You are on page 1of 15

REVIEWS

DISEASE MECHANISMS

Mechanisms underlying structural


variant formation in genomic disorders
Claudia M. B. Carvalho1,2 and James R. Lupski1,3,4,5
Abstract | With the recent burst of technological developments in genomics, and the clinical
implementation of genome-wide assays, our understanding of the molecular basis of genomic
disorders, specifically the contribution of structural variation to disease burden, is evolving
quickly. Ongoing studies have revealed a ubiquitous role for genome architecture in the
formation of structural variants at a given locus, both in DNA recombination-based processes and
in replication-based processes. These reports showcase the influence of repeat sequences on
genomic stability and structural variant complexity and also highlight the tremendous plasticity
and dynamic nature of our genome in evolution, health and disease susceptibility.

Genomic disorders
Genomic disorders are a group of diseases caused by Studying disease-causing structural variants that
Conditions that result from rearrangements of the human genome due to inherent have been classified as either extremely rare or de novo
rearrangements of the genomic instability that results in susceptibility to struc- provides a unique opportunity to glean insights into
genome rather than base pair tural variation mutagenesis1. In the context of human mutational mechanisms. Such studies have revealed
changes of DNA, and in which
disease, structural variants include deletions, duplica- complex exonic, genic and chromosomal rearrange-
genomic instability results
from the endogenous tions, triplications, added amplifications (for example, ments that can be generated in a single mutagenic
genome architecture. quadruplications) and other large-scale (that is, from event, for example, in disease-associated loci at 17p11.2
~50–200 bp, the size of an average exon, to megabases of and 17p12 (REF. 6), in MECP2 duplication syndrome
Structural variants DNA) copy number variants (CNVs) that are not resolved (OMIM 300260)7 and due to chromothripsis-like events
Variants that include copy
number variants and copy
by chromosome karyotype studies (<5 Mb), as well as in multiple congenital anomalies8,9. Other studies have
number neutral inversions, copy number-neutral inversions, insertions and trans- shown that the formation of structural variants can be
insertions and translocations in locations. Structural variants differ from the concept accompanied by additional genome modification that
a personal genome compared of single nucleotide polymorphisms (SNPs) or single may result in a disease trait. For example, CNVs and
with a reference genome.
­nucleotide variants (SNVs), which only change a s­ ingle SNVs can be generated concomitantly 10, and SNVs cre-
base or a few bases2, as the former requires the disrup- ated during mutagenic repair can potentially affect the
tion of the sugar–phosphate backbone of DNA and function of genes that do not map within the CNV. Also,
1
Department of Molecular involve many base pairs. The definition of a structural CNVs can be followed by extended regions of absence of
and Human Genetics, Baylor variant can overlap with the concept of small insertions heterozygosity (AOH)11,12. If the AOH that is generated
College of Medicine, Houston,
and deletions (indels), which were previously defined from template switching between homologues versus sis-
Texas 77030, USA.
2
Centro de Pesquisas René as variants of <10,000 bp in length3. However, in this ter chromatids occurs at an imprinted locus, disease can
Rachou – FIOCRUZ, Belo Review we consider indels as <50–100 bp in size, that result. Alternatively, the AOH region may encompass a
Horizonte, MG 30190–002, is, a variant size that can be detected within a single variant in a gene for a recessive disease and reduce it to
Brazil. next-generation sequencing read. homo­zygosity when only one parent is a ­carrier, thus
3
Department of Pediatrics,
Baylor College of Medicine,
Structural variants result from different muta- ­distorting Mendelian expectations.
Houston, Texas 77030, USA. tional mechanisms, including DNA recombination-, The accurate detection of the complete mutagenic
4
Human Genome Sequencing replication- and repair-associated processes. Current event at the single-base-pair level requires either the use
Center, Baylor College of approaches to model the formation of structural vari­ of Sanger sequencing, along with techniques that enable
Medicine, Houston, Texas
ants in the human genome include: the use of model large-scale CNV identification, or the use of composite
77030, USA.
5
Texas Children’s Hospital, organisms4; the use of in vitro human cells subjected to pipelines for CNV analysis of next-generation sequen­cing
Houston, Texas 77030, USA. stress5; and the direct observation of human genomic data13. Therefore, using combined molecular analytic
Correspondence to J.R L.  alterations, or rearrangement end products, that convey tools is necessary to delineate the entire range of vari­
jlupski@bcm.tmc.edu a disease trait. The manifested trait or genomic dis­order ation that is associated with a particular structural v­ ariant
doi:10.1038/nrg.2015.25 enables one to both ascertain the mutational event and in an individual personal genome. Methods that can be
Published online 29 Feb 2016 distinguish the affected individual from the population. combined include fluorescence in situ hybridization

224 | APRIL 2016 | VOLUME 17 www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

(FISH), array comparative genomic hybridization (aCGH), that is, such repeats are organized in hierarchical groups
Copy number variants SNP arrays, next-generation sequencing and Sanger of direct and inversely orientated sequences as opposed
(CNVs). Alteration in copy sequencing13,14. Potential implications for the evolution to simple interchromosomal and intrachromosomal
number (gain or loss) of a locus of genes and genomes remain to be further explored, SDs27–29. The genome-wide distribution patterns of large
resulting in deviation from the
normal diploid state.
but mutational signatures — that is, the ‘scars’ left in LCRs show that they overlap with regions that frequently
the genome by DNA repair and replication mechanisms undergo genomic rearrangements that are associated
Single nucleotide variants after structural variant mutagenesis has occurred — can with disease22. This finding can be explained by the fact
(SNVs). A single site in a DNA partially explain the complex pattern of poly­morphic that the large LCRs that flank a unique genomic region
sequence that differs among
human structural variants that has been revealed by the negatively affect genome stability 30, making the flanked
individuals.
1000 Genomes Project15. sequences prone to DNA rearrangements via nonallelic
Chromothripsis New mechanistic discoveries in humans are elu- homologous recombination (NAHR; also known as ectopic
“Shattering of chromosomes”. cidating how the formation of structural variants can ­homolo­gous recombination) or RBMs. These charac­
A single catastrophic event re‑structure a specific region of the genome to change teristic genomic sequence features can thus be used
affecting one chromosome and
leading to complex
gene expression either locally or genome-wide16,17, to identify novel genomic loci that are susceptible to
rearrangements in cancer. including through the alteration of chromatin architec- structural variant formation. Indeed, the first d
­ efinition
ture, with pathological consequences for carriers18. These of genomic disorders1 arose from the susceptibility of
Absence of heterozygosity discoveries are also paving the way to understanding the specific variant regions that are associ­ated with LCRs
(AOH). Refers to copy number
molecular basis of human disease, including ­diseases to undergo rearrangement in certain human diseases;
neutral genomic segments that
lack heterozygosity for assayed that result from somatic mutagenesis19. for example, a 1.4 Mb duplication of a 17p12 interval
polymorphic markers. In this Review, we first discuss the types of structural causes Charcot–Marie–Tooth disease type 1A (CMT1A;
variants that cause genomic disorders, before review- OMIM 118220), and a reciprocal recombination deletion
Template switching ing in detail the recombination and replication-based product causes hereditary neuropathy with liability to
Refers to a transient
­mechanisms (RBMs) by which these arise. Our focus is on ­pressures palsies (HNPP; OMIM 162500).
dissociation of the primer
and template followed by a the pervasive role of the genomic architecture, includ-
re-association to a distinct ing DNA sequence repeat structure and orientation, in Recurrent versus nonrecurrent rearrangements
template during DNA the mechanism of formation of structural variants that The formation of a CNV depends on the joining of
replication. It can occur within
have been implicated in human genomic disorders. two formerly separated DNA segments; these break-
the same replication fork
(short-distance template We also describe the most common types of complex point junctions yield insights into the mechanisms that
switch) or between distinct rearrange­ments and review current knowledge under- cause the chromosomal structural change. Recurrent
replication forks (long-distance lying their formation. We do not review the phenotypic rearrange­ments share the same size and genomic content
template switch). consequences of specific genomic disorders; for a recent in unrelated individuals. The breakpoints map within
Array comparative genomic
review on this topic please see REF. 20. long, highly identical, flanking interspersed paralogous
hybridization repeats (FIG. 1Aa), which mostly consist of LCRs28. One or
(aCGH). Microarray-based Repeat sequences in the human genome more dosage-sensitive genes are present in‑between the
technique that measures the The genomic architecture — that is, the local organi- repeats and will be affected by the copy number change.
relative copy number of DNA
zation of repeat and other sequences, including their Pathogenic CNVs are preferential rearrangement prod-
segments.
relative orientation, size, density and distribution — is ucts at such loci31. Recurrent structural variants can also
Replication-based a key feature that helps to predict the type of structural occur within tandem paralogous repeats (FIG. 1Ab) and
mechanisms vari­ants, ­rearrangement susceptibility and potential major alter the copy number of a dosage-sensitive gene that is
(RBMs). Replicative mechanism that underlie variation at a specific locus. present within32.
non-homologous DNA repair
mechanism of single-ended,
Approximately 50% of the human genome consists of Nonrecurrent rearrangements are rearrangements
double-stranded DNA repeat sequences21, which include mobile elements such as that have a unique size and genomic content at a given
(seDNA). Alu-processed pseudogenes, simple sequence repeats, tan- locus in unrelated individuals. A dosage-sensitive gene
demly repeated sequences (which feature at centromeres, (or genes) can be identified within the structural vari-
Rearrangement
telomeres, the short arm of acrocentric chromosomes and ant smallest region of overlap (SRO) among individu-
susceptibility
Regions of the genome prone ribosomal gene clusters) and low-copy repeats (LCRs) als who share overlapping clinical phenotypes. Distinct
to structural variation such as segmental duplications (SDs). SDs have been types of nonrecurrent rearrangements are observed in
formation. computationally defined as segments of DNA that con- genomic disorders, some of which are also delimited by
tain ≥90% of sequence identity and ≥1 kb in length in the LCRs. The formation of complex patterns of rearrange-
Mobile elements
A segment of DNA capable
reference haploid genome, and they constitute approxi- ments can frequently be observed in some of these loci.
of moving into a new genomic mately 4–5% of the human genome22. One of the surpris- For example, paralogous inverted repeats mediate the
position. ing results from compara­tive genomic studies has been formation of a unique end-product pattern that con-
the observation that the human genome sequence distin- sists of an inverted triplication that is interspersed with
Paralogous sequences
guishes itself from other species genomes, such as the fly duplicated genomic segments. This structure is desig-
Homologous sequences that
arose by duplication. or worm, by a greater genome-wide presence of SDs and nated DUP–TRP/INV–DUP (duplication–inverted
other LCRs21, which indicates that these repeat sequences triplication–­duplication)7,33–36 (FIG. 1Ba). In some cases,
Nonallelic homologous greatly ­contribute to individual human variation23,24. the inverted repeat pair consists of small repeats such
recombination Most large LCRs (>10 kb) consist of a cluster of as Alu elements37, but the majority of DUP–TRP/INV–
Nonallelic pairing of
paralogous sequences and
­ aralogous sequences of diverse genomic origin. The
p DUP rearrangements described thus far have been
crossover leading to deletions, molecular characterization of such repeated regions in LCR-mediated. Genomic disorders such as MECP2
duplications and inversions. the human genome has revealed a mosaic architecture25,26; duplication syndrome and Pelizaeus–Merzbacher

NATURE REVIEWS | GENETICS VOLUME 17 | APRIL 2016 | 225


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

A Recurrent rearrangements
Aa Interspersed repeat units Ab Tandem repeat units

*
* * *

CNV1 CNV1

CNV2 CNV2

B Nonrecurrent rearrangements
Ba DUP–TRP/INV–DUP Bb DUP and DEL Bc DUP and DEL without LCR involvement

* * *

CNV1 CNV1 CNV1

CNV2 CNV2 CNV2

CNV3 CNV3 CNV3

SRO SRO
Figure 1 | Schematic representation of recurrent and nonrecurrent genomic rearrangements observed in
genomic disorders. Black lines represent the genomic segments of a given locus. Dosage-sensitive genes
Nature that are
Reviews | Genetics
involved in the rearrangement are represented by black asterisks. Paralogous repeats or low-copy repeats (LCRs) are
represented by inverted and directly oriented horizontal orange arrows. Dashed lines indicate LCR regions. Individual
genomic copy number variants (CNVs) are represented by colour-matched rectangles (deleted segments (green);
duplicated segments (red); and triplicated segments (blue)) and identified as CNV1, CNV2 and CNV3. Locations of CNV
breakpoints are indicated by grey vertical arrows. Square brackets denote regions with clustering or grouping of
breakpoint junctions. A | Recurrent rearrangements share the same size and genomic content in unrelated individuals.
More than 40 non-overlapping genomic disorders are caused by recurrent rearrangements49. Aa | In these types of
structural variants the duplication or deletion breakpoints cluster within long, highly identical flanking interspersed
paralogous repeats (represented by directly oriented horizontal orange arrows) that serve as substrates for nonallelic
homologous recombination (NAHR). Ab | Alternatively, recurrent rearrangements can occur within tandem paralogous
repeats and affect the copy number of a dosage-sensitive gene that is present within the repeat. For example, ectopic
recombination of a tandem repeat consisting of opsin genes at Xq28 can cause red–green colour blindness
(OMIM 303900, OMIM 303800) or it can lead to copy number polymorphism at this locus32. B | Nonrecurrent
rearrangements present a unique size and genomic content at a given locus in unrelated individuals. At least 70 genomic
disorders that are caused by nonrecurrent rearrangements have been described49. Ba | DUP–TRP/INV–DUP (duplication–
inverted triplication–duplication) structures are nonrecurrent but, in some cases, have a limited genomic recurrence with
two of four breakpoints mapping to an inverted repeat pair7,33–36. In these cases, the triplicated segments are inverted in
relation to the duplications. Bb | Some genomic disorders are characterized by nonrecurrent rearrangements that show
breakpoint grouping (rather than clustering as in the recurrent cases discussed above) within paralogous repeats38–42.
Bc | By contrast, some genomic disorders are characterized by nonrecurrent rearrangements without any clustering or
grouping of breakpoints43–45. Green rectangles represent deletions (DEL); red rectangles represent duplications (DUP);
blue rectangles represent triplications (TRP). INV, inversion; SRO, smallest region of overlap.

disease (PMD; OMIM 312080) are characterized by propensity for the formation of complex rearrange­ments
nonrecurrent rearrangements, mostly duplications, such as interspersed duplications or triplications, as well
that present with one breakpoint grouping within an as different meiotic versus mitotic risk (discussed below).
LCR-laden region38–42 (FIG. 1Bb). By contrast, most of
the nonrecurrent rearrangements in genomic disorders Recurrent structural variants: key mechanisms
seem to be formed without LCR involvement (FIG. 1Bc), NAHR. Recurrent structural variants often result from
although many deletion and duplication variants are NAHR between directly-oriented or inverted LCRs that
mediated by recombination between Alu elements43–45. flank unique sequence genomic regions. According to the
Thus, two general types of genomic rearrangement, ectopic synapsis model, the recurrent size and genomic
Ectopic synapsis
Chromosomal homologue
recurrent and nonrecurrent, are observed and show content generated by events that are mediated by NAHR
synapses at a nonallelic intrinsically distinct features that reflect their under­ results from the homologous recombination requirement
position. lying mechanisms of formation. They also have a distinct for an ectopic synapsis preceding an ectopic crossing over

226 | APRIL 2016 | VOLUME 17 www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Aa Ab *

Del *
+

Dup

Inversion *

B Inter-LCR distances
>98% nucleotide similarity
* 90% to 98% nucleotide similarity

LCR length

Common recurrent CNV Uncommon recurrent CNV


* *

* *

Del

+ +

Dup * * * *

H1 * No recurrent CNV causative of genomic disorder is observed in H1 haplotype

Inversion
Del
Recurrent
H2 *
+
CNVs
Dup * *

Figure 2 | Nonallelic homologous recombination mechanism. A | Nonallelic homologous recombination (NAHR)


Nature Reviews | Genetics
between directly oriented repeats generates deletions and duplications (part Aa), whereas NAHR between inverted
|
repeats generates the inversion of the segment in between the repeats (part Ab). B   The rate of NAHR may vary within a
particular locus owing to the presence of low-copy repeat (LCR) pairs with distinct susceptibility risk. For example, NAHR
rate positively correlates with LCR length and inversely correlates with inter-LCR distance, which can lead to common
recurrent and uncommon recurrent copy number variants (CNVs) at a given locus identified in different patients with a
genomic disorder30 (for example, in common recurrent and uncommon recurrent deletions and reciprocal duplications in
Smith–Magenis syndrome and Potocki–Lupski syndrome, respectively)46. C | NAHR rate may vary among individuals with
distinct genomic architecture in a given locus or a structural variant haplotype that results in directly oriented LCR. This is
exemplified by inverted structural haplotypes that present with distinct risks of undergoing NAHR owing to structural
variability in LCR content and orientation. For instance, two divergent structural variant haplotypes, H1 and H2, are
observed in 17q21.31 region but deletions that cause Koolen-De Vries syndrome (OMIM 610443)150 occur on
chromosomes carrying the H2 haplotype. Dosage-sensitive genes flanked by LCRs are represented by asterisks. LCRs are
represented by horizontal colour-matched arrows.

that is dependent on the distance and the length of the been known to contribute to human disease for more
homologous sequence46. By contrast, RBMs do not show than 40 years48 and is perhaps one of the best studied
such a requirement, therefore generating rearrangements causative mechanisms for genomic disorders30. In fact,
that vary in size and genomic content. Ectopic homolo­ large LCR pairs located <5 Mb apart can lead to localized
gous recombination47 that occurs during meiosis has genomic instability via NAHR1,30 (FIG. 2A). Approximately

NATURE REVIEWS | GENETICS VOLUME 17 | APRIL 2016 | 227


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

40 non-overlapping genomic loci are associated with of the PRDM9 homologous recombi­nation hotspot
meiotic NAHR that manifest as genomic disorders motif ­5ʹ‑­CCNCCNTNNCCNC‑3´ (REFS 56,57) are also
of microdeletion and microduplication syndromes49. positively correlated with the frequency of NAHR28.
Importantly, the duplication of a locus predisposes the Experimentally, well-studied genomic disorders such as
region to undergo triplications in subsequent genera- the 17p11.2 microdeletion Smith–Magenis syndrome
tions by more than 100‑fold using the same molecular (SMS; OMIM 182290) and its reciprocal duplication
mechanism50. Most NAHR events that are causative of Potocki–Lupski syndrome (PTLS; OMIM 610883) indi-
genomic disorders result from crossovers between LCRs cate that preferential use of a particular LCR prevails, that
located on the same chromosome, that is, intrachromo- is, common and uncommon recurrent CNVs46,58 (FIG. 2B).
somal NAHR, or between non-allelic LCRs located in Evidence also supports the definition of active (hotspots)
homologous chromosomes, that is, interchromosomal and inactive (coldspots) crossover regions or intervals
NAHR30. Rare cases of recurrent or de novo unbalanced within LCRs59.
translocations51,52 result from crossovers between repeats Inter-individual variation in the rate of NAHR-
located on nonhomologous chromosomes. mediated deletions and duplications was observed in
The definition of the molecular features of NAHR sperm-based assay studies55,60. One of the potential
has been instrumental to uncovering susceptible regions causes is a personal or population locus-specific struc-
in the human genome28,53, as well as to discovering new tural variation that predisposes individuals who carry
genomic disorders31. The reciprocal nature of the NAHR particular structural variant haplotypes to present with
model proposes that interchromatid or interchromo- distinct NAHR susceptibility (FIG. 2C). Nevertheless, the
somal crossover will concomitantly generate a deletion contribution of inter-individual structural variants to
and a duplication of the same interval; therefore, some NAHR frequency is still not well understood. A recent
genomic disorders that arise from a deletion CNV have study revealed that the rate of meiotic NAHR correlated
also had a reciprocal duplication counterpart that is asso- between identical twins and was independent of age61,
ciated with a clinical phenotype, whereas other disorders which indicates that other as yet unidentified genetic or
remain to be clinically described1,30. Remarkably, recipro- environmental factors probably have an important role
cal duplication and deletion syndromes can sometimes in the regulation of NAHR.
manifest mirror image traits for selected phenotypes, for Inversions are particularly challenging to assess
example, weight (overweight and underweight in 17p11.2 in an individual personal genome owing to the lack of
microdeletion and microduplication syndromes, respec- high-throughput genome-wide screening techniques
tively), head size (microcephaly versus macrocephaly in to assay such structures. Early evidence suggested that
1q21.1 and 16p11.2), microdeletion and microduplication mitotic NAHR-mediated inversions frequently occur in
syndromes associated with schizophrenia (del1q21.1 and somatic cells and potentially increase with age62. Meiotic
dup16p11.2) and autism (dup1q21.1 and del16p11.2) inversions may also be frequent in the human genome.
(reviewed in REF. 54). A study using fosmid paired-end sequencing data from eight
Intrachromatidal NAHR is predicted to exclusively human personal genomes from diverse populations identi­
produce deletions, whereas interchromatidal and inter- fied 50–100 large genomic inversions that are not repre-
chromosomal NAHR can mediate both deletions and sented in the human reference genome63. Furthermore,
duplications30. This hypothesis has been confirmed genome-wide identification and mapping of inverted
experimentally by analysing the rate of de novo NAHR- LCRs revealed that as much as ~12% of the genome may
generated deletions in the germ line, which indicated be susceptible to inversion that is mediated by NAHR53;
that deletions occur approximately twice as frequently inversions are therefore probably underestimated.
as duplications55, suggesting that NAHR has a prefer- In summary, evidence suggests that paralogous
ence for generating copy-number losses rather than sequence substrate size is important to ectopic synap-
copy-number gains. sis28,46, and PRDM9 variation and the PRDM9 hotspot
motif facilitate crossing over 56,57.
NAHR frequency is driven by local genomic architecture.
The complex structure of the LCR clusters suggests that Homologous versus homeologous recombination sub-
more than one repeat pair within a certain cluster can strates. Strand exchange during intrachromosomal
be used as substrates for a de novo event. Nevertheless, homologous recombination in mammalian cells pref-
the frequency of NAHR that occurs within and between erentially occurs in regions of uninterrupted sequence
LCRs mostly depends on distance, homology and size1. identity 64. A critical target size between consecutive
The distribution and relative frequencies of NAHR- nucleotide mismatches or minimum efficient process-
mediated rearrangements positively correlate with ing segments (MEPSs) seems to be required. In mam-
LCR length, whereas the frequency of rearrangement malian genomes, MEPSs were experimentally defined as
is inversely influenced by inter-LCR distances46. Such being between 134 and 232 bases in length; shortening
correlation is even stronger when both LCR length and of that segment of identity caused a sharp decrease in the
Fosmid paired-end distance are considered (ln(LCR length/inter-LCR dis- spontaneous recombination rate in mammalian cells64.
sequencing tance))28,46. Such coupled dependence suggests that, in NAHR favours recombination between substrates that
A clone-based method to
sequence the ends of
meiotic NAHR, an ectopic synapsis occurs as a precursor share perfect or near-perfect homology, as evidenced
fragments with a known to ectopic crossing-over 46. Other LCR features such as by the frequent use of LCRs ≥10 kb in length with
size range. DNA sequence identity, GC content and concentration ≥97% sequence identity 30. Nonetheless, disease-causing

228 | APRIL 2016 | VOLUME 17 www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

rearrange­ments between repeat regions with a lower per- extensive homology provided by LCRs that flank recur-
centage of sequence identity occurs in some pathogenic rent structural variants. In addition, small insertions have
CNVs (for example, HERV elements with 94–95% simi­ been identified in a number of junctions77, a characteris-
larity 65–67). HERV-mediated CNVs show enrichment for tic that is not observed in NAHR-mediated events. Such
PRDM9 hotspot motifs as well as the property of recur- features indicate that nonrecurrent structural variants
rence65,67, which suggests NAHR as their underlying are formed by molecular mechanisms that are distinct
mechanism. Other examples of short NAHR substrates from NAHR, for example, non-homologous end joining
generating CNVs include neuro­fibromatosis type I (NF1; (NHEJ) and RBMs such as break-induced replication
OMIM 162200), where strand exchange occurs between (BIR), ­microhomology-mediated break-induced replication
intervals as short as 114 bp (REF. 68), and α-thalassemia, (MMBIR), serial replication slippage (SRS) and fork ­stalling
where recombination between shorter intervals (34 bp in and template switching (FoSTeS) (reviewed in REF. 78). NHEJ
length) generates deletions of the human α‑globin gene69. generates mostly simple, blunt CNV end points and, infre-
By contrast, recombination between imperfectly quently, short homology (1–3 nucleotides); however, small
matched substrates, that is, homeologous sequences 70, deletions or the insertion of random nucleotides can also
may be mediated by mechanisms other than NAHR. be observed at breakpoint junctions79. By contrast, most
Homeologous sequences that underlie structural vari­ of the nonrecurrent structural variants that are associated
ant formation do not show enrichment for PRDM9 with genomic disorders present 2–33‑bp-long microho-
hotspot motifs and are shorter, with a lower percentage mologies at breakpoint junctions80. In addition, the inser-
Homeologous sequences
Imperfectly matched homology, than sequences used by NAHR. Structural tion of short segments (<100 bp) copied from nearby
paralogous genomic segments. variant formation events using homeologous sequences genomic regions is observed in up to 35% of nonrecurrent
mostly generate nonrecurrent rearrangements that can structural variant junctions10. These experimental obser-
Microhomologies be important contributors to genomic disorders43,44,71–73. vations support the notion that the underlying mechanism
Short stretches of shared
nucleotide identity present at
For example, recombination between retroelements, such of formation of nonrecurrent structural vari­ants involves
the junctions of rearranged as short non-autonomous ~300 bp Alu elements, lead- DNA replication in which microhomology is used as a
genomic segments. ing to CNVs is a predominant event in some genomic primer to assist replication initiation77,81.
dis­orders74, such as autosomal-­dominant spastic para-
Non-homologous end
plegia 4 (SPG4; OMIM 182601)43,72 and alveolar capil- RBMs. Evidence of DNA synthesis, which implicates
joining
(NHEJ). Double-stranded break lary dysplasia with misalignment of pulmonary veins RBMs in the formation of structural variants, has been
(DSB) mechanism of repair that (ACDMPV; OMIM  265380) 44. Alu-Alu-mediated documented in human diseases for almost a decade77,82.
processes the broken DNA rearrange­ment events have been historically attributed A key finding was the discovery that complex genomic
ends and joins to NAHR; however, the low nucleotide sequence identity rearrangements (CGRs), that is, rearrangements that
non-homologous sequences. It
repairs the programmed DSBs
shared between the Alu family members, which can be consist of more than two breakpoint junctions80, can
created in the immune system. as low as 75%, suggests that alternative recombination be formed in a single mutational event during DNA
mech­anisms also take place43. A recent report of com- repair 77,83. This contention is strongly supported by
Break-induced replication plex genomic variants such as interspersed duplications observations of a de novo nonrecurrent triplication that
(BIR). Homologous
and triplications mediated by multiple iterative template exclusively involves a single X‑chromosome inherited
recombination pathway that
repairs single-ended switches between Alu indicates that DNA synthesis can from a non-carrier father 7, and the finding that multiple,
double-stranded breaks be involved in Alu-Alu-mediated events and supports a interspersed de novo copy number-amplified segments
(seDSBs) through the role for DNA replication in their formation37. can all be stitched together in a single mutational event
establishment of a (chromoanasynthesis)8,9. Remarkably, copy number gains
unidirectional replication fork.
Nonrecurrent structural variants: key mechanisms are more frequently associated with CGRs than are copy
Microhomology-mediated At least 70 genomic disorders are known to be caused number losses46. This phenomenon may reflect the fact
break-induced replication by nonrecurrent structural variants49. This number that the formation of nonrecurrent copy number gains is
(MMBIR). RAD51‑independent is likely to be underestimated, because many of the particularly likely to be generated by RBMs rather than
break-induced replication that
genomic disorders that are associated with recurrent by any other mechanism, although this observation is
relies on microhomology to
resume replication. structural variants co‑occur with nonrecurrent events in potentially biased by the fact that the signatures of DNA
affected individuals (for selected examples see TABLE 1). synthesis are more likely to remain in rearrangement
Serial replication slippage Nonrecurrent rearrangements exhibit characteristic structures that are associated with copy number gains46.
(SRS). Multiple rounds of genomic features that are distinct from the recurrent The replicative process BIR is a homologous recombi­
slipped strand mispairing at
the replication fork.
structural variants that are mediated by NAHR (FIG. 1), nation pathway that is conserved from phage to eukar-
for example, the scattered locations of their breakpoint yotes and serves to repair single-ended double-­stranded
Fork stalling and template junctions. Therefore, nonrecurrent rearrange­ments breaks (seDSBs)84–86 (FIG. 3a). Spontaneous generation of
switching have a genomic content that is unique to the individ- such free DNA ends occurs, for example, during replica-
(FoSTeS). Mechanism of
ual in which the rearrangement was generated (FIG. 1B), tion fork collapse at stalled forks or during the S phase
template switching between
different replication forks. which makes the interpretation of potential clinical of the cell cycle, in regions where there is no replication
consequences challenging 75. Moreover, predicting the fork coming from the opposite direction (for example,
Chromoanasynthesis occurrence of nonrecurrent events genome-wide is more subtelomeric regions). In BIR, the broken chromosome
Chromosome reconstitution or difficult than that of recurrent structural variants. 3ʹ end invades a homologous template and initiates DNA
chromosome re-assortment.
Constitutive complex
The breakpoint junctions of nonrecurrent rearrange­ replication that may extend hundreds of kilobases up to
rearrangements resulting from ments can be characterized by simple blunt ends or the telomere, in which case homozygos­ity of all markers
multiple template switches. ­ icrohomologies76, which is in sharp contrast to the
m distal to the double-stranded breaks (DSBs) can follow 86.

NATURE REVIEWS | GENETICS VOLUME 17 | APRIL 2016 | 229


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

In yeast, BIR can also occur outside of the S phase, for RAD51‑dependent and RAD51‑independent BIR
example, in the G2 phase of the cell cycle. BIR requires (that is, MMBIR) seem to have different requirements
almost all S phase replication proteins to be initiated, for homology during DSB repair 91,92. MMBIR uses
including all three major DNA polymerases (Polα–­ microhomology to resume a stalled or collapsed repli-
primase, Polδ and Polε) in addition to the replicative cation fork as opposed to the longer homology tracts
nonessential ­subunit of Polδ, Pol32 (REFS 87–89). that are used in BIR92 (FIG. 3b). MMBIR has been pro-
Recent in vitro studies in human cells under replication posed as the major formative mechanism that underlies
stress have implicated BIR in the formation of genomic nonrecurrent structural variants in genomic disorders92
rearrangements, including segmental amplifications5. on the basis that three main observations could not be
Furthermore, nonrecurrent CNVs that are induced in readily explained by NAHR, NHEJ or BIR alone: first,
mammalian cells through the use of replication inhibi- the presence of microhomology in most of the break-
tors show many similarities to nonrecurrent structural point junctions, which may reflect priming DNA repli­
variants that are formed spontaneously in genomic dis- cation; second, template-driven juxtaposition of DNA
orders. Similarities include breakpoint junctions that sequences separated by large genomic distances, that
consist of microhomologies or blunt ends and contain is, long-range template switching; and third, iterative
small insertions, as well as the occurrence of complex template switches that generate breakpoint complexity
rearrangements90. Importantly, in vitro inacti­vation of and CGRs77,92.
Xrcc4, which encodes proteins required for canonical
NHEJ (c‑NHEJ), did not alter the frequency or the fea- RBMs are error prone
tures of the resulting structural variants in mouse embry- Microhomology mediates long-range template switch-
onic stem cells, which suggests that c‑NHEJ is unlikely ing. Template switching refers to a change of the single-­
to be the underlying mechanism90. Therefore, in vitro stranded DNA template during replication within the
experi­ments support the argument that RBMs under- same replication fork (short-distance template switch)
lie the formation of nonrecurrent structural variants, a or between distinct replication forks (long-distance
hypothesis that was first proposed almost 10 years ago template switch). Long-range template switching that
from the study of ­rearrangements that are causative for produces tandem amplification was observed in stress-­
genomic disorders77,82. induced Escherichia coli between segments sharing

Table 1 | Representative LCRs that mediate recurrent SVs as well as stimulate nonrecurrent genomic disorder associated SVs*
LCR Chromosome Nonrecurrent SV Size range (Mb) Associated clinical phenotype (OMIM) Refs
A, B, C 3q29 • DEL • 1.4–3.2 • 3q29 microdeletion syndrome (609425) 151
• DUP • 0.2–2.4 • 3q29 microduplication syndrome (611936)
SoS-PREP, 5q35 • DEL, DUP • 0.37–3.7 • Sotos syndrome (117550) 152,153
SoS-DREP • DEL–NML–DEL • 0.8
A, B, C 7q11.23 • DEL • 0.082–2.4 • Williams–Beuren syndrome (194050) 154–156
• DUP • 3.6 • Williams–Beuren region duplication syndrome
• TRP • 1.25 (609757)
BP3, BP4, BP5 15q13.1 • DUP–TRP/INV–DUP • 0.65 • 15q13.1 microdeletion syndrome (612001) 36
1‑8 or A‑H 22q11.2 • DEL • 0.02–4.0 • Velocardiofacial syndrome (192430) 157–159
• DEL–NML–DEL • 2.5 • DiGeorge syndrome (188400)
• 22q11.2 distal deletion syndrome (611867)
• 22q11.2 microduplication syndrome (608363)
SMS-REPs 17p11.2 • DEL, DUP, Complex • 0.41–19.6 • Smith–Magenis syndrome (182290) 6,46,160
• Potocki–Lupski syndrome (610883)
CMT1A‑REPs 17p12 • DEL, DUP, Complex • 0.009–3.4 • Charcot–Marie–Tooth disease type 1A (118220) 6,161
• Hereditary neuropathy with liability to pressure
palsies (162500)
NF1‑REP‑A, B, C 17q12 • DEL • 0.006–3.0 • Chromosome 17q11.2 deletion syndrome (613675) 162
CRI‑S232 Xp22.31 • DUP, Complex • 0.35–1.9 • X‑linked ichthyosis (308100) 105
PMDA‑D Xq22 • DUP, DEL • 0.1–11.0 • Pelizaeus–Merzbacher disease (312080) 7,33,40,42,77
• DUP–TRP/INV–DUP • Spastic paraplegia 2, X‑linked (312920)
• Complex
JA, JB, JC Xq28 • DUP, DEL • 0.040–4.0 • MECP2 duplication syndrome (300260) 38,39,41,163
• DUP–TRP/INV–DUP • X‑linked Emery–Dreifuss muscular dystrophy 1
K1, K2
• Complex (310300)
L1 (LCR1), Xq28 • DEL • 0.005–0.115 • Incontinentia pigmenti (308300) 164
L2 (LCR2)
DEL, deletion; DUP, duplication; INV, inversion; LCR, low-copy repeat; NML, normal (copy number neutral region); SV, structural variant; TRP, triplication. *Recurrent
SVs mediated by LCRs PMDA‑D, K1, K2, L1 and L2 are not causative of diseases. Distinct SV patterns of complex rearrangements, such as DUP–NML–INV/DUP,
DUP–NML–DUP, DEL–NML–DEL, are observed at the loci as described.

230 | APRIL 2016 | VOLUME 17 www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a BIR Replication Duplicated


1 2 fork collapse Nonhomologous flap clipping conservative
product
2 2 * *
3′
* 1 1/2 * 2

1 2 1/2 1/2

b MMBIR Replication
1 2 3 fork collapse
4
3′ TS1

1 2 3 4 4/2 4/2

TS2
Complex breakpoint junction
DUP TS3
+
DUP INV

1 2 3 4/2 3/1 2/4


Figure 3 | Replication-based mechanisms generate structural variants and single-nucleotide variants.
a | Break-induced replication (BIR) can be triggered by a nick on the template strand that may cause
Naturestalling
Reviewsor the
| Genetics
collapse of the replication fork. Resection of the 5ʹ end of the broken single-ended, double-stranded DNA (seDNA)
molecule exposes a 3ʹ tail that can invade an allelic (not shown) or a paralogous genomic segment with shared homology
(ectopic recombination) to prime replication. Paralogous segments are represented by horizontal orange arrows. The use
of ectopic homology to repair broken molecules by BIR can lead to structural variants (for example, duplication (shown
here), triplications, deletions and inversions). A conservative mode of repair was recently proposed for BIR that can
contribute to perpetuate repair indels or single-nucleotide variant (SNV) (black asterisks) mutations that are acquired
during the replicative repair116,117. b | Microhomology-mediated break-induced replication (MMBIR) can be triggered by a
nick on the template strand that may cause stalling or the collapse of the replication fork92. Alternatively, MMBIR can also
occur by disrupted BIR111. Resection of the 5ʹ end of the broken seDNA molecule exposes a 3ʹ tail that can anneal to a
single-strand DNA sharing microhomology (colour-matched boxes) to prime replication. The initial polymerase
extension and replication is carried out by a low processivity polymerase, rendering this repair process prone to
undergoing multiple rounds of disengagement and template switches until a fully processive replication fork is
established. Short and/or long template switches during repair can generate complexity due to the insertion of
templated segments at the rearrangement junction. DUP, duplication; INV, inversion; TS, template switches.

micro­homology 81. The presence of microhomology occurs and is followed by template switches it can lead
indicates a recombination process that does not require to the generation of complex structural variants81,92–96.
homology for strand invasion and DNA synthesis81. In Intrachromosomal and interchromosomal template
humans, structural variants that are causative of PMD switches have the potential to generate different types of
have revealed evidence for long-distance template switch- CGRs, such as the insertion of short genomic segments
ing that is mediated by microhomology, which led to the at the repair site, large-scale copy number alterations
proposal of the FoSTeS mechanistic model in genomic (for example, duplications, triplications and higher-or-
disorders77. A key observation during the formation of der amplifications) and inversions (FIG. 4). There seems
the structural variants at the PMD locus was the occur- to be a bias towards intrachromosomal template switch-
rence of multiple iterative template switches (FoSTeSX2, ing in human rearrangements, although interchromo-
FoSTeSX3, and so on) mediated by microhomology somal rearrangements do also occur 7,10,97. Importantly, if
that generated CGR, such as interspersed duplications nonallelic, homologous segments are used as templates
and triplications77. This study also showed that the use and replication extends for several kilobases78,94, tem-
of microhomology to assist repair can contribute to the plate switches will not lead to a change in copy num-
high error rate that is associ­ated with mechanisms that ber but the resulting genomic segment can show loss of
rely on a short stretch of ­homology to achieve strand ­heterozygosity, or observed AOH94.
transfer 76 (BOX 1). The molecular characterization of structural variants
in genomic disorders has revealed that apparently ‘sim-
Template switches can generate complex structural ple’ nonrecurrent rearrangements may actually consist
variants. Replicative mechanisms can undergo multi- of complex breakpoint junctions. Insertions of short
ple rounds of strand invasion; microhomology can be templated segments (<100 nucleotides) can be found in
used to resume DNA synthesis. If strand dissociation up to 35% of simple CNV junctions10,98–103. Most short

NATURE REVIEWS | GENETICS VOLUME 17 | APRIL 2016 | 231


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Microhomology-mediated templated insertions originate from nearby segments and fidelity relative to intergenerational DNA polymer-
end joining (within 300 bp), probably reflecting short-­distance tem- ases107. Distinct polymerases may contribute to template
(MMEJ). An alternative plate switches that are due to misalignment or to rep- switching via different mechanisms. For example, in
non-homologous end joining lication slippage within the same replication fork10,13,93. mammalian cell culture, Polθ introduces short templated
mechanism that repairs broken
double-stranded breaks using
Long-distance template switches introdu­cing short tem- insertions that are synthesized from mismatched primer
sequence microhomology to plate segments at the junctions occurred within <27 kb termini during DSB repair by microhomology-­mediated end
join and stabilize DNA end (REF. 10). By contrast, long-range template switches that joining (MMEJ)108. Long-range template switching, par-
intermediates. produced large-scale CGRs, such as CNVs that are vis- ticularly involving the replication of long DNA segments,
ible by high-resolution microarrays, vary in breakpoint may involve replicative polymerases; Polδ is a prominent
junction distance from a few kilobases to megabases candidate due to its lower processivity compared with
apart39,77,99,101–105. These observations may also be relevant Polε. Supporting this argument, BIR microhomolo-
for human genomic variability, as 27% of copy number gy-mediated events depend on the Polδ subunit Pol32
losses analysed in samples from the 1,000 Genomes (in yeast) or POLD3 (in humans) to initiate the repair of
Project show insertions varying from 1 to 96 nucleotides broken forks; in humans this repair can produce dupli-
at their breakpoint junctions; approximately 50% could cations as large as 200 kb (REFS 5,89). The participation of
be characterized as a templated insertion106. translesion synthesis (TLS) polymerases such as Polζ in
short-range template switching is supported by evidence
Complex rearrangements arise due to reduced proces- that fork stalling uses Polζ to overcome the impairment
sivity of DNA polymerases. The presence of de novo in progression109,110. Importantly, recent ­studies in yeast
indels, as well as short- and long-distance trans and cis have revealed that the TLS polymerases Polζ and Rev1
template switches that occur concomitantly, strongly sup- mediate template switches in BIR-defective cells using
port a faulty DNA replication process during repair by micro­homology to prime replication111. This finding sug-
RBMs6,10,13,39,77,82,104,105. In fact, the properties observed from gests that MMBIR can actually result from an interrupted
studies of breakpoint junction formation suggest that BIR111. In aggregate, these data support the idea that tem-
the polymerase used for RBM has reduced processivity plate switching events that generate CGR may result from

Box 1 | Microhomology can be added or removed during structural variant formation


A replicative mechanism was proposed by James Shapiro in 1979 to explain the semi-conservative transposition of the
bacteriophage Mu149, in which short nucleotide sequences are duplicated flanking the target site of transposition as a
result of the replication step that follows the 5ʹ strand transfer (microhomology additive mode; see the figure, part a). By
contrast, the fork stalling and template switching (FoSTeS)/microhomology-mediated break-induced replication (MMBIR)
model proposes that single-ended, double-stranded DNA (seDNA) is processed by a nuclease. A single strand end of the
3ʹ tail is then generated upon resection of the 5ʹ end anneals by virtue of Watson–Crick base pairing to a single-strand DNA
sharing microhomology (that is, template switching) and primes DNA replication. This process has the opposite effect to
the Shapiro model and results in the deletion of a short nucleotide stretch that is in common or shared between the
segments that are involved in generating the rearrangement junction, essentially reducing it from two copies to one
(microhomology subtractive mode; see the figure, part b).

a Microhomology in the Shapiro transposition model: additive

Transposable element . . . . A T G T Tn ATCC....


5′ strand
+
transfer
Reference TTCTAGGCACATTCTG

Rearrangement TTCTAGGCAC Tn GCACATTCTG


junction

b Microhomology at ‘join point’ in FoSTeS/MMBIR: subtractive

Reference ATGAATGACAGGATA...TCTAGACATATTCGA
3′ strand
transfer
Rearrangement junction ATGAATGACATATTCGA

Nature Reviews | Genetics


232 | APRIL 2016 | VOLUME 17 www.nature.com/nrg
©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a DEL b Head-to-tail tandem DUP c DUP–NML–DUP d DUP–NML–INV/DUP

A B C A B C A B C A B C
WT

TS1
TS1
TS TS TS2
TS TS2

A C A B B C A C A B C A C′ A B C
SV
jct jct jct1 jct2 jct1 jct2

DEL B DUP B DUP A, C DUP A


NML B NML B
INV/DUP C´

e DUP–TRP/INV–DUP f DUP–TRP/INV–DUP + AOH g Rolling-circle (DUP, TRP, QDR)


A B C D A B C
A B C
WT
α β γ δ

TS1 X 2
TS1
TS BIR
TS TS2
TS2

A B C B′ A B C A B C B′ α β γ δ A B C B C B A B C
SV
jct1 jct2 jct1 jct2 jct1 jct1 jct2
DUP A, C DUP A/α, C/γ DUP A
TRP/INV B´ TRP/INV B´ QDR B
AOH δ TRP C

Figure 4 | Template switches can generate distinct patterns of ‘simple’ or ‘complex’ structural variants.
Nature ReviewsAll displayed
| Genetics
structural patterns have been identified in one or more genomic disorders (TABLE 1). Top panel: black vertical line indicates a
wild-type (WT) genomic segment. Middle panel: black arrows represent segments of DNA generated upon template switches
(TSs). Red and green colour-matched boxes represent microhomology regions in WT genomic segment further involved in
annealing and resumption of replication during TS. Orange arrows represent inverted low-copy repeats (LCRs) or regions of
extended homology that can also be used during TS. Note that formation of any homology/microhomology breakpoint
junctions (represented by outlined colour-matched boxes) is accompanied by a relative net reduction of the copy number of
those homology/microhomology regions compared to WT regardless of whether the rearrangement results in gain or loss of
the flanking genomic material (see microhomology subtractive mode (BOX 1)). Bottom panel: resulting structural variant (SV).
Letters (A, B, C, D) represent chromosome alleles. Alleles subject to copy number gains in the resulting SVs are marked as bold
letters. Primed letters represent an inverted-oriented genomic segment that originates upon a TS between complementary
strands. α, β, γ, δ, represent corresponding homologue alleles. A single TS can generate deletion (DEL) (part a) or a head to tail
tandem duplication (DUP) with one junction (jct) (part b). Two TSs can generate distinct complex end products presenting with
two breakpoint junctions (jct1 and jct2). DUP–NML–DUP: copy number neutral segment (normal; NML) interspersed between
duplications (DUP) (part c). DUP–NML–INV/DUP: inverted interspersed duplications (part d). DUP–TRP/INV–DUP: inverted
triplication interspersed with duplications (part e). DUP–TRP/INV–DUP can be generated by break-induced replication (BIR),
if inverted repeats are used as substrates for TS, or microhomology-mediated BIR if microhomology is used for TS. DUP–TRP/
INV– DUP (part f) can be associated with extended regions of absence of heterozygosity (AOH) if TS occurs between
homologous chromosomes (interchromosomal TS). Rolling-circle amplification (part g) can be generated by re-replication
(for instance, the generation of two copies of jct1) culminating in the formation of multiple copies of a given segment.

the disruption of replication, dissoci­ation of replicative or near the breakpoint junction of duplicated segments
polymerases and switching to lower ­processivity poly- (~2.1 × 10−4 mutations per base pair and ~1.7 × 10−3
merases that are error prone. events per base pair, respectively) 10. Furthermore,
kataegis, which is frequently associated with genomic
RBMs show a high mutational rate. In addition to hav- rearrangements in some types of cancer such as breast
ing a high risk of undergoing template switching, DNA cancer 112,113, is hypothesized to result from susceptible
synthesis that is associated with repair can also lead to an BIR-generated single-stranded DNA (ssDNA)4.
Kataegis
Somatic single-nucleotide
increased local mutational rate10,107. The involvement of Consistent with human data, BIR in yeast is accom-
mutation clusters or mutation DNA synthesis during repair in humans is accompanied panied by a 1,000‑fold increase in the SNV mutation
showers in cis. by increased de novo SNV and indel mutation rates at rate compared with S phase replication107. A 5ʹ extensive

NATURE REVIEWS | GENETICS VOLUME 17 | APRIL 2016 | 233


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

resection generates a 3ʹ ssDNA end that will precede the features where they occurred106. For example, DNA rep-
strand invasion mediated by Rad51 (REF. 114). In BIR, lication timing differences across the genome are associ­
the combination of an extensive ssDNA in addition to ated with distinct types of CNVs. Recurrent CNVs are
leading and lagging strands that are synthesized in an more frequently observed in early replicating regions,
asynchronous way 115–117 results in the accumulation of whereas nonrecurrent CNVs are more frequently
ssDNA that is highly susceptible to unrepaired DNA observed in late replicating regions106,124. Further analysis
lesions, the formation of secondary structures and has indicated that regions of reduced rates of replication,
cleavage by endonucleases. The conservative mode of which are consistent with polymerase pausing, seem to
DNA synthesis of newly synthesized strands in BIR via a associate with breakpoint junctions of short homology-­
migrating DNA bubble (D‑loop) may contribute to the mediated CNVs125, supporting the contention that
propagation of these orientation-dependent acquired genomic regions that are prone to polymerase pausing
mutations116,117 (FIG. 3a). Importantly, the extent of repli- may be more susceptible to genomic instability. Nearby
cation inaccuracy resulting from BIR seems to be limited repeated sequences may themselves facilitate replicative
by the distance of the break-induced event to an efficient repair that can lead to nonrecurrent structural variants.
convergent fork as well as to the activity of the endonu-
clease Mus81. Mus81 cleaves the D‑loop and converts The role of repeat sequences in the formation of non­
conservative DNA synthesis into a canonical replication recurrent structural variants. Genomic architecture
fork and W‑C semi-­conservative synthesis118. Therefore, can stimu­late the occurrence of template switches near
Mus81 limits the high mutational rate of RBMs, includ- a DNA break. In yeast, the presence of repeats close to
ing template switches, to regions that are near the break- an induced break increases the frequency of the for-
point junctions of the structural variants formed by this mation of CGRs, whereas diverged human-derived
mechanism118. Switching to a semi-­conservative synthesis Alu elements inserted near seDSBs lead to increased
mode is likely to have a selective advantage by limiting template switching 118. Interestingly, a recent study in
the burst of mutations in the long genomic regions that budding yeast showed that the position of the strand
would otherwise follow. invasion during repair by RBMs is influenced by ‘micro-
homology islands’ that flank the junctions of resulting
The role of genomic architecture rearrangements126. It remains unclear how frequently
Formation of nonrecurrent structural variants is non- microhomology islands are found in nonrecurrent
random. Breakpoint junction mapping of nonrecurrent human rearrangements but it is tempting to specu-
rearrangements in genomic disorders indicates that late that repetitive e­ lements such as Alus may provide
the occurrence of structural variants in certain regions such microhomology islands that could assist strand
of the genome is unlikely to be random. Particular transfer or stimulate template switching during repair
genomic architectures, such as repeated sequences and by RBM. Moreover, a strong association between the
repetitive elements, with the potential to form non‑B locations of the breakpoint junctions of nonrecurrent
DNA structures (for example, A‑T rich palindromes, rearrangements within LCRs has been documented in
G‑quadruplexes, short inverted repeats and retrotrans- locus-specific studies (TABLE 1). This association has also
posable elements) are frequently observed in association been observed genome-wide in polymorphic structural
with the location of a template switch or structural vari- variants127; in fact, Kidd et al.128 indicate that ~20% of the
ant breakpoint 45,119–121. How such elements contribute to breakpoints of certain types of structural variants map
structural variant formation is currently being debated: within 5 kb of LCRs.
one possibility is that they render certain regions suscep- Mechanistically, how LCRs contribute to the forma-
tible to the transient formation of secondary DNA struc- tion of nonrecurrent rearrangements is not entirely clear
tures that can lead to fork collapse and, in some cases, and probably differs in nature from the homologous
to the formation of DSBs119,120. DSBs can contribute to recombination role that LCRs have in mediating recur-
the primary cause of instability in a particular genomic rent rearrangements. This is evident from a genome-wide
region; however, DSBs may not even be required to trig- study of polymorphic structural variants that showed
ger genomic stability. For example, common fragile sites that the location of nonrecurrent rearrangements hot-
(CFSs) that harbour [A]n and [TA]n repeats and quasi-­ spots differs from those of recurrent rearrangements129.
palindrome sequences such as short inverted repeats However, contrary to the challenge of predicting struc-
(4–6 nt) have been shown to perturb the progression of tural variant breakpoints using sequence motifs, LCRs
S phase Polδ replication upon which the polymerase may do allow some degree of prediction of the occurrence of
pause and dissociate from the replication machinery 122. nonrecurrent rearrangements and support the hypoth-
This event potentially causes an error-prone polymerase esis that the genomic architecture exerts multiple but
to take over to more efficiently replicate such complex distinct influences on structural variant formation
sites, thus leading to genomic instability that is ­conducive genome-wide. This is exemplified by specific CGR
to nonrecurrent genomic rearrangements122,123. structures such as DUP–TRP/INV–DUP that can be
Nonrandom formation of nonrecurrent rearrange- formed if highly identi­cal inverted LCRs are used as sub-
ments is supported by analyses of polymorphic struc- strates for BIR7,33,34,130. The presence of inverted repeats
tural variants in human populations 15,106 in which is hypothesized to provide a high degree of nucleotide
distinct signatures of the underlying structural variant sequence identity, which renders the region susceptible
mechanism can be associated with the local genomic to undergoing ectopic template switching. Such ectopic

234 | APRIL 2016 | VOLUME 17 www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a Chromoanasynthesis RBMs underlie some human translocations. Extensive


Reference genomic homology present at the telomere and subtelomere of
structure most chromosomes29 can trigger breakage–fusion–bridge
(BFB) cycles133,134 by forming dicentric chromosomes,
intrachromosomal or interchromosomal ectopic
recombination135 and secondary structures that can
render these regions prone to breaks136. NHEJ is likely
Patient CNVs to be a predominant mechanism of repair 137, although
4
Copy no.

recurrent NAHR between interchromosomal LCRs also


3
occurs51,135,138. Recent data have implicated RBMs in the
2
repair of broken telomeres on which homology, homeo­
logy or microhomology is used during repair 37,139–141,
including telomeric short inverted repeats142. Therefore,
LCRs and other types of repeats probably contribute
to the rearrangement events involving telomeres, sup-
Patient genomic porting the conclusion that genomic architecture can
structure have multiple roles in the formation of structural vari­
ants genome-wide and that RBMs might contribute to
increasing an individual’s genome complexity.

Conclusions
b High mutational load at CNV junction due to point mutations and template switches Studies of structural variation in genomic disorders
have resulted in mechanistic findings that apply to
Ancestral Duplicated model organisms and have wide implications for fields
segment segment
as diverse as human biology, genomics, molecular diag-
nostics, disease understanding and therapeutic interven-
tions based on gene dosage correction. The emerging
* * SNV picture is a perplexing fractal: large genomic-scale struc-
tural variants, that is, alterations involving megabases
Templated-insertion
of DNA, can be associated with additional complexities,
Figure 5 | Replication-based mechanisms are error prone. a | Schematic example such as other structural variants that are generated in a
Nature
of copy number variants (CNVs) in multiple, interspersed segments Reviews
within | Genetics
the same one-off event (chromoanasynthesis and chromothripsis-­
chromosome arm. Chromoanasynthesis in constitutional genomic disorders resembles like events (FIG.  5a) ) 8,9 or a CGR accompanied by
the phenomenon of chromosome catastrophes (that is, chromothripsis) initially AOH11,12. Further alterations might be revealed when
described based on the analyses of hundreds of cancer genome sequences and in these structural variants are scrutinized down to the
different types of cancers83. The chromosomal region involved in chromoanasynthesis is base-pair sequence level owing to the presence of small-
denoted by a grey box; local genomic segments involved are represented by coloured
scale changes that can act as mutational signatures, for
arrows. Array comparative genomic hybridization (aCGH) reveals genomic segments
with copy number variation. The rearranged chromosome is shown on which the extra example, insertions or deletions of short DNA segments
copy number segments (represented by colour-matched arrows) are inserted together in and/or de novo point mutations that are present adjacent
a new chromosomal position with respect to the reference genome. This clustering of to the junction of a structural variant 10,106,128,143 (FIG. 5b).
CNVs in a new position is a hallmark of chromoanasynthesis. b | Increased mutational This picture may further complicate the interpretation of
load can be observed at the breakpoint junctions of CNVs generated by replication- potential functional consequences of a mutagenic event
based mechanisms (RBMs). SNVs include point mutations and short insertions/short for an a priori apparently simple structural variant.
deletions (indels). Red rectangles represent the duplicated segment and red arrows The genomic architecture at a particular locus con-
represent the orientation of the duplicated segments. tributes to an increase in structural variant complexity,
with a prominent role for replicative mechanisms. LCRs
template switching during a replicative repair process can mediate recurrent structural variants by providing
may result in an inverted segment being formed con- substrate sequences for NAHR, and also stimulate non-
comitantly with a copy number gain (FIG. 4e). DUP–TRP/ recurrent structural variants by RBMs, where ectopic
INV–DUP is frequently observed in patients with PMD repeats can assist DNA repair by providing homology,
Breakage–fusion–bridge due to PLP1 duplication33 and is also present in ~20% homeology or microhomology sequences. RBMs under-
(BFB) cycles of individuals with MECP2 duplication syndrome7,39. lying the formation of a ‘simple’ or a ‘complex’ structural
Processes by which sister Importantly, DUP–TRP/INV–DUP can cause a much variant have important implications for human disease
chromatids that lack a
more severe clinical phenotype if the dosage-sensitive and transmission genetics. First, the occurrence of trip-
telomere (breakage) can
retrieve them by fusion and gene, that is, PLP1 or MECP2, maps within the tripli- lications can lead to more severe clinical phenotypes;
the creation of an unstable cated segment7,33,131,132. Therefore, the presence of inverted chromoanasynthesis events can cause multiple malfor-
dicentric chromosome that repeats near dosage-sensitive genes can indicate regions mations and congenital diseases, whereas segmental
will be pulled apart during in the human genome that are susceptible to DUP–TRP/ AOH may enable the expression of recessive traits and
anaphase (bridge). Eventually,
the bridge breaks and the cycle
INV–DUP formation53, which has been proved to be of regional uniparental disomy. Second, point mutations
starts again until the clinical relevance at different genomic loci34,35,130 (TABLE 1). and indels that are derived from error-prone repair that
chromosome is stabilized. is likely to arise concomitant with a disease-­associated

NATURE REVIEWS | GENETICS VOLUME 17 | APRIL 2016 | 235


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

structural variant may affect gene expression and con- ChIP-seq revealed that modifications of local chroma-
tribute to the variability of the associated genomic tin were strongly associated with the altered expression
disorder. Third, RBM-generated nonrecurrent of particular genes in 7q11.23 (REF. 16). The contribu-
rearrange­ments are hypothesized to be formed during tion of structural variants to the regulation of gene
mitosis. Therefore, these mutations may contribute to expression in cis or in trans needs to be further explored
diseases that are caused by somatic mutations, including as it may help to explain phenotypic variability in patients
cancer, developmental and neurological disorders144 and who carry overlapping non­recurrent rearrange­ments148.
to an increased probability of parental low-level mosai­ For example, new discoveries of mechanisms for struc-
cism6,145. Germline mosaicism can alter the recurrence tural variant formation in humans are elucidating how
risk for future pregnancies19. such events can re‑structure a specific region of the
Importantly, how genomic architecture contributes to genome to change the expression of transcripts locally
altered gene expression must be broadly investigated. The or genome-wide16,17, including through altering topo-
impact of a particular structural variant on the expres- logically associating domains (TADs), with pathological
sion of overlapping genes seems to go beyond simply the consequences for carriers18.
direct effect of the altered copy number. For example, In summary, the contribution of genomic architecture
lymphoblastoid cells from patients with deletions of the to structural variant formation goes beyond the definition
Williams–Beuren region (7q11.23) show dysregulation of and expansion of a group of human diseases designated
flanking genes, reflecting long-range cis-regulatory ele- as genomic disorders. It implicates particular mechanisms
ments within the structural variant146 that can also poten- for rearrangement formation and helps to guide studies
tially change the timing of gene expression147. Recently, about the potential causes of human disease, including
chromosome conformation capture (4C‑seq) and germline and somatic mutagenesis events.

1. Lupski, J. R. Genomic disorders: structural features observation relies on the potential implication 25. Kurotaki, N., Stankiewicz, P., Wakui, K., Niikawa, N.
of the genome can lead to DNA rearrangements and for human diseases that may include not only & Lupski, J. R. Sotos syndrome common deletion is
human disease traits. Trends Genet. 14, 417–422 dosage-sensitive genes but also unmasking of mediated by directly oriented subunits within inverted
(1998). recessive traits due to the extensive AOH, Sos-REP low-copy repeats. Hum. Mol. Genet. 14,
2. Zhang, F., Gu, W., Hurles, M. E. & Lupski, J. R. Copy distorting transmission genetics leading to disease 535–542 (2005).
number variation in human health, disease, and in a family with only a single carrier parent, as well 26. Park, S. S. et al. Structure and evolution of the Smith-
evolution. Annu. Rev. Genom. Hum. Genet. 10, as imprinting disease due to the presence of Magenis syndrome repeat gene clusters, SMS-REPs.
451–481 (2009). uniparental disomy (UPD). Genome Res. 12, 729–738 (2002).
3. Mills, R. E. et al. An initial map of insertion and 12. Sahoo, T. et al. Concurrent triplication and uniparental 27. Jiang, Z. et al. Ancestral reconstruction of segmental
deletion (INDEL) variation in the human genome. isodisomy: evidence for microhomology-mediated duplications reveals punctuated cores of human genome
Genome Res. 16, 1182–1190 (2006). break-induced replication model for genomic evolution. Nat. Genet. 39, 1361–1368 (2007).
4. Sakofsky, C. J. et al. Break-induced replication is rearrangements. Eur. J. Hum. Genet. 23, 61–66 28. Dittwald, P. et al. NAHR-mediated copy-number
a source of mutation clusters underlying kataegis. (2015). variants in a clinical population: mechanistic insights
Cell Rep. 7, 1640–1648 (2014). 13. Wang, Y. et al. Characterization of 26 deletion CNVs into both genomic disorders and Mendelizing traits.
5. Costantino, L. et al. Break-induced replication repair reveals the frequent occurrence of micro-mutations Genome Res. 23, 1395–1409 (2013).
of damaged forks induces genomic duplications in within the breakpoint-flanking regions and frequent 29. Linardopoulou, E. V. et al. Human subtelomeres are
human cells. Science 343, 88–91 (2014). repair of double-strand breaks by templated hot spots of interchromosomal recombination and
6. Zhang, F. et al. The DNA replication FoSTeS/MMBIR insertions derived from remote genomic regions. segmental duplication. Nature 437, 94–100 (2005).
mechanism can generate genomic, genic and exonic Hum. Genet. 134, 589–603 (2015). 30. Stankiewicz, P. & Lupski, J. R. Genome architecture,
complex rearrangements in humans. Nat. Genet. 41, 14. Coe, B. P. et al. Refining analyses of copy number rearrangements and genomic disorders. Trends Genet.
849–853 (2009). variation identifies specific genes associated with 18, 74–82 (2002).
7. Carvalho, C. M. et al. Inverted genomic segments and developmental delay. Nat. Genet. 46, 1063–1071 31. Sharp, A. J. et al. Discovery of previously unidentified
complex triplication rearrangements are mediated by (2014). genomic disorders from the duplication architecture
inverted repeats in the human genome. Nat. Genet. 15. Sudmant, P. H. et al. An integrated map of structural of the human genome. Nat. Genet. 38, 1038–1042
43, 1074–1081 (2011). variation in 2,504 human genomes. Nature 526, (2006).
8. Liu, P. et al. Chromosome catastrophes involve 75–81 (2015). The authors apply the conceptual mechanistic
replication mechanisms generating complex genomic 16. Gheldof, N. et al. Structural variation-associated understanding of NAHR to predict genomic
rearrangements. Cell. 146, 889–903 (2011). expression changes are paralleled by chromatin instability regions and define five novel genomic
Very complex rearrangements with multiple architecture modifications. PLoS ONE 8, e79973 disorders. This article provides evidence that
template switches can be formed constitutively in (2013). comprehending the rules underlying structural
a one-off event by RBM that is reminiscent of the 17. Ricard, G. et al. Phenotypic consequences of copy variation formation in the human genome is
chromothripsis events that were first described number variation: insights from Smith-Magenis and important and provides insights enabling predictions
in cancer. Potocki-Lupski syndrome mouse models. PLoS Biol. 8, of rearrangement-prone genomic regions.
9. Kloosterman, W. P. et al. Constitutional chromothripsis e1000543 (2010). 32. Nathans, J., Piantanida, T. P., Eddy, R. L., Shows, T. B.
rearrangements involve clustered double-stranded 18. Lupianez, D. G. et al. Disruptions of topological & Hogness, D. S. Molecular genetics of inherited
DNA breaks and nonhomologous repair mechanisms. chromatin domains cause pathogenic rewiring of variation in human color vision. Science 232,
Cell Rep. 1, 648–655 (2012). gene-enhancer interactions. Cell 161, 1012–1025 203–210 (1986).
10. Carvalho, C. M. et al. Replicative mechanisms for (2015). 33. Beck, C. R. et al. Complex genomic rearrangements at
CNV formation are error prone. Nat. Genet. 45, 19. Campbell, I. M., Shaw, C. A., Stankiewicz, P. & the PLP1 locus include triplication and quadruplication.
1319–1326 (2013). Lupski, J. R. Somatic mosaicism: implications for PLoS Genet. 11, e1005050 (2015).
This work revealed how apparently simple disease and transmission genetics. Trends Genet. 31, 34. Beri, S., Bonaglia, M. C. & Giorda, R. Low-copy repeats
structural variants can actually be highly complex 382–392 (2015). at the human VIPR2 gene predispose to recurrent and
and the complexity revealed by applying multiple 20. Weischenfeldt, J., Symmons, O., Spitz, F. & nonrecurrent rearrangements. Eur. J. Hum. Genet. 21,
experimental techniques to deduce structure and Korbel, J. O. Phenotypic impact of genomic structural 757–761 (2013).
understand the resultant end product of mutation. variation: insights from and for human disease. 35. Ishmukhametova, A. et al. Dissecting the structure
An unexpectedly high mutational spectrum Nat. Rev. Genet. 14, 125–138 (2013). and mechanism of a complex duplication-triplication
represented by both SNVs and template switches 21. Lander, E. S. et al. Initial sequencing and analysis of rearrangement in the DMD gene. Hum. Mutat. 34,
can be detected in up to 52% of the CNVs at the the human genome. Nature 409, 860–921 (2001). 1080–1084 (2013).
locus studied. 22. Bailey, J. A. et al. Recent segmental duplications in the 36. Soler-Alfonso, C. et al. CHRNA7 triplication associated
11. Carvalho, C. M. et al. Absence of heterozygosity due to human genome. Science 297, 1003–1007 (2002). with cognitive impairment and neuropsychiatric
template switching during replicative rearrangements. 23. Sebat, J. et al. Large-scale copy number phenotypes in a three-generation pedigree.
Am. J. Hum. Genet. 96, 555–564 (2015). polymorphism in the human genome. Science 305, Eur. J. Hum. Genet. 22, 1071–1076 (2014).
CNVs generated post-zygotically by 525–528 (2004). 37. Gu, S. et al. Alu-mediated diverse and complex
replication-based mechanisms can produce 24. Sharp, A. J. et al. Segmental duplications and copy- pathogenic copy-number variants within human
triplications that are associated with inversion number variation in the human genome. Am. J. Hum. chromosome 17 at p13.3. Hum. Mol. Genet. 24,
and long regions of AOH. The importance of this Genet. 77, 78–88 (2005). 4061–4077 (2015).

236 | APRIL 2016 | VOLUME 17 www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

38. Bauters, M. et al. Nonrecurrent MECP2 duplications and the distribution of rearrangement types and 82. Chen, J. M., Chuzhanova, N., Stenson, P. D., Ferec, C. &
mediated by genomic architecture-driven DNA breaks mechanisms in PTLS. Am. J. Hum. Genet. 86, Cooper, D. N. Complex gene rearrangements caused by
and break-induced replication repair. Genome Res. 18, 462–470 (2010). serial replication slippage. Hum. Mutat. 26, 125–134
847–858 (2008). 59. Cooper, G. M. et al. A copy number variation morbidity (2005).
39. Carvalho, C. M. et al. Complex rearrangements in map of developmental delay. Nat. Genet. 43, 838–846 83. Stephens, P. J. et al. Massive genomic rearrangement
patients with duplications of MECP2 can occur by fork (2011). acquired in a single catastrophic event during cancer
stalling and template switching. Hum. Mol. Genet. 18, 60. Lam, K. W. & Jeffreys, A. J. Processes of de novo development. Cell. 144, 27–40 (2011).
2188–2203 (2009). duplication of human alpha-globin genes. Proc. Natl Chromothripsis was defined and recognized as a
40. Inoue, K. et al. Genomic rearrangements resulting in Acad. Sci. USA 104, 10950–10955 (2007). one-off event in cancer.
PLP1 deletion occur by nonhomologous end joining 61. MacArthur, J. A. et al. The rate of nonallelic 84. Malkova, A., Ivanov, E. L. & Haber, J. E. Double-strand
and cause different dysmyelinating phenotypes in homologous recombination in males is highly variable, break repair in the absence of RAD51 in yeast:
males and females. Am. J. Hum. Genet. 71, 838–853 correlated between monozygotic twins and a possible role for break-induced DNA replication.
(2002). independent of age. PLoS Genet. 10, e1004195 Proc. Natl Acad. Sci. USA 93, 7131–7136 (1996).
41. Small, K. & Warren, S. T. Emerin deletions occurring on (2014). 85. Morrow, D. M., Connelly, C. & Hieter, P. “Break copy”
both Xq28 inversion backgrounds. Hum. Mol. Genet. 7, 62. Flores, M. et al. Recurrent DNA inversion duplication: a model for chromosome fragment
135–139 (1998). rearrangements in the human genome. Proc. Natl formation in Saccharomyces cerevisiae. Genetics 147,
42. Woodward, K. J. et al. Heterogeneous duplications in Acad. Sci. USA 104, 6099–6106 (2007). 371–382 (1997).
patients with Pelizaeus-Merzbacher disease suggest 63. Kidd, J. M. et al. Mapping and sequencing of structural 86. Malkova, A. & Ira, G. Break-induced replication:
a mechanism of coupled homologous and variation from eight human genomes. Nature 453, functions and molecular mechanism. Curr. Opin. Genet.
nonhomologous recombination. Am. J. Hum. Genet. 56–64 (2008). Dev. 23, 271–279 (2013).
77, 966–987 (2005). 64. Waldman, A. S. & Liskay, R. M. Dependence of 87. Lydeard, J. R., Jain, S., Yamaguchi, M. & Haber, J. E.
43. Boone, P. M. et al. The Alu-rich genomic architecture of intrachromosomal recombination in mammalian Break-induced replication and telomerase-independent
SPAST predisposes to diverse and functionally distinct cells on uninterrupted homology. Mol. Cell. Biol. 8, telomere maintenance require Pol32. Nature 448,
disease-associated CNV alleles. Am. J. Hum. Genet. 95, 5350–5357 (1988). 820–823 (2007).
143–161 (2014). 65. Sun, C. et al. Deletion of azoospermia factor a 88. Lydeard, J. R. et al. Break-induced replication requires all
44. Stankiewicz, P. et al. Genomic and genic deletions of (AZFa) region of human Y chromosome caused by essential DNA replication factors except those specific
the FOX gene cluster on 16q24.1 and inactivating recombination between HERV15 proviruses. Hum. Mol. for pre‑RC assembly. Genes Dev. 24, 1133–1144
mutations of FOXF1 cause alveolar capillary dysplasia Genet. 9, 2291–2296 (2000). (2010).
and other malformations. Am. J. Hum. Genet. 84, 66. Shuvarikov, A. et al. Recurrent HERV‑H‑mediated 89. Payen, C., Koszul, R., Dujon, B. & Fischer, G. Segmental
780–791 (2009). 3q13.2‑q13.31 deletions cause a syndrome of duplications arise from Pol32‑dependent repair of
45. Vissers, L. E. et al. Rare pathogenic microdeletions and hypotonia and motor, language, and cognitive delays. broken forks through two alternative replication-based
tandem duplications are microhomology-mediated Hum. Mutat. 34, 1415–1423 (2013). mechanisms. PLoS Genet. 4, e1000175 (2008).
and stimulated by local genomic architecture. 67. Campbell, I. M. et al. Human endogenous retroviral This paper shows that segmental duplications are
Hum. Mol. Genet. 18, 3579–3593 (2009). elements promote genome instability via non-allelic formed due to DNA synthesis rather than to ectopic
46. Liu, P. et al. Frequency of nonallelic homologous homologous recombination. BMC Biol. 12, 74 (2014). homologous recombination and that the underlying
recombination is correlated with length of homology: 68. Steinmann, K. et al. Type 2 NF1 deletions are highly mechanism is dependent on the nonessential Polδ
evidence that ectopic synapsis precedes ectopic unusual by virtue of the absence of nonallelic subunit, Pol32.
crossing-over. Am. J. Hum. Genet. 89, 580–588 homologous recombination hotspots and an apparent 90. Arlt, M. F., Rajendran, S., Birkeland, S. R., Wilson, T. E.
(2011). preference for female mitotic recombination. & Glover, T. W. De novo CNV formation in mouse
47. Lichten, M., Borts, R. H. & Haber, J. E. Meiotic gene Am. J. Hum. Genet. 81, 1201–1220 (2007). embryonic stem cells occurs in the absence of
conversion and crossing over between dispersed 69. Lam, K. W. & Jeffreys, A. J. Processes of copy-number Xrcc4‑dependent nonhomologous end joining.
homologous sequences occurs frequently in change in human DNA: the dynamics of α-globin gene PLoS Genet. 8, e1002981 (2012).
Saccharomyces cerevisiae. Genetics 115, 233–246 deletion. Proc. Natl Acad. Sci. USA 103, 8921–8927 91. Ira, G. & Haber, J. E. Characterization of
(1987). (2006). RAD51‑independent break-induced replication that
48. McKusick, V. A. Human genetics. Annu. Rev. Genet. 4, 70. Mezard, C., Pompon, D. & Nicolas, A. Recombination acts preferentially with short homologous sequences.
1–46 (1970). between similar but not identical DNA sequences Mol. Cell. Biol. 22, 6384–6392 (2002).
49. Vissers, L. E. & Stankiewicz, P. Microdeletion and during yeast transformation occurs within short 92. Hastings, P. J., Ira, G. & Lupski, J. R. A microhomology-
microduplication syndromes. Methods Mol. Biol. 838, stretches of identity. Cell 70, 659–670 (1992). mediated break-induced replication model for the origin
29–75 (2012). 71. Callinan, P. A. & Batzer, M. A. Retrotransposable of human copy number variation. PLoS Genet. 5,
50. Liu, P. et al. Mechanism, prevalence, and more severe elements and human disease. Genome Dyn. 1, e1000327 (2009).
neuropathy phenotype of the Charcot-Marie-Tooth 104–115 (2006). 93. Chen, J. M., Chuzhanova, N., Stenson, P. D., Ferec, C. &
type 1A triplication. Am. J. Hum. Genet. 94, 462–469 72. Boone, P. M. et al. Alu-specific microhomology- Cooper, D. N. Meta-analysis of gross insertions causing
(2014). mediated deletion of the final exon of SPAST in three human genetic disease: novel mutational mechanisms
51. Ou, Z. et al. Observation and prediction of recurrent unrelated subjects with hereditary spastic paraplegia. and the role of replication slippage. Hum. Mutat. 25,
human translocations mediated by NAHR between Genet. Med. 13, 582–592 (2011). 207–221 (2005).
nonhomologous chromosomes. Genome Res. 21, 73. Hsiao, M. C. et al. Decoding NF1 intragenic copy- 94. Smith, C. E., Llorente, B. & Symington, L. S. Template
33–46 (2011). number variations. Am. J. Hum. Genet. 97, 238–249 switching during break-induced replication. Nature
52. Robberecht, C., Voet, T., Zamani Esteki, M., (2015). 447, 102–105 (2007).
Nowakowska, B. A. & Vermeesch, J. R. 74. Deininger, P. L. & Batzer, M. A. Alu repeats and human Another important aspect of the mutagenic nature
Nonallelic homologous recombination between disease. Mol. Genet. Metab. 67, 183–193 (1999). of BIR was revealed in this work. BIR can undergo
retrotransposable elements is a driver of de novo 75. Stankiewicz, P., Pursley, A. N. & Cheung, S. W. multiple rounds of template switching during DSB
unbalanced translocations. Genome Res. 23, Challenges in clinical interpretation of microduplications repair; dispersed repetitive sequences were
411–418 (2013). detected by array CGH analysis. Am. J. Med. Genet. A observed to lead to non-reciprocal translocations.
53. Dittwald, P. et al. Inverted low-copy repeats and 152A, 1089–1100 (2010). 95. Tsaponina, O. & Haber, J. E. Frequent
genome instability—a genome-wide analysis. 76. Ottaviani, D., LeCain, M. & Sheer, D. The role of interchromosomal template switches during gene
Hum. Mutat. 34, 210–220 (2013). microhomology in genomic structural variation. conversion in S. cerevisiae. Mol. Cell 55, 615–625
54. Golzio, C. & Katsanis, N. Genetic architecture of Trends Genet. 30, 85–94 (2014). (2014).
reciprocal CNVs. Curr. Opin. Genet. Dev. 23, 240–248 77. Lee, J. A., Carvalho, C. M. & Lupski, J. R.  A DNA 96. Iraqui, I. et al. Recovery of arrested replication forks by
(2013). replication mechanism for generating nonrecurrent homologous recombination is error-prone. PLoS Genet.
55. Turner, D. J. et al. Germline rates of de novo meiotic rearrangements associated with genomic disorders. 8, e1002976 (2012).
deletions and duplications causing several genomic Cell 131, 1235–1247 (2007). 97. Sun, Z. et al. Replicative mechanisms of CNV formation
disorders. Nat. Genet. 40, 90–95 (2008). Long-range template switching leading to complex preferentially occur as intrachromosomal events:
In this paper the authors calculate the locus-specific genomic rearrangements and causing disease was evidence from Potocki-Lupski duplication syndrome.
ratio of deletions and duplications by NAHR in male reported. Fork stalling and template switching Hum. Mol. Genet. 22, 749–756 (2013).
meiosis. The observed higher ratio of deletions mechanisms were proposed for the first time to 98. Zhang, F., Carvalho, C. M. & Lupski, J. R. Complex
versus duplications, 2/1 for autosomes, correlates explain the observed unusual breakpoint complexity. human chromosomal and genomic rearrangements.
well with theoretical predictions. 78. Hastings, P. J., Lupski, J. R., Rosenberg, S. M. & Ira, G. Trends Genet. 25, 298–307 (2009).
56. Myers, S., Freeman, C., Auton, A., Donnelly, P. & Mechanisms of change in gene copy number. Nat. Rev. 99. Chanda, B. et al. A novel mechanistic spectrum
McVean, G. A common sequence motif associated with Genet. 10, 551–564 (2009). underlies glaucoma-associated chromosome 6p25 copy
recombination hot spots and genome instability in 79. Pannunzio, N. R., Li, S., Watanabe, G. & Lieber, M. R. number variation. Hum. Mol. Genet. 17, 3446–3458
humans. Nat. Genet. 40, 1124–1129 (2008). Non-homologous end joining often uses (2008).
57. Berg, I. L. et al. PRDM9 variation strongly influences microhomology: implications for alternative end joining. 100. Chauvin, A. et al. Elucidation of the complex
recombination hot-spot activity and meiotic DNA Repair (Amst.) 17, 74–80 (2014). structure and origin of the human trypsinogen locus
instability in humans. Nat. Genet. 42, 859–863 80. Liu, P., Carvalho, C. M., Hastings, P. J. & Lupski, J. R. triplication. Hum. Mol. Genet. 18, 3605–3614 (2009).
(2010). Mechanisms for recurrent and complex human genomic 101. Coccia, M. et al. X‑linked cataract and Nance-Horan
Variation within the genomic structure of PRDM9 rearrangements. Curr. Opin. Genet. Dev. 22, 211–220 syndrome are allelic disorders. Hum. Mol. Genet. 18,
was reported to correlate with the frequency of (2012). 2643–2655 (2009).
meiotic recombination in individuals, providing 81. Slack, A., Thornton, P. C., Magner, D. B., 102. Giorgio, E. et al. Analysis of LMNB1 duplications in
direct evidence for PRDM9 involvement in HR. Rosenberg, S. M. & Hastings, P. J. On the mechanism autosomal dominant leukodystrophy provides insights
58. Zhang, F. et al. Identification of uncommon recurrent of gene amplification induced under stress in into duplication mechanisms and allele-specific
Potocki-Lupski syndrome-associated duplications Escherichia coli. PLoS Genet. 2, e48 (2006). expression. Hum. Mutat. 34, 1160–1171 (2013).

NATURE REVIEWS | GENETICS VOLUME 17 | APRIL 2016 | 237


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

103. Rugless, M. J. et al. A large deletion in the human 125. Chen, L. et al. CNV instability associated with DNA inversion polymorphism. Nat. Genet. 38, 999–1001
α-globin cluster caused by a replication error is replication dynamics: evidence for replicative (2006).
associated with an unexpectedly mild phenotype. mechanisms in CNV mutagenesis. Hum. Mol. Genet. 151. Ballif, B. C. et al. Expanding the clinical phenotype
Hum. Mol. Genet. 17, 3084–3093 (2008). 24, 1574–1583 (2015). of the 3q29 microdeletion syndrome and
104. Bi, W. et al. Increased LIS1 expression affects human 126. Anand, R. P. et al. Chromosome rearrangements via characterization of the reciprocal microduplication.
and mouse brain development. Nat. Genet. 41, template switching between diverged repeated Mol. Cytogenet. 1, 8 (2008).
168–177 (2009). sequences. Genes Dev. 28, 2394–2406 (2014). 152. Mochizuki, J. et al. Alu-related 5q35 microdeletions
105. Liu, P. et al. Copy number gain at Xp22.31 includes 127. Korbel, J. O. et al. Paired-end mapping reveals in Sotos syndrome. Clin. Genet. 74, 384–391 (2008).
complex duplication rearrangements and recurrent extensive structural variation in the human genome. 153. Rosenfeld, J. A. et al. Further evidence of contrasting
triplications. Hum. Mol. Genet. 20, 1975–1988 Science 318, 420–426 (2007). phenotypes caused by reciprocal deletions and
(2011). 128. Kidd, J. M. et al. A human genome structural variation duplications: duplication of NSD1 causes growth
106. Abyzov, A. et al. Analysis of deletion breakpoints sequencing resource reveals insights into mutational retardation and microcephaly. Mol. Syndromol. 3,
from 1,092 humans reveals details of mutation mechanisms. Cell 143, 837–847 (2010). 247–254 (2013).
mechanisms. Nat. Commun. 6, 7256 (2015). 129. Lam, H. Y. et al. Nucleotide-resolution analysis of 154. Antonell, A. et al. Partial 7q11.23 deletions further
107. Deem, A. et al. Break-induced replication is highly structural variants using BreakSeq and a breakpoint implicate GTF2I and GTF2IRD1 as the main genes
inaccurate. PLoS Biol. 9, e1000594 (2011). library. Nat. Biotechnol. 28, 47–55 (2010). responsible for the Williams–Beuren syndrome
The mutagenic nature of BIR was shown in this yeast 130. Shimojima, K. et al. Pelizaeus-Merzbacher disease neurocognitive profile. J. Med. Genet. 47, 312–320
system; Polδ was shown to contribute in part to caused by a duplication-inverted triplication-duplication (2010).
these errors. in chromosomal segments including the PLP1 region. 155. Berg, J. S. et al. Speech delay and autism spectrum
108. Yousefzadeh, M. J. et al. Mechanism of suppression of Eur. J. Med. Genet. 55, 400–403 (2012). behaviors are frequently associated with duplication
chromosomal instability by DNA polymerase POLQ. 131. del Gaudio, D. et al. Increased MECP2 gene copy of the 7q11.23 Williams-Beuren syndrome region.
PLoS Genet. 10, e1004654 (2014). number as the result of genomic duplication in Genet. Med. 9, 427–441 (2007).
109. Northam, M. R., Garg, P., Baitin, D. M., Burgers, P. M. neurodevelopmentally delayed males. Genet. Med. 8, 156. Beunders, G. et al. A triplication of the Williams–
& Shcherbakova, P. V. A novel function of DNA 784–792 (2006). Beuren syndrome region in a patient with mental
polymerase ζ regulated by PCNA. EMBO J. 25, 132. Wolf, N. I. et al. Three or more copies of the retardation, a severe expressive language delay,
4316–4325 (2006). proteolipid protein gene PLP1 cause severe behavioural problems and dysmorphisms. J. Med.
110. Cherng, N. et al. Expansions, contractions, and fragility elizaeus–Merzbacher disease. Brain 128, 743–751 Genet. 47, 271–275 (2010).
of the spinocerebellar ataxia type 10 pentanucleotide (2005). 157. Cheroki, C. et al. Genomic imbalances associated
repeat in yeast. Proc. Natl Acad. Sci. USA 108, 133. McClintock, B. The behavior in successive nuclear with mullerian aplasia. J. Med. Genet. 45, 228–232
2843–2848 (2011). divisions of a chromosome broken at meiosis. (2008).
111. Sakofsky, C. J. et al. Translesion polymerases drive Proc. Natl Acad. Sci. USA 25, 405–416 (1939). 158. Nogueira, S. I. et al. Atypical 22q11.2 deletion in a
microhomology-mediated break induced replication 134. McClintock, B. The stability of broken ends of patient with DGS/VCFS spectrum. Eur. J. Med. Genet.
leading to complex chromosomal rearrangements. chromosomes in Zea Mays. Genetics 26, 234–282 51, 226–230 (2008).
Mol. Cell 60, 860–872 (2015). (1941). 159. Yamagishi, H., Garg, V., Matsuoka, R., Thomas, T.
112. Nik-Zainal, S. et al. Mutational processes molding the 135. Zuffardi, O., Bonaglia, M., Ciccone, R. & Giorda, R. & Srivastava, D. A molecular pathway revealing a
genomes of 21 breast cancers. Cell. 149, 979–993 Inverted duplications deletions: underdiagnosed genetic basis for human cardiac and craniofacial
(2012). rearrangements?? Clin. Genet. 75, 505–513 defects. Science 283, 1158–1161 (1999).
Kataegis, the phenomenon of localized (2009). 160. Potocki, L. et al. Characterization of Potocki-Lupski
hypermutational segments in cis, was defined in 136. Hannes, F. et al. Telomere healing following DNA syndrome (dup(17)(p11.2p11.2)) and delineation of a
cancer. polymerase arrest-induced breakages is likely the main dosage-sensitive critical interval that can convey an
113. Alexandrov, L. B. et al. Signatures of mutational mechanism generating chromosome 4p terminal autism phenotype. Am. J. Hum. Genet. 80, 633–649
processes in human cancer. Nature 500, 415–421 deletions. Hum. Mutat. 31, 1343–1351 (2010). (2007).
(2013). 137. Ghezraoui, H. et al. Chromosomal translocations in 161. Zhang, F. et al. Mechanisms for nonrecurrent genomic
114. Chung, W. H., Zhu, Z., Papusha, A., Malkova, A. & human cells are generated by canonical rearrangements associated with CMT1A or HNPP: rare
Ira, G. Defective resection at DNA double-strand breaks nonhomologous end-joining. Mol. Cell 55, 829–842 CNVs as a cause for missing heritability. Am. J. Hum.
leads to de novo telomere formation and enhances (2014). Genet. 86, 892–903 (2010).
gene targeting. PLoS Genet. 6, e1000948 (2010). 138. Ballif, B. C., Yu, W., Shaw, C. A., Kashork, C. D. & 162. Venturin, M. et al. Evidence for non-homologous end
115. Wilson, M. A. et al. Pif1 helicase and Polδ promote Shaffer, L. G. Monosomy 1p36 breakpoint junctions joining and non-allelic homologous recombination in
recombination-coupled DNA synthesis via bubble suggest pre-meiotic breakage-fusion-bridge cycles are atypical NF1 microdeletions. Hum. Genet. 115, 69–80
migration. Nature 502, 393–396 (2013). involved in generating terminal deletions. Hum. Mol. (2004).
116. Saini, N. et al. Migrating bubble during break-induced Genet. 12, 2153–2165 (2003). 163. Small, K., Iber, J. & Warren, S. T. Emerin deletion
replication drives conservative DNA synthesis. Nature 139. Luo, Y. et al. Diverse mutational mechanisms cause reveals a common X‑chromosome inversion
502, 389–392 (2013). pathogenic subtelomeric rearrangements. Hum. Mol. mediated by inverted repeats. Nat. Genet. 16, 96–99
117. Donnianni, R. A. & Symington, L. S. Break-induced Genet. 20, 3769–3778 (2011). (1997).
replication occurs by conservative DNA synthesis. 140. Yatsenko, S. A. et al. Human subtelomeric copy number 164. Fusco, F. et al. Genomic architecture at the
Proc. Natl Acad. Sci. USA 110, 13475–13480 (2013). gains suggest a DNA replication mechanism for Incontinentia Pigmenti locus favours de novo
118. Mayle, R. et al. Mus81 and converging forks limit the formation: beyond breakage-fusion-bridge for telomere pathological alleles through different mechanisms.
mutagenicity of replication fork breakage. Science 349, stabilization. Hum. Genet. 131, 1895–1910 (2012). Hum. Mol. Genet. 21, 1260–1271 (2012).
742–747 (2015). 141. Lowden, M. R., Flibotte, S., Moerman, D. G. &
Endonuclease Mus81 was shown to limit the Ahmed, S. DNA synthesis generates terminal Acknowledgements
mutagenic synthesis associated with BIR during duplications that seal end‑to‑end chromosome fusions. The authors thank C. R. Beck, S. Gu, P. Stankiewicz and
DNA repair. It also inhibits template switches Science 332, 468–471 (2011). P. J. Hastings for thoughtful comments and helpful discussions.
between interspersed homeologous repeats, 142. Hermetz, K. E. et al. Large inverted duplications in The authors apologize to colleagues and the authors of rele-
including human Alu. This work importantly adds to the human genome form via a fold-back mechanism. vant papers who could not be cited owing to space limitations.
our understanding of how genomic architecture PLoS Genet. 10, e1004139 (2014). The research conducted by the authors was supported in part
contributes to increased instability and which factors 143. Conrad, D. F. et al. Mutation spectrum revealed by by the US National Institutes of Neurologic Disorders and
have evolved to avoid this. breakpoint sequencing of human germline CNVs. Stroke (RO1NS058529), National Human Genome Research
119. Bacolla, A. et al. Breakpoints of gross deletions Nat. Genet. 42, 385–391 (2010). Institute/National Heart Blood Lung Institute jointly funded
coincide with non‑B DNA conformations. Proc. Natl 144. King, D. A. et al. Mosaic structural variation in children B ay l o r H o p k i n s C e n t e r fo r M e n d e l i a n G e n o m i c s
Acad. Sci. USA 101, 14162–14167 (2004). with developmental disorders. Hum. Mol. Genet. 24, (U54HG006542), National Institute of General Medical
120. Chen, J. M., Chuzhanova, N., Stenson, P. D., Ferec, C. 2733–2745 (2015). Sciences (RO1 GM106373), the Conselho Nacional de
& Cooper, D. N. Intrachromosomal serial replication 145. Poduri, A., Evrony, G. D., Cai, X. & Walsh, C. A. Somatic Desenvolvimento Científico e Tecnológico (CNPq)
slippage in trans gives rise to diverse genomic mutation, genomic variation, and neurological disease. (476217/2013‑0) and the Young Investigator fellowship
rearrangements involving inversions. Hum. Mutat. 26, Science 341, 1237758 (Science without Borders Program) grant 402520/2012‑2 to
362–373 (2005). (2013). C.M.B.C. The content is solely the responsibility of the authors
121. Bose, P., Hermetz, K. E., Conneely, K. N. & Rudd, M. K. 146. Merla, G. et al. Submicroscopic deletion in patients and does not necessarily represent the official views of the
Tandem repeats and G‑rich sequences are enriched at with Williams-Beuren syndrome influences expression NINDS, NHGRI/NHBLI, NIGMS or NIH.
human CNV breakpoints. PLoS ONE 9, e101607 levels of the nonhemizygous flanking genes.
(2014). Am. J. Hum. Genet. 79, 332–341 (2006). Competing interests statement
122. Walsh, E., Wang, X., Lee, M. Y. & Eckert, K. A. 147. Chaignat, E. et al. Copy number variation modifies The authors declare competing interests: see Web version
Mechanism of replicative DNA polymerase delta expression time courses. Genome Res. 21, 106–113 for details.
pausing and a potential role for DNA polymerase kappa (2011).
in common fragile site replication. J. Mol. Biol. 425, 148. Stranger, B. E. et al. Relative impact of nucleotide and
232–243 (2013). copy number variation on gene expression phenotypes. DATABASES
123. Northam, M. R. et al. DNA polymerases ζ and Rev1 Science 315, 848–853 (2007). OMIM: http://www.omim.org/
mediate error-prone bypass of non‑B DNA structures. 149. Shapiro, J. A. Molecular model for the transposition 117550 | 118220 | 162200 | 162500 | 182290 | 182601 | 188400 |
Nucleic Acids Res. 42, 290–306 (2014). and replication of bacteriophage Mu and other 192430 | 194050 | 300260 | 303800 | 303900 | 308100 | 308300 |
124. Koren, A. et al. Differential relationship of DNA transposable elements. Proc. Natl Acad. Sci. USA 76, 310300 | 312080 | 312920 | 608363 | 609425 | 609757 | 610443 |
replication timing to different forms of human mutation 1933–1937 (1979). 610883 | 611867 | 611936 | 612001 | 613675
and variation. Am. J. Hum. Genet. 91, 1033–1040 150. Koolen, D. A. et al. A new chromosome 17q21.31 ALL LINKS ARE ACTIVE IN THE ONLINE PDF
(2012). microdeletion syndrome associated with a common

238 | APRIL 2016 | VOLUME 17 www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.

You might also like