The task of preserving the stability of the genetic information of a cell is characterised by several important requirements (Hoeijmakers, 2009). First, genome maintenance is a largely cell-autonomous function, as each cell within an organism is obliged to prevent or repair injury of its own genome. Second, insults that endanger genome stability require a rapid response because damage to the genetic material quickly affects essential cellular functions by interfering with gene expression. Third, any response needs to be reversibly tuned to the physiological state of the cell, in particular its stage within the cell cycle. Last, but not least, mechanisms for the protection of genome integrity have to react flexibly to a large variety of damaging agents to which a cell may be exposed.

Considering these characteristics of DNA repair, it is not surprising that post-translational modifications with members of the ubiquitin family, such as ubiquitin itself and small ubiquitin-related modifier (SUMO), have been identified as key contributors to genome maintenance. The notion that ubiquitin and SUMO can rapidly and reversibly change the properties, stability or localisation of their target proteins without the need for de novo protein synthesis makes them ideal regulators for fine-tuning DNA repair and damage response pathways. Their general action as modulators of protein function goes far beyond their contribution to regulated proteolysis, for which ubiquitin first became famous (Glickman and Ciechanover, 2002).

In this Cell Science at a Glance article and the accompanying poster, I will summarise the contributions of ubiquitin and SUMO to the major pathways of genome maintenance. Several excellent reviews on this topic have been published over the past years, which give insight into the relevant mechanisms involved (Al-Hakim et al., 2010; Bergink and Jentsch, 2009; Huang and D'Andrea, 2006; Ulrich and Walden, 2010). Here, I will place more emphasis on the crosstalk between individual repair pathways and the specific contributions of selected proteins involved in ubiquitin and SUMO conjugation to the processes involved. For information about the general principles of protein ubiquitylation and SUMOylation, the reader is referred to pertinent reviews in this field (Glickman and Ciechanover, 2002; Kerscher et al., 2006).

DNA double-strand breaks (DSBs) represent perhaps the most dangerous type of DNA lesion because they have dramatic effects on all DNA transactions, including proper segregation of chromosomes during cell division (Hoeijmakers, 2009). In order to initiate repair, a series of phosphorylation events, namely the phosphorylation of the histone variant H2AX and mediator of DNA-damage checkpoint 1 (MDC1) by the checkpoint kinase ataxia telangiectasia mutated (ATM), leads to the sequential recruitment of several E3 ubiquitin ligases in higher eukaryotes, as detailed below (Bekker-Jensen and Mailand, 2011; Panier and Durocher, 2009; Tang and Greenberg, 2010). These enzymes promote extensive ubiquitylation of histone H2A and other, yet unknown, chromatin-associated proteins in a cascade that eventually results in the localisation of another RING-finger ubiquitin ligase, BRCA1 (for breast cancer 1, early onset), as well as the checkpoint protein TP53BP1 (for tumor protein p53 binding protein 1), to the break. Whereas BRCA1 is essential for initiating DSB repair by homologous recombination, TP53BP1 has been associated with repair by non-homologous end-joining (Hiom, 2010).

The mechanistic details of TP53BP1 recruitment to DSBs are poorly understood. The protein binds to methylated histones, and it is unclear how the damage-induced ubiquitylation cascade elicits this signal for chromatin association of TP53BP1. By contrast, the signalling pathway leading to BRCA1 recruitment has now been roughly elucidated. Although there are indications that the pathway is not entirely linear, the prevailing concept is based on the recognition of post-translational modifications, such as phosphate or ubiquitin moieties, by dedicated domains in the respective effector proteins, which themselves trigger the recruitment of additional enzymes for modification of other chromatin components. Specifically, phosphorylated H2AX (γ-H2AX) is recognised by tandem BRCA1 C-terminal (BRCT) motifs on MDC1. Following phosphorylation by ATM, MDC1 is then able to interact with the forkhead-associated (FHA) domains of the RING-finger E3 RNF8 (Huen et al., 2007; Kolas et al., 2007; Mailand et al., 2007). The ubiquitin moieties conjugated by RNF8 and its cognate ubiquitin-conjugating enzyme (E2) UBC13 (officially known as ubiquitin-conjugating enzyme E2N, UBE2N) to histones H2A and H2AX, and possibly other chromatin-bound proteins, are thought to be recognised by a second RING-finger E3, RNF168, through its ‘motifs interacting with ubiquitin’ (MIU) domains (Doil et al., 2009; Stewart et al., 2009). RNF168, together with UBC13, catalyses the formation of polymeric chains, linked through lysine (K) 63 of ubiquitin, that in turn recruit the adapter protein RAP80 (receptor associated protein 80, also known as UIMC1) through its ubiquitin-interacting motifs (UIM). The dimeric BRCA1–BARD1 complex is eventually recruited through the interaction of the BRCA1 tandem BRCT motifs with the phosphorylated abraxas protein (also known as FAM175A), which is itself part of a multisubunit scaffold complex (Kim et al., 2007; Liu et al., 2007; Sobhian et al., 2007; Wang et al., 2007). Although it is clear that BRCA1 is an important mediator of DNA DSB repair through homologous recombination, its relevant substrate proteins and cognate E2 enzyme(s), the structure of the resulting ubiquitin conjugates and the physiological consequences of these modifications are not yet fully understood (Hiom, 2010).

In addition to BRCA1, several other ubiquitin E3s participate in the cascade, although their roles are less well defined. On one hand, RAD18, a RING-finger E3 best known for its function in DNA damage bypass (see below), acts downstream of RNF8, presumably by being recruited to chromatin by means of a ubiquitin-binding zinc finger (UBZ) domain (Huang et al., 2009). Intriguingly, the binding of RAD18, but not its E3 ligase activity, is required for proper homologous recombination. On the other hand, HERC2, a large HECT domain family E3 protein, associates with the FHA domain of RNF8 and seems to promote interaction with UBC13 in an as yet unknown way (Bekker-Jensen et al., 2010). Finally, a number of polycomb group proteins, which mediate transcriptional repression through modulation of chromatin structure, have recently been shown to contribute to the recruitment of both BRCA1 and TP53BP1 through H2A and H2AX ubiquitylation (Gieni et al., 2011). These include the RING-finger proteins BMI1 and RNF2 (also called RING1B or RING2), which form a heterodimeric ubiquitin E3. Although the mechanism by which the complex is initially recruited is a matter of debate – it might involve the MRN (MRE11–RAD50–NBS1) damage recognition complex and/or poly-(ADP)-ribosylation (Chou et al., 2010; Ismail et al., 2010) – sustained localisation at the site of DSBs seems to require signalling by the ATM or ataxia-telangiectasia-related (ATR) kinases, H2AX phosphorylation and its ubiquitylation by RNF8 (Ginjala et al., 2011).

Negative regulation of ubiquitylation in the context of DSB repair is exerted by a number of deubiquitylating enzymes (DUBs) (Al-Hakim et al., 2010), namely BRCC36 (for BRCA1/BRCA2-containing complex, subunit 3), ubiquitin specific peptidases (USP) 3 and 16 as well as OTUB1 (for OTU domain, ubiquitin aldehyde binding 1). BRCC36 is part of the RAP80 complex and exhibits a preference for K63-polyubiquitin chains (Sobhian et al., 2007). USP3 and USP16 have both been shown to act on histone H2A and seem to downregulate the pathway at the stage of RNF8 (Cai et al., 1999; Doil et al., 2009; Joo et al., 2007; Nicassio et al., 2007), whereas OTUB1 acts further downstream and might inhibit the action of UBC13 in a non-catalytic manner (Nakada et al., 2010). The negative influence of DUBs on the formation of damage-induced BRCA1 and TP53BP1 foci indicates that cells maintain a fine balance between ubiquitylation and deubiquitylation to regulate DSB repair.

In addition to ubiquitylation, the post-translational modifications at DSBs also involve SUMOylation, which is mediated by the SUMO E3 PIAS1 and PIAS4 proteins (for protein inhibitor of activated STAT), which both associate with the single SUMO E2 UBC9 (also known as UBE2I) (Galanty et al., 2009; Morris et al., 2009). One of the substrates appears to be BRCA1 itself, whose catalytic activity is boosted by SUMO modification. Hence, SUMOylation of the ubiquitin E3 BRCA1 represents an interesting example of crosstalk between the two modifiers. In addition, the polycomb protein Pc2 (officially known as CBX4), which acts as a SUMO ligase on a number of substrates, including itself (Wotton and Merrill, 2007), is also recruited to chromatin in a damage-dependent manner (Chou et al., 2010). There are clearly additional physiologically relevant targets, whose SUMOylation is important during DSB repair, but these remain to be identified.

Apart from the direct involvement in the recruitment of repair factors, histone ubiquitylation is known to affect chromatin on a structural level, which is particularly important for the activation of transcription. In higher eukaryotes, this involves the relaxation of chromatin through monoubiquitylation of histone H2B by the heterodimeric RING finger E3 complex RNF20–RNF40 (Fierz et al., 2011; Weake and Workman, 2008). Very recently, it has been discovered that RNF20–RNF40-dependent H2B ubiquitylation is also required for efficient DSB repair, presumably by the same mechanism of chromatin decompaction (Moyal et al., 2011). As a consequence, inhibition of this modification causes defects in the recruitment of repair factors that are involved in both homologous recombination and non-homologous end-joining processes. Interestingly, this H2B-dependent contribution appears to be independent of the signalling pathway that involves RNF8-mediated H2A ubiquitylation.

As discussed above, the ubiquitylation and SUMOylation cascades elicited at DSBs culminate in the recruitment of factors that initiate the two major pathways for the repair of such lesions (i.e. homologous recombination through recruitment of BRCA1 and non-homologous end-joining through TP53BP1). Whereas this particular signalling pathway is restricted to higher eukaryotes, SUMO also modifies several core recombination factors in both higher and lower eukaryotes. Among its prominent targets is RAD52, which promotes formation of the recombinogenic RAD51 filament. In budding yeast, SUMOylation by the E3 Siz2 stabilises Rad52 (Sacher et al., 2006) while at the same time also reducing the DNA-binding and single-strand annealing activities of this protein (Altmannova et al., 2010). SUMOylation of replication protein A (RPA), a single-stranded binding complex that is essential for recombination, replication and repair, has been detected in yeast and mammalian cells (Burgess et al., 2007; Dou et al., 2010). In mammals, this modification facilitates the recruitment of RAD51 and has been shown to be counteracted by the SUMO-specific isopeptidase SENP6 (Dou et al., 2010).

As SUMOylation is often less dependent on specific ligases than ubiquitylation, the cognate E3s have not been well defined for all SUMO targets. On the basis of the phenotypes of the respective deletion mutants, it is clear that all three major SUMO E3s in budding yeast, Siz1, Siz2 and Mms21, participate in the modification of relevant substrates. Whereas the two yeast members of the PIAS family, Siz1 and Siz2, associate with DNA through SAP (for SAF-A/B, Acinus and PIAS) domains (Okubo et al., 2004), Mms21 is recruited to DNA through its association with the Smc5–Smc6 complex, a cohesin-like assembly with an important function in homologous recombination and replication fork restart in both yeast and humans (Potts and Yu, 2005; Zhao and Blobel, 2005). Although the importance of these SUMO ligases to genome maintenance by means of homologous recombination is undeniable, many of the relevant target proteins, as well as the mechanisms by which they affect the function of these targets, have yet to be elucidated.

In addition to protein modification by SUMO, the regulation of DNA repair by homologous recombination also involves ubiquitin. In fission yeast, this modifier acts in its ‘classical’ way by inducing the degradation of the recombination factor Rad54 during the G1 phase of the cell cycle, which is consistent with a downregulation of recombination activity at this stage (Trickey et al., 2008). In this case, ubiquitin is attached by the anaphase promoting complex/cyclosome (APC/C), an E3 protein involved in cell cycle regulation that acts together with its G1-specific regulator Fzr (fizzy related, also referred to as Cdh1 and Hct1).

An interesting crosstalk between ubiquitin and SUMO was uncovered by the identification of a set of ubiquitin ligases that recognise SUMOylated proteins as their substrates by means of SUMO-interaction motifs (SIMs) (Prudden et al., 2007; Sun et al., 2007; Uzunova et al., 2007; Xie et al., 2007). In yeast, one of these SUMO-targeted ubiquitin ligases, the heterodimeric Slx5–Slx8 complex in S. cerevisiae (Rfp1/2–Slx8 in S. pombe), acts as a regulator of homologous recombination, although its mechanism of action and its relevant substrate proteins remain to be identified. Bulk removal of high-molecular-mass SUMO targets might contribute to this DNA repair process (Uzunova et al., 2007).

Homologous recombination plays a major role not only in DNA DSB repair but also in promoting the restart of stalled or collapsed replication forks. In this aspect, the SUMO system again appears to be particularly important. Although it is very likely that not all SUMOylated proteins that are relevant to this process have been identified, appropriate targets include members of the family of RECQ helicases: BLM (Bloom syndrome, RecQ helicase-like) and WRN (Werner syndrome, RecQ helicase-like) in higher eukaryotes, and Sgs1 in budding yeast (Branzei et al., 2006; Eladad et al., 2005; Kawabe et al., 2000). How SUMO affects the activities of these proteins and the subsequent restart of replication forks, however, has not been fully elucidated.

Homologous recombination is just one way of dealing with DNA damage during replication. It applies to a variety of situations, including those where replication fork progression is impeded by strand breaks or interstrand crosslinks. Small lesions, such as adducts in the template DNA, which cannot be processed by the replicative polymerases, represent more subtle disturbances to fork progression and can be resolved by different processes (Friedberg, 2005). One solution to this problem, which results in error-free bypass, is template switching through a recombination event. Alternatively, specialised, error-prone DNA polymerases that can accommodate abnormal structures in their active sites can be employed in a process referred to as translesion synthesis (TLS) (Lehmann et al., 2007). Both strategies contribute to cellular damage resistance, but they also require strict regulation, as they might themselves induce genomic instability by means of larger genome rearrangments or point mutations, respectively.

Control over both bypass pathways is mediated by ubiquitylation of the sliding clamp protein proliferating cell nuclear antigen (PCNA) (Ulrich, 2009). Monoubiquitylation at K164 by a complex containing the E3 RAD18 and the E2 RAD6 enhances the affinity of a series of damage-tolerant polymerases for PCNA and thus facilitates TLS (Bienko et al., 2005; Stelter and Ulrich, 2003). Conjugation of additional ubiquitin moieties by the RING-finger E3 Rad5 and the heterodimeric E2 Ubc13–Mms2 results in a K63-linked polyubiquitin chain, which in yeast is a prerequisite for the error-free template-switching pathway (Hoege et al., 2002). In addition, the enzymes involved in polyubiquitylation have been implicated in some aspects of TLS (Gangavarapu et al., 2006; Pages et al., 2008). Intriguingly, higher eukaryotes encode two homologues of Rad5, helicase-like transcription factor (HLTF) and histone linker PHD RING helicase (SHPRH), which apparently act non-redundantly on different types of DNA damage (Lin et al., 2011). How PCNA polyubiquitylation induces template switching is not at all understood. There is even controversy about whether the downstream events initiated by the modification involve the full set of homologous recombination factors or whether a simple fork reversal might be sufficient to align the stalled primer terminus with the alternative template. Interestingly, however, both ubiquitin-dependent TLS and template switching can occur in manner that is uncoupled from the replication fork, within post-replicative gaps, which argues against the predominant use of fork reversal as a means to initiate error-free damage bypass (Daigaku et al., 2010; Karras and Jentsch, 2010).

Budding yeast PCNA is also modified by SUMO (Hoege et al., 2002). In contrast to ubiquitylation, this reaction is not damage-dependent, but occurs constitutively during replication, and also changes the affinity of the clamp for its interaction partners. Specifically, SUMOylation enhances the binding of an anti-recombinogenic helicase, Srs2, which prevents the formation of unscheduled Rad51 filaments and thereby enables the ubiquitin-dependent bypass pathway to act under conditions of DNA damage (Papouli et al., 2005; Pfander et al., 2005). Recognition of SUMOylated PCNA by Srs2 is mediated by a SIM in the C-terminus of Srs2. At the same time, PCNA SUMOylation inhibits the binding of Eco1, which is involved in the establishment of chromatid cohesion (Moldovan et al., 2006), and induces the recruitment of an alternative clamp loader, Elg1, again through a number of SIMs, which influences genome stability in a poorly defined way (Parnas et al., 2010). Intriguingly, the human homologue of Elg1, ATAD5, was found to promote deubiquitylation of PCNA independently of SUMO, by localising a complex of the ubiquitin-specific peptidase 1 (USP1) and the USP1 accessory factor 1 (UAF1) to ubiquitylated PCNA (Lee et al., 2010). SUMOylation of PCNA was also observed in chicken DT40 cells and in Xenopus laevis egg extracts (Arakawa et al., 2006; Leach and Michael, 2005); however, its functions in these system have not been determined and are unlikely to involve a vertebrate Srs2 homologue. Hence, the SUMO-dependent crosstalk with homologous recombination might be specific to budding yeast.

Repair of DNA interstrand cross-links (ICL) is particularly important for genome stability and survival, as these lesions impair strand separation not only for the purpose of replication but also for transcription. In general, ICL repair requires either homologous recombination or a combination of nucleotide excision repair and TLS. Higher eukaryotes have developed a special processing system for ICLs, the Fanconi anaemia pathway, which is named after a hereditary disease associated with defects in this process (Alpi and Patel, 2009).

The monoubiquitylation of a heterodimeric complex of the FANC (for Fanconi anemia, complementation group) proteins FANCD2 and FANCI, whose modification triggers their localisation to chromatin, is central to the Fanconi anaemia pathway (Garcia-Higuera et al., 2001; Sims et al., 2007; Smogorzewska et al., 2007). Ubiquitylation is mediated by the Fanconi core complex, whose main component is FANCL, a RING-finger E3 that cooperates with the E2 UBE2T. A complex of other chromatin-associated proteins, the so-called recognition complex, with FANCM as its central DNA-binding component, functions in directing the core complex to the appropriate sites. The events downstream of ubiquitylation are not entirely clear, but similar to the ubiquitylation cascade at DSBs described above, they result in an activation of the BRCA1/2 pathway (Wang, 2007). It has recently been shown that the FANCD2/FANCI-associated nuclease 1 (FAN1) is recruited to monoubiquitylated FANCD2 by means of a UBZ domain, which might account, at least in part, for the function of FANCD2 ubiquitylation (Kratz et al., 2010; Liu et al., 2010; MacKay et al., 2010; Smogorzewska et al., 2010).

Intriguingly, the Fanconi anaemia pathway overlaps not only with the BRCA1 pathway that operates at DSBs but also exhibits substantial crosstalk with RAD18-dependent damage bypass. On one hand, PCNA and FANCD2 are deubiquitylated by the same DUB, USP1 (Huang et al., 2006). On the other hand, it has been recently found that PCNA monoubiquitylation is actually a prerequisite for FANCD2 ubiquitylation, potentially by resulting in the recruitment and direct stimulation of FANCL by ubiquitylated PCNA (Geng et al., 2010; Park et al., 2010; Song et al., 2010). Hence, activation of the Fanconi anaemia pathway seems to be intimately coupled to TLS as a coordinated approach to processing lesions during replication.

Nucleotide excision repair (NER) is initiated independently of DNA replication on bulky lesions that distort the helix, and involves the excision of the region around the lesion as a single-stranded oligonucleotide and filling of the resulting gap by re-synthesis. A sub-pathway of NER called global genome repair (GGR) operates throughout the genome. In humans, damage recognition is mediated by XPC (for xeroderma pigmentosum, complementation group C) and – in the case of UV-induced pyrimidine dimers – the UV-damage-specific DNA-damage-binding (DDB) protein complex, which comprises DDB1 and DDB2 (Sugasawa, 2010). Whereas DDB2 binds to the damaged DNA, DDB1 forms part of a multimeric cullin-based E3 protein, the DDB1–CUL4 complex. This E3 polyubiquitylates XPC and DDB2. Whereas DDB2 ubiquitylation results in proteasomal degradation, XPC remains stable, indicating that the modification serves regulatory purposes (Sugasawa et al., 2005). By contrast, the XPC-related protein Rad4 from budding yeast is ubiquitylated and degraded upon DNA damage (Lommel et al., 2002). However, the details of how proteolytic and non-canonical ubiquitylation regulate GGR remain controversial. Even a non-catalytic role for the 19S proteasome cap in this process has been postulated (Russell et al., 1999).

As an alternative to GGR, damage recognition on the transcribed strand of expressed genes can be accomplished by RNA polymerase II, in a process called transcription-coupled repair (TCR). Although ubiquitin or SUMO have not been directly implicated in this reaction, ubiquitylation and subsequent proteasomal degradation of the large subunit of the polymerase itself has been recognised as an important way of resolving transcription complexes that are stalled by DNA damage (Svejstrup, 2010). The budding yeast HECT-like E3 Rsp5 (and its human homologue NEDD4) together with the E2s Ubc4 and Ubc5 mediate monoubiquitylation of the polymerase (Anindya et al., 2007; Beaudenon et al., 1999). Rsp5 can also produce K63-linked chains on RNA polymerase II. However, these appear to be subject to removal by the DUB Ubp2 (Harreman et al., 2009). Further ubiquitylation of the polymerase by a cullin3-based E3 complex results in K48-linked chains, which induce proteasomal degradation of the enzyme (Harreman et al., 2009; Ribar et al., 2007). It has also been shown that a second DUB, Ubp3, is involved in deubiquitylation of the polymerase (Kvint et al., 2008).

Both GGR and TCR converge on the filling of single-stranded gaps that result from the excision of the damaged stretch. Intriguingly, although these gaps are presumably free of lesions, PCNA monoubiquitylation and recruitment of the damage-tolerant DNA polymerase κ have been reported to contribute to the efficient repair of these lesions in humans (Ogi et al., 2010). In yeast, however, NER appear to be independent of TLS by this kind of polymerase. In addition, ubiquitylation of histone H2A by RNF8 and polycomb proteins have been implicated in the reaction (Bergink et al., 2006; Marteijn et al., 2009), suggesting parallels in the signalling pathways at DSBs and during NER.

Whereas NER resolves bulky lesions, small adducts such as alkylation or oxidative damage are processed by the base excision repair (BER) pathway, which is initiated by a series of lesion-specific glycosylases that excise damaged bases. The resulting abasic site is processed further by incision of the DNA backbone, followed by removal of the sugar-phosphate moiety and re-synthesis of the correct nucleotide. Although ubiquitylation has not been heavily implicated in this pathway yet, a prominent example of SUMO modification illustrates how this modification of a protein can change the properties of its targets in terms of its intramolecular interactions. Thymine-DNA glycosylase (TDG) acts on G–T mismatches in double-stranded DNA, but the enzyme is product-inhibited by strongly binding to the abasic site. In this situation, SUMOylation of TDG promotes catalytic turnover by inducing the release of the enzyme from DNA (Hardeland et al., 2002). This is mediated by a large conformational change in the enzyme (Steinacher and Schar, 2005). Accordingly, the crystal structure of a SUMO-modified TDG domain shows a conformation incompatible with DNA binding (Baba et al., 2005). Non-covalent SUMO binding might also contribute to the stimulation of catalytic turnover (Takahashi et al., 2005). Finally, the modification might also coordinate the intracellular localisation of TDG as well as the hand-over of the enzyme to the downstream factor, the apurinic and apyrimidinic endonuclease 1 (APE1) (Hardeland et al., 2002).

Although the list of ubiquitin and SUMO targets and conjugation factors involved in the maintenance of genome stability is far from complete, the examples given here illustrate the diversity of mechanisms by which members of the ubiquitin family are known to act. Both modifiers can initiate protein degradation, but they often function in a non-proteolytic manner, either by recruiting downstream effectors through dedicated recognition domains or by directly mediating changes in the properties or catalytic activities of their target proteins. Particularly intriguing is the degree of crosstalk between the individual pathways of damage processing. In addition to overlaps in the conjugation or deconjugation factors or the modification targets, there are instances where one particular modification event promotes or activates a subsequent reaction. This principle is evident in the SUMO-targeted ubiquitin ligases, but also in the activation of the BRCA1 ubiquitin ligase by its own SUMOylation and in the activation of the FANCL ubiquitin ligase by PCNA ubiquitylation. Overall, the variety of ubiquitin- and SUMO-mediated influences on DNA replication and repair reflects very well the flexibility by which our cells react to insults to their genetic information, and these processes highlight the paramount importance of the mutual interactions between the modifiers and DNA itself. As expected, most of the ubiquitylation and SUMOylation reactions occur directly on chromatin. The modifications in turn can induce changes in the affinity of the target proteins for DNA, as observed with sumoylated Rad52 or TDG. Alternatively, they might exert a direct influence on chromatin structure, such as with ubiquitylated histone H2B. It has become clear, however, that we still lack mechanistic insight into the consequences of ubiquitin and SUMO modification for many, if not most of their targets. Further studies are therefore likely to continue the discovery of new principles of how the two modifiers contribute to genome maintenance.

This article is part of a Minifocus on Ubiquitin. For further reading, please see related articles: ‘Emerging regulatory mechanisms in ubiquitin-dependent cell cycle control’ by Annamaria Mocciaro and Michael Rape (J. Cell Sci.125, 255-263). ‘The role of ubiquitylation in receptor endocytosis and endosomal sorting’ by Kaisa Haglund and Ivan Dikic (J. Cell Sci.125, 265-275). ‘Cellular functions of the DUBs’ by Michael J. Clague et al. (J. Cell Sci.125, 277-286). ‘HECT and RING finger families of E3 ubiquitin ligases at a glance’ by Meredith B. Metzger et al. (J. Cell Sci.125, 531-537). ‘Non-canonical ubiquitin-based signals for proteasomal degradation’ by Yelena Kravtsova-Ivantsiv and Aaron Ciechanover (J. Cell Sci.125, 539-548). ‘No one can whistle a symphony alone – how different ubiquitin linkages cooperate to orchestrate NF-κB activity’ by Anna C. Schmukle and Henning Walczak (J. Cell Sci.125, 549-559).

Funding

Work in our laboratory is funded by Cancer Research UK.

Al-Hakim
A.
,
Escribano-Diaz
C.
,
Landry
M. C.
,
O'Donnell
L.
,
Panier
S.
,
Szilard
R. K.
,
Durocher
D.
(
2010
).
The ubiquitous role of ubiquitin in the DNA damage response
.
DNA Repair
9
,
1229
-
1240
.
Alpi
A. F.
,
Patel
K. J.
(
2009
).
Monoubiquitylation in the Fanconi anemia DNA damage response pathway
.
DNA Repair
8
,
430
-
435
.
Altmannova
V.
,
Eckert-Boulet
N.
,
Arneric
M.
,
Kolesar
P.
,
Chaloupkova
R.
,
Damborsky
J.
,
Sung
P.
,
Zhao
X.
,
Lisby
M.
,
Krejci
L.
(
2010
).
Rad52 SUMOylation affects the efficiency of the DNA repair
.
Nucleic Acids Res.
38
,
4708
-
4721
.
Anindya
R.
,
Aygun
O.
,
Svejstrup
J. Q.
(
2007
).
Damage-induced ubiquitylation of human RNA polymerase II by the ubiquitin ligase Nedd4, but not Cockayne syndrome proteins or BRCA1
.
Mol. Cell
28
,
386
-
397
.
Arakawa
H.
,
Moldovan
G. L.
,
Saribasak
H.
,
Saribasak
N. N.
,
Jentsch
S.
,
Buerstedde
J. M.
(
2006
).
A role for PCNA ubiquitination in immunoglobulin hypermutation
.
PLoS Biol.
4
,
e366
.
Baba
D.
,
Maita
N.
,
Jee
J. G.
,
Uchimura
Y.
,
Saitoh
H.
,
Sugasawa
K.
,
Hanaoka
F.
,
Tochio
H.
,
Hiroaki
H.
,
Shirakawa
M.
(
2005
).
Crystal structure of thymine DNA glycosylase conjugated to SUMO-1
.
Nature
435
,
979
-
982
.
Beaudenon
S. L.
,
Huacani
M. R.
,
Wang
G.
,
McDonnell
D. P.
,
Huibregtse
J. M.
(
1999
).
Rsp5 ubiquitin-protein ligase mediates DNA damage-induced degradation of the large subunit of RNA polymerase II in Saccharomyces cerevisiae
.
Mol. Cell. Biol.
19
,
6972
-
6979
.
Bekker-Jensen
S.
,
Mailand
N.
(
2011
).
The ubiquitin- and SUMO-dependent signaling response to DNA double-strand breaks
.
FEBS Lett.
585
,
2914
-
2919
.
Bekker-Jensen
S.
,
Rendtlew Danielsen
J.
,
Fugger
K.
,
Gromova
I.
,
Nerstedt
A.
,
Lukas
C.
,
Bartek
J.
,
Lukas
J.
,
Mailand
N.
(
2010
).
HERC2 coordinates ubiquitin-dependent assembly of DNA repair factors on damaged chromosomes
.
Nat. Cell Biol.
12
,
80
-
86
.
Bergink
S.
,
Jentsch
S.
(
2009
).
Principles of ubiquitin and SUMO modifications in DNA repair
.
Nature
458
,
461
-
467
.
Bergink
S.
,
Salomons
F. A.
,
Hoogstraten
D.
,
Groothuis
T. A.
,
de Waard
H.
,
Wu
J.
,
Yuan
L.
,
Citterio
E.
,
Houtsmuller
A. B.
,
Neefjes
J.
, et al. 
. (
2006
).
DNA damage triggers nucleotide excision repair-dependent monoubiquitylation of histone H2A
.
Genes Dev.
20
,
1343
-
1352
.
Bienko
M.
,
Green
C. M.
,
Crosetto
N.
,
Rudolf
F.
,
Zapart
G.
,
Coull
B.
,
Kannouche
P.
,
Wider
G.
,
Peter
M.
,
Lehmann
A. R.
, et al. 
. (
2005
).
Ubiquitin-binding domains in Y-family polymerases regulate translesion synthesis
.
Science
310
,
1821
-
1824
.
Branzei
D.
,
Sollier
J.
,
Liberi
G.
,
Zhao
X.
,
Maeda
D.
,
Seki
M.
,
Enomoto
T.
,
Ohta
K.
,
Foiani
M.
(
2006
).
Ubc9- and mms21-mediated sumoylation counteracts recombinogenic events at damaged replication forks
.
Cell
127
,
509
-
522
.
Burgess
R. C.
,
Rahman
S.
,
Lisby
M.
,
Rothstein
R.
,
Zhao
X.
(
2007
).
The Slx5-Slx8 complex affects sumoylation of DNA repair proteins and negatively regulates recombination
.
Mol. Cell. Biol.
27
,
6153
-
6162
.
Cai
S. Y.
,
Babbitt
R. W.
,
Marchesi
V. T.
(
1999
).
A mutant deubiquitinating enzyme (Ubp-M) associates with mitotic chromosomes and blocks cell division
.
Proc. Natl. Acad. Sci. USA
96
,
2828
-
2833
.
Chou
D. M.
,
Adamson
B.
,
Dephoure
N. E.
,
Tan
X.
,
Nottke
A. C.
,
Hurov
K. E.
,
Gygi
S. P.
,
Colaiacovo
M. P.
,
Elledge
S. J.
(
2010
).
A chromatin localization screen reveals poly (ADP ribose)-regulated recruitment of the repressive polycomb and NuRD complexes to sites of DNA damage
.
Proc. Natl. Acad. Sci. USA
107
,
18475
-
18480
.
Daigaku
Y.
,
Davies
A. A.
,
Ulrich
H. D.
(
2010
).
Ubiquitin-dependent DNA damage bypass is separable from genome replication
.
Nature
465
,
951
-
955
.
Doil
C.
,
Mailand
N.
,
Bekker-Jensen
S.
,
Menard
P.
,
Larsen
D. H.
,
Pepperkok
R.
,
Ellenberg
J.
,
Panier
S.
,
Durocher
D.
,
Bartek
J.
, et al. 
. (
2009
).
RNF168 binds and amplifies ubiquitin conjugates on damaged chromosomes to allow accumulation of repair proteins
.
Cell
136
,
435
-
446
.
Dou
H.
,
Huang
C.
,
Singh
M.
,
Carpenter
P. B.
,
Yeh
E. T.
(
2010
).
Regulation of DNA repair through deSUMOylation and SUMOylation of replication protein A complex
.
Mol. Cell
39
,
333
-
345
.
Eladad
S.
,
Ye
T. Z.
,
Hu
P.
,
Leversha
M.
,
Beresten
S.
,
Matunis
M. J.
,
Ellis
N. A.
(
2005
).
Intra-nuclear trafficking of the BLM helicase to DNA damage-induced foci is regulated by SUMO modification
.
Hum. Mol. Genet.
14
,
1351
-
1365
.
Fierz
B.
,
Chatterjee
C.
,
McGinty
R. K.
,
Bar-Dagan
M.
,
Raleigh
D. P.
,
Muir
T. W.
(
2011
).
Histone H2B ubiquitylation disrupts local and higher-order chromatin compaction
.
Nat. Chem. Biol.
7
,
113
-
119
.
Friedberg
E. C.
(
2005
).
Suffering in silence: the tolerance of DNA damage
.
Nat. Rev. Mol. Cell Biol.
6
,
943
-
953
.
Galanty
Y.
,
Belotserkovskaya
R.
,
Coates
J.
,
Polo
S.
,
Miller
K. M.
,
Jackson
S. P.
(
2009
).
Mammalian SUMO E3-ligases PIAS1 and PIAS4 promote responses to DNA double-strand breaks
.
Nature
462
,
935
-
939
.
Gangavarapu
V.
,
Haracska
L.
,
Unk
I.
,
Johnson
R. E.
,
Prakash
S.
,
Prakash
L.
(
2006
).
Mms2-Ubc13-dependent and -independent roles of Rad5 ubiquitin ligase in postreplication repair and translesion DNA synthesis in Saccharomyces cerevisiae
.
Mol. Cell. Biol.
26
,
7783
-
7790
.
Garcia-Higuera
I.
,
Taniguchi
T.
,
Ganesan
S.
,
Meyn
M. S.
,
Timmers
C.
,
Hejna
J.
,
Grompe
M.
,
D'Andrea
A. D.
(
2001
).
Interaction of the Fanconi anemia proteins and BRCA1 in a common pathway
.
Mol. Cell
7
,
249
-
262
.
Geng
L.
,
Huntoon
C. J.
,
Karnitz
L. M.
(
2010
).
RAD18-mediated ubiquitination of PCNA activates the Fanconi anemia DNA repair network
.
J. Cell Biol.
191
,
249
-
257
.
Gieni
R. S.
,
Ismail
I. H.
,
Campbell
S.
,
Hendzel
M. J.
(
2011
).
Polycomb group proteins in the DNA damage response: a link between radiation resistance and “stemness”
.
Cell Cycle
10
,
883
-
894
.
Ginjala
V.
,
Nacerddine
K.
,
Kulkarni
A.
,
Oza
J.
,
Hill
S. J.
,
Yao
M.
,
Citterio
E.
,
van Lohuizen
M.
,
Ganesan
S.
(
2011
).
BMI1 is recruited to DNA breaks and contributes to DNA damage-induced H2A ubiquitination and repair
.
Mol. Cell. Biol.
31
,
1972
-
1982
.
Glickman
M. H.
,
Ciechanover
A.
(
2002
).
The ubiquitin-proteasome proteolytic pathway: destruction for the sake of construction
.
Physiol. Rev.
82
,
373
-
428
.
Hardeland
U.
,
Steinacher
R.
,
Jiricny
J.
,
Schär
P.
(
2002
).
Modification of the human thymine-DNA glycosylase by ubiquitin-like proteins facilitates enzymatic turnover
.
EMBO J.
21
,
1456
-
1464
.
Harreman
M.
,
Taschner
M.
,
Sigurdsson
S.
,
Anindya
R.
,
Reid
J.
,
Somesh
B.
,
Kong
S. E.
,
Banks
C. A.
,
Conaway
R. C.
,
Conaway
J. W.
, et al. 
. (
2009
).
Distinct ubiquitin ligases act sequentially for RNA polymerase II polyubiquitylation
.
Proc. Natl. Acad. Sci. USA
106
,
20705
-
20710
.
Hiom
K.
(
2010
).
Coping with DNA double strand breaks
.
DNA Repair
9
,
1256
-
1263
.
Hoege
C.
,
Pfander
B.
,
Moldovan
G. L.
,
Pyrowolakis
G.
,
Jentsch
S.
(
2002
).
RAD6-dependent DNA repair is linked to modification of PCNA by ubiquitin and SUMO
.
Nature
419
,
135
-
141
.
Hoeijmakers
J. H.
(
2009
).
DNA damage, aging, and cancer
.
N. Engl. J. Med.
361
,
1475
-
1485
.
Huang
J.
,
Huen
M. S.
,
Kim
H.
,
Leung
C. C.
,
Glover
J. N.
,
Yu
X.
,
Chen
J.
(
2009
).
RAD18 transmits DNA damage signalling to elicit homologous recombination repair
.
Nat. Cell Biol.
11
,
592
-
603
.
Huang
T. T.
,
D'Andrea
A. D.
(
2006
).
Regulation of DNA repair by ubiquitylation
.
Nat. Rev. Mol. Cell Biol.
7
,
323
-
334
.
Huang
T. T.
,
Nijman
S. M.
,
Mirchandani
K. D.
,
Galardy
P. J.
,
Cohn
M. A.
,
Haas
W.
,
Gygi
S. P.
,
Ploegh
H. L.
,
Bernards
R.
,
D'Andrea
A. D.
(
2006
).
Regulation of monoubiquitinated PCNA by DUB autocleavage
.
Nat. Cell Biol.
8
,
339
-
347
.
Huen
M. S.
,
Grant
R.
,
Manke
I.
,
Minn
K.
,
Yu
X.
,
Yaffe
M. B.
,
Chen
J.
(
2007
).
RNF8 transduces the DNA-damage signal via histone ubiquitylation and checkpoint protein assembly
.
Cell
131
,
901
-
914
.
Ismail
I. H.
,
Andrin
C.
,
McDonald
D.
,
Hendzel
M. J.
(
2010
).
BMI1-mediated histone ubiquitylation promotes DNA double-strand break repair
.
J. Cell Biol.
191
,
45
-
60
.
Joo
H. Y.
,
Zhai
L.
,
Yang
C.
,
Nie
S.
,
Erdjument-Bromage
H.
,
Tempst
P.
,
Chang
C.
,
Wang
H.
(
2007
).
Regulation of cell cycle progression and gene expression by H2A deubiquitination
.
Nature
449
,
1068
-
1072
.
Karras
G. I.
,
Jentsch
S.
(
2010
).
The RAD6 DNA damage tolerance pathway operates uncoupled from the replication fork and is functional beyond S phase
.
Cell
141
,
255
-
267
.
Kawabe
Y.
,
Seki
M.
,
Seki
T.
,
Wang
W. S.
,
Imamura
O.
,
Furuichi
Y.
,
Saitoh
H.
,
Enomoto
T.
(
2000
).
Covalent modification of the Werner's syndrome gene product with the ubiquitin-related protein, SUMO-1
.
J. Biol. Chem.
275
,
20963
-
20966
.
Kerscher
O.
,
Felberbaum
R.
,
Hochstrasser
M.
(
2006
).
Modification of proteins by ubiquitin and ubiquitin-like proteins
.
Annu. Rev. Cell Dev. Biol.
22
,
159
-
180
.
Kim
H.
,
Chen
J.
,
Yu
X.
(
2007
).
Ubiquitin-binding protein RAP80 mediates BRCA1-dependent DNA damage response
.
Science
316
,
1202
-
1205
.
Kolas
N. K.
,
Chapman
J. R.
,
Nakada
S.
,
Ylanko
J.
,
Chahwan
R.
,
Sweeney
F. D.
,
Panier
S.
,
Mendez
M.
,
Wildenhain
J.
,
Thomson
T. M.
, et al. 
. (
2007
).
Orchestration of the DNA-damage response by the RNF8 ubiquitin ligase
.
Science
318
,
1637
-
1640
.
Kratz
K.
,
Schopf
B.
,
Kaden
S.
,
Sendoel
A.
,
Eberhard
R.
,
Lademann
C.
,
Cannavo
E.
,
Sartori
A. A.
,
Hengartner
M. O.
,
Jiricny
J.
(
2010
).
Deficiency of FANCD2-associated nuclease KIAA1018/FAN1 sensitizes cells to interstrand crosslinking agents
.
Cell
142
,
77
-
88
.
Kvint
K.
,
Uhler
J. P.
,
Taschner
M. J.
,
Sigurdsson
S.
,
Erdjument-Bromage
H.
,
Tempst
P.
,
Svejstrup
J. Q.
(
2008
).
Reversal of RNA polymerase II ubiquitylation by the ubiquitin protease Ubp3
.
Mol. Cell
30
,
498
-
506
.
Leach
C. A.
,
Michael
W. M.
(
2005
).
Ubiquitin/SUMO modification of PCNA promotes replication fork progression in Xenopus laevis egg extracts
.
J. Cell Biol.
171
,
947
-
954
.
Lee
K. Y.
,
Yang
K.
,
Cohn
M. A.
,
Sikdar
N.
,
D'Andrea
A. D.
,
Myung
K.
(
2010
).
Human ELG1 regulates the level of ubiquitinated proliferating cell nuclear antigen (PCNA) through Its interactions with PCNA and USP1
.
J. Biol. Chem.
285
,
10362
-
10369
.
Lehmann
A. R.
,
Niimi
A.
,
Ogi
T.
,
Brown
S.
,
Sabbioneda
S.
,
Wing
J. F.
,
Kannouche
P. L.
,
Green
C. M.
(
2007
).
Translesion synthesis: Y-family polymerases and the polymerase switch
.
DNA Repair
6
,
891
-
899
.
Lin
J. R.
,
Zeman
M. K.
,
Chen
J. Y.
,
Yee
M. C.
,
Cimprich
K. A.
(
2011
).
SHPRH and HLTF act in a damage-specific manner to coordinate different forms of postreplication repair and prevent mutagenesis
.
Mol. Cell
42
,
237
-
249
.
Liu
T.
,
Ghosal
G.
,
Yuan
J.
,
Chen
J.
,
Huang
J.
(
2010
).
FAN1 acts with FANCI-FANCD2 to promote DNA interstrand cross-link repair
.
Science
329
,
693
-
696
.
Liu
Z.
,
Wu
J.
,
Yu
X.
(
2007
).
CCDC98 targets BRCA1 to DNA damage sites
.
Nat. Struct. Mol. Biol.
14
,
716
-
720
.
Lommel
L.
,
Ortolan
T.
,
Chen
L.
,
Madura
K.
,
Sweder
K. S.
(
2002
).
Proteolysis of a nucleotide excision repair protein by the 26 S proteasome
.
Curr. Genet.
42
,
9
-
20
.
MacKay
C.
,
Declais
A. C.
,
Lundin
C.
,
Agostinho
A.
,
Deans
A. J.
,
MacArtney
T. J.
,
Hofmann
K.
,
Gartner
A.
,
West
S. C.
,
Helleday
T.
, et al. 
. (
2010
).
Identification of KIAA1018/FAN1, a DNA repair nuclease recruited to DNA damage by monoubiquitinated FANCD2
.
Cell
142
,
65
-
76
.
Mailand
N.
,
Bekker-Jensen
S.
,
Faustrup
H.
,
Melander
F.
,
Bartek
J.
,
Lukas
C.
,
Lukas
J.
(
2007
).
RNF8 ubiquitylates histones at DNA double-strand breaks and promotes assembly of repair proteins
.
Cell
131
,
887
-
900
.
Marteijn
J. A.
,
Bekker-Jensen
S.
,
Mailand
N.
,
Lans
H.
,
Schwertman
P.
,
Gourdin
A. M.
,
Dantuma
N. P.
,
Lukas
J.
,
Vermeulen
W.
(
2009
).
Nucleotide excision repair-induced H2A ubiquitination is dependent on MDC1 and RNF8 and reveals a universal DNA damage response
.
J. Cell Biol.
186
,
835
-
847
.
Moldovan
G. L.
,
Pfander
B.
,
Jentsch
S.
(
2006
).
PCNA controls establishment of sister chromatid cohesion during S phase
.
Mol. Cell
23
,
723
-
732
.
Morris
J. R.
,
Boutell
C.
,
Keppler
M.
,
Densham
R.
,
Weekes
D.
,
Alamshah
A.
,
Butler
L.
,
Galanty
Y.
,
Pangon
L.
,
Kiuchi
T.
, et al. 
. (
2009
).
The SUMO modification pathway is involved in the BRCA1 response to genotoxic stress
.
Nature
462
,
886
-
890
.
Moyal
L.
,
Lerenthal
Y.
,
Gana-Weisz
M.
,
Mass
G.
,
So
S.
,
Wang
S. Y.
,
Eppink
B.
,
Chung
Y. M.
,
Shalev
G.
,
Shema
E.
, et al. 
. (
2011
).
Requirement of ATM-dependent monoubiquitylation of histone H2B for timely repair of DNA double-strand breaks
.
Mol. Cell
41
,
529
-
542
.
Nakada
S.
,
Tai
I.
,
Panier
S.
,
Al-Hakim
A.
,
Iemura
S.
,
Juang
Y. C.
,
O'Donnell
L.
,
Kumakubo
A.
,
Munro
M.
,
Sicheri
F.
, et al. 
. (
2010
).
Non-canonical inhibition of DNA damage-dependent ubiquitination by OTUB1
.
Nature
466
,
941
-
946
.
Nicassio
F.
,
Corrado
N.
,
Vissers
J. H.
,
Areces
L. B.
,
Bergink
S.
,
Marteijn
J. A.
,
Geverts
B.
,
Houtsmuller
A. B.
,
Vermeulen
W.
,
Di Fiore
P. P.
, et al. 
. (
2007
).
Human USP3 is a chromatin modifier required for S phase progression and genome stability
.
Curr. Biol.
17
,
1972
-
1977
.
Ogi
T.
,
Limsirichaikul
S.
,
Overmeer
R. M.
,
Volker
M.
,
Takenaka
K.
,
Cloney
R.
,
Nakazawa
Y.
,
Niimi
A.
,
Miki
Y.
,
Jaspers
N. G.
, et al. 
. (
2010
).
Three DNA polymerases, recruited by different mechanisms, carry out NER repair synthesis in human cells
.
Mol. Cell
37
,
714
-
727
.
Okubo
S.
,
Hara
F.
,
Tsuchida
Y.
,
Shimotakahara
S.
,
Suzuki
S.
,
Hatanaka
H.
,
Yokoyama
S.
,
Tanaka
H.
,
Yasuda
H.
,
Shindo
H.
(
2004
).
NMR structure of the N-terminal domain of SUMO ligase PIAS1 and its interaction with tumor suppressor p53 and A/T-rich DNA oligomers
.
J. Biol. Chem.
279
,
31455
-
31461
.
Pages
V.
,
Bresson
A.
,
Acharya
N.
,
Prakash
S.
,
Fuchs
R. P.
,
Prakash
L.
(
2008
).
Requirement of Rad5 for DNA polymerase zeta-dependent translesion synthesis in Saccharomyces cerevisiae
.
Genetics
180
,
73
-
82
.
Panier
S.
,
Durocher
D.
(
2009
).
Regulatory ubiquitylation in response to DNA double-strand breaks
.
DNA Repair
8
,
436
-
443
.
Papouli
E.
,
Chen
S.
,
Davies
A. A.
,
Huttner
D.
,
Krejci
L.
,
Sung
P.
,
Ulrich
H. D.
(
2005
).
Crosstalk between SUMO and ubiquitin on PCNA is mediated by recruitment of the helicase Srs2p
.
Mol. Cell
19
,
123
-
133
.
Park
H. K.
,
Wang
H.
,
Zhang
J.
,
Datta
S.
,
Fei
P.
(
2010
).
Convergence of Rad6/Rad18 and Fanconi anemia tumor suppressor pathways upon DNA damage
.
PLoS ONE
5
,
e13313
.
Parnas
O.
,
Zipin-Roitman
A.
,
Pfander
B.
,
Liefshitz
B.
,
Mazor
Y.
,
Ben-Aroya
S.
,
Jentsch
S.
,
Kupiec
M.
(
2010
).
Elg1, an alternative subunit of the RFC clamp loader, preferentially interacts with SUMOylated PCNA
.
EMBO J.
29
,
2611
-
2622
.
Pfander
B.
,
Moldovan
G. L.
,
Sacher
M.
,
Hoege
C.
,
Jentsch
S.
(
2005
).
SUMO-modified PCNA recruits Srs2 to prevent recombination during S phase
.
Nature
436
,
428
-
433
.
Potts
P. R.
,
Yu
H.
(
2005
).
Human MMS21/NSE2 is a SUMO ligase required for DNA repair
.
Mol. Cell. Biol.
25
,
7021
-
7032
.
Prudden
J.
,
Pebernard
S.
,
Raffa
G.
,
Slavin
D. A.
,
Perry
J. J.
,
Tainer
J. A.
,
McGowan
C. H.
,
Boddy
M. N.
(
2007
).
SUMO-targeted ubiquitin ligases in genome stability
.
EMBO J.
26
,
4089
-
4101
.
Ribar
B.
,
Prakash
L.
,
Prakash
S.
(
2007
).
ELA1 and CUL3 are required along with ELC1 for RNA polymerase II polyubiquitylation and degradation in DNA-damaged yeast cells
.
Mol. Cell. Biol.
27
,
3211
-
3216
.
Russell
S. J.
,
Reed
S. H.
,
Huang
W.
,
Friedberg
E. C.
,
Johnston
S. A.
(
1999
).
The 19S regulatory complex of the proteasome functions independently of proteolysis in nucleotide excision repair
.
Mol. Cell
3
,
687
-
695
.
Sacher
M.
,
Pfander
B.
,
Hoege
C.
,
Jentsch
S.
(
2006
).
Control of Rad52 recombination activity by double-strand break-induced SUMO modification
.
Nat. Cell Biol.
8
,
1284
-
1290
.
Sims
A. E.
,
Spiteri
E.
,
Sims
R. J.
3rd
,
Arita
A. G.
,
Lach
F. P.
,
Landers
T.
,
Wurm
M.
,
Freund
M.
,
Neveling
K.
,
Hanenberg
H.
, et al. 
. (
2007
).
FANCI is a second monoubiquitinated member of the Fanconi anemia pathway
.
Nat. Struct. Mol. Biol.
14
,
564
-
567
.
Smogorzewska
A.
,
Matsuoka
S.
,
Vinciguerra
P.
,
McDonald
E. R.
3rd
,
Hurov
K. E.
,
Luo
J.
,
Ballif
B. A.
,
Gygi
S. P.
,
Hofmann
K.
,
D'Andrea
A. D.
, et al. 
. (
2007
).
Identification of the FANCI protein, a monoubiquitinated FANCD2 paralog required for DNA repair
.
Cell
129
,
289
-
301
.
Smogorzewska
A.
,
Desetty
R.
,
Saito
T. T.
,
Schlabach
M.
,
Lach
F. P.
,
Sowa
M. E.
,
Clark
A. B.
,
Kunkel
T. A.
,
Harper
J. W.
,
Colaiacovo
M. P.
, et al. 
. (
2010
).
A genetic screen identifies FAN1, a Fanconi anemia-associated nuclease necessary for DNA interstrand crosslink repair
.
Mol. Cell
39
,
36
-
47
.
Sobhian
B.
,
Shao
G.
,
Lilli
D. R.
,
Culhane
A. C.
,
Moreau
L. A.
,
Xia
B.
,
Livingston
D. M.
,
Greenberg
R. A.
(
2007
).
RAP80 targets BRCA1 to specific ubiquitin structures at DNA damage sites
.
Science
316
,
1198
-
1202
.
Song
I. Y.
,
Palle
K.
,
Gurkar
A.
,
Tateishi
S.
,
Kupfer
G. M.
,
Vaziri
C.
(
2010
).
Rad18-mediated translesion synthesis of bulky DNA adducts is coupled to activation of the Fanconi anemia DNA repair pathway
.
J. Biol. Chem.
285
,
31525
-
31536
.
Steinacher
R.
,
Schar
P.
(
2005
).
Functionality of human thymine DNA glycosylase requires SUMO-regulated changes in protein conformation
.
Curr. Biol.
15
,
616
-
623
.
Stelter
P.
,
Ulrich
H. D.
(
2003
).
Control of spontaneous and damage-induced mutagenesis by SUMO and ubiquitin conjugation
.
Nature
425
,
188
-
191
.
Stewart
G. S.
,
Panier
S.
,
Townsend
K.
,
Al-Hakim
A. K.
,
Kolas
N. K.
,
Miller
E. S.
,
Nakada
S.
,
Ylanko
J.
,
Olivarius
S.
,
Mendez
M.
, et al. 
. (
2009
).
The RIDDLE syndrome protein mediates a ubiquitin-dependent signaling cascade at sites of DNA damage
.
Cell
136
,
420
-
434
.
Sugasawa
K.
(
2010
).
Regulation of damage recognition in mammalian global genomic nucleotide excision repair
.
Mutat Res.
685
,
29
-
37
.
Sugasawa
K.
,
Okuda
Y.
,
Saijo
M.
,
Nishi
R.
,
Matsuda
N.
,
Chu
G.
,
Mori
T.
,
Iwai
S.
,
Tanaka
K.
,
Hanaoka
F.
(
2005
).
UV-induced ubiquitylation of XPC protein mediated by UV-DDB-ubiquitin ligase complex
.
Cell
121
,
387
-
400
.
Sun
H.
,
Leverson
J. D.
,
Hunter
T.
(
2007
).
Conserved function of RNF4 family proteins in eukaryotes: targeting a ubiquitin ligase to SUMOylated proteins
.
EMBO J.
26
,
4102
-
4112
.
Svejstrup
J. Q.
(
2010
).
The interface between transcription and mechanisms maintaining genome integrity
.
Trends Biochem. Sci.
35
,
333
-
338
.
Takahashi
H.
,
Hatakeyama
S.
,
Saitoh
H.
,
Nakayama
K. I.
(
2005
).
Noncovalent SUMO-1 binding activity of thymine DNA glycosylase (TDG) is required for its SUMO-1 modification and colocalization with the promyelocytic leukemia protein
.
J. Biol. Chem.
280
,
5611
-
5621
.
Tang
J. B.
,
Greenberg
R. A.
(
2010
).
Connecting the dots: interplay between ubiquitylation and SUMOylation at DNA double strand breaks
.
Genes Cancer
1
,
787
-
796
.
Trickey
M.
,
Grimaldi
M.
,
Yamano
H.
(
2008
).
The anaphase-promoting complex/cyclosome controls repair and recombination by ubiquitylating Rhp54 in fission yeast
.
Mol. Cell. Biol.
28
,
3905
-
3916
.
Ulrich
H. D.
(
2009
).
Regulating post-translational modifications of the eukaryotic replication clamp PCNA
.
DNA Repair
8
,
461
-
469
.
Ulrich
H. D.
,
Walden
H.
(
2010
).
Ubiquitin signalling in DNA replication and repair
.
Nat. Rev. Mol. Cell Biol.
11
,
479
-
489
.
Uzunova
K.
,
Gottsche
K.
,
Miteva
M.
,
Weisshaar
S. R.
,
Glanemann
C.
,
Schnellhardt
M.
,
Niessen
M.
,
Scheel
H.
,
Hofmann
K.
,
Johnson
E. S.
, et al. 
. (
2007
).
Ubiquitin-dependent proteolytic control of SUMO conjugates
.
J. Biol. Chem.
282
,
34167
-
34175
.
Wang
B.
,
Matsuoka
S.
,
Ballif
B. A.
,
Zhang
D.
,
Smogorzewska
A.
,
Gygi
S. P.
,
Elledge
S. J.
(
2007
).
Abraxas and RAP80 form a BRCA1 protein complex required for the DNA damage response
.
Science
316
,
1194
-
1198
.
Wang
W.
(
2007
).
Emergence of a DNA-damage response network consisting of Fanconi anaemia and BRCA proteins
.
Nat. Rev. Genet.
8
,
735
-
748
.
Weake
V. M.
,
Workman
J. L.
(
2008
).
Histone ubiquitination: triggering gene activity
.
Mol. Cell
29
,
653
-
663
.
Wotton
D.
,
Merrill
J. C.
(
2007
).
Pc2 and SUMOylation
.
Biochem. Soc. Trans.
35
,
1401
-
1404
.
Xie
Y.
,
Kerscher
O.
,
Kroetz
M. B.
,
McConchie
H. F.
,
Sung
P.
,
Hochstrasser
M.
(
2007
).
The yeast Hex3-Slx8 heterodimer is a ubiquitin ligase stimulated by substrate sumoylation
.
J. Biol. Chem.
282
,
34176
-
34184
.
Zhao
X.
,
Blobel
G.
(
2005
).
A SUMO ligase is part of a nuclear multiprotein complex that affects DNA repair and chromosomal organization
.
Proc. Natl. Acad. Sci. USA
102
,
4777
-
4782
.