• No results found

VU Research Portal

N/A
N/A
Protected

Academic year: 2021

Share "VU Research Portal"

Copied!
25
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

VU Research Portal

Loss-of-function shRNA screens to identify mechanisms of PARP inhibitor resistance in BRCA1-mutated mouse mammary tumors

Xu, G.

2016

document version

Publisher's PDF, also known as Version of record

Link to publication in VU Research Portal

citation for published version (APA)

Xu, G. (2016). Loss-of-function shRNA screens to identify mechanisms of PARP inhibitor resistance in

BRCA1-mutated mouse mammary tumors.

General rights

Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain

• You may freely distribute the URL identifying the publication in the public portal ?

Take down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

E-mail address:

vuresearchportal.ub@vu.nl

(2)

175

CHAPTER 5

(3)

176

Breast Cancer Incidence and Mortality Worldwide

Breast cancer is the second most common cancer in the world and the most common cancer in women both in the developed and developing countries. Based on the statistics from GLOBOCAN 2012 (IARC) it is estimated that worldwide over 1.67 million new cancer cases are diagnosed and over 500,000 women died of breast cancer in 2012. In addition to environmental and lifestyle factors, hereditary factors also lead to breast cancer risk (Lichtenstein et al., 2000). Among the hereditary factors dysfunction of the BRCA1 or BRCA2 tumor suppressors are the most well-known genetic risk factors for the development of breast or ovarian cancer thus far.

Mechanisms of tumor formation by loss of function of BRCA1/2

Both BRCA1 and BRCA2 have been shown to play a critical role in the repair of double-stranded DNA breaks (DSBs) by homologous recombination (HR) (Jackson and Bartek, 2009). DSBs within our DNA are repaired by HR or non-homologous end joining (NHEJ). Generally, HR utilizes the intact non-homologous DNA sequence from a sister chromatid to repair DSBs in an error-free manner, whereas NHEJ directly ligates two broken ends in a more error-prone manner. Due to the essential role of HR in genome maintenance, loss of function of BRCA1/2 leads to increased breast and ovarian cancer risks. Women with a BRCA1 or BRCA2 mutation have a ~50% chance of developing breast cancer by age 70 (Antoniou et al., 2003). So far hundreds of mutations have been discovered in the BRCA1/2 gene and more and more studies revealed that a certain mutation is common to a certain population. For example, Exon 2 deletion (Peelen et al., 1997), exon 13 deletion (Petrij-Bosch et al., 1997), and 2804delAA (Verhoog et al., 2001) account for the majority of Dutch BRCA1-related breast and/or ovarian cancer patients, due to founder mutations in the Dutch population.

(4)

177

have also been shown to bind to the BRCA2 promoter and to activate the expression of BRCA2, whereas SLUG negatively regulates BRCA2 expression by binding to the BRCA2 promoter (Tripathi et al., 2005). PI3K inhibition induced ERK activation that further increases ETS1 phosphorylation and activity to suppress BRCA1/2 expression (Ibrahim et al., 2012). BRCA2 protein levels are regulated by ubiquitin-mediated proteolysis in a process that requires its association with the E3 ubiquitin ligase family member Skp2 (Moro et al., 2006; Schoenfeld et al., 2004). Both miR-182 and miR-9 target BRCA1 transcripts and down-regulate BRCA1 expression (Moskwa et al., 2011; Sun et al., 2013) and miR-1245 suppresses BRCA2 translation (Song et al., 2012). In addition to

BRCA1/2 the tumor suppressor P53 is mutated in most BRCA1/2-related breast

cancers (Holstege et al., 2009).

HR and NHEJ in eukaryotic cells

Single-stranded DNA breaks (SSBs) by endogenous or exogenous sources are among the most common products of DNA damage in eukaryotic cells. If left unrepaired, SSBs can be converted into DSBs during DNA replication, and are then repaired by error-free HR or error-prone NHEJ. Briefly and in a simplified fashion, DSBs induce ATM‐dependent phosphorylation of histone H2AX. Then MDC1 recognizes the phosphorylation and ATM also phosphorylates MDC1 to promote RNF8 recruitment. RNF8, together with RNF168 ubiquitinate histone H2A to amplify the ubiquitin‐mediated DSB signaling. Ubiquitylation leads to recruitment of multiple effectors such as 53BP1 and BRCA1. 53BP1 promotes NHEJ throughout interphase, whereas BRCA1 promotes HR in the S/G2 phase

(5)

178

be completed through resolvase-mediated resolving of the hetero-duplexed DNA structure formed by the extended hetero-duplex, and subsequent ligation by DNA Ligase-I.

Synthetic lethality to PARP inhibitors (PARPi) of BRCA1/2-mutated cancers

When two mutations have no effect on cell viability when they occur individually, but concurrent mutation of both lead to cell death this is called synthetic lethality (Dobzhansky, 1946). This concept has been referred to in the context of identifying novel anti-cancer drugs. The goal was to use a tailored inhibitor which does not kill normal cells of the body, but induces lethality of the cancer cells because they lack activity of a specific gene. One successful example of this concept is the synthetic lethality of PARP inhibitors that is seen in BRCA1/2-related cancers. Due to the crucial role of BRCA1/2 in HR, loss of its function provides a therapeutical opportunity by targeting HR-deficiency. As novel and less toxic therapeutic approach for BRCA1/2-related cancers, inhibition of poly (ADP-ribose) polymerase (PARP) was proposed (Bryant et al., 2005; Farmer et al., 2005). PARP is a family of proteins involved in a number of cellular processes such as DNA repair and programmed cell death. PARP-1 is the most abundant enzyme in this family which detects and repairs SSBs by catalyzing the poly(ADP-ribosyl)ation of other DNA-repairing enzymes (Caldecott, 2008). Blocking PARP activity results in SSBs, which are eventually converted into DSBs during DNA replication. HR-proficient cells can repair DSBs in an error-free manner whereas HR-deficient cells cannot, and die (Polyak and Garber, 2011). As a result, PARP inhibition induces synthetic lethality in BRCA1/2-related cancers.

(6)

179 Study PARP inhibitor resistance using the K14Cre; Brca1F/F; p53F/F (KB1P) mouse model

In our lab, we want to understand why not all cancer patients with BRCA1-mutated tumors benefit from PARPi-based therapy. Due to broad genetic variation of individuals and the clinical limitations of research on human patients, mouse models of cancer provide a complementary approach to investigate mechanisms of drug resistance. The group of Jos Jonkers has shown that mammary tumors “spontaneously” arising in a genetically engineered mouse model (KB1P) highly mimic the human BRCA1-associated breast cancer. In this mouse model, the Cre recombinase is specifically expressed in the mammary gland due to the Keratin14 (K14) specific promoter. Cre recombinase induces recombination between two directly repeated loxP sites to excise the DNA sequences (Liu et al., 2007). As a result there is a stochastic deletion of the

Brca1 and p53 genes in several mammary epithelial cells. Eventually, this

results in the development of spontaneous mammary gland tumors after a mean latency of about 7 months. Treatment of tumor-bearing mice with olaparib often shrinks tumors to unpalpable sizes, however, tumors were not eradicated and eventually acquired olaparib resistance (Rottenberg et al., 2008).

Possible mechanisms of PARPi resistance in BRCA1-deficient mouse mammary tumors

(7)

180

BRCA remains to be explored. In the KB1P mouse model, however, PARPi resistance cannot be explained by the restoration of BRCA1 function, because of the large intrageneic deletion from exon 5 to exon 13 of both Brca1 alleles. This offers the opportunity to identify other mechanisms of resistance in this model.

Olaparib was identified as a substrate of ABCB1(MDR1) (Lawlor et al., 2014; Oplustilova et al., 2012; Rottenberg et al., 2008), and increased expression of the mouse Abcb1a/b (Mdr1a/b) genes contributed to olaparib resistance in

Brca1-/-;p53-/- mouse mammary tumors (Jaspers et al., 2013; Rottenberg et al., 2008). Similar findings were also described in mouse BRCA2-deficient mammary tumors which acquired olaparib resistance (Hay et al., 2009). Nonetheless, upregulation of P-gp or alteration of other drug transporters has not yet been confirmed to correlate with drug resistance in cancer patients (Borst, 2012). To circumvent MDR1-mediated resistance in our mouse tumors, the Mdr1 null alleles have been introduced into the KB1P model to generate

K14cre; Brca1F/F; p53 F/F; Mdr1-/- mice (KB1PM). Although loss of MDR1 function increased the overall survival of mice bearing Brca1-/-; p53-/-; Mdr1

(8)

181

can help to rescue doomed cells. Whereas little is known about the RIF1 or PTIP status in those cancer samples, loss of 53BP1 appears to be relatively common in TN and BRCA1-mutant breast cancer samples (Bouwman et al., 2010). In particular, the groups of Jos Jonkers and Andre Nussenzweig found that loss of 53BP1 enhanced end resection at DSBs and restored BRCA1-independent HR, resulting in olaparib resistance in BRCA1-deficient mouse mammary tumors. 53bp1 is also mutated or downregulated in some of the olaparib-resistant BRCA1-deficient mouse mammary tumors, however, about 3/4 of the olaparib-resistant tumors appear to have functional 53BP1 and other mechanisms of resistance must explain the resistance (Jaspers et al., 2013).

Loss of functional screens identify Rev7 and Helb to cause PARPi resistance when inhibited

(9)

182

transduced cells. This suggests that Rev7 or Helb, in addition to 53bp1, may predict PARPi response in BRCA1-deficient mouse mammary tumors.

REV7 (also known as MAD2 mitotic arrest deficient-like 2, MAD2L2) is a chromatin binding protein involved in cellular processes such as DNA damage repair and cell cycle control. It directly binds the DNA polymerase zeta catalytic subunit REV3L together with the inserter polymerase REV1 to form the Pol ζ-REV1 complex, which is crucial for translesion DNA synthesis (TLS) (Kikuchi et al., 2012; Tomida et al., 2015). In addition, it was found to inhibit the anaphase promoting complex/cyclosome (Teichner et al., 2011) during the metaphase-to-anaphase transition, thereby delaying metaphase-to-anaphase until all chromosomes are properly aligned on the metaphase plate (Chen and Fang, 2001; Listovsky and Sale, 2013; Pfleger et al., 2001). Moreover, REV7 has been reported to be required for anaphase-promoting complex-dependent ubiquitylation and degradation of REV1 (Chun et al., 2013). As an essential result of my PhD project, we discovered a link between REV7 and HR (Xu et al., 2015). In particular, we found that loss of REV7 rescues CTIP-dependent end resection of DSBs in BRCA1-deficient cells, leading to HR partial restoration and PARP inhibitor resistance. REV7 is recruited to DSBs in a manner dependent on the H2AX–MDC1–RNF8–RNF168–53BP1 pathway, and appears to inhibit HR and promote NHEJ by counteracting end resection at DSBs. Our findings reveal a crucial function of REV7 in coordinating DSB repair pathway choices in BRCA1-deficient cells. Our data are consistent with those of Dr. Jacobs’ group at the NKI who found that REV7 accumulates at uncapped telomeres and promotes NHEJ-mediated fusion of deprotected chromosome ends and genomic instability (Boersma et al., 2015).

(10)

183

1995; Seki et al., 1995). In addition, HELB has been shown to accumulate on chromatin in an RPA-dependent manner when the cell is exposed to replication stress and it may counteract replication stress (Guler et al., 2012). The subcellular localization of HELB is cell cycle dependent (Gu et al., 2004; Spencer et al., 2013). During G1, HELB is predominantly nuclear whereas at the G1/S transition, it is phosphorylated by CDK2, which results in its nuclear export, leaving only a small subpopulation of HELB in the nucleus (Guler et al., 2012). However, the link between HELB and HR had not yet investigated before our study, although there are other DNA helicases that have been found to be involved in HR. For example, RTEL1 or RECQL5 inhibit HR and maintain genomic stability via disruption of preformed displacement loops (Barber et al., 2008; Hu et al., 2007). PARI, also containing an UvrD-like helicase domain, inhibits HR by disrupting RAD51-ssDNA HR structures in human and DT40 chicken cells (Moldovan et al., 2012). In my PhD project, we report that HELB is recruited to resected DNA ends by interacting with RPA. Importantly, HELB counteracts EXO1- and BLM-DNA2-dependent end resection in both human and mouse cells. Consistent with its novel role as a resection antagonist, loss of HELB partially restores HR and results in PARPi resistance in BRCA1-deficient tumor cells. Intriguingly, HELB is exported from the nucleus in a CDK2-dependent manner near the G1/S transition, indicating a cell cycle regulated feedback inhibition mechanism on the process of end resection mediated by HELB.

Interactions of REV7 with other recently identified players involved in the DNA damage response

(11)

184

when both Rev7 and 53bp1 were depleted (see addendum in Chapter 2). Despite this strong evidence for a downstream role of REV7, we did not observe a direct interaction between REV7 and 53BP1.

RIF1 and PTIP are two downstream effectors of 53BP1 in response to DSBs. They are recruited by phosphorylated 53BP1 to promote NHEJ and block HR. PTIP loss rescues HR in BRCA1-deficient cells and is largely dispensable for NHEJ during CSR, whereas RIF1 loss only partially rescues HR and hypersensitivity caused by PARPi treatment of BRCA1-deficient cells. We failed to detect foci of endogenous PTIP with the available antibodies. However, the Jacobs group did not find an effect on foci formation by exogenous PTIP upon

Rev7 depletion (Boersma et al., 2015). In contrast to REV7 loss, PTIP loss does

not affect class switch recombination (Callen et al., 2013). It is therefore also not plausible that REV7 acts downstream of PTIP, consistent with the findings that foci formation by exogenous REV7 was not affected by Ptip depletion (Boersma et al., 2015). We found that REV7 deficiency does not affect the foci formation by endogenous RIF1 (which requires its interaction with 53BP1), suggesting that REV7 deficiency does not affect the interaction of 53BP1 with RIF1. However, the Jacobs group observed that REV7 recruitment to the DNA damage sites is RIF1-dependent as REV7 foci formation was abolished in RIF1-/- MEF cells, suggesting that REV7 may act downstream of 53BP1 and RIF1 (Boersma et al., 2015). Regarding the recognition of DSBs we have shown that HELB is recruited to end resected ssDNA by directly interacting with RPA. However, it is not clear how REV7 recognizes DSBs in a 53BP1-dependent manner, as we failed to see a direct interaction between 53BP1 and REV7 or between REV7 and the 53BP1-RIF1 complex. REV7 recruitment to DNA damage sites by 53BP1 and RIF1 may result from an indirect interaction or an activity elicited by chromatin-bound 53BP1 protein complexes at DSB sites. These mechanisms remain to be explored.

(12)

p53-185

deficient mES cells (unpublished data from Peter Bouwman and Ewa Gogola). Using clonogenic assays, we found that REV7 or HELB loss does not cause PARPi resistance in Brca2-deficient cell lines (chapter 2 and our unpublished data), showing that HR-deficiency due to BRCA2-deficiency cannot be restored by REV7 or HELB inhibition. This result is consistent with our finding that REV7 and HELB act more upstream in the HR pathway and contribute to protecting DNA double strand breaks from end resection.

Interactions of REV7 with REV3 and REV1 do not seem to be absolutely required for its function of counteracting end-resection

REV7 contains 211 amino acids and only has one domain. To study whether end resection by REV7 requires its interaction with REV1 and REV3, we cloned two REV7 mutants C70R (REV3-interacting deficient mutant of REV7) and L186A/Q200A/Y202A (REV1-interacting deficient mutant of REV7). As described in Chapter 2, similar to wild- type REV7 (wt REV7), both mutants were still recruited to the DNA damage sites. Reconstitution of either of them into the KB1P-B11-shRev7 and KB1P-G3-shRev7 cells partially reversed the resistance to PARPi compared with wt REV7 and empty vectors, suggesting that REV1 or REV3 interaction is not essential for REV7 to inhibit end resection at DSBs.

(13)

186

we have used two different chemical inhibitors of the proteasome: MG132 and bortezomib, and examined their impact on the recruitment of REV7 to laser-generated DNA damage sites in human cells. These experiments clearly established that the recruitment of REV7 to damage sites is sensitive to proteasome inhibition, as such treatment very efficiently prevented REV7 from localizing to laser generated micro-irradiation lesions (see addendum in Chapter 2).

To investigate the function of the 53BP1-REV7 pathway in HR-proficient cells, we have shown that REV7 loss has no effect on PARPi sensitivity in p53 -/-mammary tumor cells. However, its loss inhibits NHEJ thereby inhibiting class switch recombination (CSR) in the murine CH12 B cells, which undergoes CSR from IgM to IgA at a high rate upon stimulation (Muramatsu et al., 2000). Consistently, The Jacobs group found that REV7 accumulates at uncapped telomeres and promotes NHEJ-mediated fusion of deprotected chromosome ends and genomic instability (Boersma et al., 2015). As a result, NHEJ deficiency in Rev7 depleted cells may exhibit hallmarks of defective DSB repair such as IR sensitivity. Indeed, REV7 loss sensitizes p53-/- tumor cells to γ-irradiation (our unpublished data). Interestingly, REV3-depleted cells are also more sensitive to γ-irradiation than empty vector transduced p53-/- tumor cells (our unpublished data) which is consistent with a previous report (Sharma et al., 2012). Future experiments need to address whether this is independent of each other.

Analysis of Rev7 or Helb loss in PARPi resistant Brca1-/-;p53-/- mouse mammary tumors

(14)

187

Differential Signal (DIDS) algorithm (de Ronde et al., 2013; Rottenberg et al., 2012). Using this tool, Rev7 was among the top outliers of genes that were down regulated in resistant tumors (number 23; p = 0.0008) (see addendum in Chapter 2). We found more than 50% inhibition of the Rev7 transcript levels in 6 out of 55 PARPi resistant tumors based on the RNASeq analysis, which was confirmed by Western blotting to quantitate REV7 protein levels (our unpublished data). Together, these data strongly indicate that inhibition of Rev7 confers PARPi resistance in BRCA1-deficient tumor cells. Future studies need to investigate whether loss of Helb may also explain PARPi resistance in a set of spontaneous tumors which acquired olaparib or AZD2461 resistance in vivo.

Cross-resistance of PARPi resistant tumors with a low REV7 expression

We have previously shown that 53BP1 deletion reduces the sensitivity of BRCA1-deficient cells to DNA crosslinking agent cisplatin (Bouwman et al., 2010) and leads to improved overall survival of mice harboring tumors treated with cisplatin (Jaspers et al., 2013). We hypothesized that this results from the restoration of HR. However, the group of André Nussenzweig at the NCI reported a novel role of BRCA1 in interstrand crosslink repair (Bunting et al., 2012). This function was claimed to be independent of HR, because 53BP1 deletion does not affect the sensitivity of Brca1Δ11/Δ11 cells to cisplatin. This difference with our observations may be explained by the fact that the Nussenzweig group used a different mouse model. In the mammary tumors with the exon 11 deletion of Brca1 expression of a hypomorphic BRCA1 protein is maintained. Thus, HR may not be completely abrogated in their model. In contrast, in our KB1P model there is a complete lack of BRCA1 protein expression.

(15)

188

were treated with cisplatin (see addendum in Chapter 2). These data are consistent with our model that REV7 is in the same pathway as 53BP1.

How to overcome the HELB or REV7 loss-mediated PARPi resistance in BRCA1-deficient tumors?

Our findings suggest that BRCA1-related breast cancer patients who have lost functional HELB or REV7 in their tumor cells, may not benefit from PARPi treatment. Hence, drug-sensitizing screens with or without PARPi combination using Rev7 or Helb; Brca1; p53-deficient tumor cells may provide novel information. In Chapter 2 I showed that REV7 loss-mediated PARPi resistance is reversed by ATM inhibition by KU55933 treatment in vitro. I obtained similar findings in Helb loss cells using ATM inhibition by KU55933 or KU60019 (our unpublished data). Given that the pharmacokinetic properties of KU55933 are not sufficient to inhibit ATM in vivo, we did not make an attempt to test this treatment in our mouse model. However, the DNA damage in normal tissues resulting from treatment with a potentially efficient ATM inhibitor may be greatly exacerbated when combined with PARPi; any future combinations of PARPi and ATMi should therefore target solely the tumors to avoid excessive toxicity to normal tissues.

(16)

189 Outlook

An important question is whether our findings on new mechanims of PARPi resistance can be translated into the clinic. This question should be ideally addressed using tumors from BRCA1 mutation carriers that initially responded to PARPi and eventually acquired resistance. Unfortunately, we are not aware of a clinical trial that has generated matched samples of BRCA1-mutated breast cancer patients who initially responded to PARP inhibitors and eventually acquired resistance in order to prove our findings. Since we observe that the PARPi-resistant tumors are still sensitive to platinum drugs we do not think that platinum salts are a suitable substitution. In fact, none of the components of this pathway with functions related to DSB resection/PARPi sensitivity, including

53BP1, Rev7, RIF1 or PTIP, has so far been reported to cause resistance in

patients treated with PARP inhibitors. However, it has also not been carefully analyzed in suitable matched samples. In the future it will be important to get matched samples from the new clinical trials that have been started.

Another important tool that should be implemented in the clinic is a functional assay to test HR. Determining REV7 expression on its own may not be sufficient. In our mouse model, Ewa Gogola established an in situ assay to quantify RAD51 foci formation as readout for HR. Since this requires a single high dose of radiotherapy, it may not be feasible in patients, however. Another promising approach is the use of an ex vivo RAD51 foci formation assay (Naipal et al., 2014) to measure the HR status. This may help clinicians to give personalized therapy when treating Brca1-related cancer patients with mutant or predictive biomarkers. In addition, it may provide clinicians with more options to find the actual drug resistant mechanisms and execute more effective strategies to overcome drug resistance.

(17)

190

specimens from the NKI biobank, available in the form of tissue microarrays of breast cancers (random subtypes). We investigated REV7 and 53BP1 expression of 366 patients by immunohistochemistry. In the larger data set we found that loss of REV7 expression is more frequent in triple negative breast cancer (TNBC), in which most of BRCA1-related breast cancer is found. The staining distribution is shown in Chapter 2 and the Wilcoxon rank sum test with continuity correction gives a P value of 0.009 and we also consistently observed the trend for mutually exclusive defects in 53BP1 and REV7 expression in both cohorts (see addendum in Chapter 2 and our unpublished data). In addition, although the tumor sample is small, we have found that in the NKI cohort of human breast cancer patients there were 9 BRCA1-mutation carriers and in 4 of these, REV7 expression was aberrantly lost and all of these still express 53BP1 (our unpublished data). This indicates that such a dual defect (BRCA1/REV7) also occurs in human tumors in the clinic, and suggests that this happens with a significant frequency. Fortunately, many PARPi clinical trials now mandate biopsies once resistance occurs and this should allow us further studies in the future. Finally, we have tried to analyze HELB expression by IHC, but both the commercially available and our lab-made antibodies seem to be unsuitable for this demanding application, either due to unacceptable background staining, or not being applicable on routinely processed formalin-fixed paraffin sections deparaffinized through xylene exposure, a standard procedure that destroys some antibody-recognized epitopes.

(18)

191

novel factors involved in PARPi resistance. Since targeting mRNAs with shRNAs results only in gene knockdown, CAS9-sgRNA-mediated gene knockout may help to find hits that promote PARPi resistance when the gene is completely inactivated.. Thus far, I have only employed hairpins that target the DDR. To identify the remaining mechanisms, future studies should include a whole-genome shRNA library or at least other shRNAs pools such as those targeting chromatin-remodeling genes. Because over-expression of a certain gene may also cause PARPi resistance, gain of function screens using cDNA libraries should also be employed.

References

Antoniou, A., Pharoah, P.D., Narod, S., Risch, H.A., Eyfjord, J.E., Hopper, J.L., Loman, N., Olsson, H., Johannsson, O., Borg, A., et al. (2003). Average risks of breast and ovarian cancer associated with BRCA1 or BRCA2 mutations detected in case Series unselected for family history: a combined analysis of 22 studies. Am J Hum Genet 72, 1117-1130.

Barber, L.J., Sandhu, S., Chen, L., Campbell, J., Kozarewa, I., Fenwick, K., Assiotis, I., Rodrigues, D.N., Reis Filho, J.S., Moreno, V., et al. (2013). Secondary mutations in BRCA2 associated with clinical resistance to a PARP inhibitor. The Journal of pathology 229, 422-429. Barber, L.J., Youds, J.L., Ward, J.D., McIlwraith, M.J., O'Neil, N.J., Petalcorin, M.I., Martin, J.S., Collis, S.J., Cantor, S.B., Auclair, M., et al. (2008). RTEL1 maintains genomic stability by suppressing homologous recombination. Cell 135, 261-271.

Boersma, V., Moatti, N., Segura-Bayona, S., Peuscher, M.H., van der Torre, J., Wevers, B.A., Orthwein, A., Durocher, D., and Jacobs, J.J. (2015). MAD2L2 controls DNA repair at telomeres and DNA breaks by inhibiting 5' end resection. Nature.

Borst, P. (2012). Cancer drug pan-resistance: pumps, cancer stem cells, quiescence, epithelial to mesenchymal transition, blocked cell death pathways, persisters or what? Open biology 2, 120066.

Bouwman, P., Aly, A., Escandell, J.M., Pieterse, M., Bartkova, J., van der Gulden, H., Hiddingh, S., Thanasoula, M., Kulkarni, A., Yang, Q., et al. (2010). 53BP1 loss rescues BRCA1 deficiency and is associated with triple-negative and BRCA-mutated breast cancers. Nat Struct Mol Biol 17, 688-695.

(19)

192

Bryant, H.E., Schultz, N., Thomas, H.D., Parker, K.M., Flower, D., Lopez, E., Kyle, S., Meuth, M., Curtin, N.J., and Helleday, T. (2005). Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase. Nature 434, 913-917.

Bunting, S.F., Callen, E., Kozak, M.L., Kim, J.M., Wong, N., Lopez-Contreras, A.J., Ludwig, T., Baer, R., Faryabi, R.B., Malhowski, A., et al. (2012). BRCA1 functions independently of homologous recombination in DNA interstrand crosslink repair. Mol Cell 46, 125-135.

Bunting, S.F., Callen, E., Wong, N., Chen, H.T., Polato, F., Gunn, A., Bothmer, A., Feldhahn, N., Fernandez-Capetillo, O., Cao, L., et al. (2010). 53BP1 inhibits homologous recombination in Brca1-deficient cells by blocking resection of DNA breaks. Cell 141, 243-254.

Caldecott, K.W. (2008). Single-strand break repair and genetic disease. Nature reviews Genetics 9, 619-631.

Callen, E., Di Virgilio, M., Kruhlak, M.J., Nieto-Soler, M., Wong, N., Chen, H.T., Faryabi, R.B., Polato, F., Santos, M., Starnes, L.M., et al. (2013). 53BP1 mediates productive and mutagenic DNA repair through distinct phosphoprotein interactions. Cell 153, 1266-1280.

Chapman, J.R., Barral, P., Vannier, J.B., Borel, V., Steger, M., Tomas-Loba, A., Sartori, A.A., Adams, I.R., Batista, F.D., and Boulton, S.J. (2013). RIF1 is essential for 53BP1-dependent nonhomologous end joining and suppression of DNA double-strand break resection. Mol Cell 49, 858-871.

Chen, J., and Fang, G. (2001). MAD2B is an inhibitor of the anaphase-promoting complex. Genes Dev 15, 1765-1770.

Chun, A.C., Kok, K.H., and Jin, D.Y. (2013). REV7 is required for anaphase-promoting complex-dependent ubiquitination and degradation of translesion DNA polymerase REV1. Cell Cycle 12, 365-378.

Davis, P.L., Miron, A., Andersen, L.M., Iglehart, J.D., and Marks, J.R. (1999). Isolation and initial characterization of the BRCA2 promoter. Oncogene 18, 6000-6012.

de Ronde, J.J., Rigaill, G., Rottenberg, S., Rodenhuis, S., and Wessels, L.F. (2013). Identifying subgroup markers in heterogeneous populations. Nucleic Acids Res 41, e200.

Di Virgilio, M., Callen, E., Yamane, A., Zhang, W., Jankovic, M., Gitlin, A.D., Feldhahn, N., Resch, W., Oliveira, T.Y., Chait, B.T., et al. (2013). Rif1 prevents resection of DNA breaks and promotes immunoglobulin class switching. Science 339, 711-715.

(20)

193 Edwards, S.L., Brough, R., Lord, C.J., Natrajan, R., Vatcheva, R., Levine, D.A., Boyd, J., Reis-Filho, J.S., and Ashworth, A. (2008). Resistance to therapy caused by intragenic deletion in BRCA2. Nature 451, 1111-1115.

Escribano-Diaz, C., Orthwein, A., Fradet-Turcotte, A., Xing, M., Young, J.T., Tkac, J., Cook, M.A., Rosebrock, A.P., Munro, M., Canny, M.D., et al. (2013). A cell cycle-dependent regulatory circuit composed of 53BP1-RIF1 and BRCA1-CtIP controls DNA repair pathway choice. Mol Cell

49, 872-883.

Esteller, M., Silva, J.M., Dominguez, G., Bonilla, F., Matias-Guiu, X., Lerma, E., Bussaglia, E., Prat, J., Harkes, I.C., Repasky, E.A., et al. (2000). Promoter hypermethylation and BRCA1 inactivation in sporadic breast and ovarian tumors. Journal of the National Cancer Institute 92, 564-569.

Farmer, H., McCabe, N., Lord, C.J., Tutt, A.N., Johnson, D.A., Richardson, T.B., Santarosa, M., Dillon, K.J., Hickson, I., Knights, C., et al. (2005). Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 434, 917-921.

Feng, L., Fong, K.W., Wang, J., Wang, W., and Chen, J. (2013). RIF1 counteracts BRCA1-mediated end resection during DNA repair. J Biol Chem 288, 11135-11143.

Gu, J., Xia, X., Yan, P., Liu, H., Podust, V.N., Reynolds, A.B., and Fanning, E. (2004). Cell cycle-dependent regulation of a human DNA helicase that localizes in DNA damage foci. Mol Biol Cell

15, 3320-3332.

Guler, G.D., Liu, H., Vaithiyalingam, S., Arnett, D.R., Kremmer, E., Chazin, W.J., and Fanning, E. (2012). Human DNA helicase B (HDHB) binds to replication protein A and facilitates cellular recovery from replication stress. J Biol Chem 287, 6469-6481.

Hay, T., Matthews, J.R., Pietzka, L., Lau, A., Cranston, A., Nygren, A.O., Douglas-Jones, A., Smith, G.C., Martin, N.M., O'Connor, M., et al. (2009). Poly(ADP-ribose) polymerase-1 inhibitor treatment regresses autochthonous Brca2/p53-mutant mammary tumors in vivo and delays tumor relapse in combination with carboplatin. Cancer research 69, 3850-3855.

Holstege, H., Joosse, S.A., van Oostrom, C.T., Nederlof, P.M., de Vries, A., and Jonkers, J. (2009). High incidence of protein-truncating TP53 mutations in BRCA1-related breast cancer. Cancer Res 69, 3625-3633.

(21)

194

Ibrahim, Y.H., Garcia-Garcia, C., Serra, V., He, L., Torres-Lockhart, K., Prat, A., Anton, P., Cozar, P., Guzman, M., Grueso, J., et al. (2012). PI3K inhibition impairs BRCA1/2 expression and sensitizes BRCA-proficient triple-negative breast cancer to PARP inhibition. Cancer discovery 2, 1036-1047.

Jackson, S.P., and Bartek, J. (2009). The DNA-damage response in human biology and disease. Nature 461, 1071-1078.

Jaspers, J.E., Kersbergen, A., Boon, U., Sol, W., van Deemter, L., Zander, S.A., Drost, R., Wientjens, E., Ji, J., Aly, A., et al. (2013). Loss of 53BP1 causes PARP inhibitor resistance in Brca1-mutated mouse mammary tumors. Cancer Discov 3, 68-81.

Juvekar, A., Burga, L.N., Hu, H., Lunsford, E.P., Ibrahim, Y.H., Balmana, J., Rajendran, A., Papa, A., Spencer, K., Lyssiotis, C.A., et al. (2012). Combining a PI3K inhibitor with a PARP inhibitor provides an effective therapy for BRCA1-related breast cancer. Cancer Discov 2, 1048-1063.

Kikuchi, S., Hara, K., Shimizu, T., Sato, M., and Hashimoto, H. (2012). Structural basis of recruitment of DNA polymerase zeta by interaction between REV1 and REV7 proteins. J Biol Chem 287, 33847-33852.

Lawlor, D., Martin, P., Busschots, S., Thery, J., O'Leary, J.J., Hennessy, B.T., and Stordal, B. (2014). PARP Inhibitors as P-glyoprotein Substrates. Journal of pharmaceutical sciences 103, 1913-1920.

Li, X., Stith, C.M., Burgers, P.M., and Heyer, W.D. (2009). PCNA is required for initiation of recombination-associated DNA synthesis by DNA polymerase delta. Mol Cell 36, 704-713. Lichtenstein, P., Holm, N.V., Verkasalo, P.K., Iliadou, A., Kaprio, J., Koskenvuo, M., Pukkala, E., Skytthe, A., and Hemminki, K. (2000). Environmental and heritable factors in the causation of cancer--analyses of cohorts of twins from Sweden, Denmark, and Finland. N Engl J Med 343, 78-85.

Listovsky, T., and Sale, J.E. (2013). Sequestration of CDH1 by MAD2L2 prevents premature APC/C activation prior to anaphase onset. J Cell Biol 203, 87-100.

Liu, X., Holstege, H., van der Gulden, H., Treur-Mulder, M., Zevenhoven, J., Velds, A., Kerkhoven, R.M., van Vliet, M.H., Wessels, L.F., Peterse, J.L., et al. (2007). Somatic loss of BRCA1 and p53 in mice induces mammary tumors with features of human BRCA1-mutated basal-like breast cancer. Proc Natl Acad Sci U S A 104, 12111-12116.

(22)

195 McIlwraith, M.J., Vaisman, A., Liu, Y., Fanning, E., Woodgate, R., and West, S.C. (2005). Human DNA polymerase eta promotes DNA synthesis from strand invasion intermediates of homologous recombination. Mol Cell 20, 783-792.

Moldovan, G.L., Dejsuphong, D., Petalcorin, M.I., Hofmann, K., Takeda, S., Boulton, S.J., and D'Andrea, A.D. (2012). Inhibition of homologous recombination by the PCNA-interacting protein PARI. Mol Cell 45, 75-86.

Moldovan, G.L., Madhavan, M.V., Mirchandani, K.D., McCaffrey, R.M., Vinciguerra, P., and D'Andrea, A.D. (2010). DNA polymerase POLN participates in cross-link repair and homologous recombination. Mol Cell Biol 30, 1088-1096.

Moro, L., Arbini, A.A., Marra, E., and Greco, M. (2006). Up-regulation of Skp2 after prostate cancer cell adhesion to basement membranes results in BRCA2 degradation and cell proliferation. The Journal of biological chemistry 281, 22100-22107.

Moskwa, P., Buffa, F.M., Pan, Y., Panchakshari, R., Gottipati, P., Muschel, R.J., Beech, J., Kulshrestha, R., Abdelmohsen, K., Weinstock, D.M., et al. (2011). miR-182-mediated

downregulation of BRCA1 impacts DNA repair and sensitivity to PARP inhibitors. Molecular cell

41, 210-220.

Muramatsu, M., Kinoshita, K., Fagarasan, S., Yamada, S., Shinkai, Y., and Honjo, T. (2000). Class switch recombination and hypermutation require activation-induced cytidine deaminase (AID), a potential RNA editing enzyme. Cell 102, 553-563.

Naipal, K.A., Verkaik, N.S., Ameziane, N., van Deurzen, C.H., Ter Brugge, P., Meijers, M., Sieuwerts, A.M., Martens, J.W., O'Connor, M.J., Vrieling, H., et al. (2014). Functional ex vivo assay to select homologous recombination-deficient breast tumors for PARP inhibitor treatment. Clin Cancer Res 20, 4816-4826.

Norquist, B., Wurz, K.A., Pennil, C.C., Garcia, R., Gross, J., Sakai, W., Karlan, B.Y., Taniguchi, T., and Swisher, E.M. (2011). Secondary somatic mutations restoring BRCA1/2 predict

chemotherapy resistance in hereditary ovarian carcinomas. Journal of clinical oncology : official journal of the American Society of Clinical Oncology 29, 3008-3015.

Oplustilova, L., Wolanin, K., Mistrik, M., Korinkova, G., Simkova, D., Bouchal, J., Lenobel, R., Bartkova, J., Lau, A., O'Connor, M.J., et al. (2012). Evaluation of candidate biomarkers to predict cancer cell sensitivity or resistance to PARP-1 inhibitor treatment. Cell Cycle 11, 3837-3850. Panier, S., and Durocher, D. (2013). Push back to respond better: regulatory inhibition of the DNA double-strand break response. Nat Rev Mol Cell Biol 14, 661-672.

(23)

196

mutations in BRCA1 with strong founder effects among Dutch and Belgian hereditary breast and ovarian cancer families. Am J Hum Genet 60, 1041-1049.

Petrij-Bosch, A., Peelen, T., van Vliet, M., van Eijk, R., Olmer, R., Drusedau, M., Hogervorst, F.B., Hageman, S., Arts, P.J., Ligtenberg, M.J., et al. (1997). BRCA1 genomic deletions are major founder mutations in Dutch breast cancer patients. Nat Genet 17, 341-345.

Pfleger, C.M., Salic, A., Lee, E., and Kirschner, M.W. (2001). Inhibition of Cdh1-APC by the MAD2-related protein MAD2L2: a novel mechanism for regulating Cdh1. Genes Dev 15, 1759-1764.

Polyak, K., and Garber, J. (2011). Targeting the missing links for cancer therapy. Nat Med 17, 283-284.

Rice, J.C., Ozcelik, H., Maxeiner, P., Andrulis, I., and Futscher, B.W. (2000). Methylation of the BRCA1 promoter is associated with decreased BRCA1 mRNA levels in clinical breast cancer specimens. Carcinogenesis 21, 1761-1765.

Rottenberg, S., Jaspers, J.E., Kersbergen, A., van der Burg, E., Nygren, A.O., Zander, S.A., Derksen, P.W., de Bruin, M., Zevenhoven, J., Lau, A., et al. (2008). High sensitivity of BRCA1-deficient mammary tumors to the PARP inhibitor AZD2281 alone and in combination with platinum drugs. Proc Natl Acad Sci U S A 105, 17079-17084.

Rottenberg, S., Vollebergh, M.A., de Hoon, B., de Ronde, J., Schouten, P.C., Kersbergen, A., Zander, S.A., Pajic, M., Jaspers, J.E., Jonkers, M., et al. (2012). Impact of intertumoral heterogeneity on predicting chemotherapy response of BRCA1-deficient mammary tumors. Cancer Res 72, 2350-2361.

Saitoh, A., Tada, S., Katada, T., and Enomoto, T. (1995). Stimulation of mouse DNA primase-catalyzed oligoribonucleotide synthesis by mouse DNA helicase B. Nucleic Acids Res 23, 2014-2018.

Sakai, W., Swisher, E.M., Karlan, B.Y., Agarwal, M.K., Higgins, J., Friedman, C., Villegas, E., Jacquemont, C., Farrugia, D.J., Couch, F.J., et al. (2008). Secondary mutations as a mechanism of cisplatin resistance in BRCA2-mutated cancers. Nature 451, 1116-1120.

San Filippo, J., Sung, P., and Klein, H. (2008). Mechanism of eukaryotic homologous recombination. Annu Rev Biochem 77, 229-257.

Schoenfeld, A.R., Apgar, S., Dolios, G., Wang, R., and Aaronson, S.A. (2004). BRCA2 is ubiquitinated in vivo and interacts with USP11, a deubiquitinating enzyme that exhibits

(24)

197 Seki, M., Enomoto, T., Yanagisawa, J., Hanaoka, F., and Ui, M. (1988). Further characterization of DNA helicase activity of mouse DNA-dependent adenosinetriphosphatase B (DNA helicase B). Biochemistry 27, 1766-1771.

Seki, M., Kohda, T., Yano, T., Tada, S., Yanagisawa, J., Eki, T., Ui, M., and Enomoto, T. (1995). Characterization of DNA synthesis and DNA-dependent ATPase activity at a restrictive

temperature in temperature-sensitive tsFT848 cells with thermolabile DNA helicase B. Mol Cell Biol 15, 165-172.

Sharma, S., Hicks, J.K., Chute, C.L., Brennan, J.R., Ahn, J.Y., Glover, T.W., and Canman, C.E. (2012). REV1 and polymerase zeta facilitate homologous recombination repair. Nucleic Acids Res 40, 682-691.

Song, L., Dai, T., Xie, Y., Wang, C., Lin, C., Wu, Z., Ying, Z., Wu, J., Li, M., and Li, J. (2012). Up-regulation of miR-1245 by c-myc targets BRCA2 and impairs DNA repair. Journal of molecular cell biology 4, 108-117.

Spencer, S.L., Cappell, S.D., Tsai, F.C., Overton, K.W., Wang, C.L., and Meyer, T. (2013). The proliferation-quiescence decision is controlled by a bifurcation in CDK2 activity at mitotic exit. Cell 155, 369-383.

Sun, C., Li, N., Yang, Z., Zhou, B., He, Y., Weng, D., Fang, Y., Wu, P., Chen, P., Yang, X., et al. (2013). miR-9 regulation of BRCA1 and ovarian cancer sensitivity to cisplatin and PARP

inhibition. Journal of the National Cancer Institute 105, 1750-1758.

Swisher, E.M., Sakai, W., Karlan, B.Y., Wurz, K., Urban, N., and Taniguchi, T. (2008).

Secondary BRCA1 mutations in BRCA1-mutated ovarian carcinomas with platinum resistance. Cancer research 68, 2581-2586.

Teichner, A., Eytan, E., Sitry-Shevah, D., Miniowitz-Shemtov, S., Dumin, E., Gromis, J., and Hershko, A. (2011). p31comet Promotes disassembly of the mitotic checkpoint complex in an ATP-dependent process. Proc Natl Acad Sci U S A 108, 3187-3192.

Tomida, J., Takata, K., Lange, S.S., Schibler, A.C., Yousefzadeh, M.J., Bhetawal, S., Dent, S.Y., and Wood, R.D. (2015). REV7 is essential for DNA damage tolerance via two REV3L binding sites in mammalian DNA polymerase zeta. Nucleic Acids Res 43, 1000-1011.

Tripathi, M.K., Misra, S., Khedkar, S.V., Hamilton, N., Irvin-Wilson, C., Sharan, C., Sealy, L., and Chaudhuri, G. (2005). Regulation of BRCA2 gene expression by the SLUG repressor protein in human breast cells. The Journal of biological chemistry 280, 17163-17171.

(25)

198

olaparib in patients with BRCA1 or BRCA2 mutations and advanced breast cancer: a proof-of-concept trial. Lancet 376, 235-244.

Verhoog, L.C., van den Ouweland, A.M., Berns, E., van Veghel-Plandsoen, M.M., van Staveren, I.L., Wagner, A., Bartels, C.C., Tilanus-Linthorst, M.M., Devilee, P., Seynaeve, C., et al. (2001). Large regional differences in the frequency of distinct BRCA1/BRCA2 mutations in 517 Dutch breast and/or ovarian cancer families. Eur J Cancer 37, 2082-2090.

Wu, K., Jiang, S.W., Thangaraju, M., Wu, G., and Couch, F.J. (2000). Induction of the BRCA2 promoter by nuclear factor-kappa B. The Journal of biological chemistry 275, 35548-35556. Xu, G., Chapman, J.R., Brandsma, I., Yuan, J., Mistrik, M., Bouwman, P., Bartkova, J., Gogola, E., Warmerdam, D., Barazas, M., et al. (2015). REV7 counteracts DNA double-strand break resection and affects PARP inhibition. Nature.

Referenties

GERELATEERDE DOCUMENTEN

On typical company law matters – for example employee parti- cipation, rights of inquiry, piercing the corporate veil and con- flicts of interests – the Supreme Court is

Over the years, a wide range of SLAM algorithms have been developed, ranging from filter based to graph-based approaches. Naturally, many of these treat the data they receive

In this chapter, we address the role of interest groups during the Australian national elections in 2016.We focus on the following themes: relationships between groups

To validate the loss of 53BP1 expression as a survival factor in Brca1-null cells, we tested the effects of shRNA-mediated depletion of 53BP1 in R26 CreERT2 Brca1 SCo/Δ ES

Stannous(II)octoate was selected as catalyst for the ring opening polymerization of lactide to the PVDF macroinitiator, since it is a widely used catalyst for

Next, Ito showed that for q odd the Zassenhaus group in question has to contain a normal subgroup isomorfic to PSL(2, q) with index 1 or 2.. To conclude, Suzuki dealt with the

Here we show that loss of REV7 (also known as MAD2L2) in mouse and human cell lines re-establishes CTIP-dependent end resection of DSBs in BRCA1-deficient cells,

Grounded in a standard modern rhetoric on rape and crimes of sexVirtual public and private ual assault against women which views space women as victims without any agency,