Translational research in radiation-induced DNA damage signaling and repair
Review Article

Translational research in radiation-induced DNA damage signaling and repair

Jac A. Nickoloff, Mary-Keara Boss, Christopher P. Allen, Susan M. LaRue

Department of Environmental and Radiological Health Sciences, Flint Animal Cancer Center, College of Veterinary Medicine and Biomedical Sciences, Colorado State University, Fort Collins, CO, USA

Contributions: (I) Conception and design: JA Nickoloff, SM LaRue; (II) Administrative support: None; (III) Provision of study materials or patients: None; (IV) Collection and assembly of data: None; (V) Data analysis and interpretation: All authors; (VI) Manuscript writing: All authors; (VII) Final approval of manuscript: All authors.

Correspondence to: Jac A. Nickoloff. Department of Environmental and Radiological Health Sciences, College of Veterinary Medicine and Biomedical Sciences, Colorado State University, 1618 Campus Delivery, Fort Collins, CO 80523, USA. Email: J.Nickoloff@colostate.edu.

Abstract: Radiotherapy is an effective tool in the fight against cancer. It is non-invasive and painless, and with advanced tumor imaging and beam control systems, radiation can be delivered to patients safely, generally with minor or no adverse side effects, accounting for its increasing use against a broad range of tumors. Tumors and normal cells respond to radiation-induced DNA damage by activating a complex network of DNA damage signaling and repair pathways that determine cell fate including survival, death, and genome stability. DNA damage response (DDR) proteins represent excellent targets to augment radiotherapy, and many agents that inhibit key response proteins are being combined with radiation and genotoxic chemotherapy in clinical trials. This review focuses on how insights into molecular mechanisms of DDR pathways are translated to small animal preclinical studies, to clinical studies of naturally occurring tumors in companion animals, and finally to human clinical trials. Companion animal studies, under the umbrella of comparative oncology, have played key roles in the development of clinical radiotherapy throughout its >100-year history. There is growing appreciation that rapid translation of basic knowledge of DNA damage and repair systems to improved radiotherapy practice requires a comprehensive approach that embraces the full spectrum of cancer research, with companion animal clinical trials representing a critical bridge between small animal preclinical studies, and human clinical trials.

Keywords: DNA repair; DNA damage checkpoint signaling; radiotherapy; comparative oncology


Submitted Mar 09, 2017. Accepted for publication May 25, 2017.

doi: 10.21037/tcr.2017.06.02


Introduction

Cells respond to DNA damage by activating checkpoint signaling and DNA repair pathways, collectively termed the DNA damage response (DDR), which promotes cell survival, and suppresses cancer by promoting genome stability and by triggering programmed cell death pathways. Components of the DDR are often defective in cancer, but because the DDR is a complex network of interacting/cross-talking pathways, a defect in one DDR component can be compensated by alternative pathways. The DDR is a major determinant of cancer cell responses to chemo- and radiotherapy, most of which cause DNA damage directly or indirectly, thus DDR components are enticing targets in the quest to augment cancer therapy (1-6).

Compensatory pathways within the DDR network also represent formidable obstacles to successful cancer treatment. By improving our understanding of DDR pathways, synthetic lethal relationships can be identified among different parts of the network that can be exploited to augment cancer therapy in general, and to develop personalized therapies based on knowledge of specific DDR defects in patient tumors (7-10). This information may also be used to predict how further changes in the DDR (comprising DDR component inactivation, down-regulation, or up-regulation) may compensate for prior defects and confer resistance to general or personalized therapeutics. Such predictions may permit oncologists to monitor tumor response to initial therapy and insert a new line of attack when resistance develops, or perhaps block the compensatory resistance pathway as part of the initial therapeutic strategy. This latter strategy may ultimately prove the most effective given recent insights that contrary to prior models that posited that cancer therapeutics induced mutations that confer resistance, the vast genetic heterogeneity characteristic of most solid tumors indicates that mutations that confer resistance to therapeutics are present subclonally prior to treatment (11). In this view, chemo- or radiotherapy simply selects for pre-existing resistant cancer cells, which can account for the significant rate of failed local tumor control in clinical settings. Translational research in ionizing radiation-induced DNA damage is a very active field, as the complexities of the DDR challenge researchers to define DDR defects in particular tumor types (and in specific patient tumors) and determine how to exploit these defects in personalized treatment to improve therapeutic outcomes. A particular focus is to develop rational strategies to inhibit redundant DDR pathways, so-called synthetic lethal approaches (6,9,10,12,13).


DDR pathways

The DDR includes five major DNA repair pathways (some with sub-pathways; Figure 1) that process different types of DNA lesions. DNA damage can arise spontaneously as a result of chemical lability of DNA, or is induced by reactive oxygen species produced by normal cellular metabolism, nucleases such as RAG1/2 (14) or exogenous genotoxins including most cancer chemotherapeutics and ionizing radiation. The DDR also includes two DNA damage checkpoint signaling pathways (Figure 2), one centered on ATM that responds to double-strand breaks (DSBs), and one centered on ATR that is triggered by single-stranded DNA (ssDNA) that forms when replication forks are blocked by DNA lesions (replication stress), and when broken ends of DSBs are resected (15,16). ATM and ATR are members of the PI3-kinase-like kinase (PIKK) family that also includes DNA-PK. Together, PIKKs are “early responders” to DSBs and replication stress caused by single-strand damage and intra- and inter-strand crosslinks. Once activated, PIKKs phosphorylate large networks of proteins (17-19) including the downstream effector kinases Chk1 and Chk2 that phosphorylate p53 and other targets to arrest the cell cycle in response to damage, promote DNA repair, and promote programmed cell death pathways when damage is too extensive (16,20-22). The DDR thus presents two general targets that can be manipulated for therapeutic gain: inhibiting DNA repair sensitizes cells to damage, and inhibiting checkpoint signaling prevents cell cycle arrest in response to damage, which increases replication stress (23-27). Great strides have been made in recent years to improve our understanding of DDR proteins and processes and this has been exploited to translate basic knowledge from the lab bench to preclinical models and in several notable cases to human clinical trials and clinical practice. Because the DDR is a complex network of cross-talking and redundant pathways, inhibiting a single pathway may have limited utility. For cancers with a specific defect, targeting a redundant pathway can cause synthetic lethality (9,10). This strategy is exemplified by the successful treatment of HR-defective (BRCA1/2-mutant) breast and ovarian cancers with PARP1 inhibitors (7,28). However, the highly networked DDR also provides means for tumor cells to develop resistance to targeted therapies, and strategies are being developed to identify and block resistance pathways (29-32).

Figure 1 Various types of DNA damage (boxes) are processed by five distinct repair pathways (red font). Interstrand crosslinks require multiple repair pathways.
Figure 2 DDR signaling pathways. Ionizing radiation creates DSBs and single-strand damage that triggers PIKK activation (blue), which activates downstream effector kinases (green). Black arrows indicate functional pathways, red arrows indicate target phosphorylation by protein kinases. DDR, DNA damage response; PIKK, PI3-kinase-like kinase.

Repair of IR-induced DNA damage

IR creates DNA DSBs and many types of single-strand damage (33). DSBs have gained the most attention due to their greater cytotoxicity and risk of triggering genome instability (22,34). In mammalian cells, DSBs are primarily repaired by classical non-homologous end joining (cNHEJ). cNHEJ is fast and error-prone, but it tends to produce small deletions and insertions at junctions so it prevents large-scale genome instability and suppresses cancer (35,36). DSBs are also repaired by homologous recombination (HR), which is slow and generally accurate, but it is largely limited to S/G2 phases (37). Alternative NHEJ (aNHEJ; also called microhomology-mediated end joining) can serve as a back-up to cNHEJ (38). aNHEJ is more error-prone than cNHEJ and more likely to create large-scale genome rearrangements such as translocations (38-40). Since cNHEJ is faster and more efficient than aNHEJ, the risk of genome destabilization by aNHEJ is low unless cNHEJ is defective (40-43). Single-strand damage (single-strand breaks, base damage, abasic sites) are repaired by a set of base excision repair (BER) pathways regulated by PARP1 (44), and by nucleotide excision repair (NER).

DSB repair pathway choice is primarily determined by end resection (37). cNHEJ operates on unresected ends, catalyzed by the MRE11/RAD50/NBS1 (MRN) complex which trims radiation-damaged bases at broken ends prior to joining by Ku70/Ku80, DNA-PKcs, Lig IV, and XRCC4 (45). Broken ends at IR-induced DSBs can have a variety of chemical structures, some of which require processing by the Artemis nuclease (46). End resection is initially catalyzed by CtIP; this limited 5' to 3' resection can reveal microhomologies that serve to align ends for aNHEJ (47). HR requires extensive 5' to 3' resection (hundreds to thousands of nt), catalyzed by Exo1-BLM or Dna2-BLM (48-50). The long 3' ssDNA tails are initially bound by RPA, then RPA is exchanged for RAD51, and the RAD51-ssDNA filament searches for and invades a homologous sequence elsewhere in the genome. HR typically uses the sister chromatid as a repair template, as this reduces the chance for HR-mediated genome rearrangements and large-scale loss of heterozygosity (6,51). Many proteins function with RAD51 in mediating HR, including BRCA1, BRCA2, PALB2, BCCIP, five RAD51 paralogs (XRCC2/3, RAD51B/C/D), RAD54/B, and a growing number of Fanconi anemia proteins involved in HR repair of certain types of lesions (e.g., interstrand crosslinks) (52).

IR creates DSBs by two distinct mechanisms. IR induces frank DSBs by direct energy absorption or more commonly through production of reactive oxygen species, and evidence of these DSBs (i.e., γ-H2AX foci) is apparent almost immediately after IR. IR also induces secondary DSBs several hours after IR—these appear several hours after irradiation, when replication forks encounter unrepaired single strand breaks or single-strand lesions, causing replication stress (53,54). Note that for every DSB produced by IR, approximately 30-fold more single-strand breaks are produced, along with numerous single-strand lesions such as broken rings, abasic sites, and interstrand crosslinks (55,56). Stalled forks are initially stabilized by checkpoint and repair factors including ATR, BLM, and BRCA2 (27,57-60), but if not restarted in timely manner blocked forks can be cleaved by structure-specific nucleases including the MUS81-EME2 complex and EEPD1, and possibly Metnase, causing fork collapse to DSBs (61-65). Importantly, the replication-associated, secondary DSBs induced by IR contribute to cell killing (53).

Stalled and collapsed forks are preferentially restarted via HR which maintains genome stability and suppresses cancer (66,67). However, HR is a double-edged sword: when forks are not restarted in timely manner, HR factors can mediate formation of branched structures including “chicken feet” (68) (Figure 3), a recombination intermediate in a damage tolerance pathway. Some structures are toxic recombination intermediates that can induce genome instability or cause cell death if not resolved properly (69,70). The alternative to HR to repair DSBs at collapsed forks is NHEJ (40), but because DSBs at collapsed replication forks are one-ended (Figure 3), joining of DSBs at different collapsed forks causes large-scale genome rearrangements including deletions and chromosome translocations (71-73), or cell death if essential genes are inactivated or dicentric chromosomes are produced. From the discussion above it is clear that the tumor-killing effects of IR-induced DNA damage can be augmented in several ways. Repair of frank DSBs can be suppressed by inhibiting cNHEJ; repair of replication-associated, secondary DSBs can be suppressed by inhibiting HR; and replication stress can be increased by inhibiting repair of single-strand damage, e.g., with PARP1 inhibitors (74) or by blocking replication arrest with checkpoint inhibitors, which increases encounters of replication forks with unrepaired lesions. Given that the highly networked DDR provides tumor cells with ample “escape routes”, the most effective approaches may rely on multiple layers of targeting to achieve synthetic lethality, for example, by exploiting specific tumor weaknesses such as HR defects in some breast, ovarian, and other cancers with PARP1 inhibitors, or inducing “artificial synthetic lethality” by targeting multiple repair or checkpoint pathways in combination with radiotherapy (75).

Figure 3 Sample fates of stalled replication forks. Stalled forks may regress to chicken foot structures that can serve as intermediates in an HR-dependent, lesion tolerance pathway; such structures may be cleaved, causing instability. Alternatively, stalled forks may be cleaved, causing fork collapse to one-ended DSBs. Collapsed forks may be restarted by HR, which preserves genome stability, or broken ends may be joined by NHEJ with ends from other collapsed forks, causing genome instability. HR, homologous recombination; NHEJ, non-homologous end joining.

Distinctions between low and high linear energy transfer (LET) IR-induced DNA damage and repair

DNA repair systems evolved to deal with the constant threat of DNA damage from exogenous sources (UV light, low-level background ionizing radiation, genotoxic chemicals) and from endogenous ROS produced during oxidative metabolism. For the most part, DNA damage from these sources is widely dispersed, and repair pathways are highly efficient repairing such damage (76). LET is a measure of ionization density along a radiation track, and low LET IR largely produces dispersed DNA lesions. High LET IR, such as carbon and iron ions, create densely ionizing tracks that create complex, clustered DNA lesions including clustered DSBs, and clustered single-strand damage (77-82). Clustered DNA damage is repaired slowly or not at all, and this is a key reason why high LET carbon ions have 2–3-fold higher relative biological effect than low LET IR (83,84). It has been argued that the cytotoxic effects of low LET IR are largely due to occasional clustered lesions (78,83). In this view, the increased cytotoxicity of high LET IR simply exploits the intrinsic weakness of DNA repair systems to process clustered lesions. This can be understood as a consequence of the lack of natural selective pressure to develop systems capable of efficiently repairing clustered DNA damage. Clinical success with high LET carbon ions probably also reflects the fact that high LET IR has lower dependence on oxygen (83,84), thus hypoxic tumors resistant to traditional radiotherapy (e.g., head/neck, melanoma, and pancreatic cancers) are more readily controlled by carbon ion radiotherapy (85,86). The natural resistance of hypoxic tumors to low LET IR is a growing concern given recent evidence that most solid tumors have hypoxic regions (87).

It is well established that clustered damage is repaired less efficiently than dispersed damage (78,79,81,83). Several studies indicate that HR plays a more important role in repair of damage created by high LET IR than low LET IR (82,88,89), at least in part due to increased resection of complex lesions induced by high LET IR (90), which promotes HR and suppresses cNHEJ. The difficulty in achieving proper repair of clustered DSBs does not depend on associated (clustered) single-strand damage, as clustered DSBs at engineered I-SceI nuclease sites greatly increases cell killing and causes massive genome instability (chromothripsis, chromosome translocations) that increases with increasing DSB density (91). The chromosome translocations were formed by aNHEJ, indicating that clustered DSBs are refractory to repair by cNHEJ (91). These results also indicate that high LET IR-induced chemical modifications at broken ends does not account for inhibition of cNHEJ, as I-SceI-induced DSBs have “clean” ligatable ends. It appears instead that DSB clustering per se suppresses cNHEJ.


Targeting the DDR to sensitize tumor cells to IR

Because of the cytotoxicity of DSBs, DSB repair proteins are attractive targets to sensitize tumors to radiotherapy. As cNHEJ is a primary determinant of resistance to low LET IR (X-rays, γ-rays, protons) throughout the cell cycle (45,92,93), blocking cNHEJ has long been pursued as a general strategy to radiosensitize tumors to IR, principally with inhibitors of DNA-PKcs (94,95). However, early DNA-PKcs inhibitors were relatively non-specific (96), and later versions with high specificity suffered from other problems, such as low bioavailability and rapid clearance or inactivation in vivo (97). The Turchi lab is currently investigating Ku inhibitors (3). LigIV inhibitors have also been investigated, but to date they have shown off-target effects or activities are too low to be therapeutically useful (98,99). In general, the promise of cNHEJ inhibition revealed in biochemical and cellular studies has not translated well to preclinical or clinical studies (3).

With the recognition that certain cancers harbor HR defects, the success in treating such cancers with PARP1 inhibitors, and the importance of HR for managing replication stress, including IR-induced replication stress, there is increasing interest in targeting HR, and RAD51 in particular, for therapeutic gain. RAD51 is frequently overexpressed in cancers (100), which is associated with therapeutic resistance, poor prognosis, and increased metastasis (101). Overexpression of RAD51 is likely both an adaptation to oncogenic replication stress (27,102) and a driver of tumor progression via genome destabilization (103,104). RAD51 may be targeted by genetic means (e.g., siRNA knockdown or expression of micro-RNAs that inhibit RAD51), or by using small molecules to inhibit RAD51 biochemical functions (101). Downregulation of RAD51 preferentially sensitized tumor vs. normal cells to cisplatin-induced DNA damage (105). Although this suggests similar benefits may be seen with radiotherapy, one study failed to show such benefits with glioma cells (106). This might reflect the fact that HR plays a lesser role than cNHEJ in conferring resistance to IR. The effects of RAD51 inhibition might be more pronounced if cell survival was more dependent on managing IR-induced replication stress, an idea supported by improved outcomes when RAD51 inhibition is combined with PARP1 inhibition (106).

Personalizing cancer therapy based on known defects in DDR factors through synthetic lethal approaches is a rapidly expanding research area, both in the lab and in clinical practice (2,4,6-10,12,13,107-111). However, there are additional opportunities for targeting cancer by inducing “artificial synthetic lethality”, that is by combining radiotherapy with drugs that inhibit redundant DNA repair pathways. Studies of inhibitors of heat shock protein 90 (Hsp90) illustrate the value of this approach. Hsp90 is a molecular chaperone responsible for the conformational maintenance of a number of client proteins that play key roles in cell cycle arrest, DNA damage repair, and apoptosis following radiation (112). Inhibition of Hsp90 by different substances increases radiosensitivity of various cancer cell lines; however, although Hsp90 is the target for inhibition, radiosensitization reflects effects on various downstream proteins (113). Early Hsp90 inhibitors were based on the natural compound geldanamycin and its derivatives, 17-allylamino-17-demethoxygeldanamycin (17-AAG) and 17-DMAG (114,115), and next evolved to the orally bioavailable BIIB021 (116), all of which sensitize cells to IR. Studies of 17-AAG demonstrated radiosensitization of tumor cells resulted from downregulation of HR via BRCA2 degradation, and decreased RAD51 function (117). Subsequent studies revealed that second and third generation Hsp90 inhibitors PU-H71 and TAS-116 downregulated both HR and cNHEJ factors (118-120), thus these agents inhibit both major DSB repair pathways, and therefore induce a type of artificial synthetic lethality in combination with IR. Importantly, Hsp90 adopts different conformations in tumor vs. normal cells (121), thus Hsp90 inhibitors radiosensitize tumor cells, but not normal cells (117-120). In addition, these effects are seen in both p53 WT and p53 mutant tumor cells (122), suggesting the potential for broad use in cancer therapy. The advanced Hsp90 inhibitors that block both HR and NHEJ may prove especially useful in combination with high LET IR, given the more balanced reliance on both cNHEJ and HR in repairing clustered DNA damage (123).

ATM and ATR are master regulators of DNA damage checkpoints (cell cycle arrest) through Chk2 and Chk1, respectively (15,16), and they also promote HR (124,125). Given that these kinases phosphorylate many hundreds of targets (19), it is difficult to pinpoint radiosensitization mechanisms. Nonetheless, it is well-established that ATR suppresses origin firing, protects stressed replication forks by stabilizing the replisome, and promotes fork restart through RPA phosphorylation (15), and ATM is an important regulator of the G1/S checkpoint. Thus, inhibiting checkpoint proteins can enhance IR-induced cell killing by at least two mechanisms: inhibiting DNA repair, and increasing replication stress. ATM inhibitors such as KU55933 and KU60019 sensitize cells to IR (124,126-129) and the latter has been shown to sensitize cancer (glioma) cells but not normal fibroblasts (129). VE-821 and AZD6738 are specific ATR inhibitors currently in clinical trials as mono-therapy and in combination with radiotherapy for solid tumors (110). Chk1 inhibition by siRNA increases DNA damage-induced apoptosis and radiosensitizes p53-deficient cancer cells (130), and selective Chk1 inhibitors, CEP-3891, Chir-124, and UCN-01 also enhanced cellular radiosensitivity (131). Small molecule inhibitors of Chk1 (SAR-020106), Chk2 (PV1019), and Chk1/Chk2 (AZD7762 and XL-844) have also been tested in preclinical studies in combination with IR (110). PV1019 is particularly interesting as it selectively kills or suppresses growth of cancer cells that overexpress Chk2 while protecting normal thymocytes from IR-induced apoptosis (132).


Translational research in IR-induced DNA damage: rodent models

The ability to study IR-induced DSBs in an animal model system have allowed a comprehensive evaluation of biological effects, as well as the development of novel agents for radiotherapy, radiation protection, and treatment of radiation injury. There are numerous spontaneous and experimentally induced mutation syndromes caused by the dysfunction of various components of the DSB repair pathways that manifest as certain radiation hypersensitivity and chromosomal instability phenotypes. For example, mutation of the NHEJ process critical for T- and B-lymphocyte receptor development was used to create scid mice (133,134). When scid mice were irradiated, it was discovered that myeloid cells and fibroblasts were markedly more radiosensitive than those from control mice (133). Below we highlight examples of translational studies utilizing preclinical rodent models to understand how manipulation of IR-induced DNA damage repair and signaling can lead to clinical advances in radiation oncology. As described, pathways involved in signaling and repair of IR-induced DSBs are critical targets for cytotoxicity and the outcome of radiation therapy. Translational studies using rodent models to test the efficacy of combined DDR-modulating compounds and IR have been performed with inhibitors of ATM, DNA-PKcs, Chk1/Chk2, PARP1, and Hsp90.

Mutations in ATM cause radiation hypersensitivity in patients with the autosomal recessive disorder, ataxia-telangiectasia; therefore, inhibition of ATM within tumors is a therapeutic approach designed to disrupt DSB repair and cause tumor radiosensitization. KU55933 is a relatively specific ATM inhibitor, but its analog KU60019 is more effective at blocking ATM phosphorylation (129). Using orthotopic xenograft models of glioblastoma multiforme in mice, it was demonstrated that combined IR and KU60019 significantly increased the survival of mice by 2- to 3-fold when compared with controls (135); importantly, tumor radiosensitization was more pronounced in mice harboring p53-mutant glioblastoma xenografts than mice with genetically matched p53-wild-type tumors (135). The ability of ATM inhibitors to sensitize p53-deficient cancer cells to IR is critical in light of the tumor resistance to therapy associated with p53 mutations, and the high frequency of p53 mutations in malignant tumors (136).

A novel and potent mTOR inhibitor, NVP-BEZ235, was shown to also inhibit ATM and DNA-PKcs (137), which have catalytic domains highly homologous to phosphoinositide 3-kinases (138,139). mTOR is also a PIKK family member, thus NVP-BEZ235 is a general PIKK inhibitor. NVP-BEZ235 blocks both cNHEJ and HR repair pathways and significantly attenuates DSB repair (137). In addition to reducing phosphorylation of ATM targets and G2/M cell cycle checkpoint activation, NVP-BEZ235 was shown to radiosensitize a panel of glioblastoma multiforme cell lines. Although the effects of IR and NVP-BEZ235 on tumor growth delay were not investigated, mice bearing subcutaneous glioblastoma multiforme xenografts were used to demonstrate that NVP-BEZ235 significantly impairs IR-induced DSB repair, validating the efficacy of this drug as a DNA repair inhibitor in vivo (137). However, as discussed by Goldstein and Kastan (136), a potential drawback of NVP-BEZ235 is that it also inhibits ATR, which is crucial in resolving stalled replication forks, and more generally for cell survival (140). Disruption of ATR is lethal at both the cellular and organism levels (141,142). As such, NVP-BEZ235 was predicted to induce unacceptable systemic toxicities in patients (136). Indeed, a phase II study of NVP-BEZ235 in patients with advanced pancreatic neuroendocrine tumors revealed that the drug was poorly tolerated and, because of this, it failed to reach the second stage of the study despite observed disease stability (143).

As noted above, tumor radiosensitization can be achieved via inhibition of Chk1- and/or Chk2-mediated checkpoint signaling. Rodent models have been used to demonstrate the efficacy of Chk1 inhibitors as cancer therapeutics. UCN-01 proved to be efficacious as a single agent in delaying tumor growth of orthotopic and subcutaneous xenograft glioblastoma multiforme tumor models in mice (144), and the Chk1 inhibitor MK8776 sensitized subcutaneous pancreatic cancer xenografts to combination therapy with gemcitabine and IR (145). In an alternative approach, a genetically modified mouse model revealed that Chk2−/− mice were more resistant to normal tissue radiation damage than WT mice, as characterized in splenic lymphocytes, thymocytes, and neurons of developing brains (146). As noted above, the Chk2 inhibitor PV1019 protects normal thymocytes from IR-induced apoptosis while simultaneously showing antiproliferative effects in cancer cells that express Chk2 at high levels (132), but this compound has not yet been tested in animal models.

Because IR-induced base modifications and single-strand breaks can indirectly induce toxic DSBs by blocking replication forks, inhibition of BER provides another means of radiosensitization. PARP1 promotes repair of single-strand lesions and it has long been known that PARP1 inhibition or genetic deletion enhances radiosensitivity (147-149). Genetically-modified PARP1−/− mice are hypersensitive to whole body irradiation compared to WT controls (150), and PARP1 inhibitors have been demonstrated to radiosensitize rodent xenograft models of human cervical carcinoma (151), colorectal cancer (152), lung cancer (153,154), head and neck squamous cell carcinoma (155), glioblastoma (156), and colon cancer (157). These preclinical studies were critical to the translation of PARP1 inhibitors to human clinical trials.

Hsp90 inhibitors suppress HR by downregulating RAD51 and/or BRCA2 (112,117-120). Preclinical studies of Hsp90 inhibitors have been promising. In vivo radiosensitization with various Hsp90 inhibitors has been shown in human tumor xenograft models of cervical (158), prostate (159), and head and neck squamous cell carcinoma (116). A synthetic alternative, NVP-AUY922, was also developed and shown in vivo to delay tumor growth and increase end-point survival in a head and neck squamous cell carcinoma xenograft model (112). Visual impairment has been a serious toxicity effect associated with early-generation Hsp90 inhibitors, and this has been largely overcome by the development of TAS-116 (160). Lee et al. (119) investigated the radiosensitizing effects of TAS-116 in low LET X-ray and high LET carbon ion-irradiated human cancer cells and mouse tumor xenografts. TAS-116 decreased cell survival of both X-ray and carbon ion-irradiated human cancer cell lines (HeLa and H1299 cells), and similar to other Hsp90 inhibitors, it did not affect radiosensitivity of noncancerous human fibroblasts. The combined treatment of mouse subcutaneous cervical tumor xenografts with carbon ions and TAS-116 significantly delayed tumor growth compared to controls (119). In another HR targeting approach, growth of HeLa cell tumor xenografts in mice was significantly reduced by cisplatin combined with RAD51 knockdown (105), suggesting similar effects might be obtained by combining RAD51 inhibition with radiotherapy.

While there is great value in these preclinical rodent models for studying radiation-induced DNA damage repair and therapeutic applications of DDR inhibitors, it is important to recognize the limitations of this model for translational science. First, in order to study human cancers using the standard subcutaneous xenograft model in vivo, the mouse immune system must be significantly altered, such as with athymic nude or severe combined immunodeficiency (SCID) mice. However, the immune system strongly influences overall therapeutic responses of cancer patients. Further, while murine orthotopic tumors are grown within their host tissue and mimic local tumor growth and metastasis, xenograft models usually rely on highly passaged cell lines and clonal selection may not mimic human tumors arising spontaneously in patients. Alternatives to xenograft models are genetically engineered animal models (GEM), where oncogenes are activated and/or tumor suppressor genes are inactivated to induce tumors in situ, generally via the temporally controlled and tissue-specific expression of CRE recombinase (161-163). Tumors initiated within GEM rodent models develop in a more natural microenvironment, with supporting vasculature, stromal cells, and an intact immune system; however, inducible tumors tend to be artificially homogenous, lacking the vast heterogeneity seen in natural, spontaneous cancer (164,165). Finally, murine tumor models frequently error towards false positive therapeutic results, with cancer cures in mice frequently failing to translate to humans (166). Despite these limitations, preclinical research in rodent models is a crucial and necessary step in taking positive therapeutic results from cell lines to the clinic. In the next section we discuss advantages of translational research that involves clinical trials to treat spontaneous tumors in companion animals.


Past, present and future translational research opportunities with clinical trials to treat spontaneous tumors in companion animals

The value of spontaneously occurring tumors as a tool for translational research has recently received increased attention by cancer scientists. In June 2015, the US National Academy of Science Institute of Medicine’s National Cancer Policy forum hosted a workshop on comparative oncology. Comparative oncology was defined as the study of naturally occurring cancers in animals as models for human disease. This workshop titled “The Role of Clinical Studies for Pets with Naturally Occurring Tumors in Translational Cancer Research” explored a number of topics including the rationale for clinical trials, canine tumor biology, and lessons learned from comparative oncology (167). Although the focus of the workshop was primarily on the role of comparative oncology in drug development, the take-home messages are broadly applicable to cancer therapy in general including radiotherapy.

Current interest not-withstanding, comparative oncology in the field of radiation oncology has been ongoing since the discovery of X-rays. Understanding normal tissue tolerance and regulatory factors is the cornerstone of radiation oncology. The effects of the radiation therapy should not be worse than the disease, so radiation dose is constrained by normal tissue tolerance. While speciation resulted in gradual changes over the course of evolutionary time, DNA repair is critical for the maintenance of genome integrity, and DNA repair mechanisms are highly conserved (168). Thus, the radiation sensitivity, and the radiation tolerance of normal tissues, is similar for most mammalian species. This means that radiation therapy data from humans is valuable for the treatment of veterinary patients, and vice versa. While naturally occurring tumors are used in a wide array of studies, including evaluation of chemotherapy and cancer imaging, here we focus on early studies that evaluated tumor control and radiation effects to demonstrate important principles in radiation oncology, and on more recent molecular advances that highlight the importance of comparative oncology in radiotherapy research.

Shortly after Roentgen discovered the X-ray in 1895, both human and animal skin tumors were treated empirically. With minimal scientific methodology, encouraging responses to treatment were reported in human and veterinary patients, including dogs and horses (169-171). The fields of human and veterinary radiation oncology followed parallel paths and information was frequently shared regarding treatment outcome and normal tissue effects at national and international meetings (172). Despite early excitement, radiation therapy did not provide an effective or easy treatment for cancer. Tumor control was not durable, and patients were often plagued with severe radiation effects (173). Dr. Henri Coutard was a radiation oncologist at the Curie Institute in Paris whose keen observations changed radiation oncology (174). By the 1920s he was studying fractionation schemes using a more scientific method. He evaluated the impact of dose per fraction, total dose, tumor size, field size, and overall treatment time on tumor control and adverse radiation effects. While these studies did not elucidate an understanding of underlying biology, they did provide a protocol that could be safely delivered and resulted in more durable tumor control. During the same period, Dr. Alois Pommer, a veterinarian at what was then called the Vienna Veterinary High School, received funding from the Rockefeller Institute to start a radiation therapy program for animal patients. Pommer, like Coutard, published extensively on fractionation schedules, tumor control and radiation effects (173,175). These two pioneering radiation oncologists, both of whom used orthovoltage radiation equipment, provided a template for safe treatment for future decades.

By the late 1960s scientists had a rudimentary understanding of radiation biology, based on cell culture experiments and work with rodent tumor models. The Elkind lab reported that mammalian cells in culture could repair radiation damage, explaining one of the benefits of delivering dose in fractions instead of in large single doses (176). This led to other hypotheses about mechanisms underlying the efficacy of fractionated radiation therapy such as Wither’s description of the four Rs: repair, redistribution (in the cell cycle), repopulation, and reoxygenation of hypoxic regions (177). But cells in culture are isolated systems, and induced tumors in rodents lacked the complexity of human tumors. Naturally occurring tumors in companion animals (dogs and cats) share many commonalties with human tumors (178-180). Histological appearance and behavior of some animal tumors are markedly similar. The relative size of the tumor to the body is more comparable to that in rodent tumors, yet the gross tumor size is large enough for serial sampling. The size of the patients makes sophisticated imaging, treatment and monitoring possible, using the same technology used for humans. Most importantly, the tumors share similar microenvironments. In both species tumors can be acutely or chronically hypoxic, and intra- and extracellular pH may be altered. Genetic changes observed in human tumors are seen in canine tumors, and unlike many rodent models, the immune system is intact (180).

Dr. Edward L. Gillette helped establish a Veterinary Radiation Oncology program at Colorado State University in the late 1950s, with a vision of using naturally occurring tumors as a model for human disease. In 1968 he worked with Drs. Herman Suit and Rodney Withers at M.D. Anderson Cancer Center (181). These colleagues saw the value of the dog model and they encouraged Gillette to pursue research in comparative oncology. Gillette’s work in intraoperative radiation therapy set the standard for normal tissue tolerances in human and veterinary medicine, and his work evaluating the impact of fraction size and field size helped create data on α/β ratios that are still used by both veterinary and human radiation oncologists (181-191). But it was Gillette’s work using spontaneous tumors to evaluate radiation biological principles that elevated the value of the naturally occurring tumor model. Therapeutic gain in the context of radiotherapy was based on the concept that to improve treatment outcome (tumor control) while maintaining quality of life (limited late effects), both tumor control and late effects needed to be evaluated and compared between different treatments. Tumor control without complications is the ultimate exploitation of differences in DNA repair characteristics between tumor cells and late responding normal tissues. Gillette was able to prove this principle because of the flexibility for trial design in the naturally occurring tumor model in veterinary patients. He first conducted a trial in which dogs with naturally occurring squamous cell carcinomas of the oral cavity were randomized to receive different total doses of fractionated radiation. The patients were followed through their lifetime for tumor control and late radiation effects. He used this information to determine the dose that provided the best tumor control with least complications, and these radiation dose groups were also tested with hyperthermia treatments (192,193). These studies helped inform the design of clinical trials in human patients.

These early clinical studies of normal tissue and tumor responses to radiation in naturally occurring tumors in dogs led scientists to further utilize this translational model by evaluating DNA repair characteristics in normal dogs and in canine tumors. Sequencing of the canine ATM mRNA demonstrated high homology with the human counterpart, both in the promoter and overall gene structure, facilitating comparative studies of ATM function in dogs and a potential model for ATM deficiency (194). There is significant overlap between deregulated human and canine genes in mammary tumors, as well as from normal mammary tissues (195). ATM mRNA and protein expression were shown to be downregulated in canine mammary tumors (196), consistent with frequently observed checkpoint and DNA repair defects in human tumors (2,16). Recently the molecular mechanisms of both XLF and Ku-dependent NHEJ in canines was evaluated and proposed as a platform for development of novel chemotherapies for dogs and humans (197,198). Human cancer cell lines have been extensively characterized over the years, but only recently have in-depth studies of radiation sensitivity, including analysis of radiation-induced DSB repair been undertaken. Twenty-seven established canine cell lines were evaluated and radiation-induced DSB repair was related to radiation sensitivity, as previously shown in human tumors (199). Thus, canine cancer models present many translational research opportunities to exploit fundamental knowledge about DNA repair to improve radiotherapy.


Conclusions

There was broad consensus at the Institute of Medicine’s National Cancer Policy Workshop that a more detailed characterization of the canine genome, and expansion of comparative oncology research will define similarities and differences in dog and human cancers to benefit both companion animal and human patients (167). There are broad opportunities to apply well-established clinical trials techniques to explore promising leads in radiotherapy research: To augment and personalize radiotherapy using many forms of tumor molecular profiling (transcriptomics, metabolomics, DDR network analysis, etc.) and novel combination chemo-radiotherapies. By developing tools to evaluate DDR proteins in canine cells (197) or in treated tumors, biological responses to therapy can be defined and correlated with treatment outcomes to seek improved therapeutic strategies.

There are many key frontiers in radiation oncology that can be advanced through expansion of comparative oncology research; several are highlighted here. The past decade has produced an explosion of information about DDR “strengths and weaknesses” that are just beginning to be exploited in clinical settings (1,2,5,101,107,110,111,200,201). DDR manipulation can involve radiosensitization of tumors, as well as radioprotection of normal tissue which would allow safe delivery of increased doses to tumors. Companion animals will likely play a key role at the new frontier of radio-immunotherapy with PD-1 and PD-L1 inhibitors, to unleash the power of the immune system to “clean up” micrometastases (and perhaps even well-established metastases) by enhancing the abscopal effect (201,202). Tumor imaging through advanced radiologic methods remains a critical foundation for radiotherapy, and an area with great potential for diagnostic and theranostic progress (again, best studied in large, easily imaged, animals), especially with novel PET isotopes linked to tumor-tropic biologicals or small molecules (203,204). Carbon ion radiotherapy, with >22 years human clinical practice and >20,000 human patients treated, is making great strides (85), but it is clear that to accelerate our understanding of the complexities of tumor and normal tissue responses to heavy and light ion radiation we must attack on all fronts: at the basic cell and molecular level, with preclinical small animal models, and with clinical studies of naturally occurring tumors in companion animals. This integrated approach offers the best chance to rapidly translate life-saving cancer cures to human clinical practice.


Acknowledgments

Funding: Research in JA Nickoloff’s laboratory was supported by grant NIH R01 GM084020.


Footnote

Provenance and Peer Review: This article was commissioned by the Guest Editors (Marco Durante, Giusi I. Forte, Giorgio Russo) for the series “Radiobiological models towards a personalized radiation oncology” published in Translational Cancer Research. The article has undergone external peer review.

Conflicts of Interest: All authors have completed the ICMJE uniform disclosure form (available at http://dx.doi.org/10.21037/tcr.2017.06.02). The series “Radiobiological models towards a personalized radiation oncology” was commissioned by the editorial office without any funding or sponsorship. The authors have no other conflicts of interest to declare.

Ethical Statement: The authors are accountable for all aspects of the work in ensuring that questions related to the accuracy or integrity of any part of the work are appropriately investigated and resolved.

Open Access Statement: This is an Open Access article distributed in accordance with the Creative Commons Attribution-NonCommercial-NoDerivs 4.0 International License (CC BY-NC-ND 4.0), which permits the non-commercial replication and distribution of the article with the strict proviso that no changes or edits are made and the original work is properly cited (including links to both the formal publication through the relevant DOI and the license). See: https://creativecommons.org/licenses/by-nc-nd/4.0/.


References

  1. Basu B, Yap TA, Molife LR, et al. Targeting the DNA damage response in oncology: past, present and future perspectives. Curr Opin Oncol 2012;24:316-24. [Crossref] [PubMed]
  2. Curtin NJ. DNA repair dysregulation from cancer driver to therapeutic target. Nat Rev Cancer 2012;12:801-17. [Crossref] [PubMed]
  3. Gavande NS, Vandervere-Carozza PS, Hinshaw HD, et al. DNA repair targeted therapy: The past or future of cancer treatment? Pharmacol Ther 2016;160:65-83. [Crossref] [PubMed]
  4. O'Connor MJ. Targeting the DNA damage response in cancer. Mol Cell 2015;60:547-60. [Crossref] [PubMed]
  5. Pearl LH, Schierz AC, Ward SE, et al. Therapeutic opportunities within the DNA damage response. Nat Rev Cancer 2015;15:166-80. [Crossref] [PubMed]
  6. Nickoloff JA. DNA repair dysregulation in cancer: From molecular mechanisms to synthetic lethal opportunities. In: Wondrak GT, editor. Stress Response Pathways in Cancer From Molecular Targets to Novel Therapeutics. New York: Springer, 2014:(7-28).
  7. Helleday T. The underlying mechanism for the PARP and BRCA synthetic lethality: clearing up the misunderstandings. Mol Oncol 2011;5:387-93. [Crossref] [PubMed]
  8. Killock D. Targeted therapies: DNA polymerase theta-a new target for synthetic lethality? Nat Rev Clin Oncol 2015;12:125. [Crossref] [PubMed]
  9. Rehman FL, Lord CJ, Ashworth A. Synthetic lethal approaches to breast cancer therapy. Nat Rev Clin Oncol 2010;7:718-24. [Crossref] [PubMed]
  10. Shaheen M, Allen C, Nickoloff JA, et al. Synthetic lethality: exploiting the addiction of cancer to DNA repair. Blood 2011;117:6074-82. [Crossref] [PubMed]
  11. Schmitt MW, Loeb LA, Salk JJ. The influence of subclonal resistance mutations on targeted cancer therapy. Nat Rev Clin Oncol 2016;13:335-47. [Crossref] [PubMed]
  12. Blomen VA, Májek P, Jae LT, et al. Gene essentiality and synthetic lethality in haploid human cells. Science 2015;350:1092-6. [Crossref] [PubMed]
  13. Mohni KN, Thompson PS, Luzwick JW, et al. A synthetic lethal screen identifies DNA repair pathways that sensitize cancer cells to combined ATR inhibition and cisplatin treatments. PLoS ONE 2015;10:e0125482 [Crossref] [PubMed]
  14. Chatterji M, Tsai CL, Schatz DG. New concepts in the regulation of an ancient reaction: transposition by RAG1/RAG2. Immunol Rev 2004;200:261-71. [Crossref] [PubMed]
  15. Yazinski SA, Zou L. Functions, regulation, and therapeutic implications of the ATR checkpoint pathway. Annu Rev Genet 2016;50:155-73. [Crossref] [PubMed]
  16. Smith J, Tho LM, Xu N, et al. The ATM-Chk2 and ATR-Chk1 pathways in DNA damage signaling and cancer. Adv Cancer Res 2010;108:73-112. [Crossref] [PubMed]
  17. Hurley PJ, Bunz F. ATM and ATR: components of an integrated circuit. Cell Cycle 2007;6:414-7. [Crossref] [PubMed]
  18. Mordes DA, Cortez D. Activation of ATR and related PIKKs. Cell Cycle 2008;7:2809-12. [Crossref] [PubMed]
  19. Matsuoka S, Ballif BA, Smogorzewska A, et al. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 2007;316:1160-6. [Crossref] [PubMed]
  20. Tapia-Alveal C, Calonge TM, O'Connell MJ. Regulation of Chk1. Cell Div 2009;4:8. [Crossref] [PubMed]
  21. Merry C, Fu K, Wang J, et al. Targeting the checkpoint kinase Chk1 in cancer therapy. Cell Cycle 2010;9:279-83. [Crossref] [PubMed]
  22. Mladenov E, Magin S, Soni A, et al. DNA double-strand-break repair in higher eukaryotes and its role in genomic instability and cancer: Cell cycle and proliferation-dependent regulation. Semin Cancer Biol 2016;37-38:51-64. [Crossref] [PubMed]
  23. Berti M, Vindigni A. Replication stress: getting back on track. Nat Struct Mol Biol 2016;23:103-9. [Crossref] [PubMed]
  24. Branzei D, Foiani M. The checkpoint response to replication stress. DNA Repair (Amst) 2009;8:1038-46. [Crossref] [PubMed]
  25. Gaillard H, García-Muse T, Aguilera A. Replication stress and cancer. Nat Rev Cancer 2015;15:276-89. [Crossref] [PubMed]
  26. Mazouzi A, Velimezi G, Loizou JI. DNA replication stress: causes, resolution and disease. Exp Cell Res 2014;329:85-93. [Crossref] [PubMed]
  27. Zeman MK, Cimprich KA. Causes and consequences of replication stress. Nat Cell Biol 2014;16:2-9. [Crossref] [PubMed]
  28. Livraghi L, Garber JE. PARP inhibitors in the management of breast cancer: current data and future prospects. BMC Med 2015;13:188. [Crossref] [PubMed]
  29. Choi YE, Meghani K, Brault ME, et al. Platinum and PARP inhibitor resistance due to overexpression of MicroRNA-622 in BRCA1-Mutant ovarian cancer. Cell Rep 2016;14:429-39. [Crossref] [PubMed]
  30. Jiang J, Lu Y, Li Z, et al. Ganetespib overcomes resistance to PARP inhibitors in breast cancer by targeting core proteins in the DNA repair machinery. Invest New Drugs 2017;35:251-9. [Crossref] [PubMed]
  31. Murai J, Feng Y, Yu GK, et al. Resistance to PARP inhibitors by SLFN11 inactivation can be overcome by ATR inhibition. Oncotarget 2016;7:76534-50. [PubMed]
  32. Yalon M, Tuval-Kochen L, Castel D, et al. Overcoming resistance of cancer cells to PARP-1 inhibitors with three different drug combinations. PLoS ONE 2016;11:e0155711 [Crossref] [PubMed]
  33. Alizadeh E, Orlando TM, Sanche L. Biomolecular damage induced by ionizing radiation: the direct and indirect effects of low-energy electrons on DNA. Annu Rev Phys Chem 2015;66:379-98. [Crossref] [PubMed]
  34. Mladenov E, Magin S, Soni A, et al. DNA double-strand break repair as determinant of cellular radiosensitivity to killing and target in radiation therapy. Front Oncol 2013;3:113. [Crossref] [PubMed]
  35. Sharpless NE, Ferguson DO, O'hagan RC, et al. Impaired nonhomologous end-joining provokes soft tissue sarcomas harboring chromosomal translocations, amplifications, and deletions. Mol Cell 2001;8:1187-96. [Crossref] [PubMed]
  36. Ferguson DO, Sekiguchi JM, Chang S, et al. The nonhomologous end-joining pathway of DNA repair is required for genomic stability and the suppression of translocations. Proc Natl Acad Sci U S A 2000;97:6630-3. [Crossref] [PubMed]
  37. Symington LS, Gautier J. Double-strand break end resection and repair pathway choice. Annu Rev Genet 2011;45:247-71. [Crossref] [PubMed]
  38. Iliakis G, Murmann T, Soni A. Alternative end-joining repair pathways are the ultimate backup for abrogated classical non-homologous end-joining and homologous recombination repair: Implications for the formation of chromosome translocations. Mutat Res Genet Toxicol Environ Mutagen 2015;793:166-75. [Crossref] [PubMed]
  39. Simsek D, Brunet E, Wong SY, et al. DNA ligase III promotes alternative nonhomologous end-joining during chromosomal translocation formation. PLoS Genet 2011;7:e1002080 [Crossref] [PubMed]
  40. Wray J, Williamson EA, Singh SB, et al. PARP1 is required for chromosomal translocations. Blood 2013;121:4359-65. [Crossref] [PubMed]
  41. Simsek D, Jasin M. Alternative end-joining is suppressed by the canonical NHEJ component Xrcc4-ligase IV during chromosomal translocation formation. Nat Struct Mol Biol 2010;17:410-6. [Crossref] [PubMed]
  42. Weinstock DM, Brunet E, Jasin M. Formation of NHEJ-derived reciprocal chromosomal translocations does not require Ku70. Nat Cell Biol 2007;9:978-81. [PubMed]
  43. Wray J, Williamson EA, Chester S, et al. The transposase domain protein Metnase/SETMAR suppresses chromosomal translocations. Cancer Genet Cytogenet 2010;200:184-90. [Crossref] [PubMed]
  44. Krokan HE, Bjoras M. Base excision repair. Cold Spring Harb Perspect Biol 2013;5:a012583 [Crossref] [PubMed]
  45. Kakarougkas A, Jeggo PA. DNA DSB repair pathway choice: an orchestrated handover mechanism. Br J Radiol 2014;87:20130685 [Crossref] [PubMed]
  46. Löbrich M, Jeggo PA. Harmonising the response to DSBs: a new string in the ATM bow. DNA Repair (Amst) 2005;4:749-59. [Crossref] [PubMed]
  47. Zhang Y, Jasin M. An essential role for CtIP in chromosomal translocation formation through an alternative end-joining pathway. Nat Struct Mol Biol 2011;18:80-4. [Crossref] [PubMed]
  48. Daley JM, Chiba T, Xue X, et al. Multifaceted role of the Topo IIIα-RMI1-RMI2 complex and DNA2 in the BLM-dependent pathway of DNA break end resection. Nucleic Acids Res 2014;42:11083-91. [Crossref] [PubMed]
  49. Nimonkar AV, Genschel J, Kinoshita E, et al. BLM-DNA2-RPA-MRN and EXO1-BLM-RPA-MRN constitute two DNA end resection machineries for human DNA break repair. Genes Dev 2011;25:350-62. [Crossref] [PubMed]
  50. Sturzenegger A, Burdova K, Kanagaraj R, et al. DNA2 cooperates with the WRN and BLM RecQ helicases to mediate long-range DNA end resection in human cells. J Biol Chem 2014;289:27314-26. [Crossref] [PubMed]
  51. Nickoloff JA, Brenneman MA. Analysis of recombinational repair of DNA double-strand breaks in mammalian cells with I-SceI nuclease. In: Waldman AS, editor. Genetic Recombination - Reviews and Protocols. Methods in Molecular Biology. Totowa: Humana Press, 2004:(35-52).
  52. Ceccaldi R, Sarangi P, D'andrea AD. The fanconi anaemia pathway: new players and new functions. Nat Rev Mol Cell Biol 2016;17:337-49. [Crossref] [PubMed]
  53. Groth P, Luis Orta M, Elvers I, et al. Homologous recombination repairs secondary replication induced DNA double-strand breaks after ionizing radiation. Nucleic Acids Res 2012;40:6585-94. [Crossref] [PubMed]
  54. Harper JV, Anderson JA, O'neill P. Radiation induced DNA DSBs: Contribution from stalled replication forks? DNA Repair (Amst) 2010;9:907-13. [Crossref] [PubMed]
  55. Ward JF. Complexity of damage produced by ionizing radiation. Cold Spring Harb Symp Quant Biol 2000;65:377-82. [Crossref] [PubMed]
  56. Dextraze ME, Gantchev T, Girouard S, et al. DNA interstrand cross-links induced by ionizing radiation: an unsung lesion. Mutat Res 2010;704:101-7. [Crossref] [PubMed]
  57. Lomonosov M, Anand S, Sangrithi M, et al. Stabilization of stalled DNA replication forks by the BRCA2 breast cancer susceptibility protein. Genes Dev 2003;17:3017-22. [Crossref] [PubMed]
  58. Cobb JA, Bjergbaek L, Shimada K, et al. DNA polymerase stabilization at stalled replication forks requires Mec1 and the RecQ helicase Sgs1. EMBO J 2003;22:4325-36. [Crossref] [PubMed]
  59. Branzei D, Foiani M. Maintaining genome stability at the replication fork. Nat Rev Mol Cell Biol 2010;11:208-19. [Crossref] [PubMed]
  60. Lopes M, Cotta-Ramusino C, Pellicioli A, et al. The DNA replication checkpoint response stabilizes stalled replication forks. Nature 2001;412:557-61. [Crossref] [PubMed]
  61. Pepe A, West SC. MUS81-EME2 promotes replication fork restart. Cell Rep 2014;7:1048-55. [Crossref] [PubMed]
  62. Wu Y, Lee SH, Williamson EA, et al. EEPD1 rescues stressed replication forks and maintains genome stability by promoting end resection and homologous recombination repair. PLoS Genet 2015;11:e1005675 [Crossref] [PubMed]
  63. De Haro LP, Wray J, Williamson EA, et al. Metnase promotes restart and repair of stalled and collapsed replication forks. Nucleic Acids Res 2010;38:5681-91. [Crossref] [PubMed]
  64. Kim HS, Chen Q, Kim SK, et al. The DDN catalytic motif is required for Metnase functions in NHEJ repair and replication restart. J Biol Chem 2014;289:10930-8. [Crossref] [PubMed]
  65. Kim HS, Williamson EA, Nickoloff JA, et al. Metnase mediates loading of exonuclease 1 onto single Strand overhang DNA for end resection at stalled replication forks. J Biol Chem 2017;292:1414-25. [Crossref] [PubMed]
  66. Budzowska M, Kanaar R. Mechanisms of dealing with DNA damage-induced replication problems. Cell Biochem Biophys 2009;53:17-31. [Crossref] [PubMed]
  67. Allen C, Ashley AK, Hromas R, et al. More forks on the road to replication stress recovery. J Mol Cell Biol 2011;3:4-12. [Crossref] [PubMed]
  68. Jones RM, Petermann E. Replication fork dynamics and the DNA damage response. Biochem J 2012;443:13-26. [Crossref] [PubMed]
  69. Fabre F, Chan A, Heyer WD, et al. Alternate pathways involving Sgs1/Top3, Mus81/ Mms4, and Srs2 prevent formation of toxic recombination intermediates from single-stranded gaps created by DNA replication. Proc Natl Acad Sci U S A 2002;99:16887-92. [Crossref] [PubMed]
  70. Lambert S, Carr AM. Impediments to replication fork movement: stabilisation, reactivation and genome instability. Chromosoma 2013;122:33-45. [Crossref] [PubMed]
  71. Bunting SF, Nussenzweig A. End-joining, translocations and cancer. Nat Rev Cancer 2013;13:443-54. [Crossref] [PubMed]
  72. Byrne M, Wray J, Reinert B, et al. Mechanisms of oncogenic chromosomal translocations. Ann N Y Acad Sci 2014;1310:89-97. [Crossref] [PubMed]
  73. Hromas R, Williamson E, Lee SH, et al. Preventing the chromosomal translocations that cause cancer. Trans Am Clin Climatol Assoc 2016;127:176-95. [PubMed]
  74. Rouleau M, Patel A, Hendzel MJ, et al. PARP inhibition: PARP1 and beyond. Nat Rev Cancer 2010;10:293-301. [Crossref] [PubMed]
  75. Nickoloff JA. Improving cancer therapy by combining cell biological, physical, and molecular targeting strategies. Chin J Cancer Res 2013;25:7-9. [PubMed]
  76. Friedberg EC, Elledge SJ, Lehmann AR, et al. editors. DNA Repair, Mutagenesis, and Other Responses to DNA Damage. 1st ed. NY: Cold Spring Harbor Laboratory Press, 2014.
  77. Asaithamby A, Chen DJ. Mechanism of cluster DNA damage repair in response to high-atomic number and energy particles radiation. Mutat Res 2011;711:87-99. [Crossref] [PubMed]
  78. Eccles LJ, O'Neill P, Lomax ME. Delayed repair of radiation induced clustered DNA damage: friend or foe? Mutat Res 2011;711:134-41. [Crossref] [PubMed]
  79. Okayasu R. Repair of DNA damage induced by accelerated heavy ions--a mini review. Int J Cancer 2012;130:991-1000. [Crossref] [PubMed]
  80. Schipler A, Iliakis G. DNA double-strand-break complexity levels and their possible contributions to the probability for error-prone processing and repair pathway choice. Nucleic Acids Res 2013;41:7589-605. [Crossref] [PubMed]
  81. Sutherland BM, Bennett PV, Schenk H, et al. Clustered DNA damages induced by high and low LET radiation, including heavy ions. Phys Med 2001;17:202-4. [PubMed]
  82. Hada M, Sutherland BM. Spectrum of complex DNA damages depends on the incident radiation. Radiat Res 2006;165:223-30. [Crossref] [PubMed]
  83. Lomax ME, Folkes LK, O'neill P. Biological Consequences of radiation-induced DNA damage: relevance to radiotherapy. Clin Oncol (R Coll Radiol) 2013;25:578-85. [Crossref] [PubMed]
  84. Allen C, Borak TB, Tsujii H, et al. Heavy charged particle radiobiology: using enhanced biological effectiveness and improved beam focusing to advance cancer therapy. Mutat Res 2011;711:150-7. [Crossref] [PubMed]
  85. Kamada T, Tsujii H, Blakely EA, et al. Carbon ion radiotherapy in Japan: an assessment of 20 years of clinical experience. Lancet Oncol 2015;16:e93-e100. [Crossref] [PubMed]
  86. Tsujii H, Kamada T. A review of update clinical results of Carbon ion radiotherapy. Jpn J Clin Oncol 2012;42:670-85. [Crossref] [PubMed]
  87. Hill RP, Bristow RG, Fyles A, et al. Hypoxia and predicting radiation response. Semin Radiat Oncol 2015;25:260-72. [Crossref] [PubMed]
  88. Okayasu R, Okada M, Okabe A, et al. Repair of DNA damage induced by accelerated heavy ions in mammalian cells proficient and deficient in the non-homologous end-joining pathway. Radiat Res 2006;165:59-67. [Crossref] [PubMed]
  89. Wang H, Zhang X, Wang P, et al. Characteristics of DNA-binding proteins determine the biological sensitivity to high-linear energy transfer radiation. Nucleic Acids Res 2010;38:3245-51. [Crossref] [PubMed]
  90. Yajima H, Fujisawa H, Nakajima NI, et al. The complexity of DNA double Strand breaks is a critical factor enhancing end-resection. DNA Repair (Amst) 2013;12:936-46. [Crossref] [PubMed]
  91. Schipler A, Mladenova V, Soni A, et al. Chromosome thripsis by DNA double Strand break clusters causes enhanced cell lethality, chromosomal translocations and 53BP1-recruitment. Nucleic Acids Res 2016;44:7673-90. [Crossref] [PubMed]
  92. Shrivastav M, De Haro LP, Nickoloff JA. Regulation of DNA double-strand break repair pathway choice. Cell Res 2008;18:134-47. [Crossref] [PubMed]
  93. Jeggo P, Lavin MF. Cellular radiosensitivity: how much better do we understand it? Int J Radiat Biol 2009;85:1061-81. [Crossref] [PubMed]
  94. Li YH, Wang X, Pan Y, et al. Inhibition of non-homologous end joining repair impairs pancreatic cancer growth and enhances radiation response. PLoS ONE 2012;7:e39588 [Crossref] [PubMed]
  95. Begg AC, Stewart FA, Vens C. Strategies to improve radiotherapy with targeted drugs. Nat Rev Cancer 2011;11:239-53. [Crossref] [PubMed]
  96. Collis SJ, Deweese TL, Jeggo PA, et al. The Life and death of DNA-PK. Oncogene 2005;24:949-61. [Crossref] [PubMed]
  97. Nutley BP, Smith NF, Hayes A, et al. Preclinical pharmacokinetics and metabolism of a novel prototype DNA-PK inhibitor NU7026. Br J Cancer 2005;93:1011-8. [PubMed]
  98. Chen X, Zhong S, Zhu X, et al. Rational design of human DNA ligase inhibitors that target cellular DNA replication and repair. Cancer Res 2008;68:3169-77. [Crossref] [PubMed]
  99. Chu VT, Weber T, Wefers B, et al. Increasing the efficiency of homology-directed repair for CRISPR-Cas9-induced precise gene editing in mammalian cells. Nat Biotechnol 2015;33:543-8. [Crossref] [PubMed]
  100. Raderschall E, Stout K, Freier S, et al. Elevated levels of Rad51 recombination protein in tumor cells. Cancer Res 2002;62:219-25. [PubMed]
  101. Ward A, Khanna KK, Wiegmans AP. Targeting homologous recombination, new pre-clinical and clinical therapeutic combinations inhibiting RAD51. Cancer Treat Rev 2015;41:35-45. [Crossref] [PubMed]
  102. Bartkova J, Rezaei N, Liontos M, et al. Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 2006;444:633-7. [Crossref] [PubMed]
  103. Nagathihalli NS, Nagaraju G. RAD51 as a potential biomarker and therapeutic target for pancreatic cancer. Biochim Biophys Acta 2011;1816:209-18. [PubMed]
  104. Richardson C, Stark JM, Ommundsen M, et al. Rad51 overexpression promotes alternative double-strand break repair pathways and genome instability. Oncogene 2004;23:546-53. [Crossref] [PubMed]
  105. Ito M, Yamamoto S, Nimura K, et al. Rad51 siRNA delivered by HVJ envelope vector enhances the anti-cancer effect of cisplatin. J Gene Med 2005;7:1044-52. [Crossref] [PubMed]
  106. Quiros S, Roos WP, Kaina B. Rad51 and BRCA2--New molecular targets for sensitizing glioma cells to alkylating anticancer drugs. PLoS ONE 2011;6:e27183 [Crossref] [PubMed]
  107. Brown JS, O'Carrigan B, Jackson SP, et al. Targeting DNA Repair in Cancer: Beyond PARP Inhibitors. Cancer Discov 2017;7:20-37. [Crossref] [PubMed]
  108. Jeggo PA, Pearl LH, Carr AM. DNA repair, genome stability and cancer: a historical perspective. Nat Rev Cancer 2016;16:35-42. [Crossref] [PubMed]
  109. Majidinia M, Yousefi B. DNA damage response regulation by microRNAs as a therapeutic target in cancer. DNA Repair (Amst) 2016;47:1-11. [Crossref] [PubMed]
  110. Manic G, Obrist F, Sistigu A, et al. Trial Watch: Targeting ATM-CHK2 and ATR-CHK1 pathways for anticancer therapy. Mol Cell Oncol 2015;2:e1012976 [Crossref] [PubMed]
  111. Weber AM, Ryan AJ. ATM and ATR as therapeutic targets in cancer. Pharmacol Ther 2015;149:124-38. [Crossref] [PubMed]
  112. Zaidi S, McLaughlin M, Bhide SA, et al. The HSP90 inhibitor NVP-AUY922 radiosensitizes by abrogation of homologous recombination resulting in mitotic entry with unresolved DNA damage. PLoS ONE 2012;7:e35436 [Crossref] [PubMed]
  113. Maier P, Hartmann L, Wenz F, et al. Cellular Pathways in Response to Ionizing Radiation and Their Targetability for Tumor Radiosensitization. Int J Mol Sci 2016;17: [Crossref] [PubMed]
  114. Russell JS, Burgan W, Oswald KA, et al. Enhanced cell killing induced by the combination of radiation and the heat shock protein 90 inhibitor 17-allylamino-17- demethoxygeldanamycin: a multitarget approach to radiosensitization. Clin Cancer Res 2003;9:3749-55. [PubMed]
  115. Dote H, Cerna D, Burgan WE, et al. ErbB3 expression predicts tumor cell radiosensitization induced by Hsp90 inhibition. Cancer Res 2005;65:6967-75. [Crossref] [PubMed]
  116. Yin X, Zhang H, Lundgren K, et al. BIIB021, a novel Hsp90 inhibitor, sensitizes head and neck squamous cell carcinoma to radiotherapy. Int J Cancer 2010;126:1216-25. [PubMed]
  117. Noguchi M, Yu D, Hirayama R, et al. Inhibition of homologous recombination repair in irradiated tumor cells pretreated with Hsp90 inhibitor 17-allylamino-17-demethoxygeldanamycin. Biochem Biophys Res Commun 2006;351:658-63. [Crossref] [PubMed]
  118. Lee Y, Li HK, Masaoka A, et al. The purine scaffold Hsp90 inhibitor PU-H71 sensitizes cancer cells to heavy ion radiation by inhibiting DNA repair by homologous recombination and non-homologous end joining. Radiother Oncol 2016;121:162-8. [Crossref] [PubMed]
  119. Lee Y, Sunada S, Hirakawa H, et al. TAS-116, a novel Hsp90 inhibitor, selectively enhances radiosensitivity of human cancer cells to x-rays and Carbon ion radiation. Mol Cancer Ther 2017;16:16-24. [Crossref] [PubMed]
  120. Segawa T, Fujii Y, Tanaka A, et al. Radiosensitization of human lung cancer cells by the novel purine-scaffold Hsp90 inhibitor, PU-H71. Int J Mol Med 2014;33:559-64. [PubMed]
  121. Kamal A, Thao L, Sensintaffar J, et al. A high-affinity conformation of Hsp90 confers tumour selectivity on Hsp90 inhibitors. Nature 2003;425:407-10. [Crossref] [PubMed]
  122. Fujii Y, Kato T, Kubota N, et al. p53 Independent radio-sensitization of human lymphoblastoid cell lines by Hsp90 inhibitor 17-allylamino-17-demethoxygeldanamycin. Oncol Rep 2010;23:199-203. [PubMed]
  123. Hirakawa H, Fujisawa H, Masaoka A, et al. The combination of Hsp90 inhibitor 17AAG and heavy-ion irradiation provides effective tumor control in human lung cancer cells. Cancer Med 2015;4:426-36. [Crossref] [PubMed]
  124. Shrivastav M, Miller CA, De Haro LP, et al. DNA-PKcs and ATM co-regulate DNA double-strand break repair. DNA Repair (Amst) 2009;8:920-9. [Crossref] [PubMed]
  125. Sørensen CS, Hansen LT, Dziegielewski J, et al. The cell-cycle checkpoint kinase Chk1 is required for mammalian homologous recombination repair. Nat Cell Biol 2005;7:195-201. [Crossref] [PubMed]
  126. Allen CP, Tinganelli W, Sharma N, et al. DNA damage response proteins and oxygen modulate prostaglandin E2 growth factor release in response to low and high LET ionizing radiation. Front Oncol 2015;5:260. [Crossref] [PubMed]
  127. Ashley AK, Shrivastav M, Nie J, et al. DNA-PK phosphorylation of RPA32 Ser4/Ser8 regulates replication stress checkpoint activation, fork restart, homologous recombination and mitotic catastrophe. DNA Repair (Amst) 2014;21:131-9. [Crossref] [PubMed]
  128. Liu S, Opiyo SO, Manthey K, et al. Distinct roles for DNA-PK, ATM and ATR in RPA phosphorylation and checkpoint activation in response to replication stress. Nucleic Acids Res 2012;40:10780-94. [Crossref] [PubMed]
  129. Golding SE, Rosenberg E, Valerie N, et al. Improved ATM kinase inhibitor KU-60019 radiosensitizes glioma cells, compromises insulin, AKT and ERK prosurvival signaling, and inhibits migration and invasion. Mol Cancer Ther 2009;8:2894-902. [Crossref] [PubMed]
  130. Koniaras K, Cuddihy AR, Christopoulos H, et al. Inhibition of Chk1-dependent G2 DNA damage checkpoint radiosensitizes p53 mutant human cells. Oncogene 2001;20:7453. [Crossref] [PubMed]
  131. Syljuåsen RG, Sørensen CS, Nylandsted J, et al. Inhibition of Chk1 by CEP-3891 accelerates mitotic nuclear fragmentation in response to ionizing Radiation. Cancer Res 2004;64:9035-40. [Crossref] [PubMed]
  132. Jobson AG, Lountos GT, Lorenzi PL, et al. Cellular inhibition of checkpoint kinase 2 (Chk2) and potentiation of camptothecins and radiation by the novel Chk2 inhibitor PV1019 [7-nitro-1H-indole-2-carboxylic acid {4-[1-(guanidinohydrazone)-ethyl]-phenyl}-amide. J Pharmacol Exp Ther 2009;331:816-26. [Crossref] [PubMed]
  133. Fulop GM, Phillips RA. The scid mutation in mice causes a general defect in DNA repair. Nature 1990;347:479-82. [Crossref] [PubMed]
  134. Biedermann KA, Sun JR, Giaccia AJ, et al. scid mutation in mice confers hypersensitivity to ionizing radiation and a deficiency in DNA double-strand break repair. Proc Natl Acad Sci U S A 1991;88:1394-7. [Crossref] [PubMed]
  135. Biddlestone-Thorpe L, Sajjad M, Rosenberg E, et al. ATM kinase inhibition preferentially sensitizes p53-mutant glioma to ionizing radiation. Clin Cancer Res 2013;19:3189-200. [Crossref] [PubMed]
  136. Goldstein M, Kastan MB. The DNA damage response: implications for tumor responses to radiation and chemotherapy. Annu Rev Med 2015;66:129-43. [Crossref] [PubMed]
  137. Mukherjee B, Tomimatsu N, Amancherla K, et al. The dual PI3K/mTOR inhibitor NVP-BEZ235 is a potent inhibitor of ATM- and DNA-PKCs-mediated DNA damage responses. Neoplasia 2012;14:34-43. [Crossref] [PubMed]
  138. Liu P, Cheng H, Roberts TM, et al. Targeting the phosphoinositide 3-kinase pathway in cancer. Nat Rev Drug Discov 2009;8:627-44. [Crossref] [PubMed]
  139. Abraham RT. PI 3-kinase related kinases: 'big' players in stress-induced signaling pathways. DNA Repair (Amst) 2004;3:883-7. [Crossref] [PubMed]
  140. Toledo LI, Murga M, Zur R, et al. A cell-based screen identifies ATR inhibitors with synthetic lethal properties for cancer-associated mutations. Nat Struct Mol Biol 2011;18:721-7. [Crossref] [PubMed]
  141. Brown EJ, Baltimore D. Essential and dispensable roles of ATR in cell cycle arrest and genome maintenance. Genes Dev 2003;17:615-28. [Crossref] [PubMed]
  142. Brown EJ, Baltimore D. ATR disruption leads to chromosomal fragmentation and early embryonic lethality. Genes Dev 2000;14:397-402. [PubMed]
  143. Fazio N, Buzzoni R, Baudin E, et al. A phase II study of BEZ235 in patients with everolimus-resistant, advanced pancreatic neuroendocrine tumours. Anticancer Res 2016;36:713-9. [PubMed]
  144. Signore M, Pelacchi F, di Martino S, et al. Combined PDK1 and CHK1 inhibition is required to kill glioblastoma stem-like cells in vitro and in vivo. Cell Death Dis 2014;5:e1223 [Crossref] [PubMed]
  145. Engelke CG, Parsels LA, Qian Y, et al. Sensitization of pancreatic cancer to chemoradiation by the Chk1 inhibitor Mk8776. Clin Cancer Res 2013;19:4412-21. [Crossref] [PubMed]
  146. Takai H, Naka K, Okada Y, et al. Chk2-deficient mice exhibit radioresistance and defective p53-mediated transcription. EMBO J 2002;21:5195-205. [Crossref] [PubMed]
  147. Brown DM, Evans JW, Brown JM. The influence of inhibitors of poly (ADP-ribose) polymerase on X-ray-induced potentially lethal damage repair. Br J Cancer Suppl 1984;6:27-31. [PubMed]
  148. Ben-Hur E, Utsumi H, Elkind M. Inhibitors of poly (ADP-ribose) synthesis enhance radiation response by differentially affecting repair of potentially lethal versus sublethal damage. Br J Cancer Suppl 1984;6:39. [PubMed]
  149. Ben-Hur E, Utsumi H, Elkind MM. Inhibitors of poly(ADP-ribose) synthesis enhance X-ray killing of log-phase Chinese hamster cells. Radiat Res 1984;97:546-55. [Crossref] [PubMed]
  150. De Murcia JM, Niedergang C, Trucco C, et al. Requirement of poly(ADP-ribose) polymerase in recovery from DNA damage in mice and in cells. Proc Natl Acad Sci U S A 1997;94:7303-7. [Crossref] [PubMed]
  151. Kelland LR, Tonkin KS. The effect of 3-aminobenzamide in the radiation response of three human cervix carcinoma xenografts. Radiother Oncol 1989;15:363-9. [Crossref] [PubMed]
  152. Calabrese CR, Almassy R, Barton S, et al. Anticancer chemosensitization and radiosensitization by the novel poly(ADP-ribose) polymerase-1 inhibitor AG14361. J Natl Cancer Inst 2004;96:56-67. [Crossref] [PubMed]
  153. Albert JM, Cao C, Kim KW, et al. Inhibition of poly(ADP-ribose) polymerase enhances cell death and improves tumor growth delay in irradiated lung cancer models. Clin Cancer Res 2007;13:3033-42. [Crossref] [PubMed]
  154. Senra JM, Telfer BA, Cherry KE, et al. Inhibition of PARP-1 by olaparib (AZD2281) increases the radiosensitivity of a lung tumor xenograft. Mol Cancer Ther 2011;10:1949-58. [Crossref] [PubMed]
  155. Khan K, Araki K, Wang D, et al. Head and neck cancer radiosensitization by the novel poly(ADP-ribose) polymerase inhibitor GPI-15427. Head Neck 2010;32:381-91. [PubMed]
  156. Russo AL, Kwon HC, Burgan WE, et al. In vitro and in vivo radiosensitization of glioblastoma cells by the poly (ADP-ribose) polymerase inhibitor E7016. Clin Cancer Res 2009;15:607-12. [Crossref] [PubMed]
  157. Donawho CK, Luo Y, Luo Y, et al. ABT-888, an orally active poly(ADP-ribose) polymerase inhibitor that potentiates DNA-damaging agents in preclinical tumor models. Clin Cancer Res 2007;13:2728-37. [Crossref] [PubMed]
  158. Bisht KS, Bradbury CM, Mattson D, et al. Geldanamycin and 17-allylamino-17-demethoxygeldanamycin potentiate the in vitro and in vivo radiation response of cervical tumor cells via the heat shock protein 90-mediated intracellular signaling and cytotoxicity. Cancer Res 2003;63:8984-95. [PubMed]
  159. Bull EE, Dote H, Brady KJ, et al. Enhanced tumor cell radiosensitivity and abrogation of G2 and S phase arrest by the Hsp90 inhibitor 17-(dimethylaminoethylamino)-17-demethoxygeldanamycin. Clin Cancer Res 2004;10:8077-84. [Crossref] [PubMed]
  160. Ohkubo S, Kodama Y, Muraoka H, et al. TAS-116, a highly selective inhibitor of heat shock protein 90α and β, demonstrates potent antitumor activity and minimal ocular toxicity in preclinical models. Mol Cancer Ther 2015;14:14-22. [Crossref] [PubMed]
  161. Stiedl P, Grabner B, Zboray K, et al. Modeling cancer using genetically engineered mice. Methods Mol Biol 2015;1267:3-18. [Crossref] [PubMed]
  162. Le Magnen C, Dutta A, Abate-Shen C. Optimizing mouse models for precision cancer prevention. Nat Rev Cancer 2016;16:187-96. [Crossref] [PubMed]
  163. Jeong JH. Inducible Mouse Models for Cancer Drug Target Validation. J Cancer Prev 2016;21:243-8. [Crossref] [PubMed]
  164. Fox EJ, Salk JJ, Loeb LA. Cancer genome sequencing--an interim analysis. Cancer Res 2009;69:4948-50. [Crossref] [PubMed]
  165. Navin N, Kendall J, Troge J, et al. Tumour evolution inferred by single-cell sequencing. Nature 2011;472:90-4. [Crossref] [PubMed]
  166. Vandamme TF. Use of rodents as models of human diseases. J Pharm Bioallied Sci 2014;6:2-9. [Crossref] [PubMed]
  167. Nass SJ, Gorby H, Nass SJ, et al. The role of clinical studies for pets with naturally occurring tumors in translational cancer research: workshop summary. Washington, DC: National Academies Press, 2015. Available online: http://www.nap.edu/21830
  168. DiRuggiero J, Robb FT. Early evolution of DNA repair mechanisms. In: Pouplana LR, editor. The Geneitic Code and the Origin of Life. 1st ed. New York: Kluwer Academic/Plenum, 2017:(169-82).
  169. Eberlein R. Ein versuch mit Roentgen'schen strahlen. Monatsheftu praktischa Tierhailkunda 1896:7.
  170. Eberlein R. Rontgentherapie bei Haustieren. Verh Beirichte IIRontgenkongress; Hamburg, Germany 1906.
  171. Grubbe EH. Priority in the therapeutic use of X-rays. Radiology 1933;21:156-62. [Crossref]
  172. Gillette EL. History of veterinary radiation oncology. Vet Clin North Am Small Anim Pract 1997;27:1-6. [Crossref] [PubMed]
  173. Pommer A. X-ray therapy in veterinary medicine. In: Brandly CA, Jungher EL, editors. Advances in Veterinary Science. New York: Academic Press, 1958:(98-136).
  174. Coutard H. Roentgen therapy of epitheliomas of the tonsillar region, hypopharynxlarynx from 1920 to 1926. Am J Roentgenol 1932;28:313-31.
  175. Pommer A, Maiolino A. Roentgen epilation and erythema doses of the skin in the dog. North Am Vet 1937;18:39-58.
  176. Elkind MM, Sutton H. Radiation response of mammalian cells grown in culture. 1. Repair of X-ray damage in surviving Chinese hamster cells. Radiat Res 1960;13:556-93. [Crossref] [PubMed]
  177. Withers HR. The four R's of radiotherapy. In: Lett J, Adler H, editors. Advances in Radiation Biology. New York: Academic Press, 1975.
  178. Vail DM, Macewen EG. Spontaneously occurring tumors of companion animals as models for human cancer. Cancer Invest 2000;18:781-92. [Crossref] [PubMed]
  179. Khanna C, Lindblad-Toh K, Vail D, et al. The dog as a cancer model. Nat Biotechnol 2006;24:1065-6. [Crossref] [PubMed]
  180. Khanna C, London C, Vail D, et al. Guiding the optimal translation of new cancer treatments from canine to human cancer patients. Clin Cancer Res 2009;15:5671-7. [Crossref] [PubMed]
  181. Gillette EL, Withers HR, Tannock IF. The age sensitivity of epithelial cells of mouse small intestine. Radiology 1970;96:639-43. [Crossref] [PubMed]
  182. Hoopes PJ, Gillette EL, Withrow SJ. Intraoperative irradiation of the canine abdominal aorta and vena cava. Int J Radiat Oncol Biol Phys 1987;13:715-22. [Crossref] [PubMed]
  183. Gillette EL, Powers BE, Mcchesney SL, et al. Response of aorta and branch arteries to experimental intraoperative irradiation. Int J Radiat Oncol Biol Phys 1989;17:1247-55. [Crossref] [PubMed]
  184. Withrow SJ, Gillette EL, Hoopes PJ, et al. Intraoperative irradiation of 16 spontaneously occurring canine neoplasms. Vet Surg 1989;18:7-11. [Crossref] [PubMed]
  185. Gillette SL, Gillette EL, Powers BE, et al. Ureteral injury following experimental intraoperative radiation. Int J Radiat Oncol Biol Phys 1989;17:791-8. [Crossref] [PubMed]
  186. Gillette SM, Powers BE, Orton EC, et al. Early radiation response of the canine heart and lung. Radiat Res 1991;125:34-40. [Crossref] [PubMed]
  187. Gillette SM, Gillette EL, Shida T, et al. Late radiation response of canine mediastinal tissues. Radiother Oncol 1992;23:41-52. [Crossref] [PubMed]
  188. Gillette SM, Poulson JM, Deschesne KM, et al. Response of the canine esophagus to irradiation. Radiat Res 1998;150:365-8. [Crossref] [PubMed]
  189. Powers BE, Mcchesney SL, Gillette EL. Late radiation response of the canine trachea with change in dose per fraction. Int J Radiat Oncol Biol Phys 1987;13:1673-80. [Crossref] [PubMed]
  190. Powers BE, Thames HD, Gillette SM, et al. Volume effects in the irradiated canine spinal cord: do they exist when the probability of injury is low? Radiother Oncol 1998;46:297-306. [Crossref] [PubMed]
  191. Powers BE, Gillette EL, Mcchesney SL, et al. Bone necrosis and tumor induction following experimental intraoperative irradiation. Int J Radiat Oncol Biol Phys 1989;17:559-67. [Crossref] [PubMed]
  192. Gillette EL. Clinical use of thermal enhancement and therapeutic gain for hyperthermia combined with radiation or drugs. Cancer Res 1984;44:4836s-41s. [PubMed]
  193. Gillette EL, Mcchesney SL, Dewhirst MW, et al. Response of canine oral carcinomas to heat and radiation. Int J Radiat Oncol Biol Phys 1987;13:1861-7. [Crossref] [PubMed]
  194. Gentilini F, Turba ME, Forni M, et al. Complete sequencing of full-length canine ataxia telangiectasia mutated mRNA and characterization of its putative promoter. Vet Immunol Immunopathol 2009;128:437-40. [Crossref] [PubMed]
  195. Uva P, Aurisicchio L, Watters J, et al. Comparative expression pathway analysis of human and canine mammary tumors. BMC genomics 2009;10:135. [Crossref] [PubMed]
  196. Raposo-Ferreira TM, Bueno RC, Terra EM, et al. Downregulation of ATM gene and protein expression in canine mammary tumors. Vet Pathol 2016;53:1154-9. [Crossref] [PubMed]
  197. Koike M, Yutoku Y, Koike A. Cloning, localization and focus formation at DNA damage sites of canine Ku70. J Vet Med Sci 2017;79:554-61. [Crossref] [PubMed]
  198. Koike M, Yutoku Y, Koike A. Cloning, localization and focus formation at DNA damage sites of canine XLF. J Vet Med Sci 2017;79:22-8. [Crossref] [PubMed]
  199. Maeda J, Froning CE, Brents CA, et al. Intrinsic radiosensitivity and cellular characterization of 27 canine cancer cell lines. PLoS ONE 2016;11:e0156689 [Crossref] [PubMed]
  200. Gil Del Alcazar CR, Hardebeck MC, Mukherjee B, et al. Inhibition of DNA double-strand break repair by the dual PI3K/mTOR inhibitor NVP-BEZ235 as a strategy for radiosensitization of glioblastoma. Clin Cancer Res 2014;20:1235-48. [Crossref] [PubMed]
  201. Reynders K, Illidge T, Siva S, et al. The abscopal effect of local radiotherapy: using immunotherapy to make a rare event clinically relevant. Cancer Treat Rev 2015;41:503-10. [Crossref] [PubMed]
  202. Siva S, MacManus MP, Martin RF, et al. Abscopal effects of radiation therapy: a clinical review for the radiobiologist. Cancer Lett 2015;356:82-90. [Crossref] [PubMed]
  203. Bailly C, Clery PF, Faivre-Chauvet A, et al. Immuno-PET for clinical theranostic approaches. Int J Mol Sci 2016;18:57. [Crossref] [PubMed]
  204. Goel S, England CG, Chen F, et al. Positron emission tomography and nanotechnology: A dynamic duo for cancer theranostics. Adv Drug Deliv Rev 2016; [Epub ahead of print]. [Crossref] [PubMed]
Cite this article as: Nickoloff JA, Boss MK, Allen CP, LaRue SM. Translational research in radiation-induced DNA damage signaling and repair. Transl Cancer Res 2017;6(Suppl 5):S875-S891. doi: 10.21037/tcr.2017.06.02

Download Citation