Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Hypermutation signature reveals a slippage

and realignment model of translesion synthesis by


Rev3 polymerase in cisplatin-treated yeast
Romulo Segoviaa, Yaoqing Shenb, Scott A. Lujanc, Steven J. M. Jonesb,d, and Peter C. Stirlinga,d,1
a
Terry Fox Laboratory, BC Cancer Agency, Vancouver, BC, Canada V5Z1L3; bMichael Smith Genome Sciences Centre, Vancouver, BC, Canada V5Z4S6;
c
Genome Integrity and Structural Biology Laboratory, National Institute of Environmental Health Sciences, National Institutes of Health, Research Triangle
Park, NC 27709; and dDepartment of Medical Genetics, University of British Columbia, Vancouver, BC, Canada V6T1Z3

Edited by Rodney Rothstein, Columbia University Medical Center, New York, NY, and approved February 3, 2017 (received for review November 8, 2016)

Gene–gene or gene–drug interactions are typically quantified using multiple efforts underway to inhibit additional DNA repair pro-
fitness as a readout because the data are continuous and easily teins themselves to further sensitize cancers to killing with geno-
measured in high throughput. However, to what extent fitness cap- toxic agents or to overwhelm tumor cells with already debilitated
tures the range of other phenotypes that show synergistic effects is DNA repair capacity (2).
usually unknown. Using Saccharomyces cerevisiae and focusing on a In cancer, as in model systems, there are typically surviving cells
matrix of DNA repair mutants and genotoxic drugs, we quantify 76 after a genotoxic insult. These cells bear a signature of mutations
gene–drug interactions based on both mutation rate and fitness and associated with the genotoxin they survived (5). Mutation signature
find that these parameters are not connected. Independent of fit- analysis of tumor genomes has been refined in the past 5 years, and
ness defects, we identified six cases of synthetic hypermutation, a set of 30 canonical mutation signatures is now maintained in the
where the combined effect of the drug and mutant on mutation Catalogue of Somatic Mutations in Cancer database (5, 6). In some

GENETICS
rate was greater than predicted. One example occurred when yeast cases, the etiology of mutation patterns is strongly linked to specific
lacking RAD1 were exposed to cisplatin, and we characterized this genetic mutations or environmental exposures, whereas in other
interaction using whole-genome sequencing. Our sequencing re- cases the etiology remains unknown. Studies in model organisms
sults indicate mutagenesis by cisplatin in rad1Δ cells appeared to have sought to dissect which aspects of a mutation signature are due
depend almost entirely on interstrand cross-links at GpCpN motifs. to specific deficiencies in genome maintenance factors or to specific
Interestingly, our data suggest that the following base on the tem-
chemical treatments (7–9). Indeed, the largest such study to date in
plate strand dictates the addition of the mutated base. This result
Caenorhabditis elegans characterized both a panel of mutant
differs from cisplatin mutation signatures in XPF-deficient Caeno-
strains and the effects of Aflatoxin B1, mechlorethamine, and
rhabditis elegans and supports a model in which translesion synthe-
cisplatin (9).
sis polymerases perform a slippage and realignment extension
The intention of genotoxin treatments clinically is to kill cells
across from the damaged base. Accordingly, DNA polymerase ζ ac-
rather than mutagenize them. Model organism studies have also
tivity was essential for mutagenesis in cisplatin-treated rad1Δ cells.
Together these data reveal the potential to gain new mechanistic
provided a means to map genetic networks underlying genotoxin
insights from nonfitness measures of gene–drug interactions and
sensitivity. The systematic identification of synthetic lethal in-
extend the use of mutation accumulation and whole-genome se- teractions or chemical–genetic interactions has been led by
quencing analysis to define DNA repair mechanisms. studies in budding yeast, Saccharomyces cerevisiae. Indeed, a full

|
mutator mutation signature | translesion synthesis | gene–drug Significance
|
interaction DNA repair
Cancer cells often have defects in DNA repair and are killed
effectively by drugs that damage DNA. However, surviving
G enome maintenance pathways suppress the accumulation of
mutations derived from chemical lesions or mismatches in
DNA that arise during normal metabolic processes. Despite thou-
cells can acquire additional mutations after treatment with
these genotoxic chemicals. Here we apply a simple model
sands of potentially mutagenic lesions occurring per cell per day, system to reveal synergy between specific DNA repair muta-
mitotic cell division exhibits extremely low rates of mutation (10−8– tions and genotoxic drugs that occurs independently of fitness
10−10 per bp per generation) under normal conditions (1). Cells with defects. Moreover, by analyzing the entire genome of a
defects in DNA repair pathways are generally more permissive for mutagenized cell population, we identify a signature of mu-
mutation accumulation (MA), and this phenomenon likely underlies tations that informs the mechanism of the translesion synthe-
the predisposition of individuals inheriting cellular DNA repair de- sis DNA damage tolerance pathway. Our work establishes a
fects to cancer formation (2). Increasing the rate of mutations in a conceptual framework for predicting the mutational burden of
cells surviving genotoxin treatment and adds to a growing list
cell population makes it more likely that the necessary mutations in
of examples supporting the utility of model organism mutation
oncogenes and tumor-suppressor genes will arise in a given lineage,
signature analysis for generating mechanistic insights.
leading to cellular transformation and proliferation (3).
Although DNA repair defects can predispose to cancer for- Author contributions: R.S. and P.C.S. designed research; R.S. performed research; S.A.L.
mation, when acquired somatically they may serve as an Achilles and S.J.M.J. contributed new reagents/analytic tools; R.S., Y.S., S.A.L., S.J.M.J., and P.C.S.
Heel that can be exploited for cancer therapy. This apparent di- analyzed data; and R.S. and P.C.S. wrote the paper.
chotomy is because chemotherapeutic chemicals often work by The authors declare no conflict of interest.
damaging DNA, stalling DNA replication, or disrupting mitosis, This article is a PNAS Direct Submission.
all of which could be potentially deleterious to a cell with an ac- Data deposition: The sequence reported in this paper has been deposited in the Sequence
quired DNA repair deficiency (2). Several well-documented ex- Read Archive database (accession no. SRP091984).
amples of this type of gene–drug interaction exist for the BRCA1 1
To whom correspondence should be addressed. Email: pstirling@bccrc.ca.
and 2 genes, where their mutational status can predict sensitivity This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
of tumors to cisplatin and its derivatives (4). Indeed there are now 1073/pnas.1618555114/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1618555114 PNAS | March 7, 2017 | vol. 114 | no. 10 | 2663–2668


pairwise gene–gene interaction study is now complete for both
MMR FA NER SSA
essential and nonessential yeast genes (10). In addition several BER HR MMR HR HR FA PRR TLS HR TLS HR NHEJ
thousand small molecules have been profiled for sensitivity and A



81
Δ

Δ
Δ

Δ
n1

o1

s1

u8
d1

d5

d3

d5
ph

v3
resistance across the yeast knockout (YKO) collections (11).

us
lh
WT

ap

ex

sg

yk
ra

ra

ra

ra

re
m

m
These approaches are being combined to understand the effects No drug
of chemical perturbations on genetic interaction networks and CPT
identify gene–gene synergies in drug sensitivity (12). In each of CIS
* * * * *
these studies, the primary readout for synergy between genes and ETP

chemicals is fitness, as it is quantitative, simple to measure in 5FU


*
high throughput, and informative. Nevertheless, other quantita- MMS
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1
* *
tive phenotypic readouts are possible, and the YKO collection
High fitness Low fitness
has been profiled by numerous biochemical, cytological, and



81
B

Δ
functional phenotypes (13).

Δ
Δ

Δ
n1

o1

s1

u8
d1

d5

d3

d5
ph

v3
us
lh
WT

ap

ex

sg

yk
ra

ra

ra

ra

re
m

m
Reasoning that DNA repair deficiencies would result in cell No drug
death, mutagenesis of survivors, or both after a genotoxic insult, CPT
we assessed the overlap of fitness and mutagenesis for repre- CIS
sentative chemical genotoxins in yeast cells defective for all * *
ETP
major DNA repair pathways. Quantifying growth and mutation 5FU
rates showed little overlap between these parameters and further * *
MMS
revealed cases of unexpected hypermutation. We predicted that 3.5 21 38 56 73 91 108 126 143 160
* *
there would be a pattern of mutations associated with hyper- Low rate High rate
mutagenesis and characterized that of rad1Δ and cisplatin by C 70

CAN1R rate (x10-7) per generation


whole-genome sequencing, uncovering evidence of a translesion 60
No cisplatin
synthesis (TLS) mechanism for yeast DNA polymerase ζ (Rev3). 10μM cisplatin
50
Together these data define gene–drug interactions, underscore a
mutation signature in yeast, and apply MA and whole-genome 40
sequencing to suggest DNA repair mechanisms. 30

20
Results
10
A Network of Genotoxin-Induced Fitness Defects and Mutation Rates.
To investigate gene–drug interactions, we first established a panel 0
WT rad1Δ rad2Δ rad23Δ rad26Δ
of DNA repair mutants representing the major DNA repair Nucleotide Excision Repair mutant
pathways in yeast (Table S1). Haploid yeast bearing gene deletions
impairing homologous recombination (HR), nonhomologous end Fig. 1. Gene–drug interactions measured by fitness and mutation rates.
Heat maps illustrating (A) fitness and (B) CAN1 mutation rate relative to WT.
joining (NHEJ), nucleotide-excision repair (NER), base-excision
Interactions significantly greater than expected (P < 0.05) are indicated *.
repair (BER), mismatch repair (MMR), TLS, Fanconi Anemia- The transition from yellow to blue indicates greater fitness defects or higher
like (FA), and postreplication repair (PRR) pathways were chosen mutation rates. The first yellow box is set at the WT rate; any lower rates
to represent loss or deficiencies of each pathway. To account for (i.e., in some rev3Δ- and 5FU-treated conditions) are given the same color.
the multiple steps of, and routes for, DNA double-strand break Black boxes for rad5Δ indicate not done. (C) CAN1 mutation rates of other
(DSB) repair by HR, we also included exo1Δ, rad52Δ, sgs1Δ, NER-deficient strains in cisplatin. rad1Δ and rad2Δ exhibit SHyp (P < 0.05).
mus81Δ, and mph1Δ lines, as these mutants specifically impair
end resection, strand exchange, double Holliday junction disso-
lution, resolution, or other HR steps, respectively (14). This panel low concentrations of drugs were used to allow comparison with
of wild-type (WT) and 12 mutant strains (Fig. 1) was exposed to fluctuation assays below.
five classes of DNA damaging agents, represented by the bi- Mutation rates of the same matrix were quantified at CAN1
functional alkylator cisplatin (Cis), the antimetabolite fluorouracil using a well-plate fluctuation assay (Fig. 1B and Table S2) (15). In
(5FU), the topoisomerase I inhibitor camptothecin (Cpt), the untreated cells, the baseline mutation rates matched previously
topoisomerase II inhibitor etoposide (Etp), and the monofunc- reported rates (Table S3) (7, 15–20). Again we observed that Cpt
tional alkylator methyl methanesulfonate (MMS). We screened 78 and Etp had no major mutator effects regardless of the genetic
pairwise gene–drug combinations plus WT for changes in fitness background at the given drug concentrations. On the contrary,
and, in 76 cases, were also able to measure mutation rates (rad5Δ cisplatin, 5FU, and MMS increased the mutation rates of specific
growth was impaired to a degree that prevented fluctuation mutants. When this increase in mutation rate exceeded the
analysis). expected effects of multiplying the effect of the repair deficiency
Growth was measured over 24 h, and the area under the curve and the effect of the chemical on WT, we defined it as Synthetic
was calculated and normalized to the untreated WT to measure Hypermutation (SHyp). We chose to apply the multiplicative
fitness (Fig. 1A). Overall, we found that Cpt and Etp had no model, both because it performs best across a range of genetic
significant additional effects on any strain using the concentrations interactions based on fitness (21) and because it creates a more
tested, and although 5FU inhibited growth of all lines irrespective stringent definition of a SHyp interaction (we compare with an
of the genetic background, it only showed a synergistic interaction additive definition in Fig. S1). Applying this metric, we identified
with apn1Δ. We also observed that cisplatin had synergistic gene– six cases of SHyp: mph1Δ-Cis, rad1Δ-Cis, apn1Δ-5FU, mlh1Δ-
drug interactions with rad1Δ, rad5Δ, rad52Δ, rev3Δ, and sgs1Δ and 5FU, exo1Δ-MMS, and mph1Δ-MMS (Fig. 1B and Fig. S1). Of
that MMS had synergistic interactions with rad5Δ and sgs1Δ (P < these cases, rad1Δ-Cis, apn1Δ-5FU, and mlh1Δ-5FU showed the
0.05; Fig. 1A). These interactions are consistent with, for example, most severe SHyp phenotype (17.8-, 45.1-, and 33.4-fold increase
the known roles for BER in repair of uracil in DNA or the known over WT, respectively). Importantly, cases of SHyp did not overlap
complexity of repairing cisplatin interstrand cross-links (ICLs), with fitness defects except for rad1Δ-Cis and apn1Δ-5FU, and
which involves at least HR, NER, FA, and TLS. Moreover, this conversely many strains with fitness defect exhibited no increase in
screen possibly underrepresents chemical sensitivities because very mutation rate. Indeed, survival analysis of cisplatin sensitivity of

2664 | www.pnas.org/cgi/doi/10.1073/pnas.1618555114 Segovia et al.


sgs1Δ, mus81Δ, and rad1Δ shows nearly identical viability curves A Single-cell bottlenecks
wt, rad1Δ/Δ
over cisplatin concentration but different mutational outcomes Mutation G1000
(Fig. S2). Together our data show that decreased fitness and in- accumulation
12x
creased mutability by genotoxins in DNA repair-deficient cells are Repeat 40x ... WGS
not perfectly linked. This observation is likely both an issue of dose Genotoxic Single clone
and DNA repair mechanisms wherein some gene–drug combina- 12x expansion
treatment
tions are more likely to elicit toxic or irreparable intermediates. wt, rad1Δ/Δ, 100μM
sgs1Δ/Δ, cisplatin/2hrs
mus81Δ/Δ Mutations
Cisplatin Hypermutation in Specific NER-Deficient Contexts. One of
the outcomes of identifying unexpected increases in mutation rate B C TpT

Dinucleotide substitutions
160 14 TpG
WT TpC
is the potential to use these observations to uncover previously rad1Δ/Δ 12

Number of SNVs
TpA
unrecognized modes of mutagenesis. The strongest SHyp inter- 120 sgs1Δ/Δ 10 GpC
mus81Δ/Δ CpG
actions, apn1Δ-5FU and mlh1Δ-5FU, likely reflect known links of 8 ApT
80
BER and MMR to removing fluorouracils paired with A or G 6 ApG
ApC
from DNA (22). Whether a specific mutagenesis mode and sig- 40
4 ApA
nature was responsible for SHyp in rad1Δ-Cisplatin was less clear, 2
and we decided to pursue analysis of this interaction further. Rad1 0 0
C>A C>G C>T T>A T>C T>G CpACpGCpT GpC GpT
is a nuclease involved in cleaving DNA flanking bulky lesions Type of base substitution Reference sequence
during NER and ICL repair. Recognition of these lesions can 30
D
occur during transcription or globally due to the action of up-

Number of indels
25
stream NER components that recognize distorted DNA helices 20 Insertions
(23). To determine which aspects of NER exhibit SHyp with cis- Deletions
15
platin, we tested mutation rates in three other NER mutants—
10
rad2Δ, rad26Δ, and rad23Δ. Deletion of RAD2, which, like Rad1,
5

GENETICS
encodes a nuclease required for DNA incision and is related to
human XPG, also exhibited SHyp with cisplatin (Fig. 1C). How- 0
ever, deletion of RAD26, which selectively impairs transcription-
coupled (TC) NER, or RAD23, which partially impairs global and
TC-NER but allows efficient DNA incision, had no SHyp phe- rad1Δ/Δ (cisplatin) rad1Δ/Δ (MA) WT (MA)
notype with cisplatin (Fig. 1C). Because Rad1 acts first and Rad2 Indel pattern
is required in a structural role to position Rad1 (23), loss of either
would block incision; consequently, alternative mechanisms of Fig. 2. MA and whole-genome sequencing. (A) Experimental design of
unhooking of the cisplatin damage must be error prone. drug treatment and 1,000 generation (G1000) approaches for the indicated
strains (WGS, whole-genome sequencing). (B) Summary of WGS results of
MA and Whole-Genome Sequencing. It is possible that CAN1 showed SNVs in cisplatin-treated WT, rad1Δ, sgs1Δ, and mus81Δ homozygous dip-
loid strains. SNV number and type are summarized in the bar graphs color-
variable mutation rates in repair-deficient strains due to a locus-
coded as indicated. (C) Dinucleotide substitutions in rad1Δ-Cis genomes. The
specific effect. To explore the genomic spectrum of mutations and parental dinucleotide sequences are on the x axis, and the mutated se-
identify any evidence of such bias, we pulsed diploid WT, rad1Δ, quence is represented by the color of the bar. (D) Pattern of small indels in
sgs1Δ, and mus81Δ with a higher dose of cisplatin and sequenced rad1Δ-Cis and MA genomes. Indel size and sequence are on the x axis, and
the whole genomes of 12 independent survivors for each strain (Fig. indel types are color-coded.
2A). Consistent with the CAN1 fluctuation analysis data, the rad1Δ
mutant strain acquired significantly more mutations than WT or the
other mutants (Fig. 2B). As expected, based on the preference of C. elegans RAD1)-deficient animals (9). We extracted flanking se-
cisplatin for cross-linking guanine residues, the most common mu- quences for the predominant C > N mutations in cisplatin-treated
tation types were single-nucleotide variants (SNVs) at C:G base genomes and performed pLogo and enrichment analysis (Fig. 3A
pairs. Moreover, indels seen in rad1Δ, which were primarily –1 and Fig. S4A) (25). This analysis revealed that GpCpT motifs were
deletions, occurred at C:G base pairs, and rare dinucleotide sub- favored when all C > N mutations were considered. Because cis-
stitutions were also evident at motifs containing C:G base pairs (Fig. platin is expected to preferentially cross-link guanine residues, this
2 C and D). To ensure that the rate and type of mutations were due signature suggests that ICLs are the mutagenic lesion in yeast RAD1
to cisplatin and not simply rad1Δ, we passaged diploid WT and mutants. The observed GpCpT motif differs from C. elegans, where
rad1Δ cells through 40 single-cell bottlenecks and sequenced 12 CpCpT motifs were preferentially mutated in xpf-1 worms (9).
independent whole genomes after this 1,000 generation MA ex- Because intrastrand cross-links are much more common than ICLs
periment (Fig. 2A). The mutation rate and type evident in the (26), our data suggest either that rad1Δ yeast are able to repair
rad1Δ MA lines were similar to the CAN1 rate in untreated rad1Δ these in an error-free manner, even when lacking a key NER pro-
cells and showed proportionally more mutations at T:A base pairs tein, or that they die and are removed from the experiment. Further
and no dinucleotide substitutions (Fig. S3). No correlations between pLogo analysis with fixed positions at GC showed that, for each
rad1Δ-Cis mutations with transcription, replication timing, or mutation type, adenine was significantly enriched at the –2 position
leading-strand character were seen, although a weakly significant (i.e., ApGpCpT) (Fig. 3A, Center and Fig. S4A). When we analyzed
enrichment (P = 0.0103) was seen between nucleosome-bound the 3′ position, a different picture emerged: The mutation type
DNA and mutations (24). Thus, the specific combination of RAD1 varied with the +1 base, such that C > T mutations preferentially
deletion and cisplatin treatment engages an error-prone repair occurred in GpCpT motifs, C > A mutations in GpCpA motifs, and
process that causes a global increase in specific types of mutations. C > G mutations in GpCpG motifs (Fig. 3B and Fig. S4B). Thus,
the base that is erroneously inserted at the damaged site is selected
Cisplatin-rad1Δ Mutation Signatures Suggest Templated Base by base-pairing with the 5′ base on the opposite template strand.
Substitutions. The flanking sequences around mutations can de- Interestingly, mutations at GpCpC, which would be error free if
fine a signature that is associated with a particular genetic or using the 5′ template-strand base for synthesis, occurred at only 13
chemical perturbation. Indeed, previous work in C. elegans has ex- of 327 C > N mutations, much less frequently than for GpCpT,
amined cisplatin treatment of several mutants including XPF-1 (i.e., GpCpA, and GpCpG. There was also no evidence that indels

Segovia et al. PNAS | March 7, 2017 | vol. 114 | no. 10 | 2665


A C>N without TLS polymerase deletions (28). Remarkably, downstream
n = 327
12 templating was observed at 43% of SNVs, deletion of RAD30 had no

Fold enrichment
over
over

10 effect on this frequency, whereas REV3 deletion reduced the fre-


8
+3.51 +3.49
6 quency of SNVs matching the downstream base to 29% (Fig. 4C).
-3 -2 -1 0 +1 +2 +3 0 -3 -2 -1 0 +1 +2 +3 0
-3.51 -3.49
4 Furthermore, analysis of another published Rev3 mutation set
under

under
2 brought on by a mutant replisome (pol3-Y708A) (29) also showed a
0
pattern of templated insertions [i.e., in REV3 37% of SNVs matched

CT
G C
AGCT
G
the 3′ base (n = 126), whereas in rev3Δ 22% of SNVs matched the 3′
B C>A
n = 133 base (n = 85), P = 0.0431 Fisher’s exact test]. Together these data
8
lead us to propose a model in which slippage and realignment of the

Fold enrichment
over 7
over

+3.49
6
5
DNA template by Rev3 led to the observed mutation signature and
+3.51 4 may be generalizable to both endogenous lesions and possible also
-3 -2 -1 0 +1 +2 +3 0 -3 -2 -1 0 +1 +2 +3 0 3
-3.51 2 when Rev3 is called to replicate undamaged DNA (Fig. 4D and
under

under

1
-3.49 0 Discussion).

CA
AGCA
G C
G
Discussion
C>G
n = 39 SHyp as an Alternative Measure of Phenotypic Enhancement in

Fold enrichment
10
+3.49 Genome Maintenance. As fitness datasets on gene–gene and gene–
over
over

8
+3.51
6 drug interactions accumulate (10, 11), measurements of trigenic
-3 -2 -1 0 +1 +2 +3 0 -3 -2 -1 0 +1 +2 +3 0 4 interactions (30) or the ways in which chemicals shift genetic-
2
interaction networks (12) are enhancing the wiring diagram of the
under

under

-3.51 0
-3.49
CG
AGCG
G C
cell. Damage repair and mutation is key to the chemotherapeutic
G

action of alkylating agents, and changes in DNA damage repair can


C>T have profound effects on therapeutic efficacy. Although there is
n = 155
7 prior evidence that ultrahigh mutation rates can drive severe fitness
Fold enrichment
over
over

6
5 consequences, even lethality (31), the overlap of these phenotypes
4
-3 -2 -1 0
+3.51
+1 +2 +3 0 -3 -2 -1 0
+3.49
+1 +2 +3 0 3 broadly is not known. By systematically ablating the major DNA
-3.51 -3.49 2 repair pathways in yeast and measuring the effects of diverse gen-
under

under

1
0 otoxic chemicals, our data suggest that fitness and chemically in-
CT
G C
AGCT

duced hypermutation are not necessarily linked.


G

SHyp interactions between 5FU and mlh1Δ and apn1Δ are


Fig. 3. Mutation signature analysis induced by cisplatin in rad1Δ diploid largely consistent with the role of MMR and BER, respectively,
yeast. (A) Flanking nucleotides for C > N substitutions are presented as
pLogo plots and fold enrichment. The mutated C is centered in each plot
with fixed positions at C (Left) and GC (Center). The red line indicates sig-
nificant enrichment at P < 0.05, and the font size indicates the magnitude of A C

% Potential templated SNVs


Expected fitness p=0.033
enrichment. Overrepresented motifs (over) are above, whereas un- Observed fitness in 10μM cisplatin
50%
Relative fitness compared to WT

0.009 n.s.
derrepresented (under) are below. The increasing fold enrichment for 2–3- 1.2 40%
0.030
and 4-nucleotide motifs containing a mutated C is shown (Right). All analysis 0.000
30%
1
was performed with a set of 41 mers and cropped for visualization (Materials
20%
and Methods). (B) Flanking analysis for C > A, C > G, and C > T mutations as 0.8 0.002
10%
presented in A. Note the +1 base changing with the mutation type. 0.6
0%
rad14Δ rad14Δ rad14Δ
0.4 rad30Δ rev3Δ
occurred at GpCpC motifs, as only 2 of 12 deletions at GpCpN 0.2
D 5′ GCT 3′
occurred at GpCpC (within Dataset S1). These observations further 0 3′ CGA 5′
WT rad1Δ rad1Δ rad1Δ rad1Δ non-canonical
support a model of templated TLS as discussed below. rad30Δ rev1Δ rev3Δ unhooking ? ?

DNA Polymerase ζ Drives Rad1Δ-Cisplatin SHyp. Our mutation sig-


B 5′
3′ C A
3′
5′
80 G
nature analysis suggests error-prone DNA replication is likely caus- G
70
ing mutations in rad1Δ cells treated with cisplatin. To identify the
CAN1 mutation rate*

60 slippage
relevant polymerase, we deleted all enzymatic TLS components in- 5′ G·T 3′
50 3′ 5′
dividually in the rad1Δ background. Loss of RAD30, which slightly C A
40 G Polζ
positively affected fitness compared with the rad1Δ single mutant, G
30
showed no effect on rad1Δ hypermutation (Fig. 4 A and B), however
20 realignment
additional loss of REV1 or REV3 led to inviability when treated with 5′ GT T 3′
10
10 μM cisplatin (Fig. 4A). This observation is consistent with the 3′ C A 5′
0 G Polζ
DNA damage tolerance role of REV3 and the idea that TLS is both WT rad1Δ rad1Δ WT rad1Δ rad1Δ rad1Δ
rad30Δ rev1Δ rev3Δ G
supporting viability in RAD1-deficient cells and causing a hyper- No cisplatin 10μM cisplatin 5μM cisplatin
mutation signature, as was previously reported for complex frame-
shifts at spontaneous DNA lesions (27). A lower dose of cisplatin Fig. 4. Role of TLS in hypermutation of cisplatin-treated rad1Δ cells.
still revealed SHyp with rad1Δ but showed complete suppression of (A) Fitness measurements and (B) CAN1 mutation rates (* × 10−7 per genera-
SHyp in rad1Δrev3Δ and rad1Δrev1Δ cells (Fig. 4B), indicating that tion) in rad1Δ cells lacking the indicated TLS polymerase subunit. (C) Frequency
of SNVs identical to 3′ base in URA3 from the indicated strains from ref. 28.
Polζ TLS was responsible for the mutations. To determine if en-
(D) Model for slippage and realignment of DNA polymerase ζ (Rev3) at a cis-
dogenous DNA lesions could promote 5′-templated mutations (i.e., platin-damaged base. Noncanonical and presumably not optimal unhooking
+1 3′ on the strand being replicated by the 5′ base on the opposite of the ICL creates a substrate for error-prone TLS that promotes slippage of
template strand), we scored a published dataset of URA3 mutations Polζ at the lesion and subsequent realignment upon lesion bypass leading to a
sequenced in unperturbed NER-deficient (rad14Δ) cells with or templated SNV using the 5′ base on the template strand to insert the SNV.

2666 | www.pnas.org/cgi/doi/10.1073/pnas.1618555114 Segovia et al.


in removing uracil and 5-fluorouracil from DNA (22). Although yeast N-glycosylases, such as Ntg1 and Ntg2, could act on cisplatin
whether these are separate roles in repairing different lesions or cross-links in the absence of the preferred Rad1 pathway and
due to potential cooperation of MMR and BER is unclear (22, create a unique chemical moiety that drives slippage-realignment
32). Regardless, these data show that 5FU can overwhelm the TLS by Polζ in vivo.
repair capacity of MMR or BER-deficient cells without addi-
tionally affecting the fitness of MMR impaired cells. These data SHyp and Cancer. The framework created here sought to establish
are consistent with evidence that 5FU has multiple modes of the gene–drug relationships driving hypermutation in cells sur-
action interfering with both DNA and RNA metabolism (22). On viving genotoxin treatment beginning with canonical DNA repair
the other hand, although both NER and HR are required for pathways and drugs of clinical relevance. There remains a large
repair of cisplatin ICLs and mutants in these pathways had fit- space to be explored; in yeast, there are many hundreds of genes
ness defects, rad1Δ but not sgs1Δ or other major HR mutants whose loss of function causes genome instability and hundreds
showed SHyp with cisplatin. This difference suggests that al- of others whose increased dosage causes genome instability (8,
though survival in rad1Δ cells is supported by an error-prone 41, 42). The ways in which these will combine with each other
mechanism, that same pathway is insufficient to support sgs1Δ and with various chemicals to permit mutagenesis remain
cells, which either die, likely due to toxic recombination inter- incomplete.
mediates, or repair cisplatin damage in an error-free way. Hypermutation of surviving cancer cells after genotoxic che-
motherapy could be viewed as negative if it permits the acquisition
A Mutation Signature of TLS Polymerase Slippage and Realignment. of chemoresistant mutations. However, current evidence suggests
Whole genome sequencing analysis of mutagenized model organism that resistance mutations are often preexisting in tumor cell
genomes is a powerful way to link mutation signatures to specific populations, and their emergence is a clonal selection process
causal phenomena (33). This approach has been used to shed new rather than chemotherapy-induced mutations (43). Indeed, even
light on DNA repair mechanisms (7), the effects of genotoxic for therapy-related leukemias, p53 mutations driving leukemo-
chemicals (9, 34), and the role of cytidine deaminase enzymes in genesis were recently found to be preexisting when chemotherapy
cancer (25). Characterization of our sequencing data revealed that in occurred (44). Thus, SHyp may not be a major concern for ac-
rad1Δ yeast ICLs were the predominant source of mutations, as quiring chemoresistant mutations when treating repair-deficient

GENETICS
most variants were found at GpC motifs. This signature differs from cancers with genotoxic chemotherapies. Although chemotherapy
the cisplatin mutation signature of both C. elegans and humans, has been shown to minimally impact chemoresistant mutations,
where intrastrand cross-links seem to drive mutagenesis (i.e., muta- the DNA-damaging nature of chemotherapeutic agents can in-
tions at CpC motifs) (9, 35). There are several possible reasons for crease the overall mutational load (35). We propose that this
this difference, including the complexity of the FA pathway for ICL hypermutation of surviving cells could ultimately be beneficial in
repair in humans that is only partly represented in yeast or the high contexts where an immune response to tumor neo-antigens is
divergence of REV3 itself (i.e., human REV3L protein is >twice as desired, as is the case for therapies targeting immune checkpoints,
long and exhibits only 25% similarity to yeast Rev3). Although Rev1 where higher mutational loads are known to improve outcomes
plays an important structural role, we predict Rev3 is catalyzing the (45). Understanding more about the interactions of genotoxic
mutations that we observed in our study, as the catalytic role of Rev1 chemotherapies with genome maintenance defects in cancer could
to insert C opposite the damaged guanine would be error-free and ultimately support interventions where it is desirable to increase
thus not detectable in our assay. However, another possibility is that the mutational load or reshape the mutational profile of cancer
intrastrand cross-links are not repaired in rad1Δ cells and lead to cell cells following frontline therapy.
death, effectively masking the mutations seen at these sites in other
systems. We also found evidence of Rev3 slippage and realignment Materials and Methods
in mutations from untreated cells, suggesting that endogenous, Yeast Strains and Growth Curve Analysis. Yeast strains are listed in Table S4.
noncisplatin-induced lesions can promote the same mechanism (28). Yeast were grown on SD minimal media plus histidine, uracil, leucine, lysine,
and methionine + 2% (wt/vol) dextrose (SD+5) or YPD at 30 °C. Growth
Our mutation signature analysis revealed not only a different
curves were conducted as described (41). The observed fitness, by area under
type of mutagenic lesion in yeast but also a mechanism involving the curve, was compared with the expected fitness value using a t test (41).
Polζ, Rev3, which indicates a slippage and realignment event Unless otherwise indicated, drug concentrations used were 1 μM Cpt, 10 μM
during TLS. Previous work in vitro has identified this behavior in cisplatin, 200 μM Etp, 10 μM 5FU, and 0.0005% MMS.
another TLS polymerase, human DNA polκ (36). Indeed, this
mechanism may also be similar to that proposed for Dpo4 based Fluctuation Analysis and SHyp Analysis. Fluctuation analysis at CAN1 was
on structural and in vitro studies (37). The concept of slippage as performed with ≥18 replicates per condition as described (15), except that
a source of replication errors has been around for many years, cells were grown to saturation in the presence of drugs in SD+5 for 48 h at
although never, to our knowledge, applied to slippage and re- 30 °C. Cell numbers were determined using a TC-20 cell counter (BioRad),
alignment in B-family DNA polymerase (38). We propose that and CANR colony counts were converted to mutation rates using the Ma–
Sandri–Sarkar maximum-likelihood estimator calculator (46). To calculate
(Fig. 4D), at least in the specific context of cisplatin adducts, the
expected mutation rates, we multiplied or added (Fig. S1) the fold increase
guanine lesion is bypassed by Rev3 in a manner that uses Wat- of a gene deletion or a chemical on WT cells over untreated WT cells, then
son–Crick base-pairing with the undamaged template strand expressed this value as CANR×10−7 per generation by multiplying by the
base. This phenomenon could lead to an indel, but more often untreated WT rate. The observed rates and associated confidence intervals
the newly inserted base shifts to mispair with the damaged G were measured as described (46). Expected values below the 95% confi-
residue and is extended with the correct base at the 3′ position dence interval of the observed mutation rate were considered candidate
on the newly synthesized strand. Whether suboptimal DNA in- SHyp interactions.
cision by compensatory nucleases favors this misincorporation
event is an active area of research. Indeed, the nature of DNA MA and Whole-Genome Sequencing. Diploid strains were pulsed with 100 μM
incision in the absence of Rad1 is unclear, although structure- cisplatin for 2 h before plating on YPD. Diploids were used to buffer delete-
rious mutations, and a pulse of cisplatin ensured that mutations altering the
specific nucleases such as Slx4, regulated by Slx1, Mus81, and its
mutation spectrum did not accumulate. For MA lines, single colonies were
partner Mms4, and Yen1 have activities that could in principle plated on YPD and passaged 40 times by restreaking to fresh YPD every 2–3 d
complement Rad1 loss at some structures (39). Alternatively, (∼1,000 generations). Twelve independent lines for each condition were out-
new data in Xenopus extracts implicate the N-glycosylase ac- grown for genomic DNA extraction (8). DNA was subjected to 125 bp paired-
tivity of NEIL3, an upstream BER factor, in unhooking psor- end whole-genome sequencing at Canada’s Michael Smith Genome Sciences
alen and other ICLs (40). Therefore, it is also possible that Centre using the Illumina HiSeq2000 platform. The average sequencing

Segovia et al. PNAS | March 7, 2017 | vol. 114 | no. 10 | 2667


coverage compared with the haploid reference was 129× (i.e., >60× for each distribution. pLogos were generated online at plogo.uconn.edu/ (50). As
diploid genome). Sequence files are deposited at the Sequence Read Archive foreground input n(fg), 41-mer sequences from the rad1Δ-cisplatin genomes
(www.ncbi.nlm.nih.gov/sra) (accession no. SRP091984). Reads were aligned to were used. These sequences were centered at the mutated C for all substitu-
the Saccer3 reference genome using Burrow–Wheeler Aligner 0.5.7 (47). Var- tions with fixed positions at C and GC. Unique 41-mer sequences automatically
iants were identified exactly as described previously (8). Mutations were generated by pLogo from the Saccer3 yeast genome were used as background
extracted by comparing evolved to parental genomes. Variant calling was input n(bg). Reanalysis of literature variants was performed by manually
carried out using the mpileup utility of the SAMtools package 0.1.18 (48) with scoring whether the SNV matched the base immediately 3′ of the mutation
a threshold of 20. Copy-number variants were analyzed using a locally modi- (28, 29). The proportions of these templated SNVs were compared with a two-
fied version of CNAseq (49). Additionally, we manually verified variant calls tailed Fisher exact test.
with the Integrated Genomics Viewer (Broad Institute).
ACKNOWLEDGMENTS. We thank N. O’Neil and the P.C.S. laboratory for
Mutation Signature Analysis. Flanking sequence was extracted from the reading the manuscript and Philip Hieter for strains. The work is funded
Saccer3 genome based on the mutation positions (Dataset S1). The enrichment by Canadian Institutes of Health Research (CIHR) Grant MOP-136982, Nat-
of motifs within the diploid rad1Δ-cisplatin genomes was calculated exactly as ural Sciences and Engineering Research Council of Canada Grant RGPIN
described based on a set of 41 mers (±20 bp around the SNV) (25). The sig- 2014-04490, and the British Columbia Cancer Foundation. P.C.S. is a CIHR
nificance of motif enrichments was calculated using the hypergeometric New Investigator.

1. Farlow A, et al. (2015) The spontaneous mutation rate in the fission yeast Schizo- 26. Eastman A (1987) The formation, isolation and characterization of DNA adducts
saccharomyces pombe. Genetics 201(2):737–744. produced by anticancer platinum complexes. Pharmacol Ther 34(2):155–166.
2. Lord CJ, Ashworth A (2012) The DNA damage response and cancer therapy. Nature 27. Harfe BD, Jinks-Robertson S (2000) DNA polymerase zeta introduces multiple muta-
481(7381):287–294. tions when bypassing spontaneous DNA damage in Saccharomyces cerevisiae. Mol
3. Stratton MR, Campbell PJ, Futreal PA (2009) The cancer genome. Nature 458(7239): Cell 6(6):1491–1499.
719–724. 28. Stone JE, Lujan SA, Kunkel TA, Kunkel TA (2012) DNA polymerase zeta generates
4. Lord CJ, Ashworth A (2013) Mechanisms of resistance to therapies targeting BRCA- clustered mutations during bypass of endogenous DNA lesions in Saccharomyces
mutant cancers. Nat Med 19(11):1381–1388. cerevisiae. Environ Mol Mutagen 53(9):777–786.
5. Alexandrov LB, et al.; Australian Pancreatic Cancer Genome Initiative; ICGC Breast 29. Northam MR, Robinson HA, Kochenova OV, Shcherbakova PV (2010) Participation of
Cancer Consortium; ICGC MMML-Seq Consortium; ICGC PedBrain (2013) Signatures of DNA polymerase zeta in replication of undamaged DNA in Saccharomyces cerevisiae.
mutational processes in human cancer. Nature 500(7463):415–421. Genetics 184(1):27–42.
6. Forbes SA, et al. (2010) COSMIC (the Catalogue of Somatic Mutations in Cancer): A 30. Braberg H, et al. (2014) Quantitative analysis of triple-mutant genetic interactions.
resource to investigate acquired mutations in human cancer. Nucleic Acids Res Nat Protoc 9(8):1867–1881.
38(Database issue):D652–D657. 31. Herr AJ, Kennedy SR, Knowels GM, Schultz EM, Preston BD (2014) DNA replication
7. Serero A, Jubin C, Loeillet S, Legoix-Né P, Nicolas AG (2014) Mutational landscape of error-induced extinction of diploid yeast. Genetics 196(3):677–691.
yeast mutator strains. Proc Natl Acad Sci USA 111(5):1897–1902. 32. SenGupta T, et al. (2013) Base excision repair AP endonucleases and mismatch repair
8. Stirling PC, Shen Y, Corbett R, Jones SJ, Hieter P (2014) Genome destabilizing mutator act together to induce checkpoint-mediated autophagy. Nat Commun 4:2674.
alleles drive specific mutational trajectories in Saccharomyces cerevisiae. Genetics 33. Segovia R, Tam AS, Stirling PC (2015) Dissecting genetic and environmental mutation
196(2):403–412. signatures with model organisms. Trends Genet 31(8):465–474.
9. Meier B, et al. (2014) C. elegans whole-genome sequencing reveals mutational sig- 34. Tam AS, Chu JS, Rose AM (2015) Genome-wide mutational signature of the chemo-
natures related to carcinogens and DNA repair deficiency. Genome Res 24(10): therapeutic agent mitomycin C in Caenorhabditis elegans. G3 (Bethesda) 6(1):
1624–1636. 133–140.
10. Costanzo M, et al. (2016) A global genetic interaction network maps a wiring diagram 35. Huang KK, et al. (2016) Exome sequencing reveals recurrent REV3L mutations in cis-
of cellular function. Science 353(6306):aaf1420.
platin-resistant squamous cell carcinoma of head and neck. Sci Rep 6:19552.
11. Lee AY, et al. (2014) Mapping the cellular response to small molecules using che-
36. Mukherjee P, Lahiri I, Pata JD (2013) Human polymerase kappa uses a template-
mogenomic fitness signatures. Science 344(6180):208–211.
slippage deletion mechanism, but can realign the slipped strands to favour base
12. Li X, O’Neil NJ, Moshgabadi N, Hieter P (2014) Synthetic cytotoxicity: Digenic inter-
substitution mutations over deletions. Nucleic Acids Res 41(9):5024–5035.
actions with TEL1/ATM mutations reveal sensitivity to low doses of camptothecin.
37. Ling H, Boudsocq F, Woodgate R, Yang W (2004) Snapshots of replication through an
Genetics 197(2):611–623.
abasic lesion; structural basis for base substitutions and frameshifts. Mol Cell 13(5):
13. Giaever G, Nislow C (2014) The yeast deletion collection: A decade of functional ge-
751–762.
nomics. Genetics 197(2):451–465.
38. McCulloch SD, Kunkel TA (2008) The fidelity of DNA synthesis by eukaryotic replica-
14. Jasin M, Rothstein R (2013) Repair of strand breaks by homologous recombination.
tive and translesion synthesis polymerases. Cell Res 18(1):148–161.
Cold Spring Harb Perspect Biol 5(11):a012740.
39. Muñoz-Galván S, et al. (2012) Distinct roles of Mus81, Yen1, Slx1-Slx4, and Rad1 nu-
15. Lang GI, Murray AW (2008) Estimating the per-base-pair mutation rate in the yeast
cleases in the repair of replication-born double-strand breaks by sister chromatid
Saccharomyces cerevisiae. Genetics 178(1):67–82.
exchange. Mol Cell Biol 32(9):1592–1603.
16. Huang ME, Rio AG, Nicolas A, Kolodner RD (2003) A genomewide screen in Saccha-
40. Semlow DR, Zhang J, Budzowska M, Drohat AC, Walter JC (2016) Replication-
romyces cerevisiae for genes that suppress the accumulation of mutations. Proc Natl
dependent unhooking of DNA interstrand cross-links by the NEIL3 glycosylase. Cell
Acad Sci USA 100(20):11529–11534.
17. Doerfler L, Harris L, Viebranz E, Schmidt KH (2011) Differential genetic interactions 167(2):498–511 e414.
between Sgs1, DNA-damage checkpoint components and DNA repair factors in the 41. Stirling PC, et al. (2011) The complete spectrum of yeast chromosome instability genes
maintenance of chromosome stability. Genome Integr 2:8. identifies candidate CIN cancer genes and functional roles for ASTRA complex com-
18. Kokoska RJ, Stefanovic L, DeMai J, Petes TD (2000) Increased rates of genomic dele- ponents. PLoS Genet 7(4):e1002057.
tions generated by mutations in the yeast gene encoding DNA polymerase delta or by 42. Duffy S, et al. (2016) Overexpression screens identify conserved dosage chromosome
decreases in the cellular levels of DNA polymerase delta. Mol Cell Biol 20(20): instability genes in yeast and human cancer. Proc Natl Acad Sci USA 113(36):
7490–7504. 9967–9976.
19. Schürer KA, Rudolph C, Ulrich HD, Kramer W (2004) Yeast MPH1 gene functions in an 43. Bhang HE, et al. (2015) Studying clonal dynamics in response to cancer therapy using
error-free DNA damage bypass pathway that requires genes from homologous re- high-complexity barcoding. Nat Med 21(5):440–448.
combination, but not from postreplicative repair. Genetics 166(4):1673–1686. 44. Wong TN, et al. (2015) Role of TP53 mutations in the origin and evolution of therapy-
20. Ang JS, Duffy S, Segovia R, Stirling PC, Hieter P (2016) Dosage mutator genes in related acute myeloid leukaemia. Nature 518(7540):552–555.
Saccharomyces cerevisiae: A novel mutator mode-of-action of the Mph1 DNA heli- 45. Schumacher TN, Schreiber RD (2015) Neoantigens in cancer immunotherapy. Science
case. Genetics 204(3):975–986. 348(6230):69–74.
21. Mani R, St Onge RP, Hartman JL, 4th, Giaever G, Roth FP (2008) Defining genetic 46. Hall BM, Ma CX, Liang P, Singh KK (2009) Fluctuation analysis CalculatOR: A web tool
interaction. Proc Natl Acad Sci USA 105(9):3461–3466. for the determination of mutation rate using Luria-Delbruck fluctuation analysis.
22. Wyatt MD, Wilson DM, 3rd (2009) Participation of DNA repair in the response to Bioinformatics 25(12):1564–1565.
5-fluorouracil. Cell Mol Life Sci 66(5):788–799. 47. Li H, Durbin R (2009) Fast and accurate short read alignment with Burrows-Wheeler
23. Marteijn JA, Lans H, Vermeulen W, Hoeijmakers JH (2014) Understanding nucleotide transform. Bioinformatics 25(14):1754–1760.
excision repair and its roles in cancer and ageing. Nat Rev Mol Cell Biol 15(7):465–481. 48. Li H, et al.; 1000 Genome Project Data Processing Subgroup (2009) The Sequence
24. Clausen AR, et al. (2015) Tracking replication enzymology in vivo by genome-wide Alignment/Map format and SAMtools. Bioinformatics 25(16):2078–2079.
mapping of ribonucleotide incorporation. Nat Struct Mol Biol 22(3):185–191. 49. Jones SJ, et al. (2010) Evolution of an adenocarcinoma in response to selection by
25. Chan K, et al. (2015) An APOBEC3A hypermutation signature is distinguishable from targeted kinase inhibitors. Genome Biol 11(8):R82.
the signature of background mutagenesis by APOBEC3B in human cancers. Nat Genet 50. O’Shea JP, et al. (2013) pLogo: A probabilistic approach to visualizing sequence
47(9):1067–1072. motifs. Nat Methods 10(12):1211–1212.

2668 | www.pnas.org/cgi/doi/10.1073/pnas.1618555114 Segovia et al.

You might also like