Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy...

162
AN ABSTRACT OF THE THESIS OF Kevin M. Culligan for the degree of Doctor of Philosophy in Molecular and Cellular Biology presented on June 7. 2000. Title: Mismatch Repair in Plants: Identification and Characterization of Arabidopsis thaliana MutS Homolog Proteins. Abstract approved: John B. Hays All eukaryotic organisms examined thus far encode homologs of the eubacterial DNA- mismatch-repair protein MutS. In comparison to other eukaryotic organisms, little is known about how plants combat mismatched DNA basepairs arising during the process of DNA replication. To address whether plants utilize similar MutS-homolog (MSH) proteins in mismatch-repair systems, I focused on three primary questions: i) Do plants encode homologs of MutS proteins? ii) Do these homologs fall into the same distinct subfamilies seen in other eukaryotes? iii) Do these homologs share conserved biochemical activities with their other eukaryotic counterparts? Using sets of degenerate polymerase chain reaction (PCR) primers corresponding to highly conserved MSH protein domains, I amplified and cloned segments of Arabidopsis Redacted for Privacy

Transcript of Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy...

Page 1: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

AN ABSTRACT OF THE THESIS OF

Kevin M. Culligan for the degree of Doctor of Philosophy in Molecular and Cellular

Biology presented on June 7. 2000. Title: Mismatch Repair in Plants: Identification

and Characterization of Arabidopsis thaliana MutS Homolog Proteins.

Abstract approved:

John B. Hays

All eukaryotic organisms examined thus far encode homologs of the eubacterial DNA-

mismatch-repair protein MutS. In comparison to other eukaryotic organisms, little is

known about how plants combat mismatched DNA basepairs arising during the

process of DNA replication. To address whether plants utilize similar MutS-homolog

(MSH) proteins in mismatch-repair systems, I focused on three primary questions: i)

Do plants encode homologs of MutS proteins? ii) Do these homologs fall into the

same distinct subfamilies seen in other eukaryotes? iii) Do these homologs share

conserved biochemical activities with their other eukaryotic counterparts? Using sets

of degenerate polymerase chain reaction (PCR) primers corresponding to highly

conserved MSH protein domains, I amplified and cloned segments of Arabidopsis

Redacted for Privacy

Page 2: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

thaliana cDNA. Two of these segments encoded conserved portions of Arabidopsis

thaliana MSH genes, and their full-length cDNAs were isolated from Arabidopsis

thaliana cDNA libraries. Utilizing Arabidopsis thaliana genome sequences deposited

in GenBank, I further identified two other MSH genes; their full-length cDNAs were

isolated from Arabidopsis thaliana mRNA using reverse-transcription (RT-) PCR.

Extensive phylogenetic analyses proved that the four predicted plant-protein

sequences clearly fell into the conserved MSH2, MSH3 and MSH6 sub-families seen

in other eukaryotes. Furthermore, the phylogenetic analyses revealed a novel feature

in higher plants the presence of two MSH6-like proteins, designated MSH6 and

MSH7. Combinations of Arabidopsis thaliana atMSH2, atMSH3, atMSH6, and

atMSH7 proteins, products of in vitro transcription and translation, were analyzed for

protein-protein interactions. The atMSH2 protein formed heterodimers with atMSH6,

atMSH3, or atMSH7 proteins (designated atMutSu, atMutSI3, and atMutSy,

respectively) but no other complexes were observed. The abilities of the various

heterodimers to bind to mismatched 51-mer oligoduplexes were measured by

electrophoretic mobility-shift assays. Both atMutSct and atMutS3 shared conserved

mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel

substrate specificity. These data suggest that plants utilize unique mismatch-repair

pathways to maintain the integrity of their genomes.

Page 3: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

©Copyright by Kevin M. Culligan

June 7, 2000All Rights Reserved

Page 4: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

Mismatch Repair in Plants: Identification and Characterization ofArabidopsis

thaliana MutS Homolog Proteins

by

Kevin M. Culligan

A THESIS

submitted to

Oregon State University

in partial fulfillment ofthe requirements for the

degree of

Doctor of Philosophy

Presented June 7, 2000

Commencement June 2001

Page 5: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

Doctor of Philosophy thesis of Kevin M. Culligan presented on June 7. 2000

APPROVED:

Mp'Professor, representing Molecá1'ar and Cellular Biology

//

Chair of Program in Mo!u1ar anlCellular Biology

Dean of Gradüate'School

I understand that my thesis will become part of the permanent collection of OregonState University libraries. My signature below authorizes release of my thesis to any

reader upon request.

Kevin M. lligan, Author

Redacted for Privacy

Redacted for Privacy

Redacted for Privacy

Redacted for Privacy

Page 6: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

ACKNOWLEDGEMENTS

I wish to thank my advisor, Dr. John Hays for his sincere dedication to science,

and for inspiring me to achieve my goals. I also thank those individuals who have

assisted me throughout my time at Oregon State University, including all members of

the Hays lab, my fellow MCB graduate students, my graduate committee members,

and the faculty and staff of the MCB program. The majority of my graduate thesis

work has been supported in part by grants from NSF, and a graduate student

fellowship from NIEHS.

Page 7: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

CONTRIBUTION OF AUTHORS

Dr. Gilbert Meyer-Gauen and Dr. James Lyons-Weiler were involved in the

design, analysis and writing of Chapter 3.

Page 8: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

TABLE OF CONTENTS

Page

1. Introduction .1

1.1 Long-Patch, Methyl-Directed Mismatch Repair in E. coli, an Overview .........2

1.1.1 Evidence for Mismatch Repair In Vivo .................................................... 2

1.1.2 Evidence for Mismatch Repair In Vitro ................................................... 5

1.1.3 The E. coli Model for Methyl-Directed Mismatch Repair ..................... 10

1.2 Long-Patch Mismatch Repair in Eukaryotes .................................................. 14

1.2.1 Evidence for Eukaryotic Mismatch Repair ............................................ 14

1.2.2 Identification of MutS and MutL Homologs in Eukaryotes ................. 15

1.2.3 MSH and MILH Protein Families in Eukaryotes .................................... 171.2.4 A Generalized Model for Mismatch Repair in Eukaryotes ................... 27

1.2.5 Mismatch Repair and Recombination ................................................... 30

1.2.6 Why Study Mismatch Repair in Plants7 ................................................31

2. DNA Mismatch Repair in Plants: An Arabidopsis thaliana Gene that Predicts a

Protein Belonging to the MSH2 Subfamily of Eukaryotic MutS Homologs ....... 35

2.1 Abstract .......................................................................................................... 36

2.2 Introduction ..................................................................................................... 36

2.3 Results ............................................................................................................. 37

2.3.1 Isolation and Initial Characterization of atMSH2, and a Gene

Fragment Similar to MSH6-like Protein ............................................... 37

2.3.2 Overexpression of aIMSH2 in E. coli Cells ..........................................45

2.3.3 Negative Complementation of Wild-Type E. coli Cells in the Presence

ofatMSH2 ............................................................................................. 48

Page 9: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

TABLE OF CONTENTS (Continued)

Page

2.4 Conclusions .49

2.5 Materials and Methods .................................................................................... 49

2.5.1 Growth Conditions ................................................................................. 492.5.2 PCR Techniques ..................................................................................... 51

2.5.3 Southern Hybridization .......................................................................... 52

2.5.4 Isolation of cDNAs ................................................................................53

2.5.5 Isolation of 5'-RACE Products .............................................................. 54

2.5.6 Genomic Clone Isolation .......................................................................54

2.5.7 Analysis of DNA Sequences .................................................................55

2.5.8 Overexpression of atMSH2 ...................................................................55

2.5.9 Negative Complementation Analysis ...................................................57

3. Evolutionary Origin, Diversification and Specialization of Eukaryotic MutS

Homolog Mismatch Repair Proteins .................................................................... 58

3.1 Abstract ........................................................................................................... 59

3.2 Introduction ..................................................................................................... 60

3.3 Results and Conclusions ................................................................................. 62

3.3.1 Classification of MutS-like Proteins According to SequenceOrganization and Co-occurrence with MutL Proteins .......................... 62

3.3.2 Optimization of Analysis of MutS/MSH and MSP Phylogeny ............. 71

3.3.3 Reconstruction of MutS/MSH Phylogeny ............................................. 763.3.4 Rooting the Tree ..................................................................................... 80

3.3.5 Eukaryotic MSH Gene Duplication and Specialization ......................... 82

3.4 Perspectives ..................................................................................................... 84

3.5 Methods ........................................................................................................... 87

Page 10: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

TABLE OF CONTENTS (Continued)

3.5.1 Alignments and Phylogenetic Methods ................................................. 87

3.5.2 RASA Analysis ...................................................................................... 88

4. Arabidopsis thaliana MutS-Homolog Proteins atMSH2, atMSH3, atMSH6,

and a Novel atMSH7 Protein Form Three Distinct Heterodimers with

Different Specificities for Mismatched DNA ...................................................... 90

4.1 Abstract ........................................................................................................... 91

4.2 Introduction ..................................................................................................... 92

4.3 Results ............................................................................................................. 92

4.3.1 Identification of Arabidopsis mutS-like Genes ...................................... 924.3.2 Interactions Between MSH2, MSH3, MSH6 and MSH7 Proteins in

TranslationMixtures .............................................................................. 98

4.3.3 Mature atMSH1 Forms a Homodimer ............................................... 1084.3.4 Binding of atMSH Proteins to Mismatched DNA ............................... 110

4.4 Discussion ..................................................................................................... 115

4.5 Methods ......................................................................................................... 120

4.5.1 Isolation of cDNAs .............................................................................. 1204.5.2 Sequence Alignment and Phylogenetic Analysis ................................. 1234.5.3 In Vitro Transcription and Translation ................................................. 1234.5.4 Analysis of Proteins by Sedimentation Through Sucrose Density

Gradients ............................................................................................... 124

4.5.5 Gel-Filtration Chromatography ............................................................ 124

4.5.6 Preparation of Oligomer Duplexes ..................................................... 125

4.5.7 Mobility Shift Assays ........................................................................... 126

5. Conclusions ...........................................................................................................128

Page 11: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

TABLE OF CONTENTS (Continued)

5.1 Summary .129

5.2 Recommendations for Future Research ........................................................ 130

BIBLIOGRAPHY...................................................................................................... 133

Page 12: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

LIST OF FIGURES

Figure

1. Model for initiation of mismatch repair ........................................................... 11

2. Model for excision/resynthesis steps following mismatch repair initiation ....13

3. Eukaryotic model for mismatch repair ............................................................. 28

4. Alignment of MSH2-like proteins ................................................................... 39

5. Southern analysis of genomic DNA from Arabidopsis thaliana .....................41

6. Structure of the atMSH2 gene ......................................................................... 43

7. Phylogenetic analysis of MutS homologs ........................................................ 44

8. SDS-PAGE analysis of HIS-tag purification attempts of atMSH2 ................. 47

9. Structure of MutSIMSH and MSP proteins ..................................................... 65

10. Schematic neighbor-joining (NJ)trees and conesponding RASA taxon-variance ratios using different combinations and/or regions of MutS-like

protein sequence ............................................................................................... 73

11. Neighbor-joining (NJ) tree for Dayhoff PAM distances among MutS/MSH

proteinsequences ............................................................................................ 78

12. Schematic diagram of MutS/MSH evolution ................................................... 83

13. Alignment of MSH6-like protein sequences .................................................... 94

14. Neighbor-joining tree for Dayhoff PAM distances among a representative

set of complete MutS/MSH protein sequences ................................................ 97

Page 13: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

LIST OF FIGURES (Continued)

Figure

15. SDS-PAGE analysis of human and Arabidopsis co-synthesis reaction

mixtures............................................................................................................ 99

16. Gel-filtration-chromatography analysis of Arabidopsis 35S-labeled

proteins........................................................................................................... 104

17. Kay vs. log MW (Molecular Weight) plot of standards and MSHpolypeptides analyzed by gel-filtration chromatography ............................... 107

18. Elution profile for full-length and N-terminal deletion atMSH1 translationmixtures.......................................................................................................... 109

19. Representative mobility-shift assays of co-synthesis mixtures of human

and Arabidopsis polypeptides ........................................................................ 111

20. Representative mobility-shift assays of single human and Arabidopsis

synthesismixtures .......................................................................................... 114

Page 14: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

LIST OF TABLES

Table Page

1. Eukaryotic mismatch repair proteins ............................................................... 18

2. Negative complimentation of wild-type E. coli in the presence of atMSH2 . . .50

3. MutS/MSH and MSP sequences ...................................................................... 63

4. Relative binding of MSH heterodimers to oligomer substrates ..................... 112

Page 15: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

DEDICATION

This thesis is dedicated to my family, my mother Carol, my father James, and

my brother Brian, for their love and support.

Page 16: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

Mismatch Repair in Plants: Identification and Characterization ofArabidopsis

thaliana MutS Homolog Proteins

Chapter 1

Introduction

Kevin M. Culligan

Page 17: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

2

To faithfully transmit their genetic material (DNA) to subsequent generations,

cells must constantly recognize and repair base/base mispairs (non-Watson-Crick) that

result primarily from the process of DNA replication. The "long-patch DNA

mismatch repair" pathway enhances the fidelity of DNA replication by factors of 102

to iO (Kornberg and Baker, 1992) by correcting DNA polymerase base-

misincorporation errors (mismatches). Similar systems have been identified in both

prokaryotes and eukaryotes, and now appear to be both evolutionarily and

biochemically conserved throughout both kingdoms. To date, the mismatch-repair

pathway has been studied most extensively in the model organism Escherichia coli,

and provides the most complete mismatch-repair model (Modrich, 1991).

1.1. Long-Patch, Methyl-Directed Mismatch Repair in E. ccli. an Overview

1.1.1 Evidence for Mismatch Repair In Vivo

Mismatch repair was first proposed to explain non-reciprocal transfer of

genetic information from one DNA molecular to another, known as gene conversion.

This was first observed in genetic recombination experiments involving

bacteriophages or bacterial mutagenesis studies (Hershey and Rotman, 1949 for

example). However, it was in 1964 that Robin Holliday proposed a model for the

Page 18: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

3

process of mismatch repair to explain gene conversion in fungi (Holliday, 1964). In

his model, the process of recombination was postulated to produce regions of

heterology, which must be converted in favor of one parental DNA strand or another

(gene conversion), or else post-meiotic segregation (the non-Mendelian separation of

phenotypes, usually observed in the first mitotic cell division following meiosis)

would occur at higher rates than observed (Holliday, 1964). Holliday suggested this

process is likely to be enzymatically mediated.

More direct evidence for the existence of mismatch repair came from

Escherichia coli cell transfection studies employing either X phage (Wildenberg and

Meselson, 1975; Wagner and Meselson, 1976), 4X174 (Baas and Jansz, 1972) or T7

phage (Bauer et al., 1981) DNA heteroduplexes containing defined genetic markers.

These experiments provided evidence that heteroduplex DNA is converted in favor of

one or the other parental DNA strands (observed here through known genetic markers)

before DNA replication occurs, thus suggesting E. coli cells contain specific systems

capable of recognizing heteroduplexes in DNA and repairing them. This led Wagner

and Meselson (1976) to suggest that systems in E. coli that repair heteroduplexed

DNA may in fact play a role in correcting biosynthetic errors following DNA

replication, but it was not clear how such a system could discriminate between the

newly formed daughter strand and the parental strand. Experiments employing X

Page 19: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

phage DNA heteroduplexes with defined states of d(GATC) (dam) methylation clearly

demonstrated a distinct bias for repair against DNA strands undermethylated at

d(GATC) sequences (Pukkila et al., 1983), suggesting that the state of DNA

methylation is a signal for repair of the DNA strand. Additionally, these observations

provided an explanation for the increased mutation rate observed in dam strains.

Transfection assays were further employed to identify several gene mutations

that inactivated mismatch repair in vivo. Mutant strains defective in mutH, mutL,

mutS, and uvrD function, which confer a high spontaneous mutation rate in E. coli

cells (Cox, 1975), eliminated heteroduplex correction in these assays (Nevers and

Spatz, 1975; Bauer et al., 1981; Pukkila et al., 1983). Additionally, Glickman and

Radman (1980) had found that dam strains did not grow well in the presence of the

base analog 2-aminopurine, and utilized this phenotype to isolate second site

repressors of dam strains grown in the presence of 2-aminopurine. Since 2-

aminopurine would increase mismatch repair activity due to increases in mispairs

involving 2-aminopurine (and thus increasing the possibility of dsDNA breaks),

inactivation of the mismatch-repair pathway could then suppress this phenotype. A

majority of the suppressor sites inactivated mutH, mutL, or mutS function, suggesting

the dam and mutHLS genes act within the same pathway. Glickman and Radman

(1980) hypothesized that mismatch repair, in the absence of methylation, initiates

Page 20: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

5

repair on both strands of DNA, causing double-strand breaks (overlapping mismatch

repair excision patches) in the presence of functional mismatch repair. [It was later

suggested, however, that this lethal phenotype was actually a result of activated MutH

protein making a second incision at an unmethylated d(GATC) site in the presence of

a mispaired base, thus causing a double-strand break. (see below)]

The specificity, or efficiency of repair of particular mismatched substrates

(G/T, C/C), of the mismatch-repair system was also investigated via the transfection

assay. Transition mutations (G/T, A/C) appeared to be most efficiently repaired over

transversion mutations (GIG, A/A, T/T, C/T, G/A), and almost no repair was observed

for C/C transversion mutations (Kramer et al., 1984; Dohet et al., 1985; Jones et al.,

1987). Mismatched substrates corresponding to insertion/deletion loopouts (IDLs,

TTT'/AAAA for example) up to 4 base pairs were also repaired efficiently (Dohet et

al., 1986; Dohet et al., 1987).

1.1.2 Evidence for Mismatch Repair In Vitro

Development of an In Vitro Assay System. Although studies in vivo gave

compelling evidence for mismatch repair, they provided little insight into what

specific roles the mutH, mutL, mutS or uvrD gene products play in the biochemical

process of mismatch repair. A significant first step toward elucidating the biochemical

Page 21: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

properties of these and other proteins involved in mismatch repair, was the

development of an in vitro assay. Modrich and co-workers constructed a series of

defined mismatched substrates constructed within an EcoRI site of covalently-closed-

circular bacteriophage fl DNA (Lu et al., 1983). The fi DNA contained several

d(GATC) sites, and could be prepared with either strand methylated. Repair of these

substrates would render the molecule susceptible to cleavage by the EcoRl enzyme,

and repair rates could be determined quantitatively.

Lu et al. (1983) prepared extracts from mutt mutH, mutL, mutS, or uvrlY E.

coli cells, and tested each for its ability to repair heteroduplexes in vitro. The results

indicated that extracts deficient in MutH, MutL, MutS, or UvrD proteins lacked

significant repair of heteroduplexed substrate over mut cell extracts. However,

approximately wild-type levels of repair was observed if two or more different mutant

cell extracts (mutS + mutL, for example) were mixed. These data confirmed that

MutH, MutL, MutS, and UvrD are necessary components of mismatch repair

pathways, and indicated that individual protein components could be purified from

cell extracts by complementation experiments utilizing mismatch repair in vitro

assays.

The mutH, mutL, and mutS genes were isolated from E. co/i and S.

lyphimurium and their gene products purified (Pang et al., 1985; Su and Modrich,

Page 22: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

7

1986; Welsh et at., 1987; Grilley et al., 1989). The MutS protein was shown to bind

to mispaired bases (Su and Modrich, 1986) suggesting a role in the recognition of

mispaired bases. Although no function had been determined for the rnutL gene

product, the MutL protein was found to interact with MutS protein (Grilley et at.,

1989). The MutH protein was found to possess a very weak endonuclease activity at

hemimethylated or unmethylated d(GATC) sequences in DNA, suggestive of a role in

strand discrimination (Welsh et al., 1987).

Reconstitution of the Mismatch-Repair Reaction In Vitro. Purification of

the MutH, MutL, and MutS components of mismatch repair, along with the

development of the in vitro assay, provided the opportunity to reconstruct the entire

biochemical reaction. However, purified components known to be involved in the

process of mismatch repair were not sufficient to complete the repair reaction in vitro

(Lahue et at., 1989). This led Modrich and co-workers to search for additional

components (proteins and cofactors) that would complete the reaction. Since it was

likely that the reaction involves excision and resynthesis steps (Lu et al., 1983), a

potential candidate for an additional factor was DNA polymerase I, known to be

involved in several repair-type reactions. However, polA extracts (deficient in DNA

polymerase I) exhibited normal levels of mismatch repair. Because previous studies

showed a mismatch repair requirement for polymerase III in vivo (Lu et at., 1983;

Page 23: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

Schaaper, 1988), Modrich and co-workers tested extracts of a dnazts (the dnaZ gene

encodes the t and y subunits of DNA polymerase III holoenzyme) in in vitro assays

(Lahue et al., 1989). They found that these extracts indeed showed a temperature-

sensitive mismatch-repair activity, and by adding back purified-DNA polymerase III

holoenzyme to the reaction, they could restore activity to the mutant (dnaZts) extract at

elevated temperatures. This suggested that the highly processive, replicative

polymerase III holoenzyme was involved in the mismatch-repair reaction, consistent

with the large repair tracts formed during the mismatch-repair reaction.

Although a mixture of purified MutH, MutL, MutS, SSB, helicase II (uvrD),

and DNA polymerase III holoenzyme proteins, plus MgC12, ATP, and dNTPs, could

carry out mismatch repair in vitro, the reaction was inhibited by DNA 1igaseINAD

(Lahue et al., 1989). Since it was believed that DNA ligase was needed to covalently

close the excision/resynthesis patch, it followed that perhaps additional components

must be necessary. A 55-kD protein was isolated that restores activity to reactions

containing DNA ligase. Once purified, the protein factor appeared identical, in both

size, and N-terminal protein sequence, to exonuclease I (Exol), a 3' to 5' exonuclease.

Indeed, adding purified exonuclease I to reactions (described above), including DNA

ligase and its cofactor NAD, yielded repaired, covalently-closed DNA product,

suggesting the completion of the repair reaction (Lahue et al., 1989).

Page 24: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

To confirm protein and cofactor requirements for the initiation and repair of

mismatched substrates in vitro, experiments were conducted using mixtures lacking

individual protein or cofactor components of the reaction. The results proved that

MutH, MutL, MutS, DNA polymerase III holoenzyme, SSB, exonuclease I, DNA

helicase II, DNA ligase, NAD, MgCl2, ATP, and dNTPs were necessary and

sufficient to carryout a complete methyl-directed mismatch repair reaction on

covalently closed, d(GATC) hemimethylated repair DNA substrates (Lahue et al.,

1989). These results were further confirmed by the observation that the specificities

of mismatch repair in the reconstituted system mostly coincided with the specificities

determined by in vivo experiments and by in vitro experiments employing cell extracts

(described above). Furthermore, experiments employing various repair substrates

indicated that the presence of a d(GATC) site or a nick in either of the DNA strands

was sufficient to direct repair (to that particular strand), and in the case of the nicked

substrates, the reaction could take place in the absence of MutH protein.

Since activated MutH nicks hemimethylated d(GATC) sites either 5' or 3' to

the mismatched basepair (Bruni et al., 1988), it was likely that the excision reaction

could proceed in either a 5' to 3' or 3' to 5' direction. By mapping excision tracts

(direct EM visualization of single-stranded gaps under conditions of restricted DNA

synthesis) in both circular and linear repair substrates, Modrich and co-workers found

Page 25: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

10

the gaps formed during the excision process (up to 1 kilobase in size) spanned the

shortest distance to the mispair, whether the hemimethylated d(GATC) sequence was

5' or 3' to the mispair (Grilley etal., 1993; Cooper et al., 1993). In addition to Exol,

which complimented the 3' to 5' exonuclease activity in repair reactions, both ExoVil

and RecJ (5' to 3' exonucleases) were found to compliment the 5' to 3' exonuclease

activity when the mispair is 3' to the unmethylated d(GATC) sequence (Cooper et al.,

1993).

These critical and penetrating experiments not only led to a better

understanding of methyl-directed mismatch repair in E. coli, but it allowed Modnch

and co-workers to propose a general model for the initiation and excision reactions of

the mismatch repair pathway (Modrich, 1991).

1.1.3 The E. coli Model for Methyl-Directed Mismatch Repair

Initiation of Mismatch Repair. Current models for the initiation of methyl-

directed mismatch repair in E. coli (Figure 1) postulate two main steps, involving the

MutS, MutL, and MutH proteins. The first step is recognition of mispaired bases.

MutS, acting as a homodimer, binds with varying affinity to mispaired bases in DNA

(Su and Modrich, 1986; Su et al., 1988). The second step involves strand

discrimination to initiate the excision reaction on the newly synthesized (nascent)

Page 26: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

11

CH CHI I

4,

MutS

AlPMutLMutH

ADP + P

4,

MutH incision at

d(GATC) site

CH CH,I I

Figure 1. Model for initiation of mismatch repair. Mispaired bases are recognized byMutS protein, acting as a homodimeric complex. MutH protein, the site specificendonuclease, is likely recruited by the "matchmaker" MutL protein. MutH then createsa nick in the unmethylated strand at the hemimethylated d(GATC) site.

Page 27: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

12

DNA strand. This process is believed to be mediated by the "matchmaker" protein

MutL. A mismatch-bound MutS homodimer associates with a MutL homodimer

(Grilley et al., 1989), and through the hydrolysis of ATP to ADP + P1, possibly

undergoes a translocation/search process (depicted in Figure 1 as producing an a-

shaped structure) to activate the site-specific endonuclease MutH protein to make the

excision-initiating nick in the unmethylated strand at the nearest hemimethylated

d(GATC) sequence, in either the 3' or 5' direction. (see below; Figure 1). Because

there is delay in methylation of newly replicated d(GATC) sequences by the E. coli

DNA-adenine methylase (product of the dam gene), mismatch-provoked nicking by

Mutil of the nascent strand provides essential strand specificity. Once the nick has

been made, the excision reaction can take place.

Excision/Resynthesis. The current model for the excision/resynthesis steps of

methyl-directed mismatch repair is shown in Figure 2. The excision/resynthesis

process requires DNA helicase II, SSB, exonuclease (Exol, ExoVil, or RecJ), DNA

polymerase III holoenzyme, dNTPs, and DNA ligase/NAD. To initiate the excision

process, DNA helicase II unwinds the DNA with the aid of SSB, allowing the single-

stranded DNA-specific exonucleases (RecJ or ExoVil if the nick is 5' to the mispair,

or Exol if the nick is 3' to the mispair) to remove the DNA patch just beyond the

Page 28: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

13

CH CH,3 I I

5 3'

InitiationMutS, MutL, MutH

CH CH CH, CHI- I I- I.

ExcisionExo IExo VII

MutS, MutL,or RecJHelicase II, AT?

CH,CH CHcu-I3 I-

Resynthesis

JrJr

DNA p01111 boIoeyme,SSB

CH3 CH, Cl-I, CH3I- I- I' I

Figure 2. Model for excision/resynthesis steps following mismatch repair initiation.The nick created during mismatch repair initiation provides the signal for excisionsteps involving either ExoVil, RecJ (5' to mispair) or Exol (3' to mispair) proteins.Excision is followed by resynthesis by polymerase III, and subsequent ligation.

Page 29: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

14

mispair, leaving a gap up to several kilobases (kbs) in length. Finally, DNA

polymerase III holoenzyme fills in the excision gap, followed by the covalent closing

of the remaining nick by DNA ligase.

1.2 Long-Patch Mismatch Repair in Eukaryotes

As mentioned above, one of the first observations of mismatch repair was in

eukaryotic cells, when Holliday (1964) described the process of gene conversion in

fungi. Since then, it has become apparent that eukaryotes carry out very similar

mismatch-repair pathways observed in many eubacteria. Some of the first direct

evidence for eukaryotic mismatch repair came from in vivo experiments similar to that

described for E. coli. But it was through the highly conserved structure and function

of the mismatch repair pathway, and in particular the MutS and MutL protein

sequences, that the study of eukaryotic mismatch repair has progressed so rapidly.

1.2.1 Evidence for Eukaryotic Mismatch Repair

In addition to analysis of the post-meiotic segregation observed in

Saccharornyces cerevisiae and other fungi, transformation and transfection studies

with S. cerevisiae and monkey kidney cells (respectively) were first used to

demonstrate mismatch repair in eukaryotes (Bishop et al., 1989; Brown and Jiricny,

Page 30: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

15

1988). In these studies, the efficiencies of repair of the various heteroduplexes were

very similar to those described for E. coli, suggestive of similar pathways in

prokaryotic and eukaryotic cells. In vitro assay experiments using (human) and

Drosophila melanogaster K cell line extracts (Holmes, et al., 1990) further revealed

that base-base mispairs in nicked, open circular DNA heteroduplexes, were repaired

with similar efficiencies to experiments in E. coli, with correction highly biased to the

nicked strand. Excision/resynthesis was localized to the region between the mispair

and the nick, demonstrating that mismatch recognition was associated with the repair

reaction. Another critical study revealed that eukaryotic (human) mismatch repair,

like the mismatch-repair reaction in E. coli, occurred in a bi-directional manner, i.e.

repairing mispairs located either 5' or 3' to the nick (Fang and Modrich, 1993). But

because eukaryotic cells lack d(GATC) methylation, it remained unclear what the

strand discrimination mechanism was.

1.2.2 Identification of MutS and MutL Homologs in Eukaryotes

Both in vivo and in vitro evidence suggested that analogous mismatch repair

pathways existed in both prokaryotes and eukaryotes, but it was not clear whether the

proteins involved were evolutionarily conserved. A first clue was the isolation of

prns] (for post-meiotic segregation 1) mutants in S. cerevisiae. This mutant class

Page 31: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

16

displayed an increased post-meiotic segregation phenotype, and was also found to be

defective in mismatch repair (Kramer et al., 1989a). Cloning and sequencing of the

gene proved that it encoded a ,nutL-like protein (Kramer et al., 1989b). Additionally,

a mutS-like gene was identified in the dhfr region of the human genome (Fujii and

Shimada, 1989). Based on these findings, several other laboratories set out to identify

other mutS- and inutL-like genes based on the highly conserved amino-acid sequence

found in prokaryotic MutS and MutL protein sequences, primarily by employing a

degenerate oligonucleotide primers in polymerase chain reaction (PCR) amplification

of S. cerevisiae and human cDNA. Initially, two mutS-like genes were identified in S.

cerevisiae, called MSH1 and MSH2 (for MutS Homolog) (Reenan and Kolodner,

1992), as well as a mutL-like gene, called MLHJ (for MutL Homolog) (Prolla et al.,

1994; Bronner et al, 1994). This technique, as well as searches for additional

homologs in the completely sequences S. cerevisiae genome, revealed at least six

highly conserved MSH genes (MSHJ, MSH2, MSH3, MSH4, MSH5, and MSH6), and

four highly conserved MLH/PMS genes (MLH1, MLH2, MLH3, and PMS1) in S.

cerevisiae. Several other highly conserved genes, all of which fall into the MSH and

MLH/PMS classes found in S. cerevisiae, have been identified in a variety of animals,

including the MSH2, MSH3, MSH4, MSH5, and MSH6 genes found in the human

genome. This multiplicity of mutS and mutL-like genes found in eukaryotes was

Page 32: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

17

somewhat unexpected, since only the single MutS and MutL proteins are required for

the initial stages of mismatch repair in prokaryotes. This suggests functionally

specialized roles for eukaryotic MSH and MLHJPMS proteins.

1.2.3 MSH and MLH Protein Families in Eukaryotes

The MSH Protein Family. All MSH proteins show high similarity to

prokaryotic MutS sequences. Three regions common to almost all MutS and MSH

proteins, are found in the N-terminal, middle, and C-terminal portions of the proteins.

Of these, the C-terminal region is most highly conserved. It contains the Walker-type

ATP binding domain and the helix-turn-helix domain, present in all known MutS and

MSH protein sequences. The MSH proteins generally range in size from about 95

kDa up to about 165 kDa. Both MSH and MLH (PMS) protein functions are

summarized in Table 1.

MSH1. The S. cerevisiae MSHJ gene encodes a protein required for

mitochondnal stability. Strains harboring a disrupted MSH1 gene displayed a "petite"

phenotype, indicative of yeast mitochondrial instability (Reenan and Kolodner,

1992b). In addition, the predicted MSH1 protein sequence revealed an N-terminal

mitochondrial signal peptide, suggesting that this protein is targeted to mitochondria.

The S. cerevisiae MSH1 protein binds mismatched oligonucleotide substrates, but may

Page 33: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

EU

Table 1. Eukaryotic mismatch repair proteins

Protein Present In Inferred Function(s)

MutS Homologs

MSH1 Yeast, octocoral, plants Mitochondrial errorcorrection

MSH2 Yeast, plants, animals Mitotic, meiotic errorcorrection

MSH3 Yeast, plants, animals Mitotic, meiotic errorcorrection (largeinsertion-deletionloopouts, or IDLs)

MSH4 Yeast, plants(?), animals Promotion of meiosis

MSH5 Yeast, animals Promotion of meiosis

MSH6 Yeast, plants, animals Mitotic, meiotic errorcorrection (base-basemispairs, small IDLs)

MSH7 Plants (see below) Specialized nuclearfunction for base-basemispairs? Other?(see below)

MutL Homolos

MLH1 Yeast, plants, animals Mitotic, meiotic errorcorrection, promotion ofmeiosis

MILH2 Yeast, plants(?), animals Specialized role in errorcorrection?

MLH3 Yeast, plants(?), animals Mitotic, meiotic error

correction, large IDLs,

promotion of meiosis

PMS2 (yPMSI) Yeast, plants, animals Mitotic, meiotic error

correction.

Page 34: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

19

act via a different mechanism than do other MutS and MSH proteins (Chi and

Kolodner, 1994a; Chi and Kolodner, 1994b). Although no MSH1 has yet been

reported for higher animal genomes, an MSH1-Iike gene has been identified in the

octocoral Sarcophyturn glaucum, and in Arabidopsis thaliana. Interestingly, the S.

cerevisiae and Arabidopsis thaliana MSHJ genes are encoded by their respective

nuclear genomes, while the S. glaucum MSHJ gene is encoded by the mitochondrial

genome (Reenan and Kolodner, 1992; Pont-Kingdon et al., 1995).

MSH2. As suggested by the model for eukaryotic mismatch repair, the MSH2

protein plays a central role in nuclear mismatch recognition function, as a member of

two heterodimeric complexes (see below). An MSH2 gene was first described in S.

cerevisiae (Reenan and Kolodner, 1992a), and strains defective in MSH2 function

showed a mutator phenotype, with high levels of microsatillite instability (Reenan and

Kolodner, 1992b). At about the same time, an MSH2 -like gene was identified in the

human genome (Leach et al., 1993; Fishel et al., 1993). Interestingly, it was because

of the common mutator phenotypes (microsatillite instability) shown by E. coli and S.

cerevisiae cells defective in mismatch repair, and by HNPCC (Hereditary Non-

Polyposis Colorectal Cancer) tumor cell lines, that initially led investigators to search

for mismatch repair genes in humans (discussed below). Leach et al. (1993) utilized a

map-based approach to identify a gene (hMSH2) linked to a known HNPCC locus,

Page 35: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

20

while Fishel et al. (1993) first identified hMSH2 through the use of degenerate based

PCR primer amplification, and then showed that the gene mapped to the same known

HNPCC locus. In each case, the hMSH2 gene contained an apparent loss-of-function

mutation, suggesting a link between functional mismatch repair and a predisposition to

HNPCC. Targeted mutations in mouse homologs of MSH2 provided additional

information about the activity of MSH2 in mismatch repair, and its role in cancer

susceptibility. Msh2' mice have high levels of microsatellite instability in somatic

tissues, and are predisposed to increased levels of colon cancer (de Wind et al., 1995;

Reitmair et al., 1995). Human tumor cell lines defective in hMSH2 also show high

levels of microsatellite instability.

MSH3 and MSH6 - Functional Partners of MSH2. hMSH2, a 105 kD

polypeptide, was purified from human HeLa cells as a functional complex with

hMSH6, a 160 kD polypeptide also known as GTBP (Drummond et al., 1995;

Papadopoulos et al., 1995; Palombo et al., 1995), suggesting that hMSH2 and hMSH6

function as part of a protein complex (termed MutSct). In S. cerevisiae, genetic

characterization of the yMSH3 and yMSH6 genes suggested a functional overlap in

MSH2-dependent mismatch repair (Marsischky et al., 1996). Mutations in the yMSH3

gene led to insertions and deletions in repetitive DNA sequences, while mutations in

yMSH6 led to point mutations and single-base insertions and deletions, showing

Page 36: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

21

overlapping specificity for small IDLs (Strand et al., 1995; Marsischky et al., 1996).

While both the yMSH3 and yMSH6 mutant strains displayed a mutator phenotype,

each was significantly less so than yMSH2 mutant strains (the yMSH6 mutant had a

slightly higher mutation rate than yMSH3 mutants). Furthermore, combination of the

yMSH3 and yMSH6 mutations led to a phenotype similar to both a yMSH2 single

mutant and a yMSH2/MSH3/MSH6 triple mutant, suggesting that MSH3 and MSH6

act in separate MSH2-dependent pathways. Protein-interaction studies also revealed

that yMSH2 can interact with either yMSH3 or yMSH6, but that yMSH3 and yMSH6

do not interact with each other (Marsischky et al., 1996). Msh2', Msh6' and Msh3

transgeneic mice show analogous mutation spectra to S. cerevisiae, with Msh2 mice

having the most severe phenotype that includes a predisposition to certain types of

cancer (Edelmann et. al., 1995; Edelmann et al., 2000).

The hypothesis that MSH3 and MSH6 act in separate MSH2-dependent

pathways was further supported by the purification of two heterodimeric complexes in

human cells, MutSci (MSH2.MSH6), and an additional complex, termed MutSf3,

comprised of hMSH2 and a unique 130 kD protein, determined to be a homolog of the

yMSH3 (Drummond et al., 1997)). No additional complexes (either homopolymeric,

such as MSH2.MSH2, or heteropolymeric, such as MSH3.MSH6) were identified.

The biochemical and genetic data in yeast, mice, and human cells thus suggest that the

Page 37: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

22

MutSc (MSH2.MSH6) and MutSI3 (MSH2.MSH3) heterodimers act as the primary

complexes in eukaryotic cells for the initiation of nuclear mismatch repair.

Electrophoretic-mobility-shift assays employing both yeast and human

heterodimers have shown that MSH2.MSH6 and MSH2.MSH3 bind to a wide range

of mismatched substrates, yet their functions are partially redundant (Acharya et al.,

1997; Palombo et a!, 1995), consistent with the repair efficiencies and observed

mutation spectra (described above) in S. cerevisiae and mouse mutant strains. The

MSH2.MSH6 heterodimer binds preferentially to base/base mispairs and small

insertion/deletion loopouts (IDLs), while the MSH2.MSH3 heterodimer binds

preferentially to IDL structures greater than about 2 basepairs (Habraken et al., 1996).

In addition, biochemical and genetic studies suggest recognition of a variety of DNA

lesions, including mismatches opposite UV photoproducts (Mu et al., 1997; Wang et

al., 1999) and oxidative lesions such as 8-oxo-guanine (Leadon and Avrutskaya,

1998), by MSH2.MSH6 heterodimers (and MutS homodimers), thus mismatch repair

may play a role in reducing the mutagenic potential of misincorporations by

replicative DNA polymerases opposite DNA lesions.

MSH4 and MSH5. The MSH4 and MSH5 proteins are meiosis-specific

proteins and, although homologs of MutS, appear not to be involved in mismatch

repair. S. cerevisiae mutations in the MSH4 and MSH5 genes reduce meiotic crossing-

Page 38: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

23

over 1.4- to 3.4-fold and 1.9- to 4.0-fold, respectively (Ross-Macdonald and Roeder,

1994; Hollingsworth et al., 1995). Analysis of double mutant MSH41MSH5 strains,

and MLHJYMSH4 strains, suggest that the MSH4, MSH5, and MLH1 (see below)

proteins function in the same epistasis group to promote crossing over (Hollingsworth

et al., 1995, Hunter and Borts, 1997). MSH4 and MSH5 proteins interact to form a

heterodimeric complex (Pochart et al., 1997), similarly to MSH2.MSH6 and

MSH2.MSH3, but do not recognize mismatched bases or IDL structures. Instead, the

MSH4.MSH5 complex has been hypothesized to participate in the promotion of

meiotic recombination, possibly by stabilizing recombination intermediates, and

ensuring proper chromosome disjunction (Hunter and Borts, 1997).

MLH/PMS Protein Family. Like the MSH protein family, MLH/PMS

protein members show high similarity to their prokaryotic MutL counterparts.

MLHJPMS proteins show the highest conservation in their N-terminal regions. These

proteins range in size from about 60 kD to 100 lcD. Recently, the molecular structure

for E. coli MutL has been solved by X-ray crystallography, and shows structural

similarity to DNA gyrase (Ban and Yang, 1998). The protein was also found to

possess a weak ATPase activity (Ban and Yang, 1998). Due to their high

conservation, eukaryotic MILR/PMS proteins are likely to be similar in structure and

activity.

Page 39: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

24

MLH1. Similarly to MSH heterodimers, MLH proteins form at least three

heterodimer complexes in S. cerevisiae (see below), and there are possibly more in

mice and humans. Like MSH2, MLHI plays a central role in MILH function. Because

of evidence for mutL-like genes present in yeast (pmsl), homology based searches led

to the identification of MLHJ in S. cerevisiae (Prolla et al., 1994), and mammalian

cells (Bronner et al., 1994; Papadopoulos et al., 1994). Due to the stunning

connection between MSH2 mutations and microsatillite instability seen in HNPCC

tumor cell lines, MutL homologs were logical candidates for other known HNPCC-

linked genes. Bronner et al. (1994) and Papadopoulos et al. (1994) isolated the

hMLH1 gene and showed that it mapped to a HNPCC locus on chromosome 3. Loss-

of-function mutations were identified in affected individuals from a chromosome 3-

linked HNPCC family. This, in conjunction with the isolation of the MSH2-linked

HNPCC locus, further suggested that defects in mismatch repair proteins lead to a

predisposition for certain cancers, including colon cancer. Similarly to Msh2 mice,

Mlh1 mice have high levels of microsatellite instability in somatic tissues, and are

predisposed to increased levels of colon cancer (Baker et al., 1996; Prolla et al., 1998;

Edelmann et al., 1996). S. cerevisiae cells and human tumor cell lines defective in

hMLH1 also show high levels of microsatellite instability (Prolla et al., 1994).

Page 40: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

25

PMS2 (yPMS1). Studies in S. cerevisiae suggested that yMLHI interacts with

yPMSI to form a heterodimer (Prolla et al., 1994). In human cells, an activity of

HeLa cell lines that restored mismatch repair to a MLH1-defecient cell extract proved

to be a complex of hMLH1 and hPMS2 (Li and Modrich, 1995). [For clarity, hPMS2

is considered to be a ortholog to yPMS], based on the degree of similarity between

them, compared to other MutL homologs.] These data suggested that MLH proteins,

like MSH proteins, form specific heterodimeric complexes. Recent studies in S.

cerevisiae have shown that MLH1 interacts with PMS1 and with two recently

described proteins MLH2, and MLH3 (see below), whose genes were identified in the

S. cerevisiae genome based on homology to other MLH proteins (Wang et al., 1999).

In S. cerevisiae, both mihi and pmsl mutations cause strong mutator phenotypes and

increased microsatillite instability, similarly to mutations in msh2 (Prolla et al., 1994).

Likewise, mutations in MLH1 and PMS2 (yPMS1) in knockout mice produce

comparable degrees of mutability. Similarly to Mlh1, Pms2' mice and human cell

lines show strong microsatillite instability, and are mismatch repair defective.

However, mutations in PMS2 are rarely seen in HNPCC families, and Pms2' mice

show reduced susceptibility to colon cancer (Parsons et al., 1993; Risinger et al.,

1995; Li and Modrich 1995; Boyer et al., 1995). These data suggested redundancy in

Page 41: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

MLH/PMS function, similar to that seen for MSH pathways, but PMS2 (yPMS1) may

play a more important role in mismatch repair.

MLH2. The S. cerevisiae MLH2 (mammalian PMSI) is not believed to be

involved in mismatch repair, although the protein interacts with MLH1 (Floras-Rozas

and Kolodner, 1998; Wang et al., 1999). Pms1 mice show attenuated microsatellite

instability, but exhibit no susceptibility to colon cancer (Prolla et al., 1998; Yao et al.,

1999). It is still unclear what role yMLH2 (PMS1) has in DNA repair or

recombination pathways, if any.

MLH3. The recently described MLH3 protein has been suggested to

participate in the repair of frameshift mutations in S. cerevisiae, in an MSH2/MSH3-

dependent pathway (Floras-Rozas and Kolodner, 1998). Another study suggests

MLH3 participates in promoting meiotic recombination (possibly as a heterodimer

with MILH1, and/or in association with MSH4.MSH5) in S. cerevisiae (Wang et al.,

1999). An MLH3 gene has been identified in mammals, and has been shown to be

associated with the Cssl cancer susceptibility locus in mice (Lipkin et al., 2000; see

Table 1). However, more studies are needed to determine what role MILH3 plays in

mismatch repair functions in yeast and mammals.

Page 42: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

27

1.2.4 A Generalized Model for Mismatch Repair in Eukaryotes

The initiation of mismatch repair in eukaryotes parallels the initiation reaction

of the methyl-directed mismatch-repair pathway in E. coli. However, the function of

the MutS homodimer is replaced by at least two distinct heterodimers, MSH2.MSH6

(MutSc) and MSH2.MSH3 (MutS3), each with specific mismatch recognition

functions (summarized in Figure 3). Both heterodimers require cooperative

interactions with a MLHIPMS heterodimer (instead of a homodimer of MutL in E.

coli). The primary MLHIPMS heterodimer required for the repair of mismatched

bases, whether in cooperation with MSH2.MSH6 or MSH2.MSH3, is MLH1.PMS2

(yPMS1). Other heterodimers, such as MLH1.MILH3 may participate in related

pathways, such as the MSH2.MSH3-dependent repair of larger IDLs. MSHIMLH

complexes, when activated by mismatched substrates, likely involve interactions with

the replication machinery. In fact, both MLH1 and MSH2 were shown to interact with

PCNA (Proliferating Cell Nuclear Antigen) which is known to be a required

component for DNA replication (Umar et al., 1996). These interactions, involving

ATP-dependent conformational changes of MSH heterodimers and subsequent

translocation/search processes (see below), may invoke signals to a strand

discrimination mechanism (currently unknown) at the replication fork. It has been

postulated that a signal for the newly replicated DNA strand (and hence the target for a

Page 43: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

MLH1 PN1S2 MLIII Pi\1S2-4--' : Fi I Fi : fr -' Ii I

G

T

Base-base mispairs(G/T for example)

T

Single-base insertion-deletion loopouts

(T IDL for example)

_AflG_

Larger insertion-deletion loopouts

(+AAG IDL forexample)

Figure 3. Eukaryotic model for mismatch repair. MSH2 protein forms aheterodimer with either MSH3 or MSH6 for mispair specific recognition.MSH2MSH6 heterodimers recognize base/base mismatches and small insertion-deletion loopouts (IDLs), while MSH2MSH3 heterodimers recognize both small andlarge IDLs.

Page 44: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

29

mismatch activated excision reaction) could simply be the 3' end of the growing DNA

strand at the replication fork. In this model, mismatch repair signals could pause DNA

polymerase (Longley et al., 1997), load and/or reassemble required components to

begin the excision reaction 5' or 3' towards the mispaired base(s). One likely

component, in addition to PCNA, is EXO1 (Exonucleasel), a 5' to 3' exonuclease

found to interact with MSH2 in S. cerevisiae. Mutants deficient in exol display a

mutator phenotype, and epistasis analysis suggested EXO1 functions in an MSH2-

dependent pathway (Tishkoff et al., 1997).

At least two independent models have been proposed for the initiation of

mismatch repair. The first, proposed by Modrich and co-workers, involves an ATP-

dependent translocation/search mechanism by MutS/MSH proteins (Allen et al., 1997;

Blackwell et al., 1998a; Blackwell et al., 1998b). This model suggests that ATP

hydrolysis "drives" the protein complex along the DNA helix, allowing the coupling

of mismatch recognition to loading of the excision system at the strand break that

directs repair. An alternative model has been proposed by Fishel and co-workers

(Gradia et al., 1997; Fishel 1998; Gradia et al., 1999; Gradia et al., 2000). This model

suggests that MutSa functions similarly to G-proteins (for example, Ras), such that

the particular bound nucleotide is coupled to a switch between "on" and "off' states.

An "off' state for MutSc* is an ATP-bound form that essentially has no (or very low)

Page 45: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

30

affinity for heteroduplex DNA, while an "on" state is an ADP-bound form that binds

to heteroduplex DNA with strong affinity. MutSu.ADP, upon binding a mismatch,

rapidly exchanges ADP for ATP, releases from the mismatch, and then recruits

downstream activities to the site by a "sliding clamp" mechanism. The "sliding

clamp" is postulated to resemble a circular, or donut-shaped molecule, similar to

proteins such as PCNA, that can freely diffuse along the DNA helix (Gradia et al.,

2000).

1.2.5 Mismatch Repair and Recombination

Mismatch-repair pathways have several recognized roles known to both

positively and negatively affect recombination. MSH4, MSH5, and MILH1 have been

shown to promote crossing-over essential for the formation of synaptonemal

complexes that ensure proper chromosome disjunction (discussed in section 1.2.3).

Mismatch-repair systems also preserve genetic fidelity by antagonizing genetic

recombination between imperfectly-homologous DNA sequences on the same or

different chromosomes (or on exogenous DNA fragments). This is exemplified in

inter-specific crosses of bacterial (Rayssiguier et al., 1989) and yeast cells (Hunter et

al., 1996). For example, yeast cells deficient in mismatch repair show an increase in

fertility in inter-specific crosses. Other studies in S. cerevisiae have shown an increase

Page 46: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

31

in crossing-over between partially diverged loci in mismatch repair deficient strains

over wild-type strains (Petit et al., 1991; Selva et al., 1995). The strong positive

effects of mismatch-repair deficiency on genetic exchange, on meiotic recombination

involving partially diverged chromosomes, and on recombination between imperfectly

repeated sequences in the same genome, suggest that mismatch-repair systems help

maintain inter-species barriers and protect chromosomes against rearrangements

(Rayssiguier et al., 1989).

1.2.6 Why Study Mismatch Repair in Plants?

In contrast to multicellular animals, plants lack a reserved germ line; their

gametes are formed, late in their growth cycles, by differentiation of somatic

meristematic cells. Typically, the somatic precursors of gametophytes have divided

many times, potentially subjecting their genomes to multiple rounds of spontaneous or

environmentally-induced mutagenesis (Walbot, 1985). However, plants do not seem

to show extraordinarily high mutation rates. For example, long-lived trees presumably

produce gametes from somatic cells that are themselves the products of many annual

cycles of mitotic growth. Nevertheless, their mutation rates per zygote-to-meiosis

generation, averaged over long reproductive lives, are only an order of magnitude or

so higher than those of annuals (Klekowski, 1997).

Page 47: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

32

How do plants combat the threats to genomic stability posed by somatic

mutation? Mechanisms of selection against less fit somatic cells during growth and

development have been reviewed extensively by Klekowski (1988). Diploid

meristematic cells that acquire even partially dominant deleterious mutations may

drop out of the actively dividing pool so-called "diplontic selection." Additional

sieving out of recessive mutations may occur when haploid cells compete to form

sperm or eggs. However, Klekowski has suggested that these mechanisms alone are

not powerful enough to protect plants against rapid accumulation of extraordinary

mutational loads (Klekowski, 1988; Klekowski, 1997). Genomic-fidelity functions

must therefore be at least as efficient in plants as those of microbes and animals, if not

more so. Since mismatch repair systems are highly conserved in bacteria, yeast and

animals, it is likely that plants utilize similar mismatch repair functions to maintain

genomic fidelity.

Specialized mismatch-repair functions such as positive and negative effects on

genetic recombination, may play important roles in plants. For example, the strong

positive effects of mismatch repair deficiency (discussed above) on meiotic

recombination may have implications for the fertility of hybrids between diverged

plant species and for targeted alteration of plant genes by homologous recombination.

Page 48: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

33

Isolation of plants deficient in mismatch repair could provide a means to test this

hypothesis.

Moreover, accumulating evidence for recognition of some UV photoproducts

and chemical adducts (bases damaged by oxyradicals: 8-oxoguanine, for example) in

DNA, by both bacterial (Feng et al., 1991) and human mismatch-repair proteins (Mu

et al., 1997; Wang et al., 1999; Leadon and Avruskaya, 1998), suggests that mismatch

repair may antagonize mutagenic and/or cytotoxic effects of such damages. This latter

function of mismatch repair may thus avert some genotoxic consequences of the

inevitable exposure of plants to solar UV-B radiation and environmental insults.

Therefore, my long-range goal was to determine the extent to which MutSL-

like mismatch-repair systems maintain the genetic integrity of plant genomes. In order

to do so, I focused on three primary questions: i) Do plants encode homologs of

MutS proteins? ii) Do these homologs fall into the same distinct subfamilies seen in

other eukaryotes? iii) Do these homologs share conserved biochemical activities with

their eukaryotic counterparts, including DNA substrate specificities? I have focused

on the small crucifer Arabidopsis thaliana, a green plant that provides several

advantages. For example, its small size (typically about 30 centimeters), short

generation time (six weeks), high seed set (about 10,000 seeds per plant) and ease of

mutagenesis, have made it an ideal genetic model system. Its small, yet simple

Page 49: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

34

genome has allowed the creation of detailed genetic maps, and consequently has been

chosen as the first plant genome to be completely sequenced. As a result, it is

cunently the primary focus of basic research in the plant sciences. These are obvious

advantages for the isolation and characterization of plant genes involved in mismatch

repair, and Arabidopsis should continue to be a useful model system for the study of

plant DNA repair pathways.

Page 50: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

35

Chapter 2

DNA Mismatch Repair in Plants: An Arabidopsis thaliana Gene That Predicts aProtein Belonging to the MSH2 Subfamily of Eukaryotic MutS Homologs

Kevin M. Culligan and John B. Hays

Published, in part, in Plant PhsyiologyAmerican Society of Plant Physiologists

October, 1997

Page 51: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

36

2.1 Abstract

Degenerate sets of oligomers, corresponding to highly conserved domains of

MutS-homolog (MSH) mismatch-repair proteins, were used to prime PCR

amplifications of two Arabidopsis thaliana DNA fragments found to be homologous

to eukaryotic MSH-like genes. Phylogenetic analysis of one complete gene,

designated atMSH2 , places it in the evolutionarily distinct MSH2 subfamily.

2.2 Introduction

Our long-range goal is to determine the extent to which MutSL-like mismatch-

repair systems maintain the genetic integrity of plant genomes. Here we addressed an

initial question: do plants encode homologs of MutS proteins, and if so, do these

homologs fall into the same distinct MutS-homolog subfamilies seen in other

eukaryotes? We described below the isolation of two Arabidopsis gene fragments

encoding MutS-like proteins, using polymerase-chain-reaction (PCR) amplification of

DNA or RNA with degenerate primer sets based on evolutionary-conserved MutS

amino acids. Phylogenetic analysis of the complete sequence of one gene suggests

that it is an MSH2 homolog.

Page 52: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

37

2.3 Results

2.3.1 Isolation and Initial Characterization of atMSH2, and a Gene

Fragment Similar to MSH6-like Proteins

We utilized two sets of degenerate polymerase-chain-reaction (PCR) primers,

corresponding to all possible codons for three domains of 5-6 amino acids each that

are highly conserved among eukaryotic and prokaryotic MutS-like proteins (Figure 4),

to amplify and clone segments of Arabidopsis thaliana DNA. We used

aseptically-grown seedlings, to avoid false positives from amplification of

homologous sequences encoded by organisms that might contaminate whole plants. A

0.7-kb fragment, consistently observed under a variety of PCR reaction conditions

using primers 1 and 3, was purified and itself PCR-amplified using primers 1 and 2

(Figure 4), yielding a 0.4-kb "nested" product. Similarly, cellular cDNA templates

[obtained by reverse-transcriptase (RT)-PCR] consistently yielded a 0.35-kb fragment,

which when purified and amplified with primers 1 and 2 yielded a 0.26-kb fragment,

as predicted (Figure 4).

The 0.7-kb and 0.35-kb DNA fragments proved to encode the same MutS-like

protein; hybridization analysis demonstrated them to be fragments of an Arabidopsis

thaliana gene (see below). Probing an Arabidopsis X phage cDNA library with the

Page 53: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique
Page 54: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique
Page 55: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

0.7-kb fragment yielded, as well as shorter cDNAs, a 2-kb fragment which was used in

turn to obtain a 3.1-kb cDNA from a second library. Amplification of the 5'-most

portion of the corresponding RNA, by 5-rapid amplification of cDNA ends (5'-

RACE), using a gene specific primer site approximately 1.2 kb 5' to the poly-A tail in

the 2-kb cDNA, yielded a single product of approximately 1.9-kb, again indicating the

complete cDNA length to be 3.1 kb in length.

The longest open reading frame identified, 2814-bp in length, was anticipated

to encode a protein of 937 amino acids, similar in size to other MutS homologs. The

predicted amino acid sequence showed strong similarities with the MSH2 class of

mismatch repair proteins (Figure 4); it is 37% identical to both yeast and human

MSH2 proteins. We designate this gene atMSH2.

The fact that only single Arabidopsis DNA restriction fragments hybridized to

the 0.7-kb probe (Figure 5) argues against a family of atMSH genes with similar DNA

sequences. We searched for other Arabidopsis MutS homologs among the population

of 0.26-kb and 0.35-kb DNA products obtained using degenerate primer sets 1,2 and

1,3 respectively. Ten randomly selected 0.35-kb clones proved to be atMSH2

fragments. Of ten randomly selected 0.26-kb clones, seven were atMSH2 fragments,

one encoded a protein fragment strongly resembling eukaryotic MSH6 proteins, and

two encoded a protein fragment more similar to yeast MSH2 (data not shown). The

Page 56: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

41

l2Kb-4

5 Kb *

3Kbp

1.5 Kb

atMSH2-Iike atMSH6-Iikeprobe probe

ES

c) L) L)

S

c)

Figure 5. Southern analysis of genomic DNA from Arabidopsis thaliana. A 32P-

labeled fragment (corresponding to genomic positions +2871 to +3518 ofatMSH2, left panel, or the 270 bp PCR fragment of the MSH6-like gene, rightpanel) was used to probe genomic DNA (ecotypes Columbia and Wassilewskija)digested with EcoRl or BamHI restriction endonucleases.

Page 57: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

42

latter fragment did not hybridize to Arabidopsis DNA, and may have been amplified

from contaminating fungal/microbial DNA or RNA. However, the MSH6-Iike DNA

fragment hybridized strongly to a single restriction fragment of Arabidopsis DNA,

under the same conditions that yielded a single (dissimilar) band hybridizing to the

atMSH2 probe (Figure 5). These observations are consistent with the pattern of

functionally specialized, evolutionary diverged MutS homologs observed in other

eukaryotes (Modrich et al., 1996).

Figure 6 schematically depicts the atMSH2 gene structure deduced from the

complete genomic sequence. The 12 introns in the coding sequence range in size from

80-hp to 230-hp; all introns show consensus GT/AG splice signals. A plant

consensus polyadenylation signal is present in the 3' untranslated region (Mogen et al.,

1990).

To determine the phylogenetic relationships of atMSH2 to other MSH-Iike

genes, we compared its predicted amino-acid sequence to the sequences of all known

eukaryotic MSH2, MSH3, MSH6 proteins, and to those of bacterial MutS proteins.

Figure 7 shows the resultant distance tree, and the bootstrap values for the distance

(top) and parsimony (bottom) consensus trees. Although atMSH2 branched with all

other MSH2 sequences with bootstrap confidence value of 100 parsimony trees, it

could not be unambiguously placed within this sub-group. For example, there is a low

Page 58: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

LTG TGA

-145 1 500 1000 1500 2000 2500 3000 3500 4000 4402

Figure 6. Structure of the atMSH2 gene. Comparison of genomic and cDNAsequences revealed 12 introns (shaded boxes) throughout the longest reading frame

of atMSH2 [denoted as genomic position +1 (ATG) to +4259 (TGA) of genomicsequence]. The dark box 3' to position +4402 denotes poly-A sequences observedin cDNAs. The genomic sequence has been deposited as EMBL accession number12345.

Page 59: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

ririi

MSH2 (X. Iaevis)100

100 MSH2 (H. sapiens)

MSH2 (M. musculus)

SPE1 (D. melanogaster)100

MSH2 (A. thaliana)

MSH2 (S. cerevisiae)

MSH6 (H. sapiens)

100100

MSH6 (M. musculus)100

MSH6 (S. cerevisiae)

4 (5. pombe)

MSH3 (S. cerevisiae)

MSH3 (H. sapiens)

REP3 (M. musculus)

I

MutS (A. vinlandii)

100100 MutS (S. typhimurium)

ioo 100 L MutS (E. coD)96

100

MutS (H. influenzae)

HexA (S. pneumoniae)

0.10

Figure 7. Phylogenetic analysis of MutS homologs. CLUSTAL sequencecomparisons were analyzed using two phylogenetic methods to create two distincttrees. A neighbor-joining distance tree was constructed with the Phylip distancemethod using the Dayhoff PAM matrix, and a protein parsimony tree wasconstructed using Phylip 3.5. Both methods were masked to exclude sequencegaps, and bootstrapped (100 replicates). A representative distance tree is shownabove; both trees showed very similar branching patterns. Bootstrap values(significant, above 60) are shown above and below tree branches for distance andparsimony consensus trees, respectively. Both methods used MSH1 (excludingthe mitochondrial targeting sequence, i.e. the first 21 amino acids) as an outgroup.

All analyses were performed using Genetic Data Environment (GDE).

Page 60: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

bootstrap confidence value for the S. cerevisiae MSH2 and A. thaliana MSH2 node.

Preliminary analysis of the putative atMSH6 fragment places it in the MSH6

subgroup. Furthermore, in both the distance and parsimony trees, the MSH3 and

MSH6 subgroups branched together, with a bootstrap confidence value of 90 for the

distance tree. MSH3 and MSH6 thus appear to have diverged from a common

ancestral protein (itself distinct from MSH2), consistent with the functional overlap

between the two proteins (Marsischky et al., 1996). We believe this to be the first

rigorous phylogenetic analysis of MSH2, MSH3, MSH6 and bacterial MutS

sequences.

2.3.2 Overexpression of atMSH2 in E. coli Cells

In order to determine the functional properties of atMSH2 protein, the cDNA

corresponding to the entire open reading frame of atMSH2 was inserted into a E coli

expression vector pQE7O, downstream of the T5 promoter (plasmid pQE7O::atMSH2),

and was confirmed by sequencing the vector/insertion junctions. This construct

allows induced expression in E. coli cells of atMSH2 protein in the presence of IPTG

(0.5 mM isopropyl-3-D-thiogalactopyranoside). The accumulation of atMSH2 protein

was monitored using polyacrylamide-gel electrophoresis in sodium dodecyl sulfate

(SDS-PAGE) of E. coli cell extracts, by comparing uninduced versus induced cell

Page 61: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

cultures, and looking for induced protein bands corresponding to approximately 100

kDa. However, no induction was apparent under the conditions tested, in both the

soluble and insoluble fractions, at any molecular weight visible on SDS-PAGE,

regardless of the amount andlor length of IPTG induction. This suggests that either

the expression of atMSH2 is unstable in E. coli cells (possibly being proteolytically

degraded), or that expression from the pQE7O vector is not sufficiently strong enough

to observe induction on SDS-PAGE using whole cell extracts.

We then modified the atMSH2 cDNA in pQE7O to contain a C-terminal 6X

HTS-tag (plasmid pQE7O::atMSH2his). This allowed the partial purification of the

expressed protein from E. coli extracts, and provided an opportunity to more directly

monitor expression of atMSH2. These experiments revealed that a small amount of

protein, approximately 100 kDa in size, was induced in the presence of IPTG.

However, several other protein bands of smaller molecular weight (size range 66 to 20

kDa) that co-purified with the 100 kDa protein on Ni2 affinity resin were also

induced, suggesting that the full-length atMSH2 protein was unstable and possibly

being degraded to smaller polypeptides (Figure 8). This construct was further tested

for expression in protease deficient E. coli strains, such as BL21, with similar results.

It could not be ruled out that the aberrant expression pattern could be due to other

factors, such as downstream translational starts sites for example.

Page 62: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

12345 6M

47

97.4 kt)a

66 k)a

45 kDa

31 kDa

Figure 8. SDS-PAGE analysis of HIS-tag purification attempts of atMSH2.Lane 1 and lane 6 contain Ni2 resin batch-purified protein from lysates of XL!-blue cells transformed with pQE7O::atMSH2his plasmid DNA, and inducedwith 5mM IPTG (see Methods for details). Lane 2 and lane 4 are as aboveexcept the cells were not induced with IPTG. Lane 3 is as above except theplasmid vector (pQE7O) contained no insert (atMSH2his), and was not inducedwith IPTG. 5 tL of each sample, plus marker (right-most lane designated "M"),

was loaded onto a 10% SDS-PAGE gel, ran at 200 V for 45 mm, and stainedwith Coomasie blue.

Page 63: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

2.3.3 Negative Complementation of Wild-Type E. coli Cells in thePresence of atMSH2.

Previous studies have shown that overexpression of a heterologous member of

the MutS homolog protein family causes a "dominant negative" mutator phenotype in

E. co/i (Prudhomme et al., 1991; Fishel et al., 1993). The rational is that a

heterologous MutS can bind to mispaired bases, but is unable to signal downstream

events in mismatch repair pathways, thus causing an increased likelihood of

unrepaired mismatched bases. Because some full-length (100 kDa) atMSH2 protein

induction was apparent in strains containing pQE7O::atMSH2his, we hypothesized that

this expression could cause a dominant mutator phenotype in E. co/i cells. To test this

possibility, cells containing pQE7O: :atMSH2his were analyzed for an increased rate of

accumulation of rifampicin resistance (rifR) mutations. Mutations in E. coli that result

in rifampicin resistance have been mapped to theI

subunit of RNA polymerase and

have little or no effect on cell growth or viability (Nene and Glass, 1982).

Several isolates of E. coli strains harboring either the control plasmid pQE7O

(no insert), or plasmid pQE7O::atMSH2his, were grown to saturation in the presence

of 50 jtg/ml ampicillin and kanamycin (for plasmid selection, see Methods) plus

IPTG. Dilutions of these cultures were then plated on either LB plates containing

ampicillin and kanamycin (for cell viability counts) or LB plates containing ampicillin

and kanamycin plus rifampicin (for rifR counts). However, no significant increase of

Page 64: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

rif' colonies were observed for strains harboring the overexpression construct

pQE7O::atMSH2his versus control strains harboring the control vector pQE7O (Table

2).

2.4 Conclusions

We suggest that plants employ mismatch repair systems highly homologous to

those found in other eukaryotes. In particular, their systems include at least two of the

evolutionarily distinct MutS homologs described for yeast and mammals. Analysis of

the functional properties of these proteins likely will require purification directly from

plant tissue, or careful expression of cDNAs in either heterologous cells (insect cells,

for example) or cell extracts.

2.5 Materials and Methods

2.5.1 Growth Conditions

Arabidopsis thaliana seeds (ecotype Columbia) were sterilized in 50% commercial

hypochiorite bleach, washed five times in sterile water, and aseptically grown in

250-mI flasks containing 100 ml of liquid Murashige-Skoog (MS) medium

(Murashige et al., 1962) with 0.5% sucrose (pH 5.8). After 14 days, seedlings were

Page 65: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

50

Table 2. Negative complimentation of wild-type E. coli in the presence of atMSH2

Experiment number RifR colonies/108 celisa RifR colonies/108 celisa

pQE7O only pQE7O::atMSH2

0.9 2.2

10.4

.85

4.0

2 1.0 3.2

0.6 0.8

0.8

0.6

3 2.2 1.0

1.1 9.3

1.4

1.2

4 0.9 0.5

0.3 1.4

1.0

5 0.9 2.1

0.5 0.60.4

1.0

a Revertant rates were calculated by comparing the number of colonies for each culture

on both LB and LB+rifampicin plates.

Page 66: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

51

harvested for isolation of DNA (Murray et al., 1980) or mRNA (RNeasy RNA

isolation kit, Qiagen; mRNA Separator kit, Clontech).

2.5.2 PCR Techniques

We employed degenerate primer-oligonucleotide sets corresponding to highly

conserved MutS/MSH2 protein domains (Figure 4): primer 1, TGPNM (coding

strand) 5'-AGI GGI CCI AA(T/C) ATG GG-3'; primer 2, ELGRGT (non-coding

strand) 5'-GT ICC IC(T/G) ICC IA(AIG) (T/C)TC-3'; primer 3, FATH(Y/F)H

(non-coding strand) 5'-TG (G/A)(T/A)A (G/A)TG IGT IGC (A/G)AA. Polymerase

chain reaction amplification was performed in 100 tL reaction mixes containing 10

tL of lOX Reaction Buffer (Promega), 100 jimoles each of dATP, dCTP, dGTP,

dTTP, 2 units of Taq DNA polymerase (Promega), and 20 pmoles each of degenerate

primer pairs. A PCR optimization kit (Invitrogen) was used to optimize Mg2 and pH

conditions for each primer pair. Templates for initial PCR screening reactions using

primers 1 and 3 were either 20 ng of genomic DNA or 100 ng of cDNA [produced

from purified mRNA (see above) using an RT-PCR kit (Perkin-Elmer) in accordance

with the instructions of the manufacturer]. Amplification was carried out for 30

cycles, 30 sec. at 94°C, 30 sec. at 42°C, 3 mm. at 72°C. We analyzed reaction

products by electrophoresis on 2.5% agarose gels. Where cDNAs templated 0.35-kb

Page 67: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

52

products corresponding to the expected distance between primers 1 and 3 in

MSH2 -like sequences, these products were isolated and utilized as templates in PCR

reactions, under similar conditions, using primers 1 and 2. These reactions yielded the

expected 0.26-kb "nested" products. Among individual products initially templated by

genomic DNA using primers 1 and 3, those (0.35-kb and larger) that were repeatedly

obtained under a variety of reaction conditions were isolated and analyzed with

primers 1 and 2 as above. PCR products templated by cDNA or genomic DNA using

primers 1 and 3, and authenticated using primers 1 and 2 as above, were cloned in the

T/A vector pCRII (Stratagene).

2.5.3 Southern Hybridization

Arabidopsis thaliana genomic DNA was digested with EcoRI and BamHI

restriction enzymes, and 3 tg of the resulting DNA fragments were separated by

electrophoresis in 1% agarose. Transfer to nylon paper was as described (Maniatis et

al., 1982). A 32P-labeled probe was prepared by random priming of a 0.7-kb atMSH2

clone using a DECAprime II (Ambion) DNA labeling kit. Hybridization and

subsequent washes were at 42°C and 60°C, respectively. Final wash conditions were

2x SSC buffer (Maniatis et al., 1982), 0.5% sodium dodecyl sulfate, at 60°C for 30

mm.

Page 68: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

53

2.5.4 Isolation of cDNAs

We probed approximately iO plaques of an phage/plasmid 2-YES vector

Arabidopsis thaliana (ecotype Columbia) eDNA library with a 352-bp [32P]-labeled

Arabidopsis MSH2-like DNA fragment (see above). Plaque purification, plasmid

rescue, and cDNA isolation as described (Hoffman et al., 1996) yielded three partial

cDNAs, 1-2 kb in length, encoding the same 3'-terminal MSH2-like sequences. The

2-kb partial cDNA clone was further used to probe a library of 3-6 kb Arabidopsis

thaliana cDNAs inserted into phage lambda ZAPII (Kieber et al., 1993; available

from Arabidopsis Biological Resource Center). Infection of XLI-BIue MRF' cells

with phage stocks prepared from five positive plaques, plus helper phage, yielded

phagemids selectable by growth of transfected bacteria on ampicillin (100 g/ml)

plates (Bullock et al., 1987). In two cases, digestion with EcoRI endonuclease

released 3.1-kb fragments sufficient to encode a full length MSH2-like eDNA.

2.5.5 Isolation of 5'-RACE products

5- rapid amplification of cDNA ends (5'-RACE) was performed using a

Marathon cDNA Amplification Kit (Clontech), with purified mRNA (see above) as

described by the manufacturer. The second strand-synthesis cDNA products, ligated

to blunt-ended Marathon adapters, were used as templates in 50-pt PCR reactions

Page 69: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

54

containing 5 units KlenTaq DNA polymerase mix (Clontech), 200 pmoles adapter

primer (APi), 200 pmoles atMSH2 gene-specific primer (see below), 5 L lox

reaction buffer, and dNTPs to 200 M. We employed a "touchdown" protocol to

reduce background products: 1 cycle 94°C for 1 mm.; 5 cycles 94°C for 30 sec., 72°C

for 4 mm.; 5 cycles 94°C for 30 sec., 70°C for 4 mm.; 25 cycles 94°C for 30 sec.,

55°C for 30 sec., 72°C for 4 mm. The gene-specific primer sequence (5'- ACC TCA

GAG AAG CTG GTA ACA GTC 3') corresponded to position +1763 relative to the

first ATG of the longest open reading frame in the MSH2-like 3.1-kb cDNA (see

above). The single product, a 1.9-kb fragment, was purified by gel electrophoresis and

inserted into the T/A cloning vector pCRII.

2.5.6 Genomic Clone Isolation.

Genomic clones were identified by probing a Lambda ZAPII genomic library

(Stratagene) with the 3.1-kb cDNA fragment, and phagemids were isolated as

described above. Five genomic clones spanned the entire region of the full-length

cDNA. Their coding-region sequences were identical with the cDNA sequences.

Page 70: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

55

2.5.7 Analysis of DNA Sequences

Nucleotide sequences were obtained through the Oregon State University

Central Services Laboratory using an Applied Biosystems DNA sequencer and Taq

Dye-Primer/Dye-Terminator cycle sequencing kit, as described by the manufacturer,

plus synthetic internal primers as needed. Sequence alignments and construction

were performed using GCG (Genetics Computing Group, Version 8). CLUSTAL

alignments and phylogenetic analyses were performed using Genetic Data

Environment (GDE).

2.5.8 Overexpression of atMSH2

Primer 5pMSH2 (5'-CGC CAA YTG AGG AAT CTF CCA ATG GAG GGT

AAT-3') which overlaps the initiating aIMSH2 ATG and encodes a (AGGA)

ribosomal binding site and a Muni site, and primer 3pMSH2 (5'-CTG CAT GCT CAG

GAT CCC AGA AAC TGC CTG AGC CAG-3') which overlaps the 3' termination

codon, and encodes SphI and BamHI sites, were used to PCR amplify the atMSH2

cDNA. Positive PCR clones were digested with MunI and SphI and inserted into the

EcoRl and SphI sites of pQE7O. A full-length cDNA containing atMSH2 was digested

with the BstBI and Eco4 7111 endonucleases (unique sites within the atMSH2 cDNA

and pQE7O), and inserted into the BstBI and Eco4 7111 sites of pQE7O: :atMSH2 in

Page 71: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

56

order to replace most of the PCR amplified region of atMSH2. Both 5' and 3'

sequence regions of atMSH2 outside of the BstBI and Eco4 7111 sites were confirmed

by sequencing as described above.

6X-HIS tagged (pQE7O: :atMSH2his) constructs were obtained by digesting

pQE7O: :atMSH2 with BamHI, which drops out a BamHI specific product (<0.1 kb in

size) within the 3' region of pQE7O::atMSH2. Re-ligation of the plasmid yields a

open reading frame of atMSH2 that includes the 6X HIS-tag encoding portion of

pQE7O.

Both pQE7O::atMSH2 and pQE7O::atMSH2his were transformed into E. coli

strains along with pUBS52O, a plasmid that encodes (AGA) tRNAs and confers

kanamycin resistance. Co-transformed cells were plated on LB media containing 50

tg/mI of kanamycin and ampicillin. For overexpression analysis, positive colonies

were grown in the presence of 0.5 mM IPTG (isopropyl-3-D-thiogalactopyranoside)

for various time lengths, and denatured extracts were prepared and analyzed on SDS-

PAGE.

2.5.9 Negative Complementation Analysis

Positive ampicillin and kanamycin resistant transformants were selected and

grown to saturation in LB broth containing 50 .tg/ml of ampicillin and kanamycin,

Page 72: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

57

plus 0.5 mM IPTG. These cultures were then subcultured, and grown to 108 cells/mi,

and dilutions of these cultures were plated on LB plates containing ampicillin and

kanamycin in order to determine the number of viable cells in the cultures, and on LB

plates containing ampicillin, kanamycin, plus 100 tg/ml rifampicin to detennine the

total number of spontaneous rif mutants present in the cultures. The rate of mutation

was calculated according to Lea and Coulson (1949) using r0 = M(1.24 + lnM), where

r0 is the median number of rifR mutations in an odd number of independent cultures

and M is the average number of rifR mutations per culture. M was solved by

interpolation from the known value and then used to calculate the mutation rate,

where r=M/N, and N is the final average number of viable cells.

Page 73: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

Chapter 3

Evolutionary Origin, Diversification and Specialization of Eukaryotic MutSHomolog Mismatch Repair Proteins

Kevin M. Culligan, Gilbert Meyer-Gauen, James Lyons-Weiler and John B. Hays

Published in Nucleic Acids Research

Oxford University Press

January, 2000

Page 74: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

3.1 Abstract

Most eubacteria, and all eukaryotes examined thus far, encode homologs of the

DNA-mismatch-repair protein MutS. Although eubacteria encode only one or two

MutS-like proteins, eukaryotes encode at least six distinct MutS Homolog (MSH)

proteins, corresponding to conserved (orthologous) gene families. This suggests

evolution of individual gene family lines of descent by several

duplication/specialization events. Using quantitative phylogenetic analyses (RASA,

or Relative Apparent Synapomorphy Analysis), we demonstrated that comparison of

complete MutS protein sequences, rather than highly conserved C-terminal domains

only, maximizes information about evolutionary relationships. We identified a novel,

highly-conserved middle domain, as well as delineated clearly an N-terminal domain,

previously implicated in mismatch recognition, that shows family-specific patterns of

aromatic and charged amino acids. Our final analysis, in contrast to previous analyses

of MutS-like sequences, yielded a stable phylogenetic tree consistent with the known

biochemical functions of MutS/MSH proteins, that now assigns all known eukaryotic

MSH proteins to a monophyletic group, whose branches correspond to the respective

specialized gene families. The rooted phylogenetic tree suggests their derivation from

a mitochondrial MSH 1-like protein, itself the descendent of the MutS of a symbiont in

a primitive eukaryotic precursor.

Page 75: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

3.2 Introduction

The isolation of multiple MutS homolog-encoding genes in Arabidopsis

thaliana described in Chapter 2, and identification of additional MSH genes via the

Arabidopsis thaliana genome sequencing project (see below), suggested a similar

pattern of functional diversification and specialization seen in other eukaryotes.

However, it was unclear whether these MSH genes arose from the same or different

precursors found in yeast and humans. If, for example, plant MSH genes arose via a

different precursor than other eukaryotes, functional conservation of plant and animal

MSH homologs would thus be questionable. In order to be able to confidently predict

the function of newly identified plant MSH genes, we sought to rigorously analyze the

MSH gene family, and address the evolutionary origin of the MSH gene family. In

doing so, the analysis raised several interesting questions about the evolution of MSH

genes. Was there a single, or multiple evolutionary precursor(s) of MSH subfamilies?

What was the source of the precursor(s)? What was the order of the occurrence of the

diversification! specialization steps acquisition of nuclear-replication-fidelity and

meiotic-recombination functions, shifting to heterodimeric structures, development of

substrate-recognition specificities? We have used a phylogenetic approach, in

conjunction with genetic and biochemical information, to address these questions.

This required analysis of a wide range of sequences from highly diverse organisms

Page 76: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

61

and extraction of the maximum amount of phylogenetic information. Some earlier

studies of MSH phylogenies were incomplete with respect to the range of groups of

MutS-like proteins analyzed (Culligan and Hays, 1997; Pont-Kingdon et at., 1998),

descriptions of methods (Kolodner, 1996; Fishel and Wilson, 1997), or definition of

the phylogenetic root. A more complete study focused only on a very highly

conserved MutS/MSH region (Eisen, 1998). We have used alignments and analyses

that extend over complete protein sequences, in order to capture as much phylogenetic

information as possible. We also used a tree-independent analysis of phylogenetic

signal termed RASA (Relative Apparent Synapomorphy Analysis) to identify

phylogenetically problematic sequences; some of these were then excluded from the

analysis, and as a result, we have maximized tree stability. The results of our analysis

appeared more consistent with known biochemical functions of MutS and MSH

proteins than results of previous studies. Rooted phylogenetic analysis of complete

MutS-like sequences indicate that all eukaryotic MSH proteins are monophyletic, and

originated from a eubacterial endosymbiont.

Page 77: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

62

3.3 Results and Conclusions

3.3.1 Classification of MutS-like Proteins According to SequenceOrganization and Co-occurrence with MutL Proteins

Inspection of initial alignments of all available MutS-like protein sequences

identified two clearly distinct groups of proteins. In a representative list (Table 3) of

MutS-like sequences used in this study, the two groups showed clear differences in

organization of primary structure and genomic context (Table 3, columns 5 and 6).

For example, B. subtilis contains two mutS-like genes (here called mutS and msp, see

below) and only one mutL. Each corresponding MutS-like protein sequence contains a

C-terminal conserved domain; however, the protein designated MutS contains two

other conserved domains not present in the protein designated MSP (Table 3; Figure

9A). Furthermore, in the case of H. pylon, msp is present, but both mutS and mutL are

absent. A previous analysis also identified two groups (lineages) of MutS-like

proteins (Eisen, 1998), but the proposed composition of these differs in important

ways from the two groups identified here. We argue below that because of their gross

differences in functional domain structure the two groups delineated by our analysis

most likely have different biological functions, and therefore suggest the designation

MutS/MSH, for eubacterial MutS proteins previously designated MutSi plus all

eukaryotic MSH proteins, and MSP (MutS iaralog), for the novel eubacterial open

Page 78: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

Table 3. MutS/MSH and MSP Sequences

Protein Organism' Class

Length(A.A.$)

Conserved Domains(Coordinates)5 Genomic Repertoire

ProkarvotesMutS Streptococcus pneunloniae2 MutSIMSH 844 N M C mutS, rnutL

9-47 250-342 539-782

MutS Bacillus subtiliv2 MutS/MSH 852 N M C mutS, rout], m.rp

1-39 244-336 534-776

MutS Escherichia coli3 MutS/MSH 853 N M C mutS, murL

13-5 1 267-355 552-794

MutS Haemophilus influenzae3 MutSIMSH 854 N M C mutS, mutL

13-51 267-357 554-796

MutS Azotobaeter vinelandii MutS/MSH 855 N M C mutS, mutL

10-48 264-354 551-792MutS Thermus aquaticus3 MutSIMSH 811 N M C murS, mutL

16-54 250-339 527-760

MutS Synechocyslis sp.3 MutS/MSH 912 N M C mutS, mutL, msp

62-100 334-426 623-870

MutS Rickeasiaprowazekii3 MutS/MSH 891 N M C mutS, mutL

23-61 288-379 581-826MSP Bacillus subtilis3 MSP 785 N M C mutS, mutL, msp

-- 276-513

MSP Synechocystis sp.3 MSP 822 N M C mw'S, murL, msp-- -- 279-544

MSP Helicobacterpylori MSP 762 N M C msp-- 277-510

MSHI (mtMutS)EukarvotesSarcophytom glaucum4 MutS/MSH 982 N M C mshl and other?

5-47 334-423 631-876MSHI Saccharomyces cerevisiae MutSIMSH 959 N M C mshl/2/3/4/5/6, mlhlT2/3, pros]

61-99 342-434 685-938

MSH2 Saccharomyces cerevisiae MutS/MSH 964 N M C mshl/123/4/5/6, mlhI/213, pms]19-57 295-406 620-877

MSFI2 Arabidopsis thaliana MutS/MSH 937 N M C mshl/2/3/617, mihi, pms2, other?24-6 1 294-387 599-855

Page 79: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

Table 3. (Continued)

MSH2 Xenopus leavis MutS/MSH 933 N M C ,nsh2 and other?

19-57 299-393 601-849

MSH2 Hornu sapiens MutS/MSH 934 N M C rnsh2/3/4/5/6, mihi, pn,s2, other?19-57 300-394 602-849

MSH3 Saccharomyces cerevisiae MutS/MSH 1047 N M C rush] /2/3/4/5/6, rnlhI/2/3, puts!164-202 447-539 757-1005

MSH3 (SWI4) Saccharomyees pombe MutS/MSH 993 N M C msh3 and other?

106-144 402-495 697-948

MSH3 Homo sapiens MutSIMSH 1128 N M C msh2/3/4/5/6, mihI, prns2, other?

232-270 535-628 831-1091

MSH4 Saccharomyces cerevisiae MutS/MSH 878 N M C ,nshu/2/3/4/5/6, mi/i 1/2/3, puts]

265-356 574-8 13

MSH4 Homo sapiens MutSIMSH 936 N M C mhs2/3/4/5/6, rn/hi, prns2, other?

310-402 620-858MSH5 Saccharomyces cereviswe MutSIMSH 901 N M C rnshIf2/3/4/5/6, mihi/2/3, puts]

256-348 569-843

MSH5 Homo sapiens MutS/MSH 834 N M C msh2/3/4/5/6, mihI, prns2, other?-- 231-319 529-776

MSH6 Saccharomyces cerevisiae MutSIMSH 1242 N M C rnshl/2/3/4/5/6, ,nihI/2/3, pmsi314-352 614-706 905-1164

MSH6 Arabidopsis thaliana MutSIMSH 1338 N M C rnshlT2/3/6/7, mi/il, pn,s2, other?395-433 709-802 1023-1281

MSH6 Hornr, sapiens MutSIMSH 1360 N M C msh2/3/4/5/6, mi/il, prns2, other?409-447 733-825 1057-1321

ArchaeaMutS? Met. Thermoautotrophicum MutSIMSH 647 N M C ,nutS?

186-240 402-623

'Bold-face, entire genomic sequence currently available.

2Gram-positive eubacterium.

3Gram-negative eubacterium.

4Octocoral.

5N, N-terminal conserved domain; M, middle conserved domain; C, C-terminal conserved domain; dashes (--), absence of domain;

coordinates correspond to positions of amino acids in domains.

Page 80: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

Figure 9. Structure of MutS/MSH and MSP proteins. (A) Light gray bar, N-terminal extension found in MSH3 and MSH6 proteins. Different shaded boxesrepresent regions of high homology (> 50% sequence similarity) within classes;black, MutS/MSH N-terminal domain; dark gray, MutS/MSH middle domain; lightgray, MutS/MSH C-terminal domain; medium gray, MSP C-terminal domain.Domains Dl (ATP-binding) and D2 ('DELGRG") retain strict conservation in theMutS/MSH class. I indicates 18 amino-acid insertion in the S. species MSP protein.(B) BOXSHADE output of the MutS/MSH N-terminal domain. Black boxesrepresent identical amino-acids in more than 80% of the sequences, gray boxessimilar amino-acids in over 50% of the aligned sequences based on Dayhoff'sPAM25O matrix. P (in upper bar): Amino-acid position. [1]: 17 amino-acidinsertion in the MSH4 protein of S. cerevisiae. (C) MutS/MSH N-terminalconserved domain structure. Aro, (+), (-) respectively represent aromatic (F/Y/W),positive (DIE), and negative (K/R) charged amino-acid side chains conserved inMutSIMSH sub-families; (n) represents any amino acid. The arrow denotes thePhe-39 (F39) position of T. aquaticus MutS. (D) Middle conserved domain as in(B). The arrow indicates a known dominant negative mutation (R305H) in E. coliMutS. [2]: 14 amino-acid insertion in the MSH2 protein of S. cerevisiae.

Page 81: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique
Page 82: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

67

reading frames previously designated MutS2 (Eisen, 1998; Eisen et al., 1997). Thus,

these designations discriminate between genes that function similarly to MutS, and

those that simply contain similar domains. We consider below and in the next section

whether MSP proteins should be included in our analyses that address the questions of

MSH origin and diversification.

All eubacterial MutS and eukaryotic MSH proteins (MutS/MSH class) appear

to function in DNA error-correction and/or recombination pathways. These proteins

are highly similar with respect to sequence and domain structure (Figure 9A).

Particularly well-conserved domains appear in the N-terminal, middle, and C-terminal

regions (Figure 9). The N-terminal conserved domain, found only in bacterial MutS

and eukaryotic MSH1, MSH2, MSH3, and MSH6 proteins, is predicted to closely

interact with DNA mismatches, based on photo-crosslinking and mutagenesis studies

of the Phe-39 residue in the N-terminal domain of T.aquaticus MutS protein (Malkov

et al., 1998). MutS/MSH families show different patterns of conservation of residues

close to Phe-39 (Figures 9B, 9C) (Malkov et al., 1998). Proteins that bind base/base

mismatches eubacterial MutS, and eukaryotic MSH1, MSH2, MSH6 proteins

conserve an aromatic (F/Y/W) (FlY) doublet (e.g. Phe-39lTyr-40 in T. aquaticus

MutS). In MSH3 proteins, which may participate in binding looped-out DNA strands,

the two aromatic residues alternate with two positively-charged residues Y (KIR)

Page 83: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

(Y/F) (R/K). MSH4 and MSH5 proteins, which appear to play no role in error

correction, and presumably do not recognize mismatches, show little conservation

here.

The middle domain appears in all MutSIMSH proteins, and is conserved more

or less homogeneously among all MutS/MSH families (Figure 9D). Three-

dimensional-structure analysis suggests that this domain lies on the surface of H.

sapiens MSH2 (De Las Alas et al., 1998). Mutation of a highly conserved arginine in

this domain of E. coli MutS (R305H) confers a dominant-negative phenotype (Wu and

Marinus, 1994). Although function cannot be definitively assigned to this domain, it

might be involved in protein-protein interactions, such as those between subunits of

MutS/MSH protein heterodimers, or in interaction with other components of the

mismatch-repair apparatus. Two previous studies have focused on mapping

interaction domains of human and yeast MSH heterodimers, indicating possible N-

terminal and C-terminal heterodimerization domains in MSH2, MSH3, and MSH6

(Alani, 1996; Guerrette et al., 1998). These regions do not correspond to the middle

domain identified here, thus suggesting that the middle domain is not directly involved

in heterodimer formation. However, more definitive studies will be needed to further

elucidate MSH interaction domains.

Page 84: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

The C-terminal conserved domain [Figure 9A, alignment not shown; (Eisen,

1998)] is found in all MutSIMSH sequences, and shows the highest conservation of all

three domains. It contains helix-turn-helix and nucleotide/magnesium binding

(Walker box) subdomains, and is predicted to interact with DNA and to mediate ATP

binding and hydrolysis (Kolodner, 1996; Allen et al., 1997; Fishel and Wilson, 1997;

Eisen, 1998).

No function has been identified for genes of the MSP class, which are found

only in certain eubacteria. The predicted MSP protein sequences are relatively

conserved over their entire lengths, but differ markedly from those of MutS/MSH

protein sequences (Figure 9A). First, MSP sequences lack both the N-terminal and

middle conserved domains of MutSIMSH sequences. Second, MSP sequences contain

unique terminal extensions of approximately 200 amino acids. Third, although certain

MSP domains are similar to the conserved C-terminal domains of MutS/MSH

sequences, they appear instead near the middle of MSP sequences and do not strictly

conserve the spacing of critical functional subdomains seen in MutSIMSH C-terminal

domains (Figure 9A).

Eubacterial MutS proteins have thus far been observed only in conjunction

with MutL proteins, and eukaryotic MSH proteins only in conjunction with MLII

(PMS) proteins. All eukaryotic MSH proteins, except perhaps MSH1, appear to

Page 85: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

70

interact with one or more MLH proteins. MSH2MSH3 or MSH2'MSH6

heterodimers, plus MLH1'PMS2 heterodimers, are required for mismatch-repair

functions in yeast and human cells (Marsischky et al., 1996; Li and Modrich, 1995);

direct interactions among these proteins have been demonstrated (Prolla et al., 1994;

Habraken et al., 1998). Furthermore, MSH2'MSH3 has recently been shown to

interact with a heterodimer of MLH1MLH3 (Flores-Rozas and Kolodner, 1998).

Genetic epistasis studies indicate that MSH4, MSH5, and MLH1 act in the same

meiosis-specific pathway (Hunter and Borts, 1997), perhaps again interacting with one

another in multi-protein complexes. In contrast, MSP genes are found in some

eubacteria that lack mutL-like genes. Where both MSP and mutL genes are present, a

"true" inutS gene of the inutS/MSH class is also present (e.g. Kaneko et al., 1996).

MSP proteins thus constitute a class distinct from MutSJMSH proteins.

Among the three complete archeabactenal genome sequences reported, only

that of Methanobacterium thermoautotrophicum contains a mutS -like gene (Smith et

al., 1997). This gene predicts a protein, approximately 150 amino acids shorter than

eubacterial MutS proteins, that lacks the N-terminal domain and portions of the middle

domain of MutS/MSH sequences, but shows no MSP-like C-terminal extension or

other MSP-like characteristics. M. therinoautotrophicum encodes no MutL-like

protein, so its MutS-like protein might be considered to define a third class.

Page 86: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

71

3.3.2 Optimization of Analysis of MutS/MSH and MSP Phylogeny

In order to rigorously address questions relating to MSH origin and

diversification, we initially needed to resolve two general points. First, we needed to

define analytical techniques that used as much sequence information as possible, thus

maximizing phylogenetic signal. Second, we need objective criteria to determine

whether MSP sequences, already identified as likely to constitute a distinct class of

separate origin (see above), should be included. These points prove to be interrelated:

in analysis of complete protein sequences, masking to exclude sequence gaps and

regions of ambiguous alignment between the MutS/MSH and MSP groups invariably

yielded alignments of only a highly conserved 280-residue region found in the C-

terminal regions of MutS/MSH proteins (Figure 9A), because of the marked structural

differences outside of this region. We were able to construct neighbor-joining (NJ)

minimum-evolution and parsimony trees, using (masked) alignment of 28 MutSJMSH

and MSP C-terminal protein sequences, representing all classes of MutS-like proteins

(Figure 9). However, we observed several ambiguities in each of a number of trees

generated from different sets of sequences based on this alignment (see below).

To quantitatively estimate the amount of phylogenetic signal generated by

considering only C-terminal regions, we employed RASA [Relative Apparent

Synapomorphy Analysis, (Lyons-Weiler et al., 1996; see Methods]. We used taxon-

Page 87: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

72

variance ratios to identify long branches (Lyons-Weiler and Hoelzer, 1997). In

optimizing our analyses in order to increase confidence in the resulting trees (see

below), our criteria were (i) increased values of the tRASA test statistic (a measure of

the strength of the phylogenetic signal), (ii) homogeneity of taxon-variance ratios

(absence of long-branch attraction), a measure of tree stability, and (iii) improved

bootstrap support throughout the trees. Figure 1OA shows a condensed neighbor-

joining tree, and taxon-vanance ratios, produced by analysis of C-terminal regions of a

representative set of all MutS-like proteins (including MutS/MSH and MSP proteins,

and the putative archeabactenal MutS from M. the rinoautotrophicum). This analysis

produced a tRASA value of 9.3 (p<<O.005, see Methods). The taxon-variance ratio

for the M. thermoautotrophicum MutS sequence indicated that it might be problematic

here, in the sense of causing long-branch attraction. In the tree, the MSP group

branched together with M. thermoautotrophicum MutS within the eubacterial cluster,

and the majority of the tree branches showed low bootstrap support. A second

analysis, that excluded the M. thermoautotrophicum MutS, produced a higher tRASA

value of 11.2 (p<<O.005), and the taxon-variance ratios now suggested that the H.

pylon MSP sequence was problematic (data not shown). The tree again showed low

bootstrap support for most branches. Strikingly, the MSP group now branched within

the eukanyotic MSH cluster, together with the S. glaucum mtMutS and close to the

Page 88: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

AHutS TaSwi4 ScHutS SpHutS R9NutS SBHutS SCMSP Ba

NutS BaN

5) HSPHpz HutS Mt

MSH6 ScN MSH6 Ha

MSH6 AtMSH5 ScMSH5 Ha

H MSH4 ScMSH4 HaNSH3 ScMSH3 HaMSH2 ScMBH2 HaMBH2 XlMSH2 AtMSH1 ScHutS HiHutS BCHutS Av

2 3 4

Ftv

MSH6

MSH3

MSH3 S. cerevisiae

5 6

MSH527

'°° MSH4100 _MSH2

MutS R. prowazekii

MSH1 S. cerevisiae

MutS (Gram Negative) 90 -

MutS T. aquaticus

MutS (Gram Positive)

tRASA=9.3 (p<<0.005)

NutS TaSwiS ScNutS SCNutS Rp

NutS SB

N NutS Bao MSH6 Sc

Z MSH6 Ha

H MSH6 At

MSH3 SC

MSH3 HaSI NSH2 ScN M8H2 Ha

NSH2 XlMSH2 At

Sisal SC

NutS HiNutS BCNutS Av

mtMutS S. glaucum

MSP

MutS M. the rmo.

I 2 3 4 5 6

Ftv

-MSH6100

MSH3o°

MSH2

MSHI S. cerevisiae

MutS R. pro wazekii

MutS (Gram Negative)

MutS (Gram Positive)

tRASA=16.7 (p<<0.005)

0.10H

73

Figure 10. Schematic neighbor-joining (NJ) trees and corresponding RASA taxon-variance ratios using different combinations andlor regions of MutS-like proteinsequences. For phylogenetic methods see legend to Figure 11. (A) Representative setof all known MutS-like sequences, aligned as described in the text with a 280 amino-acid C-terminal mask. (B) Entire protein regions of MutS/MSH proteins, aligned asdescribed in the text with a complete sequence mask for a core set of 19 sequences.RASA taxon-variance ratios were determined as described in Methods.

Page 89: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

74

MSH4 and MSH5 groups. In both the first (Figure 1OA) and the second case, the

MSP group branched closest to the sequence having the longest branch in the

respective tree M. the rmoautotrophicum MutS in one case and S. glaucum mtMutS in

the other. Confidence in these two trees was thus low, because of the uneven taxon-

variance ratios and low bootstrap support. In fact, other analyses using the same C-

terminal region of the alignment, and using different subsets of sequences, produced

similar outcomes (tree instability), including long-branch attraction and low bootstrap

values. These examples, representative of tree instabilities observed in analyses of

only C-terminal regions, suggest that the C-terminal region alone is not sufficient to

resolve critical branching patterns in phylogenetic analyses of MutS-like sequences.

The lower phylogenetic signal and instability of trees associated with C-terminal

analyses that include both MutS/MSH and MSP protein sequences (Figure bA), the

gross differences between MutSIMSH and MSP protein sequences (Figure bA), and

the lack of correlation between the occurrence of MSP and MutL proteins, all suggest

that MSP proteins are not closely related to eubacterial MutS proteins. We propose

that msp genes arose independently, perhaps through domain shuffling in eubacteria,

and have excluded them from analyses described below.

We were now able to include all conserved domains (N-terminal, middle, and

C-terminal) in alignments (Figure 9). We conducted RASA and phylogenetic analyses

Page 90: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

75

of complete MutS/MSH protein sequences, to determine whether this would increase

overall phylogenetic signal, bootstrap confidence values, and more homogeneous

taxon-variance ratios. In an analysis of 24 complete MutS/MSH sequences

representing all MutS/MSH groups, individual members of the MSH4 and MSH5

groups and the S. glaucum mtMutS now showed problematic taxon-variance ratios.

Removal of these sequences produced a "core" tree with no apparent long branches,

revealed in the taxon-variance analysis. Figure lOB shows the condensed neighbor-

joining distance tree, and the taxon-vanance ratios, for this core set of 19 complete

MutS/MSH protein sequences. In comparison to the C-terminal analyses (Figure

bA), the bootstrap values are significantly increased and show less ambiguity for all

branches. All branches are resolved (Figure lOB) and stable, and the tRASA value

increases to 16.7 (p<<O.005)(Figure lOB). The core tree therefore provides a

benchmark reference point when analyses are expanded to include more diverged

sequences (e.g., MSH4 and MSH5) Thus, expanded analyses in which branching of

the core sequences were significantly altered would a priori be considered

questionable.

Page 91: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

76

3.3.3 Reconstruction of MutSIMSH Phylogeny

We next addressed the original MSH origin/diversification questions, using

complete protein sequences, excluding the MSP class, and referring back to the core

branching pattern. On the basis of protein and DNA sequence data (Gray and

Spencer, 1996; Gupta and Golding, 1996) and complementary biochemistry of energy

metabolism (Martin and Muller, 1998), the eukaryotic cell has been proposed to be the

result of endosymbiosis between an archaebactenum (host) and a cx-proteobacterium

(symbiont) similar to the modern R. prowazekii (Anderson et al., 1998). Did this

eukaryotic DNA-mismatch repair gene family evolve from eubacterial and/or

archeabacterial precursors subsequent to such an event? Because archaebacteria seem

not to possess mismatch-repair pathways involving MutS and MutL proteins,

eukaryotic MSH genes most likely came from the eubacteria. In our final

phylogenetic reconstruction we used 24 complete MutS/MSH sequences, including all

groups of MutS/MSH proteins the core sequences analyzed above (Figure lOB), the

S. glaucum mtMutS, and the MSH4 and MSH5 sub-groups. The analysis yielded the

highest degree of bootstrap support and phylogenetic signal (tRASA= 14.3; p<<O.005)

of any analysis that included all groups of MutS/MSH sequences, despite the

potentially problematic MSH4/MSH5 and S. glaucum mtrnutS taxon-variance ratios.

Although the MSH4 and MSH5 sub-groups branched together (consistent with the

Page 92: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

77

apparent roles of both in meiotic recombination), the mtMutS from S. glaucum did not

now branch with (or close to) the MSH4/MSH5 group (Figure 11), in contrast to the

tree shown in Figure IOA and a tree proposed elsewhere (Eisen, 1998). Exclusion of

any or all of these sequences in this analysis did not significantly change the overall

topology of the tree, its bootstrap values, or its close similarity to the previous core

tree (Figure lOB). This stable tree thus appears to provide the best estimate of

MutSJMSH sequence relationships. The expanded neighbor-joining tree (Figure 11)

suggests the following postulated scenario for evolution of MSH proteins in

eukaryotic cells.

An engulfed eubacterial (a-proteobacterial) cell, the precursor of the

mitochondrion (Margulis, 1970; Gray and Spencer, 1996; Gupta and Golding, 1996),

was the source of the common ancestral MSH gene. The cx-proteobacterial mutS gene

could have been transferred to the nucleus after the engulfment, as were many other

(now) nuclear genes (Palmer, 1997; Martin et al., 1998). Among eukaryotic MSH

proteins, the mitochondrial MSH1 protein subfamily is the deepest branching group

(Figure 11), and the yeast MSH1 shows the highest similarity (38%) to the cx-

proteobacterium Rickettsia prowazekii MutS. The R. prowazekii MutS branches

within a dade that includes all eukaryotic MSH protein sequences in the final tree,

with a bootstrap value of 75% (Figure 11). In addition, a similar branching pattern is

Page 93: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

Msh6 H. sapiens

Msh6 A. thaiiana

Msh6 S. cerevisiae

Msh3 H. sapiens

- Swi4 S. pombe

Msh3 S. cerevisiae

Msh2 H. sapiens

Msh2 X. iaevis

Msh2 A. thaliana

Msh2 S. cerevisiae

- Msh5 H. sapiens- Msh5 S. cerevisiae

Msh4 H. sapiens

Msh4 S. cerevisiae

Mshl S. cerevisiae

mtMutS S. glaucum

MutS R. pro wazekii

MutS E. coil

MUtS H. influenzae

91MutS A. vinelandii

i MutS T. aquaticus

MutS S. species01

MutS B. subtilis

MutS S. pneumoniae

Figure 11. Neighbor-joining (NJ) tree for Dayhoff PAM distances amongMutS/MSH protein sequences. Protein parsimony trees were also constructed usingPROTPARS (Phylip 3.5), which produced very similar results (not shown). Gaps andregions of ambiguous alignment were excluded from the analysis. The horizontalscale bar indicates evolutionary distance. Numbers above each branch represent thenumber of times the branch was found in 100 bootstrap replicas. The B. subtilis and S.pneumoniae MutS protein sequences (Gram-positive eubacteria) were used as anoutgroup. The masked alignment used to generate this tree (and the tree in Fig. 2)included the N-terminal, middle, and C-terminal regions. Differential shadings reflectthe known functional role of each group: black, eubacterial mismatch repair andrecombination; medium gray, mitochondrial mismatch repair in eukaryotes; light gray,

nuclear mismatch repair/recombination in eukaryotes. All eukaryotic homologs(MSH) are encoded by the nuclear genome, except the mitochondrially-encodedmtMutS from S. glaucum.

Page 94: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

79

observed in the core tree, here with a bootstrap value of 95%. The R. prowazekii

genome does not encode an MSP protein, but it does encode a MutL protein,

suggesting the presence of a functional mismatch-repair pathway. The close

branching of R. prowazekii MutS and eukaryotic MSH1 sequences suggests that

eubacterial mutS genes acquired during endosymbiotic events are the direct ancestors

of the mitochondrial MSH1 genes, which in turn gave rise to all other MSH genes.

Since our analysis here suggests that all members of the eukaryotic MSH gene family

are monophyletic (appear to share a common ancestor) with the mitochondrial MSH]

sub-family, any putative former post-replication error-correction pathway already

present in the protoeukaryote that engulfed the a-proteobactenal mitochondrial

ancestor would seem to have disappeared.

Interestingly, the octocoral Sarcophytorn glaucum mtMutS sequence and the

yeast MSH1 sequence branch together, although they are encoded respectively by the

S. glaucum mitochondrial genome and the yeast nuclear genome (Reenan and

Kolodner, 1992; Pont-Kingdon et al., 1995). Others have proposed that the S.

glaucum mtmutS is an example of a nuclear gene transferred to the mitochondrial

genome (Pont-Kingdon et al., 1998), or that it is of MSP origin (Eisen, 1998). Our

analysis is more consistent with two different possibilities: (i) the S. glaucum mtmutS

is an MSHJ gene that originated from a a-proteobacterial mutS (not MSP) gene, as did

Page 95: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

other MSHJ genes, but in this case has remained in its original mitochondrial genome,

exemplifying an intermediate step in the transfer and evolution of MSH genes in

eukaryotes; or (ii) the S. glaucum mtmutS (mshl) has been transfened twice, the

second time from the nucleus back into the mitochondnal genome.

A phylogenetic analysis of complete MutS/MSH protein sequences (excluding

MSP sequences, but including the M. thermoautotrophicum MutS) places the single

archaebacterial MutS within the eubacterial MutS group, branching closest to the

The rmus aquaticus MutS, with a bootstrap value of 78% in the neighbor-joining tree

(data not shown). This is in disagreement with organismal and rDNA (5S)

phylogenies which place M. thermoautotrophicum in the Euryachaeota group of the

Archaea (Maidak et at., 1997). M. therinoautotrophicum MutS may therefore have

been acquired through horizontal gene transfer, as previously suggested (Eisen, 1998),

possibly from another thermophile such as The rmus aquaticus. Archeabacteria, if

capable of post-replication error correction, would seem to use a pathway independent

of MutS or MutL proteins.

3.3.4 Rooting the Tree

One problem with deep phylogenies (those that encompass highly diverse taxa)

is the selection of an appropriate root. This is usually performed using an outgroup

Page 96: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

F;'

(Lyons-Weiler et al., 1998). In the case of the MutSIMSH phylogeny, we believe the

most appropriate root to be within the eubacteria. First, as discussed above, MSHJ

genes most likely originated from eubacterial mutS genes following mitochondrial

endosymbiotic events, providing a source for the gene duplications that eventually

resulted in specialized eukaryotic nuclear MSH genes. Second, there is no evidence

the nuclear MSH genes have an origin independent of MSH1. Archeabacteria do not

seem to have a MutS/L pathway, so they are unlikely to be the source of MSH genes.

Alternatively, another eubacterium not involved in the original endosymbiotic events

might have independently transferred its mismatch-repair gene(s) to primitive

eukaryotic cells, by the mechanism recently proposed by Doolittle (1998) for example.

However, no phylogenetic relationships apparent in our analysis suggest that nuclear

MSH genes arose from any eubacterial mutS genes other than the a-proteobactenal

mutS, via its proposed descendent, the mitochondrial MSH1, regardless of where the

root is placed. Our analyses thus suggest that mutS (and mutL) genes appeared early

(eubacterial evolution) and were later transferred to eukaryotes as part of the genomes

of (Gram-negative) a-proteobacterial endosymbionts. We therefore chose the Gram-

positive MutS sequences (here B. subtilis and S. pneumonea) as the root for the tree.

Indeed, rooted RASA (Lyons-Weiler et al., 1998) of the rest of the taxa, using a

Gram-positive-eubacterial root, resulted in the highest tRASA value of any rooted

Page 97: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

analysis of the entire set of 24 MutS/MSH sequences, indicating that these bacteria are

the optimal outgroup.

3.3.5 Eukaryotic MSH Gene Duplication and Specialization

The evolution of multiple eukaryotic nuclear MSH gene families may have

begun with transfer of a copy of the post-symbiosis mitochondrial MSHJ to the

nucleus. The remaining mitochondrial MSHJ would subsequently have been lost, as is

typical when nuclear genes encode mitochondrially targeted proteins. Duplication of

the nuclear MSH gene would have allowed one to encode an MSH1 protein targeted

back to the mitochondrion, and the other to give rise to the whole set of nuclear

mismatch-repair genes, by further duplication and specialization (Figure 12). The

increase in the DNA content of the eukaryotic genome, the development of diploidy,

the appearance of multiple chromosomes, and the evolution of meiotic recombination

may have been concomitant with the evolution of specialized mismatch

repair/recombination activities. The first duplication of the nuclear MSH ancestor

appears to have yielded the predecessor of MSH3 and MSH6 genes and the

predecessor of MSH2, MSH4, and MSH5 genes (Figure 11; Figure 12). This

duplication may have been the first step towards mispair recognition by heterodimers.

The MSH3-MSH6 predecessor apparently duplicated again to give rise to MSH3 and

Page 98: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique
Page 99: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

MSH6 subfamilies. MSH3 and MSH6 retained interaction with MSH2, but evolved

specialized but overlapping recognition functions. In the MSH2-MSH4-MSH5 gene

family, an MSH4-MSH5 predecessor with specialized meiotic functions may have

diverged from MSH2, giving rise to individual MSH4 and MSH5 gene families.

3.4 Perspectives

To address questions regarding the origin and diversification of eukaryotic

MSH proteins, we have systematically optimized our analysis of complete protein

sequences. In doing so, we have obtained a comprehensive phylogenetic

reconstruction of all known eubacterial, archeabacterial and eukaryotic groups of

MutS-like sequences, and identified two broad classes of MutS-like protein sequences,

namely MutS/MSH and MSP, consistent with the biochemical function of the former

with MutL-like proteins. This approach has established a general framework to

accurately classify newly identified MutS-like genes whose functions are unknown. A

valuable byproduct of this analysis was delineation of three domains that appear in all

MutS/MSH proteins thought to be involved in error-correction; these domains should

provide useful landmarks for establishing alignments and inferring biochemical

functions for new sequences.

Page 100: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

We suggest that because the conserved C-terminal region of MutS-like

proteins does not contain enough phylogenetic information, attempts to employ this

region only in comprehensive analyses invariably yield low bootstrap support and

unstable trees. The anomalous taxon-variance ratios for certain sequences obtained by

C-terminal analyses are indicative of long-branch attraction, which leads to erroneous

trees and unlikely hypotheses about the evolution of MutS/MSH and MSP proteins.

We suggest that future phylogenetic reconstructions use the complete amino-acid

sequences of MutS/MSH proteins, but include only those proteins that do not appear

to threaten the accuracy of the tree estimate by long-branch attraction. Our final

analysis yields a phylogenetic tree, with high bootstrap support for all branches, that

for the first time assigns all known eukaryotic MSH proteins to a single family of

proteins having distinct functional subfamilies (Figure 11), and suggests a eubacterial

endosymbiotic origin for all eukaryotic MSH genes.

Preliminary phylogenetic analyses of MutLIMLH protein sequences suggest a

similar pattern of evolution for both MSH and MLH genes, each from a single

eukaryotic ancestor {MSH1 and MLH1, respectively; unpublished observations,

(Kolodner, 1996)]. Interestingly, no gene in the S. cerevisiae genome appears to

encode a mitochondrially-targeted MLH protein, and no such genes have yet been

identified in other eukaryotes. Further phylogenetic analyses may determine whether

Page 101: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

both mutS and mutL are likely to have been acquired by eukaryotes at the same time

and/or by similar mechanisms. Although S. cerevisiae MSH1 has been shown to bind

mismatched DNA substrates with affinities similar to those of other MutS/MSH

proteins (Chi and Kolodner, 1994a), it and other MSH1 proteins may function via

novel error-avoidance mechanisms independently of MutL homologs (Chi and

Kolodner, 1994b).

It would be of high interest to identify organisms with smaller (or larger) sets

of nuclear MSH and MLH genes, representing different stages in the duplication-

specialization process, or gene loss. Plants also encode highly conserved MSH2,

MSH3, and MSH6 proteins (Culligan and Hays, 1997; Bevan et al., 1998); their

sequences clearly show conserved biochemical and phylogenetic relationships to their

respective MSH subfamilies (K.M.C. and J.B.H., Chapter 4; Figure 11). This suggests

that MSH duplication-specialization events occurred before the evolution of green

plants, and that plant MSH subfamilies were not acquired from the endosymbiotic

bacteria that gave rise to chioroplasts (cyanobacteria). Although no obvious MSHJ-

like gene is apparent in the C. elegans genome (K.M.C. and J.B.H., unpublished

observations) and no MSH1 gene has yet been reported for other animals, an MSHJ-

like gene is present in the Arabidopsis thaliana genome, and shows clear phylogenetic

relationships to other MSH1 proteins (G.M.G., K.M.C. and J.B.H, unpublished

Page 102: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

EI1

observations). It remains possible that new subfamilies of MSH genes will be

identified in eukaryotes.

3.5 Methods

3.5.1 Alignments and Phylogenetic Methods

Several different MutS-like protein sequences were used to search the latest

version of the SWISSPROT protein database, using Blast 2.0 (Altschul et al., 1990).

Using a method for the creation of multiple alignments described previously (Hogweg

and Hesper, 1984), ClustalW and Blast 2.0 alignments of all available MutS-like

protein sequences were combined into a final representative alignment (available upon

request) employing Genetic Data Environment (GDE)(Smith et al., 1994), with the

PAM 250 protein similarity matrix, and used both full-length and individual domains

to identify homologous regions. We excluded some putative proteins identified in

eukaryotic genomes (i.e. C. elegans MSH2, MSH4, MSH6) but not confirmed by

cDNA sequences. Differential shadings of alignments were carried out with

BOXSHADE version 3.03. Bootstrap and phylogenetic reconstruction methods were

performed with PHYLIP version 3.51 (Felsenstein, 1989).

Page 103: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

3.5.2 RASA Analysis

Phylogenetic signal and taxon-variance ratios were determined using RASA

Version 2.2 (Lyons-Weiler et al., 1996; Lyons-Weiler and Hoelzer, 1997). RASA

(Relative Apparent Synapomorphy Analysis) software and documentation for the

Macintosh were downloaded from the internet at http://loco.biologv.unr

edu/archives/rasa/rasa.html, and used to identify long branches, as follows. First, we

calculated for every pair (i,j) of protein sequences two parameters: E, the phenetic

similarity, a measure of the similarity of the i-f pair of sequences, corresponds here to

the total number of identical amino acids (character states) at variable sites in a given

alignment of the two proteins; RASIJ, the pairwise cladistic uniqueness, which reflects

the uniqueness of the i-j pair with respect to the other sequences, corresponds to the

total number of sequences ki,j not showing the shared i-f amino acid, summed over

all positions of i-f identity. For the set of all i-j pairs, RAS,J is on the average expected

to be a linear function of E, (Lyons-Weiler et al., 1996), since the more the positions

of identity scored, the higher the total RASIJ score. Second, in order to define a

statistical measure of phylogenetic signal, the pair of RASIJ-E,J matrices were studied in

the RASA regression, and two taxon-variance terms for each protein sequence (taxon)

i were calculated. The phenetic-variance term VarE(i) is the statistical variance for the

set of E11 values for fixed i, all ji, relative to the mean of those values. The

Page 104: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

phylogenetic (cladistic) variance term VarRAS(i) reflects the summed squared

deviations of RAS,J values from the linear regression line with respect to the E,1 values,

for all (n-i) ji. Comparison of the ratios VarRAS(i)/VarE(i) for all i sequences

provides a means to diagnose long phylogenetic branches, because these two variance

measures are proportionate when the amount of branch-length heterogeneity on the

true tree is low. These measures lose proportionality for long-branch taxa (Lyons-

Weiler and Hoelzer, 1997), because the phenetic variance VarE(i) contributed by the

long-branch taxon is low, but its cladistic variance VarRAS(i) is inflated.

We also used an independent parameter, the test statistic tRASA, to measure

the phylogenetic signal itself (Lyons-Weiler et al., 1996). The tRASA statistic is

generated by using the student's t test to assess the significance of the deviation of the

observed slope of RAS,J vs. E, from a null slope, corresponding to a null hypothesis

that considers the possibility that the character states are distributed randomly among

the taxa. The null slope is calculated by assuming equiprobable distributions of E and

RL4S among all taxa. While significantly positive tRASA values are usually associated

with hierarchical patterns in a character state matrix of the type that is expected when

truly phylogenetic patterns predominate the matrix, significantly negative tRASA

values usually indicate some source of disruption of the hierarchy, such as long-branch

taxa (Lyons-Weiler and Hoelzer, 1997).

Page 105: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

Chapter 4

Arabidopsis thaliana MutS-Homolog Proteins - atMSH2, atMSH3, atMSH6, anda Novel atMSH7 Protein - Form Three Distinct Heterodimers with Different

Specificities for Mismatched DNA.

Kevin M. Culligan and John B. Hays

Published, in part, in The Plant Cell

American Society of Plant Physiologists

June, 2000

Page 106: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

91

4.1 Abstract

Arabidopsis mismatch-repair genes predict MutS -like proteins remarkably

similar to eukaryotic homologues MSH2, MSH3, and MSH6. A novel feature in

Arabidopsis is the presence of two MSH6-like proteins, designated atMSH6 and

atMSH7. Combinations of Arabidopsis atMSH2 with atMSH3, atMSH6, or atMSH7

proteins, products of in vitro transcription and translation, were analyzed for

interactions by analytical gel-filtration chromatography. The atMSH2 protein formed

heterodimers with atMSH3, atMSH6, and atMSH7, but no homodimers of single

proteins were observed. The abilities of the various heterodimers to bind to

mismatched 51-mer duplexes were measured by electrophoretic mobility-shift assays.

Similarly to the corresponding human proteins, atMSH2'atMSH3 heterodimers bound

insertion-deletion loopouts (+AAG, +T) much better than a base-base mispair (T/G),

whereas atMSH2'atMSH6 bound the (T) substrate strongly, (T/G) well, and

(+AAG) no better than an (T/A) homoduplex. However, atMSH2atMSH7 showed a

novel specificity: moderate affinity for a (T/G) substrate and weak binding of (T).

Thus, atMSH2atMSH7 may be specialized for lesions/base mispairs not tested,

and/or (T/G) mispairs in special contexts.

Page 107: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

92

4.2 Introduction

Having previously identified the Arabidopsis atMSH2 gene, we sought to

determine whether the eukaryotic pattern of interaction of MSH2 with other MSH

proteins, to form heterodimers with different substrate specificities, is conserved in

plants. A search for mutS homologs revealed not only atMSH3 and atMSH6, but also

a homolog thus far unique to plants, designated atMSH7. Interaction and binding

studies have demonstrated the potential of the respective atMSH proteins to form three

different heterodimers, each with a different substrate specificity.

4.3 Results

4.3.1 Identification of Arabidopsis MutS-Like Genes

We previously reported the cDNA and genomic sequences of an Arabidopsis

gene that predicts a protein highly similar to other eukaryotic MSH2 proteins, and the

sequence of a PCR product encoding a polypeptide fragment similar to eukaryotic

MSH6 proteins (Culligan and Hays, 1997). We now designate the gene containing

this latter sequence atMSH7 (Accession number AJ193018), for reasons described

below. Two additional mutS-like genes were identified in the Arabidopsis genome

database (Accession numbers AL022197 and AF001308), and corresponding complete

Page 108: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

93

cDNAs were obtained by reverse-transcription PCR, as described under "Methods".

The first proved highly similar to eukaryotic MSH3 genes, and the second highly

similar to MSH6 genes. The sequences of these atMSH3 and atMSH6 cDNAs agree

with cDNA sequences recently deposited in Genebank (Accession numbers AJ007791

and AJ245967, respectively) (Ade et al., 1999).

We isolated a full-length atMSH7 cDNA by screening an Arabidopsis cDNA

library. Figure 12A shows the extensive similarity of human and yeast MSH6 proteins

to Arabidopsis atMSH6, atMSH7, and a recently identified MSH6-like protein

sequence from Zea mays, previously designated MUS2 (Accession number

AJ238786), which we now designate zmMSH7 (see below). atMSH7 and zrnMSH7

are slightly shorter than the other MSH6 proteins, by 57 to 250 amino acids, and show

about 30% identity to atMSH6. In addition to three highly conserved domains

previously identified in MSH proteins (Culligan et al., 2000; Figure 13A), we identify

an additional N-terminal domain (distinct from the N-terminal mismatch-binding

domain) present in both atMSH6 and atMSH7 proteins. Figure 13B shows the

similarity of this domain to N-terminal domains of uracil-DNA glycosylase 2 (UNG2)

and p21 proteins. A recent study identified this domain in UNG2 and p21 as a site of

interaction with PCNA (Proliferating Cell Nuclear Antigen) and RPA (Replication

Protein A) (Otterlei et al., 1999).

Page 109: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique
Page 110: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique
Page 111: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

Figure 14 presents a phylogenetic tree for a representative set of MutS/MSH

protein sequences, obtained by comparing their entire amino-acid sequences, as

described elsewhere (Culligan et al., 2000), but now including atMSH3, atMSH7, and

zmMSH7. The atMSH2, atMSH3, and atMSH6 protein sequences branch with their

respective eukaryotic homolog subfamilies, but atMSH7 and zmMSH7 form a separate

subgroup within the MSH6 radiation. This suggests a common ancestor for MSH7

genes within the MSH6 subfamily, rather than origin of atMSH7 and zmMSH7 from

some other MSH gene. The significant divergence of the MSI{6 and MSH7

sequences, indicated by their relatively long branch lengths, suggests a

duplication/specialization event early, rather than recently, in the evolution of

eukaryotes. The branching pattern of the MSH6 and MSH7 protein sequences in this

tree, as well as a phylogenetic tree produced by analysis of only MSH6 and MSH7

protein sequences (data not shown), suggest that MSH7 diverged from MSH6 before

the divergence of plants and animals. However, a more definitive study of a more

diverse set of sequences will be needed to determine more precisely when the MSH7

gene subfamily diverged from the MSH6 subfamily.

Page 112: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

scMSH5

IN(1 rpMutS

I INhiMutS

ecMutS

:,..----__--__ ___ taMutS

ssMutS

spMutS

bsMutS

hsMSH5

97

Figure 14. Neighbor-joining tree for Dayhoff PAM distances among a representativeset of complete MutS/MSH protein sequences. Gaps and regions of ambiguousalignment were excluded from the analysis. Numbers above each branch represent thenumber of times the branch was found in 100 bootstrap replicas. The Gram-positiveB. subtilis and S. pneumoniae MutS sequences were used as an outgroup. Differentialshadings reflect the known functional role of each major group: light gray, eubacterialmismatch repair and recombination; medium gray, mitochondrial mismatch repair ineukaryotes; black, nuclear mismatch repair and meiotic recombination in eukaryotes.

All eukaryotic homologs (MSH) are encoded by the nuclear genome, except themitochondrially-encoded sgMSH1. bs,=B. subtilis; sp,=S. pneumoniae;ss,Synechocystis sp.; ta,=T. aquaticus; ec,= E. coli; hi,=H. influenzae; av,=A.vinelandii; rp,=R. prowazekii; sc,=S. cerevisiae; sg,=S. glaucum; sp,=S. pombe;zm,=Z. mays; at,=A. thaliana; xl,=X. leavis; hs,=H. sapiens.

Page 113: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

4.3.2 Interactions Between MSH2, MSH3, MSH6 and MSH7 Proteins inTranslation Mixtures.

Because eukaryotic MSH proteins typically act as heterodimers (Modrich and

Lahue, 1996; Acharya et al., 1996), rigorous biochemical analysis of their properties

would require purification from cells in which they may not be abundant, non-

meristematic plant tissues for example, or from heterologous cells in which

simultaneous high expression of the respective recombinant cDNAs was carefully

balanced. In order to make an initial comparison of atMSH7 with other Arabidopsis

MSH proteins, we synthesized the respective polypeptides by in vitro transcription and

translation. Figure 15 shows electrophoretic analysis of 35S-methionine-labeled

products of transcription-translation reactions optimized for balanced synthesis of

pairs of polypeptides expected to form 1:1 heterodimers.

Physical interactions between different MSH polypeptides and aggregation of

individual polypeptides to form homodimers, heterodimers, or higher-order

complexes, as well as possible interactions between MSH polypeptides and other

proteins present in the translation reaction mixtures, were assayed by sedimentation

through sucrose density gradients. Under these conditions (see Methods), no hetero-

or homodimerization of any subset of in vitro co-translated proteins was observed,

possibly because of low protein concentrations of the MSH polypeptides in the

Page 114: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

234i- 205 kD

i:- __ __ 130

4-77 kD

43kD

Figure 15. SDS-PAGE analysis of human and Arabidopsis co-synthesis reaction

mixtures. Lane!, atMSH2 + atMSH3; lane 2, atMSH2 + atMSH6; lane3,atMSH2 + atMSH7; lane 4, hMSH2 + hMSH6. 5 iL of each in vitrotranscription/translation mixture was loaded on a 7.5% SDS-PAGE gel and run at

200Vfor45 mm.

Page 115: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

100

translation mixtures. However, hetero- and homodimerization of several of the

synthesized MSH proteins were observed by gel filtration chromatography, as shown

in Figures 16A, 16B, 16C, and 16D. Since the respective contributions to various

chromatographic fractions of two different polypeptides that are synthesized and

radiolabeled together cannot be distinguished by bulk 355-measurements, fractions of

interest were further analyzed by polyacrylamide-gel electrophoresis in sodium

dodecyl sulfate (SDS-PAGE). When synthesized alone, atMSH2 eluted at a position

near to that expected for the predicted monomer, whereas most atMSH6 eluted in the

void volume, the position expected for an aggregate of roughly 1000 kD (Figure 16A).

However, when approximately equal amounts of the two polypeptides were

synthesized together (Figure 15, lane 2), almost all radioactivity eluted in intermediate

fractions, conesponding roughly to a peak at 400 kD, and containing both atMSH2

and atMSH6 polypeptides (Figure 16A). Human polypeptides hMSH2 and

hMSH6 showed similar heterodimerization when synthesized together (Figure 15,

lane 4). However, hMSH6 synthesized alone eluted at a position corresponding to its

expected monomer molecular weight. When the specificity of these interactions was

tested by synthesis of hMSH2 together with atMSH6 or atMSH2 with hMSH6, no

significant heterodimerization was apparent, i.e. the respective polypeptides eluted as

they did when synthesized alone (Figure 16D). Interestingly, gel filtration analysis of

Page 116: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

A

0

-

VA

101

Column Volume (ml)

p 343638404244464850525456atMSH6 - - - - -atMSH2 - -------atMSH6 - -(ZtMSH2 -

atMSH6atMSH*

(Continued)

Page 117: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

102

B3OOO

V0

+--.O.. atMSH2

ii ::::

U ::::: F'N / /\ \j.

Column Volume (ml)

343638404244464850525456atMSH7 - - - __ -atMSH2

wMSH7arMSH2

arMSH7

atMSH2 -.-

(Continued)

Page 118: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

C25001

20110

e 15000

On. 101)01)

I-

500))

103

Column Volume (ml)

343638404244464850525456atMSH3 - - - - -atMSH2 - - - - -atMSH3 - - ------- - -()/MSH2

atMSH3

atMSH2 * -

(Continued)

Page 119: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

D20000

15000

10000

'.4

1.4

2 5000

1 2

S a1MSH2+hMSH6

hMSH2+atMSH6

N 0 0'

Column Volume (ml)

104

Figure 16. Gel-filtration-chromatography analysis of Arabidopsis 35S-labeled proteins.

(A) Elution profiles for atMSH2, atMSH6, and atMSH2'atMSH6 synthesis mixtures,and corresponding SDS-PAGE autoradiographs of fractions. 50-jil mixture samples

were layered onto the gel-filtration column, fractionated, and analyzed by liquidscintillation and SDS-PAGE (see Methods). Fractions 32-59 for each of the threemixtures are shown in upper (elution profile) panel, and SDS-PAGE analysis ofcorresponding even-numbered fractions (34-56) are shown in lower panels. A smallamount of the original transcription/translation synthesis reaction mixture is shown inthe left-most lane (IVTT lane). Arrow heads to the left of the gel panels denote theposition expected for each polypeptide. Arrows in gray designate theoretical positionsof polypeptides not actually present. V0 arrow denotes the exclusion volume position.(B) Elution profiles and corresponding SDS-PAGE analysis for atMSH2, atMSH7,and atMSH2'atMSH7, obtained as described above. (C) Elution profiles andcorresponding SDS-PAGE analysis for atMSH2, atMSH3, and atMSH2atMSH3,obtained as described above. (D) Elution profiles for co-translation mixtures ofatMSH2+hMSH6 and hMSH2+atMSH6, as described above. Arrow #1 denotes theelution position for atMSH2+atMSH6 heterodimer [see (A)], and arrow #2 denotes the

elution position for hMSH2+hMSH6 heterodimer.

Page 120: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

105

equimolar mixtures of separately synthesized atMSH2 and atMSH6 polypeptides

showed only about 20% as much interaction as when the polypeptides were

synthesized together, as if heterodimerization were more efficient if polypeptides were

able to interact before completing post-translational intra-chain folding (or before

interacting with other proteins in the mixtures).

Figure 16B shows that interaction between atMSH2 and atMSH7 proteins is

similar to that between atMSH2 and atMSH6. All of the atMSH7 peptide synthesized

alone eluted in the void volume (apparent molecular weight greater than 1000 kDa),

but when atMSH2 and atMSH7 were synthesized together (Figure 15, lane 3), most

35S-methionine radiolabel and most atMSH2 polypeptide and atMSH7 polypeptide

eluted near the position expected for a 270-kDa heterodimer. In contrast, atMSH3

synthesized alone eluted as a very broad peak (Figure 16C), corresponding to apparent

molecular weights ranging from 102 to iO kDa. Furthermore, when atMSH2 and

atMSH3 were synthesized and analyzed together (Figure 15, lane 1), there was no

distinct heterodimer peak (Figure 16C), although there was an increase in the amount

of radioactivity in fractions between the peak shown by atMSH2 alone and the

position of the preponderance of atMSH3 synthesized alone. However, the shift in

atMSH2 polypeptide in the presence of atMSH3 to fractions corresponding to higher

molecular weights suggests that atMSH2atMSH3 heterodimers, and perhaps higher-

Page 121: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

106

molecular-weight complexes of the two polypeptides, were formed. Although these

results indicate that atMSH3 polypeptides react with atMSH2 to form specific

heterodimers more weakly, relative to interactions with itself and/or with other

proteins present in the translation mixtures, than do atMSH6 and atMSH7, the

atMSH2atMSH3 mixture showed good substrate-binding activity (see below). A

similar phenomenon might account in part for the low levels of hMSH2hMSH3,

relative to hMSH2'hMSH6, typically found in human cells (Modrich and Lahue,

1996).

In the plot of relative elution position (Kay) versus protein molecular weight

shown in Figure 17, the positions of some expected monomer and heterodimer

positions fall off the line defined by a wide range of putative globular protein

standards. Thus, hMSH2 alone elutes at a slightly lower apparent molecular weight

(-90 kDa) and the atMSH2 monomer at a higher weight (-125 kDa) than those

predicted by their amino-acid sequences (100, 105 kDa respectively). Although

atMSH2atMSH6 elutes at 400 kDa, well beyond the position predicted by its 255

kDa size, hMSH2hMSH6 elutes about as expected, as does atMSH2atMSH7. The

symmetric peaks shown by the anomalously-eluting Arabidopsis proteins, as well as

the contrasting behavior of their human homologs, suggest that their positions reflect

Page 122: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

0.45

0.4

0.35

0.3

0.25

0.2

0. 15-p

Ovalbumin (43kD)

Albumin (67kD)

Standards

hMSH2 (100 kD)

atMSH2 (106 kD)

hMSH6 (153 kD)\\Aldolase (158kD)

a SH2+7 (230 kDhMSH2+6 (253

Ferritin (44OkD)

atMSH2+6 (255kD)

Molecular Weight

107

Figure 17. Kay vs. log MW (Molecular Weight) plot of standards and MSH polypeptidesanalyzed by gel-filtration chromatography. Gray circles denote globular standards usedto calibrate the gel-filtration column. Black diamonds denote MSH polypeptides used inthis study. The predicted (theoretical) molecular weight is shown in parentheses to theright of each monomeric or heterodimeric polypeptide.

Page 123: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

specific asymmetries that affect their hydrodynamic behavior, but we cannot

rigorously rule out less specific interactions with other proteins in the translation

mixture.

4.3.3 Mature atMSH1 Forms a Homodimer

To determine whether the Arabidopsis MSH1 protein forms a homodimer

similar to that of MutS, we inserted the full-length atMSH1 cDNA (including the

predicted N-terminal mitochondrial targeting sequence; generously provided by Dr.

Gilbert Meyer-Gauen) into pGEM3Z, and assayed in vitro atMSH1 synthesis mixtures

for the formation of homodimers and/or higher order complexes, using gel-filtration

chromatography (as described above). Most radioactivity (corresponding to atMSH1)

eluted at a position corresponding to the void volume, suggesting aggregation of

atMSH1 or non-specific interactions within the lysate (Figure 18). Because N-

terminal targeting sequences are usually proteolytically removed upon processing of

proteins imported into mitochondria, it seemed possible that interactions to form

atMSH1 homodimers might be present only in "mature" (processed N-terminal

targeting sequence) proteins. To test this, we constructed a cDNA that deleted the

mitochondrial targeting sequence of atMSH1. Synthesis mixtures programmed with

the mature atMSH] construct were assayed for the formation of atMSH1 homodimers

Page 124: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

15000

10000

U

5000o

0

1 2

2

Column Volume (ml)

109

Figure 18. Elution profile for full-length and N-terminal deletion atMSH1translation mixtures. See Figure 16 for details. Arrow #1 denotes the peak elutionposition for atMSH2+atMSH6 heterodimer (see Figure 16A), and arrow #2 denotesthe peak elution position for hMSH2+hMSH6 heterodimer.

Page 125: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

110

using gel-filtration chromatography as described above. Now, most radioactivity

eluted at the higher molecular weight position, expected for homodimers of atMSH1

(Figure 18). Furthermore, SDS-PAGE analysis of these fractions revealed a shift in

elution of mature atMSH1 (versus full-length atMSH1) to fractions conesponding to

higher molecular-weight complexes. These data suggest that mature atMSHI, but not

full-length atMSHI, forms homodimers.

4.3.4 Binding of atMSH Proteins to Mismatched DNA

We used electrophoretic-mobility-shift assays to compare the abilities of MSH

proteins synthesized in vitro to bind 51 -mer homoduplexes (T/A) or heteroduplexes

(T/G, C/C base/base mispairs or "loopout" insertions of T or AAG in the bottom

strand). The relative intensities of 35S-methionine radioactivity, and knowledge of the

number of methionine residues present in each protein, were used to select synthesis

mixtures in which the pairs of peptides were equimolar. Approximately equal

theoretical amounts of the respective heterodimers were added to the various binding

reactions, and the apparent yields of DNA-protein complexes were normalized for the

predicted heterodimer levels (see Methods). Figure 19 displays representative

electropherograms for three Arabidopsis heterodimers and one human heterodimer,

and Table 4 summarizes the results of conesponding phosphorimager measurements.

Page 126: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

hMSH2hMSH6

111

Figure 19. Representative mobility-shift assays of co-synthesis reaction mixtures ofhuman and Arabidopsis polypeptides. Co-synthesis mixtures (denoted above eachgel panel) were incubated with 32P-labeled homoduplex DNA (T/A, lane 1) orheteroduplex DNA (T/G, C/C, +T IDL, +AAG IDL, lanes 2-5, respectively) andanalyzed on non-denaturing PAGE (see Methods).

Page 127: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

Table 4. Relative binding of MSH heterodimers to oligomer substrates

Heterodimer Oligomer substrate Relative binding Binding ratio

32P-labeled (arbitrary units)a (T/G):(+IDL)"

hMSH2hMSH6

atMSH2'atMSH6

atMSH2'atMSH7

atMSH2atMSH3

T/A

T/G

c/c

+T DL+AAG IDL

T/A

T/G

c/c

+T DL+AAG IDL

T/A

T/G

c/c

+T DL+AAG DL

T/A

T/G

c/c

+T DL+AAG DL

<0.101.00

<0.101.51

<0.10

<0.101.08

<0.102.39

<0.10

<0.100.93

<0.100.23

<0.10

<0.100.27

0.39

1.11

1.20

0.59 ± 0.05

0.57 ± 0.05

5.0 ± 0.5

0.19 ± 0.03

112

asubstrate..binding values correspond to areas of shifted bands (Figure 18) of therespective 32P-labeled oligomer substrates, as determined by phosphorimaging, forapproximately equal amounts of MSH polypeptides. To facilitate comparisons, allphosphorimager values were arbitrarily divided by the value for T/G substrate binding

by hMSH2'hMSH6.bMean ratios (± standard error) are based on densitometry measurements from three

independent mobility-shift-assay experiments (including the experiments

corresponding to values in column 3), calculated by dividing the total densitometry

band area of the shifted 32P-labeled (T/G) oligomer substrate by the total (+1 IDL)

substrate band area, for each respective heterodimer.

Page 128: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

113

The upper bands in each case appear to reflect specific binding by MSH proteins,

since their intensities depend on the natures of both the DNA substrate and the

particular MSH protein heterodimer. The lower bands, which might reflect binding by

proteins specific for dsDNA ends in general (Acharya et al., 1996), were seen in every

gel lane, even when unprogrammed translation mixtures or mixtures containing only

vector DNA (pGEM-3Z) were analyzed (Figure 20). Polypeptides synthesized alone

atMSH2, atMSH3, atMSH6, atMSH7, hMSH2, hMSH6 showed no specific

binding to any of the substrates tested (Figure 20).

Both the Arabidopsis and human MSH2'MSH6 heterodimers showed the

expected preference for a (T/G) base/base mispair and a one-nucleotide (+T) loopout

(Figure 19). As has been shown for the hMSH2hMSH6 heterodimer, the

atMSH2atMSH6 heterodimer bound the one-nucleotide (+T) loopout with greater

efficiency (almost 2-fold here) than the (T/G) base/base mispair (Figure 19; Table 4),

and showed negligible binding to homoduplex (T/A) DNA or to the C/C heteroduplex

or the three-nucleotide (+AAG) loopout. Similarly, the atMSH2atMSH3 heterodimer

showed the same preference for loopout substrates vs. base/base mispair substrate

reported for its human counterpart, hMSH2hMSH3: the atMSH2atMSH3

heterodimer bound best to the (+AAG) loopout and the (+T) loopout, and weakly to

the (C/C) and (T/G) base-mispairs. Again, there was little affinity for (T/A)

Page 129: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

114

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18

Figure 20. Representative mobility-shift assays of single human and Arabidopsissynthesis mixtures. Assays were performed as described in Figure 19 andMethods. Protein synthesis mixtures: lane 1-3, atMSH2; lanes 4-6, atMSH7;lanes 7-9, atMSH6; lanes 10-11, atMSH2 + atMSH6; lanes 12-13, hMSH2 +hMSH6; lanes 14-18, pGEM-3Z vector only (negative control). 32P-labeled oligosubstrates: lanes 1,4,7,10,12, and 14 contain (T/A) homoduplex 5lmers; lanes2,5,8,11,13, and 15 contain (hG) heteroduplex 5lmers; lanes 3,6,9, and 17 contain(+T) heteroduplex 51 mers; lane 16 contains C/C heteroduplex 51 mer; lane 18contains (+AAG) heteroduplex 5 imer.

Page 130: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

115

homoduplex DNA. However, the atMSH2'atMSH7 heterodimer showed a novel

substrate specificity preference for (T/G) base/base mispairs [perhaps slightly

weaker in absolute terms than (T/G) binding by MSH2MSH6 heterodimers] over

(+T) loopouts, and essentially no affinity for other substrates (Figure 19; Table 4).

4.4 Discussion

Searches for Arabidopsis MSH genes have revealed close homologs of genes

in eukaryotic organisms atMSH2, atMSH3, and atMSH6 and an additional novel

MSH6 homolog, designated here atMSH7. Analysis of interactions between

polypeptides translated in vitro demonstrates conservation of the eukaryotic pattern of

heterodimerization of MSH2 with other MSH polypeptides, here including atMSH7.

By analogy with the previous designations of MutScx and MutSi3 for MSH2MSH6 and

MSH2'MSH3 respectively, we designate the MSH2MSH7 heterodimer as MutSy.

Assays of binding to a representative set of mismatched dsDNA oligomers show the

substrate specificity of atMutScx and atMutSi3 to be very similar to the specificities of

their eukaryotic counterparts, but the specificity of atMutSy appears considerably

different from either of these.

The existence of Arabidopsis MutS-like proteins that resemble their eukaryotic

homologs, with respect to both primary structure and activity in vitro, suggests that

Page 131: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

116

plants use classical long-patch mismatch-repair systems to enhance genomic stability.

This inference is strengthened by the identification, via the Arabidopsis genome

project, of MSHJ, MLHJ and PMS2 genes whose cDNAs predict polypeptides highly

similar to their eukaryotic counterparts (G. Meyer-Gauen, A. Tones, J. Leonard,

K.M.C. and J.B.H., unpublished observations). Furthermore, a putative MSH4-like

gene is present in the Arabidopsis genome (K.M.0 and J.B.H., unpublished

observations). Whether plants use MSH proteins to promote meiosis remains to be

determined.

Our phylogenetic analysis suggests that MSH7 diverged from MSH6 early in

eukaryotic evolution, possibly before the divergence of plants and animals. However,

the yeast genome encodes no such second MSH6like protein, and none have been

identified in other organisms, C. elegans for example (K.M.0 and J.B.H., unpublished

observations). MSH7 may thus have a genomic stability role unique to plants, but we

cannot rule out the possibility that some other eukaryotes also encode MSH7 proteins.

MSH7 shows a pattern of charged and aromatic residues in the N-terminal mismatch-

binding domain similar to that of MSH6 proteins (Culligan et al. 2000; Malkov et al.,

1998) and forms heterodimers with atMSH2, as do atMSH3 and atMSH6 proteins.

Additionally, atMSH6 and atMSH7 contain N-terminal interaction domains,

previously identified in several proteins involved in DNA metabolism, including

Page 132: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

117

hMSH6 (Nicolaides et at., 1996). These domains have been shown to be putative sites

of interaction with PCNA and/or RPA (Otterlei et at., 1999). Furthermore, several

other studies in yeast and human cells have implicated PCNA as a factor involved in

both pre- and post-excision steps in mismatch repair (Umar et at., 1996; Johnson et

al., 1996; Chen et al., 1999;). Thus, both atMSH6 and atMSH7 proteins may interact

in some way with the DNA replication machinery. The atMSH2atMSH7

heterodimer (atMutSy) failed to bind strongly to DNA containing (+1) loop-out

mismatches, although these DNAs are good substrates for both MutSu and MutS3.

Even the affinity of atMutSy for (T/G) DNA appeared slightly weaker than that of

atMutSa. This suggests that a major function of atMutSy might be specialized

recognition of DNA lesions, and/or T/G mispairs in specialized contexts.

Purified human MutSu (Mu et at., 1997; Wang et at., 1999) and bacterial MutS

(H. Wang and J.B.H, unpublished observations) bind with considerable affinity to

photoproduct/base matches, e.g. A-G opposite T-T cyblobutanepyrimidine (CPD)

dimers (T<>T) or [6-4] photoproducts (T[6-4]T), but show no preference for

T<>T/AA or T[6-4]T/AA moieties. Thus, mismatch repair might correct potential

mutations caused by insertion of incorrect bases opposite photoproducts during DNA

replication, as genetic evidence suggests is the case in E. coti (Liu et at., 2000). Such

a mechanism might be important for plants, whose DNA can be efficiently replicated

Page 133: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

118

even when photoproduct levels are as high as one per kbp (Draper and Hays, 2000).

Furthermore, optimal transcription-coupled repair of CPDs and most likely, bases

damaged by oxyradicalsin particular thymine glycol, and also possibly 8-

oxoguaninerequires MSH and MILH proteins in yeast and mammalian cells (Leadon

and Avrutskaya, 1997; Leadon and Avrutskaya, 1998; P. K. Cooper and S. A.

Leadon, personal communication). Since plants are typically subject to UV irradiation

and to byproducts of oxidative metabolism, they might utilize MSH7 proteins to

facilitate transcription-coupled repair. With respect to T/G mispairs, one special case

of high potential interest for plants might be spontaneous deamination products of 5-

methylcytosine, i.e. at GIC pairs in CpG and CpNpG sequences, in which cytosines

are frequently methylated in plants (Gruenbaum et al., 1981; Richards, 1997)

The differences between the gel-filtration behavior of various Arabidopsis

MSH proteins when synthesized alone or together in transcription-translation mixtures

may reflect their properties in vivo. For example, the appearance of discrete stable

peaks of hMSH2 or atMSH2, when synthesized alone, appears consistent with the lack

of detectable atMSH2 homodimers in human cells (Drummond et al., 1997). These

monomers show no affinity for mismatched DNA when synthesized in vitro by rabbit

reticulocyte lysates (this study; Acharya et al., 1996). Although synthesis of atMSH2

together with an equal amount of atMSH6 or atMSH7 shifts all of the atMSH2 into a

Page 134: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

119

heterodimer peak, atMSH3 alone, atMSH6 alone, and atMSH7 alone elute in the

exclusion volume, suggesting a high degree of aggregation. Furthermore, both

atMSH3 and atMSH6 show a long heterogeneous tail extending down to the monomer

molecular weight. These data suggest that when atMSH3, atMSH6, or atMSH7

polypeptides, are synthesized alone, non-specific interactions within the translation

mixture may predominate. Thus, proper folding of atMSH3, atMSH6, and atMSH7

polypeptides during synthesis, and perhaps resistance to aggregation and/or

degradation, may require the presence of atMSH2 to engender immediate

heterodimerization. This notion is consistent with the inefficient interaction seen

when polypeptides synthesized separately were mixed together. In human cells, there

appears to be no free hMSH3 or hMSH6 protein not heterodimerized with hMSH2,

and heterodimerization of atMSH2 with atMSH6 appears more efficient than with

atMSH3, consistent with the preponderance of hMutScz over hMutS3 in human cells

(Drummond et al., 1997; Marra et al., 1998).

Elucidation of the roles of plant MSH2, MSH3, MSH6, and MSH7 in

maintaining genomic stability will require construction of plants lacking the respective

proteins, and application of assays for various mismatch-repair functions, including

binding to DNA containing lesions and mismatches in various sequence contexts.

Page 135: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

120

4.5 Methods

4.5.1 Isolation of cDNAs

To obtain Arabidopsis cDNAs corresponding to the coding regions of atMSH3

and atMSH6 genes, available via the Arabidopsis genome project, we employed

Reverse Transcriptase-PCR (RT-PCR). Purified mRNA (Culligan and Hays, 1997)

was amplified using an mRNA PCR Amplification Kit (Clontech), as specified by the

manufacturer. First strand cDNA synthesis products were used as template in standard

PCR reactions to amplify 5' and 3' regions of atMSH3. Primer 5P-MSH3 (5'-GGG

GTA CCA TGG GCA AGC AAA AGC AGC-3'), which overlaps the initiating ATG

and encodes a KpnI site and Kozak consensus sequence, and primer MSH3-6 (5'-AAA

TCT CAG AAA CAG CAT CAA G-3'), corresponding to position +1456 of the

cDNA, were used to amplify the 5' region of atMSH3 cDNA. Similarly, primer 3P-

MSH3 (5'-ACG CUT CGA CTC AAA ATG AAC AAG TTG G-3'), which overlaps

the 3' termination codon and encodes a SphI site, and primer MSH3-2 (5'-AAT CAG

CGC AGG TAA CTT' AGA AG-3'), corresponding to position + 1048 of the cDNA,

were used to amplify the 3' end of atMSH3 cDNA. DNA samples from individual

clones were sequenced by the Oregon State University Central Services Laboratory,

using a Taq dye-Primer/dye-terminator cycle sequencing kit (Applied Biosystems),

Page 136: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

121

and compared to genomic sequences to determine intron/exon boundaries and to verify

the accuracy of the PCR amplifications. Correct 5' and 3' cDNA sequences were

joined via a unique BamHl site present in the overlapping regions, and the product

inserted into the Kpnl and Sphl sites of plasmid pGEM-3Z (Promega).

Arabidopsis MSH6 cDNA sequences were obtained as described for atMSH3,

using primer 5P-MSH6 (5'-GTC GGA TCC GCC ATG GCT CCG TCT CGC CGA-

3'), which overlaps the initiating ATG and encodes a BamHl site and Kozak

consensus sequence, and primer MSH6-9 (5'-ACT TTG CAA ATC TAA GCA GAC

TCT A-3'), corresponding to position +2074 of the cDNA, to amplify the 5' region.

Primer 3P-MSH6 (5'-CTG GTC GAC TTA GTT GGT TAA CCG GAG-3'), which

overlaps the termination codon and contains a Sail site, and primer MSH6-8 (5'-GCT

AAG GTG ITO AGT TAT GCA ACA 0-3'), corresponding to position +1741 of the

cDNA, were used to amplify the 3' region of atMSH6. Correct 5' and 3' clones were

joined via a unique Hindill site in the overlapping region, and the product inserted into

the BamHl and Sail sites of pGEM-3Z.

A full-length atMSH7 cDNA was isolated by probing an Arabidopsis cDNA

library (size range 3-6 kb) in phage XZAPII (Kieber et al. 1993; Arabidopsis

Biological Resource Center), with a 352-bp 32P-labeled PCR fragment of atMSH7

(Culligan and Hays, 1997; Accession number AF009657). The atMSH7 cDNA was

Page 137: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

122

inserted into pGEM-3Z using unique XbaI and Hindu! sites flanking the cDNA. The

cDNA was further altered to encode a Kozak consensus sequence at the 5' end by

PCR amplification using primer 5P-MSH7 (5'-CGG GAT CCT ATA CCA TGG TGC

AGC GCC AGA GAT CG-3'), which overlaps the initiating ATG and encodes a

BamHI site, and primer MSH7-1O (5'-CGA GCT AAT AGC Till' TGC ACT GC-3'),

corresponding to position +1019 of the cDNA. The 5' portion of the cDNA was

replaced with this PCR product, utilizing unique restriction sites.

The atMSH2 cDNA (Culligan and Hays, 1997) was inserted into the EcoRl and

SphI sites of pGEM-3Z. Primer 5P-MSH2 (5'-AGC AAT TOT ATA CCA TOG

AGG GTA All TCG AG-3'), which overlaps the initiating ATG and encodes a Muni

site and Kozak consensus sequence, and primer MSH2-1300 (5'-ACC TCA GAG

AAG CTG GTA ACA GTC-3'), corresponding to position +1763 of the cDNA, were

used to modify the 5' end as described above for atMSH7.

Human MSH2 and MSH6 cDNAs, inserted into plasmids pET-3d and -29a

respectively, were generously provided by Richard Fishel (Acharya et al., 1996).

An Arabidopsis thaliana MSH1 cDNA was provided by Dr. Gilbert Meyer-

Gauen, and inserted into pGEM3Z using unique restriction sites.

Page 138: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

123

4.5.2 Sequence Alignment and Phylogenetic Analysis

Alignments of complete protein sequences, and construction of the

phylogenetic tree, were performed as described previously (Culligan et al., 2000).

4.5.3 In Vitro Transcription and Translation

Plasmids containing cDNAs (see above) were purified from E. coli lysates

using a Plasmid Maxi Kit (Qiagen). In vitro protein synthesis using TnT Quick (rabbit

reticulocyte lysate) kits (Promega) was carried out as specified by the manufacturer, in

50-tl reaction volumes containing 35S-methionine (New England Nuclear). Each

reaction was optimized by varying the amount of template DNA from 0.5 to 2 tg,

and/or varying the Mg2 and K ion concentrations. Reactions were incubated at 30°C

for 90 minutes. For co-synthesis (in vitro transcription/translation of two different

DNA template in the same mixture), a series of mixtures containing slightly different

ratios of the two templates were incubated, and products analyzed by polyacrylamide-

gel electrophoresis in sodium dodecyl sulfate (SDS-PAGE), and visualized by

autoradiography. Mixtures containing equimolar ratios of each polypeptide, as

determined by densitometry with normalization for the respective numbers of

methionines, were used for further analyses. Mixtures were diluted 20-fold in buffer

A (50mM potassium phosphate, pH 7.5; 50 mM NaCl; 0.1 mM EDTA; 1mM DTT;

Page 139: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

124

5% glycerol) and centrifuged through a Centricon- 100 microconcentrator (Millipore)

until volumes were reduce 20-fold. The reconcentrated mixtures were again analyzed

by SDS-PAGE and densitometry, to determine the respective relative amounts of pairs

of polypeptides (potential heterodimers; see Figure 15), and used for subsequent gel-

filtration or mobility-shift-assay experiments.

4.5.4 Analysis of Proteins by Sedimentation Through Sucrose DensityGradients

The various co-synthesis mixtures were centrifuged through a linear 6-20%

sucrose gradient containing buffer A (see above). Samples (100 jtl) were mixed with

approximately 5 tl of 50% glycerol, and layered onto gradients. Centrifugation was

performed in a Beckman SW5O.1 swinging bucket rotor at 45,000 rpm for 20 hrs at

4°C. Fractions (4 drops) were collected from the bottom of the gradient, and used to

determine the amount of incorporated 35S-methionine present, by liquid scintillation

analysis, and/or analyzed by SDS-PAGE and autoradiography.

4.5.5 Gel-Filtration Chromatography

In vitro protein-synthesis mixtures were applied to an 0.7 x 50 cm column of

Sephacryl S300, previously equilibrated with buffer B (50 mM potassium phosphate,

pH 7.5; 150 mM NaC1; 0.1 mM EDTA; 1 mM DTT; 5% glycerol) and calibrated with

Page 140: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

125

ovalbumin (43 kD), albumin (67 kD), aldolase (158 kD), and ferritin (440 kD).

Fractions (225 tl) were collected at an elution rate of 0.08 ml/min. Aliquots were

used to determine the amount of incorporated 35S-methionine present, by liquid

scintillation analysis, and/or analyzed by SDS-PAGE and autoradiography.

4.5.6 Preparation of Oligomer Duplexes

DNA substrates were prepared and purified as described previously (Wang et

al., 1999). Briefly, purified top-strand 51-mers (5'-AAT GGT TAG CAA TCA TAG

TGG CAA GTT GGA GTC AAT CGT CTC TCG TTA TTC-3') were 5'-end-Iabeled

with y32P-ATP in 50-jfl reactions containing 20 pmoles of oligomer, 40 pmoles y32P-

ATP, 10 units T4 polynucleotide kinase, and 1X Kinase Buffer (Gibco BRL), and

incubated for 30 minutes at 37°C. The reaction was terminated by heating to 70°C for

10 minutes. We added, to the 32P-labeled top strand, 20 pmoles of a particular bottom

strand (51-mer T/A, 5'-GAA TAA CGA GAG ACG ATF GAC TCC AAC 1TG CCA

CTA TGA TTG CTA ACC ATT-3'; 51-mer T/G, 5'- GAA TAA CGA GAG ACG

ATT GAC TCC GAC TTG CCA CTA TGA TTG CTA ACC ATT-3'; 51-mer C/C,

5'-GAA TAA CGA GAG ACG ATT GAC TCC AAC TTC CCA CTA TGA TTG

CTA ACC A11I'-3'; 50-mer +T, 5'-GAA TAA CGA GAG ACG AU GAC TCC ACT

TGC CAC TAT GAT TGC TAA CCA TT-3'; 48-mer +AAG, 5'-GAA TAA CGA

Page 141: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

126

GAG ACG All: GAC TCC AAG CCA CTA TGA 11TG CTA ACC ATT-3'), heated

the mixtures to 85°C for 5 minutes, and allowed them to cool to room temperature

over a period of at least six hours. After addition of 0.2 volumes of BND-cellulose,

5M NaCl was added to a final concentration of 1M. Mixtures were incubated 5

minutes, then layered onto Sephadex G-50 Nick Spin Columns (Amersham

Pharmacia Biotech) previously equilibrated with TE buffer, and samples recovered as

specified by the manufacturer. To confirm the absence of single-stranded unannealed

oligomers, a small aliquot of the recovered oligomer sample was retreated with

polynucleotide kinase and 32P-ATP and electrophoresed on 15% polyacrylamide gels,

under conditions which resolve double-stranded from single-stranded 51-mers.

4.5.7 Mobility Shift Assays

Mobility-shift assays were carried out essentially as described previously

(Acharya et al., 1996) in 30-tl reactions containing 50 mM potassium phosphate (pH

7.5), 50 mM NaCl, 0.1 mM EDTA, 1 mM DIT, 5 mM AMP, 10% glycerol, 2 .tg poly

(dI'dC), 0.5 pmoles 32P-labeled oligo, and 8-12 tl of in vitro transcription/translation

mixtures. These mixtures were incubated at 25°C for 30 minutes, loaded onto 4%

polyacrylamide gels containing 2% glycerol and TAE buffer, and electrophoresed at

50 mA for four hours with buffer recirculation. The gels were dried onto 3MM

Page 142: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

127

Whatman paper and visualized by autoradiography at room temperature, or analyzed

using a Molecular Dynamics Phosphorimager.

Page 143: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

128

Chapter 5

Conclusions

Kevin M. Culligan

Page 144: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

129

5.1 Summary

Upon initiation of this work, most knowledge of eukaryotic mismatch repair

systems was limited to either yeast or mammalian cells. There was essentially no

information as to whether plants use conserved mismatch-repair systems found in

other eukaryotes. Because fungi are evolutionarily more distant from plants than

animals, it seemed possible that the early eukaryotic cell(s) that gave rise to green

plants could have encoded fewer MSH and/or MLH proteins, thus producing a

different, yet related, set of MSH and MLH proteins. Thus, we first attempted to

identify a subset of MSH genes that would suggest whether similar eukaryotic

mismatch repair systems were also found in plants. If so, then additional experiments

could be conducted to determine if these genes encoded conserved proteins with

conserved functions.

The work described in this thesis clearly suggest that plants require conserved

mismatch-repair systems to maintain the stability of their genomes: (1) The results

described in Chapter 2 suggested that plants encode at least two MSH proteins, and

these proteins clearly fell into their respective MSH subfamilies based on sequence

conservation. Therefore, plants do encode evolutionarily conserved MSH proteins.

(2) Through the use of rigorous phylogenetic techniques, we were able to construct

evolutionary trees that suggested all known MSH proteins (both plants and animals)

Page 145: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

130

are related to a common eukaryotic MSH ancestor (Chapter 3). Interestingly, this

common ancestor is MSHJ, which itself appears to be the direct descendent of the

MutS of c*-proteobacteria hypothesized to have given rise to eukaryotic mitochondria.

Elucidation of this rigorous phylogenetic analysis further allowed the development of

techniques to accurately predict MSH subfamilies and possible functions of newly

discovered MSH genes. Indeed, these techniques were used to predict the novel

MSH7 subfamily, thus far unique to plants. (3) Chapter 4 describes the initial

biochemical characterization of Arabidopsis MSH proteins. The results clearly

suggest that plants retain specific biochemical functions found in other eukaryotic

cells (formation of MutSc and MutSI3 heterodimers and their respective binding

activities), but plants encode an additional heterodimer with a novel function.

Therefore, plants may require enhanced genomic fidelity functions beyond other

eukaryotes (yeast and mammalian cells, for example).

5.2 Recommendations for Future Research

Further rigorous experimentation will be required in order to unambiguously

determine the biochemical functions of the Arabidopsis MSH proteins. In vitro co-

translated MSH heterodimers can be further tested for binding to substrates containing

various mismatches/lesions, including all possible base-base mismatches, 5-

Page 146: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

131

methylcytosine CpG islands, cyclobutane pyrimidine dimers, 8-oxo-guanine, and so

on. However, due to the heterogeneous nature of the in vitro extract mixtures, near

absolute binding efficiencies will be difficult to determine accurately. Thus,

purification of Arabidopsis MSH heterodimers remains most attractive. Human

MutSu has been purified directly from mismatch repair proficient cell lines

(Drummond et al., 1994), as well as from recombinant sources, such as from insect

cells (Gradia et al., 1997). However, purification of Arabidopsis MSH heterodimers

from plant tissue may prove difficult. A recent study (Ade et al., 1999) suggests that

MSH transcripts are detectable (on northern-blot analysis) in suspension cultures, but

are undetectable in non-meristematic tissues, such as leaves. A more plausible

purification approach would be carefully balanced, simultaneous high expression of

the respective recombinant cDNAs in heterologous cells (insect cells for example).

Another purification alternative may be to produce the proteins in vitro, using a novel

wheat germ lysate-translation system (Madin et al., 2000).

Information gained from biochemical characterization will fit well with

ongoing experiments in this laboratory. We are currently developing GUS-reversion

assay systems to determine mutation rates in planta, which will be instrumental in

correlating in vitro biochemical observations with those in vivo. Briefly, plants

transformed with T-DNAs harboring out-of-frame GUS constructs [ATG-(G)7-GUS

Page 147: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

132

for example] are scored for their frequency of reverent blue sectors, in stained tissues,

to determine spontaneous mutation rates. To study mismatch repair effects on

mutation frequencies, we seek to identify Arabidopsis plants defective in mismatch

repair. We are currently searching for null (T-DNA tagged) mutations in MSH and

MLH-like proteins, by PCR-screening Arabidopsis T-DNA tagged libraries.

Furthermore, we are constructing "dominant-negative" lines by expressing altered

Arabidopsis cDNAs that mimic known dominant-negative mutations in other

organisms. These dominant negative lines are advantageous because, unlike some

anti-sense constructs, they should be stable and engender tightly-defined phenotypes,

and they are easily introgressed into other Arabidopsis lines, such as the quartet (qrt)

mutant background, which is currently being used in Arabidopsis for tetrad analysis

(Preuss et al., 1994; Copenhaver et al., 1998). Furthermore, the dominant-negative

approach will allow the creation of mismatch-repair-deficiency phenotypes in other

plants, such as important crop species.

Page 148: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

133

BIBLIOGRAPHY

Acharya, S., Wilson, T., Gradia, S., Kane, M. F., Guerrette, S., Marsischky, G. T.,

Kolodner, R., and Fishel, R. 1996. hMSH2 forms specific mispair-bindingcomplexes with hMSH3 and hMSH6. Proc. Nati. Acad. Sci. USA 93, 13629-

13634.

Ade, J., Beizile, F., Philippe, H., and Doutriaux, M.P. 1999. Four mismatch repairparalogues coexist in Arabidopsis thaliana: AtMSH2, AtMSH3, AtMSH6-1and AtMSH6-2. Mol. Gen. Genet. 262, 239-249.

Allen, D. J., Makhov, A., Grilley, M., Taylor, J., Thresher, R., Modrich, P., and

Griffith, J. D. 1997. MutS mediates heteroduplex loop formation by a

translocation mechanism. EMBO J. 16,4467-4476.

Altschul, S. F., Gish, W., Miller, W., Myers, E. W. and Lipman, D. J. 1990. Basiclocal alignment search tool. J. Mo!. Biol. 215, 403-410.

Andersson, S.G., Zomorodipour, A., Andersson, J.O., Sicheritz-Ponten, T., Alsmark,

U.C., Podowski, R.M., Naslund, A.K., Eriksson, A.S., Winkler, H.H., and

Kurland, C.G. 1998. The genome sequence of Rickettsia prowazekii and theorigin of mitochondria. Nature 396, 133-140.

Baas, P.D., and Jansz, H.S. 1972. Asymmetric information transfer during phi X174DNA replication. J. Mol. Biol. 63:557-568.

Bauer, J., Krammer, G., and Knippers, R. 1981. Asymmetric repair of bacteriophageT7 heteroduplex DNA. Mol. Gen. Genet. 181:541-547.

Bevan, M., Bancroft, I., Bent, E., Love, K., Goodman, H., Dean, C., Bergkamp, R.,

Dirkse, W., Van Staveren, M., Stiekema, W., et al. 1998. Analysis of 1.9 Mbof contiguous sequence from chromosome 4 of Arabidopsis thaliana. Nature

391, 485-488.

Page 149: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

134

Bronner, C.E., Baker, S.M., Momson, P.T., Warren, G., Smith, L.G., Lescoe, M.K.,Kane, M., Earabino, C., Lipford, J., Lindblom, A., et al. 1994. Mutation inthe DNA mismatch repair gene homologue hMLH1 is associated with

hereditary non-polyposis colon cancer. Nature. 368:258-261.

Brown, T.C., and Jiricny, J. 1988. Different base/base mispairs are corrected with

different efficiencies and specificities in monkey kidney cells. Cell 54:705-

711.

Bruni, R., Martin, D., and Jiricny, J. 1988. d(GATC) sequences influenceEscherichia coli mismatch repair in a distance-dependent manner frompositions both upstream and downstream of the mismatch. Nucleic Acids Res.

16: 4875-4890.

Bullock, W.O., Fernandez, J.M., and Short, J.M. 1987. XL 1-BLUE: A high efficiency

plasmid transforming recA E. co/i strain with beta-galactosidase selection.Biotechniques. 5: 376-379.

Chen, C., Merrill, B.J., Lau, P.J., Hoim, C., and Kolodner, RD. 1999. Saccharomycescerevisiae p0130 (proliferating cell nuclear antigen) mutations impair

replication fidelity and mismatch repair. Mo!. Cell Biol. 19, 7801-7815

Chi, N.-W., and Kolodner, R. D. 1994a. Purification and characterization of MSH1, ayeast mitochondrial protein that binds to DNA mismatches. J. Biol. Chem.

269, 29984-29992.

Chi, N.-W., and Kolodner, R. D. 1994b. The effect ofDNA mismatches on the

ATPase activity of MSH1, a protein in yeast mitochondria that recognizes

DNA mismatches. J. Biol. Chem. 269, 29993-29997.

Cooper, D.L., Lahue, R.S., and Modrich, P. 1993. Methyl-directed mismatch repair isbidirectional. J. Biol. Chem. 268:11823-11829.

Copenhaver, G.P., Browne, W.E., and Preuss, D. 1998. Assaying genome-widerecombination and centromere functions with Arabidopsis tetrads. Proc. Nat!.

Acad. Sci. U. S. A. 95(1):247-252.

Page 150: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

135

Cox, E.C. 1976. Bacterial mutator genes and the control of spontaneous mutation.

Annu Rev Genet. 10:135-156.

Culligan, K. M., and Hays, J. B. 1997. DNA mismatch repair in plants. AnArabidopsis thaliana gene that predicts a protein belonging to the MSH2

subfamily of eukaryotic MutS homologs. Plant Physiol. 115, 833-839.

Culligan, K.M., Meyer-Gauen, G., Lyons-Weiler, J., and Hays, J.B. 2000.Evolutionary origin, diversification, and specialization of eukaryotic MutS-Homolog mismatch-repair proteins. Nucleic Acids Res. 28, 463-47 1.

De Las Alas, M. M., De Bruin, R. A. M., Eyck, L. T., Los, G., and Howell, S. B.1998. Prediction-based threading of the hMSH2 DNA mismatch repair

protein. FASEB J. 12: 653-663.

de Vries, S.S., Baart, E.B., Dekker, M., Siezen, A., de Rooij, D.G., de Boer P., and te

Riele, H. 1999. Mouse MutS-like protein MshS is required for properchromosome synapsis in male and female meiosis. Genes Dev. 13: 523-53 1.

de Wind, N., Dekker, M., Bems, A., Radman, M., and te Riele, H. 1995. Inactivationof the mouse Msh2 gene results in mismatch repair deficiency, methylationtolerance, hyperrecombination, and predisposition to cancer. Cell 82:32 1-330.

DePamphilis, M.L., and Wassarman, P.M. 1980. Replication of eukaryoticchromosomes: a close-up of the replication fork. Annu. Rev. Biochem. 49:627-

666.

Dohet, C., Dzidic, S., Wagner, R., and Radman, M. 1987. Large non-homology inheteroduplex DNA is processed differently than single base pair mismatches.

Mol Gen Genet. 206:181-184.

Dohet, C., Wagner, R., and Radman, M. 1986. Methyl-directed repair of frameshiftmutations in heteroduplex DNA. Proc. Natl. Acad. Sci. U S A. 83:3395-3397.

Doolittle, W. F. 1998. You are what you eat: a gene transfer ratchet could account for

bacterial genes in eukaryotic nuclear genomes. Trends Genet. 14, 307-311.

Page 151: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

136

Draper, C.K., and Hays, J.B. 2000. Replication of chloroplast, mitochondrial andnuclear DNA during growth of unirradiated and UVB-irradiated Arabidopsis

leaves. Plant Journal, in press.

Drummond, J.T., Anthony, A., Brown, R., and Modrich, p. 1996. Cisplatin andadriamycin resistance are associated with MutLu and mismatch repair

deficiency in an ovarian tumor cell line. J. Biol. Chem. 271, 19645-19648.

Drummond, J.T., Genschel, J., Wolf, E., and Modrich, p. 1997. DHFR/MSH3amplification in methotrexate-resistant cells alters the hMutSa/hMutSI3 ratio

and reduces the efficiency of base-base mismatch repair. Proc. Nati. Acad.

Sci. USA 94, 10144-10149.

Drummond, J.T., Li, G-M., Longley, M.J., and Modrich, p. 1995. Isolation of anhMSH2-p 160 heterodimer that restores DNA mismatch repair to tumor cells.

Science 268:1909-1912.

Duckett, D.R., Drummond, J.T., Murchie, A.I.H., Reardon, J., Sancar, A., Lilley,D.M.J., and Modrich, p. 1996. Human MutSa recognizes damaged DNA

base pairs containing 06-methylguanine, 04-methylthymine, or the cisplatin-

d(GpG) adduct. Proc. Nail. Acad. Sci. USA 93, 6443-6447.

Echols, H., Goodman, M.F. 1991. Fidelity mechanisms in DNA replication. Annu Rev

Biochem. 60: 477-511.

Edelmann, W., Umar, A., Yang, K., Heyer, J., Kucherlapati, M., Lia, M., Kneitz, B.,Avdievich, E., Fan, K., Wong, E., Crouse, G., Kunkel, T., Lipkin, M.,Kolodner, R.D., and Kucherlapati, R. 2000. The DNA mismatch repair genesMsh3 and Msh6 cooperate in intestinal tumor suppression. Cancer Res.

sI:{II:{YJ

Edelmann, W., Yang, K., Umar, A., Heyer, J., Lau, K., Fan, K., Liedtke, W., Cohen,

P.E., Kane, M.F., Lipford, J.R., Yu, N., Crouse, G.F., Pollard, J.W., Kunkel,

T., Lipkin, M., Kolodner, R., and Kucherlapati, R. 1997. Mutation in the

mismatch repair gene Msh6 causes cancer susceptibility. Cell 91:467-477.

Page 152: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

137

Edenberg, H.J., Anderson, S., and DePamphilis, M.L. 1978. Involvement of DNApolymerase alpha in simian virus 40 DNA replication. J. Biol. Chem.

253:3273-3280.

Eisen, J. A. 1998. A phylogenomic study of the MutS family of proteins. Nucleic

Acids Res. 26, 429 1-4300.

Eisen, J. A., Kaiser, D., and Myers, R. M. 1997. Gastrogenomic delights: a movable

feast. Nature Med. 3:1076-1078.

Fang, W.H., and Modrich, P. 1993. Human strand-specific mismatch repair occursby a bidirectional mechanism similar to that of the bacterial reaction. J. Biol.

Chem. 268: 11838-11844.

Fang, W.H., and Modrich, P. 1993. Human strand-specific mismatch repair occurs bya bidirectional mechanism similar to that of the bacterial reaction. J. Biol.

Chem. 268:11838-11844.

Felsenstein, J. 1989. PHYLIP, Phylogeny Inference Package (Version 3.2) Cladistics

5: 164-166.

Feng W-Y., Lee, E., Hays, J.B. 1991. Recombinagenic processing of UV-lightphotoproducts in nonreplicating phage DNA by the Escherichia co/i methyl-directed mismatch repair system. Genetics. 129: 1007-1020.

Fishel, R., and Wilson, T. 1997. MutS homologs in mammalian cells. Curr. Opin.Genet. Dev. 7: 105-113.

Fishel, R., Lescoe, M.K., Rao, M.R., Copeland, N.G., Jenkins, N.A., Garber, J., Kane,M., and Kolodner, R. 1993. The human mutator gene homolog MSH2 and itsassociation with hereditary nonpolyposis colon cancer. Cell 75:1027-1038.

Fujii, H., and Shimada, T. 1989. Isolation and characterization of cDNA clones

derived from the divergently transcribed gene in the region upstream from the

human dihydrofolate reductase gene. J. Biol. Chem. 264:10057-10064.

Page 153: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

138

Glickman, B.W., and Radman, M. 1980. Escherichia coli mutator mutants deficientin methylation-instructed DNA mismatch correction. Proc. Nati. Acad. Sci. U

SA. 77:1063-1067.

Gradia, S., Acharya, S., and Fishel, R. 1997. The human mismatch recognitioncomplex hMSH2-hMSH6 functions as a novel molecular switch. Cell 91:995-

1005.

Gradia, S., Subramanian, D., Wilson, T., Acharya, S., Makhov, A., Griffith, J., and

Fishel, R. 1999. hMSH2-hMSH6 forms a hydrolysis-independent slidingclamp on mismatched DNA. Mol. Cell 3:255-261.

Gray, M. W., and Spencer, D. F. 1996. in Evolution ofMicrobial Life (eds Roberts,

D. McL., Sharp, P., Alderson, G. & Collins, M.). Cambridge University Press,

pp. 109-126.

Grilley, M., Griffith, J., and Modrich, P. 1993. Bidirectional excision in methyl-directed mismatch repair. J. Biol. Chem. 268:11830-11837.

Grilley, M., Welsh, K.M., Su, S.S., and Modrich, P. 1989. Isolation andcharacterization of the Escherichia coil mutL gene product. J. Biol. Chem.

264:1000-1004.

Gruenbaum, Y., Naveh-Many, T., Cedar, H., and Razin, A. 1981. Sequencespecificity of methylation in higher plant DNA. Nature 292:860-862.

Gupta, R. S., and Golding, G. B. 1996. The origin of the eukaryotic cell. Trends

Biochem. Sci. 21:166-171.

Habraken, Y., Sung, P., Prakash, L. and Prakash, S. 1998. ATP-dependent assemblyof a ternary complex consisting of a DNA mismatch and the yeast MSH2-MSH6 and MLH1-PMS1 protein complexes. J. Biol. Chem. 273:9837-984 1.

Habraken, Y., Sung, P., Prakash, L., and Prakash, S. 1996. Binding ofinsertion/deletion DNA mismatches by the heterodimer of yeast mismatchrepair proteins MSH2 and MSH3. Current Biology. 6: 1185-1187.

Page 154: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

139

Hoffman, P.D., Batschauer, A., and Hays, J.B. 1996. PHH1, a novel gene fromArabidopsis thaliana that encodes a protein similar to plant blue-lightphotoreceptors and microbial photolyases. Mol Gen Genet. 253: 259-265.

Hogweg, P., and Hesper, B. J. 1984. The alignment of sets of sequences and theconstruction of phylogenetic trees: an integrated method. J. Mol. Evol. 20: 175-

186.

Hollingsworth, N. M., Ponte, L., and Halsey, C. 1995. MSH5, a novel MutS homolog,facilitates meiotic reciprocal recombination between homologs inSaccharomyces cerevisiae but not mismatch repair. Genes Dev. 9:1728-1739.

Hollingsworth, N. M., Ponte, L., and Halsey, C. 1995. MSH5, a novel MutS homolog,facilitates meiotic reciprocal recombination between homologs inSaccharomyces cerevisiae but not mismatch repair. Genes Dev. 9:1728-1739.

Holmes, J. Jr., Clark, S., and Modrich, P. 1990. Strand-specific mismatch correctionin nuclear extracts of human and Drosophila melanogaster cell lines. Proc.

Nati. Acad. Sci. USA. 87:5837-5841.

Hunter, N., and Borts, R. H. 1997. MIhI is unique among mismatch repair proteins inits ability to promote crossing-over during meiosis. Genes Dev. 11:1573-1582.

Hunter, N., Chambers, S.R., Louis E.J., and Borts, RH. 1996. The mismatch-repairsystem contributes to meiotic sterility in an interspecies yeast hybrid. EMBO

J. 15:1726-1733.

Johnson, R.E., Kovvali, G.K., Guzder, S.N., Amin, N.S., Holm, C., Habraken, Y.,Sung, P., Prakash, L., and Prakash, S. 1996. Evidence for involvement ofyeast proliferating cell nuclear antigen in DNA mismatch repair. J. Biol.

Chem. 271:27987-27990.

Johnson, R.E., Kovvali, G.K., Prakash, L., and Prakash, S. 1996. Requirement of the

yeast MSH3 and MSH6 genes for MSH2-dependent genomic stability. J. Biol.

Chem. 271: 7285-7288.

Jones, M., Wagner, R., and Radman, M. 1987. Mismatch repair and recombination in

E. coli. Cell 50:621-626.

Page 155: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

140

Kaneko, T., Sato, S., Kotani, H., Tanaka, A., Asamizu, E., Nakamura, Y., Miyajima,N., Hirosawa, M., Sugiura, M., Sasamoto, S., Kimura, T., Hosouchi, T.,Matsuno, A., Muraki, A., Nakazaki, N., Naruo, K., Okumura, S., Shimpo, S.,

Takeuchi, C., Wada, T., Watanabe, A., Yamada, M., Yasuda, M., and Tabata,

S. 1996. Sequence analysis of the genome of the unicellular cyanobacteriumSynechocystis sp. strain PCC6803. II. Sequence determination of the entiregenome and assignment of potential protein-coding regions. DNA Res. 3:109-

136.

Kat, A., Thilly, W.G., Fang, W.-h., Longley, M.J., Li, G.-M., and Modrich, P. 1993.An alkylation-tolerant, mutator human cell line is deficient in strand-specific

mismatch repair. Proc. Nati. Acad. Sci. USA 90:6424-6428.

Kieber, J.J., Rothenberg, M., Roman, G., Feldmann, K.A., and Ecker, J.R. 1993.CTR 1, a negative regulator of the ethylene response pathway in Arabidopsis,

encodes a member of the raf family of protein kinases. Cell 72:427-441.

Klekowski, E.J. 1988. Mutation, Developmental Selection and Plant Evolution.Columbia University Press, New York, New York.

Klekowski, E.J. 1997. Somatic mutation theory of clonality. In The Ecology andEvolution of Clonal Plants. (H. deKroon and J. van Groenendael, Eds.).Backbuys Publishers, Leyden, The Netherlands.

Kolodner, R. 1996. Biochemistry and genetics of eukaryotic mismatch repair. Genes

Dev. 10:1433-1442.

Kolodner, R., and Marsischky, G.T. 1999. Eukaryotic DNA mismatch repair. Curr.

Opin. Genet. Dev. 9:89-96.

Kornberg, A., and Baker, T.A. 1992. DNA Replication, pp. 497-498, W.H. Freeman

& Co., New York.

Kramer, B., Kramer, W., and Fritz, H.J. 1984. Different base/base mismatches arecorrected with different efficiencies by the methyl-directed DNA mismatch-

repair system of E. coli. Cell 3 8:879-887.

Page 156: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

141

Kramer, B., Kramer, W., Williamson, M.S., and Fogel, S. 1989. Heteroduplex DNAcorrection in Saccharomyces cerevisiae is mismatch specific and requires

functional PMS genes. Mol. Cell Biol. 9:4432-4440.

Lahue, R.S., Au, K.G., and Modrich, p. 1989. DNA mismatch correction in a defined

system. Science 245:160-164.

Lea, D.E., and Coulson, C.A. 1949. The distribution of numbers of mutants in

bacterial populations. J. Genet. 49: 264-285.

Leach, F.S., Nicolaides, N.C., Papadopoulos, N., Liu, B., Jen, J., Parsons, R.,

Peltomaki, P., Sistonen, P., Aaltonen, L.A.,Nystrom-Lahti, M., et al. 1993.

Mutations of a mutS homolog in hereditary nonpolyposis colorectal cancer.

Cell 75:1215-1225.

Leadon, S.A., and Avrutskaya, A.V. 1997. Differential involvement of the human

mismatch repair proteins, hMILH1 and hMSH2, in transcription-coupled repair.

Cancer Res. 57: 3784-3791.

Leadon, S.A., and Avrutskaya, A.V. 1998. Requirement for DNA mismatch repairproteins in the transcription-coupled repair of thymine glycols in

Saccharomyces cerevisiae. Mutat. Res. 407:177-187.

Li, G.-M., and Modrich, p. 1995. Restoration of mismatch repair to nuclear extractsof H6 colorectal tumor cells by a heterodimer of human MutL homologs. Proc.

Nati. Acad. Sci. USA 92:1950-1954.

Li, G.-M., Wang, H., and Romano, L.J. 1996. Human MutSa specifically binds to

DNA containing aminofluorene and acetylaminofluorene adducts. J. Biol.

Chem. 271:24084-24088.

Liu, H., Hewitt, S., and Hays, J.B. 1999. Antagonism of ultraviolet-light mutagenesisby the MutHLS mismatch repair system of Escherichia coli. Genetics

154:503-5 12.

Loeb, L.A. 1991. Mutator phenotype may be required for multistage carcinogenesis.

Cancer Res. 51: 3075-3079.

Page 157: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

142

Longley, M.J., Pierce, A.J., and Modrich, P. 1997. DNA polymerase delta is requiredfor human mismatch repair in vitro. JBiol Chem. 272: 10917-10921.

Lu, A.L., Clark, S., and Modnch, P. 1983. Methyl-directed repair of DNA base-pairmismatches in vitro. Proc. Nati. Acad. Sci. U S A. 80:4639-4643.

Lu, A.L., Welsh, K., Clark, S., Su, S.S., and Modrich, P. 1984. Repair of DNA base-pair mismatches in extracts of Escherichia coli. Cold Spring Harb. Symp.

Quant. Biol. 49:589-596.

Lyons-Weiler, J. and Hoelzer, G. A. 1997. Escaping from the Felsenstein zone bydetecting long branches in phylogenetic data. Mol. Phylogen. Evol. 8:375-384.

Lyons-Weiler, J., Hoelzer, G. A., and Tausch, R. J. 1998. Optimal outgroup analysis.

Biol. J. Linn. Soc. 64:493-511.

Lyons-Weiler, J., Hoelzer, G. A., and Tausch, R. J. 1996. Relative apparentsynapomorphy analysis (RASA). I: The statistical measurement of

phylogenetic signal. Mol. Biol. Evol. 13:749-757.

Madin, K., Sawasaki, T., Ogasawara, T., and Endo, Y. 2000. A highly efficient androbust cell-free protein synthesis system prepared from wheat embryos: plantsapparently contain a suicide system directed at ribosomes. Proc. Nail. Acad.

Sci. U S A. 97:559-564.

Maidak, B. L., Olsen, g. J., Larsen, N., Overbeek, R., McCaughey, M. J., and Woese,C. R. 1997. The RDP (Ribosomal Database Project). Nucleic Acids Res.

25: 109-111.

Malkov, V. A., Biswas, I., Camerini-Otero, R. D., and Hsieh, P. 1998. Photocross-linking of the NH2-terminal region of Taq MutS protein to the major groove ofa heteroduplex DNA. J. Biol. Chem. 272:238 1 1-238 17.

Maniatis T.,, Fritsch E.F., Sambrook, J., 1982. Molecular Cloning A Laboratory

Manual. Cold Spring Harbor Laboratory, Cold Spring Harbor, New York.

Margulis, L. (1970) Origin ofEukaryotic Cells (Yale Univ. Press, New Haven, CT).

Page 158: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

143

Marra, G., laccarino, I., Leuien, T., Roscilli, G., Delmastro, P., and Jiricny, J. 1998.Mismatch repair deficiency associated with overexpression of the MSH3 gene.

Proc. Nat!. Acad. Sci. USA 95:8568-8573.

Marsischky, G. T., Filosi, N., Kane, M. F., and Kolodner, R. 1996. Redundancy ofSaccharomyces cerevisiae MSH3 and MSH6 in MSH2-dependent mismatch

repair. Genes Dev. 10:407-420.

Martin, W., and Muller, M. 1998. The hydrogen hypothesis for the first eukaryote.

Nature 392:37-41.

Martin, W., Stoebe, B., Goremykin, V., Hansmann, S., Hasegawa, M., and Kowallik,

K. V. 1998. Gene transfer to the nucleus and the evolution of chioroplasts.

Nature 393:162-165.

McGraw, B.R., and Marinus, M.G. 1980. Isolation and characterization of Dam+revertants and suppressor mutations that modify secondary phenotypes of dam-

3 strains of Escherichia coli K-12. Mo!. Gen. Genet. 178:309-3 15.

Modrich, P. 1991. Mechanisms and biological effects of mismatch repair. Annu. Rev

Genet. 25, 229-253.

Modrich, P., and Lahue, R. 1996. Mismatch repair in replication fidelity, geneticrecombination, and cancer biology. Annu. Rev. Biochem. 65:101-133.

Mogen, B., Graybosch, R., MacDonald, M., and Hunt, A.G. 1990. Upstream elementsother than AAUAAA are required for efficient messenger RNA 3'-endformation in plants. The Plant Cell. 2: 1261-1272.

Mu, D., Tursun, M., Duckett, D.R., Drummond, J.T., Modrich, P., and Sancar, A.1997. Recognition and repair of compound DNA lesions (base damage andmismatch) by human mismatch repair and excision repair systems. Mo!. Cell.

Biol. 17:760-769.

Murashige, T., and Skoog, F. 1962. A revised medium for rapid growth and bioassaywith tobacco tissue culture. Physiol Plant. 15:473-497.

Page 159: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

144

Murray, M.G., and Thompson, W.F. 1980. Rapid isolation of high molecular weightplant DNA. Nucleic Acids Res. 8: 4321-4325.

Nene, V., and Glass, R.E. 1982. Genetic studies of the p3-subunit of Escherichia coli

DNA polymerase I. The effect of known, single amino acid substitutions in an

essential protein. Mol. Gen. Genet. 239:97-108.

Nevers, P., and Spatz, H.C. 1975. Escherichia coli mutants uvrD and uvrE deficientin gene conversion of lambda-heteroduplexes. Mol. Gen. Genet. 139:233-243.

Nicolaides, N.C., Palombo, F., Kinzler, K.W., Vogelstein, B., and Jiricny, J. 1996.

Molecular cloning of the N-terminus of GTBP. Genomics 31:395-397.

Otterlei , M., Warbrick, E., Nageihus, T.A., Haug, T, Slupphaug, G., Akbari, M.,Aas, P.A., Steinsbekk, K., Bakke, 0., and Krokan, H.E. 1999. Post-replicativebase excision repair in replication foci. EMBO J. 18:3834-3844.

Palmer, J. D. 1997. Organelle genomes: going, going, gone! Science 275:790-791.

Palombo ,F., Gallinari, P., laccarino, I., Lettieri, T., Hughes, M., D'Arrigo, A., Truong,

0., Hsuan, J.J., and Jiricny, J. 1995. GTBP, a 160-kilodalton protein essentialfor mismatch-binding activity in human cells. Science 268: 1912-1914.

Palombo, F., laccarino, I., Nakajima, E., Ikejma, M., Shimada, T., and Jiricny, J. 1996.hMutSI3, a heterodimer of hMSH2 and hMSH3, binds to insertion-deletion

loops in DNA. Current Biology 6:1181-1184.

Papadopoulos, N., Nicolaides, N.C., Liu, B., Parsons, R., Lengauer, C., Palombo, F.,

D'Arrigo, A., Markowitz, S., Willson, J.K., Kinzler, K.W., et al. 1995.

Mutations of GTBP in genetically unstable cells. Science 268:1915-1917.

Petit, M.-A., Dimpfl, J., Radman, M., and Echols, E. 1991. Control of large

chromosomal duplications in Escherichia coli by the mismatch repair system.

Genetics 129:327-332.

Pochart, P., Woltering, D., and Hollingsworth, N.M. 1997. Conserved propertiesbetween functionally distinct MutS homologs in yeast. J. Biol. Chem.

272:30345-30349.

Page 160: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

145

Pont-Kingdon G., Okada N.A., Macfarlane J.L., Beagley C.T., Watkins-Sims C.D.,Cavalier-Smith T., Clark-Walker G.D., and Woistenholme D.R. 1998.Mitochondrial DNA of the coral Sarcophyton glaucum contains a gene for ahomologue of bacterial MutS: a possible case of gene transfer from the nucleus

to the mitochondnon. J. Mol. Evol. 46:419-431.

Pont-Kingdon, G. A., Okada, N. A., Macfarlane, J. L., Beagley, C. T., Woistenholme,D. R., Cavalier-Smith, T., and Clark-Walker, G. D. 1995. A coral

mitochondrial mutS gene. Nature 375:109-111.

Preuss, D., Rhee, S.Y., and Davis, R.W. 1994. Tetrad analysis possible in

Arabidopsis with mutation of the QUARTET (QRT) genes. Science 264:1458-

1460.

Prolla, T. A., Pang, Q., Alani, E., Kolodner, R. D., and Liskay, R. M. 1994. MILHI,PMS 1, and MSH2 interactions during the initiation of DNA mismatch repair in

yeast. Science 265:1091-1093.

Prolla, T. A., Pang, Q., Alani, E., Kolodner, R. D., and Liskay, R. M. 1994. MLH1,PMS 1, and MSH2 interactions during the initiation of DNA mismatch repair in

yeast. Science 265:1091-1093.

Prolla, T.A., Christie, D.M., and Liskay, R.M. 1994. Dual requirement in yeast DNAmismatch repair for MLH1 and PMS1, two homologs of the bacterial mutL

gene. Mol. Cell Biol. 14:407-4 15.

Pukkila, P.J., Peterson, J., Herman, G., Modrich, P., and Meselson, M. 1983. Effects

of high levels of DNA adenine methylation on methyl-directed mismatchrepair in Escherichia coli. Genetics 104:571-582.

Radman, M., Dohet, C., Bourgingnon, M-F., Doubleday, O.P., and Lecomte, P. 1981.

High fidelity devices in the reproduction of DNA. In Chromosome Damageand DNA Repair (E. Seeberg and K. Kieppe, Eds.). Plenum Publishing Corp.,

New York.

Page 161: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

146

Rayssiguier, C., Thaler, D.S., and Radman, M. 1989. The barrier to recombinationbetween Escherichia coli and Salmonella lyphimurium is disrupted inmismatch-repair mutants. Nature 342:396-401.

Reenan, R. A., and Kolodner, R. D. 1992a. Characterization of insertion mutations in

the Saccharomyces cerevisiae MSH1 and MSH2 genes: evidence for separate

mitochondrial and nuclear functions. Genetics 132:975-985.

Reenan, R. A., and Kolodner, R. D. 1992b. Isolation and characterization of twoSaccharomyces cerevisiae genes encoding homologs of the bacterial HexA and

MutS mismatch repair proteins. Genetics 132:963-973.

Reitmair, A.H., Schmits, R., Ewel, A., Bapat, B., Redston, M., Mitri, A., Waterhouse,P., Mittrucker, H.W., Wakeham, A., Liu, B., et al. 1995. MSH2 deficientmice are viable and susceptible to lymphoid tumors. Nat. Genet. 11:64-70.

Richards, E. J. 1997. DNA methylation and plant development. Trends Genet.

13:319-323.

Ross-Macdonald, P., and Roeder, G. S. 1994. Mutation of a meiosis-specific MutShomolog decreases crossing over but not mismatch correction. Cell 79:1069-

1080.

Schaaper, R.M. 1988. Mechanisms of mutagenesis in the Escherichia coli mutatormutD5: role of DNA mismatch repair. Proc. Nail. Acad. Sci. USA 85:8126-

8 130.

Smith, D. R. et al. 1997. Complete genome sequence of Methanobacteriumthermoautotrophicum deltaH: functional analysis and comparative genomics.

J. Bacteriol. 179:7135-7155.

Strand, M., Earley, M.C., Crouse, G.F., and Petes, T.D. 1995. Mutations in the MSH3

gene preferentially lead to deletions within tracts of simple repetitive DNA in

Saccharomyces cerevisiae. Proc. Nat!. Acad. Sci. U S A 92:10418-10421.

Su, S.S., and Modrich, P. 1986. Escherichia coli mutS-encoded protein binds tomismatched DNA base pairs. Proc. Nati. Acad. Sci. USA 83:5057-5061.

Page 162: Redacted for Privacy - CORE · mismatch specificities with their eukaryotic counterparts, but MutSy showed a novel substrate specificity. These data suggest that plants utilize unique

147

Su, S.S., Lahue, R.S., Au, K.G., and Modrich, P. 1988. Mispair specificity of methyl-directed DNA mismatch correction in vitro. J. Biol. Chem. 263:6829-6835.

Umar, A., Buermeyer, A.B., Simon, J.A., Thomas, D.C., Clark, A.B., Liskay, R.M.,and Kunkel, T.A. 1996. Requirement for PCNA in DNA mismatch repair at a

step preceding DNA resynthesis. Cell 87:65-73.

Umar, A., Risinger, J.I., Glaab, W.E., Tindall, K.R., Barrett, J.C., and Kunkel, T.A.1998. Functional overlap in mismatch repair by human MSH3 and MSH6.

Genetics 148:1637-1646.

Wagner, R. Jr., and Meselson, M. 1976. Repair tracts in mismatched DNAheteroduplexes. Proc. Nat!. Acad. Sci. USA 73:4135-4139.

Walbot, V. 1985. On the life strategies of plants and animals. Trends Genet. 1:165-169.

Wang, H., Lawrence, C.W., Li, G.-M., and Hays, J.B. 1999. Specific binding ofhuman MSH2MSH6 mismatch-repair protein heterodimers to DNAincorporating thymine- or uracil-containing UV light photoproducts opposite

mismatched bases. J. Biol. Chem. 274:16894-16900.

Welsh, K.M., Lu, A.L., Clark, S., and Modrich P. 1987. Isolation andcharacterization of the Escherichia coli mutH gene product. J. Biol. Chem.

262:15624-15629.

Wildenberg, J., and Meselson, M. 1975. Mismatch repair in heteroduplex DNA.

Proc. Natl. Acad. Sci. U S A. 72:2202-2206.

Woese, C. R., Gilbert, W. & Gillevet, P. M. 1994. The Genetic Data Environment:An expandable GUI for multiple sequence analysis. CABIOS 10:671-675.

Wu, T.-H., and Marinus, M. G. 1994. Dominant negative mutator mutations in the

mutS gene of Escherichia coli. J. Bacteriol. 176:5393-5400.