Hostname: page-component-848d4c4894-pftt2 Total loading time: 0 Render date: 2024-05-02T03:55:49.540Z Has data issue: false hasContentIssue false

Mismatch repair is a double-edged sword in the battle against microsatellite instability

Published online by Cambridge University Press:  05 September 2022

Carson J. Miller
Affiliation:
Laboratory of Cell and Molecular Biology, National Institute of Diabetes, Digestive and Kidney Diseases, National Institutes of Health, Bethesda, MD, USA
Karen Usdin*
Affiliation:
Laboratory of Cell and Molecular Biology, National Institute of Diabetes, Digestive and Kidney Diseases, National Institutes of Health, Bethesda, MD, USA
*
Author for correspondence: Karen Usdin, E-mail: karenu@nih.gov
Rights & Permissions [Opens in a new window]

Abstract

Roughly 3% of the human genome consists of microsatellites or tracts of short tandem repeats (STRs). These STRs are often unstable, undergoing high-frequency expansions (increases) or contractions (decreases) in the number of repeat units. Some microsatellite instability (MSI) is seen at multiple STRs within a single cell and is associated with certain types of cancer. A second form of MSI is characterised by expansion of a single gene-specific STR and such expansions are responsible for a group of 40+ human genetic disorders known as the repeat expansion diseases (REDs). While the mismatch repair (MMR) pathway prevents genome-wide MSI, emerging evidence suggests that some MMR factors are directly involved in generating expansions in the REDs. Thus, MMR suppresses some forms of expansion while some MMR factors promote expansion in other contexts. This review will cover what is known about the paradoxical effect of MMR on microsatellite expansion in mammalian cells.

Type
Review
Creative Commons
Creative Common License - CCCreative Common License - BY
This is a work of the US Government and is not subject to copyright protection within the United States. Published by Cambridge University Press
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © National Institutes of Health, 2022

Introduction

Tracts containing tandem arrays of short perfect- or near-perfect repeat units are common in the human genome (Ref. Reference Subramanian1). These short tandem repeats (STRs), or microsatellites, consist of repeat units that are generally ~1–6 nucleotides long. The STRs are found in promoters, exons, introns, as well as in intergenic regions where they can impact gene expression in a myriad of different ways, including affecting promoter activity, RNA polymerase processivity, splicing, translation rates and protein function (Ref. Reference Usdin2). Many STRs are polymorphic, giving rise to expansions, or increases in the number of repeat units; as well as contractions, or loss of repeat units. Such STRs are sometimes referred to as variable number tandem repeats (VNTRs). STRs can be a significant source of human genetic variation and the instability of some of these tracts can have biological consequences because of their intrinsic effects on gene expression (Refs Reference Usdin2, Reference Marshall3). In addition, many of these sequences form secondary structures that are thought to make them difficult to replicate (Refs Reference Usdin and Woodford4Reference Shastri11). This can result in the generation of chromosome abnormalities of different kinds (reviewed in Ref. Reference Lokanga, Kumari and Usdin12).

Two major classes of human STR expansions are known: the first class is associated with genome-wide microsatellite instability (MSI), while the second class is associated with expansions of a specific microsatellite. Genome-wide MSI is associated with a predisposition to certain cancers including haematological malignancies as well as certain colon, urothelial, hepatobiliary, pancreatic, bladder, kidney, prostate, endometrial, ovarian and breast cancers (Refs Reference Grindedal13Reference Win20). In contrast, the locus-specific expansions define the repeat expansion disorders (REDs), a group of 40+ human genetic disorders that are primarily neurological or neurodevelopmental in nature. Diseases in this group include Huntington's disease (HD), caused by expansion of CAG/CTG-STR in the first exon of the huntingtin (HTT) gene (Ref. 21); Friedreich ataxia (FRDA), caused by a GAA/TTC-STR in intron 1 of the frataxin (FXN) gene (Ref. Reference Campuzano22); myotonic dystrophy type 1 (DM1), caused by expansion of a CTG/CAG-STR in the 3′ UTR of the DMPK gene (Refs Reference Brook23, Reference Mahadevan24); and the Fragile X-related disorders (FXDs), resulting from expansion of a CGG/CCG-STR in the 5′ UTR of the FMR1 gene (Refs Reference Verkerk25Reference Brunberg28). While it was initially thought that STR expansions in the REDs occurred by a mechanism similar to cancer-associated MSI, emerging evidence suggests these two types of STR expansions have completely different molecular mechanisms. Genome-wide MSI results from errors arising during DNA synthesis that normally would be repaired by the MMR machinery, that is, MMR factors all act anti-mutagenically at these loci to suppress expansions. In contrast, STR expansions in the REDs actually require certain components of the MMR machinery, that is, these MMR factors can also act pro-mutagenically. While work in model systems suggests that other mechanisms of STR expansion may be possible (e.g. Ref. Reference Kononenko29), this review will focus on what is currently known about STR expansions arising from either the pro- or anti-mutagenic roles of MMR.

Cancer-associated MSI

During DNA replication two types of errors can be introduced into DNA: mismatches and insertions or deletions (INDELs). Mismatches arise from insertion of the incorrect base in the daughter strand by the DNA polymerase. Most of these mismatches are removed by the proofreading function of the polymerase, but those that escape this proofreading will cause point mutations if the daughter strand is replicated before the mismatch is removed (Refs Reference Kunkel and Erie30, Reference Lujan31). INDELs are thought to result primarily from the dissociation of the DNA polymerase from the template thus creating an opportunity for two strands of DNA to slip relative to one another. Dissociation might be exacerbated by an encounter with impediments to replication fork progression such as those formed by unusual DNA structures, strongly bound proteins or collisions with a transcription complex, while mispriming may be favoured when one of the DNA strands forms a stable secondary structure (Ref. Reference Gadgil32). When strand-slippage occurs within a STR, out-of-register reannealing can occur with priming from the slipped position as illustrated in Figure 1. This results in either a loop out of the template strand or a loop out on the nascent strand depending on whether reannealing occurs 5′ or 3′ on the template. Failure to remove the loop out leads to expansions if the loop out is in the nascent strand and contractions if it is in the template strand (Refs Reference Ellegren33Reference Streisinger36).

Fig. 1. A model for MSI in which loop outs generated by strand-slippage of the nascent strand during DNA replication escape MMR and result in the incorporation of a small number of additional units, indicated by the green boxes, or the loss of a small number of repeat units, indicated by the dotted triangle. Whether repeats are gained or lost depends on whether repriming occurs further 3′ on the template resulting in nascent strand loop outs or further 5′ on the template resulting in loop outs being formed on the template strand.

Repair of these replication errors is carried out primarily by the MMR machinery that travels behind the replication complex. In eukaryotes, recognition of mismatches and INDELs during MMR is accomplished by either of two MutS complexes, both of which are heterodimers of homologous proteins: MSH2/MSH6 in the MutSα complex and MSH2/MSH3 in the MutSβ complex. MutSα is primarily involved in the recognition of mismatches and 1 base INDELs (Refs Reference Drummond37Reference Warren39). MutSβ, on the other hand, is involved primarily in the repair of larger loops (Refs Reference Gupta, Gellert and Yang40Reference Palombo42). (A third MutS complex found in mammals, MutSγ, is a MSH4/MSH5 heterodimer that functions almost exclusively in meiotic crossover resolution (Ref. Reference Snowden43).) After lesion binding, MutSα and MutSβ recruit a member of the MutL family of proteins. Mammals have three different MutL complexes. Like the MutS complexes, each of the MutL complexes are heterodimers; consisting of MLH1 bound to either PMS2, MLH3 or PMS1 to form MutLα, MutLγ or MutLβ, respectively (Refs Reference Flores-Rozas and Kolodner44Reference Pluciennik46). MutLα is the most abundant of the MutL complexes and is responsible for most MMR. MutLγ plays a minor role in MMR, primarily in cooperation with MutSβ (Refs Reference Cannavo47, Reference Roesner48). MutLα and MutLγ are nucleases that introduce nicks into the MMR template, a critical step in the repair process. The role of MutLβ in MMR is unclear (Refs Reference Prolla49, Reference Räschle50). While PMS1 is more abundant than MLH3, it lacks the DQHA(X)2E(X)4E nuclease motif present in both PMS2 and MLH3 and its loss is not associated with increased MSI in mice (Ref. Reference Prolla49). Multiple lines of evidence suggest that after initial mismatch binding by MutSα, additional MutSα complexes load onto the DNA, followed by recruitment of multiple MutLα proteins (Refs Reference Bradford51, Reference Hombauer52). A similar situation may apply to MutSβ-directed recruitment of MutLγ, since formation of MutLγ polymers on the mismatch template has also been shown to be important for proper MutLγ-mediated repair (Ref. Reference Manhart53). Excision of the nicked strand is carried out by a 5′ to 3′ exonuclease such as exonuclease 1 (EXO1) (Refs Reference Wei54, Reference Genschel, Bazemore and Modrich55) or FAN1 (Ref. Reference Kratz56). Strand-displacement synthesis by Pol δ can also remove the nicked strand. This is followed by repair synthesis by Pol δ, with sealing of the remaining nick by DNA ligase I to complete the repair process.

While this process is relatively efficient, strand-slippage occurs so frequently that some MSI occurs even in the presence of the normal MMR machinery. Mean rates of ~10−5 to 10−7 MSI events per locus per cell generation have been reported in human cells with functional MMR, orders of magnitude higher than the mutation rate seen in unique sequence (Refs Reference Hile, Yan and Eckert57Reference Hatch and Farber59). Loss of MMR results in rates of MSI that can be 2–3 orders of magnitude higher (Ref. Reference Christopher60). The wide variation in mutation rates of different STRs is related in part to the size of the repeat units, their sequence composition and the size and purity of the repeat tract (Refs Reference Eckert and Hile61, Reference Campregher62). The likelihood of instability at a specific microsatellite is also related to the normal target of the dysfunctional MMR gene. So, mutations in MSH2, MLH1 and PMS2 increase instability of microsatellites containing mononucleotide, dinucleotide and tetranucleotide repeat units; MSH6 mutations affect microsatellites with mononucleotide and some dinucleotide repeat units; and MSH3 mutations affect dinucleotide and tetranucleotide containing repeat units, but not ones consisting of mononucleotide repeat units.

More than 90% of MSI events involve the gain or loss of a single repeat unit with a very limited number of mutations involving multiple units (Ref. Reference Brinkmann63). MSI often exhibits an expansion bias (Refs Reference Ellegren64Reference Twerdi, Boyer and Farber67). This bias is reduced at very large microsatellites (Refs Reference Ellegren64, Reference Sun65), perhaps reflecting the formation of stable secondary structures and the resultant difficulties associated with replication of the region. This could result in a dependency on proteins such as the Werner's syndrome helicase (WRN) to remove the secondary structure thus allowing replication to proceed (Ref. Reference van Wietmarschen68). It has also been suggested that these structures promote error-prone DNA synthesis resulting in mutations that affect the purity of the repeat tract (Ref. Reference Murat, Guilbaud and Sale10). This in turn would reduce the likelihood of further expansion. While most MSI events of this kind involve a single repeat unit, MSI with an expansion bias could over time result in the large microsatellites that accumulate in cancer cell lines such as HCT116 and KM12 that lack MLH1 (as well as MSH3 in the case of HCT116) (Ref. Reference van Wietmarschen68).

MSI in the REDs

In contrast, studies in RED patient cohorts using genome-wide association (GWA) or the testing of candidate MMR gene polymorphisms suggest that functional MMR components are required for some, if not all, STR expansions that cause the REDs (Refs 69Reference Hwang74). This is consistent with evidence from mouse and human cell models of a number of these disorders that shows a requirement for MutSβ and MutLγ (reviewed in Ref. Reference Zhao75). Canonical MMR per se is unlikely to be responsible for these expansions since mutations in EXO1 and FAN1, 5′-3′ exonucleases that act downstream of the MutS and MutL proteins in normal MMR, protect against expansion in mouse or human tissue culture models (Refs Reference McAllister73, Reference Zhao and Usdin76Reference Goold79), and GWA studies data are consistent with a protective role for FAN1 in reducing expansions in humans (Refs 69, Reference McAllister73). Furthermore, Lig4, the ligase required for non-homologous end joining (NHEJ), a form of double-strand break repair (DSBR), also protects against expansion in a mouse model of the FXDs (Ref. Reference Gazy80). This suggests that NHEJ competes with the expansion pathway for access to a common DSB intermediate.

As with cancer-associated MSI, the extent of expansion in the REDs is related in part to the length and purity of the repeat tract (Refs Reference Wright81Reference Nolin88). Mathematical modelling of human somatic expansions and empirical observations of both germline and somatic expansions over time in mice support the idea that most expansion events involve the addition of 1–2 repeat units (Refs Reference Møllersen89, Reference Zhao and Usdin90). As with expansions arising in MMR-deficient cells, this can result in the production of much larger alleles over time. However, in some cell types the MMR factor-dependent expansions occur at frequencies that are orders of magnitude higher than the MSI occurring in the absence of MMR at the same locus (Ref. Reference Miller91).

The mechanism responsible for this high-frequency expansion process is not fully understood. Clues to what this process may be include the fact that expansions can occur in post-mitotic cells such as oocytes and neurons (Refs Reference Zhao, Lu and Usdin78, Reference Yrigollen92Reference Swami95). Thus, these events can be independent of chromosomal replication. Furthermore, the fact that the STR in the X-linked FMR1 gene that causes the FXDs only expands when it is on the active X chromosome indicates that STR expansion requires transcription or transcriptionally-competent chromatin (Ref. Reference Lokanga96). A role for oxidative damage is suggested by the fact that the loss of OGG1 and NEIL1, DNA glycosylases involved in the base excision repair of oxidative damage, decreases the expansion frequency in a HD mouse model (Refs Reference Kovtun97, Reference Møllersen98) and exogenous sources of oxidative stress increase the expansion frequency in some mouse and tissue culture models (Refs Reference Entezam99, Reference Jonson100). However, antioxidants only have a minimal effect on expansion (Refs Reference Møllersen101, Reference Gomes-Pereira and Monckton102) and even in the absence of OGG1 and NEIL1 many expansions are still seen (Refs Reference Kovtun97, Reference Møllersen98). Thus, endogenous oxidative stress may not be the only trigger for the expansion process or even the most important one.

Although MutLγ is the least abundant of all the MutL complexes, its nuclease activity is required for repeat expansion (Refs Reference Hayward, Steinbach and Usdin103Reference Roy105). Thus, expansion either involves a substrate that is bound preferentially by MutSβ/MutLγ or MutLγ cleavage is uniquely able to generate an intermediate that can be processed to generate an expansion. Interestingly, MutLγ has been shown to cut the DNA strand opposite to the mismatch in vitro (Ref. Reference Kadyrova106) and MutLγ is required during meiosis for the resolution of Holliday junctions (HJs) (Ref. Reference Zakharyevich107). Loop outs formed within the STR by both DNA strands might resemble such a four-way junction. Such structures could potentially arise any time the repeat tract was unpaired since out-of-register reannealing could occur particularly if one or both strands formed stable secondary structures as many STRs do (reviewed in Refs Reference Zhao75, Reference Mirkin108). Cleavage of the opposite strand at each of the loop outs could then result in a staggered DSB. Interestingly, we have shown that EXO1, which plays a structural role in determining the orientation of cleavage of HJs (Ref. Reference Cannavo47), also plays a nuclease-independent role in protecting against repeat expansion (Ref. Reference Zhao77). This raises the possibility that cleavage of the expansion intermediate may result in a DSB that is prone to expand and one that is not. A model for repeat expansion that accommodates these observations is shown in Figure 2. In this model expansions arise when out-of-register reannealing of the DSB occurs. This leaves a gap of a small number of repeat units that is then repaired by gap-filling. The net effect is that a small number of repeats are added to the repaired allele.

Fig. 2. A model for repeat expansions and contraction in the REDs in which loop outs are formed on one or both strands during transcription or at other times that the DNA was unpaired. These loop outs are then bound by MutS and MutL proteins. Cleavage by MutLγ results in the formation of a staggered DSB with 5′ overhangs. Out-of-register reannealing of the DSB can produce a substrate for simple gap filling which results in the addition of repeat units. Exonucleolytic processing of the DSB can result in products with shorter 5′ overhangs or blunt ends. These products may then be processed, perhaps by NHEJ or gap-filling, to generate the loss of repeat units as indicated by the dotted triangle. The extent of contraction would depend on the amount of exonucleolytic cleavage that occurs prior to repair.

In addition to MutSβ and MutLγ, MutLβ has also been shown to be required for expansion in embryonic stem cells from a mouse FXD model (Ref. Reference Latham87). An active role for MutLβ in generating expansions in HD is also suggested by the fact that PMS1 variants predicted to be deleterious are most frequently associated with a later age at onset/less severe phenotype in HD cohorts (Ref. Reference McAllister73). However, since MutLβ lacks a nuclease and is not required for MMR, how it contributes to expansions is unclear. This is not the only unresolved issue; although MutSα contributes to expansions in FXD and FRDA mice and FRDA iPSCs (Refs Reference Lujan31, Reference Bourn109, Reference Zhao110), little if any effect of the loss of MSH6 was seen in DM1 mouse model (Ref. Reference Foiry111) or in a human cell model system of FRDA (Ref. Reference Halabi112). Furthermore, reducing MutLα levels also has different effects in various systems. In a mouse model of the FXDs, MutLα is required for expansions in ESCs (Ref. Reference Latham87), while in a mouse model of DM1 loss of MutLα only resulted in a 50% decrease in expansions (Ref. Reference Gomes-Pereira113). Reduced PMS2 caused no change in the expansion frequency in a human cell model of FRDA (Ref. Reference Jonson100), whereas a mouse model of FRDA lacking Pms2 showed an increase in expansions (Ref. Reference Bourn109). The latter is consistent with the observation that a missense mutation in PMS2 correlates with an earlier age at onset in HD (Ref. 69). These differences do not necessarily mean that the expansion mechanisms in these diseases are fundamentally different. Since multiple MutS and MutL complexes are involved in binding to a mismatch, a case can be made that MutSα and MutLα are able to act in an auxiliary capacity to promote expansions when the essential factors, such as MutSβ and MutLγ, and perhaps MutLβ, are limiting (Ref. Reference Zhao75).

Contractions are also seen in the REDs-associated STRs and their mouse models, although their aetiology is less well understood. A bimodal distribution of contractions is seen in the germline of a mouse model of the FXDs, with some contractions involving the loss of just 1–2 repeat units whilst others involve the loss of much larger numbers of repeat units (Ref. Reference Zhao114). These larger contractions are sometimes difficult to discern in somatic cells because, unlike expansions, the contraction products do not seem to fall into a single size class. Curiously, loss of MSH3 results in a decrease not only in expansions, but also in the number of large contractions that are observed, with the decrease in these events being associated with a corresponding increase in the number of small contractions (Ref. Reference Zhao114). One model consistent with these observations is that large contractions represent a second possible outcome of the events that give rise to expansions, with contractions arising from DSBs that undergo some exonucleolytic cleavage prior to DSBR as illustrated in Figure 2.

Concluding remarks

Thus, MMR factors can both suppress and promote MSI. Which MSI pathway predominates is likely to depend on a variety of cell-type specific factors including the frequency of cell division and the relative levels of factors that promote or suppress each type of MSI. For example, while MMR is important for preventing genome-wide MSI in the colon, as evidenced by the high frequency of MSI-high colonic tumours in individuals with germline mutations in MMR proteins, MSI-high tumours originating in neurons are rare (Refs Reference Aarnio115Reference Guerrini-Rousseau117). The low level of MSI-high tumours in neurons may reflect in part the fact that neurons are non-dividing and thus likely to rarely generate the substrates for the MMR pathway. The high tumour incidence in the colon might reflect the consequences of exposure to dietary mutagens in rapidly dividing cells. In contrast, repeat expansion in neurons, particularly striatal neurons, occurs at high frequency in both mouse models of REDs and in REDs patients (Refs Reference Shelbourne118Reference Gonitel120). The high frequency of STR expansion in neurons of REDs patients may reflect the high levels of factors such as MutSβ, OGG1 and NEIL1 that promote expansion, and low levels of proteins such as EXO1, that protect against them (Ref. Reference Zhao77). The high levels of transcription of the affected genes in neurons may also contribute to the incidence of these expansions, by increasing the opportunity for formation of the substrates upon which the expansion process acts.

The paradoxical effect of MMR proteins on MSI is particularly apparent at the Fmr1 locus in embryonic stem cells from a mouse model of the FXDs. Consistent with the role of functional MMR components in generating expansions, a high rate of expansions is seen at this locus in MMR-proficient cells derived from these mice. For example, most alleles with ~280 repeat units gained an additional 19 repeats over a 52-day period in cells wildtype for MLH3 (Ref. Reference Latham87). In contrast, in cells with a similar repeat number that lacked MLH3, the modal size of the allele actually decreased by one repeat over the same period. Thus, the same microsatellite expands rapidly in MMR-proficient cells and contracts more slowly in MMR-deficient ones. The fact that MMR-deficiency results in contractions at this locus rather than expansions serves to emphasise the fundamentally different events occurring at this locus.

In the case of MSI-high tumours, the expanded microsatellites themselves might be a source of vulnerability that could be exploited for therapeutic purposes. Since expansion results in a dependence on DNA helicases such as WRN (Ref. Reference van Wietmarschen68), it may be possible to selectively eliminate the cancer cells using a synthetic lethal approach that targets these enzymes (Ref. Reference Chan121). In the case of many REDs, a growing body of evidence suggests that somatic expansion of the disease-associated STR significantly worsens the age at onset and/or disease severity (Refs 69, Reference Bettencourt70, Reference Flower72, Reference McAllister73). Since most of these diseases are severely life-limiting and lack any effective treatment or cure, there is an interest in exploring ways to reduce this MSI in somatic cells. This approach has additional appeal in that any success in this regard would be relevant to multiple REDs. Of course, given the requirement of many MMR factors for protecting against cancer, targeting these factors to reduce expansion poses a challenge. However, MSH3 and MLH3 are not major players in MMR and may thus be acceptable targets particularly for those diseases with a high early mortality. For example, tail vein injection of a splice switching oligonucleotide that favours the production of an MLH3 isoform lacking the nuclease domain has already been shown to reduce expansion in some peripheral tissues of a mouse model of HD (Ref. Reference Møllersen101). Since the absence of PMS1 is not associated with tumour susceptibility or any other obvious phenotype in mice (Ref. Reference Prolla49), depletion of this factor may be an even more attractive approach. While delivery of therapeutics to deep brain regions such as the striatum is not trivial, implanted intracerebroventricular devices have been used successfully for decades to deliver chemotherapeutic agents to treat central nervous system malignancies (Ref. Reference Triarico122). This experience could perhaps be leveraged for the ongoing delivery of MMR-targeting molecules to treat REDs.

Financial support

This work was made possible by funding from the Intramural Program of the National Institute of Diabetes, Kidney and Digestive Diseases to KU (DK057808).

Conflict of interest

None.

References

Subramanian, S et al. (2003) SSRD: simple sequence repeats database of the human genome. Comparative and Functional Genomics 4, 342345.CrossRefGoogle ScholarPubMed
Usdin, K (2008) The biological effects of simple tandem repeats: lessons from the repeat expansion diseases. Genome Research 18, 10111019.Google ScholarPubMed
Marshall, JN et al. (2021) Variable number tandem repeats – their emerging role in sickness and health. Experimental Biology and Medicine (Maywood) 246, 13681376.Google ScholarPubMed
Usdin, K and Woodford, KJ (1995) CGG repeats associated with DNA instability and chromosome fragility form structures that block DNA synthesis in vitro. Nucleic Acids Research 23, 42024209.Google ScholarPubMed
Yudkin, D et al. (2014) Chromosome fragility and the abnormal replication of the FMR1 locus in fragile X syndrome. Human Molecular Genetics 23, 29402952.Google ScholarPubMed
Voineagu, I et al. (2009) Replisome stalling and stabilization at CGG repeats, which are responsible for chromosomal fragility. Nature Structural & Molecular Biology 16, 226228.CrossRefGoogle ScholarPubMed
Krasilnikova, MM and Mirkin, SM (2004) Replication stalling at Friedreich's ataxia (GAA)n repeats in vivo. Molecular and Cellular Biology 24, 22862295.Google ScholarPubMed
Follonier, C et al. (2013) Friedreich's ataxia-associated GAA repeats induce replication-fork reversal and unusual molecular junctions. Nature Structural & Molecular Biology 20, 486494.Google ScholarPubMed
Thys, RG and Wang, YH (2015) DNA Replication dynamics of the GGGGCC repeat of the C9orf72 gene. Journal of Biological Chemistry 290, 2895328962.Google ScholarPubMed
Murat, P, Guilbaud, G and Sale, JE (2020) DNA polymerase stalling at structured DNA constrains the expansion of short tandem repeats. Genome Biology 21, 209.CrossRefGoogle ScholarPubMed
Shastri, N et al. (2018) Genome-wide identification of structure-forming repeats as principal sites of fork collapse upon ATR inhibition. Molecular Cell 72, 222238, e11.CrossRefGoogle ScholarPubMed
Lokanga, RA, Kumari, D and Usdin, K (2021) Common threads: aphidicolin-inducible and folate-sensitive fragile sites in the human genome. Frontiers in Genetics 12, 708860.Google ScholarPubMed
Grindedal, EM et al. (2009) Germ-line mutations in mismatch repair genes associated with prostate cancer. Cancer Epidemiology, Biomarkers & Prevention 18, 24602467.CrossRefGoogle ScholarPubMed
Grindedal, EM et al. (2009) High risk of endometrial cancer in colorectal cancer kindred is pathognomonic for MMR-mutation carriers. Familial Cancer 8, 145151.CrossRefGoogle ScholarPubMed
Peltomaki, P et al. (1993) Microsatellite instability is associated with tumors that characterize the hereditary non-polyposis colorectal carcinoma syndrome. Cancer Research 53, 58535855.Google ScholarPubMed
Ionov, Y et al. (1993) Ubiquitous somatic mutations in simple repeated sequences reveal a new mechanism for colonic carcinogenesis. Nature 363, 558561.CrossRefGoogle ScholarPubMed
Thibodeau, SN, Bren, G and Schaid, D (1993) Microsatellite instability in cancer of the proximal colon. Science 260, 816819.Google ScholarPubMed
Wimmer, K and Kratz, CP (2010) Constitutional mismatch repair-deficiency syndrome. Haematologica 95, 699701.Google ScholarPubMed
Buerki, N et al. (2012) Evidence for breast cancer as an integral part of Lynch syndrome. Genes Chromosomes & Cancer 51, 8391.CrossRefGoogle ScholarPubMed
Win, AK et al. (2012) Risks of primary extracolonic cancers following colorectal cancer in lynch syndrome. Journal of the National Cancer Institute 104, 13631372.CrossRefGoogle ScholarPubMed
The Huntington's Disease Collaborative Research Group (1993) A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington's disease chromosomes. Cell 72, 971983.CrossRefGoogle Scholar
Campuzano, V et al. (1996) Friedreich's ataxia: autosomal recessive disease caused by an intronic GAA triplet repeat expansion. Science 271, 14231427.CrossRefGoogle ScholarPubMed
Brook, JD et al. (1992) Molecular basis of myotonic dystrophy: expansion of a trinucleotide (CTG) repeat at the 3′ end of a transcript encoding a protein kinase family member. Cell 69, 385.Google ScholarPubMed
Mahadevan, M et al. (1992) Myotonic dystrophy mutation: an unstable CTG repeat in the 3′ untranslated region of the gene. Science 255, 12531255.CrossRefGoogle ScholarPubMed
Verkerk, AJ et al. (1991) Identification of a gene (FMR-1) containing a CGG repeat coincident with a breakpoint cluster region exhibiting length variation in fragile X syndrome. Cell 65, 905914.CrossRefGoogle ScholarPubMed
Yu, S et al. (1991) Fragile X genotype characterized by an unstable region of DNA. Science 252, 11791181.CrossRefGoogle ScholarPubMed
Allingham-Hawkins, DJ et al. (1999) Fragile X premutation is a significant risk factor for premature ovarian failure: the International Collaborative POF in Fragile X study – preliminary data. American Journal of Medical Genetics 83, 322325.3.0.CO;2-B>CrossRefGoogle ScholarPubMed
Brunberg, JA et al. (2002) Fragile X premutation carriers: characteristic MR imaging findings of adult male patients with progressive cerebellar and cognitive dysfunction. AJNR. American Journal of Neuroradiology 23, 17571766.Google ScholarPubMed
Kononenko, AV et al. (2018) Mechanisms of genetic instability caused by (CGG)n repeats in an experimental mammalian system. Nature Structural & Molecular Biology 25, 669676.CrossRefGoogle Scholar
Kunkel, TA and Erie, DA (2015) Eukaryotic mismatch repair in relation to DNA replication. Annual Review of Genetics 49, 291313.CrossRefGoogle ScholarPubMed
Lujan, SA et al. (2012) Mismatch repair balances leading and lagging strand DNA replication fidelity. PLoS Genetics 8, e1003016.CrossRefGoogle ScholarPubMed
Gadgil, R et al. (2017) Replication stalling and DNA microsatellite instability. Biophysical Chemistry 225, 3848.CrossRefGoogle ScholarPubMed
Ellegren, H (2004) Microsatellites: simple sequences with complex evolution. Nature Reviews Genetics 5, 435445.CrossRefGoogle ScholarPubMed
Schlötterer, C and Tautz, D (1992) Slippage synthesis of simple sequence DNA. Nucleic Acids Research 20, 211215.CrossRefGoogle ScholarPubMed
Viguera, E, Canceill, D and Ehrlich, SD (2001) Replication slippage involves DNA polymerase pausing and dissociation. EMBO Journal 20, 25872595.CrossRefGoogle ScholarPubMed
Streisinger, G et al. (1966) Frameshift mutations and the genetic code. This paper is dedicated to Professor Theodosius Dobzhansky on the occasion of his 66th birthday. Cold Spring Harbor Symposia on Quantitative Biology 31, 7784.CrossRefGoogle Scholar
Drummond, JT et al. (1995) Isolation of an hMSH2-p160 heterodimer that restores DNA mismatch repair to tumor cells. Science 268, 19091912.CrossRefGoogle ScholarPubMed
Palombo, F et al. (1995) GTBP, a 160-kilodalton protein essential for mismatch-binding activity in human cells. Science 268, 19121914.CrossRefGoogle ScholarPubMed
Warren, JJ et al. (2007) Structure of the human MutSalpha DNA lesion recognition complex. Molecular Cell 26, 579592.CrossRefGoogle ScholarPubMed
Gupta, S, Gellert, M and Yang, W (2011) Mechanism of mismatch recognition revealed by human MutSβ bound to unpaired DNA loops. Nature Structural & Molecular Biology 19, 7278.CrossRefGoogle ScholarPubMed
Habraken, Y et al. (1996) Binding of insertion/deletion DNA mismatches by the heterodimer of yeast mismatch repair proteins MSH2 and MSH3. Current Biology 6, 11851187.CrossRefGoogle ScholarPubMed
Palombo, F et al. (1996) hMutSbeta, a heterodimer of hMSH2 and hMSH3, binds to insertion/deletion loops in DNA. Current Biology 6, 11811184.CrossRefGoogle ScholarPubMed
Snowden, T et al. (2004) hMSH4-hMSH5 recognizes Holliday junctions and forms a meiosis-specific sliding clamp that embraces homologous chromosomes. Molecular Cell 15, 437451.CrossRefGoogle ScholarPubMed
Flores-Rozas, H and Kolodner, RD (1998) The Saccharomyces cerevisiae MLH3 gene functions in MSH3-dependent suppression of frameshift mutations. Proceedings of the National Academy of Sciences of the USA 95, 12404–9.CrossRefGoogle ScholarPubMed
Kadyrov, FA et al. (2006) Endonucleolytic function of MutLalpha in human mismatch repair. Cell 126, 297308.Google ScholarPubMed
Pluciennik, A et al. (2010) PCNA function in the activation and strand direction of MutLα endonuclease in mismatch repair. Proceedings of the National Academy of Sciences of the USA 107, 1606616071.CrossRefGoogle ScholarPubMed
Cannavo, E et al. (2020) Regulation of the MLH1-MLH3 endonuclease in meiosis. Nature 586, 618622.CrossRefGoogle ScholarPubMed
Roesner, LM et al. (2013) Stable expression of MutLγ in human cells reveals no specific response to mismatched DNA, but distinct recruitment to damage sites. Journal of Cellular Biochemistry 114, 24052414.CrossRefGoogle ScholarPubMed
Prolla, TA et al. (1998) Tumour susceptibility and spontaneous mutation in mice deficient in Mlh1, Pms1 and Pms2 DNA mismatch repair. Nature Genetics 18, 276279.CrossRefGoogle ScholarPubMed
Räschle, M et al. (1999) Identification of hMutLbeta, a heterodimer of hMLH1 and hPMS1. Journal of Biological Chemistry 274, 3236832375.CrossRefGoogle ScholarPubMed
Bradford, KC et al. (2020) Dynamic human MutSalpha-MutLalpha complexes compact mismatched DNA. Proceedings of the National Academy of Sciences of the USA 117, 1630216312.CrossRefGoogle ScholarPubMed
Hombauer, H et al. (2011) Visualization of eukaryotic DNA mismatch repair reveals distinct recognition and repair intermediates. Cell 147, 10401053.CrossRefGoogle ScholarPubMed
Manhart, CM et al. (2017) The mismatch repair and meiotic recombination endonuclease Mlh1-Mlh3 is activated by polymer formation and can cleave DNA substrates in trans. PLoS Biology 15, e2001164.CrossRefGoogle ScholarPubMed
Wei, K et al. (2003) Inactivation of exonuclease 1 in mice results in DNA mismatch repair defects, increased cancer susceptibility, and male and female sterility. Genes & Development 17, 603614.CrossRefGoogle ScholarPubMed
Genschel, J, Bazemore, LR and Modrich, P (2002) Human exonuclease I is required for 5′ and 3′ mismatch repair. Journal of Biological Chemistry 277, 1330213311.CrossRefGoogle ScholarPubMed
Kratz, K et al. (2021) FANCD2-associated nuclease 1 partially compensates for the lack of exonuclease 1 in mismatch repair. Molecular and Cellular Biology 41, e0030321.CrossRefGoogle ScholarPubMed
Hile, SE, Yan, G and Eckert, KA (2000) Somatic mutation rates and specificities at TC/AG and GT/CA microsatellite sequences in nontumorigenic human lymphoblastoid cells. Cancer Research 60, 16981703.Google ScholarPubMed
Eckert, KA, Yan, G and Hile, SE (2002) Mutation rate and specificity analysis of tetranucleotide microsatellite DNA alleles in somatic human cells. Molecular Carcinogenesis 34, 140150.CrossRefGoogle ScholarPubMed
Hatch, SB and Farber, RA (2004) Mutation rates in the complex microsatellite MYCL1 and related simple repeats in cultured human cells. Mutation Research 545, 117126.Google ScholarPubMed
Christopher, J et al. (2019) Quantifying microsatellite mutation rates from intestinal stem cell dynamics in Msh2-deficient murine epithelium. Genetics 212, 655665.CrossRefGoogle ScholarPubMed
Eckert, KA and Hile, SE (2009) Every microsatellite is different: intrinsic DNA features dictate mutagenesis of common microsatellites present in the human genome. Molecular Carcinogenesis 48, 379388.CrossRefGoogle ScholarPubMed
Campregher, C et al. (2010) The nucleotide composition of microsatellites impacts both replication fidelity and mismatch repair in human colorectal cells. Human Molecular Genetics 19, 26482657.CrossRefGoogle ScholarPubMed
Brinkmann, B et al. (1998) Mutation rate in human microsatellites: influence of the structure and length of the tandem repeat. American Journal of Human Genetics 62, 14081415.CrossRefGoogle ScholarPubMed
Ellegren, H (2000) Heterogeneous mutation processes in human microsatellite DNA sequences. Nature Genetics 24, 400402.Google ScholarPubMed
Sun, JX et al. (2012) A direct characterization of human mutation based on microsatellites. Nature Genetics 44, 11611165.CrossRefGoogle ScholarPubMed
Wierdl, M, Dominska, M and Petes, TD (1997) Microsatellite instability in yeast: dependence on the length of the microsatellite. Genetics 146, 769779.CrossRefGoogle ScholarPubMed
Twerdi, CD, Boyer, JC and Farber, RA (1999) Relative rates of insertion and deletion mutations in a microsatellite sequence in cultured cells. Proceedings of the National Academy of Sciences of the USA 96, 28752879.Google Scholar
van Wietmarschen, N et al. (2020) Repeat expansions confer WRN dependence in microsatellite-unstable cancers. Nature 586, 292298.Google ScholarPubMed
Genetic Modifiers of Huntington's Disease Consortium (2019) CAG repeat not polyglutamine length determines timing of Huntington's disease onset. Cell 178, 887900.CrossRefGoogle Scholar
Bettencourt, C et al. (2016) DNA repair pathways underlie a common genetic mechanism modulating onset in polyglutamine diseases. Annals of Neurology 79, 983990.CrossRefGoogle ScholarPubMed
Lee, JM et al. (2017) A modifier of Huntington's disease onset at the MLH1 locus. Human Molecular Genetics 26, 38593867.CrossRefGoogle ScholarPubMed
Flower, M et al. (2019) MSH3 modifies somatic instability and disease severity in Huntington's and myotonic dystrophy type 1. Brain 142, 18761886.Google Scholar
McAllister, B et al. (2022) Exome sequencing of individuals with Huntington's disease implicates FAN1 nuclease activity in slowing CAG expansion and disease onset. Nature Neuroscience 25, 446457.CrossRefGoogle ScholarPubMed
Hwang, YH et al. (2022) Both cis and trans-acting genetic factors drive somatic instability in female carriers of the FMR1 premutation. Scientific Reports 12, 10419.CrossRefGoogle ScholarPubMed
Zhao, X et al. (2021) Modifiers of somatic repeat instability in mouse models of Friedreich ataxia and the fragile X-related disorders: implications for the mechanism of somatic expansion in Huntington's disease. Journal of Huntington's Disease 10, 149163.CrossRefGoogle ScholarPubMed
Zhao, XN and Usdin, K (2018) FAN1 protects against repeat expansions in a Fragile X mouse model. DNA Repair (Amst) 69, 15.CrossRefGoogle Scholar
Zhao, X et al. (2018) MutLgamma promotes repeat expansion in a Fragile X mouse model while EXO1 is protective. PLoS Genetics 14, e1007719.CrossRefGoogle Scholar
Zhao, X, Lu, H and Usdin, K (2021) FAN1′s protection against CGG repeat expansion requires its nuclease activity and is FANCD2-independent. Nucleic Acids Research 49, 1164311652.Google ScholarPubMed
Goold, R et al. (2021) FAN1 controls mismatch repair complex assembly via MLH1 retention to stabilize CAG repeat expansion in Huntington's disease. Cell Reports 36, 109649.Google ScholarPubMed
Gazy, I et al. (2019) Double-strand break repair plays a role in repeat instability in a fragile X mouse model. DNA Repair (Amst) 74, 6369.Google Scholar
Wright, GEB et al. (2019) Length of uninterrupted CAG, independent of polyglutamine size, results in increased somatic instability, hastening onset of Huntington disease. American Journal of Human Genetics 104, 11161126.CrossRefGoogle ScholarPubMed
Findlay Black, H et al. (2020) Frequency of the loss of CAA interruption in the HTT CAG tract and implications for Huntington disease in the reduced penetrance range. Genetics in Medicine 22, 21082113.CrossRefGoogle ScholarPubMed
Wright, GEB et al. (2020) Interrupting sequence variants and age of onset in Huntington's disease: clinical implications and emerging therapies. The Lancet. Neurology 19, 930939.CrossRefGoogle ScholarPubMed
Cumming, SA et al. (2018) De novo repeat interruptions are associated with reduced somatic instability and mild or absent clinical features in myotonic dystrophy type 1. European Journal of Human Genetics 26, 16351647.CrossRefGoogle ScholarPubMed
Overend, G et al. (2019) Allele length of the DMPK CTG repeat is a predictor of progressive myotonic dystrophy type 1 phenotypes. Human Molecular Genetics 28, 22452254.CrossRefGoogle ScholarPubMed
Morales, F et al. (2020) Longitudinal increases in somatic mosaicism of the expanded CTG repeat in myotonic dystrophy type 1 are associated with variation in age-at-onset. Human Molecular Genetics 29, 24962507.CrossRefGoogle ScholarPubMed
Latham, GJ et al. (2014) The role of AGG interruptions in fragile X repeat expansions: a twenty-year perspective. Frontiers in Genetics 5, 244.Google ScholarPubMed
Nolin, SL et al. (2015) Fragile X full mutation expansions are inhibited by one or more AGG interruptions in premutation carriers. Genetics in Medicine 17, 358364.CrossRefGoogle ScholarPubMed
Møllersen, L et al. (2010) Continuous and periodic expansion of CAG repeats in Huntington's disease R6/1 mice. PLoS Genetics 6, e1001242.CrossRefGoogle ScholarPubMed
Zhao, XN and Usdin, K (2018) Timing of expansion of fragile X premutation alleles during intergenerational transmission in a mouse model of the fragile X-related disorders. Frontiers in Genetics 9, 314.Google Scholar
Miller, CJ et al. (2020) All three mammalian MutL complexes are required for repeat expansion in a mouse cell model of the Fragile X-related disorders. PLoS Genetics 16, e1008902.CrossRefGoogle Scholar
Yrigollen, CM et al. (2014) AGG interruptions and maternal age affect FMR1 CGG repeat allele stability during transmission. Journal of Neurodevelopmental Disorders 6, 24.CrossRefGoogle ScholarPubMed
Kovalenko, M et al. (2012) Msh2 acts in medium-spiny striatal neurons as an enhancer of CAG instability and mutant huntingtin phenotypes in Huntington's disease knock-in mice. PLoS ONE 7, e44273.CrossRefGoogle ScholarPubMed
Kennedy, L and Shelbourne, PF (2000) Dramatic mutation instability in HD mouse striatum: does polyglutamine load contribute to cell-specific vulnerability in Huntington's disease? Human Molecular Genetics 9, 25392544.Google ScholarPubMed
Swami, M et al. (2009) Somatic expansion of the Huntington's disease CAG repeat in the brain is associated with an earlier age of disease onset. Human Molecular Genetics 18, 30393047.Google ScholarPubMed
Lokanga, AR et al. (2014) X inactivation plays a major role in the gender bias in somatic expansion in a mouse model of the fragile X-related disorders: implications for the mechanism of repeat expansion. Human Molecular Genetics 23, 49854994.Google Scholar
Kovtun, IV et al. (2007) OGG1 initiates age-dependent CAG trinucleotide expansion in somatic cells. Nature 447, 447452.CrossRefGoogle ScholarPubMed
Møllersen, L et al. (2012) Neil1 is a genetic modifier of somatic and germline CAG trinucleotide repeat instability in R6/1 mice. Human Molecular Genetics 21, 49394947.CrossRefGoogle ScholarPubMed
Entezam, A et al. (2010) Potassium bromate, a potent DNA oxidizing agent, exacerbates germline repeat expansion in a fragile X premutation mouse model. Human Mutation 31, 611616.Google Scholar
Jonson, I et al. (2013) Oxidative stress causes DNA triplet expansion in Huntington's disease mouse embryonic stem cells. Stem Cell Research 11, 12641271.CrossRefGoogle ScholarPubMed
Møllersen, L et al. (2016) Effects of anthocyanins on CAG repeat instability and behaviour in Huntington's disease R6/1 mice. PLoS Currents 8.Google ScholarPubMed
Gomes-Pereira, M and Monckton, DG (2020) Chronic exposure to cadmium and antioxidants does not affect the dynamics of expanded CAG*CTG trinucleotide repeats in a mouse cell culture system of unstable DNA. Frontiers in Cellular Neuroscience 14, 606331.CrossRefGoogle ScholarPubMed
Hayward, BE, Steinbach, PJ and Usdin, K (2020) A point mutation in the nuclease domain of MLH3 eliminates repeat expansions in a mouse stem cell model of the Fragile X-related disorders. Nucleic Acids Research 48, 78567863.CrossRefGoogle Scholar
Halabi, A, Fuselier, KTB and Grabczyk, E (2018) GAA*TTC repeat expansion in human cells is mediated by mismatch repair complex MutLgamma and depends upon the endonuclease domain in MLH3 isoform one. Nucleic Acids Research 46, 40224032.CrossRefGoogle ScholarPubMed
Roy, JCL et al. (2021) Somatic CAG expansion in Huntington's disease is dependent on the MLH3 endonuclease domain, which can be excluded via splice redirection. Nucleic Acids Research 49, 39073918.CrossRefGoogle ScholarPubMed
Kadyrova, LY et al. (2020) Human MutLgamma, the MLH1-MLH3 heterodimer, is an endonuclease that promotes DNA expansion. Proceedings of the National Academy of Sciences of the USA 117, 35353542.CrossRefGoogle ScholarPubMed
Zakharyevich, K et al. (2012) Delineation of joint molecule resolution pathways in meiosis identifies a crossover-specific resolvase. Cell 149, 334347.CrossRefGoogle ScholarPubMed
Mirkin, SM (2006) DNA structures, repeat expansions and human hereditary disorders. Current Opinion in Structural Biology 16, 351358.Google ScholarPubMed
Bourn, RL et al. (2012) Pms2 suppresses large expansions of the (GAA.TTC)n sequence in neuronal tissues. PLoS ONE 7, e47085.CrossRefGoogle ScholarPubMed
Zhao, XN et al. (2016) A MutSbeta-dependent contribution of MutSalpha to repeat expansions in fragile X premutation mice? PLoS Genetics 12, e1006190.Google ScholarPubMed
Foiry, L et al. (2006) Msh3 is a limiting factor in the formation of intergenerational CTG expansions in DM1 transgenic mice. Human Genetics 119, 520526.CrossRefGoogle ScholarPubMed
Halabi, A et al. (2012) DNA mismatch repair complex MutSbeta promotes GAA.TTC repeat expansion in human cells. Journal of Biological Chemistry 287, 2995829967.Google ScholarPubMed
Gomes-Pereira, M et al. (2004) Pms2 is a genetic enhancer of trinucleotide CAG.CTG repeat somatic mosaicism: implications for the mechanism of triplet repeat expansion. Human Molecular Genetics 13, 18151825.CrossRefGoogle ScholarPubMed
Zhao, XN et al. (2015) Mutsbeta generates both expansions and contractions in a mouse model of the Fragile X-associated disorders. Human Molecular Genetics 24, 70877096.Google Scholar
Aarnio, M et al. (1999) Cancer risk in mutation carriers of DNA-mismatch-repair genes. International Journal of Cancer 81, 214218.3.0.CO;2-L>CrossRefGoogle ScholarPubMed
Therkildsen, C et al. (2015) Glioblastomas, astrocytomas and oligodendrogliomas linked to Lynch syndrome. European Journal of Neurology 22, 717724.Google ScholarPubMed
Guerrini-Rousseau, L et al. (2019) Constitutional mismatch repair deficiency-associated brain tumors: report from the European C4CMMRD consortium. Neurooncology Advances 1, vdz033.Google ScholarPubMed
Shelbourne, PF et al. (2007) Triplet repeat mutation length gains correlate with cell-type specific vulnerability in Huntington disease brain. Human Molecular Genetics 16, 11331142.CrossRefGoogle ScholarPubMed
Lee, JM et al. (2011) Quantification of age-dependent somatic CAG repeat instability in Hdh CAG knock-in mice reveals different expansion dynamics in striatum and liver. PLoS ONE 6, e23647.Google ScholarPubMed
Gonitel, R et al. (2008) DNA instability in postmitotic neurons. Proceedings of the National Academy of Sciences of the USA 105, 34673472.Google ScholarPubMed
Chan, EM et al. (2019) WRN helicase is a synthetic lethal target in microsatellite unstable cancers. Nature 568, 551556.CrossRefGoogle ScholarPubMed
Triarico, S et al. (2019) Improving the brain delivery of chemotherapeutic drugs in childhood brain tumors. Cancers (Basel) 11, 824.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1. A model for MSI in which loop outs generated by strand-slippage of the nascent strand during DNA replication escape MMR and result in the incorporation of a small number of additional units, indicated by the green boxes, or the loss of a small number of repeat units, indicated by the dotted triangle. Whether repeats are gained or lost depends on whether repriming occurs further 3′ on the template resulting in nascent strand loop outs or further 5′ on the template resulting in loop outs being formed on the template strand.

Figure 1

Fig. 2. A model for repeat expansions and contraction in the REDs in which loop outs are formed on one or both strands during transcription or at other times that the DNA was unpaired. These loop outs are then bound by MutS and MutL proteins. Cleavage by MutLγ results in the formation of a staggered DSB with 5′ overhangs. Out-of-register reannealing of the DSB can produce a substrate for simple gap filling which results in the addition of repeat units. Exonucleolytic processing of the DSB can result in products with shorter 5′ overhangs or blunt ends. These products may then be processed, perhaps by NHEJ or gap-filling, to generate the loss of repeat units as indicated by the dotted triangle. The extent of contraction would depend on the amount of exonucleolytic cleavage that occurs prior to repair.