Download as pdf or txt
Download as pdf or txt
You are on page 1of 41

14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

Official reprint from UpToDate®

www.uptodate.com
© 2021 UpToDate, Inc. and/or its affiliates. All Rights Reserved.

Molecular genetics of colorectal cancer


Authors: Harold Frucht, MD, Aimee L Lucas, MD, MS
Section Editors: Richard M Goldberg, MD, Benjamin A Raby, MD, MPH
Deputy Editor: Diane MF Savarese, MD

All topics are updated as new evidence becomes available and our peer review process is complete.

Literature review current through: Aug 2021. | This topic last updated: Jan 21, 2021.

INTRODUCTION

Colorectal cancer (CRC) is a common disease. Approximately 149,500 new cases are
diagnosed each year in the United States, of which 104,270 originate in the colon and the
rest originate in the rectum [1]. Annually, approximately 52,980 Americans die of CRC,
accounting for approximately 9 percent of all cancer deaths; in the United States, CRC ranks
third in both incidence and cause of cancer death in both men and women. Global, country-
specific data on incidence and mortality are available from the World Health Organization
(WHO) GLOBOCAN database.

The risk factors for CRC are both environmental and inherited. The mode of presentation of
CRC follows one of three patterns that are reflective of these differing risk factors: sporadic,
inherited, and familial:

● Sporadic disease, in which there is no family history, accounts for approximately 70


percent of all CRCs. It is most common over the age of 50, and dietary and
environmental factors have been etiologically implicated. (See "Colorectal cancer:
Epidemiology, risk factors, and protective factors".)

● Fewer than 10 percent of patients have a true inherited predisposition to CRC, and these
cases are subdivided according to whether or not colonic polyps are a major disease
manifestation. The diseases with polyposis include familial adenomatous polyposis
(FAP), MUTYH-associated polyposis (MAP), and the hamartomatous polyposis syndromes
(eg, Peutz-Jeghers, juvenile polyposis [2], phosphatase and tensin homolog [PTEN]
hamartoma tumor [Cowden] syndrome), while those without polyposis are referred to
as hereditary nonpolyposis CRC (HNPCC; Lynch syndrome). These conditions are all
associated with a high risk of developing CRC. In many cases, the causative genetic

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&sel… 1/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

mutation has been identified, and a test is available. (See "Clinical manifestations and
diagnosis of familial adenomatous polyposis" and "MUTYH-associated polyposis" and
"Peutz-Jeghers syndrome: Clinical manifestations, diagnosis, and management" and
"Juvenile polyposis syndrome" and "PTEN hamartoma tumor syndromes, including
Cowden syndrome" and "Lynch syndrome (hereditary nonpolyposis colorectal cancer):
Clinical manifestations and diagnosis".)

● The third and least well understood pattern is known as "familial" CRC, which accounts
for up to 25 percent of cases. Affected patients have a family history of CRC, but the
pattern is not consistent with one of the inherited syndromes described above.
Individuals from these families are at increased risk of developing CRC, although the risk
is not as high as with the inherited syndromes. Having a single affected first-degree
relative (ie, parent, child, sibling) increases the risk of developing CRC 1.7-fold over that
of the general population. The risk is further increased if two first-degree relatives have
CRC or if the index case is diagnosed before age 55. (See "Colorectal cancer:
Epidemiology, risk factors, and protective factors".)

Some of these patients may have familial CRC type X, in which clinical criteria are met for
Lynch syndrome but in the absence of an identified germline mutation in one of the
mismatch repair genes, the genetic hallmark of Lynch syndrome [3]. The term familial
CRC syndrome is probably a misnomer as it is likely that these patients have a currently
unidentified inherited genetic mutation. However, data suggest that individuals with
CRC arising in the context of familial CRC type X do not have outcomes that are as
favorable as those seen in individuals with CRC in the setting of Lynch syndrome [4].
(See 'Mismatch repair genes' below.)

In general, the mechanisms underlying familial clustering of CRC in the absence of a


discernible inherited syndrome remain incompletely understood.

Our level of understanding of the molecular events underlying CRC is far greater than for
other common solid tumors. Specific germline mutations are responsible for the inherited
CRC syndromes, while a stepwise accumulation of somatic mutations is thought to underlie
most sporadic cases. In contrast, the genetic abnormalities underlying familial CRC remain
incompletely understood.

This topic will review the major genetic aspects of colorectal carcinogenesis, with particular
emphasis on sporadic CRC. Inherited conditions that significantly increase the risk of CRC
(eg, FAP and Lynch syndrome) and the role of genetic testing and screening for patients with
an inherited predisposition are discussed elsewhere. (See "Clinical manifestations and
diagnosis of familial adenomatous polyposis" and "Lynch syndrome (hereditary
nonpolyposis colorectal cancer): Clinical manifestations and diagnosis" and "Familial
adenomatous polyposis: Screening and management of patients and families" and "Lynch
https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&sel… 2/41
14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

syndrome (hereditary nonpolyposis colorectal cancer): Cancer screening and management"


and "Juvenile polyposis syndrome".)

MOLECULAR PATHOGENESIS OF COLORECTAL CANCER

Specific genetic changes are thought to drive the transformation from normal colonic
epithelium to invasive cancer. Genetic mutations can be inherited or acquired. Any genetic
mutation that occurs at or before fertilization of the ovum is termed a germline mutation
and can be transmitted from parent to offspring as an inherited defect. If the mutation
occurs spontaneously in the sperm, ovum, or zygote, the affected person's parents do not
manifest the cancer phenotype, but future progeny may inherit the de novo mutation. More
commonly, a spontaneous mutation appears in a cell during the growth and/or development
of a particular tissue or organ; this is called a somatic mutation. Because these mutations
often confer a selective growth advantage, they result in preferential proliferation of the cell
containing the mutated genetic material (clonal evolution) [5].

The clonal nature of tumors is a critical feature of the somatic mutation/clonal evolution
theory of human carcinogenesis. According to this model, the growth advantage acquired by
a single mutated cell allows its progeny to outnumber those of neighboring cells. From
within this clonal population, a single cell acquires a second mutation, providing an
additional growth advantage that allows further clonal expansion. Subsequent waves of
clonal expansion are driven by the sequential acquisition of more mutations, further cellular
disorganization, and eventually the ability to invade and metastasize.

The adenoma-carcinoma sequence — Most human CRCs are thought to arise from


adenomas (adenomatous polyps) that become dysplastic ( figure 1). Adenomatous polyps
form in the colon when normal mechanisms regulating epithelial renewal are disrupted.
Surface cells lining the intestine are continuously lost into the bowel lumen due to apoptosis
and exfoliation, and must be continuously replaced. Typically, proliferation occurs exclusively
at the crypt base. As cells move towards the luminal surface, they cease proliferating and
terminally differentiate. This ordered process is increasingly disrupted as adenomas increase
in size, become dysplastic, and eventually attain invasive potential.

The hypothesis that invasive CRCs develop from intermediate precancerous precursors is
supported by pathologic, epidemiologic, and observational clinical data (see "Overview of
colon polyps", section on 'Adenomatous polyps'). Summarized briefly:

● Early carcinomas are frequently seen within large adenomatous polyps, and areas of
adenomatous change can often be found surrounding human CRCs.

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&sel… 3/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

● Adenomas and carcinomas are found in similar distributions throughout the large
bowel, and adenomas are typically observed 10 to 15 years prior to the onset of cancer
in both sporadic and familial cases.

● In animal models, adenomas develop before carcinomas, and carcinomas develop


exclusively in adenomatous tissue.

● The ability to reduce the incidence of CRC through removal of polyps has been shown in
controlled trials in humans [6].

In 1990, Fearon and Vogelstein described the molecular basis for CRC as a multistep process
in which each accumulated genetic event conferred a selective growth advantage to the
colonic epithelial cell [5]. Later studies have served to further refine their hypothesis.

According to the Vogelstein model, germline or somatic mutations are required for
malignant transformation, and the accumulation of multiple genetic mutations rather than
their sequence determines the biological behavior of the tumor ( figure 2). Germline
mutations underlie the common inherited syndromes (eg, familial adenomatous polyposis
[FAP], Lynch syndrome), while sporadic cancers result from the stepwise accumulation of
multiple somatic mutations. Mutations in the adenomatous polyposis coli (APC) gene, a
feature common to both inherited and sporadic CRCs, occur early in the process, while p53
tumor suppressor gene mutations generally occur late.

In addition to point mutations, other genetic changes that are implicated in human
tumorigenesis include altered DNA methylation, and gene rearrangements, amplifications,
overexpression, and deletions.

Serrated polyp pathway — The majority of CRCs are believed to progress through an


adenoma-carcinoma sequence. However, more recent evidence increasingly supports the
existence of an alternative route for colorectal carcinogenesis through serrated polyps, a
group that encompasses a morphological spectrum, including hyperplastic polyps, mixed
hyperplastic polyp/adenoma, and serrated adenomas. (See 'Hypermethylation phenotype
(CIMP+) pathway' below and "Overview of colon polyps", section on 'Sessile serrated polyps
and traditional serrated adenomas'.)

Molecular pathways to colorectal tumorigenesis — There appear to be at least three


molecular pathways leading to colorectal tumorigenesis ( table 1): the chromosomal
instability (CIN) pathway, which is typified by the inherited condition FAP; the mutator-
phenotype/DNA mismatch repair pathway, which is implicated in the inherited condition
Lynch syndrome as well as in a proportion of sporadic CRCs in which there is loss of DNA
mismatch repair protein function; and the hypermethylation phenotype
hyperplastic/serrated polyp pathway, which is characterized by a high frequency of

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&sel… 4/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

methylation of some CpG islands (CpG island hypermethylation phenotype [CIMP]-positive)


[7].

The chromosomal instability (APC) pathway — CIN results from "gain of function"


mutations [5]. These may result either in activation of growth promoting pathways including
oncogenes or diminished activity of tumor suppressor genes or apoptotic pathways.

These tumors may be inherited (as with FAP) or sporadic, and they are characterized by
gross chromosomal abnormalities including deletions, insertions, and loss of heterozygosity.

The mutator phenotype/mismatch repair pathway — Another pathway leading from


adenomatous polyps, and probably from serrated adenomas, to invasive cancer has also
been described, the mismatch repair pathway. This pathway is involved in CRCs arising in
association with the inherited condition Lynch syndrome. (See "Lynch syndrome (hereditary
nonpolyposis colorectal cancer): Clinical manifestations and diagnosis", section on
'Genetics'.)

The key element of this pathway is dysfunction of DNA mismatch repair (MMR) enzymes,
resulting from germline mutations in one of several different DNA mismatch repair genes,
most commonly MLH1 or MSH2. Cells with deficient DNA repair capacity due to silencing of
MMR genes accumulate DNA errors throughout the genome. The biologic "footprint" is the
accumulation of abnormalities in short sequences of nucleotide bases that are repeated
dozens to hundreds of times within the genome; these are called microsatellites, and the
tumors are described as having the phenotype of high levels of microsatellite instability
(MSI-H).

Besides being the biologic hallmark of Lynch syndrome, high levels of MSI are also found in
approximately 15 percent of sporadic CRCs. However, in most of these cases, gene silencing
is not due to a specific MMR mutation, but to an epigenetic phenomenon, hypermethylation
of the gene promoter for the MMR enzyme (usually MLH1), which leads to transcriptional
silencing of gene expression [8]. (See 'Mismatch repair genes' below.)

Hypermethylation phenotype (CIMP+) pathway — Epigenetic alterations such as DNA


hypomethylation and loss of imprinting, as well as DNA hypermethylation, can silence the
expression of certain genes, including MMR enzymes [8-11]. CRCs that have a particularly
high frequency of methylation of some CpG islands (in which a cytosine [C] base is followed
immediately by a guanine [G] base that are linked with a phosphodiester bond [CpG]) are
referred to as CIMP+ tumors [8]. The defect may result in hypermethylation of the promoter
region of MMR enzymes such as MLH1 and silencing of gene expression [12].

Activating mutations in the BRAF gene (most of which are in the V600E codon) occur almost
exclusively in MSI-H, CIMP+ CRCs that do not carry mutations in KRAS [13,14]. Lynch-related

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&sel… 5/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

CRCs present only with KRAS and not BRAF mutations [13,14]. BRAF V600E mutations are
particularly prevalent in smokers with sporadic CRCs [15]. The presence of a BRAF V600E
mutation appears to abrogate the favorable prognosis that is typically associated with MSI-H
tumors [16,17]. Whether the adverse prognosis associated with V600E BRAF mutations is also
seen in individuals with non-V600 BRAF mutations is unclear. (See "Pathology and prognostic
determinants of colorectal cancer", section on 'RAS and BRAF'.)

Sporadic CRCs with a high degree of microsatellite instability and BRAF mutations are a
clinically distinct subgroup that is widely considered to develop from serrated polyps (
table 1) [7,18,19]. (See "Overview of colon polyps", section on 'Sessile serrated polyps and
traditional serrated adenomas'.)

SPECIFIC MOLECULAR ABNORMALITIES

The following sections will describe the major abnormalities in oncogenes, tumor
suppressor genes, mismatch repair (MMR) genes, and epigenetic phenomena such as DNA
hypomethylation and hypermethylation that underlie CRC tumorigenesis, including the
importance of each factor in cell cycle control and the molecular and clinical consequences
of individual mutations. This discussion is meant as an overview of the molecular
pathogenesis of CRC rather than an exhaustive survey, which can be found elsewhere [20].

Oncogenes — Oncogenes are homologs of normal cellular genes that participate in cell


growth pathways and cell cycle regulation. A mutational change in an oncogene leads to
constitutive activation of the gene, which then results in uncontrolled cellular proliferation
[21,22]. Because the normal gene function is activated, these are referred to as gain of
function mutations.

Among the oncogenes implicated in sporadic CRC are RAS, SRC, MYC, and the human
epidermal growth factor receptor 2 (HER2; previously called HER2/neu or ERBB-2); the most
important is RAS ( table 2) [23-27].

RAS — The RAS oncogene exists as three cellular variants, HRAS, KRAS, and NRAS. Although
all three oncogenes, when mutated, have the ability to transform normal cells, KRAS is the
most frequently mutated in human CRC [28-30]. The importance of RAS to colorectal
tumorigenesis is underscored by the finding that CRC cells in which a mutated RAS gene has
been removed or replaced lose their ability to form tumors in nude mice [31].

The RAS oncogenes encode a family of small proteins with homology to G-proteins that
regulate cellular signal transduction by acting as a one-way switch for the transmission of
extracellular growth signals to the nucleus [32]. These proteins normally cycle between an
inactive guanosine diphosphate (GDP)-bound state and an active guanosine triphosphate

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&sel… 6/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

(GTP)-bound state. RAS mutations, typically point mutations, leave the protein resistant to
GTP hydrolysis by GTPase, resulting in a constitutively active GTP-bound protein and a
continuous growth stimulus. Data from animal models suggest that RAS mutations may
contribute to colorectal tumorigenesis by activating cancer stem cells that have already been
activated by adenomatous polyposis coli (APC) mutations [33].

RAS mutations are found in up to 50 percent of sporadic CRCs and 50 percent of colonic
adenomas larger than 1 cm; they are rarely seen in smaller adenomas [28,34]. At least two
reports suggest that they are more common in proximal colon cancers than in more distal
colorectal primaries [35,36]. KRAS has also been implicated in the process of tumor invasion
and metastasis [37,38].

The lack of mutations in smaller adenomas suggests that RAS mutations are acquired during
later adenoma progression [39]. However, RAS mutations are not limited to dysplastic colonic
lesions. Up to 100 percent of nondysplastic aberrant crypt foci (ACF, believed to be the first
intermediate between normal colonic mucosa and the adenomatous polyp) and 25 percent
of hyperplastic polyps have RAS mutations, but their significance is unclear [28,31,32,40].

The identification of RAS mutations in CRC is of potential clinical relevance for both screening
and therapy:

● The detection of RAS mutations in fecal material is a potentially sensitive screening


method for the early diagnosis of CRC. Detection of KRAS mutations, in combination with
aberrantly methylated BMP3 and NDRG4 promoter areas, beta-actin, and a test for stool
hemoglobin are included in current multi-target stool DNA testing for colon cancer
(Cologuard) [41]. The utility of this test for CRC screening is discussed elsewhere. (See
"Tests for screening for colorectal cancer", section on 'Multitarget stool DNA tests with
fecal immunochemical testing'.)

● Posttranslational modification of the RAS protein by the enzyme farnesyl transferase is


necessary for activation, a fact that has been exploited for therapeutic purposes. The
therapeutic potential of agents that target the RAS signal transduction pathway (eg,
farnesyl transferase inhibitors) is being explored in patients with CRC whose tumors
contain RAS mutations.

● The presence of a RAS mutation in CRC is significantly associated with the absence of
response to agents targeting the epidermal growth factor receptor (EGFR) such as
cetuximab. (See "Systemic therapy for nonoperable metastatic colorectal cancer:
Selecting the initial therapeutic approach", section on 'RAS mutations'.)

Tumor suppressor genes — In contrast to oncogenes, tumor suppressor genes normally


have an inhibitory influence on the cell cycle. Once these genes are deleted or their function

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&sel… 7/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

reduced, normal control mechanisms are no longer operative, and growth proceeds
unchecked. At the cellular level, tumor suppressor genes act in a recessive fashion, meaning
that the function of the normal protein is lost only when both copies (alleles) of the gene are
inactivated by point mutations, rearrangements, or deletions.

Tumor suppressor genes were first described by Knudson in the context of childhood
retinoblastoma (RB), which is caused by mutational inactivation of the retinoblastoma (RB1)
gene, and presents either as hereditary or sporadic disease [42,43]. A "two-hit" model was
proposed to explain the different clinical features of these two presentations ( figure 3).
(See "Retinoblastoma: Clinical presentation, evaluation, and diagnosis", section on
'Pathogenesis'.)

● The inherited form of RB requires a germline mutation that can be either inherited or de
novo (ie, the result of a new germline mutation), plus a second somatic mutation,
occurring later in development, that affects the remaining RB1 allele within retinal cells
[22,42]. Affected individuals are at risk for multifocal and bilateral tumors.

● By contrast, in the nonhereditary form of RB, both allelic mutations arise spontaneously
in a single retinal cell (ie, both are somatic mutations). The resulting phenotype is that of
a unifocal, unilateral tumor that presents at a later age than the inherited variant and
without heritable transmission to later offspring [43].

The first molecular evidence for the involvement of tumor suppressor genes in CRC came
from the study of allelic loss, in which large chromosomal deletions were detected using
polymorphic markers that distinguish the two alleles present in the germline. When
comparing tumor alleles with those present in normal tissue, deletions were identified as
"loss of heterozygosity" (LOH).

In early studies of CRC, LOH for chromosomes 5q, 8p, 17p, or 18q was detected in 36, 50, 73,
and 75 percent of cases, respectively [34,44]. Presumptive tumor suppressor genes were
subsequently identified on 5q (the location of the APC gene), 18q (the location of the deleted
in colon cancer [DCC] gene, and the SMAD4 and SMAD2 genes), and 17p (location of the TP53
gene).

APC gene — Perhaps the most critical gene in the early development of CRC is the
adenomatous polyposis coli (APC) gene ( figure 4). Somatic mutations in both alleles are
present in 80 percent of sporadic CRCs, and a single germline mutation in this gene is
responsible for familial adenomatous polyposis (FAP), a dominantly inherited syndrome
characterized by the development of hundreds to thousands of colorectal polyps by the
second or third decade of life. (See "Clinical manifestations and diagnosis of familial
adenomatous polyposis", section on 'Genetics'.)

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&sel… 8/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

A germline APC mutation is also thought to contribute to the development of familial CRC in
Ashkenazi Jews [22,45,46]. A thymine to adenine transversion at nucleotide 3920 in the APC
gene that resulted in a substitution of lysine for isoleucine at codon 1307 (I1307K) is found in
6 percent of all persons of Ashkenazi Jewish descent but in a higher frequency of Ashkenazi
Jews with both a personal and family history of CRC (28 percent) [45]. This mutation was
previously thought to represent a polymorphism.

The recognition of the importance of the APC gene began with genetic studies linking
inheritance of the FAP syndrome to chromosome 5q21 and the subsequent identification of
germline mutations involving a gene at this locus, the APC gene. The earliest malignant
lesions in these patients, dysplastic aberrant crypt foci (microadenomas) and small
adenomatous polyps, have lost the second APC allele (through deletion or somatic
mutation), suggesting that loss of APC expression is a very early event in CRC tumorigenesis
[47,48].

The function of the APC gene product and the mechanism whereby the abnormal gene
promotes tumor formation are beginning to be understood. An important clue was the
observation that most sporadic CRCs with normal or wild-type APC had mutations in the
CTNNB1 gene, which encodes beta-catenin, a protein involved in the same signaling cascade
as the APC gene product, the Wnt (Wingless-type) signaling pathway [49,50]. It is now
hypothesized that most sporadic CRCs are initiated by activating Wnt pathway mutations,
characterized by the stabilization of beta-catenin and constitutive transcription by a beta-
catenin/T-cell factor (Tcf)-4 complex.

The Wnt pathway is an evolutionarily conserved signal transduction pathway that is


necessary for embryonic development [50-52]. It also plays a central role in supporting
intestinal epithelial renewal, an important fact since CRC is thought to originate in the
expansion of colonic crypt cells.

The basic features of the Wnt signaling pathway are depicted in the figure ( figure 5). The
normal APC protein appears to prevent the accumulation of cytosolic and nuclear beta-
catenin by mediating its phosphorylation and resultant degradation. The majority of
mutations in the APC gene (both germline and somatic) lead to premature truncation of the
APC protein and loss of its beta-catenin regulatory domains ( figure 4). Loss of functional
APC (as well as mutations in the beta-catenin gene) results in the nuclear accumulation of
beta-catenin, which binds and activates the transcription factor Tcf-4 [53-55].

It is proposed that beta-catenin/Tcf-4 acts as a switch controlling proliferation versus


differentiation in the intestinal crypt epithelial cells [56,57]. Activation of this pathway
prevents the cells from either entering G1 arrest or undergoing terminal differentiation and
induces resistance to apoptosis [58]. The end result is cellular proliferation. In addition,
because several other cell signaling pathways converge with the Wnt pathway, it represents
https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&sel… 9/41
14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

a "final common pathway" through which multiple abnormalities affecting other cellular
signaling pathways may result in the same carcinogenic result.

In addition to the typical loss of function tumor suppression mechanism of carcinogenesis,


more recent evidence suggests that C-terminally truncated APC may also have gain of
function properties; this mechanism of carcinogenesis promotion is typically attributed to
oncogenes and not tumor suppressor genes ( figure 4). Studies have reported that a
truncated APC gene promotes cell survival through regulation of the BCL2 gene [59,60],
stimulates cell migration through mediation of a guanine nucleotide exchange factor termed
Asef [61], and activates proliferation of human colonic epithelial cells in culture [62,63].

Other mechanisms may also contribute to the tumorigenic potential of APC mutations.
Mutations in APC (but not CTNNB1) are associated with chromosomal instability [64],
predisposing the cell to "hits" in other genes that may contribute to tumor progression and
malignant transformation.

TP53 gene — The TP53 gene on chromosome 17p is the most commonly mutated gene in
human cancer. In approximately 50 to 70 percent of CRCs, TP53 inactivation occurs by a
mutation of one allele followed by loss of the remaining wild type gene [34,65-68]. 17p
sequences are lost in as many as 75 percent of CRCs, while they are rarely lost in adenomas
and aberrant crypt foci, suggesting that loss of p53 function represents a relatively late
event in colorectal tumorigenesis [5,34,69]. In keeping with this hypothesis, a large
international study of 3583 CRCs found an increase in the frequency of TP53 mutations with
advancing disease stage [68].

The normal "wild-type" TP53 gene produces a DNA-binding protein p53 that acts as a
transcriptional activator of growth inhibitory genes. Wild-type p53 may be particularly critical
when cells are under stress. Normally, cells arrest their growth in response to DNA
damaging agents and other stressors (eg, hypoxia) via induction/activation of p53 [70-72].
Once activated, p53 induces a variety of growth-limiting responses, including cell cycle arrest
(in order to facilitate DNA repair), apoptosis, senescence, and differentiation. p53 produces
these responses largely by altering the expression of a number of target genes, at least 20 of
which have been described as being under transcriptional control of p53. Because of its
central role in preventing the propagation of cells with DNA damage, p53 has been referred
to as the "guardian of the genome" [73].

Although inactivation of the p53 pathway appears to be a late event in the majority of
human CRCs, it may represent an earlier event in inflammatory bowel disease-related CRC.
(See "Surveillance and management of dysplasia in patients with inflammatory bowel
disease", section on 'Molecular pathogenesis'.)

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 10/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

Given the association of p53 pathway inactivation with CRC, it is unclear why patients with
the Li Fraumeni syndrome (a condition caused by a germline mutation in TP53 in which
patients frequently develop carcinomas, sarcomas, and leukemias) are not at particularly
increased risk of developing CRCs. However, germline TP53 mutations may be associated
with early onset CRC. In a report derived from the population- and clinic-based Colon Cancer
Family Registry, germline TP53 mutations were identified in 6 of 457 individuals (1.3 percent)
diagnosed with CRC at age 40 or younger [74]. This frequency is roughly comparable to the
prevalence of germline APC mutations in CRC. (See "Li-Fraumeni syndrome".)

The identification of TP53 mutations in an individual CRC is of potential clinical significance,


prognostically and therapeutically. In many but not all studies, patients whose tumors
harbor TP53 mutations have worse outcomes and shorter survival than those without such
mutations. At least some of the discordant results may be due to the fact that the prognostic
influence of TP53 abnormalities appears to depend on tumor site, type of mutation, and the
use of adjuvant therapy [68]. This topic is discussed in detail elsewhere. (See "Pathology and
prognostic determinants of colorectal cancer", section on 'Molecular factors'.)

From a therapeutic standpoint, it is hoped that p53 may prove to be a highly selective and
effective target for intervention. Several new therapies under study for advanced disease
specifically target TP53-mutant cells, while others seek to correct the TP53 mutations directly
or to restore the integrity of the p53 pathway [60,75-77].

Chromosome 18q: The DCC, SMAD4, and SMAD2 genes — As with the TP53 and APC
genes, the first evidence of a tumor suppressor gene on chromosome 18q came from
studies of allelic loss in CRC. In an early study, one copy of 18q was lost in 73 percent of
sporadic CRCs and 47 percent of large adenomas with foci of invasive cancer, but in fewer
than 15 percent of less advanced adenomas [34]. In 1989, a candidate gene termed the
"deleted in colorectal carcinoma" (DCC) gene was identified at 18q21 [78], and point
mutations in the DCC gene have been identified in CRCs [79,80].

Gene mutations presumably lead to a loss of expression of the DCC protein, which is thought
to have a role in cell-cell or cell-matrix interactions [78,81,82]. DCC is normally expressed in
many tissues, including the colonic mucosa, although its normal function has been difficult
to elucidate because of its large size and the lack of expression in CRCs [81].

Loss of DCC expression may have prognostic value, particularly in patients with early stage
CRC. Five-year survival rates seem to be worse for patients with stage II (node-negative (
table 3)) CRCs that lack DCC expression compared with those that express it [83]. For
patients with DCC-negative stage II disease, prognosis more closely approximates that of
patients with more advanced stage III (node-positive) disease [84].

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 11/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

The ultimate benefit of this information may be the identification of a subgroup of patients
with stage II colon cancer who might benefit from adjuvant chemotherapy. While appealing,
there are no prospective data that currently support the validity of this strategy. (See
"Adjuvant therapy for resected stage III (node-positive) colon cancer".)

A second tumor suppressor gene at 18q was identified during the course of investigation of
allelic losses in pancreatic cancer, termed the DPC4 (deleted in pancreatic cancer) gene, now
redesignated SMAD4 [34]. (See "Molecular pathogenesis of exocrine pancreatic cancer",
section on 'Tumor suppressor genes'.)

The SMAD4 gene encodes a protein that may be important to the signaling pathway of the
transforming growth factor beta (TGF-beta) superfamily of signaling polypeptides. TGF-beta
suppresses the growth of most normal cells by binding to type I (TGFBR1) and type II
(TGFBR2) transmembrane receptors, but many cancer cells are resistant to this growth-
suppressive effect. In CRC cells, SMAD4 is required for TGF-beta signaling, and at least in
vitro, reintroduction of the gene (via transfection of an intact chromosome 18) is associated
with restoration of TGF-beta sensitivity [85,86].

Mutations in SMAD4 or a third putative tumor suppressor gene that also maps to 18q
(SMAD2) have been found in a subset of sporadic CRCs [63,87-89]. Perhaps more importantly,
germline mutations in SMAD4 (and in BMPR1A [ALK3], a gene that codes for a member of the
TGF-beta receptor superfamily, which is located upstream from SMAD4 [90]) have also been
identified in patients with juvenile polyposis. These patients develop multiple juvenile polyps
that are distinct from adenomas, and they are at an increased risk for invasive CRCs. (See
"Juvenile polyposis syndrome", section on 'Genetics'.)

TGF-beta signaling — As noted above, one mechanism by which CRC cells escape the
normal inhibitory influence of transforming growth factor beta (TGF-beta) is through SMAD4
mutations, which interfere with the production of a protein that is required for TGF-beta
signaling. Other possible mechanisms of interference with normal TGF-beta signaling in CRC
cells have also been identified, including inactivating mutations in the TGF-beta receptor,
type II (TGFBR2) gene [91] and molecular changes that result in the redirection of TGF-beta
growth inhibitory signals into growth stimulatory signals.

BRCA1 and BRCA2 genes — Breast cancer associated (BRCA) 1 and 2 are tumor
suppressor genes known to function in homologous recombination [92,93]. BRCA1 operates
in both DNA repair and checkpoint activation, and it has a role that is upstream to that of
BRCA2, which functions in the primary mechanism of homologous recombination. Germline
mutations in BRCA1 and BRCA2 are predominantly associated with breast and ovarian
cancers. BRCA2 and, to a lesser extent, BRCA1 have been associated with pancreatic cancers,
prostate cancers, gastric cancers, and CRCs. Three specific founder mutations in BRCA1/2
have been shown to be frequent in individuals of Ashkenazi descent. (See "Cancer risks and
https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 12/41
14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

management of BRCA1/2 carriers without cancer", section on 'Cancer risks in BRCA1/2


carriers' and "Genetic testing and management of individuals at risk of hereditary breast and
ovarian cancer syndromes", section on 'Population-based testing for those of Ashkenazi
Jewish descent'.)

Although the data regarding the risk of CRC in BRCA carriers have been inconsistent, a
prospective study of BRCA1 mutation carriers suggested an increased risk of CRC in women
age 30 to 49, which returned to population-level risk after age 50; no increased risk was
noted for BRCA2 mutation carriers [94]. Similarly, a meta-analysis of 18 cohort and case-
control studies suggested that carriers of BRCA1 (but not BRCA2) have a higher risk for CRC
(odds ratio [OR] 1.49, 95% CI 1.14-1.85) [94-96].

It is often difficult to determine genotype/phenotype causation, as opposed to association,


unless molecular genetic analysis is undertaken [97]. A conclusion to support BRCA1/2
causation, as opposed to association, with CRC, as of yet, has no supportive molecular data.
This issue will become more common as multigene genetic testing becomes predominant.
One study of multigene panel testing in individuals suspected of having Lynch syndrome by
National Comprehensive Cancer Network (NCCN) criteria revealed that 5.6 percent of
individuals (71 of 1260) had non-Lynch cancer predisposition germline mutations [98,99].
Fifteen of these individuals had BRCA1/2 mutations; 93 percent met NCCN criteria for Lynch
syndrome genetic testing, yet only 33 percent met NCCN criteria for BRCA1/2 genetic testing.

A decision regarding referral for genetic testing in individuals with CRC should take cancer-
related risk factors into account, including personal history of cancer, familial history of
cancer, and any prior genetic testing of family members. (See "Genetic testing and
management of individuals at risk of hereditary breast and ovarian cancer syndromes",
section on 'Criteria for genetic risk evaluation'.)

Given the lack of information on the age of onset of CRC in at-risk individuals, it is difficult to
make recommendations about whether or not BRCA1 mutation carriers should initiate CRC
screening at an earlier age than is typically recommended. This subject is discussed in detail
elsewhere. (See "Colorectal cancer: Epidemiology, risk factors, and protective factors",
section on 'Hereditary CRC syndromes'.)

Mismatch repair genes — MMR genes are responsible for correcting the ubiquitous
nucleotide base mispairings and small insertions or deletions that occur during DNA
replication [100-103]. Several of these genes exist, including hMSH2 (human mutS homolog
2), hMLH1 (human mutL homolog 1), hPMS1 and hPMS2 (human postmeiotic segregation 1
and 2), hMSH6 (human mutS homolog 6), and hMLH3, an MMR gene that interacts with
MLH1.

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 13/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

Germline mutations in one of the MMR genes appear to be the underlying genetic defect in
most kindreds with Lynch syndrome, and loss of expression of MMR genes can also be found
in approximately 15 percent of sporadic CRCs [104-106]. The biologic footprint of tumors
that have deficient MMR is a high level of microsatellite instability (MSI-H). (See
'Microsatellite instability high versus low' below.)

However, sporadic tumors with defective expression of MMR genes do not contain MMR
gene mutations; instead, they have epigenetic changes (acquired hypermethylation of the
promoters of both alleles of MLH1 gene) that silence gene expression. (See 'Epigenetic
alterations affecting mismatch repair genes' below.)

Another group is described as having "Lynch-like" syndrome; these are patients who have
MSI-H tumors and loss of expression of one or more of the MMR genes but have neither a
germline MMR mutation or promoter hypermethylation of the MLH1 gene. At least some of
these patients have biallelic somatically acquired MMR gene mutations [107,108]. The risk of
CRC appears to be lower in these families than it is in families with Lynch syndrome but
higher than that of families with sporadic CRC [109]. (See "Lynch syndrome (hereditary
nonpolyposis colorectal cancer): Clinical manifestations and diagnosis", section on
'Differential diagnosis'.)

These patients are not to be confused with another subgroup of patients who meet clinical
criteria for Lynch syndrome (eg, the Amsterdam criteria ( table 4)) but do not have a
discernible mutation in one of the MMR genes (ie, they are MMR-proficient). This subset of
patients is sometimes referred to as familial colorectal cancer syndrome type X (FCCTX), to
distinguish them from Lynch syndrome [3,110-112]. In contrast to Lynch syndrome, these
patients do not carry a risk for extracolonic tumors, and the histopathologic features that
characterize Lynch-related colon cancers are absent [112,113]. (See "Lynch syndrome
(hereditary nonpolyposis colorectal cancer): Clinical manifestations and diagnosis", section
on 'Differential diagnosis'.)

Although FCCTX is termed "familial," this may be a misnomer, as these patients typically
present as isolated cases, likely caused by an as-of-yet undefined germline mutation. At least
some data suggest that gain of genetic material on chromosome 20q and loss on
chromosome 18q may serve to discriminate between CRCs associated with FCCTX and Lynch
syndrome [114,115].

Testing strategies for Lynch syndrome in patients with CRC are discussed in detail elsewhere.
(See "Lynch syndrome (hereditary nonpolyposis colorectal cancer): Clinical manifestations
and diagnosis", section on 'Identification of individuals at risk for Lynch syndrome'.)

Microsatellite instability high versus low — Cells that are MMR deficient accumulate
DNA errors throughout the genome [116]. The biologic "footprint" of an MMR defect is the

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 14/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

accumulation of abnormalities in short sequences of nucleotide bases that are repeated


dozens to hundreds of times within the genome; these are called microsatellites [116].
Several critical growth regulatory genes (eg, the TGFRB2, BAX, insulin-like growth factor 2
receptor [IGF2R]) contain microsatellites in the promoter region and are therefore
susceptible to frameshift mutations. This leaves the cell vulnerable to mutations in these
genes controlling cell growth, and this phenomenon is termed MSI [91,116,117].

Many but not all tumors that contain MMR mutations can be identified by the presence of
MSI-H. The majority of patients with Lynch syndrome have MSI-H tumors. In addition,
approximately 15 percent of sporadic tumors are MSI-H. (See "Lynch syndrome (hereditary
nonpolyposis colorectal cancer): Clinical manifestations and diagnosis".)

In contrast to microsatellite-stable CRCs, sporadic tumors with MSI-H have characteristic


clinicopathologic features. They tend to occur in the proximal colon, have a greater
mucinous component, contain lymphocytic infiltration, and are more often poorly
differentiated. Interestingly, the tendency to have a lymphocytic infiltrate likely reflects
immune activation from T-cells directed against tumor-specific carboxy-terminal frameshift
peptides that are associated with microsatellite instability [118].

Although tumors in Lynch syndrome tend to be poorly differentiated, the presence of MSI
mitigates the adverse prognostic impact of this feature, and in fact, MSI-H tumors are
associated with longer survival in both Lynch syndrome and sporadic cases, for unclear
reasons. This topic is discussed in detail elsewhere. (See "Pathology and prognostic
determinants of colorectal cancer", section on 'Histologic type, grade of differentiation, and
presence of mucin' and "Pathology and prognostic determinants of colorectal cancer",
section on 'Mismatch repair deficiency'.)

Most laboratories use a panel of several microsatellite loci when testing for MSI [119]. A
panel consisting of three dinucleotide repeats and two mononucleotide repeats has been
proposed as a standard test for MSI; a tumor is called MSI-H when at least two (40 percent)
are affected by instability [104]. While most tumors show either a high degree of instability
or no unstable markers, a minority display instability in <40 percent of the markers studied.
These tumors are referred to as MSI-low (MSI-L). While almost all MSI-H tumors are MMR
deficient, most MSI-L tumors have no MMR defect [120,121]. (See "Lynch syndrome
(hereditary nonpolyposis colorectal cancer): Clinical manifestations and diagnosis", section
on 'Genetics'.)

The significance of the MSI-L phenotype is incompletely understood [120,122,123]. It is not


associated with silencing of MLH1 or any of the other known DNA MMR genes, but there is
evidence to suggest that MSI-L is a nonrandom phenomenon with a biologic basis [123-125]:

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 15/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

● Some data suggest that MSI-L in sporadic CRC reflects loss of the MutS Homologue 3
(MSH3) gene [123].

● The MSI-L phenotype has also been associated with methylation of the promoter for the
DNA repair gene O-6 methylguanine DNA methyltransferase (MGMT) [126]. It has been
hypothesized that loss of expression of the MGMT gene results in accumulation of
methyl G:T mismatches and excess stress on the MMR system, which ultimately leads to
MSI-L [126,127]. While rare in Lynch syndrome, MGMT promoter methylation or loss of
MGMT gene expression occurs in up to 25 percent of sessile serrated adenomas, 78
percent of dysplastic serrated adenomas, and 50 percent of serrated adenocarcinomas
[127,128]. These findings suggest that the factors leading to the MSI-L phenotype may
be important in the serrated neoplasia/CIMP pathway ( table 1).

In any case, accumulating evidence suggests that the MSI-L phenotype is associated with a
poor clinical outcome in both stage II and III CRC [123,129-131].

Epigenetic alterations affecting mismatch repair genes — As noted above, mutations


and allelic loss of one of the MMR genes are responsible for the MSI-H phenotype in most
cases of Lynch syndrome. In contrast, hypermethylation of the promoter region of some
MMR genes and/or DNA hypomethylation with loss of imprinting (ie, silencing of gene
expression) is thought to underlie cases of sporadic CRC that display the MSI-H phenotype
[100,132-137]. Epigenetic inactivation of the second normal MMR gene allele may also play a
role in individuals with Lynch syndrome, in whom the second allele must be inactivated in
order to progress to cancer [138,139]:

● DNA hypermethylation specifically targets CpG dinucleotides, which are present in the
promoters of many genes (including the MMR gene hMLH1). Although the stimulus that
drives hypermethylation remains unknown, methylated CpG is bound by a family of
proteins known as methyl-CpG binding domain proteins. These proteins, in turn, form a
multiprotein complex that alters chromatin conformation and silences expression [140-
142]. (See 'Hypermethylation phenotype (CIMP+) pathway' above.)

● Another mechanism of epigenetic inactivation is DNA hypomethylation and loss of


imprinting (LOI) [11]. Imprinting refers to the selective loss of expression of parent
lineage-specific genes due to selective methylation of one allele. (See "Inheritance
patterns of monogenic disorders (Mendelian and non-Mendelian)", section on 'Parent-
of-origin effects (imprinting)'.)

This pattern is set in the zygote and maintained during development to suppress the
expression of the maternal or paternal copy of an allele. LOI for the insulin-like growth
factor 2 gene (IGF-2) has been highly associated with MMR-deficient CRC [136,137,143].
However, LOI for IGF2 is found more often in the normal colonic mucosa and peripheral

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 16/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

blood lymphocytes of patients with CRC compared with those without the disease
[136,144]. These findings have led some to hypothesize that LOI for IGF2 represents a
risk factor for CRC rather than a somatic defect underlying tumorigenesis [144].

● Large deletions in the 3' end of the epithelial cell adhesion molecule (EPCAM) gene leads
to hypermethylation of the adjacent MSH2 gene on chromosome location 2p21, which
has the effect of silencing expression of the MSH2 gene [145-147]. Up to 6 percent of
Lynch patients have an EPCAM epigenetic mutation. (See "Lynch syndrome (hereditary
nonpolyposis colorectal cancer): Clinical manifestations and diagnosis", section on
'Genetics'.)

MUTYH defects and familial colorectal cancer — A small proportion of patients with
multiple colorectal adenomas and a family history of CRC have germline mutations (often
biallelic) in the base excision repair gene mutY homolog (MUTYH), sometimes in conjunction
with somatic mutations in the APC gene [148-150]. These mutations predispose patients to
recessive inheritance of multiple colonic adenomas and a variant form of FAP, frequently
referred to as MUTYH-associated polyposis (MAP). In one series of 152 patients with multiple
adenomas seen at one institution, 7.5 percent of those without a germline APC mutation
were found to have two separate germline MUTYH mutations [148]. These findings have
implications for screening strategies in patients suspected of having FAP, which in most
cases is inherited in an autosomal dominant pattern. (See 'APC gene' above.)

Perhaps more importantly, an increasing number of reports suggest that germline MUTYH
mutations may account for a substantial fraction of familial CRCs that occur in the absence
of a dominantly inherited familial syndrome [150-155]. In a population-based series that
screened 1238 patients diagnosed with CRC over a three-year period and 1255 healthy age-
and gender-matched control subjects without a personal history of cancer for germline
MUTYH mutations (Y165C and G382D), mutation carriers were significantly more likely to
develop CRC and were more likely to have first- or second-degree relatives with CRC [152].

The phenotype is variable, ranging from 10 or more polyps to widespread polyposis with
associated CRC. The risk of CRC among carriers of biallelic mutations appears to be high; it
was close to 100 percent by age 60 in one study [156], and 43 percent in another [157]. Some
data suggest that heterozygotes have a slightly increased risk of CRC [152,158,159], while
other series report no increased risk in those with monoallelic mutations [157].

The prognosis may be more favorable in CRCs that develop in the setting of MAP compared
with those that arise in the general population [160]. In a retrospective European cohort
study comparing 147 patients with MAP-related CRC and 272 population-based matched
control patients with CRC, survival was significantly better for the patients with MAP-
associated CRC, even after adjustment for differences in age, stage, sex, subsite, country,

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 17/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

and year of diagnosis (hazard ratio [HR] for death 0.48, 95% CI 0.32 to 0.72). Though
compelling, these findings need to be replicated in independent, prospective studies.

Modifier genes — In addition to the genes described, several other genes seem to be
important in colorectal carcinogenesis, although their exact roles and mechanisms of
tumorigenesis have not been fully determined.

COX-2 — A substantial body of evidence supports a protective effect of aspirin and other
cyclooxygenase (COX) inhibitors on the development of CRC. Furthermore, one nonspecific
COX inhibitor, sulindac, can cause polyp regression in patients with FAP. The mechanism
underlying these effects is not well understood, but they suggest a role for the COX-2 gene,
which is upregulated in CRC cells, in colorectal tumorigenesis. This is an area of active
investigation and is discussed in detail elsewhere. (See "NSAIDs (including aspirin): Role in
prevention of colorectal cancer".)

PPAR gene — The peroxisome proliferator-activating receptor (PPAR) gene has been


implicated in colorectal carcinogenesis. The PPAR gene encodes a family of nuclear receptors
that function as transcriptional regulators for proteins controlling lipid metabolism and cell
growth. Activation of these receptors inhibits cell growth and promotes differentiation in a
variety of epithelial cell types, including CRC cells [161]. Preliminary studies demonstrate that
PPAR gene is downstream from the APC gene and may also be involved in the COX pathway
[162-165].

Loss of function mutations in PPAR has been described in sporadic CRCs [166]. Furthermore,
there is some evidence that abnormalities in the PPAR genes are responsible for the
increased frequency of CRCs and adenomas in patients with acromegaly. (See "Colorectal
cancer: Epidemiology, risk factors, and protective factors", section on 'Other risk factors'.)

SUMMARY

● Colorectal cancer (CRC) represents an ideal model for the study of the molecular
pathogenesis of cancer due to the accessibility of tissue for biopsy and the clear
progression from normal colonic epithelium to invasive cancer via an intermediate
precursor, the adenomatous polyp. A multistep process of specific genetic changes is
thought to drive the transformation from normal colonic epithelium to invasive cancer.
Single, specific germline mutations underlie the common inherited syndromes (eg,
adenomatous polyposis coli [APC], Lynch syndrome), while sporadic cancers result from
the stepwise accumulation of multiple somatic mutations. Mutations in the APC gene
occur early, while others, such as mutations of the TP53 suppressor gene, generally
occur late in the process ( figure 2). (See 'Molecular pathogenesis of colorectal cancer'
above.)

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 18/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

The concept of microsatellite instability (MSI) provides an additional layer of complexity


to this model. The MSI-high (MSI-H) phenotype is associated with Lynch syndrome and is
also observed in approximately 10 to 15 percent of sporadic CRCs. (See 'The mutator
phenotype/mismatch repair pathway' above.)

● The available data suggest that most colorectal neoplasias begin with inactivation
(through a germline or sporadic mutation) of the APC gene, whether or not MSI is
present. However, the genetic events may thereafter diverge, depending on the
mechanism underlying the genetic instability (ie, chromosomal instability, germline
mutations in DNA mismatch repair [MMR] enzymes, CpG island hypermethylation
phenotype [CIMP+]). This has led to the molecular classification of CRC into three
carcinogenic pathways ( table 1) [7]. (See 'Molecular pathways to colorectal
tumorigenesis' above.)

Compared with patients with chromosomal instability (which includes those with familial
adenomatous polyposis [FAP] and the majority of sporadic CRCs), the MSI-H or
"mutator" pathway involves a unique destabilizing mechanism and inactivation of a
different set of genes (which encode the DNA MMR proteins). The CIMP+ pathway
involves hypermethylation of the gene promoter for a DNA MMR gene. Regardless of
the molecular pathway involved, the final result is the same.

● In addition to providing insight as to the biology of all CRCs, the identification of specific
genetic mutations responsible for colorectal tumorigenesis has had a direct influence on
clinical care. Patients at highest risk for developing CRC can be identified via genetic
testing for specific germline mutations. (See 'Tumor suppressor genes' above and
'Mismatch repair genes' above and 'MUTYH defects and familial colorectal cancer'
above.)

Furthermore, new molecular screening methods for early detection of CRC via detection
of mutations in fecal material are under study. In addition, these mutations are also
being examined as prognostic markers and as potential therapeutic targets. (See 'RAS'
above.)

ACKNOWLEDGMENT

The editorial staff at UpToDate would like to acknowledge Paula M Calvert, MD, who
contributed to an earlier version of this topic review.

Use of UpToDate is subject to the Subscription and License Agreement.

REFERENCES
https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 19/41
14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

1. Siegel RL, Miller KD, Fuchs HE, Jemal A. Cancer Statistics, 2021. CA Cancer J Clin 2021;
71:7.

2. Wirtzfeld DA, Petrelli NJ, Rodriguez-Bigas MA. Hamartomatous polyposis syndromes:


molecular genetics, neoplastic risk, and surveillance recommendations. Ann Surg Oncol
2001; 8:319.

3. Lindor NM. Familial colorectal cancer type X: the other half of hereditary nonpolyposis
colon cancer syndrome. Surg Oncol Clin N Am 2009; 18:637.
4. Choi YH, Lakhal-Chaieb L, Kröl A, et al. Risks of Colorectal Cancer and Cancer-Related
Mortality in Familial Colorectal Cancer Type X and Lynch Syndrome Families. J Natl
Cancer Inst 2019; 111:675.
5. Fearon ER, Vogelstein B. A genetic model for colorectal tumorigenesis. Cell 1990; 61:759.

6. Zauber AG, Winawer SJ, O'Brien MJ, et al. Colonoscopic polypectomy and long-term
prevention of colorectal-cancer deaths. N Engl J Med 2012; 366:687.

7. Noffsinger AE. Serrated polyps and colorectal cancer: new pathway to malignancy. Annu
Rev Pathol 2009; 4:343.

8. Weisenberger DJ, Siegmund KD, Campan M, et al. CpG island methylator phenotype
underlies sporadic microsatellite instability and is tightly associated with BRAF mutation
in colorectal cancer. Nat Genet 2006; 38:787.

9. Das PM, Singal R. DNA methylation and cancer. J Clin Oncol 2004; 22:4632.

10. Shen L, Kondo Y, Rosner GL, et al. MGMT promoter methylation and field defect in
sporadic colorectal cancer. J Natl Cancer Inst 2005; 97:1330.
11. van Engeland M, Derks S, Smits KM, et al. Colorectal cancer epigenetics: complex
simplicity. J Clin Oncol 2011; 29:1382.

12. Goel A, Nagasaka T, Arnold CN, et al. The CpG island methylator phenotype and
chromosomal instability are inversely correlated in sporadic colorectal cancer.
Gastroenterology 2007; 132:127.

13. Rajagopalan H, Bardelli A, Lengauer C, et al. Tumorigenesis: RAF/RAS oncogenes and


mismatch-repair status. Nature 2002; 418:934.
14. Domingo E, Niessen RC, Oliveira C, et al. BRAF-V600E is not involved in the colorectal
tumorigenesis of HNPCC in patients with functional MLH1 and MSH2 genes. Oncogene
2005; 24:3995.
15. Samowitz WS, Albertsen H, Sweeney C, et al. Association of smoking, CpG island
methylator phenotype, and V600E BRAF mutations in colon cancer. J Natl Cancer Inst
2006; 98:1731.

16. French AJ, Sargent DJ, Burgart LJ, et al. Prognostic significance of defective mismatch
repair and BRAF V600E in patients with colon cancer. Clin Cancer Res 2008; 14:3408.

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 20/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

17. Ogino S, Nosho K, Kirkner GJ, et al. CpG island methylator phenotype, microsatellite
instability, BRAF mutation and clinical outcome in colon cancer. Gut 2009; 58:90.
18. Spring KJ, Zhao ZZ, Karamatic R, et al. High prevalence of sessile serrated adenomas
with BRAF mutations: a prospective study of patients undergoing colonoscopy.
Gastroenterology 2006; 131:1400.

19. Chan TL, Zhao W, Leung SY, et al. BRAF and KRAS mutations in colorectal hyperplastic
polyps and serrated adenomas. Cancer Res 2003; 63:4878.

20. Lynch JP, Hoops TC. The genetic pathogenesis of colorectal cancer. Hematol Oncol Clin
North Am 2002; 16:775.
21. Sherr CJ. Cancer cell cycles. Science 1996; 274:1672.

22. Cancer genetics and cancer predisposition testing. In: American Society of Clinical Oncol
ogy Curriculum, American Society of Clinical Oncology (Ed), Alexandria 1998.
23. Forgacs I. Oncogenes and gastrointestinal cancer. Gut 1988; 29:417.

24. Cartwright C. Intestinal cell growth control: role of Src tyrosine kinases.
Gastroenterology 1998; 114:1335.

25. Hamilton SR. The molecular genetics of colorectal neoplasia. Gastroenterology 1993;
105:3.

26. Kapitanović S, Radosević S, Kapitanović M, et al. The expression of p185(HER-2/neu)


correlates with the stage of disease and survival in colorectal cancer. Gastroenterology
1997; 112:1103.

27. Irby RB, Mao W, Coppola D, et al. Activating SRC mutation in a subset of advanced
human colon cancers. Nat Genet 1999; 21:187.

28. Takayama T, Ohi M, Hayashi T, et al. Analysis of K-ras, APC, and beta-catenin in aberrant
crypt foci in sporadic adenoma, cancer, and familial adenomatous polyposis.
Gastroenterology 2001; 121:599.

29. Shibata D, Schaeffer J, Li ZH, et al. Genetic heterogeneity of the c-K-ras locus in
colorectal adenomas but not in adenocarcinomas. J Natl Cancer Inst 1993; 85:1058.

30. Tortola S, Marcuello E, González I, et al. p53 and K-ras gene mutations correlate with
tumor aggressiveness but are not of routine prognostic value in colorectal cancer. J Clin
Oncol 1999; 17:1375.
31. Shirasawa S, Furuse M, Yokoyama N, Sasazuki T. Altered growth of human colon cancer
cell lines disrupted at activated Ki-ras. Science 1993; 260:85.
32. Bourne HR, Sanders DA, McCormick F. The GTPase superfamily: conserved structure and
molecular mechanism. Nature 1991; 349:117.

33. Moon BS, Jeong WJ, Park J, et al. Role of oncogenic K-Ras in cancer stem cell activation
by aberrant Wnt/β-catenin signaling. J Natl Cancer Inst 2014; 106:djt373.

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 21/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

34. Vogelstein B, Fearon ER, Hamilton SR, et al. Genetic alterations during colorectal-tumor
development. N Engl J Med 1988; 319:525.

35. Frattini M, Balestra D, Suardi S, et al. Different genetic features associated with colon
and rectal carcinogenesis. Clin Cancer Res 2004; 10:4015.

36. Harada K, Hiraoka S, Kato J, et al. Genetic and epigenetic alterations of Ras signalling
pathway in colorectal neoplasia: analysis based on tumour clinicopathological features.
Br J Cancer 2007; 97:1425.
37. Giehl K. Oncogenic Ras in tumour progression and metastasis. Biol Chem 2005; 386:193.
38. Miranda E, Destro A, Malesci A, et al. Genetic and epigenetic changes in primary
metastatic and nonmetastatic colorectal cancer. Br J Cancer 2006; 95:1101.
39. Pretlow TP, Brasitus TA, Fulton NC, et al. K-ras mutations in putative preneoplastic
lesions in human colon. J Natl Cancer Inst 1993; 85:2004.

40. Losi L, Roncucci L, di Gregorio C, et al. K-ras and p53 mutations in human colorectal
aberrant crypt foci. J Pathol 1996; 178:259.

41. Imperiale TF, Ransohoff DF, Itzkowitz SH, et al. Fecal DNA versus fecal occult blood for
colorectal-cancer screening in an average-risk population. N Engl J Med 2004; 351:2704.
42. Knudson AG Jr. Mutation and cancer: statistical study of retinoblastoma. Proc Natl Acad
Sci U S A 1971; 68:820.
43. Knudson AG Jr. Hereditary cancer, oncogenes, and antioncogenes. Cancer Res 1985;
45:1437.
44. Vogelstein B, Fearon ER, Kern SE, et al. Allelotype of colorectal carcinomas. Science 1989;
244:207.

45. Laken SJ, Petersen GM, Gruber SB, et al. Familial colorectal cancer in Ashkenazim due to
a hypermutable tract in APC. Nat Genet 1997; 17:79.

46. Drucker L, Shpilberg O, Neumann A, et al. Adenomatous polyposis coli I1307K mutation
in Jewish patients with different ethnicity: prevalence and phenotype. Cancer 2000;
88:755.
47. Spirio LN, Samowitz W, Robertson J, et al. Alleles of APC modulate the frequency and
classes of mutations that lead to colon polyps. Nat Genet 1998; 20:385.
48. Lamlum H, Ilyas M, Rowan A, et al. The type of somatic mutation at APC in familial
adenomatous polyposis is determined by the site of the germline mutation: a new facet
to Knudson's 'two-hit' hypothesis. Nat Med 1999; 5:1071.
49. Su LK, Vogelstein B, Kinzler KW. Association of the APC tumor suppressor protein with
catenins. Science 1993; 262:1734.
50. Bienz M, Clevers H. Linking colorectal cancer to Wnt signaling. Cell 2000; 103:311.

51. Fearnhead NS, Britton MP, Bodmer WF. The ABC of APC. Hum Mol Genet 2001; 10:721.
https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 22/41
14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

52. Uthoff SM, Eichenberger MR, McAuliffe TL, et al. Wingless-type frizzled protein receptor
signaling and its putative role in human colon cancer. Mol Carcinog 2001; 31:56.
53. Korinek V, Barker N, Morin PJ, et al. Constitutive transcriptional activation by a beta-
catenin-Tcf complex in APC-/- colon carcinoma. Science 1997; 275:1784.

54. Morin PJ, Sparks AB, Korinek V, et al. Activation of beta-catenin-Tcf signaling in colon
cancer by mutations in beta-catenin or APC. Science 1997; 275:1787.

55. Goss KH, Groden J. Biology of the adenomatous polyposis coli tumor suppressor. J Clin
Oncol 2000; 18:1967.
56. van de Wetering M, Sancho E, Verweij C, et al. The beta-catenin/TCF-4 complex imposes
a crypt progenitor phenotype on colorectal cancer cells. Cell 2002; 111:241.
57. Shih IM, Wang TL, Traverso G, et al. Top-down morphogenesis of colorectal tumors. Proc
Natl Acad Sci U S A 2001; 98:2640.
58. Kim PJ, Plescia J, Clevers H, et al. Survivin and molecular pathogenesis of colorectal
cancer. Lancet 2003; 362:205.

59. Brocardo M, Lei Y, Tighe A, et al. Mitochondrial targeting of adenomatous polyposis coli
protein is stimulated by truncating cancer mutations: regulation of Bcl-2 and
implications for cell survival. J Biol Chem 2008; 283:5950.

60. Watanabe T, Sullenger BA. Induction of wild-type p53 activity in human cancer cells by
ribozymes that repair mutant p53 transcripts. Proc Natl Acad Sci U S A 2000; 97:8490.
61. Kawasaki Y, Senda T, Ishidate T, et al. Asef, a link between the tumor suppressor APC
and G-protein signaling. Science 2000; 289:1194.
62. Zhang L, Theodoropoulos PC, Eskiocak U, et al. Selective targeting of mutant
adenomatous polyposis coli (APC) in colorectal cancer. Sci Transl Med 2016; 8:361ra140.
63. MacGrogan D, Pegram M, Slamon D, Bookstein R. Comparative mutational analysis of
DPC4 (Smad4) in prostatic and colorectal carcinomas. Oncogene 1997; 15:1111.

64. Fodde R, Kuipers J, Rosenberg C, et al. Mutations in the APC tumour suppressor gene
cause chromosomal instability. Nat Cell Biol 2001; 3:433.

65. Soussi T. The p53 tumor suppressor gene: from molecular biology to clinical
investigation. Ann N Y Acad Sci 2000; 910:121.
66. Baker SJ, Preisinger AC, Jessup JM, et al. p53 gene mutations occur in combination with
17p allelic deletions as late events in colorectal tumorigenesis. Cancer Res 1990;
50:7717.
67. Kikuchi-Yanoshita R, Konishi M, Ito S, et al. 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
1992; 52:3965.

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 23/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

68. Russo A, Bazan V, Iacopetta B, et al. The TP53 colorectal cancer international
collaborative study on the prognostic and predictive significance of p53 mutation:
influence of tumor site, type of mutation, and adjuvant treatment. J Clin Oncol 2005;
23:7518.
69. Kirsch DG, Kastan MB. Tumor-suppressor p53: implications for tumor development and
prognosis. J Clin Oncol 1998; 16:3158.
70. Kastan MB, Onyekwere O, Sidransky D, et al. Participation of p53 protein in the cellular
response to DNA damage. Cancer Res 1991; 51:6304.
71. Kuerbitz SJ, Plunkett BS, Walsh WV, Kastan MB. Wild-type p53 is a cell cycle checkpoint
determinant following irradiation. Proc Natl Acad Sci U S A 1992; 89:7491.

72. Woods DB, Vousden KH. Regulation of p53 function. Exp Cell Res 2001; 264:56.
73. Lane DP. Cancer. p53, guardian of the genome. Nature 1992; 358:15.
74. Yurgelun MB, Masciari S, Joshi VA, et al. Germline TP53 Mutations in Patients With Early-
Onset Colorectal Cancer in the Colon Cancer Family Registry. JAMA Oncol 2015; 1:214.
75. Hamid O, Varterasian ML, Wadler S, et al. Phase II trial of intravenous CI-1042 in
patients with metastatic colorectal cancer. J Clin Oncol 2003; 21:1498.

76. Warren RS, Kirn DH. Liver-directed viral therapy for cancer p53-targeted adenoviruses
and beyond. Surg Oncol Clin N Am 2002; 11:571.
77. Raj K, Ogston P, Beard P. Virus-mediated killing of cells that lack p53 activity. Nature
2001; 412:914.
78. Fearon ER, Cho KR, Nigro JM, et al. Identification of a chromosome 18q gene that is
altered in colorectal cancers. Science 1990; 247:49.
79. Hedrick L, Cho KR, Fearon ER, et al. The DCC gene product in cellular differentiation and
colorectal tumorigenesis. Genes Dev 1994; 8:1174.

80. Chan SS, Zheng H, Su MW, et al. UNC-40, a C. elegans homolog of DCC (Deleted in
Colorectal Cancer), is required in motile cells responding to UNC-6 netrin cues. Cell
1996; 87:187.

81. Thiagalingam S, Lengauer C, Leach FS, et al. Evaluation of candidate tumour suppressor
genes on chromosome 18 in colorectal cancers. Nat Genet 1996; 13:343.

82. Cho KR, Oliner JD, Simons JW, et al. The DCC gene: structural analysis and mutations in
colorectal carcinomas. Genomics 1994; 19:525.
83. Popat S, Houlston RS. A systematic review and meta-analysis of the relationship
between chromosome 18q genotype, DCC status and colorectal cancer prognosis. Eur J
Cancer 2005; 41:2060.
84. Sun XF, Rütten S, Zhang H, Nordenskjöld B. Expression of the deleted in colorectal
cancer gene is related to prognosis in DNA diploid and low proliferative colorectal

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 24/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

adenocarcinoma. J Clin Oncol 1999; 17:1745.


85. Goyette MC, Cho K, Fasching CL, et al. Progression of colorectal cancer is associated
with multiple tumor suppressor gene defects but inhibition of tumorigenicity is
accomplished by correction of any single defect via chromosome transfer. Mol Cell Biol
1992; 12:1387.
86. Reiss M, Santoro V, de Jonge RR, Vellucci VF. Transfer of chromosome 18 into human
head and neck squamous carcinoma cells: evidence for tumor suppression by
Smad4/DPC4. Cell Growth Differ 1997; 8:407.

87. Riggins GJ, Thiagalingam S, Rozenblum E, et al. Mad-related genes in the human. Nat
Genet 1996; 13:347.

88. Eppert K, Scherer SW, Ozcelik H, et al. MADR2 maps to 18q21 and encodes a TGFbeta-
regulated MAD-related protein that is functionally mutated in colorectal carcinoma. Cell
1996; 86:543.
89. Xie W, Rimm DL, Lin Y, et al. Loss of Smad signaling in human colorectal cancer is
associated with advanced disease and poor prognosis. Cancer J 2003; 9:302.
90. Zhou XP, Woodford-Richens K, Lehtonen R, et al. Germline mutations in BMPR1A/ALK3
cause a subset of cases of juvenile polyposis syndrome and of Cowden and Bannayan-
Riley-Ruvalcaba syndromes. Am J Hum Genet 2001; 69:704.
91. Markowitz S, Wang J, Myeroff L, et al. Inactivation of the type II TGF-beta receptor in
colon cancer cells with microsatellite instability. Science 1995; 268:1336.
92. Moynahan ME, Chiu JW, Koller BH, Jasin M. Brca1 controls homology-directed DNA
repair. Mol Cell 1999; 4:511.

93. Moynahan ME, Pierce AJ, Jasin M. BRCA2 is required for homology-directed repair of
chromosomal breaks. Mol Cell 2001; 7:263.

94. Phelan CM, Iqbal J, Lynch HT, et al. Incidence of colorectal cancer in BRCA1 and BRCA2
mutation carriers: results from a follow-up study. Br J Cancer 2014; 110:530.
95. Oh M, McBride A, Yun S, et al. BRCA1 and BRCA2 Gene Mutations and Colorectal Cancer
Risk: Systematic Review and Meta-analysis. J Natl Cancer Inst 2018; 110:1178.
96. Garcia JM, Rodriguez R, Dominguez G, et al. Prognostic significance of the allelic loss of
the BRCA1 gene in colorectal cancer. Gut 2003; 52:1756.
97. Lucas AL, Shakya R, Lipsyc MD, et al. High prevalence of BRCA1 and BRCA2 germline
mutations with loss of heterozygosity in a series of resected pancreatic adenocarcinoma
and other neoplastic lesions. Clin Cancer Res 2013; 19:3396.
98. Yurgelun MB, Allen B, Kaldate RR, et al. Identification of a Variety of Mutations in Cancer
Predisposition Genes in Patients With Suspected Lynch Syndrome. Gastroenterology
2015; 149:604.

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 25/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

99. National Comprehensive Cancer Network (NCCN). NCCN clinical practice guidelines in o
ncology. https://www.nccn.org/professionals/physician_gls (Accessed on August 19, 202
1).

100. Chung DC, Rustgi AK. DNA mismatch repair and cancer. Gastroenterology 1995;
109:1685.
101. Papadopoulos N, Nicolaides NC, Liu B, et al. Mutations of GTBP in genetically unstable
cells. Science 1995; 268:1915.
102. Baker SM, Bronner CE, Zhang L, et al. Male mice defective in the DNA mismatch repair
gene PMS2 exhibit abnormal chromosome synapsis in meiosis. Cell 1995; 82:309.
103. Papadopoulos N, Nicolaides NC, Wei YF, et al. Mutation of a mutL homolog in hereditary
colon cancer. Science 1994; 263:1625.

104. Boland CR, Thibodeau SN, Hamilton SR, et al. A National Cancer Institute Workshop on
Microsatellite Instability for cancer detection and familial predisposition: development
of international criteria for the determination of microsatellite instability in colorectal
cancer. Cancer Res 1998; 58:5248.

105. Syngal S, Weeks JC, Schrag D, et al. Benefits of colonoscopic surveillance and
prophylactic colectomy in patients with hereditary nonpolyposis colorectal cancer
mutations. Ann Intern Med 1998; 129:787.

106. Shibata D, Peinado MA, Ionov Y, et al. Genomic instability in repeated sequences is an
early somatic event in colorectal tumorigenesis that persists after transformation. Nat
Genet 1994; 6:273.

107. Sourrouille I, Coulet F, Lefevre JH, et al. Somatic mosaicism and double somatic hits can
lead to MSI colorectal tumors. Fam Cancer 2013; 12:27.
108. Mensenkamp AR, Vogelaar IP, van Zelst-Stams WA, et al. Somatic mutations in MLH1
and MSH2 are a frequent cause of mismatch-repair deficiency in Lynch syndrome-like
tumors. Gastroenterology 2014; 146:643.
109. Rodríguez-Soler M, Pérez-Carbonell L, Guarinos C, et al. Risk of cancer in cases of
suspected lynch syndrome without germline mutation. Gastroenterology 2013; 144:926.
110. Ku CS, Cooper DN, Wu M, et al. Gene discovery in familial cancer syndromes by exome
sequencing: prospects for the elucidation of familial colorectal cancer type X. Mod
Pathol 2012; 25:1055.
111. Klarskov L, Holck S, Bernstein I, Nilbert M. Hereditary colorectal cancer diagnostics:
morphological features of familial colorectal cancer type X versus Lynch syndrome. J Clin
Pathol 2012; 65:352.
112. Lindor NM, Rabe K, Petersen GM, et al. Lower cancer incidence in Amsterdam-I criteria
families without mismatch repair deficiency: familial colorectal cancer type X. JAMA
2005; 293:1979.
https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 26/41
14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

113. Shiovitz S, Copeland WK, Passarelli MN, et al. Characterisation of familial colorectal
cancer Type X, Lynch syndrome, and non-familial colorectal cancer. Br J Cancer 2014;
111:598.

114. Therkildsen C, Jönsson G, Dominguez-Valentin M, et al. Gain of chromosomal region 20q


and loss of 18 discriminates between Lynch syndrome and familial colorectal cancer. Eur
J Cancer 2013; 49:1226.

115. Middeldorp A, van Eijk R, Oosting J, et al. Increased frequency of 20q gain and copy-
neutral loss of heterozygosity in mismatch repair proficient familial colorectal
carcinomas. Int J Cancer 2012; 130:837.
116. Ionov Y, Peinado MA, Malkhosyan S, et al. Ubiquitous somatic mutations in simple
repeated sequences reveal a new mechanism for colonic carcinogenesis. Nature 1993;
363:558.
117. Fujiwara T, Stolker JM, Watanabe T, et al. Accumulated clonal genetic alterations in
familial and sporadic colorectal carcinomas with widespread instability in microsatellite
sequences. Am J Pathol 1998; 153:1063.
118. Schwitalle Y, Kloor M, Eiermann S, et al. Immune response against frameshift-induced
neopeptides in HNPCC patients and healthy HNPCC mutation carriers. Gastroenterology
2008; 134:988.
119. Hatch SB, Lightfoot HM Jr, Garwacki CP, et al. Microsatellite instability testing in
colorectal carcinoma: choice of markers affects sensitivity of detection of mismatch
repair-deficient tumors. Clin Cancer Res 2005; 11:2180.
120. de la Chapelle A, Hampel H. Clinical relevance of microsatellite instability in colorectal
cancer. J Clin Oncol 2010; 28:3380.
121. Mueller J, Gazzoli I, Bandipalliam P, et al. Comprehensive molecular analysis of
mismatch repair gene defects in suspected Lynch syndrome (hereditary nonpolyposis
colorectal cancer) cases. Cancer Res 2009; 69:7053.

122. Thibodeau SN, French AJ, Cunningham JM, et al. Microsatellite instability in colorectal
cancer: different mutator phenotypes and the principal involvement of hMLH1. Cancer
Res 1998; 58:1713.

123. Haugen AC, Goel A, Yamada K, et al. Genetic instability caused by loss of MutS
homologue 3 in human colorectal cancer. Cancer Res 2008; 68:8465.

124. Halford S, Sasieni P, Rowan A, et al. Low-level microsatellite instability occurs in most
colorectal cancers and is a nonrandomly distributed quantitative trait. Cancer Res 2002;
62:53.
125. Halford SE, Sawyer EJ, Lambros MB, et al. MSI-low, a real phenomenon which varies in
frequency among cancer types. J Pathol 2003; 201:389.

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 27/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

126. Whitehall VL, Walsh MD, Young J, et al. Methylation of O-6-methylguanine DNA
methyltransferase characterizes a subset of colorectal cancer with low-level DNA
microsatellite instability. Cancer Res 2001; 61:827.

127. Dong SM, Lee EJ, Jeon ES, et al. Progressive methylation during the serrated neoplasia
pathway of the colorectum. Mod Pathol 2005; 18:170.

128. Jass JR, Baker K, Zlobec I, et al. Advanced colorectal polyps with the molecular and
morphological features of serrated polyps and adenomas: concept of a 'fusion' pathway
to colorectal cancer. Histopathology 2006; 49:121.

129. Nazemalhosseini Mojarad E, Kashfi SM, Mirtalebi H, et al. Low Level of Microsatellite
Instability Correlates with Poor Clinical Prognosis in Stage II Colorectal Cancer Patients. J
Oncol 2016; 2016:2196703.
130. Wright CM, Dent OF, Newland RC, et al. Low level microsatellite instability may be
associated with reduced cancer specific survival in sporadic stage C colorectal
carcinoma. Gut 2005; 54:103.
131. Kohonen-Corish MR, Daniel JJ, Chan C, et al. Low microsatellite instability is associated
with poor prognosis in stage C colon cancer. J Clin Oncol 2005; 23:2318.
132. Weinberg RA. Oncogenes and tumor suppressor genes. CA Cancer J Clin 1994; 44:160.

133. Veigl ML, Kasturi L, Olechnowicz J, et al. Biallelic inactivation of hMLH1 by epigenetic
gene silencing, a novel mechanism causing human MSI cancers. Proc Natl Acad Sci U S
A 1998; 95:8698.

134. Herman JG, Umar A, Polyak K, et al. Incidence and functional consequences of hMLH1
promoter hypermethylation in colorectal carcinoma. Proc Natl Acad Sci U S A 1998;
95:6870.

135. Cunningham JM, Kim CY, Christensen ER, et al. The frequency of hereditary defective
mismatch repair in a prospective series of unselected colorectal carcinomas. Am J Hum
Genet 2001; 69:780.

136. Cui H, Horon IL, Ohlsson R, et al. Loss of imprinting in normal tissue of colorectal cancer
patients with microsatellite instability. Nat Med 1998; 4:1276.
137. Nakagawa H, Chadwick RB, Peltomaki P, et al. Loss of imprinting of the insulin-like
growth factor II gene occurs by biallelic methylation in a core region of H19-associated
CTCF-binding sites in colorectal cancer. Proc Natl Acad Sci U S A 2001; 98:591.
138. Kane MF, Loda M, Gaida GM, et al. Methylation of the hMLH1 promoter correlates with
lack of expression of hMLH1 in sporadic colon tumors and mismatch repair-defective
human tumor cell lines. Cancer Res 1997; 57:808.
139. Esteller M, Fraga MF, Guo M, et al. DNA methylation patterns in hereditary human
cancers mimic sporadic tumorigenesis. Hum Mol Genet 2001; 10:3001.

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 28/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

140. Baylin SB, Esteller M, Rountree MR, et al. Aberrant patterns of DNA methylation,
chromatin formation and gene expression in cancer. Hum Mol Genet 2001; 10:687.
141. Tycko B. Epigenetic gene silencing in cancer. J Clin Invest 2000; 105:401.

142. Herman JG, Baylin SB. Gene silencing in cancer in association with promoter
hypermethylation. N Engl J Med 2003; 349:2042.
143. Rhee I, Bachman KE, Park BH, et al. DNMT1 and DNMT3b cooperate to silence genes in
human cancer cells. Nature 2002; 416:552.
144. Cui H, Cruz-Correa M, Giardiello FM, et al. Loss of IGF2 imprinting: a potential marker of
colorectal cancer risk. Science 2003; 299:1753.
145. Kuiper RP, Vissers LE, Venkatachalam R, et al. Recurrence and variability of germline
EPCAM deletions in Lynch syndrome. Hum Mutat 2011; 32:407.

146. Ligtenberg MJ, Kuiper RP, Chan TL, et al. Heritable somatic methylation and inactivation
of MSH2 in families with Lynch syndrome due to deletion of the 3' exons of TACSTD1.
Nat Genet 2009; 41:112.

147. Niessen RC, Hofstra RM, Westers H, et al. Germline hypermethylation of MLH1 and
EPCAM deletions are a frequent cause of Lynch syndrome. Genes Chromosomes Cancer
2009; 48:737.

148. Sieber OM, Lipton L, Crabtree M, et al. Multiple colorectal adenomas, classic
adenomatous polyposis, and germ-line mutations in MYH. N Engl J Med 2003; 348:791.
149. Sampson JR, Dolwani S, Jones S, et al. Autosomal recessive colorectal adenomatous
polyposis due to inherited mutations of MYH. Lancet 2003; 362:39.
150. Wang L, Baudhuin LM, Boardman LA, et al. MYH mutations in patients with attenuated
and classic polyposis and with young-onset colorectal cancer without polyps.
Gastroenterology 2004; 127:9.
151. Ricciardiello L, Goel A, Mantovani V, et al. Frequent loss of hMLH1 by promoter
hypermethylation leads to microsatellite instability in adenomatous polyps of patients
with a single first-degree member affected by colon cancer. Cancer Res 2003; 63:787.
152. Croitoru ME, Cleary SP, Di Nicola N, et al. Association between biallelic and monoallelic
germline MYH gene mutations and colorectal cancer risk. J Natl Cancer Inst 2004;
96:1631.
153. Enholm S, Hienonen T, Suomalainen A, et al. Proportion and phenotype of MYH-
associated colorectal neoplasia in a population-based series of Finnish colorectal cancer
patients. Am J Pathol 2003; 163:827.

154. Kambara T, Whitehall VL, Spring KJ, et al. Role of inherited defects of MYH in the
development of sporadic colorectal cancer. Genes Chromosomes Cancer 2004; 40:1.

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 29/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

155. Al-Tassan N, Chmiel NH, Maynard J, et al. Inherited variants of MYH associated with
somatic G:C-->T:A mutations in colorectal tumors. Nat Genet 2002; 30:227.
156. Farrington SM, Tenesa A, Barnetson R, et al. Germline susceptibility to colorectal cancer
due to base-excision repair gene defects. Am J Hum Genet 2005; 77:112.

157. Lubbe SJ, Di Bernardo MC, Chandler IP, Houlston RS. Clinical implications of the
colorectal cancer risk associated with MUTYH mutation. J Clin Oncol 2009; 27:3975.

158. Webb EL, Rudd MF, Houlston RS. Colorectal cancer risk in monoallelic carriers of MYH
variants. Am J Hum Genet 2006; 79:768.
159. Croitoru ME, Cleary SP, Berk T, et al. Germline MYH mutations in a clinic-based series of
Canadian multiple colorectal adenoma patients. J Surg Oncol 2007; 95:499.
160. Nielsen M, van Steenbergen LN, Jones N, et al. Survival of MUTYH-associated polyposis
patients with colorectal cancer and matched control colorectal cancer patients. J Natl
Cancer Inst 2010; 102:1724.
161. Gupta RA, Brockman JA, Sarraf P, et al. Target genes of peroxisome proliferator-activated
receptor gamma in colorectal cancer cells. J Biol Chem 2001; 276:29681.
162. He TC, Chan TA, Vogelstein B, Kinzler KW. PPARdelta is an APC-regulated target of
nonsteroidal anti-inflammatory drugs. Cell 1999; 99:335.

163. Yang WL, Frucht H. Activation of the PPAR pathway induces apoptosis and COX-2
inhibition in HT-29 human colon cancer cells. Carcinogenesis 2001; 22:1379.
164. Dobbie Z, Muller PY, Heinimann K, et al. Expression of COX-2 and Wnt pathway genes in
adenomas of familial adenomatous polyposis patients treated with meloxicam.
Anticancer Res 2002; 22:2215.
165. Gupta RA, Tan J, Krause WF, et al. Prostacyclin-mediated activation of peroxisome
proliferator-activated receptor delta in colorectal cancer. Proc Natl Acad Sci U S A 2000;
97:13275.
166. Sarraf P, Mueller E, Smith WM, et al. Loss-of-function mutations in PPAR gamma
associated with human colon cancer. Mol Cell 1999; 3:799.
Topic 2485 Version 48.0

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 30/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

GRAPHICS

Pathogenesis of colorectal cancer: adenoma-carcinoma


sequence

Adapted from O'Brien, MJ, Winawer, SJ, Waye, JB. Colorectal polyps. In: Management of
Gastrointestinal Diseases, Winawer, SJ (Ed). Gower Medical, New York, 1992.

Graphic 56774 Version 1.0

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 31/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

A genetic model for colorectal tumorigenesis

LOH: loss of heterozygosity; DCC: deleted in colon cancer gene; APC: adenomatous polyposis coli gene; ACF:
aberrant crypt foci; MMR: DMA mismatch repair enzyme.

Modified from: Lynch, JP, Hoops, TC. The genetic pathogenesis of colorectal cancer. Hematol Oncol Clin North Am
2002; 16:775.

Graphic 55601 Version 1.0

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 32/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

Molecular classification of colorectal carcinoma

Chromosomal instability Mismatch repair Serrated/CIMP Hybrid


Heredity pathway pathway pathway pathway

Hereditary and sporadic Hereditary Hereditary and sporadic Sporadic

CIMP status Negative Negative High High Low

MSI status MSS MSI-H MSI-H MSI-L MSI-L or MSS

Chromosomal Present Absent Absent Absent Present


instability

KRAS mutation +++ +/- --- --- +++

BRAF mutation --- --- +++ +++ ---

MLH1 status Normal Mutation Methylated Partial Normal


methylation

MGMT --- --- +/- +++ +++


methylation

CIMP: CpG island methylator phenotype; MSS: microsatellite stability; MSI: microsatellite instability; MSI-H: high-level microsatellite
instability; MSI-L: low-level microsatellite instability; MGMT: O-6-methlyguanine DNA methyltransferase; +++: present; +/-: might or
might not be present; ---: absent.

Noffsinger AE. Serrated polyps and colorectal cancer: New pathway to malignancy. Annu Rev Pathol 2009; 4:343. Reprinted, with
permission, from the Annual Review of Pathology. Copyright © 2009 Volume 4, 2009 by Annual Reviews. www.annualreviews.org.

Graphic 72239 Version 2.0

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 33/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

Genetic mutations that are associated with an increased risk of colorectal cancer

Syndrome/disease Responsible gene Mode of acquisition

FAP APC Germline (inherited)

HNPCC (Lynch syndrome) MMR genes


hMSH2
hMLH1
hPMS1
hPMS2
hMSH3
hMSH6
EPCAM (epigenetic alteration)

Sporadic tumor Tumor suppressor genes Somatic (acquired)


TP53
DCC
APC

Oncogenes
MYC
RAS
SRC
HER2

MMR genes (epigenetic change)

FAP: familial adenomatous polyposis; APC: adenomatous polyposis coli gene; HNPCC: hereditary nonpolyposis colorectal cancer;
MMR: mismatch repair; hMSH: human mutS homolog; hMLH1: human mutL homolog 1; hPMS: human postmeiotic segregation;
EPCAM: epithelial cell adhesion molecule; DCC: deleted in colorectal carcinoma; HER2: human epidermal growth factor receptor 2.

Graphic 60613 Version 7.0

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 34/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

Genetics of retinoblastoma formation

The RB1 gene is located on human chromosome 13. In hereditary retinoblastoma, the
fertilized egg carries 1 defective copy of the RB1 gene and all retinal cells in this
offspring carry only a single functional RB1 gene copy. If this surviving copy is
eliminated in a retinal cell by a nonheritable mutation (the second hit of the Knudson
2-hit hypothesis), the cell will lack RB1 gene function and will proliferate into a tumor.
In nonheritable retinoblastoma, the fertilized egg is genetically wild type at the RB1
locus. In the retina of this offspring, retinoblastoma development requires 2 successive
nonheritable mutations striking both copies of the RB1 gene in the retinal precursor
cells. Because only a single nonheritable mutation is needed to eliminate RB1 function
in familial cases, multiple cells in both eyes are affected. In contrast, the 2 nonheritable
mutations required in nonheritable disease are unlikely to affect a single cell lineage,
yielding at most 1 tumor.

Graphic 73602 Version 5.0

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 35/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

Structure and major functions of both the full-length and truncated adenomatous polyposis co

(A) Full-length APC proteins contain multiple domains, including oligomerization domain, an armadillo repeat-domain, a 15- or
domain important for binding to beta-catenin, SAMP repeats for axin binding, a basic domain for microtubule binding, and C-te
that bind to EB1 protein. Due to its numerous interactions with a variety of proteins, APC is involved in cellular processes relate
migration,

adhesion, proliferation, differentiation, and chromosomal segregation. Most APC mutations occur within MCR.

(B) The C-terminally truncated proteins present in colorectal cancer lack the domains that are required for binding to microtubu
beta-catenin, thus leading to the induction of chromosomal instability, activation of proliferation, and inhibition of differentiatio
may have dominant properties that lead to stronger stimulation of cell migration and promotion of cellular survival.

EB1: end-binding protein 1; DLG: discs large gene; IQGAP1: IQ-motif-containing GTPase activation protein 1.

Reproduced from: Zhang L, Shay JW. Multiple Roles of APC and its Therapeutic Implications in Colorectal Cancer. J Natl Cancer Inst 2017; 109(8):
permission of Oxford University Press. Copyright © 2017 Oxford University Press.

Graphic 116584 Version 5.0

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 36/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

The APC gene, and possible pathogenetic mechanisms in familial


adenomatous polyposis

The APC (adenomatous polyposis coli) gene modulates β-catenin, Tcf transcriptional activation, and
Wnt signal transduction.

(A) In the presence of wildtype APC or in the absence of Wnt ligand, β-catenin is localized to the
adherens junction where it is associated with E-cadherin, α-catenin, p120cas, and indirectly with the
cytoskeleton. GSK3β phosphorylates β-catenin in a complex that contains β-catenin, APC, and axin
family members, and β-catenin is rapidly degraded by ubiquination at the proteosome.

(B) When APC is mutated, β-catenin accumulates in the cytoplasm and the nucleus. Similarly, binding
of Wnt ligand to its receptor, known as frizzled, inactivates the GSK3β kinase through dishevelled,
generating a cytosolic pool of β-catenin. β-catenin is associated with members of the Tcf family of
transcription factors and modulates the transcription of target genes with Tcf recognition sequences.
In some instances, β-catenin increases transcription of target genes by competing for Tcf binding
with corepressors, such as Groucho and CREB-binding protein (CBP), to relieve transcriptional
repression.

Reproduced with permission from: Goss KH, Groden J. Biology of the adenomatous polyposis coli tumor supressor. J
Clin Oncol 2000; 18:1967. Copyright © 2000 American Society of Clinical Oncology.

Graphic 72022 Version 3.0

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 37/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

Colorectal cancer TNM staging AJCC UICC 8th edition

Primary tumor (T)


T category T criteria

TX Primary tumor cannot be assessed

T0 No evidence of primary tumor

Tis Carcinoma in situ, intramucosal carcinoma (involvement of lamina propria with no extension
through muscularis mucosae)

T1 Tumor invades the submucosa (through the muscularis mucosa but not into the muscularis
propria)

T2 Tumor invades the muscularis propria

T3 Tumor invades through the muscularis propria into pericolorectal tissues

T4 Tumor invades* the visceral peritoneum or invades or adheres ¶ to adjacent organ or


structure

T4a Tumor invades* through the visceral peritoneum (including gross perforation of the bowel
through tumor and continuous invasion of tumor through areas of inflammation to the
surface of the visceral peritoneum)

T4b Tumor directly invades* or adheres ¶ to adjacent organs or structures

* Direct invasion in T4 includes invasion of other organs or other segments of the colorectum as a result of direct
extension through the serosa, as confirmed on microscopic examination (for example, invasion of the sigmoid colon by a
carcinoma of the cecum) or, for cancers in a retroperitoneal or subperitoneal location, direct invasion of other organs or
structures by virtue of extension beyond the muscularis propria (ie, respectively, a tumor on the posterior wall of the
descending colon invading the left kidney or lateral abdominal wall; or a mid or distal rectal cancer with invasion of
prostate, seminal vesicles, cervix, or vagina).

¶ Tumor that is adherent to other organs or structures, grossly, is classified cT4b. However, if no tumor is present in the
adhesion, microscopically, the classification should be pT1-4a depending on the anatomical depth of wall invasion. The V
and L classification should be used to identify the presence or absence of vascular or lymphatic invasion whereas the PN
prognostic factor should be used for perineural invasion.

Regional lymph nodes (N)


N category N criteria

NX Regional lymph nodes cannot be assessed

N0 No regional lymph node metastasis

N1 One to three regional lymph nodes are positive (tumor in lymph nodes measuring ≥0.2 mm),
or any number of tumor deposits are present and all identifiable lymph nodes are negative

N1a One regional lymph node is positive

N1b Two or three regional lymph nodes are positive

N1c No regional lymph nodes are positive, but there are tumor deposits in the:
Subserosa
Mesentery
Nonperitonealized pericolic, or perirectal/mesorectal tissues

N2 Four or more regional nodes are positive

N2a Four to six regional lymph nodes are positive

N2b Seven or more regional lymph nodes are positive

Distant metastasis (M)


M category M criteria

M0 No distant metastasis by imaging, etc; no evidence of tumor in distant sites or organs. (This
category is not assigned by pathologists.)

M1 Metastasis to one or more distant sites or organs or peritoneal metastasis is identified

M1a Metastasis to one site or organ is identified without peritoneal metastasis

M1b Metastasis to two or more sites or organs is identified without peritoneal metastasis

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 38/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

M1c Metastasis to the peritoneal surface is identified alone or with other site or organ metastases

Prognostic stage groups


When T is... And N is... And M is... Then the stage group is...

Tis N0 M0 0

T1, T2 N0 M0 I

T3 N0 M0 IIA

T4a N0 M0 IIB

T4b N0 M0 IIC

T1-T2 N1/N1c M0 IIIA

T1 N2a M0 IIIA

T3-T4a N1/N1c M0 IIIB

T2-T3 N2a M0 IIIB

T1-T2 N2b M0 IIIB

T4a N2a M0 IIIC

T3-T4a N2b M0 IIIC

T4b N1-N2 M0 IIIC

Any T Any N M1a IVA

Any T Any N M1b IVB

Any T Any N M1c IVC

TNM: Tumor, Node, Metastasis; AJCC: American Joint Committee on Cancer; UICC: Union for International Cancer Control.

Used with permission of the American College of Surgeons, Chicago, Illinois. The original source for this information is the AJCC Cancer
Staging Manual, Eighth Edition (2017) published by Springer International Publishing. Corrected at 4th printing, 2018.

Graphic 111438 Version 9.0

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 39/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

Amsterdam II criteria for Lynch syndrome

There should be at least three relatives with any Lynch syndrome-associated cancer (colorectal cancer, cancer of the
endometrium, small bowel, ureter, or renal pelvis)

One should be a first-degree relative of the other two

At least two successive generations should be affected

At least one should be diagnosed before age 50

Familial adenomatous polyposis should be excluded in the colorectal cancer case(s), if any

Tumors should be verified by pathological examination

Adapted from Vasen HF, Watson P, Mecklin JP, Lynch HT. New clinical criteria for hereditary nonpolyposis colorectal cancer (HNPCC, Lynch
syndrome) proposed by the International Collaborative group on HNPCC. Gastroenterology 1999; 116:1453.

Graphic 59832 Version 7.0

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 40/41


14/9/21 9:00 Molecular genetics of colorectal cancer - UpToDate

Contributor Disclosures
Harold Frucht, MD Nothing to disclose Aimee L Lucas, MD, MS Nothing to disclose Richard M
Goldberg, MD Equity Ownership/Stock Options: Advanced Chemotec Inc [Pancreatic cancers].
Consultant/Advisory Boards: Taiho [Drug development]; Merck [GI cancers]; Novartis [New drug
development]; Amgen [Paid lecture and travel]; Genentech [Expert testimony]; AdaptImmune; Astra
Zeneca; Bayer; G1 Therapeutics [Service on a DSMB]; Advanced Chemotec Inc. Other Financial Interest:
Genentech [Biosimilar]; Taiho [Expert testimony]. Benjamin A Raby, MD, MPH Consultant/Advisory
Boards: Sanofi [Asthma]; Genzyme [Asthma]; Regeneron [Asthma]; Teva [Asthma]. Diane MF Savarese,
MD Nothing to disclose

Contributor disclosures are reviewed for conflicts of interest by the editorial group. When found, these
are addressed by vetting through a multi-level review process, and through requirements for
references to be provided to support the content. Appropriately referenced content is required of all
authors and must conform to UpToDate standards of evidence.

Conflict of interest policy

https://www.uptodate.com/contents/molecular-genetics-of-colorectal-cancer/print?search=Colorectal cancer genetics&source=search_result&s… 41/41

You might also like