DNA mismatch repair

Last updated
Diagram of DNA mismatch repair pathways. The first column depicts mismatch repair in eukaryotes, while the second depicts repair in most bacteria. The third column shows mismatch repair, to be specific in E. coli. DNA mismatch repair.png
Diagram of DNA mismatch repair pathways. The first column depicts mismatch repair in eukaryotes, while the second depicts repair in most bacteria. The third column shows mismatch repair, to be specific in E. coli .
Micrograph showing loss of staining for MLH1 in colorectal adenocarcinoma in keeping with DNA mismatch repair (left of image) and benign colorectal mucosa (right of image). Colorectal adenocarcinoma with MMR - MLH1 -- high mag.jpg
Micrograph showing loss of staining for MLH1 in colorectal adenocarcinoma in keeping with DNA mismatch repair (left of image) and benign colorectal mucosa (right of image).

DNA mismatch repair (MMR) is a system for recognizing and repairing erroneous insertion, deletion, and mis-incorporation of bases that can arise during DNA replication and recombination, as well as repairing some forms of DNA damage. [1] [2]

Contents

Mismatch repair is strand-specific. During DNA synthesis the newly synthesised (daughter) strand will commonly include errors. In order to begin repair, the mismatch repair machinery distinguishes the newly synthesised strand from the template (parental). In gram-negative bacteria, transient hemimethylation distinguishes the strands (the parental is methylated and daughter is not). However, in other prokaryotes and eukaryotes, the exact mechanism is not clear. It is suspected that, in eukaryotes, newly synthesized lagging-strand DNA transiently contains nicks (before being sealed by DNA ligase) and provides a signal that directs mismatch proofreading systems to the appropriate strand. This implies that these nicks must be present in the leading strand, and evidence for this has recently been found. [3] Recent work [4] has shown that nicks are sites for RFC-dependent loading of the replication sliding clamp, proliferating cell nuclear antigen (PCNA), in an orientation-specific manner, such that one face of the donut-shape protein is juxtaposed toward the 3'-OH end at the nick. Loaded PCNA then directs the action of the MutLalpha endonuclease [5] to the daughter strand in the presence of a mismatch and MutSalpha or MutSbeta.

Any mutational event that disrupts the superhelical structure of DNA carries with it the potential to compromise the genetic stability of a cell. The fact that the damage detection and repair systems are as complex as the replication machinery itself highlights the importance evolution has attached to DNA fidelity.

Examples of mismatched bases include a G/T or A/C pairing (see DNA repair). Mismatches are commonly due to tautomerization of bases during DNA replication. The damage is repaired by recognition of the deformity caused by the mismatch, determining the template and non-template strand, and excising the wrongly incorporated base and replacing it with the correct nucleotide. The removal process involves more than just the mismatched nucleotide itself. A few or up to thousands of base pairs of the newly synthesized DNA strand can be removed.

Mismatch repair proteins

DNA mismatch repair protein, C-terminal domain
PDB 1h7u EBI.jpg
hpms2-atpgs
Identifiers
SymbolDNA_mis_repair
Pfam PF01119
Pfam clan CL0329
InterPro IPR013507
PROSITE PDOC00057
SCOP2 1bkn / SCOPe / SUPFAM
Available protein structures:
Pfam   structures / ECOD  
PDB RCSB PDB; PDBe; PDBj
PDBsum structure summary

Mismatch repair is a highly conserved process from prokaryotes to eukaryotes. The first evidence for mismatch repair was obtained from S. pneumoniae (the hexA and hexB genes). Subsequent work on E. coli has identified a number of genes that, when mutationally inactivated, cause hypermutable strains. The gene products are, therefore, called the "Mut" proteins, and are the major active components of the mismatch repair system. Three of these proteins are essential in detecting the mismatch and directing repair machinery to it: MutS, MutH and MutL (MutS is a homologue of HexA and MutL of HexB).

MutS forms a dimer (MutS2) that recognises the mismatched base on the daughter strand and binds the mutated DNA. MutH binds at hemimethylated sites along the daughter DNA, but its action is latent, being activated only upon contact by a MutL dimer (MutL2), which binds the MutS-DNA complex and acts as a mediator between MutS2 and MutH, activating the latter. The DNA is looped out to search for the nearest d(GATC) methylation site to the mismatch, which could be up to 1 kb away. Upon activation by the MutS-DNA complex, MutH nicks the daughter strand near the hemimethylated site. MutL recruits UvrD helicase (DNA Helicase II) to separate the two strands with a specific 3' to 5' polarity. The entire MutSHL complex then slides along the DNA in the direction of the mismatch, liberating the strand to be excised as it goes. An exonuclease trails the complex and digests the ss-DNA tail. The exonuclease recruited is dependent on which side of the mismatch MutH incises the strand – 5' or 3'. If the nick made by MutH is on the 5' end of the mismatch, either RecJ or ExoVII (both 5' to 3' exonucleases) is used. If, however, the nick is on the 3' end of the mismatch, ExoI (a 3' to 5' enzyme) is used.

The entire process ends past the mismatch site - i.e., both the site itself and its surrounding nucleotides are fully excised. The single-strand gap created by the exonuclease can then be repaired by DNA Polymerase III (assisted by single-strand-binding protein), which uses the other strand as a template, and finally sealed by DNA ligase. DNA methylase then rapidly methylates the daughter strand.

MutS homologs

When bound, the MutS2 dimer bends the DNA helix and shields approximately 20 base pairs. It has weak ATPase activity, and binding of ATP leads to the formation of tertiary structures on the surface of the molecule. The crystal structure of MutS reveals that it is exceptionally asymmetric, and, while its active conformation is a dimer, only one of the two halves interacts with the mismatch site.

In eukaryotes, MutShomologs form two major heterodimers: Msh2/Msh6 (MutSα) and Msh2/Msh3 (MutSβ). The MutSα pathway is involved primarily in base substitution and small-loop mismatch repair. The MutSβ pathway is also involved in small-loop repair, in addition to large-loop (~10 nucleotide loops) repair. However, MutSβ does not repair base substitutions.

MutL homologs

MutL also has weak ATPase activity (it uses ATP for purposes of movement). It forms a complex with MutS and MutH, increasing the MutS footprint on the DNA.

However, the processivity (the distance the enzyme can move along the DNA before dissociating) of UvrD is only ~40–50 bp. Because the distance between the nick created by MutH and the mismatch can average ~600 bp, if there is not another UvrD loaded the unwound section is then free to re-anneal to its complementary strand, forcing the process to start over. However, when assisted by MutL, the rate of UvrD loading is greatly increased. While the processivity (and ATP utilisation) of the individual UvrD molecules remains the same, the total effect on the DNA is boosted considerably; the DNA has no chance to re-anneal, as each UvrD unwinds 40-50 bp of DNA, dissociates, and then is immediately replaced by another UvrD, repeating the process. This exposes large sections of DNA to exonuclease digestion, allowing for quick excision (and later replacement) of the incorrect DNA.

Eukaryotes have five MutLhomologs designated as MLH1, MLH2, MLH3, PMS1, and PMS2. They form heterodimers that mimic MutL in E. coli. The human homologs of prokaryotic MutL form three complexes referred to as MutLα, MutLβ, and MutLγ. The MutLα complex is made of MLH1 and PMS2 subunits, the MutLβ heterodimer is made of MLH1 and PMS1, whereas MutLγ is made of MLH1 and MLH3. MutLα acts as an endonuclease that introduces strand breaks in the daughter strand upon activation by mismatch and other required proteins, MutSα and PCNA. These strand interruptions serve as entry points for an exonuclease activity that removes mismatched DNA. Roles played by MutLβ and MutLγ in mismatch repair are less-understood.

MutH: an endonuclease present in E. coli and Salmonella

MutH is a very weak endonuclease that is activated once bound to MutL (which itself is bound to MutS). It nicks unmethylated DNA and the unmethylated strand of hemimethylated DNA but does not nick fully methylated DNA. Experiments have shown that mismatch repair is random if neither strand is methylated.[ citation needed ] These behaviours led to the proposal that MutH determines which strand contains the mismatch. MutH has no eukaryotic homolog. Its endonuclease function is taken up by MutL homologs, which have some specialized 5'-3' exonuclease activity. The strand bias for removing mismatches from the newly synthesized daughter strand in eukaryotes may be provided by the free 3' ends of Okazaki fragments in the new strand created during replication.

PCNA β-sliding clamp

PCNA and the β-sliding clamp associate with MutSα/β and MutS, respectively. Although initial reports suggested that the PCNA-MutSα complex may enhance mismatch recognition, [6] it has been recently demonstrated [7] that there is no apparent change in affinity of MutSα for a mismatch in the presence or absence of PCNA. Furthermore, mutants of MutSα that are unable to interact with PCNA in vitro exhibit the capacity to carry out mismatch recognition and mismatch excision to near wild type levels. Such mutants are defective in the repair reaction directed by a 5' strand break, suggesting for the first time MutSα function in a post-excision step of the reaction.

Clinical significance

Inherited defects in mismatch repair

Mutations in the human homologues of the Mut proteins affect genomic stability, which can result in microsatellite instability (MSI), implicated in some human cancers. In specific, the hereditary nonpolyposis colorectal cancers (HNPCC or Lynch syndrome) are attributed to damaging germline variants in the genes encoding the MutS and MutL homologues MSH2 and MLH1 respectively, which are thus classified as tumour suppressor genes. One subtype of HNPCC, the Muir-Torre Syndrome (MTS), is associated with skin tumors. If both inherited copies (alleles) of a MMR gene bear damaging genetic variants, this results in a very rare and severe condition: the mismatch repair cancer syndrome (or constitutional mismatch repair deficiency, CMMR-D), manifesting as multiple occurrences of tumors at an early age, often colon and brain tumors. [8]

Epigenetic silencing of mismatch repair genes

Sporadic cancers with a DNA repair deficiency only rarely have a mutation in a DNA repair gene, but they instead tend to have epigenetic alterations such as promoter methylation that inhibit DNA repair gene expression. [9] About 13% of colorectal cancers are deficient in DNA mismatch repair, commonly due to loss of MLH1 (9.8%), or sometimes MSH2, MSH6 or PMS2 (all ≤1.5%). [10] For most MLH1-deficient sporadic colorectal cancers, the deficiency was due to MLH1 promoter methylation. [10] Other cancer types have higher frequencies of MLH1 loss (see table below), which are again largely a result of methylation of the promoter of the MLH1 gene. A different epigenetic mechanism underlying MMR deficiencies might involve over-expression of a microRNA, for example miR-155 levels inversely correlate with expression of MLH1 or MSH2 in colorectal cancer. [11]

Cancers deficient in MLH1
Cancer typeFrequency of deficiency in cancerFrequency of deficiency in adjacent field defect
Stomach32% [12] [13] 24%-28%
Stomach (foveolar type tumors)74% [14] 71%
Stomach in high-incidence Kashmir Valley73% [15] 20%
Esophageal73% [16] 27%
Head and neck squamous cell carcinoma (HNSCC)31%-33% [17] [18] 20%-25%
Non-small cell lung cancer (NSCLC)69% [19] 72%
Colorectal10% [10]

MMR failures in field defects

A field defect (field cancerization) is an area of epithelium that has been preconditioned by epigenetic or genetic changes, predisposing it towards development of cancer. As pointed out by Rubin " ...there is evidence that more than 80% of the somatic mutations found in mutator phenotype human colorectal tumors occur before the onset of terminal clonal expansion." [20] [21] Similarly, Vogelstein et al. [22] point out that more than half of somatic mutations identified in tumors occurred in a pre-neoplastic phase (in a field defect), during growth of apparently normal cells.

MLH1 deficiencies were common in the field defects (histologically normal tissues) surrounding tumors; see Table above. Epigenetically silenced or mutated MLH1 would likely not confer a selective advantage upon a stem cell, however, it would cause increased mutation rates, and one or more of the mutated genes may provide the cell with a selective advantage. The deficientMLH1 gene could then be carried along as a selectively near-neutral passenger (hitch-hiker) gene when the mutated stem cell generates an expanded clone. The continued presence of a clone with an epigenetically repressed MLH1 would continue to generate further mutations, some of which could produce a tumor.

MSI and immune checkpoint blockade response

MMR and mismatch repair mutations were initially observed to associate with immune checkpoint blockade efficacy in a study examining responders to anti-PD1. [23] The association between MSI positivity and positive response to anti-PD1 was subsequently validated in a prospective clinical trial and approved by the FDA. [24]



MMR components in humans

In humans, seven DNA mismatch repair (MMR) proteins (MLH1, MLH3, MSH2, MSH3, MSH6, PMS1 and PMS2) work coordinately in sequential steps to initiate repair of DNA mismatches. [25] In addition, there are Exo1-dependent and Exo1-independent MMR subpathways. [26]

Other gene products involved in mismatch repair (subsequent to initiation by MMR genes) in humans include DNA polymerase delta, PCNA, RPA, HMGB1, RFC and DNA ligase I, plus histone and chromatin modifying factors. [27] [28]

In certain circumstances, the MMR pathway may recruit an error-prone DNA polymerase eta (POLH). This happens in B-lymphocytes during somatic hypermutation, where POLH is used to introduce genetic variation into antibody genes. [29] However, this error-prone MMR pathway may be triggered in other types of human cells upon exposure to genotoxins [30] and indeed it is broadly active in various human cancers, causing mutations that bear a signature of POLH activity. [31]

MMR and mutation frequency

Recognizing and repairing mismatches and indels is important for cells because failure to do so results in microsatellite instability (MSI) and an elevated spontaneous mutation rate (mutator phenotype). In comparison to other cancer types, MMR-deficient (MSI) cancer has a very high frequency of mutations, close to melanoma and lung cancer, [32] cancer types caused by much exposure to UV radiation and mutagenic chemicals.

In addition to a very high mutation burden, MMR deficiencies result in an unusual distribution of somatic mutations across the human genome: this suggests that MMR preferentially protects the gene-rich, early-replicating euchromatic regions. [33] In contrast, the gene-poor, late-replicating heterochromatic genome regions exhibit high mutation rates in many human tumors. [34]

The histone modification H3K36me3, an epigenetic mark of active chromatin, has the ability to recruit the MSH2-MSH6 (hMutSα) complex. [35] Consistently, regions of the human genome with high levels of H3K36me3 accumulate less mutations due to MMR activity. [31]

Loss of multiple DNA repair pathways in tumors

Lack of MMR often occurs in coordination with loss of other DNA repair genes. [9] For example, MMR genes MLH1 and MLH3 as well as 11 other DNA repair genes (such as MGMT and many NER pathway genes) were significantly down-regulated in lower grade as well as in higher grade astrocytomas, in contrast to normal brain tissue. [36] Moreover, MLH1 and MGMT expression was closely correlated in 135 specimens of gastric cancer and loss of MLH1 and MGMT appeared to be synchronously accelerated during tumor progression. [37]

Deficient expression of multiple DNA repair genes is often found in cancers, [9] and may contribute to the thousands of mutations usually found in cancers (see Mutation frequencies in cancers).

Aging

A popular idea, that has failed to gain significant experimental support, is the idea that mutation, as distinct from DNA damage, is the primary cause of aging. Mice defective in the mutL homolog Pms2 have about a 100-fold elevated mutation frequency in all tissues, but do not appear to age more rapidly. [38] These mice display mostly normal development and life, except for early onset carcinogenesis and male infertility.

See also

Related Research Articles

<span class="mw-page-title-main">DNA repair</span> Cellular mechanism

DNA repair is a collection of processes by which a cell identifies and corrects damage to the DNA molecules that encodes its genome. In human cells, both normal metabolic activities and environmental factors such as radiation can cause DNA damage, resulting in tens of thousands of individual molecular lesions per cell per day. Many of these lesions cause structural damage to the DNA molecule and can alter or eliminate the cell's ability to transcribe the gene that the affected DNA encodes. Other lesions induce potentially harmful mutations in the cell's genome, which affect the survival of its daughter cells after it undergoes mitosis. As a consequence, the DNA repair process is constantly active as it responds to damage in the DNA structure. When normal repair processes fail, and when cellular apoptosis does not occur, irreparable DNA damage may occur, including double-strand breaks and DNA crosslinkages. This can eventually lead to malignant tumors, or cancer as per the two-hit hypothesis.

<span class="mw-page-title-main">Neoplasm</span> Abnormal mass of tissue as a result of abnormal growth or division of cells

A neoplasm is a type of abnormal and excessive growth of tissue. The process that occurs to form or produce a neoplasm is called neoplasia. The growth of a neoplasm is uncoordinated with that of the normal surrounding tissue, and persists in growing abnormally, even if the original trigger is removed. This abnormal growth usually forms a mass, when it may be called a tumour or tumor.

<span class="mw-page-title-main">Mismatch repair cancer syndrome</span> Medical condition

Mismatch repair cancer syndrome (MMRCS) is a cancer syndrome associated with biallelic DNA mismatch repair mutations. It is also known as Turcot syndrome and by several other names.

<span class="mw-page-title-main">Base excision repair</span> DNA repair process

Base excision repair (BER) is a cellular mechanism, studied in the fields of biochemistry and genetics, that repairs damaged DNA throughout the cell cycle. It is responsible primarily for removing small, non-helix-distorting base lesions from the genome. The related nucleotide excision repair pathway repairs bulky helix-distorting lesions. BER is important for removing damaged bases that could otherwise cause mutations by mispairing or lead to breaks in DNA during replication. BER is initiated by DNA glycosylases, which recognize and remove specific damaged or inappropriate bases, forming AP sites. These are then cleaved by an AP endonuclease. The resulting single-strand break can then be processed by either short-patch or long-patch BER.

<span class="mw-page-title-main">Microsatellite instability</span> Condition of genetic hypermutability

Microsatellite instability (MSI) is the condition of genetic hypermutability that results from impaired DNA mismatch repair (MMR). The presence of MSI represents phenotypic evidence that MMR is not functioning normally.

<span class="mw-page-title-main">Muir–Torre syndrome</span> Medical condition

Muir–Torre syndrome is a rare hereditary, autosomal dominant cancer syndrome that is thought to be a subtype of HNPCC. Individuals are prone to develop cancers of the colon, genitourinary tract, and skin lesions, such as keratoacanthomas and sebaceous tumors. The genes affected are MLH1, MSH2, and more recently, MSH6, and are involved in DNA mismatch repair.

<span class="mw-page-title-main">MSH2</span> Protein-coding gene in the species Homo sapiens

DNA mismatch repair protein Msh2 also known as MutS homolog 2 or MSH2 is a protein that in humans is encoded by the MSH2 gene, which is located on chromosome 2. MSH2 is a tumor suppressor gene and more specifically a caretaker gene that codes for a DNA mismatch repair (MMR) protein, MSH2, which forms a heterodimer with MSH6 to make the human MutSα mismatch repair complex. It also dimerizes with MSH3 to form the MutSβ DNA repair complex. MSH2 is involved in many different forms of DNA repair, including transcription-coupled repair, homologous recombination, and base excision repair.

<span class="mw-page-title-main">MLH1</span> Protein-coding gene in the species Homo sapiens

DNA mismatch repair protein Mlh1 or MutL protein homolog 1 is a protein that in humans is encoded by the MLH1 gene located on chromosome 3. It is a gene commonly associated with hereditary nonpolyposis colorectal cancer. Orthologs of human MLH1 have also been studied in other organisms including mouse and the budding yeast Saccharomyces cerevisiae.

<span class="mw-page-title-main">MSH6</span> Protein-coding gene in the species Homo sapiens

MSH6 or mutS homolog 6 is a gene that codes for DNA mismatch repair protein Msh6 in the budding yeast Saccharomyces cerevisiae. It is the homologue of the human "G/T binding protein," (GTBP) also called p160 or hMSH6. The MSH6 protein is a member of the Mutator S (MutS) family of proteins that are involved in DNA damage repair.

<span class="mw-page-title-main">PMS2</span> Protein-coding gene in the species Homo sapiens

Mismatch repair endonuclease PMS2 is an enzyme that in humans is encoded by the PMS2 gene.

<span class="mw-page-title-main">O-6-methylguanine-DNA methyltransferase</span> Mammalian protein found in Homo sapiens

O6-alkylguanine DNA alkyltransferase (also known as AGT, MGMT or AGAT) is a protein that in humans is encoded by the O6-methylguanine DNA methyltransferase (MGMT) gene. O6-methylguanine DNA methyltransferase is crucial for genome stability. It repairs the naturally occurring mutagenic DNA lesion O6-methylguanine back to guanine and prevents mismatch and errors during DNA replication and transcription. Accordingly, loss of MGMT increases the carcinogenic risk in mice after exposure to alkylating agents. The two bacterial isozymes are Ada and Ogt.

<span class="mw-page-title-main">Exonuclease 1</span> Protein-coding gene in the species Homo sapiens

Exonuclease 1 is an enzyme that in humans is encoded by the EXO1 gene.

<span class="mw-page-title-main">MSH3</span> Protein-coding gene in the species Homo sapiens

DNA mismatch repair protein, MutS Homolog 3 (MSH3) is a human homologue of the bacterial mismatch repair protein MutS that participates in the mismatch repair (MMR) system. MSH3 typically forms the heterodimer MutSβ with MSH2 in order to correct long insertion/deletion loops and base-base mispairs in microsatellites during DNA synthesis. Deficient capacity for MMR is found in approximately 15% of colorectal cancers, and somatic mutations in the MSH3 gene can be found in nearly 50% of MMR-deficient colorectal cancers.

<span class="mw-page-title-main">MBD4</span> Protein-coding gene in the species Homo sapiens

Methyl-CpG-binding domain protein 4 is a protein that in humans is encoded by the MBD4 gene.

<span class="mw-page-title-main">PMS1</span> Protein-coding gene in humans

PMS1 protein homolog 1 is a protein that in humans is encoded by the PMS1 gene.

<span class="mw-page-title-main">MLH3</span> Protein-coding gene in the species Homo sapiens

DNA mismatch repair protein Mlh3 is a protein that in humans is encoded by the MLH3 gene.

Synthetic lethality is defined as a type of genetic interaction where the combination of two genetic events results in cell death or death of an organism. Although the foregoing explanation is wider than this, it is common when referring to synthetic lethality to mean the situation arising by virtue of a combination of deficiencies of two or more genes leading to cell death, whereas a deficiency of only one of these genes does not. In a synthetic lethal genetic screen, it is necessary to begin with a mutation that does not result in cell death, although the effect of that mutation could result in a differing phenotype, and then systematically test other mutations at additional loci to determine which, in combination with the first mutation, causes cell death arising by way of deficiency or abolition of expression.

Genome instability refers to a high frequency of mutations within the genome of a cellular lineage. These mutations can include changes in nucleic acid sequences, chromosomal rearrangements or aneuploidy. Genome instability does occur in bacteria. In multicellular organisms genome instability is central to carcinogenesis, and in humans it is also a factor in some neurodegenerative diseases such as amyotrophic lateral sclerosis or the neuromuscular disease myotonic dystrophy.

<span class="mw-page-title-main">Cancer epigenetics</span> Field of study in cancer research

Cancer epigenetics is the study of epigenetic modifications to the DNA of cancer cells that do not involve a change in the nucleotide sequence, but instead involve a change in the way the genetic code is expressed. Epigenetic mechanisms are necessary to maintain normal sequences of tissue specific gene expression and are crucial for normal development. They may be just as important, if not even more important, than genetic mutations in a cell's transformation to cancer. The disturbance of epigenetic processes in cancers, can lead to a loss of expression of genes that occurs about 10 times more frequently by transcription silencing than by mutations. As Vogelstein et al. points out, in a colorectal cancer there are usually about 3 to 6 driver mutations and 33 to 66 hitchhiker or passenger mutations. However, in colon tumors compared to adjacent normal-appearing colonic mucosa, there are about 600 to 800 heavily methylated CpG islands in the promoters of genes in the tumors while these CpG islands are not methylated in the adjacent mucosa. Manipulation of epigenetic alterations holds great promise for cancer prevention, detection, and therapy. In different types of cancer, a variety of epigenetic mechanisms can be perturbed, such as the silencing of tumor suppressor genes and activation of oncogenes by altered CpG island methylation patterns, histone modifications, and dysregulation of DNA binding proteins. There are several medications which have epigenetic impact, that are now used in a number of these diseases.

DNA methylation in cancer plays a variety of roles, helping to change the healthy cells by regulation of gene expression to a cancer cells or a diseased cells disease pattern. One of the most widely studied DNA methylation dysregulation is the promoter hypermethylation where the CPGs islands in the promoter regions are methylated contributing or causing genes to be silenced.

References

  1. Iyer RR, Pluciennik A, Burdett V, Modrich PL (February 2006). "DNA mismatch repair: functions and mechanisms". Chemical Reviews. 106 (2): 302–23. doi:10.1021/cr0404794. PMID   16464007.
  2. Larrea AA, Lujan SA, Kunkel TA (May 2010). "SnapShot: DNA mismatch repair". Cell. 141 (4): 730–730.e1. doi: 10.1016/j.cell.2010.05.002 . PMID   20478261. S2CID   26969788.
  3. Heller RC, Marians KJ (December 2006). "Replisome assembly and the direct restart of stalled replication forks". Nature Reviews. Molecular Cell Biology. 7 (12): 932–43. doi:10.1038/nrm2058. PMID   17139333. S2CID   27666329.
  4. Pluciennik A, Dzantiev L, Iyer RR, Constantin N, Kadyrov FA, Modrich P (September 2010). "PCNA function in the activation and strand direction of MutLα endonuclease in mismatch repair". Proceedings of the National Academy of Sciences of the United States of America. 107 (37): 16066–71. doi: 10.1073/pnas.1010662107 . PMC   2941292 . PMID   20713735.
  5. Kadyrov FA, Dzantiev L, Constantin N, Modrich P (July 2006). "Endonucleolytic function of MutLalpha in human mismatch repair". Cell. 126 (2): 297–308. doi: 10.1016/j.cell.2006.05.039 . PMID   16873062. S2CID   15643051.
  6. Flores-Rozas H, Clark D, Kolodner RD (November 2000). "Proliferating cell nuclear antigen and Msh2p-Msh6p interact to form an active mispair recognition complex". Nature Genetics. 26 (3): 375–8. doi:10.1038/81708. PMID   11062484. S2CID   20861705.
  7. Iyer RR, Pohlhaus TJ, Chen S, Hura GL, Dzantiev L, Beese LS, Modrich P (May 2008). "The MutSalpha-proliferating cell nuclear antigen interaction in human DNA mismatch repair". The Journal of Biological Chemistry. 283 (19): 13310–9. doi: 10.1074/jbc.M800606200 . PMC   2423938 . PMID   18326858.
  8. Online Mendelian Inheritance in Man (OMIM): 276300
  9. 1 2 3 Bernstein C, Bernstein H (May 2015). "Epigenetic reduction of DNA repair in progression to gastrointestinal cancer". World Journal of Gastrointestinal Oncology. 7 (5): 30–46. doi: 10.4251/wjgo.v7.i5.30 . PMC   4434036 . PMID   25987950.
  10. 1 2 3 Truninger K, Menigatti M, Luz J, Russell A, Haider R, Gebbers JO, et al. (May 2005). "Immunohistochemical analysis reveals high frequency of PMS2 defects in colorectal cancer". Gastroenterology. 128 (5): 1160–71. doi: 10.1053/j.gastro.2005.01.056 . PMID   15887099.
  11. Valeri N, Gasparini P, Fabbri M, Braconi C, Veronese A, Lovat F, et al. (April 2010). "Modulation of mismatch repair and genomic stability by miR-155". Proceedings of the National Academy of Sciences of the United States of America. 107 (15): 6982–7. Bibcode:2010PNAS..107.6982V. doi: 10.1073/pnas.1002472107 . PMC   2872463 . PMID   20351277.
  12. Kupčinskaitė-Noreikienė R, Skiecevičienė J, Jonaitis L, Ugenskienė R, Kupčinskas J, Markelis R, et al. (2013). "CpG island methylation of the MLH1, MGMT, DAPK, and CASP8 genes in cancerous and adjacent noncancerous stomach tissues". Medicina. 49 (8): 361–6. doi: 10.3390/medicina49080056 . PMID   24509146.
  13. Waki T, Tamura G, Tsuchiya T, Sato K, Nishizuka S, Motoyama T (August 2002). "Promoter methylation status of E-cadherin, hMLH1, and p16 genes in nonneoplastic gastric epithelia". The American Journal of Pathology. 161 (2): 399–403. doi:10.1016/S0002-9440(10)64195-8. PMC   1850716 . PMID   12163364.
  14. Endoh Y, Tamura G, Ajioka Y, Watanabe H, Motoyama T (September 2000). "Frequent hypermethylation of the hMLH1 gene promoter in differentiated-type tumors of the stomach with the gastric foveolar phenotype". The American Journal of Pathology. 157 (3): 717–22. doi:10.1016/S0002-9440(10)64584-1. PMC   1949419 . PMID   10980110.
  15. Wani M, Afroze D, Makhdoomi M, Hamid I, Wani B, Bhat G, et al. (2012). "Promoter methylation status of DNA repair gene (hMLH1) in gastric carcinoma patients of the Kashmir valley" (PDF). Asian Pacific Journal of Cancer Prevention. 13 (8): 4177–81. doi: 10.7314/apjcp.2012.13.8.4177 . PMID   23098428.
  16. Chang Z, Zhang W, Chang Z, Song M, Qin Y, Chang F, et al. (January 2015). "Expression characteristics of FHIT, p53, BRCA2 and MLH1 in families with a history of oesophageal cancer in a region with a high incidence of oesophageal cancer". Oncology Letters. 9 (1): 430–436. doi:10.3892/ol.2014.2682. PMC   4246613 . PMID   25436004.
  17. Tawfik HM, El-Maqsoud NM, Hak BH, El-Sherbiny YM (2011). "Head and neck squamous cell carcinoma: mismatch repair immunohistochemistry and promoter hypermethylation of hMLH1 gene". American Journal of Otolaryngology. 32 (6): 528–36. doi:10.1016/j.amjoto.2010.11.005. PMID   21353335.
  18. Zuo C, Zhang H, Spencer HJ, Vural E, Suen JY, Schichman SA, et al. (October 2009). "Increased microsatellite instability and epigenetic inactivation of the hMLH1 gene in head and neck squamous cell carcinoma". Otolaryngology–Head and Neck Surgery. 141 (4): 484–90. doi:10.1016/j.otohns.2009.07.007. PMID   19786217. S2CID   8357370.
  19. Safar AM, Spencer H, Su X, Coffey M, Cooney CA, Ratnasinghe LD, et al. (June 2005). "Methylation profiling of archived non-small cell lung cancer: a promising prognostic system". Clinical Cancer Research. 11 (12): 4400–5. doi: 10.1158/1078-0432.CCR-04-2378 . PMID   15958624.
  20. Rubin H (March 2011). "Fields and field cancerization: the preneoplastic origins of cancer: asymptomatic hyperplastic fields are precursors of neoplasia, and their progression to tumors can be tracked by saturation density in culture". BioEssays. 33 (3): 224–31. doi:10.1002/bies.201000067. PMID   21254148. S2CID   44981539.
  21. Tsao JL, Yatabe Y, Salovaara R, Järvinen HJ, Mecklin JP, Aaltonen LA, et al. (February 2000). "Genetic reconstruction of individual colorectal tumor histories". Proceedings of the National Academy of Sciences of the United States of America. 97 (3): 1236–41. Bibcode:2000PNAS...97.1236T. doi: 10.1073/pnas.97.3.1236 . PMC   15581 . PMID   10655514.
  22. Vogelstein B, Papadopoulos N, Velculescu VE, Zhou S, Diaz LA, Kinzler KW (March 2013). "Cancer genome landscapes". Science. 339 (6127): 1546–58. Bibcode:2013Sci...339.1546V. doi:10.1126/science.1235122. PMC   3749880 . PMID   23539594.
  23. Rizvi, Naiyer; Hellmann, Matthew; Snyder, Alexandra; Kvistborg, Pia; Makarov, Vladimir; Havel, Jonathan; Lee, William; Yuan, Jianda; Wong, Phillip; Ho, Teresa; Miller, Martin; Rekhtman, Natasha; Moreira, Andra; Ibrahim, Fawzia; Bruggeman, Cameron; Gasmi, Billel; Zappasodi, Roberta; Maeda, Yuka; Sander, Chris; Garon, Edward; Merghoub, Taha; Wolchok, Jedd; Schumacher, Ton; Timothy, Chan (2015). "Mutational landscape determines sensitivity to PD-1 blockade in non-small cell lung cancer". Science. 6230 (348): 124-128. doi:10.1126/science.aaa1348. PMC   4993154 . PMID   25765070.
  24. Center for Drug Evaluation and Research. "Approved Drugs –FDA grants accelerated approval to pembrolizumab for first tissue/site agnostic indication". www.fda.gov. Retrieved 2017-05-24.
  25. Pal T, Permuth-Wey J, Sellers TA (August 2008). "A review of the clinical relevance of mismatch-repair deficiency in ovarian cancer". Cancer. 113 (4): 733–42. doi:10.1002/cncr.23601. PMC   2644411 . PMID   18543306.
  26. Goellner EM, Putnam CD, Kolodner RD (August 2015). "Exonuclease 1-dependent and independent mismatch repair". DNA Repair. 32: 24–32. doi:10.1016/j.dnarep.2015.04.010. PMC   4522362 . PMID   25956862.
  27. Li GM (January 2008). "Mechanisms and functions of DNA mismatch repair". Cell Research. 18 (1): 85–98. doi: 10.1038/cr.2007.115 . PMID   18157157.
  28. Li GM (July 2014). "New insights and challenges in mismatch repair: getting over the chromatin hurdle". DNA Repair. 19: 48–54. doi:10.1016/j.dnarep.2014.03.027. PMC   4127414 . PMID   24767944.
  29. Chahwan R, Edelmann W, Scharff MD, Roa S (August 2012). "AIDing antibody diversity by error-prone mismatch repair". Seminars in Immunology. 24 (4): 293–300. doi:10.1016/j.smim.2012.05.005. PMC   3422444 . PMID   22703640.
  30. Hsieh P (September 2012). "DNA mismatch repair: Dr. Jekyll and Mr. Hyde?". Molecular Cell. 47 (5): 665–6. doi:10.1016/j.molcel.2012.08.020. PMC   3457060 . PMID   22980456.
  31. 1 2 Supek F, Lehner B (July 2017). "Clustered Mutation Signatures Reveal that Error-Prone DNA Repair Targets Mutations to Active Genes". Cell. 170 (3): 534–547.e23. doi: 10.1016/j.cell.2017.07.003 . hdl: 10230/35343 . PMID   28753428.
  32. Tuna M, Amos CI (November 2013). "Genomic sequencing in cancer". Cancer Letters. 340 (2): 161–70. doi:10.1016/j.canlet.2012.11.004. PMC   3622788 . PMID   23178448.
  33. Supek F, Lehner B (May 2015). "Differential DNA mismatch repair underlies mutation rate variation across the human genome". Nature. 521 (7550): 81–4. Bibcode:2015Natur.521...81S. doi:10.1038/nature14173. PMC   4425546 . PMID   25707793.
  34. Schuster-Böckler B, Lehner B (August 2012). "Chromatin organization is a major influence on regional mutation rates in human cancer cells". Nature. 488 (7412): 504–7. Bibcode:2012Natur.488..504S. doi:10.1038/nature11273. PMID   22820252. S2CID   205229634.
  35. Li F, Mao G, Tong D, Huang J, Gu L, Yang W, Li GM (April 2013). "The histone mark H3K36me3 regulates human DNA mismatch repair through its interaction with MutSα". Cell. 153 (3): 590–600. doi:10.1016/j.cell.2013.03.025. PMC   3641580 . PMID   23622243.
  36. Jiang Z, Hu J, Li X, Jiang Y, Zhou W, Lu D (December 2006). "Expression analyses of 27 DNA repair genes in astrocytoma by TaqMan low-density array". Neuroscience Letters. 409 (2): 112–7. doi:10.1016/j.neulet.2006.09.038. PMID   17034947. S2CID   54278905.
  37. Kitajima Y, Miyazaki K, Matsukura S, Tanaka M, Sekiguchi M (2003). "Loss of expression of DNA repair enzymes MGMT, hMLH1, and hMSH2 during tumor progression in gastric cancer". Gastric Cancer. 6 (2): 86–95. doi: 10.1007/s10120-003-0213-z . PMID   12861399.
  38. Narayanan L, Fritzell JA, Baker SM, Liskay RM, Glazer PM (April 1997). "Elevated levels of mutation in multiple tissues of mice deficient in the DNA mismatch repair gene Pms2". Proceedings of the National Academy of Sciences of the United States of America. 94 (7): 3122–7. Bibcode:1997PNAS...94.3122N. doi: 10.1073/pnas.94.7.3122 . PMC   20332 . PMID   9096356.

Further reading