yeast Rad51 and Rad54 proteins. Nature 393: 91-94.

234

Petukhova G, Van Komen S, Vergano S, Klein H, Sung P. 1999. Yeast Rad54 Promotes Rad51-dependent Homologous DNA Pairing via ATP Hydrolysis-driven Change in DNA Double Helix Conformation. J. Biol. Chem. 274: 29453-29462. Pittman DL, Schimenti JC. 2000. Midgestation lethality in mice deficient for the RecA- related gene, Rad51d/Rad51l3. genesis 26: 167-173. Pittman DL, Weinberg LR, Schimenti JC. 1998. Identification, Characterization, and Genetic Mapping ofRad51d,a New Mouse and HumanRAD51/RecA-Related Gene. Genomics 49: 103-111. Prakash S, Johnson RE, Prakash L. 2005. EUKARYOTIC TRANSLESION SYNTHESIS DNA POLYMERASES: Specificity of Structure and Function. Annu. Rev. Biochem. 74: 317-353. Pugh BF, Cox MM. 1987. Stable binding of recA protein to duplex DNA. Unraveling a paradox. J. Biol. Chem. 262: 1326-1336. Qian X, He Y, Ma X, Fodje MN, Grochulski P, Luo Y. 2006a. Calcium stiffens archaeal Rad51 recombinase from Methanococcus voltae for homologous recombination. J. Biol. Chem. 281: 39380-39387. Qian X, He Y, Wu Y, Luo Y. 2006b. Asp302 determines potassium dependence of a RadA recombinase from Methanococcus voltae. J. Mol. Biol. 360: 537-547. Qian X, Wu Y, He Y, Luo Y. 2005. Crystal structure of Methanococcus voltae RadA in complex with ADP: hydrolysis-induced conformational change. Biochemistry (Mosc). 44: 13753-13761. Ranatunga W, Jackson D, Lloyd JA, Forget AL, Knight KL, Borgstahl GEO. 2001. Human RAD52 Exhibits Two Modes of Self-association. J. Biol. Chem. 276: 15876-15880. Rehrauer WM, Kowalczykowski SC. 1993. Alteration of the nucleoside triphosphate (NTP) catalytic domain within Escherichia coli recA protein attenuates NTP hydrolysis but not joint molecule formation. J. Biol. Chem. 268: 1292-1297. Reid S, Schindler D, Hanenberg H, Barker K, Hanks S, Kalb R, Neveling K, Kelly P, Seal S, Freund M et al. 2007. Biallelic mutations in PALB2 cause Fanconi anemia subtype FA-N and predispose to childhood cancer. Nat. Genet. 39: 162-164. Reisman D, Glaros S, Thompson EA. 2009. The SWI/SNF complex and cancer. Oncogene 28: 1653-1668. Renodon-Corniere A, Takizawa Y, Conilleau S, Tran V, Iwai S, Kurumizaka H, Takahashi M. 2008. Structural analysis of the human Rad51 protein-DNA complex filament by tryptophan fluorescence scanning analysis: transmission of allosteric effects between ATP binding and DNA binding. J. Mol. Biol. 383: 575- 587. Reymer A, Frykholm K, Morimatsu K, Takahashi M, Norden B. 2009. Structure of human Rad51 protein filament from molecular modeling and site-specific linear dichroism spectroscopy. Proc. Natl. Acad. Sci. U. S. A. 106: 13248-13253. Rice JC, Allis CD. 2001. Code of silence. Nature 414: 258-261. Rice MC, Smith ST, Bullrich F, Havre P, Kmiec EB. 1997. Isolation of human and mouse genes based on homology to REC2, a recombinational repair gene from the fungus Ustilago maydis. Proc. Natl. Acad. Sci. U. S. A. 94: 7417-7422.

235

Richardson C, Stark JM, Ommundsen M, Jasin M. 2004. Rad51 overexpression promotes alternative double-strand break repair pathways and genome instability. Oncogene 23: 546-553. Rijkers T, Van Den Ouweland J, Morolli B, Rolink AG, Baarends WM, Van Sloun PPH, Lohman PHM, Pastink A. 1998. Targeted Inactivation of Mouse RAD52 Reduces Homologous Recombination but Not Resistance to Ionizing Radiation. Mol. Cell. Biol. 18: 6423-6429. Ristic D, Modesti M, van der Heijden T, van Noort J, Dekker C, Kanaar R, Wyman C. 2005. Human Rad51 filaments on double- and single-stranded DNA: correlating regular and irregular forms with recombination function. Nucleic Acids Res 33: 3292-3302. Ritchie MF, Zhou Y, Soboloff J. 2011. Transcriptional mechanisms regulating Ca2+ homeostasis. Cell Calcium 49: 314-321. Rogakou EP, Boon C, Redon C, Bonner WM. 1999. Megabase chromatin domains involved in DNA double-strand breaks in vivo. J. Cell Biol. 146: 905-916. Rogakou EP, Nieves-Neira W, Boon C, Pommier Y, Bonner WM. 2000. Initiation of DNA fragmentation during apoptosis induces phosphorylation of H2AX histone at serine 139. J. Biol. Chem. 275: 9390-9395. Rogakou EP, Pilch DR, Orr AH, Ivanova VS, Bonner WM. 1998. DNA double-stranded breaks induce histone H2AX phosphorylation on serine 139. J. Biol. Chem. 273: 5858-5868. Roy A, Kucukural A, Zhang Y. 2010. I-TASSER: a unified platform for automated protein structure and function prediction. Nat. Protocols 5: 725-738. Saha A, Wittmeyer J, Cairns BR. 2002. Chromatin remodeling by RSC involves ATP- dependent DNA translocation. Genes Dev. 16: 2120-2134. Saha A, Wittmeyer J, Cairns BR. 2006. Chromatin remodelling: the industrial revolution of DNA around histones. Nat Rev Mol Cell Biol 7: 437-447. San Filippo J, Sung P, Klein H. 2008. Mechanism of eukaryotic homologous recombination. Annu. Rev. Biochem. 77: 229-257. Sauvageau S, Stasiak AZ, Banville I, Ploquin M, Stasiak A, Masson J-Y. 2005. Fission Yeast Rad51 and Dmc1, Two Efficient DNA Recombinases Forming Helical Nucleoprotein Filaments. Mol. Cell. Biol. 25: 4377-4387. Savitsky K, Bar Shira A, Gilad S, Rotman G, Ziv Y, Vanagaite L, Tagle DA, Smith S, Uziel T, Sfez S et al. 1995. A single ataxia telangiectasia gene with a product similar to PI-3 kinase [see comments]. Science 268: 1749-1753. Saydam N, Kanagaraj R, Dietschy T, Garcia PL, Pena-Diaz J, Shevelev I, Stagljar I, Janscak P. 2007. Physical and functional interactions between Werner syndrome helicase and mismatch-repair initiation factors. Nucleic Acids Res 35: 5706-5716. Schaetzlein S, Kodandaramireddy NR, Ju Z, Lechel A, Stepczynska A, Lilli DR, Clark AB, Rudolph C, Kuhnel F, Wei K et al. 2007. Exonuclease-1 Deletion Impairs DNA Damage Signaling and Prolongs Lifespan of Telomere-Dysfunctional Mice. Cell 130: 863-877. Schild D, Lio Y-c, Collins DW, Tsomondo T, Chen DJ. 2000a. Evidence for Simultaneous Protein Interactions between Human Rad51 Paralogs. J. Biol. Chem. 275: 16443-16449.

236

Schild D, Lio YC, Collins DW, Tsomondo T, Chen DJ. 2000b. Evidence for simultaneous protein interactions between human Rad51 paralogs. J. Biol. Chem. 275: 16443-16449. Schild D, Wiese C. 2010. Overexpression of RAD51 suppresses recombination defects: a possible mechanism to reverse genomic instability. Nucleic Acids Res 38: 1061- 1070. Schlacher K, Christ N, Siaud N, Egashira A, Wu H, Jasin M. 2011. Double-Strand Break Repair-Independent Role for BRCA2 in Blocking Stalled Replication Fork Degradation by MRE11. Cell 145: 529-542. Schoenmakers EFPM, Huysmans C, Van de Ven WJM. 1999. Allelic Knockout of Novel Splice Variants of Human Recombination Repair Gene RAD51B in t(12;14) Uterine Leiomyomas. Cancer Res. 59: 19-23. Scully R, Chen J, Ochs RL, Keegan K, Hoekstra M, Feunteun J, Livingston DM. 1997. Dynamic Changes of BRCA1 Subnuclear Location and Phosphorylation State Are Initiated by DNA Damage. Cell 90: 425-435. Sehorn MG, Sigurdsson S, Bussen W, Unger VM, Sung P. 2004. Human meiotic recombinase Dmc1 promotes ATP-dependent homologous DNA strand exchange. Nature 429: 433-437. Seitz EM, Haseltine CA, Kowalczykowski SC. 2001. DNA recombination and repair in the Archaea. in Adv. Appl. Microbiol. (ed. B Paul), pp. 101-169. Academic Press. Shah PP, Zheng X, Epshtein A, Carey JN, Bishop DK, Klein HL. 2010. Swi2/Snf2- Related Translocases Prevent Accumulation of Toxic Rad51 Complexes during Mitotic Growth. Mol. Cell 39: 862-872. Shaked H, Avivi-Ragolsky N, Levy AA. 2006. Involvement of the Arabidopsis SWI2/SNF2 Chromatin Remodeling Gene Family in DNA Damage Response and Recombination. Genetics 173: 985-994. Shan Q, Cox MM, Inman RB. 1996. DNA strand exchange promoted by RecA K72R. Two reaction phases with different Mg2+ requirements. J. Biol. Chem. 271: 5712- 5724. Sharan SK, Kuznetsov SG. 2007. Resolving RAD51C function in late stages of homologous recombination. Cell Div 2: 15. Sharan SK, Morimatsu M, Albrecht U, Lim D-S, Regel E, Dinh C, Sands A, Eichele G, Hasty P, Bradley A. 1997. Embryonic lethality and radiation hypersensitivity mediated by Rad51 in mice lacking Brca2. Nature 386: 804-810. Sharples GJ, Lloyd RG. 1991. Resolution of Holliday junctions in Escherichia coli: identification of the ruvC gene product as a 19-kilodalton protein. J. Bacteriol. 173: 7711-7715. Shen X, Mizuguchi G, Hamiche A, Wu C. 2000. A chromatin remodelling complex involved in transcription and DNA processing. Nature 406: 541-544. Shen X, Ranallo R, Choi E, Wu C. 2003. Involvement of actin-related proteins in ATP- dependent chromatin remodeling. Mol. Cell 12: 147-155. Shen Z, Cloud KG, Chen DJ, Park MS. 1996. Specific Interactions between the Human RAD51 and RAD52 Proteins. J. Biol. Chem. 271: 148-152.

237

Shi I, Hallwyl SC, Seong C, Mortensen U, Rothstein R, Sung P. 2009. Role of the Rad52 amino-terminal DNA binding activity in DNA strand capture in homologous recombination. J. Biol. Chem. 284: 33275-33284. Shibata T, Cunningham RP, DasGupta C, Radding CM. 1979. Homologous pairing in genetic recombination: complexes of recA protein and DNA. Proc. Natl. Acad. Sci. U. S. A. 76: 5100-5104. Shim EY, Chung W-H, Nicolette ML, Zhang Y, Davis M, Zhu Z, Paull TT, Ira G, Lee SE. 2010. Saccharomyces cerevisiae Mre11/Rad50/Xrs2 and Ku proteins regulate association of Exo1 and Dna2 with DNA breaks. EMBO J. 29: 3370-3380. Shim EY, Hong SJ, Oum JH, Yanez Y, Zhang Y, Lee SE. 2007. RSC mobilizes nucleosomes to improve accessibility of repair machinery to the damaged chromatin. Mol. Cell. Biol. 27: 1602-1613. Shim EY, Ma JL, Oum JH, Yanez Y, Lee SE. 2005. The yeast chromatin remodeler RSC complex facilitates end joining repair of DNA double-strand breaks. Mol. Cell. Biol. 25: 3934-3944. Shim KS, Schmutte C, Tombline G, Heinen CD, Fishel R. 2004. hXRCC2 enhances ADP/ATP processing and strand exchange by hRAD51. J. Biol. Chem. 279: 30385-30394. Shim KS, Schmutte C, Yoder K, Fishel R. 2006. Defining the salt effect on human RAD51 activities. DNA Repair (Amst) 5: 718-730. Shin DS, Pellegrini L, Daniels DS, Yelent B, Craig L, Bates D, Yu DS, Shivji MK, Hitomi C, Arvai AS et al. 2003. Full-length archaeal Rad51 structure and mutants: mechanisms for RAD51 assembly and control by BRCA2. EMBO J. 22: 4566-4576. Shinohara A, Ogawa H, Ogawa T. 1992. Rad51 protein involved in repair and recombination in S. cerevisiae is a RecA-like protein. Cell 69: 457-470. Shinohara A, Ogawa T. 1998. Stimulation by Rad52 of yeast Rad51- mediated recombination. Nature 391: 404-407. Shinohara A, Shinohara M, Ohta T, Matsuda S, Ogawa T. 1998. Rad52 forms ring structures and co-operates with RPA in single-strand DNA annealing. Genes Cells 3: 145-156. Shivji MKK, Mukund SR, Rajendra E, Chen S, Short JM, Savill J, Klenerman D, Venkitaraman AR. 2009. The BRC repeats of human BRCA2 differentially regulate RAD51 binding on single- versus double-stranded DNA to stimulate strand exchange. Proc. Natl. Acad. Sci. U. S. A. 106: 13254-13259. Shu Z, Smith S, Wang L, Rice MC, Kmiec EB. 1999. Disruption of muREC2/RAD51L1 in Mice Results in Early Embryonic Lethality Which Can Be Partially Rescued in a p53-/- Background. Mol. Cell. Biol. 19: 8686-8693. Siaud N, Dray E, Gy I, Gerard E, Takvorian N, Doutriaux M-P. 2004. Brca2 is involved in meiosis in Arabidopsis thaliana as suggested by its interaction with Dmc1. EMBO J. 23: 1392-1401. Sigurdsson S, Trujillo K, Song B, Stratton S, Sung P. 2001a. Basis for avid homologous DNA strand exchange by human Rad51 and RPA. J. Biol. Chem. 276: 8798-8806.

238

Sigurdsson S, Van Komen S, Bussen W, Schild D, Albala JS, Sung P. 2001b. Mediator function of the human Rad51B-Rad51C complex in Rad51/RPA-catalyzed DNA strand exchange. Genes Dev. 15: 3308-3318. Sigurdsson S, Van Komen S, Petukhova G, Sung P. 2002. Homologous DNA Pairing by Human Recombination Factors Rad51 and Rad54. J. Biol. Chem. 277: 42790- 42794. Simon M, North JA, Shimko JC, Forties RA, Ferdinand MB, Manohar M, Zhang M, Fishel R, Ottesen JJ, Poirier MG. 2011. Histone fold modifications control nucleosome unwrapping and disassembly. Proc. Natl. Acad. Sci. U. S. A. 108: 12711-12716. Singleton MR, Dillingham MS, Wigley DB. 2007. Structure and Mechanism of Helicases and Nucleic Acid Translocases. Annu. Rev. Biochem. 76: 23-50. Singleton MR, Wentzell LM, Liu Y, West SC, Wigley DB. 2002. Structure of the single- strand annealing domain of human RAD52 protein. Proc. Natl. Acad. Sci. U. S. A. 99: 13492-13497. Sinha M, Peterson CL. 2008. A Rad51 presynaptic filament is sufficient to capture nucleosomal homology during recombinational repair of a DNA double-strand break. Mol. Cell 30: 803-810. Sinha M, Watanabe S, Johnson A, Moazed D, Peterson CL. 2009. Recombinational repair within heterochromatin requires ATP-dependent chromatin remodeling. Cell 138: 1109-1121. Snow R. 1967. Mutants of yeast sensitive to ultraviolet light. J. Bacteriol. 94: 571-575. Sobhian B, Shao G, Lilli DR, Culhane AC, Moreau LA, Xia B, Livingston DM, Greenberg RA. 2007. RAP80 Targets BRCA1 to Specific Ubiquitin Structures at DNA Damage Sites. Science 316: 1198-1202. Solinger JA, Heyer W-D. 2001. Rad54 protein stimulates the postsynaptic phase of Rad51 protein-mediated DNA strand exchange. Proc. Natl. Acad. Sci. U. S. A. 98: 8447-8453. Solinger JA, Kiianitsa K, Heyer WD. 2002. Rad54, a Swi2/Snf2-like recombinational repair protein, disassembles Rad51:dsDNA filaments. Mol. Cell 10: 1175-1188. Sonoda E, Sasaki MS, Buerstedde JM, Bezzubova O, Shinohara A, Ogawa H, Takata M, Yamaguchi-Iwai Y, Takeda S. 1998. Rad51-deficient vertebrate cells accumulate chromosomal breaks prior to cell death. EMBO J. 17: 598-608. Spell RM, Jinks-Robertson S. 2004. Examination of the Roles of Sgs1 and Srs2 Helicases in the Enforcement of Recombination Fidelity in Saccharomyces cerevisiae. Genetics 168: 1855-1865. Spycher C, Miller ES, Townsend K, Pavic L, Morrice NA, Janscak P, Stewart GS, Stucki M. 2008. Constitutive phosphorylation of MDC1 physically links the MRE11- RAD50-NBS1 complex to damaged chromatin. J. Cell Biol. 181: 227-240. Stahl FW. 1979. Genetic Recombination. W.H. Freeman and Co., San Francisco. Stark JM, Hu P, Pierce AJ, Moynahan ME, Ellis N, Jasin M. 2002. ATP Hydrolysis by Mammalian RAD51 Has a Key Role during Homology-directed DNA Repair. J. Biol. Chem. 277: 20185-20194.

239

Stasiak AZ, Larquet E, Stasiak A, Müller S, Engel A, Van Dyck E, West SC, Egelman EH. 2000. The human Rad52 protein exists as a heptameric ring. Curr. Biol. 10: 337-340. Stern M, Jensen R, Herskowitz I. 1984. Five SWI genes are required for expression of the HO gene in yeast. J. Mol. Biol. 178: 853-868. Stewart GS, Panier S, Townsend K, Al-Hakim AK, Kolas NK, Miller ES, Nakada S, Ylanko J, Olivarius S, Mendez M et al. 2009. The RIDDLE syndrome protein mediates a ubiquitin-dependent signaling cascade at sites of DNA damage. Cell 136: 420-434. Story R, Bishop D, Kleckner N, Steitz T. 1993. Structural relationship of bacterial RecA proteins to recombination proteins from bacteriophage T4 and yeast. Science 259: 1892-1896. Story RM, Steitz TA. 1992. Structure of the recA protein-ADP complex. Nature 355: 374-376. Strahl BD, Allis CD. 2000. The language of covalent histone modifications. Nature 403: 41-45. Strathern JN, Klar AJS, Hicks JB, Abraham JA, Ivy JM, Nasmyth KA, McGill C. 1982. Homothallic switching of yeast mating type cassettes is initiated by a double- stranded cut in the MAT locus. Cell 31: 183-192. Strom L, Karlsson C, Lindroos HB, Wedahl S, Katou Y, Shirahige K, Sjogren C. 2007. Postreplicative formation of cohesion is required for repair and induced by a single DNA break. Science 317: 242-245. Strom L, Lindroos HB, Shirahige K, Sjogren C. 2004. Postreplicative recruitment of cohesin to double-strand breaks is required for DNA repair. Mol. Cell 16: 1003- 1015. Stucki M. 2009. Histone H2A.X Tyr142 phosphorylation: a novel sWItCH for apoptosis? DNA Repair (Amst) 8: 873-876. Stucki M, Clapperton JA, Mohammad D, Yaffe MB, Smerdon SJ, Jackson SP. 2005. MDC1 directly binds phosphorylated histone H2AX to regulate cellular responses to DNA double-strand breaks. Cell 123: 1213-1226. Sugawara N, Wang X, Haber JE. 2003. In vivo roles of Rad52, Rad54, and Rad55 proteins in Rad51-mediated recombination. Mol. Cell 12: 209-219. Sugiyama T, Kantake N. 2009. Dynamic regulatory interactions of rad51, rad52, and replication protein-a in recombination intermediates. J. Mol. Biol. 390: 45-55. Sugiyama T, Zaitseva EM, Kowalczykowski SC. 1997. A single-stranded DNA-binding protein is needed for efficient presynaptic complex formation by the Saccharomyces cerevisiae Rad51 protein. J. Biol. Chem. 272: 7940-7945. Sung P. 1994. Catalysis of ATP-dependent homologous DNA pairing and strand exchange by yeast RAD51 protein. Science 265: 1241-1243. Sung P. 1997a. Function of Yeast Rad52 Protein as a Mediator between Replication Protein A and the Rad51 Recombinase. J. Biol. Chem. 272: 28194-28197. Sung P. 1997b. Yeast Rad55 and Rad57 proteins form a heterodimer that functions with replication protein A to promote DNA strand exchange by Rad51 recombinase. Genes Dev. 11: 1111-1121.

240

Sung P, Robberson DL. 1995. DNA strand exchange mediated by a RAD51-ssDNA nucleoprotein filament with polarity opposite to that of RecA. Cell 82: 453-461. Sung P, Stratton SA. 1996. Yeast Rad51 Recombinase Mediates Polar DNA Strand Exchange in the Absence of ATP Hydrolysis. J. Biol. Chem. 271: 27983-27986. Svendsen JM, Harper JW. 2010. GEN1/Yen1 and the SLX4 complex: Solutions to the problem of Holliday junction resolution. Genes Dev. 24: 521-536. Svendsen JM, Smogorzewska A, Sowa ME, O'Connell BC, Gygi SP, Elledge SJ, Harper JW. 2009. Mammalian BTBD12/SLX4 assembles a Holliday junction resolvase and is required for DNA repair. Cell 138: 63-77. Swagemakers SMA, Essers J, de Wit J, Hoeijmakers JHJ, Kanaar R. 1998. The Human Rad54 Recombinational DNA Repair Protein Is a Double-stranded DNA- dependent ATPase. J. Biol. Chem. 273: 28292-28297. Symington LS. 2002. Role of RAD52 Epistasis Group Genes in Homologous Recombination and Double-Strand Break Repair. Microbiol. Mol. Biol. Rev. 66: 630-670. Symington LS, Gautier J. 2011. Double-Strand Break End Resection and Repair Pathway Choice. Annu. Rev. Genet. 45: 247-271. Symington LS, Heyer WD. 2006. Some disassembly required: role of DNA translocases in the disruption of recombination intermediates and dead-end complexes. Genes Dev. 20: 2479-2486. Symington LS, Holloman WK. 2008. Resolving resolvases: the final act? Mol. Cell 32: 603-604. Symington LS, Kolodner R. 1985. Partial purification of an enzyme from Saccharomyces cerevisiae that cleaves Holliday junctions. Proc. Natl. Acad. Sci. U. S. A. 82: 7247-7251. Takahagi M, Iwasaki H, Nakata A, Shinagawa H. 1991. Molecular analysis of the Escherichia coli ruvC gene, which encodes a Holliday junction-specific endonuclease. J. Bacteriol. 173: 5747-5753. Takata M, Sasaki MS, Sonoda E, Fukushima T, Morrison C, Albala JS, Swagemakers SMA, Kanaar R, Thompson LH, Takeda S. 2000. The Rad51 Paralog Rad51B Promotes Homologous Recombinational Repair. Mol. Cell. Biol. 20: 6476-6482. Takata M, Sasaki MS, Sonoda E, Morrison C, Hashimoto M, Utsumi H, Yamaguchi-Iwai Y, Shinohara A, Takeda S. 1998. Homologous recombination and non- homologous end-joining pathways of DNA double-strand break repair have overlapping roles in the maintenance of chromosomal integrity in vertebrate cells. EMBO J. 17: 5497-5508. Takata M, Sasaki MS, Tachiiri S, Fukushima T, Sonoda E, Schild D, Thompson LH, Takeda S. 2001. Chromosome Instability and Defective Recombinational Repair in Knockout Mutants of the Five Rad51 Paralogs. Mol. Cell. Biol. 21: 2858-2866. Tamkun JW. 1995. The role of brahma and related proteins in transcription and development. Curr. Opin. Genet. Dev. 5: 473-477. Tan TL, Kanaar R, Wyman C. 2003. Rad54, a Jack of all trades in homologous recombination. DNA Repair (Amst) 2: 787-794.

241

Tan TLR, Essers J, Citterio E, Swagemakers SMA, de Wit J, Benson FE, Hoeijmakers JHJ, Kanaar R. 1999. Mouse Rad54 affects DNA conformation and DNA- damage-induced Rad51 foci formation. Curr. Biol. 9: 325-328. Tarsounas M, Davies D, West SC. 2003. BRCA2-dependent and independent formation of RAD51 nuclear foci. Oncogene 22: 1115-1123. Tavtigian SV, Simard J, Rommens J, Couch F, Shattuck Eidens D, Neuhausen S, Merajver S, Thorlacius S, Offit K, Stoppa Lyonnet D et al. 1996a. The complete BRCA2 gene and mutations in chromosome 13q-linked kindreds [see comments]. Nat. Genet. 12: 333-337. Tavtigian SV, Simard J, Rommens J, Couch F, Shattuck-Eidens D, Neuhausen S, Merajver S, Thorlacius S, Offit K, Stoppa-Lyonnet D et al. 1996b. The complete BRCA2 gene and mutations in chromosome 13q-linked kindreds. Nat. Genet. 12: 333-337. Tebbs RS, Zhao Y, Tucker JD, Scheerer JB, Siciliano MJ, Hwang M, Liu N, Legerski RJ, Thompson LH. 1995. Correction of chromosomal instability and sensitivity to diverse mutagens by a cloned cDNA of the XRCC3 DNA repair gene. Proc. Natl. Acad. Sci. U. S. A. 92: 6354-6358. Thacker J, Tambini CE, Simpson PJ, Tsui LC, Scherer SW. 1995. Localization to chromosome 7q36.1 of the human XRCC2 gene, determining sensitivity to DNA- damaging agents. Hum. Mol. Genet. 4: 113-120. Thoma F, Koller T. 1977. Influence of histone H1 on chromatin structure. Cell 12: 101- 107. Thoma NH, Czyzewski BK, Alexeev AA, Mazin AV, Kowalczykowski SC, Pavletich NP. 2005. Structure of the SWI2/SNF2 chromatin-remodeling domain of eukaryotic Rad54. Nat Struct Mol Biol 12: 350-356. Thompson LH, Schild D. 1999. The contribution of homologous recombination in preserving genome integrity in mammalian cells. Biochimie 81: 87-105. -. 2001. Homologous recombinational repair of DNA ensures mammalian chromosome stability. Mutation Research/Fundamental and Molecular Mechanisms of Mutagenesis 477: 131-153. Thorslund T, Esashi F, West SC. 2007. Interactions between human BRCA2 protein and the meiosis-specific recombinase DMC1. EMBO J. 26: 2915-2922. Thorslund T, McIlwraith MJ, Compton SA, Lekomtsev S, Petronczki M, Griffith JD, West SC. 2010. The breast cancer tumor suppressor BRCA2 promotes the specific targeting of RAD51 to single-stranded DNA. Nat Struct Mol Biol 17: 1263-1265. Tjeertes JV, Miller KM, Jackson SP. 2009. Screen for DNA-damage-responsive histone modifications identifies H3K9Ac and H3K56Ac in human cells. EMBO J. 28: 1878-1889. Tombline G, Fishel R. 2002. Biochemical characterization of the human RAD51 protein. I. ATP hydrolysis. J. Biol. Chem. 277: 14417-14425. Tombline G, Shim KS, Fishel R. 2002. Biochemical characterization of the human RAD51 protein. II. Adenosine nucleotide binding and competition. J. Biol. Chem. 277: 14426-14433.

242

Tsubouchi H, Roeder GS. 2003. The Importance of Genetic Recombination for Fidelity of Chromosome Pairing in Meiosis. Developmental Cell 5: 915-925. Tsukuda T, Fleming AB, Nickoloff JA, Osley MA. 2005. Chromatin remodelling at a DNA double-strand break site in Saccharomyces cerevisiae. Nature 438: 379-383. Tsutsui Y, Morishita T, Iwasaki H, Toh H, Shinagawa H. 2000. A Recombination Repair Gene of Schizosaccharomyces pombe, rhp57, Is a Functional Homolog of the Saccharomyces cerevisiae RAD57 Gene and Is Phylogenetically Related to the Human XRCC3 Gene. Genetics 154: 1451-1461. Tsuzuki T, Fujii Y, Sakumi K, Tominaga Y, Nakao K, Sekiguchi M, Matsushiro A, Yoshimura Y, MoritaT. 1996. Targeted disruption of the Rad51 gene leads to lethality in embryonic mice. Proc. Natl. Acad. Sci. U. S. A. 93: 6236-6240. Ulyanova NP, Schnitzler GR. 2005. Human SWI/SNF generates abundant, structurally altered dinucleosomes on polynucleosomal templates. Mol. Cell. Biol. 25: 11156- 11170. Unal E, Arbel-Eden A, Sattler U, Shroff R, Lichten M, Haber JE, Koshland D. 2004. DNA damage response pathway uses histone modification to assemble a double- strand break-specific cohesin domain. Mol. Cell 16: 991-1002. van Attikum H, Fritsch O, Gasser SM. 2007. Distinct roles for SWR1 and INO80 chromatin remodeling complexes at chromosomal double-strand breaks. EMBO J. 26: 4113-4125. van Attikum H, Fritsch O, Hohn B, Gasser SM. 2004. Recruitment of the INO80 complex by H2A phosphorylation links ATP-dependent chromatin remodeling with DNA double-strand break repair. Cell 119: 777-788. Van Komen S, Petukhova G, Sigurdsson S, Stratton S, Sung P. 2000. Superhelicity- Driven Homologous DNA Pairing by Yeast Recombination Factors Rad51 and Rad54. Mol. Cell 6: 563-572. Van Komen S, Petukhova G, Sigurdsson S, Sung P. 2002. Functional cross-talk among Rad51, Rad54, and replication protein A in heteroduplex DNA joint formation. J. Biol. Chem. 277: 43578-43587. Vaz F, Hanenberg H, Schuster B, Barker K, Wiek C, Erven V, Neveling K, Endt D, Kesterton I, Autore F et al. 2010. Mutation of the RAD51C gene in a Fanconi anemia-like disorder. Nat. Genet. 42: 406-409. Veaute X, Jeusset J, Soustelle C, Kowalczykowski SC, Le Cam E, Fabre F. 2003. The Srs2 helicase prevents recombination by disrupting Rad51 nucleoprotein filaments. Nature 423: 309-312. Wang B, Elledge SJ. 2007. Ubc13/Rnf8 ubiquitin ligases control foci formation of the Rap80/Abraxas/Brca1/Brcc36 complex in response to DNA damage. Proc. Natl. Acad. Sci. U. S. A. 104: 20759-20763. Wang B, Matsuoka S, Ballif BA, Zhang D, Smogorzewska A, Gygi SP, Elledge SJ. 2007. Abraxas and RAP80 form a BRCA1 protein complex required for the DNA damage response. Science 316: 1194-1198. Wang X, Haber JE. 2004. Role of Saccharomyces single-stranded DNA-binding protein RPA in the strand invasion step of double-strand break repair. PLoS Biol 2: E21. Ward JD, Muzzini DM, Petalcorin MI, Martinez-Perez E, Martin JS, Plevani P, Cassata G, Marini F, Boulton SJ. 2010. Overlapping mechanisms promote postsynaptic

243

RAD-51 filament disassembly during meiotic double-strand break repair. Mol. Cell 37: 259-272. Weinstock GM, McEntee K, Lehman IR. 1981a. Hydrolysis of nucleoside triphosphates catalyzed by the recA protein of Escherichia coli. Characterization of ATP hydrolysis. J. Biol. Chem. 256: 8829-8834. -. 1981b. Hydrolysis of nucleoside triphosphates catalyzed by the recA protein of Escherichia coli. Steady state kinetic analysis of ATP hydrolysis. J. Biol. Chem. 256: 8845-8849. Welz-Voegele C, Jinks-Robertson S. 2008. Sequence divergence impedes crossover more than noncrossover events during mitotic gap repair in yeast. Genetics 179: 1251- 1262. West SC. 1997. Processing of recombination intermediates by the RuvABC proteins. Annu. Rev. Genet. 31: 213-244. -. 2003. Molecular views of recombination proteins and their control. Nat Rev Mol Cell Biol 4: 435-445. -. 2009. The search for a human Holliday junction resolvase. Biochem. Soc. Trans. 37: 519-526. West SC, Korner A. 1985. Cleavage of cruciform DNA structures by an activity from Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. U. S. A. 82: 6445-6449. Whitby MC, Dixon J. 1997. A new Holliday junction resolving enzyme from Schizosaccharomyces pombe that is homologous to CCE1 from Saccharomyces cerevisiae. J. Mol. Biol. 272: 509-522. Wiese C, Dray EØ, Groesser T, San Filippo J, Shi I, Collins DW, Tsai M-S, Williams GJ, Rydberg B, Sung P et al. 2007. Promotion of Homologous Recombination and Genomic Stability by RAD51AP1 via RAD51 Recombinase Enhancement. Mol. Cell 28: 482-490. Williams AB, Michael WM. 2010. Eviction notice: new insights into Rad51 removal from DNA during homologous recombination. Mol. Cell 37: 157-158. Wold MS, Kelly T. 1988. Purification and characterization of replication protein A, a cellular protein required for in vitro replication of simian virus 40 DNA. Proc. Natl. Acad. Sci. U. S. A. 85: 2523-2527. Wolner B, Peterson CL. 2005. ATP-dependent and ATP-independent roles for the Rad54 chromatin remodeling enzyme during recombinational repair of a DNA double strand break. J. Biol. Chem. 280: 10855-10860. Wolner B, van Komen S, Sung P, Peterson CL. 2003. Recruitment of the Recombinational Repair Machinery to a DNA Double-Strand Break in Yeast. Mol. Cell 12: 221-232. Wong AKC, Pero R, Ormonde PA, Tavtigian SV, Bartel PL. 1997. RAD51 Interacts with the Evolutionarily Conserved BRC Motifs in the Human Breast Cancer Susceptibility Gene brca2. J. Biol. Chem. 272: 31941-31944. Wong JMS, Ionescu D, Ingles CJ. 2003. Interaction between BRCA2 and replication protein A is compromised by a cancer-predisposing mutation in BRCA2. Oncogene 22: 28-33.

244

Wooster R, Bignell G, Lancaster J, Swift S, Seal S, Mangion J, Collins N, Gregory S, Gumbs C, Micklem G et al. 1995. Identification of the breast cancer susceptibility gene BRCA2. Nature 378: 789-792. Wooster R, Neuhausen S, Mangion J, Quirk Y, Ford D, Collins N, Nguyen K, Seal S, Tran T, Averill D et al. 1994. Localization of a breast cancer susceptibility gene, BRCA2, to chromosome 13q12-13. Science 265: 2088-2090. Wu D, Topper LM, Wilson TE. 2008a. Recruitment and dissociation of nonhomologous end joining proteins at a DNA double-strand break in Saccharomyces cerevisiae. Genetics 178: 1237-1249. Wu L, Davies SL, North PS, Goulaouic H, Riou JF, Turley H, Gatter KC, Hickson ID. 2000. The Bloom's syndrome gene product interacts with topoisomerase III. J. Biol. Chem. 275: 9636-9644. Wu L, Hickson ID. 2006. DNA helicases required for homologous recombination and repair of damaged replication forks. Annu. Rev. Genet. 40: 279-306. Wu Y, He Y, Moya IA, Qian X, Luo Y. 2004. Crystal structure of archaeal recombinase RADA: a snapshot of its extended conformation. Mol. Cell 15: 423-435. Wu Y, Kantake N, Sugiyama T, Kowalczykowski SC. 2008b. Rad51 protein controls Rad52-mediated DNA annealing. J. Biol. Chem. 283: 14883-14892. Wu Y, Qian X, He Y, Moya IA, Luo Y. 2005. Crystal structure of an ATPase-active form of Rad51 homolog from Methanococcus voltae. Insights into potassium dependence. J. Biol. Chem. 280: 722-728. Xia B, Sheng Q, Nakanishi K, Ohashi A, Wu J, Christ N, Liu X, Jasin M, Couch FJ, Livingston DM. 2006. Control of BRCA2 Cellular and Clinical Functions by a Nuclear Partner, PALB2. Mol. Cell 22: 719-729. Xia F, Taghian DG, DeFrank JS, Zeng Z-C, Willers H, Iliakis G, Powell SN. 2001. Deficiency of human BRCA2 leads to impaired homologous recombination but maintains normal nonhomologous end joining. Proc. Natl. Acad. Sci. U. S. A. 98: 8644-8649. Xiao A, Li H, Shechter D, Ahn SH, Fabrizio LA, Erdjument-Bromage H, Ishibe- Murakami S, Wang B, Tempst P, Hofmann K et al. 2009. WSTF regulates the H2A.X DNA damage response via a novel tyrosine kinase activity. Nature 457: 57-62. Xu F, Zhang K, Grunstein M. 2005. Acetylation in histone H3 globular domain regulates gene expression in yeast. Cell 121: 375-385. Xu Y, Sun Y, Jiang X, Ayrapetov MK, Moskwa P, Yang S, Weinstock DM, Price BD. 2010. The p400 ATPase regulates nucleosome stability and chromatin ubiquitination during DNA repair. J. Cell Biol. 191: 31-43. Yamada K, Miyata T, Tsuchiya D, Oyama T, Fujiwara Y, Ohnishi T, Iwasaki H, Shinagawa H, Ariyoshi M, Mayanagi K et al. 2002. Crystal structure of the RuvA-RuvB complex: a structural basis for the Holliday junction migrating motor machinery. Mol. Cell 10: 671-681. Yamaguchi-Iwai Y, Sonoda E, Buerstedde J-M, Bezzubova O, Morrison C, Takata M, Shinohara A, Takeda S. 1998. Homologous Recombination, but Not DNA Repair, Is Reduced in Vertebrate Cells Deficient in RAD52. Mol. Cell. Biol. 18: 6430- 6435.

245

Yang H, Jeffrey PD, Miller J, Kinnucan E, Sun Y, Thoma NH, Zheng N, Chen PL, Lee WH, Pavletich NP. 2002. BRCA2 function in DNA binding and recombination from a BRCA2-DSS1-ssDNA structure. Science 297: 1837-1848. Yang H, Li Q, Fan J, Holloman WK, Pavletich NP. 2005. The BRCA2 homologue Brh2 nucleates RAD51 filament formation at a dsDNA-ssDNA junction. Nature 433: 653-657. Yin J, Sobeck A, Xu C, Meetei AR, Hoatlin M, Li L, Wang W. 2005. BLAP75, an essential component of Bloom's syndrome protein complexes that maintain genome integrity. EMBO J. 24: 1465-1476. Yu VPCC, Koehler M, Steinlein C, Schmid M, Hanakahi LA, van Gool AJ, West SC, Venkitaraman AR. 2000. Gross chromosomal rearrangements and genetic exchange between nonhomologous chromosomes following BRCA2 inactivation. Genes Dev. 14: 1400-1406. Zaitseva EM, Zaitsev EN, Kowalczykowski SC. 1999. The DNA binding properties of Saccharomyces cerevisiae Rad51 protein. J. Biol. Chem. 274: 2907-2915. Zakharyevich K, Ma Y, Tang S, Hwang PY-H, Boiteux S, Hunter N. 2010. Temporally and Biochemically Distinct Activities of Exo1 during Meiosis: Double-Strand Break Resection and Resolution of Double Holliday Junctions. Mol. Cell 40: 1001-1015. Zhang Z, Fan H-Y, Goldman JA, Kingston RE. 2007. Homology-driven chromatin remodeling by human RAD54. Nat Struct Mol Biol 14: 397-405. Zhao GY, Sonoda E, Barber LJ, Oka H, Murakawa Y, Yamada K, Ikura T, Wang X, Kobayashi M, Yamamoto K et al. 2007. A critical role for the ubiquitin- conjugating enzyme Ubc13 in initiating homologous recombination. Mol. Cell 25: 663-675. Zhou Q, Kojic M, Cao Z, Lisby M, Mazloum NA, Holloman WK. 2007. Dss1 Interaction with Brh2 as a Regulatory Mechanism for Recombinational Repair. Mol. Cell. Biol. 27: 2512-2526. Zhu Z, Chung W-H, Shim EY, Lee SE, Ira G. 2008. Sgs1 Helicase and Two Nucleases Dna2 and Exo1 Resect DNA Double-Strand Break Ends. Cell 134: 981-994. Zickler D, Kleckner N. 1999. Meiotic chromosomes: integrating structure and function. Annu. Rev. Genet. 33: 603-754.

246