75
views
0
recommends
+1 Recommend
0 collections
    0
    shares
      • Record: found
      • Abstract: found
      • Article: not found

      Homologous recombination-deficient tumors are hyper-dependent on POLQ-mediated repair

      research-article

      Read this article at

      ScienceOpenPublisherPMC
      Bookmark
          There is no author summary for this article yet. Authors can add summaries to their articles on ScienceOpen to make them more accessible to a non-specialist audience.

          Abstract

          Large-scale genomic studies have shown that half of epithelial ovarian cancers (EOCs) have alterations in genes regulating homologous recombination (HR) repair 1 . Loss of HR accounts for the genomic instability of EOCs and for their cellular hyper-dependence on alternative poly-ADP ribose polymerase (PARP)-mediated DNA repair mechanisms 2-5 . Previous studies have implicated the DNA polymerase POLQ 6 in a pathway required for the repair of DNA double-strand breaks 7-9 , referred to as the error-prone microhomology-mediated end-joining (MMEJ) pathway 10-13 . Whether POLQ interacts with canonical DNA repair pathways to prevent genomic instability remains unknown. Here we report an inverse correlation between HR activity and POLQ expression in EOCs. While knockdown of POLQ in HR-proficient cells up-regulates HR activity and RAD51 nucleofilament assembly, knockdown of POLQ in HR-deficient EOCs enhances cell death. Consistent with these results, genetic inactivation of an HR gene (Fancd2) and Polq in mice results in embryonic lethality. Moreover, POLQ contains RAD51 binding motifs, and it blocks RAD51-mediated recombination. Our results reveal a synthetic lethal relationship between the HR pathway and POLQ-mediated repair in EOCs, and identify POLQ as a novel druggable target for cancer therapy. To examine changes in polymerase activity between tumors and normal tissues, we screened polymerase gene expression profiles in a broad number of cancers (Supplementary Table 1). Gene set enrichment analysis (GSEA) revealed specific and recurrent overexpression of POLQ in EOCs (Extended Data Fig. 1a-c). POLQ was up-regulated in a grade-dependent manner and its expression positively correlated with numerous mediators of HR (Extended Data Fig. 1d-j). Since POLQ has been suggested to play a role in DNA repair 7-10 , we investigated a potential role for POLQ in HR repair. To test the relationship between POLQ expression and HR, we used a cell-based assay which measures the efficiency of recombination of two GFP alleles (DR-GFP) 14 . Knockdown of POLQ with siRNA (Extended Data Fig. 2a) resulted in an increase in HR efficiency, similar to that observed by depleting the anti-recombinases PARI or BLM 15,16 . Depletion of POLQ caused a significant increase in basal and radiation (IR)-induced RAD51 foci (Fig. 1a, b and Extended Data Fig. 2b-d), and depletion of POLQ in 293T cells conferred cellular hypersensitivity to mitomycin C (MMC) and an increase in MMC-induced chromosomal aberrations (Extended Data Fig. 2e, f). These findings suggest that human POLQ inhibits HR and participates in the maintenance of genome stability. Given that POLQ shares structural homology with coexpressed RAD51-binding ATPases (Extended Data Fig. 1k, l), we hypothesized that POLQ might regulate HR through an interaction with RAD51. Indeed, RAD51 was detected in Flag-tagged POLQ immunoprecipitates, and purified full-length Flag-POLQ bound recombinant human RAD51 (Fig. 1c, d). Pull-down assays with recombinant GST-RAD51 and in vitro translated POLQ truncation mutants defined a region of POLQ binding to RAD51 spanning amino acid 847-894 (Fig. 1e, f and Extended Data Fig. 2g, h). Sequence homology of POLQ with the RAD51 binding domain of C. elegans RFS-1 17 identified a second binding region (Extended Data Fig. 2i). Peptides arrays narrowed down the RAD51 binding activity of POLQ to three distinct motifs (Fig. 1g and Extended Data Fig. 2j). Substitution arrays confirmed the interaction and highlighted the importance of the 847–894 POLQ region as both necessary and sufficient for RAD51 binding (Extended Data Fig. 3a, b). Taken together, these results indicate that POLQ is a RAD51-interacting protein that regulates HR. In order to address the role of POLQ in HR regulation, we assessed the ability of wild-type (WT) or mutant POLQ to complement the siPOLQ-dependent increase in RAD51 foci. Full-length wild-type POLQ fully reduced IR-induced RAD51 foci, unlike POLQ mutated at ATPase catalytic residues (A-dead) or POLQ lacking interaction with RAD51 (ΔRAD51) (Fig. 2a, b). Expression of a POLQ mutant lacking the polymerase domain (ΔPol1) was sufficient to decrease IR-induced RAD51 foci, suggesting that the N-terminal half of POLQ is sufficient to disrupt RAD51 foci (Fig. 2b and Extended Data Fig. 3c, d). We next measured the ability of wild-type or mutant POLQ to complement the siPOLQ-dependent increase in HR efficiency. Again, expression of full-length POLQ or ΔPol1 decreased the recombination frequency when compared to cells expressing other POLQ constructs, suggesting that the N-terminal half of POLQ containing the RAD51 binding domain and the ATPase domain is needed to inhibit HR (Fig. 2c and Extended Data Fig. 3e). A purified recombinant POLQ fragment (ΔPol2) from insect cells exhibited low levels of basal ATPase activity, as previously reported 18 (Fig. 2d, e). POLQ ATPase activity was selectively stimulated by the addition of single-stranded DNA (ssDNA) or fork DNA (Fig. 2e and Extended Data Fig. 4a). Electrophoretic mobility gel shift assays (EMSA) showed specific binding of POLQ to ssDNA (Fig. 2f and Extended Data Fig. 4b). We incubated ΔPol2 with ssDNA and measured RAD51-ssDNA nucleofilament assembly. Interestingly, RAD51-ssDNA assembly was reduced by wild-type ΔPol2 but not by A-dead or ΔRAD51, indicating that POLQ negatively affects RAD51-ssDNA assembly through its RAD51 binding and ATPase activities (Fig. 2g and Extended Data Fig. 4c-f). Furthermore, POLQ decreased the efficiency of D-loop formation, confirming that POLQ is a negative regulator of HR (Fig. 2h and Extended Data Fig. 4g-j). Since POLQ is up-regulated in subgroups of cancers associated with HR deficiency (Fig. 3a) and POLQ activity shows specificity for replicative stress-mediated structures (ss and fork DNA) (Fig. 2e, f), we examined the cellular functions of POLQ under replicative stress. Subcellular fractionation revealed that POLQ is enriched in chromatin in response to ultraviolet (UV) light; and RAD51 binding by POLQ was enhanced by UV exposure, suggesting that POLQ regulates HR in cells under replicative stress (Extended Data Fig. 5a, b). POLQ-depleted cells were hypersensitive to cellular stress and DNA damage along with an exacerbated checkpoint activation and increased γH2AX phosphorylation (Fig. 3b, c). Furthermore, the cell cycle progression of POLQ-depleted cells was impaired after DNA damage (Fig. 3d, e). To determine the role of POLQ in replication dynamics, single-molecule analyses were performed on extended DNA fibers 19 . Abnormalities in replication fork progression were observed in POLQ-depleted cells (Fig. 3f, g and Extended Data Fig. 5c, d). These results suggest that POLQ maintains genomic stability at stalled or collapsed replication forks by promoting fork restart. To examine the regulation of POLQ, we quantified POLQ expression by RT-qPCR. POLQ was selectively up-regulated in HR-deficient ovarian cancer cell lines. Complementation of a BRCA1 or FANCD2-deficient cell lines, restored normal HR function and reduced POLQ expression to normal levels. Conversely, siRNA-mediated inhibition of HR genes increased POLQ expression (Extended Data Fig. 5e, f). POLQ expression was significantly higher in subgroups of cancers with HR deficiency and a high genomic instability pattern 20 (Fig. 3a and Extended Data Fig. 5g). Patients with high POLQ expression had a better response to platinum chemotherapy, a surrogate for HR deficiency, suggesting that POLQ expression inversely correlates with HR activity and may be useful as a biomarker for platinium sensitivity (Extended Data Fig. 5h, i). Together, these data indicate that increased POLQ expression is driven by HR deficiency. To assess the possible synthetic lethality between HR genes and POLQ, we generated an HR-deficient ovarian tumor cell line, A2780-shFANCD2 cells (Extended Data Fig. 6a-c). These cells, and the parental A2780 cells, were subjected to POLQ depletion, and survival following exposure to cytotoxic drugs was measured. POLQ depletion reduced the survival of HR-deficient cells exposed to inhibitors of PARP (PARPi), cisplatin (CDDP), or MMC (Extended Data Fig. 6d-f). POLQ inhibition impaired the survival of BRCA1-deficient tumors (MDA-MB-436) after PARPi treatment but had no effect on the complemented line (MDA-MB-436 + BRCA1) (Fig. 4a). POLQ-depleted cells were hypersensitive to ATM inhibition, known to create an HR defect phenotype 21 . Chromosomal breakage, checkpoint activation, and γH2AX phosphorylation in response to MMC were exacerbated by POLQ depletion (Fig. 4b and Extended Data Fig. 6g, h). Furthermore, a whole-genome shRNA screen performed on HR-deficient (FANCA-/-) fibroblasts showed that shRNAs targeting POLQ impair cell survival in MMC (Extended Data Fig. 6i), suggesting that HR-deficient cells cannot survive in the absence of POLQ. Next, we investigated the interaction between the HR and POLQ pathways in vivo by interbreeding Fancd2+/− and Polq+/− mice. Although Fancd2-/- and Polq-/- mice are viable and exhibit subtle phenotypes 7,22 , viable Fancd2−/−Polq−/− mice were uncommon from these matings (Extended Data Fig. 7a). The only surviving Fancd2−/−Polq−/− pups exhibited severe congenital malformations and were either found dead or died prematurely. Fancd2−/−Polq−/− embryos showed severe congenital malformations, and mouse embryonic fibroblasts (MEFs) generated from Fancd2−/−Polq−/− embryos showed hypersensitivity to PARPi (Fig. 4c and Extended Data Fig. 7b-e). These data suggest that loss of the HR and POLQ repair pathways in vivo results in embryonic lethality. Since xenografts of tumors cells expressing shRNAs against both FANCD2 and POLQ did not stably propagate in mice (Extended Data Fig. 7f), we xenotransplanted A2780-shFANCD2 cells expressing either doxycycline-inducible POLQ or Scr shRNA in athymic nude mice. POLQ depletion significantly impaired tumor growth after PARPi treatment (Fig. 4d, e and Extended Data Fig. 7g, h). Moreover, mice bearing POLQ-depleted tumors had a survival advantage following PARPi treatment compared to control mice (Fig. 4f). POLQ-depleted HR-deficient tumor cells also exhibited decreased survival in in vivo dual-colour competition experiments (Extended Data Fig. 7i-l). Collectively, these data confirm that HR-deficient tumors are hypersensitive to inhibition of POLQ-mediated repair. To understand which functions of POLQ are required for resistance to DNA-damaging agents, we performed a series of complementation studies in HR-deficient cells. Expression of full-length POLQ or ΔPol1, but not ΔRAD51, in HR-deficient POLQ-depleted cells treated with PARPi or MMC was able to rescue toxicity, suggesting that the anti-recombinase activity of POLQ maintains the genomic stability of HR-deficient cells (Fig. 4g, h and Extended Data Fig. 8a, b). Moreover, the toxicity induced by loss of POLQ in HR-deficient cells was rescued by depletion of RAD51 showing that, in the absence of POLQ, RAD51 is toxic to HR-deficient cells (Fig. 4i). These results suggest a role for POLQ in limiting toxic HR events 23 (Extended Data Fig. 4c-f) and may explain why HR-deficient cells overexpress and depend on an anti-recombinase for survival. High mutation rates have been observed in HR-deficient tumors 24 . Previous studies have shown that POLQ is an error-prone polymerase 25,26 that participates in alternative end-joining (alt-EJ) 10 . Therefore, we assessed the role of POLQ in error-prone DNA repair in human cancer cells. POLQ inhibition reduced alt-EJ efficiency in U2OS cells, similar to the reduction observed following depletion of PARP1, another critical factor in end-joining 27,28 (Extended Data Fig. 9a). Expression of full-length POLQ, ΔRAD51, or A-dead, but not the ΔPol1 mutant, complemented the cells, suggesting that the polymerase domain of POLQ is required for end-joining (Extended Data Fig. 9b). GFP-tagged full-length POLQ formed foci after UV treatment in a PARP-dependent manner (Extended Data Fig. 9c). POLQ inhibition reduced the mutation frequency induced by UV light, and tumors with high POLQ expression harbored more somatic point mutations than those with lower POLQ levels (Extended Data Fig. 9d, e). These results suggest that POLQ contributes to the mutational signature observed in some HR-deficient tumors 29 . In human cancers, a deficiency in one DNA repair pathway can result in cellular hyper-dependence on a second compensatory DNA repair pathway 4 . Here, we show that POLQ is overexpressed in EOCs and other tumors with HR defects 30 . Wild-type POLQ limits RAD51-ssDNA nucleofilament assembly (Extended Data Fig. 10a) and promotes alt-EJ (Fig. 4j). We demonstrate that HR-deficient tumors are hypersensitive to inhibition of POLQ-mediated repair. Therefore, POLQ appears to channel DNA repair by antagonizing HR and promoting PARP1-dependent error-prone repair (Extended Data Fig. 10b). These results offer a potential new therapeutic target for cancers with inactivated HR. Methods Bioinformatic analysis Gene Set Enrichment Analysis algorithm (GSEA, www.broadinstitute.org) was performed for the datasets summarized in Supplementary Table 1. TransLesion Synthesis (TLS) and polymerase gene sets are described in Supplementary Table 3. Row expression data were downloaded from Gene Expression Omnibus (GEO). Quantile normalizations were performed using the RMA routine through GenePattern. GSEA was run using GenePattern (www.broadinstitute.org) and corresponding P values were computed using 2,000 permutations. The DNA repair gene set used in Extended Data Figure 1g has been determined according to a list of 151 DNA genes previously used 31 . GSEA analysis for 151 repair genes has been performed on the ovarian serous datasets (GSE14001, GSE14007, GSE18520, GSE16708, GSE10971). The list of 20 genes shown in Extended Data Figure 1g represents the top 20 expressed gene in cancer samples (median of the 5 datasets). The waterfall plot in Extended Data Figure 1h was generated as follows: the 20 genes defined in (g) were used as a gene set; GSEA for indicated data sets was performed and the nominal P values were plotted. Supervised analysis of gene expression for GSE9891 was performed with respect to differential expression that differentiated the third of tumors with highest POLQ expression from the 2 third with lowest POLQ levels. A list of the 200 most differentially expressed probe sets between the 2 groups (Supplementary Table 2) with false discovery rate <0.05 was analyzed for biological pathways (hypergeometrical test; www.broadinstitute.org). TCGA datasets were accessed through the public TCGA data portal (www.tcga-data.nci.nih.gov). Figure 3a reflects POLQ gene expression in the ovarian carcinoma dataset GSE9891, uterine carcinoma TCGA and breast carcinoma TCGA. Normalization of POLQ expression values across datasets was performed using z-score transformation. POLQ expression values were subdivided in subgroups reflecting the stage of the disease (for GSE9891: grade 3 ovarian serous carcinoma, n=143 compared to type 1 (grade 1) ovarian cancers, n=20; for uterine: serous like tumors, n=60 compared to the rest of the tumors, n=172; for breast: basal like breast carcinoma, n=80 compared to the rest of the tumors, n=421). Progression-free survival curves were generated by the Kaplan-Meier method and differences between survival curves were assessed for statistical significance with the log-rank test. In the absence of a clinically defined cutoff point for POLQ expression levels we divided patients into 2 groups: those with POLQ mRNA levels equal to or above the median (POLQ high group) and those with values below the median (POLQ low group). We then analyzed the correlation of POLQ with outcome in each group. Patients with CCNE amplification (resistant to CDDP) were excluded from the analysis. For mutation count, we accessed data from tumors included in the TCGA datasets for which gene expression and whole-exome DNA sequencing was available. Data were accessed through the public TCGA data portal and the cBioPortal for Cancer Genomics (www.cbioportal.org). For each TCGA dataset, non-synonymous mutation count was assessed in tumors with the highest POLQ expression (top 33%) and compared to tumors with low POLQ expression (the remaining, 67%). In the uterine TCGA 20 , we curated all tumors except the ultra and hyper-mutated group (i.e, POLE and MSI tumors). In the breast TCGA 32 , all tumors were analyzed. In the ovarian TCGA 1 , we curated tumors harbouring molecular alterations (via mutation and epigenetic silencing) of the HR pathway. Plasmid construction To facilitate subcloning, a silent mutation (A390A) was introduced into the POLQ gene sequence to remove the unique Xho1 cutting site. Full-length or truncated POLQ cDNA were PCR-amplified and subcloned into pcDNA3-N-Flag, pFastBac-C-Flag, pOZ-C-Flag-HA, or GFP-C1 vectors to generate the various constructs. Point mutations and loop deletions were introduced by QuikChange II XL Site-Directed Mutagenesis Kit (Agilent Technologies) and confirmed by DNA sequencing. For POLQ rescue experiments (Fig. 4g, h and Extended Data Fig. 3d, e), POLQ cDNA constructs resistant to siPOLQ1 were generated into the pOZ-C-Flag-HA vector and the construct were stably expressed in indicated cell line by retroviral transduction. The POLQ ATPase catalytically-dead mutant (A-dead) was generated by mutating the walker A and B motifs (K121A and D216A, E217A, respectively). pOZ-C-Flag-HA POLQ constructs were generated for retroviral transduction, and stable cells were selected using magnetic Dynabeads (Life Technologies) conjugated to the IL2R antibody (Millipore). SiRNA and shRNA sequence information For siRNA-mediated knockdown, the following target sequences were used: POLQ (Qiagen POLQ_1 used as siPOLQ1 and Qiagen POLQ_6 used as siPOLQ2); BRCA1 (Qiagen BRCA1_13); PARP1 (Qiagen PARP1_6); REV1 (5′-CAGCGCAUCUGUGCCAAAGAA-TT-3′); BRCA2 (5′-GAAGAAUGCAGGUUUAAUATT-3′); BLM (5′-AUCAGCUAGAGGCGAUCAATT-3′); FANCD2 (5′-GGAGAUUGAUGGUCUACUATT-3′) and PARI (5′-AGGACACAUGUAAAGGGAUUGUCUATT-3′). AllStars negative control siRNA (Qiagen) served as the negative control. ShRNAs targeting human FANCD2 was previously generated in the pTRIP/DU3-MND-GFP vector 33 . ShRNAs targeting human POLQ (CGGGCCTCTTTAGATATAAAT), human BRCA2 (AAGAAGAATGCAGGTTTAATA) or Control (Scr, scramble) were generated in the pLKO-1 vector. POLQ (V2THS_198349) and non-silencing TRIPZ-RFP doxycycline-inducible shRNA were purchased from Open Biosystems. All shRNAs were transduced using lentivirus. Immunoblot analysis, fractionation and pull-down assays Cells were lysed with 1 % NP40 lysis buffer (1 % NP40, 300 mM NaCl, 0.1 mM EDTA, 50 mM Tris [pH 7.5]) supplemented with protease inhibitor cocktail (Roche), resolved by NuPAGE (Invitrogen) gels, and transferred onto nitrocellulose membrane, followed by detection using the LAS-4000 Imaging system (GE Healthcare Life Sciences). For immunoprecipitation, cells were lysed with 300 mM NaCl lysis buffer, and the lysates were diluted to 150 mM NaCl before immunoprecipitation. Lysates were incubated with anti-Flag agarose resin (Sigma) followed by washes with 150 mM NaCl buffer. In vitro transcription and translation reactions were carried out using the TNT T7 Quick Coupled Transcription-Translation System (Promega). For cellular fractionation, cells were incubated with low salt permeabilization buffer (10 mM Tris [pH 7.3], 10 mM KCl 1.5 mM MgCl2) with protease inhibitor on ice for 20 minutes. Following centrifugation, nuclei were resuspended in 0.2 M HCl and the soluble fraction was neutralized with 1 M Tris-HCl [pH 8.0]. Nuclei were lysed in 150 mM NaCl and following centrifugation, the chromatin pellet was digested by micrococcal nuclease (Roche) for 5 minutes at room temperature. Recombinant GST-RAD51 and GST-PCNA fusion protein were expressed in BL21 strain and purified using glutathione-Sepharose beads (GE Healthcare) as previously described 15 . Beads with equal amount of GST or GST-RAD51 were incubated with in vitro–translated Flag-tagged POLQ variants in 150 mM NaCl lysis buffer. Antibodies and chemicals Antibodies used in this study included: anti-PCNA (PC-10), anti-FANCD2 (FI-17), anti-RAD51 (H-92), anti-GST (B14), and Histone H3 (FL-136) and anti-vinculin (H-10) (Santa Cruz); anti-Flag (M2) (Sigma); anti-pS317CHK1 (2344), anti-pT68CHK2 (2661) (Cell signalling); anti-pS824KAP-1 (A300-767A) (Bethyl); anti-pS317γH2AX (05636) (Millipore); anti-pS15p53 (ab1431) and anti-POLQ (ab80906) (abcam); anti-BrdU (555627) (BD Pharmingen). Mitomycin C (MMC), cis-diamminedichloroplatinum(II) (Cisplatin, CDDP), and Hydroxyurea (HU) were purchased from Sigma. The PARPi rucaparib (AG-014699) was purchased from Selleckchem and ABT-888 from AbbVie. Rucaparib was used for all in vitro assays and ABT-888 was used for all in vivo experiments. Chromosomal breakage analysis 293T and Vu 423 cells were twice-transfected with siRNAs for 48 hours and incubated for 48 hours with or without the indicated concentrations of MMC. For complementation studies on 293T shFANCD2, POLQ cDNA constructs were transfected 24 hours after the first siRNA transfection. Cells were exposed for 2 hours to 100 ng/ml of colcemid and treated with a hypotonic solution (0.075 M KCl) for 20 minutes and fixed with 3:1 methanol/acetic acid. Slides were stained with Wright's stain and 50 metaphase spreads were scored for aberrations. The relative number of chromosomal breaks was calculated relative to control cells (si Scr). For clarity of the Figure 4b, radial figures were excluded from the analysis. Reporter assays and immunofluorescence HR and alt-EJ efficiency was measured using the DR-GFP (HR efficiency) and the alt-EJ reporter assay, performed as previously described 14,27,34 . Briefly, 48 hours before transfection of SceI cDNA, U20S-DR-GFP cells were transfected with indicated siRNA or PARPi (1 μM). The HR activity was determined by FACS quantification of viable GFP-positive cells 96 hours after SceI was transfected. For RAD51 immunofluorescence experiments, cells were transfected with indicated siRNA 48 hours before treatment with HU (2 mM) or IR (10 Gy). For complementation studies, POLQ cDNA constructs were either transfected 24 hours after siRNA transfection (Fig. 2b, 2c and Extended Data Fig. 9b) or stably expressed in indicated cell line (Extended Data Fig. 3d, e). 6 hours after HU or IR treatment, cells were fixed with 4% paraformaldehyde for 10 minutes at room temperature, followed by extraction with 0.3% Triton X-100 for 10 minutes on ice. Antibody staining was performed at room temperature for 1 hour. For quantification of RAD51 foci in BrdU positive cells, cells were transfected with indicated siRNA 48 hours before treatment with IR (10 Gy). 2 hours after IR treatment, cells were treated with BrdU pulse (10 μM) for 2 hours and subsequently fixed with 4% paraformaldehyde and stained for RAD51 as described above. Cells were then fixed in ethanol (4°C, over night), treated with 1.5 M HCL for 30 minutes and stained for BrdU antibody. The relative number of cells with more than 10 RAD51 foci was calculated relative to control cells (si Scr). Statistical differences between cells transfected with siRNAs (si POLQ1, si POLQ2, si BRCA2, si PARI or si BLM relative to control (si Scr) were assessed. For GFP fluorescence, cells were grown on coverslip, treated with UV (24 hours after GFP-POLQ transfection; 20 J/m2), fixed with 4% paraformaldehyde for 10 min at 25 °C 4 hours after the UV treatment, washed three times with PBS and mounted with DAPI-containing mounting medium (Vector Laboratories). When indicated cells were treated with PARPi (1 μM) 24 hours before GFP-POLQ transfection. Images were captured using a Zeiss AX10 fluorescence microscope and AxioVision software. Cells with GFP foci were quantified by counting number of cells with more than five foci. At least 150 cells were counted for each sample. Cell survival assays For assessing cellular cytotoxicity, cells were seeded into 96-well plates at a density of 1000 cells/well. Cytotoxic drugs were serially diluted in media and added to the wells. At 72 hours, CellTiter-Glo reagent (Promega) was added to the wells and the plates were scanned using a luminescence microplate reader. Survival at each drug concentration was plotted as a percentage of the survival in drug-free media. Each data point on the graph represents the average of three measurements, and the error bars represent the standard deviation. For clonogenic survival, 1000 cells/well were seeded into six-well plates and treated with cytotoxic drugs the next day. For MMC and PARPi, cells were treated continuously with indicated drug concentrations. For CDDP, cells were treated for 24 hours and cultured for 14 days in drug-free media. Colony formation was scored 14 days after treatment using 0.5% (w/v) crystal violet in methanol. Survival curves were expressed as a percentage ± s.e.m. over three independent experiments of colonies formed relative to the DMSO-treated control. Cell cycle analysis A2780 cells expressing Scr or POLQ shRNA were synchronized by a double thymidine block (Sigma) and subsequently exposed to MMC (1 μg/ml for 2 hours), IR (10 Gy) or HU (2 mM, over night). At the indicated time points following drug release, cells were fixed in chilled 70% ethanol, stored overnight at -20°C, washed with PBS, and resuspended in propidium iodide. A fraction of those cells was analyzed by immunoblotting for DNA damage response proteins. The immunoblot analysis of γH2AX shows staining after 0, 24, 48 and 72 hours of HU treatment. For proliferation experiments, cells were incubated with 5-ethynyl-2′-deoxyuridine (EdU) (10 μM) for 1 hour at each time point after MMC exposure (1 μg/ml for 2 hours). Cells were washed and resuspended in culture medium for 2 hours prior to be analyzed by flow cytometry. Edu Staining was performed using the Click-iT EdU kit (Life Technologies). DNA Fiber Analysis A2780 cells expressing Scr or POLQ shRNA were incubated with 25 μM chlorodeoxyuridine (CldU) (Sigma, C6891) for 20 minutes. Cells were then treated with 2 mM hydroxyurea (HU) for 2 hours and incubated in 250 μM iododeoxyuridine (ldU) (Sigma, I7125) for 25 minutes after washout of the drug. Spreading of DNA fibers on glass slides was done as reported 19 . Glass slides were then washed in distilled water and in 2.5 M HCl for 80 minutes followed by three washes in PBS. The slides were incubated for 1 hour in blocking buffer (PBS with 1% BSA and 0.1% NP40) and then for 2 hours in rat anti-BrdU antibody (1:250, Abcam, ab6326). After washing with blocking buffer the slides were incubated for 2 hours in goat anti-rat Alexa 488 antibody (1:1000, Life Technologies, A-11006). The slides were then washed with PBS and 0.1% NP40 and then incubated for 2 hours with mouse anti-BrdU antibody diluted in blocking buffer (1:100, BD Biosciences, 347580). Following an additional wash with PBS and 0.1% NP40, the fibers were stained for 2 hours with chicken anti-mouse Alexa 594 (1:1000, Life Technologies, A-21201). At least 150 fibers were counted per condition. Pictures were taken with an Olympus confocal microscope and the fibers were analyzed by ImageJ software. The number of stalled or collapsed forks were measured by DNA fibers that had incorporated only CIdU. Stalled or collapsed forks counted in POLQ-depleted cells is expressed as fold-change after HU treatment relative to the fold-change observed in control cells, which was arbitrarily set to 1. SupF mutagenesis assay 293T cells twice-transfected with siRNAs for 48 hours were then transfected with undamaged or damaged (UVC, 1,000 J/m2) pSP189 plasmids using GeneJuice (Novagen). After 48 hours, plasmid DNA was isolated with a miniprep kit (Promega) and digested with DpnI. After ethanol precipitation, extracted plasmids were transformed into the β-galactosidase–MBM7070 indicator strain through electroporation (GenePulsor X Cell; Bio-Rad) and plated onto LB plates containing 1 mM IPTG, 100 μg/ml 5-bromo-4-chloro-3-indolyl-β-D-galactopyranoside and 100 μg/ml ampicillin. White and blue colonies were scored using ImageJ software, and the mutation frequency was calculated as the ratio of white (mutant) to total (white plus blue) colonies. POLQ gene expression RNA samples extracted using the TRIzol Reagent (Invitrogen) were reverse transcribed using the Transcriptor Reverse Transcriptaze kit (Roche) and oligo dT primers. The resulting cDNA was use to analyzed POLQ expression by RT-qPCR using with QuantiTect SYBRGreen (Qiagen), in an iCycler machine (Bio-Rad). POLQ gene expression values were normalized to expression of the housekeeping gene GAPDH, using the ΔCT method and are shown on a log2 scale. The primers used for POLQ are as follows: POLQ primer 1 (Forward: 5′-TATCTGCTGGAACTTTTGCTGA-3′; Reverse: 5′-CTCACACCATTTCTTTGATGGA-3′); POLQ primer 2 (Forward: 5′-CTACAAGTGAAGGGAGATGAGG-3′; Reverse: 5′-TCAGAGGGTTTCACCAATCC-3′). POLQ purification from insect SF9 cells A POLQ fragment (ΔPol2) containing the ATPase domain with a RAD51 binding site (amino acids 1 to 1000) was cloned into pFastBac-C-Flag and purified from baculovirus-infected SF9 insect cells as previously described 35 . Briefly, SF9 cells were seeded in 15-cm dishes at 80-90% confluency and infected with baculovirus. Three days post-infection, cells were harvested and lysed in 500 mM NaCl lysis buffer (500 mM NaCl, 0.01 % NP40, 0.2 mM EDTA, 20% Glycerol, 1 mM DTT, 0.2 mM PMSF, 20 mM Tris [pH 7.6]) supplemented with Halt protease inhibitor cocktail (Thermo Scientific) and Calpain I inhibitor (Roche) and the protein was eluted in lysis buffer supplemented with 0.2 mg/ml of Flag peptide (Sigma). The protein was concentrated in lysis buffer using 10 kDa centrifugal filters (Amicon). The protein was quantified by comparing its staining intensity (Coomassie-R250) with that of BSA standards in a 7% tris-glycine SDS-PAGE gel. Purified protein was flash-frozen in small aliquots in liquid nitrogen and stored at -80°C. Radiometric ATPase assay Each 10 μl reaction consisted of 200 nM ATP, reaction buffer (20 mM Tris-HCl [pH 7.6], 5 mM MgCl2, 0.05 mg/ml BSA, 1 mM DTT), and 5 μCi of [γ-32P]-ATP. For corresponding reactions, ssDNA, dsDNA, and forked DNA were added to the reaction in excess at a final concentration of 600 nM. Once all of the non-enzymatic reagents were combined, recombinant POLQ was added to start the ATPase reaction. After incubation for 90 minutes at room temperature, stop buffer (125 mM EDTA [pH 8.0]) was added and approximately ∼0.05 μCi was spotted onto PEI-coated thin-layer chromatography (TLC) plates (Sigma). Unhydrolyzed [γ-32P]-ATP was separated from the released inorganic phosphate [32Pi] with 1 M acetic acid, 0.25 M lithium chloride as the mobile phase. TLC plates were exposed to a phosphor screen and imaged with the BioRad Imager PMC. ssDNA, dsDNA, and forked DNA were generated as previously described 35 . To remove any contaminating ssDNA, dsDNA and forked DNA were gel purified after annealing. Spots corresponding to [γ-32P]-ATP and the released inorganic phosphate [32Pi] were quantified (in units of pixel intensity) and the fraction of ATP hydrolyzed calculated for each POLQ concentration. Electrophoretic Mobility Gel Shift Assay (EMSA) Binding of POLQ to ssDNA was assessed using EMSA. 60-mer single-stranded DNA (ssDNA) or double-stranded DNA (dsDNA) oligonucleotides (5 nM) were incubated with increasing amount of POLQ (0, 5, 10, 50, or 100 nM) in 10 μl of binding buffer (20 mM HEPES-K+, [pH 7.6], 5 mM magnesium acetate, 0.1 μg/μl BSA, 5% glycerol, 1 mM DTT, 0.2 mM EDTA, and 0.01% NP-40) for one hour on ice. POLQ protein was added at a 10-fold dilution so that the final salt concentration was approximately 50 mM NaCl. The ssDNA probes are 5′ fluorescently-labeled with IRDye-700 (IDT). After incubation, the samples were analyzed on a 5% native polyacrylamide/0.5 × TBE gel at 4°C. A fluorescent imager (Li-Cor) was used to visualize the samples in the gel. RAD51 purification Human GST-RAD51 was purified from bacteria as described 36 . Xenopus RAD51 (xRAD51) was purified as follow. N-terminally His-tagged SUMO-RAD51 was expressed in BL21 pLysS cells. Three hours after induction with 1 mM IPTG cells were harvested and resuspended in Buffer A (50 mM Tris-Cl [pH 7.5], 350 mM NaCl, 25% Sucrose, 5 mM β-mercaptoethanol, 1 mM PMSF and 10 mM imidazole). Cells were lysed by supplementation with Triton X-100 (0.2% final concentration), three freeze-thaw cycles and sonication (20 pulses at 40% efficiency). Soluble fraction was separated by centrifugation and incubated with 2 mL of Ni-NTA resin (Qiagen) for 1 hour at 4°C. After washing the resin with 100 mL of wash buffer (Buffer A supplemented with 1 M NaCl, final concentration) the salt concentration was brought down to 350 mM. His-SUMO-RAD51 was eluted with a linear gradient of imidazole from 10 mM - 300 mM in Buffer A. Eluted fractions were analyzed by SDS-PAGE. His-SUMO-RAD51 containing fractions were pooled and supplemented with Ulp1 protease to cleave the His-SUMO tag and dialyzed overnight into Buffer B (50 mM Tris-Cl [pH 7.5], 350 mM NaCl, 25% Sucrose, 10% Glycerol, 5 mM β-mercaptoethanol, 10 mM imidazole and 0.05% Triton X-100). The dialyzed fraction was incubated with Ni-NTA resin for 1 hour at 4°C and the RAD51 containing flow-through fraction was collected and dialyzed overnight into Buffer C (100 mM Potassium phosphate [pH 6.8], 150 mM NaCl, 10% Glycerol, 0.5 mM DTT and 0.01% Triton-X). RAD51 was further purified by Hydroxyapatite (Bio-Rad) chromatography. After washing with ten column volumes of Buffer C, RAD51 was eluted with a linear gradient of Potassium phosphate [pH 6.8] from 100 mM - 800 mM. RAD51 containing fractions were analyzed by SDS-PAGE and dialyzed into storage buffer (20 mM HEPES-KOH [pH 7.4], 150 mM NaCl, 10% Glycerol, 0.5 mM DTT). Purified protein was flash-frozen in small aliquots in liquid nitrogen and stored at -80°C. D-loop assay D-loop formation assays were performed using xRAD51 and conducted as previously described 37 . Briefly, nucleofilaments were first formed by incubating RAD51 (1 μM) with end-labeled 90-mer ssDNA (3 μM nt) at 37 °C for 10 minutes in reaction buffer containing 20 mM HEPES-KOH [pH 7.4], 1 mM ATP, 1 mM Mg(Cl)2, 1 mM DTT, BSA (100 μg/mL), 20 mM phosphocreatine and creatine phosphokinase (20 μg/mL). After the 10 minutes incubation increasing amounts of POLQ (0, 0.1, 0.5, or 1.0 μM) and RPA (200 nM) were added and incubated for an additional 15 minutes at 37°C. Reaction was then supplemented with 1 mM CaCl2 followed by further incubation at 37°C for 15 minutes. D-loop formation was initiated by the addition of supercoiled dsDNA (pBS-KS (-), 79 μM bp) and incubation at 37°C for 15 minutes. D-loops were analyzed by electrophoresis on a 0.9% agarose gel after deproteinization. Gel was dried and exposed to a PhosphoImager (GE Healthcare) screen for quantification. Substitution peptide arrays and RAD51-ssDNA filament experiments Subtitution peptide arrays were performed as previously described 17 . RAD51 displacement assays were performed as follow. Binding reactions (10 μl) contained 5′-32P-end-labelled DNA substrates (0.5 ng of 60 mer ssDNA) and various amounts of human RAD51 and/or POLQ in binding buffer (40 mM Tris-HCl [pH 7.5], 50 mM NaCl, 10 mM KCl, 2 mM DTT, 5 mM ATP, 5 mM MgCl2, 1 mM DTT, 100 mg/ml BSA) were conducted at room temperature. After 5 minutes incubation with POLQ and a further 5 minutes incubation with RAD51 or vice versa, an equimolar amount of cold DNA substrate was added to the reaction. Products were then analyzed by electrophoresis through 10% PAGE (200V for 40 min in 0.5×Tris-borate-EDTA buffer) and visualized by autoradiography. Interbreeding of the Fancd2 and Polq mice For the characterization of Fancd2/Polq conditional knockouts, we crossed C57BL/6J mice (Jackson Laboratory). Fancd2+/-Polq+/+ mice, previously generated in our laboratory 22 , were crossed with Fancd2+/+Polq+/- mice 7 to generate Fancd2+/-Polq+/- mice. These double heterozygous mice were then interbred, and the offspring from these mating pairs were genotyped using PCR primers for Fancd2 and Polq. A statistical comparison of the observed with the predicted genotypes was performed using a 2-sided Fisher's exact test. Primary MEFs were generated from E13.5 to E15 embryos and cultured in RPMI supplemented with 15% fetal bovine serum and 1% penicillin-streptomycin. All data generated in the study were extracted from experiments performed on primary MEFs from passage 1 to passage 4. The primers used for mice genotyping are as follows: Fancd2 PCR primers OST2cF (5′-CATGCATATAGGAACCCGAAGG-3′), OST2aR (5′-CAGGACCTTTGGAGAAGCAG-3′) and LTR2bF (5′-GGCGTTACTTAAGCTAGCTTG-3′); Polq PCR primers IMR5973 (5′-TGCAGTGTACAGATGTTACTTTT-3′), IMR 5974 (5′-TGGAGGTAGCATTTCTTCTC-3′), IMR 5975 (5′-TCACTAGGTTGGGGTTCTC-3′) and IMR 5976 (5′-CATCAGAAGCTGACTCTAGAG-3′). Specific PCR conditions are available upon request. Studies of xenograft-bearing CrTac:NCr-Foxn1nu mice The Animal Resource Facility at The Dana-Farber Cancer Institute approved all housing situations, treatments and experiments using mice. No more than five mice were housed per air-filtered cage with ad libitum access to standard diet and water, and were maintained in a temperature and light-controlled animal facility under pathogen-free conditions. All mice described in this text were drug and procedure naïve before the start of the experiments. For every xenograft study, we subcutaneously implanted approximately 1.0 × 106 A2780 cells (1:1 in Matrigel Matrix, BD Biosciences) into both flanks of 6-8 week old female CrTac:NCr-Foxn1nu mice (Taconic). Doxycycline (Sigma) was added to the food (625 PPM) and bi-weekly (Tuesday and Friday) to the water (200 μg/ml) for mice bearing tumors that reached 100-200 mm3. Roughly one week (5-6 days) after the addition of Doxycycline to the diet, mice were randomized to twice daily treatment schedules with vehicle (0.9% NaCl) or PARPi (ABT-888; 50 mg per kg body weight) by oral gavage administration for the indicated number of weeks. Overall survival was determined using Kaplan-Meier analyses performed with Log-Rank tests to assess differences in median survival for each shRNA condition (shScr or shPOLQ) and each treatment condition (vehicle or PARPi) (GraphPad Prism 6 Software). For competition assays, A2780 cells expressing FANCD2-GFP shRNA (GFP cells) or a combination of FANCD2-GFP shRNA with (doxycycline inducible) Scr-RFP or POLQ-RFP shRNA (GFP-RFP cells) were mixed at an equal ratio of GFP to GFP-RFP cells, and thereafter injected into nude mice given doxycycline-containing diets and treated with either vehicle or PARPi or CDDP. For competition assays, mice received identical doxycycline and PARPi drug treatment. For the Cisplatin competition assay, mice were randomized into semi-weekly treatment regimens with vehicle (0.9% NaCl) or CDDP (5 mg per kg body weight) by intraperitoneal injection. After three to four weeks of treatment, mice were euthanized and tumors were grown in vitro, in the presence of doxycycline (2 μg/ml for 4 days). The relative ratio of GFP to GFP-RFP cells was determined by FACS analysis. Tumor volumes were calculated bi-weekly using caliper measurements (length × width2)/2. Growth curves were plotted as the mean tumor volume (mm3) for each treatment group; relative tumor volume (RTV) indicates change in tumor volume at a given time point relative to that at the day before initial dosing (=1). Mice were unbiasedly assigned into different treatment groups. Drug treatment and outcome assessment was performed in a blinded manner. Mice were monitored every day and euthanized by CO2 inhalation when tumor size (≥2 cm), tumor status (necrosis/ulceration) or body weight loss (≥20%) reached ethical endpoint, according to the rules of the Animal Resource Facility at The Dana-Farber Cancer Institute. Immunohistochemical staining We stained formalin-fixed paraffin-embedded sections of harvested xenografts with antibodies specific for γ-H2AX (pSer139) (Upstate Biotechnology) and Ki67 (Dako). At least two xenografts were scored for each treatment. Tumors were collected three weeks after treatment. At least five 40× fields were scored. The mean ± s.e.m. percentage of positive cells from five images in each treatment group was calculated. Statistical analysis Unless stated otherwise, all data are represented as mean ± s.e.m. over at least three independent experiments, and significance was calculated using the Student's t test. Asterisks indicate statistically significant (*, P < 0.05; **, P < 10-2; ***, P < 10-3) values. All the in vivo experiments were run with at least 6 tumors from 6 mice for each condition. Extended Data Extended Data Figure 1 POLQ is highly expressed in epithelial ovarian cancers (EOCs) and POLQ expression correlates with expression of HR genes Gene set enrichment analysis (GSEA) for expression of TransLesion Synthesis (TLS) (a) and polymerase (b) genes between primary cancers and control samples in 28 independent datasets from 19 different cancers types. Enrichment values (represented as a single dot for each gene in a defined dataset) were determined using the rank metric score to compare expression values between cancers and control samples. Dots above the dashed line reflect enrichment in cancer samples, whereas dots below the dashed line show gene expression enriched in control samples. Datasets were ranked based on the amplitude of the rank metric score and plotted as shown. c, POLQ gene expression in 40 independent datasets from 19 different cancer types. For each dataset, POLQ values were expressed as fold-change differences relative to the mean expression in control samples, which was arbitrarily set to 1. d, POLQ expression correlates with tumor grade and MKi67 gene expression in the ovarian TCGA (n=494 patients with ovarian carcinoma (grade 1, n=5; grade 2, n=61; grade 3, n=428) and control samples, n=8). e, POLQ expression correlates with tumor grade MKi67 gene expression in the ovarian dataset GSE9891 (n=251 patients with ovarian serous and endometrious carcinoma for which grade status was available (grade 1, n=20; grade 2, n=88; grade 3, n=143)). Statistical correlation was assessed using the Pearson test (for d: r=0.65, P < 10-3; for e: r=0.77, P < 10-3). f, Top-ranked biological pathways differentially expressed between samples expressing high levels of POLQ (high POLQ, 1st 33%, n=95) relative to samples with low POLQ expression (low POLQ, 67%, n=190) on the ovarian dataset GSE9891 (n=285 patients with ovarian carcinoma). Significance values were determined by the hypergeometrical test using the 200 most differentially expressed probesets between the 2 groups (high POLQ and low POLQ). g, GSEA for expression of DNA repair genes between primary cancers and control samples in 5 independent ovarian cancer datasets. A representative heat map showing differential gene expression between ovarian cancers and controls is shown from GSE14407. For each dataset, DNA repair genes were ranked based on the metric score reflecting their enrichment in cancer samples. The top 20 DNA repair genes primarily expressed in cancer samples compared to control samples is shown on the right. h, GSEA for the top 20 DNA repair genes defined in (g) between primary cancers and control samples in 40 independent cancer datasets. The nominal P-value was used as a measure of the expression enrichment in cancer samples and represented as a waterfall plot. When the gene set expression was enriched in control samples, the P-value was arbitrarily set to 1. i, POLQ expression correlates with RAD51 and FANCD2 gene expression in 285 samples from the ovarian dataset GSE9891. Statistical correlation was assessed using the Pearson test (r=0.71, P < 10-3). j, Top 10 genes that most closely correlated with POLQ expression (gene neighbors analysis) for 1046 cell lines from the CCLE collection. DNA repair activity for these genes is indicated in the table. Increased HR gene expression is known to positively correlate with improved response to platinum based chemotherapy (a surrogate of HR deficiency) and thus can be predictive of decreased HR activity 31,38 . Conceptually, a state of HR deficiency may lead to compensatory increased expression of other HR genes. k, Top-ranked Gene Ontology (GO) terms for the molecular functions encoded by the top 20 DNA repair genes defined in Extended Data Figure 1g. l, Schematic representation of POLQ domain structure with the helicases (BLM, RECQL4, RAD54B and RAD54L) that co-expressed with POLQ (from Extended Data Figure 1g). Conserved amino-acid sequences of ATP binding and hydrolysis motifs (namely Walker A and B) are indicated. Cox plots in c that show twenty-fifth to seventy-fifth percentiles, with lines indicating the median, and whiskers indicating the smallest and largest values. For d and e (top panels), each dot represents the expression value from one patient, brackets show mean ± s.e.m. Extended Data Figure 2 POLQ is a RAD51-interacting protein required for maintenance of genomic stability a, siRNA sequences (siPOLQ1 and siPOLQ2) efficiently down-regulate exogenously transfected POLQ protein. POLQ levels were detected by immunoblotting with Flag or POLQ antibody (left) and by RT-qPCR using 2 different sets of POLQ primers (right). The asterisk on the immunoblot indicates a non-specific band. Expression was normalized using GAPDH as a reference gene. POLQ gene expression values are displayed as fold-change differences relative to the mean expression in control cells, which was arbitrarily set to 1. b, Quantification of baseline and HU-induced RAD51 foci in U2OS cells transfected with indicated siRNA. c, Quantification of baseline and HU-induced γH2AX foci in U2OS cells transfected with indicated siRNA. d, Quantification of IR-induced RAD51 foci in BrdU-positive U2OS cells transfected with indicated siRNA. e, POLQ inhibition by siRNA induced a decrease in the cellular survival of 293T cells treated with MMC in a 3-day survival assay. f, Quantification of chromosomal aberrations in 293T cells transfected with indicated siRNA. g, Schematic representation of POLQ truncation proteins used for RAD51 interaction studies. h, Endogenous RAD51 co-precipitates with Flag-tagged POLQ-ΔPol1 (POLQ-1-1416) but not POLQ-1633-Cter, each stably expressed in HeLa cells. i, Sequence alignment between the RAD51-interacting motifs of C. elegans RFS-1 and human POLQ. j, Schematic of POLQ domain structure with its homologs HELQ and POLN. All data show mean ± s.e.m. Extended Data Figure 3 Characterization of RAD51-interacting motifs in POLQ a, Substitution peptide array probed with recombinant RAD51 and analyzed by immunoblotting. A 20-mer peptide spanning each of the RAD51 binding sites (shown in Fig. 1g) were created in which each amino acid of the original peptide was mutated to each of the 20 amino acids and RAD51 binding activity was tested. The amino acid change for each of the amino acids of the RAD51 interacting domain of POLQ is shown on the right. Ponceau staining was used to visualize position of the peptides within the array. b, GST-RAD51 pull-down with in vitro-translated POLQ proteins missing indicated amino acids. c, Schematic of POLQ mutants used in complementation studies. d, Quantification of IR-induced RAD51 foci in U2OS cells stably integrated with empty vector (EV) or POLQ- ΔPol1 cDNA, that is refractory to siPOLQ1. Cells were transfected with indicated siRNA and subsequently treated with IR. The number of cells with more than 10 RAD51 foci was calculated relative to control cells (si Scr). e, DR-GFP assay in U2OS cells stably integrated with empty vector (EV) or indicated POLQ cDNA constructs refractory to siPOLQ1 and transfected with indicated siRNA. All data show mean ± s.e.m. Extended Data Figure 4 POLQ is an ATPase that suppresses RAD51-ssDNA nucleofilament assembly and formation of RAD51-dependent D-loop structures a, Representative ΔPol2 WT radiometric ATPase assay. b, Gel mobility shift assays with ΔPol2 WT and ssDNA. c, Coomassie-stained gel showing the purified ΔPol2-A-dead fragment. d, Representative ΔPol2-A-dead radiometric ATPase assay. e, Quantification of ΔPol2-A-dead ATPase activity. (ssDNA: single-stranded DNA; dsDNA: double-stranded DNA). f, Assembly/disruption of RAD51-ssDNA filaments in the presence of increasing amounts of ΔPol2 WT. The order in which each component was added to the reaction is noted above. g, Schematics of the formation of RAD51-dependent D-loop structures. h Formation of RAD51-containing D-loop structures following the addition of increasing amounts of ΔPol2 WT. i, Fraction of D-loop formed following the addition of increasing amounts of ΔPol2 WT. j, Effect of POLQ expression levels and HR status on tumor sensitivity to cisplatin or PARPi. Data in i shows mean ± s.e.m. Extended Data Figure 5 POLQ functions under replicative stress and is induced by HR deficiency a, POLQ recruitment to the chromatin is enhanced by UV treatment. HeLa cells stably integrated with either Flag-tagged ΔPol1 or POLQ-1633-Cter (Extended Data Fig. 2g) were subjected to UV treatment. Cells were collected at indicated time points after UV treatment and IPs were performed on nuclear and chromatin fractions. b, HeLa cells stably integrated with ΔPol1 were treated with UV and harvested at indicated time points following UV exposure. POLQ and RAD51 co-precipitation is enhanced by UV treatment. c, Quantification of DNA fiber lengths isolated from WT or Polq-/- MEFs. d, Quantification of DNA fiber lengths isolated from WT or Polq-/- MEFs transfected with either EV, or POLQ cDNA constructs. e, POLQ gene expression was analyzed by RT-qPCR in HR-deficient ovarian cancer cell lines (PEO-1 and UWB1-289) compared with other ovarian cancer cell lines, HeLa (cervical cancer) cells and 293T (transformed human embryonic kidney) cells. Expression was normalized using GAPDH gene as a reference. POLQ expression values are displayed as fold-change relative to the mean expression in HR-proficient control cells, which was arbitrarily set to 1. f, POLQ gene expression analysis (RT-qPCR) in 293T cells transfected with siRNA targeting FANCD2, BRCA1 or BRCA2 (left panel) and in corrected PD20 cells (PD20 + FANCD2) relative to FANCD2-deficient cells (PD20) (right panel). Expression was normalized using GAPDH gene as a reference. POLQ expression values are presented as fold-change relative to the mean expression in control cells, which was arbitrarily set to 1. g, POLQ gene expression in 5 datasets of serous epithelial ovarian carcinoma (frequently associated with an HR deficiency) and 1 dataset of clear cell ovarian carcinoma (subgroup not associated with HR alterations). For each dataset, POLQ expression values are displayed as fold-change differences relative to the mean expression in control samples, which was arbitrarily set to 1. h, Progression-free survival (PFS) after first line platinum chemotherapy for patients with ovarian carcinoma (ovarian carcinoma TCGA). Statistical significance was assessed by the Log-Rank test (P < 10-2). i, Effect of siPOLQ and the different POLQ cDNA constructs on HR read-out. NA: not applicable. Box plots in c, d, and g show twenty-fifth to seventy-fifth percentiles, with lines indicating the median, and whiskers indicating the smallest and largest values. Data in e and f show mean ± s.e.m. Extended Data Figure 6 POLQ inhibition sensitizes HR-deficient tumors to cytotoxic drug exposure Clonogenic formation of A2780 cells expressing Scrambled (Scr) shRNA or shRNAs against FANCD2 or BRCA2 with increasing amounts of MMC (a), UV (b) or IR (c). Clonogenic formation of A2780 cells expressing Scrambled (Scr) or FANCD2 shRNA, together with shRNA targeting POLQ, in increasing concentrations of CDDP (d), MMC (e) or PARPi (f). g, Inhibition of POLQ reduces the survival of A2780 cells after 3 days of continuous exposure to the ATM inhibitor Ku55933. h, Immunoblot analyses in A2780 cells expressing FANCD2 shRNA together with siRNA targeting POLQ or Scr at 24 hours after indicated MMC pulse treatment. i, FANCA-deficient fibroblasts (GM6418) were infected with a whole-genome shRNA library and treated with MMC for 7 days. The fold-change enrichment of each shRNA after MMC treatment was determined by sequencing relative to the infected cells before treatment. TP53 depletion is known to improve survival of FANCA-/- cells 33 . WRN depletion has recently been shown to be synthetically lethal with HR deficiency 39 . Each column represents the mean of at least 2 independent shRNAs. All data show mean ± s.e.m. Extended Data Figure 7 HR and POLQ repair pathways are synthetical lethal in vivo a, Genotypes frequencies of offspring from interbred Fancd2+/−Polq+/− mice. Ψ: four Fancd2-/-Polq-/- offsprings were observed with several congenital malformations and premature death within 48 hours of birth. b, Description of Fancd2−/−Polq−/− offspring generated in the study. The offspring presented congenital malformations (i.e., eye defects) together with reduced size and body weight. The arrow shows absence of the right eye. c, Genotypes frequencies of E13.5 to E15 embryos (13.5 to 15 days post coitum) from interbred Fancd2+/−Polq+/− mice. d, Description of congenital malformations and their measured frequencies observed in E13.5 to E15 Fancd2−/−Polq−/− embryos generated in the study. e, Clonogenic formation of WT, Fancd2-/- , Polq-/- and Fancd2−/−Polq−/− MEFs with increasing concentrations of PARPi. f, A2780 cells were transduced with indicated shRNAs and xenotransplanted into both flanks of athymic nude mice. The tumor volumes for individual mice were measured biweekly for 8 weeks. Each group represents n ≥ 5 tumors from n ≥ 5 mice. g, Ki67 and γH2AX quantification in tumors treated with either vehicle or PARPi. h, Representative Ki67 and γH2AX staining of A2780-shFANCD2 xenografts expressing sh Scr or sh POLQ in athymic nude mice, treated with either vehicle or PARPi. Scale bars, 100 μM. i, In vivo competition assay design. j, Tumor chimerism post xenotransplantation for indicated conditions. k, Representative flow cytometry analysis of tumors before xenotransplantation (post FACS sorting) or after xenotransplantation (post transplant, PARPi). The percentage of GFP-RFP cells is indicated. l, Tumor chimerism post xenotransplantation for indicated conditions. For Data in j and l, each circle represents data from one tumor and each group represents n ≥ 7 tumors from n ≥ 6 mice. Brackets show mean ± s.e.m. Data in e-g show mean ± s.e.m. For f each group represents n ≥ 6 tumors from n ≥ 6 mice. Extended Data Figure 8 POLQ is required for HR-deficient cell survival and limits the formation of RAD51 structures in HR-deficient cells a, Clonogenic formation of Fancd2-/-Polq-/- MEFs transfected with full-length POLQ cDNA constructs in the presence of increasing concentrations of PARPi. b, Chromosome breakage analysis of FANCD2-depleted cells that were first transfected with the indicated siRNA and full-length POLQ cDNA constructs refractory to siPOLQ1 and then exposed to MMC. c, DR-GFP assay in U2OS cells transfected with indicated siRNA. d, Quantification of baseline and IR-induced RAD51 foci in U2OS cells transfected with indicated siRNA. e, RAD51 recruitment to chromatin is enhanced by UV treatment. Vu423 cells (BRCA2-/-) were collected at indicated time points after UV treatment and immunoblotting performed on the cytoplasmic, nuclear and chromatin fractions. f, RAD51 recruitment to chromatin in Vu423 cells (BRCA2-/-) transfected with indicated siRNA. Histone H3 was used as a control for chromatin fractionation. All data show mean ± s.e.m. Extended Data Figure 9 POLQ participates in error-prone DNA repair a, End-joining reporter assay in U2OS cells transfected with indicated siRNA and/or treated with PARPi. b, End-joining reporter assay in U2OS cells transfected with indicated siRNA and POLQ cDNA constructs refractory to siPOLQ1. c, UV damage-induced POLQ foci formation in U2OS cells. POLQ foci were abolished by pre-treatment with PARPi. d, Mutation frequency was determined in damaged supF plasmid, recovered from siRNA-treated 293T cells. e, Non-synonymous mutation count in ovarian, uterine and breast TCGA. All data show mean ± s.e.m. Extended Data Figure 10 Model depicting the role of POLQ in DNA repair a, Mechanistic model for how POLQ limits RAD51-ssDNA filament assembly. According to this model, the ATPase domain of POLQ may prevent the assembly of RAD51 monomers into RAD51 polymers, perhaps by depleting local ATP concentrations. The RAD51 binding domains in the central region of POLQ may then sequester the RAD51 monomers, preventing filament assembly. b, I. Under physiological conditions, POLQ expression is low and its impact on repqir of DNA double-strand breaks (DSB) is limited. II. When HR deficiency occurs, POLQ is then highly expressed and channels DSB repair toward alt-EJ. III. In the case of an HR-defect, the loss of POLQ leads to cell death through the persistence of toxic RAD51 intermediates and inhibition of alt-EJ. Supplementary Material table 1 table 2 table 3

          Related collections

          Most cited references28

          • Record: found
          • Abstract: found
          • Article: not found

          Alternative-NHEJ Is a Mechanistically Distinct Pathway of Mammalian Chromosome Break Repair

          Introduction Faithful repair of DNA damage is essential to suppress genetic instability and tumorigenesis. Conversely, the efficacy of cancer therapies that utilize DNA damaging agents is likely limited by the ability of cancer cells to repair such damage. One form of DNA damage that is prone to causing mutations is a chromosomal double-strand break (DSB), which can result from DNA replication, reactive oxygen species, radiation therapy, and some types of chemotherapy [1]. Characterizing the factors and pathways of DSB repair is important to understand the process of mutagenesis during cancer development and treatment. Non-homologous end joining (NHEJ) is a major pathway of DSB repair, in which the ends are ligated without the use of extensive homology. NHEJ appears to comprise both classical-NHEJ and alternative-NHEJ (alt-NHEJ). Classical-NHEJ requires a number of factors important for V(D)J recombination, including the KU70/80 heterodimer (KU), XRCC4, Ligase IV, and DNA-PKcs [2],[3]. Also, classical-NHEJ is predicted to result in minimal processing of the DSB during repair [3],[4]. In contrast, alt-NHEJ appears to be independent of the above factors, and often results in a deletion with microhomology at the repair junction [4]–[12]. Genetic rearrangements consistent with alt-NHEJ have been observed in chromosomal translocations associated with both spontaneous and therapy-related cancer [13], and in reversion mutations of BRCA2 following DNA damage caused by PARP-inhibition [14]. Thus, alt-NHEJ-derived mutations appear to be associated with cancer development and may result from some cancer therapeutics. In contrast to the NHEJ pathways, homology-directed repair (HDR/GC) and single-strand annealing (SSA) employ significant degrees of homology [15]. HDR/GC utilizes a homologous template for gene conversion (GC) through strand-invasion and nascent DNA synthesis. HDR/GC is most precise when the identical sister chromatid is used as the template for repair. Thus, factors that are important for HDR/GC might be expected to be genome stabilizing. In contrast to HDR/GC, SSA involves annealing of homologous single strands to bridge the ends of the DSB, resulting in a deletion between the repeats. Such deletions have been observed between homologous segments of ALU elements in germ-line mutations of several tumor suppressor genes [16]. It is not clear to what degree alt-NHEJ is mechanistically distinct from SSA or even HDR/GC in mammalian cells. We sought to examine this mechanistic distinction by developing an assay for alt-NHEJ repair of a chromosomal DSB, where the predominant repair product is a 35-nucleotide (nt) deletion with 8 nt of microhomology at the repair junction. We have used this assay, along with a novel method for inducible control of the I-SceI endonuclease in stable cell lines, for a comparative genetic analysis of alt-NHEJ, SSA, and HDR/GC. From these studies, we found that alt-NHEJ shares KU/CtIP-mediated regulation of end-processing in common with SSA and HDR/GC, but involves a unique mechanism for completion of repair with regards to the role of ERCC1, RAD52, and RAD51. Results We have sought to understand the genetic relationship between multiple pathways of DSB repair in mammalian cells, since individual repair pathways show a different propensity for mutagenesis. For this, we used a series of chromosome-integrated reporters to monitor the repair of DSBs induced by the I-SceI endonuclease. Each individual reporter is designed such that repair of I-SceI-induced DSBs by a specific pathway restores a GFP expression cassette. Such repair can then be scored in individual cells as green fluorescence using flow cytometric analysis (FACS). In each reporter-containing cell line, the generation of GFP+ cells is confirmed to be absolutely dependent upon expression of I-SceI (data not shown). Total-NHEJ Results in a Variety of Repair Products We have developed two GFP-based chromosomal reporters to measure NHEJ. The first reporter, EJ5-GFP, detects multiple classes of NHEJ events, and thus can be considered an assay for total-NHEJ. We have presented this reporter mostly to provide context for the other reporter (EJ2-GFP), which is designed to monitor only alt-NHEJ events. EJ5-GFP is modeled after other reporters for NHEJ [4],[6],[9], in that it measures repair between two tandem endonuclease cut sites. Specifically, EJ5-GFP contains a promoter that is separated from a GFP coding cassette by a puro gene that is flanked by two I-SceI sites that are in the same orientation (Figure 1A). Once the puro gene is excised by NHEJ repair of the two I-SceI-induced DSBs, the promoter is joined to the rest of the expression cassette, leading to restoration of the GFP+ gene. Since the two I-SceI-induced DSBs have complementary 3′ overhangs, such NHEJ could potentially restore an I-SceI site. Alternatively, NHEJ could fail to restore the I-SceI site, leading to an I-SceI-resistant site. In addition, a restored I-SceI site could also be re-cleaved and repaired to result in an I-SceI-resistant site. 10.1371/journal.pgen.1000110.g001 Figure 1 Total-NHEJ repair between two tandem I-SceI sites results in a variety of products. (A) EJ5-GFP is shown along with two classes of NHEJ repair products that can restore a GFP expression cassette: one that restores an I-SceI site (I-SceI+), and one that is I-SceI–resistant (I-SceI-). (B) Restoration of the I-SceI site is common in wild-type cells, but undetectable in KU-deficient cells. EJ5-GFP was integrated into HEK293, wild-type ES, and Ku70-/- ES cells. Following transient I-SceI expression in each of these cell lines, GFP+ cells were sorted, and the GFP genes were amplified from these samples using primers depicted in (A). Shown are these products digested with I-SceI or left uncut. (C) The overall frequency of total-NHEJ is unaffected by KU deficiency. Shown is the frequency of repair of EJ5-GFP resulting in GFP+ cells from wild-type and Ku70-/- ES cells transfected in parallel with an I-SceI expression vector. Also shown are Ku70-/- ES cells cotransfected with both I-SceI and KU70 expression vectors. To determine the relative contribution of these different NHEJ products from repair of EJ5-GFP in mammalian cells, we integrated EJ5-GFP into both wild-type mouse embryonic stem (ES) cells, as well as transformed human embryonic kidney (HEK293) cells (see Materials and Methods). Following transient expression of I-SceI in these cell lines and sorting GFP+ cells, we amplified the GFP genes and digested the products with I-SceI. From this analysis, we found evidence of I-SceI-restoration in approximately 40% of the total products from both ES and HEK293 cells (Figure 1B). Regarding the other events (60%), we cloned I-SceI-resistant products from the ES cell line sample and sequenced individual clones (Table S1). Based on these sequences, the I-SceI-resistant NHEJ products showed deletions between 8–27 nucleotides, where the majority of clones (11/12) showed 2–4 nucleotides of microhomology at the junctions. Thus, NHEJ repair of the EJ5-GFP reporter results in either restoration of the I-SceI site, or generation of deletion NHEJ events, often with microhomology at the junctions. In previous studies with similar NHEJ reporters, KU-deficient cells showed a defect in restoration of the I-SceI site [4],[6]. To test this notion further, we integrated EJ5-GFP into Ku70-/- ES cells. We then transfected this line with an I-SceI expression vector, and subsequently measured the frequency of NHEJ events that resulted in a GFP+ gene, and quantified the restoration of the I-SceI site as described above. In these experiments, we found that Ku70-/- ES cells showed approximately equivalent overall frequencies of repair relative to wild-type cells (2.2% and 2.5% respectively, Figure 1C). However, PCR analysis of the repair products in the Ku70-/- cells showed only I-SceI-resistant products (Figure 1B). These results suggest that restoration of the I-SceI site during NHEJ repair is absolutely KU-dependent, but that I-SceI-resistant NHEJ events are KU-independent. In summary, EJ5-GFP provides an assessment of total-NHEJ events, which comprises both KU-dependent restoration of an I-SceI site, as well as deletion products with some evidence of microhomology at the junctions. Alt-NHEJ Is Suppressed by KU in Mammalian Cells We have chosen to focus on the subset of total-NHEJ events that show evidence of microhomology at the junctions, also called alt-NHEJ events. For this, we developed a novel reporter (EJ2-GFP), which is designed so the GFP+ products would predominantly reflect a discrete alt-NHEJ event. This reporter contains a single expression cassette for an N-terminal tag (NLS/Zinc-finger, [17]) fused to GFP, except the coding sequence is disrupted between the tag and GFP by an I-SceI site followed by stop codons in all three reading frames (Figure 2A). As well, the I-SceI site and stop codons are flanked by 8 nts. of microhomology, which if annealed during alt-NHEJ would restore the coding frame between the tag and GFP, and cause a 35 nt deletion. This alt-NHEJ repair product also generates an XCM1 restriction site. 10.1371/journal.pgen.1000110.g002 Figure 2 Alt-NHEJ is suppressed by KU. (A) EJ2-GFP is shown with 3 NHEJ products that are found to result in GFP+ cells. Shown in the ovals are the relative contributions of these products, based on the analysis shown in (B). The predominant GFP+ product, labeled Xcm1+, uses 8 nts of homology flanking the I-SceI site to generate an XCM1 site, resulting in a 35 nt deletion. (B) Analysis of EJ2-GFP repair products that restore the GFP+ gene. EJ2-GFP was integrated into HEK293, wild-type ES, and Ku70-/- ES cells. Using the primers shown in (A), the GFP genes were amplified from the parental ES EJ2-GFP cell line, and also from sorted GFP+ cells from each of the above cell lines following transient I-SceI expression. Shown are these amplification products, which were either uncut or cut with XCM1. (C) Repair by alt-NHEJ (EJ2-GFP) is suppressed by KU. Shown are the frequencies of alt-NHEJ repair, following transient I-SceI expression, for the wild-type and Ku70-/- EJ2-GFP cell lines, along with the Ku70-/- line co-transfected with an expression vector for KU70. Also shown are parallel experiments with the SA-GFP and DR-GFP reporters (see Figure 3). Asterisks denote a statistical difference in repair efficiency between Ku70-/- versus both wild-type, as well as Ku70-/- with transient expression of KU70 (p<0.0005). We determined the contribution of the XCM1+ product relative to total GFP+ repair products of EJ2-GFP, as integrated in ES and HEK293 cells. For this, we sorted GFP+ cells that resulted from I-SceI expression, amplified the GFP genes by PCR, and digested the amplification products with XCM1. From these experiments, we found that the XCM1+ product accounts for approximately 85% of the total repair products in both ES and 293 cells (Figure 2A and 2B). In addition to this predominant repair event, GFP+ products derived from EJ2-GFP also include a few minor repair events, which we identified by sequencing of cloned PCR products (Figure 2A and 2B; see Table S2). For instance, one minor repair event involves a 23 nt deletion with no evidence of microhomology at the junctions (approximately 10% of total events). The final set of events showed larger deletions, which ranged between 140–350 nt and showed microhomology at the repair junctions (approximately 5%). The larger deletion products apparently restore a GFP+ cassette because the GFP start codon was placed proximal to the transcription start site (unpublished data). In summary, while GFP+ products derived from EJ2-GFP can include some minor repair events, the predominant event (XCM1+) involves 8 nt of microhomology and a 35 nt deletion, which is characteristic of alt-NHEJ [4]. From previous studies [4],[6], and the above experiments with EJ5-GFP (Figure 1B and 1C), alt-NHEJ appears to be KU-independent. To investigate this notion further, we compared the efficiency of EJ2-GFP repair in wild-type and Ku70-/- ES cells following transfections with an I-SceI expression vector. We found that the Ku70-/- cells exhibited a 4-fold increase in the restoration of the GFP+ gene over wild-type cells, and that this increase was reversed by co-transfection of a KU70 expression vector (Figure 2C). Furthermore, analysis of GFP+ products from Ku70-/- showed a similar pattern as wild-type cells, in that the XCM1+ alt-NHEJ product was predominant (Figure 2B). Thus, the alt-NHEJ repair events measured by EJ2-GFP are not only KU-independent, but also appear to be inhibited by KU. In relation to other pathways, KU also suppresses HDR/GC and SSA, as described previously [18], and as confirmed in parallel experiments with EJ2-GFP (Figure 2C, see Figure 3). Given that KU-deficiency can lead to elevated DSB end-processing [19],[20], these results raise the possibility that alt-NHEJ, SSA, and HDR/GC share such end-processing as a common intermediate. 10.1371/journal.pgen.1000110.g003 Figure 3 An inducible system for I-SceI in stable cell lines used to show that siRNA-mediated disruption of CtIP affects multiple repair pathways. (A) Shown is the structure of the DR-GFP reporter along with the HDR/GC repair product that results in GFP+ cells, as described previously in ES cells [18]. (B) System for inducible control of I-SceI in stable cell lines. Cell lines were established with ES cells and HEK293 cells that contain the DR-GFP reporter and stable expression of the TAM-I-SceI-TAM (TST) fusion protein. These cell lines were either left untreated, or treated with 4-hydroxytamoxifen (4OHT) for a limited time (8 h for ES, 24 h for HEK293), and analyzed 3 d after starting the treatment. Shown are flow cytometric (FACS) profiles of 105 cells, where green fluorescence is plotted on the y-axis and auto orange fluorescence is on the x-axis. (C) Shown is the structure of SA-GFP reporter along with the GFP+ product of SSA repair. As discussed previously, HDR/GC associated with crossing over does not likely contribute significantly to this assay [18]. (D) CtIP promotes alt-NHEJ, SSA and HDR, but is dispensable for total-NHEJ. HEK293 cell lines with individual reporters were exposed to control siRNA (siCTRL), a pool of three CtIP-targeting siRNAs (siCTIP-p), or a distinct single CtIP-targeting siRNA (siCTIP-1). Subsequently, I-SceI was activated by 4OHT, and repair was measured as in (B). Shown are repair frequencies relative to the mean value of siCTRL samples treated in parallel. Asterisks denote a statistical difference from siCTRL with the substrates EJ2-GFP, SA-GFP, DR-GFP, and EJ5-GFP for both siCTIP-p (p<0.0001, p = 0.0012, p<0.0001, and p = 0.0009, respectively) and siCTIP-1 (p = 0.0021, p = 0.0002, p<0.0001, and p = 0.0023, respectively). Inducible System for I-SceI in Stable Cell Lines To continue to test the above hypothesis, we sought to perform siRNA experiments in HEK293 cells with the DSB reporters. However, we first would like to describe a novel technological approach for such siRNA experiments. In general, use of I-SceI-based reporters for such experiments would require transfection of the siRNA followed by a second transfection of the I-SceI-expression vector. Such serial transfections appear to cause increased toxicity, which can lead to variability between experiments (unpublished observations). Thus, we have developed a method for inducible activation of I-SceI in stable mammalian cell lines to bypass the need for a second transfection during siRNA experiments. Specifically, we used a mutant form of the estrogen receptor ligand binding domain, where in the absence of 4-hydroxytamoxifen (4OHT), this domain (TAM) appears to restrict access of fused proteins to chromosomes, while addition of 4OHT releases this restriction [21]. We made a series of expression vectors for fusion proteins between the TAM-domain and I-SceI (see Figure S1), and we chose to continue with an expression vector for TAM fused to both ends of I-SceI: TAM-I-SceI-TAM (TST). We generated stable cell lines expressing the TST fusion protein using a wild-type ES cell line and an HEK293 cell line, each containing an integrated copy of the DR-GFP reporter (see Materials and Methods). Repair of DR-GFP by the HDR/GC pathway results in the restoration of a GFP gene (Figure 3A), as previously described in mouse embryonic stem (ES) cells [18]. Following establishment of the TST-expressing cell lines, we analyzed 4OHT-dependent activation of I-SceI, as measured by GFP+ cells. From these experiments, we found low background levels of GFP+ cells from untreated samples, whereas 4OHT treatment resulted in an approximate 50-fold and 10-fold induction of GFP+ cells in ES cells and HEK293 cells, respectively (Figure 3B). Furthermore, we found that the low background levels were stable for at least 4-6 weeks of continuous culture (unpublished data). To measure not only HDR/GC, but also other repair pathways using this method, we subsequently developed similar TST stable cell lines with HEK293 cells containing the EJ2-GFP, EJ5-GFP, and SA-GFP reporters (see Materials and Methods). The SA-GFP reporter measures SSA (Figure 3C), as previously described in ES cells [18]. As discussed in this previous report, while it is formally possible that HDR/GC associated with crossing-over (CO) could also result in a GFP+ product from SA-GFP, two lines of evidence strongly suggest that CO events provide a negligible contribution to this assay. For one, multiple independent analyses have shown that CO during DSB repair in mammalian cells occurs at a frequency of less than 1% of the efficiency of the GFP+ repair events measured by SA-GFP [18],[22],[23]. As well, disruption of strand-exchange factors (BRCA2/RAD51) causes a significant increase in the efficiency of GFP+ repair of SA-GFP [18], which is inconsistent with a CO mechanism. In summary, we have generated HEK293 cell lines with stable expression of an inducible I-SceI (TST) and four different reporters to measure alt-NHEJ (EJ2-GFP), total-NHEJ (EJ5-GFP), SSA (SA-GFP), and HDR/GC (DR-GFP). The End-Processing Factor CtIP Promotes alt-NHEJ, SSA, and HDR/GC, But Is Dispensable for the Absolute Levels of total-NHEJ As described above, we sought to examine whether DSB end-processing may be a common mechanistic step in alt-NHEJ, SSA, and HDR/GC. For this, we focused on the factor CtIP [24], which is important for processing DSBs into ssDNA, detected as RPA-foci in mammalian cells following DNA damage [25],[26]. Regarding repair pathways, CtIP appears important for HDR/GC in both human cells and S. pombe, but is dispensable for plasmid end joining in S. pombe [25],[27]. We tested the hypothesis that CtIP in mammalian cells promotes not only HDR/GC, but also alt-NHEJ and SSA. For this, we performed siRNA knock-down of CtIP in the relevant HEK293 cell lines with individual reporters and stable expression of the inducible I-SceI protein (TST). We knocked-down CtIP levels using two different siRNA reagents: a pool of three siRNA duplexes (siCTIP-p), and a previously described single unique siRNA duplex (siCTIP-1)[25], (see Materials and Methods and Figure S1D). We compared these CtIP-depleted cells to control cells transfected with a non-targeting siRNA (siCTRL). We transfected each set of siRNAs into the HEK293 cell lines 48h prior to induction of I-SceI with 4OHT. The induction with 4OHT continued for 24h, and we assayed repair frequencies (%GFP+ cells) 3d after the start of the 4OHT treatment. We confirmed reduction in CtIP mRNA for both siCTIP-p and siCTIP-1 by RT-PCR of RNA isolated from parallel transfections at the onset of 4OHT addition (see Materials and Methods; unpublished data). From these experiments, we observed that HDR/GC, alt-NHEJ, and SSA were all significantly reduced in CtIP-depleted cells treated with either siCTIP-p (1.9-fold, 1.7-fold and 1.6-fold, respectively; Figure 3D) or siCTIP-1 (1.9-fold for each pathway; Figure 3D). In contrast, the absolute level of total-NHEJ was slightly increased in CtIP-depleted cells using either siCTIP-p or siCTIP-1 (1.3-fold and 1.4-fold, respectively; Figure 3D). Thus, CtIP appears to promote HDR/GC, alt-NHEJ, and SSA, but is dispensable for total-NHEJ. We suggest that CtIP-mediated DSB end-processing is important to generate ssDNA for the later steps of repair by HDR/GC, alt-NHEJ, and SSA. Alt-NHEJ Is Mechanistically Distinct from SSA and HDR/GC During Late Steps of Repair We also considered how alt-NHEJ might diverge from SSA and HDR/GC at later mechanistic steps. In particular, we addressed how factors important for completion of SSA may influence alt-NHEJ, since both pathways often involve annealing of flanking homology and subsequent processing of non-homologous single-stranded tails. Regarding SSA, these annealing and processing steps appear to be promoted by RAD52 and ERCC1, since cells deficient in these factors show a decreased level of SSA [18], and also because these factors possess relevant in vitro activities. RAD52 can function in vitro to directly promote homologous strand annealing, and also to mediate RAD51 function during strand exchange [28]. Though, only the strand annealing activity would be expected to be important for SSA in mammalian cells, since RAD51 appears to inhibit SSA [18]. ERCC1/XPF is a structure-specific endonuclease that catalyzes 5′ excision during nucleotide excision repair [29]. In addition, this complex shows efficient cleavage of 3′ overhangs, which could promote processing of non-homologous single-stranded tails during DSB repair [30]. Furthermore, ERCC1/XPF can form a complex with RAD52, which may suggest that single-strand tail processing and annealing may be coordinated by this complex during repair [31]. To directly examine the role of RAD52 and ERCC1 in alt-NHEJ, we integrated EJ2-GFP into Rad52-/- and Ercc1-/- ES cells (see Materials and Methods), and determined the fold-change in repair resulting from complementation with the relevant expression vector (i.e. RAD52 or ERCC1). Specifically, we transfected cells with an I-SceI expression vector along with either the relevant complementation vector or empty vector, and then assayed repair three days later as in Figure 2C. As well, previously described Rad52-/- and Ercc1-/- ES cell lines with DR-GFP and SA-GFP [18] were transfected in parallel. These experiments showed that the efficiency of SSA (SA-GFP) increased upon complementation with each of the relevant expression vectors (3.8-fold for ERCC1, Figure 4A; 1.9-fold for RAD52, Figure 4B). In contrast, the efficiency of alt-NHEJ and HDR/GC only slightly increased by complementation with the expression vector for ERCC1 (1.5-fold, and 1.4-fold, respectively; Figure 4A), and mildly decreased by complementation with RAD52 (1.4-fold reduced, and 2-fold reduced, respectively; Figure 4B). These absolute measurements of alt-NHEJ could include any of the products that result in a GFP+ gene (see Figure 2A). Notably, while ERCC1 complementation promotes each pathway to some extent, the effect is significantly greater for SSA, as compared to alt-NHEJ and HDR/GC (2.5-fold and 2.7-fold, respectively). These results indicate that alt-NHEJ is mechanistically distinct from SSA, in that this pathway is both less dependent upon ERCC1 and is not promoted by RAD52. 10.1371/journal.pgen.1000110.g004 Figure 4 The roles of ERCC1, RAD52, and RAD51 during alt-NHEJ, HDR/GC, and SSA. (A) While ERCC1 significantly promotes SSA, it plays a minor role in HDR/GC and alt-NHEJ. Ercc1-/- ES cell lines with EJ2-GFP, SA-GFP, and DR-GFP were transfected with an I-SceI expression vector, along with either an expression vector for ERCC1 or empty vector (EV). Shown are the levels of repair relative to the mean value of a parallel set of EV transfections, which allows a direct comparison of the effect of complementation on the different reporters. Asterisks denote a statistical difference in repair relative to EV (alt-NHEJ and SSA, p<0.0001; DR-GFP, p = 0.0066), and the dagger denotes a statistical difference in the level of complementation relative to SA-GFP (p<0.0001). (B) RAD52 promotes SSA but not HDR/GC or alt-NHEJ. Rad52-/- ES cell lines with the reporters shown in (A) were transfected with an I-SceI expression vector, along with either an expression vector for RAD52 or empty vector. Shown are levels of repair as described in (A). Asterisks denote a statistical difference in repair relative to EV (alt-NHEJ, p = 0.0003; SA-GFP and DR-GFP, p<0.0001). (C) RAD51 promotes HDR/GC, inhibits SSA, and plays no clear role in alt-NHEJ. Wild-type ES cell lines with each of the reporters were cotransfected with an I-SceI expression vector along with either an expression vector for a BRC3 peptide derived from BRCA2, an expression vector for RAD51-K133R, or EV [18]. Shown are levels of repair calculated relative to EV as in (A). Asterisks denote a statistical difference in repair relative to EV (SSA, p<0.016; DR-GFP, p<0.0008). Finally, since the above studies showed several mechanistic similarities between alt-NHEJ and HDR/GC, we next considered a probable mechanistic distinction between these pathways. Namely, we suspected that alt-NHEJ might not require RAD51-mediated strand-exchange. To examine this, we used two dominant negative inhibitors of RAD51: BRC3 and RAD51-K133R. BRC3 is a short peptide derived from BRCA2 that can inhibit RAD51 function [32]. RAD51-K133R is a mutant peptide defective in ATP-hydrolysis that results in hyper-stable strand invasion intermediates [33]. We tested the effect of these peptides on repair of the EJ2-GFP, DR-GFP, and SA-GFP reporters in otherwise wild-type ES cells. For each cell line, we co-transfected the I-SceI expression vector along with vectors expressing either BRC3 or RAD51-K133R, and compared the efficiency of repair relative to cells transfected with I-SceI and empty vector. From these experiments, BRC3 and RAD51-K133R resulted in a 2.3-fold and 6-fold decrease in HDR/GC, respectively, and a 1.4-fold and 1.8-fold increase in SSA, respectively (Figure 4C), which is consistent with previous results [18]. In contrast, from parallel transfections with the EJ2-GFP ES cell line, BRC3 and RAD51-K133R showed no significant effect on alt-NHEJ repair (Figure 4C). Thus, alt-NHEJ is distinct from HDR/GC and SSA, in that it is not affected by disruption of RAD51 function. In summary, alt-NHEJ shows a number of mechanistic distinctions from SSA and HDR/GC during later steps of repair. Discussion Chromosomal DSBs can be repaired by a variety of pathways with distinct mechanistic requirements and potentials for mutagenesis. Given the role of mutagenesis during cancer development and treatment, it will be important to understand the mechanistic overlap of these pathways in detail. To this end, we have identified some mechanistic commonalities and differences between three DSB repair pathways: alt-NHEJ, SSA, and HDR/GC (Figure 5). 10.1371/journal.pgen.1000110.g005 Figure 5 Model for the mechanistic relationships between alt-NHEJ, SSA, and HDR/GC. Individual genetic factors, shown in ovals, are placed in the pathways based on the genetic analysis presented here, and other studies discussed in the text. End processing steps are shown as 5′ to 3′ resection, which need not be the precise mechanism in mammalian cells. The lengths of homologous annealing and 3′ end cleavage are modeled as being less extensive for both alt-NHEJ and HDR/GC relative to SSA. To begin with, each of these pathways appears to be similarly affected by factors implicated in the control of DSB end-processing, in that they are each suppressed by KU and promoted by CtIP. Such DSB end-processing probably involves 5′ to 3′ resection, as has been directly observed to extend several kilobases in S. cerevisiae [34]; however, the precise nature and extent of DSB end-processing has yet to be determined in vertebrate cells. For example, it is possible that ssDNA could be formed via chromatin remodeling followed by unwinding by a DNA helicase. In any case, activation of end-processing likely requires bypassing KU-mediated protection of DSB ends [19],[20],[35]. Such bypass may be initiated by disrupting the binding of KU with DNA. Alternatively, as KU is removed from DSBs, factors could increase the probability that KU-free ends are then processed, for example, by promoting open chromatin structures [36], and/or by activating the end-processing machinery. CtIP could function via any of these mechanisms during early steps of repair to promote HDR/GC, alt-NHEJ, and SSA. However, its ability to bind to the MRE11 complex and promote its nuclease activity, suggests it may directly promote the end-processing machinery to generate ssDNA [25],[26]. Alternatively, since CtIP is also a transcription factor [37], it could conceivably promote DSB end-processing by opening chromatin or affecting some other upstream process. Notably, CtIP is cell cycle regulated in mammalian cells and in S. pombe, showing its highest levels in S-phase through G2/M [27],[38],[39]. Thus, repair pathways that are promoted by CtIP, including alt-NHEJ, might be expected to be more prevalent in these later stages of the cell cycle. In general, further characterization of the nature and mechanism of end-processing in mammalian cells will lead to insight into the role of CtIP in regulating this process during repair. Along these lines, our findings that CtIP promotes repair of both EJ2-GFP and SA-GFP, which involve deletions of 35 nt and 2.7 kb, respectively, suggests that CtIP-mediated DSB end-processing can extend over a relatively wide-range of sizes. Following DSB end-processing that results in ssDNA as described above, the mechanisms of alt-NHEJ, SSA, and HDR/GC appear to diverge based on how the ssDNA is utilized during repair. For example, such ssDNA could allow either annealing of flanking homology for alt-NHEJ and SSA, or RAD51-mediated strand exchange during HDR/GC. Consistent with this notion, inhibiting RAD51 function disrupts only HDR/GC, such that RAD51 assembly on ssDNA likely commits repair to HDR/GC versus other pathways of repair. Considering the mechanisms of annealing and 3′ end-processing, we have observed that alt-NHEJ is slightly inhibited by RAD52, and is only moderately promoted by ERCC1. In contrast, SSA is significantly promoted by both of these factors. This mechanistic distinction may result from variations in the distance between homologous sequences, the length of the homology, and/or the absolute requirement for homologous annealing. For instance, RAD52 may play a specific role for annealing extensive regions of homology, and hence only promote SSA. This mechanism is supported by in vitro studies of RAD52, showing that its preferred binding substrate appears to be long stretches of ssDNA, though some binding to small regions of ssDNA can also be observed [40]. Similarly, the specific role for ERCC1 during SSA could reflect a necessity for this factor in cleaving particularly long 3′ single-stranded tails; however, inconsistent with this model, ERCC1/XPF shows significant cleavage activity on short (15 nt) single stranded tails [30]. Then again, alt-NHEJ may only rarely involve processing of 3′ single-stranded tails, and thus may often involve other intermediate structures that could be cleaved by a different nuclease complex. Notably, with regard to each of these mechanistic steps of alt-NHEJ, mammalian cells show both similarities and differences with yeast. For instance, our findings with KU/CtIP in mammalian cells are consistent with experiments in S. cerevisiae that showed KU-independence [6] and SAE2-activation [41] of alt-NHEJ, the latter of which may be relevant to mammalian cells assuming that SAE2 is a homologue of CtIP [25]. Regarding later steps of alt-NHEJ, the XPF homologue (RAD10) in S. cerevisiae is critical for this process [6], whereas RAD52 appears dispensable [42]. Thus, apart from the increased dependence on ERCC1/XPF for alt-NHEJ in yeast, these findings are similar to our results with EJ2-GFP in mammalian cells. In contrast, an S. pombe study on alt-NHEJ showed the opposite of the S. cerevisiae results, in that XPF (Rad16) appears dispensable, and RAD52 (Rad22) is important [12]. However, these S. pombe experiments were plasmid-based and involved microhomology very close to the end of the DSB. Similarly, a plasmid-based alt-NHEJ assay in S. cerevisiae also showed activation of repair by RAD52 [43]. In general, these distinctions highlight the notion that the mechanism of alt-NHEJ may be distinct between mammalian cells and yeast, but may also be affected by the length of homology, the distance separating the homologous segments, and/or the context of a DSB in a plasmid versus a chromosome. Reflecting such differences, alt-NHEJ pathways have been categorized using multiple names, each of which reflect certain features of a defined set of repair events: micro-SSA, microhomology-mediated end-joining (MMEJ), KU-independent end-joining, and backup-NHEJ (B-NHEJ) [4]–[12]. While it may be beneficial to find consensus on a particular term, the diversity of terminology also suggest the presence of multiple subclasses of NHEJ events. The predominant event measured by EJ2-GFP, described here as alt-NHEJ, is most similar to MMEJ, in that this product is KU-independent, shows evidence of microhomology at the junction, and results in a deletion. In contrast, other events could be mechanistically more akin to SSA or micro-SSA, with respect to extent of homology, the distance between homologous sequences, and RAD52/ERCC1-dependence. Furthermore, some repair events, while KU-independent, lack evidence of microhomology [4],[9], such that so-called KU-independent NHEJ or B-NHEJ may reflect a larger class of events relative to only MMEJ. Further analysis of the mechanisms of this variety of repair events will continue to clarify the subclasses of NHEJ. Among these different subclasses of NHEJ, alt-NHEJ/MMEJ appears to play a significant role in the etiology of mutations that arise during cancer development and treatment. For instance, a screen for PARP-inhibitor resistant BRCA2-mutant cells revealed a set of reversion mutations that are consistent with alt-NHEJ [14]. Thus, combination of PARP-inhibition and simultaneous disruption of alt-NHEJ may be effective in eliminating PARP-inhibitor resistant cancer cells. As supported by our findings with EJ2-GFP, a target for such therapy may include CtIP [37], whereas disruption of KU-dependent NHEJ pathways would be predicted to be ineffective. Though, PARP has been shown to play a role in plasmid-based NHEJ assays [11], such that it would be important to ensure that alt-NHEJ is targeted separately from PARP function. Similar to the BRCA2 example, tumors deficient in ERCC1 would also be predicted to be relatively proficient at repair of DSBs by alt-NHEJ, which is consistent with the notion that DSB-inducing agents may be less effective on these tumors than interstrand crosslinking agents [44]. Finally, since alt-NHEJ appears to play a significant role in therapy-induced oncogenic chromosomal translocations [13], targeting this pathway, again perhaps via CtIP, may enhance the efficacy of such therapy. In summary, further analysis of the mechanisms and mutagenic potential of individual DSB repair pathways will continue to inform the development of therapeutic approaches to cancer treatment. Materials and Methods Plasmids and Cell Lines The expression vector for the fusion protein of TAM-I-SceI-TAM (TST) was generated by PCR amplification of the TAM domain from TAM-CRE [21], and the I-SceI coding sequence from pCBASce, which were cloned in frame into pCAGGS-BSKX [45], as shown in Figure S1. The EJ2SceGFP gene (EJ2-GFP) was generated by cloning gcctagggataacagggtaattagatgacaagcc into the XCM1 site of pCAGGS-NZEGFP [46]. EJ2SceGFP was then cloned into pim-DR-GFP [47], and downstream of pgk-puro to generate pim-EJ2-GFP and EJ2-GFP-Puro, respectively. For EJ5-GFP, first an I-SceI site was cloned between the AgeI and BclI sites of pim-EJ2-GFP (EJ5sceGFP), and also at the HindIII site of pgk-puro (puroSce). Then, an EcoRI/I-SceI fragment of puroSce was cloned into EJ5SceGFP, followed by cloning an I-SceI site into the EcoRI site of this vector. Pim-EJ5-GFP was then completed by replacement of an EcoRI fragment that was lost in the previous step. ES cells were cultured as previously described [45], and HEK293 cells (HEK293-A7, New England Biolabs) were cultured according to the directions of the supplier, except we used DMEM high-glucose without phenol red containing Hepes buffer (Invitrogen). HEK293 cells were grown on plates treated with 0.01% poly-lysine (Sigma). Mouse ES cell lines with DR-GFP and SA-GFP targeted to hprt or Pim1 were described previously [18], [45]–[47]. Pim-EJ2-GFP was used to target the Pim1 locus of AB2.2 wild-type ES cells [48], and Ku70-/- ES cells [49], using methods previously described [45], except targeting was detected by PCR. Pim-DR-GFP, Hprt-SA-GFP, and EJ2-GFP-Puro were randomly integrated into HEK293 cells by electroporation with 1×107 cells suspended in 800 µl PBS in a 0.4 cm cuvette, followed by pulsing the cells at 250 V, 950 µF, and selecting single clones with 3 µg/ml puromycin. Similarly, EJ2-GFP-Puro was randomly integrated into Ercc1-/- and Rad52-/- ES cells as above, except using electroporation conditions of 680 V and 10 µF. Integration of an intact copy of each randomly integrated reporter was confirmed in single clones by Southern blot analysis with a GFP fragment as the probe (data not shown). Stable cell lines expressing TST were generated by electroporation as described above, except with voltages varying between 200–250V, with 20–30 µg of TST expression plasmid and a selection plasmid. We used two different selection cassettes, with 10 µg of pgk-bsd (gift from Dr. Pentao Liu) for the HEK293 and 5 µg of pmc1neo for the ES cells. Clones were selected in the relevant antibiotic for 6–10d at 400 µg/ml G418 or 5–10mg/ml blasticidin (Invitrogen). Individual selected clones were screened for significant induction of GFP+ cells following 24h treatment with 0.3 µM and 3 µM 4-hydroxytamoxifen (4OHT, dissolved in ethanol, Sigma) for ES and HEK293 cells, respectively. Repair Assays To measure the repair by transient transfection, 2.5×104 cells/cm2 were plated and transfected the next day with 0.8 µg/ml of pCBASce mixed with 3.6 µl/ml of Lipofectamine 2000 (Invitrogen) along with a variety of other vectors. The KU and RAD52 expression vectors were added at 0.8 µg/ml, the ERCC1 vector was added at 0.4 µg/ml, the RAD51-K133R vector was added at 0.1 µg/ml, and the BRC3 vector was added at 0.2 µg/ml. For each experiment, an equivalent amount of empty vector (pCAGGS-BSKX) was included in the parallel transfections. Each of these expression vectors have been previously described [18]. GFP positive cells were quantified by flow cytometric analysis (FACS) 3d after transfection on a Cyan ADP (Dako). Amplification of PCR products from sorted GFP+ cells, associated restriction digests, and quantification of bands were performed using the primers KNDRF and KNDRR as previously described for analysis of DR-GFP [50]. To measure repair using the inducible I-SceI protein (TST) in combination with siRNA-mediated inhibition of CtIP, HEK293 cell lines with each of the reporters and stable expression of TST were first plated on 24 well plates at 105 cells/well. The following day, the wells were transfected with 70nM siRNA duplex mixed with 4ul/ml of Lipofectamine 2000 in Optimem (Invitrogen). After 4.5h, transfection complexes were diluted two-fold with media without antibiotics, and 48h after the initiation of transfection, 4OHT was added at 3 µM for 24h. Three days after 4OHT was added, the percentage of GFP+ cells was analyzed by FACS as described above. Knockdown of CtIP levels using the various siRNAs was confirmed by RT-PCR from RNA samples isolated from parallel transfections at the time of 4OHT addition (data not shown). Amplification product was quantified at the threshold cycle by including SYBR green in the PCR reaction and using an iQ5 cycler for real-time analysis at the end of each cycle (BioRad). Products were normalized relative to a primer set directed against actin. Sequences of the siRNAs siCtIP-p (Santa Cruz Biotechnology), and siCtIP-1 [25], and primers for RT-PCR are shown in Figure S1D. Repair frequencies are the mean of at least three transfections or four 4OHT treatments, and error bars represent the standard deviation from the mean. For some experiments, repair frequencies are shown relative to samples co-transfected with I-SceI and an empty vector (EV). For this calculation, the percentage of GFP+ cells from each sample was divided by the mean value of the EV samples treated in the parallel experiment. Similarly, to calculate the fold-difference in repair between siRNA-treated and control-siRNA treated cells, the percentage of GFP+ cells from each sample was divided by the mean value of control-siRNA samples from the parallel experiment. Statistical analysis was performed using the unpaired t-test. Supporting Information Figure S1 Details of TAM-I-SceI fusion proteins. (A) Shown is a schematic for control of TAM-I-SceI fusion proteins using the hormone 4OHT. (B) Shown are the primer sequences used to generate expression vectors for TAM-I-SceI fusions from the parent vectors TAM-CRE, along with pCBASce and pCAGGS-BSKX. For TAM-I-SceI (TS): a PCR product of TAM-CRE, using Scetam1 and Scetam7, was cloned into EcoRI/SalI sites of pCAGGS-BSKX, followed by insertion of a BbsI/AvrII fragment of pCBASce. For SceTAM (ST), a PCR product of pCBASce, using Scetam3 and Scetam4, was cloned into EcoRI/BglII sites of pCAGGS-BSKX, followed by insertion into the BglII/XhoI sites of this vector with a PCR product of TAM-CRE using Scetam5 and Scetam8 digested with BamHI/SalI. For TamSceTam (TST): a PCR product of TAM-CRE, using Scetam1 and Scetam7, was cloned into EcoRI/BbsI sites of ST. (C) We tested each of the ST, TS, and TST plasmids by transient transfection into the DR-GFP ES cell line, followed by treatment with 4OHT for 24 h, or untreated. I-SceI activity is measured by induction of HDR/GC, as determined 3 d after the 4OHT treatment. In these experiments, we found that each of the plasmids conferred approximately equivalent I-SceI activity in the presence of 4OHT, while the TST fusion showed the lowest background activity in the absence of 4OHT. (D) Shown are the relevant sequences for the CtIP siRNA experiments, as described in Materials and Methods. (0.59 MB TIF) Click here for additional data file. Table S1 Repair junctions for EJ2-GFP. PCR products shown in Figure 1C from ES cells were cloned into the PCR2.1 vector (Invitrogen) according to the manufacturer's instructions, and 12 individual clones with detectable inserts were sequenced using the M13F primer. Shown is the sequence surrounding the two I-SceI sites (bold) in the parental EJ5-GFP reporter, along with repair products from sorted GFP+ cells. Products that were identified in multiple independent clones are noted in parentheses. Microhomology found at or near the junctions is underlined, and the length of microhomology is noted. The sizes of the deletions relative to the I-SceI+ product are also shown, starting from the 3′ end of the coding strand (shown as ATAA/ in the I-SceI+ product). (0.04 MB DOC) Click here for additional data file. Table S2 Repair junctions for EJ2-GFP. PCR products shown in Figure 2B from ES cells were cloned into the PCR2.1 vector (Invitrogen) according to the manufacturer's instructions, and individual clones with detectable inserts were sequenced using the M13F primer. Shown is the sequence surrounding the I-SceI site (bold) in the parental EJ2-GFP reporter, along with various repair products from sorted GFP+ cells. The sequence of the XCM1+ product was confirmed in a clone generated from the uncut PCR product. The sequence of the 23 nt deletion product was found in 3 clones from the XCM1-resistant PCR product, where the junction is marked by a hyphen for clarity. Regarding the larger deletions, 7 clones in total were sequenced, and one product was found twice, as noted in the parentheses. Microhomology found at or near the junctions is underlined, and the length of microhomology is noted, where a discontinuous tract of homology is noted as dis. The sizes of the deletions from the I-SceI cut site are also shown, starting from the 3′ end of the coding strand (shown as ATAA/ in the parent reporter). (0.04 MB DOC) Click here for additional data file.
            Bookmark
            • Record: found
            • Abstract: found
            • Article: not found

            Srs2 and Sgs1-Top3 suppress crossovers during double-strand break repair in yeast.

            Very few gene conversions in mitotic cells are associated with crossovers, suggesting that these events are regulated. This may be important for the maintenance of genetic stability. We have analyzed the relationship between homologous recombination and crossing-over in haploid budding yeast and identified factors involved in the regulation of crossover outcomes. Gene conversions unaccompanied by a crossover appear 30 min before conversions accompanied by exchange, indicating that there are two different repair mechanisms in mitotic cells. Crossovers are rare (5%), but deleting the BLM/WRN homolog, SGS1, or the SRS2 helicase increases crossovers 2- to 3-fold. Overexpressing SRS2 nearly eliminates crossovers, whereas overexpression of RAD51 in srs2Delta cells almost completely eliminates the noncrossover recombination pathway. We suggest Sgs1 and its associated topoisomerase Top3 remove double Holliday junction intermediates from a crossover-producing repair pathway, thereby reducing crossovers. Srs2 promotes the noncrossover synthesis-dependent strand-annealing (SDSA) pathway, apparently by regulating Rad51 binding during strand exchange.
              Bookmark
              • Record: found
              • Abstract: found
              • Article: not found

              Repair of double-strand breaks by end joining.

              Nonhomologous end joining (NHEJ) refers to a set of genome maintenance pathways in which two DNA double-strand break (DSB) ends are (re)joined by apposition, processing, and ligation without the use of extended homology to guide repair. Canonical NHEJ (c-NHEJ) is a well-defined pathway with clear roles in protecting the integrity of chromosomes when DSBs arise. Recent advances have revealed much about the identity, structure, and function of c-NHEJ proteins, but many questions exist regarding their concerted action in the context of chromatin. Alternative NHEJ (alt-NHEJ) refers to more recently described mechanism(s) that repair DSBs in less-efficient backup reactions. There is great interest in defining alt-NHEJ more precisely, including its regulation relative to c-NHEJ, in light of evidence that alt-NHEJ can execute chromosome rearrangements. Progress toward these goals is reviewed.
                Bookmark

                Author and article information

                Journal
                0410462
                6011
                Nature
                Nature
                Nature
                0028-0836
                1476-4687
                29 January 2015
                02 February 2015
                12 February 2015
                12 August 2015
                : 518
                : 7538
                : 258-262
                Affiliations
                [1 ]Department of Radiation Oncology, Dana-Farber Cancer Institute, Harvard Medical School
                [2 ]Department of Biological Chemistry & Molecular Pharmacology, Harvard Medical School
                [3 ]Department of Molecular & Cellular Biology, Harvard University, Cambridge
                [4 ]Howard Hughes Medical Institute, Department of Biological Chemistry and Molecular Pharmacology, Harvard Medical School
                [5 ]Howard Hughes Medical Institute, Division of Genetics, Brigham and Women's Hospital
                [6 ]DNA Damage Response Laboratory, Cancer Research UK, London Research Institute, Clare Hall
                [7 ]Department of Medical Oncology, Medical Gynecologic Oncology Program, Dana-Farber Cancer Institute, Harvard Medical School
                Author notes
                Correspondence and requests for materials should be addressed to A.D.D ( alan_dandrea@ 123456dfci.harvard.edu )
                Article
                NIHMS651721
                10.1038/nature14184
                4415602
                25642963
                7e401d34-125b-4196-84bb-7f1ac1c4740a
                History
                Categories
                Article

                Uncategorized
                Uncategorized

                Comments

                Comment on this article