<<

Commentary 2107 cohesion – rings, knots, orcs and fellowship

Laura A. Díaz-Martínez1,*, Juan F. Giménez-Abián2 and Duncan J. Clarke3 1Department of Pharmacology, UT-Southwestern Medical Center, 6001 Forest Park Rd, Dallas, TX75390, USA 2Proliferación Celular, CSIC, Ramiro de Maeztu 9, 28040-Madrid, Spain 3Department of Genetics, Cell Biology and Development, University of Minnesota, 420 Washington Avenue SE, Minneapolis, MN55455, USA *Author for correspondence (e-mail: [email protected])

Accepted 6 May 2008 Journal of Cell Science 121, 2107-2114 Published by The Company of Biologists 2008 doi:10.1242/jcs.029132

Summary Sister- cohesion is essential for accurate chromosome the onset of allows sister-chromatid separation. segregation. A key discovery towards our understanding of However, recent studies have revealed activities that provide sister-chromatid cohesion was made 10 years ago with the cohesion in the absence of . Here we review these identification of . Since then, cohesins have been shown advances and propose an integrative model in which to be involved in cohesion in numerous organisms, from yeast to chromatid cohesion is a result of the combined activities of mammals. Studies of the composition, regulation and structure multiple cohesion mechanisms. of the cohesin complex led to a model in which cohesin loading during S-phase establishes cohesion, and cohesin cleavage at Key words: Cohesin, Chromosome cohesion, Catenations

Introduction et al., 1997), as measured using the LacO/LacI-GFP system Accurate is the crux of – failure (Straight et al., 1996). A cohered locus is normally visualized as a causes genetic disorders, spontaneous abortions and cancer (Kops discrete fluorescent spot in early mitosis, but this was observed to et al., 2005; Rubio et al., 2005; Shah et al., 2003; Weaver and prematurely separate into two discrete spots in cohesin mutants. Cleveland, 2006; Weaver et al., 2007). To allow accurate Second, the association of cohesin with DNA mirrors the cohesion segregation, sister must remain cohered from their cycle: cohesin is bound to from S-phase, when inception, which occurs in S-phase, until anaphase (Fig. 1). Sister- cohesion is established, until anaphase, when cohesion is abolished chromatid cohesion has been proposed to depend on a group of (Michaelis et al., 1997). Other experiments indicate that cleavage proteins called cohesins, which were first identified in yeast (Guacci of Mcd1 is necessary for loss of cohesion. cannot

Journal of Cell Science et al., 1997; Michaelis et al., 1997), that form a tetra-subunit complex separate in yeast cells that express non-cleavable Mcd1 that is of Mcd1 (also known as Rad21 in Schizosaccharomyces pombe and resistant to Esp1 () (Uhlmann et al., 1999). Conversely, Scc1 in ), Smc1, Smc3 and Scc3 (also artificial Mcd1 cleavage (not by Esp1) leads to premature sister- known as SA1 and SA2 in mammals). This ‘cohesin complex’ can chromatid separation (Uhlmann et al., 2000). Together, these studies form a 35-nm ring that tethers sister duplexes. Two main models form the foundation of a compelling model in which cohesion have been proposed to explain how this ring provides cohesion: the depends on the cohesin complex and cleavage of the ring results ring might entrap both duplexes (Haering et al., 2002; Nasmyth in loss of cohesion. and Haering, 2005) or two rings might interact, each one entrapping Here we review the role of cohesin in sister-chromatid cohesion a duplex (Anderson et al., 2002; Guacci, 2007; Huang et al., 2005). and discuss recent discoveries of cohesin-independent mechanisms Conserved orthologs of the cohesins are found in metazoans of cohesion. We will not, however, discuss recent evidence (Darwiche et al., 1999; Losada et al., 1998; Rollins et al., 1999; indicating that the regulation of the onset of anaphase by the spindle- Sumara et al., 2000), indicating that cohesin might be the universal assembly checkpoint is more complex than had been previously mechanism of sister-chromatid cohesion. Accessory proteins appreciated (Chestukhin et al., 2003; Gimenez-Abian et al., 2005a; regulate the loading of cohesin onto chromosomes (reviewed in Gimenez-Abian et al., 2005b; Papi et al., 2005; Zur and Brandeis, Huang and Laurent, 2004; Lee and Orr-Weaver, 2001; Nasmyth 2001) because this topic has been reviewed in detail elsewhere and Schleiffer, 2004; Riedel et al., 2004; Skibbens, 2005; Uhlmann, (Clarke et al., 2005). In this Commentary, we propose a model in 2004) and a protease, separase, cleaves Mcd1, presumably leading which cohesion depends on multiple mechanisms that collaborate to sister-chromatid separation (Ciosk et al., 1998; Uhlmann et al., to ensure cohesion along the length of the whole chromosome. 1999; Uhlmann et al., 2000). Thus, it is thought that cohesin-loading factors establish cohesion and separase triggers anaphase. Cohesion without cohesin in yeast The first evidence that protein complexes provide cohesion came Although it has been widely documented in budding yeast that from studies of circular minichromosomes in budding yeast; these cohesins can provide cohesion, the penetrance of the loss-of- minichromosomes remain cohered in the absence of DNA cohesion phenotype in cohesin mutants depends on the locus that intertwinings (Guacci et al., 1994; Koshland and Hartwell, 1987). is observed (Antoniacci and Skibbens, 2006; Baetz et al., 2004; Two lines of investigation have indicated that cohesin provides this Ciosk et al., 2000; D’Amours et al., 2004; Guacci et al., 1997; activity. First, mutant cohesins have been found to cause an Lam et al., 2006; Mayer et al., 2001; Michaelis et al., 1997; Strom increased distance between sister loci (Guacci et al., 1997; Michaelis et al., 2007; Sullivan et al., 2004; Suter et al., 2004; Toth et al., 2108 Journal of Cell Science 121 (13)

2007; Losada et al., 2005; Rankin et al., 2005; Toyoda and Yanagida, 2006; Watrin et al., 2006) cells. Typically, increased distance between sister chromatids was observed in these experiments (Losada et al., 2005; Toyoda and Yanagida, 2006), but none of these approaches resulted in complete separation of the ion chromatids. Briefly, Losada and co-workers observed that ~30% of Cohes Chromosome Accurate biorientation segregation mitotic chromosomes that were assembled in cohesin-depleted No cohes Xenopus egg extracts had unpaired regions and that the distance ion between sister chromatids was increased, but without complete separation (Losada et al., 2005). These findings were corroborated by Kenney and Heald with the additional important observation that separation in cohesin-depleted Xenopus egg extracts is promoted only upon bipolar attachment to the mitotic spindle

No biorientation (Kenney and Heald, 2006). It could be argued that these depleted extracts have small amounts of cohesin, which are sufficient for Fig. 1. Sister-chromatid cohesion is required for accurate chromosome centromere cohesion. However, Kenney and Heald simultaneously segregation. During a mitotic , the genome is duplicated in S phase depleted two cohesin subunits almost to completion. and each identical copy is then segregated into the daughter cells. In These data partly agree with phenotypes that have been observed , this process is complex owing to the fragmentation of the genome in Scc1–/–/– chicken cells (the chicken Scc1 gene is located on in several chromosomes. Eukaryotic cells have evolved a mechanism, termed sister-chromatid cohesion, that keeps the two copies of a chromosome (sister chromosome 2, which is trisomic in DT40 cells), which exhibit an chromatids) together from the moment of duplication to the onset of anaphase. increased distance between, but retained association of, the sister This mechanism ensures the accurate segregation of one and only one copy of chromatids (Sonoda et al., 2001). Unlike in the Xenopus egg extracts, each chromosome to each daughter cell. When sister-chromatid cohesion is some cohesion was retained even after chromosomes attached to defective, mitotic processes such as chromosome biorientation and chromosome segregation are disrupted, resulting in aneuploidy, a hallmark of the spindle because the Scc1-depleted chicken cells aligned most most cancers. of their chromosomes into a plate. Consistent with the Xenopus data, chromosome segregation was abnormal (Sonoda et al., 2001; Vagnarelli et al., 2004). Chromosome congression to a metaphase plate followed by aberrant chromosome segregation 1999) (reviewed in Koshland and Guacci, 2000; Nasmyth and were also observed in cohesin-depleted Drosophila cells (Vass et al., Schleiffer, 2004; Skibbens, 2000). Only telomeres lose cohesion in 2003). Cells with separated sister chromatids and aberrant anaphases 100% of cohesin-mutant cells, whereas loci at the chromosome arms were observed after Rad21 knockdown but not after depletion of and pericentromere remain cohered in 40-75% of cells and the stromal antigen (SA), another cohesin subunit. SA-depleted cells ribosomal DNA (rDNA) locus remains cohered in the majority of had cohered sister chromatids (although the distance between the cohesin-mutant cells (Fig. 2). This indicates that cohesin is sufficient two chromatids was slightly increased) and were able to perform

Journal of Cell Science to provide cohesion in yeast, because it seems to be the only anaphase normally (Vass et al., 2003). mechanism that operates at telomeres, but also that cohesin- In human cells, the specific conditions in which cohesin independent mechanisms might contribute to cohesion at other loci. depletions have been performed have led to different interpretations In a scenario in which cohesins were the sole mechanism of the data. Here we describe our attempt to unify these observations. providing cohesion, non-functional cohesin or the absence of a Several reports describe premature loss of cohesion in 40-80% of cohesin subunit should lead to loss of cohesion upon replication of Rad21-depleted mitotic cells (Losada et al., 2005; Rankin et al., any locus. In studies that timed the loss of cohesion in mcd1 mutants, 2005; Toyoda and Yanagida, 2006; Watrin et al., 2006). In all of some cells did lose cohesion as soon as DNA replication was these studies, following Rad21 depletion using RNAi in observed (Noble et al., 2006). However, despite synchronous DNA asynchronous cultures, mitotic cells were isolated and sister- replication within populations of yeast cells in such experiments, chromatid separation scored as a percentage of mitotic cells (rather loss of cohesion typically occurred asynchronously in mcd1 mutants than as a percentage of total cells). In our analyses of cohesin- after the release from G1 arrest. The maximum percentage of mcd1 depleted cells, using the same small interfering (si)RNAs for RNAi, cells that displayed a loss-of-cohesion phenotype was only achieved we observed that ~4% of the total cells both arrested in mitosis and after all the cells reached G2 (Michaelis et al., 1997; Noble et al., had separated sister chromatids (Diaz-Martinez et al., 2007). This 2006). The fact that loss of cohesion in cohesin-defective cells can small percentage of the total cells corresponds to the 40-80% of be delayed for some time after replication suggests that other the mitotic cells described above. mechanisms of cohesion exist. The key issues thus become identifying the source of these arrested mitotic cells that have separated sister chromatids and Cohesin in other eukaryotes determining at which point cohesion was lost. By examining Cohesins are conserved from yeast to humans, which suggests that complete cell cycles in synchronized cells, we observed that ~100% their role in cohesion is conserved. Efforts to directly assess the of cohesin-depleted cells reach mitosis with normally cohered sister role of cohesin in sister-chromatid cohesion in higher eukaryotes chromatids and that all of these are able to form a metaphase plate include the depletion of cohesins from Xenopus laevis egg extracts (Diaz-Martinez et al., 2007). Some cells appeared to lose cohesion (Kenney and Heald, 2006; Losada et al., 1998), an inducible knock- asynchronously, usually after a prolonged metaphase delay, and out of Rad21 in chicken cells (Sonoda et al., 2001), and the use of some of these cells then arrested in mitosis (Diaz-Martinez et al., RNA interference (RNAi) to deplete cohesins from Drosophila 2007). These arrested cells presumably account for the previous melanogaster (Vass et al., 2003) and human (Diaz-Martinez et al., observations of mitotic cells with separated sister chromatids that Multiple cohesion mechanisms 2109

remains unclear, but one possibility is that cleavage of Rad21 has Telomere separation: a signaling role that serves to inactivate the spindle-assembly 1 86% mcd1-1 (NOC) checkpoint, triggering coordinated and efficient loss of cohesion. 0 k b

3 BMH1-locus separation: Upon checkpoint silencing, multiple pathways that drive 24% mcd1-1 (NOC)2 synchronous anaphase might become active. Consistent with the observation that most cohesin-depleted VTC3-locus separation: human cells retain centromere cohesion until metaphase or 56% mcd1-1 (NOC)2 anaphase, the frequency of cells with separated sister chromatids Centromere separation: following RNAi of cohesin dramatically diminishes to less than

~50% mcd1-1 (NOC)3 20% of mitotic cells after 3 hours in colcemid (Inoue et al., 2007) 2

53% mcd1-1 (NOC) and 4% after 16 hours (Diaz-Martinez et al., 2007). By contrast, 40 k b

k b the percentage of cells with separated sister chromatids that is 3 2 URA3-locus separation: observed after a long mitotic arrest in the presence of nocodazole <5% wild type (NOC)4 20% wild type (cdc16)4 remains high (>65% after 16 hours) in cells that have been depleted 25% wild type (NOC)5 of the centromere-guardian Shugoshin (Sgo1). Therefore, Sgo1, a 15% scc1-73 (NOC)6 protein that is thought to protect centromeric cohesin from removal, 2 5 k b 40% mcd1-1 (NOC) might regulate cohesin and additional cohesion factors at the 3 40% smc1-259 (cdc20)7 centromere. Together, these results indicate that sister-chromatid b 40% scc1-73 (NOC)8 cohesion can be maintained in a manner that is independent of 50% scc1-73 (NOC)4 129 k cohesin, but that cohesin is needed for faithful anaphase. Thus, 60% smc1-259 (cdc16)4 80% scc1-73 (cdc16)4 studies in yeast cells and higher eukaryotes have left open the

400 k b 400 90% scc1-73 (cdc20)9 possibility that cohesin and additional factors collaborate to ensure the association between sister chromatids until anaphase. HOR2-locus separation: 27% mcd1-1 (NOC)2 Cohesin-independent cohesion DNA catenations rDNA-locus separation: 57% mcd1-1 (cdc20) 2 DNA catenation (the physical intertwining of sister chromatids) was <10% mcd1-1 (NOC)10 the first mechanism of cohesion to be formally proposed (Murray <10% scc1-73 (cdc20)11 and Szostak, 1985). Catenations are a byproduct of semi- conservative replication (Sundin and Varshavsky, 1980) and Fig. 2. Contribution of cohesin to cohesion in budding yeast depends on the physically couple replication to cohesion – one being the inevitable locus involved. A summary of studies in which a loss of cohesion was consequence of the other (Fig. 3). By contrast, cohesin-mediated observed at different loci in cohesin mutants is shown. These results indicate cohesion is biochemically coupled to replication, because genomic that the specific contribution of the cohesin complex to the cohesion of any DNA can be replicated to completion in the absence of cohesin. two sister-chromatid loci is locus dependent. All the data refer to the loss of Journal of Cell Science cohesion that has been measured in G2-M cells and the methods that were Logically, cells probably co-evolved their mechanisms of genome used to arrest the cells in mitosis are shown in brackets. Note that cdc16 and replication and sister-chromatid cohesion. As a mechanism of cdc20 cells arrest in mitosis in the presence of a functional spindle, whereas cohesion, catenation fulfils the expected physical dependency on nocodazole (NOC) disrupts the spindle. The chromosome depicted in the replication, whereas cohesin-mediated cohesion does not. figure is not drawn to scale and does not represent a particular yeast The removal of catenations is performed by chromosome: the loci that are listed belong to different chromosomes. The α colors reflect the level of penetrance of the loss-of-cohesion phenotype when type II enzymes (Top2 in budding yeast and topoisomerase II cohesin mutations are present. Loci listed in red boxes show more loss of and IIβ in higher eukaryotes), the activity of which is required for cohesion, meaning that cohesion at these loci depends mostly on the cohesin chromosome individualization (Gimenez-Abian et al., 2000), sister- complex. Loci in the green boxes show an ~50% loss of cohesion, suggesting chromatid resolution (Gimenez-Abian et al., 1995) and chromosome that cohesin partially contributes to cohesion at these loci. Loci in the blue boxes show a very low loss of cohesion after cohesin mutation, meaning that segregation (Clarke et al., 1993; Holm et al., 1985; Holm et al., chromosome cohesion at these loci depends mainly on mechanisms other than 1989; Sakaguchi and Kikuchi, 2004; Uemura et al., 1987; Uemura cohesin. Studies referenced in this schematic: 1(Antoniacci and Skibbens, and Yanagida, 1984). That catenations can maintain an association 2006), 2(Lam et al., 2006), 3(Guacci et al., 1997), 4(Michaelis et al., 1997), between chromatids is evident from the failed sister-chromatid 5 6 7 8 (Mayer et al., 2001), (Suter et al., 2004), (Strom et al., 2007), (Toth et al., separation, which leads to a cut (cells untimely torn) phenotype in 1999), 9(Ciosk et al., 2000), 10(D’Amours et al., 2004), 11(Sullivan et al., 2004). the absence of Top2 (Holm et al., 1985; Uemura et al., 1987). What appear to be catenations that link sister-chromatid are present in human cells that are arrested in mitosis by are seen after depletion of cohesin (Losada et al., 2005; Rankin poisons (Bickmore and Oghene, 1996). Furthermore, some evidence et al., 2005; Toyoda and Yanagida, 2006; Watrin et al., 2006). indicates that catenation-mediated cohesion is sufficient for cohesion However, because such cells first form metaphase plates before independent of cohesin (Toyoda and Yanagida, 2006; Vagnarelli sister-chromatid separation (Diaz-Martinez et al., 2007), we suggest et al., 2004; Diaz-Martinez et al., 2006), and it has been suggested that this phenotype is not the result of a complete failure in the that the removal of both cohesin and catenations contributes to establishment of cohesion, but is due to uncoordinated loss of regulate sister-chromatid separation (Diaz-Martinez et al., 2006; cohesion during anaphase. We propose that cohesin accounts for Kenney and Heald, 2006; Toyoda and Yanagida, 2006). synchronous sister-chromatid separation at anaphase but is not The requirement of topoisomerase II at different stages of the required for centromere cohesion in early mitosis. The mechanism chromosome cycle is presumably why it has been an effective target by which cohesin promotes synchronous sister-chromatid separation of anti-cancer drugs (Nelson et al., 1984; Tewey et al., 1984). 2110 Journal of Cell Science 121 (13)

regulate cohesion solely via Top2. Smt4 is required for cohesion in the context of a circular minichromosome that lacks catenations, Semi- indicating that there is another Smt4 target that contributes to conservative cohesion. replication of DNA Interestingly, Smt4-dependent cohesion shares three important features with cohesin-mediated cohesion. First, as observed in cohesin mutants, cohesion is lost only in a fraction of smt4 cells. Catenated Second, similar to cohesin-mediated cohesion, Smt4-dependent sister cohesion varies from locus to locus (Bachant et al., 2002). Consistent chromatids with a sufficiency for cohesin-dependent cohesion at telomeres (Antoniacci and Skibbens, 2006), smt4 mutants do not show loss Decatenation of cohesion at telomeres (Bachant et al., 2002), whereas cohesion by topo- isomerase II at the URA3 locus is partially lost both in smt4 mutants (Bachant et al., 2002) and in cohesin mutants (Weitzer et al., 2003). These results suggest that these two mechanisms of cohesion have additive Decatenated roles in sister-chromatid cohesion. A systematic analysis to sister understand the specific contributions of each cohesion mechanism chromatids at different loci is needed. Third, the separation that is observed after smt4 or cohesin inactivation occurs after the S-phase is Fig. 3. The formation of catenations and their resolution by topoisomerase II. completed, at least in a subset of cells. These results suggest that Catenations are a byproduct of DNA replication and they form at the points at sister-chromatid cohesion at a specific locus is the result of a which two replication forks collide. Removal of catenations is performed by complex balance between different cohesion mechanisms – absence type-II , which produce a double-strand break in one of of any one of these mechanisms might lead to weakened cohesion the chromatids and pass the other through the break. Two resolved sister that, although able to initially provide association and perhaps chromatids are the result of the strand-passing process and re-ligation reaction. provide cohesion in most of the cells during an unperturbed mitosis, is not sufficient to maintain pairing after a prolonged mitotic arrest. However, the idea that there is a progressive removal of catenations Studies to determine whether this is true, and to test the specific during the cell cycle led to the idea that, although catenations remain contribution of each cohesion mechanism at different loci, are at centromeres until anaphase, they are maintained as a consequence needed. of cohesion and their removal is not regulated. This hypothesis is A cohesion-related role for sumoylation of topoisomerase II has supported by the dispensability of catenations for minichromosome also been revealed in higher eukaryotes. Depletion of the SUMO E3- segregation in yeast (Guacci et al., 1994; Koshland and Hartwell, ligase PIASγ from human cells leads to prolonged metaphase arrest 1987). However, as pointed out by Guacci et al. (Guacci et al., 1994), and lack of enrichment of topoisomerase IIα at centromeres (Diaz- these data do not rule out a function of catenation in the Martinez et al., 2006). Xenopus PIASγ is required for the formation

Journal of Cell Science establishment of cohesion. Moreover, minichromosomes are of a unique SUMO focus at the inner centromere (Azuma et al., 2005) segregated 100-fold less efficiently than endogenous yeast and is responsible for topoisomerase-IIα sumoylation. Disrupting chromosomes (10–3 versus 10–5 errors per ) (Koshland PIASγ activity in egg extracts leads to failed chromosome segregation and Hartwell, 1987). Whether this decrease in the faithfulness of (Azuma et al., 2005). Topoisomerase IIα is sumoylated in human chromosome segregation is due to reduced cohesion in the absence cells (Mao et al., 2003) and evidence suggests that PIASγ is the of catenations remains unknown. relevant E3 ligase (Diaz-Martinez et al., 2006; Agostinho et al., 2008). The view that decatenation is unregulated has begun to change Possible redundancy or cooperation between SUMO ligases is in light of studies indicating that topoisomerase IIα is regulated suggested by the dispensability of PIASγ in mice (Wong et al., 2004), by sumoylation during mitosis in yeast (Bachant et al., 2002; which in turn rely on the SUMO-ligase activity of RanBP2 for Takahashi et al., 2006) (reviewed in Porter and Farr, 2004), Xenopus topoisomerase-II enrichment at the centromere (Dawlaty et al., (Azuma et al., 2005; Azuma et al., 2003), mice (Dawlaty et al., 2008). Interestingly, regardless of the particular SUMO ligase that is 2008) and human (Diaz-Martinez et al., 2006) cells. In budding involved, sumoylation of topoisomerase II is required for its yeast, mutation of the SUMO-isopeptidase Smt4 leads to premature enrichment at the centromere and defects in this process can result sister-chromatid separation. This cohesion defect is probably caused in failure of sister-chromatid segregation (Azuma et al., 2005; Diaz- by defects in the regulation of decatenation, because Top2 Martinez et al., 2006; Dawlaty et al., 2008). Furthermore, PIASγ sumoylation was increased in smt4 mutants, a non-sumoylatable most probably regulates a cohesin-independent form of cohesion, top2-SNM mutant rescued the cohesion defect of smt4 mutants because the centromeres of human cells simultaneously depleted of (Bachant et al., 2002) and constitutive Top2 sumoylation induced PIASγ and the cohesin-protector Sgo1 remain cohered but lack enrichment of Top2 at pericentromeric regions (Takahashi et al., cohesin at the centromere (Diaz-Martinez et al., 2006). 2006). These results suggest that sumoylation increases Top2 These data have revealed a mechanism that specifically regulates activity at the centromeres by promoting its recruitment to these topoisomerase IIα at the centromere during mitosis in metazoans. loci, which leads to premature decatenation. Thus, centromeric catenations might be actively maintained until That Smt4-dependent regulation of cohesion is independent of anaphase, in a checkpoint-dependent manner, by modulation of cohesin is indicated by the unchanged distribution of cohesin in the activities of SUMO ligases. However, in addition to smt4 mutants and the sister-chromatid separation observed in these topoisomerase II, other proteins that have cohesion-related mutants in the absence of separase (Bachant et al., 2002). Despite functions, such as Pds5 (Stead et al., 2003) and Ycs4 (D’Amours these intriguing data, other facts indicate that Smt4 is unlikely to et al., 2004), are known to be sumoylated. Therefore, the SUMO Multiple cohesion mechanisms 2111

PP2A

CK2 P SUMO PIAS␥

Catenated sister chromatids Inter-chromatid catenations Topoisomerase II␣

PP2A

Plk1 ? SUMO Pds5 Aurora-B P SA1/2 Haspin Rad21 Separase mc1 S

mc3 S Inter-chromatid Sister cohesin chromatids

Cohesin complex

Fig. 4. Multiple mechanisms of cohesion. Cohesion between any sister-chromatid loci is the result of cooperation between at least two mechanisms of cohesion: DNA catenations (top) and cohesin (bottom). On one hand, the cohesin complex – formed by Smc1, Smc3, Mcd1/Rad21/Scc1 and Scc3/SA1/SA2 – is thought to tether the two sister chromatids together by physical entrapment. DNA catenations, on the other hand, provide sister-chromatid cohesion by the intertwinement of the two chromatids. The contribution of each mechanism at a specific locus might be influenced by factors such as the spacing of catenations, the location of cohesin-binding regions, structure and changes in chromatid cohesion that are induced after DNA replication (e.g. DNA damage or de novo cohesin loading) (Kim et al., 2002; Nagao et al., 2004; Potts et al., 2006; Sjogren and Nasmyth, 2001; Strom et al., 2004; Strom et al., 2007; Unal et al., 2007). The existence of different cohesion mechanisms is advantageous because it allows differential regulation of cohesion at specific chromosome regions. Both cohesin and catenations are subject to complex regulatory mechanisms and have to be concertedly removed during anaphase, possibly by post-translational modifications such as phosphorylation and sumoylation. Some of these regulatory mechanisms are depicted here.

Journal of Cell Science ligases have the potential to drive different aspects of sister- occurred in ~30% of cells. Interestingly, orc2 mcd1 and orc5 mcd1 chromatid separation and perhaps direct the concerted dissolution of mutants showed an additive defect (Shimada and Gasser, 2007), different cohesion mechanisms (Fig. 4). Further understanding suggesting that ORC and cohesin act in parallel to provide cohesion. of these enzymes will clarify their roles in sister-chromatid cohesion. ORC-mediated cohesion can be dynamically re-established in G2-M after loss of cohesion has occurred (Shimada and Gasser, ORC and 2007), but this cohesion depends on cohesin, suggesting that some During DNA replication, the cohesin loading factors interact with degree of sister-chromatid association is required. This is consistent components of the replication machinery such as PCNA and RFC with a model in which tight sister-chromatid association depends (reviewed in Guacci, 2007; Skibbens et al., 2007); thus, components on cooperative activities. of the replication fork might activate these loading mechanisms. In A role of ORC in mitosis seems to be conserved in metazoans, addition, replication proteins might have a direct role in cohesion. because Orc2 and Orc5 mutants in Drosophila cells (Pflumm and The origin recognition complex (ORC) comprises six subunits, Botchan, 2001), as well as Orc2- or Orc6-depleted human cells marks replication origins in eukaryotic cells and facilitates the (Prasanth et al., 2004; Prasanth et al., 2002), have mitotic defects binding of other replication proteins (Quintana and Dutta, 1999). that increase ploidy, induce mitotic arrest and spindle abnormalities, ORC mutants in budding yeast arrest in mitosis with a normal 2C and affect chromosome congression and condensation (Pflumm and DNA content and show an increased loss of centromeric plasmids Botchan, 2001; Prasanth et al., 2004; Prasanth et al., 2002), which (Dillin and Rine, 1998; Loo et al., 1995). Genetic interactions is consistent with the localization of ORC to centromeres and between ORC genes and MCD1 have been described in budding centrosomes during mitosis (Prasanth et al., 2004; Prasanth et al., yeast, and mutations in orc5 enhance the cohesion defect of mcd1 2002). However, none of these studies has yet described defects in (scc1-73) cells (Suter et al., 2004). Cells that are depleted of Orc2 cohesion (Pflumm and Botchan, 2001; Prasanth et al., 2004; lose cohesion even though the binding of cohesin to chromosomes Prasanth et al., 2002). does not change. Increased targeting of ORC to the URA3 locus Recent studies have also linked the condensin complex decreased the loss of cohesion that was induced by MCD1 mutation with cohesion. Drosophila mutants of a condensin subunit (Shimada and Gasser, 2007). Thus, ORC-mediated cohesion is (Cap-G) have cohesion defects at centromeres, whereas the sufficient for cohesion when ORC is present in high quantities, and chromosome arms remain cohered even in anaphase, which it does not require cohesin. However, the loss of cohesion after causes chromosome missegregation (Dej et al., 2004). In yeast, Orc2 depletion, similar to that seen after cohesin depletion, only mutations in condensin subunits have been associated with loss 2112 Journal of Cell Science 121 (13)

of cohesion (Lam et al., 2006; Vas et al., 2007) as well as with As is evident from the yeast data, the contribution of different lack of separation at the rDNA locus (D’Amours et al., 2004; cohesion mechanisms varies from locus to locus, perhaps being Freeman et al., 2000; Strunnikov et al., 1995). Whether this is influenced by the locations in which catenations arise during due to independent roles of condensin, namely cohesion and replication, and the sites of cohesin, condensin and ORC binding, condensation, or two manifestations of the same activity that lead as well as the surrounding chromatin environment. to opposite effects, perhaps owing to the special structure of the rDNA locus, remains to be tested. The contribution of condensin Different needs, different solutions to cohesion, similar to the contribution of cohesin, varies from The existence of collaborative mechanisms of cohesion makes the locus to locus and the effects are additive in a condensin-cohesin need for multiple pathways that inactivate cohesion in anaphase double mutant (Lam et al., 2006). The possibility that condensin obligatory. Two events that are necessary for sister-chromatid maintains catenation-mediated cohesion remains to be tested, but separation, namely cohesin removal (promoted by separase and the condensin has been shown to be involved in recruiting kinases of the pathway) and decatenation (promoted by topoisomerase II to the chromosome scaffold (Coelho et al., 2003). topoisomerase II), have been widely documented, but further This suggests a link between condensin and catenations. research is needed to understand the mechanism that regulates their Alternatively, the role of condensin in cohesion might be a concerted activity in anaphase. Two types of mechanism for secondary consequence of its role in condensation. This might be the concerted removal of multiple cohesion devices can be particularly important in regions that have complex chromatin envisioned: a parallel and independent activation of all cohesion- structures such as centromeres and rDNA. removal mechanisms upon the generation of a single ‘go’ signal, or a network-type regulation in which activation of one separation Multiple mechanisms of cohesion mechanism influences the activity of the others. It is also possible An integrative model of cohesion that a mixture of these two mechanisms exists so that a single ‘go’ Based on the data discussed above, we propose a model in which signal serves as the initial activator of some or all of these pathways, cohesion is the result of a complex balance and coordination which, in turn, regulate the other pathways. Interestingly, recent between several mechanisms of cohesion. Two main mechanisms, research has shown that some cohesion mechanisms might share a provided by DNA catenations and cohesin, result in cohesion that common regulatory theme: sumoylation of both topoisomerase II is distributed along the whole chromosome in all eukaryotes and the cohesin-associated protein Pds5 have been shown to that have been studied (Fig. 4). However, these mechanisms are regulate cohesion (Azuma et al., 2005; Azuma et al., 2003; Bachant subject to differential regulation at centromeres, arms, telomeres et al., 2002; Diaz-Martinez et al., 2006; Stead et al., 2003). These and perhaps other loci. Both centromeric cohesin and catenations results suggest that SUMO-related processes might be responsible are maintained in a spindle-checkpoint-dependent manner in higher for tipping the balance in favor of loss of cohesion during anaphase. eukaryotes. Similarly, cohesin and catenations are removed from The existence of multiple mechanisms of cohesion creates the chromosomes coordinately in anaphase – complete and efficient difficulty of ensuring their concerted deactivation to allow complete removal of both of these cohesion systems is required for the sister-chromatid segregation during mitosis. However, it can also accurate segregation of chromosomes. explain the ability of cells to provide differential cohesion dynamics

Journal of Cell Science We propose that catenations are the default or primary within a chromosome. Examples of the locus-specific timing of mechanism of cohesion because they are physically coupled to separation exist in nature, such as the centromere-breathing replication. Cohesin is then loaded onto replicated (already phenomenon (the brief and reversible separation of centromeres catenated) sister chromatids. This process occurs concurrently with upon attachment to the spindle) and the late separation of the rDNA replication, because cohesin-loading mechanisms are active during locus in budding yeast. Separation of the rDNA locus seems to be S phase. But, we argue that the establishment of cohesin-mediated a clear example of specialized cohesion because cohesion at this cohesion in S phase might be a dynamic process. The observation locus is independent of cohesin and its separation depends on a that activation of cohesin-loading mechanisms during G2-M after functional condensin complex (D’Amours et al., 2004). Although a single double-strand break leads to genome-wide de novo parallel separation (at centromeres and chromosome arms) of the establishment of cohesion (Strom et al., 2007; Unal et al., 2007) sister chromatids during anaphase is the norm in human cells during raises the possibility that a similar mechanism operates during S undisturbed mitoses (Gimenez-Abian et al., 2005b), ~4% of HeLa phase. In this scenario, activation of cohesin loading would lead cells separate the centromeres before the chromosome arms to loading of cohesin at all loci. Some of these cohesin complexes (Gimenez-Abian et al., 2005b). This suggests that the mechanisms will land in unreplicated regions and will later be removed to allow of chromosome-arm and centromere separation, although normally passage of the replication machinery. Once that region is replicated acting in concert, can be uncoupled. This differential regulation is and held close to its sister chromatid (perhaps aided by perhaps more dramatically revealed by the separation of catenations), cohesin can be loaded again, this time providing chromosome arms but not centromeres during prolonged arrest in cohesion. It can be envisioned that this genome-wide loading of the presence of microtubule poisons (Gimenez-Abian et al., 2004). cohesin is a dynamic event that occurs throughout S phase. The observations that expression of non-cleavable cohesin or The alternative model is that cohesin loading is restricted to the depletion of separase results in failure to separate the chromosome replication forks. arms but that centromeres can separate (Diaz-Martinez et al., 2007; In addition to catenations and cohesin, other mechanisms of Gimenez-Abian et al., 2005a; Gimenez-Abian et al., 2005b; Papi cohesion that are contributed to by ORC and condensin might exist. et al., 2005; Yalon et al., 2004) suggest that cohesin-mediated However, further research is needed to determine whether these are cohesion, although present at the centromere, might be primarily bona fide cohesion factors or whether they are regulators of cohesin responsible for the cohesion of chromosome arms, whereas and catenations. ORC and condensin could also act by affecting centromeric cohesion could be mediated by a cohesin-independent cohesion indirectly – by providing specific chromatin structures. mechanism, both factors being synchronously removed in anaphase. Multiple cohesion mechanisms 2113

It is noteworthy that a step-wise loss of cohesion is observed during centromeres requires RanBP2-mediated SUMOylation of topoisomerase IIalpha. Cell 133, 103-115. , with cohesion at the chromosome arms being lost Dej, K. J., Ahn, C. and Orr-Weaver, T. L. (2004). Mutations in the Drosophila condensin during anaphase I, allowing the separation of homologous subunit dCAP-G: defining the role of condensin for chromosome condensation in mitosis chromosomes, followed by the dissolution of centromeric cohesion and gene expression in interphase. Genetics 168, 895-906. Diaz-Martinez, L. A., Gimenez-Abian, J. F., Azuma, Y., Guacci, V., Gimenez-Martin, during anaphase II to separate sister chromatids. The existence of G., Lanier, L. M. and Clarke, D. J. (2006). PIASgamma is required for faithful two mechanisms of cohesion that are differentially regulated at the chromosome segregation in human cells. PLoS ONE 1, e53. arms and centromeres could account for this step-wise regulation Diaz-Martinez, L. A., Gimenez-Abian, J. F. and Clarke, D. J. (2007). Cohesin is dispensable for centromere cohesion in human cells. PLoS ONE 2, e318. of cohesion during meiosis. Dillin, A. and Rine, J. (1998). Roles for ORC in M phase and S phase. Science 279, 1733- 1737. Freeman, L., Aragon, A. L. and Strunnikov, A. (2000). The condensin complex governs Conclusions chromosome condensation and mitotic transmission of rDNA. J. Cell Biol. 149, 811- A complex picture has emerged in which diverse mechanisms 824. collaborate to provide cohesion along the chromosome. These Gimenez-Abian, J. F., Clarke, D. J., Mullinger, A. M., Downes, C. S. and Johnson, R. T. (1995). A postprophase topoisomerase II-dependent chromatid core separation step mechanisms of cohesion include, but might not be limited to, in the formation of metaphase chromosomes. J. Cell Biol. 131, 7-17. cohesin and DNA catenations. Renewed effort to understand the Gimenez-Abian, J. F., Clarke, D. J., Devlin, J., Gimenez-Abian, M. I., De la Torre, nature, regulation and complex interactions of the multiple C., Johnson, R. T., Mullinger, A. M. and Downes, C. S. (2000). Premitotic chromosome individualization in mammalian cells depends on topoisomerase II activity. Chromosoma mechanisms of cohesion is needed to attain a comprehensive 109, 235-244. knowledge of these systems. Gimenez-Abian, J. F., Sumara, I., Hirota, T., Hauf, S., Gerlich, D., de la Torre, C., Ellenberg, J. and Peters, J. M. (2004). Regulation of sister chromatid cohesion between chromosome arms. Curr. Biol. 14, 1187-1193. We thank V. Guacci, H. Yu, J. Bachant, P. R. Potts and M. Potts for Gimenez-Abian, J. F., Diaz-Martinez, L. A., Waizenegger, I. C., Gimenez-Martin, G. critical reading of this manuscript, and members of the Clarke and Yu and Clarke, D. J. (2005a). Separase is required at multiple pre-anaphase cell cycle labs for discussions. L.A.D.-M. was partially funded by CONACyT stages in human cells. Cell Cycle 4, 1576-1584. and BC043245 from the Department of Defense (BCRP). Work in the Gimenez-Abian, J. F., Diaz-Martinez, L. A., Wirth, K. G., Andrews, C. A., Gimenez- Martin, G. and Clarke, D. J. (2005b). Regulated separation of sister centromeres Clarke Lab is funded by NIH grant CA099033. depends on the spindle assembly checkpoint but not on the anaphase promoting complex/cyclosome. Cell Cycle 4, 1561-1575. References Guacci, V. (2007). Sister chromatid cohesion: the cohesin cleavage model does not ring Agostinho, M., Santos, V., Ferreira, F., Costa, R., Cardoso, J., Pinheiro, I., Rino, true. Genes Cells 12, 693-708. J., Jaffray, E., Hay, R. T. and Ferreira, J. (2008). Conjugation of human Guacci, V., Hogan, E. and Koshland, D. (1994). Chromosome condensation and sister topoisomerase 2 alpha with small ubiquitin-like modifiers 2/3 in response to chromatid pairing in budding yeast. J. Cell Biol. 125, 517-530. topoisomerase inhibitors: cell cycle stage and chromosome domain specificity. Guacci, V., Koshland, D. and Strunnikov, A. (1997). A direct link between sister chromatid Cancer Res. 68, 2409-2418. cohesion and chromosome condensation revealed through the analysis of MCD1 in S. Anderson, D. E., Losada, A., Erickson, H. P. and Hirano, T. (2002). Condensin and cerevisiae. Cell 91, 47-57. cohesin display different arm conformations with characteristic hinge angles. J. Cell Haering, C. H., Lowe, J., Hochwagen, A. and Nasmyth, K. (2002). Molecular architecture Biol. 156, 419-424. of SMC proteins and the yeast cohesin complex. Mol. Cell 9, 773-788. Antoniacci, L. M. and Skibbens, R. V. (2006). Sister-chromatid telomere cohesion is Holm, C., Goto, T., Wang, J. C. and Botstein, D. (1985). DNA topoisomerase II is required nonredundant and resists both spindle forces and telomere motility. Curr. Biol. 16, 902- at the time of mitosis in yeast. Cell 41, 553-563. 906. Holm, C., Stearns, T. and Botstein, D. (1989). DNA topoisomerase II must act at mitosis Azuma, Y., Arnaoutov, A. and Dasso, M. (2003). SUMO-2/3 regulates topoisomerase II to prevent and chromosome breakage. Mol. Cell. Biol. 9, 159-168. in mitosis. J. Cell Biol. 163, 477-487. Huang, C. E., Milutinovich, M. and Koshland, D. (2005). Rings, bracelet or snaps: Azuma, Y., Arnaoutov, A., Anan, T. and Dasso, M. (2005). PIASy mediates SUMO-2 fashionable alternatives for Smc complexes. Philos. Trans. R. Soc. Lond. B Biol. Sci. Journal of Cell Science conjugation of Topoisomerase-II on mitotic chromosomes. EMBO J. 24, 2172-2182. 360, 537-542. Bachant, J., Alcasabas, A., Blat, Y., Kleckner, N. and Elledge, S. J. (2002). The SUMO- Huang, J. and Laurent, B. C. (2004). A role for the RSC chromatin remodeler in regulating 1 isopeptidase Smt4 is linked to centromeric cohesion through SUMO-1 modification cohesion of sister chromatid arms. Cell Cycle 3, 973-975. of DNA topoisomerase II. Mol. Cell 9, 1169-1182. Inoue, A., Li, T., Roby, S. K., Valentine, M. B., Inoue, M., Boyd, K., Kidd, V. J. and Baetz, K. K., Krogan, N. J., Emili, A., Greenblatt, J. and Hieter, P. (2004). The ctf13- Lahti, J. M. (2007). Loss of ChlR1 helicase in mouse causes lethality due to the 30/CTF13 genomic haploinsufficiency modifier screen identifies the yeast chromatin accumulation of aneuploid cells generated by cohesion defects and placental remodeling complex RSC, which is required for the establishment of sister chromatid malformation. Cell Cycle 6, 1646-1654. cohesion. Mol. Cell. Biol. 24, 1232-1244. Kenney, R. D. and Heald, R. (2006). Essential roles for cohesin in and spindle Bickmore, W. A. and Oghene, K. (1996). Visualizing the spatial relationships between function in Xenopus egg extracts. J. Cell Sci. 119, 5057-5066. defined DNA sequences and the axial region of extracted metaphase chromosomes. Cell Kim, J. S., Krasieva, T. B., LaMorte, V., Taylor, A. M. and Yokomori, K. (2002). Specific 84, 95-104. recruitment of human cohesin to laser-induced DNA damage. J. Biol. Chem. 277, 45149- Chestukhin, A., Pfeffer, C., Milligan, S., DeCaprio, J. A. and Pellman, D. (2003). 45153. Processing, localization, and requirement of human separase for normal anaphase Kops, G. J., Weaver, B. A. and Cleveland, D. W. (2005). On the road to cancer: aneuploidy progression. Proc. Natl. Acad. Sci. USA 100, 4574-4579. and the mitotic checkpoint. Nat. Rev. Cancer 5, 773-785. Ciosk, R., Zachariae, W., Michaelis, C., Shevchenko, A., Mann, M. and Nasmyth, K. Koshland, D. and Hartwell, L. H. (1987). The structure of sister minichromosome DNA (1998). An ESP1/PDS1 complex regulates loss of sister chromatid cohesion at the before anaphase in Saccharomyces cerevisiae. Science 238, 1713-1716. metaphase to anaphase transition in yeast. Cell 93, 1067-1076. Koshland, D. E. and Guacci, V. (2000). Sister chromatid cohesion: the beginning of a Ciosk, R., Shirayama, M., Shevchenko, A., Tanaka, T., Toth, A., Shevchenko, A. and long and beautiful relationship. Curr. Opin. Cell Biol. 12, 297-301. Nasmyth, K. (2000). Cohesin’s binding to chromosomes depends on a separate complex Lam, W. W., Peterson, E. A., Yeung, M. and Lavoie, B. D. (2006). Condensin is required consisting of Scc2 and Scc4 proteins. Mol. Cell 5, 243-254. for chromosome arm cohesion during mitosis. Genes Dev. 20, 2973-2984. Clarke, D. J., Johnson, R. T. and Downes, C. S. (1993). Topoisomerase II inhibition Lee, J. Y. and Orr-Weaver, T. L. (2001). The molecular basis of sister-chromatid cohesion. prevents anaphase chromatid segregation in mammalian cells independently of the Annu. Rev. Cell Dev. Biol. 17, 753-777. generation of DNA strand breaks. J. Cell Sci. 105, 563-569. Loo, S., Fox, C. A., Rine, J., Kobayashi, R., Stillman, B. and Bell, S. (1995). The origin Clarke, D. J., Diaz-Martinez, L. A. and Gimenez-Abian, J. F. (2005). Anaphase recognition complex in silencing, cell cycle progression, and DNA replication. Mol. promoting complex or cyclosome? Cell Cycle 4, 1585-1592. Biol. Cell 6, 741-756. Coelho, P. A., Queiroz-Machado, J. and Sunkel, C. E. (2003). Condensin-dependent Losada, A., Hirano, M. and Hirano, T. (1998). Identification of Xenopus SMC protein localisation of topoisomerase II to an axial chromosomal structure is required for sister complexes required for sister chromatid cohesion. Genes Dev. 12, 1986-1997. chromatid resolution during mitosis. J. Cell Sci. 116, 4763-4776. Losada, A., Yokochi, T. and Hirano, T. (2005). Functional contribution of Pds5 to cohesin- D’Amours, D., Stegmeier, F. and Amon, A. (2004). Cdc14 and condensin control the mediated cohesion in human cells and Xenopus egg extracts. J. Cell Sci. 118, 2133- dissolution of cohesin-independent chromosome linkages at repeated DNA. Cell 117, 2141. 455-469. Mao, Y., Abrieu, A. and Cleveland, D. W. (2003). Activating and silencing the mitotic Darwiche, N., Freeman, L. A. and Strunnikov, A. (1999). Characterization of the checkpoint through CENP-E-dependent activation/inactivation of BubR1. Cell 114, 87- components of the putative mammalian sister chromatid cohesion complex. Gene 233, 98. 39-47. Mayer, M. L., Gygi, S. P., Aebersold, R. and Hieter, P. (2001). Identification of Dawlaty, M. M., Malureanu, L., Jeganathan, K. B., Kao, E., Sustmann, C., Tahk, S., RFC(Ctf18p, Ctf8p, Dcc1p): an alternative RFC complex required for sister chromatid Shuai, K., Grosschedl, R. and van Deursen, J. M. (2008). Resolution of sister cohesion in S. cerevisiae. Mol. Cell 7, 959-970. 2114 Journal of Cell Science 121 (13)

Michaelis, C., Ciosk, R. and Nasmyth, K. (1997). Cohesins: chromosomal proteins that Strom, L., Karlsson, C., Lindroos, H. B., Wedahl, S., Katou, Y., Shirahige, K. and prevent premature separation of sister chromatids. Cell 91, 35-45. Sjogren, C. (2007). Postreplicative formation of cohesion is required for repair and Murray, A. W. and Szostak, J. W. (1985). Chromosome segregation in mitosis and meiosis. induced by a single DNA break. Science 317, 242-245. Annu. Rev. Cell Biol. 1, 289-315. Strunnikov, A. V., Hogan, E. and Koshland, D. (1995). SMC2, a Saccharomyces cerevisiae Nagao, K., Adachi, Y. and Yanagida, M. (2004). Separase-mediated cleavage of cohesin gene essential for chromosome segregation and condensation, defines a subgroup within at interphase is required for DNA repair. Nature 430, 1044-1048. the SMC family. Genes Dev. 9, 587-599. Nasmyth, K. and Schleiffer, A. (2004). From a single double helix to paired double helices Sullivan, M., Higuchi, T., Katis, V. L. and Uhlmann, F. (2004). Cdc14 phosphatase induces and back. Philos. Trans. R. Soc. Lond. B Biol. Sci. 359, 99-108. rDNA condensation and resolves cohesin-independent cohesion during budding yeast Nasmyth, K. and Haering, C. H. (2005). The structure and function of SMC and kleisin anaphase. Cell 117, 471-482. complexes. Annu. Rev. Biochem. 74, 595-648. Sumara, I., Vorlaufer, E., Gieffers, C., Peters, B. H. and Peters, J. M. (2000). Nelson, E. M., Tewey, K. M. and Liu, L. F. (1984). Mechanism of antitumor drug action: Characterization of vertebrate cohesin complexes and their regulation in prophase. J. poisoning of mammalian DNA topoisomerase II on DNA by 4’-(9-acridinylamino)- Cell Biol. 151, 749-762. methanesulfon-m-anisidide. Proc. Natl. Acad. Sci. USA 81, 1361-1365. Sundin, O. and Varshavsky, A. (1980). Terminal stages of SV40 DNA replication proceed Noble, D., Kenna, M. A., Dix, M., Skibbens, R. V., Unal, E. and Guacci, V. (2006). via multiply intertwined catenated dimers. Cell 21, 103-114. Intersection between the regulators of sister chromatid cohesion establishment and Suter, B., Tong, A., Chang, M., Yu, L., Brown, G. W., Boone, C. and Rine, J. (2004). maintenance in budding yeast indicates a multi-step mechanism. Cell Cycle 5, 2528- The origin recognition complex links replication, sister chromatid cohesion and 2536. transcriptional silencing in Saccharomyces cerevisiae. Genetics 167, 579-591. Papi, M., Berdougo, E., Randall, C. L., Ganguly, S. and Jallepalli, P. V. (2005). Multiple Takahashi, Y., Yong-Gonzalez, V., Kikuchi, Y. and Strunnikov, A. (2006). SIZ1/SIZ2 roles for separase auto-cleavage during the G2/M transition. Nat. Cell Biol. 7, 1029- control of chromosome transmission fidelity is mediated by the sumoylation of 1035. topoisomerase II. Genetics 172, 783-794. Pflumm, M. F. and Botchan, M. R. (2001). Orc mutants arrest in metaphase with Tewey, K. M., Chen, G. L., Nelson, E. M. and Liu, L. F. (1984). Intercalative antitumor abnormally condensed chromosomes. Development 128, 1697-1707. drugs interfere with the breakage-reunion reaction of mammalian DNA topoisomerase Porter, A. C. and Farr, C. J. (2004). Topoisomerase II: untangling its contribution at the II. J. Biol. Chem. 259, 9182-9187. centromere. Chromosome Res. 12, 569-583. Toth, A., Ciosk, R., Uhlmann, F., Galova, M., Schleiffer, A. and Nasmyth, K. (1999). Potts, P. R., Porteus, M. H. and Yu, H. (2006). Human SMC5/6 complex promotes sister Yeast cohesin complex requires a conserved protein, Eco1p(Ctf7), to establish cohesion chromatid homologous recombination by recruiting the SMC1/3 cohesin complex to between sister chromatids during DNA replication. Genes Dev. 13, 320-333. double-strand breaks. EMBO J. 25, 3377-3388. Toyoda, Y. and Yanagida, M. (2006). Coordinated requirements of human topo II and Prasanth, S. G., Prasanth, K. V. and Stillman, B. (2002). Orc6 involved in DNA cohesin for metaphase centromere alignment under Mad2-dependent spindle checkpoint replication, chromosome segregation, and cytokinesis. Science 297, 1026-1031. surveillance. Mol. Biol. Cell 17, 2287-2302. Prasanth, S. G., Prasanth, K. V., Siddiqui, K., Spector, D. L. and Stillman, B. (2004). Uemura, T. and Yanagida, M. (1984). Isolation of type I and II DNA topoisomerase mutants Human Orc2 localizes to centrosomes, centromeres and heterochromatin during from fission yeast: single and double mutants show different phenotypes in cell growth chromosome inheritance. EMBO J. 23, 2651-2663. and chromatin organization. EMBO J. 3, 1737-1744. Quintana, D. G. and Dutta, A. (1999). The metazoan origin recognition complex. Front. Uemura, T., Ohkura, H., Adachi, Y., Morino, K., Shiozaki, K. and Yanagida, M. (1987). Biosci. 4, D805-D815. DNA topoisomerase II is required for condensation and separation of mitotic Rankin, S., Ayad, N. G. and Kirschner, M. W. (2005). Sororin, a substrate of the anaphase- chromosomes in S. pombe. Cell 50, 917-925. promoting complex, is required for sister chromatid cohesion in vertebrates. Mol. Cell Uhlmann, F. (2004). The mechanism of sister chromatid cohesion. Exp. Cell Res. 296, 80- 18, 185-200. 85. Riedel, C. G., Gregan, J., Gruber, S. and Nasmyth, K. (2004). Is chromatin remodeling Uhlmann, F., Lottspeich, F. and Nasmyth, K. (1999). Sister-chromatid separation at required to build sister-chromatid cohesion? Trends Biochem. Sci. 29, 389-392. anaphase onset is promoted by cleavage of the cohesin subunit Scc1. Nature 400, 37-42. Rollins, R. A., Morcillo, P. and Dorsett, D. (1999). Nipped-B, a Drosophila homologue Uhlmann, F., Wernic, D., Poupart, M. A., Koonin, E. V. and Nasmyth, K. (2000). of chromosomal adherins, participates in activation by remote enhancers in the cut and Cleavage of cohesin by the CD clan protease separin triggers anaphase in yeast. Cell Ultrabithorax genes. Genetics 152, 577-593. 103, 375-386. Rubio, C., Pehlivan, T., Rodrigo, L., Simon, C., Remohi, J. and Pellicer, A. (2005). Unal, E., Heidinger-Pauli, J. M. and Koshland, D. (2007). DNA double-strand breaks Embryo aneuploidy screening for unexplained recurrent miscarriage: a minireview. Am. trigger genome-wide sister-chromatid cohesion through Eco1 (Ctf7). Science 317, 245- J. Reprod. Immunol. 53, 159-165. 248. Sakaguchi, A. and Kikuchi, A. (2004). Functional compatibility between isoform alpha Vagnarelli, P., Morrison, C., Dodson, H., Sonoda, E., Takeda, S. and Earnshaw, W. and beta of type II DNA topoisomerase. J. Cell Sci. 117, 1047-1054. C. (2004). Analysis of Scc1-deficient cells defines a key metaphase role of vertebrate

Journal of Cell Science Shah, K., Sivapalan, G., Gibbons, N., Tempest, H. and Griffin, D. K. (2003). The genetic cohesin in linking sister . EMBO Rep. 5, 167-171. basis of infertility. Reproduction 126, 13-25. Vas, A. C., Andrews, C. A., Kirkland Matesky, K. and Clarke, D. J. (2007). In vivo Shimada, K. and Gasser, S. M. (2007). The origin recognition complex functions in sister- analysis of chromosome condensation in Saccharomyces cerevisiae. Mol. Biol. Cell 18, chromatid cohesion in Saccharomyces cerevisiae. Cell 128, 85-99. 557-568. Sjogren, C. and Nasmyth, K. (2001). Sister chromatid cohesion is required for Vass, S., Cotterill, S., Valdeolmillos, A. M., Barbero, J. L., Lin, E., Warren, W. D. and postreplicative double-strand break repair in Saccharomyces cerevisiae. Curr. Biol. 11, Heck, M. M. (2003). Depletion of Drad21/Scc1 in Drosophila cells leads to instability 991-995. of the cohesin complex and disruption of mitotic progression. Curr. Biol. 13, 208-218. Skibbens, R. V. (2000). Holding your own: establishing sister chromatid cohesion. Genome Watrin, E., Schleiffer, A., Tanaka, K., Eisenhaber, F., Nasmyth, K. and Peters, J. M. Res. 10, 1664-1671. (2006). Human Scc4 is required for cohesin binding to chromatin, sister-chromatid Skibbens, R. V. (2005). Unzipped and loaded: the role of DNA helicases and RFC clamp- cohesion, and mitotic progression. Curr. Biol. 16, 863-874. loading complexes in sister chromatid cohesion. J. Cell Biol. 169, 841-846. Weaver, B. A. and Cleveland, D. W. (2006). Does aneuploidy cause cancer? Curr. Opin. Skibbens, R. V., Maradeo, M. and Eastman, L. (2007). Fork it over: the cohesion Cell Biol. 18, 658-667. establishment factor Ctf7p and DNA replication. J. Cell Sci. 120, 2471-2477. Weaver, B. A., Silk, A. D., Montagna, C., Verdier-Pinard, P. and Cleveland, D. W. Sonoda, E., Matsusaka, T., Morrison, C., Vagnarelli, P., Hoshi, O., Ushiki, T., Nojima, (2007). Aneuploidy acts both oncogenically and as a tumor suppressor. Cancer Cell 11, K., Fukagawa, T., Waizenegger, I. C., Peters, J. M. et al. (2001). Scc1/Rad21/Mcd1 25-36. is required for sister chromatid cohesion and kinetochore function in vertebrate cells. Weitzer, S., Lehane, C. and Uhlmann, F. (2003). A model for ATP hydrolysis-dependent Dev. Cell 1, 759-770. binding of cohesin to DNA. Curr. Biol. 13, 1930-1940. Stead, K., Aguilar, C., Hartman, T., Drexel, M., Meluh, P. and Guacci, V. (2003). Pds5p Wong, K. A., Kim, R., Christofk, H., Gao, J., Lawson, G. and Wu, H. (2004). Protein regulates the maintenance of sister chromatid cohesion and is sumoylated to promote inhibitor of activated STAT Y (PIASy) and a splice variant lacking exon 6 enhance the dissolution of cohesion. J. Cell Biol. 163, 729-741. sumoylation but are not essential for embryogenesis and adult life. Mol. Cell. Biol. 24, Straight, A. F., Belmont, A. S., Robinett, C. C. and Murray, A. W. (1996). GFP tagging 5577-5586. of budding yeast chromosomes reveals that protein-protein interactions can mediate sister Yalon, M., Gal, S., Segev, Y., Selig, S. and Skorecki, K. L. (2004). Sister chromatid chromatid cohesion. Curr. Biol. 6, 1599-1608. separation at human telomeric regions. J. Cell Sci. 117, 1961-1970. Strom, L., Lindroos, H. B., Shirahige, K. and Sjogren, C. (2004). Postreplicative Zur, A. and Brandeis, M. (2001). Securin degradation is mediated by fzy and fzr, and is recruitment of cohesin to double-strand breaks is required for DNA repair. Mol. Cell required for complete chromatid separation but not for cytokinesis. EMBO J. 20, 792- 16, 1003-1015. 801.