Skip to main content
Wiley Open Access Collection logoLink to Wiley Open Access Collection
. 2022 Jan 10;20(2):347–361. doi: 10.1111/vco.12795

DNA damage response proteins in canine cancer as potential research targets in comparative oncology

Beatriz Hernández‐Suárez 1,, David A Gillespie 2, Aleksandra Pawlak 1
PMCID: PMC9304296  PMID: 34923737

Abstract

The DNA damage response (DDR) is a complex signal transduction network that is activated when endogenous or exogenous genotoxins damage or interfere with the replication of genomic DNA. Under such conditions, the DDR promotes DNA repair and ensures accurate replication and division of the genome. High levels of genomic instability are frequently observed in cancers and can stem from germline loss‐of‐function mutations in certain DDR genes, such as BRCA1, BRCA2, and p53, that form the basis of human cancer predisposition syndromes. In addition, mutation and/or aberrant expression of multiple DDR genes are frequently observed in sporadic human cancers. As a result, the DDR is considered to represent a viable target for cancer therapy in humans and a variety of strategies are under investigation. Cancer is also a significant cause of mortality in dogs, a species that offers certain advantages for experimental oncology. Domestic dogs present numerous inbred lines, many of which display predisposition to specific forms of cancer and the study of which may provide insight into the biological basis of this susceptibility. In addition, clinical trials are possible in dogs and may lead to therapeutic insights that could ultimately be extended to humans. Here we review what is known specifically about the DDR in dogs and discuss how this knowledge could be extended and exploited to advance experimental oncology in this species.

Keywords: BRCA1, BRCA2, canine cancer, p53, Rad51, TopBP1


Abbreviations

53BP1

p53‐binding protein 1

9‐1‐1

Rad9‐Rad1‐Hus1

ATM

autosomal‐recessive ataxia‐telangiectasia mutated

ATR

autosomal‐recessive ataxia‐telangiectasia mutated and Rad3‐related

ATRIP

ATR interacting protein

BACH1

BRCA1‐associated C‐terminal helicase

BCRT

BRCA1 carboxyl terminal

BRC

BRCA2 repeats

BRCA1

breast cancer 1

BRCA2

breast cancer 2

CC3

cleaved caspase 3

CDKs

cyclin‐dependent kinases

cDNA

complementary DNA

Chk1

check point kinase 1

Chk2

check point kinase 2

CML

chronic myelogenous leukaemia

CMT

canine mammary tumour

DBD

DNA binding domain in p53

DNA

deoxyribonucleic acid

DNA‐PKcs

DNA dependent protein kinases

DDR

DNA damage response

EDD

E3 identified by differential display

GST

specific glutathione‐S‐transferase

γH2AX

phosphorylated‐gamma histone H2AX

H2AX

histone H2AX

HCC

hepatocellular carcinoma

HR

homologous recombination

IR

irradiation

MDC1

mediator of DNA damage protein 1

MDM2

murine double minute 2

MMR

mismatch repair

MRN

Mre11‐Rad50‐NBS1

MVC

minute virus of canine

NHEJ

non‐homologous end joining

NOS

nitrogen species

p21

cyclin‐dependent kinase inhibitor p21

p27

cyclin‐dependent kinase inhibitor 1B

p53

tumour protein p53

PALB2

partner and localizer Of BRCA2

PCNA

proliferating cell nuclear antigen

PIKK

phosphatidyl inositol 3′ kinase‐related kinases

PTEN

phosphatidylinositol 3,4,5‐trisphosphate 3‐phosphatase

Rad51

Rad51 recombinase

RAD9

Rad9 checkpoint clamp component of 9‐1‐1 clamp

ROS

reactive oxygen species

RPA

replication protein A

SNP

single nucleotide polymorphisms

SSPC

single‐strand conformational polymorphism

TopBP1

topoisomerase IIβ binding protein 1

tp53

gen of tumour protein p53

UV

ultraviolet light

1. INTRODUCTION

1.1. DNA damage response and its role in cancer

The DNA damage response (DDR) is a signalling and effector pathway designed to ensure the genetic stability of eukaryotic cells. It is a complex, multi‐faceted process that is activated in response to DNA damage inflicted by both endogenous sources, such as reactive oxygen species (ROS), reactive nitrogen species (NOS), or exogenous sources, such as ultraviolet light (UV) or irradiation (IR). 1 A variety of different mechanisms can contribute to genomic instability. 2 One is enzymatic deamination, which converts 5‐methylcytidine to a thymidine residue through removal of the amine group. Replication of the resulting T/ G mismatch is mutagenic, and breast and other cancers have been found to exhibit mutation or aberrant expression of cytosine deaminases of the AID/ APOBEC families. 3 Another mechanism is microsatellite instability, in which expansions of repeated sequences of DNA occur during replication. Microsatellite instability is linked to mutations in the mismatch repair (MMR) system 4 and are found in hereditable cancers such as Lynch syndrome. 5 Macroscopic chromosome translocations, an abnormal rearrangement of one chromosome segment to another, are frequently observed in lymphoma. 6 Sporadic DNA damage lesions can also arise as a result of errors during DNA replication. 7 Somatically acquired mutations affecting DNA repair and genome stability genes are frequently observed in cancers, and it is considered that they play an important role in the development of the corresponding sporadic cancers. 8 , 9

After a DNA damage lesion occurs, a cascade of highly coordinated signalling events begins which activate cell cycle checkpoints, promote DNA repair, and eliminate cells with irreparable lesions via programmed cell death or apoptosis. 10 This cascade (Figure 1) starts with the recognition of damage by the Mre11‐Rad50‐NBS1 (MRN) complex 11 and Rad9‐Rad1‐Hus1 (9‐1‐1) complexes, which act as damage sensors and help in the recruitment of additional DDR proteins to the damage site. These sensor molecules send signals to recruit the phosphatidyl inositol 3′ kinase‐related kinases (PIKK), ataxia‐telangiectasia mutated (ATM), ataxia‐telangiectasia mutated and Rad3‐related (ATR) and/or DNA dependent protein kinases (DNA‐PKcs), which then phosphorylate and activate other transducers and effector proteins. One important target of ATM kinase is histone H2AX, which in its phosphorylated form (γH2AX) serves as a platform for the recruitment of DDR proteins at the damage site. 12 As a result, γH2AX is a well‐known DNA damage marker widely used in DDR research that can be used to monitor and quantify damage lesions and also to detect DNA fragmentation arising from programmed cell death or apoptosis. 13

FIGURE 1.

FIGURE 1

DDR signalling cascade. When DNA damage occurs, the cascade of DDR proteins starts. First, sensor proteins (MRN and 9‐1‐1 complexes) recognize the damage and RPA binds the single stranded DNA (ssDNA). ATM/ATR kinases are recruited to DNA damage site by interacting with the sensor proteins. Those kinases immediately phosphorylate histone H2AX at serine 139 (γH2AX), which helps to attract repair factors. Then, the mediator proteins (BRCA1, MDC1 and TopBP1) stabilize the protein interactions and increase the damage signalling. First, the transducer kinases ATM and ATR activate the effector kinases, Chk1/Chk2, by phosphorylation, which by activation of p53 or inhibition of CDKs induce cell cycle arrest or apoptosis if the damage is irreparable. The mediators BRCA1 and 53BP1 compete to promote different pathways: BRCA1 promotes HR repair pathways, while 53BP1 favours NHEJ pathway.Note: The principal kinases of the DDR are coloured blue. Other DDR proteins discussed at length in this review are depicted in green (BRCA1, BRCA2) or orange (Rad51, TopBP1), while functional components not discussed in detail are grey. Black arrows symbolize activation, and red lines symbolize inhibition

During this initial step other important proteins such as p53‐binding protein 1 (53BP1), Mediator of DNA Damage Checkpoint protein 1 (MDC1), Topoisomerase IIβ binding protein 1 (TopBP1) and Breast Cancer 1 (BRCA1) amplify the chromatin modifications and help in activation of the checkpoint effector kinases: Checkpoint kinase 1 (Chk1) and Checkpoint kinase 2 (Chk2). These kinases control multiple cell cycle checkpoints and promote DNA repair by modulating the activity of various effectors, such as cyclin‐dependent kinases (CDKs), tumour suppressor protein 53 (p53), and Rad51. 14 Chk1 and Chk2 function to induce cell cycle arrest, DNA repair, chromatin assembly, transcriptional and posttranscriptional regulation of gene expression and cell death through apoptosis. 15 , 16 , 17 , 18 , 19 , 20 , 21 Other relevant proteins that participate in this cascade are Breast Cancer Associated 2 (BRCA2), whose principal role is to promote DNA repair by homologous recombination (HR), 22 and cyclin/CDK complexes and CDK inhibitors, such as p21 CIP1, whose role is to arrest the damaged cell at different points of the cell cycle. 23 , 24 Collectively, the DDR/checkpoint system acts to prevent cell division with damaged or partially replicated DNA and to promote accurate repair to enable the damaged cell to survive without permanent genetic damage or mutation. Evidence suggests that these processes play an important role in preventing neoplastic transformation and cell death under conditions of genomic or oncogenic stress. 25 A simplified diagram of the DDR signalling cascade is shown in Figure 1.

1.2. DNA damage response in canine cancers

The DDR system is highly conserved and its organization and functional components are very similar in all mammalian species. 20 , 21 The similarity between dogs and humans in this context is special because certain important proteins respond in the same way to DNA damage compared to other species. 26 One example is shown in a study where the expression of p53 protein was monitored at various times after DNA damage in different mammalian species. 27 The authors generated human, mouse, dog, monkey, and rat kidney cell lines expressing a fluorescent p53. After irradiation, they observed p53 protein levels to increase and decrease in an oscillatory pattern over a period of 10 h in all species. The oscillatory profile of response of p53 protein was however found to be much slower and more similar between dogs and humans than in rodents. 27

The literature lacks extensive information on the course of DDR in dogs and details on the role of its individual components. A number of papers describe selected proteins of the DDR pathway in dogs, however there are no studies that document the overall DNA repair process. In addition, although some studies have focused on the description of selected DDR proteins in relation to mutations, changes in the expression level, and their use in comparative oncology, in general they do not consider how knowledge of the canine DDR system could be used to develop new anticancer therapies. 28 , 29 One study on canine cell lines infected with a canine minute virus (MVC) 30 provides molecular insights into how the DDR functions in dogs. In the case of the DDR induced by MCV, ATM is known to be necessary to induce the G2/M arrest signal. 30 However, the proximal effector kinase(s) that mediate the cell cycle arrest remains unclear. Another recent study examining MMR deficiency in canine tumours, has shown that oral malignant melanoma and hepatocellular carcinoma (HCC) lack expression of one or more MMR proteins more frequently than other cancer types in dogs. 31 An interesting study showed differential rates of DNA repair rate in embryonic fibroblasts from humans, dogs, mice and rats. 32 As predicted, the DNA repair rate increased in the species with longer life span, suggesting a positive relationship between DNA repair proficiency and longevity. Non‐Homologous End Joining (NHEJ) activity in humans and dogs was the highest both in vitro and in vivo, as longer life span may require a more functional DDR to allow cells extended replicative potential. 32

Some studies investigating DDR proteins in dogs show important similarities in function and expression with their human homologues. A comparison in Table 1 clearly shows that many of the most important proteins involved in the DDR, such as TopBP1, p53, Rad51, BRCA1, BRCA2, Chk1a, Chk2a, PTEN, PCNA, p21, p27 and Cyclin A, are highly conserved between humans and dogs, indicating that the dogs can be research models for the investigation of DDR phenomena. Combining these discoveries with the fact that humans and dogs have the highest DNA repair rates, emphasizes again the potential of using of dogs as a research model to uncover biological insights that could be translated to humans.

TABLE 1.

Summary of similarities and differences in DDR proteins between humans and dogs

Protein Dogs Humans
TopBP1
  • Similar function in both species

  • Overexpressed in malignancy

  • Cytoplasmic localization in malignancy 33 , 34 , 35

  • Inhibition of p53 pathways when TopBP1 is overexpressed 36

p53
  • 78.4% shared protein identity

  • Similar function in both species

  • Role in malignant transformation when mutated

  • Highly conserved DNA binding region and C‐ and N‐termini

  • Alteration present in different types of cancers

  • Low‐frequency inherited mutations in cancer 37 , 38 , 39 , 40 , 41 , 42 , 43 , 44 , 45

Rad51
  • 99% gene homology

  • Interaction with BRCA2 and BRCT domains

  • Overexpressed in cancer

  • Presence of polymorphisms in tumour 24 , 46 , 47 , 48 , 49

  • Interaction with BRCA2 C‐terminal domain 50

  • Validated primers 48 , 49

BRCA1
  • 84% of gene identity

  • BCR conserved regions share 77% homology

  • Mutations increased risk of mammary tumour

  • Increased malignancy when expression decrease 46 , 51 , 52 , 53

BRCA2
  • 68% of protein homology

  • Mutations increased risk of mammary tumour

  • Increased malignancy when expression increased 46 , 48 , 52 , 54 , 55

  • Validated primers 46

  • Its loss triggers p53 mutations 56

Chk1 a
  • Inhibition suppresses proliferation 57

Chk2 a
  • Mutation in cancer 58

PTEN
  • Lack of or reduced expression correlated with malignancy 29 , 59

PCNA
  • Increased in proliferating tumours 60

p21
p27
Cyclin A
  • Overexpression correlated with carcinogenesis 23

a

These proteins have not been deeply studied in dogs. For TopBP1, p53, Rad51, BRCA1, and BRCA2, studies that have validated antibody reagents for protein detection, or PCR primers for mRNA quantification, with appropriate positive and negative controls are indicated.

An additional advantage of the dog model is the fact that in both species, humans and dogs, development of tumours is spontaneous, showing higher incidence with age. Heterogeneous course of disease in different patients with an analogous metastatic behaviour, comparable response to antineoplastic therapies, and alteration in DDR pathway are further similarities. 28 , 63 , 64 Additionally, over 360 genetic disorders related to cancer were described in dogs, and this constitutes the largest set of natural genetic disorders in a non‐human species. 65 These findings clearly justify the need for close cooperation between veterinary and human oncologists in the field of genetic instability and particularly DDR disorders. The increasing availability of a wide panel of various canine cancer cell lines hugely facilitates such studies. 66 , 67 , 68 , 69

Due to significance of DDR disorders in the cancer development in dogs, this review discusses the key components of the DDR, presents their physiological role and the abnormalities associated with these proteins found in canine cancers. We consider findings on the potential causes of genetic instability in canine cancers and indicate potential new research directions. These directions may contribute to the development of new anti‐cancer therapies in veterinary medicine, and in view of the potential role of the dog as a model for the study of human diseases, also in human medicine.

2. SELECTED DDR PROTEINS AND THEIR POTENTIAL ROLE IN CANINE RESEARCH

2.1. BRCA1 and BRCA2

Inherited germ‐line mutations affecting the BRCA1 and BRCA2 genes are associated with the development of familial cancer in women. BRCA1 and BRCA2 are tumour suppressors discovered in the early 1990s as genes conferring breast cancer susceptibility. 70 , 71 Their principal role is to maintain genomic stability by promoting the error‐free repair of DNA double strand breaks (DSBs) by HR. 22 Although BRCA1 and BRCA2 were initially described as breast cancer predisposition genes, and germ‐line mutations in either gene markedly increase the risk of this type of cancer in affected individuals, it is known that inherited or sporadic mutations in BRCA1 and BRCA2 can also contribute to the development of other cancers, for example, ovarian cancer, pancreatic cancer, melanoma, or even prostate and breast cancer in men. 72 BRCA1 and BRCA2 were also reported to play a role in some other genetic syndromes such as Fanconi anaemia. 73

The important role of these proteins in genomic stability is related to their influence on DNA repair, replication, and transcription. 56 BRCA1 is a multi‐functional protein that plays multiple regulatory roles through interactions with other important components involved in each of these processes. Thus, BRCA1 interacts with different DDR proteins such as TopBP1 33 , 74 or the MRN complex to impede its binding to DNA, it co‐localizes with γH2AX after DNA damage, controls the activity of c‐Abl (a protein kinase involved in apoptosis), and is hyperphosphorylated by DDR kinases in response to DNA damage. 22 , 56 Interestingly, it was shown to have a role in G2/M checkpoint by activating Chk1. 75 BRCA1 possess a highly conserved RING domain in its structure, 53 through which it gains its E3 ubiquitin ligase activity. 76 By means of its C‐terminus, BRCA1 is capable of binding some elements of the transcriptional machinery (RNA polymerase II or helicase A), which explains its role in transcription control. 56

BRCA2 is a large protein that contains eight clustered BRC motifs that are highly conserved in mammalian species. In fact, the BRC motifs constitute the regions of BRCA2 that are the most highly conserved between humans and dogs, with 70% identity. 77 In addition, sequence analysis of germline BRCA2 mutations associated with familial breast and ovarian cancer showed that these occur most frequently in exon 11 (where the BRC motifs are located). 78

BRCA1 has two roles in HR repair (Figure 2). First, it promotes end resection at the damage site by inhibiting the activity of 53BP1 and consequently the NHEJ pathway. 74 , 79 Second, it promotes Rad51 to replace RPA bound to ssDNA generated by strand resection. To do this, BRCA1 acts in conjunction with BRCA2, which binds Rad51 and promotes its translocation from the cytoplasm to the nucleus. 80 BRCA1 forms a complex with PALB2 that acts as a link between BRCA1 and BRCA2, forming the BRCA1‐PALB2‐BRCA2‐Rad51 foci at the sites of DNA damage that are needed to start HR repair. 79

FIGURE 2.

FIGURE 2

The role of BRCA1/2 in DNA repair via homologous recombination. After a DNA lesion occurs, sensor proteins attract 53BP1 to start the NHEJ. However, when BRCA1 is recruited to the DNA lesion sites, it inhibits 53BP1‐TopBP1 interactions and promotes HR pathway instead. BCRA1 helps in the recruitment of Rad51. Then Rad51 forms a complex with PALB2 and BRCA2, and the complex is relocated from the cytoplasm to the nucleus. There, Rad51 binds DNA lesion site by interacting with RPA and replacing it. Finally, Rad51 forms nucleoprotein filaments on ssDNA that create a more stable platform for the repair machinery

BRCA2's primary role in HR is through its interaction with Rad51 (more information about this interaction is provided in the section on Rad51), to regulate the subcellular localization of Rad51 (Figure 2). 46 , 50 , 80 BRCA2 binds Rad51 from its BRC repeats in exon 11 and from its COOH‐terminal domain encoded by exon 27 (both exons are highly conserved regions). The importance of the BRCA2‐Rad51 interaction for genomic stability was supported by a study of mutant mice bearing a deletion of BRCA2 in exon 27, which exhibited a high incidence of spontaneous cancer. 81 HR also seems to be controlled by p53 interaction with BRCA2, where p53 binds to the BRCA2‐Rad51 complex to repress HR. It is not clear yet if this suppression happens because p53 disrupts formation of the BRCA2‐Rad51 complex or because p53 prevents binding of the complex to DNA. 82 It is known that BRCA2 deficiency results in chromosome instability, which is related to the poorly understood role of BRCA2 in stabilization of stalled replication forks. 83

Studies in which BRCA genes were depleted or produced defective proteins, identified the most important roles of these proteins in DNA repair system. The loss of BRCA1 not only results in defective repair, but it also boosts apoptosis, genetic instability, and tumorigenesis. 84 BRCA2 deficiency results in elevated levels of chromosome breaks, probably due to its role in stabilization of stalled replication forks, but more interesting is that its loss triggers tp53 mutations, which can contribute to cancer progression. 56

The canine BRCA1 gene was described for the first time in a comparative study in 1996, 53 when BRCA1 genes from human, mouse and dog were sequenced and compared. This revealed an 84% identity of the canine gene with its human homologue. 53 Subsequent sequencing of the canine BRCA2 gene showed a greater homology between humans and dogs than between humans and mice, with a 76% cDNA sequence homology and 68% BRCA2 protein homology. 48 Due to the important role of BRCA1 and BRCA2 in breast cancer in humans, most of the studies in dogs focused on mammary tumours. A study on gene expression showed that decreased BRCA2 levels are related to tumour development. 85 These experiments identified different splicing variants of BRCA2, one of which could contribute to a reduction in BRCA2 protein levels. This particular splice variant encodes a form of BRCA2 that is unable to interact with a stabilizer protein, DSS1 (deleted in split hand/split foot). Thus, this transcript encodes an unstable and dysfunctional isoform of BRCA2 protein. 85 Further studies of single nucleotide polymorphisms (SNPs) in canine BRCA genes have shown that variants of both BRCA1 and BRCA2 genes can be associated with canine mammary tumour (CMT) development as occurs in humans, whilst other SNPs appear only in tumour samples and never in control normal tissue. 52 , 86 , 87

The canine homologues of BRCA genes are also linked to spontaneous cancer and metastases. 88 Nieto and colleagues 51 analysed BRCA1 expression levels in mammary dysplasia and tumours. They showed that in normal mammary gland the expression of BRCA1 was exclusively nuclear, whereas in a neoplastic gland BRCA1 protein was also observed in the cytoplasm. The relationship between BRCA1 expression and malignancy was also analysed and, as expected, malignancy increased when the levels of BRCA1 decreased. 51 In contrast, another study demonstrated that although BRCA1 levels did not correlate with malignancy, overexpression of RAD51 and BRCA2 was observed in lymph node metastases. 46 The association of high BRCA2‐Rad51 expression with malignancy likely relates to their function in DNA repair by HR.

The clinical and molecular similarities existing between the mammary tumours in canines and humans, makes the dog a useful model to study mammary cancer. 89 The observed similarities also apply to the relationship between mammary tumours and DNA damage because mammary tumour development in dogs and women has been associated with deregulation of BRCA1/2 gene function. 85 , 90 , 91 As dogs have a long history of inbreeding with low levels of genetic variation, it has been suggested that canine mammary tumour (CMT) development in a single breed should have a more defined homogenous origin compared with human, which could permit identification of breed specific CMT risk factors. 92 Therefore, the breed predisposition to CMT has been used to study BRCA2 mutations. 80 , 93 One study conducted in Sweden regarding the English Springer Spaniel breed, which has the highest incidence of CMT in that country, showed that mutations in BRCA1 or BRCA2 increased the risk of the cancer development fourfold. In addition, whereas the frequency of BRCA1 mutations rate seemed to be higher in malignant samples, the frequency of BRCA2 mutations was similar in malignant and benign tumour samples. 52 However, this study did not assess the functional effects of BRCA1/2 gene mutations and further investigations are required to understand the molecular mechanisms responsible for the observed phenomenon in this breed. When considering the aetiology and pathogenesis of CMT, or even genetic/breed predispositions, it must not be forgotten that steroid hormones (oestrogen and progesterone) and their receptors play significant roles in mammary tumour development. 94 As with early pregnancy/oophorectomy in women, early spaying in dogs is linked with lower disease incidence, 95 therefore, any research describing the incidence of CMT must be considered in relation to the current ovariohysterectomy trends in a given country.

2.2. p53

p53 was discovered in 1979 through studies of SV40 antigen T oncoproteins and is widely known as the “guardian of the genome” due to its function in maintaining genomic integrity. 96 , 97 In its different roles the protein: (a) Acts as a transcription factor for different genes to regulate the cell cycle (Figure 3); (b) Triggers apoptosis in response to oncogene activation or cell proliferation under stress conditions; (c) Silences repetitive sequences of DNA that could be cancer promoters; and (d) Helps in prevention of bacterial and viral infections acting as a part of the innate immune system. 14 It is also involved in anti‐angiogenesis and autophagy. 98 , 99 , 100 , 101

FIGURE 3.

FIGURE 3

Function of p53 in response to DNA damage. After DNA damage p53 is phosphorylated and consequently activated. p53 triggers the cell cycle arrest by inducing expression of p21, which binds and inhibits the activity of cyclin/CDK complexes, such as cyclin E/CDK2 and cyclin A/CDK2 that are necessary for the transition from G1 into S‐phase of the cell cycle. p53 can also trigger apoptosis by inducing the expression of other target proteins such as Puma, Noxa, Bax, and Bak, which are pro‐apoptotic proteins. 101 A major regulator of p53 activity is a ubiquitin ligase MDM2, which ubiquitinates and targets p53 for rapid degradation via the proteasome in the absence of cellular stress

The activity of p53 is regulated by multiple post‐translational modifications. Site‐specific phosphorylation catalysed by multiple protein kinases, including ATM and Chk2, can modulate p53 stability and gene‐regulatory functions in diverse ways, depending on the site of modification. p53 is degraded by proteasomes after being ubiquitinated by MDM2, a ubiquitin ligase. 102 , 103

After more than 40 years of research on p53, its importance in cancer is clear and a loss of wild type tp53 expression or tumour suppressor function is frequently linked to malignant transformation. 41 In addition, when tp53 is mutated, it can gain oncogenic functions and facilitate cell transformation and tumorigenesis. 38 “Gain of function” mutations can alter p53 molecular functions in complex ways: (a) Through a change of protein function by disruption of protein–protein interactions (e.g., disruption of ATM‐MRN complex formation by capturing MRN through interaction with MRE11); (b) By increasing or inhibiting the activity of certain transcription factors (e.g., mutated p53 can bind to p53‐related p63/p73 proteins, forming aggregates and preventing their activation in response to DNA damage and consequently suppressing p63/p73‐dependent transcription of apoptotic or growth‐inhibitory genes such as Bax or p21); (c) Changes in binding properties to other DNA regions not related to gene expression (e.g., matrix elements that participate in chromatin remodelling). 104 Aberrations of p53 expression or function appear in more than 50% of human cancers, and 90% of p53 mutations occur in so called hot‐spots, often in the DNA binding domain encoded by exons 5–8. 37

In veterinary medicine tp53 is an important research topic, and as in humans, it has a role in malignant transformation when mutated. 105 Veldhoen and Milner isolated the canine version of p53 in 1998 and presented its homology analysis with other mammalian species. Canine p53 protein shares 86.3% homology with feline, 72.3% with murine, and 78.4% with human protein. The central core of the protein, the DNA binding region, together with regions at the C‐ and N‐termini are highly conserved between species 106 and these are also the regions prone to mutations, in both humans and dogs. 39

Several studies have documented altered p53 expression in different types of veterinary cancers. Takeda et al. analysed mammary tumours, squamous carcinomas and basal cell tumours in dogs and cats documenting p53 protein expression levels by immunohistochemistry. 23 , 107 Detectable expression of p53 protein by this method was found in 24.6% and 16.3% of mammary tissue samples in dogs and cats respectively, with a higher percentage of expression in the malignant versus benign samples, indicating that p53 overexpression can be related to malignancy. In the case of squamous carcinoma samples, the percentage of cells that expressed p53 was high, 37.5% in dogs and 40% in cats, suggesting an association between high levels of the protein and tumorigenesis. 107 However, no p53 overexpression was detected in the basal cell tumour samples. 23 , 107

Another study analysed 170 samples of CMT which were classified into three different groups depending on the malignancy grade. 44 The first analysis was a histological classification. The authors compared the histological grade with the malignancy level of the tumour, observing that higher histological grade was significantly correlated with higher malignancy. Then, sections of each CMT sample were stained with a validated antibody against p53. They observed that p53 was detectably expressed only in a minority of samples, 8 out of 170. A correlation between p53 expression and higher histological grade was observed; but there was no correlation between p53 and malignancy. The conclusion was that p53 expression is related to proliferation of the tumour. 44

In a more recent study, 35 samples of intestinal cancer were analysed by immunohistochemistry. Histopathology analysis showed that 20 of them were malignant tumours, and the remaining 15 were benign. Interestingly, the malignant samples expressed higher levels of p53 compared to the benign ones. Again, in this study, a validated antibody was used to detect p53, and we can add that we have tested the same antibody and observed p53 expression by western blot in canine cell extracts (BHS and AP, unpublished results). Furthermore, sequencing analysis revealed that 3 of the malignant tissues analysed carried a tp53 mutation. 108

Another immunohistochemistry analysis performed by Kumaraguruparan and colleagues involved 30 samples of mammary tumours of dogs and humans. 60 The study compared the expression level of p53 and Bcl‐2 (an antiapoptotic protein) in tumour and adjacent non‐tumour tissues. Both proteins were found to be overexpressed in the tumour as compared with the adjacent normal tissue, 78% and 75% of the samples presented elevated levels of Bcl‐2 expression in tumour tissue in humans and dogs respectively, and 75% and 73% of the samples in the case of p53. 60 The authors concluded that the overexpression of both proteins denoted an apoptosis‐resistant phenotype in canine and human cells. This finding paved a path for a concept in which CMT can be used as a model for breast cancer study in humans. A more specific analysis on the mutational status of canine p53 protein and mutations in exons 5–8 of tp53 in 20 mammary canine tumours (12 malignant and 8 benign) 109 confirmed mutations in 33% of the examined tumour samples. The malignant tumours contained four missense and two nonsense mutations, while in the benign tissues two missense and one silent mutation were detected. Five out of six missense mutations were located in the highly conserved regions corresponding to the DNA‐binding domain in humans. These regions are often called “hot‐spots”. Researchers did a follow‐up of the dogs and found that four of them suffered from tumour recurrence after surgery and died. Interestingly, three out of the four deceased dogs exhibited mutant tp53. 109

Another study examined p53 in multicentric lymphoma in 28 dogs. 110 A total of 19 B‐cell and 9 T‐cell type lymphomas were assessed, and found to show differential expression of p53, with higher levels in T‐cell than B‐cell lymphoma as has also been observed in humans. Perhaps surprisingly, levels of p53 expression were similar in all samples within each tumour type, although elevated p53 expression is also rare in human lymphoma. 110 Even though tp53 seems not to be overexpressed in lymphoma, it can be used as a prognostic factor as demonstrated by the genetic study of Koshino et al. 111 They performed a PCR SSPC (single‐strand conformational polymorphism) to analyse tp53 mutation in 43 dogs with high‐grade lymphoma before and after treatment. They found that only 16% (7 out of 43 dogs) presented a tp53 mutation. After the chemotherapy, 88% of the non‐mutation cases responded well to the treatment, versus only 33% of the mutated cases, meaning that tp53 mutations had a negative prognostic significance.

tp53 mutation can also be inherited as in the human multi‐cancer susceptibility Li‐Fraumeni syndrome. 40 Veldhoen and colleagues investigated tp53 mutations in eight canine lymphomas, comparing tumour and non‐tumour tissue. Out of eight dogs only three expressed a mutant tp53, and interestingly one of them presented a heterozygous mutation (tp53+/−) both in the tumour and in normal somatic tissue, indicative of germline transmission. This study was the first evidence of heritable tp53 mutations in dogs. 42 Another study focusing on germline mutations of tp53 involved 10 dogs with cancer. Three of the animals presented a tp53 mutation, including one with heterozygous mutation in normal somatic tissues, confirming heritability of tp53 mutations. 43 Germline mutations of tp53 were also found in multicancer‐like syndrome and were suspected to be the cause of the tumour. 58 The mutations analysed in the study were those of tp53 and Chk2, where the latter were associated with failure in G2 arrest and genome instability.

A recent innovative study reanalyzed all the data from the 684 whole genome and exome sequences of canine tumours available to date, in order to quantify tumour mutational burden. 112 The authors performed a comparison between tumour and normal samples from each tumour type analysed, a breed validation in order to detect germline mutations associated to a specific breed, and finally, a human‐dog comparison. They observed that there is a relation between somatic mutations and the type of the tumour; T‐cell lymphoma, osteosarcoma, oral melanoma and hemangiosarcoma presented the highest tumour mutational burden. The analysis also showed that the mutation burden was similar among breeds for a given tumour type, meaning that variation is primarily tumour type‐specific rather than related to breed. The overall tumour mutational burden values were slightly lower in dogs compared to humans. Another interesting fact is that the tumour mutational burden is correlated with tp53 mutations. tp53 mutation have been found in 16.7% of the analysed canine tumours, and it was shown that tp53 mutations were preferentially present in tumours bearing a relatively higher tumour mutational burden. In the analysis comparing breeds, tp53 mutations seems to be more frequent in Golden Retrievers, Maltese or Rottweiler, compared to others. However similar oncogenic pathway alterations were observed in canine and human cancers of a given type, suggesting that the evolution of cancer is similar in dogs and humans. 112

Many immunohistochemistry analyses have been done to study p53, but there is the potential for conflict regarding the reliability of this technique when it comes to study of mutations of tp53. Strong immuno‐expression of p53 could represent accumulation of non‐functional but stable mutant protein, whereas absence of detectable expression could be considered incorrectly as the normal, vanishingly low levels of wildtype tp53, when actually p53 expression is truly absent because of missense mutations. 44 Despite this caveat, the results obtained by immunohistochemistry to detect tp53 mutations discussed in the following studies, 44 , 60 , 110 correlate with results obtained by PCR 108 , 111 and sequencing analysis. 112 All come to the same conclusion: p53 mutation can be used as a prognostic factor that its related to proliferation and malignancy of tumours. The novel technique of CRISPR/Cas9 has been used to developed a tp53 knockout canine cell line, 113 which will mark a milestone in p53 study in canine cancer. In sum, the results of various studies clearly indicate that p53 plays a key role in human and animal carcinogenesis. Both the structure and function of this protein (also in its mutated form) showed considerable inter‐species homology.

2.3. TopBP1

TopBP1 is a protein with multiple roles in the DDR. It encompasses eight hydrophobic multiple protein–protein interaction domains similar to BRCT (BRCA1 carboxyl terminal), which make it interesting due to the structural similarities to BRCA1. In fact, TopBP1 and BRCA1 share 35% of sequence homology. 33 , 35 To achieve its diverse functions, TopBP1 interacts with and regulates the activity of a wide range of different DDR proteins (e.g., BRCA1, 53BP1, p53, MDC1, ATR, BACH1 (BRCA1‐associated C‐terminal helicase), RPA/RAD51, RAD9 (Rad9 Checkpoint clamp component of 9‐1‐1 clamp: Rad9‐Hus1‐Rad1) by binding through BRCT domains to phosphorylated sites within these partner proteins (Figure 4). 25 , 36 , 114 , 115 The domain of TopBP1 which interacts with and activates ATR (the ATR Activation Domain, AAD) is distinct from the eight BRCTs domains. 116 Importantly, TopBP1 binds not only to multiple interacting partner proteins but also to DNA. 25

FIGURE 4.

FIGURE 4

TopBP1 interactions. TopBP1 contains 8 BRCT domains through which it binds different DDR components, and one AAD through which it interacts with ATR kinase. After DNA damage, MDC1 is recruited to the damage site. 114 It binds γH2AX at serine 139 and helps in the recruitment of MRN complex. 121 It also binds the 5th BRCT domain of TopBP1, which is recruited to this DNA damage site by interaction with MRN and Rad9. 122 ATR is recruited to the RPA‐coated ssDNA, but it is not enough to be activated. TopBP1 interacts with ATR through its ADD domain helping ATR activation in a ATRIP dependent manner. 121 , 123 This is how TopBP1 participates in the activation of ATR‐Chk1 pathway as a response to DNA damage. 124 , 125 Interestingly, it was shown that 53BP1 can bind TopBP1 along with Rad9, and cooperate in the activation of ATR to control G1/S checkpoint 126

TopBP1 is also involved in HR activity. BRCA2 recruits Rad51 to the RPA‐coated ssDNA, but TopBP1 is needed for the formation of Rad51 foci. The mechanism is not clear but available data suggest that BCRT 7/8 domains are essential for filament formation. 117 Another way of HR regulation is through the interaction of TopBP1 with 53BP1 and BRCA1. When both proteins bind to TopBP1, BRCA1 inhibits 53BP1 and resection of DNA ends can occur. In cancer cells without BRCA1, the TopBP1‐53BP1 interaction is stabilized and impairs HR, which is likely to result in genomic instability. 74 The authors suggest that the interaction between BRCA1 and TopBP1 is promoted by ATR, which promotes HR under replication stress conditions.

Another interesting function of TopBP1 is the regulation of p53 protein. The interaction between BRCT 7/8 of TopBP1 with DBD of p53 inhibits p53 activation. In this way, p53 cannot induce cell cycle arrest or apoptosis, and damaged cells continue to survive and replicate with potential for malignant transformation. 36

Expression of TopBP1 is expected in nucleus, as it is a protein involved in replication and DDR signalling. However, in breast cancer, TopBP1 expression has also been observed in the cytoplasm compared to normal breast tissue. 33 , 118 A histological study on the expression level of TopBP1 in breast samples from 12 healthy human patients and 61 carcinomas 33 showed that both overexpression and downregulation of the protein were related to cancer. TopBP1 occurred not only in nucleus but also in cytoplasm in cancer cells, suggesting that this mis‐localization could be related to malignancy. TopBP1 overexpression also affects p53 function. When TopBP1 is overexpressed, the G1/S phase checkpoint and apoptosis control mediated by p53 is inhibited, what promotes cancer development. 36 TopBP1 overexpression could be caused by EDD (E3 identified by differential display) protein, a ubiquitin conjugating the enzyme that controls the localization of TopBP1, which is frequently altered in different types of cancers. 119 , 120 A schematic description of TopBP1 protein interactions is shown in Figure 4.

The importance of TopBP1 in veterinary medicine is a contemporary and interesting research topic. Immunochemistry studies of TopBP1 in dogs and cats with mammary cancer, 34 , 35 confirmed its expression in all samples but the reaction with the polyclonal antibody against the human protein was much greater in the malignant tissues. 34 As already mentioned, TopBP1 appears in the cytoplasm of malignant but not of normal cells. The pattern of TopBP1 staining in samples for both humans and dogs was similar. In both cases, normal and benign tissues showed nuclear TopBP1, whereas in malignant tissue both nuclear and cytoplasmic staining was observed. In both studies, the incidence of aberrant expression of TopBP1 was found to be statistically significant in malignant samples compared to normal and benign tissue. 33 , 34 Based on these results, it can be concluded that TopBP1 protein is similar in structure in humans, dogs and cats (cross reaction with human antibody) and probably performs similar functions and undergoes similar changes in canine, feline and human cancers: its overexpression and cytoplasmic localization is related to increased malignancy of mammary tumours.

The role of TopBP1 in canine cancer is probably as significant as in human but thus far relatively little research has been performed on this subject. Even though TopBP1 has been proved to be related with malignancy, it is not simply a proliferation marker. The previous cited studies compare TopBP1 expression with Ki67, 33 , 34 which is a nuclear protein frequently used as a marker of proliferation in cancer. 127 Immunohistochemistry staining showed that TopBP1 was expressed in all benign and malignant samples, but only samples expressing Ki67 were proliferating tumours. 33 , 34 More research is needed both on the structure and function of TopBP1 in animal cancer.

2.4. Rad51

Rad51 is a recombinase of the RecA family of highly conserved proteins that share a common protein fold in their catalytic domain. All members are DNA‐dependent ATPases and can create nucleoprotein filaments on DNA that activate the catalytic activity of the DNA bound proteins. 128 Simultaneously, as a component of the DDR, Rad51 plays a central role in HR pathway. The principal function of Rad51 is first to stabilize a DNA chain broken due to direct damage or replication errors, and second, to recognize homologous DNA present in a sister chromatid to start pairing and strand exchange. 128 The requirement for a sister chromatid for DSB repair via HR probably explains why high levels of Rad51 are expressed in cells in the S and S/G2 phases. Rad51 is further stabilized upon formation of the nucleoprotein filament, which can show two conformations (open and closed), making it flexible and dynamic. The regulation of Rad51 activity in DDR is mediated by its interaction with other important proteins of this system. p53, by binding to Rad51, can inhibit strand DNA exchanges and regression of stalled replication forks. 47 BRCA2 is directly related to the regulation and coordination of HR repair mediated by binding Rad51 (Figure 2). Specifically, BRCA2 binds Rad51 through the repeated BRC domains. This way it helps Rad51 both to relocate from the cytoplasm to the nucleus under the conditions of damage and to load Rad51 onto DNA to form the nucleoprotein filament. During nucleoprotein filament formation BRCA2 helps Rad51 to displace RPA from single stranded DNA. 129 , 130 Other important interactions of Rad51 involve protein kinases. For example, ATM activates c‐Abl which then regulates Rad51 activity. In this context, it is interesting that the oncogenic version of c‐Abl, BCR‐ABL fusion kinase (expressed in chronic myelogenous leukaemia [CML]) also phosphorylates Rad51.

The level of Rad51 expression is generally higher in human tumour cell lines and in primary tumours than in normal tissues and it is associated with genomic instability and resistance to chemo‐ and radiotherapies. 131 Also, Rad51 shows sequence polymorphism in different tumours. 132 There are two possible explanations for the elevated Rad51 levels observed in tumour cells. The first is that increased Rad51 expression contributes to malignant transformation, and the second that it reflects the higher proliferation rates of malignant cells, since Rad51 levels are maximal during S and S/G2 under normal conditions. 47

The canine RAD51 gene cloned and sequenced in 2001 showed a very high homology with human and murine genes. 48 As murine Rad51 was known to interact with BRCA2 through its C‐terminus and not only through BRC domains as in humans, the canine Rad51 was tested for its capacity to bind both, the BRC domains and BRCA2 C‐terminus. It was found that a deletion of BRCA2 C‐terminal domain could increase genomic instability and predispose to cancers in dogs (see BRCA2). 50 This study also demonstrated that Rad51 bound strongly to BRC 1, 2 and 4, moderately strongly to BRC 8, and weakly to BRC 3, 5 and 7, while no interaction was observed with BRC 6 domain. Tests were also performed to check the importance of C‐terminus in the strength of Rad51‐BRCA2 interaction, where surprisingly, BRC 3, 5 and 8 deletions resulted in stronger binding. Based on these experiments, the authors hypothesized that the strength and correct arrangement of those domains could be related to HR repair proficiency. In other experiments, the effects of BRC3 and BRC4 polymorphisms on the interaction with Rad51 were investigated. To that end, the BRC domains of BRCA2 gene were sequenced from 236 dogs of five different breeds, three with a high risk of CMT, one with a low risk, and one previously not researched. Two polymorphisms were detected, T1425P and K1435R (described before). The first polymorphism did not appear in Labrador retriever breed, which is interesting, as this breed was the one with low risk of CMT. However, it appeared in 26 other dogs (two were homozygotes). The second polymorphism appeared in all breeds, in 96 dogs (26 homozygotes), but was predominant in Chihuahuas, one of the breed with high risk of CMT. 55 , 133 Another study analysed the relationship between RAD51 mutations (A209S and T225S) and PALB2 protein. A two‐hybrid assay indicated that the interaction of mutant type Rad51 with PALB2 was weaker than that of wild type Rad51. The conclusion was that RAD51 mutations affected oligomerization of the protein and that this could be the reason for its attenuated interaction with PALB2, which could promote cancer in dogs. 49

Ozmen et al. investigated genetic variations in exons 11 and 27 of BRCA2 gene (a region corresponding to Rad51 binding site), and variations in exons 6 to 9 from RAD51 gene (corresponding to PALB2 and BRCA2 interaction regions) in canine mammary gland. The most prevalent polymorphism found was T1425P in BRC3. This polymorphism was also found by Ochiai et al. and suggested to be associated with CMT. 55 In the second part of the study, a total of nine variations of canine RAD51 gene in exon 7, exon 8, intron 7 and intron 8 were identified. Four of them were non‐synonymous and altered the protein sequence. Interestingly, all these potentially significant variations were located in the region binding PALB2 and none in that binding BRCA2. 80 , 134

Other genetic variations associated with CMT consist of three single nucleotide polymorphisms (SNP), two in RAD51 gene and one in STK11 (serine/threonine kinase involved in cell growth control). This analysis included 373 dogs of which 212 suffered from CMT. The incidence of the SNPs was as follows: 50.9% in dogs with CMT and 35.4% in healthy individuals for RAD51‐SNP1, 42% in CMT and 31.9% in healthy individuals for RAD51‐SNP2, and 40.3% in CMT and 26.7% in healthy individuals for STK11. The SNPs in both genes were located in the intronic regions, but neither association was significant enough to provide strong evidence for a causal role in tumour development. 135

Due to the important functional connections between BRCA2 and Rad51, all studies concerning Rad51 in dogs are carried out in animals suffering from CMT. Some data suggest a possible relation between RAD51 alterations, genomic instability, and predisposition to CMT, but so far RAD51 variations have not been confirmed as a primary cause of cancer.

3. CLINICAL ASPECTS OF DNA DAMAGE RESEARCH IN CANCER IN DOGS

Research on this issue is important both because of the role of DNA damage in the pathogenesis of various types of cancer and because DDR disorders are a potential therapeutic target for cancer therapy. Examples of how the assessment of DNA damage can help in the clinical treatment of cancer in dogs is shown by the recent results of a number of scientific studies.

It is well known that some dogs breeds are predisposed to lymphoma 136 , 137 , 138 , 139 and genetic disease studies in dogs are especially powerful, due to dogs' relative inbreeding and the associated lack of genetic heterogeneity. 140 , 141 It was hypothesized that lymphoma susceptibility may be associated to breed‐related increase in DNA damage and a study to evaluate this hypothesis is the research done by Thamm et al. seeking to explain the cause of the higher incidence of lymphoma in Golden retrievers. 142 The subject of the article is a pilot study on 21 Golden retrievers with lymphoma, 20 age‐matched healthy Golden retrievers and 20 age‐matched healthy mixed‐breed dogs, evaluating DNA repair capability following exposure to either ionizing radiation (IR) or the chemical mutagen bleomycin. The research shows inter‐individual variation in DNA repair capacity, evaluated in stimulated canine lymphocytes exposed in vitro utilizing the G2 chromosomal radiosensitivity assay to quantify chromatid‐type aberrations (gaps and breaks). 143 Surprisingly, the results of the study point to more individual (rather than breed) susceptibility, but at the same time suggests that deficiencies in heritable factors related to DNA repair capabilities may be involved in the development of canine lymphoma. These studies set the stage for larger confirmatory studies, as well as candidate‐based approaches to probe specific genetic susceptibility factors.

Another study trying to explain the same genetic/breed predisposition to lymphoma was carried out on boxers. 144 Research aimed to evaluate whether boxer dogs have more endogenous DNA damage in peripheral leukocytes, than age‐matched non‐boxers, and whether DNA damage is associated with specific Glutathione‐S‐transferase (GST) alleles. Authors found no difference in leukocyte DNA damage, as measured by the comet assay, between boxers and age‐matched non‐boxer dogs, nor did they see an association with DNA damage and advancing age within the boxer breed. Observed lack of correlation may be explained by the fact that the experimental set‐up was different from that used in the Golden retriever study and relied only on an assessment of spontaneous DNA damage. If the response to induced DNA damage ex vivo in boxers versus non‐boxers was measured it might possibly uncover DNA repair defects that are masked in a population with heterogeneous exposures.

Another aspect of the clinical use of DNA damage research is shown in the article describing how fish oil supplementation (1000 mg; containing 232 mg EPA and 136 mg DHA), could affect DNA damage in PBMC of healthy dogs. 145 In this study, also no DNA damage was induced, and the percentage of cells bearing spontaneous DNA damage was assessed using comet assay. The study showed that fish oil supplementation not only does not induce DNA damage in PBMC, but actually reduces it. An interesting continuation of this study seems to be to investigate whether a similar effect of fish oil may be observed in conditions of induction of DNA damage in the previously mentioned Golden Retrievers and Boxers. This could be of key importance for an introduction of fish oils as an adjuvant therapy in the treatment of lymphoma in those breeds where the defective response to DNA damage is considered to be one of the predisposing causes.

The role of diet in the prevention of certain types of cancer has also been shown in a study using the dog as a model for prostate cancer research. 146 Authors presented the first evidence that prostatic DNA damage measured by comet assay may serve as a functional marker of selenium's anticarcinogenic effect on the prostate. Results suggest that measurement of selenium concentration can provide a non‐invasive method for titrating and individualizing optimal selenium intake required for prostate cancer protection.

Another example of the potential clinical usefulness of DNA damage assessment may be a study which aimed to determine whether healthy people and dogs in the same households share urinary exposures to potentially mutagenic chemical carcinogens which can lead to the development of urothelial carcinoma. 147 Although voided urothelial cell yields were inadequate to quantify DNA damage, research showed that healthy humans and pet dogs have shared urinary exposures to known mutagenic chemicals, with significantly higher levels in dogs. Correlation studies between mutagenic chemicals found in urine and their effects on DNA damage induction may provide key information on the pathogenesis of urothelial carcinoma in humans and dogs.

Assessment of the DNA damage may also be used in clinical trials to elucidate the mechanism of action of tested agents. Such an application is presented by Dull et al. in their article describing an immunofluorescence assay that distinguishes between apoptosis and drug‐induced DSBs by measuring coexpression of γH2AX and membrane blebbing‐associated cleaved caspase 3 (CC3) to indicate apoptosis, and γH2AX in the absence of CC3 blebbing to indicate drug‐induced DNA damage. 13 Because the primary pharmacodynamic endpoint for genotoxic agents is induction of markers of DNA damage repair such assays may have broad clinical and preclinical applicability and be of fundamental importance in the development of new therapies in veterinary oncology.

In veterinary medicine, we have more and more reports on preclinical and clinical trials using DNA damage determinations. 148 , 149 , 150 , 151 , 152 , 153 It seems that it is only a matter of time until knowledge about disorders of DNA damage repair in canine cancer cells will be introduced into clinical use to target specific defects with maximally effective therapies. In human medicine, an excellent example of such an approach is the introduction of PARP inhibitors to the treatment of BRCA‐dependent breast and ovarian cancer in women. 154 , 155 , 156 Thanks to the dissemination of knowledge about the possibilities of DDR research in dogs, similar targeted therapies could be also introduced in veterinary medicine.

4. CONCLUSIONS AND FUTURE DIRECTIONS

Cancer is a genetic disease linked to genomic instability in the incipient cancer cell. Failures in DDR systems are likely to contribute to cancer development, which makes them an interesting and important topic in cancer research. The findings we discuss here clearly show similarities between canine and human DDR proteins and justify the use of dogs as valuable models for DDR study. Further studies on the role and significance of key DDR components: BRCA1, BRCA2, p53, TopBP1 and Rad51 proteins in canine tumours will provide the missing information. The importance of such research is significant: if the DDR could be targeted for cancer therapy in dogs, this could aid development of analogous treatments in human oncology.

CONFLICT OF INTEREST

The authors declare no conflicts of interest.

AUTHOR CONTRIBUTIONS

Conceptualization, Aleksandra Pawlak; writing—original draft preparation, Beatriz Hernández‐Suárez and Aleksandra Pawlak; writing—review and editing, Aleksandra Pawlak and D.G.; supervision, Aleksandra Pawlak and David A. Gillespie; All authors have read and agreed to the published version of the manuscript.

ACKNOWLEDGEMENTS

The publication financed by the project “UPWR 2.0:international and interdisciplinary programme of development of Wrocław University of Environmental and Life Sciences”, co‐financed by the European Social Fund under the Operational Program Knowledge Education Development, under contract No. POWR.03.05.00‐00‐Z062/18 of 4 June 2019 and by the Polish National Agency for Academic Exchange under Grant No. PPI/APM/2019/1/00044/U/00001. David A Gillespie is an Augustín de Betancourt Investigador of the Universidad de La Laguna.

Hernández‐Suárez B, Gillespie DA, Pawlak A. DNA damage response proteins in canine cancer as potential research targets in comparative oncology. Vet Comp Oncol. 2022;20(2):347‐361. doi: 10.1111/vco.12795

Funding information European Social Fund under the Operational Program Knowledge Education Development, Grant/Award Number: POWR.03.05.00‐00‐Z062/18; Polish National Agency for Academic Exchange, Grant/Award Number: PPI/APM/2019/1/00044/U/00001; UPWR 2.0: International and Interdisciplinary Programme of Development of Wrocław University of Environmental and Life Sciences

REFERENCES

  • 1. Tian H, Gao Z, Li HZ, et al. DNA damage response ‐ a double‐edged sword in cancer prevention and cancer therapy. Cancer Lett. 2015;358:8‐16. doi: 10.1016/j.canlet.2014.12.038 [DOI] [PubMed] [Google Scholar]
  • 2. Jackson M, Marks L, May GHW, Wilson JB. The genetic basis of disease. Essays Biochem. 2018;62(5):643‐723. doi: 10.1042/EBC20170053 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3. Kanu N, Cerone MA, Goh G, et al. DNA replication stress mediates APOBEC3 family mutagenesis in breast cancer. Genome Biol. 2016;17(1):185. doi: 10.1186/s13059-016-1042-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4. Deshpande M, Romanski PA, Rosenwaks Z, Gerhardt J. Gynecological cancers caused by deficient mismatch repair and microsatellite instability. Cancers (Basel). 2020;12(11):3319. doi: 10.3390/cancers12113319 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5. Yao Y, Dai W. Genomic instability and cancer. J Carcinog Mutagen. 2014;05(02):1‐3. doi: 10.4172/2157-2518.1000165 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6. Panagopoulos I, Gorunova L, Johannsdottir IMR, et al. Chromosome translocation t(14;21)(q11;q22) activates both OLIG1 and OLIG2 in pediatric T‐cell lymphoblastic malignancies and May signify adverse prognosis. Cancer Genom Proteom. 2020;17(1):41‐48. doi: 10.21873/cgp.20166 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Corcoran NM, Clarkson MJ, Stuchbery R, Hovens CM. Molecular pathways: targeting DNA repair pathway defects enriched in metastasis. Clin Cancer Res. 2004;16(6):629‐633. doi: 10.1158/1078-0432.CCR-15-1050 [DOI] [PubMed] [Google Scholar]
  • 8. Charames G, Bapat B. Genomic instability and cancer. Curr Mol Med. 2003;3(7):589‐596. doi: 10.2174/1566524033479456 [DOI] [PubMed] [Google Scholar]
  • 9. Schmitt MW, Prindle MJ, Loeb LA. Implications of genetic heterogeneity in cancer. Ann N Y Acad Sci. 2012;1267(1):110‐116. doi: 10.1111/j.1749-6632.2012.06590.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10. O'Connor MJ. Targeting the DNA damage response in cancer. Mol Cell. 2015;60(4):547‐560. doi: 10.1016/j.molcel.2015.10.040 [DOI] [PubMed] [Google Scholar]
  • 11. McGowan CH, Russell P. The DNA damage response: sensing and signaling. Curr Opin Cell Biol. 2004;16(6):629‐633. doi: 10.1016/j.ceb.2004.09.005 [DOI] [PubMed] [Google Scholar]
  • 12. Podhorecka M, Skladanowski A, Bozko P. H2AX phosphorylation: its role in DNA damage response and cancer therapy. J Nucl Acids. 2010;2010:1‐9. doi: 10.4061/2010/920161 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13. Dull AB, Wilsker D, Hollingshead M, et al. Development of a quantitative pharmacodynamic assay for apoptosis in fixed tumor tissue and its application in distinguishing cytotoxic drug—induced DNA double strand breaks from DNA double strand breaks associated with apoptosis. Oncotarget. 2018;9(24):7104‐7116. doi: 10.18632/oncotarget.24936 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14. Joerger AC, Fersht AR. The p53 pathway: origins, inactivation in cancer, and emerging therapeutic approaches. Annu Rev Biochem. 2016;85(1):375‐404. doi: 10.1146/annurev-biochem-060815-014710 [DOI] [PubMed] [Google Scholar]
  • 15. Malaquin N, Carrier‐Leclerc A, Dessureault M, Rodier F. DDR‐mediated crosstalk between DNA‐damaged cells and their microenvironment. Front Genet. 2015;6(2):1‐8. doi: 10.3389/fgene.2015.00094 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Morandell S, Yaffe MB. Exploiting synthetic lethal interactions between DNA damage signaling, checkpoint control, and p53 for targeted cancer therapy. Progress in Molecular Biology and Translational Science; 2012;110:289‐314. doi: 10.1016/B978-0-12-387665-2.00011-0 [DOI] [PubMed] [Google Scholar]
  • 17. Goldstein M, Kastan MB. The DNA damage response: implications for tumor responses to radiation and chemotherapy. Annu Rev Med. 2015;66(1):129‐143. doi: 10.1146/annurev-med-081313-121208 [DOI] [PubMed] [Google Scholar]
  • 18. Pilié PG, Tang C, Mills GB, Yap TA. State‐of‐the‐art strategies for targeting the DNA damage response in cancer. Nat Rev Clin Oncol. 2019;16(2):81‐104. doi: 10.1038/s41571-018-0114-z [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19. Sancar A, Lindsey‐Boltz LA, Ünsal‐Kaçmaz K, Linn S. Molecular mechanisms of mammalian DNA repair and the DNA damage checkpoints. Annu Rev Biochem. 2004;73(1):39‐85. doi: 10.1146/annurev.biochem.73.011303.073723 [DOI] [PubMed] [Google Scholar]
  • 20. Niida H, Nakanishi M. DNA damage checkpoints in mammals. Mutagenesis. 2006;21(1):3‐9. doi: 10.1093/mutage/gei063 [DOI] [PubMed] [Google Scholar]
  • 21. Lou Z, Chen J. Mammalian DNA damage response pathway. Genome Instability in Cancer Development. Vol 570. Springer‐Verlag; 2005:425‐455. doi: 10.1007/1-4020-3764-3_15 [DOI] [PubMed] [Google Scholar]
  • 22. Gudmundsdottir K, Ashworth A. The roles of BRCA1 and BRCA2 and associated proteins in the maintenance of genomic stability. Oncogene. 2006;25(43):5864‐5874. doi: 10.1038/sj.onc.1209874 [DOI] [PubMed] [Google Scholar]
  • 23. Murakami Y, Tateyama S, Rungsipipat A, Uchida K, Yamaguchi R. Immunohistochemical analysis of Cyclin a, Cyclin D1 and P53 in mammary tumors, squamous cell carcinomas and basal cell tumors of dogs and cats. J Vet Med Sci. 2000;62(7):743‐750. doi: 10.1292/jvms.62.743 [DOI] [PubMed] [Google Scholar]
  • 24. Klopfleisch R, Gruber AD. Differential expression of cell cycle regulators p21, p27 and p53 in metastasizing canine mammary adenocarcinomas versus normal mammary glands. Res Vet Sci. 2009;87(1):91‐96. doi: 10.1016/j.rvsc.2008.12.010 [DOI] [PubMed] [Google Scholar]
  • 25. Choi JH, Lindsey‐Boltz LA, Sancar A. Cooperative activation of the ATR checkpoint kinase by TopBP1 and damaged DNA. Nucleic Acids Res. 2009;37(5):1501‐1509. doi: 10.1093/nar/gkn1075 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26. Zhang J, Chen X, Kent MS, Rodriguez CO, Chen X. Establishment of a dog model for the p53 family pathway and identification of a novel isoform of p21 cyclin‐dependent kinase inhibitor. Mol Cancer Res. 2009;7(1):67‐78. doi: 10.1158/1541-7786.MCR-08-0347 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. Stewart‐Ornstein J, Cheng HW, (Jacky), Lahav G. Conservation and divergence of p53 oscillation dynamics across species. Cell Syst. 2017;5(4):410‐417. doi: 10.1016/j.cels.2017.09.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Grosse N, Van Loon B, Rohrer BC. DNA damage response and DNA repair ‐ dog as a model? BMC Cancer. 2014;14(1):1‐8. doi: 10.1186/1471-2407-14-203 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Klopfleisch R, von Euler H, Sarli G, Pinho SS, Gärtner F, Gruber AD. Molecular carcinogenesis of canine mammary tumors: news from an old disease. Vet Pathol. 2011;48(1):98‐116. doi: 10.1177/0300985810390826 [DOI] [PubMed] [Google Scholar]
  • 30. Luo Y, Chen AY, Qiu J. Bocavirus infection induces a DNA damage response that facilitates viral DNA replication and mediates cell death. J Virol. 2011;85(1):133‐145. doi: 10.1128/jvi.01534-10 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Inanaga S, Igase M, Sakai Y, et al. Mismatch repair deficiency in canine neoplasms. Vet Pathol. 2021;19:1058‐1063. doi: 10.1177/03009858211022704 [DOI] [PubMed] [Google Scholar]
  • 32. Park SH, Kang HJ, Kim HS, et al. Higher DNA repair activity is related with longer replicative life span in mammalian embryonic fibroblast cells. Biogerontology. 2011;12(6):565‐579. doi: 10.1007/s10522-011-9355-2 [DOI] [PubMed] [Google Scholar]
  • 33. Going JJ, Nixon C, Dornan ES, Boner W, Donaldson MM, Morgan IM. Aberrant expression of TopBP1 in breast cancer. Histopathology. 2007;50(4):418‐424. doi: 10.1111/j.1365-2559.2007.02622.x [DOI] [PubMed] [Google Scholar]
  • 34. Morris JS, Nixon C, Bruck A, Nasir L, Morgan IM, Philbey AW. Immunohistochemical expression of TopBP1 in feline mammary neoplasia in relation to histological grade, Ki67, ERα and p53. Vet J. 2008;175:218‐226. doi: 10.1016/j.tvjl.2007.01.006 [DOI] [PubMed] [Google Scholar]
  • 35. Morris JS, Nixon C, King OJA, Morgan IM, Philbey AW. Expression of TopBP1 in canine mammary neoplasia in relation to histological type, Ki67, ERα and p53. Vet J. 2009;179(3):422‐429. doi: 10.1016/j.tvjl.2007.10.025 [DOI] [PubMed] [Google Scholar]
  • 36. Liu K, Bellam N, Lin H‐Y, et al. Regulation of p53 by TopBP1: a potential mechanism for p53 inactivation in cancer. Mol Cell Biol. 2009;29(10):2673‐2693. doi: 10.1128/mcb.01140-08 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Broustas CG, Lieberman HB. DNA damage response genes and the development of cancer metastasis. Radiat Res. 2014;181(2):111‐130. doi: 10.1667/rr13515.1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. Goh AM, Coffill CR, Lane DP. The role of mutant p53 in human cancer. J Pathol. 2011;223(2):116‐126. doi: 10.1002/path.2784 [DOI] [PubMed] [Google Scholar]
  • 39. Greenblatt MS, Bennett WP, Hollstein M, Harris CC. Mutations in the p53 tumor suppressor gene: clues to cancer etiology and molecular pathogenesis. Cancer Res. 1994;54(18):4855‐4878. [PubMed] [Google Scholar]
  • 40. Frebourg T, Barbier N, Yan YX, et al. Germ‐line p53 mutations in 15 families with Li‐Fraumeni syndrome. Am J Hum Genet. 1995;56(3):608‐615. [PMC free article] [PubMed] [Google Scholar]
  • 41. Petitjean A, Mathe E, Kato S, et al. Impact of mutant p53 functional properties on TP53 mutation patterns and tumor phenotype: lessons from recent developments in the IARC TP53 database. Hum Mutat. 2007;28(6):622‐629. doi: 10.1002/humu.20495 [DOI] [PubMed] [Google Scholar]
  • 42. Veldhoen N, Stewart J, Brown R, Milner J. Mutations of the p53 gene in canine lymphoma and evidence for germ line p53 mutations in the dog. Oncogene. 1998;16(2):249‐255. doi: 10.1038/sj.onc.1201489 [DOI] [PubMed] [Google Scholar]
  • 43. Veldhoen N, Watterson J, Brash M, Milner J. Identification of tumour‐associated and germ line p53 mutations in canine mammary cancer. Br J Cancer. 1999;81(3):409‐415. doi: 10.1038/sj.bjc.6690709 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44. Brunetti B, Bacci B, Angeli C, Benazzi C, Muscatello LV. p53, ER, and Ki67 expression in canine mammary carcinomas and correlation with pathological variables and prognosis. Vet Pathol. 2021;58(2):325‐331. doi: 10.1177/0300985820973462 [DOI] [PubMed] [Google Scholar]
  • 45. Russell DS, Jaworski L, Kisseberth WC. Immunohistochemical detection of p53, PTEN, Rb, and p16 in canine osteosarcoma using tissue microarray. J Vet Diagnos Investig. 2018;30(4):504‐509. doi: 10.1177/1040638718770239 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46. Klopfleisch R, Gruber AD. Increased expression of BRCA2 and RAD51 in lymph node metastases of canine mammary adenocarcinomas. Vet Pathol. 2009;46(3):416‐422. doi: 10.1354/vp.08-VP-0212-K-FL [DOI] [PubMed] [Google Scholar]
  • 47. Richardson C. RAD51, genomic stability, and tumorigenesis. Cancer Lett. 2005;218:127‐139. doi: 10.1016/j.canlet.2004.08.009 [DOI] [PubMed] [Google Scholar]
  • 48. Ochiai K, Morimatsu M, Tomizawa N, Syuto B. Cloning and sequencing full length of canine Brca2 and Rad51 cDNA. J Vet Med Sci. 2001;63(10):1103‐1108. doi: 10.1292/jvms.63.1103 [DOI] [PubMed] [Google Scholar]
  • 49. Uemura M, Ochiai K, Morimatsu M, et al. The canine RAD51 mutation leads to the attenuation of interaction with PALB2. Vet Comp Oncologia. 2019;18(2):247‐255. doi: 10.1111/vco.12542 [DOI] [PubMed] [Google Scholar]
  • 50. Ochiai K, Morimatsu M, Yoshikawa Y, Syuto B, Hashizume K. Brca2 C‐terminus interacts with Rad51 and contributes to nuclear focus formation in double‐strand break repair of DNA. Biomed Res. 2004;25(6):269‐275. doi: 10.2220/biomedres.25.269 [DOI] [Google Scholar]
  • 51. Nieto A, Pérez‐Alenza MD, Del Castillo N, Tabanera E, Castaño M, Peña L. BRCA1 expression in canine mammary dysplasias and tumours: relationship with prognostic variables. J Comp Pathol. 2003;128(4):260‐268. doi: 10.1053/jcpa.2002.0631 [DOI] [PubMed] [Google Scholar]
  • 52. Rivera PJ, Melin M, Biagi T, et al. Mammary tumor development in dogs is associated with BRCA1 and BRCA2. Cancer Res. 2009;69(22):8770‐8774. doi: 10.1158/0008-5472.CAN-09-1725 [DOI] [PubMed] [Google Scholar]
  • 53. Szabo CI, Wagner LA, Francisco LV, et al. Human, canine and murine BRCA1 genes: sequence comparison among species. Hum Mol Genet. 1996;5(9):1289‐1298. doi: 10.1093/hmg/5.9.1289 [DOI] [PubMed] [Google Scholar]
  • 54. Klopfleisch R, Schütze M, Gruber AD. RAD51 protein expression is increased in canine mammary carcinomas. Vet Pathol. 2010;47(1):98‐101. doi: 10.1177/0300985809353310 [DOI] [PubMed] [Google Scholar]
  • 55. Ochiai K, Ishiguro‐Oonuma T, Yoshikawa Y, et al. Polymorphisms of canine BRCA2 BRC repeats affecting interaction with RAD51. Biomed Res. 2015;36(2):155‐158. doi: 10.2220/biomedres.36.155 [DOI] [PubMed] [Google Scholar]
  • 56. Yoshida K, Miki Y. Role of BRCA1 and BRCA2 as regulators of DNA repair, transcription, and cell cycle in response to DNA damage. Cancer Sci. 2004;95(11):866‐871. doi: 10.1111/j.1349-7006.2004.tb02195.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57. Takeuchi Y, Fujino Y, Watanabe M, et al. Screening of therapeutic targets for canine mast cell tumors from a variety of kinase molecules. J Vet Med Sci. 2011;73(10):1295‐1302. doi: 10.1292/jvms.11-0093 [DOI] [PubMed] [Google Scholar]
  • 58. Marfe G, De Martino L, Tafani M, et al. A multicancer‐like syndrome in a dog characterized by p53 and cell cycle‐checkpoint kinase 2 (CHK2) mutations and Sirtuin gene (SIRT1) down‐regulation. Res Vet Sci. 2012;93:240‐245. doi: 10.1016/j.rvsc.2011.07.030 [DOI] [PubMed] [Google Scholar]
  • 59. Ressel L, Millanta F, Caleri E, Innocenti VM, Poli A. Reduced PTEN protein expression and its prognostic implications in canine and feline mammary tumors. Vet Pathol. 2009;46(5):860‐868. doi: 10.1354/vp.08-VP-0273-P-FL [DOI] [PubMed] [Google Scholar]
  • 60. Kumaraguruparan R, Prathiba D, Nagini S. Of humans and canines: immunohistochemical analysis of PCNA, Bcl‐2, p53, cytokeratin and ER in mammary tumours. Res Vet Sci. 2006;81:218‐224. doi: 10.1016/j.rvsc.2005.08.002 [DOI] [PubMed] [Google Scholar]
  • 61. Abukhdeir AM, Park BH. p21 and p27: roles in carcinogenesis and drug resistance. Expert Rev Mol Med. 2008;10:e19. doi: 10.1017/S1462399408000744 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62. Klopfleisch R, Schütze M, Gruber AD. Loss of p27 expression in canine mammary tumors and their metastases. Res Vet Sci. 2010;88:300‐303. doi: 10.1016/j.rvsc.2009.08.007 [DOI] [PubMed] [Google Scholar]
  • 63. Pawlak A, Obmińska‐Mrukowicz B, Rapak A. The dog as a model for comparative studies of lymphoma and leukemia in humans. Postepy Hig Med Dosw. 2013;67:471‐480. doi: 10.5604/17322693.1050411 [DOI] [PubMed] [Google Scholar]
  • 64. Pawlak A, Henklewska M. The role of Bcl‐xL protein research in veterinary oncology. Int J Mol Sci. 2020;21(7):2511. doi: 10.3390/ijms21072511 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65. Ostrander EA, Kruglyak L. Unleashing the canine genome. Genome Res. 2000;10(9):1271‐1274. doi: 10.1101/gr.155900 [DOI] [PubMed] [Google Scholar]
  • 66. Rütgen BC, Hammer SE, Gerner W, et al. Establishment and characterization of a novel canine B‐cell line derived from a spontaneously occurring diffuse large cell lymphoma. Leuk Res. 2010;34(7):932‐938. doi: 10.1016/j.leukres.2010.01.021 [DOI] [PubMed] [Google Scholar]
  • 67. Dhawan D, Ramos‐Vara JA, Stewart JC, Zheng R, Knapp DW. Canine invasive transitional cell carcinoma cell lines: in vitro tools to complement a relevant animal model of invasive urinary bladder cancer. Urol Oncol Semin Orig Investig. 2009;27(3):284‐292. doi: 10.1016/j.urolonc.2008.02.015 [DOI] [PubMed] [Google Scholar]
  • 68. Pawlak A, Ziolo E, Kutkowska J, et al. A novel canine B‐cell leukaemia cell line. Establishment, characterisation and sensitivity to chemotherapeutics. Vet Comp Oncol. 2017;15(4):1218‐1231. doi: 10.1111/vco.12257 [DOI] [PubMed] [Google Scholar]
  • 69. Grudzien M, Pawlak A, Kutkowska J, et al. A newly established canine NK ‐type cell line and its cytotoxic properties. Vet Comp Oncol. 2021;49(4):1‐11. doi: 10.1111/vco.12695 [DOI] [PubMed] [Google Scholar]
  • 70. Miki Y, Swensen J, Shattuck‐Eidens D, et al. A strong candidate for the breast and ovarian cancer susceptibility gene BRCA1. Science. 1994;266(5182):66‐71. doi: 10.1126/science.7545954 [DOI] [PubMed] [Google Scholar]
  • 71. Wooster R, Bignell G, Lancaster J, et al. Identification of the breast cancer susceptibility gene BRCA2. Nature. 1995;378(6559):789‐792. doi: 10.1038/378789a0 [DOI] [PubMed] [Google Scholar]
  • 72. Narod SA, Salmena L. BRCA1 and BRCA2 mutations and breast cancer. Discov Med. 2011;12(66):445‐453. [PubMed] [Google Scholar]
  • 73. Litman R, Gupta R, Brosh R Jr, Cantor S. BRCA‐FA pathway as a target for anti‐tumor drugs. Anticancer Agents Med Chem. 2008;8(4):426‐430. doi: 10.2174/187152008784220285 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74. Liu Y, Cussiol JR, Dibitetto D, et al. TOPBP1Dpb11 plays a conserved role in homologous recombination DNA repair through the coordinated recruitment of 53BP1Rad9. J Cell Biol. 2017;216(3):623‐639. doi: 10.1083/jcb.201607031 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75. Yarden RI, Pardo‐Reoyo S, Sgagias M, Cowan KH, Brody LC. BRCA1 regulates the G2/M checkpoint by activating Chk1 kinase upon DNA damage. Nat Genet. 2002;30(3):285‐289. doi: 10.1038/ng837 [DOI] [PubMed] [Google Scholar]
  • 76. Clark SL, Rodriguez AM, Snyder RR, Hankins GDV, Boehning D. Structure‐function of the tumor suppressor BRCA1. Comput Struct Biotechnol J. 2012;1(1):e201204005. doi: 10.5936/csbj.201204005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77. Bignell G. The BRC repeats are conserved in mammalian BRCA2 proteins. Hum Mol Genet. 1997;6(1):53‐58. doi: 10.1093/hmg/6.1.53 [DOI] [PubMed] [Google Scholar]
  • 78. Arai M, Utsunomiya J, Miki Y. Familial breast and ovarian cancers. Int J Clin Oncol. 2004;9(4):270‐282. doi: 10.1007/s10147-004-0423-3 [DOI] [PubMed] [Google Scholar]
  • 79. Prakash R, Zhang Y, Feng W, Jasin M. Homologous recombination and human health: the roles of BRCA1, BRCA2, and associated proteins. Cold Spring Harb Perspect Biol. 2015;7(4):a016600. doi: 10.1101/cshperspect.a016600 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80. Ozmen O, Kul S, Risvanli A, Ozalp G, Sabuncu A, Kul O. Somatic SNPs of the BRCA2 gene at the fragments encoding RAD51 binding sites of canine mammary tumors. Vet Comp Oncol. 2017;15(4):1479‐1486. doi: 10.1111/vco.12293 [DOI] [PubMed] [Google Scholar]
  • 81. McAllister KA, Bennett LM, Houle CD, et al. Cancer susceptibility of mice with a homozygous deletion in the COOH‐terminal domain of the Brca2 gene. Cancer Res. 2002;62(4):990‐994. [PubMed] [Google Scholar]
  • 82. Rajagopalan S, Andreeva A, Rutherford TJ, Fersht AR. Mapping the physical and functional interactions between the tumor suppressors p53 and BRCA2. Proc Natl Acad Sci. 2010;107(19):8587‐8592. doi: 10.1073/pnas.1003689107 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83. Lomonosov M. Stabilization of stalled DNA replication forks by the BRCA2 breast cancer susceptibility protein. Genes Dev. 2003;17(24):3017‐3022. doi: 10.1101/gad.279003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84. Deng C‐X, Scott F. Role of the tumor suppressor gene Brca1 in genetic stability and mammary gland tumor formation. Oncogene. 2000;19(8):1059‐1064. doi: 10.1038/sj.onc.1203269 [DOI] [PubMed] [Google Scholar]
  • 85. Yoshikawa Y, Morimatsu M, Ochiai K, et al. Reduced canine BRCA2 expression levels in mammary gland tumors. BMC Vet Res. 2015;11(1):1‐8. doi: 10.1186/s12917-015-0483-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86. Enginler SO, Akiş I, Toydemir TSF, et al. Genetic variations of BRCA1 and BRCA2 genes in dogs with mammary tumours. Vet Res Commun. 2014;38(1):21‐27. doi: 10.1007/s11259-013-9577-7 [DOI] [PubMed] [Google Scholar]
  • 87. Sun W, Yang X, Qiu H, et al. Relationship between three novel snps of brca1 and canine mammary tumors. J Vet Med Sci. 2015;77(11):1541‐1543. doi: 10.1292/jvms.15-0044 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88. Pinho SS, Carvalho S, Cabral J, Reis CA, Gärtner F. Canine tumors: a spontaneous animal model of human carcinogenesis. Transl Res. 2012;159:165‐172. doi: 10.1016/j.trsl.2011.11.005 [DOI] [PubMed] [Google Scholar]
  • 89. Abdelmegeed S, Mohammed S. Canine mammary tumors as a model for human disease (review). Oncol Lett. 2018;2:8195‐8205. doi: 10.3892/ol.2018.8411 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90. Söderlund K, Skoog L, Fornander T, Askmalm MS. The BRCA1/BRCA2/Rad51 complex is a prognostic and predictive factor in early breast cancer. Radiother Oncol. 2007;84(3):242‐251. doi: 10.1016/j.radonc.2007.06.012 [DOI] [PubMed] [Google Scholar]
  • 91. Uva P, Aurisicchio L, Watters J, et al. Comparative expression pathway analysis of human and canine mammary tumors. BMC Genom. 2009;10:1‐20. doi: 10.1186/1471-2164-10-135 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92. Gray M, Meehan J, Martínez‐Pérez C, et al. Naturally‐occurring canine mammary tumors as a translational model for human breast cancer. Front. Oncol. 2020;10:617. doi: 10.3389/fonc.2020.00617 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93. Thumser‐Henner P, Nytko KJ, Bley CR. Mutations of BRCA2 in canine mammary tumors and their targeting potential in clinical therapy. BMC Vet Res. 2020;16:30. doi: 10.1186/s12917-020-2247-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94. MacEwen EG, Patnaik AK, Harvey HJ, Panko WB. Estrogen receptors in canine mammary tumors. Cancer Res. 1982;42(6):1‐2259. [PubMed] [Google Scholar]
  • 95. Pichon M, Broet P, Magdelenat H, et al. Prognostic value of steroid receptors after long‐term follow‐up of 2257 operable breast cancers. Br J Cancer. 1996;73(12):1545‐1551. doi: 10.1038/bjc.1996.291 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96. Lane DP. P53, Guardian of the genome. Nature. 1992;358(6381):15‐16. doi: 10.1038/358015a0 [DOI] [PubMed] [Google Scholar]
  • 97. Linzer DIH, Levine AJ. Characterization of a 54K Dalton cellular SV40 tumor antigen present in SV40‐transformed cells and uninfected embryonal carcinoma cells. Cell. 1979;17(1):43‐52. doi: 10.1016/0092-8674(79)90293-9 [DOI] [PubMed] [Google Scholar]
  • 98. Levine AJ. p53, the cellular gatekeeper for growth and division. Cell. 1997;88(3):323‐331. doi: 10.1016/S0092-8674(00)81871-1 [DOI] [PubMed] [Google Scholar]
  • 99. Williams AB, Schumacher B. p53 in the DNA‐damage‐repair process. Cold Spring Harb Perspect Med. 2016;6(5):1‐15. doi: 10.1101/cshperspect.a026070 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100. Levine AJ. P53 and the immune response: 40 years of exploration—a plan for the future. Int J Mol Sci. 2020;21(2):541. doi: 10.3390/ijms21020541 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101. Chen J. The cell‐cycle arrest and apoptotic functions of p53 in tumor initiation and progression. Cold Spring Harb Perspect Med. 2016;6(3):a026104. doi: 10.1101/cshperspect.a026104 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102. Wade M, Li Y‐C, Wahl GM. MDM2, MDMX and p53 in oncogenesis and cancer therapy. Nat Rev Cancer. 2013;13(2):83‐96. doi: 10.1038/nrc3430 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103. Linares LK, Hengstermann A, Ciechanover A, Muller S, Scheffner M. HdmX stimulates Hdm2‐mediated ubiquitination and degradation of p53. Proc Natl Acad Sci. 2003;100(21):2009‐2014. doi: 10.1073/pnas.2030930100 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104. Muller PAJ, Vousden KH. P53 mutations in cancer. Nat Cell Biol. 2013;15(1):2‐8. doi: 10.1038/ncb2641 [DOI] [PubMed] [Google Scholar]
  • 105. Setoguchi A, Sakai T, Okuda M, et al. Aberrations of the p63 tumor suppressor gene in various tumors in dogs. Am J Vet Res. 2001;62(3):433‐439. doi: 10.2460/ajvr.2001.62.433 [DOI] [PubMed] [Google Scholar]
  • 106. Veldhoen N, Milner J. Isolation of canine p53 cDNA and detailed characterization of the full length canine p53 protein. Oncogene. 1998;16(8):1077‐1084. doi: 10.1038/sj.onc.1201863 [DOI] [PubMed] [Google Scholar]
  • 107. Rungsipipat A, Tateyama S, Yamaguchi R, Uchida K, Miyoshi N, Hayashi T. Immunohistochemical analysis of c‐yes and c‐erbB‐2 oncogene products and p53 tumor suppressor protein in canine mammary tumors. J Vet Med Sci. 1999;61(1):27‐32. doi: 10.1292/jvms.61.27 [DOI] [PubMed] [Google Scholar]
  • 108. Cho SH, Seung BJ, Kim SH, Lim HY, Sur JH. Overexpression and mutation of p53 exons 4–8 in canine intestinal adenocarcinoma. J Comp Pathol. 2020;175:79‐84. doi: 10.1016/j.jcpa.2019.12.008 [DOI] [PubMed] [Google Scholar]
  • 109. Lee CH, Kweon OK. Mutations of p53 tumor suppressor gene in spontaneous canine mammary tumors. J Vet Sci. 2002;3(4):321‐325. doi: 10.4142/jvs.2002.3.4.321 [DOI] [PubMed] [Google Scholar]
  • 110. Sokołowska J, Cywińska A, Malicka E. P53 expression in canine lymphoma. J Vet Med Ser A Physiol Pathol Clin Med. 2005;52(4):172‐175. doi: 10.1111/j.1439-0442.2005.00707.x [DOI] [PubMed] [Google Scholar]
  • 111. Koshino A, Goto‐Koshino Y, Setoguchi A, Ohno K, Tsujimoto H. Mutation of p53 gene and its correlation with the clinical outcome in dogs with lymphoma. J Vet Intern Med. 2016;30(1):223‐229. doi: 10.1111/jvim.13807 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112. Alsaihati BA, Ho K‐L, Watson J, et al. Canine tumor mutational burden is correlated with TP53 mutation across tumor types and breeds. Nat Commun. 2021;12(1):4670. doi: 10.1038/s41467-021-24836-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113. Eun K, Park MG, Jeong YW, et al. Establishment of TP53‐knockout canine cells using optimized CRIPSR/Cas9 vector system for canine cancer research. BMC Biotechnol. 2019;19(1):1. doi: 10.1186/s12896-018-0491-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114. Wang J, Gong Z, Chen J. MDC1 collaborates with TopBP1 in DNA replication checkpoint control. J Cell Biol. 2011;193(2):267‐273. doi: 10.1083/jcb.201010026 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115. Yamane K, Tsuruo T. Conserved BRCT regions of TopBP1 and of the tumor suppressor BRCA1 bind strand breaks and termini of DNA. Oncogene. 1999;18(37):5194‐5203. doi: 10.1038/sj.onc.1202922 [DOI] [PubMed] [Google Scholar]
  • 116. Kumagai A, Lee J, Yoo HY, Dunphy WG. TopBP1 activates the ATR‐ATRIP complex. Cell. 2006;124(5):943‐955. doi: 10.1016/j.cell.2005.12.041 [DOI] [PubMed] [Google Scholar]
  • 117. Moudry P, Watanabe K, Wolanin KM, et al. TOP BP1 regulates RAD51 phosphorylation and chromatin loading and determines PARP inhibitor sensitivity. J Cell Biol. 2016;212(3):281‐288. doi: 10.1083/jcb.201507042 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118. Forma E, Krzeslak A, Bernaciak M, Romanowicz‐Makowska H, Brys M. Expression of TopBP1 in hereditary breast cancer. Mol Biol Rep. 2012;39(7):7795‐7804. doi: 10.1007/s11033-012-1622-z [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119. Clancy JL, Henderson MJ, Russell AJ, et al. EDD, the human orthologue of the hyperplastic discs tumour suppressor gene, is amplified and overexpressed in cancer. Oncogene. 2003;22(32):5070‐5081. doi: 10.1038/sj.onc.1206775 [DOI] [PubMed] [Google Scholar]
  • 120. Honda Y, Tojo M, Matsuzaki K, et al. Cooperation of HECT‐domain ubiquitin ligase hHYD and DNA topoisomerase II‐binding protein for DNA damage response. J Biol Chem. 2002;277(5):3599‐3605. doi: 10.1074/jbc.M104347200 [DOI] [PubMed] [Google Scholar]
  • 121. Thada V, Cortez D. Common motifs in ETAA1 and TOPBP1 required for ATR kinase activation. J Biol Chem. 2019;294(21):8395‐8402. doi: 10.1074/jbc.RA119.008154 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122. Lee J, Kumagai A, Dunphy WG. The Rad9‐Hus1‐Rad1 checkpoint clamp regulates interaction of TopBP1 with ATR. J Biol Chem. 2007;282(38):8036‐8044. doi: 10.1074/jbc.M704635200 [DOI] [PubMed] [Google Scholar]
  • 123. Mordes DA, Glick GG, Zhao R, Cortez D. TopBP1 activates ATR through ATRIP and a PIKK regulatory domain. Genes Dev. 2008;22:1478‐1489. doi: 10.1101/gad.1666208 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124. Ohashi E, Takeishi Y, Ueda S, Tsurimoto T. Interaction between Rad9–Hus1–Rad1 and TopBP1 activates ATR–ATRIP and promotes TopBP1 recruitment to sites of UV‐damage. DNA Rep. 2014;21:1‐11. doi: 10.1016/j.dnarep.2014.05.001 [DOI] [PubMed] [Google Scholar]
  • 125. Forma E, Brys M. TopBP1 in DNA damage response. DNA Repair. InTech; 2011. doi: 10.5772/22274 [DOI] [Google Scholar]
  • 126. Bigot N, Day M, Baldock RA, Watts FZ, Oliver AW, Pearl LH. Phosphorylation‐mediated interactions with TOPBP1 couple 53BP1 and 9‐1‐1 to control the G1 DNA damage checkpoint. Elife. 2019;8:510032. doi: 10.7554/eLife.44353 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127. Yang C, Zhang J, Ding M, et al. Ki67 targeted strategies for cancer therapy. Clin Transl Oncol. 2018;20(5):570‐575. doi: 10.1007/s12094-017-1774-3 [DOI] [PubMed] [Google Scholar]
  • 128. Morrical SW. DNA‐pairing and annealing processes in homologous recombination and homology‐directed repair. Cold Spring Harb Perspect Biol. 2015;7(2):a016444. doi: 10.1101/cshperspect.a016444 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129. Van Der Zon NL, Kanaar R, Wyman C. Variation in RAD51 details a hub of functions: opportunities to advance cancer diagnosis and therapy [version 2; peer review: 4 approved]. F1000Research. 2018;7:1453. doi: 10.12688/f1000research.15650.2 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130. Jensen RB, Carreira A, Kowalczykowski SC. Purified human BRCA2 stimulates RAD51‐mediated recombination. Nature. 2010;467(7316):678‐683. doi: 10.1038/nature09399 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131. Gachechiladze M, Škarda J, Soltermann A, Joerger M. RAD51 as a potential surrogate marker for DNA repair capacity in solid malignancies. Int J Cancer. 2017;141(7):1286‐1294. doi: 10.1002/ijc.30764 [DOI] [PubMed] [Google Scholar]
  • 132. Cheng D, Shi H, Zhang K, Yi L, Zhen G. RAD 51 gene 135G/C polymorphism and the risk of four types of common cancers: a meta‐analysis. Diagn Pathol. 2014;9(1):18. doi: 10.1186/1746-1596-9-18 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133. Ochiai K, Yoshikawa Y, Oonuma T, Tomioka Y, Hashizume K, Morimatsu M. Interactions between canine RAD51 and full length or truncated BRCA2 BRC repeats. Vet J. 2011;190:293‐295. doi: 10.1016/j.tvjl.2010.11.001 [DOI] [PubMed] [Google Scholar]
  • 134. Ozmen O, Kul S, Risvanli A, Ozalp G, Sabuncu A, Kul O. Single nucleotide variations of the canine RAD51 domains, which directly binds PALB2 and BRCA2. Jpn J Vet Res. 2017;65(2):75‐82. doi: 10.14943/jjvr.65.2.75 [DOI] [Google Scholar]
  • 135. Canadas A, Santos M, Nogueira A, et al. Canine mammary tumor risk is associated with polymorphisms in RAD51 and STK11 genes. J Vet Diagnos Invest. 2018;30(5):733‐738. doi: 10.1177/1040638718789231 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136. Edwards DS, Henley WE, Harding EF, Dobson JM, Wood JLN. Breed incidence of lymphoma in a UKpopulation of insured dogs. Vet Comp Oncol. 2003;1(4):200‐206. doi: 10.1111/j.1476-5810.2003.00025.x [DOI] [PubMed] [Google Scholar]
  • 137. Pastor M, Chalvet‐Monfray K, Marchal T, et al. Genetic and environmental risk indicators in canine non‐Hodgkin's lymphomas: breed associations and geographic distribution of 608 cases diagnosed throughout France over 1 year. J Vet Intern Med. 2009;23(2):301‐310. doi: 10.1111/j.1939-1676.2008.0255.x [DOI] [PubMed] [Google Scholar]
  • 138. Elvers I, Turner‐Maier J, Swofford R, et al. Exome sequencing of lymphomas from three dog breeds reveals somatic mutation patterns reflecting genetic background. Genome Res. 2015;25(11):1634‐1645. doi: 10.1101/gr.194449.115 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139. Jankowska U, Jagielski D, Czopowicz M, Sapierzyński R. The animal‐dependent risk factors in canine T‐cell lymphomas. Vet Comp Oncol. 2017;15(2):307‐314. doi: 10.1111/vco.12164 [DOI] [PubMed] [Google Scholar]
  • 140. Leroy G. Genetic diversity, inbreeding and breeding practices in dogs: results from pedigree analyses. Vet J. 2011;189(2):177‐182. doi: 10.1016/j.tvjl.2011.06.016 [DOI] [PubMed] [Google Scholar]
  • 141. Wade CM. Inbreeding and genetic diversity in dogs: results from DNA analysis. Vet J. 2011;189:183‐188. doi: 10.1016/j.tvjl.2011.06.017 [DOI] [PubMed] [Google Scholar]
  • 142. Thamm DH, Grunerud KK, Rose BJ, Vail DM, Bailey SM. DNA repair deficiency as a susceptibility marker for spontaneous lymphoma in Golden retriever dogs: a case‐control study. PLoS One. 2013;8(7):1‐8. doi: 10.1371/journal.pone.0069192 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143. Sanford KK, Parshad R. The contribution of deficient DNA repair to chromosomal radiosensitivity of CHO cells after G2 irradiation. Cancer Genet Cytogenet. 1999;108(1):38‐41. doi: 10.1016/S0165-4608(98)00108-3 [DOI] [PubMed] [Google Scholar]
  • 144. Craun K, Ekena J, Sacco J, Jiang T, Motsinger‐Reif A, Trepanier LA. Genetic and environmental risk for lymphoma in boxer dogs. J Vet Intern Med. 2020;34(5):2068‐2077. doi: 10.1111/jvim.15849 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145. Pellegrino FJ, Risso A, Corrada Y, Gambaro RC, Seoane AI. Influence of dietary fish oil supplementation on DNA damage in peripheral blood lymphocytes of nine healthy dogs. Vet Rec Open. 2021;8(1):e12. doi: 10.1002/vro2.12 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146. Waters DJ, Shen S, Glickman LT, et al. Prostate cancer risk and DNA damage: translational significance of selenium supplementation in a canine model. Carcinogenesis. 2005;26(7):1256‐1262. doi: 10.1093/carcin/bgi077 [DOI] [PubMed] [Google Scholar]
  • 147. Craun K, Luethcke KR, Shafer M, et al. Environmental chemical exposures in the urine of dogs and people sharing the same households. J Clin Transl Sci. 2021;5(1):e54. doi: 10.1017/cts.2020.548 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148. Pawlak A, Henklewska M, Suárez BH, et al. Chalcone Methoxy derivatives exhibit Antiproliferative and Proapoptotic activity on canine lymphoma and leukemia cells. Molecules. 2020;25(19):4362. doi: 10.3390/molecules25194362 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149. Pawlak A, Bajzert J, Bugiel K, et al. Ubiquitin‐specific protease 7 as a potential therapeutic target in dogs with hematopoietic malignancies. J Vet Intern Med. 2021;35(2):1041‐1051. doi: 10.1111/jvim.16082 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150. Henklewska M, Pawlak A, Li R‐F, Yi J, Zbyryt I, Obmińska‐Mrukowicz B. Benzyl Isothiocyanate, a vegetable‐derived compound, induces apoptosis via ROS accumulation and DNA damage in canine lymphoma and leukemia cells. Int J Mol Sci. 2021;22(21):11772. doi: 10.3390/ijms222111772 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151. Tierny D, Serres F, Segaoula Z, et al. Phase i clinical pharmacology study of F14512, a new polyamine‐vectorized anticancer drug, in naturally occurring canine lymphoma. Clin Cancer Res. 2015;21(23):5314‐5323. doi: 10.1158/1078-0432.CCR-14-3174 [DOI] [PubMed] [Google Scholar]
  • 152. Mellinas‐Gomez M, Spanswick VJ, Paredes‐Moscosso SR, et al. Activity of the DNA minor groove cross‐linking agent SG2000 (SJG‐136) against canine tumours. BMC Vet Res. 2015;11(1):215. doi: 10.1186/s12917-015-0534-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153. York D, Withers SS, Watson KD, Seo KW, Rebhun RB. Enrofloxacin enhances the effects of chemotherapy in canine osteosarcoma cells with mutant and wild‐type p53. Vet Comp Oncol. 2017;15(3):1087‐1100. doi: 10.1111/vco.12250 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154. Faraoni I, Compagnone M, Lavorgna S, et al. BRCA1, PARP1 and γH2AX in acute myeloid leukemia: role as biomarkers of response to the PARP inhibitor olaparib. Biochim Biophys Acta ‐ Mol Basis Dis. 2015;1852(3):462‐472. doi: 10.1016/j.bbadis.2014.12.001 [DOI] [PubMed] [Google Scholar]
  • 155. Gadducci A, Guerrieri ME. PARP inhibitors alone and in combination with other biological agents in homologous recombination deficient epithelial ovarian cancer: from the basic research to the clinic. Crit Rev Oncol Hematol. 2017;114:153‐165. doi: 10.1016/j.critrevonc.2017.04.006 [DOI] [PubMed] [Google Scholar]
  • 156. Park HJ, Bae JS, Kim KM, et al. The PARP inhibitor olaparib potentiates the effect of the DNA damaging agent doxorubicin in osteosarcoma. J Exp Clin Cancer Res. 2018;37(1):107. doi: 10.1186/s13046-018-0772-9 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Veterinary and Comparative Oncology are provided here courtesy of Wiley

RESOURCES