The N-Terminal DNA-Binding Domain of Rad52 Promotes RAD51-Independent Recombination in Saccharomyces cerevisiae (original) (raw)

Journal Article

,

Department of Biology

, Graduate School of Science, Osaka University, Toyonaka, Osaka 560-0043, Japan

Search for other works by this author on:

,

Department of Biology

, Graduate School of Science, Osaka University, Toyonaka, Osaka 560-0043, Japan

Search for other works by this author on:

,

Department of Biology

, Graduate School of Science, Osaka University, Toyonaka, Osaka 560-0043, Japan

Search for other works by this author on:

,

Department of Biology

, Graduate School of Science, Osaka University, Toyonaka, Osaka 560-0043, Japan

Search for other works by this author on:

Department of Biology

, Graduate School of Science, Osaka University, Toyonaka, Osaka 560-0043, Japan

Japanese Science and Technology (JST)

, Toyonaka, Osaka 560-0043, Japan

Corresponding author: Department of Biology, Graduate School of Science, Osaka University, 1-1 Machikaneyama, Toyonaka, Osaka 560-0043, Japan. E-mail: ashino@bio.sci.osaka-u.ac.jp

Search for other works by this author on:

1

Present address: Research Institute for Radiation Biology and Medicine, Hiroshima University, Hiroshima 734-8553, Japan.

Author Notes

Published:

01 December 2003

Cite

Mariko Tsukamoto, Kentaro Yamashita, Toshiko Miyazaki, Miki Shinohara, Akira Shinohara, The N-Terminal DNA-Binding Domain of Rad52 Promotes _RAD51_-Independent Recombination in Saccharomyces cerevisiae, Genetics, Volume 165, Issue 4, 1 December 2003, Pages 1703–1715, https://doi.org/10.1093/genetics/165.4.1703
Close

Navbar Search Filter Mobile Enter search term Search

Abstract

In Saccharomyces cerevisiae, the Rad52 protein plays a role in both _RAD51_-dependent and _RAD51_-independent recombination pathways. We characterized a rad52 mutant, rad52-329, which lacks the C-terminal Rad51-interacting domain, and studied its role in _RAD51_-independent recombination. The rad52-329 mutant is completely defective in mating-type switching, but partially proficient in recombination between inverted repeats. We also analyzed the effect of the rad52-329 mutant on telomere recombination. Yeast cells lacking telomerase maintain telomere length by recombination. The rad52-329 mutant is deficient in _RAD51_-dependent telomere recombination, but is proficient in _RAD51_-independent telomere recombination. In addition, we examined the roles of other recombination genes in the telomere recombination. The _RAD51_-independent recombination in the rad52-329 mutant is promoted by a paralogue of Rad52, Rad59. All components of the Rad50-Mre11-Xrs2 complex are also important, but not essential, for _RAD51_-independent telomere recombination. Interestingly, RAD51 inhibits the _RAD51_-independent, _RAD52_-dependent telomere recombination. These findings indicate that Rad52 itself, and more precisely its N-terminal DNA-binding domain, promote an essential reaction in recombination in the absence of RAD51.

HOMOLOGOUS recombination is important in genome stability. Dysfunction results in genome instability, which is often associated with the onset of cancer. Recombination is required for the repair of double-strand breaks (DSBs) as well as for the segregation of homologous chromosomes during meiosis. Recently it has been shown that recombination is also involved in maintaining telomere length.

In Saccharomyces cerevisiae, genes in the RAD52 epistasis group (RAD50, -51, -52, -54, -55, -57, -59, TID1/RDH54, MRE11, and XRS2) are involved in recombination (Pâques and Haber 1999; Symington 2002). So far, two pathways have been identified: RAD51 dependent and RAD51 independent. RAD51 encodes a bacterial RecA homolog, which works during the search for homology and the exchange of strands (Shinohara et al. 1992; Sung 1994). The _RAD51_-dependent pathway is a major conservative repair pathway for DSBs (Pâques and Haber 1999), but in some circumstances, the nonconservative _RAD51_-independent pathway operates. The _RAD51_-dependent pathway requires most of the members of the RAD52 epistasis group, but the _RAD51_-independent pathway requires only a few members.

In the absence of RAD51, a single DSB on one chromosome in diploid cells is repaired by _RAD51_-independent recombination, termed break-induced replication (BIR; Malkova et al. 1996; Kraus et al. 2001). A centromere-proximal DSB can be repaired by invasion of the broken end into a homologous region on another chromosome (Figure 1). A replication fork can be established, and the entire chromosome arm can be copied, resulting in the duplication of a large portion of the chromosome arm. It depends totally on RAD52 (Malkova et al. 1996) and also involves the Mre11-Rad50-Xrs2 (MRX) complex, Rad59 (a paralogue of Rad52), and Tid1/Rdh54 (Signon et al. 2001). In addition, BIR initiated by a DSB at the MAT locus requires a _cis_-acting sequence (Malkova et al. 2001). Although BIR was originally described for the rad51 mutant, it also occurs in wild-type (WT) cells (Morrow et al. 1997; Kraus et al. 2001). RAD59 was originally identified as a gene necessary for recombination between inverted repeats in cells lacking RAD51 (Bai and Symington 1996), which is thought to occur through a BIR event coupled with single-strand annealing (Kang and Symington 2000).

Recombination is necessary for telomere elongation in cells lacking telomerase (Kass-Eisler and Greider 2000; McEachern et al. 2000) and for the contraction of telomeres in wild-type cells (Bucholc et al. 2001). In S. cerevisiae, telomeres consist of ∼350 bp of TG1–3 repeats and subtelomeric Y′ sequences (long and short Y′) next to the repeats. Telomerase, a reverse transcriptase composed of a catalytic protein, Est2, and other proteins, as well as an RNA component encoded by the TLC1 gene, performs de novo synthesis of the repeats (Singer and Gottschling 1994). If any components of the telomerase are missing, telomeres are gradually shortened and most cells die. However, survivors arise and can elongate their telomeres by recombination (Lundblad and Blackburn 1993; Singer and Gottschling 1994). Two types of survivors are generated (Lundblad and Blackburn 1993; Le et al. 1999; Teng and Zakian 1999). Type I survivors often amplify the subtelomeric Y′ sequence, whereas type II survivors are generated by amplification of TG1–3 repeats. Type I survivors use _RAD51_-dependent recombination, while type II survivors use _RAD51_-independent recombination (Le et al. 1999; Teng and Zakian 1999). In this article, these are termed type I and type II telomere recombination, respectively. The mechanism of telomere recombination is proposed to be similar to that of BIR (Kass-Eisler and Greider 2000; Kraus et al. 2001).

—A model for BIR in the absence of RAD51. After the resection of a DSB end, ssDNA is used for homology search and strand exchange, resulting in the formation of a D-loop. The 3′ end of the invaded strand becomes a substrate for leading-strand synthesis. How a subsequent lagging strand is synthesized is unknown (see Kraus  et al. 2001 for more detail). A newly synthesized strand is shown as a gray line.

Figure 1.

—A model for BIR in the absence of RAD51. After the resection of a DSB end, ssDNA is used for homology search and strand exchange, resulting in the formation of a D-loop. The 3′ end of the invaded strand becomes a substrate for leading-strand synthesis. How a subsequent lagging strand is synthesized is unknown (see Kraus et al. 2001 for more detail). A newly synthesized strand is shown as a gray line.

The RAD52 gene is required for most recombination in yeast (Symington 2002). Rad52 consists of at least two domains: the N-terminal DNA-binding domain and the C-terminal Rad51-interacting domain (Shinohara et al. 1992; Milne and Weaver 1993; Shinohara and Ogawa 1998). The N-terminal domain is conserved among species while the C-terminal domain is not (Muris et al. 1994). The Rad59 protein shares homology with the N terminus of Rad52 and lacks the C-terminal domain (Bai and Symington 1996). In vivo, Rad52 and Rad59 form a complex (Davis and Symington 2001). However, genetic analysis shows that Rad59 plays a less significant role than Rad52 in recombination. Purified yeast and human Rad52 proteins bind to both single-stranded DNA (ssDNA) and double-stranded DNA (dsDNA; Mortensen et al. 1996; Shinohara et al. 1998; Parsons et al. 2000) and form ring-like structures by themselves and on the DNAs (Shinohara et al. 1998; Van Dyck et al. 1998; Stasiak et al. 2000; Kagawa et al. 2002; Singleton et al. 2002). The N-terminal domain of Rad52 is responsible for the DNA binding and ring formation. Rad52 stimulates Rad51-mediated recombination (Sung 1997; New et al. 1998; Shinohara and Ogawa 1998) and annealing of complementary ssDNAs in vitro (Mortensen et al. 1996; Shinohara et al. 1998; Sugiyama et al. 1998).

Here, we show that the N-terminal DNA-binding domain of Rad52 facilitates _RAD51_-independent recombination at telomeres. Genetic analyses using various mutants support the view that the telomere recombination is mechanistically related to BIR. The _RAD51_-independent recombination pathway is stimulated by the elimination of RAD51.

MATERIALS AND METHODS

Strains and plasmids: All yeast strains and their genotypes are shown in Table 1. Strains used for telomere analysis (except JHUY564; kindly provided by Carol Greider) were derivatives of S288C. Recombination between inverted repeats was measured in the background of W303. The rad50, rad51, rad52, rad59, mre11, and xrs2 deletion strains were constructed by one-step gene replacement using a fragment containing rad50::hisG-URA3-hisG (Alani et al. 1989), rad51::hisG-URA3-hisG (Shinohara et al. 1992), rad52::hisG-URA3-hisG (a gift from T. Ogawa), rad59::hisG-URA3-hisG, mre11::hisG-URA3-hisG (Johzuka and Ogawa 1995), and xrs2::KanMX6, respectively. rad59::hisG-URA3-hisG was constructed by inserting a _Bam_HI-_Bgl_II fragment containing hisG-URA3-hisG from pNKY51 (Alani et al. 1987) into the _Bgl_II-_Eco_47III site of the RAD59 gene. xrs2::KanMX6 was constructed as follows: A fragment containing the XRS2 promoter was amplified by PCR using primers (X2-1, 5′-CCGCTCGAGA GAGGACACCAAAG, and X2-2, 5′-GTACTACCCACATATGT TTATAGTTATC) and cloned into the _Xho_I-_Eco_RV site of pBluescript II SK+ (pMS235). A _Pst_I-_Bam_HI fragment of YCp50-XRS2 (a gift from Jim Haber) was inserted into the _Pst_I-_Bam_HI site of pMS235, resulting in the plasmid pMS236. The plasmid pMS237 was constructed by cloning a _Nde_I-_Spe_I fragment containing KanMX6 from pMJ476 (Wach et al. 1994) into the _Nde_I-_Spe_I site of pMS236. The rad52-329 allele was constructed by replacing a _Bam_HI-_Sal_I fragment of the RAD52 gene by a PCR fragment amplified using the following oligonucleotides: oligo(A) (5′-CGggatccCTGAAACGCTTCCTGGCCG), which contains a termination codon (underlined) next to the _Bam_HI site followed by the sequence downstream from the authentic stop codon of RAD52, and oligo(B) (5′-GGgtcgacGT CCAAGAAATACATTGG), which contains the _Sal_I site down-stream of RAD52. A fragment containing the rad52-329 allele is inserted into the _Sma_I-_Eco_RI site of YIplac195 (Gietz and Sugino 1988), resulting in the plasmid pMT001. Parental strains were transformed with pMT001 digested with _Bst_EII and selected on a synthetic dextrose (SD) medium plate lacking uracil. Ura+ transformants were picked and then selected on a plate containing 5-fluoroorotic acid for Ura–. PCR and Southern blotting confirmed genotypes of the mutants.

The rad52-329 and rad51 deletion alleles were introduced into YDB057 or yAR071 as described above. The rad52 deletion was introduced into yAR71 by adaptamer-mediated PCR using the KlURA3 gene (Reid et al. 2002).

Strain Genotype Reference
YMT013 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 This study
YMT026 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD52/rad52-329 This study
YMT050 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD51/rad51::hisG This study
YMT038 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD52/rad52:: hisG-URA3-hisG This study
YMT178 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/his3 TLC1/tlc1::LEU2 RAD51/rad51::hisGRAD52/rad52::hisG-URA3-hisG This study
YMT075 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD51/rad51::hisG RAD52/rad52-329 This study
YMT106 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG This study
YMT101 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG RAD52/rad52-329 This study
YMT116 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG RAD51/rad51::hisG This study
YMT111 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG RAD51/rad51::hisG RAD52/rad52-329 This study
YMT335 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD50/rad50::hisG-URA3-hisG RAD52/rad52-329 This study
YMT381 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 MRE11/mre11::hisG-URA3-hisG This study
YMT385 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 XRS2/xrs2::KanMX6 This study
JHUY563 MATadhml::ADE1/dhml::ADE1 dhmr::ADE1/dhmr::ADE1 ade1/ade1 leu2/leu2 lys5/lys5 ura3/ura3 RAD51/rad51::LEU2 tlc1:URA3/TLC1 rad59::kanMX4/RAD59 Chen et al. (2001)
YDB057 ho HML::HMRa-_(URA3) MAT_α matdist::tetO(112)-(NAT1) HMRa ade3::GAL10-HO ade1-110 trp1 leu2::TetR-GFP-(LEU2) T. Miyazaki and A. Shinohara (unpublished data)
YMT546 A YDB057 derivative with rad52-329 This study
yAR71 MATa trp1-1 leu2-3,112 his3-11,15 ura3-1 ade2-ura3::hisG can1-100 spo13:hisG his3::ade2-5_′Δ_-TRP1-ade2-n Rattray and Symington (1994)
YMT353 A yAR71 derivative with rad52-329 This study
YMT362 A yAR71 derivative with rad52::KlURA3 This study
YMT367 A yAR71 derivative with rad51::hisG-URA3-hisG This study
Strain Genotype Reference
YMT013 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 This study
YMT026 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD52/rad52-329 This study
YMT050 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD51/rad51::hisG This study
YMT038 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD52/rad52:: hisG-URA3-hisG This study
YMT178 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/his3 TLC1/tlc1::LEU2 RAD51/rad51::hisGRAD52/rad52::hisG-URA3-hisG This study
YMT075 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD51/rad51::hisG RAD52/rad52-329 This study
YMT106 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG This study
YMT101 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG RAD52/rad52-329 This study
YMT116 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG RAD51/rad51::hisG This study
YMT111 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG RAD51/rad51::hisG RAD52/rad52-329 This study
YMT335 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD50/rad50::hisG-URA3-hisG RAD52/rad52-329 This study
YMT381 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 MRE11/mre11::hisG-URA3-hisG This study
YMT385 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 XRS2/xrs2::KanMX6 This study
JHUY563 MATadhml::ADE1/dhml::ADE1 dhmr::ADE1/dhmr::ADE1 ade1/ade1 leu2/leu2 lys5/lys5 ura3/ura3 RAD51/rad51::LEU2 tlc1:URA3/TLC1 rad59::kanMX4/RAD59 Chen et al. (2001)
YDB057 ho HML::HMRa-_(URA3) MAT_α matdist::tetO(112)-(NAT1) HMRa ade3::GAL10-HO ade1-110 trp1 leu2::TetR-GFP-(LEU2) T. Miyazaki and A. Shinohara (unpublished data)
YMT546 A YDB057 derivative with rad52-329 This study
yAR71 MATa trp1-1 leu2-3,112 his3-11,15 ura3-1 ade2-ura3::hisG can1-100 spo13:hisG his3::ade2-5_′Δ_-TRP1-ade2-n Rattray and Symington (1994)
YMT353 A yAR71 derivative with rad52-329 This study
YMT362 A yAR71 derivative with rad52::KlURA3 This study
YMT367 A yAR71 derivative with rad51::hisG-URA3-hisG This study
Strain Genotype Reference
YMT013 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 This study
YMT026 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD52/rad52-329 This study
YMT050 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD51/rad51::hisG This study
YMT038 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD52/rad52:: hisG-URA3-hisG This study
YMT178 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/his3 TLC1/tlc1::LEU2 RAD51/rad51::hisGRAD52/rad52::hisG-URA3-hisG This study
YMT075 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD51/rad51::hisG RAD52/rad52-329 This study
YMT106 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG This study
YMT101 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG RAD52/rad52-329 This study
YMT116 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG RAD51/rad51::hisG This study
YMT111 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG RAD51/rad51::hisG RAD52/rad52-329 This study
YMT335 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD50/rad50::hisG-URA3-hisG RAD52/rad52-329 This study
YMT381 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 MRE11/mre11::hisG-URA3-hisG This study
YMT385 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 XRS2/xrs2::KanMX6 This study
JHUY563 MATadhml::ADE1/dhml::ADE1 dhmr::ADE1/dhmr::ADE1 ade1/ade1 leu2/leu2 lys5/lys5 ura3/ura3 RAD51/rad51::LEU2 tlc1:URA3/TLC1 rad59::kanMX4/RAD59 Chen et al. (2001)
YDB057 ho HML::HMRa-_(URA3) MAT_α matdist::tetO(112)-(NAT1) HMRa ade3::GAL10-HO ade1-110 trp1 leu2::TetR-GFP-(LEU2) T. Miyazaki and A. Shinohara (unpublished data)
YMT546 A YDB057 derivative with rad52-329 This study
yAR71 MATa trp1-1 leu2-3,112 his3-11,15 ura3-1 ade2-ura3::hisG can1-100 spo13:hisG his3::ade2-5_′Δ_-TRP1-ade2-n Rattray and Symington (1994)
YMT353 A yAR71 derivative with rad52-329 This study
YMT362 A yAR71 derivative with rad52::KlURA3 This study
YMT367 A yAR71 derivative with rad51::hisG-URA3-hisG This study
Strain Genotype Reference
YMT013 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 This study
YMT026 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD52/rad52-329 This study
YMT050 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD51/rad51::hisG This study
YMT038 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD52/rad52:: hisG-URA3-hisG This study
YMT178 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/his3 TLC1/tlc1::LEU2 RAD51/rad51::hisGRAD52/rad52::hisG-URA3-hisG This study
YMT075 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD51/rad51::hisG RAD52/rad52-329 This study
YMT106 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG This study
YMT101 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG RAD52/rad52-329 This study
YMT116 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG RAD51/rad51::hisG This study
YMT111 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD59/rad59::hisG-URA3-hisG RAD51/rad51::hisG RAD52/rad52-329 This study
YMT335 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 RAD50/rad50::hisG-URA3-hisG RAD52/rad52-329 This study
YMT381 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 MRE11/mre11::hisG-URA3-hisG This study
YMT385 MATaura3/ura3 leu2/leu2 TRP1/trp1 his3/HIS3 TLC1/tlc1::LEU2 XRS2/xrs2::KanMX6 This study
JHUY563 MATadhml::ADE1/dhml::ADE1 dhmr::ADE1/dhmr::ADE1 ade1/ade1 leu2/leu2 lys5/lys5 ura3/ura3 RAD51/rad51::LEU2 tlc1:URA3/TLC1 rad59::kanMX4/RAD59 Chen et al. (2001)
YDB057 ho HML::HMRa-_(URA3) MAT_α matdist::tetO(112)-(NAT1) HMRa ade3::GAL10-HO ade1-110 trp1 leu2::TetR-GFP-(LEU2) T. Miyazaki and A. Shinohara (unpublished data)
YMT546 A YDB057 derivative with rad52-329 This study
yAR71 MATa trp1-1 leu2-3,112 his3-11,15 ura3-1 ade2-ura3::hisG can1-100 spo13:hisG his3::ade2-5_′Δ_-TRP1-ade2-n Rattray and Symington (1994)
YMT353 A yAR71 derivative with rad52-329 This study
YMT362 A yAR71 derivative with rad52::KlURA3 This study
YMT367 A yAR71 derivative with rad51::hisG-URA3-hisG This study

Liquid assay for telomere recombination: Diploid cells were transformed with a tlc1::LEU2 fragment from a plasmid, pBS/tlc1::LEU2 (a generous gift from F. Ishikawa), and selected on an SD plate lacking leucine. Diploid cells heterologous for tlc1::LEU2 were sporulated and dissected. Haploid cells with an appropriate genotype were selected and grown to saturation in YPDA medium at 30°. Every 24 hr, the cell density was measured by counting cells with a hemocytometer, and then the cultures were diluted with fresh YPDA medium to a density of 105 cells/ml. This cycle was repeated for 10–16 days. At various time points, cells were collected for Southern blotting. Several independent isolates were analyzed for each mutant.

Single-colony streak assay: Haploid cells were selected as described above and streaked on a YPDA plate. After incubation for 48 hr at 30°, single colonies were picked up and restreaked on a fresh YPDA plate. This restreaking was repeated 10 times to permit loss of viability and appearance of survivors. Single colonies from streak 7 were grown to saturation in YPDA medium, and cells were collected for DNA analysis.

Southern blotting of telomeres: Genomic DNAs were digested with _Xho_I and separated in a 0.8% agarose gel. DNA was transferred onto a Hybond N membrane (Amersham, Buckinghamshire, UK) and UV crosslinked. The membrane was then hybridized with a 32P-labeled, random-primed _Xho_I, _Eco_RI-fragment from a plasmid, pYNH3 (provided by F. Ishikawa). For characterization of the type II telomere pattern, DNAs were digested with _Alu_I, _Hae_III, _Hin_fI, and _Msp_I, subjected to electrophoresis in a 1.2% agarose gel, and then analyzed by Southern blotting. Blots were visualized using the phosphorimager BAS3000 (Fuji).

Physical analysis of mating-type switching: Cells were grown to a density of 1 × 107 cells/ml in YP-raffinose medium and HO endonuclease was induced by addition of 2% galactose. Strains were treated after 60 min, collected, and resuspended with YPDA medium to repress the GAL10::HO gene. At intervals, aliquots of cells were collected for DNA analysis. The DNA was digested with _Sty_I and subjected to electrophoresis in a 0.8% agarose gel for 18 hr at 10 V/cm. DNA fragments were transferred onto a nylon membrane and hybridized with a 32P-labeled probe. The probe was made using a PCR-amplified fragment: forward, 5′-TATGGCTATACCCTTATC, and reverse, 5′-GCATTTGAGTGGATACGC.

—Kinetics of mating-type switching. (A) A schematic representation of the MATa and MATα loci indicating the location of the StyI sites and the hybridization probe. (B) HO endonuclease produces a 0.7-kb fragment from the 1.9-kb MATα fragment. A 1.0-kb StyI fragment is produced when the mating type switches from MATα to MATa. Cells were exposed to galactose for 1 hr and incubated further in medium containing glucose. Time “0” is the time when the medium was exchanged. DNAs were isolated from cultures at indicated times and analyzed by Southern blotting. A nonspecific band from the rad52-329 allele is indicated by an asterisk. Wild type (YDB057), uninduced, lane 1; wild type, induced, lanes 2–11; rad52-329 (YMT546), uninduced, lane 12; rad52-329, induced, lanes 13–22.

Figure 2.

—Kinetics of mating-type switching. (A) A schematic representation of the MATa and _MAT_α loci indicating the location of the _Sty_I sites and the hybridization probe. (B) HO endonuclease produces a 0.7-kb fragment from the 1.9-kb _MAT_α fragment. A 1.0-kb _Sty_I fragment is produced when the mating type switches from _MAT_α to MATa. Cells were exposed to galactose for 1 hr and incubated further in medium containing glucose. Time “0” is the time when the medium was exchanged. DNAs were isolated from cultures at indicated times and analyzed by Southern blotting. A nonspecific band from the rad52-329 allele is indicated by an asterisk. Wild type (YDB057), uninduced, lane 1; wild type, induced, lanes 2–11; rad52-329 (YMT546), uninduced, lane 12; rad52-329, induced, lanes 13–22.

Determination of recombination frequencies: Recombination frequencies were determined as described previously (Rattray and Symington 1994). Single colonies were grown on YPDA plates for 2–3 days. More than 15 pink colonies were resuspended in water and plated at the appropriate dilutions to determine total cell number and the number of Ade+ prototrophs. Median mitotic recombination frequencies were determined and rates (events/cell/generation) were calculated according to the following formula: rate = (0.4343 × median frequency)/(log N – log _N_0), where N is the number of cells present in the colony and _N_0 (number of initial cells) = 1.

RESULTS

To examine the role of N- and C-terminal domains of Rad52 protein, we analyzed the effect of a C-terminal deletion mutant called rad52-329 on recombination. The rad52-329 allele, deleting 173 amino acids of a C-terminal Rad51-interacting domain of Rad52, encodes a hypothetical 296-amino-acid protein when it is translated from the third codon of the open reading frame (Adzuma et al. 1984). Western blotting shows that this mutant cell expresses Rad52 protein with the expected size of 32 kD (data not shown). The mutant shows mild sensitivity to DNA-damaging agents (data not shown), as do other C-terminal deletion mutants of RAD52 such as rad52-327 (Adzuma et al. 1984; Boundy-Mills and Livingston 1993; Asleson et al. 1999).

Repair of HO-induced DSBs is defective in a rad52-329 mutant: To confirm that the rad52-329 mutant is indeed defective in _RAD51_-dependent recombination, we analyzed the mating-type switching, a gene conversion event completely dependent on RAD51 function (Pâques and Haber 1999). Repair of an HO-induced DSB at the MAT locus is monitored at the DNA level after induction of HO endonuclease for 1 hr. To measure the formation of recombination products, the DNA samples were digested with _Sty_I, which cut within Ya, but not Yα sequences (Figure 2A). The appearance of a 1.0-kb _Sty_I fragment is indicative of DSB repair from the HMRa locus. In the wild-type strain, the repair is efficient and completed within 150 min (Figure 2B). In the rad52-329 mutant, the DSB was introduced with similar kinetics as in wild type, but no recombinant molecule was formed. The DSB disappeared during further incubation, possibly due to the extensive degradation of the DSB ends. Consistent with this, we observed that the amount of a 2.3-kb DNA fragment next to the DSB fragment decreased in the mutant (lanes 19–22). These results indicate that the rad52-329 mutant is completely defective in the _RAD51_-dependent gene conversion at the MAT locus.

—Recombination between inverted repeats. (A) A schematic representation of the recombination substrate (for more detail, see Rattray and Symington 1994). (B) Colonies from wild-type (yAR71), rad52-329 (YMT353), rad52 deletion (YMT362), and rad51 deletion (YMT367) cells were suspended in water and plated on YPDA and SD lacking adenine. Recombination rates were calculated as described in materials and methods. At least 15 independent colonies were analyzed for each strain and the median values were used for the calculation.

Figure 3.

—Recombination between inverted repeats. (A) A schematic representation of the recombination substrate (for more detail, see Rattray and Symington 1994). (B) Colonies from wild-type (yAR71), rad52-329 (YMT353), rad52 deletion (YMT362), and rad51 deletion (YMT367) cells were suspended in water and plated on YPDA and SD lacking adenine. Recombination rates were calculated as described in materials and methods. At least 15 independent colonies were analyzed for each strain and the median values were used for the calculation.

rad52-329 reduces recombination between inverted repeats: We also analyzed recombination between inverted repeats in the rad52-329 mutant. A previously described recombination substrate was used (Rattray and Symington 1994). The substrate is located on chromosome XV and consists of inverted heteroalleles of the ADE2 gene (Figure 3A). Both alleles are inactive, but the substrate can be rearranged by recombination events to form a functional ADE2 gene. Using this substrate, the rate of recombination was determined to be 3.78 × 10–5 events/cell/generation in wild type and <4.2 × 10–8 in a rad52 deletion mutant (Figure 3B), consistent with the previous report (Rattray and Symington 1994). The rad52-329 mutant reduces the rate fourfold relative to wild type. This rate is comparable to that in the rad51 mutant, but is much higher than that in the rad52 deletion. Thus, the rad52-329 mutant is partially defective in recombination between the repeats. Furthermore, we found that recombinants in the rad52-329 mutant were often accompanied with inversion of an intervening marker, TRP1, as seen in rad51 (data not shown). The inversion occurs in a nonconservative, _RAD51_-independent recombination (Bai et al. 1999; Kang and Symington 2000). Thus, the rad52-329 mutant is proficient in _RAD51_-independent recombination between the repeats.

In the absence of telomerase, the rad52-329 mutant initially shows senescence but later elongates its telomeres: In the absence of the TLC1 gene, which encodes an RNA component of yeast telomerase, cells initially loose viability and later subsets of the cells survive by elongating their telomeres by recombination (Lundblad and Blackburn 1993). We compared viability and telomere lengths of a set of isogenic mutants (in the background of S288C), including tlc1, tlc1 rad52, and tlc1 rad52-329. Cell viability was measured by serial streaking and serial dilution assays. In the serial dilution assay, cell viability was measured by diluting liquid cultures to 105 cells/ml, allowing them to regrow for 24 hr, and counting the cell density (Singer and Gottschling 1994; Le et al. 1999). Wild-type cells reach 108 cells/ml in 24 hr, but tlc1 mutant cells that have lost viability or grow more slowly will reach only lower densities (see Figure 4). The ability to generate survivors was defined as the cells' ability to recover upon reaching their minimum growth rate after 6–9 days in culture. Since each cell seems to follow a unique fate in the telomere metabolism, we analyzed several independent cultures for each mutant, which were derived from independent isolates in the tetrad analysis of parental diploids.

The tlc1 mutant's growth rate in the S288C background declined gradually over 7 days and then increased as reported in other backgrounds (Lundblad and Blackburn 1993; Singer and Gottschling 1994). The tlc1 rad52-329 mutant's growth rate declined faster than that of the tlc1 mutant, although more slowly than that of the tlc1 rad52 mutant (Figure 4A). The tlc1 rad52 mutant produces few survivors (see below), but tlc1 rad52-329 generates survivors after 6 days.

We next constructed a tlc1 rad52-329 rad51 triple mutant and compared it to the tlc1 rad51 and tlc1 rad52-329 double mutants (Figure 4A). tlc1 rad51 shows an accelerated decline and generates survivors after 6 days as reported previously (Le et al. 1999; Chen et al. 2001). tlc1 rad52-329 rad51 exhibits a survival curve similar to those of tlc1 rad51 and tlc1 rad52-329, suggesting that rad52-329 and rad51 mutant cells without telomerase expand their telomeres via the same pathway.

The rad52-329 mutant generates only type II survivors: To examine which recombination pathway gives rise to survivors, telomere structures were analyzed. Genomic DNAs were digested with _Xho_I and analyzed by Southern blotting using a telomere probe (Figure 5). In the absence of telomerase, two distinct types of survivors are seen, as shown by the pattern of _Xho_I restriction fragments. In type I survivors, a short _Xho_I fragment <1 kb and strong amplified 5.2- and/or 6.7-kb Y′ fragments are formed. Type II survivors have long tracts of telomeric TG1–3 repeats and some amplified Y′ fragments; the amplification of Y′ in type II depends on the isolates examined (Teng and Zakian 1999). In these type II survivors, the multiple distinct bands of telomeres between 1 and 6 kb or more (sometimes smeary) represent individual chromosomes with different lengths of TG1–3 repeats. In the liquid assay, type II survivors grow faster than type I survivors and are expected to dominate (Teng et al. 2000). Consistent with this, tlc1 survivors predominantly showed the type II telomere pattern (Figure 5A). The tlc1 rad52-329 and tlc1 rad52-329 rad51 survivors also showed the type II telomere pattern, as did the survivors in the tlc1 rad51 mutant (Figure 5, B–D).

—The rad52-329 and rad59 mutants elongate telomeres in the absence of telomerase. Cells were grown to saturation, counted every 24 hr with a hemocytometer, and then diluted to 105 cells/ml. The curves were the average of cultures from independent isolates with each genotype; “n” refers to the number of independent isolates analyzed. (A) Open circles, WT (n = 3); solid circles, tlc1 (n = 11); open triangles, tlc1 rad52-329 (n = 7); solid triangles, tlc1 rad52 (n = 20); open squares, tlc1 rad51 (n = 12); solid squares, tlc1 rad51 rad52-329 (n = 5). (B) Open circles, WT; solid circles, tlc1; open triangles, tlc1 rad59 fast-growing survivors (n = 6); solid triangles, tlc1 rad59 slow-growing survivors (n = 3). (C) Open circles, WT; solid circles, tlc1; open triangles, tlc1 rad59 fast-growing survivors; solid triangles, tlc1 rad51 rad59 (n = 3); open squares, tlc1 rad51; solid squares, tlc1 rad51 rad52-329 rad59 (n = 5). (D) Open circles, WT; solid circles, tlc1; open triangles, tlc1 rad52-329; solid triangles, tlc1 rad52-329 rad59 (n = 1); open squares, tlc1 rad59 slow-growing survivors.

Figure 4.

—The rad52-329 and rad59 mutants elongate telomeres in the absence of telomerase. Cells were grown to saturation, counted every 24 hr with a hemocytometer, and then diluted to 105 cells/ml. The curves were the average of cultures from independent isolates with each genotype; “_n_” refers to the number of independent isolates analyzed. (A) Open circles, WT (n = 3); solid circles, tlc1 (n = 11); open triangles, tlc1 rad52-329 (n = 7); solid triangles, tlc1 rad52 (n = 20); open squares, tlc1 rad51 (n = 12); solid squares, tlc1 rad51 rad52-329 (n = 5). (B) Open circles, WT; solid circles, tlc1; open triangles, tlc1 rad59 fast-growing survivors (n = 6); solid triangles, tlc1 rad59 slow-growing survivors (n = 3). (C) Open circles, WT; solid circles, tlc1; open triangles, tlc1 rad59 fast-growing survivors; solid triangles, tlc1 rad51 rad59 (n = 3); open squares, tlc1 rad51; solid squares, tlc1 rad51 rad52-329 rad59 (n = 5). (D) Open circles, WT; solid circles, tlc1; open triangles, tlc1 rad52-329; solid triangles, tlc1 rad52-329 rad59 (n = 1); open squares, tlc1 rad59 slow-growing survivors.

Since liquid cultures select the fastest-growing cells, we also analyzed survivors from single colonies on plates. After survivors were generated, a single colony was picked and cultured overnight, and its genomic DNA was analyzed (Table 2). In 36 independent colonies of a tlc1 survivor, both type I and type II survivors were detected. The ratio of type I to type II is 0.9, which is less than that reported previously (Teng and Zakian 1999; Chen et al. 2001). On the other hand, the tlc1 rad52-329, tlc1 rad51, and tlc1 rad52-329 rad51 mutants give rise to type II survivors alone, even when analyzed by the plate assay. The fact that the rad52-329 mutant carries out type II telomere recombination shows that the C-terminal domain of Rad52 is dispensable for type II telomere recombination.

The tlc1 rad59 mutant generates both type I and type II survivors: The structure of Rad52-329 is very similar to that of Rad59 (Bai and Symington 1996). We therefore examined the effect of a rad59 null mutation on telomere recombination in the absence of TLC1 and compared phenotypes to those of the tlc1 rad52-329 mutant. Previously, Chen et al. (2001) reported that the tlc1 rad59 double mutant generates only type I survivors. However, in the liquid assay, we found that it can form both type I and type II survivors. Cultures from independent isolates of the tlc1 rad59 double mutant follow either of two fates (Figure 4B). On the basis of their growth curves and telomere structures, we refer to fastgrowing and slow-growing tlc1 rad59 survivors (Figure 4B), which were also verified by Southern blotting to show type II and type I telomere patterns, respectively (Figure 5E). The fast-growing cells generated survivors after 8 days, while the slow-growing cells did so after 9 days. Moreover, the slow-growing survivors did not reach normal growth rate during subsequent culture. The presence of the type II telomere pattern in fast-growing tlc1 rad59 survivors was confirmed by digesting genomic DNA with four 4-base recognition restriction enzymes (Figure 6). This enabled us to measure the length of telomeric TG1–3 repeats, since the enzymes digest most of the genome into small pieces, leaving the repeat intact (Teng et al. 2000). The fast-growing tlc1 rad59 survivors produced various lengths of longer TG1–3 repeats, as seen in tlc1 cells. These results indicate that RAD59 is not essential for efficient type II telomere recombination.

—The rad52-329 mutant generates type II survivors. An independent isolate of each mutant was grown in YPDA medium. On the days indicated at the top of each lane, genomic DNA was extracted and structures of telomeres were determined. Genomic DNA was digested with XhoI. Southern blotting was carried out as described in materials and methods. Positions of long Y′ (6.7 kb) and short Y′ (5.2 kb) fragments are shown on the right. In the tlc1 mutant culture (A, lanes 4 and 5), some amplification of the short Y′ fragment is seen in addition to the type II telomere pattern. Loss of the short Y′ fragment was observed in fast-growing tlc1 rad59 and tlc1 rad51 rad59 rad52-329 mutant cultures, while the loss of long Y′ was in some tlc1 cultures (E, lane 2). These losses of Y′ fragments are dependent upon the isolates examined. (A) WT, lane 1; tlc1, lanes 2–5. (B) tlc1 rad52-329. (C) tlc1 rad51. (D) tlc1 rad51 rad52-329. (E) WT, lane 1; tlc1, lane 2; fast-growing tlc1 rad59 survivors, lanes 3–7; slow-growing tlc1 rad59 survivors, lanes 8–12. (F) WT, lane 1; tlc1, lane 2; tlc1 rad51 rad59 rad52-329, lanes 3–6.

Figure 5.

—The rad52-329 mutant generates type II survivors. An independent isolate of each mutant was grown in YPDA medium. On the days indicated at the top of each lane, genomic DNA was extracted and structures of telomeres were determined. Genomic DNA was digested with _Xho_I. Southern blotting was carried out as described in materials and methods. Positions of long Y′ (6.7 kb) and short Y′ (5.2 kb) fragments are shown on the right. In the tlc1 mutant culture (A, lanes 4 and 5), some amplification of the short Y′ fragment is seen in addition to the type II telomere pattern. Loss of the short Y′ fragment was observed in fast-growing tlc1 rad59 and tlc1 rad51 rad59 rad52-329 mutant cultures, while the loss of long Y′ was in some tlc1 cultures (E, lane 2). These losses of Y′ fragments are dependent upon the isolates examined. (A) WT, lane 1; tlc1, lanes 2–5. (B) tlc1 rad52-329. (C) tlc1 rad51. (D) tlc1 rad51 rad52-329. (E) WT, lane 1; tlc1, lane 2; fast-growing tlc1 rad59 survivors, lanes 3–7; slow-growing tlc1 rad59 survivors, lanes 8–12. (F) WT, lane 1; tlc1, lane 2; tlc1 rad51 rad59 rad52-329, lanes 3–6.

In the liquid assay, 6 of 9 tlc1 rad59 cultures from independent isolates generate type II survivors, while the remaining 3 generate type I survivors (Table 3). However, the ratio might be an overestimate since the liquid assay favors fast-growing type II cells. Indeed, in the plate assay, only 2 of 10 independent tlc1 rad59 surviving colonies show the type II telomere pattern (Table 2). Therefore, the presence of the type II telomere pattern in tlc1 rad59 without growth selection confirms the above idea that the rad59 mutant without telomerase is proficient in type II recombination, but with a more reduced rate than that of the tlc1 strain.

We further confirmed that RAD59 is not essential for type II recombination by analyzing survivors in a tlc1 rad51 rad59 triple mutant, which eliminates the type I pathway. In the liquid assay, three of nine cultures from independent isolates of the tlc1 rad51 rad59 triple mutant generated survivors (Figure 4C), while the remaining six did not generate any survivors (Table 3). On the other hand, all isolates of the tlc1 rad51 and tlc1 rad59 double mutants produced survivors. Southern analysis confirmed the type II telomere pattern in the survivors (data now shown). These results are consistent with the hypothesis that RAD59 facilitates efficient _RAD52_-dependent type II recombination.

Our results on the presence of a type II telomere pattern in the tlc1 rad59 survivors are in contrast to those reported previously (Chen et al. 2001). Since the strain backgrounds are different between previous and current studies, we reanalyzed a tlc1 rad59 strain used by

TABLE 2

Distribution of survivor types of various mutants lacking telomerase in the plate assay

No. of colonies_a_ (% of total)
Strain (relevant genotype) Total studied Type I Type II P value_b_
tlc1 36 17 (47) 19 (53)
tlc1 rad52 NA NA NA NA
tlc1 rad51 11 0 (0) 11 (100) 0.0043
tlc1 rad52-329 14 0 (0) 14 (100) 0.0016
tlc1 rad51 rad52-329 8 0 (0) 8 (100) 0.013
tlc1 rad59 10 8 (80) 2 (20) 0.65
tlc1 rad50 14 12 (86) 2 (14) 0.013
No. of colonies_a_ (% of total)
Strain (relevant genotype) Total studied Type I Type II P value_b_
tlc1 36 17 (47) 19 (53)
tlc1 rad52 NA NA NA NA
tlc1 rad51 11 0 (0) 11 (100) 0.0043
tlc1 rad52-329 14 0 (0) 14 (100) 0.0016
tlc1 rad51 rad52-329 8 0 (0) 8 (100) 0.013
tlc1 rad59 10 8 (80) 2 (20) 0.65
tlc1 rad50 14 12 (86) 2 (14) 0.013

NA, not available.

a

The telomere structure of independent colonies after streak 7 was determined as described in materials and methods.

b

P values were obtained by the chi-square independent test.

TABLE 2

Distribution of survivor types of various mutants lacking telomerase in the plate assay

No. of colonies_a_ (% of total)
Strain (relevant genotype) Total studied Type I Type II P value_b_
tlc1 36 17 (47) 19 (53)
tlc1 rad52 NA NA NA NA
tlc1 rad51 11 0 (0) 11 (100) 0.0043
tlc1 rad52-329 14 0 (0) 14 (100) 0.0016
tlc1 rad51 rad52-329 8 0 (0) 8 (100) 0.013
tlc1 rad59 10 8 (80) 2 (20) 0.65
tlc1 rad50 14 12 (86) 2 (14) 0.013
No. of colonies_a_ (% of total)
Strain (relevant genotype) Total studied Type I Type II P value_b_
tlc1 36 17 (47) 19 (53)
tlc1 rad52 NA NA NA NA
tlc1 rad51 11 0 (0) 11 (100) 0.0043
tlc1 rad52-329 14 0 (0) 14 (100) 0.0016
tlc1 rad51 rad52-329 8 0 (0) 8 (100) 0.013
tlc1 rad59 10 8 (80) 2 (20) 0.65
tlc1 rad50 14 12 (86) 2 (14) 0.013

NA, not available.

a

The telomere structure of independent colonies after streak 7 was determined as described in materials and methods.

b

P values were obtained by the chi-square independent test.

Chen and colleagues. The tlc1 rad59 double mutant formed survivors slightly later than the tlc1 mutant (Figures 7A and 8), consistent with the previous results (Chen et al. 2001). By analyzing a telomere structure in individual cultures, we found that 2 of 17 _tlc1 rad59_-independent isolates clearly showed a band pattern of type II (lane 1 in Figure 7B). The remaining 15 isolates exhibited a band pattern typical for type I survivors, but some of them contained faint bands of amplified TG1–3 repeats (lanes 8 and 9 in Figure 7B). These results are consistent with the conclusion that Rad59 is necessary for most, but not all, type II recombination.

—The tlc1 rad59 mutant generates both type I and type II survivors. Genomic DNA was prepared from each strain and digested with AluI, HaeIII, HinfI, and MspI. Southern blotting was performed as described in materials and methods. The numbers at the top of the lanes represent the number of days that cells were grown. WT, lane 1; tlc1, lane 2; fast-growing tlc1 rad59 survivors, lanes 3–6; slow-growing tlc1 rad59 survivors, lanes 7–10.

Figure 6.

—The tlc1 rad59 mutant generates both type I and type II survivors. Genomic DNA was prepared from each strain and digested with _Alu_I, _Hae_III, _Hin_fI, and _Msp_I. Southern blotting was performed as described in materials and methods. The numbers at the top of the lanes represent the number of days that cells were grown. WT, lane 1; tlc1, lane 2; fast-growing tlc1 rad59 survivors, lanes 3–6; slow-growing tlc1 rad59 survivors, lanes 7–10.

Rad52-329 alone promotes type II telomere recombination: To examine whether the N-terminal portion of Rad52 alone has an ability to carry out recombination, we next analyzed the tlc1 rad52-329 rad59 triple mutant. In the liquid assay, only 1 of 11 independent isolates generated fast-growing survivors (Figure 4D), which were of type II (data not shown), while the remaining 10 did not generate any survivors. This indicates that the N-terminal domain of Rad52 can perform type II telomere recombination in the absence of RAD59, although the recombination in the mutant is very inefficient. However, a tlc1 rad51 rad52-329 rad59 quadruple mutant generates survivors more efficiently than the tlc1 rad52-329 rad59 triple mutant; five of seven cultures from independent isolates of the quadruple mutant gave rise to fast-growing survivors (Figure 4C and Table 3). The difference between the tlc1 rad52-329 rad59 triple mutant and the tlc1 rad51 rad52-329 rad59 quadruple mutant is statistically significant (P < 0.05, Student's _t_-test). All the survivors show the type II telomere pattern (Figure 5F). This supports the conclusion that Rad52-329 can recombine telomeres in the absence of Rad51 and Rad59. It also suggests that Rad51 inhibits type II telomere recombination.

The tlc1 rad52 double mutant generates few survivors (Lundblad and Blackburn 1993): Of 25 independent isolates, 1 gave rise to survivors, and, on the basis of Southern blotting, these appeared to be type II (Table 3 and data not shown). However, these survivors are not typical type II, since they exhibit very slow growth, more like type I survivors. This suggests that even in the absence of RAD52, Rad51 may elongate the telomeres very inefficiently. Alternatively, it is possible to postulate a third minor telomere elongation pathway, which is independent of both RAD51 and RAD52. Since RAD51 inhibits _RAD51_-independent recombination, we reanalyzed

TABLE 3

Summary of the liquid assay for telomere recombination in various mutants lacking TLC1 and various recombination genes

Strain (relevant genotype) Senescence_a_ No. of isolates generating survivors_b_ Telomere pattern_c_
tlc1 + 11/11 II
tlc1 rad52 +++ 1/25_d_ II?d
tlc1 rad51 +++ 12/12 II
tlc1 rad52-329 ++ 7/7 II
tlc1 rad51 rad52 +++ 0/25 NA
tlc1 rad51 rad52-329 +++ 5/5 II
tlc1 rad59 + 9/9 I (3/9) and II (6/9)e
tlc1 rad51 rad59 +++ 3/9 II
tlc1 rad59 rad52-329 +++ 1/11 II
tlc1 rad51 rad59 rad52-329 +++ 5/7 II
tlc1 rad50 +++ 7/7 I and II
tlc1 rad50 rad52-329 +++ 2/11 II
tlc1 mre11 +++ 14/14 I and II
tlc1 xrs2 +++ 14/14 I and II
Strain (relevant genotype) Senescence_a_ No. of isolates generating survivors_b_ Telomere pattern_c_
tlc1 + 11/11 II
tlc1 rad52 +++ 1/25_d_ II?d
tlc1 rad51 +++ 12/12 II
tlc1 rad52-329 ++ 7/7 II
tlc1 rad51 rad52 +++ 0/25 NA
tlc1 rad51 rad52-329 +++ 5/5 II
tlc1 rad59 + 9/9 I (3/9) and II (6/9)e
tlc1 rad51 rad59 +++ 3/9 II
tlc1 rad59 rad52-329 +++ 1/11 II
tlc1 rad51 rad59 rad52-329 +++ 5/7 II
tlc1 rad50 +++ 7/7 I and II
tlc1 rad50 rad52-329 +++ 2/11 II
tlc1 mre11 +++ 14/14 I and II
tlc1 xrs2 +++ 14/14 I and II

NA, not available.

a

“ + ,” “ + +,” and “ + + +” indicate mild, moderate, and rapid senescence, respectively. See Figure 4 for comparison.

b

Cultures of independent isolates from each strain were analyzed. The number of isolates producing survivors per total isolates are shown; e.g., 1/11 means that only one isolate generated survivors and the remaining 10 isolates did not generate any survivors during culture.

c

Telomere pattern was determined by Southern blotting, as shown in Figure 5.

d

Oneof 25 tlc1 rad52 isolates generated the type II telomere pattern. This survivor was unusual; e.g., it grew much more slowly than other type II survivors.

e

Parentheses show factions of isolates with type I and type II telomere patterns.

TABLE 3

Summary of the liquid assay for telomere recombination in various mutants lacking TLC1 and various recombination genes

Strain (relevant genotype) Senescence_a_ No. of isolates generating survivors_b_ Telomere pattern_c_
tlc1 + 11/11 II
tlc1 rad52 +++ 1/25_d_ II?d
tlc1 rad51 +++ 12/12 II
tlc1 rad52-329 ++ 7/7 II
tlc1 rad51 rad52 +++ 0/25 NA
tlc1 rad51 rad52-329 +++ 5/5 II
tlc1 rad59 + 9/9 I (3/9) and II (6/9)e
tlc1 rad51 rad59 +++ 3/9 II
tlc1 rad59 rad52-329 +++ 1/11 II
tlc1 rad51 rad59 rad52-329 +++ 5/7 II
tlc1 rad50 +++ 7/7 I and II
tlc1 rad50 rad52-329 +++ 2/11 II
tlc1 mre11 +++ 14/14 I and II
tlc1 xrs2 +++ 14/14 I and II
Strain (relevant genotype) Senescence_a_ No. of isolates generating survivors_b_ Telomere pattern_c_
tlc1 + 11/11 II
tlc1 rad52 +++ 1/25_d_ II?d
tlc1 rad51 +++ 12/12 II
tlc1 rad52-329 ++ 7/7 II
tlc1 rad51 rad52 +++ 0/25 NA
tlc1 rad51 rad52-329 +++ 5/5 II
tlc1 rad59 + 9/9 I (3/9) and II (6/9)e
tlc1 rad51 rad59 +++ 3/9 II
tlc1 rad59 rad52-329 +++ 1/11 II
tlc1 rad51 rad59 rad52-329 +++ 5/7 II
tlc1 rad50 +++ 7/7 I and II
tlc1 rad50 rad52-329 +++ 2/11 II
tlc1 mre11 +++ 14/14 I and II
tlc1 xrs2 +++ 14/14 I and II

NA, not available.

a

“ + ,” “ + +,” and “ + + +” indicate mild, moderate, and rapid senescence, respectively. See Figure 4 for comparison.

b

Cultures of independent isolates from each strain were analyzed. The number of isolates producing survivors per total isolates are shown; e.g., 1/11 means that only one isolate generated survivors and the remaining 10 isolates did not generate any survivors during culture.

c

Telomere pattern was determined by Southern blotting, as shown in Figure 5.

d

Oneof 25 tlc1 rad52 isolates generated the type II telomere pattern. This survivor was unusual; e.g., it grew much more slowly than other type II survivors.

e

Parentheses show factions of isolates with type I and type II telomere patterns.

survivors of a rad51 rad52 double mutant in the absence of TLC1. No survivors were recovered in the liquid assay (0/25).

The tlc1 rad52-329 rad50 mutant generates type II survivors: Next, we analyzed the tlc1 rad52-329 rad50 triple mutant. RAD50 is required, but not essential for type II recombination (Le et al. 1999; Chen et al. 2001). As shown previously, all independent isolates of the tlc1 rad50 double mutant generated slow-growing survivors with type I telomere pattern (data not shown). On the other hand, 2 of 14 isolates of the tlc1 rad52-329 rad50 triple mutant generated survivors, while the remaining 12 isolates did not generate any survivors (Figure 9A). The survivors in the triple mutant exhibited a typical type II telomere pattern (Figure 9B). Thus, we concluded that Rad52-329 protein can carry out type II telomere recombination even in the absence of RAD50 function.

—The tlc1 rad59 mutant in a different strain background also generates type II survivors. JHUY564 was sporulated and survival of haploid colonies with a relevant genotype was analyzed in the liquid assay. The survival curves of each mutant (A) and Southern blots of telomere structure at 14 days in seven individual isolates (B) were obtained as described in materials and methods. (A) Solid circles, tlc1; open triangles, tlc1 rad59 (an average of 14 independent isolates). (B) Telomere structure in tlc1 rad59 survivors, lanes 1–9. Some amplified TG1–3 repeats in type I survivors are indicated by arrowheads.

Figure 7.

—The tlc1 rad59 mutant in a different strain background also generates type II survivors. JHUY564 was sporulated and survival of haploid colonies with a relevant genotype was analyzed in the liquid assay. The survival curves of each mutant (A) and Southern blots of telomere structure at 14 days in seven individual isolates (B) were obtained as described in materials and methods. (A) Solid circles, tlc1; open triangles, tlc1 rad59 (an average of 14 independent isolates). (B) Telomere structure in tlc1 rad59 survivors, lanes 1–9. Some amplified TG1–3 repeats in type I survivors are indicated by arrowheads.

—The tlc1 rad59 mutants in different strain backgrounds show different kinetics for generation of survivors. Cells were grown to saturation, counted every 24 hr with a hemocytometer, and then diluted to 105 cells/ml. The curves from eight independent isolates are shown. (A) tlc1 rad59 in S288C background. Open circles, triangles, and squares indicate fast-growing survivors; solid circles, triangles, and squares indicate slow-growing survivors. (B) tlc1 rad59 in JHUY564 background.

Figure 8.

—The tlc1 rad59 mutants in different strain backgrounds show different kinetics for generation of survivors. Cells were grown to saturation, counted every 24 hr with a hemocytometer, and then diluted to 105 cells/ml. The curves from eight independent isolates are shown. (A) tlc1 rad59 in S288C background. Open circles, triangles, and squares indicate fast-growing survivors; solid circles, triangles, and squares indicate slow-growing survivors. (B) tlc1 rad59 in JHUY564 background.

The tlc1 mre11 and tlc1 xrs2 mutants are similar to the tlc1 rad50 mutant: Rad50 forms a complex with Mre11 and Rad50 (Usui et al. 1998). Previous reports showed some functional differences in telomere maintenance among components of the complex (Boulton and Jackson 1998; Bucholc et al. 2001). Different from the previous report (Boulton and Jackson 1998), all three mutant cells in the S288C background show a similar growth rate even at a high temperature such as 37° (data not shown). We also studied the effect of mre11 and xrs2 null mutations on telomere recombination (Figure 10). Both tlc1 mre11 and tlc1 xrs2 double mutants share the same phenotype as the tlc1 rad50 strain of generating survivors more slowly than the tlc1 mutant (Figure 10A). The mutants generate survivors more slowly than tlc1 alone. Most survivors from tlc1 mre11 or tlc1 xrs2 show the type I telomere pattern, but a few show the type II telomere pattern (Figure 10B and Table 3). Thus, all three members of the MRX complex are necessary for efficient type II recombination.

DISCUSSION

Previous genetic studies of yeast lacking telomerase have defined two recombination pathways (Kass-Eisler and Greider 2000). The formation of type I survivors requires the _RAD51_-dependent pathway, which involves RAD52, RAD54, RAD55, and RAD57 as well as RAD51. The formation of type II survivors requires the _RAD51_-independent pathway, which involves RAD50, RAD52, RAD59, MRE11, and XRS2. Recently, other factors that promote type II telomere extension have been identified: SGS1, MEC1, and TEL1 (Cohen and Sinclair 2001; Huang et al. 2001; Johnson et al. 2001; Tsai et al. 2002). Among the genes involved in the _RAD51_-independent pathway, three components of the MRX complex are not essential for the formation of type II survivors (Le et al. 1999; Chen et al. 2001). In contrast to the previous report (Chen et al. 2001), we showed that the rad59 mutation reduces type II telomere recombination, but does not eliminate it. TID1/RDH54 is not necessary for telomere recombination (M. Tsukamoto and A. Shinohara, unpublished results). Thus, only RAD52 is an absolute requirement for type II telomere recombination among the RAD52 epistasis group.

One of the key questions in _RAD51_-independent recombination is to find which protein(s) is responsible for the search for homologous sequences and for strand invasion. It is proposed that type II survivors are generated through rolling-circle replication primed by an intrachromosomal telomeric D-loop (Chen et al. 2001; Natarajan and McEachern 2002). We speculate that Rad52 might catalyze the formation of the intrachromosomal D-loop. Since the N-terminal domain of Rad52 catalyzes annealing of complementary ssDNAs (Mortensen et al. 1996; Shinohara and Ogawa 1998; Kagawa et al. 2001), this annealing activity might be involved in D-loop formation.

Our results described here indicate that Rad59 and the MRX complex are important, but not essential, for type II recombination. There might be multiple _RAD51_-independent recombination pathways at telomeres: RAD59 dependent and RAD59 independent or MRX dependent and MRX independent. Alternatively, there is one _RAD51_-independent recombination pathway in which Rad59 and the MRX complex facilitate the recombination catalyzed by Rad52. The genetic requirement for telomere recombination supports the view that type II telomere recombination is mechanistically similar to BIR. However, recombination at telomeres is slightly different from BIR events at other genomic loci. BIR initiated by DSB at the MAT locus requires a specific _cis_-acting sequence (Malkova et al. 2001). Chromosome ends might contain such an element that facilitates BIR. Alternatively, a constraint, which enforces the requirement for the BIR facilitator on internal chromosomal loci, might be alleviated at the end of chromosomes.

—The tlc1 rad52-329 rad50 mutant generates type II survivors. (A) The survival curves of each mutant. Solid circles, tlc1 (n = 11); open circles, tlc1 rad50 (n = 5); solid triangles, tlc1 rad50 rad52-329 (n = 2); “n” refers to the number of independent isolates analyzed. (B) Southern blots of telomere structure in tlc1 rad52-329 rad50 survivors showing the type II telomere pattern.

Figure 9.

—The tlc1 rad52-329 rad50 mutant generates type II survivors. (A) The survival curves of each mutant. Solid circles, tlc1 (n = 11); open circles, tlc1 rad50 (n = 5); solid triangles, tlc1 rad50 rad52-329 (n = 2); “_n_” refers to the number of independent isolates analyzed. (B) Southern blots of telomere structure in tlc1 rad52-329 rad50 survivors showing the type II telomere pattern.

Since telomeres are composed of irregular TG1–3 repeats (Shampay et al. 1984), there is a high potential for mismatches and short homology between terminal sequences when the sequences recombine. Thus, it is thought that _RAD51_-independent telomere recombination is homeologous BIR, which depends on a short homology. Interestingly, deletion of a mismatch repair gene, MSH2, alleviates the growth defect of telomerase-deficient cells, suggesting that homeologous recombination is suppressed in cells without telomerase (Rizki and Lundblad 2001). On the other hand, longer-sequence homology is necessary for type I recombination. Consistent with this idea, Haber and his colleagues showed that a minimum homology length required for the _RAD51_-independent recombination is 30 bp, while that for the _RAD51_-dependent recombination is 100 bp (Ira and Haber 2002).

_RAD51_-dependent and _RAD51_-independent telomere recombination pathways seem to be partially exclusive of each other. Amplification of Y′ elements is rarely seen in type II survivors, while amplification of TG1–3 repeats is not observed in type I survivors of the tlc1. It is possible to consider that a factor, which determines the choice of telomere recombination pathways, might fluctuate in each tlc1 isolate (or cell). Such a factor might be a positive and/or negative regulator for the recombination pathways. Our results that Rad51 has a positive effect on type I, but a negative one on type II recombination, suggest that Rad51 might be a protein whose amounts in a cell determine the pathway choice (see below). Furthermore, it is shown that Cdc13, a telomere end-binding protein, or the Ku complex suppresses _RAD51_-dependent telomere recombination (Grandin and Charbonneau 2003), suggesting that the composition of telomere ends determines a cell's ability to choose a telomere recombination pathway. We showed that the tlc1 rad59 mutants in different strain backgrounds exhibit similar, but distinct, phenotypes in the choice of recombination pathways to elongate telomeres. These results suggest that a genetic difference between strains and/or even in a cell determines which recombination pathway is predominantly used to elongate telomeres in the absence of telomerase.

Telomere elongation by recombination is considered to be a rare event. However, given that all 32 ends of chromosomes should be elongated in survivors of tlc1 cells, we might underestimate the rate of recombination per chromosome end. Further analysis is necessary to determine how frequent telomere recombination is. Interestingly, in some insects such as the mosquito Anopheles and the dipteran Chironomus, telomere recombination is thought to be the sole mechanism for maintaining the repeats at chromosome ends (Lopez et al. 1996; Roth et al. 1997). At least in these organisms, recombination seems to be an efficient process for telomere elongation.

—tlc1 mre11 and tlc1 xrs2 are similar to tlc1 rad50. The survival curves of each mutant (A) and telomere structures (B) were obtained as described above. (A) Solid circles, tlc1 (n = 11); open circles, tlc1 rad50 (n = 7); open triangles, tlc1 mre11 (n = 14); solid triangles, tlc1 xrs2 (n = 14); “n” refers to the number of independent isolates analyzed. (B) tlc1 mre11 type I survivors, lanes 1–5; tlc1 mre11 type II survivors, lanes 6–10; tlc1 xrs2 type I survivors, lanes 11–15; tlc1 xrs2 type II survivors, lanes 16–20.

Figure 10.

tlc1 mre11 and tlc1 xrs2 are similar to tlc1 rad50. The survival curves of each mutant (A) and telomere structures (B) were obtained as described above. (A) Solid circles, tlc1 (n = 11); open circles, tlc1 rad50 (n = 7); open triangles, tlc1 mre11 (n = 14); solid triangles, tlc1 xrs2 (n = 14); “_n_” refers to the number of independent isolates analyzed. (B) tlc1 mre11 type I survivors, lanes 1–5; tlc1 mre11 type II survivors, lanes 6–10; tlc1 xrs2 type I survivors, lanes 11–15; tlc1 xrs2 type II survivors, lanes 16–20.

Footnotes

Communicating editor: L. Symington

Acknowledgement

We are grateful to Akiyo Yamazaki and Harumi Kato for construction of strains and plasmids; to Hisao Masukata, Takuro Nakagawa, and Neil Hunter for helpful discussion; to Lorraine Symington for unpublished results; and to Jim Haber, Fuyuki Ishikawa, Tomoko Ogawa, and Rodney Rothstein for plasmids and Carol Greider and Lorraine Symington for strains. This work was supported by grants from the Ministry of Education, Science and Culture of Japan to Priority Area to A.S.

LITERATURE CITED

Adzuma

K

,

Ogawa

T

,

Ogawa

H

,

1984

Primary structure of the RAD52 gene in Saccharomyces cerevisiae

.

Mol. Cell. Biol.

4

:

2735

2744

.

Alani

E

,

Cao

L

,

Kleckner

N

,

1987

A method for gene disruption that allows repeated use of URA3 selection in the construction of multiply disrupted yeast strains

.

Genetics

116

:

541

545

.

Alani

E

,

Subbiah

S

,

Kleckner

N

,

1989

The yeast RAD50 gene encodes a predicted 153-kD protein containing a purine nucleotide-binding domain and two large heptad-repeat regions

.

Genetics

122

:

47

57

.

Asleson

E N

,

Okagaki

R J

,

Livingston

D M

,

1999

A core activity associated with the N terminus of the yeast RAD52 protein is revealed by RAD51 overexpression suppression of C-terminal rad52 truncation alleles

.

Genetics

153

:

681

692

.

Bai

Y

,

Symington

L S

,

1996

A Rad52 homolog is required for _RAD51_-independent mitotic recombination in Saccharomyces cerevisiae

.

Genes Dev.

10

:

2025

2037

.

Bai

Y

,

Davis

A P

,

Symington

L S

,

1999

A novel allele of RAD52 that causes severe DNA repair and recombination deficiencies only in the absence of RAD51 or RAD59

.

Genetics

153

:

1117

1130

.

Boulton

S J

,

Jackson

S P

,

1998

Components of the Ku-dependent non-homologous end-joining pathway are involved in telomeric length maintenance and telomeric silencing

.

EMBO J.

17

:

1819

1828

.

Boundy-Mills

K L

,

Livingston

D M

,

1993

A Saccharomyces cerevisiae RAD52 allele expressing a C-terminal truncation protein: activities and intragenic complementation of missense mutations

.

Genetics

133

:

39

49

.

Bucholc

M

,

Park

Y

,

Lustig

A J

,

2001

Intrachromatid excision of telomeric DNA as a mechanism for telomere size control in Saccharomyces cerevisiae

.

Mol. Cell. Biol.

21

:

6559

6573

.

Chen

Q

,

Ijpma

A

,

Greider

C W

,

2001

Two survivor pathways that allow growth in the absence of telomerase are generated by distinct telomere recombination events

.

Mol. Cell. Biol.

21

:

1819

1827

.

Cohen

H

,

Sinclair

D A

,

2001

Recombination-mediated lengthening of terminal telomeric repeats requires the Sgs1 DNA helicase

.

Proc. Natl. Acad. Sci. USA

98

:

3174

3179

.

Davis

A P

,

Symington

L S

,

2001

The yeast recombinational repair protein Rad59 interacts with Rad52 and stimulates single-strand annealing

.

Genetics

159

:

515

525

.

Gietz

R D

,

Sugino

A

,

1988

New yeast-Escherichia coli shuttle vectors constructed with in vitro mutagenized yeast genes lacking six-base pair restriction sites

.

Gene

74

:

527

534

.

Grandin

N

,

Charbonneau

M

,

2003

The Rad51 pathway of telomerase-independent maintenance of telomeres can amplify TG1–3 equences in yku and cdc13 mutants of Saccharomyces cerevisiae

.

Mol. Cell. Biol.

23

:

3721

3734

.

Huang

P

,

Pryde

F E

,

Lester

D

,

Maddison

R L

,

Borts

R H

et al. .,

2001

SGS1 is required for telomere elongation in the absence of telomerase

.

Curr. Biol.

11

:

125

129

.

Ira

G

,

Haber

J E

,

2002

Characterization of _RAD51_-independent break-induced replication that acts preferentially with short homologous sequences

.

Mol. Cell. Biol.

22

:

6384

6392

.

Johnson

F B

,

Marciniak

R A

,

McVey

M

,

Stewart

S A

,

Hahn

W C

et al. .,

2001

The Saccharomyces cerevisiae WRN homolog Sgs1p participates in telomere maintenance in cells lacking telomerase

.

EMBO J.

20

:

905

913

.

Johzuka

K

,

Ogawa

H

,

1995

Interaction of Mre11 and Rad50: two proteins required for DNA repair and meiosis-specific double-strand break formation in Saccharomyces cerevisiae

.

Genetics

139

:

1521

1532

.

Kagawa

W

,

Kurumizaka

H

,

Ikawa

S

,

Yokoyama

S

,

Shibata

T

,

2001

Homologous pairing promoted by the human Rad52 protein

.

J. Biol. Chem.

276

:

35201

35208

.

Kagawa

W

,

Kurumizaka

H

,

Ishitani

R

,

Fukai

S

,

Nureki

O

et al. .,

2002

Crystal structure of the homologous-pairing domain from the human Rad52 recombinase in the undecameric form

.

Mol. Cell

10

:

359

371

.

Kang

L E

,

Symington

L S

,

2000

Aberrant double-strand break repair in rad51 mutants of Saccharomyces cerevisiae

.

Mol. Cell. Biol.

20

:

9162

9172

.

Kass-Eisler

A

,

Greider

C W

,

2000

Recombination in telomere-length maintenance

.

Trends Biochem. Sci.

25

:

200

204

.

Kraus

E

,

Leung

W Y

,

Haber

J E

,

2001

Break-induced replication: a review and an example in budding yeast

.

Proc. Natl. Acad. Sci. USA

98

:

8255

8262

.

Le

S

,

Moore

J K

,

Haber

J E

,

Greider

C W

,

1999

RAD50 and RAD51 define two pathways that collaborate to maintain telomeres in the absence of telomerase

.

Genetics

152

:

143

152

.

Lopez

C C

,

Nielsen

L

,

Edstrom

J E

,

1996

Terminal long tandem repeats in chromosomes from Chironomus pallidivittatus

.

Mol. Cell. Biol.

16

:

3285

3290

.

Lundblad

V

,

Blackburn

E H

,

1993

An alternative pathway for yeast telomere maintenance rescues _est1_-senescence

.

Cell

73

:

347

360

.

Malkova

A

,

Ivanov

E L

,

Haber

J E

,

1996

Double-strand break repair in the absence of RAD51 in yeast: a possible role for break-induced DNA replication

.

Proc. Natl. Acad. Sci. USA

93

:

7131

7136

.

Malkova

A

,

Signon

L

,

Schaefer

C B

,

Naylor

M L

,

Theis

J F

et al. .,

2001

_RAD51_-independent break-induced replication to repair a broken chromosome depends on a distant enhancer site

.

Genes Dev.

15

:

1055

1060

.

McEachern

M J

,

Krauskopf

A

,

Blackburn

E H

,

2000

Telomeres and their control

.

Annu. Rev. Genet.

34

:

331

358

.

Milne

G T

,

Weaver

D T

,

1993

Dominant negative alleles of RAD52 reveal a DNA repair/recombination complex including Rad51 and Rad52

.

Genes Dev.

7

:

1755

1765

.

Morrow

D M

,

Connelly

C

,

Hieter

P

,

1997

“Break copy” duplication: a model for chromosome fragment formation in Saccharomyces cerevisiae

.

Genetics

147

:

371

382

.

Mortensen

U H

,

Bendixen

C

,

Sunjevaric

I

,

Rothstein

R

,

1996

DNA strand annealing is promoted by the yeast Rad52 protein

.

Proc. Natl. Acad. Sci. USA

93

:

10729

10734

.

Muris

D F

,

Bezzubova

O

,

Buerstedde

J M

,

Vreeken

K

,

Balajee

A S

et al. .,

1994

Cloning of human and mouse genes homologous to RAD52, a yeast gene involved in DNA repair and recombination

.

Mutat. Res.

315

:

295

305

.

Natarajan

S

,

McEachern

M J

,

2002

Recombinational telomere elongation promoted by DNA circles

.

Mol. Cell. Biol.

22

:

4512

4521

.

New

J H

,

Sugiyama

T

,

Zaitseva

E

,

Kowalczykowski

S C

,

1998

Rad52 protein stimulates DNA strand exchange by Rad51 and replication protein A

.

Nature

391

:

407

410

.

Pâques

F

,

Haber

J E

,

1999

Multiple pathways of recombination induced by double-strand breaks in Saccharomyces cerevisiae

.

Microbiol. Mol. Biol. Rev.

63

:

349

404

.

Parsons

C A

,

Baumann

P

,

Van Dyck

E

,

West

S C

,

2000

Precise binding of single-stranded DNA termini by human RAD52 protein

.

EMBO J.

19

:

4175

4181

.

Rattray

A J

,

Symington

L S

,

1994

Use of a chromosomal inverted repeat to demonstrate that the RAD51 and RAD52 genes of Saccharomyces cerevisiae have different roles in mitotic recombination

.

Genetics

138

:

587

595

.

Reid

R J

,

Lisby

M

,

Rothstein

R

,

2002

Cloning-free genome alterations in Saccharomyces cerevisiae using adaptamer-mediated PCR

.

Methods Enzymol.

350

:

258

277

.

Rizki

A

,

Lundblad

V

,

2001

Defects in mismatch repair promote telomerase-independent proliferation

.

Nature

411

:

713

716

.

Roth

C W

,

Kobeski

F

,

Walter

M F

,

Biessmann

H

,

1997

Chromosome end elongation by recombination in the mosquito Anopheles gambiae

.

Mol. Cell. Biol.

17

:

5176

5183

.

Shampay

J

,

Szostak

J W

,

Blackburn

E H

,

1984

DNA sequences of telomeres maintained in yeast

.

Nature

310

:

154

157

.

Shinohara

A

,

Ogawa

T

,

1998

Stimulation by Rad52 of yeast Rad51-mediated recombination

.

Nature

391

:

404

407

.

Shinohara

A

,

Ogawa

H

,

Ogawa

T

,

1992

Rad51 protein involved in repair and recombination in S. cerevisiae is a RecA-like protein

.

Cell

69

:

457

470

.

Shinohara

A

,

Shinohara

M

,

Ohta

T

,

Matsuda

S

,

Ogawa

T

,

1998

Rad52 forms ring structures and co-operates with RPA in single-strand DNA annealing

.

Genes Cells

3

:

145

156

.

Signon

L

,

Malkova

A

,

Naylor

M L

,

Klein

H

,

Haber

J E

,

2001

Genetic requirements for _RAD51_- and _RAD54_-independent break-induced replication repair of a chromosomal double-strand break

.

Mol. Cell. Biol.

21

:

2048

2056

.

Singer

M S

,

Gottschling

D E

,

1994

TLC1: template RNA component of Saccharomyces cerevisiae telomerase

.

Science

266

:

404

409

.

Singleton

M R

,

Wentzell

L M

,

Liu

Y

,

West

S C

,

Wigley

D B

,

2002

Structure of the single-strand annealing domain of human RAD52 protein

.

Proc. Natl. Acad. Sci. USA

99

:

13492

13497

.

Stasiak

A Z

,

Larquet

E

,

Stasiak

A

,

Muller

S

,

Engel

A

et al. .,

2000

The human Rad52 protein exists as a heptameric ring

.

Curr. Biol.

10

:

337

340

.

Sugiyama

T

,

New

J H

,

Kowalczykowski

S C

,

1998

DNA annealing by RAD52 protein is stimulated by specific interaction with the complex of replication protein A and single-stranded DNA

.

Proc. Natl. Acad. Sci. USA

95

:

6049

6054

.

Sung

P

,

1994

Catalysis of ATP-dependent homologous DNA pairing and strand exchange by yeast RAD51 protein

.

Science

265

:

1241

1243

.

Sung

P

,

1997

Function of yeast Rad52 protein as a mediator between replication protein A and the Rad51 recombinase

.

J. Biol. Chem.

272

:

28194

28197

.

Symington

L S

,

2002

Role of RAD52 epistasis group genes in homologous recombination and double-strand break repair

.

Microbiol. Mol. Biol. Rev.

66

:

630

670

.

Teng

S C

,

Zakian

V A

,

1999

Telomere-telomere recombination is an efficient bypass pathway for telomere maintenance in Saccharomyces cerevisiae

.

Mol. Cell. Biol.

19

:

8083

8093

.

Teng

S C

,

Chang

J

,

McCowan

B

,

Zakian

V A

,

2000

Telomerase-independent lengthening of yeast telomeres occurs by an abrupt Rad50p-dependent, Rif-inhibited recombinational process

.

Mol. Cell

6

:

947

952

.

Tsai

Y L

,

Tseng

S F

,

Chang

S H

,

Lin

C C

,

Teng

S C

,

2002

Involvement of replicative polymerases, Tel1p, Mec1p, Cdc13p, and the Ku complex in telomere-telomere recombination

.

Mol. Cell. Biol.

22

:

5679

5687

.

Usui

T

,

Ohta

T

,

Oshiumi

H

,

Tomizawa

J

,

Ogawa

H

et al. .,

1998

Complex formation and functional versatility of Mre11 of budding yeast in recombination

.

Cell

95

:

705

716

.

Van Dyck

E

,

Hajibagheri

N M

,

Stasiak

A

,

West

S C

,

1998

Visualisation of human rad52 protein and its complexes with hRad51 and DNA

.

J. Mol. Biol.

284

:

1027

1038

.

Wach

A

,

Brachat

A

,

Pohlmann

R

,

Philippsen

P

,

1994

New heterologous modules for classical or PCR-based gene disruptions in Saccharomyces cerevisiae

.

Yeast

10

:

1793

1808

.

Author notes

1

Present address: Research Institute for Radiation Biology and Medicine, Hiroshima University, Hiroshima 734-8553, Japan.

© Genetics 2003

Citations

Views

Altmetric

Metrics

Total Views 301

226 Pageviews

75 PDF Downloads

Since 3/1/2021

Month: Total Views:
March 2021 1
April 2021 6
May 2021 11
June 2021 9
August 2021 5
September 2021 4
October 2021 8
November 2021 10
January 2022 5
February 2022 7
March 2022 2
April 2022 9
May 2022 5
June 2022 1
July 2022 10
August 2022 22
September 2022 9
October 2022 2
November 2022 1
December 2022 6
January 2023 3
February 2023 4
March 2023 10
April 2023 4
May 2023 4
June 2023 3
July 2023 6
August 2023 12
September 2023 3
October 2023 7
November 2023 6
December 2023 8
January 2024 9
February 2024 8
March 2024 6
April 2024 15
May 2024 14
June 2024 10
July 2024 17
August 2024 7
September 2024 11
October 2024 1

×

Email alerts

Citing articles via

More from Oxford Academic