# 114500

COLORECTAL CANCER; CRC


Alternative titles; symbols

COLON CANCER


Phenotype-Gene Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
Gene/Locus Gene/Locus
MIM number
1p36.13 {?Colorectal cancer, susceptibility to} 114500 AD, SMu 3 PLA2G2A 172411
1p13.2 Colorectal cancer, somatic 114500 3 NRAS 164790
2q13 Colorectal cancer with chromosomal instability, somatic 114500 3 BUB1 602452
3p22.1 Colorectal cancer, somatic 114500 3 CTNNB1 116806
3q26.32 Colorectal cancer, somatic 114500 3 PIK3CA 171834
4p16.3 Colorectal cancer, somatic 114500 3 FGFR3 134934
4q31.3 {Colorectal cancer, susceptibility to} 114500 AD, SMu 3 TLR2 603028
5q22.2 Colorectal cancer, somatic 114500 3 APC 611731
5q22.2 Colorectal cancer, somatic 114500 3 MCC 159350
7q11.23 Colon cancer, somatic 114500 3 PTPN12 600079
7q34 Colorectal cancer, somatic 114500 3 BRAF 164757
8p22 Colorectal cancer, somatic 114500 3 DLC1 604258
8p22 Colorectal cancer, somatic 114500 3 PDGFRL 604584
8q22.1 Colon cancer, somatic 114500 3 RAD54B 604289
11p11.2 Colon cancer, somatic 114500 3 PTPRJ 600925
11q13.3 {Colorectal cancer, susceptibility to} 114500 AD, SMu 3 CCND1 168461
14q24.3 Colorectal cancer, somatic 114500 3 MLH3 604395
14q32.33 Colorectal cancer, somatic 114500 3 AKT1 164730
15q15.1 Colorectal cancer, somatic 114500 3 BUB1B 602860
17p13.1 {Colorectal cancer} 114500 AD, SMu 3 TP53 191170
17p11.2 Colorectal cancer, somatic 114500 3 FLCN 607273
17q24.1 Colorectal cancer, somatic 114500 3 AXIN2 604025
18q21.2 Colorectal cancer, somatic 114500 3 DCC 120470
19q13.33 Colorectal cancer, somatic 114500 3 BAX 600040
20q11.23 Colon cancer, advanced, somatic 114500 3 SRC 190090
20q13.2 {Colon cancer, susceptibility to} 114500 AD, SMu 3 AURKA 603072
22q13.2 Colorectal cancer, somatic 114500 3 EP300 602700
Clinical Synopsis
 

INHERITANCE
- Somatic mutation
- Autosomal dominant
NEOPLASIA
- Colorectal cancer

TEXT

A number sign (#) is used with this entry because mutations in several different genes have been identified in colorectal cancer (CRC).


Description

Colorectal cancer is a heterogeneous disease that is common in both men and women. In addition to lifestyle and environmental risk factors, gene defects can contribute to an inherited predisposition to CRC. CRC is caused by changes in different molecular pathogenic pathways, such as chromosomal instability, CpG island methylator phenotype, and microsatellite instability. Chromosome instability is the most common alteration and is present in almost 85% of all cases (review by Schweiger et al., 2013).

Genetic Heterogeneity of Colorectal Cancer

Mutations in a single gene result in a marked predisposition to colorectal cancer in familial adenomatous polyposis (FAP; 175100) and hereditary nonpolyposis colorectal cancer (HNPCC; see 120435). FAP is caused by mutations in the APC gene (611731), whereas HNPCC is caused by mutations in several genes, including MSH2 (609309), MLH1 (120436), PMS1 (600258), PMS2 (600259), MSH6 (600678), TGFBR2 (190182), and MLH3 (604395). Epigenetic silencing of MSH2 results in a form of HNPCC (see HNPCC8, 613244). Other colorectal cancer syndromes include autosomal recessive adenomatous polyposis (608456), which is caused by mutations in the MUTYH gene (604933), and oligodontia-colorectal cancer syndrome (608615), which is caused by mutations in the AXIN2 gene (604025). Colorectal cancer is a predominant form of cancer in tumor predisposition syndrome-4 (TPDS4; 609265), which is caused by mutation in the CHEK2 gene (604373). A germline mutation in the PLA2G2A gene (172411) was identified in a patient with colorectal cancer.

Germline susceptibility loci for colorectal cancer have also been identified. CRCS1 (608812) is conferred by mutation in the GALNT12 gene (610290) on chromosome 9q22; CRCS2 (611469) maps to chromosome 8q24; CRCS3 (612229) is conferred by variation in the SMAD7 gene (602932) on chromosome 18; CRCS4 (601228) is conferred by variation on 15q that causes increased and ectopic expression of the GREM1 gene (603054); CRCS5 (612230) maps to chromosome 10p14; CRCS6 (612231) maps to chromosome 8q23; CRCS7 (612232) maps to chromosome 11q23; CRCS8 (612589) maps to chromosome 14q22; CRCS9 (612590) maps to 16q22; CRCS10 (612591) is conferred by mutation in the POLD1 gene (174761) on chromosome 19q13; CRCS11 (612592) maps to chromosome 20p12; and CRCS12 (615083) is conferred by mutation in the POLE gene (174762) on chromosome 12q24.

Somatic mutations in many different genes, including KRAS (190070), PIK3CA (171834), BRAF (164757), CTNNB1 (116806), FGFR3 (134934), AXIN2 (604025), AKT1 (164730), MCC (159350), MYH11 (160745), PARK2 (602544), RNF43 (612482), and BUB1 (601452), have been identified in colorectal cancer.


Clinical Features

Colon cancer is a well-known feature of familial polyposis coli. Cancer of the colon occurred in 7 members of 4 successive generations of the family reported by Kluge (1964), leading him to suggest a simple genetic basis for colonic cancer independent of polyposis. The combination of colonic and endometrial cancer has been observed in many families (e.g., Williams, 1978).

Sivak et al. (1981) studied a kindred with the familial cancer syndrome in which every confirmed affected member had at least 1 primary carcinoma of the colon. The average age at which cancer appeared was 38 years. Multiple primary neoplasms occurred in 23% of cancer patients.

Budd and Fink (1981) reported a family with a high frequency of mucoid colonic carcinoma. Since endometrial carcinoma, atypical endometrial hyperplasia, uterine leiomyosarcoma, bladder transitional carcinoma, and renal cell carcinoma also occurred in the family, this may be the same disorder as the Lynch cancer family syndrome type II (120435).

Bamezai et al. (1984) reported an Indian Sikh kindred in which 8 persons suffered from cancer of the cecum, not associated with polyposis.

Burt et al. (1985) studied a large Utah kindred called to attention because of occurrence of colorectal cancer in a brother, a sister, and a nephew. No clear inheritance pattern was discernible until systematic screening was undertaken for colonic polyps using flexible proctosigmoidoscopy. One or more adenomatous polyps were found in 41 of 191 family members (21%) and 12 of 132 controls (9%)--p less than 0.005. Pedigree analysis showed best fit with autosomal dominant inheritance. Cannon-Albright et al. (1988) extended the studies with investigations of 33 additional kindreds. The kindreds were selected through either a single person with an adenomatous polyp or a cluster of relatives with colonic cancer. The kindreds all had common colorectal cancers, not the rare inherited condition of familial polyposis coli or nonpolyposis inherited colorectal cancer. Likelihood analysis strongly supported dominant inheritance of a susceptibility to colorectal adenomas and cancers, with a gene frequency of 19%. According to the most likely genetic model, adenomatous polyps and colorectal cancers occur only in genetically susceptible persons; however, the 95% confidence interval for this proportion was 53 to 100%.

Ponz de Leon et al. (1992) analyzed data on 605 families of probands with colorectal cancer in the province of Modena in Italy. Among the 577 presumed nonpolyposis cases, both parents had colorectal cancer in 11, one parent in 130, and neither parent in 436. Segregation was compatible with dominant transmission of susceptibility to cancer.

Mecklin (1987) investigated the frequency of hereditary colorectal cancer among all colorectal cancer patients diagnosed in 1 Finnish county during the 1970s. The cancer family syndrome type of hereditary nonpolyposis colorectal carcinoma emerged as the most common verifiable risk factor, involving between 3.8 and 5.5% of all colorectal cancer patients. The frequencies of familial adenomatosis and ulcerative colitis were 0.2% and 0.6%, respectively. The observed frequency is probably an underestimate. The patients with cancer family syndrome were young, accounting for 29 to 39% of the patients under 50 years of age, and their tumors were located predominantly (65%) in the right hemicolon.


Pathogenesis

The state of DNA methylation appears to play a role in genetic instability in colorectal cancer cells. Lengauer et al. (1997) noted that DNA methylation is essential in prokaryotes, dispensable in lower eukaryotes (such as Saccharomyces cerevisiae) yet present and presumably important in mammals. Many cancers have been shown to have a global hypomethylation of DNA compared with normal tissues. Treatment of cells or animals with 5-azacytidine (5-aza-C), a demethylating agent that irreversibly inactivates methyltransferase (see 156569), is oncogenic in vitro and in vivo. Conversely, other studies showed that hypermethylation of specific sequences found in some tumors can be associated with the inactivation of tumor suppressor gene expression. Mice genetically deficient in methyltransferase are resistant to colorectal tumorigenesis initiated by mutation of the APC (611731) tumor suppressor gene, and treatment of these mice with 5-aza-C enhances the resistance (Laird et al., 1995).

Lengauer et al. (1997) reported a striking difference in the expression of exogenously introduced retroviral genes in various colorectal cancer cell lines. Extinguished expression was associated with DNA methylation and could be reversed by treatment with the demethylating agent 5-aza-C. A striking correlation between genetic instability and methylation capacity suggested that methylation abnormalities may play a role in the chromosome segregation processes in cancer cells. It has been speculated that genetic instability is necessary for a tumor to accumulate the numerous genetic alterations that accompany carcinogenesis. There appeared to exist 2 pathways of genetic instability in colorectal cancer. The first is found in about 15% of tumors and involves point mutations, microdeletions, and microinsertions associated with deficiency of mismatch repair (MMR). The second is found in MMR-proficient cells and involves gains and losses of whole chromosomes. Lengauer et al. (1997) suggested that methylation abnormalities are intrinsically and directly involved in the generation of the second type of instability, thus allowing for the selection of methylation-negative cells during the clonal evolution of tumors. The hypothesis was supported by the observation that demethylation is associated with chromosomal aberrations, including mitotic dysfunction and translocation, and was consistent with the hypothesis relating methylation and aneuploidy put forward by Thomas (1995). Jones and Gonzalgo (1997) commented on altered DNA methylation and genome instability as a new pathway to cancer.

In a second report, Lengauer et al. (1997) showed that tumors without microsatellite instability exhibit a striking defect in chromosome segregation, resulting in gains or losses in excess of 10(-2) per chromosome per generation. This form of chromosomal instability reflected a continuing cellular defect that persisted throughout the lifetime of the tumor cell and was not simply related to chromosome number. While microsatellite instability is a recessive trait, chromosomal instability appeared to be dominant. The data indicated that persistent genetic instability may be critical for the development of all colorectal cancers, and that this instability can arise through 2 distinct pathways.

Adenocarcinoma of the small intestine is rare in the general population, but its histologic features are similar to those of the much more common colorectal adenocarcinoma, and it is seen as part of the HNPCC tumor predisposition spectrum. Wheeler et al. (2002) examined the possible role of mismatch repair defects in the pathogenesis of sporadic small intestinal adenocarcinoma. The replication error status was determined in a total of 21 nonfamilial, nonampullary small intestinal adenocarcinomas: only 1 tumor was scored as replication error-positive. This tumor showed normal immunostaining for MLH1 (see 120436) and MSH2. The authors commented that this result may reflect an epigenetic change in the tumor rather than germline mutation in a mismatch repair gene, and concluded that mismatch repair defects were unlikely to contribute significantly to the genetic pathway leading to sporadic small intestinal adenocarcinoma.

Vilar and Gruber (2010) reviewed the role of microsatellite instability (MSI) in the development of CRC. They stated that approximately 15% of CRCs display MSI owing either to epigenetic silencing of MLH1 or to a germline mutation in one of the mismatch repair genes MLH1, MSH2, MSH6, or PMS2. They noted that MSI tumors have a better prognosis than microsatellite stable CRCs, but that MSI cancers do not necessarily have the same response to the chemotherapeutic strategies used to treat microsatellite stable tumors.

Batlle et al. (2005) showed that although Wnt (see 164820) signaling remains constitutively active, most human colorectal cancers lose expression of EphB (see 600600) at the adenoma-carcinoma transition. They found that loss of EphB expression strongly correlated with degree of malignancy. Furthermore, reduction of EphB activity accelerated tumorigenesis in the colon and rectum of Apc(Min/+) mice (see 611731), and resulted in formation of aggressive adenocarcinomas. Batlle et al. (2005) concluded that loss of EphB expression represents a critical step in colorectal cancer progression.

By microdissection of bifurcating colonic crypts and sequencing of the entire mitochondrial genome in all of the cells, Greaves et al. (2006) demonstrated that stochastic mutations in mtDNA resulting in phenotypic cytochrome c oxidase (COX) deficiency of were identical in both arms of a crypt that was bifurcating. Furthermore, they showed that patches of neighboring crypts deficient in COX also shared identical mitochondrial mutations, and that these patches increased in size with age, indicating that crypt fission is a mechanism by which mutations can spread within the colon.

Xia et al. (2012) showed that prostaglandin E2 (PGE2) silences certain tumor suppressor and DNA repair genes through DNA methylation to promote tumor growth. Their findings uncovered a theretofore unrecognized role for PGE2 in the promotion of tumor progression, and provided a rationale for considering the development of a combination treatment using PTGS2 (600262) inhibitors and demethylating agents for the prevention or treatment of colorectal cancer.

Seshagiri et al. (2012) systematically analyzed more than 70 pairs of primary human colon tumors using next-generation sequencing to characterize their exomes, transcriptomes, and copy number alterations. They identified 36,303 protein-altering somatic changes that included several novel recurrent mutations in the Wnt pathway gene TCF7L2 (602228), chromatin-remodeling genes such as TET2 (612839) and TET3 (613555), and receptor tyrosine kinases including ERBB3 (190151). The analysis for significantly mutated cancer genes identified 23 candidates, including the cell cycle checkpoint kinase ATM (607585). Copy number and RNA-seq data analysis identified amplifications and corresponding overexpression of IGF2 in a subset of colon tumors. Furthermore, using RNA-seq data, Seshagiri et al. (2012) identified multiple fusion transcripts including recurrent gene fusions involving R-spondin family members RSPO2 (610575) and RSPO3 (610574) that together occur in 10% of colon tumors. The RSPO fusions were mutually exclusive with APC (611731) mutations, indicating that they probably have a role in the activation of Wnt signaling and tumorigenesis. Consistent with this, Seshagiri et al. (2012) showed that RSPO fusion proteins were capable of potentiating Wnt signaling.

Grivennikov et al. (2012) investigated mechanisms responsible for tumor-elicited inflammation in a mouse model of colorectal tumorigenesis which, like human colorectal cancer, exhibits upregulation of IL23 (605580) and IL17 (603149). They showed that IL23 signaling promotes tumor growth and progression, and development of tumoral IL17 response. IL23 is mainly produced by tumor-associated myeloid cells that are likely to be activated by microbial products, which penetrate the tumors but not adjacent tissue. Both early and late colorectal neoplasms exhibit defective expression of several barrier proteins. Grivennikov et al. (2012) proposed that barrier deterioration induced by colorectal cancer-initiating genetic lesions results in adenoma invasion by microbial products that trigger tumor-elicited inflammation, which in turn drives tumor growth.

Huber et al. (2012) described the crucial role of IL22BP (606648) in controlling tumorigenesis and epithelial cell proliferation in the colon. IL22BP is highly expressed by dendritic cells in the colon in steady-state conditions. Sensing of intestinal tissue damage via the NLRP3 (606416) or NLRP6 (609650) inflammasomes led to an IL18 (600953)-dependent downregulation of IL22BP, thereby increasing the ratio of IL22 (605330)/IL22BP. IL22, which is induced during intestinal tissue damage, exerted protective properties during the peak of damage, but promoted tumor development if uncontrolled during the recovery phase. Thus, the IL22-IL22BP axis critically regulates intestinal tissue repair and tumorigenesis in the colon.

Vermeulen et al. (2013) quantified the competitive advantage during tumor development of Apc (611731) loss, Kras (190070) activation, and p53 (191170) mutations in the mouse intestine. Their findings indicated that the fate conferred by these mutations is not deterministic, and many mutated stem cells are replaced by wildtype stem cells after biased but still stochastic events. Furthermore, Vermeulen et al. (2013) found that p53 mutations display a condition-dependent advantage, and especially in colitis-affected intestines, clones harboring mutations in this gene were favored. Vermeulen et al. (2013) concluded that their work confirmed the notion that the tissue architecture of the intestine suppresses the accumulation of mutated lineages.

Liu et al. (2015) demonstrated that genomic deletion of TP53 frequently encompasses essential neighboring genes, rendering cancer cells with hemizygous TP53 deletion vulnerable to further suppression of such genes. The authors identified POLR2A (180660) as such a gene that is almost always codeleted with TP53 in human cancers. It encodes the largest and catalytic subunit of the RNA polymerase II complex, which is specifically inhibited by alpha-amanitin. Liu et al. (2015) analyzed the Cancer Genome Atlas (TCGA) and Cancer Cell Line Encyclopedia (CCLE) databases, which revealed that expression levels of POLR2A are tightly correlated with its gene copy numbers in human colorectal cancer. Suppression of POLR2A with alpha-amanitin or siRNAs selectively inhibits the proliferation, survival, and tumorigenic potential of colorectal cancer cells with hemizygous TP53 loss in a p53-independent manner. Clinical applications of alpha-amanitin had been limited owing to its liver toxicity; however, Liu et al. (2015) found that alpha-amanitin-based antibody-drug conjugates (Moldenhauer et al., 2012) are highly effective therapeutic agents with reduced toxicity. Liu et al. (2015) showed that low doses of alpha-amanitin-conjugated anti-epithelial cell adhesion molecule (EpCAM; 185535) antibody led to complete tumor regression in mouse models of human colorectal cancer with hemizygous deletion of POLR2A.


Clinical Management

Various laboratory, clinical, and epidemiologic evidence suggested that calcium may help prevent colorectal adenomas. Baron et al. (1999) conducted a randomized, double-blind trial of the effect of supplementation with calcium carbonate on the recurrence of colorectal adenomas. They found a significant, though moderate, reduction in the risk of recurrent colorectal adenomas in the supplemented group.

In randomized trials of aspirin to determine its efficacy in prevention of colorectal adenomas, Sandler et al. (2003) and Baron et al. (2003) studied patients with either previous colorectal cancer or recent histologically documented adenomas, respectively. Both studies found that aspirin was associated with a significant reduction in the incidence of colorectal adenomas.

Inhibition of the BRAF(V600E) (164757.0001) oncoprotein by the small-molecule drug PLX4032 (vemurafenib) is highly effective in the treatment of melanoma. However, colon cancer patients harboring the same BRAF(V600E) oncogenic lesion have poor prognosis and show only a very limited response to this drug. To investigate the cause of this limited therapeutic effect in BRAF(V600E) mutant colon cancer, Prahallad et al. (2012) performed an RNA interference-based genetic screen in human cells to search for kinases whose knockdown synergizes with BRAF(V600E) inhibition. They reported that blockade of the epidermal growth factor receptor (EGFR; 131550) shows strong synergy with BRAF(V600E) inhibition. Prahallad et al. (2012) found in multiple BRAF(V600E) mutant colon cancers that inhibition of EGFR by the antibody drug cetuximab or the small-molecule drugs gefitinib or erlotinib is strongly synergistic with BRAF(V600E) inhibition, both in vitro and in vivo. Mechanistically, Prahallad et al. (2012) found that BRAF(V600E) inhibition causes a rapid feedback activation of EGFR, which supports continued proliferation in the presence of BRAF(V600E) inhibition. Melanoma cells express low levels of EGFR and are therefore not subject to this feedback activation. Consistent with this, Prahallad et al. (2012) found that ectopic expression of EGFR in melanoma cells is sufficient to cause resistance to PLX4032. Prahallad et al. (2012) concluded that BRAF(V600E) mutant colon cancers (approximately 8 to 10% of all colon cancers) might benefit from combination therapy consisting of BRAF and EGFR inhibitors.

Development of Resistance to Chemotherapeutic Agents

Antibodies against EGFR, cetuximab and panitumumab, are widely used to treat colorectal cancer. Unfortunately, patients eventually develop resistance to these agents. Montagut et al. (2012) described an acquired EGFR ectodomain mutation (S492R) that prevents cetuximab binding and confers resistance to cetuximab. Cells with this mutation, however, retain binding to and are growth inhibited by panitumumab. Two of 10 subjects studied with metastatic colon cancer progression after cetuximab treatment acquired this mutation. One subject with cetuximab resistance harboring the S492R mutation responded to treatment with panitumumab.

Misale et al. (2012) showed that molecular alterations (in most instances point mutations) of KRAS (190070) are causally associated with the onset of acquired resistance to anti-EGFR treatment in colorectal cancers. Expression of mutant KRAS under the control of its endogenous gene promoter was sufficient to confer cetuximab resistance, but resistant cells remained sensitive to combinatorial inhibition of EGFR and mitogen-activated protein kinase kinase (MEK; see 176872). Analysis of metastases from patients who developed resistance to cetuximab or panitumumab showed the emergence of KRAS amplification in one sample and acquisition of secondary KRAS mutations in 60% (6 out of 10) of the cases. KRAS mutant alleles were detectable in the blood of cetuximab-treated patients as early as 10 months before radiographic documentation of disease progression. Misale et al. (2012) concluded that their results identified KRAS mutations as frequent drivers of acquired resistance to cetuximab in colorectal cancers, indicated that the emergence of KRAS mutant clones can be detected noninvasively months before radiographic progression, and suggested early initiation of a MEK inhibitor as a rational strategy for delaying or reversing drug resistance.

Diaz et al. (2012) determined whether mutant KRAS DNA could be detected in the circulation of 28 patients receiving monotherapy with panitumumab, a therapeutic anti-EGFR antibody. They found that 9 out of 24 (38%) patients whose tumors were initially KRAS wildtype developed detectable mutations in KRAS in their sera, 3 of which developed multiple different KRAS mutations. The appearance of these mutations was very consistent, generally occurring between 5 and 6 months following treatment. Mathematical modeling indicated that the mutations were present in expanded subclones before the initiation of panitumumab treatment. Diaz et al. (2012) concluded that the emergence of KRAS mutations is a mediator of acquired resistance to EGFR blockade and that these mutations can be detected in a noninvasive manner. The results also explained why solid tumors develop resistance to targeted therapies in a highly reproducible fashion.

Among 512 patients who had metastatic colorectal cancer without RAS (KRAS or NRAS, 164790) mutations, Douillard et al. (2013) found that progression-free survival was 10.1 months with the combination of panitumumab-FOLFOX4 (oxaliplatin, fluorouracil, and leucovorin) versus 7.9 months with FOLFOX4 alone (hazard ratio for progression or death with combination therapy, 0.72; 95% CI 0.58 to 0.90; p = 0.004). Overall survival was 26.0 months in the panitumumab-FOLFOX4 group versus 20.2 months in the FOLFOX4-alone group (hazard ratio for death, 0.78; 95% CI 0.62-0.99; p = 0.04). A total of 108 patients (17%) with nonmutated KRAS exon 2 had other RAS mutations. These mutations were associated with inferior progression-free survival and overall survival with panitumumab-FOLFOX4 treatment, which was consistent with the findings in patients with KRAS mutations in exon 2. BRAF mutations were a negative prognostic factor.


Diagnosis

Prediction of Colorectal Cancer Risk

Loss of imprinting, an epigenetic alteration affecting the insulin-like growth factor II gene (IGF2; 147470), is found in normal colonic mucosa of about 30% of colorectal cancer patients, but it is found in only 10% of healthy individuals. In a pilot study to investigate the utility of loss of imprinting as a marker of colorectal cancer risk, Cui et al. (2003) evaluated 172 patients at a colonoscopy clinic. The adjusted odds ratio for loss of imprinting in lymphocytes was 5.15 for patients with a positive family history (95% CI, 1.70-16.96; p = 0.002), 3.46 for patients with adenomas (95% CI, 1.14-11.37; p = 0.026), and 21.7 for patients with colorectal cancer (95% CI, 3.48-153.6; p = 0.0005). Loss of imprinting can be assayed with a DNA-based blood test, and Cui et al. (2003) concluded that it may be a valuable predictive marker of an individual's risk for colorectal cancer.


Mapping

To identify susceptibility genes for familial colorectal neoplasia, Daley et al. (2008) conducted a comprehensive, genomewide linkage scan of 194 kindreds. Clinical information (histopathology, size and number of polyps, and other primary cancers) was used in conjunction with age at onset and family history for classification of the families into 5 phenotypic subgroups (severe histopathology, oligopolyposis, young colon/breast and multiple cancer) before analysis. By expanding the traditional affected sib pair design to include unaffected and discordant sib pairs, analytical power and robustness to type I error were increased. Linkage peaks of interest were identified at several sites. At marker D1S1665 (1p31.1), there was strong evidence for linkage in the multiple cancer subgroup (p = 0.00007). For 15q14-q22, a linkage peak was identified in the full sample, oligopolyposis, and young phenotypes. This region includes the locus associated with hereditary mixed polyposis syndrome (HMPS; 601228) in families of Ashkenazi descent. Daley et al. (2008) provided compelling evidence linking this region in families of European descent with oligopolyposis and/or young age at onset (51 years or younger) phenotypes. They found linkage to BRCA2 (600185) in the colon/breast phenotypic subgroup and identified a second locus in the region of D21S1437 segregating with, but distinct from, BRCA2. Linkage to 17p13.3 at marker D17S1308 in the breast/colon subgroup identified HIC1 (603825) as a candidate gene. The study demonstrated that using clinical information, unaffected sibs, and family history can increase the analytic power of a linkage study.

Associations Pending Confirmation

In a large kindred with excess colorectal cancer, Neklason et al. (2010) performed 2 separate genomewide scans and additional fine mapping and identified a single major locus on chromosome 13q22.1-q31.3 that segregated with adenomatous polyps and colon cancer, for which they obtained a nonparametric linkage score of 24 (lod score of 2.99; p = 0.001) at D13S251. Haplotype analysis identified a 21-Mb interval encompassing a nonrecombinant region bounded by rs2077779 and rs2351871 and containing 27 genes. Sequencing of 8 candidate genes failed to identify a clearly deleterious mutation. Neklason et al. (2010) noted that chromosome 13q is commonly gained and overexpressed in colon cancers and correlates with metastasis, suggesting the presence of an important cancer progression gene, and stated that evaluation of tumors from the kindred revealed a gain of chromosome 13q as well.


Cytogenetics

Bass et al. (2011) reported whole-genome sequencing from 9 individuals with colorectal cancer, including primary colorectal tumors and matched adjacent nontumor tissues, at an average of 30.7x and 31.9x coverage, respectively. They identified an average of 75 somatic rearrangements per tumor, including complex networks of translocations between pairs of chromosomes. Eleven rearrangements encode predicted in-frame fusion proteins, including a fusion of VTI1A (614316) and TCF7L2 (602278) found in 3 out of 97 colorectal cancers. Although TCF7L2 encodes TCF4, which cooperates with beta-catenin (116806) in colorectal carcinogenesis, the fusion lacks the TCF4 beta-catenin-binding domain. Bass et al. (2011) found a colorectal carcinoma cell line harboring the fusion gene to be dependent on VTI1A-TCF7L2 for anchorage-independent growth using RNA interference-mediated knockdown.


Molecular Genetics

In the DNA from 1 colon and 2 lung carcinoma cell lines, Perucho et al. (1981) demonstrated the same or closely related transforming elements. By DNA-mediated gene transfer, mouse fibroblasts could be morphologically transformed and rendered tumorigenic in nude mice.

In preliminary observations, Pathak and Goodacre (1986) found deletion of 12p in colorectal cancer specimens.

Fearon et al. (1987) studied the clonal composition of human colorectal tumors. Using X-linked RFLPs, they showed that all 50 tumors from females showed a monoclonal pattern of X-chromosome inactivation; these tumors included 20 carcinomas and 30 adenomas of either familial or spontaneous type. In over 75% of carcinomas examined, somatic loss of chromosome 17p sequences was found; such loss was rare in adenomas. Fearon et al. (1987) suggested that a gene on the short arm of chromosome 17 may be associated with progression from the benign to the malignant state.

By a combination of DNA hybridization analyses and tissue sectioning techniques, Bos et al. (1987) demonstrated that RAS gene mutations occur in over a third of colorectal cancers, that most of the mutations are at codon 12 of the KRAS gene (190070), and that the mutations usually precede the development of malignancy.

In 38 tumors from 25 patients with familial polyposis coli, and in 20 sporadic colon carcinomas, Okamoto et al. (1988) found frequent occurrence of allele loss on chromosome 22, with some additional losses on chromosomes 5, 6, 12q, and 15. The DNA probe C11p11, which has been found to be linked to familial polyposis coli, also detected frequent allele loss in both familial and sporadic colon carcinomas but not in benign adenomas. In a more extensive study, Vogelstein et al. (1988) studied the interrelationships of the 4 alterations demonstrated in colorectal cancer (RAS gene mutations and deletions of chromosome 5, 17 and 18 sequences) and determined their occurrence with respect to different stages of colorectal tumorigenesis. They found RAS gene mutations frequently in adenomas, this being the first demonstration of such in benign human tumors. In adenomas greater than 1 cm in size, the prevalence was similar to that observed in carcinomas (58% and 47%, respectively). Sequences on chromosome 5 that are linked to familial adenomatous polyposis were seldom lost in adenomas from such patients. Therefore, the Knudson model is unlikely to be applicable to the adenoma/carcinoma sequence in this disorder. Chromosome 18 sequences were lost frequently in colon carcinomas (73%) and in advanced adenomas (47%), but only occasionally in earlier stage adenomas (11-13%); see 120470. Chromosome 17 sequences were usually lost only in carcinomas (75%). The results suggested a model wherein the steps required for malignancy involve the activation of a dominantly acting oncogene coupled with the loss of several genes that normally suppress tumorigenesis.

Wildrick and Boman (1988) found deletion of the glucocorticoid receptor locus (138040), located on 5q, in colorectal cancers.

Law et al. (1988) examined the question of whether the gene for familial polyposis coli on chromosome 5 may be the site of changes leading to colorectal cancer in the general population, analogous to recessive tumor genes in retinoblastoma and Wilms tumor. To avoid error in interpretation of allelic loss from a study of nonhomogeneous samples, tumor cell populations were first microdissected from 24 colorectal carcinomas, an additional 9 cancers were engrafted in nude mice, and nuclei were flow-sorted in an additional 2. Of 31 cancers informative for chromosome 5 markers, only 6 (19%) showed loss of heterozygosity of chromosome 5 alleles, compared to 19 of 34 (56%) on chromosome 17, and 17 of 33 (52%) on chromosome 18. Law et al. (1988) concluded that FPC is a true dominant for adenomatosis but not a common recessive gene for colon cancer, and that simple mendelian models involving loss of alleles at a single locus may be inappropriate for understanding common human solid tumors.

Vogelstein et al. (1989) examined the extent and variation of allelic loss for polymorphic DNA markers in every nonacrocentric autosomal arm in 56 paired colorectal carcinoma and adjacent normal colonic mucosa specimens. They referred to the analysis as an allelotype, in analogy with a karyotype. Three major conclusions were drawn from the study: (1) Allelic deletions are remarkably common; 1 of the alleles of each polymorphic marker tested was lost in at least some tumors, and some tumors lost more than half of their parental alleles. (2) In addition to allelic deletions, new DNA fragments not present in normal tissue were identified in 5 carcinomas; these new fragments contained repeated sequences (of the variable-number-of-tandem-repeat type). (3) Patients with more than the median percentage of allelic deletions had a considerably worse prognosis than did the other patients, although the stage and size of the primary tumors were very similar in the 2 groups.

Delattre et al. (1989) reviewed the 3 general types of genetic alterations in colorectal cancer: (1) change in DNA content of the malignant cells as monitored by flow cytometry; (2) specific loss of genetic material, i.e., a complete loss of chromosome 18 and a structural rearrangement of chromosome 17 leading most often to the loss of 1 short arm, and loss of part of 5q as demonstrated by loss of heterozygosity; and (3) in nearly 40% of tumors, activation by point mutation of RAS oncogenes (never HRAS, rarely NRAS, and most frequently KRAS). In KRAS, with 1 exception, the activation has always occurred by a change in the coding properties of the twelfth or thirteenth codon. In studies of the multiple genetic alterations in colorectal cancer, Delattre et al. (1989) found that deletions and mitotic abnormalities occurred more frequently in distal than in proximal tumors. The frequency of KRAS mutations did not differ between proximal and distal cancers.

In studies of 15 colorectal tumors, Konstantinova et al. (1991) found rearrangements of the short arm of chromosome 17, leading to deletion of this arm or part of it in 12; in 2 others, one of the homologs of pair 17 was lost. One chromosome 18 was lost in 12 out of 13 cases with fully identified numerical abnormalities; chromosome 5, in 6 tumors; and other chromosomes in lesser numbers of cases. See 120470 for a discussion of a gene on chromosome 18 called DCC ('deleted in colorectal cancer') that shows mutations, including point mutations, in colorectal tumor tissue; also see 164790 for a discussion of a mutation in the NRAS oncogene in colorectal cancer.

On the basis of complex segregation analysis of a published series of consecutive pedigrees ascertained through patients undergoing treatment for colorectal cancer, Houlston et al. (1992) concluded that a dominant gene (or genes) with a frequency of 0.006 with a lifetime penetrance of 0.63 is likely. The gene was thought to account for 81% of colorectal cancer in patients under 35 years of age; however, by age 65, about 85% appeared to be phenocopies.

Fearon and Vogelstein (1990) reviewed the evidence supporting their multistep genetic model for colorectal tumorigenesis. They suggested that multiple mutations lead to a progression from normal epithelium to metastatic carcinoma through hyperplastic epithelium--early adenoma--intermediate adenoma--late adenoma--and carcinoma. The genes in which mutations occur at steps in this process include APC (611731) on chromosome 5, KRAS on chromosome 12, TP53 (191170) on 17p, and DCC on chromosome 18. Other genes that have been demonstrated or suspected of involvement in colorectal cancer include MSH2 (609309) on chromosome 2 and the DRA candidate colon tumor-suppressor gene (126650) on chromosome 7. Sarraf et al. (1999) presented evidence that colon cancer in humans is associated with loss-of-function mutations in the PPARG gene (601487).

Kikuchi-Yanoshita et al. (1992) presented evidence that genetic changes in both alleles of the TP53 gene through mutation and LOH, which result in abnormal protein accumulation, are involved in the conversion of adenoma to early carcinoma in both familial adenomatous polyposis and in nonfamilial polyposis cases.

Kinzler and Vogelstein (1996) gave a review of hereditary colorectal cancer and the multistep process of carcinogenesis that typically develops over decades and appears to require at least 7 genetic events for completion. They stated that the genetic defect in FAP involves the rate of tumor initiation by targeting the gatekeeper function of the APC gene. In contrast, the defect in HNPCC largely affects tumor aggression by targeting the genome guardian function of DNA repair.

Rajagopalan et al. (2002) systematically evaluated mutation in BRAF (164757) and KRAS (190070) in 330 colorectal tumors. There were 32 mutations in BRAF, 28 with a V600E mutation (164757.0001) and 1 each with the R462I (164757.0002), I463S (164757.0003), G464E (164757.0004), or K601E (164757.0005) mutations. All but 2 mutations seemed to be heterozygous, and in all 20 cases for which normal tissue was available, the mutations were shown to be somatic. In the same set of tumors there were 169 mutations in KRAS. No tumor exhibited mutations in both BRAF and KRAS. There was also a striking difference in the frequency of BRAF mutations between cancers with and without mismatch repair deficiency. All but 1 of the 15 BRAF mutations identified in mismatch repair deficient cases resulted in a V600E substitution. Rajagopalan et al. (2002) concluded their results provide strong support for the hypothesis that BRAF and KRAS mutations are equivalent in their tumorigenic effects. Both genes seem to be mutated at a similar phase of tumorigenesis, after initiation but before malignant conversion. Moreover, no tumor concurrently contained both BRAF and KRAS mutations.

To determine whether carriers of BLM (604610) mutations are at increased risk of colorectal cancer, Gruber et al. (2002) genotyped 1,244 cases of colorectal cancer and 1,839 controls, both of Ashkenazi Jewish ancestry, to estimate the relative risk of colorectal cancer among carriers of the BLM(Ash) founder mutation. Ashkenazi Jews with colorectal cancer were more than twice as likely to carry the BLM(Ash) (604610.0001) mutation than Ashkenazi Jewish controls without colorectal cancer (odds ratio = 2.45, 95% confidence interval 1.3 to 4.8; p = 0.0065). Gruber et al. (2002) verified that the APC I1307K mutation (611731.0029) did not confound their results.

Lynch and de la Chapelle (2003) provided a general discussion of hereditary colorectal cancer. They presented a flow diagram of the breakdown of 1,044 unselected consecutive patients with colorectal cancer. Tumors from 129 patients (12%) were positive for microsatellite instability; 28 of these patients were positive for germline mutations in MLH1 or MSH2, giving HNPCC a 2.7% frequency among the 1,044 patients. In the 88% of the patients whose tumors had no microsatellite instability, no mutations were found in MLH1 or MSH2.

Bardelli et al. (2003) used high-throughput sequencing technologies and bioinformatics to investigate how many or how often members of the tyrosine kinase family were altered in any particular cancer type. The protein kinase complement of the human genome (the 'kinome') can be organized into a dendrogram containing 9 broad groups of genes. Bardelli et al. (2003) selected 1 major branch of this dendrogram, containing 3 of the 9 groups, including the 90 tyrosine kinase genes (TK group), the 43 tyrosine kinase-like genes (TKL group), and the 5 receptor guanylate cyclase genes (RGC group), for mutation analysis. The 819 exons containing the kinase domains from the annotated TK, TKL, and RGC genes were screened from 35 colorectal cancer cell lines and were directly sequenced. Fourteen genes had somatic mutations within their kinase domains. Bardelli et al. (2003) analyzed these 14 genes for mutations in another 147 colorectal cancers and identified 46 mutations, 2 of which were synonymous; the remainder were either nonsynonymous or splice site alterations. All of these mutations were found to be somatic in the cancers that could be assessed by sequencing DNA from matched normal tissue. Seven genes were mutated in more than 1 tumor in the cohort: NTRK3 (191316), FES (190030), KDR (191306), EPHA3 (179611), NTRK2 (600456), MLK4, and GUCY2F (300041).

Samuels et al. (2004) examined the sequences of 117 exons that encode the predicted kinase domains of 8 phosphatidylinositol-3 kinase genes and 8 PI3K-like genes in 35 colorectal cancers. PIK3CA (171834) was the only gene with somatic mutations. Subsequent sequence analysis of all coding exons of PIK3CA in 199 additional colorectal cancers revealed mutations in a total of 74 tumors (32%). Samuels et al. (2004) also evaluated 76 premalignant colorectal tumors; only 2 mutations were found, both in very advanced tubulovillous adenomas greater than 5 cm in diameter. Thus, Samuels et al. (2004) concluded that PIK3CA mutations generally arise late in tumorigenesis, just before or coincident with invasion. Mutations in PIK3CA were also identified in 4 of 15 glioblastomas (27%), 3 of 12 gastric cancers (25%), 1 of 12 breast cancers (8%), and 1 of 24 lung cancers (4%). No mutations were observed in 11 pancreatic cancers or 12 medulloblastomas. In total, 92 mutations were observed, all of which were determined to be somatic in the cancers that could be assessed. Samuels et al. (2004) concluded that the sheer number of mutations observed in this gene strongly suggests that they are functionally important. Furthermore, most of the mutations were nonsynonymous and occurred in the PI3K helical and kinase domains, suggesting functional significance.

Clear-cut inherited mendelian traits, such as FAP or HNPCC, account for less than 4% of colorectal cancers. Another 20% of all colorectal cancers are thought to occur in individuals with a significant inherited multifactorial susceptibility to colorectal cancer that is not obviously familial. Incompletely penetrant, comparatively rare missense variants in the APC gene (611731) have been described in patients with multiple colorectal adenomas. For example, the I1307K mutation in the APC gene, which is found in Ashkenazi Jewish populations with an incidence of approximately 6%, confers a significantly increased risk of developing multiple adenomas and colorectal cancer. The glu1317-to-gln mutation in the APC gene (E1317Q; 611731.0036), which is found in non-Jewish Caucasian populations at a low frequency, similarly appears to confer a significantly increased risk of multiple adenomatous polyps. These variants represent a category of variation that has been suggested, generally, to account for a substantial fraction of such multifactorial inherited susceptibility to colorectal cancer. Fearnhead et al. (2004) explored this rare variant hypothesis for multifactorial inheritance using multiple colorectal adenomas as the model. Patients with multiple adenomas were screened for germline variants in a panel of candidate genes. Germline DNA was obtained from 124 patients with 3 to 100 histologically proven synchronous or metachronous adenomatous polyps. All patients were tested for the APC gene variants I1307K and E1317Q and for variants in the AXIN1 (603816), CTNNB1, MLH1, and MSH2 genes. The control group consisted of 483 randomly selected individuals. Potentially pathogenic germline variants were found in 30 of 124 patients (24.9%), compared with 55 of 483 controls (approximately 12%). This overall difference was highly significant, suggesting that many rare variants collectively contribute to inherited susceptibility to colorectal adenomas.

Parsons et al. (2005) selected 340 genes encoding serine/threonine kinases from the human genome and analyzed them for mutations in the kinase domain in tumors from colorectal cancer patients. A total of 23 changes, including 20 nonsynonymous point mutations, 1 insertion, and 1 splice site alteration, were identified. The gene mutations affected 8 different proteins: 6 were in mitogen-activated protein kinase kinase-4 (MKK4/JNKK1; 601335), 6 in myosin light-chain kinase-2 (MYLK2; 606566), 3 in phosphoinositide-dependent protein kinase-1 (PDK1; 605213, of which 2 mutations affected the same residue in the kinase domain), 2 in p21-activated kinase-4 (PAK4; 605451), 2 in v-akt murine thymoma viral oncogene homolog-2 kinase (AKT2; 164731), and 2 in MAP/microtubule affinity-regulating kinase-3 (MARK3; 602678); there was 1 alteration in cell-division cycle-7 kinase (CDC7; 603311) and another in a hypothetical casein kinase (PDIK1L). Eighteen of the 23 somatic mutations occurred at evolutionarily conserved residues. MKK4/JNKK1 is altered in a variety of tumor types, but no mutations in any of the other genes had theretofore been found in colorectal cancers. Three of the altered genes, PDK1, AKT2, and PAK4, encode proteins involved in the phosphatidylinositol-3-hydroxykinase pathway, and 2 of these (AKT2 and PAK4) are overexpressed in human cancers. Overall, nearly 40% of colorectal tumors had alterations in 1 of 8 PI(3)K-pathway genes.

Boraska Jelavic et al. (2006) studied genotype and allele frequencies of the GT microsatellite repeat polymorphism in intron 2 of the TLR2 gene (603028.0002) in 89 Croatian patients with sporadic colorectal cancer and 88 Croatian sex- and age-matched controls. The frequency of TLR2 alleles with 20 and 21 GT repeats was decreased (p = 0.0044 and p = 0.001, respectively) and the frequency of the allele with 31 GT repeats was increased (p = 0.0147) in patients versus controls. The authors also found that the gly299 allele of the TLR4 gene (603030.0001) was more frequent in colorectal cancer patients than controls (p = 0.0269).

Sjoblom et al. (2006) determined the sequence of well-annotated human protein-coding genes in 2 common tumor types. Analysis of 13,023 genes in 11 breast and 11 colorectal cancers revealed that individual tumors accumulate an average of about 90 mutant genes, but that only a subset of these contribute to the neoplastic process. Using stringent criteria to delineate this subset, Sjoblom et al. (2006) identified 189 genes (average of 11 per tumor) that were mutated at significant frequency. The vast majority of these were not known to be genetically altered in tumors and were predicted to affect a wide range of cellular functions, including transcription, adhesion, and invasion. Sjoblom et al. (2006) concluded that their data defined the genetic landscape of 2 human cancer types, provided new targets for diagnostic and therapeutic intervention, and opened fertile avenues for basic research in tumor biology.

Forrest and Cavet (2007), Getz et al. (2007), and Rubin and Green (2007) commented on the article by Sjoblom et al. (2006), citing statistical problems that, if addressed, would result in the identification of far fewer genes with significantly elevated mutation rates. Parmigiani et al. (2007) responded that the conclusions of the above authors were inaccurate because they were based on analyses that did not fully take into account the experimental design and other critical features of the Sjoblom et al. (2006) study.

To catalog the genetic changes that occur during tumorigenesis, Wood et al. (2007) isolated DNA from 11 breast and 11 colorectal tumors and determined the sequences of the genes in the Reference Sequence database in these samples. Based on analysis of exons representing 20,857 transcripts from 18,191 genes, Wood et al. (2007) concluded that the genomic landscapes of breast and colorectal cancers are composed of a handful of commonly mutated gene 'mountains' and a much larger number of gene 'hills' that are mutated at low frequency. Wood et al. (2007) described statistical and bioinformatic tools that may help identify mutations with a role in tumorigenesis. The gene mountains comprise well-known cancer genes such as APC (611731), KRAS (190070), and TP53 (191170). Furthermore, Wood et al. (2007) observed that most tumors accumulated approximately 80 mutations, and that the majority of these were harmless. Fewer than 15 mutations are likely to be responsible for driving the initiation, progression, or maintenance of the tumor.

Alhopuro et al. (2008) identified somatic mutations in the MYH11 gene in 56 (56%) of 101 samples of colorectal cancer tissue showing microsatellite instability. All 56 mutations were within a mononucleotide repeat of 8 cytosines (C8) in the last exon of the MYH11 SM2 isoform, which is susceptible to mutations under microsatellite instability, and were predicted to lead to a frameshift and elongation of the protein. All mutations were found within epithelial cells. Analysis of microsatellite stable tumors identified 2 somatic mutations in the same tumor that were not in the C8 repeat. Functional expression studies of the mutant proteins showed unregulated actin-activated motor activity.

McMurray et al. (2008) showed that a large proportion of genes controlled synergistically by loss-of-function p53 and Ras activation are critical to the malignant state of murine and human colon cells. Notably, 14 of 24 'cooperation response genes' were found to contribute to tumor formation in gene perturbation experiments. In contrast, only 1 of 14 perturbations of the genes responding in a nonsynergistic manner had a similar effect. McMurray et al. (2008) concluded that synergistic control of gene expression by oncogenic mutations thus emerges as an underlying key to malignancy, and provides an attractive rationale for identifying intervention targets in gene networks downstream of oncogenic gain- and loss-of-function mutations.

To help distinguish between driver and passenger mutations in colorectal cancer, Starr et al. (2009) used a transposon-based genetic screen in mice to identify candidate genes. Mice harboring mutagenic 'Sleeping Beauty' (SB) transposons were crossed with mice expressing SB transposase in gastrointestinal tract epithelium. Most of the offspring developed intestinal lesions including intraepithelial neoplasia, adenomas, and adenocarcinomas. Analysis of over 16,000 transposon insertions identified 77 candidate CRC genes, 60 of which are mutated and/or dysregulated in human CRC and thus are most likely to drive tumorigenesis. The genes included APC, PTEN (601728), and SMAD4 (600993). The screen also identified 17 candidate genes that had not been implicated in CRC, including POLI (605252), PTPRK (602545), and RSPO2 (610575).

In colonocytes from COX-deficient crypts from 2 patients with colon cancer, Greaves et al. (2006) identified 2 missense mutations in the MTCO1 gene (see 516030.0010 and 516030.0011, respectively).

Using high-throughput screening of 14,662 human protein coding transcripts, Sjoblom et al. (2006) found that the PKHD1 gene (606702) was the seventh most common somatically mutated gene in colorectal cancer. Germline mutations in the PKHD1 gene cause autosomal recessive polycystic kidney disease (263200). Ward et al. (2011) observed an association between the common T36M PKHD1 allele (606702.0001) and protection against colorectal cancer. Germline heterozygosity for the mutant allele was found in 0.42% of 3,603 healthy European controls and in 0.027% of 3,767 patients with colorectal cancer (p = 0.0002; odds ratio of 0.072). The authors postulated that reduced PKHD1 activity may enhance mitotic instability, which may inhibit carcinogenesis.

Dorard et al. (2011) identified a mutant of HSP110 (see 610703), which they called HSP110-delta-E9, in colorectal cancer showing microsatellite instability (MSI CRC), generated from an aberrantly spliced mRNA and lacking the HSP110 substrate-binding domain. This mutant was expressed at variable levels in almost all MSI CRC cell lines and primary tumors tested. HSP110-delta-E9 impaired both the normal cellular localization of HSP110 and its interaction with other HSPs, thus abrogating the chaperone activity and antiapoptotic function of HSP110 in a dominant-negative manner. HSP110-delta-E9 overexpression caused the sensitization of cells to anticancer agents such as oxaliplatin and 5-fluorouracil, which are routinely prescribed in the adjuvant treatment of people with colorectal cancer. The survival and response to chemotherapy of subjects with colorectal cancer showing microsatellite instability was associated with the tumor expression level of HSP110-delta-E9. Dorard et al. (2011) concluded that HSP110 may thus constitute a major determinant for both prognosis and treatment response in colorectal cancer.

The Cancer Genome Atlas Network (2012) conducted a genome-scale analysis of 276 colorectal carcinoma samples analyzing exome sequence, DNA copy number, promoter methylation, and mRNA and microRNA expression. A subset of these samples (97) underwent low-depth-of-coverage whole-genome sequencing. In total, 16% of colorectal carcinomas were found to be hypermutated: three-quarters of these had the expected high microsatellite instability, usually with hypermethylation and MLH1 silencing, and one-quarter had somatic mismatch-repair gene and polymerase epsilon mutations. Excluding the hypermutated cancers, colon and rectal cancers were found to have considerably similar patterns of genomic alteration. Twenty-four genes were significantly mutated. In addition to the expected APC, TP53, SMAD4, PIK3CA, and KRAS mutations, the authors found frequent mutations in ARID1A (603024), SOX9 (608160), and FAM123B (300647). Recurrent copy number alterations included potentially drug-targetable amplification of ERBB2 (164870) and amplification of IGF2 (147470). Recurrent chromosomal translocations included the fusion of NAV2 (607026) and WNT pathway member TCF7L1 (604652). Integrative analyses suggested new markers for aggressive colorectal carcinoma and an important role for MYC-directed transcriptional activation and repression.

By whole-exome sequencing, Segui et al. (2015) identified a heterozygous nonsense mutation in the FAN1 gene (613534) in a Spanish family (family 1) that met the Amsterdam criteria for risk of hereditary nonpolyposis CRC. The mutation was present in 2 affected brothers and an affected son, as well as that son's unaffected 47-year-old sister and his unaffected 21-year-old son. The variant was not found in the ESP or 1000 Genomes Project databases or in 1,648 alleles of Spanish origin, including 286 sporadic CRC patients. Analysis of FAN1 in another 176 mismatch repair (MMR)-proficient Amsterdam-positive families revealed affected individuals from 4 families (families 2 to 5) with a FAN1 mutation, including 1 nonsense and 3 missense variants. Only 2 unaffected family members were tested, and 1 was a mutation carrier. In vitro analysis of 1 of the missense variants suggested that it causes a DNA interstrand crosslink repair defect. Whole-exome sequencing of the tumor from the proband in family 1 showed a somatic mutation burden corresponding to that of nonhypermutant CRCs, and no clear evidence of somatic FAN1 second hits was obtained. The authors concluded that FAN1 is implicated in the inherited susceptibility to CRC.


REFERENCES

  1. Alhopuro, P., Phichith, D., Tuupanen, S., Sammalkorpi, H., Nybondas, M., Saharinen, J., Robinson, J. P., Yang, Z., Chen, L.-Q., Orntoft, T., Mecklin, J.-P., Jarvinen, H., and 12 others. Unregulated smooth-muscle myosin in human intestinal neoplasia. Proc. Nat. Acad. Sci. 105: 5513-5518, 2008. [PubMed: 18391202, images, related citations] [Full Text]

  2. Bamezai, R., Singh, G., Khanna, N. N., Singh, S. Genetics of site specific colon cancer: a family study. Clin. Genet. 26: 129-132, 1984. [PubMed: 6088133, related citations] [Full Text]

  3. Bardelli, A., Parsons, D. W., Silliman, N., Ptak, J., Szabo, S., Saha, S., Markowitz, S., Willson, J. K. V., Parmigiani, G., Kinzler, K. W., Vogelstein, B., Velculescu, V. E. Mutational analysis of the tyrosine kinome in colorectal cancers. Science 300: 949 only, 2003. [PubMed: 12738854, related citations] [Full Text]

  4. Baron, J. A., Beach, M., Mandel, J. S., van Stolk, R. U., Haile, R. W., Sandler, R. S., Rothstein, R., Summers, R. W., Snover, D. C., Beck, G. J., Bond, J. H., Greenberg, E. R. Calcium supplements for the prevention of colorectal adenomas. New Eng. J. Med. 340: 101-107, 1999. [PubMed: 9887161, related citations] [Full Text]

  5. Baron, J. A., Cole, B. F., Sandler, R. S., Haile, R. W., Ahnen, D., Bresalier, R., McKeown-Eyssen, G., Summers, R. W., Rothstein, R., Burke, C. A., Snover, D. C., Church, T. R., and 12 others. A randomized trial of aspirin to prevent colorectal adenomas. New Eng. J. Med. 348: 891-899, 2003. [PubMed: 12621133, related citations] [Full Text]

  6. Bass, A. J., Lawrence, M. S., Brace, L. E., Ramos, A. H., Drier, Y., Cibulskis, K., Sougnez, C., Voet, D., Saksena, G., Sivachenko, A., Jing, R., Parkin, M., and 28 others. Genomic sequencing of colorectal adenocarcinomas identifies a recurrent VTI1A-TCF7L2 fusion. Nature Genet. 43: 964-968, 2011. [PubMed: 21892161, images, related citations] [Full Text]

  7. Batlle, E., Bacani, J., Begthel, H., Jonkeer, S., Gregorieff, A., van de Born, M., Malats, N., Sancho, E., Boon, E., Pawson, T., Gallinger, S., Pals, S., Clevers, H. EphB receptor activity suppresses colorectal cancer progression. Nature 435: 1126-1130, 2005. Note: Erratum: Nature 436: 881 only, 2005. [PubMed: 15973414, related citations] [Full Text]

  8. Boraska Jelavic, T., Barisic, M., Drmic Hofman, I., Boraska, V., Vrdoljak, E., Peruzovic, M., Hozo, I., Puljiz, Z., Terjic, J. Microsatelite (sic) GT polymorphism in the toll-like receptor 2 is associated with colorectal cancer. Clin. Genet. 70: 156-160, 2006. [PubMed: 16879199, related citations] [Full Text]

  9. Bos, J. L., Fearon, E. R., Hamilton, S. R., Verlaan-de Vries, M., van Boom, J. H., van der Eb, A. J., Vogelstein, B. Prevalence of ras gene mutations in human colorectal cancers. Nature 327: 293-297, 1987. [PubMed: 3587348, related citations] [Full Text]

  10. Budd, D. C., Fink, D. L. Mucoid colonic carcinoma as an autosomal-dominant inherited syndrome. Arch. Surg. 116: 901-905, 1981. [PubMed: 7259491, related citations] [Full Text]

  11. Burt, R. W., Bishop, D. T., Cannon, L. A., Dowdle, M. A., Lee, R. G., Skolnick, M. H. Dominant inheritance of adenomatous colonic polyps and colorectal cancer. New Eng. J. Med. 312: 1540-1544, 1985. [PubMed: 4000184, related citations] [Full Text]

  12. Cancer Genome Atlas Network. Comprehensive molecular characterization of human colon and rectal cancer. Nature 487: 330-337, 2012. [PubMed: 22810696, images, related citations] [Full Text]

  13. Cannon-Albright, L. A., Skolnick, M. H., Bishop, T., Lee, R. G., Burt, R. W. Common inheritance of susceptibility to colonic adenomatous polyps and associated colorectal cancers. New Eng. J. Med. 319: 533-537, 1988. [PubMed: 2841598, related citations] [Full Text]

  14. Cui, H., Cruz-Correa, M., Giardiello, F. M., Hutcheon, D. F., Kafonek, D. R., Brandenburg, S., Wu, Y., He, X., Powe, N. R., Feinberg, A. P. Loss of IGF2 imprinting: a potential marker of colorectal cancer risk. Science 299: 1753-1755, 2003. [PubMed: 12637750, related citations] [Full Text]

  15. Daley, D., Lewis, S., Platzer, P., MacMillen, M., Willis, J., Elston, R. C., Markowitz, S. D., Wiesner, G. L. Identification of susceptibility genes for cancer in a genome-wide scan: results from the colon neoplasia sibling study. Am. J. Hum. Genet. 82: 723-736, 2008. [PubMed: 18313025, images, related citations] [Full Text]

  16. Delattre, O., Olschwang, S., Law, D. J., Melot, T., Remvikos, Y., Salmon, R. J., Sastre, X., Validire, P., Feinberg, A. P., Thomas, G. Multiple genetic alterations in distal and proximal colorectal cancer. Lancet 334: 353-356, 1989. Note: Originally Volume II. [PubMed: 2569552, related citations] [Full Text]

  17. Diaz, L. A., Jr., Williams, R. T., Wu, J., Kinde, I., Hecht, J. R., Berlin, J., Allen, B., Bozic, I., Reiter, J. G., Nowak, M. A., Kinzler, K. W., Oliner, K. S., Vogelstein, B. The molecular evolution of acquired resistance to targeted EGFR blockade in colorectal cancers. Nature 486: 537-540, 2012. [PubMed: 22722843, images, related citations] [Full Text]

  18. Dorard, C., de Thonel, A., Collura, A., Marisa, L., Svrcek, M., Lagrange, A., Jego, G., Wanherdrick, K., Joly, A. L., Buhard, O., Gobbo, J., Penard-Lacronique, V., and 16 others. Expression of a mutant HSP110 sensitizes colorectal cancer cells to chemotherapy and improves disease prognosis. Nature Med. 17: 1283-1289, 2011. [PubMed: 21946539, related citations] [Full Text]

  19. Douillard, J.-Y., Oliner, K. S., Siena, S., Tabernero, J., Burkes, R., Barugel, M., Humblet, Y., Bodoky, G., Cunningham, D., Jassem, J., Rivera, F., Kocakova, I., and 10 others. Panitumumab-FOLFOX4 treatment and RAS mutations in colorectal cancer. New Eng. J. Med. 369: 1023-1034, 2013. [PubMed: 24024839, related citations] [Full Text]

  20. Fearnhead, N. S., Wilding, J. L., Winney, B., Tonks, S., Bartlett, S., Bicknell, D. C., Tomlinson, I. P. M., Mortensen, N. J. M., Bodmer, W. F. Multiple rare variants in different genes account for multifactorial inherited susceptibility to colorectal adenomas. Proc. Nat. Acad. Sci. 101: 15992-15997, 2004. [PubMed: 15520370, related citations] [Full Text]

  21. Fearon, E. R., Hamilton, S. R., Vogelstein, B. Clonal analysis of human colorectal tumors. Science 238: 193-197, 1987. [PubMed: 2889267, related citations] [Full Text]

  22. Fearon, E. R., Vogelstein, B. A genetic model for colorectal tumorigenesis. Cell 61: 759-767, 1990. [PubMed: 2188735, related citations] [Full Text]

  23. Forrest, W. F., Cavet, G. Comment on 'The consensus coding sequences of human breast and colorectal cancers.' (Abstract) Science 317: 1500 only, 2007.

  24. Getz, G., Hofling, H., Mesirov, J. P., Golub, T. R., Meyerson, M., Tibshirani, R., Lander, E. S. Comment on 'The consensus coding sequences of human breast and colorectal cancers.' (Abstract) Science 317: 1500 only, 2007.

  25. Greaves, L. C., Preston, S. L., Tadrous, P. J., Taylor, R. W., Barron, M. J., Oukrif, D., Leedham, S. J., Deheragoda, M., Sasieni, P., Novelli, M. R., Jankowski, J. A. Z., Turnbull, D. M., Wright, N. A., McDonald, S. A. C. Mitochondrial DNA mutations are established in human colonic stem cells, and mutated clones expand by crypt fission. Proc. Nat. Acad. Sci. 103: 714-719, 2006. [PubMed: 16407113, images, related citations] [Full Text]

  26. Grivennikov, S. I., Wang, K., Mucida, D., Stewart, C. A., Schnabl, B., Jauch, D., Taniguchi, K., Yu, G.-Y., Osterreicher, C. H., Hung, K. E., Datz, C., Feng, Y., and 9 others. Adenoma-linked barrier defects and microbial products drive IL-23/IL-17-mediated tumour growth. Nature 491: 254-258, 2012. [PubMed: 23034650, images, related citations] [Full Text]

  27. Gruber, S. B., Ellis, N. A., Scott, K. K., Almog, R., Kolachana, P., Bonner, J. D., Kirchhoff, T., Tomsho, L. P., Nafa, K., Pierce, H., Low, M., Satagopan, J., and 12 others. BLM heterozygosity and the risk of colorectal cancer. Science 297: 2013 only, 2002. Note: Erratum: Science 298: 751 only, 2002. [PubMed: 12242432, related citations] [Full Text]

  28. Houlston, R. S., Collins, A., Slack, J., Morton, N. E. Dominant genes for colorectal cancer are not rare. Ann. Hum. Genet. 56: 99-103, 1992. [PubMed: 1503398, related citations] [Full Text]

  29. Huber, S., Gagliani, N., Zenewicz, L. A., Huber, F. J., Bosurgi, L., Hu, B., Hedl, M., Zhang, W., O'Connor, W., Jr., Murphy, A. J., Valenzuela, D. M., Yancopoulos, G. D., Booth, C. J., Cho, J. H., Ouyang, W., Abraham, C., Flavell, R. A. IL-22BP is regulated by the inflammasome and modulates tumorigenesis in the intestine. Nature 491: 259-263, 2012. [PubMed: 23075849, images, related citations] [Full Text]

  30. Jones, P. A., Gonzalgo, M. L. Altered DNA methylation and genome instability: a new pathway to cancer? Proc. Nat. Acad. Sci. 94: 2103-2105, 1997. [PubMed: 9122155, related citations] [Full Text]

  31. Kikuchi-Yanoshita, R., Konishi, M., Ito, S., Seki, M., Tanaka, K., Maeda, Y., Iino, H., Fukayama, M., Koike, M., Mori, T., Sakuraba, H., Fukunari, H., Iwama, T., Miyaki, M. Genetic changes of both p53 alleles associated with the conversion from colorectal adenoma to early carcinoma in familial adenomatous polyposis and non-familial adenomatous polyposis patients. Cancer Res. 52: 3965-3971, 1992. [PubMed: 1319835, related citations]

  32. Kinzler, K. W., Vogelstein, B. Lessons from hereditary colorectal cancer. Cell 87: 159-170, 1996. [PubMed: 8861899, related citations] [Full Text]

  33. Kluge, T. Familial cancer of the colon. Acta Chir. Scand. 127: 392-398, 1964. [PubMed: 14149641, related citations]

  34. Konstantinova, L. N., Fleischman, E. W., Knisch, V. I., Perevozchikov, A. G., Kopnin, B. P. Karyotype pecularities (sic) of human colorectal adenocarcinomas. Hum. Genet. 86: 491-496, 1991. [PubMed: 2016090, related citations] [Full Text]

  35. Laird, P. W., Jackson-Grusby, L., Fazeli, A., Dickinson, S. L., Jung, W. E., Li, E., Weinberg, R. A., Jaenisch, R. Suppression of intestinal neoplasia by DNA hypomethylation. Cell 81: 197-205, 1995. [PubMed: 7537636, related citations] [Full Text]

  36. Law, D. J., Olschwang, S., Monpezat, J. P., Lefrancois, D., Jagelman, D., Petrelli, N. J., Thomas, G., Feinberg, A. P. Concerted nonsyntenic allelic loss in human colorectal carcinoma. Science 241: 961-965, 1988. [PubMed: 2841761, related citations] [Full Text]

  37. Lengauer, C., Kinzler, K. W., Vogelstein, B. DNA methylation and genetic instability in colorectal cancer cells. Proc. Nat. Acad. Sci. 94: 2545-2550, 1997. [PubMed: 9122232, images, related citations] [Full Text]

  38. Lengauer, C., Kinzler, K. W., Vogelstein, B. Genetic instability in colorectal cancers. Nature 386: 623-627, 1997. [PubMed: 9121588, related citations] [Full Text]

  39. Liu, Y., Zhang, X., Han, C., Wan, G., Huang, X., Ivan, C., Jiang, D., Rodriguez-Aguayo, C., Lopez-Berestein, G., Rao, P. H., Maru, D. M., Pahl, A., He, X., Sood, A. K., Ellis, L. M., Anderl, J., Lu, X. TP53 loss creates therapeutic vulnerability in colorectal cancer. Nature 520: 697-701, 2015. Note: Erratum: Nature 597: E6, 2021. [PubMed: 25901683, images, related citations] [Full Text]

  40. Lovett, E. Familial cancer of the gastro-intestinal tract. Brit. J. Surg. 63: 19-22, 1976. [PubMed: 1267869, related citations] [Full Text]

  41. Lovett, E. Family studies in cancer of the colon and rectum. Brit. J. Surg. 63: 13-18, 1976. [PubMed: 1267868, related citations] [Full Text]

  42. Lynch, H. T., de la Chapelle, A. Hereditary colorectal cancer. New Eng. J. Med. 348: 919-932, 2003. [PubMed: 12621137, related citations] [Full Text]

  43. Mathis, V. M. Familiaeres Colon Karzinom. Ein Stammbaum aus dem Kanton Aargau. Schweiz. Med. Wschr. 92: 1673-1678, 1962. [PubMed: 13933637, related citations]

  44. McMurray, H. R., Sampson, E. R., Compitello, G., Kinsey, C., Newman, L., Smith, B., Chen, S.-R., Klebanov, L., Salzman, P., Yakovlev, A., Land, H. Synergistic response to oncogenic mutations defines gene class critical to cancer phenotype. Nature 453: 1112-1116, 2008. [PubMed: 18500333, images, related citations] [Full Text]

  45. Mecklin, J.-P. Frequency of hereditary colorectal carcinoma. Gastroenterology 93: 1021-1025, 1987. [PubMed: 2820826, related citations] [Full Text]

  46. Misale, S., Yaeger, R., Hobor, S., Scala, E., Janakiraman, M., Liska, D., Valtorta, E., Schiavo, R., Buscarino, M., Siravegna, G., Bencardino, K., Cercek, A., and 14 others. Emergence of KRAS mutations and acquired resistance to anti-EGFR therapy in colorectal cancer. Nature 486: 532-536, 2012. [PubMed: 22722830, images, related citations] [Full Text]

  47. Moldenhauer, G., Salnikov, A. V., Luttgau, S., Herr, I., Anderl, J., Faulstich, H. Therapeutic potential of amanitin-conjugated anti-epithelial cell adhesion molecule monoclonal antibody against pancreatic carcinoma. J. Nat. Cancer Inst. 104: 622-634, 2012. [PubMed: 22457476, related citations] [Full Text]

  48. Montagut, C., Dalmases, A., Bellosillo, B., Crespo, M., Pairet, S., Iglesias, M., Salido, M., Gallen, M., Marsters, S., Tsai, S. P., Minoche, A., Seshagiri, S., Serrano, S., Himmelbauer, H., Bellmunt, J., Rovira, A., Settleman, J., Bosch, F., Albanell, J. Identification of a mutation in the extracellular domain of the epidermal growth factor receptor conferring cetuximab resistance in colorectal cancer. Nature Med. 18: 221-223, 2012. Note: Erratum: Nature Med. 18: 1445 only, 2012. [PubMed: 22270724, related citations] [Full Text]

  49. Neklason, D. W., Tuohy, T. M., Stevens, J., Otterud, B., Baird, L., Kerber, R. A., Samowitz, W. S., Kuwada, S. K., Leppert, M. F., Burt, R. W. Colorectal adenomas and cancer link to chromosome 13q22.1-13q31.3 in a large family with excess colorectal cancer. J. Med. Genet. 47: 692-699, 2010. [PubMed: 20522424, images, related citations] [Full Text]

  50. Okamoto, M., Sasaki, M., Sugio, K., Sato, C., Iwama, T., Ikeuchi, T., Tonomura, A., Sasazuki, T., Miyaki, M. Loss of constitutional heterozygosity in colon carcinoma from patients with familial polyposis coli. Nature 331: 273-277, 1988. [PubMed: 2827040, related citations] [Full Text]

  51. Parmigiani, G., Lin, J., Boca, S. M., Sjoblom, T., Jones, S., Wood, L. D., Parsons, D. W., Barber, T., Buckhaults, P., Markowitz, S. D., Park, B. H., Bachman, K. E., Papadopoulos, N., Vogelstein, B., Kinzler, K. W., Velculescu, V. E. Response to comments on 'The consensus coding sequences of human breast and colorectal cancers.' (Abstract) Science 317: 1500 only, 2007.

  52. Parsons, D. W., Wang, T.-L., Samuels, Y., Bardelli, A., Cummins, J. M., DeLong, L., Silliman, N., Ptak, J., Szabo, S., Willson, J. K. V., Markowitz, S., Kinzler, K. W., Vogelstein, B., Lengauer, C., Velculescu, V. E. Mutations in a signalling pathway. (Letter) Nature 436: 792 only, 2005. [PubMed: 16094359, related citations] [Full Text]

  53. Pathak, S., Goodacre, A. Specific chromosome anomalies and predisposition to human breast, renal cell, and colorectal carcinoma. Cancer Genet. Cytogenet. 19: 29-36, 1986. [PubMed: 3940177, related citations] [Full Text]

  54. Perucho, M., Goldfarb, M., Shimizu, K., Lama, C., Fogh, J., Wigler, M. Human-tumor-derived cell lines contain common and different transforming genes. Cell 27: 467-476, 1981. [PubMed: 6101201, related citations] [Full Text]

  55. Ponz de Leon, M., Scapoli, C., Zanghieri, G., Sassatelli, R., Sacchetti, C., Barrai, I. Genetic transmission of colorectal cancer: exploratory data analysis from a population based registry. J. Med. Genet. 29: 531-538, 1992. [PubMed: 1518021, related citations] [Full Text]

  56. Prahallad, A., Sun, C., Huang, S., Di Nicolantonio, F., Salazar, R., Zecchin, D., Beijersbergen, R. L., Bardelli, A., Bernards, R. Unresponsiveness of colon cancer to BRAF(V600E) inhibition through feedback activation of EGFR. Nature 483: 100-103, 2012. [PubMed: 22281684, related citations] [Full Text]

  57. Rajagopalan, H., Bardelli, A., Lengauer, C., Kinzler, K. W., Vogelstein, B., Velculescu, V. E. RAF/RAS oncogenes and mismatch-repair status. (Letter) Nature 418: 934 only, 2002. [PubMed: 12198537, related citations] [Full Text]

  58. Rubin, A. F., Green, P. Comment on 'The consensus coding sequences of human breast and colorectal cancers.' (Abstract) Science 317: 1500 only, 2007.

  59. Samuels, Y., Wang, Z., Bardelli, A., Silliman, N., Ptak, J., Szabo, S., Yan, H., Gazdar, A., Powell, S. M., Riggins, G. J., Willson, J. K. V., Markowitz, S., Kinzler, K. W., Vogelstein, B., Velculescu, V. E. High frequency of mutations of the PIK3CA gene in human cancers. Science 304: 554 only, 2004. [PubMed: 15016963, related citations] [Full Text]

  60. Sandler, R. S., Halabi, S., Baron, J. A., Budinger, S., Paskett, E., Keresztes, R., Petrelli, N., Pipas, J. M., Karp, D. D., Loprinzi, C. L., Steinbach, G., Schilsky, R. A randomized trial of aspirin to prevent colorectal adenomas in patients with previous colorectal cancer. New Eng. J. Med. 348: 883-890, 2003. Note: Erratum: New Eng. J. Med. 348: 1939 only, 2003. [PubMed: 12621132, related citations] [Full Text]

  61. Sarraf, P., Mueller, E., Smith, W. M., Wright, H. M., Kum, J. B., Aaltonen, L. A., de la Chapelle, A., Spiegelman, B. M., Eng, C. Loss-of-function mutations in PPAR-gamma associated with human colon cancer. Molec. Cell 3: 799-804, 1999. [PubMed: 10394368, related citations] [Full Text]

  62. Schweiger, M.-R., Hussong, M., Rohr, C., Lehrach, H. Genomics and epigenomics of colorectal cancer. Wiley Interdiscip. Rev. Syst. Biol. Med. 5: 205-219, 2013. [PubMed: 23325509, related citations] [Full Text]

  63. Segui, N., Mina, L. B., Lazaro, C., Sanz-Pamplona, R., Pons, T., Navarro, M., Bellido, F., Lopez-Doriga, A., Valdes-Mas, R., Pineda, M., Guino, E., Vidal, A., and 21 others. Germline mutations in FAN1 cause hereditary colorectal cancer by impairing DNA repair. Gastroenterology 149: 563-566, 2015. [PubMed: 26052075, related citations] [Full Text]

  64. Seshagiri, S., Stawiski, E. W., Durinck, S., Modrusan, Z., Storm, E. E., Conboy, C. B., Chaudhuri, S., Guan, Y., Janakiraman, V., Jaiswal, B. S., Guillory, J., Ha, C., and 15 others. Recurrent R-spondin fusions in colon cancer. Nature 488: 660-664, 2012. [PubMed: 22895193, images, related citations] [Full Text]

  65. Sivak, M. V., Jr., Sivak, D. S., Braun, W. A., Sullivan, B. H., Jr. A linkage study of HLA and inherited adenocarcinoma of the colon. Cancer 48: 76-81, 1981. [PubMed: 7237394, related citations] [Full Text]

  66. Sjoblom, T., Jones, S., Wood, L. D., Parsons, D. W., Lin, J., Barber, T. D., Mandelker, D., Leary, R. J., Ptak, J., Silliman, N., Szabo, S., Buckhaults, P., and 17 others. The consensus coding sequences of human breast and colorectal cancers. Science 314: 268-274, 2006. [PubMed: 16959974, related citations] [Full Text]

  67. Starr, T. K., Allaei, R., Silverstein, K. A. T., Staggs, R. A., Sarver, A. L., Bergemann, T. L., Gupta, M., O'Sullivan, M. G., Matise, I., Dupuy, A. J., Collier, L. S., Powers, S., Oberg, A. L., Asmann, Y. W., Thibodeau, S. N., Tessarollo, L., Copeland, N. G., Jenkins, N. A., Cormier, R. T., Largaespada, D. A. A transposon-based genetic screen in mice identifies genes altered in colorectal cancer. Science 323: 1747-1750, 2009. [PubMed: 19251594, related citations] [Full Text]

  68. Thomas, J. H. Genomic imprinting proposed as a surveillance mechanism for chromosome loss. Proc. Nat. Acad. Sci. 92: 480-482, 1995. [PubMed: 7831314, related citations] [Full Text]

  69. Vermeulen, L., Morrissey, E., van der Heijden, M., Nicholson, A. M., Sottoriva, A., Buczacki, S., Kemp, R., Tavare, S., Winton, D. J. Defining stem cell dynamics in models of intestinal tumor initiation. Science 342: 995-998, 2013. [PubMed: 24264992, related citations] [Full Text]

  70. Vilar, E., Gruber, S. B. Microsatellite instability in colorectal cancer--the stable evidence. Nature Rev. Clin. Oncol. 7: 153-162, 2010. [PubMed: 20142816, images, related citations] [Full Text]

  71. Vogelstein, B., Fearon, E. R., Hamilton, S. R., Kern, S. E., Preisinger, A. C., Leppert, M., Nakamura, Y., White, R., Smits, A. M. M., Bos, J. L. Genetic alterations during colorectal-tumor development. New Eng. J. Med. 319: 525-532, 1988. [PubMed: 2841597, related citations] [Full Text]

  72. Vogelstein, B., Fearon, E. R., Kern, S. E., Hamilton, S. R., Preisinger, A. C., Nakamura, Y., White, R. Allelotype of colorectal carcinomas. Science 244: 207-211, 1989. [PubMed: 2565047, related citations] [Full Text]

  73. Ward, C. J., Wu, Y., Johnson, R. A., Woollard, J. R., Bergstralh, E. J., Cicek, M. S., Bakeberg, J., Rossetti, S., Heyer, C. M., Petersen, G. M., Lindor, N. M., Thibodeau, S. N., Harris, P. C., Torres, V. E., Hogan, M. C., Boardman, L. A. Germline PKHD1 mutations are protective against colorectal cancer. Hum. Genet. 129: 345-349, 2011. [PubMed: 21274727, related citations] [Full Text]

  74. Wheeler, J. M. D., Warren, B. F., Mortensen, N. J. M., Kim, H. C., Biddolph, S. C., Elia, G., Beck, N. E., Williams, G. T., Shepherd, N. A., Bateman, A. C., Bodmer, W. F. An insight into the genetic pathway of adenocarcinoma of the small intestine. Gut 50: 218-223, 2002. [PubMed: 11788563, images, related citations] [Full Text]

  75. Wildrick, D. M., Boman, B. M. Chromosome 5 allele loss at the glucocorticoid receptor locus in human colorectal carcinomas. Biochem. Biophys. Res. Commun. 150: 591-598, 1988. [PubMed: 3342039, related citations] [Full Text]

  76. Williams, C. Management of malignancy in 'cancer families'. Lancet 311: 198-199, 1978. Note: Originally Volume I. [PubMed: 74618, related citations] [Full Text]

  77. Wood, L. D., Parsons, D. W., Jones, S., Lin, J., Sjoblom, T., Leary, R. J., Shen, D., Boca, S. M., Barber, T., Ptak, J., Silliman, N., Szabo, S., and 30 others. The genomic landscapes of human breast and colorectal cancers. Science 318: 1108-1113, 2007. [PubMed: 17932254, related citations] [Full Text]

  78. Xia, D., Wang, D., Kim, S.-H., Katoh, H., DuBois, R. N. Prostaglandin E2 promotes intestinal tumor growth via DNA methylation. Nature Med. 18: 224-226, 2012. [PubMed: 22270723, images, related citations] [Full Text]


Marla J. F. O'Neill - updated : 07/14/2021
Marla J. F. O'Neill - updated : 09/02/2016
Ada Hamosh - updated : 6/25/2015
Ada Hamosh - updated : 12/6/2013
Ada Hamosh - updated : 10/23/2013
Cassandra L. Kniffin - updated : 2/18/2013
Ada Hamosh - updated : 12/4/2012
Ada Hamosh - updated : 9/18/2012
Ada Hamosh - updated : 9/5/2012
Ada Hamosh - updated : 8/10/2012
Ada Hamosh - updated : 7/17/2012
Ada Hamosh - updated : 6/26/2012
Ada Hamosh - updated : 3/15/2012
Ada Hamosh - updated : 3/14/2012
Ada Hamosh - updated : 12/12/2011
Cassandra L. Kniffin - updated : 4/20/2011
Marla J. F. O'Neill - updated : 12/1/2010
Carol A. Bocchini - updated : 11/4/2010
Marla J. F. O'Neill - updated : 10/5/2009
Ada Hamosh - updated : 9/14/2009
Ada Hamosh - updated : 6/18/2009
Ada Hamosh - updated : 7/29/2008
Ada Hamosh - updated : 7/18/2008
Cassandra L. Kniffin - updated : 4/28/2008
Ada Hamosh - updated : 2/14/2008
Ada Hamosh - updated : 1/9/2008
Victor A. McKusick - updated : 11/20/2007
Ada Hamosh - updated : 10/31/2006
Marla J. F. O'Neill - updated : 9/22/2006
Ada Hamosh - updated : 9/8/2005
Ada Hamosh - updated : 7/27/2005
Victor A. McKusick - updated : 4/15/2005
Ada Hamosh - updated : 4/30/2004
Ada Hamosh - updated : 5/29/2003
Ada Hamosh - updated : 4/3/2003
Victor A. McKusick - updated : 3/14/2003
Ada Hamosh - updated : 9/30/2002
George E. Tiller - updated : 9/26/2002
Ada Hamosh - updated : 9/17/2002
Paul Brennan - updated : 3/19/2002
Paul Brennan - updated : 3/13/2002
Paul Brennan - updated : 3/6/2002
George E. Tiller - updated : 6/19/2001
Stylianos E. Antonarakis - updated : 7/20/1999
Victor A. McKusick - updated : 2/9/1999
Victor A. McKusick - updated : 4/21/1997
Creation Date:
Victor A. McKusick : 6/4/1986
carol : 10/10/2023
carol : 10/09/2023
carol : 11/05/2021
carol : 07/15/2021
alopez : 07/14/2021
carol : 04/22/2020
carol : 09/13/2019
carol : 09/02/2016
carol : 05/04/2016
alopez : 11/24/2015
alopez : 7/7/2015
alopez : 6/25/2015
alopez : 12/6/2013
alopez : 10/23/2013
tpirozzi : 10/1/2013
carol : 3/11/2013
carol : 2/19/2013
ckniffin : 2/18/2013
alopez : 12/6/2012
terry : 12/4/2012
alopez : 9/19/2012
terry : 9/18/2012
alopez : 9/5/2012
carol : 8/10/2012
terry : 8/10/2012
terry : 7/27/2012
alopez : 7/19/2012
terry : 7/17/2012
alopez : 6/26/2012
terry : 6/26/2012
alopez : 3/15/2012
alopez : 3/14/2012
alopez : 12/19/2011
terry : 12/12/2011
carol : 9/7/2011
wwang : 5/2/2011
ckniffin : 4/20/2011
ckniffin : 4/20/2011
carol : 4/20/2011
wwang : 12/3/2010
carol : 12/2/2010
terry : 12/1/2010
carol : 11/4/2010
carol : 3/19/2010
alopez : 2/4/2010
ckniffin : 1/15/2010
wwang : 10/14/2009
terry : 10/5/2009
wwang : 9/29/2009
alopez : 9/14/2009
wwang : 7/29/2009
alopez : 6/24/2009
terry : 6/18/2009
wwang : 2/13/2009
ckniffin : 2/9/2009
terry : 1/12/2009
terry : 1/9/2009
carol : 9/19/2008
alopez : 8/18/2008
terry : 7/29/2008
wwang : 7/18/2008
wwang : 6/9/2008
ckniffin : 4/28/2008
carol : 2/15/2008
alopez : 2/15/2008
terry : 2/14/2008
ckniffin : 2/5/2008
carol : 1/31/2008
ckniffin : 1/28/2008
ckniffin : 1/28/2008
alopez : 1/28/2008
terry : 1/9/2008
alopez : 12/7/2007
terry : 11/20/2007
alopez : 9/27/2007
alopez : 8/31/2007
alopez : 11/3/2006
alopez : 11/3/2006
terry : 10/31/2006
alopez : 10/9/2006
wwang : 9/22/2006
wwang : 5/17/2006
carol : 4/14/2006
alopez : 12/5/2005
alopez : 9/9/2005
terry : 9/8/2005
alopez : 7/28/2005
terry : 7/27/2005
carol : 6/3/2005
mgross : 4/15/2005
mgross : 4/14/2005
mgross : 4/14/2005
mgross : 4/13/2005
tkritzer : 2/11/2005
alopez : 4/30/2004
terry : 4/30/2004
carol : 7/10/2003
mgross : 5/29/2003
joanna : 5/29/2003
terry : 5/29/2003
terry : 4/3/2003
carol : 3/21/2003
tkritzer : 3/18/2003
terry : 3/14/2003
alopez : 9/30/2002
alopez : 9/30/2002
tkritzer : 9/30/2002
cwells : 9/26/2002
alopez : 9/17/2002
alopez : 3/19/2002
alopez : 3/13/2002
alopez : 3/6/2002
cwells : 6/20/2001
cwells : 6/19/2001
carol : 10/20/2000
carol : 8/12/1999
mgross : 7/20/1999
mgross : 2/16/1999
mgross : 2/15/1999
terry : 2/9/1999
alopez : 6/27/1997
jenny : 4/21/1997
terry : 4/14/1997
terry : 12/10/1996
terry : 12/9/1996
carol : 5/31/1994
terry : 5/13/1994
mimadm : 4/9/1994
warfield : 4/6/1994
carol : 2/24/1993
carol : 10/12/1992

# 114500

COLORECTAL CANCER; CRC


Alternative titles; symbols

COLON CANCER


SNOMEDCT: 363406005;   ICD10CM: C18, C18.9;   ICD9CM: 153, 153.9;   DO: 9256;  


Phenotype-Gene Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
Gene/Locus Gene/Locus
MIM number
1p36.13 {?Colorectal cancer, susceptibility to} 114500 Autosomal dominant; Somatic mutation 3 PLA2G2A 172411
1p13.2 Colorectal cancer, somatic 114500 3 NRAS 164790
2q13 Colorectal cancer with chromosomal instability, somatic 114500 3 BUB1 602452
3p22.1 Colorectal cancer, somatic 114500 3 CTNNB1 116806
3q26.32 Colorectal cancer, somatic 114500 3 PIK3CA 171834
4p16.3 Colorectal cancer, somatic 114500 3 FGFR3 134934
4q31.3 {Colorectal cancer, susceptibility to} 114500 Autosomal dominant; Somatic mutation 3 TLR2 603028
5q22.2 Colorectal cancer, somatic 114500 3 APC 611731
5q22.2 Colorectal cancer, somatic 114500 3 MCC 159350
7q11.23 Colon cancer, somatic 114500 3 PTPN12 600079
7q34 Colorectal cancer, somatic 114500 3 BRAF 164757
8p22 Colorectal cancer, somatic 114500 3 DLC1 604258
8p22 Colorectal cancer, somatic 114500 3 PDGFRL 604584
8q22.1 Colon cancer, somatic 114500 3 RAD54B 604289
11p11.2 Colon cancer, somatic 114500 3 PTPRJ 600925
11q13.3 {Colorectal cancer, susceptibility to} 114500 Autosomal dominant; Somatic mutation 3 CCND1 168461
14q24.3 Colorectal cancer, somatic 114500 3 MLH3 604395
14q32.33 Colorectal cancer, somatic 114500 3 AKT1 164730
15q15.1 Colorectal cancer, somatic 114500 3 BUB1B 602860
17p13.1 {Colorectal cancer} 114500 Autosomal dominant; Somatic mutation 3 TP53 191170
17p11.2 Colorectal cancer, somatic 114500 3 FLCN 607273
17q24.1 Colorectal cancer, somatic 114500 3 AXIN2 604025
18q21.2 Colorectal cancer, somatic 114500 3 DCC 120470
19q13.33 Colorectal cancer, somatic 114500 3 BAX 600040
20q11.23 Colon cancer, advanced, somatic 114500 3 SRC 190090
20q13.2 {Colon cancer, susceptibility to} 114500 Autosomal dominant; Somatic mutation 3 AURKA 603072
22q13.2 Colorectal cancer, somatic 114500 3 EP300 602700

TEXT

A number sign (#) is used with this entry because mutations in several different genes have been identified in colorectal cancer (CRC).


Description

Colorectal cancer is a heterogeneous disease that is common in both men and women. In addition to lifestyle and environmental risk factors, gene defects can contribute to an inherited predisposition to CRC. CRC is caused by changes in different molecular pathogenic pathways, such as chromosomal instability, CpG island methylator phenotype, and microsatellite instability. Chromosome instability is the most common alteration and is present in almost 85% of all cases (review by Schweiger et al., 2013).

Genetic Heterogeneity of Colorectal Cancer

Mutations in a single gene result in a marked predisposition to colorectal cancer in familial adenomatous polyposis (FAP; 175100) and hereditary nonpolyposis colorectal cancer (HNPCC; see 120435). FAP is caused by mutations in the APC gene (611731), whereas HNPCC is caused by mutations in several genes, including MSH2 (609309), MLH1 (120436), PMS1 (600258), PMS2 (600259), MSH6 (600678), TGFBR2 (190182), and MLH3 (604395). Epigenetic silencing of MSH2 results in a form of HNPCC (see HNPCC8, 613244). Other colorectal cancer syndromes include autosomal recessive adenomatous polyposis (608456), which is caused by mutations in the MUTYH gene (604933), and oligodontia-colorectal cancer syndrome (608615), which is caused by mutations in the AXIN2 gene (604025). Colorectal cancer is a predominant form of cancer in tumor predisposition syndrome-4 (TPDS4; 609265), which is caused by mutation in the CHEK2 gene (604373). A germline mutation in the PLA2G2A gene (172411) was identified in a patient with colorectal cancer.

Germline susceptibility loci for colorectal cancer have also been identified. CRCS1 (608812) is conferred by mutation in the GALNT12 gene (610290) on chromosome 9q22; CRCS2 (611469) maps to chromosome 8q24; CRCS3 (612229) is conferred by variation in the SMAD7 gene (602932) on chromosome 18; CRCS4 (601228) is conferred by variation on 15q that causes increased and ectopic expression of the GREM1 gene (603054); CRCS5 (612230) maps to chromosome 10p14; CRCS6 (612231) maps to chromosome 8q23; CRCS7 (612232) maps to chromosome 11q23; CRCS8 (612589) maps to chromosome 14q22; CRCS9 (612590) maps to 16q22; CRCS10 (612591) is conferred by mutation in the POLD1 gene (174761) on chromosome 19q13; CRCS11 (612592) maps to chromosome 20p12; and CRCS12 (615083) is conferred by mutation in the POLE gene (174762) on chromosome 12q24.

Somatic mutations in many different genes, including KRAS (190070), PIK3CA (171834), BRAF (164757), CTNNB1 (116806), FGFR3 (134934), AXIN2 (604025), AKT1 (164730), MCC (159350), MYH11 (160745), PARK2 (602544), RNF43 (612482), and BUB1 (601452), have been identified in colorectal cancer.


Clinical Features

Colon cancer is a well-known feature of familial polyposis coli. Cancer of the colon occurred in 7 members of 4 successive generations of the family reported by Kluge (1964), leading him to suggest a simple genetic basis for colonic cancer independent of polyposis. The combination of colonic and endometrial cancer has been observed in many families (e.g., Williams, 1978).

Sivak et al. (1981) studied a kindred with the familial cancer syndrome in which every confirmed affected member had at least 1 primary carcinoma of the colon. The average age at which cancer appeared was 38 years. Multiple primary neoplasms occurred in 23% of cancer patients.

Budd and Fink (1981) reported a family with a high frequency of mucoid colonic carcinoma. Since endometrial carcinoma, atypical endometrial hyperplasia, uterine leiomyosarcoma, bladder transitional carcinoma, and renal cell carcinoma also occurred in the family, this may be the same disorder as the Lynch cancer family syndrome type II (120435).

Bamezai et al. (1984) reported an Indian Sikh kindred in which 8 persons suffered from cancer of the cecum, not associated with polyposis.

Burt et al. (1985) studied a large Utah kindred called to attention because of occurrence of colorectal cancer in a brother, a sister, and a nephew. No clear inheritance pattern was discernible until systematic screening was undertaken for colonic polyps using flexible proctosigmoidoscopy. One or more adenomatous polyps were found in 41 of 191 family members (21%) and 12 of 132 controls (9%)--p less than 0.005. Pedigree analysis showed best fit with autosomal dominant inheritance. Cannon-Albright et al. (1988) extended the studies with investigations of 33 additional kindreds. The kindreds were selected through either a single person with an adenomatous polyp or a cluster of relatives with colonic cancer. The kindreds all had common colorectal cancers, not the rare inherited condition of familial polyposis coli or nonpolyposis inherited colorectal cancer. Likelihood analysis strongly supported dominant inheritance of a susceptibility to colorectal adenomas and cancers, with a gene frequency of 19%. According to the most likely genetic model, adenomatous polyps and colorectal cancers occur only in genetically susceptible persons; however, the 95% confidence interval for this proportion was 53 to 100%.

Ponz de Leon et al. (1992) analyzed data on 605 families of probands with colorectal cancer in the province of Modena in Italy. Among the 577 presumed nonpolyposis cases, both parents had colorectal cancer in 11, one parent in 130, and neither parent in 436. Segregation was compatible with dominant transmission of susceptibility to cancer.

Mecklin (1987) investigated the frequency of hereditary colorectal cancer among all colorectal cancer patients diagnosed in 1 Finnish county during the 1970s. The cancer family syndrome type of hereditary nonpolyposis colorectal carcinoma emerged as the most common verifiable risk factor, involving between 3.8 and 5.5% of all colorectal cancer patients. The frequencies of familial adenomatosis and ulcerative colitis were 0.2% and 0.6%, respectively. The observed frequency is probably an underestimate. The patients with cancer family syndrome were young, accounting for 29 to 39% of the patients under 50 years of age, and their tumors were located predominantly (65%) in the right hemicolon.


Pathogenesis

The state of DNA methylation appears to play a role in genetic instability in colorectal cancer cells. Lengauer et al. (1997) noted that DNA methylation is essential in prokaryotes, dispensable in lower eukaryotes (such as Saccharomyces cerevisiae) yet present and presumably important in mammals. Many cancers have been shown to have a global hypomethylation of DNA compared with normal tissues. Treatment of cells or animals with 5-azacytidine (5-aza-C), a demethylating agent that irreversibly inactivates methyltransferase (see 156569), is oncogenic in vitro and in vivo. Conversely, other studies showed that hypermethylation of specific sequences found in some tumors can be associated with the inactivation of tumor suppressor gene expression. Mice genetically deficient in methyltransferase are resistant to colorectal tumorigenesis initiated by mutation of the APC (611731) tumor suppressor gene, and treatment of these mice with 5-aza-C enhances the resistance (Laird et al., 1995).

Lengauer et al. (1997) reported a striking difference in the expression of exogenously introduced retroviral genes in various colorectal cancer cell lines. Extinguished expression was associated with DNA methylation and could be reversed by treatment with the demethylating agent 5-aza-C. A striking correlation between genetic instability and methylation capacity suggested that methylation abnormalities may play a role in the chromosome segregation processes in cancer cells. It has been speculated that genetic instability is necessary for a tumor to accumulate the numerous genetic alterations that accompany carcinogenesis. There appeared to exist 2 pathways of genetic instability in colorectal cancer. The first is found in about 15% of tumors and involves point mutations, microdeletions, and microinsertions associated with deficiency of mismatch repair (MMR). The second is found in MMR-proficient cells and involves gains and losses of whole chromosomes. Lengauer et al. (1997) suggested that methylation abnormalities are intrinsically and directly involved in the generation of the second type of instability, thus allowing for the selection of methylation-negative cells during the clonal evolution of tumors. The hypothesis was supported by the observation that demethylation is associated with chromosomal aberrations, including mitotic dysfunction and translocation, and was consistent with the hypothesis relating methylation and aneuploidy put forward by Thomas (1995). Jones and Gonzalgo (1997) commented on altered DNA methylation and genome instability as a new pathway to cancer.

In a second report, Lengauer et al. (1997) showed that tumors without microsatellite instability exhibit a striking defect in chromosome segregation, resulting in gains or losses in excess of 10(-2) per chromosome per generation. This form of chromosomal instability reflected a continuing cellular defect that persisted throughout the lifetime of the tumor cell and was not simply related to chromosome number. While microsatellite instability is a recessive trait, chromosomal instability appeared to be dominant. The data indicated that persistent genetic instability may be critical for the development of all colorectal cancers, and that this instability can arise through 2 distinct pathways.

Adenocarcinoma of the small intestine is rare in the general population, but its histologic features are similar to those of the much more common colorectal adenocarcinoma, and it is seen as part of the HNPCC tumor predisposition spectrum. Wheeler et al. (2002) examined the possible role of mismatch repair defects in the pathogenesis of sporadic small intestinal adenocarcinoma. The replication error status was determined in a total of 21 nonfamilial, nonampullary small intestinal adenocarcinomas: only 1 tumor was scored as replication error-positive. This tumor showed normal immunostaining for MLH1 (see 120436) and MSH2. The authors commented that this result may reflect an epigenetic change in the tumor rather than germline mutation in a mismatch repair gene, and concluded that mismatch repair defects were unlikely to contribute significantly to the genetic pathway leading to sporadic small intestinal adenocarcinoma.

Vilar and Gruber (2010) reviewed the role of microsatellite instability (MSI) in the development of CRC. They stated that approximately 15% of CRCs display MSI owing either to epigenetic silencing of MLH1 or to a germline mutation in one of the mismatch repair genes MLH1, MSH2, MSH6, or PMS2. They noted that MSI tumors have a better prognosis than microsatellite stable CRCs, but that MSI cancers do not necessarily have the same response to the chemotherapeutic strategies used to treat microsatellite stable tumors.

Batlle et al. (2005) showed that although Wnt (see 164820) signaling remains constitutively active, most human colorectal cancers lose expression of EphB (see 600600) at the adenoma-carcinoma transition. They found that loss of EphB expression strongly correlated with degree of malignancy. Furthermore, reduction of EphB activity accelerated tumorigenesis in the colon and rectum of Apc(Min/+) mice (see 611731), and resulted in formation of aggressive adenocarcinomas. Batlle et al. (2005) concluded that loss of EphB expression represents a critical step in colorectal cancer progression.

By microdissection of bifurcating colonic crypts and sequencing of the entire mitochondrial genome in all of the cells, Greaves et al. (2006) demonstrated that stochastic mutations in mtDNA resulting in phenotypic cytochrome c oxidase (COX) deficiency of were identical in both arms of a crypt that was bifurcating. Furthermore, they showed that patches of neighboring crypts deficient in COX also shared identical mitochondrial mutations, and that these patches increased in size with age, indicating that crypt fission is a mechanism by which mutations can spread within the colon.

Xia et al. (2012) showed that prostaglandin E2 (PGE2) silences certain tumor suppressor and DNA repair genes through DNA methylation to promote tumor growth. Their findings uncovered a theretofore unrecognized role for PGE2 in the promotion of tumor progression, and provided a rationale for considering the development of a combination treatment using PTGS2 (600262) inhibitors and demethylating agents for the prevention or treatment of colorectal cancer.

Seshagiri et al. (2012) systematically analyzed more than 70 pairs of primary human colon tumors using next-generation sequencing to characterize their exomes, transcriptomes, and copy number alterations. They identified 36,303 protein-altering somatic changes that included several novel recurrent mutations in the Wnt pathway gene TCF7L2 (602228), chromatin-remodeling genes such as TET2 (612839) and TET3 (613555), and receptor tyrosine kinases including ERBB3 (190151). The analysis for significantly mutated cancer genes identified 23 candidates, including the cell cycle checkpoint kinase ATM (607585). Copy number and RNA-seq data analysis identified amplifications and corresponding overexpression of IGF2 in a subset of colon tumors. Furthermore, using RNA-seq data, Seshagiri et al. (2012) identified multiple fusion transcripts including recurrent gene fusions involving R-spondin family members RSPO2 (610575) and RSPO3 (610574) that together occur in 10% of colon tumors. The RSPO fusions were mutually exclusive with APC (611731) mutations, indicating that they probably have a role in the activation of Wnt signaling and tumorigenesis. Consistent with this, Seshagiri et al. (2012) showed that RSPO fusion proteins were capable of potentiating Wnt signaling.

Grivennikov et al. (2012) investigated mechanisms responsible for tumor-elicited inflammation in a mouse model of colorectal tumorigenesis which, like human colorectal cancer, exhibits upregulation of IL23 (605580) and IL17 (603149). They showed that IL23 signaling promotes tumor growth and progression, and development of tumoral IL17 response. IL23 is mainly produced by tumor-associated myeloid cells that are likely to be activated by microbial products, which penetrate the tumors but not adjacent tissue. Both early and late colorectal neoplasms exhibit defective expression of several barrier proteins. Grivennikov et al. (2012) proposed that barrier deterioration induced by colorectal cancer-initiating genetic lesions results in adenoma invasion by microbial products that trigger tumor-elicited inflammation, which in turn drives tumor growth.

Huber et al. (2012) described the crucial role of IL22BP (606648) in controlling tumorigenesis and epithelial cell proliferation in the colon. IL22BP is highly expressed by dendritic cells in the colon in steady-state conditions. Sensing of intestinal tissue damage via the NLRP3 (606416) or NLRP6 (609650) inflammasomes led to an IL18 (600953)-dependent downregulation of IL22BP, thereby increasing the ratio of IL22 (605330)/IL22BP. IL22, which is induced during intestinal tissue damage, exerted protective properties during the peak of damage, but promoted tumor development if uncontrolled during the recovery phase. Thus, the IL22-IL22BP axis critically regulates intestinal tissue repair and tumorigenesis in the colon.

Vermeulen et al. (2013) quantified the competitive advantage during tumor development of Apc (611731) loss, Kras (190070) activation, and p53 (191170) mutations in the mouse intestine. Their findings indicated that the fate conferred by these mutations is not deterministic, and many mutated stem cells are replaced by wildtype stem cells after biased but still stochastic events. Furthermore, Vermeulen et al. (2013) found that p53 mutations display a condition-dependent advantage, and especially in colitis-affected intestines, clones harboring mutations in this gene were favored. Vermeulen et al. (2013) concluded that their work confirmed the notion that the tissue architecture of the intestine suppresses the accumulation of mutated lineages.

Liu et al. (2015) demonstrated that genomic deletion of TP53 frequently encompasses essential neighboring genes, rendering cancer cells with hemizygous TP53 deletion vulnerable to further suppression of such genes. The authors identified POLR2A (180660) as such a gene that is almost always codeleted with TP53 in human cancers. It encodes the largest and catalytic subunit of the RNA polymerase II complex, which is specifically inhibited by alpha-amanitin. Liu et al. (2015) analyzed the Cancer Genome Atlas (TCGA) and Cancer Cell Line Encyclopedia (CCLE) databases, which revealed that expression levels of POLR2A are tightly correlated with its gene copy numbers in human colorectal cancer. Suppression of POLR2A with alpha-amanitin or siRNAs selectively inhibits the proliferation, survival, and tumorigenic potential of colorectal cancer cells with hemizygous TP53 loss in a p53-independent manner. Clinical applications of alpha-amanitin had been limited owing to its liver toxicity; however, Liu et al. (2015) found that alpha-amanitin-based antibody-drug conjugates (Moldenhauer et al., 2012) are highly effective therapeutic agents with reduced toxicity. Liu et al. (2015) showed that low doses of alpha-amanitin-conjugated anti-epithelial cell adhesion molecule (EpCAM; 185535) antibody led to complete tumor regression in mouse models of human colorectal cancer with hemizygous deletion of POLR2A.


Clinical Management

Various laboratory, clinical, and epidemiologic evidence suggested that calcium may help prevent colorectal adenomas. Baron et al. (1999) conducted a randomized, double-blind trial of the effect of supplementation with calcium carbonate on the recurrence of colorectal adenomas. They found a significant, though moderate, reduction in the risk of recurrent colorectal adenomas in the supplemented group.

In randomized trials of aspirin to determine its efficacy in prevention of colorectal adenomas, Sandler et al. (2003) and Baron et al. (2003) studied patients with either previous colorectal cancer or recent histologically documented adenomas, respectively. Both studies found that aspirin was associated with a significant reduction in the incidence of colorectal adenomas.

Inhibition of the BRAF(V600E) (164757.0001) oncoprotein by the small-molecule drug PLX4032 (vemurafenib) is highly effective in the treatment of melanoma. However, colon cancer patients harboring the same BRAF(V600E) oncogenic lesion have poor prognosis and show only a very limited response to this drug. To investigate the cause of this limited therapeutic effect in BRAF(V600E) mutant colon cancer, Prahallad et al. (2012) performed an RNA interference-based genetic screen in human cells to search for kinases whose knockdown synergizes with BRAF(V600E) inhibition. They reported that blockade of the epidermal growth factor receptor (EGFR; 131550) shows strong synergy with BRAF(V600E) inhibition. Prahallad et al. (2012) found in multiple BRAF(V600E) mutant colon cancers that inhibition of EGFR by the antibody drug cetuximab or the small-molecule drugs gefitinib or erlotinib is strongly synergistic with BRAF(V600E) inhibition, both in vitro and in vivo. Mechanistically, Prahallad et al. (2012) found that BRAF(V600E) inhibition causes a rapid feedback activation of EGFR, which supports continued proliferation in the presence of BRAF(V600E) inhibition. Melanoma cells express low levels of EGFR and are therefore not subject to this feedback activation. Consistent with this, Prahallad et al. (2012) found that ectopic expression of EGFR in melanoma cells is sufficient to cause resistance to PLX4032. Prahallad et al. (2012) concluded that BRAF(V600E) mutant colon cancers (approximately 8 to 10% of all colon cancers) might benefit from combination therapy consisting of BRAF and EGFR inhibitors.

Development of Resistance to Chemotherapeutic Agents

Antibodies against EGFR, cetuximab and panitumumab, are widely used to treat colorectal cancer. Unfortunately, patients eventually develop resistance to these agents. Montagut et al. (2012) described an acquired EGFR ectodomain mutation (S492R) that prevents cetuximab binding and confers resistance to cetuximab. Cells with this mutation, however, retain binding to and are growth inhibited by panitumumab. Two of 10 subjects studied with metastatic colon cancer progression after cetuximab treatment acquired this mutation. One subject with cetuximab resistance harboring the S492R mutation responded to treatment with panitumumab.

Misale et al. (2012) showed that molecular alterations (in most instances point mutations) of KRAS (190070) are causally associated with the onset of acquired resistance to anti-EGFR treatment in colorectal cancers. Expression of mutant KRAS under the control of its endogenous gene promoter was sufficient to confer cetuximab resistance, but resistant cells remained sensitive to combinatorial inhibition of EGFR and mitogen-activated protein kinase kinase (MEK; see 176872). Analysis of metastases from patients who developed resistance to cetuximab or panitumumab showed the emergence of KRAS amplification in one sample and acquisition of secondary KRAS mutations in 60% (6 out of 10) of the cases. KRAS mutant alleles were detectable in the blood of cetuximab-treated patients as early as 10 months before radiographic documentation of disease progression. Misale et al. (2012) concluded that their results identified KRAS mutations as frequent drivers of acquired resistance to cetuximab in colorectal cancers, indicated that the emergence of KRAS mutant clones can be detected noninvasively months before radiographic progression, and suggested early initiation of a MEK inhibitor as a rational strategy for delaying or reversing drug resistance.

Diaz et al. (2012) determined whether mutant KRAS DNA could be detected in the circulation of 28 patients receiving monotherapy with panitumumab, a therapeutic anti-EGFR antibody. They found that 9 out of 24 (38%) patients whose tumors were initially KRAS wildtype developed detectable mutations in KRAS in their sera, 3 of which developed multiple different KRAS mutations. The appearance of these mutations was very consistent, generally occurring between 5 and 6 months following treatment. Mathematical modeling indicated that the mutations were present in expanded subclones before the initiation of panitumumab treatment. Diaz et al. (2012) concluded that the emergence of KRAS mutations is a mediator of acquired resistance to EGFR blockade and that these mutations can be detected in a noninvasive manner. The results also explained why solid tumors develop resistance to targeted therapies in a highly reproducible fashion.

Among 512 patients who had metastatic colorectal cancer without RAS (KRAS or NRAS, 164790) mutations, Douillard et al. (2013) found that progression-free survival was 10.1 months with the combination of panitumumab-FOLFOX4 (oxaliplatin, fluorouracil, and leucovorin) versus 7.9 months with FOLFOX4 alone (hazard ratio for progression or death with combination therapy, 0.72; 95% CI 0.58 to 0.90; p = 0.004). Overall survival was 26.0 months in the panitumumab-FOLFOX4 group versus 20.2 months in the FOLFOX4-alone group (hazard ratio for death, 0.78; 95% CI 0.62-0.99; p = 0.04). A total of 108 patients (17%) with nonmutated KRAS exon 2 had other RAS mutations. These mutations were associated with inferior progression-free survival and overall survival with panitumumab-FOLFOX4 treatment, which was consistent with the findings in patients with KRAS mutations in exon 2. BRAF mutations were a negative prognostic factor.


Diagnosis

Prediction of Colorectal Cancer Risk

Loss of imprinting, an epigenetic alteration affecting the insulin-like growth factor II gene (IGF2; 147470), is found in normal colonic mucosa of about 30% of colorectal cancer patients, but it is found in only 10% of healthy individuals. In a pilot study to investigate the utility of loss of imprinting as a marker of colorectal cancer risk, Cui et al. (2003) evaluated 172 patients at a colonoscopy clinic. The adjusted odds ratio for loss of imprinting in lymphocytes was 5.15 for patients with a positive family history (95% CI, 1.70-16.96; p = 0.002), 3.46 for patients with adenomas (95% CI, 1.14-11.37; p = 0.026), and 21.7 for patients with colorectal cancer (95% CI, 3.48-153.6; p = 0.0005). Loss of imprinting can be assayed with a DNA-based blood test, and Cui et al. (2003) concluded that it may be a valuable predictive marker of an individual's risk for colorectal cancer.


Mapping

To identify susceptibility genes for familial colorectal neoplasia, Daley et al. (2008) conducted a comprehensive, genomewide linkage scan of 194 kindreds. Clinical information (histopathology, size and number of polyps, and other primary cancers) was used in conjunction with age at onset and family history for classification of the families into 5 phenotypic subgroups (severe histopathology, oligopolyposis, young colon/breast and multiple cancer) before analysis. By expanding the traditional affected sib pair design to include unaffected and discordant sib pairs, analytical power and robustness to type I error were increased. Linkage peaks of interest were identified at several sites. At marker D1S1665 (1p31.1), there was strong evidence for linkage in the multiple cancer subgroup (p = 0.00007). For 15q14-q22, a linkage peak was identified in the full sample, oligopolyposis, and young phenotypes. This region includes the locus associated with hereditary mixed polyposis syndrome (HMPS; 601228) in families of Ashkenazi descent. Daley et al. (2008) provided compelling evidence linking this region in families of European descent with oligopolyposis and/or young age at onset (51 years or younger) phenotypes. They found linkage to BRCA2 (600185) in the colon/breast phenotypic subgroup and identified a second locus in the region of D21S1437 segregating with, but distinct from, BRCA2. Linkage to 17p13.3 at marker D17S1308 in the breast/colon subgroup identified HIC1 (603825) as a candidate gene. The study demonstrated that using clinical information, unaffected sibs, and family history can increase the analytic power of a linkage study.

Associations Pending Confirmation

In a large kindred with excess colorectal cancer, Neklason et al. (2010) performed 2 separate genomewide scans and additional fine mapping and identified a single major locus on chromosome 13q22.1-q31.3 that segregated with adenomatous polyps and colon cancer, for which they obtained a nonparametric linkage score of 24 (lod score of 2.99; p = 0.001) at D13S251. Haplotype analysis identified a 21-Mb interval encompassing a nonrecombinant region bounded by rs2077779 and rs2351871 and containing 27 genes. Sequencing of 8 candidate genes failed to identify a clearly deleterious mutation. Neklason et al. (2010) noted that chromosome 13q is commonly gained and overexpressed in colon cancers and correlates with metastasis, suggesting the presence of an important cancer progression gene, and stated that evaluation of tumors from the kindred revealed a gain of chromosome 13q as well.


Cytogenetics

Bass et al. (2011) reported whole-genome sequencing from 9 individuals with colorectal cancer, including primary colorectal tumors and matched adjacent nontumor tissues, at an average of 30.7x and 31.9x coverage, respectively. They identified an average of 75 somatic rearrangements per tumor, including complex networks of translocations between pairs of chromosomes. Eleven rearrangements encode predicted in-frame fusion proteins, including a fusion of VTI1A (614316) and TCF7L2 (602278) found in 3 out of 97 colorectal cancers. Although TCF7L2 encodes TCF4, which cooperates with beta-catenin (116806) in colorectal carcinogenesis, the fusion lacks the TCF4 beta-catenin-binding domain. Bass et al. (2011) found a colorectal carcinoma cell line harboring the fusion gene to be dependent on VTI1A-TCF7L2 for anchorage-independent growth using RNA interference-mediated knockdown.


Molecular Genetics

In the DNA from 1 colon and 2 lung carcinoma cell lines, Perucho et al. (1981) demonstrated the same or closely related transforming elements. By DNA-mediated gene transfer, mouse fibroblasts could be morphologically transformed and rendered tumorigenic in nude mice.

In preliminary observations, Pathak and Goodacre (1986) found deletion of 12p in colorectal cancer specimens.

Fearon et al. (1987) studied the clonal composition of human colorectal tumors. Using X-linked RFLPs, they showed that all 50 tumors from females showed a monoclonal pattern of X-chromosome inactivation; these tumors included 20 carcinomas and 30 adenomas of either familial or spontaneous type. In over 75% of carcinomas examined, somatic loss of chromosome 17p sequences was found; such loss was rare in adenomas. Fearon et al. (1987) suggested that a gene on the short arm of chromosome 17 may be associated with progression from the benign to the malignant state.

By a combination of DNA hybridization analyses and tissue sectioning techniques, Bos et al. (1987) demonstrated that RAS gene mutations occur in over a third of colorectal cancers, that most of the mutations are at codon 12 of the KRAS gene (190070), and that the mutations usually precede the development of malignancy.

In 38 tumors from 25 patients with familial polyposis coli, and in 20 sporadic colon carcinomas, Okamoto et al. (1988) found frequent occurrence of allele loss on chromosome 22, with some additional losses on chromosomes 5, 6, 12q, and 15. The DNA probe C11p11, which has been found to be linked to familial polyposis coli, also detected frequent allele loss in both familial and sporadic colon carcinomas but not in benign adenomas. In a more extensive study, Vogelstein et al. (1988) studied the interrelationships of the 4 alterations demonstrated in colorectal cancer (RAS gene mutations and deletions of chromosome 5, 17 and 18 sequences) and determined their occurrence with respect to different stages of colorectal tumorigenesis. They found RAS gene mutations frequently in adenomas, this being the first demonstration of such in benign human tumors. In adenomas greater than 1 cm in size, the prevalence was similar to that observed in carcinomas (58% and 47%, respectively). Sequences on chromosome 5 that are linked to familial adenomatous polyposis were seldom lost in adenomas from such patients. Therefore, the Knudson model is unlikely to be applicable to the adenoma/carcinoma sequence in this disorder. Chromosome 18 sequences were lost frequently in colon carcinomas (73%) and in advanced adenomas (47%), but only occasionally in earlier stage adenomas (11-13%); see 120470. Chromosome 17 sequences were usually lost only in carcinomas (75%). The results suggested a model wherein the steps required for malignancy involve the activation of a dominantly acting oncogene coupled with the loss of several genes that normally suppress tumorigenesis.

Wildrick and Boman (1988) found deletion of the glucocorticoid receptor locus (138040), located on 5q, in colorectal cancers.

Law et al. (1988) examined the question of whether the gene for familial polyposis coli on chromosome 5 may be the site of changes leading to colorectal cancer in the general population, analogous to recessive tumor genes in retinoblastoma and Wilms tumor. To avoid error in interpretation of allelic loss from a study of nonhomogeneous samples, tumor cell populations were first microdissected from 24 colorectal carcinomas, an additional 9 cancers were engrafted in nude mice, and nuclei were flow-sorted in an additional 2. Of 31 cancers informative for chromosome 5 markers, only 6 (19%) showed loss of heterozygosity of chromosome 5 alleles, compared to 19 of 34 (56%) on chromosome 17, and 17 of 33 (52%) on chromosome 18. Law et al. (1988) concluded that FPC is a true dominant for adenomatosis but not a common recessive gene for colon cancer, and that simple mendelian models involving loss of alleles at a single locus may be inappropriate for understanding common human solid tumors.

Vogelstein et al. (1989) examined the extent and variation of allelic loss for polymorphic DNA markers in every nonacrocentric autosomal arm in 56 paired colorectal carcinoma and adjacent normal colonic mucosa specimens. They referred to the analysis as an allelotype, in analogy with a karyotype. Three major conclusions were drawn from the study: (1) Allelic deletions are remarkably common; 1 of the alleles of each polymorphic marker tested was lost in at least some tumors, and some tumors lost more than half of their parental alleles. (2) In addition to allelic deletions, new DNA fragments not present in normal tissue were identified in 5 carcinomas; these new fragments contained repeated sequences (of the variable-number-of-tandem-repeat type). (3) Patients with more than the median percentage of allelic deletions had a considerably worse prognosis than did the other patients, although the stage and size of the primary tumors were very similar in the 2 groups.

Delattre et al. (1989) reviewed the 3 general types of genetic alterations in colorectal cancer: (1) change in DNA content of the malignant cells as monitored by flow cytometry; (2) specific loss of genetic material, i.e., a complete loss of chromosome 18 and a structural rearrangement of chromosome 17 leading most often to the loss of 1 short arm, and loss of part of 5q as demonstrated by loss of heterozygosity; and (3) in nearly 40% of tumors, activation by point mutation of RAS oncogenes (never HRAS, rarely NRAS, and most frequently KRAS). In KRAS, with 1 exception, the activation has always occurred by a change in the coding properties of the twelfth or thirteenth codon. In studies of the multiple genetic alterations in colorectal cancer, Delattre et al. (1989) found that deletions and mitotic abnormalities occurred more frequently in distal than in proximal tumors. The frequency of KRAS mutations did not differ between proximal and distal cancers.

In studies of 15 colorectal tumors, Konstantinova et al. (1991) found rearrangements of the short arm of chromosome 17, leading to deletion of this arm or part of it in 12; in 2 others, one of the homologs of pair 17 was lost. One chromosome 18 was lost in 12 out of 13 cases with fully identified numerical abnormalities; chromosome 5, in 6 tumors; and other chromosomes in lesser numbers of cases. See 120470 for a discussion of a gene on chromosome 18 called DCC ('deleted in colorectal cancer') that shows mutations, including point mutations, in colorectal tumor tissue; also see 164790 for a discussion of a mutation in the NRAS oncogene in colorectal cancer.

On the basis of complex segregation analysis of a published series of consecutive pedigrees ascertained through patients undergoing treatment for colorectal cancer, Houlston et al. (1992) concluded that a dominant gene (or genes) with a frequency of 0.006 with a lifetime penetrance of 0.63 is likely. The gene was thought to account for 81% of colorectal cancer in patients under 35 years of age; however, by age 65, about 85% appeared to be phenocopies.

Fearon and Vogelstein (1990) reviewed the evidence supporting their multistep genetic model for colorectal tumorigenesis. They suggested that multiple mutations lead to a progression from normal epithelium to metastatic carcinoma through hyperplastic epithelium--early adenoma--intermediate adenoma--late adenoma--and carcinoma. The genes in which mutations occur at steps in this process include APC (611731) on chromosome 5, KRAS on chromosome 12, TP53 (191170) on 17p, and DCC on chromosome 18. Other genes that have been demonstrated or suspected of involvement in colorectal cancer include MSH2 (609309) on chromosome 2 and the DRA candidate colon tumor-suppressor gene (126650) on chromosome 7. Sarraf et al. (1999) presented evidence that colon cancer in humans is associated with loss-of-function mutations in the PPARG gene (601487).

Kikuchi-Yanoshita et al. (1992) presented evidence that genetic changes in both alleles of the TP53 gene through mutation and LOH, which result in abnormal protein accumulation, are involved in the conversion of adenoma to early carcinoma in both familial adenomatous polyposis and in nonfamilial polyposis cases.

Kinzler and Vogelstein (1996) gave a review of hereditary colorectal cancer and the multistep process of carcinogenesis that typically develops over decades and appears to require at least 7 genetic events for completion. They stated that the genetic defect in FAP involves the rate of tumor initiation by targeting the gatekeeper function of the APC gene. In contrast, the defect in HNPCC largely affects tumor aggression by targeting the genome guardian function of DNA repair.

Rajagopalan et al. (2002) systematically evaluated mutation in BRAF (164757) and KRAS (190070) in 330 colorectal tumors. There were 32 mutations in BRAF, 28 with a V600E mutation (164757.0001) and 1 each with the R462I (164757.0002), I463S (164757.0003), G464E (164757.0004), or K601E (164757.0005) mutations. All but 2 mutations seemed to be heterozygous, and in all 20 cases for which normal tissue was available, the mutations were shown to be somatic. In the same set of tumors there were 169 mutations in KRAS. No tumor exhibited mutations in both BRAF and KRAS. There was also a striking difference in the frequency of BRAF mutations between cancers with and without mismatch repair deficiency. All but 1 of the 15 BRAF mutations identified in mismatch repair deficient cases resulted in a V600E substitution. Rajagopalan et al. (2002) concluded their results provide strong support for the hypothesis that BRAF and KRAS mutations are equivalent in their tumorigenic effects. Both genes seem to be mutated at a similar phase of tumorigenesis, after initiation but before malignant conversion. Moreover, no tumor concurrently contained both BRAF and KRAS mutations.

To determine whether carriers of BLM (604610) mutations are at increased risk of colorectal cancer, Gruber et al. (2002) genotyped 1,244 cases of colorectal cancer and 1,839 controls, both of Ashkenazi Jewish ancestry, to estimate the relative risk of colorectal cancer among carriers of the BLM(Ash) founder mutation. Ashkenazi Jews with colorectal cancer were more than twice as likely to carry the BLM(Ash) (604610.0001) mutation than Ashkenazi Jewish controls without colorectal cancer (odds ratio = 2.45, 95% confidence interval 1.3 to 4.8; p = 0.0065). Gruber et al. (2002) verified that the APC I1307K mutation (611731.0029) did not confound their results.

Lynch and de la Chapelle (2003) provided a general discussion of hereditary colorectal cancer. They presented a flow diagram of the breakdown of 1,044 unselected consecutive patients with colorectal cancer. Tumors from 129 patients (12%) were positive for microsatellite instability; 28 of these patients were positive for germline mutations in MLH1 or MSH2, giving HNPCC a 2.7% frequency among the 1,044 patients. In the 88% of the patients whose tumors had no microsatellite instability, no mutations were found in MLH1 or MSH2.

Bardelli et al. (2003) used high-throughput sequencing technologies and bioinformatics to investigate how many or how often members of the tyrosine kinase family were altered in any particular cancer type. The protein kinase complement of the human genome (the 'kinome') can be organized into a dendrogram containing 9 broad groups of genes. Bardelli et al. (2003) selected 1 major branch of this dendrogram, containing 3 of the 9 groups, including the 90 tyrosine kinase genes (TK group), the 43 tyrosine kinase-like genes (TKL group), and the 5 receptor guanylate cyclase genes (RGC group), for mutation analysis. The 819 exons containing the kinase domains from the annotated TK, TKL, and RGC genes were screened from 35 colorectal cancer cell lines and were directly sequenced. Fourteen genes had somatic mutations within their kinase domains. Bardelli et al. (2003) analyzed these 14 genes for mutations in another 147 colorectal cancers and identified 46 mutations, 2 of which were synonymous; the remainder were either nonsynonymous or splice site alterations. All of these mutations were found to be somatic in the cancers that could be assessed by sequencing DNA from matched normal tissue. Seven genes were mutated in more than 1 tumor in the cohort: NTRK3 (191316), FES (190030), KDR (191306), EPHA3 (179611), NTRK2 (600456), MLK4, and GUCY2F (300041).

Samuels et al. (2004) examined the sequences of 117 exons that encode the predicted kinase domains of 8 phosphatidylinositol-3 kinase genes and 8 PI3K-like genes in 35 colorectal cancers. PIK3CA (171834) was the only gene with somatic mutations. Subsequent sequence analysis of all coding exons of PIK3CA in 199 additional colorectal cancers revealed mutations in a total of 74 tumors (32%). Samuels et al. (2004) also evaluated 76 premalignant colorectal tumors; only 2 mutations were found, both in very advanced tubulovillous adenomas greater than 5 cm in diameter. Thus, Samuels et al. (2004) concluded that PIK3CA mutations generally arise late in tumorigenesis, just before or coincident with invasion. Mutations in PIK3CA were also identified in 4 of 15 glioblastomas (27%), 3 of 12 gastric cancers (25%), 1 of 12 breast cancers (8%), and 1 of 24 lung cancers (4%). No mutations were observed in 11 pancreatic cancers or 12 medulloblastomas. In total, 92 mutations were observed, all of which were determined to be somatic in the cancers that could be assessed. Samuels et al. (2004) concluded that the sheer number of mutations observed in this gene strongly suggests that they are functionally important. Furthermore, most of the mutations were nonsynonymous and occurred in the PI3K helical and kinase domains, suggesting functional significance.

Clear-cut inherited mendelian traits, such as FAP or HNPCC, account for less than 4% of colorectal cancers. Another 20% of all colorectal cancers are thought to occur in individuals with a significant inherited multifactorial susceptibility to colorectal cancer that is not obviously familial. Incompletely penetrant, comparatively rare missense variants in the APC gene (611731) have been described in patients with multiple colorectal adenomas. For example, the I1307K mutation in the APC gene, which is found in Ashkenazi Jewish populations with an incidence of approximately 6%, confers a significantly increased risk of developing multiple adenomas and colorectal cancer. The glu1317-to-gln mutation in the APC gene (E1317Q; 611731.0036), which is found in non-Jewish Caucasian populations at a low frequency, similarly appears to confer a significantly increased risk of multiple adenomatous polyps. These variants represent a category of variation that has been suggested, generally, to account for a substantial fraction of such multifactorial inherited susceptibility to colorectal cancer. Fearnhead et al. (2004) explored this rare variant hypothesis for multifactorial inheritance using multiple colorectal adenomas as the model. Patients with multiple adenomas were screened for germline variants in a panel of candidate genes. Germline DNA was obtained from 124 patients with 3 to 100 histologically proven synchronous or metachronous adenomatous polyps. All patients were tested for the APC gene variants I1307K and E1317Q and for variants in the AXIN1 (603816), CTNNB1, MLH1, and MSH2 genes. The control group consisted of 483 randomly selected individuals. Potentially pathogenic germline variants were found in 30 of 124 patients (24.9%), compared with 55 of 483 controls (approximately 12%). This overall difference was highly significant, suggesting that many rare variants collectively contribute to inherited susceptibility to colorectal adenomas.

Parsons et al. (2005) selected 340 genes encoding serine/threonine kinases from the human genome and analyzed them for mutations in the kinase domain in tumors from colorectal cancer patients. A total of 23 changes, including 20 nonsynonymous point mutations, 1 insertion, and 1 splice site alteration, were identified. The gene mutations affected 8 different proteins: 6 were in mitogen-activated protein kinase kinase-4 (MKK4/JNKK1; 601335), 6 in myosin light-chain kinase-2 (MYLK2; 606566), 3 in phosphoinositide-dependent protein kinase-1 (PDK1; 605213, of which 2 mutations affected the same residue in the kinase domain), 2 in p21-activated kinase-4 (PAK4; 605451), 2 in v-akt murine thymoma viral oncogene homolog-2 kinase (AKT2; 164731), and 2 in MAP/microtubule affinity-regulating kinase-3 (MARK3; 602678); there was 1 alteration in cell-division cycle-7 kinase (CDC7; 603311) and another in a hypothetical casein kinase (PDIK1L). Eighteen of the 23 somatic mutations occurred at evolutionarily conserved residues. MKK4/JNKK1 is altered in a variety of tumor types, but no mutations in any of the other genes had theretofore been found in colorectal cancers. Three of the altered genes, PDK1, AKT2, and PAK4, encode proteins involved in the phosphatidylinositol-3-hydroxykinase pathway, and 2 of these (AKT2 and PAK4) are overexpressed in human cancers. Overall, nearly 40% of colorectal tumors had alterations in 1 of 8 PI(3)K-pathway genes.

Boraska Jelavic et al. (2006) studied genotype and allele frequencies of the GT microsatellite repeat polymorphism in intron 2 of the TLR2 gene (603028.0002) in 89 Croatian patients with sporadic colorectal cancer and 88 Croatian sex- and age-matched controls. The frequency of TLR2 alleles with 20 and 21 GT repeats was decreased (p = 0.0044 and p = 0.001, respectively) and the frequency of the allele with 31 GT repeats was increased (p = 0.0147) in patients versus controls. The authors also found that the gly299 allele of the TLR4 gene (603030.0001) was more frequent in colorectal cancer patients than controls (p = 0.0269).

Sjoblom et al. (2006) determined the sequence of well-annotated human protein-coding genes in 2 common tumor types. Analysis of 13,023 genes in 11 breast and 11 colorectal cancers revealed that individual tumors accumulate an average of about 90 mutant genes, but that only a subset of these contribute to the neoplastic process. Using stringent criteria to delineate this subset, Sjoblom et al. (2006) identified 189 genes (average of 11 per tumor) that were mutated at significant frequency. The vast majority of these were not known to be genetically altered in tumors and were predicted to affect a wide range of cellular functions, including transcription, adhesion, and invasion. Sjoblom et al. (2006) concluded that their data defined the genetic landscape of 2 human cancer types, provided new targets for diagnostic and therapeutic intervention, and opened fertile avenues for basic research in tumor biology.

Forrest and Cavet (2007), Getz et al. (2007), and Rubin and Green (2007) commented on the article by Sjoblom et al. (2006), citing statistical problems that, if addressed, would result in the identification of far fewer genes with significantly elevated mutation rates. Parmigiani et al. (2007) responded that the conclusions of the above authors were inaccurate because they were based on analyses that did not fully take into account the experimental design and other critical features of the Sjoblom et al. (2006) study.

To catalog the genetic changes that occur during tumorigenesis, Wood et al. (2007) isolated DNA from 11 breast and 11 colorectal tumors and determined the sequences of the genes in the Reference Sequence database in these samples. Based on analysis of exons representing 20,857 transcripts from 18,191 genes, Wood et al. (2007) concluded that the genomic landscapes of breast and colorectal cancers are composed of a handful of commonly mutated gene 'mountains' and a much larger number of gene 'hills' that are mutated at low frequency. Wood et al. (2007) described statistical and bioinformatic tools that may help identify mutations with a role in tumorigenesis. The gene mountains comprise well-known cancer genes such as APC (611731), KRAS (190070), and TP53 (191170). Furthermore, Wood et al. (2007) observed that most tumors accumulated approximately 80 mutations, and that the majority of these were harmless. Fewer than 15 mutations are likely to be responsible for driving the initiation, progression, or maintenance of the tumor.

Alhopuro et al. (2008) identified somatic mutations in the MYH11 gene in 56 (56%) of 101 samples of colorectal cancer tissue showing microsatellite instability. All 56 mutations were within a mononucleotide repeat of 8 cytosines (C8) in the last exon of the MYH11 SM2 isoform, which is susceptible to mutations under microsatellite instability, and were predicted to lead to a frameshift and elongation of the protein. All mutations were found within epithelial cells. Analysis of microsatellite stable tumors identified 2 somatic mutations in the same tumor that were not in the C8 repeat. Functional expression studies of the mutant proteins showed unregulated actin-activated motor activity.

McMurray et al. (2008) showed that a large proportion of genes controlled synergistically by loss-of-function p53 and Ras activation are critical to the malignant state of murine and human colon cells. Notably, 14 of 24 'cooperation response genes' were found to contribute to tumor formation in gene perturbation experiments. In contrast, only 1 of 14 perturbations of the genes responding in a nonsynergistic manner had a similar effect. McMurray et al. (2008) concluded that synergistic control of gene expression by oncogenic mutations thus emerges as an underlying key to malignancy, and provides an attractive rationale for identifying intervention targets in gene networks downstream of oncogenic gain- and loss-of-function mutations.

To help distinguish between driver and passenger mutations in colorectal cancer, Starr et al. (2009) used a transposon-based genetic screen in mice to identify candidate genes. Mice harboring mutagenic 'Sleeping Beauty' (SB) transposons were crossed with mice expressing SB transposase in gastrointestinal tract epithelium. Most of the offspring developed intestinal lesions including intraepithelial neoplasia, adenomas, and adenocarcinomas. Analysis of over 16,000 transposon insertions identified 77 candidate CRC genes, 60 of which are mutated and/or dysregulated in human CRC and thus are most likely to drive tumorigenesis. The genes included APC, PTEN (601728), and SMAD4 (600993). The screen also identified 17 candidate genes that had not been implicated in CRC, including POLI (605252), PTPRK (602545), and RSPO2 (610575).

In colonocytes from COX-deficient crypts from 2 patients with colon cancer, Greaves et al. (2006) identified 2 missense mutations in the MTCO1 gene (see 516030.0010 and 516030.0011, respectively).

Using high-throughput screening of 14,662 human protein coding transcripts, Sjoblom et al. (2006) found that the PKHD1 gene (606702) was the seventh most common somatically mutated gene in colorectal cancer. Germline mutations in the PKHD1 gene cause autosomal recessive polycystic kidney disease (263200). Ward et al. (2011) observed an association between the common T36M PKHD1 allele (606702.0001) and protection against colorectal cancer. Germline heterozygosity for the mutant allele was found in 0.42% of 3,603 healthy European controls and in 0.027% of 3,767 patients with colorectal cancer (p = 0.0002; odds ratio of 0.072). The authors postulated that reduced PKHD1 activity may enhance mitotic instability, which may inhibit carcinogenesis.

Dorard et al. (2011) identified a mutant of HSP110 (see 610703), which they called HSP110-delta-E9, in colorectal cancer showing microsatellite instability (MSI CRC), generated from an aberrantly spliced mRNA and lacking the HSP110 substrate-binding domain. This mutant was expressed at variable levels in almost all MSI CRC cell lines and primary tumors tested. HSP110-delta-E9 impaired both the normal cellular localization of HSP110 and its interaction with other HSPs, thus abrogating the chaperone activity and antiapoptotic function of HSP110 in a dominant-negative manner. HSP110-delta-E9 overexpression caused the sensitization of cells to anticancer agents such as oxaliplatin and 5-fluorouracil, which are routinely prescribed in the adjuvant treatment of people with colorectal cancer. The survival and response to chemotherapy of subjects with colorectal cancer showing microsatellite instability was associated with the tumor expression level of HSP110-delta-E9. Dorard et al. (2011) concluded that HSP110 may thus constitute a major determinant for both prognosis and treatment response in colorectal cancer.

The Cancer Genome Atlas Network (2012) conducted a genome-scale analysis of 276 colorectal carcinoma samples analyzing exome sequence, DNA copy number, promoter methylation, and mRNA and microRNA expression. A subset of these samples (97) underwent low-depth-of-coverage whole-genome sequencing. In total, 16% of colorectal carcinomas were found to be hypermutated: three-quarters of these had the expected high microsatellite instability, usually with hypermethylation and MLH1 silencing, and one-quarter had somatic mismatch-repair gene and polymerase epsilon mutations. Excluding the hypermutated cancers, colon and rectal cancers were found to have considerably similar patterns of genomic alteration. Twenty-four genes were significantly mutated. In addition to the expected APC, TP53, SMAD4, PIK3CA, and KRAS mutations, the authors found frequent mutations in ARID1A (603024), SOX9 (608160), and FAM123B (300647). Recurrent copy number alterations included potentially drug-targetable amplification of ERBB2 (164870) and amplification of IGF2 (147470). Recurrent chromosomal translocations included the fusion of NAV2 (607026) and WNT pathway member TCF7L1 (604652). Integrative analyses suggested new markers for aggressive colorectal carcinoma and an important role for MYC-directed transcriptional activation and repression.

By whole-exome sequencing, Segui et al. (2015) identified a heterozygous nonsense mutation in the FAN1 gene (613534) in a Spanish family (family 1) that met the Amsterdam criteria for risk of hereditary nonpolyposis CRC. The mutation was present in 2 affected brothers and an affected son, as well as that son's unaffected 47-year-old sister and his unaffected 21-year-old son. The variant was not found in the ESP or 1000 Genomes Project databases or in 1,648 alleles of Spanish origin, including 286 sporadic CRC patients. Analysis of FAN1 in another 176 mismatch repair (MMR)-proficient Amsterdam-positive families revealed affected individuals from 4 families (families 2 to 5) with a FAN1 mutation, including 1 nonsense and 3 missense variants. Only 2 unaffected family members were tested, and 1 was a mutation carrier. In vitro analysis of 1 of the missense variants suggested that it causes a DNA interstrand crosslink repair defect. Whole-exome sequencing of the tumor from the proband in family 1 showed a somatic mutation burden corresponding to that of nonhypermutant CRCs, and no clear evidence of somatic FAN1 second hits was obtained. The authors concluded that FAN1 is implicated in the inherited susceptibility to CRC.


See Also:

Lovett (1976); Lovett (1976); Mathis (1962)

REFERENCES

  1. Alhopuro, P., Phichith, D., Tuupanen, S., Sammalkorpi, H., Nybondas, M., Saharinen, J., Robinson, J. P., Yang, Z., Chen, L.-Q., Orntoft, T., Mecklin, J.-P., Jarvinen, H., and 12 others. Unregulated smooth-muscle myosin in human intestinal neoplasia. Proc. Nat. Acad. Sci. 105: 5513-5518, 2008. [PubMed: 18391202] [Full Text: https://doi.org/10.1073/pnas.0801213105]

  2. Bamezai, R., Singh, G., Khanna, N. N., Singh, S. Genetics of site specific colon cancer: a family study. Clin. Genet. 26: 129-132, 1984. [PubMed: 6088133] [Full Text: https://doi.org/10.1111/j.1399-0004.1984.tb00802.x]

  3. Bardelli, A., Parsons, D. W., Silliman, N., Ptak, J., Szabo, S., Saha, S., Markowitz, S., Willson, J. K. V., Parmigiani, G., Kinzler, K. W., Vogelstein, B., Velculescu, V. E. Mutational analysis of the tyrosine kinome in colorectal cancers. Science 300: 949 only, 2003. [PubMed: 12738854] [Full Text: https://doi.org/10.1126/science.1082596]

  4. Baron, J. A., Beach, M., Mandel, J. S., van Stolk, R. U., Haile, R. W., Sandler, R. S., Rothstein, R., Summers, R. W., Snover, D. C., Beck, G. J., Bond, J. H., Greenberg, E. R. Calcium supplements for the prevention of colorectal adenomas. New Eng. J. Med. 340: 101-107, 1999. [PubMed: 9887161] [Full Text: https://doi.org/10.1056/NEJM199901143400204]

  5. Baron, J. A., Cole, B. F., Sandler, R. S., Haile, R. W., Ahnen, D., Bresalier, R., McKeown-Eyssen, G., Summers, R. W., Rothstein, R., Burke, C. A., Snover, D. C., Church, T. R., and 12 others. A randomized trial of aspirin to prevent colorectal adenomas. New Eng. J. Med. 348: 891-899, 2003. [PubMed: 12621133] [Full Text: https://doi.org/10.1056/NEJMoa021735]

  6. Bass, A. J., Lawrence, M. S., Brace, L. E., Ramos, A. H., Drier, Y., Cibulskis, K., Sougnez, C., Voet, D., Saksena, G., Sivachenko, A., Jing, R., Parkin, M., and 28 others. Genomic sequencing of colorectal adenocarcinomas identifies a recurrent VTI1A-TCF7L2 fusion. Nature Genet. 43: 964-968, 2011. [PubMed: 21892161] [Full Text: https://doi.org/10.1038/ng.936]

  7. Batlle, E., Bacani, J., Begthel, H., Jonkeer, S., Gregorieff, A., van de Born, M., Malats, N., Sancho, E., Boon, E., Pawson, T., Gallinger, S., Pals, S., Clevers, H. EphB receptor activity suppresses colorectal cancer progression. Nature 435: 1126-1130, 2005. Note: Erratum: Nature 436: 881 only, 2005. [PubMed: 15973414] [Full Text: https://doi.org/10.1038/nature03626]

  8. Boraska Jelavic, T., Barisic, M., Drmic Hofman, I., Boraska, V., Vrdoljak, E., Peruzovic, M., Hozo, I., Puljiz, Z., Terjic, J. Microsatelite (sic) GT polymorphism in the toll-like receptor 2 is associated with colorectal cancer. Clin. Genet. 70: 156-160, 2006. [PubMed: 16879199] [Full Text: https://doi.org/10.1111/j.1399-0004.2006.00651.x]

  9. Bos, J. L., Fearon, E. R., Hamilton, S. R., Verlaan-de Vries, M., van Boom, J. H., van der Eb, A. J., Vogelstein, B. Prevalence of ras gene mutations in human colorectal cancers. Nature 327: 293-297, 1987. [PubMed: 3587348] [Full Text: https://doi.org/10.1038/327293a0]

  10. Budd, D. C., Fink, D. L. Mucoid colonic carcinoma as an autosomal-dominant inherited syndrome. Arch. Surg. 116: 901-905, 1981. [PubMed: 7259491] [Full Text: https://doi.org/10.1001/archsurg.1981.01380190035008]

  11. Burt, R. W., Bishop, D. T., Cannon, L. A., Dowdle, M. A., Lee, R. G., Skolnick, M. H. Dominant inheritance of adenomatous colonic polyps and colorectal cancer. New Eng. J. Med. 312: 1540-1544, 1985. [PubMed: 4000184] [Full Text: https://doi.org/10.1056/NEJM198506133122403]

  12. Cancer Genome Atlas Network. Comprehensive molecular characterization of human colon and rectal cancer. Nature 487: 330-337, 2012. [PubMed: 22810696] [Full Text: https://doi.org/10.1038/nature11252]

  13. Cannon-Albright, L. A., Skolnick, M. H., Bishop, T., Lee, R. G., Burt, R. W. Common inheritance of susceptibility to colonic adenomatous polyps and associated colorectal cancers. New Eng. J. Med. 319: 533-537, 1988. [PubMed: 2841598] [Full Text: https://doi.org/10.1056/NEJM198809013190902]

  14. Cui, H., Cruz-Correa, M., Giardiello, F. M., Hutcheon, D. F., Kafonek, D. R., Brandenburg, S., Wu, Y., He, X., Powe, N. R., Feinberg, A. P. Loss of IGF2 imprinting: a potential marker of colorectal cancer risk. Science 299: 1753-1755, 2003. [PubMed: 12637750] [Full Text: https://doi.org/10.1126/science.1080902]

  15. Daley, D., Lewis, S., Platzer, P., MacMillen, M., Willis, J., Elston, R. C., Markowitz, S. D., Wiesner, G. L. Identification of susceptibility genes for cancer in a genome-wide scan: results from the colon neoplasia sibling study. Am. J. Hum. Genet. 82: 723-736, 2008. [PubMed: 18313025] [Full Text: https://doi.org/10.1016/j.ajhg.2008.01.007]

  16. Delattre, O., Olschwang, S., Law, D. J., Melot, T., Remvikos, Y., Salmon, R. J., Sastre, X., Validire, P., Feinberg, A. P., Thomas, G. Multiple genetic alterations in distal and proximal colorectal cancer. Lancet 334: 353-356, 1989. Note: Originally Volume II. [PubMed: 2569552] [Full Text: https://doi.org/10.1016/s0140-6736(89)90537-0]

  17. Diaz, L. A., Jr., Williams, R. T., Wu, J., Kinde, I., Hecht, J. R., Berlin, J., Allen, B., Bozic, I., Reiter, J. G., Nowak, M. A., Kinzler, K. W., Oliner, K. S., Vogelstein, B. The molecular evolution of acquired resistance to targeted EGFR blockade in colorectal cancers. Nature 486: 537-540, 2012. [PubMed: 22722843] [Full Text: https://doi.org/10.1038/nature11219]

  18. Dorard, C., de Thonel, A., Collura, A., Marisa, L., Svrcek, M., Lagrange, A., Jego, G., Wanherdrick, K., Joly, A. L., Buhard, O., Gobbo, J., Penard-Lacronique, V., and 16 others. Expression of a mutant HSP110 sensitizes colorectal cancer cells to chemotherapy and improves disease prognosis. Nature Med. 17: 1283-1289, 2011. [PubMed: 21946539] [Full Text: https://doi.org/10.1038/nm.2457]

  19. Douillard, J.-Y., Oliner, K. S., Siena, S., Tabernero, J., Burkes, R., Barugel, M., Humblet, Y., Bodoky, G., Cunningham, D., Jassem, J., Rivera, F., Kocakova, I., and 10 others. Panitumumab-FOLFOX4 treatment and RAS mutations in colorectal cancer. New Eng. J. Med. 369: 1023-1034, 2013. [PubMed: 24024839] [Full Text: https://doi.org/10.1056/NEJMoa1305275]

  20. Fearnhead, N. S., Wilding, J. L., Winney, B., Tonks, S., Bartlett, S., Bicknell, D. C., Tomlinson, I. P. M., Mortensen, N. J. M., Bodmer, W. F. Multiple rare variants in different genes account for multifactorial inherited susceptibility to colorectal adenomas. Proc. Nat. Acad. Sci. 101: 15992-15997, 2004. [PubMed: 15520370] [Full Text: https://doi.org/10.1073/pnas.0407187101]

  21. Fearon, E. R., Hamilton, S. R., Vogelstein, B. Clonal analysis of human colorectal tumors. Science 238: 193-197, 1987. [PubMed: 2889267] [Full Text: https://doi.org/10.1126/science.2889267]

  22. Fearon, E. R., Vogelstein, B. A genetic model for colorectal tumorigenesis. Cell 61: 759-767, 1990. [PubMed: 2188735] [Full Text: https://doi.org/10.1016/0092-8674(90)90186-i]

  23. Forrest, W. F., Cavet, G. Comment on 'The consensus coding sequences of human breast and colorectal cancers.' (Abstract) Science 317: 1500 only, 2007.

  24. Getz, G., Hofling, H., Mesirov, J. P., Golub, T. R., Meyerson, M., Tibshirani, R., Lander, E. S. Comment on 'The consensus coding sequences of human breast and colorectal cancers.' (Abstract) Science 317: 1500 only, 2007.

  25. Greaves, L. C., Preston, S. L., Tadrous, P. J., Taylor, R. W., Barron, M. J., Oukrif, D., Leedham, S. J., Deheragoda, M., Sasieni, P., Novelli, M. R., Jankowski, J. A. Z., Turnbull, D. M., Wright, N. A., McDonald, S. A. C. Mitochondrial DNA mutations are established in human colonic stem cells, and mutated clones expand by crypt fission. Proc. Nat. Acad. Sci. 103: 714-719, 2006. [PubMed: 16407113] [Full Text: https://doi.org/10.1073/pnas.0505903103]

  26. Grivennikov, S. I., Wang, K., Mucida, D., Stewart, C. A., Schnabl, B., Jauch, D., Taniguchi, K., Yu, G.-Y., Osterreicher, C. H., Hung, K. E., Datz, C., Feng, Y., and 9 others. Adenoma-linked barrier defects and microbial products drive IL-23/IL-17-mediated tumour growth. Nature 491: 254-258, 2012. [PubMed: 23034650] [Full Text: https://doi.org/10.1038/nature11465]

  27. Gruber, S. B., Ellis, N. A., Scott, K. K., Almog, R., Kolachana, P., Bonner, J. D., Kirchhoff, T., Tomsho, L. P., Nafa, K., Pierce, H., Low, M., Satagopan, J., and 12 others. BLM heterozygosity and the risk of colorectal cancer. Science 297: 2013 only, 2002. Note: Erratum: Science 298: 751 only, 2002. [PubMed: 12242432] [Full Text: https://doi.org/10.1126/science.1074399]

  28. Houlston, R. S., Collins, A., Slack, J., Morton, N. E. Dominant genes for colorectal cancer are not rare. Ann. Hum. Genet. 56: 99-103, 1992. [PubMed: 1503398] [Full Text: https://doi.org/10.1111/j.1469-1809.1992.tb01136.x]

  29. Huber, S., Gagliani, N., Zenewicz, L. A., Huber, F. J., Bosurgi, L., Hu, B., Hedl, M., Zhang, W., O'Connor, W., Jr., Murphy, A. J., Valenzuela, D. M., Yancopoulos, G. D., Booth, C. J., Cho, J. H., Ouyang, W., Abraham, C., Flavell, R. A. IL-22BP is regulated by the inflammasome and modulates tumorigenesis in the intestine. Nature 491: 259-263, 2012. [PubMed: 23075849] [Full Text: https://doi.org/10.1038/nature11535]

  30. Jones, P. A., Gonzalgo, M. L. Altered DNA methylation and genome instability: a new pathway to cancer? Proc. Nat. Acad. Sci. 94: 2103-2105, 1997. [PubMed: 9122155] [Full Text: https://doi.org/10.1073/pnas.94.6.2103]

  31. Kikuchi-Yanoshita, R., Konishi, M., Ito, S., Seki, M., Tanaka, K., Maeda, Y., Iino, H., Fukayama, M., Koike, M., Mori, T., Sakuraba, H., Fukunari, H., Iwama, T., Miyaki, M. Genetic changes of both p53 alleles associated with the conversion from colorectal adenoma to early carcinoma in familial adenomatous polyposis and non-familial adenomatous polyposis patients. Cancer Res. 52: 3965-3971, 1992. [PubMed: 1319835]

  32. Kinzler, K. W., Vogelstein, B. Lessons from hereditary colorectal cancer. Cell 87: 159-170, 1996. [PubMed: 8861899] [Full Text: https://doi.org/10.1016/s0092-8674(00)81333-1]

  33. Kluge, T. Familial cancer of the colon. Acta Chir. Scand. 127: 392-398, 1964. [PubMed: 14149641]

  34. Konstantinova, L. N., Fleischman, E. W., Knisch, V. I., Perevozchikov, A. G., Kopnin, B. P. Karyotype pecularities (sic) of human colorectal adenocarcinomas. Hum. Genet. 86: 491-496, 1991. [PubMed: 2016090] [Full Text: https://doi.org/10.1007/BF00194640]

  35. Laird, P. W., Jackson-Grusby, L., Fazeli, A., Dickinson, S. L., Jung, W. E., Li, E., Weinberg, R. A., Jaenisch, R. Suppression of intestinal neoplasia by DNA hypomethylation. Cell 81: 197-205, 1995. [PubMed: 7537636] [Full Text: https://doi.org/10.1016/0092-8674(95)90329-1]

  36. Law, D. J., Olschwang, S., Monpezat, J. P., Lefrancois, D., Jagelman, D., Petrelli, N. J., Thomas, G., Feinberg, A. P. Concerted nonsyntenic allelic loss in human colorectal carcinoma. Science 241: 961-965, 1988. [PubMed: 2841761] [Full Text: https://doi.org/10.1126/science.2841761]

  37. Lengauer, C., Kinzler, K. W., Vogelstein, B. DNA methylation and genetic instability in colorectal cancer cells. Proc. Nat. Acad. Sci. 94: 2545-2550, 1997. [PubMed: 9122232] [Full Text: https://doi.org/10.1073/pnas.94.6.2545]

  38. Lengauer, C., Kinzler, K. W., Vogelstein, B. Genetic instability in colorectal cancers. Nature 386: 623-627, 1997. [PubMed: 9121588] [Full Text: https://doi.org/10.1038/386623a0]

  39. Liu, Y., Zhang, X., Han, C., Wan, G., Huang, X., Ivan, C., Jiang, D., Rodriguez-Aguayo, C., Lopez-Berestein, G., Rao, P. H., Maru, D. M., Pahl, A., He, X., Sood, A. K., Ellis, L. M., Anderl, J., Lu, X. TP53 loss creates therapeutic vulnerability in colorectal cancer. Nature 520: 697-701, 2015. Note: Erratum: Nature 597: E6, 2021. [PubMed: 25901683] [Full Text: https://doi.org/10.1038/nature14418]

  40. Lovett, E. Familial cancer of the gastro-intestinal tract. Brit. J. Surg. 63: 19-22, 1976. [PubMed: 1267869] [Full Text: https://doi.org/10.1002/bjs.1800630104]

  41. Lovett, E. Family studies in cancer of the colon and rectum. Brit. J. Surg. 63: 13-18, 1976. [PubMed: 1267868] [Full Text: https://doi.org/10.1002/bjs.1800630103]

  42. Lynch, H. T., de la Chapelle, A. Hereditary colorectal cancer. New Eng. J. Med. 348: 919-932, 2003. [PubMed: 12621137] [Full Text: https://doi.org/10.1056/NEJMra012242]

  43. Mathis, V. M. Familiaeres Colon Karzinom. Ein Stammbaum aus dem Kanton Aargau. Schweiz. Med. Wschr. 92: 1673-1678, 1962. [PubMed: 13933637]

  44. McMurray, H. R., Sampson, E. R., Compitello, G., Kinsey, C., Newman, L., Smith, B., Chen, S.-R., Klebanov, L., Salzman, P., Yakovlev, A., Land, H. Synergistic response to oncogenic mutations defines gene class critical to cancer phenotype. Nature 453: 1112-1116, 2008. [PubMed: 18500333] [Full Text: https://doi.org/10.1038/nature06973]

  45. Mecklin, J.-P. Frequency of hereditary colorectal carcinoma. Gastroenterology 93: 1021-1025, 1987. [PubMed: 2820826] [Full Text: https://doi.org/10.1016/0016-5085(87)90565-8]

  46. Misale, S., Yaeger, R., Hobor, S., Scala, E., Janakiraman, M., Liska, D., Valtorta, E., Schiavo, R., Buscarino, M., Siravegna, G., Bencardino, K., Cercek, A., and 14 others. Emergence of KRAS mutations and acquired resistance to anti-EGFR therapy in colorectal cancer. Nature 486: 532-536, 2012. [PubMed: 22722830] [Full Text: https://doi.org/10.1038/nature11156]

  47. Moldenhauer, G., Salnikov, A. V., Luttgau, S., Herr, I., Anderl, J., Faulstich, H. Therapeutic potential of amanitin-conjugated anti-epithelial cell adhesion molecule monoclonal antibody against pancreatic carcinoma. J. Nat. Cancer Inst. 104: 622-634, 2012. [PubMed: 22457476] [Full Text: https://doi.org/10.1093/jnci/djs140]

  48. Montagut, C., Dalmases, A., Bellosillo, B., Crespo, M., Pairet, S., Iglesias, M., Salido, M., Gallen, M., Marsters, S., Tsai, S. P., Minoche, A., Seshagiri, S., Serrano, S., Himmelbauer, H., Bellmunt, J., Rovira, A., Settleman, J., Bosch, F., Albanell, J. Identification of a mutation in the extracellular domain of the epidermal growth factor receptor conferring cetuximab resistance in colorectal cancer. Nature Med. 18: 221-223, 2012. Note: Erratum: Nature Med. 18: 1445 only, 2012. [PubMed: 22270724] [Full Text: https://doi.org/10.1038/nm.2609]

  49. Neklason, D. W., Tuohy, T. M., Stevens, J., Otterud, B., Baird, L., Kerber, R. A., Samowitz, W. S., Kuwada, S. K., Leppert, M. F., Burt, R. W. Colorectal adenomas and cancer link to chromosome 13q22.1-13q31.3 in a large family with excess colorectal cancer. J. Med. Genet. 47: 692-699, 2010. [PubMed: 20522424] [Full Text: https://doi.org/10.1136/jmg.2009.076091]

  50. Okamoto, M., Sasaki, M., Sugio, K., Sato, C., Iwama, T., Ikeuchi, T., Tonomura, A., Sasazuki, T., Miyaki, M. Loss of constitutional heterozygosity in colon carcinoma from patients with familial polyposis coli. Nature 331: 273-277, 1988. [PubMed: 2827040] [Full Text: https://doi.org/10.1038/331273a0]

  51. Parmigiani, G., Lin, J., Boca, S. M., Sjoblom, T., Jones, S., Wood, L. D., Parsons, D. W., Barber, T., Buckhaults, P., Markowitz, S. D., Park, B. H., Bachman, K. E., Papadopoulos, N., Vogelstein, B., Kinzler, K. W., Velculescu, V. E. Response to comments on 'The consensus coding sequences of human breast and colorectal cancers.' (Abstract) Science 317: 1500 only, 2007.

  52. Parsons, D. W., Wang, T.-L., Samuels, Y., Bardelli, A., Cummins, J. M., DeLong, L., Silliman, N., Ptak, J., Szabo, S., Willson, J. K. V., Markowitz, S., Kinzler, K. W., Vogelstein, B., Lengauer, C., Velculescu, V. E. Mutations in a signalling pathway. (Letter) Nature 436: 792 only, 2005. [PubMed: 16094359] [Full Text: https://doi.org/10.1038/436792a]

  53. Pathak, S., Goodacre, A. Specific chromosome anomalies and predisposition to human breast, renal cell, and colorectal carcinoma. Cancer Genet. Cytogenet. 19: 29-36, 1986. [PubMed: 3940177] [Full Text: https://doi.org/10.1016/0165-4608(86)90368-7]

  54. Perucho, M., Goldfarb, M., Shimizu, K., Lama, C., Fogh, J., Wigler, M. Human-tumor-derived cell lines contain common and different transforming genes. Cell 27: 467-476, 1981. [PubMed: 6101201] [Full Text: https://doi.org/10.1016/0092-8674(81)90388-3]

  55. Ponz de Leon, M., Scapoli, C., Zanghieri, G., Sassatelli, R., Sacchetti, C., Barrai, I. Genetic transmission of colorectal cancer: exploratory data analysis from a population based registry. J. Med. Genet. 29: 531-538, 1992. [PubMed: 1518021] [Full Text: https://doi.org/10.1136/jmg.29.8.531]

  56. Prahallad, A., Sun, C., Huang, S., Di Nicolantonio, F., Salazar, R., Zecchin, D., Beijersbergen, R. L., Bardelli, A., Bernards, R. Unresponsiveness of colon cancer to BRAF(V600E) inhibition through feedback activation of EGFR. Nature 483: 100-103, 2012. [PubMed: 22281684] [Full Text: https://doi.org/10.1038/nature10868]

  57. Rajagopalan, H., Bardelli, A., Lengauer, C., Kinzler, K. W., Vogelstein, B., Velculescu, V. E. RAF/RAS oncogenes and mismatch-repair status. (Letter) Nature 418: 934 only, 2002. [PubMed: 12198537] [Full Text: https://doi.org/10.1038/418934a]

  58. Rubin, A. F., Green, P. Comment on 'The consensus coding sequences of human breast and colorectal cancers.' (Abstract) Science 317: 1500 only, 2007.

  59. Samuels, Y., Wang, Z., Bardelli, A., Silliman, N., Ptak, J., Szabo, S., Yan, H., Gazdar, A., Powell, S. M., Riggins, G. J., Willson, J. K. V., Markowitz, S., Kinzler, K. W., Vogelstein, B., Velculescu, V. E. High frequency of mutations of the PIK3CA gene in human cancers. Science 304: 554 only, 2004. [PubMed: 15016963] [Full Text: https://doi.org/10.1126/science.1096502]

  60. Sandler, R. S., Halabi, S., Baron, J. A., Budinger, S., Paskett, E., Keresztes, R., Petrelli, N., Pipas, J. M., Karp, D. D., Loprinzi, C. L., Steinbach, G., Schilsky, R. A randomized trial of aspirin to prevent colorectal adenomas in patients with previous colorectal cancer. New Eng. J. Med. 348: 883-890, 2003. Note: Erratum: New Eng. J. Med. 348: 1939 only, 2003. [PubMed: 12621132] [Full Text: https://doi.org/10.1056/NEJMoa021633]

  61. Sarraf, P., Mueller, E., Smith, W. M., Wright, H. M., Kum, J. B., Aaltonen, L. A., de la Chapelle, A., Spiegelman, B. M., Eng, C. Loss-of-function mutations in PPAR-gamma associated with human colon cancer. Molec. Cell 3: 799-804, 1999. [PubMed: 10394368] [Full Text: https://doi.org/10.1016/s1097-2765(01)80012-5]

  62. Schweiger, M.-R., Hussong, M., Rohr, C., Lehrach, H. Genomics and epigenomics of colorectal cancer. Wiley Interdiscip. Rev. Syst. Biol. Med. 5: 205-219, 2013. [PubMed: 23325509] [Full Text: https://doi.org/10.1002/wsbm.1206]

  63. Segui, N., Mina, L. B., Lazaro, C., Sanz-Pamplona, R., Pons, T., Navarro, M., Bellido, F., Lopez-Doriga, A., Valdes-Mas, R., Pineda, M., Guino, E., Vidal, A., and 21 others. Germline mutations in FAN1 cause hereditary colorectal cancer by impairing DNA repair. Gastroenterology 149: 563-566, 2015. [PubMed: 26052075] [Full Text: https://doi.org/10.1053/j.gastro.2015.05.056]

  64. Seshagiri, S., Stawiski, E. W., Durinck, S., Modrusan, Z., Storm, E. E., Conboy, C. B., Chaudhuri, S., Guan, Y., Janakiraman, V., Jaiswal, B. S., Guillory, J., Ha, C., and 15 others. Recurrent R-spondin fusions in colon cancer. Nature 488: 660-664, 2012. [PubMed: 22895193] [Full Text: https://doi.org/10.1038/nature11282]

  65. Sivak, M. V., Jr., Sivak, D. S., Braun, W. A., Sullivan, B. H., Jr. A linkage study of HLA and inherited adenocarcinoma of the colon. Cancer 48: 76-81, 1981. [PubMed: 7237394] [Full Text: https://doi.org/10.1002/1097-0142(19810701)48:1<76::aid-cncr2820480116>3.0.co;2-r]

  66. Sjoblom, T., Jones, S., Wood, L. D., Parsons, D. W., Lin, J., Barber, T. D., Mandelker, D., Leary, R. J., Ptak, J., Silliman, N., Szabo, S., Buckhaults, P., and 17 others. The consensus coding sequences of human breast and colorectal cancers. Science 314: 268-274, 2006. [PubMed: 16959974] [Full Text: https://doi.org/10.1126/science.1133427]

  67. Starr, T. K., Allaei, R., Silverstein, K. A. T., Staggs, R. A., Sarver, A. L., Bergemann, T. L., Gupta, M., O'Sullivan, M. G., Matise, I., Dupuy, A. J., Collier, L. S., Powers, S., Oberg, A. L., Asmann, Y. W., Thibodeau, S. N., Tessarollo, L., Copeland, N. G., Jenkins, N. A., Cormier, R. T., Largaespada, D. A. A transposon-based genetic screen in mice identifies genes altered in colorectal cancer. Science 323: 1747-1750, 2009. [PubMed: 19251594] [Full Text: https://doi.org/10.1126/science.1163040]

  68. Thomas, J. H. Genomic imprinting proposed as a surveillance mechanism for chromosome loss. Proc. Nat. Acad. Sci. 92: 480-482, 1995. [PubMed: 7831314] [Full Text: https://doi.org/10.1073/pnas.92.2.480]

  69. Vermeulen, L., Morrissey, E., van der Heijden, M., Nicholson, A. M., Sottoriva, A., Buczacki, S., Kemp, R., Tavare, S., Winton, D. J. Defining stem cell dynamics in models of intestinal tumor initiation. Science 342: 995-998, 2013. [PubMed: 24264992] [Full Text: https://doi.org/10.1126/science.1243148]

  70. Vilar, E., Gruber, S. B. Microsatellite instability in colorectal cancer--the stable evidence. Nature Rev. Clin. Oncol. 7: 153-162, 2010. [PubMed: 20142816] [Full Text: https://doi.org/10.1038/nrclinonc.2009.237]

  71. Vogelstein, B., Fearon, E. R., Hamilton, S. R., Kern, S. E., Preisinger, A. C., Leppert, M., Nakamura, Y., White, R., Smits, A. M. M., Bos, J. L. Genetic alterations during colorectal-tumor development. New Eng. J. Med. 319: 525-532, 1988. [PubMed: 2841597] [Full Text: https://doi.org/10.1056/NEJM198809013190901]

  72. Vogelstein, B., Fearon, E. R., Kern, S. E., Hamilton, S. R., Preisinger, A. C., Nakamura, Y., White, R. Allelotype of colorectal carcinomas. Science 244: 207-211, 1989. [PubMed: 2565047] [Full Text: https://doi.org/10.1126/science.2565047]

  73. Ward, C. J., Wu, Y., Johnson, R. A., Woollard, J. R., Bergstralh, E. J., Cicek, M. S., Bakeberg, J., Rossetti, S., Heyer, C. M., Petersen, G. M., Lindor, N. M., Thibodeau, S. N., Harris, P. C., Torres, V. E., Hogan, M. C., Boardman, L. A. Germline PKHD1 mutations are protective against colorectal cancer. Hum. Genet. 129: 345-349, 2011. [PubMed: 21274727] [Full Text: https://doi.org/10.1007/s00439-011-0950-8]

  74. Wheeler, J. M. D., Warren, B. F., Mortensen, N. J. M., Kim, H. C., Biddolph, S. C., Elia, G., Beck, N. E., Williams, G. T., Shepherd, N. A., Bateman, A. C., Bodmer, W. F. An insight into the genetic pathway of adenocarcinoma of the small intestine. Gut 50: 218-223, 2002. [PubMed: 11788563] [Full Text: https://doi.org/10.1136/gut.50.2.218]

  75. Wildrick, D. M., Boman, B. M. Chromosome 5 allele loss at the glucocorticoid receptor locus in human colorectal carcinomas. Biochem. Biophys. Res. Commun. 150: 591-598, 1988. [PubMed: 3342039] [Full Text: https://doi.org/10.1016/0006-291x(88)90434-2]

  76. Williams, C. Management of malignancy in 'cancer families'. Lancet 311: 198-199, 1978. Note: Originally Volume I. [PubMed: 74618] [Full Text: https://doi.org/10.1016/s0140-6736(78)90627-x]

  77. Wood, L. D., Parsons, D. W., Jones, S., Lin, J., Sjoblom, T., Leary, R. J., Shen, D., Boca, S. M., Barber, T., Ptak, J., Silliman, N., Szabo, S., and 30 others. The genomic landscapes of human breast and colorectal cancers. Science 318: 1108-1113, 2007. [PubMed: 17932254] [Full Text: https://doi.org/10.1126/science.1145720]

  78. Xia, D., Wang, D., Kim, S.-H., Katoh, H., DuBois, R. N. Prostaglandin E2 promotes intestinal tumor growth via DNA methylation. Nature Med. 18: 224-226, 2012. [PubMed: 22270723] [Full Text: https://doi.org/10.1038/nm.2608]


Contributors:
Marla J. F. O'Neill - updated : 07/14/2021
Marla J. F. O'Neill - updated : 09/02/2016
Ada Hamosh - updated : 6/25/2015
Ada Hamosh - updated : 12/6/2013
Ada Hamosh - updated : 10/23/2013
Cassandra L. Kniffin - updated : 2/18/2013
Ada Hamosh - updated : 12/4/2012
Ada Hamosh - updated : 9/18/2012
Ada Hamosh - updated : 9/5/2012
Ada Hamosh - updated : 8/10/2012
Ada Hamosh - updated : 7/17/2012
Ada Hamosh - updated : 6/26/2012
Ada Hamosh - updated : 3/15/2012
Ada Hamosh - updated : 3/14/2012
Ada Hamosh - updated : 12/12/2011
Cassandra L. Kniffin - updated : 4/20/2011
Marla J. F. O'Neill - updated : 12/1/2010
Carol A. Bocchini - updated : 11/4/2010
Marla J. F. O'Neill - updated : 10/5/2009
Ada Hamosh - updated : 9/14/2009
Ada Hamosh - updated : 6/18/2009
Ada Hamosh - updated : 7/29/2008
Ada Hamosh - updated : 7/18/2008
Cassandra L. Kniffin - updated : 4/28/2008
Ada Hamosh - updated : 2/14/2008
Ada Hamosh - updated : 1/9/2008
Victor A. McKusick - updated : 11/20/2007
Ada Hamosh - updated : 10/31/2006
Marla J. F. O'Neill - updated : 9/22/2006
Ada Hamosh - updated : 9/8/2005
Ada Hamosh - updated : 7/27/2005
Victor A. McKusick - updated : 4/15/2005
Ada Hamosh - updated : 4/30/2004
Ada Hamosh - updated : 5/29/2003
Ada Hamosh - updated : 4/3/2003
Victor A. McKusick - updated : 3/14/2003
Ada Hamosh - updated : 9/30/2002
George E. Tiller - updated : 9/26/2002
Ada Hamosh - updated : 9/17/2002
Paul Brennan - updated : 3/19/2002
Paul Brennan - updated : 3/13/2002
Paul Brennan - updated : 3/6/2002
George E. Tiller - updated : 6/19/2001
Stylianos E. Antonarakis - updated : 7/20/1999
Victor A. McKusick - updated : 2/9/1999
Victor A. McKusick - updated : 4/21/1997

Creation Date:
Victor A. McKusick : 6/4/1986

Edit History:
carol : 10/10/2023
carol : 10/09/2023
carol : 11/05/2021
carol : 07/15/2021
alopez : 07/14/2021
carol : 04/22/2020
carol : 09/13/2019
carol : 09/02/2016
carol : 05/04/2016
alopez : 11/24/2015
alopez : 7/7/2015
alopez : 6/25/2015
alopez : 12/6/2013
alopez : 10/23/2013
tpirozzi : 10/1/2013
carol : 3/11/2013
carol : 2/19/2013
ckniffin : 2/18/2013
alopez : 12/6/2012
terry : 12/4/2012
alopez : 9/19/2012
terry : 9/18/2012
alopez : 9/5/2012
carol : 8/10/2012
terry : 8/10/2012
terry : 7/27/2012
alopez : 7/19/2012
terry : 7/17/2012
alopez : 6/26/2012
terry : 6/26/2012
alopez : 3/15/2012
alopez : 3/14/2012
alopez : 12/19/2011
terry : 12/12/2011
carol : 9/7/2011
wwang : 5/2/2011
ckniffin : 4/20/2011
ckniffin : 4/20/2011
carol : 4/20/2011
wwang : 12/3/2010
carol : 12/2/2010
terry : 12/1/2010
carol : 11/4/2010
carol : 3/19/2010
alopez : 2/4/2010
ckniffin : 1/15/2010
wwang : 10/14/2009
terry : 10/5/2009
wwang : 9/29/2009
alopez : 9/14/2009
wwang : 7/29/2009
alopez : 6/24/2009
terry : 6/18/2009
wwang : 2/13/2009
ckniffin : 2/9/2009
terry : 1/12/2009
terry : 1/9/2009
carol : 9/19/2008
alopez : 8/18/2008
terry : 7/29/2008
wwang : 7/18/2008
wwang : 6/9/2008
ckniffin : 4/28/2008
carol : 2/15/2008
alopez : 2/15/2008
terry : 2/14/2008
ckniffin : 2/5/2008
carol : 1/31/2008
ckniffin : 1/28/2008
ckniffin : 1/28/2008
alopez : 1/28/2008
terry : 1/9/2008
alopez : 12/7/2007
terry : 11/20/2007
alopez : 9/27/2007
alopez : 8/31/2007
alopez : 11/3/2006
alopez : 11/3/2006
terry : 10/31/2006
alopez : 10/9/2006
wwang : 9/22/2006
wwang : 5/17/2006
carol : 4/14/2006
alopez : 12/5/2005
alopez : 9/9/2005
terry : 9/8/2005
alopez : 7/28/2005
terry : 7/27/2005
carol : 6/3/2005
mgross : 4/15/2005
mgross : 4/14/2005
mgross : 4/14/2005
mgross : 4/13/2005
tkritzer : 2/11/2005
alopez : 4/30/2004
terry : 4/30/2004
carol : 7/10/2003
mgross : 5/29/2003
joanna : 5/29/2003
terry : 5/29/2003
terry : 4/3/2003
carol : 3/21/2003
tkritzer : 3/18/2003
terry : 3/14/2003
alopez : 9/30/2002
alopez : 9/30/2002
tkritzer : 9/30/2002
cwells : 9/26/2002
alopez : 9/17/2002
alopez : 3/19/2002
alopez : 3/13/2002
alopez : 3/6/2002
cwells : 6/20/2001
cwells : 6/19/2001
carol : 10/20/2000
carol : 8/12/1999
mgross : 7/20/1999
mgross : 2/16/1999
mgross : 2/15/1999
terry : 2/9/1999
alopez : 6/27/1997
jenny : 4/21/1997
terry : 4/14/1997
terry : 12/10/1996
terry : 12/9/1996
carol : 5/31/1994
terry : 5/13/1994
mimadm : 4/9/1994
warfield : 4/6/1994
carol : 2/24/1993
carol : 10/12/1992