<<

UNIVERSITY OF CINCINNATI

DATE: 3-22-03

I, Gregory T. Langland , hereby submit this as part of the requirements for the degree of:

DOCTORATE OF PHILOSOPHY (Ph.D.)

in:

The Department of Molecular Genetics, Biochemistry and

Microbiology of the College of Medicine

It is entitled:

INTERACTION BETWEEN THE BLM AND THE DNA MISMATCH REPAIR , MLH1

Approved by:

Joanna Groden Ph.D.

Richard Wenstrup M.D.

Jim Stringer Ph.D.

Kathleen Dixon Ph.D.

Peter Stambrook Ph.D.

ii

INTERACTION BETWEEN THE BLM HELICASE AND THE DNA MISMATCH REPAIR PROTEIN, MLH1

A dissertation submitted to the

Division of Research and Advanced Studies of the University of Cincinnati

in partial fulfillment of the requirements for the degree of

DOCTORATE OF PHILOSOPHY (Ph.D.)

In the Department of Molecular Genetics, Biochemistry & Microbiology of the College of Medicine

2002

by

Gregory T. Langland

B.S., University of Cincinnati, 1992

Committee Chair: Joanna Groden, Ph.D.

iii

`Abstract

Bloom’s syndrome (BS) is a rare autosomal recessive disorder that greatly predisposes affected individuals to . Such individuals also are small in size, sensitive to the sun, have immune dysfunction and gross genomic instability. The cytogenetics of BS cells have been extensively studied and have shown increased levels of , quadriradial formations, telomeric associations and breakage. The responsible for BS has been positionally cloned and and encodes a RecQ helicase family member with strand displacement activity that is dependent on ATP and Mg2+.

In order to have a greater understanding of BLM helicase function in the cell in regards to DNA replication, recombination and repair, we identified protein- partners of BLM. The C-terminus of BLM identified the DNA mismatch repair protein MLH1 from a yeast two-hybrid screen. In vitro and in vivo immunoprecipitations confirmed the interaction between these two .

Using an in vitro mismatch repair assay, BS cell extracts were tested for their ability to correct a single nucleotide mismatch. The BS cell extracts were able to remove the single nucleotide mismatch from the plasmid DNA, demonstrating that the BLM-MLH1 interaction is not necessary to correct a single nucleotide mismatch.

To test the hypothesis that MLH1 may regulate the substrate specificity or helicase activity, two different experiments were performed. Gel-shifts were performed plus or minus the presence of MLH1 with different DNA substrates, however MLH1 had no effect on BLM’s ability to bind the different DNA

iv

substrates. Helicase assays then were performed which demonstrated that

MLH1 or the mutL heterodimer modulates the enzymatic activity of BLM by stimulating BLM’s strand displacement activity on the double-overhang (DO) substrate.

Finally, we performed experiments with the supF20 mutagenesis system and demonstrated that extracts from BS cells are unable to utilize micro- elements within the supF20 gene to restore supF function following the induction of a double strand break (DSB). Additional experiments with the pUC18 mutagenesis system demonstrate that although the efficiency and fidelity of DSB repair by BS extracts are comparable to those of normal extracts when ligatable ends are present, a significant 5-fold increase in with BS extracts is observed when terminal phosphates are removed from the DNA substrate that needs repair. Mutant plasmids recovered following DSB repair by

BS extracts contain smaller deletions within the lacZα gene not commonly recovered from normal extracts. Colorectal cancer cell lineHCT116 extracts lacking MLH1 were also examined although the efficiency and fidelity of end- joining was similar to control extracts. This suggests that the BLM-MLH1 interaction is not necessary for proper end-joining.

In summary this work demonstrates that BS cells lacking the BLM helicase process DSBs differently than normal cells and strongly suggests a role for BLM in aligning micro-homology elements during recombinational events in DSB repair. Disruption of the BLM helicase may lead to replication fork collapses, improper processing of DSBs, genomic instability and ultimately cancer.

v

vi

Acknowledgments

Someone once told me anything worth having you will have to work hard and fight for. The completion of my doctoral degree has been no different. It’s more than classwork and benchwork. It has been a learning experience, but let’s just keep it at that. I would like to that everybody that supported me through this effort. Many thanks go to Rick for encouraging me to enter graduate school and gave me the support and encouragement needed to further my career in science. I would like to thank Joanna for having me as a student and all of my committee members for their support and their signature, especially Dr. Peter Stambrook, Dr. Sohaib Khan and my mother for the financial support. Chris H. for showing me that you can give a great thirty-minute seminar with virtually no data. Chris T. for teaching how to make money in graduate school via day-trading and/or fantasy football. Kathy, who will always be my favorite post-doc (Oh, sorry now Assistant Professor). Jenn who Chris T. and I will be fighting for when we have our own labs. Will and Amod will always be remembered as DS-1 and DS-2, they made it fun to be at work. Greg B. for being there and conducting triage after a committee meeting or seminar. I definitely have to thank Al for purifying that nasty protein that I never could. I guess that ten years of protein purification definitely means you have skills. Both Tims were entertaining but I preferred the second one. Chelsea, thanks for the movies and the lunches. James for being there through thick and thin. Lisa, Heather, Mike B.,Dirk, Kevin and Mike for entertaining my wife while I was in graduate school. And last but definitely not least I have to thank Rachel for all the love and support.

Now for the questions I hear every day:

When are you are graduating ??

Where are you going ??

Is Rachel moving with you ??

To Be Continued……

vii

Acknowledgements vii

Abstract iv

Table of Contents 1

Abbreviations 3

List of Figures 5

List of Tables 7

Chapter 1 – Literature Review 8 Cancer and Genomic Instability 8 Bloom’s syndrome 11 Patient phenotype 15 Positional cloning of BLM 17 Functional motifs of the BLM helicase 19 BLM and PML bodies 23 BLM homologs and orthologs 24 The BLM protein exists as an oligomeric form 26 BLM and WRN knockout mice 27 DNA replication 29 Sensitivity of BS to DNA damaging agents 31 The BASC complex 35 Protein-partners of the BLM helicase 36

Chapter 2 – Thesis Rationale 39 Examining the role of the BLM helicase in mismatch repair and double-strand break repair.

Chapter 3 – Material and Methods 40 Cell Culture 40 Reagents and Enzymes 40 Nuclear extract preparation 41 Expression construct generation and characterization 42 Yeast-two hybrid screening 42 Isolation and renaturation of BLM-C 43 IVTT immunoprecipitations 44 Mixed lysate immunoprecipitations 45 In vivo immunoprecipitations 46 Mismatch repair assay 46 supF20 double-strand break repair assay 47 pUC18 double-strand break repair assay 48 Expression and purification of yBLM 49 Preparation of helicase substrates 51

1

Helicase assays 54 Gel-shift assays 55

Chapter 4 – The BLM helicase interacts with MLH1 56 BLM identifies MLH1 in a yeast two-hybrid screen 56 IVTT BLM-C and full-length MLH1 interact 56 Far western assays confirm the interaction between 59 MLH1 and BLM-C BLM and MLH1 interact in vivo 63 DNA mismatch repair activities of BS and control cell 64 extracts are equivalent.

Chapter 5– Stimulation of BLM helicase activity by mismatch repair proteins Mismatch repair in E. coli 66 BLM purification and characterization 69 Gel-shift experiments 73 Helicase assays 75

Chapter 6 – The BLM helicase is necessary for normal 82 double-strand break repair Double-strand break repair in mammalian cells 82 In vitro end-joining assay using the supF20 vector 83 In vitro end-joining assay using the pUC18 vector 83 Sequence analysis of pUC18 mutants 90 Examination of the efficiency and fidelity of 94 End-joining from HCT116 cell extracts.

Chapter 7- Conclusions and Future Directions 97

Chapter 8- References 113

2

Abbreviations aa- amino acid

AT- ataxia telangiectasia

BLM- Bloom’s syndrome gene

BLM- Bloom’s syndrome protein bp-

BS- Bloom’s syndrome

DNA-PK- DNA protein kinase

DO- double overhang substrate

DSB- double-strand break

ENU- N-ethyl-nitrosurea

HU- hydroxyurea

IDL- insertion/deletion loop

IP- immunoprecipitation

IVTT- in vitro transcription/translation

MLH- mutL homologue

MMC- mitomycin C

NE- nuclear extract

NLS- nuclear localization signal nt.- nucleotide

PML- promyelocytic leukemia

PMS- postmeiotic segregation increased

RPA-

SCE- sister-chromatid exchange

SCP- somatic crossover point

SDS-PAGE- sodium dodecyl sulfate- polyacrylamide gel electrophoresis

UDS- unscheduled DNA synthesis

WRN- Werner’s syndrome gene

3

WRN- Werner’s syndrome protein

WS- Werner’s syndrome

4

List of Figures Page No.

Figure 1. DNA repair act as caretakers of the genome. 10

Figure 2. Individual affected by Bloom’s syndrome. 12

Figure 3. BS cells are characterized by high levels of genomic 14 instability.

Figure 4. Functional motifs of the BLM helicase. 21

Figure 5. BLM identifies MLH1 in a yeast two-hybrid screen 57

Figure 6. Immunoprecipitations of in vitro transcribed and 58 translated (IVTT) protein products demonstrate the interaction between the C-terminus of BLM and MLH1.

Figure 7. Mixed lysate immunoprecipitation demonstrates the 61 interaction between full-length BLM and MLH1 or RPA.

Figure 8. Far western assays demonstrate the interaction 62 between the BLM-C terminus and MLH1.

Figure 9. BLM and MLH1 interact in vivo. 63

Figure10. DNA mismatch repair activities of BS and 65 HeLa cell extracts are equivalent.

Figure 11. Co-localization of BLM and hMLH1 in the nucleus 68 of WI-38/VA-13 cells.

Figure 12. The combinatorial specificites of MutS- and 69 MutL-related heterocomplexes.

Figure 13. Silver stain and western blot of BLM fractions 71 purified by nickel chelation affinity chromatography.

Figure 14. The recombinant BLM protein has helicase activity 72 that is ATP and Mg2+dependent.

Figure 15. The BLM helicase has a higher affinity for 4 nt. and 74 12 nt. insertion/deletion loops,

5

Figure 16. Effect of MLH1 on BLM helicase activity. 77

Figure 17. Effect of MLH1/PMS2 heterodimer on BLM helicase 78 activity.

Figure 18. The MLH1/PMS2 heterodimer stimulates BLM 80 helicase activity in vitro.

Figure 19. Effect of MLH1or MLH1/PMS2 heterodimer on 81 strand displacement activity.

Figure 20. Plasmid substrates used in the DSB repair assays. 84

Figure 21. The efficiency and fidelity of double-strand DNA 88 break repair using normal, corrected BS, BS and AT nuclear extracts.

Figure 22. Deletion mutations in plasmids recovered from 91 the in vitro DSB repair assay using pUC18 EcoRI-CIP as the substrate.

Figure 23. The efficiency of double-strand DNA break repair 95 using normal, corrected BS, BS, AT and HCT116 nuclear extracts.

Figure 24. The fidelity of double-strand DNA break repair 96 using normal, corrected BS, BS, AT and HCT116 nuclear extracts.

Figure25. Deletion mutations in plasmids recovered from 97 the in vitro DSB repair assay using pUC18 EcoRI-CIP as the substrate.

Figure 26. Repeat instability with novel larger and smaller alleles. 103

Figure 27. Proposed pathway for the recombinational 106 DNA repair of replication forks

Figure 28. Proposed model of the BLM helicase 108 and its interaction with mismatch repair proteins correcting a lesion on the leading strand.

Figure 29. Model for the role of BLM in DSB repair 111

6

List of Tables Page

Table 1- DNA substrates used in gel-shift and helicase assays 53

Table 2- The efficiency and fidelity of DNA DSB repair of 85 linearized supF20 by nuclear extracts from control and BS cells.

7

Chapter One. Literature Review

Cancer and Genomic Instability

Cancer occurs when normal cellular control mechanisms do not function properly leading to uncontrolled proliferation and subsequent invasion of neighboring tissues. It spreads through the body and sets up secondary growth sites in a process known as metastasis (Hanahan D and Weinberg RA, 2000).

Many different cell types from different organs can develop . Normal cells become malignant by the aberrant expression of proto-oncogenes or the loss of function of tumor suppressor genes. Most oncogenes are derived from normal growth control genes such as growth factors and receptors, signal transduction proteins, transcription factors and cell cycle control proteins (Alberts et al 2002).

Expression patterns of cellular oncogenes and tumor suppressor genes can be altered by local alterations of DNA structure, by chromosomal translocations or by extensive reduplication of DNA regions including an oncogene. Understanding the molecular events responsible for these genetic aberrations is difficult because the processes of eukaryotic DNA replication, repair and recombination are not well understood. Several rare genetic disorders feature both genomic instability and a predisposition to cancer as hallmark features (Figure 1). “Reverse genetics” has proven useful in the identification of the defective genes in these genetic instability syndromes and has shown that

8

many of these disease genes encoded proteins implicated in DNA metabolism

9

Figure 1. DNA repair genes act as caretakers of the genome. Defects in DNA repair are associated with genomic instability and gross chromosomal rearrangements, leading to mutations in genes such as and cancer formation. Adapted from Khanna and Jackson 2001.

10

(Ellis 1997). For example, are mutated in persons with rare inherited human disorders that are characterized by chromosome instability and cancer such as Bloom’s syndrome, Cockayne’s syndrome, trichothiodystrophy, Werner’s syndrome and xeroderma pigmentosum. In order to understand how aberrant

DNA metabolism can lead to cancer formation in a cell, Bloom’s syndrome is an ideal disorder to investigate the relationship between rates of both genomic instability and cancer formation.

Bloom’s syndrome

Bloom’s syndrome (BS) is a rare autosomal recessive disorder first described by David Bloom, a New York dermatologist in the 1950’s (Bloom

1966). The major clinical manifestations are small stature, sun sensitive redness of the face, immunodeficiency, male infertility and the development of cancers of many different types (Figure 2). One of the defining features of the disease is the presence of chromosomal aberrations in BS cells (German 1995). Other genomic instability syndromes such as Fanconi anemia, ataxia telangiectasia and Werner’s syndrome also have elevated levels of chromosome gaps, breaks and rearrangements. Increased rates of both sister-chromatid exchange and quadriradial formations are seen only in BS cells and not in cells derived from any of these other genomic instability syndromes.

11

Figure 2. The clinical presentation of Bloom’s syndrome (BS). Persons with BS are small in size, sun-sensitive, immunodeficient and predisposed to cancer.

12

Examination of BS gave key insights to the pathological nature of the syndrome (Figure 3). Similar to chromosomes from persons with other diseases that lead to genomic instability, BS chromosomes show increased amounts of gaps, breaks and structurally rearranged chromosomes. The two hallmark cytogenetic features of BS most likely arise due to increased rates of genetic exchange in regions of high homology. These exchanges may be between a chromatid of each of the two homologues of a chromosome pair (e.g., between the two chromosome 4s), where the points of exchange are homologous. Following such an exchange, the lesion detectable microscopically at metaphase is a symmetrical four-armed entity known as a quadriradial (Qr).

These exchanges may also be intrachromosomal, occurring between the two sister chromatids of one chromosome; such exchanges are known as sister chromatid exchange (SCE). These SCEs are detectable in a metaphase spread by appropriate cell staining and microscopy. Normal cells display less than ten

SCEs per metaphase spread, while BS cells typically display 50 to 100 SCEs per metaphase spread.

13

Increased chromosome gaps and breaks.

Quadriradial formations.

SCE

Increased sister chromatid exchange.

NO SCE

Figure 3. BS cells are characterized by genomic instability.

14

It has been suggested that the different cytogenetic anomalies seen in cells from BS persons are related to the phase of the cell cycle when the chromosome break occurs. If a break occurs in G1, the broken chromosome end would not unite with the remainder of the chromosome thus forming a gap or break. Breaks in sister chromatids during the S/G2 phase are thought to be responsible for the increased SCE seen in BS cells (Kuhn and Therman 1986).

Two loci have been studied in detail to determine if mutations are also increased in BS cells. The locus encoding HPRT on the X chromosome and the locus that determines the MN blood type on chromosome 4 showed increased numbers of mutations (Tachibana, Tatsumi et al. 1996). Other evidence for genomic instability is the detection of excessive mutations at regions in the genome consisting of highly repetitive sequence (Therman, Otto et al. 1981).

BS phenotype

Persons with BS have small body size with normal proportioning. The face is narrow with prominent ears and nasal features. Hypersensitivity to sunlight to areas of exposed skin during infancy and patchy areas of hyper- and hypo-pigmentation of the skin are common in BS. These pigmentation changes suggest somatic mosaicism in the melanocytes. The voice of a BS patient is high pitched and somewhat coarse and squeaky. There may be a slight excess of minor anatomic anomalies in BS persons, including anomalous digits, wedges

15

of altered color of the iris, localized areas of white hair and obstructing anomalies of the urethra (German 1995).

Many individuals with BS experience a variable degree of vomiting and diarrhea during infancy, often producing severe dehydration. Immunodeficiency of a generalized type is common in BS and shows a wide range of variation in severity ranging from mild or symptomatic to severe. Viral and bacterial infections are common in the respiratory tract and ear canals of BS patients.

Diabetes mellitus has been diagnosed in twenty of the 165 persons in the BS

Registry, and in all cases is late-onset or noninsulin-dependent type II diabetes.

Reduced fertility is seen in both men and women with BS. In males, the testes are abnormally small and display a complete failure of .

Women with BS have an abnormally shortened menstrual life with chronic fertility problems. It is interesting to note that three BS females have given birth to normal children.

BS individuals generally have a restricted intellectual ability. Intelligence is generally average to low average, but can vary greatly. The intellectual ability of a person with BS is not predictable as some persons are mentally deficient while others are normal. Some BS individuals have completed college and obtained professional degrees. However even in the BS persons with normal intelligence, there is a noticeable short attention span.

BS persons are much more susceptible to neoplasia development than normal individuals of the same age. This predisposition to cancer is thought to be responsible for the reduced life span of individuals with BS, which is 28 years.

16

The most impressive aspects of neoplasia in BS are the increased rates at which both benign and malignant neoplasias arise, the wide variety of anatomic sites and cellular types affected, and the exceptionally early age at which time the neoplasias develop. Leukemias and carcinomas are usually responsible for deaths in BS persons. Lymphomas are the most prevalent type of hematopoietic malignancy, while solid tumors are more likely to occur in the colon, skin and breast.

Positional cloning of BLM

Traditional linkage analysis allows scientists to map genes associated with genetic diseases. However, rare autosomal recessive disorders do not often affect large numbers of individuals which hamper these types of statistical analyses. Lander and Botstein first suggested that recessive disorders could be mapped efficiently by studying DNA from consanguineous relationships. This specialized form of gene linkage is known as homozygosity mapping and was used to analyze consanguinitous families with BS. Linkage was found with loci on chromosome 15q (Ellis, Roe et al. 1994). A polymorphic tetranucleotide repeat present in an intron of the FES gene was homozygous in 25 of 26 individuals with BS whose parents were consanguineous. Woodrage et. al. described a patient with features of both BS and Prader-Willi syndrome

(Woodage, et al. 1994). Genetic analysis determined that the ’s were isodisomic, suggesting that there had been a failure of chromosome separation in leading to two copies of a chromosome 15 with a mutation

17

of BLM. McDaniel and Schultz utilized BS cells as recipients for microcell- mediated chromosome transfer to show that the transfer of human chromosome

15 suppressed the elevated SCE levels in this BLM-deficient cell line (McDaniel and Schultz 1992).

The genetic instability of BS includes increased frequency of somatic recombinational events. Ellis et al. made a key observation in 1995 that helped map the BLM gene in a more timely manner (Ellis, et al. 1995). Studying SCE in

BS cell lines showed high levels of SCE in lymphoblastoid cell lines. However, a small sub-number of these cell lines demonstrated low SCE. Analysis of these high and low cell lines at polymorphic loci proximal and distal to the putative BLM locus showed that loci distal to BLM had become homozygous in low SCE lines compared to high SCE lines from the same person, while loci proximal to BLM remained heterozygous. These observations suggested that a recombinational event had occurred in the low SCE population resulting in a functional BLM gene and low SCE. Ellis et al. used the low SCE cell lines with a reduction to homozygosity to map BLM by a novel approach known as somatic crossover point (SCP) mapping. The gene responsible for BS was mapped by determining the genotype of low SCE from five different BS persons. The strategy identified the most proximal heterozygous locus and the distal locus reduced to homozygosity in the low SCE cell lines.

A strong candidate for BLM was identified by direct selection of a cDNA derived from this minimally defined 250 kB region identified by somatic crossover point mapping. Analysis of this candidate gene revealed a 4437 bp cDNA

18

putatively encoding a 1417-amino acid(aa) peptide with homology to the RecQ helicases, a subfamily of DExH box-containing DNA and RNA helicases. The presence of chain-terminating mutations in this candidate gene in BS persons permitted the conclusion that this gene was the disease gene for BS. Mutational analysis of the first thirteen unrelated persons with BS identified 7 unique mutations. Four of the seven mutations resulted in premature termination of the protein suggested the cause of BS is loss of enzymatic function of the BLM gene product.

Functional motifs of the BLM helicase

Analysis of the peptide sequence of the BLM gene product shows many different motifs (Figure 3). The center of the protein includes a helicase domain with seven-conserved helicase motifs present in other molecules with defined helicase activities. Very little is know about the function of these helicase motifs, with the exception of the Walker A and B sites necessary for ATP-binding and subsequent hydrolysis. All ATP-dependent helicases possess the

(GXGXGK[T/S]) which forms a P-loop necessary for ATP-binding. The second

Walker motif is also known as the DEAD or DExH box and contains the essential aspartate (D) that interacts with ATP via Mg2+ (Patel and Picha et al. 2000). Site- directed mutagenesis of either the Walker A and B sites demonstrated that these motifs were necessary for ATP-binding, hydrolysis and enzymatic function of many helicases. It should be noted that the use of primary sequence to detect

19

helicase motifs does not always identify polypeptides that possess strand displacement activity in vitro.

20

1417 1 ACIDIC MOTIFS ACIDIC nuclear staining. nuclear cells shows punctateGFP-C-terminus of BLM transfected intocos functional nuclearlocalization sequence is C-terminal.. Walker A and B boxes for ATP-binding and hydrolysis.The domain residesinthe center of the protein and contains the Figure 4. Functional motifs of helicase. the BLM POL II POL RNA HELICASE DOMAIN RecQ-Ct GFP-BLM-C HRDC The helicase The helicase NLS

21

Ian Hickson’s laboratory was the first to show the BLM protein has helicase activity in vitro using a classic in vitro strand displacement assay. Single

-stranded DNA is purified and a complementary radiolabeled oligo is annealled to the M13 template (Karow, et al. 1997). This DNA duplex substrate is mixed the potential helicase with ATP and co-factors crucial for the enzymatic reaction.

Helicase activity is determined by detecting the displaced oligo by native gel electrophoresis and autoradiography. Utilizing this strand displacement assay, purified human BLM protein (hBLM) was shown to possess a 3’ to 5’ DNA-DNA strand displacement activity that was dependent on ATP and Mg2++. The

ATPase activity of this helicase was stimulated by the presence of DNA.

Mutations made in the Walker A site resulted in a loss of helicase activity. This directionality is characteristic of the RecQ subfamily of helicases; E. coli RecQ, yeast Sgs-1 and the human WRN helicase also act on DNA in a 3’ to 5’ dependent manner with respect to the labeled oligo (Chakraverty and Hickson

1999). The BLM helicase also acts upon non-canonical Watson-Crick structures such as G4 DNA which is present in G-rich motifs present in the rDNA gene clusters, the immunoglobulin heavy chain switch regions and telomeric repeats

(Sun, et al. 1998).

Many nuclear proteins contain short amino acid sequences that direct them to the nucleus of a cell. These nuclear localization sequences (NLS) interact with specific nuclear import receptors to permit access to the nucleus.

For a putative NLS to be considered functional, it must pass several tests.

Deletions or mutations of nuclear localization signal in a protein should show that

22

the protein is excluded from the nucleus or show strictly cytoplasmic staining by immunohistochemistry or immunofluorescence. Attaching a putative NLS to a large non-nuclear protein such as β-glucosidase and showing nuclear import of the chimera suggests the NLS is functional. The bipartite (two stretches of positively charged residues) NLS in the carboxy-terminal domain of BLM was shown to be functional (Figure 4). (Heppner-Goss personal communication) At least two other putative nuclear localization sequences are present in the BLM protein (one in the helicase domain and the other in the amino terminal domain, although there is no experimental evidence that suggests these NLS are functional (Kaneko, Orii et al. 1997).

BLM and PML-bodies

The BLM helicase localizes to discrete nuclear bodies commonly known as PML (promyelocytic leukemia) bodies, a distinctive class of subnuclear domain (Zhong, et al. 1999). These domains appear by indirect immunofluorescence microscopy as punctate speckles and can be referred to as

Kremer bodies, ND 10 or PODs (PML oncogenic domains). PML was identified by analysis of a chromosome translocation breakpoint in leukemic cells from APL

(acute promyelocytic leukemia) patients. Gene-targeted disruption of PML demonstrated that PML functions as a growth and tumor suppressor in vivo and is important for multiple apoptotic pathways. BLM protein is diffusely localized in cells null for PML. In contrast, PML does form discrete foci without BLM. SCE

23

rates in cells null for PML displayed a two-fold increase when compared to control cells (Zhong, et al. 1999).

BLM homologs and orthologs

To understand the function(s) of BLM helicase, we can examine the function of RecQ subfamily members in DNA metabolism. The founding member of the RecQ subfamily of DNA/RNA helicases is RecQ, a 70kD polypeptide originally isolated from E. coli mutants in the RecF recombination pathway

(Umezu, et al. 1990). Mutations in the RecF pathway cause a recombination deficiency and increased ultraviolet (UV) sensitivity in a recBC shcB background.

Promoter studies of the E. coli RecQ helicase suggested that the E. coli RecQ promoter is normally repressed by the LexA repressor binding sites ten nucleotides downstream of the transcription start site. RecQ in E. coli was inducible by both UV light and treatment with mitomycin C (MMC) (Bierne,

Seigneur et al. 1997). The E. coli RecQ promoter possesses lexA repressor binding sites and is inducible by both UV light and MMC, suggesting that SOS regulation mediates RecQ (Mendonca, et al. 1995). The E. coli

RecQ protein possesses an ATP-dependent and 3’ to 5’ helicase activity in vitro.

Its helicase activity was stimulated by several single stranded-binding proteins

(SSB’s), including T4 gene 32 protein and E. coli SSB (Umezu and Nakayama

1993).

RECQL/DNA helicase Q1 is the human homologue to the E.coli RecQ protein (Puranam and Blackshear 1994). These two proteins are equivalent in

24

size and share homology at the amino acid level. Two isoforms of RECQL have been cloned in mice (Puranam, et al. 1995). The alpha form is expressed in a ubiquitous manner, while the beta form is expressed specifically in the testis. In humans there are at least four different isoforms of RECQL. The chromosomal location and genetic sequence of RECQL have been determined but few expression studies are reported in the literature. It is interesting to note that mutations in RECQL4 have been genetically linked to a rare autosomal recessive disorder associated with genomic instability known as Rothmund-Thomson syndrome (Kitao, et al. 1999).

The BLM helicase is similar in molecular weight to the RecQ-like helicase associated with Werner’s syndrome (WRN). The seven-helicase motifs present in the WRN helicase domain reside roughly in the center of the molecule with other significant functional motives in either the N- or C-terminal domain. The location of the NLS in the C-terminal portion of the polypeptide prevents truncated molecules from nuclear entry. Homozygous mutations of the WRN helicase result in the rare autosomal recessive disorder Werner’s syndrome (WS) associated with premature aging (Epstein and Motulsky 1996). Even though both BS and WS are caused by the loss of function of a RecQ helicase, the phenotype of each rare disorder is very different.

Sgs-1 (suppressor of slow growth phenotype) was originally identified in

Saccharomyces cerevisiae because it rescues the slow growth phenotype of the topIII mutant (Watt, et al. 1996). Yeast lacking functional topoisomerase III display a pleiotropic phenotype including slow growth, G2/M arrest and genomic

25

instability. Mutations of Sgs-1 suppressed this phenotype in the top3-deficient yeast. Analysis of the region responsible for the suppression of the top3 mutants revealed a gene encoding a RecQ helicase. Further studies revealed that Sgs-1 also interacts with topoisomerase II in yeast and has strand displacement activity in vitro (Watt, et al. 1995).

Rqh-1 is the fission yeast homologue of the BLM helicase. Loss of Rqh-1 function results in yeast that are more sensitive to UV, gamma radiation and the

DNA synthesis inhibitor, hydroxyurea (Stewart, et al. 1997). Overexpression of

Rqh-1 results in cells that are more sensitive to both UV and gamma radiation, and loss of S and G2-phase cell cycle checkpoint control. These results suggest that this helicase is important in both DNA metabolism and cell cycle progression.

The BLM protein exists in an oligomeric form

The quaternary structure of this recombinant protein has been determined using size-exclusion chromatography and electron microscopy (Karow, et al.

1999). Size exclusion chromatography demonstrated that the major ATPase activity eluted at an estimated molecular weight of 800 kD to 1 MegaD.

Antibodies to BLM identified the BLM helicase in these same fractions. Electron microscopy showed that the BLM helicase forms hexameric ring structures with an overall diameter of 13nm with a central hole of 3.5nm diameter. A fourfold square form was also detected, but may represent one of several oligomeric species. The fourfold symmetric images may represent an oligomeric form

26

distinct from the hexameric ring such as a tetramer or octamer. It was also suggested this image could be the result of the side view of a two-tiered hexameric barrel structure such as the bacteriophage T7 helicase/primase. Most helicases characterized to date act as dimers or hexamers. However, BLM is the first eukaryotic oligomeric helicase. More importantly, this suggests the BLM helicase may form an oligomer and encircle a continuous DNA strand at any internal site without the requirement for a single stranded DNA end. This function may be crucial for the BLM helicase to function in DNA replication to open replication sequences or to aid the transcriptional machinery of the cell by melting promoter sequences.

Blm and Wrn “knockout” mice

Phil Leder’s group was the first to use gene-targetting to target the endogenous Blm gene in mouse (Chester, et al. 1998). A site upstream of the helicase domain was targeted for disruption by homologous recombination in embryonic stem cells. This would disable both the helicase and NLS activity from the protein. Unlike the patient, a mouse without a functional BLM gene results in an embryonic lethal phenotype. Specifically, Blm-

/- embryos do not survive past 13.5 days past conception. These embryos also are smaller than wild type controls and exhibit developmental delays and increased rates of apoptosis. As an indication of DNA damage micronuclei were present in the white blood cells of these embryos. Sister chromatid exchange in

BLM-/- embryonic fibroblasts was increased seven fold demonstrating that loss of

27

helicase function leads to genomic instability, increased rates of apoptosis and embryonic lethality.

Interestingly, this group has also targeted the Werner’s gene in mice (Wu,

He et al. 1998). Homologous recombination generated the mouse helicase domain replaced by a neo-cassette. The targeting strategy employed could potentially give rise to two differently processed mutant WRN transcripts. The first transcript would produce a truncated WRN protein plus the neo cassette.

The second mutant transcript would code for the WRN protein minus helicase motifs three and four. Both of these WRN proteins are presumed to lack helicase activity. The knock out had a mildly increased rate of embryonic lethality and sensitivity to topoisomerase inhibitors. However, mice lacking functional WRN helicase display a relatively normal phenotype during the first year of life.

DNA replication

The unusual cytogenetic anomalies associated with Bloom’s syndrome suggested that there was some inherent defect in DNA metabolism present in cells derived from BS patients. Many early studies involved determining if there are notable differences in DNA regulation between normal and BS cells. Many subtle differences were found in regards to DNA regulation. DNA fork displacement rates were measured in three BS cell lines and compared to normal fibroblasts (Kapp 1982). In each of the BS cell lines tested, the DNA fork displacement rate was less than 65% of the values in the normal fibroblasts.

Nucleotide content of normal and BS cells were measured and the ATP/ADP

28

ratio was three times lower in the BS cells compared to control (Kenne and

Akerblom 1990). This altered nucleotide pool was hypothesized to be responsible for the genetic abnormalities of BS due to the lack of energy for DNA synthesis/repair enzymes and could be responsible for the phenotypic slow growth.

Size analysis of genomic DNA showed differences in the sizes of the DNA replication intermediates between normal and BS cells suggesting some inherent defect in DNA replication (Ockey and Saffhill 1986; Lonn, Lonn et al. 1990).

Although it should be noted that these differences in DNA replication intermediates were also detected in low SCE populations of BS cells suggesting the molecular mechanism responsible for these intermediates is not the absence of the BLM helicase but of another mutation that was generated in this mutator phenotype.

The ability of BS cells to join double strand breaks (DSB) in vivo was measured by Runger and Kraemer in 1989 (Runger and Kraemer 1989).

Linearized plasmid DNA was transfected in either BS cells or normal cells as a control. BS cells have a 3 fold lower rate of ligation efficiency when compared to the control cells. Mutation frequency was 2 to 21 fold higher in the BS cells as compared to control, and analysis of these mutations revealed point mutations to be predominating factor causing loss of transgene function with the occurrence of insertions and deletions as well.

DSB plays a crucial role in normal immunoglobulin gene processing.

Combinatorial joining occurs between the V,D, and J segments and variability in

29

the position of joining and nucleotide addition in heavy chain genes present putative molecular mechanisms in DNA regulation that could be altered in BS cells. These altered processes help explain the genomic instability and the impaired immune function of patients with BS.

Finally, it has been demonstrated that cells exfoliated from BS patients are ten times more likely to be micronucleated (Rosin and German 1985). Micronuclei arise when some chromosome fragment lacking a kinetochore necessarily lags at anaphase to be encompassed with its own private nuclear membrane at the time that the two main groups of telophase chromosomes are surrounded by the major nuclear membrane of each of the daughter cells.

These data suggests the BLM helicase have functional roles in DNA replication and DSB repair. Differences in DNA fork displacement rates and DSB repair functions point to specific pathways where the BLM helicase could play a functional role. Since this helicase is implicated to be important in DNA replication and repair and lack of this helicase leads to cancer, many studies have focussed on response to DNA damaging agents and subsequent cell-cycle progression.

Sensitivity of BS cells to DNA damaging agents

Many studies have utilized the BS cell lines to examine their sensitivities to different DNA damaging agents and measuring sister chromatid-exchange, cell cycle-progression and the expression patterns of proteins shown to be important in cell cycle progression and genomic stability such as p53. Giannelli

30

and others decided to study the effects of both UV irradiation and specific DNA polymerase inhibitors on unscheduled DNA synthesis (UDS) in 1982 (Giannelli,

Botcherby et al. 1982). This study helped lead the way to study DNA damaging agents and cell cycle progression in BS cells. UV irradiation induced more UDS in ten BS cell lines compared to control. Aphidocolin could suppress this UDS in nine of the ten BS cell lines (Krepinsky, Rainbow et al. 1980). This suggests one possible function of the BLM helicase is to somehow suppress DNA replication at extraneous sites, possibly where DNA damage has occurred. The results of the study provided the field with two directions. The first was to examine the sensitivity of BS cells to other DNA damaging agents. Secondly, the cells could be examined to determine if cell cycle-checkpoints and other proteins important in monitoring cell cycle-progression and genomic stability are functioning correctly.

Several studies have been conducted to measure the sensitivity of BS cells to the DNA damaging agent mitomycin C (MMC). MMC is an alkylating agent, which produces bulky monoadducts and interstrand crosslinks. Several studies have been performed with BS cells to measure their sensitivity to MMC by monitoring their viability and rates of SCE. Shiraishi and Sandberg were the first group to show BS lymphocytes were more sensitive to MMC by measuring viability and SCE (Shiraishi and Sandberg 1978). BS lymphocytes were much more sensitive to MMC and SCE increased at least two fold compared to normal lymphocytes. However, another group had performed similar experiments and

BS fibroblasts were much more sensitive to MMC in the cell viability assays but

31

did not present the increased rates of SCE (Hook, Kwok et al. 1984). It was suggested that the differences reported could be due to the different BS cells used. The first group used lymphocytes while the second group used fibroblasts.

Another possibility is that is considerable heterogeneity between the different BS patients, which is present in the varied phenotype of this disorder.

BS cells have also been shown to be sensitive to ethylating agents such as N-ethyl-nitrosourea (ENU) (Kurihara, Inoue et al. 1987). Unlike the modest increase of SCE due to MMC exposure, ENU shows a greatly enhanced effect on the BS cell lines compared to control. BS cells have been shown to be four times as sensitive to the lethal effects of ENU and SCE increased 60 fold in BS cell lines.

The first study to address specific role of BLM in a cell cycle-specific manner utilized the fact that gamma radiation of mammalian cells in G2 phase just prior to mitosis produces chromatid aberrations such as breaks and gaps (Aurias,

Antoine et al. 1985). It was hypothesized that there would be more breaks and gaps in skin fibroblasts from individuals with different types of familial cancer and from fibroblasts derived from individuals with genetic diseases that predispose them to cancer. Gamma radiation was used to induce breaks in the DNA. Cells were trapped in mitosis one and a half hours later using colcemid (this ensured that the cells counted for the study were in the G2 phase when exposed to the radiation), and mitotic cells were stained and chromatid breaks and gaps were counted. There were increased rates of chromatid breaks and gaps in the fibroblasts derived from familial cancer as well as the cancer prone syndromes

32

compared to control fibroblasts. Specifically, this suggests that the BLM helicase may play an active role in in the G2 phase of the cell cycle.

The relationship between ultraviolet (UV) light and its effects on BS cells has been intimately studied due to sun sensitivity present in the affected individual (Evans, Adams et al. 1978). BS cells were ten times more sensitive to

UV light as measured by cell viability assays and SCE increased ten fold when exposed (Krepinsky, Rainbow et al. 1980). To help explain the increased genomic instability following exposure to UV light, the incision step utilized by the cell to remove the photoadduct was measured in BS cells compared to normal fibroblasts. Incision products were detected less frequently in BS cells suggesting a defect in either the detection of the photoadduct or the actual incision step itself is somehow inhibited or delayed due to the lack of the BLM helicase.

Absence of functional p53 or other proteins important to genome integrity and cell cycle progression could help explain the sensitivity of BS cells to DNA damaging agents. It has been well established that p53 is activated at both transcriptional and post-transcriptional levels by a variety of DNA damaging agents. Activation of p53 by DNA damage arrests the cells in the G1 phase of the cell cycle; preventing DNA replication from occurring on damaged templates.

Cells mutant in p53 function will progress normally through the cell cycle entering

S-phase even in the presence of damaged DNA. Van Laar et. Al. ascertained the p53 status in cell lines derived from hereditary disorders associated with UV

33

sensitivity which included BS (van Laar, Steegenga et al. 1994). First, this study gives a FACS analysis profile of BS cells compared to controls. BS cell lines show similar percentages of cells in different phases of the cell cycle. Nine out of ten BS cell lines arrest in G1 when exposed to UV light suggesting functional p53 is present. Thus the G1 checkpoint of the cell cycle is functional in most BS cell lines. It is interesting to note that rqh1 in fission yeast regulates cell cycle checkpoint control and is homologous to the BLM helicase.

More recently Judith Campisi’s group has studied the BLM helicase and its response to DNA damage (Bischof, Kim et al. 2001). The BLM helicase resides in PML-bodies during late S and G2 phases of the cell cycle. BLM co- localized with RAD51 and RPA after cells were treated with ionizing radiation.

Finally, they observed a G2 delay where BLM protein levels accumulated that was both ATM and p53 independent. Other groups have reported that BLM protein levels decrease when exposed to UVC and observe BLM phosphorylation when cells are treated with HU or UVC (Ababou, Dumaire et al. 2002). Pichierri and Franchitto showed that functional BLM is required for the localization of the

RAD50-MRE11-NBS1 complex to sites of replication arrest but is not essential in the activation of BRCA1 either after stalled replication forks or gamma irradatioin.

This group reported that the BLM protein is phosphorylated after replication arrest in an Ataxia and ATR-dependent manner, which is contrary to the Campisi report (Franchitto and Pichierri 2002).

34

The BRCA1-associated genome surveillance complex

The BLM helicase exists in an oligomeric form and is known to exist in a large 2-mega Dalton with other proteins intimately involved in DNA repair. This complex has been named BASC (BRCA1-associated genome surveillance complex) (Wang, Cortez et al. 2000). Members of this complex include BRCA1,

RAD50-MRE11-NBS1, MSH2 and 6, MLH1, ATM and BLM. Members of this complex can be co-immunoprecipitated from HeLa nuclear extracts and co- localize when cells are treated with specific DNA damaging agents. Both HU and ionizing radiation greatly increased the number of foci that contained both

BRCA1 and BLM compared to untreated cells. The BLM helicase also colocalized with the RAD50-MRE11-NBS1 complex when cells were exposed to

HU strongly suggesting that all these proteins work in unison to resolve these aberrant DNA structures.

It is interesting to note that many of the proteins associated with BASC possess the ability to bind abnormal DNA structures, such as double-strand breaks, base-pair mismatches, Holliday junctions, cruciform DNA and telomere repeat sequences. Therefore it is possible that these proteins may act as sensors for aberrant DNA structures that arise during DNA replication, recombination and repair. RAD50-MRE11-NBS1 and ATM may detect double- strand breaks and the mismatch repair proteins may detect distortions in the helix due to the chemical modifications of bases or the presence of IDLs

(Insertion/deletion loops). The MSH2-MSH6 heterodimer not only binds to mismatched DNA but also has affinity for Holliday junctions thereby acting as

35

sensors of recombination and replication fork damage. BLM has affinity for forked DNA, synthetic Holliday junctions and telomeric G4 DNA.

BASC also has signal transducers such as the ATM kinase and ATR.

This would allow for a rapid response to activate cell-cycle checkpoints through phosphorylation and activation of p53 and Chk proteins.

Protein partnering

Protein partners can provide essential clues to the function of a specific gene in the cell. The previous section of this review discussed the proteins that

BLM existed in a complex. This section of the review is dedicated to proteins that directly interact with BLM or possess a functional interaction.

The first protein shown to co-localize and co-immunoprecipitate with the

BLM protein was human topoisomerase IIIα (Johnson, Lombard et al. 2000; Wu,

Davies et al. 2000). Topoisomerases are enzymes known to relieve torsional stress introduced into DNA when complemetary strands are separated by helicases. This interaction was expected, since and yeast TOP3α have been shown to interact both physically and genetically. BLM and human topoisomerase IIIα are able to bind to each other in vitro showing that the interaction is direct. Topo IIIα has two different binding regions in the BLM helicase. The N-terminal binding region is aa 1-212 and the C-terminal binding region is aa 1266-1417. BLM and yeast TOPIII were shown to interact genetically similar to the SGS1 TOPIII interaction. The slow growth phenotype seen in the top3 background can be generated by a BLM contruct with the

36

topoisomerase interaction domains deleted in a sgs1 backgound. These data strongly suggest that topoisomerases are required to interact with helicases directly to resolve the topological substrates that the helicases generate during unwinding duplex DNA.

Replication protein A (RPA) is the eukaryotic version of bacterial single- stranded binding protein and is intimately associated with many different aspects of DNA metabolism (Brosh, Li et al. 2000). Far western analysis showed a direct interaction between the BLM helicase and 70-kDa subunit of RPA. RPA stimulates BLM helicase activity and is necessary for BLM to be processive on long tracts of DNA (> 250bp) RPA and BLM colocalize in meiotic prophase of mouse spermatocytes further suggesting an in vivo role for this interaction

(Walpita, Plug et al. 1999).

The RAD51 protein is key to homologous recombination in eukaryotes and

Hickson’s group was the first to detect the interaction between the BLM helicase and RAD51 via a yeast two-hybrid screen (Wu, Davies et al. 2001). The C- terminal domain of BLM (aa 966-1417) interacted with human RAD51 protein.

Far western blots determined that RAD51 interacted with both the N-terminus of

BLM (aa 1-212) and C-terminus (aa 1317-1417). BLM and RAD51 co-localize to discrete foci in response to gamma-irradiation. BLM and RAD51 exist as a complex in cells arrested with aphidicolin. SGS1 and yeast RAD51 were shown to interact demonstrating that this interaction is evolutionarily conserved.

p53 and BLM interact both in vitro and in vivo and p53 mediated apoptosis is reduced in BS cells (Wang, Tseng et al. 2001). Restoring BLM expression to

37

BS cells allows p53-mediated apoptosis to occur. p53 also attenuates BLM’s ability to unwind Holliday junctions in vitro. p53 also has been shown to associate with recombinative repair complexes during S phase with MSH2, however the presence of BLM has not been examined (Zink, Mayr et al. 2002).

Helicases are important tools for the cell to regulate DNA replication, repair, recombination and gene transcription. Understanding the function of these enzymes and how the cell regulates their activities will provide key insights into how mutations can spontaneously arise due to errors in nucleic acid metabolism, which in turn may lead to cancer formation. Many helicases have been identified by either traditional biochemical purification or by positional cloning. Many of these helicases unwind duplex nucleic acid molecules in vitro, but very little is known about their specific functions or activities in the cell. It is currently hypothesized that different families of helicases function by a common molecular mechanism and different cell types will utilize the mechanistic features of the particular helicase. An example of this would be the use of different types of helicases in the immune system to regulate and generate immunoglobulin diversity through genetic means. Therefore understanding the regulation of the

BLM helicase will provide clues about how the cell regulates the function of BLM and possibly other eukaryotic RecQ helicases, that when mutated lead to rare autosomal recessive disorders associated with increased incidents of cancer formation and genomic instability.

38

Chapter Two

Thesis Rationale: Examining the Role of the BLM helicase in Mismatch repair and Double-Strand Break Repair.

The aim of this work is to determine the role of the BLM helicase in regards to DNA repair. Many other helicases play active roles in DNA repair.

RecQ helicase is responsible for regulating recombination in E. coli via the RecF repair pathway. Another example is the UvrD helicase which is necessary for strand displacement during mismatch repair in E. coli . Eukaryotic helicases also play functional roles in DNA repair.

Absence of RecQ helicases in mammalian cells results in increased levels of recombination and increased mutation frequency. WS cells are sensitive to the

DNA damaging agent 4-nitroquinone and BS cells are mildly sensitive to a variety of DNA-damaging agents. The BLM helicase exists in a 2-megadalton complex with other proteins implicated in mismatch repair and double-strand break repair.

Therefore we hypothesized that the BLM helicase may play crucial roles in mismatch and double-strand break repair. To further understand the role of the

BLM helicase in DNA repair, we performed yeast-two hybrid screens to identify direct protein partners of the BLM helicase that could directly implicate this RecQ helicase in specific pathways of DNA repair and we performed helicase an in vitro double-strand break and mismatch repair assays to further understand the role of the BLM helicase in DNA repair.

39

Chapter Three

Materials and Methods

Cell Culture

HCT116 cells were purchased from the American Type Culture Collection

(ATCC). HCT116 cells were cultured in McCoy’s 5a Medium (Life Technologies,

Inc.). The SV40-transformed BS fibroblast cell line, GM08505C, was purchased from Coriell Laboratories (Camden, NJ). This cell line was derived from an

Ashkenazi Jewish female, homozygous for the BLMASH allele (a 6 bp deletion/7 bp insertion at nucleotide 2281). Cells were grown in Dulbecco’s Modified

Minimal Essential Medium (Life Technologies), supplemented with 10% fetal bovine serum (FBS, Hyclone). The SV40-transformed ataxia telangiectasia (AT) fibroblast cell line, AT5BIVA, was purchased from the Human Genetics Mutant

Cell Repository (Camden, NJ). The SV40-transformed normal human fetal lung fibroblast cell line (W138VA13) was purchased from American Type Culture

Collection (Rockville, MD) and used as a control. Both AT and normal cells were grown in Eagle’s Minimal Essential Medium (Life Technologies), supplemented with 10% FBS, 100 U/ml penicillin,100µg/ml streptomycin.

Reagents and Enzymes

Restriction endonucleases and T4 DNA ligase were purchased from New

England Biolabs. PCR was performed using Pwo polymerase from Roche

Molecular Biochemicals. Oligonucleotides were synthesized on an Applied

40

Biosystems synthesizer and DNA sequencing was performed at the University of

Cincinnati DNA Core Laboratory. Plasmids were purified using Qiagen purification kits. In vitro transcription and translation (IVTT) reactions were performed using the Promega TNT-coupled rabbit reticulocyte lysates system.

Radiolabeled [35S]methionine from PerkinElmer Life Sciences was used to label proteins. To detect BLM by western blot, a polyclonal antibody to BLM was purchased from Novus. To immunoprecipitate BLM, goat polyclonal antibodies from Santa Cruz, Both BLM-1 and 2, were used, while MLH1 western analyses and immunoprecipitations were performed with monoclonal antibodies from

PharMingen. An α-replication protein A 70-kDa subunit antibody (Ab-1) and protein A/G-agarose were purchased from Calbiochem.

Nuclear Extracts

Nuclear extracts were prepared from cell lines based on the modified method of Lopez and Coppey as described in Li et al. with all the steps performed at 4°C. 1 x 108 cells were harvested, washed in PBS, and resuspended in 4 mL of hypotonic buffer [20 mM Hepes (pH 7.5), 5 mM KCl, 1.5 mM MgCl2 and 1mM DTT]. Cells were gently lysed using a Dounce homogenizer and lysis was monitored by trypan blue staining. The released intact nuclei were recovered by centrifugation at 2000 x g for 1 min. The nuclei were extensively washed before the nuclear envelope was broken by three cycles of freezing and thawing. Debris was pelleted and soluble nuclear protein was recovered from the supernatant by ammonium sulfate precipitation. The

41

precipitate was resuspended in 50 mM Tris (pH 7.5), 0.1 mM EDTA, 10 mM 2- mercaptoethanol, 0.1 mM phenylmethylsulfonyl fluoride, and 10% glycerol and dialyzed overnight against the same buffer. Nuclear extracts were snap frozen and stored in small aliquots at –80°C.

Expression construct generation and characterization

BLM-pET constructs for protein expression were derived by PCR using the pOPRSVI-BLM construct as a template (Ellis, Proytcheva et al. 1999). The pET30A-BLM-C construct was generated by PCR amplification with Pwo polymerase and cloning of BLM nucleotides 2931-3995 directionally with the

EcoRI and BamHI sites present in the polylinker. The pET24D-BLM-C construct was generated by PCR amplification and cloning of BLM nucleotides 3063-4325 into the NcoI site in the polylinker. The pET24D-BLM-N (nucleotides 75-1838) and the pET24D-BLM-H (nucleotides1839-3077) constructs were generated by

PCR amplification of BLM and cloning into the NheI site. Full-length MLH1 was generated by PCR from the plasmid identified in the yeast two-hybrid screen and subsequently was cloned into the pRSETB plasmid. IVTT reactions were analyzed by western analysis with antibodies specific for either BLM or MLH1.

Products from IVTT reactions were also verified by immunoprecipitation.

Yeast-two hybrid screening

A yeast two-hybrid screen was performed using the Matchmaker Two-

Hybrid System from Clontech. DNA encoding the C-terminus of BLM

42

(nucleotides 3108-4319) was cloned into the pAS2 vector by PCR. The C- terminus of BLM was used as bait to screen a human B-lymphocyte cDNA library also purchased from Clontech. Transformants were plated onto a trp-, leu-, his- plus 50 mM 3-AT plates. Plates were incubated at 30°C for 10 days and the growth of the yeast monitored. β-galactosidase assays for potential interacting clones were performed according to the instructions provided by Clontech

Matchmaker Two-Hybrid system.

Isolation and renaturation of BLM-C

BLM-C protein was expressed from pET30A-BLM-C in E. coli BL21

(DE3) pLysS induced with 0.5 mM isopropyl-1-thiogalactopyranoside, collected by centrifugation, rapidly frozen and stored at –80°C. All subsequent procedures were performed at 4°C. Cells were thawed and resuspended in ice cold 10 mM

Tris-HCl pH 7.5, 150 mM NaCl, 1 mM KCl supplemented with a bacterial protease inhibitor cocktail (Sigma). Cells were passed through a french press once at 20,000 pounds per square inch. Cell lysates were centrifuged at 15000 x g for one hour. Since the majority of the BLM-C protein resided in the insoluable fraction, the pellet was resuspended in his-binding buffer (5 mM imidazole, 500 mM NaCl with 8 M urea, centrifuged as before and passed through a 0.4 micron filter before application to a Ni(II)-agarose column (His-Bind Resin, Novagen).

Protein was eluted with his-binding buffer containing 8 M urea and 250 mM imidazole. Fractions were separated by 15% sodium dodecyl sulfate- polyacrylamide gel electrophoresis (SDS-PAGE) and stained with Coomassie

43

blue to determine yield and purity. Fractions that were more than 80% pure were dialyzed in a stepwise fashion in 10 mM Tris-HCl pH 7.5, 150 mM NaCl, 10% glycerol with decreasing amounts of urea to obtain soluable protein.

Approximately 30% of the total protein remained in solution after removal of urea.

Far western analysis

Twenty-five µg of purified BLM-C terminus protein was resolved by 15%

SDS-PAGE. The protein was transferred to PVDF membrane (Millipore) in transfer buffer at 90V for 1 hour. The membrane was stained with Ponceau S to detect the protein and this portion of the membrane was excised. The membrane and bound protein were blocked in 5% non-fat dry milk (NFDM) in tris-buffered saline (TBS) for 1 hour and incubated with 25 µg of nuclear extracts for 1 hour on ice. The membrane was washed four times with TBS-Tween 20

(0.1%) and the bound proteins eluted by boiling in 1X SDS-PAGE loading buffer.

These samples were separated by SDS-PAGE, transferred to PVDF membrane and probed with an α-MLH1 antibody (Pharmigen) to detect MLH1.

IVTT immunoprecipitations

IVTT (Promega) reactions with [35S]-methionine were performed with pET24D-BLM-C or pRSETB-MLH1 using purified DNA (Qiagen) according to the manufacturer’s recommendations. Thirty µl of each IVTT reaction was added to a total volume of 500 µl of binding buffer (20 mM Tris pH 7.5, 10% glycerol, 150 mM NaCl, 5 mM EDTA, 1 mM dithiothreitol, 0.1% Tween-20, 1 mM

44

phenylmethylsulfonyl fluoride and mammalian protease inhibitors (Sigma)).

Proteins were incubated for 1 hour to allow binding before they were precleared with 25 µl of protein A/G beads. Five µg of primary antibody and 25 µl of protein

A/G were used in each reaction. The beads were washed four times with binding buffer and then boiled. Proteins were separated by 10% SDS-PAGE. The [35S] signal was intensified with Enhance (NEN-DuPont) following the instructions of the manufacturer.

Mixed lysate immunoprecipitations

Full-length human BLM cDNA with a hexahistidine tag at the 5’ end was cloned into the pFASTBAC-HTc vector purchased from Gibco-BRL to create the plasmid pFASTBAC-HTc-BLM . Virus production for insect cell infection was performed according to the protocol provided by the Bac-To-Bac Baculovirus

Expression Systems (Gibco-BRL). Insect cells (Sf21) were infected with the viral stock pFASTBAC-HTc-hBLM with a multiplicity of infection of 2. Infected insect cells were incubated at 30° C for 48 hours and lysed in 50 mM Tris-HCl pH 8.5, 5 mM 2-mercaptoethanol, 100 mM KCl, 1mM PMSF, 1% Nonidet P-40 at 4°C.

Insoluable material was removed by centrifugation at 10,000g for 30 minutes at

4° C. Insect cell lysates were mixed with K562 human cell nuclear extracts (NE), which contain MLH1. Equal volumes of insect cell and K562 NE were precleared with pansorbin A (Calbiochem). Either 10 µg of α-MLH1 (Pharmigen) or 10 µg of

α-BLM-1 (Santa Cruz) antibody was used to immunoprecipitate MLH1 or BLM respectively. Fifty µl of protein A/G agarose was added, incubated for 1 hour and

45

washed extensively with IP wash buffer (20 mM Tris-HCl pH 8.0, 0.1 M NaCl, 1 mM EDTA, 1% NP-40). The precipitated proteins were eluted into 1X SDS-

PAGE buffer and separated by 10% SDS-PAGE. Western blot analysis was performed to detect the presence of BLM, MLH1 or replication protein A (70kD subunit) in the immunoprecipitates.

In vivo immunoprecipitations

Five mg of K562 NE was precleared with 50 µl of protein A/G beads.

Fifty µg of BLM affinity-purified antibodies (Santa Cruz BLM-1 and -2) were added to the precleared K562 NE and rotated for 2 hours at 4° C. One hundred fifty µl of protein A/G agarose beads was added to the mixture and rotated for an additional 2 hours. The immunoprecipitates were washed with 1 ml of IP wash buffer four times. The precipitated proteins were eluted into 1X SDS-PAGE buffer and separated by 8% SDS-PAGE. Western blot analysis was performed to detect BLM and MLH1.

Mismatch repair assay

Preparation of cell free extracts and mismatched substrates and procedures for measuring mismatch repair were as described (Umar, Boyer et al.

1994). Heteroduplex substrates for repair studies contained the mismatch or unpaired base and a nick in the (-) strand at position –264 (3’ –nicked substrate) or +276 (5’ –nicked substrate), where position +1 is the first transcribed nucleotide of the lacZα gene. Repair reactions (25 µl) contained 30 mM 4-(2-

46

hydroxyethyl)-1-piperazine-ethanesulfonic acid (pH 7.8); 7 mM MgCl2; 200 µM each CTP, GTP,UTP; 4 mM ATP; 100 µM each dCTP, dATP, dGTP, dTTP; 40 mM creatine phosphate; 100 mg/ml creatine phosphokinase; 15 mM sodium phosphate pH 7.5; 1 fmol of substrate DNA; and 50 µg of extract proteins.

Reactions were incubated for 15 minutes at 37°C. Substrate DNA was recovered and introduced into E. coli NR9162 (mutS) via electroporation and plated to score plaques as described [20]. Repair efficiency is expressed in percent as 100 X (1 minus the ratio of the percentages of mixed bursts obtained from extract-treated and untreated samples).

supF20 DSB Repair Assay

SupF20 was a generous gift from Michael Seidman and Michael Lin

(National Institute on Aging, National Institutes of Health, Baltimore, MD).

SupF20 is a derivative of the pZ189 plasmid and carries an ampicillin resistance gene and a modified Escherichia coli supF suppressor tRNA gene that serves as a mutagenesis marker. An insertion, including a BssHII restriction endonuclease site, in the 5’ region of the tRNA sequence destroys supF function. Function can only be restored when a recombinational event across a duplicated microhomology patch results in a specific 11bp deletion. In this study, a DSB was generated in purified supF20 DNA at the BssHII restriction endonuclease site. Linear DNA was separated from circular plasmid by agarose gel electrophoresis and purified by organic extraction. To assess the DSB repair efficiency of nuclear extracts, 50 µg of protein from each extract were incubated

47

for one hour at 30°C with 1µg of linear supF20 in 50 µL of DSB repair buffer [65.5 mM Tris (ph 7.5), 10 mM MgSO4, 1 mM ATP, 91 nM EDTA and 9.1% glycerol].

The reaction was stopped by the addition of EDTA to a final concentration of 20 mM and by incubation in RNAse A (Life Technologies, Inc.) and proteinase K

(Roche). DNA was purified by organic extraction and 1-10ng were used to transform E. coli MBM7070, a strain carrying an amber mutation in the lacZα gene. This mutation can be overcome by functional supF resulting in β- galactosidase synthesis. Transformants were selected on LB plates containing

100 µg/mL ampicillin and 40 µg/mL X-gal. Transformation efficiency, reflecting

DSB repair efficiency, was expressed as the total number of colonies per ug of

DNA. Fidelity of repair was determined by calculating the frequency of supF20 mutants over the total number of transformed colonies. SupF20 mutants with restored supF function were identified as blue colonies on X-gal-containing plates.

pUC18 DSB Repair Assay

pUC18 carries an ampicillin resistance gene and the lacZα gene. A DSB was generated in purified pUC18 DNA at the EcoRI restriction endonuclease site, disrupting the lacZα gene. Linear DNA was separated from circular plasmid by electrophoresis and purified from agarose by organic extraction. For experiments removing the terminal phosphate at the DSB site, DNA was incubated with 1U alkaline calf intestinal phosphatase (Life Technologies, Inc.) for one hour at 37°C. To assess the DSB repair efficiency of the nuclear

48

extracts, 50 µg of protein from each extract were incubated for one hour at 30°C with 1 µg of linear pUC18 in a 50 µL reaction mixture of DSB repair buffer. The reaction was stopped and DNA purified as in the supF20 DSB repair assay above. DNA (1-10 ng) was used to transform E. coli DH5α. Transformants were selected on LB plates containing 100 µg/ml of ampicillin and 40 µg/mL of X-gal.

Transformation efficiency, reflecting DSB repair efficiency, was expressed as the total number of colonies per µg of DNA. Fidelity of repair was determined by calculating the frequency of lacZα mutants over the total number of transformed colonies. LacZα mutants were identified as white or light blue colonies on X-gal containing plates. Statistical significance was determined using the students’ t test. Mutant colonies were picked and restreaked onto fresh LB plates containing 100 µg/mL ampicillin and 40 µg/mL X-gal. DNA was purified from overnight cultures using a Qiagen miniprep kit and sequenced using an 18 nt primer (TGAGAGTGCACCATATGC) located at nucleotide 172 of the pUC18c plasmid. Sequencing was performed by the Univeristy DNA Core Laboratory at the College of Medicine (Cincinnati, OH) using an Applied Biosystem Inc. 377 automated sequencer and dye terminator sequencing chemistry.

yBLM expression and purification

The yeast construct pJK1 (to express hBLM in yeast) was a gift from Ian

Hickson (Karow, Chakraverty et al. 1997) and yeast strain JEL1 (MATα leu2 trp1 ura3-52 prb1-1122 pep4-3 dhis3::PGAL10-GAL4) was provided by Jim Wang.

(Austin, Marsh et al. 1995) Briefly, the human BLM cDNA was cloned into the

49

pYES2 (Invitrogen). The N-terminus of BLM was modified to add a yeast Kozak consensus sequence and the first five codons of the TOP2 open reading frame.

The C-terminus was modified to add a hexahistidine tag to aid in purification.

JEL1 yeast transformed with pJK1 were grown in –Ura yeast media containing 2% glucose at 30°C to an OD600 of 0.6. Yeast were centrifuged at

3000 x g for six minutes, resuspended in growth media –Ura plus 2% galactose of equal volume and grown for 24 hours at 30°C. The yeast cells were pelleted as before. All subsequent steps were performed at 4°C. The yeast pellet was resuspended in Buffer A: 50 mM KPO4 pH=7.0 500 mM KCl 10% glycerol.

Mammalian protease inhibitor cocktail was added at 1/500 (Sigma P-8340).

Typically, 35 mL of buffer A was used to resuspend a one liter culture of yeast.

To lyse the yeast cells a French press was employed. Yeast were loaded into cold French press and 20,000 psi was applied. The French press was opened and the lysate was collected in a dropwise fashion. The yeast lysate was centrifuged at 20,000 x g for 30 minutes to remove insoluable material. While the sample was spinning, NTA resin (Qiagen) was prepared. The column was charged with five bed volumes of 1X charge buffer (50 mM NiSO4). The column was equilibrated with 3 bed volumes of 1X binding buffer (Buffer A plus 50 mM imidizole). The supernatent was mixed with the equilibrated resin and allowed to bind for 2-5 hours. The sample was returned to the column and the flow through was collected. All fractions were saved including flow through for western analysis to determine if BLM protein was present. The column was washed with

1X binding buffer until baseline OD280 is achieved. BLM was eluted as a single

50

step with Buffer A plus 500 mM imidizole. One mL fractions of this single step elution were collected. The fractions containing BLM protein were dialyzed into

Buffer Z: 60 mM Tris-HCl pH=7.5, 100 mM KCl, 1mM EDTA, 1mM DTT and 10% glycerol. The protein sample was then concentrated on a YM30 spin column

(Millipore), aliquoted, snap frozen in liquid nitrogen and stored at –80°C.

Preparation of helicase substrates

Single-stranded DNA oligomers were 5'-end-labeled with 32P according to

Sambrook et al. To prepare the partial DNA duplex for helicase assays, the labeled oligomer was mixed with a 2-fold molar excess of a complementary

unlabeled DNA oligomer in 10 mM Tris-HCl buffer, pH 8.0, 10 mM MgCl2; 1mM

EDTA with a mineral oil overlay to prevent evaporation; following denaturation at

100 C for 15 min, the DNA was allowed to anneal by slow cooling to room temperature. The annealed substrate was resolved from the unannealed radioactive oligo by agarose gel electrophoresis (4% non-denaturing 1X TBE (90 mM Tris base, 90 mM boric acid, 1.0 mM EDTA)). After adequate separation, the annealed substrate was excised from the gel and extracted from the gel by the

‘freeze-squeeze’ method. The gel slice was sliced into small pieces with a razor blade and smashed up in between parafilm. The gel material is placed in a microfuge tube with 500 µl of phenol/chloroform and 300 µl of TE. This tube is vortexed and placed at –80°C until completely frozen. The tube was spun at

10,000 x g for 20 minutes at room temperature. The aqueous phase was

51

removed and extracted again with chloroform/isoamyl alcohol 49:1. The DNA was precipitated with sodium acetate/ethanol and resuspended in TE.

52

Table I- Synthetic substrates used in gel-shift and helicase assays

DO substrate -also referred to as double overhang substrate or WRN substrate has both 5’ and 3’ single stranded regions

20 mer 5'-d(CGCTAGCAATATTCTGCAGC)-3' 5’ 3’ 3’ * 5’ 46 mer 5'-d(GCGCGGAAGCTTGGCTGCAGAATATTGCTAGCGGGAATTCGGCGCG)-3'

Blunt end duplex- 50mer

5’-GACGCTGCCGAATTCTGGCTTGCTAGGACATCTTTGCCCACGTTGACCCG-3’

5’-CGGGTCAACGTGGGCAAAGATGTCCTAGCAAGCCAGAATTCGGCAGCGTC-3’

G G/T mismatch- 50mer T

5’-GACGCTGCCGAATTCTGGCTTGCTAGGACATCTTTGCCCACGTTGACCCG-3’

5’-CGGGTCAACGTGGGCAAAGATGTCTTAGCAAGCCAGAATTCGGCAGCGTC-3’

4 nt. IDL loop- 50mer

5’-GACGCTGCCGAATTCTGGCTTGCTAGGACATCTTTGCCCACGTTGACCCG-3’

5’-CGGGTCAACGTGGGCAAAGATGTCCGGGGTAGCAAGCCAGAATTCGGCAGCGTC-3’

12 nt. IDL loop- 50mer

5’-GACGCTGCCGAATTCTGGCTTGCTAGGACATCTTTGCCCACGTTGACCCG-3’

5’-CGGGTCAACGTGGGCAAAGATGTCCGAAAGGGCCAATAGCAAGCCAGAATTCGGCAGCGTC-3’ Y-Fork- 50mer

5’-GACGCTGCCGAATTCTGGCTTGCTAGGACATCTTTGCCCACGTTGACCCG-3’

5’-CGGGTCAACGTGGGCAAAGATGTCCTAGCAATGTTATCGTCGATGACGTC-3’

Synthetic X-junction- 50mer

5’-CGGGTCAACGTGGGCAAAGATGTCCTAGCAAGCCAGAATTCGGCAGCGTC-3’ 5’-GACGCTGCCGAATTCTGGCTTGCTAGGACATCTTTGCCCACGTTGACCCG-3’ 5’-GACGTCATAGACGATTACATTGCTAGGACATGCTGTCTAGAGACTATCGC-3’ 5’-GCGATAGTCTCTAGACAGCATGTCCTAGCAAGCCAGAATTCGGCAGCGTC-3’

53

Helicase assays

DNA helicase activity was detected by the displacement of a 32P-labeled

5'-20-mer oligonucleotide 5'-d(CGCTAGCAATATTCTGCAGC)-3' from its partial duplex with the complementary unlabeled 46-mer 5'- d(GCGCGGAAGCTTGGCTGCAGAATATTGCTAGCGGGAATTCGGCGCG)-3'.

This substrate has been named the DO substrate (see Table I)Reaction mixtures

contained in a final volume of 50 µl, 50 mM Tris-HCl buffer, pH 7.5, 5 mM MgCl2,

5 mM ATP, 100 µg/mL of BSA, 50 ng of labeled DNA substrate, and 20 ng of

BLM protein. Other proteins were added into the helicase assay as described either in the figure or text. HMLH1 was a gift from Tom Kunkel and MLH1/PMS2 was a gift from Michael Liskay. DNA/protein mixtures were incubated at 37°C for

30 min, and DNA unwinding was terminated by rapid cooling on ice and by the addition of 10 µl of a solution of 2.0% SDS, 50 mM EDTA, 3.0% bromphenol blue, 3.0% xylene cyanol, 40% glycerol. The displaced single-stranded oligonucleotide was separated from the partial DNA duplex substrate by electrophoresis through a non-denaturing 12% polyacrylamide gel in TBE buffer

(90 mM Tris base, 90 mM boric acid, 1.0 mM EDTA) at room temperature for 2 h ours under 100V (8V/cm). Labeled DNA bands were visualized by autoradiography.

54

Gel-shift Assays

The DNA-binding reactions (20 µl) contained 20 mM triethanolamine-HCl,

pH 7.5,2mM MgCl2, 1mM ATP S, 0.1 µg/ml BSA, and 1 mM dithiothreitol, and protein concentration as indicated in the figures. Reaction mixtures were incubated at room temperature for 20 min and fixed in the presence of 0.25% glutaraldehyde for 10 min at 37°C. The products were separated by electrophoresis through 4% nondenaturing polyacrylaminde gels at 4°C for 3 hours and visualized using a PhosphorImager or film autoradiography.

55

Chapter 4 – The BLM helicase interacts with MLH1

BLM identifies MLH1 in a yeast two-hybrid screen

In order to help elucidate the functional role of the BLM helicase in regards to DNA metabolism, a yeast two-hybrid screen was performed with different domains of the BLM helicase. The N-terminal and C-terminal domains were cloned into a GAL4-yeast two-hybrid vector and used to screen a human B- lymphocyte cDNA library (Fig. 5A). The N-terminus of BLM auto-activated, however the C-terminus of BLM identified five clones, two of which contained a full-length cDNA encoding the DNA mismatch repair protein MLH1 (Fig. 5B).

IVTT BLM-C and full-length MLH1 interact

IVTT products were mixed and immunoprecipitated to investigate the interaction between BLM and MLH1. MLH1 is present when mixed with the C- terminus of BLM and immunoprecipitated with an antibody specific for BLM

(Figure 6). The C-terminus of BLM is detectable when mixed with MLH1 and immunoprecipitated with an MLH1-specific antibody. The N-terminal and helicase domains of the BLM helicase were unable to co-precipitate with MLH1.

These data provide further evidence that BLM and MLH1 interact in vitro through the C-terminus of BLM.

56

Figure 5A

11417HELICASE DOMAIN NLS 1036 1417

SEGMENT OF BLM USED FOR YEAST TWO-HYBRID SCREEN Figure 5B GAL4-DBD- GAL4-DBD BLM-C

GAL4-AD-PROTHYMOSIN 1 GAL4-AD-PROTHYMOSIN 2 GAL4-AD-MLH1-1 GAL4-AD-MLH1-2

POSITIVE NEGATIVE CONTROL CONTROL

Figure 5. BLM identifies MLH1 in a yeast two-hybrid screen. A, schematic of BLM (aa 1-1417) showing its helicase domain and the aa segment(1036-1417) used in the yeast two-hybrid screen that identified MLH1. The location of a functional bipartite nuclear localization signal (NLS) aa (1336-1349) is also shown. B, yeast plate demonstrates the interaction bwtween theC-terminus of BLM and two clones containing MLH1. β-galactosidase activity Is shown from yeast co-transformed with pAS2 (top left quadrant) or pAS2-BLM-C (top right quadrant) and pGAL4-AD-PROTHYMOSIN or pGAL4-AD-MLH1 (two full-length prothymosin and two full-length MLH1 clones) from a human B-lymphocyte library. Positive (bottom left quadrant) and negative (bottom right quadrant) controls for β-galactosidase activity are shown.

57

1 M H L G L G C B M g - 1 - Ig - I α H g α m M L L H H H H B M T T T - - I I IT I T T W T T W W W V P IV I IP IP IP I

90 kD 45 kD

1 2 3 4 5 6 7 8

Figure 6. Immunoprecipitations of in vitro transcribed and translated (IVTT) protein products demonstrate the interaction between the C- terminus of BLM and MLH1. Labeled [35S]methionine IVTT protein mixes of BLM-C and MLH1 were immunoprecipitated with antibodies to BLM or MLH1. Lanes 1 and 2 represent input controls of 10 ul of IVTT-BLM-C or 10 ul of IVTT-MLH1, respectively. Lane 3 contains 50 ul of IVTT-BLM-C immunoprecipitated with α-MLH1. Lane 4 contains 50 ul of IVTT-MLH1 immunoprecipitated with α -BLM. Lane 5 contains a mixture of 50 ul of IVTT- MLH1 and 50 ul of IVTT-BLM-C immunoprecipitated with α-BLM. Lane 6 contains a mixture of 50 ul of IVTT-MLH1 and 50 ul of IVTT-BLM-C immunoprecipitated with goat IgG. Lane seven contains a mixture of 50 ul of IVTT-MLH1 and 50 ul of IVTT-BLM-C immunoprecipitated with α-MLH1. Lane eight contains a mixture of 50 ul of IVTT-MLH1 and 50 ul of IVTT-BLM-C immunoprecipitated with mouse IgG.

58

To further study this in vitro interaction, lysates from insect cells that expressed full-length BLM were mixed with lysates mixed from the human erythroleukemia cell line K562. Immunoprecipitation with an α-MLH1 antibody, but not nonspecific IgG, immunoprecipitated BLM (Figure 7A). Conversely, immunoprecipitation with an α-BLM antibody demonstrated that MLH1 is associated with BLM (Fig. 7B). As a positive control, BLM was immunoprecipitated with α-BLM and probed for the replication protein A (RPA)

70-kDa subunit, a known protein partner of the BLM helicase. RPA was detected in the α-BLM immunoprecipitation but not in the IgG control immunoprecitates

(Fig. 7C).

Far western assays confirm the interaction between MLH1 and BLM-C

Far western assays were performed to confirm the in vitro interaction of the C-terminus of BLM and MLH1. Protein induction studies in bacteria show that the C-terminus of BLM was expressed at high levels in E. coli but was only present in the insoluble fraction. The C-terminus of BLM was consequently isolated under denaturing conditions using nickel chromatography and then slowly dialyzed into a tris-buffered solution. Purified fractions were analyzed by

SDS-PAGE followed by Coomassie blue staining to identify several fractions that were more than 90% pure. Far western assays then confirmed that the C- terminus of BLM and MLH1 interact in vitro. Briefly, the C-terminus of BLM was bound to PVDF membrane and incubated in K562 nuclear extracts (25 ug) for 1 hr. The membrane was then was in TBS-T and the bound proteins were eluted

59

into SDS-PAGE reducing buffer. These eluted samples were separated by 10%

SDS-PAGE and transferred to PVDF membrane. Western blot analysis with an

α-MLH1 antibody detected a 90-kDa band (as well as some smaller degradation products) that co-migrates with both the nuclear extract control and IVTT MLH1

(Fig.8). Although high salt washes greatly diminished the interaction between the two proteins, membrane alone and membrane coated with a nonspecific protein

(cytochrome C) did not bind MLH1 demonstrating the specificity of the BLM-

MLH1 interaction. MLH1 was also expressed in vitro by IVTT and was capable of binding to the immobilized C-terminus of BLM (Fig. 8).

60

A , , M M 1 M L L H L B G B L B + Ig + M E E - + m α N N E 2 H 2 H N 6 IT 6 IT 2 5 5 6 W W 5 K K K IP IP α -BLM 205 kD

1 2 3 M G L Ig -B g α

M G L TH TH Ig B I I g - W W α P P I I ITH TH , , M I M M L W W L L B P B B I IP + , , + + E E E E E N N N N N 2 2 2 2 2 6 6 6 6 6 5 5 5 5 5 K ,K K K K B α-MLH1 90 kD

C α -RPA 70 kD

Figure 7. Mixed lysate immunoprecipitation demonstrates the interaction between full-length BLM and MLH1 or RPA. Lysates from insect cells infected with baculovirus expressing full-length BLM were mixed with K562 NE. Immunoprecipitated proteins were resolved on 10% SDS-PAGE and transferred to membrane. The first lane of each western blot contains K562 NE and insect cell lysates infected with BLM baculovirus as a control to show the size and specificity of α-BLM (Novus), α-MLH1 (PharMingen) and α-RPA (Santa Cruz) antibodies. In A, the second lane contains mixed lysates immunoprecipitated with an IgG control while the third lane contains mixed lysates immunoprecipitated with α-MLH1. In B, the second lane contains K562 NE immunoprecipitated with goat IgG, the third lane contains K562 NE immunoprecipitated with α-BLM (Santa Cruz), the fourth lane contains mixed lysates immunoprecipitated with goat IgG and the fifth lane contains mixed lysates immunoprecipitated with α-BLM (Santa Cruz). In B, the western blot was probed with α-MLH1. Figure 7C shows the same western blot as B stripped and probed with α-RPA.

61

H AS l W 1 aC LH N 1 -M M LH m M 00 1 - T 5 LH VT th M T +I L i - VT C O w T I E TR E E T + M N N N IV E O O + + + N R C C A H C - - -C BR C E TT M M M O N V L L LM E T I B B B M CY MLH1 90 kD

45 kD

1 2 3 4 5 6 7

Figure 8. Far western assays demonstrate the interaction between the BLM-C terminus and MLH1. Lane 1 contains 25 µg of K562 nuclear extract (NE) to show the size of MLH1. The sample in lane 2 is IVTT-MLH1, which includes 10 µl of the IVTT-MLH1 mixture used in these experiments. Lane 3 shows proteins bound to BLM-C protein following incubation in 25 µg of K562 NE. Lane 4 shows protein bound to BLM-C protein following incubation with 25 µg of K562 NE and a 500 mM NaCl wash. Lane 5 shows protein bound to BLM-C protein following incubation with 10 µl of IVTT-MLH1. Lane 6 shows protein bound to PVDF membrane blocked with 5% NFDM following incubation with 10 µl of IVTT-MLH1. Lane 7 shows protein bound to cytochrome C following incubation with 10 µl of IVTT-MLH1. Samples were resolved by 10% SDS-PAGE and visualized by autoradiography.

62

To test for in vivo interactions, K562 nuclear extracts were used for immunoprecipitation with α-BLM or α-MLH1 antibodies. MLH1 was co- immunoprecipitated with BLM from K562 nuclear extracts but was not present in the IgG control immunprecipitates (Figure 9). Endogenous BLM could not be detected when nuclear extracts were immunoprecipitated with antibodies that recognize MLH1. This may be due to the low expression levels of BLM, its cell cycle-specific regulation or the BLM-MLH1 interaction masks the MLH1 antibody recognition site. It should be noted that Wang et al. have immunoprecipitated

MLH1 and identified BLM, although the reverse experiment was not performed.

M G L , g I , -B E g E α N E H N H 2 T 2 N 6 I 6 IT 2 5 5 6 K W K W 5 P P K I I

-90 kD MLH1

1 2 3

Figure 9. BLM and MLH1 interact in vivo. K562 nuclear extracts were immunoprecipitated with α-BLM antibodies or a goat IgG control. Lane 1 contains lysates from K562 NE showing the size of MLH1. Lane 2 contains an immunoprecipitation from the same extracts using a goat IgG as a negative control. Lane 3 contains an immunoprecipitation from the same extracts using an α-BLM antibody.

63

DNA mismatch repair activities of BS and control cell extracts are equivalent.

Finally, the functional signifiance of the interaction between BLM and

MLH1 was tested by examining the ability of BS cell extracts to carry out DNA mismatch repair by measuring their ability to repair mispaired substrates in vitro.

M13mp2 DNA was used as a repair substrate and contained a covalently closed

(+) strand and a (-) strand with a nick (to direct repair to this strand) located several hundred base pairs away from the mispair in the lacZα complementation gene. The (+) strand encodes one plaque phenotype and the (-) stand encodes the other phenotype (either colorless or blue). If the unrepaired heteroduplex is introduced into an E. coli strain deficient in methyl-directed mismatch repair, plaques will have a mixed phenotype due to expression of both strands.

However, repair in a repair-proficient human cell extract will reduce the percentage of mixed plaques and increase the ratio of the (+) strand phenotype relative to that of the (-) strand phenotype as the nick directs repair to the (-) strand.

BS cell extracts repair a G-G mispair as efficiently as a HeLa cell extract, which is repair proficient (Figure 10). Repair is observed regardless of whether the nick is 3’ or 5’ to the mismatch, consistent with the bi-directional repair capability of the human mismatch repair system. The change in the ratio of the blue to colorless plaques indicates that repair is specific for the (-) strand as directed by the nick in that strand. The BS extract also repairs substates containing an A-C mismatch or either of two different unrepaired nucleotides. In contrast to these results, extracts of cell lines exhibiting microsatellite instability

64

and having mutations in any of four mismatch repair genes (MLH1, PMS2, MSH2 or MSH6) are uniformly deficient in strand-specific mismatch repair. These results suggest that BLM does not directly function in mismatch repair pathway but may play crucial roles in the processing of heteroduplex formations during replication, recombination or other types of DNA repair.

G:G (3' nick) G:G (5' nick)

r r 80

i 80 i

a

a

p 60 p 60

e

e

R

R

40 40

% 20 % 20 0 0 HeLa BS HeLa BS

G:G (3' nick) G:G (5' nick) 2.5 4 2.0 3 1.5 2 1.0 1 Blue/white Blue/white 0.5 0 0 Mock HeLa BS Mock HeLa BS

Figure 10. DNA mismatch repair activities of BS and HeLa cell extracts are equivalent. A, Repair efficiency in percent of a G-G mispair at postion 88 in the lacZα-complementation gene in extracts of BS or HeLa cells. B, The ratio of pure blue to pure colorless plaques from extracts of BS or HeLa cells. The results reflect counting more than 500 plaques per variable. In addition to the results shown, repair in an extract from BS cells was also observed for the following substrates, with the (+) strand listed first, the (-) strand listed next (where a dash indicates a missing nucleotide) and the position of the mismatch listed last: C*A at 87, 25% repair; *T at 91, 86% repair; C*at 132-136, 82% repair. These substrates all contained a 3’ nick at position –264.

65

Chapter 5- The BLM helicase is stimulated by mismatch repair proteins

One of the protein partners of the BLM helicase is the mismatch repair protein MLH1 (Langland, Kordich et. al 2001, Pedrazzi, et al 2001). These two proteins colocalize in VA13 fibroblasts (Figure 11). MLH1 is known to interact with the C-terminus of BLM, but the functional significance is not known. The mismatch repair system is intact in BS cells, at least at correcting single nucleotide mismatches (Langland, Kordich et. al 2001, Pedrazzi, et. al 2001).

However, the heterodimeric complexes of MutS- and MutL related proteins that have been identified from studies of DNA mismatch repair have also been implicated in the processing of recombination intermediates (Nakagawa, 1999 ).

Several yeast mismatch repair genes are required for maintaining normal levels of crossing over but not for gene conversion events. The genes that are responsible for regulating these crossover events include MLH1, MLH3, MSH4,

MSH5, the ZIP genes and EXO1. Specifically, most of these proteins are important in meiotic crossing over. Mutations in MLH1 and MSH5 in mice result in male and female infertility.

Mismatch repair proteins have also been implicated in the processing of 5’ tailed structures generated by single-stranded annealing recombination and directly recognize and bind to Holliday junctions in vitro. MSH2 and p53 associate with recombinative repair complexes in vivo during S-phase and BLM may be present as well (Zink, Mayr et al. 2002). Figure 12 nicely summarizes the different DNA substrates for the different mismatch repair proteins.

66

Mismatch repair in E.coli

In E. coli mismatch repair is the primary mechanism for repair of replication errors and also prevents recombination between highly divergent DNA sequences. Therefore a functional mismatch repair system in E. coli ensures the precision of chromosomal replication and maintains genomic stability. In human cells, functional loss of the mismatch repair system results in genomic instability and hereditary colon cancer, which underscores the importance of this DNA repair pathway.

The sequence of biochemical reactions that define the mismatch repair pathway in E. coli has been well described and the proteins responsible for each step are known. The MutS dimer recognizes and binds to the mismatch. The

MutL dimer binds to the MutS-DNA complex and the DNA is looped out in an active search for the nearest d(GATC) methylation site. Once this methylation site is found, the mismatch repair complex stimulates MutH to generate a nick on the unmethylated strand at the hemimethylated d(GATC) site. UvrD helicase and the appropriate exonuclease then excise the error-containing DNA. The gap is then filled by DNA polymerase III and DNA ligase seals the nick (Marti, Kunz et al

2002).

Of particular interest is the fact that the UvrD helicase unwinds duplex

DNA in a 3’ to 5’ direction as does the BLM helicase. UvrD interacts physically with MutL and the MutL protein stimulates the helicase activity of UvrD more than

10-fold on a conventional helicase substrate (Hall, Jorden et al. 1998).

67

Figure 11. Co-localisation of BLM and hMLH1 in the nucleus of WI-38/VA-13 cells. Indirect immunofluorescence of BLM (green) and hMLH1 (red) is shown in WI-38/VA-13 cells. The yellow colour results from overlap of the red and green foci. Nuclear DNA was revealed by staining with Hoechst 33258. Adapted from Pedrazzi, G. et al 2001.

68

2

2 2 2

Figure 12. The combinatorial specificites of MutS- and MutL-related heterocomplexes. (a) Base/base mispairs; (b) insertion/deletion loops; (c) 5’ tailed structures generatedby single-strand DNA annealing (SSA) recombination; (d) Holliday junctions

Adapted from Nakagawa et al. 1999

69

Since BLM and MLH1 interact directly, we hypothesized that this nteraction may regulate the substrate specificity or helicase activity similarly to how MutL stimulates UvrD helicase activity.

BLM purification and characterization

To test this hypothesis full-length human BLM protein must be expressed and purified. A yeast system was used similar to that of Ian Hickson with slight modifications described in the Materials and Methods of this thesis. Figure 13 shows the results of the purification and our BLM fractions were more than 90% pure by silver stain and Coomassie. Now that the protein is relatively pure it also has to have enzymatic activity (helicase activity). RecQ helicases possess strand displacement activity that is ATP and Mg2+ dependent. Figure 14 shows the results of the strand displacement assay. In this assay 5 ng of rBLM protein is added to 50 ng of purified substrate and incubated in 50 mM Tris-HCl, pH 7.5,

5 mM MgCl2, 5 mM ATP, 100 µg/ml bovine serum albumin, 50 mM NaCl at 37 C for thirty minutes. The reaction was terminated by the addition of SDS 2% and

EDTA 10mM final concentration.

70

MW markers Load 200 kD Flow through Flow through trailing 127 kD 150 mM wash 450 mM elute 81 kD 450 mM elute trailing Additional 15 mL elute

37 kD HeLa nuclear extract

200 kD BLM protein

127 kD

81 kD

37 kD

Figure 13. Silver stain and western blot of BLM fractions purified by nickel chelation affinity chromatography. Samples were resolved by 8% SDS- PAGE. Western blot was probed with α-BLM (Novus Biologicals).

71

1 2 3 4 5 6 7

* * 3’ 5’ +

3’ 5’ Lane 1 small oligo Lane 2 large oligo Lane 3 annealed DO substrate Lane 4 BLM - ATP Lane 5 BLM + ATP Lane 6 BLM – Mg2+ Lane 7 boiled DO substrate

Figurer 14. The recombinant BLM protein has helicase activity that is ATP and Mg2+ dependent. Lane 1 is P32 labeled small oligo (20mer). Lane 2 is p32 labelled large oligo. Lane 3 is annealed substrate alone. Lane 4 is 5ng of BLM incubated with 50ng of labelled substrate but no ATP is present. Lane 5 is 5ng of BLM incubated with 50ng of labelled substrate but ATP is present. Lane 6 is 5ng of BLM incubated with 50ng of labelled substrate ATP is present but no divalent cations. Lane 7 is 50 ng of boiled DNA substrate. The samples were resolved on 12% non-denaturing 1XTBE gels and visuallized using a phosphoimager.

72

Gel-shift experiments

To test the hypothesis that MLH1 regulates the substrate specificity of the

BLM helicase, we performed gel shift experiments with different types of substrates. The substrates used for the gel-shift experiments were as follows:

(1) double-stranded DNA duplex, (2) single nucleotide mismatch, (3) 4 nt. IDL,

(4) 12 nt. IDL, (5) forked substrate and (6) a synthetic . Lane one shows where the free DNA (0.1 fmol) probe migrates on the non-denaturing acrylamide gel. Lane two is the DNA probe plus BLM protein (5 nM). Lane three is the DNA probe plus empty vector control and lane four is the DNA probe (0.1 fmol), BLM protein (5 nM) and MLH1 (50 nM). The results from Figure 15 show that the BLM helicase has a relatively low affinity for the blunt-ended double- stranded duplex and the addition of MLH1 did not affect BLM’s binding affinity for these substrates. The BLM protein had a higher affinity for 4 nt. IDL, 12 nt. IDL, the forked substrate and the X-junction (synthetic Holliday junction). However,

MLH1 again had no effect on BLM’s affinity for these substrates. Since, MLH1 had no observable effect on BLM’s substrate specificity, we decided to determine if MLH1 could regulate the helicase activity of BLM since mutL is known to stimulate UvrD and many of BLM’s protein-partners are known to modulate its enzymatic (unwinding) activity.

73

G T 1 2 3 4 1 2 3 4

Free probe

4nt. IDL 12 nt. IDL 1 2 3 4 1 2 3 4 Protein/DNA complex

Free probe

1 2 3 4 1 2 3 4

1. Free probe 2. Probe + BLM 3. Probe + EV 4. Probe + BLM and MLH1

Figure 15. The BLM helicase has a higher affinity for 4 nt. and 12 nt. insertion/deletion loops, stalled replication forks and synthetic Holliday junctions. The DNA-binding reactions (20 µl) contained 20 mM triethanolamine-HCl,

pH 7.5, 2 mM MgCl2, 1 mM ATP S, 0.1 µg/ml BSA, and 1 mM dithiothreitol, and protein concentration as indicated in the figures. Reaction mixtures were incubated at room temperature for 20 min and fixed in the presence of 0.25% glutaraldehyde for 10 min at 37 °C. The products were separated by electrophoresis through 4% nondenaturing polyacrylaminde gels at 4 °C for 3 h, and visualized using a PhosphorImager or film autoradiography.

74

Helicase assays

In order to test the hypothesis that MLH1 modulates the helicase activity of BLM, helicase assays were performed with and without the presence of MLH1.

50 uL reaction mixtures containing 50 mM Tris-HCl pH=7.5, 5mM MgCl2, 5 mM

ATP, 100ug/ml BSA, 50 mM NaCl, 1 fmol DNA substrate, 6 nM BLM and increasing amounts of MLH1 (a generous gift from Tom Kunkel Laboratory of

Molecular Genetics and Laboratory of Structural Biology, NIEHS, National

Institutes of Health, Research Triangle Park, North Carolina 27709, USA). As shown in Figure 16 MLH1 stimulates the helicase activity of BLM on the double overhang (DO) substrate when compared to the buffer control. Stimulation of helicase activity was observed when the concentrations of MLH1 were between

60 and 240 nM of MLH1 suggesting that relatively high amounts of MLH1 are needed to stimulate the helicase activity of BLM. This may be reflective to the relative levels of MLH1 and BLM in vivo, since MLH1 is a much more abundant molecule than BLM in the cell. Since, MLH1 heterodimerizes with PMS2 in mammalian cells a much more biologically significant experiment would be to use the MLH1/PMS2 heterodimer which was a gift from Guy Tomer and Michael

Liskay of the Department of Molecular and Medical Genetics, Oregon Health and

Science University, Portland, Oregon 97201, USA. Like MLH1 alone,

MLH1/PMS2 was able to stimulate the helicase activity of BLM on the DO substrate (Figure 17). Figure 18 shows the stimulation of unwinding by the

MLH1/PMS2 heterodimer compared to buffer alone by graphing the data using the Imagquant/ Molecular Dynamics to quantitate the % unwinding and Microsoft

75

Excel to plot the data. There is a four-fold stimulation with the addition of MLH1-

PMS2 heterodimer when compared to buffer alone. Figure 19 are control experiments to show that the MLH1 or MLH1/PMS2 heterodimer used in our experiments do not possess any contaminating factors or intrinsic strand- displacement activites. BLM was able to unwind the DO substrate but no detectable strand-displacement activity was observed in either MLH1 or

MLH1/PMS2.

This makes a very interesting observation that like mutL and the UvrD helicase, the BLM helicase is stimulated by both MLH1 and MLH1/PMS2 heterodimer. Since BLM is an enzyme it is not surprising that many of its protein-partners will regulate its enzymatic function in some manner. RPA has been shown to facilitate BLM and WRN helicase processivity on long DNA substrates and it has also been demonstrated that p53 inhibits BLM helicase activity in vitro.

76

BLM - + + + + + + - 6 nM

MLH1 ------nM

Buffer control 3’ * 5’ *

1 2 3 4 5 6 7 8

BLM - + + + + + + - 6 nM MLH1 - - 15 30 60 120 240 - nM

Increasing 3’ * 5’ amounts of * MLH1 1 2 3 4 5 6 7 8

Figure 16. Effect of MLH1 on BLM helicase activity. The helicase assay used 6 nM BLM on the double overhang substrate (1 fmol). The first lane is substrate alone. The second lane is BLM without ATP. The third through seventh lanes either have a buffer control or increasing amounts of MLH1 (15-240 nM). Lane eight is a heat-denatured control.

77

BLM - + + + + + + + - 6 nM MLH1/PMS2 ------

3’ * 5’ Buffer control *

1 2 3 4 5 6 7 8 9

BLM - + + + + + + + - 6 nM

MLH1/PMS2 - - 15 30 60 120240 580 - nM

3’ * 5’ Increasing * amounts of MLH1/PMS2

1 2 3 4 5 6 7 8 9

Figure 17. Effect of MLH1/PMS2 heterodimer on BLM helicase activity. The helicase assay used 6 nM BLM on the duplex substrate (1 fmol). The first lane is substrate alone. The second lane is BLM without ATP. The third through eighth lanes either have a buffer control or increasing amounts of MLH1/PMS2 (15-240 nM). Lane eight is a heat-denatured control.

78

Figure 18. The MLH1/PMS2 heterodimer stimulates BLM helicase activity in vitro.

110

100

90

80

70 % Unwinding 60

50 Buffer alone 40

30

20 0 40 80 120 160 200 60 nM 120 nM 240 nM 580 nM Increasing amounts of MLH1/PMS2

79

3’ * 5’

*

1 2 3 4 5 6

Figure 19. Effect of MLH1or MLH1/PMS2 heterodimer on strand displacement activity. The first lane is substrate alone (pmol). The second lane is BLM (6 nM) without ATP. The third lane is BLM helicase plus ATP. The fourth lane is MLH1(240 nM) from Kunkel. The fifth lane is MLH1/PMS2 (240 nM) from Liskay. Lane six is a heat denatured control.

80

Chapter Six – The BLM Helicase is necessary for normal double-strand break repair

Two major pathways are responsible for the repair of DSBs in mammalian cells, nonhomologous end joining and DSB repair via homologous recombination. (Kanaar, Hoeijmakers et al. 1998; Khanna and Jackson 2001)

Runger and Kraemer were the first to suggest that there may be an alteration in the ability of BS cells to repair DSBs in DNA, because the ability of BS cells in culture to rejoin linearized plasmid was reduced and more error-prone than control cells. (Runger and Kraemer 1989) Interestingly, current reports suggest the BLM helicase may play functional roles in both types of DSB repair.

Mutations in Dmblm, which lead to partial male sterility, can be partially complemented by overexpression of -70, a protein well known for its role in nonhomologous end joining. (Kusano, Johnson-Schlitz et al. 2001) Aditionally,

WRN, the helicase deficient in Werner’s syndrome, directly interacts with Ku-70 to regulate its intrinsic exonuclease activity. (Yannone, Roy et al. 2001) These findings suggest mammalian RecQ helicases have functional roles in nonhomologous end joining. BLM is a member of the BRCA 1-associated surveillance complex and may play a role in DSB repair by homologous recombination because it interacts directly with Rad51, a protein known to function in homologous recombination. (Wang, Cortez et al. 2000; Wu, Davies et al. 2001) A greater number of DSBs occur in chicken DT40 cells with homozygous mutations in BLM and Rad54, compared to control cells, whereas in

81

response to agents that induce DSBs, BLM colocalizes with the Mre11, p95 and

Rad50 complex. (Wang, Cortez et al. 2000)

In vitro end-joining assay using the supF20 vector

In a preliminary experiment, the in vitro DSB repair efficiency of BS nuclear extracts was assessed by examining the transformation capacity of

BssHII-digested supF20 vector (see Figure 20A). Restoration of supF function in the supF20 plasmid requires a specific 11 base-pair deletion event.

Transformation efficiency of linear supF20 was the same order of magnitude for plasmids incubated in BS, corrected BS and control nuclear extracts (Table 1), suggesting that the plasmid was re-circularized by nuclear extracts in vitro and that BS nuclear extracts were indeed capable of repairing DSBs. However, no restoration of supF function was observed during DSB repair using BS nuclear extracts compared to 0.256% using control extracts and 0.176% using corrected cell extracts (Table 1), strongly suggesting that BLM may be required for the alignment of the micro-homology elements and subsequent 11 base-pair deletion event.

In vitro end-joining assay using the pUC18 vector

Therefore, we decided to examine the DSB repair ability and fidelity of BS nuclear extracts using a pUC18/lacZ reporter plasmid (Figure 20B). Similar to the supF20 repair assay, DSB repair efficiency using pUC18/lacZ is determined by restoration of the transformation capacity of EcoRI-digested plasmid. As controls for the effectiveness of the assay, nuclear extracts from a normal, a

82

corrected BS and an ataxia telangiectasia (AT) cell line were used. A three-fold higher mutation rate than normal has recently been demonstrated for AT extracts incubated with pUC18 lacking the 5' phosphate group at the EcoRI-DSB site. (Li,

Carty et al. 2001) Successful transformation of bacteria followed incubation of linear pUC18 in nuclear extracts of control, corrected BS, BS and AT cells

6 occurred at a rate of approximately 10 colonies/µg DNA for each extract.

83

A. BssHII-digested supF20

ampr

supF20 pBRori

supF MCS

7bp insertion GGGT GGGT

BssHII site

B. EcoRI-digested pUC18

ampr

ori pUC18

AGCTCG 3’ 5’* AATTCGAGC TCGAGCTTAA* 5’ 3’ GCTCG β-galactosidase

* -represents terminal phosphates which can be removed by treatment of the plasmid with calf intestinal phosphatase

Figure 20. Plasmid substrates used in the DSB repair assays. (A) supf20 reverts to an active SupF gene through the deletion of 11bp of DNA between the microhomology elements GGGT. supf20 was linearized with BssHII which cuts between the two regions of microhomology. Precise deletion of the 11bp at this site with subsequent ligation restores SupF activity. These events are scored as blue colonies in the indicator bacterial strain on X-gal-containing plates. (B) pUC18 was digested with EcoRI and treated with calf intestinal phosphatase for experiments where terminal phosphates are removed. Error-free end-joining results in functional β-galactosidase and can be scored as blue colonies in the bacterial indicator strain on X-gal-containing plates. White colonies are scored and can be recovered as mutants.

84

Table 2. The efficiency and fidelity of DNA DSB repair of linearized supF20 by nuclear extracts from control and BS cells.

Exp. No. Nuclear ATP Blue/ Transformation Relative Mutation extract total no. efficiency transformation frequencyb colonies (colonies/µg DNA) efficiencya (%)

1 Control - 0/5211 1.39 x 107 0

+ 12/4730 3.78 x 107 2.72 0.254

2 Control - 0/9434 1.57 x 107 0

+ 27/10431 3.48 x 107 2.23 0.259

1 BS - 0/3310 6.62 x 106 0

+ 0/4831 3.10 x 107 4.70 0

2 BS - 0/10219 8.50 x 106 0

+ 0/8348 2.78 x 107 3.27 0

1 Corrected - 0/1456 1.50 x 107 0 BS line + 6/3440 3.40 x 107 2.26 0.174

2 Corrected - 0/1096 1.10 x 107 0 BS line + 8/4432 4.40 x 107 4.00 0.180

Combined Control + 39/15,161 3.63 x 107 2.48 0.256 average

Combined BS + 0/13,179 2.94 x 107 2.88 0 average Combined Corrected + 14/7872 3.90 x 107 3.00 0.177 average BS line

a Relative transformation efficiencies were determined in relation to parallel controls in which ATP was omitted from the nuclear extract mix. b Mutation frequency was calculated as the number of blue colonies over the total number of colonies x 100%.

C Combined average represents the total number of colonies from experiments 1 and 2. The transformation efficiency, relative transformation efficiency and mutation frequency are the average of the two experiments.

85

Thus, BS nuclear extracts were capable of repairing DSBs that retained ligatable ends at an efficiency similar to control, corrected BS and AT extracts, confirming the observations using the supF20 plasmid.

The DNA termini generated by restriction endonuclease digestion differ from those generated by oxidative damage, in that the DNA ends of the former are directly ligatable. To examine the effectiveness of DSB repair when no ligatable ends are present, terminal phosphates were removed from the digested pUC18 substrate by incubation in calf intestinal phosphatase prior to incubation in nuclear extracts. Without the terminal phosphate at the DSB site, transformation efficiency dropped by approximately an order of magnitude for all three extracts. No difference in the relative transformation efficiencies was observed when pUC18 with (Fig. 21A) or without terminal phosphates (Fig. 21B) was used as a substrate. Relative transformation efficiencies were determined in relation to controls in which ATP was omitted from the nuclear extract mix. Thus,

DSB repair efficiency of the four nuclear extracts were similar.

The frequency of lacZα mutation was then examined to determine whether the contribution of error-prone repair pathways was similar for extracts from each cell type. The relative frequencies of lacZα mutants were significantly higher

(p<0.05) in experiments where linear pUC18 without the terminal phosphate group was incubated in BS or AT nuclear extract than the frequency of lacZα mutants from incubations in control or corrected BS nuclear extract (Fig.21D).

There were no significant differences in mutation frequencies when the DNA

DSB substrate retained the terminal phosphate (Fig. 21C).

86

C Relative Transformation A Relative Mutation Efficiency Figure 21 Figure 4.0 1.5 2.5 3.5 Frequency 0.5 0 1 2 3 1 2 0 3 4 6 5 Normal Normal Corrected Corrected BS BS BS BS AT AT B D Relative Mutation Relative Transformation Frequency Efficiency 10 16 18 10 12 14 2 0 4 6 8 0 8 2 4 6 Normal omlB AT BS Normal Corrected Corrected BS BS BS AT

87

Figure 21. The efficiency and fidelity of double-strand DNA break repair using normal, corrected BS, BS and AT nuclear extracts. A DSB was generated in pUC18 DNA at the EcoRI-site, disrupting the lacZ gene. Linear

DNA was untreated, thus retaining terminal phosphate groups (A, C), or treated with calf intestinal phosphatase to remove them (B, D). 1µg of linear pUC18 was incubated with nuclear extracts of control, corrected BS, BS and AT cells in a

DSB repair assay, then used to transform bacteria (A and B). Transformation efficiency, reflecting DSB repair efficiency, was determined by calculating the total number of colonies per µg DNA. The fold increase in transformation efficiency in relation to controls in which ATP was omitted from the DSB repair assay is plotted (C and D). Mutation frequency, reflecting the fidelity of DSB repair, was determined by calculating the frequency of lacZ mutants over the total number of transformed colonies. The fold increase in mutation frequency in relation to parallel controls in which ATP was omitted from the assay is plotted.

88

Thus, repair of DSBs that did not have a terminal phosphate group was more error-prone in BS extracts than in extracts from normal or corrected BS cells.

Sequence analysis of pUC18 mutants

Sequence analysis of some of the pUC18 mutants recovered in this study is shown in Figure 22. The majority of mutations generated by both the normal, corrected BS, BS and AT nuclear extracts were deletions, most of which spanned the EcoRI-induced DSB site. The majority of repair catalyzed in normal extracts and all the repair catalyzed by AT extracts utilize small, single or double nucleotide regions of homology at the ends of the deletion, a mechanism that is observed using similar assays. (Thacker, Chalk et al. 1992) Conversely, such short repeat-mediated repair was rarely observed in the mutants generated by the BS extracts. A six base-pair, CGAATT, deletion was frequently observed in

BS extracts, occurring in 14 out of 29 mutants generated by BS extracts repairing pUC18 without terminal phosphates. In contrast, this deletion was not observed in mutants generated by normal or corrected BS extracts. This six base-pair deletion did not arise via a micro-homology-mediated repair mechanism.

Additionally, the distribution of deletion size changes with or without BLM in the extracts. Deletions from the control or corrected BS extracts show a wide range of sizes ranging from 1 to 250bp in length. Deletions from BS extracts were generally smaller than the other cell types (averaging 6 to 10bps); AT deletions were on average 25bp in length. These data suggest that the BLM helicase may

89

be responsible for unwinding the termini at DSBs to facilitate the access of the break to other proteins involved in DSB repair or to search for microhomologies.

90

Figure 22.

TAGAGGATCCCCGGGTACCGAGCTCGAATTCGTAATCATGGTCATAGCTGTTTCCTGTGTGAAATTGTTATCCGCT

NORMAL: CCC______CCC (1, 241bp) C______C (1, 92bp) CCGG______CCGG (1, 89bp) CG______CG (1, 76bp) CCG______CCG (1, 61bp) GC______GC (1, 58bp) CT______CT (1, 48bp) C______C (1, 37bp) AgaggatccccgggtaccgagctcgaA (1, 30bp) AGCTcgaattcgtaatcatggtcatAGCT (7, 25bp) CGggtaccgagctcgaattCG (1, 19bp) CGggtaccgagctCG (1, 12bp) TCgtaatcatggTC (2, 12bp)

CORRECTED BS: A______A (1, 262bp) C______T (1, 240bp) G______C (1, 108bp) C______T (1, 103bp) A______G (1, 75bp) C______T (1, 70bp) GaggatccccgggtaccgagctcgaattcgT (1, 29bp) AccgagctcgaattcgtaA (2, 17bp) GgtaccgagctcgaatccC (1, 17bp) GgtaccgagctcgaA (2, 13bp) GgtaccgagctcgA (1, 12bp) TcgaattcgT (2, 8bp) TcgtaatcA (3, 7bp) CgA (3, 1bp) BS: G______A (1, 229bp) G______G (1, 191bp) A______T (1, 48bp) CgaattcgtaaT (1, 10bp) TtcgtaatC (5, 7bp) TcgaattC (14, 6bp) GattcgT (2, 5bp) CtcgaaT (1, 5bp) CgA (3,1bp) AT:

CC______CC (1, 37bp) G______G (1, 30bp) TCgaattcgtaatcatggtcatagctgttTC (1, 29bp) GggtaccgagctcgaattcgtaatcatG (1, 29bp) AGCTcgaattcgtaatcatggtcatAGCT (19, 25bp) CgagctcgaattcgtaatC (1,18bp) GTAccgagctcgaattcGTA (1,17bp) GaattcgtaatcaatG (1,14bp) TCgtaatcatggTC (1,12bp)

91

Figure 22. Deletion mutations in plasmids recovered from the in vitro DSB repair assay using pUC18 EcoRI-CIP as the substrate. Mutants generated from the incubation of EcoRI-cut CIP pUC18 in nuclear extracts are shown. Part of the original pUC18 sequence flanking the EcoRI-cut site is displayed at the top of the figure with the EcoRI site italized for comparison. The lower case letters represent the nucleotides deleted. Nucleotides are not shown for deletions larger than 30 bp. The letters in bold on the sides of the deletions represent microhomologies. Numbers in parentheses show the number of times each mutation was observed and the deletion size, respectively. These data were collected from two independent pUC18 DSB repair assays.

92

In this study, we have demonstrated no reduction in the efficiency with which BS nuclear extracts repair a restriction enzyme-induced DSB in two different plasmid substrates. This result is in contrast to the decrease in the efficiency of DSB repair of BS cells reported by Runger and Kraemer. (Runger and Kraemer 1989) In their experiments, linear DNA was transfected into BS fibroblasts, which were then cultured for two to three days before DNA recovery and bacterial transformation. It is possible that the efficiency of plasmid replication rather than repair was being measured. As BS cells exhibit a tremendously increased rate of sister chromatid exchange and an altered progression of the replication fork (Ellis, Lennon et al. 1995; German 1995), a defect in DNA replication may explain these experimental differences.

Although the DSB repair efficiency of BS extracts in our experiments was comparable to the efficiency of normal extracts, a significant decrease in the fidelity of repair was observed. This finding suggests that the absence of the BLM protein, a 3' to 5' helicase, may impede the unwinding of the duplex DNA near the DSB. Indeed, the predominance of short deletions in the mutants obtained following incubation in BS extracts, in contrast to the longer deletions obtained following incubation in normal and corrected BS nuclear extracts, supports this hypothesis.

Examination of the efficiency and fidelity of end-joining in HCT116 cell extracts. \ Since the BLM helicase interacts with the mismatch repair protein MLH1, we hypothesized that this interaction may be necessary for BLM to process the ends of the plasmid DNA. If this is indeed the case cells deficient in MLH1

93

should have a higher relative mutation frequency like BS cells and have deletions that are typically smaller than control cells. To test this hypothesis we used the

HCT116 cell line since it expresses no detectable MLH1. For these set of experiments pUC18 was digested with EcoRV resulting in a blunt end and EcoRI leaving a sticky end. Both of these substrates were treated with CIP to remove 5’ terminal phosphates, since a difference in mutation frequency was observed in both the BS and AT NEs. We first examined the relative transformation efficiency from the different cell lines. The end-joining efficiency or relative transformation efficiency was equivalent in all the cell lines tested including the

HCT116 cell line (Figure 23) with both substrates tested. Cells deficient in the mismatch repair protein MLH1 are competant are religating the plasmid ends compared to our control and patient derived cell lines.

94

4.0

3.5

3 2.5

2 1.5 Efficiency 1 0.5 Relative Transformation 0 Control Corrected BS AT HCT116 BS

pUC18-EcoRV-CIP

10

8

6

4 Efficiency

2 Relative Transformation 0 Control Corrected BS AT HCT116 BS

pUC18-EcoRI-CIP

Figure 23. The efficiency of double-strand DNA break repair using normal, corrected BS, BS, AT and HCT116 nuclear extracts. A DSB was generated in pUC18 DNA at either the EcoRVor EcoRI-site, disrupting the lacZ gene. Linear DNA was treated with calf intestinal phosphatase to remove the 5’ terminal phosphate. Linear pUC18 (1 µg) was incubated with nuclear extracts of control, corrected BS, BS, AT and HCT116 cells in a DSB repair assay, then used to transform bacteria. Transformation efficiency, reflecting DSB repair efficiency, was determined by calculating the total number of colonies per µg DNA. The fold increase in transformation efficiency in relation to controls in which ATP was omitted from the DSB repair assay is plotted.

95

6

5

4

3

2 Frequency

1 Relative Mutation

0 Normal Corrected BS AT HCT116 BS

pUC18-EcoRV-CIP 18 16 14 12 10 8

Frequency 6

Relative Mutation 4 2 0 Normal Corrected BS AT HCT116 BS

pUC18-EcoRI-CIP

Figure 24. The fidelity of double-strand DNA break repair using normal, corrected BS, BS, AT and HCT116 nuclear extracts. A DSB was generated in pUC18 DNA at either the EcoRVor EcoRI-site, disrupting the lacZ gene. Linear DNA was treated with calf intestinal phosphatase to remove the 5’ terminal phosphate. Linear pUC18 (1 µg) was incubated with nuclear extracts of control, corrected BS, BS, AT and HCT116 cells in a DSB repair assay, then used to transform bacteria. Mutation frequency, reflecting the fidelity of DSB repair, was determined by calculating the frequency of lacZα mutants over the total number of transformed colonies. The fold increase in mutation frequency in relation to parallel controls in which ATP was omitted from the assay is plotted; bars, ±SD.

96

Next we examined the frequency of lacZα mutation between the different cell-types to see if the lack of MLH1 in the HCT116 cell line results in a similar mutation frequency as cells deficient in BLM. If the mutation frequencies were similar, this would suggest that the two proteins functioned in the same pathway in regards to nonhomologous end-joining. However, this was not the case. Only the BS and AT cell lines had higher levels of relative mutation frequency when compared to control lines (Figure 24). We sequenced the junctions of twenty mutants derived from the HCT116 cell line and when compared to the BS and corrected BS cell line the deletions were much larger than those from the BS cell line (Figure 25). These data suggest that MLH1 and BLM do not interact in a functional manner in regards to nonhomologous end-joining.

Chapter Seven- Conclusions and future directions

Bloom’s syndrome is a rare autosomal recessive disorder that results in small size, sun sensitivity, immunodeficiency, genomic instability and a gross predisposition to cancer. The gene when mutated that is responsible for Bloom’s syndrome has been positionally cloned and is a RecQ helicase family member.

At least three other human RecQ family members have been identified and when mutated predispose those individuals to cancer formation. Mutations in the

Werner’s helicase lead to premature aging and cancer formation is a RecQ helicase family member with very similar biochemical properties to the BLM helicase. However, the patient phenotype between an individual with Bloom’s and an individual affected by Werner’s syndrome is very different. The difference

97

in patient phenotype is probably due to differences in patterns of expression and differences in protein partners.

The Werner’s helicase is expressed at low levels in most tissue types and

BLM is expressed at low levels in most tissue types with the exception of the thymus, spleen and testes. It is interesting to note that there is DNA recombinational events occurring in certain cells of these tissues and secondly these are highly proliferative tissues, which gives two reasonable explanations why these tissues have relatively high levels of BLM. BLM and WRN are also expressed differently at different phases of the cell cycle. The Werner’s protein is expressed at low levels throughout all phases of the cell-cycle, while the BLM protein rises four-fold as cells enter S-phase and is rapidly phosphorylated and degraded as the cell enters mitosis (Dutertre, Sekhri et al. 2002). The observation that the BLM helicase is expressed in cell-cycle dependent fashion provides different avenues of investigation to explain these initial observations.

The tissue-specific expression of BLM suggests there may be tissue-specific transcription factors responsible for the higher levels of expression in the thymus, testes and spleen. The promoter and gene sequence of BLM in human has been cloned but remains virtually uncharacterized in the literature to date. The second observation that BLM is phosphorylated in M-phase and is rapidly degraded also leaves one to wonder what specific kinase is responsible for this phosphorylation event and its subsequent degradation. BLM is known to be phosphorylated by the ATM kinase in response to DNA damage (Beamish, Kedar et al. 2002) but

98

the mitosis-specific phosphorylation has been shown to be independent of the PI-

3 related kinases.

It is interesting to note that the BLM protein and the WRN protein have overlapping protein-partners and also have unique protein-partners in the cell.

Biochemically, both the BLM and WRN helicases bind and unwind similar substrates (Mohaghegh, Karow et al 2001); therefore to understand the exact pathways of DNA metabolism it is will be vital to determine the unique protein- partners for each of these helicases. BLM, WRN and RecQ1 all interact with hRPA and this interaction stimulates their helicase activity and makes them more processive on substrates longer than 300 bps in length. The interaction between

RecQ helicases and RPA is conserved through its orthologs (Sgs1 interacts with yeast RPA) and is specific since mammalian RecQ helicases are not stimulated by E. coli SSBs.

BLM and WRN also both interact physically and are regulated by p53. Curt Harris’ group has shown that both BLM and WRN proteins interact with p53 both in vitro and in vivo (Yang, et al. 2002). The in vivo interaction of these two RecQ helicases could only be observed when cells were exposed to ionizing radiation, strongly suggesting a post-translation modification is necessary for this interaction to occur. p53 attenuated the unwinding ability of both the BLM and WRN helicases of synthetic X-junctions in vitro. The C- terminal p53 peptide (aa 373-383) inhibited BLM and WRN helicase activities, interestingly phosphorylation at Ser376 or Ser378 of p53 blocks this inhibition. p53,

Rad51 and BLM colocalized to nuclear foci when cells were treated with HU,

99

suggesting that all of these proteins are necessary for the restarting of stalled replication forks and the recombinational repair of DSBs.

Using a yeast-two hybrid system we screened a B-cell cDNA library to search for novel protein-partners of the BLM helicase. The C-terminus of BLM identified two full-length clones of the mismatch repair protein MLH1. Far western assays and in vitro transcription-translation immunoprecipitations were performed to confirm the in vitro interaction between the C-terminus of BLM and full-length MLH1. In vivo imunoprecipitations confirmed the interaction between these two proteins and another group reported the interaction between BLM and

MLH1 and also showed that the two proteins colocalized in the nucleus. MLH1is thought to act as a scafolding protein to bring various components of the eukaryotic mismatch repair together. The obvious question to ask was the BLM helicase necessary for a functional mismatch repair system ? Utilizing an in vitro mismatch repair assay, nuclear extracts were made from control cells and BS cells to test the ability of these extracts to remove and correct a single nucleotide mismatch (G/G). These experiments were performed in collaboration with Tom

Kunkel’s laboratory and BS nuclear extracts were able to repair the single nucleotide mismatch just as efficiently as the control cells. Pedrazzi et al. 2001 also tested BS cell lines for their ability to correct a single G/T mismatch and showed very similar results as our group. However, this is only one type substrate that the mismatch repair system recognizes and corrects. We our currently testing via Tom Kunkel’s laboratory small 3 and 4 nt. IDLs to see if BS extracts are capable at correcting these types of aberrant substrates. BS cells

100

have a mild degree of microsatellite instability, from secondary clones created by

Joel Straughen (manuscript in preparation) Figure 26. This microsatellite instability may be the result of certain types of substrates (ie small IDLs) escaping the mismatch repair system due to the lack of BLM to process them, their presence in Holliday junctions which is where mismatches are known to evade the mismatch repair system or they could be the result of unequal SCEs.

101

15 16 17 18 19 20 21 22 23

Figure 26. Repeat instability with novel larger and smaller alleles. DNA from secondary clones of BS cells, Primers that flank the dinucleotide repeat sequence D18S50 were used, with one of the primers 32P-radiolabeled at its 5’ end. The resulting PCR reaction was run on a nondenaturing polyacrylamide gel. The gels were dried and autoradiographed. Lanes are labeled to indicate the secondary clones analyzed.

102

Since the BS cell extracts have a functional mismatch repair system, we hypothesized that the BLM-MLH1 interaction may regulate the either the substrate specificity or the helicase activity or both of BLM. We created different synthetic substrates using oligos and performed gel-shifts plus or minus the presence of MLH1. BLM’s affinity for these different DNA substrates did not change with the addition for MLH1. There are several possible explanations for these results. MLH1 may not regulate the substrate specificity of the BLM helicase. The stochiometry between BLM and MLH1 may not be correct to regulate substrate specificity or the other mutL and mutS heterodimers must also be present for this function to occur.

We hypothesized that MLH1 may modulate the helicase activity of BLM.

Several other protein-partners of BLM modulate helicase activity of BLM such as

RPA and p53. Secondly, mutL stimulates UvrD helicase activity in vitro. So we performed helicase assays with and without either MLH1and the human mutL heterodimer. Both MLH1 and the mutL heterodimer were able to stimulate the helicase activity on the DO substrate ~ 4 fold over buffer alone. Neither MLH1 or

MLH1/PMS2 possessed any intrinsic helicase activity either from contaminating proteins or from these proteins themselves having strand-displacement activity.

It would be interesting to see if this stimulation increased on different substrates such as X-junctions or IDLs or if the addition of the mutS heterodimers increased the stimulation of the unwinding capability of BLM. The mechanism of how

MLH1 or the eukaryotic mutL heterodimer stimulates BLM’s helicase activity is not known. It is possible that MLH1 binds to the unwound DNA and keeps it from

103

re-annealing similar to RPA. However, this is unlikely since RPA has a much higher affinity for single-stranded DNA than MLH1. The interaction between BLM and MLH1 could induce a confirmational change in BLM that makes the enzyme more processive. Removing the section of BLM that is necessary for its interaction with MLH1 or using peptides or antibodies that negated the interaction between these two proteins would allow us to determine if MLH1 caused a confirmational change in BLM that is responsible for its increase in helicase activity.

We have shown that the BLM helicase is stimulated but the mutL heterodimer in vitro and Curt Harris’ group has shown that p53 inhibits the helicase activity of both BLM and WRN on X-junction substrates. We must remember that in vivo MLH1 associates with PMS2 on most occasions and with the mutS heterodimer MSH2/6. This is particularing intriguing because MSH2/6, p53 and BLM all have high affinities for Holliday junctions (synthetic X-junctions).

Cells treated with HU in S-phase show a complete co-localization of MSH2 and p53 (Zink et al 2002). These co-localization data strongly suggest that these two proteins along with BLM participate in the resolution of stalled replication forks via replication fork regression and/or homologous recombination and Holliday junction resolution. BLM could play a role in recombinational repair via a role in replication fork regression similar to RecA and RecG in E. coli (Figure 27) (Robu et al 2001, McGlynn and Lloyd 2002). In the RecQ field this is the most popular theory for the function of BLM in regards to DNA replication, recombination and repair. Without BLM to aid in the replication fork restarts (via replication fork

104

regression) there would be increased levels of double-strand breaks in the cell and this could easily account of the increased levels of sister-chromatid exchanges in BS cells.

105

Figure 27. Proposed pathway for the recombinational DNA repair of replication forks

DNA replication fork

DNA lesion

Replication fork demise

BLM-mediated fork regression

Replication

Reverse branch migration

Replication restart BLM helicase

106

However, in the absence of either p53 or MSH2 there are increased levels of genomic instability, hyper-recombination and sister-chromatid exchanges which is similar to what occurs in BS cells (De Wind, Dekker et al. 1995)

(Livingstone, White et al. 1992). Therefore it is quite possible that BLM, p53,

MLH1/PMS2, MSH2 and 6 all act together in S-phase cells to aid in the resolution of stalled replication forks. In Figure 27 either p53 or MSH2 could help resolve different intermediates generated by replication fork regression, or one could envision the following model (Figure 28). p53 recognizes a stalled replication fork that has been converted into a Holliday junction; BLM and the mismatch repair proteins arrive at the heteroduplex , if there is significant heterology in the duplex, it is destabilized with the aid of BLM and the invading 3’ strand is removed. p53 also could help control the correct alignment of the 3’ invading end since it forms complexes with polymerase α and β and excises mismatched nucleotides in a replication assay (Huang 1998). Once the stalled replication fork is resolved through homologous recombination, BLM, the mismatch repair proteins and p53 disband, DNA synthesis continues and the cell-cycle progresses.

107

DNA nick

Cleavage of chickenfoot Replication fork collapse

Double strand break and end-processing exonucleases degrade lesion

Strand invasion Mediated and recombination by Rad51

p53 binds to Holliday junction

p53 activates an intra-S phase checkpoint and recruits the BLM helicase and mismatch repair proteins

Replication fork restarts after Holliday junction is resolved, proteins disband and cell enters back into S-phase

Figure 27. Proposed model of the BLM helicase and its interaction with mismatch repair proteins andp53 correcting a lesion on the lagging strand.

108

There are numerous experiments one can perform with the BLM helicase, mutL-mutS heterodimer and p53. Biochemical studies with all of these proteins should be conducted with respect to different DNA substrates such as bubbles and loops, p53 and mutL/mutS could be placed in the helicase assay at the same time. Co-localization studies should be performed with cell lines that are deficient in p53, BLM or the mismatch repair proteins to determine if any of these proteins are crucial to the others in recruitment to stalled replication forks.

Mechanistically, it isn’t clear whether p53 recruits BLM to the stalled replication fork or if BLM recruits p53. It is most probable that all of these proteins work together to resolve the Holliday junction or the chickenfoot structure. If these aberrant substrates are not resolved, p53 is not released from the aberrant DNA structure and the cell either enters apoptosis or senecenses.

We have also performed in vitro end-joining assays to determine if the

BLM helicase is involved in double-strand break repair. Using the supF20 system we determined that the BLM helicase is necessary for the use of small microhomology elements present in the substrate. The pUC18 end-joining assay allows one to measure the efficiency and fidelity of end-joining. BS extracts were just as efficient at rejoining the restriction enzyme digested plasmid as control extracts. These results are in contrast to Runger et al who transfected cut plasmid into BS cells and control cells and detected a lower end-joining efficiency and Gaymes et al who reported that BS cells have a higher end-joining efficiency compared to control cells. These differences could be due to how the nuclear

109

extracts are prepared or the temperature and conditions of the assay. Gaymes et al 2002 increase in end-joining efficiency is ~ two-fold and this could be partially due to variation in the assay. The fidelity of repair in BS cells is not as accurate in BS extracts as compared to control extracts and there is a 5–fold increase in mutation frequency in BS extracts. Most of the mutations in BS extracts were small deletions compared to the deletions from control extracts although there were a subset of BS deletions that were relatively large in size.

Very few of the BS deletions utilized elements of microhomology which supports the supF20 data strongly suggesting that BLM is necessary for the alignments of microhomology elements close to the break. HCT116 extracts were also tested and the mutation frequency was similar to the control cells. This suggests that the BLM-MLH1 interaction isn’t necessary for proper end-joining.

Figure 29 illustrates three potential pathways of end-joining that we are observing as deletions from our extracts. Small deletions may be Ku and DNA-

PK quickly resealing the double–strand break. This is a very efficient process and very little DNA is lost in the process. This is the major pathway observed in the BS extracts. The second pathway involves the usage of microhomology elements and is rarely observed in the BS extracts strongly suggesting that BLM is necessary to help align these elements of microhomology. The large deletions which were greater than 200bp were observed in all the extracts but were a very small percentage in the BS extracts. This suggests that it is a separate error- prone pathway that probably isn’t utilized as much by the cell.

110

Figure 29. Model for the role of BLM in DSB repair

ampr

ori pUC18

AGCTCG 3’ 5’* AATTCGAGC TCGAGCTTAA* 5’ 3’ GCTCG β-galactosidase

Medium size deletions Large deletions Small deletions utilizing microhomology

Ku heterodimer bind to the DNA ends Strand exposure via the BLM helicase and the RAD 50/Mre11 NBS1 complex Ku recruits

DNA-PKcs Ends of DNA

Excessive Reanneal at repeats Digestion by Unknown Xrcc4 and exonuclease ligase IV seal the nick Snip ends Repair synthesis Ligation

Ligase IV

Very efficient process Very little DNA is lost Deletion size is less than 10bp Large deletions No microhomology Medium size deletions needed requires the BLM helicase BLM helicase for microhomology element usage Ku 70/80 heterodimer

Rad50, Nbs1, Mre11 complex

DNA-PKcs Xrcc4 and ligase IV DNA with microhomology

111

These data suggest that BLM plays a role in double-strand break repair and further studies should pursue this idea. Our laboratory has preliminary evidence that BLM is phosphorylated by DNA-PK in the presence of DNA in vitro. If this site is also phosphorylated in vivo, than its functional significance can be investigated by mutating the phosphorylation site to alanine and transfected into

BS cells to see if cells expressing mutated BLM are more sensitive to DNA damaging agents. Maria Jasin’s vector system could also be employed to determine if BLM is involved in homology-directed repair.

In conclusion, BLM is a RecQ helicase with many different functions in

DNA replication, recombination and repair. We were the first group to define the mismatch repair protein MLH1 as a protein-partner of the BLM helicase. MLH1 and the human mutL heterodimer are able to stimulate the helicase activity of

BLM. It is also known that MSH2 and p53 (two other components of the BASC complex) colocalize at stalled replication forks during S-phase. Future studies should be directed to more precisely determine the roles BLM, p53, and MSH2 play in the reinitiation of stalled replication forks. Both RecQ and RecA promote the regression of stalled replication forks in vitro (Robu et al 2001) (McGlynn and

Lloyd 2002). Similar studies could be performed with purified BLM. We have also shown a functional role for BLM in the end-joining process. BLM is phosphorylated by DNA-PK in vitro and if BLM is phosphorylated in vivo by DNA-

PK this will open up another avenue to study BLM, its phosphorylation status and the sensitivity of BS cells to DNA damaging agents when BLM cannot be phosphorylated at sites specific for DNA-PK.

112

Chapter Eight- References

Ababou, M., Dumaire, V., Lecluse, Y., Amor-Gueret, M. (2002). "Bloom's

syndrome protein response to ultraviolet-C radiation and hydroxyurea-

mediated DNA synthesis inhibition." Oncogene 21(13): 2079-88.

Aurias, A., Antoine, J.L., Assathiany, R., Odievre, M., Dutrillaux, B. (1985).

"Radiation sensitivity of Bloom's syndrome lymphocytes during S and G2

phases." Cancer Genet Cytogenet 16(2): 131-6.

Austin, C.A., Marsh, K.L., Wasserman, R.A., Willmore, E., Sayer, P.J., Wang,

J.C., Fisher, L.M. (1995). "Expression, domain structure, and enzymatic

properties of an active recombinant human DNA topoisomerase II beta." J

Biol Chem 270(26): 15739-46.

Bierne, H., Seigneur, M., Ehrlich, S.D., Michel, B. (1997). "uvrD mutations

enhance tandem repeat deletion in the Escherichia coli chromosome via

SOS induction of the RecF recombination pathway." Mol Microbiol 26(3):

557-67.

Bischof, O., Kim, S.H., Irving, J., Beresten, S., Ellis, N.A., Campisi, J. (2001).

"Regulation and localization of the in response to

DNA damage." J Cell Biol 153(2): 367-80.

113

Bloom, D. (1966). "[Congenital telangiectatic erythema syndrome and stunted

growth]." Arch Argent Dermatol 16(1): 26-8.

Brosh, R.M. Jr, Li, J.L., Kenny, M.K., Karow, J.K., Cooper, M.P., Kureekattil,

R.P., Hickson, I.D., Bohr, V.A. (2000). "Replication protein A physically

interacts with the Bloom's syndrome protein and stimulates its helicase

activity." J Biol Chem 275(31): 23500-8.

Chakraverty, R. K. and I. D. Hickson (1999). "Defending genome integrity during

DNA replication: a proposed role for RecQ family helicases." Bioessays

21(4): 286-94.

Chester, N., Kuo, F., Kozak, C., O'Hara, C.D., Leder, P. (1998). "Stage-specific

apoptosis, developmental delay, and embryonic lethality in mice

homozygous for a targeted disruption in the murine Bloom's syndrome

gene." Genes Dev 12(21): 3382-93.

Ellis, N. A. (1997). "DNA helicases in inherited human disorders." Curr Opin

Genet Dev 7(3): 354-63.

114

Ellis, N.A., Groden, J., Ye T.Z., Straughen, J., Lennon, D.J., Ciocci, S.,

Proytcheva, M., German, J. (1995). "The Bloom's syndrome gene product

is homologous to RecQ helicases." Cell 83(4): 655-66.

Ellis, N.A., Lennon, D.J., Proytcheva, M., Alhadef,f B., Henderson, E.E., German,

J. (1995). "Somatic intragenic recombination within the mutated locus BLM

can correct the high sister-chromatid exchange phenotype of Bloom

syndrome cells." Am J Hum Genet 57(5): 1019-27.

Ellis, N.A., Proytcheva, M., Sanz, M.M., Ye, T.Z., German, J. (1999).

"Transfection of BLM into cultured bloom syndrome cells reduces the

sister-chromatid exchange rate toward normal." Am J Hum Genet 65(5):

1368-74.

Ellis, N.A., Roe, A.M., Kozloski, J., Proytcheva, M., Falk, C., German, J. (1994).

"Linkage disequilibrium between the FES, D15S127, and BLM loci in

Ashkenazi Jews with Bloom syndrome." Am J Hum Genet 55(3): 453-60.

Epstein, C. J. and A. G. Motulsky (1996). "Werner syndrome: entering the

helicase era." Bioessays 18(12): 1025-7.

115

Evans, H.J., Adams, A.C., Clarkson, J.M., German, J. (1978). "Chromosome

aberrations and unscheduled DNA synthesis in X- and UV-irradiated

lymphocytes from a boy with Bloom's syndrome and a man with

xeroderma pigmentosum." Cytogenet Cell Genet 20(1-6): 124-40.

Franchitto, A. and Pichierri, P. (2002). "Bloom's syndrome protein is required for

correct relocalization of RAD50/MRE11/NBS1 complex after replication

fork arrest." J Cell Biol 157(1): 19-30.

German, J. (1995). "Bloom's syndrome." Dermatol Clin 13(1): 7-18.

Giannelli, F., Botcherby, P.K., Avery, J.A. (1982). "The effect of aphidicolin on the

rate of DNA replication and unscheduled DNA synthesis of Bloom

syndrome and normal fibroblasts." Hum Genet 60(4): 357-9.

Hall, M.C., Jordan, J.R., Matson, S.W. (1998) Evidence for a physical interaction

between the Escherichia coli methyl-directed mismatch repair proteins

MutL and UvrD. EMBO J. 17(5):1535-41

Hook GJ, Kwok E, Heddle JA. (1984). "Sensitivity of Bloom syndrome fibroblasts

to mitomycin C." Mutat Res 131(5-6): 223-30.

116

Johnson, F.B., Lombard, D.B., Neff, N.F., Mastrangelo, M.A., Dewolf, W., Ellis,

N.A., Marciniak, R.A., Yin, Y., Jaenisch, R., Guarente, L. (2000).

"Association of the Bloom syndrome protein with topoisomerase IIIalpha in

somatic and meiotic cells." Cancer Res 60(5): 1162-7.

Kanaar, R., Hoeijmakers, J.H., van Gent, D.C. (1998). "Molecular mechanisms of

DNA double strand break repair." Trends Cell Biol 8(12): 483-9.

Kaneko, H., Orii K.O., Matsui, E., Shimozawa, N., Fukao, T., Matsumoto, T.,

Shimamoto, A., Furuichi, Y., Hayakawa, S., Kasahara, K., Kondo, N.

(1997). "BLM (the causative gene of Bloom syndrome) protein

translocation into the nucleus by a nuclear localization signal." Biochem

Biophys Res Commun 240(2): 348-53.

Kapp, L. N. (1982). "DNA fork displacement rates in Bloom's syndrome

fibroblasts." Biochim Biophys Acta 696(2): 226-7.

Karow, J.K., Chakraverty, R.K., Hickson, I.D. (1997). "The Bloom's syndrome

gene product is a 3'-5' DNA helicase." J Biol Chem 272(49): 30611-4.

Karow, J.K., Newman, R.H., Freemont, P.S., Hickson, I.D. (1999). "Oligomeric

ring structure of the Bloom's syndrome helicase." Curr Biol 9(11): 597-600.

117

Kenne, K. and Akerblom, L. (1990). "Deoxyribonucleoside triphosphate pool

levels in three cell strains of human chromosome instability syndromes:

ataxia telangiectasia (GM2052), Bloom's syndrome (GM1492), and

Fanconi's anemia (GM368)." Cancer Biochem Biophys 11(1): 69-77.

Khanna, K. K. and Jackson, S.P. (2001). "DNA double-strand breaks: signaling,

repair and the cancer connection." Nat Genet 27(3): 247-54.

Kitao, S., Shimamoto, A., Goto, M., Miller, R.W., Smithson, W.A., Lindor, N.M.,

Furuichi, Y. (1999). "Mutations in RECQL4 cause a subset of cases of

Rothmund-Thomson syndrome." Nat Genet 22(1): 82-4.

Krepinsky, A.B., Rainbow, A.J., Heddle, J.A. (1980). "Studies on the ultraviolet

light sensitivity of Bloom's syndrome fibroblasts." Mutat Res 69(2): 357-68.

Kuhn, E. M. and Therman, E. (1986). "Cytogenetics of Bloom's syndrome."

Cancer Genet Cytogenet 22(1): 1-18.

Kurihara, T., Inoue, M., Tatsumi, K. (1987). "Hypersensitivity of Bloom's

syndrome fibroblasts to N-ethyl-N-nitrosourea." Mutat Res 184(2): 147-51.

118

Kusano, K., Johnson-Schlitz, D.M., Engels, W.R. (2001). "Sterility of Drosophila

with mutations in the Bloom syndrome gene--complementation by Ku70."

Science 291(5513): 2600-2.

Langland G, Kordich J, Creaney J, Goss KH, Lillard-Wetherell K, Bebenek K,

Kunkel TA, Groden J. (2001). The Bloom's syndrome protein (BLM)

interacts with MLH1 but is not required for DNA mismatch repair. J Biol

Chem 2001 276(32):30031-5

Li, Y., Carty, M.P., Oakley, G.G., Seidman, M.M., Medvedovic, M., Dixon, K.

(2001). "Expression of ATM in ataxia telangiectasia fibroblasts rescues

defects in DNA double-strand break repair in nuclear extracts." Environ

Mol Mutagen 37(2): 128-40.

Lonn, U., Lonn, S., Nylen, U., Winblad, G., German, J. (1990). "An abnormal

profile of DNA replication intermediates in Bloom's syndrome." Cancer

Res 50(11): 3141-5.

Marti, T.M., Kunz, C., Fleck, O. (2002). DNA mismatch repair and mutation

avoidance pathways. J Cell Physiol 191(1):28-41

119

McDaniel, L. D. and Schultz, R.A. (1992). "Elevated sister chromatid exchange

phenotype of Bloom syndrome cells is complemented by human

chromosome 15." Proc Natl Acad Sci U S A 89(17): 7968-72.

McGlynn,P.and Lloyd, R.G. (2002). "Genome stability and the processing of

damaged replication forks by RecG" Trends Genet 2002 Aug;18(8):413-9

Mendonca, V.M., Klepin, H.D., Matson, S.W. (1995). "DNA helicases in

recombination and repair: construction of a delta uvrD delta helD delta

recQ mutant deficient in recombination and repair." J Bacteriol 177(5):

1326-35.

Neff, N.F., Ellis, N.A., Ye, T.Z., Noonan, J., Huang, K., Sanz, M., Proytcheva, M.

(1999). "The DNA helicase activity of BLM is necessary for the correction

of the genomic instability of bloom syndrome cells." Mol Biol Cell 10(3):

665-76.

Ockey, C. H. and Saffhill, R. (1986). "Delayed DNA maturation, a possible cause

of the elevated sister-chromatid exchange in Bloom's syndrome."

Carcinogenesis 7(1): 53-7.

Patel, SS, and Picha, KM. (2000). Structure and function of hexameric helicases.

Annu Rev Biochem 69:651-97.

120

Pedrazzi, G., Perrera, C., Blaser, H., Kuster, P., Marra, G., Davies, S.L., Ryu,

G.H., Freire, R., Hickson, I.D., Jiricny, J., Stagljar I. (2001). Direct

association of Bloom's syndrome gene product with the human mismatch

repair protein MLH1. Nucleic Acids Res 29(21):4378-86

Puranam, K. L. and Blackshear, P.J. (1994). "Cloning and characterization of

RECQL, a potential human homologue of the Escherichia coli DNA

helicase RecQ." J Biol Chem 269(47): 29838-45.

Puranam, K.L., Kennington, E., Sait, S.N., Shows, T.B., Rochelle, J.M., Seldin,

M.F., Blackshear, P.J. (1995). "Chromosomal localization of the gene

encoding the human DNA helicase RECQL and its mouse homologue."

Genomics 26(3): 595-8.

Rosin, M. P. and German, J. (1985). "Evidence for chromosome instability in vivo

in Bloom syndrome: increased numbers of micronuclei in exfoliated cells."

Hum Genet 71(3): 187-91.

121

Runger, T. M. and Kraemer, K.H. (1989). "Joining of linear plasmid DNA is

reduced and error-prone in Bloom's syndrome cells." Embo J 8(5): 1419-

25.

Robu, M.E., Inman, R.B. and Cox, M.M. (2001). "RecA protein promotes the

regession of stalled replication forks in vitro. PNAS 98(15): 8211-8218.

Shiraishi, Y. and. Sandberg A.A. (1978). "Effects of mitomycin C on sister

chromatid exchange in normal and Bloom's syndrome cells." Mutat Res

49(2): 233-8.

Stewart, E., Chapman, C.R., Al-Khodairy, F., Carr, A.M., Enoch, T. (1997).

"rqh1+, a fission yeast gene related to the Bloom's and Werner's

syndrome genes, is required for reversible S phase arrest." Embo J

16(10): 2682-92.

Sun, H., Karow, J.K., Hickson, .I.D., Maizels, N. (1998). "The Bloom's syndrome

helicase unwinds G4 DNA." J Biol Chem 273(42): 27587-92.

Tachibana, A., Tatsumi, K., Masui, T., Kato, T. (1996). "Large deletions at the

HPRT locus associated with the mutator phenotype in a Bloom's

syndrome lymphoblastoid cell line." Mol Carcinog 17(1): 41-7.

122

Thacker, J., Chalk, J., Ganesh, A., North, P. (1992). "A mechanism for deletion

formation in DNA by human cell extracts: the involvement of short

sequence repeats." Nucleic Acids Res 20(23): 6183-8.

Otto, P.G., Otto, P.A., Therman, E. (1981). "Mitotic recombination and

segregation of satellites in Bloom's syndrome." Chromosoma 82(5): 627-

36.

Umar, A., Boyer, J.C., Thomas, D.C., Nguyen, D.C., Risinger, J.I., Boyd, J.,

Ionov, Y., Perucho, M., Kunkel, T.A. (1994). "Defective mismatch repair in

extracts of colorectal and cell lines exhibiting

microsatellite instability." J Biol Chem 269(20): 14367-70.

Umezu, K. and Nakayama, H. (1993). "RecQ DNA helicase of Escherichia coli.

Characterization of the helix-unwinding activity with emphasis on the effect

of single-stranded DNA-binding protein." J Mol Biol 230(4): 1145-50.

Umezu, K., Nakayama, K., Nakayama, H. (1990). "Escherichia coli RecQ protein

is a DNA helicase." Proc Natl Acad Sci U S A 87(14): 5363-7.

van Laar, T., Steegenga, W.T., Jochemsen, A.G., Terleth, C., van der Eb, A.J.

(1994). "Bloom's syndrome cells GM1492 lack detectable p53 protein but

123

exhibit normal G1 cell-cycle arrest after UV irradiation." Oncogene 9(3):

981-3.

Walpita, D., Plug, A.W., Neff, N.F., German, J., Ashley, T. (1999). "Bloom's

syndrome protein, BLM, colocalizes with replication protein A in meiotic

prophase nuclei of mammalian spermatocytes." Proc Natl Acad Sci U S A

96(10): 5622-7.

Wang, X.W., Tseng, A., Ellis, N.A., Spillare, E.A., Linke, S.P., Robles, AI, Seker,

H., Yang, Q., Hu, P., Beresten, S., Bemmels, N.A., Garfield, S., Harris,

C.C. (2001). "Functional interaction of p53 and BLM DNA helicase in

apoptosis." J Biol Chem 276(35): 32948-55.

Wang, Y., Cortez, D., Yazdi, P., Neff, N., Elledge, S.J., Qin, J. (2000). "BASC, a

super complex of BRCA1-associated proteins involved in the recognition

and repair of aberrant DNA structures." Genes Dev 14(8): 927-39.

Watt, P.M., Louis, E.J., Borts, R.H., Hickson, I.D (1996). "SGS1, a homologue of

the Bloom's and Werner's syndrome genes, is required for maintenance of

genome stability in ." Genetics 144(3): 935-45.

124

Watt, P.M., Louis, E.J., Borts, R.H., Hickson, I.D. (1995). "Sgs1: a eukaryotic

homolog of E. coli RecQ that interacts with topoisomerase II in vivo and is

required for faithful chromosome segregation." Cell 81(2): 253-60.

Woodage, T., Prasad, M., Dixon, J.W., Selby, R.E, Romain, D.R., Columbano-

Green, L.M., Graham, D., Rogan, P.K., Seip, J.R., Smith, A., et al. (1994).

"Bloom syndrome and maternal uniparental disomy for chromosome 15."

Am J Hum Genet 55(1): 74-80.

Wu, J., He, J., Mountz, J.D. (1998). "Effect of age and apoptosis on the mouse

homologue of the huWRN gene." Mech Ageing Dev 103(1): 27-44.

Wu, L., Davies, S.L., Levitt, N.C., Hickson, I.D. (2001). "Potential role for the BLM

helicase in recombinational repair via a conserved interaction with

RAD51." J Biol Chem 276(22): 19375-81.

Wu, L., Davies, S.L., North, P.S., Goulaouic, H., Riou, J.F., Turley, H., Gatter,

K.C., Hickson, I.D. (2000). "The Bloom's syndrome gene product interacts

with topoisomerase III." J Biol Chem 275(13): 9636-44.

Yannone, S.M., Roy, S., Chan, D.W., Murphy, M.B., Huang, S., Campisi, J.,

Chen, D.J. (2001). "Werner syndrome protein is regulated and

125

phosphorylated by DNA-dependent protein kinase." J Biol Chem 276(41):

38242-8.

Zhong, S., Hu, P., Ye, T.Z., Stan, R., Ellis, N.A., Pandolfi, P.P. (1999). "A role for

PML and the nuclear body in genomic stability." Oncogene 18(56): 7941-

7.

Zink, D., Mayr, C., Janz, C and Wiesmuller, L. (2002). "Association of p53 and

MSH2 with recombinative repair complexes during S phase." Oncogene

21(31): 4788-800.

126