You are on page 1of 8

Review

The complex basis underlying common


fragile site instability in cancer
Efrat Ozeri-Galai*, Assaf C. Bester* and Batsheva Kerem
Department of Genetics, The Alexander Silberman Institute of Life Sciences, Edmond J. Safra Campus, The Hebrew University,
Jerusalem 91904, Israel

Common fragile sites (CFSs) were characterized almost of tumor suppressor genes and genomic amplification of
30 years ago as sites undergoing genomic instability in oncogenes [16–18]. In addition, CFSs are preferential hot-
cancer. Recently, in vitro studies have found that onco- spots for integrations of viral DNA, which can lead to cancer
gene-induced replication stress leads to CFS instability. development [19–22]. Recently, several studies have dem-
In vivo, CFSs were found to be preferentially unstable onstrated that oncogene expression leads to replication
during early stages of cancer development and to leave a stress [23–26]. Indeed, owing to the sensitivity of CFSs to
unique signature of instability. It is now increasingly replication stress, the genomic instability at CFSs precedes
clear that, along the spectrum of replication features the instability in other genomic regions [26]. In recent years,
characterizing CFSs, failure of origin activation is a com- studies have revealed a role for checkpoint and DNA dam-
mon feature. This and other features of CFSs, together age response (DDR) proteins in the mechanisms controlling
with the replication stress characterizing early stages of CFS stability (Table 1).
cancer development, lead to incomplete replication that In this review, we focus on recent results that shed new
results in genomic instability preferentially at CFSs. light on the sequence and replication features of CFSs and
Here, we review the shared and unique characteristics the mechanisms leading to their instability under replica-
of CFSs, their underlying causes and their implications, tion stress, such as during cancer development. Specifical-
particularly with respect to the development of cancer. ly, we discuss new findings from studies using the high-
resolution DNA combing approach to analyze the replica-
The involvement of CFSs in cancer tion dynamics along CFS regions; we also examine the role
CFSs were first described in 1984 as gaps and constrictions of CFSs in genomic instability in cancer.
in metaphase chromosomes of cells grown under mild
replication stress conditions [1]. The expression level of Sequence characteristics of CFSs
a CFS is measured by the frequency of the gaps and For many years, no specific DNA sequences characterizing
constrictions at that site on metaphase chromosomes. CFSs were found. However, examination of the DNA se-
CFSs are present in all individuals and are considered quence and structural characteristics showed that they are
to be part of the normal chromosomal structure. For some enriched in sequences with high DNA helix flexibility in
CFSs in humans, an ortholog has been found in mice [2–5] the twist angle [27,28]. These flexible sequences were
and primates [6]. Several studies suggested that CFSs are found to comprise interrupted AT-dinucleotide repeats
conserved even in the yeast Saccharomyces cerevisiae. (AT-rich) with a potential to form stable DNA secondary
These studies found regions in the yeast genome that structures that might lead to inhibition of DNA replication
are slow replicating [7] or are potential impediments to [29] (Box 1). Recent studies described below in more detail
replication forks [8,9]. At present, 87 CFSs are listed in the demonstrate that these AT-rich flexibility sequences can
NCBI gene database (http://www.ncbi.nlm.nih.gov/gene), lead to replication arrest and DNA breakage.
but the exact number depends on the inducer, cell type and An in vitro study of the replication pattern along DNA
method of analysis. One study analyzed 25 000 meta- fragments from the CFS FRA16D showed that the DNA
phases in lymphocytes and reported 230 CFSs; however, polymerase is arrested near AT-rich sequences [30]. Inter-
most of these sites were infrequently expressed [10]. estingly, the addition of the Werner helicase (Box 2), which
In vitro, most CFSs are induced using low concentrations is implicated in the resolution of DNA secondary structures,
of aphidicolin, an inhibitor of polymerase a, d and e [11,12]. alleviated the polymerase stalling at these sequences [30].
Under these conditions, CFSs are hot spots for sister chro- This result supports the role of secondary structures formed
matid exchange and show high frequency of translocations at AT-rich sequences in DNA replication perturbation. A
and deletions in somatic cell hybrid systems [13]. Thus, DNA second study analyzed the fragility of sequences from
breakage at CFSs leads to the genomic instability of these FRA16D in a yeast system. In this study, a YAC-containing
sites. In vivo, CFSs correlate with chromosomal breakpoints the sequence of the FRA16D region was found to break at
in tumors [14,15], and were found to be involved in deletions AT-rich sequences, even in the absence of replication stress.
Moreover, using a two-dimensional gel, it was demonstrated
Corresponding author: Kerem, B. (batshevak@mail.savion.huji.ac.il) that AT repeats can lead to replication fork arrest [31].
Keywords: fragile sites; replication; cancer; genomic instability; DNA damage;
dormant origins.
These two studies demonstrate that AT-rich sequences
*
These authors contributed equally to this manuscript. are difficult to replicate, probably as a result of secondary
0168-9525/$ – see front matter ß 2012 Published by Elsevier Ltd. doi:10.1016/j.tig.2012.02.006 Trends in Genetics, June 2012, Vol. 28, No. 6 295
Review Trends in Genetics June 2012, Vol. 28, No. 6

Table 1. Proteins involved in CFS stability under replication stress conditions


Protein General known Suggested function in CFS stability Experimental method Effect on CFS expression Refs
function under replication stress conditions (increase vs decrease)
ATR DNA damage Replication stress response and Downregulation, knockout, Increase [79,80]
transducer maintenance of stalled fork stability kinase-dead ATR, Seckel
syndrome cells
CHK1 DNA damage Replication stress response Downregulation Increase [81]
checkpoint
ATM DNA damage DSB repair Gene mutation Increase only with [82]
transducer ATR downregulation
BRCA1 DNA damage G2–M checkpoint Gene mutation, knockout, Increase [83]
mediator downregulation
FANCD2 DNA repair DNA repair Gene mutation, downregulation Increase [84]
XLF DNA repair Non-homologous end-joining (NHEJ) Gene mutation, downregulation Increase [85]
RAD51 DNA repair Homologous recombination (HR) Downregulation Increase [86]
DNA-PK DNA repair NHEJ Downregulation Increase [86]
Ligase IV DNA repair NHEJ Downregulation Increase [86]
SMC1 Cohesin subunit Postreplicative DSB repair Downregulation Increase [87]
Claspin S-phase checkpoint Maintaining stalled fork stability Downregulation Increase [88]
HUS1 Cell cycle checkpoint Intra S-phase checkpoint and Knockout Increase [89]
DNA repair
WRN Helicase Resolving DNA secondary structures Gene mutation, downregulation Increase [90]
BLM Helicase Resolving DNA secondary structures Gene mutation Increase [91]
Topo I Topoisomerase Regulating the coupling between the Inhibition by betulinic acid Decrease [92]
helicase and polymerase and/or camptothecin
Topo I Topoisomerase Preventing interference between Downregulation Increase [93]
replication and transcription

structure formation that subsequently leads to DNA break- Replication features of CFSs
age. Furthermore, an indication for the role of AT-rich Replication timing studies
sequences in DNA instability in vivo was found when Numerous mechanisms underlying CFS sensitivity to rep-
sequences at recurrent cancer breakpoints occurring in lication stress have been proposed years ago, including late
CFSs were shown to harbor DNA flexibility (AT-rich) timing of replication, widely separated origins and DNA
sequences [32–34]. In addition, proteins that are important sequences that are unusually difficult to replicate [36].
for the resolution of DNA secondary structures, such as Analyzing the unique replication pattern of CFSs is im-
helicases and topoisomerase I, were shown to modulate portant for understanding the replication features leading
CFS stability (Box 2 and Table 1). Recently, direct evidence to their sensitivity to replication stress, which results in
for replication fork arrest at AT-rich sequences was provided their instability. The replication pattern of several CFSs
by analyzing the endogenous FRA16C region in human cells was initially analyzed using replication timing methods
[35], discussed in more detail below. and showed a variety of replication patterns for different
CFSs. FRA3B [36,37] and FRA16D [38] start replication
very late during S-phase in unperturbed cells. A mild
Box 1. Sequence characteristics of CFSs replication stress generated by aphidicolin treatment
CFSs are mapped based on their appearance on metaphase leads, in some cases, to incomplete replication even in
chromosomes. Most CFSs have only been mapped cytogenetically, G2 cells [36,38,39]. Exploring the replication time on
and only 35 of them were mapped more accurately by molecular metaphase chromosomes showed that the replication of
cloning using fluorescence in situ hybridization (FISH) analysis [71]. FRA3B is asynchronous, and that only the late-replicating
A large number of cloned CFSs map to large genomic regions of
several hundred kb. Several studies have suggested that, although
allele showed gaps and constrictions [37].
CFSs map to large regions, there are ‘hot spots’ of instability that In other fragile sites (FRA7H [40], FRA7G [16], FRA1H
can be defined as the core of the fragile region [32,57,67,94]. For and FRA2G [41]), replication starts during early to mid
most CFSs, the core of fragility has not yet been identified, which S-phase. In these sites, a significant delay in replication
makes the determination of DNA sequences specific to CFSs a timing between adjacent regions was found. These results
difficult task. Although CFSs do not share specific sequences,
analysis of the flexibility of the twist angle between each two base
indicate that there is an intrinsic delay in the replication
pairs along the DNA sequence revealed high flexibility sequences in progression in these fragile sites. For FRA1H and FRA2G,
the CFSs FRA3B, FRA7H and FRA7G [28]. The high flexibility it was found that by late S-phase, approximately half of the
sequences that were analyzed for sequence composition are rich fragile site regions were still unreplicated [41]. The repli-
with interrupted runs of AT-dinucleotide repeats [29]. In subsequent
cation delay at FRA7H and FRA7G was shown to further
years, more cloned CFSs were analyzed and also found to be
enriched for AT-rich flexible sequences [27]. In silico and in vitro increase under replication stress by the addition of aphi-
analyses revealed that AT-rich sequences longer than 200 bp have dicolin treatment [16,40]. The delay in replication at these
the potential to form stable secondary structures [29,95]. The fragile sites could result from slow progression of the
structural characteristics of AT-rich flexibility sequences can lead replication fork throughout the region or from stalling of
to replication perturbation and thus CFS instability.
the replication fork at specific sites.
296
Review Trends in Genetics June 2012, Vol. 28, No. 6

Box 2. Proteins involved in resolution of DNA secondary FRA3B


structures required for CFS stability FRA3B is mapped over a large genomic region exceeding 3
Mb. Within this region lies the large tumor suppressor
Several proteins that are important for the resolution of DNA
secondary structures or for the stability of stalled forks have been gene, fragile histidine triad (FHIT), extending over 1.5 Mb.
shown to play a role in CFS stability. Downregulation of the Werner DNA combing analysis of 1.6 Mb overlapping FHIT dem-
helicase, which is implicated in the resolution of secondary onstrated that the replication rate and fork stalling fre-
structures leading to replication fork arrest, leads to gaps and quency along this site are similar to that along the whole
constrictions on metaphase chromosomes even in unstressed cells,
and it is important to maintain fragile site stability under normal
genome [45]. A correlation between the origin pattern and
conditions [90]. Another helicase implicated in CFS expression is the expression level of FRA3B was found. Analysis in
BLM, the helicase deficient in Bloom’s syndrome. Spontaneously lymphocyte cells, where FRA3B is highly expressed,
occurring chromosome aberrations in Bloom’s syndrome cells were showed that a large region (700 kb) within the fragile site
found to correlate with bands harboring fragile sites, oncogenes and had no origins of replication, even following replication
breakpoints involved in cancer rearrangements [91]. Another key
enzyme that is important for the integrity of DNA replication is
stress. By contrast, in fibroblasts, low FRA3B expression
topoisomerase I (Topo I). In vitro studies showed that aphidicolin levels correlate with multiple initiation and termination
treatment leads to uncoupling between the polymerase and the events that are evenly distributed all along the fragile
helicase, resulting in long stretches of single-stranded DNA (ssDNA) region. Thus, the paucity in origins along FRA3B plays
that can form secondary structures, such as hairpins or cruciforms,
a role in its instability.
at AT-rich sequences. Partial inhibition of Topo I can alleviate the
polymerase–helicase uncoupling and so reduce the probability of These results are also consistent with findings from a
the formation of secondary structures. Indeed, inhibition of Topo I genome-wide analysis of replication timing, in both lym-
results in a reduction of gaps and breaks at CFSs induced by phocytes and fibroblasts, using the Repli-Seq technique
aphidicolin [92]. Another study investigated the function of Topo I in [46]. The Repli-seq method does not allow mapping of exact
preventing the formation of RNA–DNA hybrids during transcription,
origin locations, but it can distinguish between origin-poor
which probably prevents the interference between replication and
transcription [93]. Interestingly, the results show that Topo I regions and regions with evenly distributed active origins.
depletion by short hairpin RNA induced an increase in the frequency Analysis of CFS expression showed that the repertoire of
of the CFS expression level. The different effects of the various CFSs expressed in fibroblasts is different from that of
functions of Topo I on CFS expression still need to be investigated. lymphocytes [47]. Although most of the highly expressed
Downregulation of claspin, a mediator protein involved in Chk1
activation, leads to CFS instability, even under normal growth
CFSs in fibroblasts were previously described in lympho-
conditions [88]. The yeast homolog of claspin, Mrc1, localizes to cytes, their expression levels were much lower [10,47]. The
replication forks during normal S-phase [96]. Mrc1 may form a initiation profiles from the Repli-seq analysis for another
pausing complex with Tof1 that binds to stalled forks and maintains major fragile site in lymphocytes, FRA16D, and the two
their stability by preventing progression of the replication fork strongest CFSs in fibroblasts showed patterns similar to
complex upon arrest of DNA synthesis [96]. In addition, it was
demonstrated that, in an Mrc1 mutant, repeat sequences that form
FRA3B. The expression level of FRA16D was also lower in
DNA secondary structures, show higher levels of instability [97]. fibroblasts compared with lymphocytes [45]. Nevertheless,
These results suggest that the role of claspin in maintaining the FRA3B and FRA16D were still among the 12 most
stability of forks stalled at DNA secondary structures is important for expressed CFSs in fibroblasts, representing 2.4% and
CFS stability. Other proteins involved in CFS stability are listed in
5.5% of the breaks, respectively [47]. The expression of
Table 1 (main text).
FRA3B and FRA16D in cells in which the origin density in
the fragile region is similar to the whole genome indicates
that features of these fragile sites other than paucity of
These replication timing analyses suggest that late or origins must contribute to their instability.
delayed replication of the fragile region is part of the The origins of replication at FRA3B were also assessed
mechanism leading to the gaps and constrictions at meta- using nascent strand analysis followed by microarray
phase chromosomes of cells grown under replication stress quantification [48]. This study found four inefficient origins
conditions. Because the cell cycle checkpoints are per- in a region of 50 kb within the 700 kb origin-poor region in
turbed in many cancers, the unreplicated regions within lymphocytes. Thus, both studies suggest that, in lympho-
CFSs cannot lead to cell cycle arrest; thus, the cell proceeds cytes, there is a large genomic region within FRA3B that
to mitosis with incomplete DNA replication at CFSs, has very low efficiency of origin activation. The combina-
resulting in chromosomal instability. However, many tion of late replication with low efficiency of origins could
regions in the genome replicate very late during S-phase, explain the incomplete replication of the FRA3B region in
and replication of at least 1% of the genome extends into G2 G2–M, leading to its instability.
[42], indicating that late replication itself is insufficient to
generate fragility. FRA16C
A recent DNA combing study analyzing the replication
DNA combing and nascent strand studies pattern along 600 kb within the CFS FRA16C identified
Only in recent years, with the development of powerful four AT-rich flexibility sequences over a length of 400 bp
techniques for the analysis of DNA replication, has a more [35]. The analysis revealed a slower replication rate and
detailed description of CFS replication features emerged. shorter origin distance at the fragile site region compared
To date, the replication dynamics along three CFSs, with the whole genome under normal growth conditions,
FRA3B, FRA16C and FRA6E, have been studied using indicating that additional origins are activated. Because
two of these novel methods: DNA combing [43] and nascent dormant origins are normally activated following replica-
strand DNA analysis [44]. tion stress, the replication pattern of FRA16C indicates
297
Review Trends in Genetics June 2012, Vol. 28, No. 6

that replication of this region is inherently perturbed.


However, following aphidicolin treatment, no additional Late rep
rich lica
origin activation was detected at the FRA16C region. In AT- tio
s n
fact, the interorigin distance in FRA16C under normal ne

ge

Or
growth conditions was shorter than the interorigin dis-

igin
ge
Lar
tance in the whole genome under replication stress. These

pauci
results imply that the FRA16C region activates all avail-

ation
able dormant origins under normal growth conditions.

ty
Under normal growth conditions as well as under repli-

eplic

FAD
cation stress, the FRA16C region shows higher levels of

dr

O
replication fork arrest compared with the whole genome.

Pe
b
rtu
These arrest sites were not randomly distributed but were

rt
ub
Pe ed
preferentially close to the AT-rich sequences. These results on DD
Transcripti R
support a model in which the replication at FRA16C is
intrinsically perturbed, even under normal growth condi-
tions, owing to AT-rich sequences. The stability of the TRENDS in Genetics

fragile site under normal conditions is maintained by Figure 1. The molecular characteristics of common fragile sites (CFSs) and the
dormant origin activation, which allows the completion interactions among them. The main characteristics of CFSs are presented around
of DNA replication. Under replication stress conditions, the circle. Lines in the interior of the circle connect these different characteristics.
The color code indicates characteristics that have been shown to contribute
the replication is further perturbed and more replication together to the expression of CFSs. Abbreviations: DDR, DNA damage response;
forks arrest at AT-rich sequences. However, no additional FADO, failure in activation of dormant origins.
origins are available, leading to replication delay and
fragile site instability. The replication dynamics in this
CFS demonstrate the role of the AT-rich sequences in the One characteristic of CFSs contributing to instability is
replication perturbation along CFSs. This replication per- colocalization with very large genes. Some of the large CFS
turbation, together with the inability to activate dormant genes [such as FHIT and WW domain containing oxidore-
origins under replication stress, probably leads to incom- ductase (WWOX)] play important roles in cancer develop-
plete replication, destabilizing the CFS. ment. It has been proposed that interference between the
replication and transcription processes of these large genes
FRA6E could explain CFS instability. Evidence supporting this
FRA6E harbors the large parkinson protein 2 (PARK2) hypothesis was found in a recent study that showed a
gene, which extends over 1.3 Mb. The replication pattern of correlation between the instability of CFSs and the expres-
approximately 1 Mb of FRA6E containing part of PARK2 sion level of the underlying large genes. Moreover, the
was analyzed using DNA combing [49]. The analyzed levels of RNA:DNA hybrids formed at CFSs harboring
region of this fragile site contains relatively long AT-rich large genes were shown to modulate CFS instability
sequences (400–1383 bp), along which the replication rate [50]. However, given that only approximately half of the
of the fragile site was slower, the origin distance was CFS regions are associated with very large genes, this
shorter and a higher frequency of fork arrests was found mechanism cannot explain the fragility of all CFSs [51].
compared with the whole genome. A comparison between Several recent studies focused on the chromatin struc-
the replication dynamics along FRA6E and a specific ture of CFSs indicated that they are characterized by a
early-replicating non-fragile region showed no significant specific nucleosome binding pattern and a compact chro-
difference in these parameters. These results suggest that matin structure, both of which can influence the replication
features other than replication dynamics, such as late pattern of CFSs (Box 3). However, these are not common
replication, are important components of CFS identity. features of all CFSs. Thus, condensed chromatin may
In addition, an approximately 500-kb region of FRA6E explain the instability of a subtype of CFSs or contribute
contained few activated origins compared with the flank- to CFS instability together with an additional feature, such
ing regions. These results suggest that both replication as fork stalling or origin paucity.
arrests and paucity of origin activation together lead to the The DNA replication studies described above demon-
replication stress sensitivity of FRA6E. Therefore, FRA6E strate that, in all studied CFSs, the replication pattern
represents a combination of the replication patterns of the differs from that of the whole genome. However, there are
two CFSs described above. Further high-resolution analy- clearly distinct replication features for different fragile
sis of FRA6E as well as other CFSs is required to evaluate sites, all resulting in high sensitivity to replication stress.
the entire spectrum of replication patterns leading to the We propose that the key characteristics of the CFS region
replication stress sensitivity of CFSs. underlying their replication traits can best be viewed as a
range. At one end are fragile sites harboring a large region
Mechanisms leading to CFS instability with inefficient origins. At the other end are fragile sites
Several lines of evidence suggest that no single mechanism containing sequences that form secondary structures im-
can account for the instability of all CFSs. Different poten- peding the replication fork progression. These sequences
tial characteristics of CFSs have been proposed over the lead to intrinsic replication perturbation and dormant
years, but none are common to all of the cloned CFSs origin activation, even under normal growth conditions.
(Figure 1). It is possible that there are also some CFSs between these
298
Review Trends in Genetics June 2012, Vol. 28, No. 6

Box 3. Chromatin organization at CFSs CFSs are involved in amplifications of genomic regions
Chromatin structure plays a role in replication, transcription and
along the chromosome, leading to multiple gene copies and,
repair processes [98], and so may impact CFS stability. The hence, elevated expression of those genes [60]. The ampli-
chromatin structure is influenced by nucleosome position, histones fication of MET, which involves breakpoints in FRA7G, as
and other DNA-binding proteins as well as by histone modifications, well as oncogene amplifications in FRA11F, FRA2H,
such as specific methylations and acetylations. FRA2S, FRA2G, FRA2C and FRA7I [61–64], were shown
In humans, most nucleosomes are not found at fixed sites but are
randomly distributed. However, analysis of the nucleosome posi-
to be involved in cancer development. CFSs are also in-
tions at the breakpoint regions of FRA3B showed that nucleosomes volved in the formation of translocations. One example
at these sites are located in fixed positions (termed ‘phased’ is the translocation between 3p14.2 and 8q24.1,
nucleosomes) [99]. DNA sequences associated in vivo with the t(3;8)(pl4.2;q24.13), identified in a family with high inci-
phased nucleosomes do not lead to phased nucleosomes in in vitro dence of clear renal cell carcinoma (RCC) [65]. The trans-
reconstitution experiments. These results suggest that other factors
contribute to the arrangement of nucleosomes at FRA3B [99]. This
location breakpoint on chromosome 3 is located in the
pattern of nucleosomes dictates which sequences will always be tumor-suppressor FHIT, close to the core of FRA3B
either wrapped around the histones or in the linker regions. For [66,67]. Recently, the breakpoints leading to a RET–PTC
example, if the origins of replication are always wrapped around the translocation, which is frequently found in papillary thy-
histones, this could interfere with the binding of origin complexes
roid carcinoma, were located in the CFSs FRA10C and
and delay replication initiation [98].
In another study, CFSs were found to be characterized by FRA10G [68]. Furthermore, treatment of a thyroid epithe-
hypoacetylation of histones. Analysis of six of the molecularly cloned lial cell line with a CFS inducer leads to spontaneous RET–
CFSs showed that five had lower levels of acetylated histones PTC rearrangements in vitro [68]. Altogether, a large
compared with the average in the genome implying that the number of studies, mainly analyzing chromosomal insta-
chromatin structure was relatively condensed. Moreover, relaxing
bility in tumor cells, indicate that CFSs are unstable
the chromatin by treatment with the histone deacetylase (HDAC)
inhibitor, TSA, or with 5-Aza, an inhibitor of DNA methylation, led to a during cancer development. However, until recently it
decrease in CFS expression [100]. This result supports the role of remained unclear whether genomic instability of CFSs
condensed chromatin structure in CFS instability. High levels of in cancer results simply from their inherent instability
acetylation have been associated with access to factors needed to or if it is a result of positive selection driven by altered
promote the binding of pre-replication complexes to replication
origins [101]. In addition, relaxed chromatin is important for enabling
cancer genes mapped within CFSs, leading to their over-
efficient repair of DNA damage [102]. Thus, the compact nature of the representation.
chromatin at CFSs might impair origin assembly and/or DNA damage One way to answer this question is to look at analyses of
repair, leading to CFS expression. The results of these two studies can oncogene amplification in a variety of cancers. The break-
help define the characteristics leading to the unique replication age–fusion–bridge (BFB) model predicts that amplification
pattern of CFSs and their sensitivity to replication stress.
is achieved by cycles of chromosomal breaks and fusions,
leading to amplification of genomic regions close to the
two extremes. These fragile sites might contain both diffi- chromosomal breakpoints [60]. According to this model,
cult to replicate sequences and inefficient origins. Howev- double-strand breaks (DSBs) could occur anywhere be-
er, in all cases, the CFS region fails to activate additional tween the centromere and the amplified region. In vitro
origins under replication stress conditions, leading to rep- studies showed that the chromosomal breakpoints, located
lication delay and, in some cases, incomplete replication in at the boundaries of BFB amplifications, are set by breaks
G2 cells. From the CFSs analyzed so far, it seems that, within CFSs [69]. Evidence for a role of CFSs in the BFB
despite the different replication features, late and/or mechanism leading to oncogene amplification in vivo was
delayed replication and limited origin activation are found in a cell line derived from human gastric carcinoma
shared by all CFSs and result in their genomic instability. [70]. Analysis of the breakpoints that led to MET amplifi-
cation revealed a ‘ladder-like’ structure and an inverted
Genome instability of CFSs during cancer development repeat organization of the amplicon, with FRA7G at its
Molecular mapping of fragile sites reveals that genomic boundaries, as predicted by the BFB model [16]. MET
instability in CFSs alters the expression of important amplification induced by FRA7G instability was later
cancer genes in different cancer types. FRA3B, the most found in primary esophageal adenocarcinoma tumor cells
studied CFS, maps to intron 5 of the tumor suppressor gene [18]. Similarly, additional fragile sites (FRA11F, FRA2H,
FHIT [52]. This CFS is altered in a variety of human FRA2S, FRA2C and FRA7I) are involved in BFB-type
cancers, including lung, breast, head and neck, stomach oncogene amplifications in different cancer types [61–64].
and pancreatic carcinomas [53]. Interestingly, integrations The finding of fragile sites at the boundary of the
of human papilloma virus (HPV) in cervical cancer cells amplicons shows that the DSBs involved in these BFB
were also found in FRA3B [54]. FRA16D also harbors the amplifications occur preferentially in the fragile regions.
very large tumor-suppressor gene, WWOX [32,55]. Dele- Furthermore, these results indicate that oncogene ampli-
tions leading to loss of heterozygosity (LOH) in FRA16D fications are indeed the outcome of an intrinsic instability
were found in different cancer types [56]. Deletions and conferred by the fragile site sequences, because, in the case
LOH in other CFSs are also associated with cancer, includ- of intrachromosomal amplification of large genomic
ing FRA6E [57] and FRA9E [58]. A more recent analysis of regions, the breaks are distant from the targeted genes
LOH in 20 patients with premalignant Barrett’s esophagus and so are not affected by selection [16].
revealed copy number losses associated with many CFSs Additional corroboration for the supposition that the
(FRA3B, FRA9A/9C, FRA4D, FRA5E, FRA1K, FRAXC, unstable nature of CFSs drives chromosomal instability in
FRA16D and FRA12B) [59]. cancer emerged recently from an elegant study [71] in
299
Review Trends in Genetics June 2012, Vol. 28, No. 6

which a large-scale analysis of 746 cancer cell lines new model in which replication stress is an important
revealed an instability signature in CFSs that differed mechanism leading to genomic instability during early
significantly from the signature normally found in cancer stages of cancer development. This proposes new targets
genes [71]. The deletions in recessive cancer genes were for cancer therapy based on modulation of the replication
enriched with homozygous deletions that confer selective stress response [76].
growth advantage, whereas CFS deletions were character- For many years, the mechanism leading to the instabil-
ized by small hemizygous deletions, indicating an in- ity of CFSs in cancer remained largely unknown. However,
creased local rate of DNA breakage. Importantly, the recent work has shown that activation of oncogenes leads
hemizygous signature of the fragile sites was also found to perturbed DNA replication, resulting in hyper-replica-
in fragile sites harboring tumor suppressor genes, such as tion, slow replication fork progression and activation of
FRA3B and FRA16D, that did not result in homozygous dormant origins of replication [23,24,77,78]. This onco-
inactivation of the underlying genes. However, by chance, gene-induced replication stress might explain the prefer-
some regions were converted to homozygous deletions by a ential genomic instability in CFSs. Importantly, oncogene-
small or large deletion on the other parental chromosome induced replication stress was found to result from insuffi-
[71]. This study suggests that the instability in fragile sites cient nucleotide synthesis required to support normal DNA
is the result of their intrinsic features regardless of their replication, which is consistent with in vitro DNA breaks in
gene content. CFSs induced by perturbing the nucleotide balance and
Another type of cancer-related instability involves the biosynthesis through application of hydroxyurea or BrdU,
integration of foreign DNA, especially viral DNA, into the as well as through folate deprivation [23]. Understanding
genome. Such integrations can act as a causative agent or a the molecular basis characterizing the different fragile
cofactor in tumorigenicity. HPV, the most important can- sites will shed light on chromosomal instability in cancer.
cer-related virus, is integrated in the genome of most
cervical cancer cells [72]. The first reported HPV integra- Concluding remarks
tion into a CFS was in FRA3B, near the RCC translocation Recent advances in technologies enabling whole-genome
breakpoint [54]. Further analyses revealed preferential instability and DNA replication analyses have led to a
HPV integrations into CFSs throughout the genome in breakthrough in understanding the mechanisms leading
primary cervical tumors [19–22]. Because HPV integrates to the instability of CFSs during the early stages of cancer
passively into the genome and is not directed by viral development. These analyses revealed that replication
genes, it was important to determine the molecular basis stress that occurs during cancer development leads to
for the preferential integration of HPV into CFSs. Recent- preferential instability along the replication-sensitive
ly, the expression of the HPV-16 E6 and E7 oncogenes was CFS regions. Instability in CFSs in cancer may lead to
found to confer replication stress leading to DNA DSBs and deletions, translocations, amplifications and integrations
genomic instability preferentially at CFSs [23]. This sug- of viral DNA. Furthermore, it is now increasingly clear that
gests that, following HPV infection, the host cells suffer CFSs are a collection of heterogeneous sites along the
from replication stress, leading preferentially to DSBs in genome that share features leading to their sensitivity
CFSs. The integration of the viral DNA is a result of to replication stress. However, various mechanisms might
incorrect ligation of the two double-stranded ends of the underlie the basis of this sensitivity in different subgroups
broken host chromosome and the viral DNA. of CFSs (Figure 1). It will be important to investigate the
Another type of viral integration that increases the risk interplay between the different mechanisms and their
of cancer development is the integration of retroviral vec- synergistic effects on CFS stability in vitro and in vivo.
tors used in gene therapy trials [73]. A large-scale analysis Further studies are needed to unfold the complex basis of
of a retroviral vector based on the murine leukemia virus CFS stability and its relevance to cancer development and
revealed preferential integrations of the retroviral DNA other human diseases.
into CFSs [74], indicating perturbed replication following
retroviral infection. Altogether, findings in recent years References
indicate that CFSs are preferentially unstable during early 1 Glover, T.W. et al. (1984) DNA polymerase alpha inhibition by
stages of cancer development. Because instability in CFSs aphidicolin induces gaps and breaks at common fragile sites in
alters the expression of important cancer genes, it is human chromosomes. Hum Genet. 67, 136–142
2 Glover, T.W. et al. (1998) The murine Fhit gene is highly similar to its
important to understand the molecular mechanisms lead- human orthologue and maps to a common fragile site region. Cancer
ing to instability of CFSs during cancer development. Res. 58, 3409–3414
3 Shiraishi, T. et al. (2001) Sequence conservation at human and mouse
CFSs are preferentially unstable during early stages of orthologous common fragile regions, FRA3B/FHIT and Fra14A2/Fhit.
cancer development Proc. Natl. Acad. Sci. U.S.A. 98, 5722–5727
4 Krummel, K.A. et al. (2002) The common fragile site FRA16D and its
To shed light on the events leading to genomic instability in associated gene WWOX are highly conserved in the mouse at Fra8E1.
cancer, several studies focused on the early stages of cancer Genes Chromosomes Cancer 34, 154–167
development. Large-scale LOH and comparative genome 5 Rozier, L. et al. (2004) Characterization of a conserved aphidicolin-
hybridization analyses during early stages of different sensitive common fragile site at human 4q22 and mouse 6C1: possible
association with an inherited disease and cancer. Oncogene 23, 6872–
cancers revealed that the instability in CFSs precedes
6880
the instability in other genomic regions [25,59]. Similar 6 Ruiz-Herrera, A. et al. (2004) Conservation of aphidicolin-induced
results were obtained in genome-wide analyses in precan- fragile sites in Papionini (Primates) species and humans.
cerous experimental models [75]. These findings suggest a Chromosome Res. 12, 683–690

300
Review Trends in Genetics June 2012, Vol. 28, No. 6

7 Cha, R.S. and Kleckner, N. (2002) ATR homolog Mec1 promotes fork 33 Mitsui, J. et al. (2010) Mechanisms of genomic instabilities underlying
progression, thus averting breaks in replication slow zones. Science two common fragile-site-associated loci, PARK2 and DMD, in germ
297, 602–606 cell and cancer cell lines. Am. J. Hum. Genet. 87, 75–89
8 Lemoine, F.J. et al. (2005) Chromosomal translocations in yeast 34 Debacker, K. et al. (2007) FRA18C: a new aphidicolin-inducible fragile
induced by low levels of DNA polymerase a model for chromosome site on chromosome 18q22, possibly associated with in vivo
fragile sites. Cell 120, 587–598 chromosome breakage. J. Med. Genet. 44, 347–352
9 Admire, A. et al. (2006) Cycles of chromosome instability are 35 Ozeri-Galai, E. et al. (2011) Failure of origin activation in response to
associated with a fragile site and are increased by defects in DNA fork stalling leads to chromosomal instability at fragile sites. Mol. Cell
replication and checkpoint controls in yeast. Genes Dev. 20, 159–173 43, 122–131
10 Mrasek, K. et al. (2010) Global screening and extended nomenclature 36 Le Beau, M.M. et al. (1998) Replication of a common fragile site,
for 230 aphidicolin-inducible fragile sites, including 61 yet unreported FRA3B, occurs late in S phase and is delayed further upon induction:
ones. Int. J. Oncol. 36, 929–940 implications for the mechanism of fragile site induction. Hum. Mol.
11 Cheng, C.H. and Kuchta, R.D. (1993) DNA polymerase epsilon: Genet. 7, 755–761
aphidicolin inhibition and the relationship between polymerase 37 Wang, L. et al. (1999) Allele-specific late replication and fragility of the
and exonuclease activity. Biochemistry 32, 8568–8574 most active common fragile site, FRA3B. Hum. Mol. Genet. 8, 431–437
12 Ikegami, S. et al. (1978) Aphidicolin prevents mitotic cell division by 38 Palakodeti, A. et al. (2004) The role of late/slow replication of the
interfering with the activity of DNA polymerase-alpha. Nature 275, FRA16D in common fragile site induction. Genes Chromosomes
458–460 Cancer 39, 71–76
13 Arlt, M.F. et al. (2006) Common fragile sites as targets for 39 Wang, L. et al. (1998) Frequent homozygous deletions in the FRA3B
chromosome rearrangements. DNA Repair 5, 1126–1135 region in tumor cell lines still leave the FHIT exons intact. Oncogene
14 Hecht, F. and Hecht, B.K. (1984) Fragile sites and chromosome 16, 635–642
breakpoints in constitutional rearrangements I. Amniocentesis. 40 Hellman, A. et al. (2000) Replication delay along FRA7H, a common
Clin. Genet. 26, 169–173 fragile site on human chromosome 7, leads to chromosomal
15 Yunis, J.J. (1984) Fragile sites and predisposition to leukemia and instability. Mol. Cell. Biol. 20, 4420–4427
lymphoma. Cancer Genet. Cytogenet. 12, 85–88 41 Pelliccia, F. et al. (2008) Replication timing of two human common
16 Hellman, A. et al. (2002) A role for common fragile site induction in fragile sites: FRA1H and FRA2G. Cytogenet. Genome Res. 121, 196–
amplification of human oncogenes. Cancer Cell 1, 89–97 200
17 Kotzot, D. et al. (2000) Parental origin and mechanisms of formation of 42 Widrow, R.J. et al. (1998) Very late DNA replication in the human cell
cytogenetically recognisable de novo direct and inverted duplications. cycle. Proc. Natl. Acad. Sci. U.S.A. 95, 11246–11250
J. Med. Genet. 37, 281–286 43 Bensimon, A. et al. (1994) Alignment and sensitive detection of DNA
18 Miller, C.T. et al. (2006) Genomic amplification of MET with by a moving interface. Science 265, 2096–2098
boundaries within fragile site FRA7G and upregulation of MET 44 Giacca, M. et al. (1997) Mapping replication origins by quantifying
pathways in esophageal adenocarcinoma. Oncogene 25, 409–418 relative abundance of nascent DNA strands using competitive
19 Thorland, E.C. et al. (2003) Common fragile sites are preferential polymerase chain reaction. Methods 13, 301–312
targets for HPV16 integrations in cervical tumors. Oncogene 22, 45 Letessier, A. et al. (2011) Cell-type-specific replication initiation
1225–1237 programs set fragility of the FRA3B fragile site. Nature 470, 120–123
20 Thorland, E.C. et al. (2000) Human papillomavirus type 16 46 Hansen, R.S. et al. (2010) Sequencing newly replicated DNA reveals
integrations in cervical tumors frequently occur in common fragile widespread plasticity in human replication timing. Proc. Natl. Acad.
sites. Cancer Res. 60, 5916–5921 Sci. U.S.A. 107, 139–144
21 Yu, T. et al. (2005) The role of viral integration in the development of 47 Le Tallec, B. et al. (2011) Molecular profiling of common fragile sites in
cervical cancer. Cancer Genet. Cytogenet. 158, 27–34 human fibroblasts. Nat. Struct. Mol. Biol. 18, 1421–1423
22 Matovina, M. et al. (2009) Identification of human papillomavirus type 48 Palakodeti, A. et al. (2010) Impaired replication dynamics at the
16 integration sites in high-grade precancerous cervical lesions. FRA3B common fragile site. Hum. Mol. Genet. 19, 99–110
Gynecol. Oncol. 113, 120–127 49 Palumbo, E. et al. (2010) Replication dynamics at common fragile site
23 Bester, A.C. et al. (2011) Nucleotide deficiency promotes genomic FRA6E. Chromosoma 119, 575–587
instability in early stages of cancer development. Cell 145, 435–446 50 Helmrich, A. et al. (2011) Collisions between replication and
24 Di Micco, R. et al. (2006) Oncogene-induced senescence is a DNA transcription complexes cause common fragile site instability at
damage response triggered by DNA hyper-replication. Nature 444, the longest human genes. Mol. Cell 44, 966–977
638–642 51 McAvoy, S. et al. (2007) Non-random inactivation of large common
25 Bartkova, J. et al. (2005) DNA damage response as a candidate anti- fragile site genes in different cancers. Cytogenet. Genome Res. 118,
cancer barrier in early human tumorigenesis. Nature 434, 864–870 260–269
26 Gorgoulis, V.G. et al. (2005) Activation of the DNA damage checkpoint 52 Zimonjic, D.B.B. et al. (1997) Positions of chromosome 3p14.2 fragile
and genomic instability in human precancerous lesions. Nature 434, sites (FRA3B) within the FHIT gene. Cancer Res. 57, 1166–1170
907–913 53 Pichiorri, F. et al. (2008) Molecular parameters of genome instability:
27 Lukusa, T. and Fryns, J.P. (2008) Human chromosome fragility. roles of fragile genes at common fragile sites. J. Cell. Biochem. 104,
Biochim. Biophys. Acta 1779, 3–16 1525–1533
28 Mishmar, D. et al. (1998) Molecular characterization of a common 54 Wilke, C.M. et al. (1996) FRA3B extends over a broad region and
fragile site (FRA7H) on human chromosome 7 by the cloning of a contains a spontaneous HPV16 integration site: direct evidence for
simian virus 40 integration site. Proc. Natl. Acad. Sci. U.S.A. 95, the coincidence of viral integration sites and fragile sites. Hum. Mol.
8141–8146 Genet. 5, 187–195
29 Zlotorynski, E. et al. (2003) Molecular basis for expression of common 55 Bednarek, A.K. et al. (2001) WWOX, the FRA16D gene, behaves as a
and rare fragile sites. Mol. Cell. Biol. 23, 7143–7151 suppressor of tumor growth. Cancer Res. 61, 8068–8073
30 Shah, S.N. et al. (2010) DNA structure and the Werner protein 56 O’Keefe, L.V. and Richards, R.I. (2006) Common chromosomal fragile
modulate human DNA polymerase delta-dependent replication sites and cancer: focus on FRA16D. Cancer Lett. 232, 37–47
dynamics within the common fragile site FRA16D. Nucleic Acids 57 Denison, S.R. et al. (2003) Characterization of FRA6E and its
Res. 38, 1149–1162 potential role in autosomal recessive juvenile parkinsonism and
31 Zhang, H. and Freudenreich, C.H. (2007) An AT-rich sequence in ovarian cancer. Genes Chromosomes Cancer 38, 40–52
human common fragile site FRA16D causes fork stalling and 58 Callahan, G. et al. (2003) Characterization of the common fragile site
chromosome breakage in S. cerevisiae. Mol. Cell 27, 367–379 FRA9E and its potential role in ovarian cancer. Oncogene 22, 590–601
32 Ried, K. et al. (2000) Common chromosomal fragile site FRA16D 59 Lai, L.A. et al. (2010) Deletion at fragile sites is a common and early
sequence: identification of the FOR gene spanning FRA16D and event in Barrett’s esophagus. Mol. Cancer Res. 8, 1084–1094
homozygous deletions and translocation breakpoints in cancer 60 Albertson, D.G. (2006) Gene amplification in cancer. Trends Genet. 22,
cells. Hum. Mol. Genet. 9, 1651–1663 447–455

301
Review Trends in Genetics June 2012, Vol. 28, No. 6

61 Reshmi, S.C. et al. (2007) Relationship between FRA11F and 11q13 82 Ozeri-Galai, E. et al. (2008) Interplay between ATM and ATR in
gene amplification in oral cancer. Genes Chromosomes Cancer 46, the regulation of common fragile site stability. Oncogene 27, 2109–
143–154 2117
62 Pelliccia, F. et al. (2010) Breakages at common fragile sites set 83 Arlt, M.F. et al. (2004) BRCA1 is required for common-fragile-site
boundaries of amplified regions in two leukemia cell lines K562 – stability via its G2/M checkpoint function. Mol. Cell. Biol. 24, 6701–
molecular characterization of FRA2H and localization of a new CFS 6709
FRA2S. Cancer Lett. 299, 37–44 84 Howlett, N.G. et al. (2005) The Fanconi anemia pathway is
63 Ciullo, M. et al. (2002) Initiation of the breakage-fusion-bridge required for the DNA replication stress response and for the
mechanism through common fragile site activation in human regulation of common fragile site stability. Hum. Mol. Genet. 14,
breast cancer cells: the model of PIP gene duplication from a break 693–701
at FRA7I. Hum. Mol. Genet. 11, 2887–2894 85 Schwartz, M. et al. (2009) Impaired replication stress response in cells
64 Blumrich, A. et al. (2011) The FRA2C common fragile site maps to the from immunodeficiency patients carrying Cernunnos/XLF mutations.
borders of MYCN amplicons in neuroblastoma and is associated with PLoS ONE 4, e4516
gross chromosomal rearrangements in different cancers. Hum. Mol. 86 Schwartz, M. et al. (2005) Homologous recombination and
Genet. 20, 1488–1501 nonhomologous end-joining repair pathways regulate fragile site
65 Cohen, A.J. et al. (1979) Hereditary renal-cell carcinoma associated stability. Genes Dev. 19, 2715–2726
with a chromosomal translocation. N. Engl. J. Med. 301, 592–595 87 Musio, A. et al. (2005) SMC1 involvement in fragile site expression.
66 Paradee, W. et al. (1996) A 350-kb cosmid contig in 3p14.2 that crosses Hum. Mol. Genet. 14, 525–533
the t(3;8) hereditary renal cell carcinoma translocation breakpoint 88 Focarelli, M.L. et al. (2009) Claspin inhibition leads to fragile site
and 17 aphidicolin-induced FRA3B breakpoints. Genomics 35, 87–93 expression. Genes Chromosomes Cancer 48, 1083–1090
67 Ohta, M. et al. (1996) The FHIT gene, spanning the chromosome 89 Zhu, M. and Weiss, R.S. (2007) Increased common fragile site
3p14.2 fragile site and renal carcinoma-associated t(3;8) breakpoint, expression, cell proliferation defects, and apoptosis following
is abnormal in digestive tract cancers. Cell 84, 587–597 conditional inactivation of mouse Hus1 in primary cultured cells.
68 Gandhi, M. et al. (2010) DNA breaks at fragile sites generate Mol. Biol. Cell 18, 1044–1055
oncogenic RET/PTC rearrangements in human thyroid cells. 90 Pirzio, L.M. et al. (2008) Werner syndrome helicase activity is
Oncogene 29, 2272–2280 essential in maintaining fragile site stability. J. Cell Biol. 180,
69 Coquelle, A. et al. (1997) Expression of fragile sites triggers 305–314
intrachromosomal mammalian gene amplification and sets 91 Fundia, A. et al. (1995) Non-random distribution of spontaneous
boundaries to early amplicons. Cell 89, 215–225 chromosome aberrations in two Bloom Syndrome patients.
70 Motoyama, T. et al. (1986) Comparison of seven cell lines derived from Hereditas 122, 239–243
human gastric carcinomas. Acta Pathol. Jpn 36, 65–83 92 Arlt, M.F. and Glover, T.W. (2010) Inhibition of topoisomerase I
71 Bignell, G.R. et al. (2010) Signatures of mutation and selection in the prevents chromosome breakage at common fragile sites. DNA
cancer genome. Nature 463, 893–898 Repair 9, 678–689
72 Lehoux, M. et al. (2009) Molecular mechanisms of human 93 Tuduri, S. et al. (2009) Topoisomerase I suppresses genomic
papillomavirus-induced carcinogenesis. Public Health Genomics 12, instability by preventing interference between replication and
268–280 transcription. Nat. Cell Biol. 11, 1315–1324
73 Hacein-Bey-Abina, S. et al. (2003) LMO2-associated clonal T cell 94 Sozzi, G. et al. (1996) The FHIT gene 3p14.2 is abnormal in lung
proliferation in two patients after gene therapy for SCID-X1. cancer. Cell 85, 17–26
Science 302, 415–419 95 Inagaki, H. et al. (2009) Chromosomal instability mediated by non-B
74 Bester, A.C. et al. (2006) Fragile sites are preferential targets for DNA: cruciform conformation and not DNA sequence is responsible
integrations of MLV vectors in gene therapy. Gene Therapy 13, 1057– for recurrent translocation in humans. Genome Res. 19, 191–198
1059 96 Katou, Y. et al. (2003) S-phase checkpoint proteins Tof1 and
75 Tsantoulis, P.K. et al. (2008) Oncogene-induced replication stress Mrc1 form a stable replication-pausing complex. Nature 424,
preferentially targets common fragile sites in preneoplastic lesions. 1078–1083
A genome-wide study. Oncogene 27, 3256–3264 97 Freudenreich, C.H. and Lahiri, M. (2004) Structure-forming CAG/
76 Murga, M. et al. (2011) Exploiting oncogene-induced replicative stress CTG repeat sequences are sensitive to breakage in the absence of
for the selective killing of Myc-driven tumors. Nat. Struct. Mol. Biol. Mrc1 checkpoint function and S-phase checkpoint signaling:
18, 1331–1335 implications for trinucleotide repeat expansion diseases. Cell Cycle
77 Dominguez-Sola, D. et al. (2007) Non-transcriptional control of DNA 3, 1370–1374
replication by c-Myc. Nature 448, 445–451 98 Wang, Y-H. (2006) Chromatin structure of human chromosomal
78 Bartkova, J. et al. (2006) Oncogene-induced senescence is part of the fragile sites. Cancer Lett. 232, 70–78
tumorigenesis barrier imposed by DNA damage checkpoints. Nature 99 Mulvihill, D.J. and Wang, Y-H. (2004) Two breakpoint clusters at
444, 633–637 fragile site FRA3B form phased nucleosomes. Genome Res. 14, 1350–
79 Casper, A.M. et al. (2002) ATR regulates fragile site stability. Cell 111, 1357
779–789 100 Jiang, Y. et al. (2009) Common fragile sites are characterized by
80 Casper, A.M. et al. (2004) Chromosomal instability at common fragile histone hypoacetylation. Hum. Mol. Genet. 18, 4501–4512
sites in Seckel syndrome. Am. J. Hum. Genet. 75, 654–660 101 Iizuka, M. et al. (2006) Regulation of replication licensing by
81 Durkin, S.G. et al. (2006) Depletion of CHK1, but not CHK2, induces acetyltransferase Hbo1. Mol. Cell. Biol. 26, 1098–1108
chromosomal instability and breaks at common fragile sites. 102 Cann, K.L and Dellaire, G. (2011) Heterochromatin and the DNA
Oncogene 25, 4381–4388 damage response: the need to relax. Biochem. Cell Biol. 89, 45–60

302

You might also like