mitotic recombination in yeast: elements controlling its incidence

24
Review Mitotic recombination in yeast: elements controlling its incidence Andre ´s Aguilera*, Sebastia ´n Cha ´vez and Francisco Malago ´n Departamento de Gene ´tica, Facultad de Biologia, Universidad de Sevilla, Avd. Reina Mercedes 6, 41012 Sevilla, Spain * Correspondence to: A. Aguilera, Departamento de Gene ´tica, Facultad de Biologia, Universidad de Sevilla, Avd. Reina Mercedes 6, 41012 Sevilla, Spain. E-mail: [email protected] Received: 8 February 2000 Abstract Mitotic recombination is an important mechanism of DNA repair in eukaryotic cells. Given the redundancy of the eukaryotic genomes and the presence of repeated DNA sequences, recombination may also be an important source of genomic instability. Here we review the data, mainly from the budding yeast S. cerevisiae, that may help to understand the spontaneous origin of mitotic recombination and the different elements that may control its occurrence. We cover those observations suggesting a putative role of replication defects and DNA damage, including double-strand breaks, as sources of mitotic homologous recombination. An important part of the review is devoted to the experimental evidence suggesting that transcription and chromatin structure are important factors modulating the incidence of mitotic recombination. This is of great relevance in order to identify the causes and risk factors of genomic instability in eukaryotes. Copyright # 2000 John Wiley & Sons, Ltd. Keywords: Saccharomyces cerevisiae; homologous recombination; hyper-recombination; double-strand breaks; transcription; chromatin structure; replication defects; DNA damage; DNA repair Contents Introduction ....................... 731 Mitotic recombination as a source of genetic instability ....................... 732 DNA damage as a source of mitotic recombination ............................... 733 Replication defects as a source of mitotic homo- logous recombination events .......... 738 DNA damage and replication defects as a source of recombination in other organisms ...... 739 Effect of transcription on mitotic recombination ............................... 739 Role of chromatin structure on mitotic recom- bination ........................ 744 Concluding remarks .................. 747 References ........................ 747 Introduction As a consequence of the importance of homologous recombination in the generation of genetic variation and its essential role in sexual reproduction, a large amount of experimental data have emerged throughout the years regarding conjugational recombination in bacteria and meiotic recombina- tion in eukaryotes, particularly fungi. They have made a great contribution to the understanding of the molecular mechanisms of recombination and its biological significance. 117,122,151,167,179,193,197,200,202 However, beyond its role in sexual reproduction, homologous recombination has a great relevance in vegetative growth as a DNA repair mechanism. Mitotic recombination is a ubiquitous process in eukaryotes. It is the basis for gene targeting, for some human genomic disorders causing genetic diseases, 136 and it has been used regularly as a tool to investigate other biological processes, such as the genetic basis of development using genetic mosaics obtained by somatic recombination. 64 Although with different biological meaning and genetic control, both mitotic and meiotic recombi- nation use common factors and steps (reviewed in 76, 105, 159, 167, 169, 179, 193, 200, 202). The genetic and physical analysis of mitotic and meiotic recombination in S. cerevisiae has provided one Yeast Yeast 2000; 16: 731–754. Copyright # 2000 John Wiley & Sons, Ltd.

Upload: andres-aguilera

Post on 06-Jun-2016

212 views

Category:

Documents


0 download

TRANSCRIPT

Review

Mitotic recombination in yeast: elements controllingits incidence

AndreÂs Aguilera*, SebastiaÂn ChaÂvez and Francisco MalagoÂnDepartamento de GeneÂtica, Facultad de Biologia, Universidad de Sevilla, Avd. Reina Mercedes 6, 41012 Sevilla, Spain

*Correspondence to:A. Aguilera, Departamento deGeneÂtica, Facultad de Biologia,Universidad de Sevilla, Avd. ReinaMercedes 6, 41012 Sevilla, Spain.E-mail: [email protected]

Received: 8 February 2000

Abstract

Mitotic recombination is an important mechanism of DNA repair in eukaryotic cells.

Given the redundancy of the eukaryotic genomes and the presence of repeated DNA

sequences, recombination may also be an important source of genomic instability. Here we

review the data, mainly from the budding yeast S. cerevisiae, that may help to understand

the spontaneous origin of mitotic recombination and the different elements that may

control its occurrence. We cover those observations suggesting a putative role of

replication defects and DNA damage, including double-strand breaks, as sources of

mitotic homologous recombination. An important part of the review is devoted to the

experimental evidence suggesting that transcription and chromatin structure are important

factors modulating the incidence of mitotic recombination. This is of great relevance in

order to identify the causes and risk factors of genomic instability in eukaryotes.

Copyright # 2000 John Wiley & Sons, Ltd.

Keywords: Saccharomyces cerevisiae; homologous recombination; hyper-recombination;

double-strand breaks; transcription; chromatin structure; replication defects; DNA

damage; DNA repair

Contents

Introduction . . . . . . . . . . . . . . . . . . . . . . . 731Mitotic recombination as a source of genetic

instability . . . . . . . . . . . . . . . . . . . . . . . 732DNA damage as a source of mitotic recombination

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 733Replication defects as a source of mitotic homo-

logous recombination events . . . . . . . . . . 738DNA damage and replication defects as a source of

recombination in other organisms . . . . . . 739Effect of transcription on mitotic recombination

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 739Role of chromatin structure on mitotic recom-

bination . . . . . . . . . . . . . . . . . . . . . . . . 744Concluding remarks . . . . . . . . . . . . . . . . . . 747References . . . . . . . . . . . . . . . . . . . . . . . . 747

Introduction

As a consequence of the importance of homologousrecombination in the generation of genetic variationand its essential role in sexual reproduction, a large

amount of experimental data have emergedthroughout the years regarding conjugationalrecombination in bacteria and meiotic recombina-tion in eukaryotes, particularly fungi. They havemade a great contribution to the understanding ofthe molecular mechanisms of recombination and itsbiological signi®cance.117,122,151,167,179,193,197,200,202

However, beyond its role in sexual reproduction,homologous recombination has a great relevance invegetative growth as a DNA repair mechanism.Mitotic recombination is a ubiquitous process ineukaryotes. It is the basis for gene targeting, forsome human genomic disorders causing geneticdiseases,136 and it has been used regularly as a toolto investigate other biological processes, such as thegenetic basis of development using genetic mosaicsobtained by somatic recombination.64

Although with different biological meaning andgenetic control, both mitotic and meiotic recombi-nation use common factors and steps (reviewed in76, 105, 159, 167, 169, 179, 193, 200, 202). Thegenetic and physical analysis of mitotic and meioticrecombination in S. cerevisiae has provided one

YeastYeast 2000; 16: 731±754.

Copyright # 2000 John Wiley & Sons, Ltd.

important hallmark for all types of homologous

recombination: its initiation by a double-strand

break (DSB). Meiotic recombination is initiated by

a DSB211,212 catalysed by Spo11p.11,101 The DSB is

basically processed by a recombinational repair

mechanism215 that was proposed after performing

experiments of gene targeting in mitotically dividing

yeast cells using gapped plasmids 163. A great deal

of experimental evidence has been provided on the

ability of a DSB to induce a mitotic recombination

event that proceeds via a mechanism that shares all

de®ned hallmarks of homologous recombination:

strand invasion, heteroduplex formation, mismatch

repair, DNA synthesis, and the generation of

Holliday junctions75,105,164,167,193 (Figure 1). Never-

theless, our knowledge on the spontaneous origin of

mitotic recombination is very meagre.Our purpose is to review the data, mainly from

the budding yeast S. cerevisiae, that may help to

understand the spontaneous origin of mitotic

recombination and the different factors that maycontrol its incidence. Most of this review will bebased on results reporting a stimulation of mitoticrecombination (hyper-recombination), some ofwhich were partially reviewed previously.5 A list ofall known S. cerevisiae genes from which, to ourknowledge, mutations with a signi®cant sponta-neous hyper-recombination phenotype have beenreported, is shown in Table 1. We will cover thoseobservations suggesting a putative role of replica-tion defects and DNA damages as sources ofmitotic homologous recombination. However, themain focus of this review will be the experimentalevidence suggesting that transcription and chroma-tin structure are important factors modulating theincidence of mitotic recombination, since there areno recent reviews devoted to this subject. We willcover very few aspects of the mechanisms of eitherhomologous recombination or DSB repair in yeast,as a considerable number of recent reviews areavailable.75,77,159,164,167,197

Mitotic recombination as a source ofgenetic instability

Recombinational repair of a DNA sequence uses anhomologous partner to resynthesize the damagedregion. In a diploid yeast cell, the homologouspartner can be either the homologous chromosomeor the sister chromatid, whereas in haploid yeastcells the homologue should be a sister. As aconsequence, recombination should not lead toany genetic rearrangement. It should have nogenetic consequences if using a sister chromatid or

Figure 1. DSB repair model for recombination. The classicmodel by Szostak et al215. involves the invasion of the 3k-endsinto the homologous DNA molecule to form two hetero-duplex regions and two Holliday junctions. Repair of theheteroduplex determines whether or not a gene conversionevent takes place. Single-strand annealing between the twoinvading and resynthesized DNA strands can re-establish theoriginal molecular con®guration without involvement ofcrossing over.167,216 Resolution of the Holliday junctiondetermines whether or not a crossing-over occurs167,215

Figure 2. Recombination events can occur between DNArepeats located in the same chromatid (a), in homologouschromosomes (b), in non-homologous chromosomes (c) orin sister chromatids (d)

732 A. Aguilera, S. ChaÂvez and F. MalagoÂn

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

it should only lead to a mitotic recombination event

between homologous chromosomes, usually a gene

conversion unassociated with crossing-over.As evidenced by the sequencing of the yeast

genome, S. cerevisiae contains a signi®cant amount

of duplicated chromosomal regions and stretches of

repeated DNA sequences that are potential sub-

strates of recombination192,237 (Figure 2). Since

Jackson and Fink95 and Klein and Petes106 reported

that recombination could occur between repeated

copies of a gene arti®cially introduced into one

chromosome, many different types of DNA repeat

constructs have been used to study mitotic recom-

bination in yeast (reviewed in 169). Particularly

interesting was the observation that recombination

could also occur between ectopic copies of one gene

arti®cially introduced into non-homologous chro-

mosomes,96 as was previously reported for the

naturally occurring CYC1 gene S. cerevisiae43 or

between the naturally occurring copies of tRNA

genes in the ®ssion yeast Schizosaccharomyces

pombe.152

Contrary to recombination between homologous

chromosomes and sister chromatids, recombination

between repeated DNA sequences located in either

the same or different chromosomes can have drastic

consequences for the stability of the genome. It can

lead to gross chromosomes rearrangments such as

deletions, inversions or translocations,5,107 even

though many such rearrangements may occur by

aberrant replication or by non-homologous end-

joining (NHEJ).33,79

Among the different types of rearrangements that

can be generated by interchanges between repeated

DNA sequence, deletions occurring between long

interspersed repeats deserve particular attention.

They can arise by recombination events such as

cross-over between two repeats, unequal sister

chromatid exchange or unequal sister chromatid

gene conversion181 (Figure 3), that presumably

follow a standard mechanism of DSB recombina-

tional repair. In addition, deletions could theoreti-

cally also occur by non-conservative events such as

one-ended invasion10,174 or single-strand annealing

(SSA)54,55,129,165 (Figure 4). Sister-chromatid repli-

cation slippage could also lead to deletions.133 It is

likely that these speci®c mechanisms responsible for

deletions between direct repeats can explain the

differential effect of some mutations on deletions

vs. other types of exchanges.

DNA damage as a source of mitoticrecombination

Double-strand breaks

Different experimental evidence suggests that DSBs

act as initiators of mitotic recombination. Thus, it is

well known that DSB-inducing agents such as

X-rays and c-rays induce recombination in

yeast.44,45,84 Since the work by Orr-Weaver et al.,163

indicating that a plasmid linearized with a restric-

tion endonuclease is very ef®ciently integrated into

the chromosome by homologous recombination,

and the ®nding that MAT switching occurred by a

DSB catalysed by the HO endonuclease,205 many

other studies have been performed in the yeast

S. cerevisiae, in which DSBs have been produced

in vitro by restriction endonuclease120,147,162,174

or in vivo by the site-speci®c endonucleases

HO54,55,157,165,182,183,208 and I-SceI173, the Flp

recombinase175 and restriction endonucleases.125 In

Figure 3. Recombination events responsible for deletionsbetween non-tandem direct repeats: intrachromatid cross-ing-over, unequal sister chromatid reciprocal exchange andunequal sister chromatid gene conversion, all of themoccurring presumably by the DSB-repair model shown inFigure 1

Mitotic recombination in yeast: control of incidence 733

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

Table 1. S. cerevisiae genes for which mutations with a spontaneous hyper-rec phenotype have been described

Gene Function of gene product References

CDC5 Ser/Thr-protein kinase required for exit from mitosis 82

CDC8 Thymidylate kinase 82CDC6 Initiation of DNA replication 25,82

CDC9 DNA ligase 61,177

CDC14 Protein phosphatase involved in cell cycle control 82CDC20 WD-40 protein involved in microtubule function 82

CDC23 Component of anaphase promoting complex, APC 82

CDC45 Initiation of DNA replication 86

CDC73 RNA pol II-associated protein 195CTF4 (CHL15) DNA polymerase a binding protein 115

DNA2 ssDNA-dependent ATPase, DNA helicase and endonuclease 53

DUN1 Protein kinase involved in regulation of DNA repair genes 49

HPR1 Transcription and genomic stability 2,3HSM3 MSH1 homologue 51

MCM1 Transcription factor of the MADS box family 42

MCM2 Component of the MCM complex of initiation of replication 239

MCM3 Component of the MCM complex of initiation of replication 239MCM4 (CDC54) Component of the MCM complex of initiation of replication 86

MCM5 (CDC46) Component of the MCM complex of initiation of replication 86

MCM7 (CDC47) Component of the MCM complex of initiation of replication 86MEC1 DNA damage checkpoint protein 226

MMS21 DNA-repair protein 150,177

MRE11 (RAD58) DSB-repair ssDNA endo- and dsDNA exonuclease 6

PAF1 RNA pol II associated protein 195PAT1 Topoisomerase II-associated protein 230

PKC1 Protein kinase C1 88

PMS1 mutL-like DNA mismatch repair protein 235

POL1 (CDC17) DNA polymerase a 2,82,134POL3 (CDC2) DNA polymerase h large subunit 2,68,82

POL30 PCNA 30

POL31 (HYS2) DNA polymerase h small subunit 210PRI1 DNA primase small subunit 132

PRI2 DNA primase large subunit 132

RAD2 NER ssDNA endonuclease 119

RAD3 DNA helicase of TFIIH involved in NER and transcription 66,138,149RAD5 Snf2p family of DNA helicases 128

RAD6 Ubiquitin conjugating enzyme involved in error-prone repair 103

RAD18 DNA-binding protein involved in DNA repair 14,128,143

RAD9 DNA damage checkpoint protein 48RAD24 DNA damage checkpoint protein 118

RAD27 (RTH1) ssDNA FLAP endonuclease I 201,214,222,226

RAD50 DSB-repair protein 141

RAD51 RecA-like DNA strand exchange protein 1,127,144RAD52 DSB recombinational repair protein 8,140

RAD54 DSB-repair DNA-dependent ATPase 144

RAD55 DSB-repair DNA-binding protein 144RAD57 DSB-repair protein 144

RFA1 ssDNA-binding replication protein 198,199

RFC2 Replication Factor C 160

SGS1 RecQ-like DNA helicase 63,232SIR2 DNA-silencing protein 69

SOH1 Putatively related with RNA pol II transcription 46

SPO11 DSB transesterase of initiation of meiotic recombination 24

SPT4 Chromatin-related protein involved in transcription 139SPT6 Chromatin-related protein involved in transcription 139

SRS2 (HPR5) DNA helicase 2,166

SSL1 Component of TFIIH 138

734 A. Aguilera, S. ChaÂvez and F. MalagoÂn

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

all cases, reciprocal exchange and gene conversion,as well as deletions, have been detected. This,together with data indicating that meiotic recombi-nation was initiated by DSBs,76,105,179,193,202 has ledto the belief that DSBs are the main event initiatinga recombination reaction that would proceed viathe DSB repair mechanism (Figure 1).

Once end-joining was shown to be an alternativemechanism of DSB repair in mammalian cells,similar evidence was provided for this type ofrepair in yeast, especially after the identi®cation of

the yeast homologues of the mammalian Ku70 andKu80 proteins (reviewed in 33, 50, 99, 126, 224,233). Despite the different relevance of recombina-tion and end-joining in yeast vs. higher eukaryotes,it seems clear that both types of mechanisms areused for DSB repair, the importance of eachpathway being dependent on the stage of the cellcycle in which the DSB is repaired.87,100 Inexponentially growing yeast cells, recombinationseems to be the most prominent DSB-repair path-way, as deduced from the strong sensitivity to DSB-

Figure 4. Non-conservative intramolecular mechanisms of recombination leading to deletions: single-strand annealing andone-ended invasion cross-over

Table 1. Continued

Gene Function of gene product References

SSL2 (RAD25) DNA helicase of TFIIH 119THO2 Transcription and genomic stability 172

TOP1 DNA topoisomerase I 31,124

TOP2 DNA topoisomerase II 31

TOP3 DNA topoisomerase III 9,229TRF4 Required for cell viability in the absence of Topo I 185

XRS2 DSB-repair protein 92

Mitotic recombination in yeast: control of incidence 735

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

inducing agents of recombinational-repair mutantsvs. end-joining mutants. Indeed, X-ray sensitivity ofKu mutants are only clearly observed in doublemutant backgrounds containing an additionalmutation in a recombinational repair gene, such asrad52.17,18,135,148,194 The question arises as towhether or not the primary event leading to aspontaneous recombination event in mitosis is aDSB or another type of DNA damage or replica-tion error.

In mitosis, evidence that recombination is spon-taneously initiated by DSBs is provided by the factthat all genes required for spontaneous mitoticrecombination (RAD50, RAD51, RAD52, RAD54,RAD55, RAD57, RAD59, XRS2, MRE11 andRFA1) are required for the repair of DSBs inducedby c-ray.84,169

The involvement of DSBs in spontaneous recom-bination may also explain the hyper-recombinationphenotypes conferred by mutations in RAD50,MRE11, XRS2. Since religation of DSBs by end-joining requires these genes, it is likely that moreevents are shunted into the recombinational repairpathway in their absence.6,28,92,223 The possibilitythat hyper-recombination in these mutants isproduced as a consequence of a failure to repairdamage from a sister chromatid has also beenraised.167 In any case, the explanation may not bethat simpleÐotherwise hyper-recombination shouldhave been found associated with mutations inHDF1 and HDF2, encoding the Ku70 and Ku80yeast homologues.

Interestingly, rad51, rad54, rad55, rad57 and rfa1mutants are hyper-recombinant for direct repeatrecombination leading to deletions.1,127,144,198 Sincethe wild-type copies of these genes are required forrecombination, these results suggest that deletionsmust occur by a speci®c type of mechanismindependent of those gene products. This may besingle-strand annealing (SSA). Therefore, DSBsthat are not repaired by a gene conversion/recipro-cal exchange type of homologous recombination(see Figure 1) can be shunted into repair by a strandresection mechanism that does not require aRAD51-dependent strand exchange reaction, pre-sumably single-strand annealing (SSA) (Figure 4).This would explain the speci®c hyper-rec phenotypefor deletions in these rad mutants. Interestingly,RAD59, another gene involved in a RAD51-inde-pendent pathway of DSB repair, is involved in theformation of deletions between direct repeats.93

Mutations in RAD59 confer a slight increasein recombination between homologous chromo-somes.7 This has also been shown for a particularrad52 allele that shares other phenotypes withRAD59-loss-of-function mutations.8 Therefore, itseems that DSBs can also be channelled intodifferent DNA recombination repair pathways,either RAD51- or RAD59-dependent. In addition,recombinagenic breaks might also be channelledinto recombinational repair pathways not leading toa detectable genetic exchange. Thus, to explain thehyper-recombination phenotypes of the checkpointmutations rad9, Fasullo et al.48 have proposed thatthe RAD9 checkpoint may channel the repair ofDSBs into sister-chromatid exchange (SCE), there-fore impeding a detectable mitotic recombinationevent between two homologues. Such an impedi-ment should be released in rad9 mutants.

Single-strand breaks

Whether any type of DNA break can eventuallylead to mitotic recombination without being pre-viously processed into a DSB has been addressed intwo different studies. By using a gIIp-based system,the group of J. Strathern has shown that single-strand DNA breaks also induce recombination.206

However, since the DNA molecule where the breakwas produced acted as receptor of information,their results suggested that the recombinationevents proceeded according to a DSB recombina-tional repair model.215 It is likely that single-strandbreaks were processed into DSBs before recombina-tion took place. Galli and Schiestl59 observed that,whereas gIIp-induced single-strand breaks had noeffect on recombination between direct repeats inthe G1 or G2 stage of the cell cycle, recombinationwas increased when induced during the S phase.Instead, I-SceI-induced DSBs increased recombina-tion in both G1 and G2. Again, these results suggestthat the single-strand break induces recombinationif processed into a DSB, and that processing into aDSB especially occurs when a replication forkarrives at the site of the nick during S phase.

DNA damage other than DNA breaks

Many pieces of evidences suggest that DNAdamage other than DSBs can be channelled intothe recombinational repair pathway. This can bededuced from many hyper-recombination mutantsso far analysed in yeast (Table 1). Thus, sponta-

736 A. Aguilera, S. ChaÂvez and F. MalagoÂn

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

neous DNA damage susceptible to nucleotideexcision repair (NER) or error-prone repair mayexplain hyper-recombination in a number ofmutants affected in NER and error-prone repair, ifthe unrepaired damage is channelled into recombi-national repair. This can be the case of the hyper-rec phenotypes of the NER mutations in RAD3,149

RAD25119 or SSL1138 or other repair mutationssuch as hpr5/srs22,166,180 rad1814,128,143 or rad6.103

Increased heteroallelic gene conversion caused bydun1 has also been explained by proposing thatDNA-damage-induced gene expression mediated byDUN1 channels the repair of UV damage into anon-recombinagenic repair pathway.49 It is likelythat such damage is processed into a DNA breakbefore being channelled into recombination.

The work of Jinks-Robertson's group213 isparticularly revealing for the evaluation of theimportance of DNA damage other than DSB as asource of mitotic recombination. By inactivatingNER or base excision repair (BER) in yeast cellsby mutations, a strong hyper-rec phenotype isobserved that is synergistic when both BER andNER pathways are simultaneously inactivated. Thehyper-rec phenotype is enhanced if oxidizing agents,such as hydrogen peroxide or menadione, are addedto the cells. In all cases, recombination is dependenton the RAD52 gene. These results suggest that anytype of spontaneous damage may lead to DNAbreaks that need to be repaired by homologousrecombination. It is likely that, in wild-type strains,part of the spontaneous damage produced bymethylation, oxidation and other chemical reactionsescape their main routes of repair, such as BER orNER. The subsequent action of DNA replication,or the partial action of an endonuclease at the siteof the damage, may lead to a subsequent recombi-nagenic DNA break.

The induction of recombination by differentDNA-damaging agents such as UV, alkylatingagents, cysplatin, etc. can be explained in similarterms.20,34,37,41,80 They produce a DNA lesion thatmay be processed into a DNA gap or break. Insome cases, replication may be involved. Thus, UV-induced thymidine dimers or 3-methyladenine, thatcannot be traversed by the replicative DNA poly-merase, can become a DSB during replication.Indeed, the block of the replication fork has beensuggested to be one of the causes of increasedmitotic recombination after 8-methoxypsoralen andUVA irradiation,37 or after cisplatin treatment.80

DSBs are observed after post-treatment incubationand may be responsible for the hyper-recombination observed at the ARG4 locus inmitosis. This is consistent with the observationindicating that interstrand cross-links requireRAD51 to be repaired94 or that the resistance tothe alkylating agent methyl methane sulphonate(MMS) requires all the RAD50 series of genesinvolved in recombinational repair.41,84,169 DSBscould also be generated by the un®nished action ofexcision repair as an intermediate in the reaction.However, since the replication fork is blocked bythe adducts, it is likely that a large proportion ofthose DSBs occur by blockage of the replicativeDNA polymerase (see below). Indeed, the inductionof recombination by UV, MMS, EMS and 4-NQOhas been shown to depend on cell division, suggest-ing (once more) that the recombinagenic event mustbe produced after the replicative polymeraseattempts to go through the unrepaired lesion.60

Mismatches

Mismatches may also stimulate mitotic recombina-tion, as can be deduced from the observation thatplasmid integration is stimulated by mismatches.240

It has been proposed that mismatches couldpotentially lead to recombinagenic structuresduring mismatch repair (MMR) such as single-strand gaps or denatured DNA regions.240 On theother hand, the meiotic studies of Borts andHaber,15 using multiple heterozygosities, suggeststhat a heteroduplex containing multiple mismatchesmay trigger new recombination events. Therefore,it is possible that mismatches can initiate recom-bination events, either directly or after processinginto a DNA break. Nevertheless, we still needmore experimental evidence to con®rm such apossibility.

A different hyper-recombination phenotyperelated to mismatches is that caused by mutationsin PMS1235 and other genes involved in MMR suchas MSH2.30,97,191 The mismatch repair proteinsrecognize mismatches formed in the heteroduplexDNA occurring during recombination and act as abarrier for recombination. In heteroduplexesformed between homeologous sequences, thosebarriers would be very abundant; thus mismatchrepair would play an anti-recombinogenic role. As aconsequence, the inactivation of mismatch-repairgenes would allow heteroduplex to form between

Mitotic recombination in yeast: control of incidence 737

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

homeologous DNA sequences and recombinationto be completed, explaining the speci®c enhance-ment of recombination between homeologous DNAsequences. This enhancement of recombination canbe explained by extension of on-going recombina-tion events, rather than by a stimulation ofinitiation events.

Replication defects as a source ofmitotic homologous recombinationevents

Replication defects may be one of the main sourcesof mitotic recombination. The accumulation ofeither blocked replication forks or DNA gaps byincomplete replication trigger a post-replicativerecombinational repair pathway that uses a homo-logous partner to bypass the unreplicated region.Thus, it has been shown, in yeast, that the inhibitorof replication hydroxyurea (HU) induces recombi-nation in dividing cells but not in G1- or G2-arrested yeast cells.58 This suggests that the inhibi-tion of replication must lead to recombinageniclesions, presumably DNA gaps or breaks. It is likelythat single-stranded DNA breaks caused as aconsequence of HU treatment frequently lead toDSBs.145,146 The increase in short repeat intrachro-mosomal recombination observed at palindromicstructures may also be the consequence of thedif®culties of the replicative polymerase to progressthrough stem structures.67,85,131,184

Good evidence that replication errors can be animportant source of mitotic recombination comesfrom the analysis of mutants in the DNA replica-tion machinery (Table 1). They show a strongincrease in mitotic recombination between eitherhomologous chromosomes or repeats, whetherintra- or inter-chromosomal. The ®rst replicationmutations shown to lead to strong hyper-recombination phenotypes in yeast were those inthe structural genes of DNA ligase I,61,177 thymidy-late kinase82 or the large subunits of DNApolymerases I2,82,134 and III.2,68,82 Subsequently,hyper-recombination has also been shown formutations in the small subunits of DNA polymer-ase III,210 or the structural genes of the two primasesubunits132 or the replication factors A198 and C.160

Mutations in genes involved in initiation of replica-tion have also been shown to lead to spontaneoushyper-recombination. This is the case of mutations

in MCM2, MCM3, MCM4, MCM5 and

MCM7,86,239 CDC6,25,82 CDC4586 or the DNA

helicase gene DNA2 with a putative role in replica-

tion.53

It is likely that stimulation of recombination in

all replication-defective mutants is provided by the

accumulation of gaps during the S phase of the cell

cycle. Those gaps can become DSBs by subsequent

nicking of the DNA template (Figure 5). This may

also be the case for the mec1 and pol30 alleles that

are lethal in a rad52 background. They accumulate

single-strand intermediates that presumably become

DSBs.145,146 Thus, the 3k-end would be able to

initiate a strand invasion process into a homologous

sequence that would result in a recombination

event. The recent work on mutations of RAD27, a

ssDNA Flap endonuclease, involved in the removal

of Okazaki fragments is very illustrative in this

sense. These mutations show a strong hyper-rec

phenotype178,201,222,226 and are synthetically lethal

with any mutation of the RAD52 series of genes

involved in DSB recombinational repair.214,222

After displacement of the Okazaki fragment by the

DNA polymerase, lack of removal of the 5k ¯ap

structure by Rad27p results in a single-strand gap

susceptible to breakage on the template strand,

resulting in a DSB.222

The recent work on E. coli by the group of B.

Michel showing that the RuvB helicase (involved in

Figure 5. DSB formation as a consequence of replicationfork blockage121. Either a nick occurs in the template of thelagging strand or the two nascent DNA strands re-anneal toform a Holliday junction with a DSB end

738 A. Aguilera, S. ChaÂvez and F. MalagoÂn

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

the migration of Holliday junctions) is required forthe repair of replication blocks, suggests that Holli-day junctions may be one of the intermediatesformed when replication is impeded.190 The nascentleading and lagging strands can reanneal, by actionof RuvB, to form a Holliday junction and a double-strand end that could initiate a recombination eventthat would require resolution of the Hollidayjunction (Figure 5). Although no evidence of thisprocess has been reported yet in yeast, it isparticularly interesting the accumulation of Holli-day junctions in the rDNA region that are formedin the absence of the strand-invasion proteinRad51p during the S phase of the cell cycle.243

Such Holliday junctions can be explained perfectlyas intermediates of DNA replication. The results ofKadyk and Hartwell,98 indicating that most recom-binational repair occurs between sister chromatids,is consistent with the idea that an important sourceof spontaneous recombination may come fromreplication. It is likely that some of the replicativeerrors are shunted into homologous recombinationevents that use a homologous chromosome as atemplate.

Lack of topoisomerase activity is another impor-tant source of mitotic recombination in yeastthat might be related to replication. Hyper-recombination has been observed in top1, top231

and top3 mutants9,229 as well as trf4,185 pat1230 orsgs1232. Whose gene products have a functional orphysical relationship with Top1, Top2 and Top3,respectively. Consistent with these data, DACA (N-2[-dimethylamino)ethyl]acridine-4-carboxamide), aninhibitor of topoisomerase II, is a recombinagenicdrug to which rad52 cells are very sensitive.52 It ispossible that topoisomerases are involved in DNAreplication termination, as suggested for topoisome-rase III.62 However, it is also likely that the hyper-rec phenotype associated with top mutations isrelated to the role of topoisomerases in otheraspects of DNA metabolism, such as transcriptionor DNA topology by itself (see the section on DNAstructure and topological constraints).

DNA damage and replication defects asa source of recombination in otherorganisms

Studies on the different factors inducing recombina-tion in vegetative cells have been reviewed in

other organisms from bacteria to mammaliancells.26,65,90,137,207,219 Data on mitotic homologoushyper-recombination in eukaryotes other than theyeast S. cerevisiae are very scarce. In the ®ssionyeast Sz. pombe, there are only few hyper-recmutants described.73,74,196,220 Studies on recombina-tion induced by DNA-damage-inducing agents,such as X-rays, or hyper-recombination conferredby mutations in genes involved in DNA repairpathways, other than recombination repair, in E.coli, are abundant. Documentation can be found insome recent reviews.32,197

The role of replication defects in recombinationhas been extensively studied in prokaryotes. The®rst hyper-recombination mutants of E. coli werefound to be affected in different DNA replicationfunctions, including DNA ligase, polymerases orthymidylate kinase, more than 20 yearsago.114,142,242 Indeed, many of the studies per-formed in the yeast S. cerevisiae owe much to theprevious work on E. coli. Information on hyper-recombination in E. coli can be found in severalreviews.12,97,110,111,122,137,197 In general, there is astrong parallelism between hyper-recombination inbacteria and yeast. This parallelism suggests thatyeast studies should become an important referencein order to understand the factors that stimulate theincidence of mitotic recombination and geneticinstability in higher eukaryotes.

Effect of transcription on mitoticrecombination

A connection between mitotic recombination andgene transcription has been well established in yeastas well as prokaryotes and higher eukaryotes.In yeast, both RNA polymerase I- and poly-merase II-driven transcription are known to beinvolved in the induction of mitotic recombinationand genetic rearrangementsÐdeletions, reciprocalexchanges, gene conversions and microsatellitedestabilization.

Mating-type switching is probably a clear exam-ple in which, given the developmental control oftranscription and recombination of the genesinvolved (MAT, HML, HMR and HO), the inter-connection between both processes has to be tightlycontrolled. It has been well established that silen-cing impedes recombination at the HML and HMRloci.77,167 Recently, it has been shown that there is

Mitotic recombination in yeast: control of incidence 739

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

an in¯uence of transcription on recombination thatis exerted at a distance, independently of thetranscriptional status of the donor and acceptorsequences, by the action of a recombinationalenhancer regulated at the transcriptional level.238

As this aspect of the in¯uence of transcription onrecombination has been recently reviewed,78 we willfocus on other examples, which deal with systemsthat show different recombination levels dependingon their transcriptional state.

RNA polI stimulated recombination: HOT1

The ®rst evidence for transcription-induced recom-bination in yeast was obtained after the isolation ofthe recombination hotspot HOT1102 as a cis-actingelement able to stimulate both intra- and inter-chromosomal recombination. HOT1 corresponds toa 4.6 kb rDNA fragment containing the 3k-end ofthe 25S gene, the 5S gene, the two non-transcribedspacers and the 5k end of the 35S rRNA. Roeder'sgroup,228 using his4 direct repeat intrachromosomalassays, demonstrated that a 570 bp fragment con-taining the initiation site of the 35S rRNAprecursor (I element) and an enhancer of transcrip-tion by RNA polymerase I (E element) were enoughto stimulate recombination. Both I and E elementshave identical behaviour in transcription andrecombination assays. That transcription, promotedby HOT1, is indeed responsible for the stimulationof recombination, has been shown by the facts thatdeletion mutations in the E or I element abolishboth the transcription promotion and recombinatorstimulation204 and that HOT1 activity is lost in aRNA polymerase I mutant incapable of transcrip-tion of the 35S RNA gene.89 It is certainly possiblethat transcription-driven recombination is an indir-ect consequence of the effect of transcription onDNA, caused by the unwinding of the DNA duplexand changes in the local supercoiling or in chroma-tin structure. Interestingly, when Voelkel-Meimanet al.228 inserted an RNA pol I transcriptiontermination between HOT1 and adjacent sequences,the stimulation of recombination was abolished.This indicates that transcription needs to proceedthrough at least one repeat unit to promotestimulation of recombination. This result excludesthe possibility that putative changes in chromatinstructure occurring and/or propagating from thepromoter region are the only factor responsible forrecombination enhancement.

A better knowledge of the biological elementscontained in the HOT1 sequence has provided newclues for the understanding of its recombination-stimulating activity. One of the non-transcribedspacers (NTS1) contains a site called RFB thatblocks the replication fork in the direction oppositeto that of transcription of the 35S RNA gene.21,130

Kobayashi et al.108 has found that the minimalregion essential for the replication block activity liesin the same E element required for transcriptionand recombination enhancement. Interestingly,Huang and Keil89 have shown, by deletion andsite-directed mutation analysis, a good correlationbetween recombination and transcription in theHOT1 region, except for a deletion of a segmentthat is exactly the same as RFB and that greatlyreduces HOT1 activity, with little effect on tran-scription. Kobayashi and Horiuchi109 have identi-®ed FOB1 as a gene encoding a trans-factorrequired for RFB activity. Mutations in this geneabolish the Hot1 phenotype. As they suggest,Fob1p is likely involved in blocking the replicationfork through the RFB site, resulting in enhance-ment of recombination. However, how the RFBactivity is related to transcription to stimulaterecombination remains to be elucidated (see later).Finally, an important feature of the HOT1 sequenceis its ability to greatly stimulate deletions (100-fold).This contrasts with its weak effect on gene conver-sions.228 In any case, the in¯uence of HOT1 onrecombination has been investigated outside of itsrDNA original context.

Stimulation of recombination by RNA polII-mediated transcription

Reports showing that RNA polymerase II-dependent transcription also stimulates recombina-tion appeared after the identi®cation of HOT1.Rothstein's group showed that recombinationevents in directly repeated sequences of the GAL10gene lead to deletions, but not to gene conversions.The frequency of these deletions is increased byRNA pol II-driven transcription.218 As in the caseof HOT1, the level of stimulation of recombinationcorrelates with the level of expression of the repeatsystem. This suggests that recombination may beinitiated in either one of the repeats or theintervening region. The RAD52 gene and, to alesser extent, RAD1 are required, although therecombination pathway that remains in a rad52

740 A. Aguilera, S. ChaÂvez and F. MalagoÂn

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

rad1 double mutant is positively affected bytranscription as well. It is likely that recombinationis stimulated at the level of initiation and not byimproving the ef®ciency of the recombinationmachinery.217

Another relevant case of RNA pol II transcrip-tion induced recombination is the induction ofTy ectopic recombination, which is mainly non-reciprocal. When transcription of a marked Ty copyis under the control of the GAL1 promoter, its rateof recombination is highly stimulated by transcrip-tional induction. This stimulation is synergistic withthe effect caused by increased levels of donor TycDNA.155 Again, RAD1 and RAD52 are needed forthe transcriptional enhancement of recombination,but mutations in these genes do not completelyabolish the different types of events seen followingtranscription.156

Since transcription driven by both RNA poly-merases I and II can stimulate different forms ofrecombination, an interesting question is whethertranscription at any promoter is able to stimulate acertain type of exchange event. A study by Brattyet al.19 suggests that this may not be the case.Recombination of a chromosomal ade 1 allele witha plasmid-borne ADE1 ORF was stimulated whentranscription of the latter was induced by the GAL1promoter, but not when induced by the promoter-enhancer region of the rDNA locus (RNA pol I) orthe ADH1 promoter (RNA pol II).19 These resultspoint to an in¯uence of the promoter sequence onthe recombination event, at least for plasmid-chromosome exchange.

Recombination associated to RNA polII-transcription elongation: HPR1 and THO2

Although numerous examples exist connectingtranscription and recombination, the nature ofthat connection has not yet been elucidated. Onepossibility is that there is an indirect effect oftranscription on recombination by increasing theaccessibility of DNA to damaging agents or to therecombination machinery. The identi®cation andanalysis of hyper-recombination mutants haveprovided new clues for the understanding of sucha connection. One particularly relevant mutant ishpr1. Originally identi®ed in a genetic screen ofhyper-rec mutants, hpr1 shows a strong increase inthe rate of deletions occurring at direct repeats withno effect on gene conversion or sensitivity to DNA-

damaging agents, as shown for HOT1- and GAL10-stimulated recombination.3,4 The null hpr1 mutantshows pleiotropic phenotypes, such as temperature-sensitivity and reduced levels of transcription ofdifferent reporter constructs,46,241 which suggests aninvolvement of Hpr1p in transcription. Consis-tently, suppressors of the ts phenotype have beenmapped in structural genes related to RNA poly-merase II-dependent transcription.47,225 The isola-tion of mutations in the genes SRB2 and HRS1/MED3 as suppresors of the hyper-recombinationphenotype of hpr1D provided a ®rst link betweenthe phenotypes of transcription and recombina-tion.170,187,188 Both SRB2 and HRS1 encode sub-units of the mediator complex of the RNApolymerase II holoenzyme,113,153,171 and are fullyrequired for the hyper-rec phenotype conferred byhpr1D.187

The transcription-recombination link in hpr1appeared stronger after experiments showing thatthe hyper-rec phenotype of the hpr1 mutants isdetected in direct repeats containing an interveningsequence that is transcribed. Insertion of the CYC1transcription terminator downstream of the DNArepeat that is transcribed, so that transcriptioncannot get through the intervening region, abolishesthe hyper-recombination phenotype.176 Theseresults suggest a role of transcription elongation,rather than initiation, in stimulating recombination.Indeed, a transcriptional analysis of hpr1 mutantsindicates that Hpr1p has a functional role intranscriptional elongation.29 The importance oftranscription elongation impairment on recombina-tion is con®rmed by the fact that the hyper-recphenotype of hpr1 strongly depends on the natureof the transcribed intervening sequence. Recombi-nation levels are extremely high (almost 100%frequencies) when the RNA pol II proceeds throughDNA sequences in which transcription elongation isstrongly impaired (i.e. lacZ). Indeed, there is acomplete correlation between the reluctance of aparticular DNA sequence to be transcribed in hpr1mutants and the ability of such a sequence toinduce recombination when inserted between directrepeats.29

The conclusions reached by studying hpr1 strainshave been corroborated by the identi®cation andcharacterization of mutations in THO2, a geneisolated as a high-copy suppressor of hpr1. As is thecase of hpr1D, tho2D is strongly hyper-rec, it showsan even stronger impairment of transcriptional

Mitotic recombination in yeast: control of incidence 741

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

elongation than hpr1 and indicates a completecorrelation between impairment of transcriptionand recombination in a given sequence.172 Byconstructing recombination systems under the tran-scriptional control of the GAL1-regulated promoterlocated outside of the repeats, it has also beendemonstrated that the hyper-rec effect of tho2 isonly observed when transcription is active.172

In addition to the stimulation of deletions causedby hpr1 and tho2 mutants, other forms of geneticinstability (such as plasmid loss) have beenobserved in association with the impairment oftranscription elongation. In both types of mutants,the stability of centromeric plasmids is reducedwhen transcription of a plasmid-borne gene isactivated, their levels of instability being dependenton the sensitivity of the transcribed sequence to thehpr1 and tho2 mutation.29,172 These results can beinterpreted as if the hyper-recombination phenotypewere not mainly the consequence of an increase inthe recruitment of the recombination machinerybut, rather, an increase in DNA damage orreplication errors able to initiate recombinationwhen occurring between repeats. If no repeats arepresent, the stability of the replication unit becomescompromised.

Integrated view of transcription-associatedrecombination

Can transcription elongation impairment explainother cases of transcription-induced recombination?Additional studies would be necessary to give ade®nitive answer, but some details suggest that thismight be the case. In all cited cases, transcription isnot marginally present in one locus. Instead, theelongating RNA polymerase traverses either onerepeat unit or the whole recombination system; thisis the case for HOT1.228 In the GAL10-dependentrecombination assay of Thomas and Rothstein,218

the promoter lies between the repeats and istranscribed outwards from the repeat unit towhich it is fused. Interestingly, in all cases inwhich recombination was stimulated, an mRNAwas found that initiated at the GAL10 promoterand entered the bacterial DNA sequences of theintervening region. Although the importance of thistranscript has not been evaluated in connectionwith the stimulation of deletions, it is noteworthythe parallelism between the presence of such atranscript and the need for transcription to occur

through the sequences ¯anked by the repeats inhpr1 and tho2 in all cases in which recombination isstimulated. Another important coincidence ofHOT1, GAL10 and hpr1- and tho2-dependentrecombination is the strong stimulation of dele-tions, whereas little or no effect is observed on geneconversion. We now know that deletions occurringbetween direct repeats are the result of a non-conservative recombination event, presumablysingle-strand annealing or one-ended cross-over.This has been shown for hpr1D mutants186 and itis presumably the situation for the other casesreported. The most important feature that differ-entiates deletions from other recombination events,such as gene conversions or reciprocal exchanges, isthat they can be initiated in the heterologous region¯anked by the repeats.54,55,147,174,208 As observed inhpr1 and tho2D strains, transcription-inducedrecombination is mostly initiated at the region¯anked by the repeats, where transcription elonga-tion impairment is believed to occur. This could bethe case of HOT1- and GAL10-promoted recombi-nation, even though initiation of recombination atthe DNA repeat that is transcribed also occurs.Finally, in all cases, there is a similar dependencyon RAD1, RAD10 and RAD52 of the induceddeletion events.156,186,204,217 Taken together, thesedata suggest that the type of recombinationmechanism triggered by transcription could be thesame in all cases reported.

It has been proposed that stalled or blockedtranscription complexes can trigger recombinationevents.29,176 It might confer nuclease hypersensitivesites as a consequence of an open chromatinstructure or partially unwound DNA, it mightfacilitate the attack of DNA damaging agents, orit may interfere with the replication machinery.More experiments are required to solve this puzzle.However, it is worth discussing the hypothesis thattranscription-associated recombination might beinduced by an arrest of the replication fork aftercolliding with a stalled or blocked transcriptioncomplex29,176 (see Figure 5). Although originallyproposed for hpr1 recombination events,176 thishypothesis could be extended to the other casesreported. As mentioned earlier, HOT1 activityrequires RFB replication fork blocking activity.Indeed, mutations in the FOB1 gene, required forRFB activity, abolish the Hot1 phenotype.38,89 Thiscertainly suggest that HOT1-induced recombinationis mediated by replication fork blocking. Transcrip-

742 A. Aguilera, S. ChaÂvez and F. MalagoÂn

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

tion elongation may ®rst provoke the collision with

the replication fork and the RFB activity would

then be required for blockage of replication, so that

a deletion would occur by either single-strand

annealing, one-ended invasion or even sister chro-

matid replication slippage.Other types of genetic rearrangements can be

potentiated by transcription in yeast. The most

relevant is the destabilization of simple repeti-

tive DNA sequences. GAL1-induced transcription

through a poly GT tract destabilizes it four- to

nine-fold.234 The effect of transcription is signi®cant

even in msh2 and pms1 mismatch repair-de®cient

mutants, indicating that the observed phenomenon

cannot be explained by a reduction in the ef®ciency

of DNA mismatch repair, but is very likely due to

an increase in the error rate of DNA polymerase. It

would certainly be interesting to know whether

microsatellite instability and deletions of long DNA

repeats are the result of the same molecular event in

connection with transcription. As suggested for

hpr1- and tho2-stimulated recombination events,

the interference of the transcription machinery

with the replication fork during elongation is an

attractive alternative to explain microsatellite

instability.

Transcription-induced recombination in otherorganisms

The interconnection between transcription and

recombination is not exclusive to yeast. On the

contrary, it seems to be a general phenomenon

observed from bacteria to mammals. Some impor-

tant examples are mentioned in this review. The

®rst description of a transcription-induced recombi-

nation process was the Rpo-mediated recombina-

tion of phage lambda in Escherichia coli.91 Rpo

recombination is induced by RNA polymerase, is

recA-independent, and is distinct from Red, Int,

RecBC and RecE pathways. The cross-over occurs

within a narrow region of the phage genome that is

actively transcribed. If the transcribed region is

extended by a rho mutation, affecting transcription

termination, the cross-over region is extended in

parallel, indicating that RNA chain elongation is

the transcription step required for recombination in

bacteria also.91 Other examples of transcription-

facilitated homologous recombination have been

described by studying specialized transduction by

phage l and generalized transduction by phages T1and T4.39,40

Deletions produced by illegitimate recombinationin E. coli are also induced by transcription. pBR332derivatives carrying a pTac promoter suffer dele-tions at a 10x2 rate. Deletion formation is stronglydependent on the strength of the promoter, thetranscript length, and the relative orientation of thepromoter and the origin of replication.227 Indeed,B. Michel's group has suggested that these are theresult of a collision between converging replicationand transcription machineries.227

In the yeast Sz. pombe, transcription driven fromthe ADH1 promoter has also been shown tostimulate recombination,72 consistent with theresults of S. cerevisiae. In mammalian cells, thereare different examples of transcription-inducedrecombination. Thus, it has been shown that atranscribed target site improves the chances of genetargeting in human cells221 or that two integratedcopies of neomycin genes recombine, in Chinesehamster ovary cells, between three- and seven-foldmore frequently when one of the repeats istranscriptionally induced. Interestingly, both directand inverted repeat systems are stimulated and, inthis case, most of the events are gene conver-sions.158

The most relevant case of the in¯uence oftranscription on recombination in animals is prob-ably the assemble of the variable region ofinmunoglobulins and class-switching. The transcrip-tional activity of ¯anking sequences plays a directrole in regulating V(D)J recombination.13 Theelimination of a transcriptional enhancer and apromoter causes a 20- to 100-fold reduction in thefrequency of rearrangements, whereas the elimina-tion of a transcriptional silencer causes a ®ve-foldincrease.123,161 The molecular mechanism by whichtranscription stimulates V(D)J recombination isnot clear, but it has been proposed that transcrip-tional activity increases the accessibility of RAGrecombinases to their substrate sites (RSS). Con-sistent with this hypothesis, a study on the cleavageaccessibility of RSS in different cell types anddevelopmental stages shows that cell-type speci®cchromatin structure determines the targeting ef®-ciency of the RAG proteins and that chromatinstructure plays a role in the allelic exclusionmechanism.203 Class switch recombination has alsobeen shown to be enhanced by transcription.36 Inthis case, RNA: DNA hybrid formation has been

Mitotic recombination in yeast: control of incidence 743

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

proposed to be an intermediate in the process ofstimulation.35

Therefore, it is clear that the transcriptionactivity and/or the transcription machinery providea scenario in which DNA exchanges are stronglyfacilitated in organisms from bacteria to mammals.However, since neither transcription and recombi-nation in vivo involves naked DNA but, rather, ahigher order protein±DNA structure (chromatin ineukaryotes), we cannot ignore the fact that the ®rstconsequence of transcription is the remodelling ofchromatin structure. Chromatin structure affectsrecombination in both mitosis and meiosis. We willreview these effects in mitosis only. As we will see, itis certainly possible that the effects of chromatinand transcription on mitotic recombination areconsequences of similar types of DNA disturbance.

Role of chromatin structure on mitoticrecombination

As already mentioned, transcription-driven recom-bination may be an indirect consequence of theeffect of transcription on chromatin. Thus, unwind-ing of DNA, or changes in the local supercoilingcaused by transcription, might be responsible forrecombination stimulation. Since recombinationoccurs between DNA molecules organized inchromatin, differences in chromatin structure canexplain the opposite behaviour of two giventranscription units. It is plausible that a change inchromatin structure facilitates the access of recom-bination proteins, leads to hypersensitivity tonucleases and endogenous DNA-damaging agents,or stimulates a pairing reaction. Chromatin struc-ture is therefore an important factor to beconsidered when trying to analyse the effectof transcription on recombination and geneticrearrangement.

Although invoked frequently to explain results,evidence connecting chromatin organization andrecombination is scarce. One good example is HOendonuclease-induced mating-type switching. Haberand collaborators have investigated the require-ments for the different elements of the recombina-tion apparatus in this process of gene conversion.They found that, when the HML or HMR donorsadopt the usual silenced and inaccessible chromatinstructure, the gene conversion event requiredRad51p, Rad52p, Rad54p, Rad55p and Rad57p.

However, when the donor is not silenced andlocated on a plasmid, Rad52p alone is enough tocarry out the recombination event.209 They con-clude that the other Rad proteins, not needed whenthe donor is episomal, are required to facilitatestrand invasion into the otherwise innaccesibledonor sequences. This result re¯ects different needsfor the recombination machinery to complete arecombination event that depends on the chromatinstatus of the DNA.

SIR2

A very good example of the role of chromatinstructure on recombination is the effect of tran-scriptional silencing on the reduction of bothmitotic and meiotic intrachromosomal recombina-tion between the tandem repeats of the rDNA locus.The key protein mediating this repression is Sir2p,which is also involved in maintaining transcrip-tional repression at the silent mating type loci andtelomeres.69A sir2 mutant shows a 10- and 15-foldincrease in recombination that is speci®c for rDNAduplication. This increase in recombination isstrictly dependent on Rad52p (and on Rad50p inmeiosis), whereas the basal recombination in SIR2strains is independent of those functions, suggestingthat Sir2p is involved in excluding the rDNA locusfrom the recombination apparatus.69 Consistentwith this idea, Esposito and collaborators havealso demostrated, by assays of sensitivity to micro-ccocal nuclease and dam methyltransferase, thatSir2p is required for a more closed chromatinstructure in two regions of the rDNA repeats.56

rDNA silencing and chromatin accessibility respondto SIR2 dosage and, at the same time, silencinginversely correlates with recombination, indicatinga double effect of closed chromatin in repressingboth transcription and recombination.56 The differ-ent requirements of basal and sir2-induced recom-bination for RAD52 argue against transcriptionalactivity as an intermediate between chromatinstructure and recombination. In accordance withthis, the sir2 mutation seems not to in¯uence HOT1activity.

SPT/SIN

In the previous cases, chromatin structures seems toin¯uence recombination by conditioning the ef®-ciency of the cellular machinary for DNA transac-tions within the locus. Both cases re¯ect how

744 A. Aguilera, S. ChaÂvez and F. MalagoÂn

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

silenced DNA (heterochromatin) is a barrier toboth transcription and recombination. A differentstudy using intrachromosomal repeat system sug-gests that chromatin alterations in the euchromatinmight also be responsible for higher levels ofrecombination. MalagoÂn and Aguilera139 measuredthe recombination rates of nine different mutantsaffected in SWI/SNF and SPT/SIN genes. TheSWI/SNF genes code for the subunits of one ofthe main chromatin remodelling complexes neededfor transcriptional induction of a broad set ofgenes,168 whereas SPT/SIN genes encode proteinsthat in¯uence chromatin structure in a negative wayfor transcription.236 The spt4 and spt6 mutationshave a hyper-recombination phenotype that coversreciprocal exchanges, gene conversions and dele-tions between direct repeats. The broad range ofrecombination events suggests that the effect of themutations is exerted at initiation and that ade®ciently packed chromatin can be more accessibleto recombinagenic agents.139 An interesting ques-tion, yet to be answered, is whether the hyper-receffect of spt4 and spt6 is mediated by transcription,as Spt4p and Spt6p have been involved in chroma-tin-mediated repression of transcription.236 SPT6has been shown to contact with histone H3 and tocontrol chromatin structure.16 In addition, Winstonand colaborators have found a direct relationshipbetween SPT4, SPT5 and SPT6 and transcriptionelongation by RNA polymerase II in vivo.83 Adifferent question emerging from these results iswhether the hyper-rec phenotype of the spt muta-tions is also mediated by transcription elongation.In this sense, it is noteworthy that the hyper-recmutation hpr1 also displays genetic interactionswith mutants affected in chromatin structure.Thus, deletion of SIN1, encoding an HMG1-likeprotein, partially suppresses the transcriptionalphenotype of hpr1, and elevated gene dosage ofindividual histones shows severe growth defects incombination with hpr1.241

DNA structure and topological constraints

Another aspect of DNA structure that has beenrelated to mitotic recombination is the topologicalstate of chromatin. The ®rst genetic approach tothis ®eld was made by Fink and collaborators.31

They found that mitotic recombination in therDNA locus, but not in RNA pol II-transcribedloci, is stimulated in the absence of topoisomerase

activities. Both top1 and top2 mutants show 50- to200-fold higher frequencies of recombination rela-tive to wild types. The same results were basicallyobtained by Kim and Wang104 who, in addition,observed that the deleted copies of rDNA in a top1top2 mutant accumulate as extrachromosomal ringsthat can be integrated back into the chromosome iftopoisomerase activity is recovered. Both groupspostulate the accumulation of topological stressresulting from a high transcriptional activity as apossible explanation for the observed hyper-recombination. Topoisomerase activity (either I orII) is, in fact, essential for RNA synthesis,22 as it isrequired for transcriptional elongation by RNApolymerase I.189 This requirement is strengthened ina mutant lacking subunit A34.5 of RNA polymer-ase I, as it becomes quasi-essential in a top1background.57

Mutants affected in TOP1 and/or TOP2 genesare not reported to be hyper-rec in systemstranscribed by RNA polymerase II. This is inagreement with the observation that topoisomerasesI and II are not essential for RNA polymeraseII-dependent transcription.22 However DNA-supercoiling can be accumulated to a certainextent by RNA polymerase II activity in themutant background,23 suggesting that accumulationof DNA supercoiling per se is not suf®cient toinduce recombination. One additional factor thatdifferentiates the effect of topoisomerase mutantson transcription by RNA polymerases I and II isthe behaviour of the chromatin templates. In theabsence of topoisomerases I and II, a severetranscription-dependent chromatin transition canbe detected in rDNA, as analysed by nucleasedigestion and psoralen cross-linking, but not in agene strongly transcribed by RNA polymerase II.27

A disrupted chromatin structure under topologicalstress might therefore be the optimum scenario fortranscription-stimulated recombination. The differ-ent behaviour of RNA polymerases I and II in thisrespect opens the possibility of the existence ofredundant topological activities operating in RNApolymerase II-dependent transcription.

It is certainly signi®cant that the third eukaryotictopoisomerase, Top3p, was uncovered by mutationsconferring hyper-recombination between directrepeats in yeast. Repeated sequences and therDNA region are very instable in top3 mutants aswell. TOP3 codes for a topoisomerase similar tobacterial type I enzyme.229 Mutants affected in

Mitotic recombination in yeast: control of incidence 745

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

TOP3 show a very poor growth phenotype that canbe suppressed by mutations in a second gene, SGS1,encoding a reverse gyrase activity.63 Sgs1p interactsphysically with Top3p and Top2p, and sgs1 muta-tions also suppress the instability in the rDNA locuscaused by top3. Based on this observation, it hasbeen proposed that Top3p acts on DNA substratesmade by Sgs1p.62 SGS1 is, in fact, homologous totwo human genes involved in genome disordersleading to premature aging: the Bloom's and theWerner's syndrome genes.231,232 Mutations in SGS1also show a broad hyper-rec phenotype, as theypresent an increased frequency of interchromoso-mal homologous recombination, intrachromosomalexcision recombination, and ectopic recombina-tion.63,232 Altogether, these results indicate that thecomplex formed by Top3p and Sgs1p is veryimportant for maintaining genome integrity. HowTop3p and Sgs1p, as well the other topoisomerases,affect recombination is yet to be elucidated. Thework of Harmon et al.81 on the E. coli homologuesTopIII and RecQ suggests that the catenation/decatenation activity of Top3 would be required tosuppress recombination between repeats. TheSgs1p±Top3p complex would suppress recombina-tion by its strand passage activity.63,81 When tworeplication forks collide, the unreplicated interforkregion would become highly underwound andinaccessible to Top3p making the helicase activityof Sgs1p necessary for Top3 to access the regionand decatenate the replicated DNA.

Finally, speci®c DNA regions have been shownto act as mitotic hot-spots for recombination. Thisis the case of the 1.5 kb region found by Neitz andCarbon.154 This is an AT-rich DNA sequence that,when introduced into a linear plasmid, causes theaccumulation of linear forms of the plasmids,suggesting that it may act as a target for DSBs,whether or not mediated by nucleases. It wouldcertainly be interesting to know which particularfeature makes this sequence behave as a hot-spotfor mitotic recombination.

Effect of chromatin and DNA structures inrecombination in other organisms

Not many examples are available about the in¯u-ence of chromatin organization on recombinationin non-yeast systems. Among those we have alreadymentioned, the study on the cleavage accessibility ofRSS in different cell types and developmental stages

in V(D)J recombination shows that cell-type speci®cchromatin structure determines the targeting ef®-ciency of the RAG proteins.203

The possibility that supercoiling can generaterecombinagenic DNA structures has also beenstudied in E. coli. Kohwi and Panchenko112 haveshown that direct repeat sequences separated byeither 200 bp or 1000 bp and containing 25±30 bppoly(dG)-poly(dC) sequences in the 5k region of apTac promoter induce homologous recombinationwhen transcription is activated. In this scenario,negative supercoiling is accumulated at the 5k regionof the promoter, inducing the formation of a triple-helix structure at the poly(dG)-poly(dC) tract. Theypostulate that such a triple structure causesenhancement of recombination, because no hyper-recombination was observed when the dG tract wasnot present. The intramolecular triples would foldDNA, bringing the repeated DNA sequences intoproximity, so facilitating recombination. Althoughthis model may be valid for proximal direct repeatsin bacteria, it is not clear whether it can explain thecases of transcription-induced recombination inyeast, in which direct repeats are separated bymore than 10 kb and are placed in the context ofchromatin.

Finally, although eubacterial chromatin is quitedifferent compared with the eukaryotic DNAorganization, published mechanistic informationabout the effect of nucleosomes on recombinationreactions comes from in vitro experiments withreconstituted nucleosomes and bacterial recom-bination proteins. Thus, Kotani and Kmiec116

have observed that transcription activates RecA-promoted homologous pairing of nucleosomalDNA. On the other hand, Grigoriev and Hsieh70

have shown that a histone octamer blocks branchmigration of a Holliday junction when reconstitutedonto an appropriate DNA molecule. They subse-quently demonstrated that the Holliday junctioncannot migrate through the nucleosomal core unlessDNA±histone interactions are completely dis-rupted. Grigoriev and Hsieh70 propose that, duringrecombination, enzymes are required not only toaccelerate the intrinsic branch migration but also tofacilitate the passage through chromatin. They haverecently shown that the E. coli RuvAB protein isable to drive the migration of a Holliday junctionthrough a nucleosome in the presence of ATP.71

These results predict that eukaryotic cells havedeveloped biochemical functions to allow recombi-

746 A. Aguilera, S. ChaÂvez and F. MalagoÂn

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

nation to proceed through chromatin. It is certainlyattractive to think that the transcriptional machin-ery may well do such a job.

Concluding remarks

Mitotic recombination is one form of DNA repairthat preferentially uses sister chromatids to repairDNA breaks, many of which may be generatedduring replication or by DNA damage. The analysisof hyper-recombination mutations and the effect ofDNA-damaging agents on recombination is a verygood approach to understanding the molecularbasis of the origin of mitotic recombination.Initiation of recombination in mitosis may beaffected by many elements and processes such asNER, BER, replication, transcription and chroma-tin structure. Particularly intriguing is the connec-tion with transcription and chromatin structure; itis important that these processes be deciphered atthe molecular level to obtain a complete knowledgeof the mechanisms of recombination and theirrelevance in the stability of the genome in yeastand other organisms. The systematic functionalanalysis of the yeast genome provides a novelapproach to identifying new hyper-recombinationmutants. However, we still need much more geneticand molecular analyses to understand both hyper-recombination and the molecular basis of initiationof spontaneous mitotic recombination. This is ofgreat relevance in order to identify the causes andrisk factors of genomic instability in eukaryotes.And yeast are, certainly, the best eukaryotic systemfor studying mitotic recombination in vivo.

Acknowledgements

We would like to thank F. Prado for reading the manuscript

and W. Reven for style correction. This work has been

funded by the EU (BIO4-CT97-2294) and the Ministry of

Education and Culture of Spain (Grants PB96-1350 and

BIO98-1363-CE).

References

1. Aguilera A. Genetic evidence for different RAD52-depen-

dent intrachromosomal recombination pathways in Sac-

charomyces cerevisiae. Curr Genet 1995; 27: 298±305.

2. Aguilera A, Klein HL. Genetic control of intrachromoso-

mal recombination in Saccharomyces cerevisiae. I. Isolation

and genetic characterization of hyper-recombination muta-

tions. Genetics 1988; 119: 779±790.

3. Aguilera A, Klein HL. Genetic and molecular analysis of

recombination events in Saccharomyces cerevisiae occurring

in the presence of the hyper-recombination mutation hpr1.

Genetics 1989; 122: 503±517.

4. Aguilera A, Klein HL. HPR1, a novel yeast gene that

prevents intrachromosomal excision recombination, shows

carboxy-terminal homology to the Saccharomyces cerevisiae

TOP1 gene. Mol Cell Biol 1990; 10: 1439±1451.

5. Aguilera A, Klein HL. Hyperrecombination mutations in

Saccharomyces cerevisiae. Methods Mol Genet 1994; 3:

107±130.

6. Ajimura M, Leem SH, Ogawa H. Identi®cation of new

genes required for meiotic recombination in Saccharomyces

cerevisiae. Genetics 1993; 133: 51±66.

7. Bai Y, Symington LS. A Rad52 homolog is required for

RAD51-independent mitotic recombination in Saccharo-

myces cerevisiae. Genes Dev 1996; 10: 2025±2037.

8. Bai Y, Davis AP, Symington LS. A novel allele of RAD52

that causes severe DNA repair and recombination de®cien-

cies only in the absence of RAD51 or RAD59. Genetics

1999; 153: 1117±1130.

9. Bailis AM, Arthur L, Rothstein R. Genome rearrangement

in top3 mutants of Saccharomyces cerevisiae requires a

functional RAD1 excision repair gene. Mol Cell Biol 1992;

12: 4988±4993.

10. Belmaaza A, Chartrand P. One-sided invasion events in

homologous recombination at double-strand breaks. Mutat

Res 1994; 314: 199±208.

11. Bergerat A, de Massy B, Gadelle D, Varoutas PC, Nicolas

A, Forterre P. An atypical topoisomerase II from Archaea

with implications for meiotic recombination. Nature 1997;

386: 414±417.

12. Bierne H, Michel B. When replication forks stop. Mol

Microbiol 1994; 13: 17±23.

13. Blackwell TK, Moore MW, Yancopoulos GD, Suh H,

Lutzker S, Selsing E, Alt FW. Recombination between

immunoglobulin variable region gene segments is enhanced

by transcription. Nature 1986; 324: 585±589.

14. Boram WR, Roman H. Recombination in Saccharomyces

cerevisiae: a DNA repair mutation associated with elevated

mitotic gene conversion. Proc Natl Acad Sci USA 1976; 73:

2828±2832.

15. Borts RH, Haber JE. Meiotic recombination in yeast:

alteration by multiple heterozygosities. Science 1987; 237:

1459±1465.

16. Bortvin A, Winston F. Evidence that Spt6p controls

chromatin structure by a direct interaction with histones.

Science 1996; 272: 1473±1476.

17. Boulton SJ, Jackson SP. Identi®cation of a Saccharomyces

cerevisiae Ku80 homologue: roles in DNA double strand

break rejoining and in telomeric maintenance. Nucleic Acids

Res 1996; 24: 4639±4648.

18. Boulton SJ, Jackson SP. Saccharomyces cerevisiae Ku70

potentiates illegitimate DNA double-strand break repair

and serves as a barrier to error-prone DNA repair

pathways. EMBO J 1996; 15: 5093±5103.

19. Bratty J, Ferbeyre G, Molinaro C, Cedergren R. Stimula-

tion of mitotic recombination upon transcription from the

Mitotic recombination in yeast: control of incidence 747

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

yeast GAL1 promoter but not from other RNA polymerase

I, II and III promoters. Curr Genet 1996; 30: 381±388.

20. Brennan RJ, Swoboda BE, Schiestl RH. Oxidative muta-

gens induce intrachromosomal recombination in yeast.

Mutat Res 1994; 308: 159±167.

21. Brewer BJ, Fangman WL. A replication fork barrier at the

3k end of yeast ribosomal RNA genes. Cell 1988; 55:

637±643.

22. Brill SJ, DiNardo S, Voelkel-Meiman K, Sternglanz R.

Need for DNA topoisomerase activity as a swivel for DNA

replication for transcription of ribosomal RNA. Nature

1987; 326: 414±416.

23. Brill SJ, Sternglanz R. Transcription-dependent DNA

supercoiling in yeast DNA topoisomerase mutants. Cell

1988; 54: 403±411.

24. Bruschi CV, Esposito M. Enhancement of spontaneous

mitotic recombination by the meiotic mutant spo11-1 in

Saccharomyces cerevisiae. Proc Natl Acad Sci USA 1983;

80: 7566±7570.

25. Bruschi CV, McMillan JN, Coglievina M, Esposito MS.

The genomic instability of yeast cdc6-1/cdc6-1 mutants

involves chromosome structure and recombination. Mol

Gen Genet 1995; 249: 8±18.

26. Carroll D. Modes of DNA repair in Xenopus oocytes, eggs

and extracts. In DNA Damage and Repair, vol I: DNA

repair in prokaryotes and lower eukaryotes, Nickoloff JA,

Hoekstra MF, (eds). Humana Press: Totowa, NJ; 1998;

597±616.

27. Cavalli G, Bachmann D, Thoma F. Inactivation of

topoisomerases affects transcription-dependent chromatin

transitions in rDNA but not in a gene transcribed by RNA

polymerase II. EMBO J 1996; 15: 590±597.

28. Chamankhah M, Xiao W. Molecular cloning and genetic

characterization of the Saccharomyces cerevisiae NGS1/

MRE11 gene. Curr Genet 1998; 34: 368±374.

29. ChaÂvez S, Aguilera A. The yeast HPR1 gene has a

functional role in transcriptional elongation that uncovers

a novel source of genome instability. Genes Dev 1997; 11:

3459±3470.

30. Chen C, Merrill BJ, Lau PJ, Holm C, Kolodner RD.

Saccharomyces cerevisiae pol30 (proliferating cell nuclear

antigen) mutations impair replication ®delity and mismatch

repair. Mol Cell Biol 1999; 19: 7801±7815.

31. Christman MF, Dietrich FS, Fink GR. Mitotic recombina-

tion in the rDNA of S. cerevisiae is suppressed by the

combined action of DNA topoisomerases I and II. Cell

1988; 55: 413±425.

32. Cox MM. A broadening view of recombinational DNA

repair in bacteria. Genes Cells 1998; 3: 65±78.

33. Critchlow SE, Jackson SP. DNA end-joining: from yeast to

man. Trends Biochem Sci 1998; 23: 394±398.

34. Cundari E, Vellosi R, Galli A, Bronzetti G. Inducibility of

gene conversion in Saccharomyces cerevisiae treated with

MMS. Mutat Res 1986; 174: 271±274.

35. Daniels GA, Lieber MR. RNA : DNA complex formation

upon transcription of immunoglobulin switch regions:

implications for the mechanism and regulation of class

switch recombination. Nucleic Acids Res 1995; 23:

5006±5011.

36. Daniels GA, Lieber MR. Strand speci®city in the transcrip-

tional targeting of recombination at immunoglobulin switch

sequences. Proc Natl Acad Sci USA 1995; 92: 5625±5629.

37. Dardalhon M, de Massy B, Nicolas A, Averbeck D. Mitotic

recombination and localized DNA double-strand breaks are

induced after 8-methoxypsoralen and UVA irradiation in

Saccharomyces cerevisiae. Curr Genet 1998; 34: 30±42.

38. Defossez PA, Prusty R, Kaeberlein M, Lin SJ, Ferrigno P,

Silver PA, Keil RL, Guarente L. Elimination of replication

block protein Fob1 extends the life span of yeast mother

cells. Mol Cell 1999; 3: 447±455.

39. Dul JL, Drexler H. Transcription stimulates recombination.

II. Generalized transduction of Escherichia coli by phages

T1 and T4. Virology 1988; 162: 471±477.

40. Dul JL, Drexler H. Transcription stimulates recombination.

I. Specialized transduction of Escherichia coli by lambda trp

phages. Virology 1988; 162: 466±470.

41. Eckardt-Schupp F, Klaus C. Radiation inducible DNA

repair processes in eukaryotes. Biochimie 1999; 81: 161±171.

42. Elble R, Tye BK. Chromosome loss, hyperrecombination,

and cell cycle arrest in a yeast mcm1 mutant. Mol Biol Cell

1992; 3: 971±980.

43. Ernst JF, Stewart JW, Sherman F. The cyc1-11 mutation in

yeast reverts by recombination with a nonallelic gene:

composite genes determining the iso-cytochromes c. Proc

Natl Acad Sci USA 1981; 78: 6334±6338.

44. Esposito MS, Wagstaff JE. Mechanisms of mitotic recom-

bination. In The Molecular Biology of the Yeast Saccharo-

myces: Life Cycle and Inheritance, Strathern JN, Jones EW,

Broach JR (eds). Cold Spring Harbor Laboratory Press:

New York; 1981; 341-370.

45. Fabre F, Roman H. Genetic evidence for inducibility of

recombination competence in yeast. Proc Natl Acad Sci

USA 1977; 74: 1667±1671.

46. Fan HY, Klein HL. Characterization of mutations that

suppress the temperature-sensitive growth of the hpr1D

mutant of Saccharomyces cerevisiae. Genetics 1994; 137:

945±956.

47. Fan HY, Cheng KK, Klein HL. Mutations in the RNA

polymerase II transcription machinery suppress the hyper-

recombination mutant hpr1D of Saccharomyces cerevisiae.

Genetics 1996; 142: 749±759.

48. Fasullo M, Bennett T, AhChing P, Koudelik J. The

Saccharomyces cerevisiae RAD9 checkpoint reduces the

DNA damage-associated stimulation of directed transloca-

tions. Mol Cell Biol 1998; 18: 1190±1200.

49. Fasullo M, Koudelik J, AhChing P, Giallanza P, Cera C.

Radiosensitive and mitotic recombination phenotypes of

the Saccharomyces cerevisiae dun1 mutant defective in DNA

damage-inducible gene expression. Genetics 1999; 152:

909±919.

50. Featherstone C, Jackson SP. DNA double-strand break

repair. Curr Biol 1999; 9: 759±761.

51. Fedorova IV, Gracheva LM, Kovaltzova SV, Evstuhina

TA, Alekseev SY, Korolev VG. The yeast HSM3 gene acts

in one of the mismatch repair pathways. Genetics 1998; 148:

963±973.

52. Ferguson LR, Turner PM, Baguley BC. Induction of

mitotic crossing-over by the topoisomerase II poison

DACA (N-[2-dimethylamino)ethyl]acridine-4-carboxamide)

in Saccharomyces cerevisiae. Mutat Res 1993; 289: 157±163.

53. Fiorentino DF, Crabtree GR. Characterization of Saccha-

748 A. Aguilera, S. ChaÂvez and F. MalagoÂn

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

romyces cerevisiae dna2 mutants suggests a role for the

helicase late in S phase. Mol Biol Cell 1997; 8: 2519±2537.

54. Fishman-Lobell J, Haber JE. Removal of non-homologous

DNA ends in double-strand break recombination. Science

1992; 258: 480±484.

55. Fishman-Lobell J, Rudin N, Haber JE. Two alternative

pathways of double-strand break repair that are kinetically

separable and independently modulated. Mol Cell Biol

1992; 12: 1292±1303.

56. Fritze CE, Verschueren K, Strich R, Easton Esposito R.

Direct evidence for SIR2 modulation of chromatin structure

in yeast rDNA. EMBO J 1997; 16: 6495±6509.

57. Gadal O, Mariotte-Labarre S, Chedin S, Quemeneur E,

Carles C, Sentenac A, Thuriaux P. A34.5, a non-essential

component of yeast RNA polymerase I, cooperates with

subunit A14 and DNA topoisomerase I to produce a

functional rDNA synthesis machine. Mol Cell Biol 1997; 17:

1787±1795.

58. Galli A, Schiestl RH. Hydroxyurea induces recombination

in dividing but not in G1 and G2 cell cycle arrested yeast

cells. Mutat Res 1996; 354: 69±75.

59. Galli A, Schiestl RH. Effects of DNA double-strand and

single-strand breaks on intrachromosomal recombination

events in cell-cycle-arrested yeast cells. Genetics 1998; 149:

1235±1250.

60. Galli A, Schiestl RH. Cell division transforms mutagenic

lesions into deletion-recombinagenic lesions in yeast cells.

Mutat Res 1999; 429: 13±26.

61. Game JC, Johnston LH, von Borstel RC. Enhanced mitotic

recombination in a ligase-defective mutant of the yeast

Saccharomyces cerevisiae. Proc Natl Acad Sci U S A 1979;

76: 4589±4592.

62. Gangloff S, Lieber MR, Rothstein R. Transcription,

topoisomerases and recombination. Experientia 1994; 50:

261±269.

63. Gangloff S, McDonald JP, Bendixen C, Arthur L,

Rothstein R. The yeast type I topoisomerase Top3 interacts

with Sgs1, a DNA helicase homolog: a potential eukaryotic

reverse gyrase. Mol Cell Biol 1994; 14: 8391±8398.

64. Gehring WJ. Genetics Mosaics and Cell Differentiation.

Springer-Verlag: New York; 1978.

65. Gloor GB, Lankenau DH. Gene conversion in mitotically

dividing cells: a view from Drosophila. Trends Genet 1998;

14: 43±46.

66. Golin JE, Esposito MS. Evidence for joint genic control of

spontaneous mutation and genetic recombination during

mitosis in Saccharomyces. Mol Gen Genet 1977; 150:

127±135.

67. Gordenin DA, Lobachev KS, Degtyareva NP, Malkova

ALPerkinsE, Resnick MA. Inverted DNA repeats: a source

of eukaryotic genomic instability. Mol Cell Biol 1993; 13:

5315±5322.

68. Gordenin DA, Proscyavichus YY, Malkova AL, Tro®mova

MV, Peterzen A. Yeast mutants with increased bacterial

transposon Tn5 excision. Yeast 1991; 7: 37±50.

69. Gottlieb S, Esposito RE. A new role for a yeast transcrip-

tional silencer gene, SIR2, in regulation of recombination in

ribosomal DNA. Cell 1989; 56: 771±776.

70. Grigoriev M, Hsieh P. A histone octamer blocks branch

migration of a Holliday junction. Mol Cell Biol 1997; 17:

7139±7150.

71. Grigoriev M, Hsieh P. Migration of a Holliday junction

through a nucleosome directed by the E. coli RuvAB motor

protein. Mol Cell 1998; 2: 373±381.

72. Grimm C, Schaer P, Munz P, Kohli J. The strong ADH1

promoter stimulates mitotic and meiotic recombination at

the ADE6 gene of Schizosaccharomyces pombe. Mol Cell

Biol 1991; 11: 289±298.

73. Grossenbacher-Grunder AM, Thuriaux P. Spontaneous

and UV-induced recombination in radiation-sensitive

mutants of Schizosaccharomyces pombe. Mutat Res 1981;

81: 37±48.

74. Gysler-Junker A, Bodi Z, Kohli J. Isolation and character-

ization of Schizosaccharomyces pombe mutants affected in

mitotic recombination. Genetics 1991; 128: 495±504.

75. Haber JE. In vivo biochemistry: physical monitoring of

recombination induced by site-speci®c endonucleases. Bio-

essays 1995; 17: 609±620.

76. Haber JE. A super new twist on the initiation of meiotic

recombination. Cell 1997; 89: 163±166.

77. Haber JE. Mating-type gene switching in Saccharomyces

cerevisiae. Ann Rev Genet 1998; 32: 561±599.

78. Haber JE. A locus control region regulates yeast recombi-

nation. Trends Genet 1998; 14: 317±321.

79. Haber JE. Sir-Ku-itous routes to make ends meet. Cell

1999; 97: 829±832.

80. Hannan MA, Zimmer SG, Hazle J. Mechanisms of cisplatin

(cis-diamminodichloroplatinum II)-induced cytotoxicity

and genotoxicity in yeast. Mutat Res 1984; 127: 23±30.

81. Harmon FG, DiGate RJ, Kowalczykowski SC. RecQ

helicase and topoisomerase III comprise a novel DNA

strand passage function: a conserved mechanism for control

of DNA recombination. Mol Cell 1999; 3: 611±620.

82. Hartwell LH, Smith D. Altered ®delity of mitotic chromo-

some transmission in cell cycle mutants of S. cerevisiae.

Genetics 110: 381±395.

83. Hartzog GA, Wada T, Handa H, Winston F. Evidence that

Spt4, Spt5, and Spt6 control transcription elongation by

RNA polymerase II in Saccharomyces cerevisiae. Genes Dev

1998; 12: 357±369.

84. Haynes RH, Kunz BA. DNA repair and mutagenesis in

yeast. In The Molecular Biology of the Yeast Saccharo-

myces: Life Cycle and Inheritance, Strathern JN, Jones EW,

Broach JR (eds). Cold Spring Harbor Laboratory Press:

New York; 1981; 371±414.

85. Henderson ST, Petes TD. Instability of a plasmid-borne

inverted repeat in Saccharomyces cerevisiae. Genetics 1993;

133: 57±62.

86. Hennessy KM, Lee A, Chen E, Botstein D. A group of

interacting yeast DNA replication genes. Genes Dev 1991; 5:

958±969.

87. Hiom K. Dna repair: Rad52Ðthe means to an end. Curr

Biol 1999; 9: 446±448.

88. Huang KN, Symington LS. Mutation of the gene encoding

protein kinase C1 stimulates mitotic recombination in

Saccharomyces cerevisiae. Mol Cell Biol 1994; 14:

6039±6045.

89. Huang GS, Keil RL. Requirements for activity of the yeast

mitotic recombination hotspot HOT1: RNA polymerase I

and multiple cis-acting sequences. Genetics 1995; 141:

845±855.

90. Humayun MZ. SOS and Mayday: multiple inducible

Mitotic recombination in yeast: control of incidence 749

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

mutagenic pathways in Escherichia coli. Mol Microbiol

1998; 30: 905±910.

91. Ikeda H, Matsumoto T. Transcription promotes recA-

independent recombination mediated by DNA-dependent

RNA polymerase in Escherichia coli. Proc Natl Acad Sci

U S A 1979; 76: 4571±4575.

92. Ivanov EL, Korolev VG, Fabre F. XRS2, a DNA repair

gene of Saccharomyces cerevisiae, is needed for meiotic

recombination. Genetics 1992; 132: 651±664.

93. Jablonovich Z, Liefshitz B, Steinlauf R, Kupiec M.

Characterization of the role played by the RAD59 gene of

Saccharomyces cerevisiae in ectopic recombination. Curr

Genet 1999; 36: 13±20.

94. Jachymczyk WJ, von Borstel RC, Mowat MR, Hastings PJ.

Repair of interstrand cross-links in DNA of Saccharomyces

cerevisiae requires two systems for DNA repair: the RAD3

system and the RAD51 system. Mol Gen Genet 1981; 182:

196±205.

95. Jackson JA, Fink GR. Gene conversion between duplicated

genetic elements in yeast. Nature 1981; 292: 306±311.

96. Jinks-Robertson S, Petes TD. High-frequency meiotic gene

conversion between repeated genes on nonhomologous

chromosomes in yeast. Proc Natl Acad Sci U S A 1985; 82:

3350±3354.

97. Jiricny J. Replication errors: cha(lle)nging the genome.

EMBO J 1998; 17: 6427±6436.

98. Kadyk LC, Hartwell LH. Sister chromatids are preferred

over homologs as substrates for recombinational repair in

Saccharomyces cerevisiae. Genetics 1992; 132: 387±402.

99. Kanaar R, Hoeijmakers JH. Recombination and joining:

different means to the same ends. Genes Funct 1997; 1:

165±174.

100. Kanaar R, Hoeijmakers JH, van Gent DC. Molecular

mechanisms of DNA double strand break repair. Trends

Cell Biol 1998; 8: 483±489.

101. Keeney S, Giroux CN, Kleckner N. Meiosis-speci®c DNA

double-strand breaks are catalyzed by Spo11, a member of

a widely conserved protein family. Cell 1997; 88: 375±384.

102. Keil RL, Roeder GS. Cis-acting, recombination-stimulating

activity in a fragment of the ribosomal DNA of S.

cerevisiae. Cell 1984; 39: 377±386.

103. Kern R, Zimmermann FK. The in¯uence of defects in

excision and error prone repair on spontaneous and

induced mitotic recombination and mutation in Saccharo-

myces cerevisiae. Mol Gen Genet 1978; 161: 81±88.

104. Kim RA, Wang JC. A subthreshold level of DNA

topoisomerases leads to the excision of yeast rDNA as

extrachromosomal rings. Cell 1989; 57: 975±985.

105. Kleckner N. Meiosis: how could it work?. Proc Natl Acad

Sci U S A 1996; 93: 8167±8174.

106. Klein HL, Petes TD. Intrachromosomal gene conversion in

yeast. Nature 1981; 289: 144±148.

107. Klein HL. Genetic control of intrachromosomal recombi-

nation. Bioessays 1995; 17: 147±159.

108. Kobayashi T, Hidaka M, Nishizawa M, Horiuchi T.

Identi®cation of a site required for DNA replication fork

blocking activity in the rDNA gene cluster in Saccharo-

myces cerevisiae. Mol Gen Genet 1992; 233: 355±362.

109. Kobayashi T, Horiuchi T. A yeast gene product, Fob1

protein, required for both replication fork blocking and

recombinational hotspot activities. Genes Cells 1996; 1:

465±474.

110. Kogoma T. Recombination by replication. Cell 1996; 85:

625±627.

111. Kogoma T. Stable DNA replication: interplay between

DNA replication, homologous recombination, and tran-

scription. Microbiol Mol Biol Rev 1997; 61: 212±238.

112. Kohwi Y, Panchenko Y. Transcription-dependent recombi-

nation induced by triple-helix formation. Genes Dev 1993; 7:

1766±1778.

113. Koleske AJ, Young RA. An RNA polymerase II holo-

enzyme responsive to activators. Nature 1994; 368: 466±469.

114. Konrad EB. Method for the isolation of Escherichia coli

mutants with enhanced recombination between chromoso-

mal duplications. J Bacteriol 1977; 130: 167±172.

115. Kouprina N, Kroll E, Bannikov V, Bliskovsky V, Gizatullin

R, Kirillov A, Shestopalov B, Zakharyev V, Hieter P,

Spencer F, et al. CTF4(CHL15) mutants exhibit defective

DNA metabolism in the yeast Saccharomyces cerevisiae.

Mol Cell Biol 1992; 12: 5736±5747.

116. Kotani H, Kmiec EB. Transcription activates RecA-

promoted homologous pairing of nucleosomal DNA. Mol

Cell Biol 1994; 14: 1949±1955.

117. Kowalczykowski SC, Dixon DA, Eggleston AK, Lauder

SD, Rehrauer WM. Biochemistry of homologous recombi-

nation in Escherichia coli. Microbiol Rev 1994; 58: 401±465.

118. Kowalski S, Laskowski W. The effect of three rad genes on

survival, inter and intragenic mitotic recombination in

Saccharomyces. Mol Gen Genet 1975; 136: 75±86.

119. Kozhina T, Kozhin S, Stepanova V, Yarovoy B, Donich V,

Fedorova I, Korolev V. UVS112Ða gene involved in

excision repair of yeast. Yeast 1995; 11: 1129±1138.

120. Kunes S, Botstein D, Fox MS. Transformation of yeast

with linearized plasmid DNA. Formation of inverted

dimers and recombinant plasmid products. J Mol Biol

1985; 184: 375±387.

121. Kuzminov A. Instability of inhibited replication forks in E.

coli. Bioessays 1995; 17: 733±741.

122. Kuzminov A. Recombinational repair of DNA damage in

Escherichia coli and bacteriophage lambda. Microbiol Mol

Biol Rev 1999; 63: 751±813.

123. Lauster R, Reynaud CA, Martensson IL, Peter A, Bucchini

D, Jami J, Weill JC. Promoter, enhancer and silencer

elements regulate rearrangement of an immunoglobulin

transgene. EMBO J 1993; 12: 4615±4623.

124. Levin NA, Bjornsti MA, Fink GR. A novel mutation in

DNA topoisomerase I of yeast causes DNA damage and

RAD9-dependent cell cycle arrest. Genetics 1993; 133:

799±814.

125. Lewis LK, Kirchner JM, Resnick MA. Requirement for

end-joining and checkpoint functions, but not RAD52-

mediated recombination, after EcoRI endonuclease clea-

vage of Saccharomyces cerevisiae DNA. Mol Cell Biol 1998;

18: 1891±1902.

126. Lieber MR, Grawunder U, Wu X, Yaneva M. Tying loose

ends: roles of Ku and DNA-dependent protein kinase in the

repair of double-strand breaks. Curr Opin Genet Dev 1997;

7: 99±104.

127. Liefshitz B, Parket A, Maya R, Kupiec M. The role of

DNA repair genes in recombination between repeated

sequences in yeast. Genetics 1995; 140: 1199±1211.

750 A. Aguilera, S. ChaÂvez and F. MalagoÂn

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

128. Liefshitz B, Steinlauf R, Friedl A, Eckardt-Schupp F,

Kupiec M. Genetic interactions between mutants of the

`error-prone' repair group of Saccharomyces cerevisiae and

their effect on recombination and mutagenesis. Mutat Res

1998; 407: 135±145.

129. Lin FL, Sperle K, Sternberg N. Model for homologous

recombination during transfer of DNA into mouse L cells:

role for DNA ends in the recombination process. Mol Cell

Biol 1984; 4: 1020±1034.

130. Linskens MH, Huberman JA. Organization of replication

of ribosomal DNA in Saccharomyces cerevisiae. Mol Cell

Biol 1988; 8: 4927±4935.

131. Lobachev KS, Shor BM, Tran HT, Taylor W, Keen JD,

Resnick MA, Gordenin DA. Factors affecting inverted

repeat stimulation of recombination and deletion in

Saccharomyces cerevisiae. Genetics 1998; 148: 1507±1524.

132. Longhese MP, Jovine L, Plevani P, Lucchini G. Condi-

tional mutations in the yeast DNA primase genes affect

different aspects of DNA metabolism and interactions in

the DNA polymerase a-primase complex. Genetics 1993;

133: 183±191.

133. Lovett ST, Drapkin PT, Sutera VAJr, Gluckman-Peskind

TJ. A sister-strand exchange mechanism for recA-

independent deletion of repeated DNA sequences in

Escherichia coli. Genetics 1993; 135: 631±642.

134. Lucchini G, Falconi MM, Pizzagalli A, Aguilera A, Klein

HL, Plevani P. Nucleotide sequence and characterization of

temperature-sensitive pol1 mutants of Saccharomyces cere-

visiae. Gene 1990; 90: 99±104.

135. Lustig AJ. The Kudos of non-homologous end-joining. Nat

Genet 1999; 23: 130±131.

136. Lupski JR. Genomic disorders: structural features of the

genome can lead to DNA rearrangements and human

disease traits. Trends Genet 1998; 14: 417±422.

137. Mahajan SK. Pathways of homologous recombination in

Escherichia coli. In Genetic Recombination, Kucherlapatti R,

Smith GR, (eds). ASM: Washington, DC; 1988; 89±140.

138. Maines S, Negritto MC, Wu X, Manthey GM, Bailis AM.

Novel mutations in the RAD3 and SSL1 genes perturb

genome stability by stimulating recombination between

short repeats in Saccharomyces cerevisiae. Genetics 1998;

150: 963±976.

139. MalagoÂn F, Aguilera A. Differential intrachromosomal

hyper-recombination phenotype of spt4 and spt6 mutants

of S. cerevisiae. Curr Genet 1996; 30: 101±106.

140. Malone RE, Montelone BA, Edwards C, Caney K,

Hoekstra MF. A reexamination of the role of the RAD52

gene in spontaneous recombination. Curr Genet 1988; 14:

211±223.

141. Malone RE, Ward T, Lin S, Waring J. The RAD50 gene, a

member of the double strand break repair epistasis group, is

not required for spontaneous mitotic recombination in

yeast. Curr Genet 1990; 18: 111±116.

142. Marinus MG, Konrad EB. Hyper-recombination in dam

mutants of Escherichia coli K-12. Mol Gen Genet 1976; 149:

273±277.

143. Mayer VW, Goin CJ. Semidominance of rad18-2 for several

phenotypic characters in Saccharomyces cerevisiae. Genetics

1984; 106: 577±589.

144. McDonald JP, Rothstein R. Unrepaired heteroduplex DNA

in Saccharomyces cerevisiae is decreased in RAD1 RAD52-

independent recombination. Genetics 1994; 137: 393±405.

145. Merrill BJ, Holm C. The RAD52 recombinational repair

pathway is essential in pol30 (PCNA) mutants that

accumulate small single-stranded DNA fragments during

DNA synthesis. Genetics 1998; 148: 611±624.

146. Merrill BJ, Holm C. A requirement for recombinational

repair in Saccharomyces cerevisiae is caused by DNA

replication defects of mec1 mutants. Genetics 1999; 153:

595±605.

147. Mezard C, Nicolas A. Homologous, homeologous, and

illegitimate repair of double-strand breaks during transfor-

mation of a wild-type strain and a rad52 mutant strain of

Saccharomyces cerevisiae. Mol Cell Biol 1994; 14:

1278±1292.

148. Milne GT, Jin S, Shannon KB, Weaver DT. Mutations in

two Ku homologs de®ne a DNA end-joining repair path-

way in Saccharomyces cerevisiae. Mol Cell Biol 1996; 16:

4189±4198.

149. Montelone BA, Hoekstra MF, Malone RE. Spontaneous

mitotic recombination in yeast: the hyper-recombinational

rem1 mutations are alleles of the RAD3 gene. Genetics 1988;

119: 289±301.

150. Montelone BA, Koelliker KJ. Interactions among muta-

tions affecting spontaneous mutation, mitotic recombina-

tion and DNA repair in yeast. Curr Genet 1995; 27:

102±109.

151. Mosig G. Recombination and recombination-dependent

DNA replication in bacteriophage T4. Ann Rev Genet

1998; 32: 379±413.

152. Munz P, Amstutz H, Kohli J, Leupold U. Recombination

between dispersed serine tRNA genes in Schizosaccharo-

myces pombe. Nature 1982; 300: 225±231.

153. Myers LC, Gustafsson CM, Bushnell DA, Lui M,

Erdjument-Bromage H, Tempst P, Kornberg RD. The

Med proteins of yeast and their function through the

RNA polymerase II carboxy-terminal domain. Genes Dev

1998; 12: 45±54.

154. Neitz M, Carbon J. Characterization of a centromere-linked

recombination hot spot in Saccharomyces cerevisiae. Mol

Cell Biol 1987; 7: 3871±3879.

155. Nevo-Caspi Y, Kupiec M. Transcriptional induction of Ty

recombination in yeast. Proc Natl Acad Sci U S A 1994; 91:

12711±12715.

156. Nevo-Caspi Y, Kupiec M. Induction of Ty recombination

in yeast by cDNA and transcription: role of the RAD1 and

RAD52 genes. Genetics 1996; 144: 947±955.

157. Nickoloff JA, Singer JD, Hoekstra MF, Heffron F.

Double-strand breaks stimulate alternative mechanisms of

recombination repair. J Mol Biol 1989; 207: 527±541.

158. Nickoloff JA. Transcription enhances intrachromosomal

homologous recombination in mammalian cells. Mol Cell

Biol 1992; 12: 5311±5318.

159. Nickoloff JA, Hoekstra MF. Double-strand break repair

and recombination in Escherichia coli. In DNA Damage and

Repair, vol I: DNA Repair in Prokaryotes and Lower

Eukaryotes, Nickoloff JA, Hoekstra MF, (eds). Humana

Press: Totowa, NJ; 1998; 335±362.

160. Noskov VN, Araki H, Sugino A. The RFC2 gene, encoding

the third-largest subunit of the replication factor C

Mitotic recombination in yeast: control of incidence 751

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

complex, is required for an S-phase checkpoint in Saccharo-

myces cerevisiae. Mol Cell Biol 1998; 18: 4914±4923.

161. Oltz EM, Alt FW, Lin WC, Chen J, Taccioli G, Desiderio

S, Rathbun G. A V(D)J recombinase-inducible B-cell line:

role of transcriptional enhancer elements in directing

V(D)J recombination. Mol Cell Biol 1993; 13: 6223±6230.

162. Orr-Weaver TL, Nicolas A, Szostak JW. Gene conversion

adjacent to regions of double-strand break repair. Mol Cell

Biol 1988; 8: 5292±5298.

163. Orr-Weaver TL, Szostak JW, Rothstein RJ. Yeast trans-

formation: a model system for the study of recombination.

Proc Natl Acad Sci U S A 1981; 78: 6354±6358.

164. Osman F, Subramani S. Double-strand break-induced

recombination in eukaryotes. Prog Nucleic Acid Res Mol

Biol 1998; 58: 263±299.

165. Ozenbergen BA, Roeder GS. A unique pathway of double-

strand break repair operates in tandemly repeated genes.

Mol Cell Biol 1991; 11: 1222±1231.

166. Palladino F, Klein HL. Analysis of mitotic and meiotic

defects in Saccharomyces cerevisiae SRS2 DNA helicase

mutants. Genetics 1992; 132: 23±37.

167. PaÃques F, Haber JE. Multiple pathways of recombination

induced by double-strand breaks in Saccharomyces cerevi-

siae. Microbiol Mol Biol Rev 1999; 63: 349±404.

168. Peterson CL, Tamkun JW. The SWI±SNF complex: a

chromatin remodeling machine?. Trends Biochem Sci 1995;

20: 143±146.

169. Petes TD, Malone RE, Symington LS. Recombination in

yeast. In The Molecular and Cellular Biology of the

Yeast Saccharomyces: Genome Dynamics, vol I: Protein

Synthesis and Energetics, Broach JR, Pringle JR, Jones EW

(eds). Cold Spring Harbor Laboratory Press: New York;

1991; 407±521.

170. Piruat JI, Aguilera A. Mutations in the yeast SRB2 general

transcription factor suppress hpr1-induced recombination

and show defects in DNA repair. Genetics 1996; 143:

1533±1542.

171. Piruat JI, ChaÂvez S, Aguilera A. The yeast HRS1 gene is

involved in positive and negative regulation of transcription

and shows genetic characteristics similar to SIN4 and

GAL11. Genetics 1997; 147: 1585±1594.

172. Piruat JI, Aguilera A. A novel yeast gene, THO2, is

involved in RNA pol II transcription and provides new

evidence for transcriptional elongation-associated recombi-

nation. EMBO J 1998; 17: 4859±4872.

173. Plessis A, Perrin A, Haber JE, Dujon B. Site-speci®c

recombination determined by I-SceI, a mitochondrial

group I intron-encoded endonuclease expressed in the

yeast nucleus. Genetics 1992; 130: 451±460.

174. Prado F, Aguilera A. Role of reciprocal exchange, one-

ended invasion crossover and single-strand annealing on

inverted and direct repeat recombination in yeast: different

requirements for the RAD1, RAD10, and RAD52 genes.

Genetics 1995; 139: 109±123.

175. Prado F, GonzaÂlez-Barrera S, Aguilera A. RAD52-depen-

dent and -independent homologous recombination initiated

by Flp recombinase at a single FRT site ¯anked by direct

repeats. Mol Gen Genet 2000; 263: 73±80.

176. Prado F, Piruat JI, Aguilera A. Recombination between

DNA repeats in yeast hpr1D cells is linked to transcription

elongation. EMBO J 1997; 16: 2826±2835.

177. Prakash S, Prakash L. Increased spontaneous mitotic

segregation in MMS-sensitive mutants of Saccharomyces

cerevisiae. Genetics 1977; 87: 229±236.

178. Richard GF, Dujon B, Haber JE. Double-strand break

repair can lead to high frequencies of deletions within short

CAG/CTG trinucleotide repeats. Mol Gen Genet 1999; 261:

871±882.

179. Roeder GS. Meiotic chromosomes: it takes two to tango.

Genes Dev 1997; 11: 2600±2621.

180. Rong L, Palladino F, Aguilera A, Klein HL. The hyper-

gene conversion hpr5-1 mutation of Saccharomyces cerevi-

siae is an allele of the SRS2/RADH gene. Genetics 1991;

127: 75±85.

181. Rothstein R, Helms C, Rosenberg N. Concerted deletions

and inversions are caused by mitotic recombination between

delta sequences in Saccharomyces cerevisiae. Mol Cell Biol

1987; 7: 1198±1207.

182. Rudin N, Haber JE. Ef®cient repair of HO-induced

chromosomal breaks in Saccharomyces cerevisiae by recom-

bination between ¯anking homologous sequences. Mol Cell

Biol 1988; 8: 3918±3928.

183. Rudin N, Sugarman E, Haber JE. Genetic and physical

analysis of double-strand break repair and recombination in

Saccharomyces cerevisiae. Genetics 1989; 122: 519±534.

184. Ruskin B, Fink GR. Mutations in POL1 increase the

mitotic instability of tandem inverted repeats in Saccharo-

myces cerevisiae. Genetics 1993; 133: 43±56.

185. Sadoff BU, Heath-Pagliuso S, Castano IB, Zhu Y, Kieff

FS, Christman MF. Isolation of mutants of Saccharomyces

cerevisiae requiring DNA topoisomerase I. Genetics 1995;

141: 465±479.

186. Santos-Rosa H, Aguilera A. Increase in incidence of

chromosome instability and non-conservative recombina-

tion between repeats in Saccharomyces cerevisiae hpr1D

strains. Mol Gen Genet 1994; 245: 224±236.

187. Santos-Rosa H, Aguilera A. Isolation and genetic analysis

of extragenic suppressors of the hyper-deletion phenotype

of the Saccharomyces cerevisiae hpr1D mutation. Genetics

1995; 139: 57±66.

188. Santos-Rosa H, Clever B, Heyer WD, Aguilera A. The

yeast HRS1 gene encodes a polyglutamine-rich nuclear

protein required for spontaneous and hpr1-induced dele-

tions between direct repeats. Genetics 1996; 142: 705±716.

189. Schultz MC, Brill SJ, Ju Q, Sternglanz R, Reeder RH.

Topoisomerases and yeast rRNA transcription: negative

supercoiling stimulates initiation and topoisomerase activity

is required for elongation. Genes Dev 1992; 6: 1332±1341.

190. Seigneur M, Bidnenko V, Ehrlich SD, Michel B. RuvAB

acts at arrested replication forks. Cell 1998; 95: 419±430.

191. Selva EM, New L, Crouse GF, Lahue RS. Mismatch

correction acts as a barrier to homeologous recombination

in Saccharomyces cerevisiae. Genetics 1995; 139: 1175±1188.

192. Seoighe C, Wolfe KH. Extent of genomic rearrangement

after genome duplication in yeast. Proc Natl Acad Sci

U S A 1998; 95: 4447±4452.

193. Shinohara A, Ogawa T. Homologous recombination and

the roles of double-strand breaks. Trends Biochem Sci 1995;

20: 387±391.

194. Siede W, Friedl AA, Dianova I, Eckardt-Schupp F,

Friedberg EC. The Saccharomyces cerevisiae Ku autoanti-

752 A. Aguilera, S. ChaÂvez and F. MalagoÂn

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

gen homologue affects radiosensitivity only in the absence

of homologous recombination. Genetics 1996; 142: 91±102.

195. Shi X, Chang M, Wolf AJ, Chang CH, Frazer-Abel AA,

Wade PA, Burton ZF, Jaehning JA. Cdc73p and Paf1p are

found in a novel RNA polymerase II-containing complex

distinct from the Srbp-containing holoenzyme. Mol Cell

Biol 1997; 17: 1160±1169.

196. Sipickzi M, Grossenbacher-Grunder AM, Bodi Z. Recom-

bination and mating-type switching in a ligase-defective

mutant of Schizosaccharomyces pombe. Mol Gen Genet

1999; 220: 307±313.

197. Smith GR. DNA double-strand break repair and recombi-

nation in Escherichia coli. In DNA Damage and Repair, vol.

I: DNA Repair in Prokaryotes and Lower Eukaryotes,

Nickoloff JA, Hoekstra MF (eds). Humana Press: Totowa,

NJ; 1998; 135±162.

198. Smith J, Rothstein R. A mutation in the gene encoding the

Saccharomyces cerevisiae single-stranded DNA-binding

protein Rfa1 stimulates a RAD52-independent pathway for

direct-repeat recombination. Mol Cell Biol 1995; 15:

1632±1641.

199. Smith J, Rothstein R. An allele of RFA1 suppresses

RAD52-dependent double-strand break repair in Saccharo-

myces cerevisiae. Genetics 1999; 151: 447±458.

200. Smith KN, Nicolas A. Recombination at work for meiosis.

Curr Opin Genet Dev 1998; 8: 200±211.

201. Sommers CH, Miller EJ, Dujon B, Prakash S, Prakash L.

Conditional lethality of null mutations in RTH1

that encodes the yeast counterpart of a mammalian

5k- to 3k-exonuclease required for lagging strand DNA

synthesis in reconstituted systems. J Biol Chem 1995; 270:

4193±4196.

202. Stahl F. Meiotic recombination in yeast: coronation of the

double-strand-break repair model. Cell 1996; 87: 965±968.

203. Stanhope-Baker P, Hudson KM, Shaffer AL, Constanti-

nescu A, Schlissel MS. Cell type-speci®c chromatin struc-

ture determines the targeting of V(D)J recombinase activity

in vitro. Cell 1996; 85: 887±897.

204. Stewart SE, Roeder GS. Transcription by RNA polymerase

I stimulates mitotic recombination in Saccharomyces cere-

visiae. Mol Cell Biol 1989; 9: 3464±3472.

205. Strathern JN, Klar AJ, Hicks JB, Abraham JA, Ivy JM,

Nasmyth KA, McGill C. Homothallic switching of yeast

mating type cassettes is initiated by a double-stranded cut in

the MAT locus. Cell 1982; 31: 183±192.

206. Strathern JN, Weinstock KG, Higgins DR, McGill CB. A

novel recombinator in yeast based on gene II protein from

bacteriophage f1. Genetics 1991; 127: 61±73.

207. Subramani S, Seaton BL. Homologous recombination in

mitotically dividing mammalian cells. In Genetic Recombi-

nation, Kucherlapatti R, Smith GR (eds). ASM: Washing-

ton, DC; 1988; 549±574.

208. Sugawara N, Haber JE. Characterization of double-strand

break-induced recombination: homology requirements and

single-stranded DNA formation. Mol Cell Biol 1992; 12:

563±575.

209. Sugawara N, Ivanov EL, Fishman-Lobell J, Ray BL, Wu

X, Haber JE. DNA structure-dependent requirements for

yeast RAD genes in gene conversion. Nature 1995; 373:

84±86.

210. Sugimoto K, Sakamoto Y, Takahashi O, Matsumoto K.

HYS2, an essential gene required for DNA replication in

Saccharomyces cerevisiae. Necleic Acids Res 1995; 23:

3493±3500.

211. Sun H, Treco D, Schultes NP, Szostak JW. Double-strand

breaks at an initiation site for meiotic gene conversion.

Nature 1989; 338: 87±90.

212. Sun H, Treco D, Szostak JW. Extensive 3k-overhanging,

single-stranded DNA associated with the meiosis-speci®c

double-strand breaks at the ARG4 recombination initiation

site. Cell 1991; 64: 1155±1161.

213. Swanson RL, Morey NJ, Doetsch PW, Jinks-Robertson S.

Overlapping speci®cities of base excision repair, nucleotide

excision repair, recombination, and translesion synthesis

pathways for DNA base damage in Saccharomyces cerevi-

siae. Mol Cell Biol 1999; 19: 2929±2935.

214. Symington LS. Homologous recombination is required for

the viability of rad27 mutants. Nucleic Acids Res 1998; 26:

5589±5595.

215. Szostak JW, Orr-Weaver TL, Rothstein RJ, Stahl FW. The

double-strand-break repair model for recombination. Cell

1983; 33: 25±35.

216. Thaler DS, Stahl MM, Stahl FW. Tests of the double-

strand-break repair model for Red-mediated recombination

of phage l and plasmid ldv. Genetics 1987; 116: 501±511.

217. Thomas BJ, Rothstein R. The genetic control of direct-

repeat recombination in Saccharomyces: the effect of rad52

and rad1 on mitotic recombination at GAL10, a transcrip-

tionally regulated gene. Genetics 1989; 123: 725±738.

218. Thomas BJ, Rothstein R. Elevated recombination rates in

transcriptionally active DNA. Cell 1989; 56: 619±630.

219. Thompson L. Mammalian cell mutations affecting recom-

bination. In Genetic Recombination, Kucherlapatti R, Smith

GR (eds). ASM: Washington, DC; 1988; 597±620.

220. Thuriaux P. Direct selection of mutants in¯uencing gene

conversion in the yeast Schizosaccharomyces pombe. Mol

Gen Genet 1985; 199: 365±371.

221. Thyagarajan B, Johnson BL, Campbell C. The effect of

target site transcription on gene targeting in human cells in

vitro. Nucleic Acids Res 1995; 23: 2784±2790.

222. Tishkoff DX, Filosi N, Gaida GM, Kolodner RD. A novel

mutation avoidance mechanism dependent on S. cerevisiae

RAD27 is distinct from DNA mismatch repair. Cell 1997;

88: 253±263.

223. Tsubouchi H, Ogawa H. A novel mre11 mutation impairs

processing of double-strand breaks of DNA during both

mitosis and meiosis. Mol Cell Biol 1998; 18: 260±268.

224. Tsukamoto Y, Ikeda H. Double-strand break repair

mediated by DNA end-joining. Genes Cells 1998; 3:

135±144.

225. Uemura H, Pandit S, Jigami Y, Sternglanz R. Mutations in

GCR3, a gene involved in the expression of glycolytic genes

in Saccharomyces cerevisiae, suppress the temperature-

sensitive growth of hpr1 mutants. Genetics 1996; 142:

1095±1103.

226. Vallen EA, Cross FR. Mutations in RAD27 de®ne a

potential link between G1 cyclins and DNA replication.

Mol Cell Biol 1995; 15: 4291±4302.

227. Vilette D, Ehrlich SD, Michel B. Transcription-induced

deletions in Escherichia coli plasmids. Mol Microbiol 1995;

17: 493±504.

228. Voelkel-Meiman K, Keil RL, Roeder GS. Recombination-

Mitotic recombination in yeast: control of incidence 753

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.

stimulating sequences in yeast ribosomal DNA correspond

to sequences regulating transcription by RNA polymerase I.

Cell 1987; 48: 1071±1079.

229. Wallis JW, Chrebet G, Brodsky G, Rolfe M, Rothstein R.

A hyper-recombination mutation in S. cerevisiae identi®es a

novel eukaryotic topoisomerase. Cell 1989; 58: 409±419.

230. Wang X, Watt PM, Borts RH, Louis EJ, Hickson ID. The

topoisomerase II-associated protein, Pat1p, is required for

maintenance of rDNA locus stability in Saccharomyces

cerevisiae. Mol Gen Genet 1999; 261: 831±840.

231. Watt PM, Louis EJ, Borts RH, Hickson ID. Sgs1: a

eukaryotic homolog of E. coli RecQ that interacts with

topoisomerase II in vivo and is required for faithful

chromosome segregation. Cell 1995; 81: 253±260.

232. Watt PM, Hickson ID, Borts RH, Louis EJ. SGS1, a

homologue of the bloom's and Werner's syndrome genes, is

required for maintenance of genome stability in Saccharo-

myces cerevisiae. Genetics 1996; 144: 935±945.

233. Weaver DT. What to do to an end: DNA double-strand-

break repair. Trends Genet 1995; 11: 388±392.

234. Wierdl M, Greene CN, Datta A, Jinks-Robertson S, Petes

TD. Destabilization of simple repetitive DNA sequences by

transcription in yeast. Genetics 1996; 143: 713±721.

235. Williamson MS, Game JC, Fogel S. Meiotic gene conver-

sion mutants in Saccharomyces cerevisiae. I. Isolation and

characterization of pms1-1 and pms1-2. Genetics 1985; 110:

609±646.

236. Winston F, Carlson M. Yeast SNF/SWI transcriptional

activators and the SPT/SIN chromatin connection. Trends

Genet 1992; 8: 387±391.

237. Wolfe KH, Shields DC. Molecular evidence for an ancient

duplication of the entire yeast genome. Nature 1997; 387:

708±713.

238. Wu C, Weiss K, Yang C, Harris MA, Tye BK, Newlon CS,

Simpson RT, Haber JE. Mcm1 regulates donor preference

controlled by the recombination enhancer in Saccharomyces

mating-type switching. Genes Dev 1998; 12: 1726±1737.

239. Yan H, Gibson S, Tye BK. Mcm2 and Mcm3, two proteins

important for ARS activity, are related in structure and

function. Genes Dev 1991; 5: 944±957.

240. Zgaga Z, Chanet R, Radman M, Fabre F. Mismatch-

stimulated plasmid integration in yeast. Curr Genet 1991;

19: 329±332.

241. Zhu Y, Peterson CL, Christman MF. HPR1 encodes a

global positive regulator of transcription in Saccharomyces

cerevisiae. Mol Cell Biol 1995; 15: 1698±1708.

242. Zieg J, Maples VF, Kushner SR. Recombinant levels of

Escherichia coli K-12 mutants de®cient in various replica-

tion, recombination, or repair genes. J Bacteriol 1978; 134:

958±966.

243. Zou H, Rothstein R. Holliday junctions accumulate in

replication mutants via a RecA homolog-independent

mechanism. Cell 1997; 90: 87±96.

754 A. Aguilera, S. ChaÂvez and F. MalagoÂn

Copyright # 2000 John Wiley & Sons, Ltd. Yeast 2000; 16: 731±754.