Next Article in Journal
PGC-1s in the Spotlight with Parkinson’s Disease
Previous Article in Journal
Diphenyl-Methane Based Thyromimetic Inhibitors for Transthyretin Amyloidosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting Genome Stability in Melanoma—A New Approach to an Old Field

Department of Molecular Biology of Cancer, Medical University of Lodz, 92-215 Lodz, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(7), 3485; https://doi.org/10.3390/ijms22073485
Submission received: 4 March 2021 / Revised: 24 March 2021 / Accepted: 25 March 2021 / Published: 28 March 2021
(This article belongs to the Section Molecular Oncology)

Abstract

:
Despite recent groundbreaking advances in the treatment of cutaneous melanoma, it remains one of the most treatment-resistant malignancies. Due to resistance to conventional chemotherapy, the therapeutic focus has shifted away from aiming at melanoma genome stability in favor of molecularly targeted therapies. Inhibitors of the RAS/RAF/MEK/ERK (MAPK) pathway significantly slow disease progression. However, long-term clinical benefit is rare due to rapid development of drug resistance. In contrast, immune checkpoint inhibitors provide exceptionally durable responses, but only in a limited number of patients. It has been increasingly recognized that melanoma cells rely on efficient DNA repair for survival upon drug treatment, and that genome instability increases the efficacy of both MAPK inhibitors and immunotherapy. In this review, we discuss recent developments in the field of melanoma research which indicate that targeting genome stability of melanoma cells may serve as a powerful strategy to maximize the efficacy of currently available therapeutics.

1. Introduction

The genome of all living cells is constantly exposed to insults that generate DNA modifications, frequently causing DNA damage and affecting the ability of cells to survive and divide. DNA lesions include modified or mismatched bases, bulky DNA adducts, single-strand and double-strand breaks as well as crosslinks [1,2]. In the course of evolution, a large network of repair mechanisms has developed to repair all types of DNA damage: direct damage reversal [3], mismatch repair (MMR) [4], base excision repair (BER) [5], nucleotide excision repair (NER) [6], homologous repair (HR) [7], non-homologous end joining (NHEJ) [8] and the Fanconi anemia pathway (FA) [9]. Deficiencies in the repair mechanisms lead to several pathologies, including cancer development, neurological disorders and premature aging [10,11,12].
Melanoma originates in pigment-producing melanocytes and is one of the few malignancies with a continuously rising global incidence [13,14]. It is projected to be the fifth and sixth most common cancer in men and women, respectively, and accounts for the vast majority of skin cancer-related deaths [14,15]. Melanoma is one of the most highly mutated cancers, which is in large part attributable to UV light-induced cytidine to thymidine (C>T) transitions [16,17]. When looking at a general landscape of cancer development and progression, genomic instability can be viewed as a metastatic fuel [18,19]. In several types of cancer, genetic instability contributes to the acquisition of a phenotype needed for colonization of distant organs [20,21,22,23,24,25,26] and metastatic progression correlates with an increase in both mutation burden and alteration of genes involved in DNA damage response [22,27,28,29,30,31]. Somewhat surprisingly, despite the immense frequency of genetic alterations in melanoma, it is not associated with somatic defects in DNA repair [32]. On the contrary, melanomas overexpress some of the elements of the DNA repair machinery [33]. Furthermore, the ability of melanoma cells to give rise to distant metastases may rely on a certain level of genetic stability, as evidenced by increased expression of DNA repair associated genes in metastatic tumors, compared to primary lesions [34].
Historically, the DNA repair capacity of melanoma cells is considered a potential factor of profound lack of success of systemic treatments. While treatment of melanoma patients has been revolutionized with molecularly targeted therapeutics against the most frequently altered signaling cascade in melanoma—the RAS/RAF/MEK/ERK (MAPK) pathway, the majority of patients relapse within months [35,36,37,38]. Several lines of investigation indicate that treatment of melanoma cells with MAPK inhibitors uncovers DNA damage-associated vulnerabilities in melanoma cells that could be exploited therapeutically [39,40,41,42]. Furthermore, it is becoming increasingly recognized that the efficacy of immunotherapy can be substantially improved by disrupting genome integrity in melanoma cells [43,44,45,46]. In this review, we discuss recent findings concerning the mechanisms that preserve genomic stability of melanoma cells following therapy and present the discourse about the therapeutic potential of targeting DNA repair to improve the survival of melanoma patients.

2. Conventional Chemotherapy and Radiotherapy

Genome stability of cancer cells has been a focus of anti-cancer therapy for over a century, as evidenced by the universal use of DNA-damaging chemotherapeutics against virtually all cancer types. Historically, a number of DNA-damaging drugs were used against melanoma, including platinum-based drugs (cisplatin, carboplatin) and alkylating agents (dacarbazine, temozolomide, vincristine, vinblastine, carmustine, fotemustine, paclitaxel) [47]. To date, dacarbazine (DTIC) remains the only FDA-approved chemotherapeutic for the treatment of melanoma [48], although it provides hardly any clinical benefit, as complete responses were observed in less than 5% of patients [49,50]. Similar response rates were obtained with temozolomide (TMZ), an oral analog of DTIC capable of penetrating the blood-brain barrier [51]. DTIC and TMZ are pro-drugs converted to a DNA methylating product via enzymatic conversion in the liver or spontaneous conversion in all tissues, respectively. Methylation of DNA occurs primarily at O6 and N7 positions on guanine, a reaction which can be directly reversed by methyl guanine methyl transferase (MGMT). While correlation between TMZ resistance and MGMT expression is well established [52,53,54], the results of clinical evaluation of MGMT inhibitors in combination with TMZ were unsatisfactory, which could be explained by the lack of specificity of both drugs and therefore high off-target toxicity that likely precluded administration of effective doses [55]. In general, cytotoxic chemotherapy is largely ineffective against melanoma. Early efforts to increase the efficacy of chemotherapy against melanoma include polytherapy, such as BOLD (bleomycin, vincristine [Oncovin], lomustine, dacarbazine) and CVD (cisplatin, vinblastine, dacarbazine) [56,57], or the addition of immunotherapies such as IFN-α or IL-2, as well as anti-estrogens [58,59,60,61]. Neither attempt, however, has improved objective response, but instead decreased quality of life due to severe side effects of treatment [50]. Despite its profound inefficiency against melanoma, chemotherapy retains a role in palliative care of patients who no longer benefit from other treatment modalities [15]. Similarly, ionizing radiation (IR), with certain exceptions, is rarely used in melanoma patients as a single treatment [15,62,63]. Due to the lack of significant therapeutic response of melanoma patients to ionizing radiation and chemotherapy, the scientific interest in targeting melanoma cell genome integrity has been limited overall. However, recent advances in molecular mechanisms facilitating survival of melanoma cells treated with the current standard-of-care therapeutics reveal potential in disrupting genome stability as a strategy to increase drug efficacy, which will be discussed in detail below.

3. Mitogen-Activated Protein Kinase Inhibitors

The present-day melanoma treatment paradigm is in large part dictated by the prevalence of oncogenic activation of the mitogen-activated protein kinase (MAPK) pathway [64]. Because of its fundamental role in melanoma development, the MAPK pathway has been the subject of intensive research, which led to the development of small molecule inhibitors of the BRAFV600 mutant protein (vemurafenib, dabrafenib, encorafenib) and the downstream kinases MEK1/2 (trametinib, cobimetinib, binimetinib) [15,35,36,65,66]. The introduction of targeted therapeutics marked a breakthrough in treatment of patients with advanced melanoma: for the first time in decades, a newly developed treatment modality significantly improved melanoma patient survival [65,66]. However, even with combined inhibition of BRAF and MEK, the development of drug resistance is practically inevitable and most patients relapse within a year [35,36]. It is now clear that pharmaceutical inhibition of the MAPK pathway alone will not provide the anticipated cure for melanoma patients and that additional treatment options must be identified.
Large efforts are being directed towards deciphering the molecular determinants of drug resistance. The results of these efforts constantly add to a growing list of mechanisms that underlie the ability of melanoma cells to withstand treatment (recently reviewed in [67]. One of the reasons behind limited therapeutic success is profound intra-tumoral diversity, which results in the coexistence of distinct, phenotypically diverse subpopulations of melanoma cells that exhibit variable sensitivity to treatment [68,69,70,71]. In general, anti-cancer treatments are primarily effective against rapidly proliferating cells, which may lead to the selection of slow-cycling, treatment-resistant cells of a dedifferentiated phenotype that are capable of repopulating the tumor mass [72,73,74,75,76,77,78,79,80]. While the stem-like, tumor-initiating cells comprise a heterogeneous subcategory of cancer cells, it has been observed that their stemness and intrinsic drug resistance depend on a robust DNA damage response and genome stability [81,82,83,84,85,86]. In melanoma, MAPK inhibition has been shown to be less effective against cells with high expression of stem-like markers such as nerve growth factor receptor (NGFR, also known as CD271), JARID1B or AXL [78,87,88]. Interestingly, all three markers partake in the protection of genome integrity.
In addition to its established role in melanoma initiation and metastasis [89,90], NGFR has been shown to regulate genes involved in mitotic stability and DNA repair [91,92,93]. NGFR-responsive genes include Never in Mitosis (NIMA) Related Kinase 2 (NEK2), DNA Topoisomerase II Alpha TOP2A, and RAD51 Associated Protein 1 (RAD51AP1) [92]. NEK2 and TOP2A are nuclear enzymes essential for high-fidelity chromosome separation: NEK2 phosphorylates centrosomal proteins promote splitting of mitotic centrosomes while TOP2A resolves topological states of DNA during replication, transcription and repair [94,95]. RAD51AP1 is a DNA binding protein required for RAD51-mediated homologous repair and FA pathway of DNA damage response [96,97]. Furthermore, gene set enrichment analysis (GSEA) revealed that NGFR-positive melanoma cells were particularly enriched for genes participating in NER and DNA replication [93]. Importantly, knockdown of NGFR increased the frequency of DNA damage in melanoma cells and enhanced their sensitivity to fotemustine [91,92]. Such observations indicate that NGFR not only acts as a phenotypic marker of treatment-resistant cells, but also actively protects melanoma genome integrity, contributing to drug resistance.
Treatment of melanoma cells and tumors with MAPK inhibitors has been shown to induce selection of a preexisting slow-cycling subpopulation of cells with high expression of JARID1B [78,84,98]. JARID1B is a histone demethylase that plays an important role in genome stability: upon double strand DNA break, JARID1B is recruited to the damage site and demethylates H3K4me3, and thereby loosens local chromatin structure and enables Ku70/80 or BRCA1 to assemble at the damage site for NHEJ- or HR-mediated repair, respectively [99]. Pharmacological or genetic inhibition of JARID1B has been shown to induce the accumulation of H3K4me3 at the site of irradiation-induced DNA damage, impair recruitment of DNA repair machinery and compromise repair of the damage [100]. Notably, JARID1B was recently identified as a marker of cross-resistance in melanoma cells treated with BRAF inhibitors and radiation [101].
AXL receptor tyrosine kinase, whose expression is associated with stem-like properties, up-regulation of DNA repair genes and resistance to DNA damaging agents in many cancer types [102,103,104,105,106], has also been reported to be overexpressed in some MAPKi-resistant melanoma cells [88,107]. Furthermore, inhibition of AXL has been shown to aid DNA repair-targeting treatment in several cancers [104,105,108]. In melanoma, AXL inhibition, together with checkpoint kinase 1 and 2 (CHK1/CHK2) silencing, reduced the expression of DNA damage repair proteins and increased apoptosis [109]. CHK1 and CHK2 are central to proper DNA damage response, as they are activated by ataxia telangiectasia and Rad3-related (ATR) protein in response to single-strand break or ataxia-telangiectasia mutated (ATM) kinase, when double strand DNA damage occurs [110]. Although Müller and colleagues reported an inverse correlation between AXL and MITF levels in drug-resistant melanoma [88], drug resistance is not exclusively determined by AXLhigh/MITFlow phenotype [67,111]. Melanoma cells may also acquire resistance to MAPK inhibitors through a mechanism dependent on elevated expression of MITF [112,113]. Interestingly, MITF has been observed to promote chemoresistance [114] and genomic and transcriptomic homeostasis [115]. Furthermore, MITF was found to directly regulate the transcription of BRCA1 and FANCA together with a number of genes responsible for DNA repair and replication, including RAD51L3, RAD54, POLM, DNA ligase 1 (LIG1), EME1 and TERT [116,117]. Overall, increased expression of proteins involved, at least in part, in the protection of genome stability in melanoma appears to be a common feature of various mechanisms of drug resistance.
Accumulating evidence suggests that disrupting DNA repair may be an efficient strategy against melanoma. For instance, entinostat, a class I histone deacetylase (HDAC) inhibitor, reduced the expression of RAD51 and FANCD2, which, in turn, sensitized melanoma cells to TMZ and allowed for synthetic lethal targeting of PARP with olaparib [118]. Furthermore, a recent study discovered that dual inhibition of BRAFV600E and MEK with dabrafenib and trametinib potently suppressed the expression of several genes of the homologous repair pathway in a subset of melanoma cell lines, including BRIP1, BRCA2, EME1 and RBBP8 [41]. Maertens and colleagues found that combined suppression of BRAF/MEK together with entinostat resulted in additional downregulation of NHEJ related genes, including XRCC4, XRCC5, XRCC6, PNKP and PARP3, and reduced melanoma cell survival. In search of a predictive biomarker of this drug-sensitivity, the authors found that MGMT expression correlated with sensitivity to combined MAPK and HDAC inhibition [41]. Although MGMT plays a vital role in DNA repair by direct reversion of methyl guanine adducts generated by alkylating agents such as DTIC and TMZ [119], in this setting MGMT expression was deemed a passive biomarker that marked a phenotypic subset of melanomas, without having a functional role in drug sensitivity or resistance [41]. Overall, the study uncovered latent defects in DNA repair of MGMT-expressing melanoma cells exploitable only during MAPK inhibition, as treatment with PARP inhibitor alone was insufficient to exert cytotoxic effects in MGMT-expressing melanomas. Arguably, these findings could have a profound clinical application. Clinical assessment of MGMT promoter methylation, which is currently routinely employed to predict sensitivity of glioblastoma to TMZ [120], could be used in melanoma patients to predict responsiveness to this triple drug combination. Clinical reports indicate that MGMT promoter is methylated, and thus suppressed, in about 21.5–35% of metastatic melanomas [119,121,122,123], which entails that inhibition of BRAF, MEK and HDAC could potentially benefit a large number of melanoma patients. Compatible with findings by Maertens and colleagues, our group has recently reported that vemurafenib and trametinib reduced the expression of BRIP1, BRCA1 and BRCA2 in 4 out of 5 patient-derived BRAFV600E cell lines [42]. In addition, we found that insulin modulated the response of melanoma cells to MAPK inhibition by attenuating drug-induced reduction in the expression of DNA repair genes, deregulation of glutathione homeostasis and increase in DNA damage, which altogether led to substantially reduced cytotoxicity of vemurafenib and trametinib [42]. The role of MAPK-targeted therapeutics in deregulation of DNA repair was also confirmed in other cancer types: BRAFV600E inhibition using vemurafenib was shown to suppress double strand break repair and sensitize thyroid cancer cells to ionizing radiation [124], and MEK inhibition using trametinib induced similar effects in pancreatic cancer cells [125]. Altogether, these reports point to a promising direction in melanoma treatment that could potentially combine MAPK inhibitors with various DNA repair-targeting drugs that have already been FDA-approved.
Once melanoma cells become resistant to MAPK inhibitors, effective therapeutic options are scarce. This prompted intense investigation into specific targeting of MAPKi resistant melanoma cells. One approach to battling therapeutic resistance aims to exploit the apparent drug addiction of melanoma cells: reportedly, melanomas adapted to MAPK inhibition exhibit loss-of-fitness upon drug withdrawal [126,127]. Drug-resistant melanomas are known to circumvent MAPK inhibition and reactivate ERK activity [128,129] and recently, the supra-basal ERK activity induced by drug withdrawal has been reported to be a targetable vulnerability of drug-resistant cells [39]. Depending on the degree of ERK hyperactivation, it caused cell cycle deceleration or DNA damage and parthanatos-related cell death [39]. Pharmacological suppression of DNA repair with ATM and PARP inhibitors augmented the drug addiction phenomenon, leading to caspase-dependent cell death in slow-cycling cells with innately weak phospho-ERK rebound or accelerated parthanatos-related death in cells with excessive ERK activity [39]. These findings suggest there is a therapeutic window for targeting DNA repair in patients who progressed on MAPK inhibitors and no longer receive treatment, regardless of the initial mechanisms underlying drug resistance (Figure 1). Another approach against drug-resistant cancer cells is to identify and target potential vulnerabilities that developed in parallel with drug resistance. For instance, melanoma cells resistant to MAPK inhibitors exhibited reduced expression of TAp73alfa, which binds to p53-responsive genes and modulates DNA repair and genomic stability [40]. Downregulated levels of TAp73alfa resulted in reduced NER capacity in melanoma cells and enhanced susceptibility to DNA cross-linking agents as consequences of accumulating DNA damage [40]. Moreover, it was found that drug-resistant melanoma cells rely on RAD51 expression for survival and can be successfully targeted with Rad51 inhibitors [33]. Rad51 is an indispensable protein of the homology-directed DNA repair pathway and is critical for the maintenance of genomic stability [7]. Its expression in melanoma cells is in part regulated by Elk1 as an outlet of MAPK signaling [33,130]. Inhibition of RAD51 led to the accumulation of DNA damage and enhanced drug efficacy in drug-naïve, vemurafenib-resistant and vemurafenib and trametinib double-resistant melanoma cells, as well as in a xenograft of melanoma cells isolated from a patient with developed resistance to vemurafenib [33]. Notably, Rad51 inhibitors showed no apparent toxicity in mice, which warrants further research into clinical application of Rad51 inhibitors in the treatment of cancers, including melanoma [33]. Drug-resistant melanoma cells have also been shown to lose phosphatase and tensin homolog (PTEN) expression to maintain the pro-survival signaling of the PI3K/AKT pathway [131,132]. Interestingly, cells deficient in PTEN exhibited compromised chromosomal stability, DNA replication and NHEJ repair [133,134,135,136], which makes PTEN-deficient melanomas potential candidates for synthetic lethal targeting of the redundant DNA repair proteins. Enhanced lethality in PTEN-null melanoma tumors has been reported following PARP, RAD51 and ATR inhibition [137,138]. Notably, loss of PTEN expression is frequently observed in melanomas irrespective of MAPKi resistance. Although genetic alterations in PTEN are found in only about 15% of metastatic melanomas [139], epigenetic silencing of PTEN is common at different stages of the disease [140,141].
Lastly, several groups have proposed to target DNA repair proteins in melanoma cells independently of MAPK inhibitors. Santamaria and colleagues reported that melanoma cells depend on the expression of lysyl oxidase-like 3 (LOXL3) for survival [142]. Gain- and loss-of-function studies revealed that LOXL3 contributed to melanomagenesis, protected genome integrity of melanoma cells and physically interacted with other proteins involved in genome integrity maintenance, such as BRCA2 and DNA mismatch repair protein MSH2. Loss of LOXL3 resulted in aberrant DNA damage response, characterized by inefficient ATR activation and a substantial decrease in protein levels of BRCA1, BRCA2, MSH2 and Rad51, altogether leading to the accumulation of double-strand DNA breaks and aberrant mitosis [142]. Another group has reported that disruption of DNA repair and epigenetic state by knockdown of a BER enzyme thymine DNA glycosylase (TDG) may be a potent strategy against melanoma, and furthermore, the group also evaluated candidate inhibitors of TDG for the treatment of melanoma [143]. Finally, CC-115, a dual inhibitor of mTOR and DNA-PK, has been shown to be cytotoxic and radiosensitizing in melanoma cells and is currently under clinical evaluation [144].

4. Immunotherapy

As with MAPK pathway inhibitors, the past decade brought groundbreaking advances in immunotherapy for melanoma patients. The therapeutic paradigm has shifted from a relatively ineffective cytokine-based treatment to antibody-mediated immune checkpoint inhibition that provides exceptionally durable responses [145]. Currently used immunotherapeutics target cytotoxic T-lymphocyte-associated antigen 4 (CTLA-4) or programmed cell death receptor 1 (PD-1) and prevent them from binding to CD80/CD86 or PD-L1, respectively (Figure 2). These particular ligand-receptor interactions constrain lymphocyte activation in a non-redundant manner [146,147]. Nivolumab and pembrolizumab, which are both anti-PD-1 antibodies, demonstrate superior clinical benefit over CTLA-4 blocking ipilimumab and remain the standard of care for melanoma patients whose tumors do not harbor BRAFV600 mutations [148,149]. Although immunotherapy provides unprecedentedly durable benefit in some patients, a substantial percentage of patients relapse within 2 years of treatment and approximately half of patients do not achieve significant therapeutic response at all [150,151,152,153]. Therefore, the ultimate challenge of immunotherapy is to sensitize drug-refractory melanomas that evade immune recognition and prevent relapse in immunotherapy-responsive individuals.
Molecular determinants of effective immunosurveillance include various interdependent host and tumor cell factors, which have been described in greater detail elsewhere [154,155,156,157]. In short, anti-tumor activity of the immune system depends on the presence of tumor-specific antigens on the melanoma cell surface, immune infiltration of the tumor and activation of the immune cell response [158]. The expression and altered repertoire of antigens is driven by large quantities of somatic mutations, which in melanoma, on average, is greater than in many other tumors [159,160]. The mutation burden, however, varies between patients, which renders some melanoma tumors less immunogenic [160,161,162,163]. Furthermore, for neoantigens to generate immunological response, the cell needs to properly process and present neoepitopes by the major histocompatibility complex class I (MHC class I), which is often downregulated in melanomas [164,165]. A critical prerequisite for successful immunotherapy is the presence of immune cells inside the tumor [166,167]. Finally, the outcome of melanoma and immune cell encounters depends on sufficient activation of the immune cells present in the tumor microenvironment, which melanoma cells prohibit by upregulating the expression of programmed death-ligand 1 (PD-L1) [168].
Considering that all the aforementioned factors are indispensable for full anti-tumor immunological response, it has been increasingly recognized that DNA damage, particularly when induced by ionizing radiation, may serve as an immunotherapy-augmenting treatment [169]. Ionizing radiation induces a variety of DNA lesions that have the potential to be cytotoxic or mutagenic, both of which can stimulate an innate and adaptive immunological response [170,171] (Figure 2). Damaged DNA increases the likelihood of neoantigen formation, enhances the expression of MHC class I loaded with unique, irradiation-induced peptides and as a result, alerts the immune system of melanoma cells [172,173]. Furthermore, irradiation can lead to an immunogenic cell death, associated with the release of damage-associated molecular pattern (DAMP) molecules which boost immunological defense [174]. Radiation can induce the accumulation of cytosolic DNA, either in the form of micronuclei generated from chromosomal aberrations or small double-stranded DNA fragments that leak through a ruptured nuclear envelope [175]. Cytosolic DNA can then trigger inflammatory signaling in cancer cells or can be released and sensed by immune cells within the tumor microenvironment. It has been shown that ionizing radiation enhances melanoma-derived DNA uptake by antigen presenting cells (APC) and promotes anti-tumor immunity [176,177]. Mechanistically, the presence of cytosolic, tumor-derived DNA in immune cells is recognized by cyclic GMP-AMP synthase (cGAS) which activates stimulator of interferon (IFN) genes (STING), leading to IFN-beta production and CD8+ T cell priming [178]. Importantly, anti-tumor efficacy of radiotherapy has been shown to substantially rely on host cell STING signaling, rather than direct cytotoxicity of ionizing radiation in melanoma cells [179].
With all the immunostimulatory effects of DNA damage, it is worth emphasizing that radiotherapy as a single agent is not effective against melanoma, at least not in terms of cure [180]. As with chemotherapy, the efficacy of radiotherapy relies on the DNA repair capacity of targeted tumor cells and is substantially more effective in cancers with deficiencies in DNA repair [181]. Nonetheless, radiation temporarily shrinks melanoma tumors, which denotes some degree of cancer cell death, and this in turn can be leveraged with immunotherapy [48]. Most importantly, radiotherapy has been shown to increase tumor immune infiltration, which is also demonstrated by increased expression of PD-L1 on the surface of melanoma cells as an adaptive mechanism against immune cell mobilization inside the tumor mass [182,183]. Given that the preexisting lymphocyte infiltration is crucial for immunotherapy efficacy [158], this otherwise drug-efficacy-lowering PD-L1 expression ultimately results in a substantially more effective tumor cell killing after anti-PD-1 treatment. Furthermore, radiotherapy combined with immunotherapeutics has been shown to generate tumor antigen-specific memory T-cell responses and limit T-cell exhaustion [182,184]. Radiation can also induce an immune-mediated phenomenon called the abscopal effect, in which irradiation of one tumor leads to the regression of a non-irradiated tumor at a distance from the irradiated site [185]. Mechanistically, this effect is based on DNA damage-induced mobilization of systemic immunosurveillance, particularly mediated by cGAS-STING signaling [186]. Several clinical reports have demonstrated the abscopal effect in melanoma patients who received radiotherapy prior to treatment with anti-CLTA-4 [186,187,188] and anti-PD-1 antibodies [189,190,191]. However, the sequence and timing of treatment is believed to be a crucial determinant of the efficacy of this therapeutic strategy. Dovedi and colleagues found that irradiation-induced CD8+ T cell infiltration sharply decreased after 7 days post radiation, which abrogated the synergistic effect of radiotherapy and PD-1/PD-L1 blockade [182]. Another study found that inflammatory signaling after radiotherapy-induced DNA damage was dependent on mitotic progression, which generated micronuclei and stimulated cGAS-STING signaling. Both inhibiting mitosis as well as the loss of STING signaling impaired interferon signaling and the loss of STING signaling limited the abscopal effect of radiation [186]. Hence, allowing time for tumor cells to divide was necessary for the immunostimulatory effects. Most clinical reports have postulated that combined treatment provides superior benefit when it is not substantially stretched over time [192]. Concurrent administration of anti-CTLA-4 or anti-PD-1 treatment in patients who underwent stereotactic radiosurgery of metastatic lesions in the brain significantly improved median tumor reduction comparing to non-concurrent immunotherapy [193]. Furthermore, overall survival of patients who received ipilimumab within 14 days of radiotherapy was significantly greater than in patients who underwent radiotherapy 4 months prior to immunotherapy [194]. Altogether, substantial benefit of combining immunotherapy with radiation for melanoma patients has been proven [43,44,45,46]. On the other hand, a retrospective study comparing immunotherapy with or without preceding radiotherapy reported no evidence of favorable outcome of pretreatment radiotherapy. However, the study was not set up to determine the time-from-radiotherapy impact on the efficacy of immunotherapy, and included stage IV patients who underwent radiotherapy at a non-specified time prior to immunotherapy administration [195]. This study, however, is a good example of the need of strict control of treatment schedule.
Overall, as evidenced by the use of radiotherapy, DNA damage exposes tumor cells to the immune system and stimulates T cell activation. This is further illustrated by the association between defects in DNA repair and the efficacy of immunotherapy [196,197], particularly in tumors with mutations in MMR genes, in which exceptional responses to pembrolizumab have been observed [198,199,200]. Notably, genomic landscape of triple wild-type melanomas (lacking mutations in RAS, RAF and NF1 genes), which account for 5–10% of melanomas, was recently characterized as being enriched for mutations in DNA repair-related genes [201]. Although overall rare in melanoma, mutations in DNA repair-related genes were also associated with enhanced response to immunotherapy. For instance, an unprecedented overall survival rate exceeding 7 years was reported for a patient with metastatic melanoma lacking the expression of MMR gene MSH6 [202], while in a preclinical melanoma model, knockdown of MSH2 in poorly immunogenic tumors markedly improved the response to anti-PD-1 immunotherapy [203]. Furthermore, presumed loss-of-function mutations in BRCA2 were found to be significantly more frequent in anti-PD-1 treatment responsive tumors [152]. It logically follows that targeting DNA repair together with irradiation of melanoma tumors could amplify the immunogenic DNA damage. Combination of radiotherapy with AsiDNA, a DNA repair trapping molecule that hijacks HR and NHEJ proteins, has been well tolerated in mice [204] and encouraged phase I clinical evaluation of this treatment modality [205]. Tourneau and colleagues reported complete responses in 30% of patients and speculated that immunogenicity-induced abscopal effect contributed to the systemic response to locally injected treatment [205]. Simultaneous targeting of DNA repair with immunotherapy has also been successful in BRCA-deficient cancers, which serves as an example of how decreased genome stability synergizes with immune checkpoint blockade [206,207]. Granted, cells deficient in HR due to mutations in BRCA genes are 100–1000-fold more sensitive to pharmaceutical inhibition of PARP; however, a study analyzing PARP inhibition in combination with AsiDNA in HR-proficient tumors showed that this drug combination can recapitulate synthetic lethality in tumors, regardless of their HR mutation status [208].

5. Conclusions and Future Directions

MAPK pathway antagonists and immune checkpoint inhibitors will likely remain the cornerstone of melanoma treatment. However, it is clear that additional targets need to be identified to maximize treatment efficacy, eliminate subpopulations of melanoma cells that are intrinsically resistant to treatment and prohibit the development of drug resistance. A growing body of evidence supports the hypothesis that melanoma cells depend on a certain level of genome stability when undergoing drug treatment and that inducing DNA damage may augment the efficacy of MAPK inhibition and immune checkpoint blockade. While anti-tumor effects of BRAF and MEK inhibitors could be enhanced with drugs targeting proteins involved in DNA repair, immunotherapy may draw benefit from directly interfering with the DNA integrity. Several clinical trials are under way to evaluate the safety and efficacy of treatment regimens that combine MAPK or immune checkpoint inhibitors with chemotherapy and radiotherapy or drugs that potentially destabilize genome integrity (Table 1). Such drugs include inhibitors of histone deacetylases (HBI-8000, 4SC-202 or entinostat), PARP (olaparib, niraparib and talazoparib) and ATR (ceralasertib/AZD6738). However, along with measuring response rates to treatment, still many clinical trials are prominently focused on assessing dosage limited toxicities and incidences of adverse effects. Nonetheless, the outcome of these studies may pave the way for research into targeting melanoma genome stability in order to enhance the efficacy of current first-line treatment options. The majority of ongoing phase 2 and 3 clinical trials test the combination of immunotherapy with genome-targeting treatments (Table 1). Notably, these studies are not powered to establish the optimal treatment schedule, which may be a crucial determinant of the treatment outcome. Therefore, it would be beneficial to set out clinical evaluation aimed specifically to elucidate the best radiation and immunotherapy timeline. On the other hand, only a few of the current clinical trials examine the therapeutic potential of MAPK inhibitors combined with radiation, one evaluates combination of MAPKi with an AXL inhibitor and none involve drugs targeting proteins of the DNA repair pathways or histone deacetylases (Table 1). This is likely because preclinical data that points to an exploitable dependency on genome stability in melanoma cells treated with MAPK inhibitors is fairly recent and not yet extensively explored. In summary, combining current anti-melanoma drugs with therapeutic agents that destabilize the melanoma cell genome and interfere with DNA repair offers great promise in melanoma treatment and further studies on both preclinical and clinical levels are warranted.

6. Methodology

To comprehensively cover studies relevant to the investigated topic, a literature search was conducted using Pubmed, Scopus and Web of Science, based on the following keywords, alone or in combination: melanoma, DNA damage, DNA repair, targeted therapy, immunotherapy, chemotherapy, radiotherapy, genome stability. Preference was given to papers published within the past 10 years. All information presented in this review has been thoroughly examined and discussed between the authors.
A summary of clinical trials presented in Table 1 was prepared based on clinicaltrials.gov (accessed on 23 March 2021) database [209] using “melanoma” as the keyword and filtered by: ”recruiting”, ”not yet recruiting” and ”active, not recruiting” to focus only on ongoing studies. The initial list yielded 923 hits. The search was narrowed down to clinical trials involving skin cutaneous melanoma and excluded phase 1 clinical trials to focus on studies measuring the outcome of treatment instead of only the appropriate dosage of treatment. Finally, studies that involved a combination of immune checkpoint or MAPK inhibitors with genotoxic agents as well as studies evaluating drugs targeting DNA repair alone or in combination with immune checkpoint or MAPK inhibitors were selected.

Author Contributions

Conceptualization, M.O.; writing—original draft preparation, M.O.; writing—review and editing, M.W. Both authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank Malgorzata Czyz for helpful remarks and discussion.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roos, W.P.; Thomas, A.D.; Kaina, B. DNA damage and the balance between survival and death in cancer biology. Nat. Rev. Cancer 2016, 16, 20–33. [Google Scholar] [CrossRef] [PubMed]
  2. Chatterjee, N.; Walker, G.C. Mechanisms of DNA damage, repair, and mutagenesis. Environ. Mol. Mutagen. 2017, 58, 235–263. [Google Scholar] [CrossRef] [Green Version]
  3. Yi, C.; He, C. DNA repair by reversal of DNA damage. Cold Spring Harb. Perspect. Biol. 2013, 5, a012575. [Google Scholar] [CrossRef] [PubMed]
  4. Li, G.M. Mechanisms and functions of DNA mismatch repair. Cell Res. 2008, 18, 85–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Robertson, A.B.; Klungland, A.; Rognes, T.; Leiros, I. DNA repair in mammalian cells: Base excision repair: The long and short of it. Cell. Mol. Life Sci. 2009, 66, 981–993. [Google Scholar] [CrossRef]
  6. Marteijn, J.A.; Lans, H.; Vermeulen, W.; Hoeijmakers, J.H. Understanding nucleotide excision repair and its roles in cancer and ageing. Nat. Rev. Mol. Cell. Biol. 2014, 15, 465–481. [Google Scholar] [CrossRef] [PubMed]
  7. Wright, W.D.; Shah, S.S.; Heyer, W.D. Homologous recombination and the repair of DNA double-strand breaks. J. Biol. Chem. 2018, 293, 10524–10535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Davis, A.J.; Chen, D.J. DNA double strand break repair via non-homologous end-joining. Transl. Cancer Res. 2013, 2, 130–143. [Google Scholar] [CrossRef]
  9. Walden, H.; Deans, A.J. The Fanconi anemia DNA repair pathway: Structural and functional insights into a complex disorder. Annu. Rev. Biophys. 2014, 43, 257–278. [Google Scholar] [CrossRef]
  10. Friedberg, E.C. DNA damage and repair. Nature 2003, 421, 436–440. [Google Scholar] [CrossRef]
  11. McKinnon, P.J. DNA repair deficiency and neurological disease. Nat. Rev. Neurosci. 2009, 10, 100–112. [Google Scholar] [CrossRef] [Green Version]
  12. Tiwari, V.; Wilson, D.M., 3rd. DNA Damage and Associated DNA Repair Defects in Disease and Premature Aging. Am. J. Hum. Genet. 2019, 105, 237–257. [Google Scholar] [CrossRef] [Green Version]
  13. Dimitriou, F.; Krattinger, R.; Ramelyte, E.; Barysch, M.J.; Micaletto, S.; Dummer, R.; Goldinger, S.M. The World of Melanoma: Epidemiologic, Genetic, and Anatomic Differences of Melanoma across the Globe. Curr. Oncol. Rep. 2018, 20, 87. [Google Scholar] [CrossRef]
  14. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2020. CA Cancer J. Clin. 2020, 70, 7–30. [Google Scholar] [CrossRef]
  15. American Cancer Society. Available online: https://www.cancer.org/cancer/melanoma-skin-cancer.html (accessed on 10 February 2021).
  16. Cancer Genome Atlas Network. Genomic Classification of Cutaneous Melanoma. Cell 2015, 161, 1681–1696. [Google Scholar] [CrossRef] [Green Version]
  17. Birkeland, E.; Zhang, S.; Poduval, D.; Geisler, J.; Nakken, S.; Vodak, D.; Meza-Zepeda, L.A.; Hovig, E.; Myklebost, O.; Knappskog, S.; et al. Patterns of genomic evolution in advanced melanoma. Nat. Commun. 2018, 10, 2665. [Google Scholar] [CrossRef] [PubMed]
  18. Fidler, I.J.; Kripke, M.L. The challenge of targeting metastasis. Cancer Metastasis Rev. 2015, 34, 635–641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Bakhoum, S.F.; Ngo, B.; Laughney, A.M.; Cavallo, J.A.; Murphy, C.J.; Ly, P.; Shah, P.; Sriram, R.K.; Watkins, T.B.K.; Taunk, N.K.; et al. Chromosomal instability drives metastasis through a cytosolic DNA response. Nature 2018, 553, 467–472. [Google Scholar] [CrossRef] [Green Version]
  20. Campbell, P.J.; Yachida, S.; Mudie, L.J.; Stephens, P.J.; Pleasance, E.D.; Stebbings, L.A.; Morsberger, L.A.; Latimer, C.; McLaren, S.; Lin, M.L.; et al. The patterns and dynamics of genomic instability in metastatic pancreatic cancer. Nature 2010, 467, 1109–1113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Yachida, S.; Jones, S.; Bozic, I.; Antal, T.; Leary, R.; Fu, B.; Kamiyama, M.; Hruban, R.H.; Eshleman, J.R.; Nowak, M.A.; et al. Distant metastasis occurs late during the genetic evolution of pancreatic cancer. Nature 2010, 467, 1114–1117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Bertucci, F.; Ng, C.K.Y.; Patsouris, A.; Droin, N.; Piscuoglio, S.; Carbuccia, N.; Soria, J.C.; Dien, A.T.; Adnani, Y.; Kamal, M.; et al. Genomic characterization of metastatic breast cancers. Nature 2019, 569, 560–564. [Google Scholar] [CrossRef]
  23. Yates, L.R.; Knappskog, S.; Wedge, D.; Farmery, J.H.R.; Gonzalez, S.; Martincorena, I.; Alexandrov, L.B.; Van Loo, P.; Haugland, H.K.; Lilleng, P.K.; et al. Genomic Evolution of Breast Cancer Metastasis and Relapse. Cancer Cell. 2017, 32, 169–184.e7. [Google Scholar] [CrossRef] [Green Version]
  24. Jamal-Hanjani, M.; Wilson, G.A.; McGranahan, N.; Birkbak, N.J.; Watkins, T.B.K.; Veeriah, S.; Shafi, S.; Johnson, D.H.; Mitter, R.; Rosenthal, R.; et al. Tracking the Evolution of Non-Small-Cell Lung Cancer. N. Engl. J. Med. 2017, 376, 2109–2121. [Google Scholar] [CrossRef] [Green Version]
  25. Turajlic, S.; Xu, H.; Litchfield, K.; Rowan, A.; Chambers, T.; Lopez, J.I.; Nicol, D.; O’Brien, T.; Larkin, J.; Horswell, S.; et al. Tracking Cancer Evolution Reveals Constrained Routes to Metastases: TRACERx Renal. Cell 2018, 173, 581–594.e12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Casuscelli, J.; Weinhold, N.; Gundem, G.; Wang, L.; Zabor, E.C.; Drill, E.; Wang, P.I.; Nanjangud, G.J.; Redzematovic, A.; Nargund, A.M.; et al. Genomic landscape and evolution of metastatic chromophobe renal cell carcinoma. JCI Insight. 2017, 2, e92688. [Google Scholar] [CrossRef] [PubMed]
  27. Angus, L.; Smid, M.; Wilting, S.M.; van Riet, J.; Van Hoeck, A.; Nguyen, L.; Nik-Zainal, S.; Steenbruggen, T.G.; Tjan-Heijnen, V.C.G.; Labots, M.; et al. The genomic landscape of metastatic breast cancer highlights changes in mutation and signature frequencies. Nat. Genet. 2019, 51, 1450–1458. [Google Scholar] [CrossRef]
  28. Mei, P.; Freitag, C.E.; Wei, L.; Zhang, Y.; Parwani, A.V.; Li, Z. High tumor mutation burden is associated with DNA damage repair gene mutation in breast carcinomas. Diagn. Pathol. 2020, 15, 50. [Google Scholar] [CrossRef]
  29. Bashashati, A.; Ha, G.; Tone, A.; Ding, J.; Prentice, L.M.; Roth, A.; Rosner, J.; Shumansky, K.; Kalloger, S.; Senz, J.; et al. Distinct evolutionary trajectories of primary high-grade serous ovarian cancers revealed through spatial mutational profiling. J. Pathol. 2013, 231, 21–34. [Google Scholar] [CrossRef]
  30. Quigley, D.A.; Dang, H.X.; Zhao, S.G.; Lloyd, P.; Aggarwal, R.; Alumkal, J.J.; Foye, A.; Kothari, V.; Perry, M.D.; Bailey, A.M.; et al. Genomic Hallmarks and Structural Variation in Metastatic Prostate Cancer. Cell 2018, 174, 758–769.e9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Barroso-Sousa, R.; Jain, E.; Cohen, O.; Kim, D.; Buendia-Buendia, J.; Winer, E.; Lin, N.; Tolaney, S.M.; Wagle, N. Prevalence and mutational determinants of high tumor mutation burden in breast cancer. Ann. Oncol. 2020, 31, 387–394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Lawrence, M.S.; Stojanov, P.; Polak, P.; Kryukov, G.V.; Cibulskis, K.; Sivachenko, A.; Carter, S.L.; Stewart, C.; Mermel, C.H.; Roberts, S.A.; et al. Mutational heterogeneity in cancer and the search for new cancer-associated genes. Nature 2013, 499, 214–218. [Google Scholar] [CrossRef] [PubMed]
  33. Makino, E.; Fröhlich, L.M.; Sinnberg, T.; Kosnopfel, C.; Sauer, B.; Garbe, C.; Schittek, B. Targeting Rad51 as a strategy for the treatment of melanoma cells resistant to MAPK pathway inhibition. Cell Death Dis. 2020, 11, 581. [Google Scholar] [CrossRef] [PubMed]
  34. Kauffmann, A.; Rosselli, F.; Lazar, V.; Winnepenninckx, V.; Mansuet-Lupo, A.; Dessen, P.; van den Oord, J.J.; Spatz, A.; Sarasin, A. High expression of DNA repair pathways is associated with metastasis in melanoma patients. Oncogene 2008, 27, 565–573. [Google Scholar] [CrossRef] [Green Version]
  35. Long, G.V.; Stroyakovskiy, D.; Gogas, H.; Levchenko, E.; de Braud, F.; Larkin, J.; Garbe, C.; Jouary, T.; Hauschild, A.; Grob, J.J.; et al. Dabrafenib and trametinib versus dabrafenib and placebo for Val600 BRAF-mutant melanoma: A multicentre, double-blind, phase 3 randomised controlled trial. Lancet 2015, 386, 444–451. [Google Scholar] [CrossRef]
  36. Robert, C.; Karaszewska, B.; Schachter, J.; Rutkowski, P.; Mackiewicz, A.; Stroiakovski, D.; Lichinitser, M.; Dummer, R.; Grange, F.; Mortier, L.; et al. Improved overall survival in melanoma with combined dabrafenib and trametinib. N. Engl. J. Med. 2015, 372, 30–39. [Google Scholar] [CrossRef] [Green Version]
  37. Luke, J.J.; Flaherty, K.T.; Ribas, A.; Long, G.V. Targeted agents and immunotherapies: Optimizing outcomes in melanoma. Nat. Rev. Clin. Oncol. 2017, 14, 463–482. [Google Scholar] [CrossRef] [Green Version]
  38. Long, G.V.; Eroglu, Z.; Infante, J.; Patel, S.; Daud, A.; Johnson, D.B.; Gonzalez, R.; Kefford, R.; Hamid, O.; Schuchter, L.; et al. Long-Term Outcomes in Patients With BRAF V600-Mutant Metastatic Melanoma Who Received Dabrafenib Combined With Trametinib. J. Clin. Oncol. 2018, 36, 667–673. [Google Scholar] [CrossRef] [PubMed]
  39. Hong, A.; Moriceau, G.; Sun, L.; Lomeli, S.; Piva, M.; Damoiseaux, R.; Holmen, S.L.; Sharpless, N.E.; Hugo, W.; Lo, R.S. Exploiting Drug Addiction Mechanisms to Select against MAPKi-Resistant Melanoma. Cancer Discov. 2018, 8, 74–93. [Google Scholar] [CrossRef] [Green Version]
  40. Makino, E.; Gutmann, V.; Kosnopfel, C.; Niessner, H.; Forschner, A.; Garbe, C.; Sinnberg, T.; Schittek, B. Melanoma cells resistant towards MAPK inhibitors exhibit reduced TAp73 expression mediating enhanced sensitivity to platinum-based drugs. Cell Death Dis. 2018, 9, 930. [Google Scholar] [CrossRef]
  41. Maertens, O.; Kuzmickas, R.; Manchester, H.E.; Emerson, C.E.; Gavin, A.G.; Guild, C.J.; Wong, T.C.; De Raedt, T.; Bowman-Colin, C.; Hatchi, E.; et al. MAPK Pathway Suppression Unmasks Latent DNA Repair Defects and Confers a Chemical Synthetic Vulnerability in BRAF-, NRAS-, and NF1-Mutant Melanomas. Cancer Discov. 2019, 9, 526–545. [Google Scholar] [CrossRef] [Green Version]
  42. Osrodek, M.; Rozanski, M.; Czyz, M. Insulin Reduces the Efficacy of Vemurafenib and Trametinib in Melanoma Cells. Cancer Manag. Res. 2020, 12, 7231–7250. [Google Scholar] [CrossRef]
  43. Hiniker, S.M.; Chen, D.S.; Reddy, S.; Chang, D.T.; Jones, J.C.; Mollick, J.A.; Swetter, S.M.; Knox, S.J. A systemic complete response of metastatic melanoma to local radiation and immunotherapy. Transl. Oncol. 2012, 5, 404–407. [Google Scholar] [CrossRef] [Green Version]
  44. Mohiuddin, M.; Park, H.; Hallmeyer, S.; Richards, J. High-Dose Radiation as a Dramatic, Immunological Primer in Locally Advanced Melanoma. Cureus 2015, 7, e417. [Google Scholar] [CrossRef] [Green Version]
  45. Haymaker, C.L.; Kim, D.; Uemura, M.; Vence, L.M.; Phillip, A.; McQuail, N.; Brown, P.D.; Fernandez, I.; Hudgens, C.W.; Creasy, C.; et al. Metastatic Melanoma Patient Had a Complete Response with Clonal Expansion after Whole Brain Radiation and PD-1 Blockade. Cancer Immunol. Res. 2017, 5, 100–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Hiniker, S.M.; Reddy, S.A.; Maecker, H.T.; Subrahmanyam, P.B.; Rosenberg-Hasson, Y.; Swetter, S.M.; Saha, S.; Shura, L.; Knox, S.J. A Prospective Clinical Trial Combining Radiation Therapy With Systemic Immunotherapy in Metastatic Melanoma. Int. J. Radiat. Oncol. Biol. Phys. 2016, 96, 578–588. [Google Scholar] [CrossRef] [Green Version]
  47. Bomar, L.; Senithilnathan, A.; Ahn, C. Systemic Therapies for Advanced Melanoma. Dermatol. Clin. 2019, 37, 409–423. [Google Scholar] [CrossRef] [PubMed]
  48. National Cancer Institute. Available online: https://www.cancer.gov/types/skin/patient/melanoma-treatment-pdq (accessed on 25 February 2021).
  49. Gogas, H.J.; Kirkwood, J.M.; Sondak, V.K. Chemotherapy for metastatic melanoma: Time for a change? Cancer 2007, 109, 455–464. [Google Scholar] [CrossRef] [PubMed]
  50. Wilson, M.A.; Schuchter, L.M. Chemotherapy for Melanoma. Cancer Treat. Res. 2016, 167, 209–229. [Google Scholar] [CrossRef]
  51. Middleton, M.R.; Grob, J.J.; Aaronson, N.; Fierlbeck, G.; Tilgen, W.; Seiter, S.; Gore, M.; Aamdal, S.; Cebon, J.; Coates, A.; et al. Randomized phase III study of temozolomide versus dacarbazine in the treatment of patients with advanced metastatic malignant melanoma. J. Clin. Oncol. 2000, 18, 158–166. [Google Scholar] [CrossRef] [PubMed]
  52. Kokkinakis, D.M.; Ahmed, M.M.; Chendil, D.; Moschel, R.C.; Pegg, A.E. Sensitization of pancreatic tumor xenografts to carmustine and temozolomide by inactivation of their O6-Methylguanine-DNA methyltransferase with O6-benzylguanine or O6-benzyl-2’-deoxyguanosine. Clin. Cancer Res. 2003, 9, 3801–3807. [Google Scholar] [PubMed]
  53. Nagane, M.; Kobayashi, K.; Ohnishi, A.; Shimizu, S.; Shiokawa, Y. Prognostic significance of O6-methylguanine-DNA methyltransferase protein expression in patients with recurrent glioblastoma treated with temozolomide. Jpn. J. Clin. Oncol. 2007, 37, 897–906. [Google Scholar] [CrossRef] [Green Version]
  54. Kitange, G.J.; Carlson, B.L.; Schroeder, M.A.; Grogan, P.T.; Lamont, J.D.; Decker, P.A.; Wu, W.; James, C.D.; Sarkaria, J.N. Induction of MGMT expression is associated with temozolomide resistance in glioblastoma xenografts. Neuro-Oncology 2009, 11, 281–291. [Google Scholar] [CrossRef] [Green Version]
  55. Ranson, M.; Hersey, P.; Thompson, D.; Beith, J.; McArthur, G.A.; Haydon, A.; Davis, I.D.; Kefford, R.F.; Mortimer, P.; Harris, P.A.; et al. Randomized trial of the combination of lomeguatrib and temozolomide compared with temozolomide alone in chemotherapy naive patients with metastatic cutaneous melanoma. J. Clin. Oncol. 2007, 25, 2540–2545. [Google Scholar] [CrossRef] [Green Version]
  56. Eggermont, A.M.; Kirkwood, J.M. Re-evaluating the role of dacarbazine in metastatic melanoma: What have we learned in 30 years? Eur. J. Cancer 2004, 40, 1825–1836. [Google Scholar] [CrossRef] [PubMed]
  57. Luke, J.J.; Schwartz, G.K. Chemotherapy in the management of advanced cutaneous malignant melanoma. Clin. Dermatol. 2013, 31, 290–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Hansson, J.; Aamdal, S.; Bastholt, L.; Brandberg, Y.; Hernberg, M.; Nilsson, B.; Stierner, U.; von der Maase, H.; Nordic Melanoma Cooperative Group. Two different durations of adjuvant therapy with intermediate-dose interferon alfa-2b in patients with high-risk melanoma (Nordic IFN trial): A randomised phase 3 trial. Lancet Oncol. 2011, 12, 144–152. [Google Scholar] [CrossRef]
  59. Garbe, C.; Radny, P.; Linse, R.; Dummer, R.; Gutzmer, R.; Ulrich, J.; Stadler, R.; Weichenthal, M.; Eigentler, T.; Ellwanger, U.; et al. Adjuvant low-dose interferon α2a with or without dacarbazine compared with surgery alone: A prospective-randomized phase III DeCOG trial in melanoma patients with regional lymph node metastasis. Ann. Oncol. 2008, 19, 1195–1201. [Google Scholar] [CrossRef] [PubMed]
  60. Cocconi, G.; Bella, M.; Calabresi, F.; Tonato, M.; Canaletti, R.; Boni, C.; Buzzi, F.; Ceci, G.; Corgna, E.; Costa, P. Treatment of metastatic malignant melanoma with dacarbazine plus tamoxifen. N. Engl. J. Med. 1992, 327, 516–523. [Google Scholar] [CrossRef]
  61. Falkson, C.I.; Ibrahim, J.; Kirkwood, J.M.; Coates, A.S.; Atkins, M.B.; Blum, R.H. Phase III trial of dacarbazine versus dacarbazine with interferon alpha-2b versus dacarbazine with tamoxifen versus dacarbazine with interferon alpha-2b and tamoxifen in patients with metastatic malignant melanoma: An Eastern Cooperative Oncology Group study. J. Clin. Oncol. 1998, 16, 1743–1751. [Google Scholar] [CrossRef]
  62. Rogers, S.J.; Puric, E.; Eberle, B.; Datta, N.R.; Bodis, S.B. Radiotherapy for Melanoma: More than DNA Damage. Dermatol. Res. Pract. 2019, 2019, 9435389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Farshad, A.; Burg, G.; Panizzon, R.; Dummer, R. A retrospective study of 150 patients with lentigo maligna and lentigo maligna melanoma and the efficacy of radiotherapy using Grenz or soft X-rays. Br. J. Dermatol. 2002, 146, 1042–1046. [Google Scholar] [CrossRef]
  64. Savoia, P.; Fava, P.; Casoni, F.; Cremona, O. Targeting the ERK Signaling Pathway in Melanoma. Int. J. Mol. Sci. 2019, 20, 1483. [Google Scholar] [CrossRef] [Green Version]
  65. Chapman, P.B.; Hauschild, A.; Robert, C.; Haanen, J.B.; Ascierto, P.; Larkin, J.; Dummer, R.; Garbe, C.; Testori, A.; Maio, M.; et al. Improved survival with vemurafenib in melanoma with BRAF V600E mutation. N. Engl. J. Med. 2011, 364, 2507–2516. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Flaherty, K.T.; Infante, J.R.; Daud, A.; Gonzalez, R.; Kefford, R.F.; Sosman, J.; Hamid, O.; Schuchter, L.; Cebon, J.; Ibrahim, N.; et al. Combined BRAF and MEK inhibition in melanoma with BRAF V600 mutations. N. Engl. J. Med. 2012, 367, 1694–1703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Kozar, I.; Margue, C.; Rothengatter, S.; Haan, C.; Kreis, S. Many ways to resistance: How melanoma cells evade targeted therapies. Biochim. Biophys. Acta Rev. Cancer 2019, 1871, 313–322. [Google Scholar] [CrossRef]
  68. Romano, E.; Pradervand, S.; Paillusson, A.; Weber, J.; Harshman, K.; Muehlethaler, K.; Speiser, D.; Peters, S.; Rimoldi, D.; Michielin, O. Identification of multiple mechanisms of resistance to vemurafenib in a patient with BRAFV600E-mutated cutaneous melanoma successfully rechallenged after progression. Clin. Cancer Res. 2013, 19, 5749–5757. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Tirosh, I.; Izar, B.; Prakadan, S.M.; Wadsworth, M.H., 2nd; Treacy, D.; Trombetta, J.J.; Rotem, A.; Rodman, C.; Lian, C.; Murphy, G.; et al. Dissecting the multicellular ecosystem of metastatic melanoma by single-cell RNA-seq. Science 2016, 352, 189–196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Shannan, B.; Perego, M.; Somasundaram, R.; Herlyn, M. Heterogeneity in Melanoma. Cancer Treat. Res. 2016, 167, 1–15. [Google Scholar] [CrossRef] [PubMed]
  71. Ahn, A.; Chatterjee, A.; Eccles, M.R. The Slow Cycling Phenotype: A Growing Problem for Treatment Resistance in Melanoma. Mol. Cancer Ther. 2017, 16, 1002–1009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Bhatia, R.; Holtz, M.; Niu, N.; Gray, R.; Snyder, D.S.; Sawyers, C.L.; Arber, D.A.; Slovak, M.L.; Forman, S.J. Persistence of malignant hematopoietic progenitors in chronic myelogenous leukemia patients in complete cytogenetic remission following imatinib mesylate treatment. Blood 2003, 101, 4701–4707. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Dean, M.; Fojo, T.; Bates, S. Tumour stem cells and drug resistance. Nat. Rev. Cancer 2005, 5, 275–284. [Google Scholar] [CrossRef]
  74. Dylla, S.J.; Beviglia, L.; Park, I.K.; Chartier, C.; Raval, J.; Ngan, L.; Pickell, K.; Aguilar, J.; Lazetic, S.; Smith-Berdan, S.; et al. Colorectal cancer stem cells are enriched in xenogeneic tumors following chemotherapy. PLoS ONE 2008, 3, e2428. [Google Scholar] [CrossRef]
  75. Singh, A.; Settleman, J. EMT, cancer stem cells and drug resistance: An emerging axis of evil in the war on cancer. Oncogene 2010, 29, 4741–4751. [Google Scholar] [CrossRef] [Green Version]
  76. Wu, F.H.; Mu, L.; Li, X.L.; Hu, Y.B.; Liu, H.; Han, L.T.; Gong, J.P. Characterization and functional analysis of a slow-cycling subpopulation in colorectal cancer enriched by cell cycle inducer combined chemotherapy. Oncotarget 2017, 8, 78466–78479. [Google Scholar] [CrossRef] [PubMed]
  77. Shlush, L.I.; Mitchell, A.; Heisler, L.; Abelson, S.; Ng, S.W.K.; Trotman-Grant, A.; Medeiros, J.J.F.; Rao-Bhatia, A.; Jaciw-Zurakowsky, I.; Marke, R.; et al. Tracing the origins of relapse in acute myeloid leukaemia to stem cells. Nature 2017, 547, 104–108. [Google Scholar] [CrossRef]
  78. Roesch, A.; Vultur, A.; Bogeski, I.; Wang, H.; Zimmermann, K.M.; Speicher, D.; Körbel, C.; Laschke, M.W.; Gimotty, P.A.; Philipp, S.E.; et al. Overcoming intrinsic multidrug resistance in melanoma by blocking the mitochondrial respiratory chain of slow-cycling JARID1B(high) cells. Cancer Cell 2013, 23, 811–825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Hartman, M.L.; Rozanski, M.; Osrodek, M.; Zalesna, I.; Czyz, M. Vemurafenib and trametinib reduce expression of CTGF and IL-8 in V600EBRAF melanoma cells. Lab. Investig. 2017, 97, 217–227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Moore, N.; Houghton, J.; Lyle, S. Slow-cycling therapy-resistant cancer cells. Stem Cells Dev. 2012, 21, 1822–1830. [Google Scholar] [CrossRef] [Green Version]
  81. Desai, A.; Webb, B.; Gerson, S.L. CD133+ cells contribute to radioresistance via altered regulation of DNA repair genes in human lung cancer cells. Radiother. Oncol. 2014, 110, 538–545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Lim, Y.C.; Roberts, T.L.; Day, B.W.; Harding, A.; Kozlov, S.; Kijas, A.W.; Ensbey, K.S.; Walker, D.G.; Lavin, M.F. A role for homologous recombination and abnormal cell-cycle progression in radioresistance of glioma-initiating cells. Mol. Cancer Ther. 2012, 11, 1863–1872. [Google Scholar] [CrossRef] [Green Version]
  83. Srivastava, A.K.; Han, C.; Zhao, R.; Cui, T.; Dai, Y.; Mao, C.; Zhao, W.; Zhang, X.; Yu, J.; Wang, Q.E. Enhanced expression of DNA polymerase eta contributes to cisplatin resistance of ovarian cancer stem cells. Proc. Natl. Acad. Sci. USA 2015, 112, 4411–4416. [Google Scholar] [CrossRef] [Green Version]
  84. Liu, Y.; Burness, M.L.; Martin-Trevino, R.; Guy, J.; Bai, S.; Harouaka, R.; Brooks, M.D.; Shang, L.; Fox, A.; Luther, T.K.; et al. RAD51 Mediates Resistance of Cancer Stem Cells to PARP Inhibition in Triple-Negative Breast Cancer. Clin. Cancer Res. 2017, 23, 514–522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Vitale, I.; Manic, G.; De Maria, R.; Kroemer, G.; Galluzzi, L. DNA Damage in Stem Cells. Mol. Cell 2017, 66, 306–319. [Google Scholar] [CrossRef]
  86. Schulz, A.; Meyer, F.; Dubrovska, A.; Borgmann, K. Cancer Stem Cells and Radioresistance: DNA Repair and Beyond. Cancers 2019, 11, 862. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Lehraiki, A.; Cerezo, M.; Rouaud, F.; Abbe, P.; Allegra, M.; Kluza, J.; Marchetti, P.; Imbert, V.; Cheli, Y.; Bertolotto, C.; et al. Increased CD271 expression by the NF-kB pathway promotes melanoma cell survival and drives acquired resistance to BRAF inhibitor vemurafenib. Cell Discov. 2015, 1, 15030. [Google Scholar] [CrossRef] [PubMed]
  88. Müller, J.; Krijgsman, O.; Tsoi, J.; Robert, L.; Hugo, W.; Song, C.; Kong, X.; Possik, P.A.; Cornelissen-Steijger, P.D.; Geukes Foppen, M.H.; et al. Low MITF/AXL ratio predicts early resistance to multiple targeted drugs in melanoma. Nat. Commun. 2014, 5, 5712. [Google Scholar] [CrossRef]
  89. Boiko, A.D.; Razorenova, O.V.; van de Rijn, M.; Swetter, S.M.; Johnson, D.L.; Ly, D.P.; Butler, P.D.; Yang, G.P.; Joshua, B.; Kaplan, M.J.; et al. Human melanoma-initiating cells express neural crest nerve growth factor receptor CD271. Nature 2010, 466, 133–137. [Google Scholar] [CrossRef]
  90. Civenni, G.; Walter, A.; Kobert, N.; Mihic-Probst, D.; Zipser, M.; Belloni, B.; Seifert, B.; Moch, H.; Dummer, R.; van den Broek, M.; et al. Human CD271-positive melanoma stem cells associated with metastasis establish tumor heterogeneity and long-term growth. Cancer Res. 2011, 71, 3098–3109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Redmer, T.; Welte, Y.; Behrens, D.; Fichtner, I.; Przybilla, D.; Wruck, W.; Yaspo, M.L.; Lehrach, H.; Schäfer, R.; Regenbrecht, C.R. The nerve growth factor receptor CD271 is crucial to maintain tumorigenicity and stem-like properties of melanoma cells. PLoS ONE 2014, 9, e92596. [Google Scholar] [CrossRef] [Green Version]
  92. Redmer, T.; Walz, I.; Klinger, B.; Khouja, S.; Welte, Y.; Schäfer, R.; Regenbrecht, C. The role of the cancer stem cell marker CD271 in DNA damage response and drug resistance of melanoma cells. Oncogenesis 2017, 6, e291. [Google Scholar] [CrossRef] [Green Version]
  93. Filipp, F.V.; Li, C.; Boiko, A.D. CD271 is a molecular switch with divergent roles in melanoma and melanocyte development. Sci. Rep. 2019, 9, 7696. [Google Scholar] [CrossRef] [Green Version]
  94. Fang, Y.; Zhang, X. Targeting NEK2 as a promising therapeutic approach for cancer treatment. Cell Cycle 2016, 15, 895–907. [Google Scholar] [CrossRef] [Green Version]
  95. Zhang, M.; Liang, C.; Chen, Q.; Yan, H.; Xu, J.; Zhao, H.; Yuan, X.; Liu, J.; Lin, S.; Lu, W.; et al. Histone H2A phosphorylation recruits topoisomerase IIα to centromeres to safeguard genomic stability. EMBO J. 2020, 39, e101863. [Google Scholar] [CrossRef]
  96. Pires, E.; Sung, P.; Wiese, C. Role of RAD51AP1 in homologous recombination DNA repair and carcinogenesis. DNA Repair 2017, 59, 76–81. [Google Scholar] [CrossRef]
  97. Liang, F.; Miller, A.S.; Longerich, S.; Tang, C.; Maranon, D.; Williamson, E.A.; Hromas, R.; Wiese, C.; Kupfer, G.M.; Sung, P. DNA requirement in FANCD2 deubiquitination by USP1-UAF1-RAD51AP1 in the Fanconi anemia DNA damage response. Nat. Commun. 2019, 10, 2849. [Google Scholar] [CrossRef] [PubMed]
  98. Roesch, A.; Fukunaga-Kalabis, M.; Schmidt, E.C.; Zabierowski, S.E.; Brafford, P.A.; Vultur, A.; Basu, D.; Gimotty, P.; Vogt, T.; Herlyn, M. A temporarily distinct subpopulation of slow-cycling melanoma cells is required for continuous tumor growth. Cell 2010, 141, 583–594. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Li, X.; Liu, L.; Yang, S.; Song, N.; Zhou, X.; Gao, J.; Yu, N.; Shan, L.; Wang, Q.; Liang, J.; et al. Histone demethylase KDM5B is a key regulator of genome stability. Proc. Natl. Acad. Sci. USA 2014, 111, 7096–7101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Bayo, J.; Tran, T.A.; Wang, L.; Pena-Llopis, S.; Das, A.K.; Martinez, E.D. Jumonji inhibitors overcome radioresistance in cancer through changes in H3K4 methylation at double-strand breaks. Cell. Rep. 2018, 25, 1040–1050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Shannan, B.; Matschke, J.; Chauvistré, H.; Vogel, F.; Klein, D.; Meier, F.; Westphal, D.; Bruns, J.; Rauschenberg, R.; Utikal, J.; et al. Sequence-dependent cross-resistance of combined radiotherapy plus BRAFV600E inhibition in melanoma. Eur. J. Cancer 2019, 109, 137–153. [Google Scholar] [CrossRef]
  102. Zhu, C.; Wei, Y.; Wei, X. AXL receptor tyrosine kinase as a promising anti-cancer approach: Functions, molecular mechanisms and clinical applications. Mol. Cancer 2019, 18, 153. [Google Scholar] [CrossRef] [Green Version]
  103. Brand, T.M.; Iida, M.; Stein, A.P.; Corrigan, K.L.; Braverman, C.M.; Coan, J.P.; Pearson, H.E.; Bahrar, H.; Fowler, T.L.; Bednarz, B.P.; et al. AXL Is a Logical Molecular Target in Head and Neck Squamous Cell Carcinoma. Clin. Cancer Res. 2015, 21, 2601–2612. [Google Scholar] [CrossRef] [Green Version]
  104. Balaji, K.; Vijayaraghavan, S.; Diao, L.; Tong, P.; Fan, Y.; Carey, J.P.; Bui, T.N.; Warner, S.; Heymach, J.V.; Hunt, K.K.; et al. AXL Inhibition Suppresses the DNA Damage Response and Sensitizes Cells to PARP Inhibition in Multiple Cancers. Mol. Cancer Res. 2017, 15, 45–58. [Google Scholar] [CrossRef] [Green Version]
  105. Kariolis, M.S.; Miao, Y.R.; Diep, A.; Nash, S.E.; Olcina, M.M.; Jiang, D.; Jones, D.S., 2nd; Kapur, S.; Mathews, I.I.; Koong, A.C.; et al. Inhibition of the GAS6/AXL pathway augments the efficacy of chemotherapies. J. Clin. Investig. 2017, 127, 183–198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Wang, C.; Jin, H.; Wang, N.; Fan, S.; Wang, Y.; Zhang, Y.; Wei, L.; Tao, X.; Gu, D.; Zhao, F.; et al. Gas6/Axl Axis Contributes to Chemoresistance and Metastasis in Breast Cancer through Akt/GSK-3β/β-catenin Signaling. Theranostics 2016, 6, 1205–1219. [Google Scholar] [CrossRef]
  107. Zuo, Q.; Liu, J.; Huang, L.; Qin, Y.; Hawley, T.; Seo, C.; Merlino, G.; Yu, Y. AXL/AKT axis mediated-resistance to BRAF inhibitor depends on PTEN status in melanoma. Oncogene 2018, 37, 3275–3289. [Google Scholar] [CrossRef]
  108. Sen, T.; Tong, P.; Diao, L.; Li, L.; Fan, Y.; Hoff, J.; Heymach, J.V.; Wang, J.; Byers, L.A. Targeting AXL and mTOR Pathway Overcomes Primary and Acquired Resistance to WEE1 Inhibition in Small-Cell Lung Cancer. Clin. Cancer Res. 2017, 23, 6239–6253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Flem-Karlsen, K.; McFadden, E.; Omar, N.; Haugen, M.H.; Øy, G.F.; Ryder, T.; Gullestad, H.P.; Hermann, R.; Mælandsmo, G.M.; Flørenes, V.A. Targeting AXL and the DNA Damage Response Pathway as a Novel Therapeutic Strategy in Melanoma. Mol. Cancer Ther. 2020, 19, 895–905. [Google Scholar] [CrossRef] [Green Version]
  110. Smith, J.; Tho, L.M.; Xu, N.; Gillespie, D.A. The ATM-Chk2 and ATR-Chk1 pathways in DNA damage signaling and cancer. Adv. Cancer Res. 2010, 108, 73–112. [Google Scholar] [CrossRef] [PubMed]
  111. Hartman, M.L.; Sztiller-Sikorska, M.; Gajos-Michniewicz, A.; Czyz, M. Dissecting Mechanisms of Melanoma Resistance to BRAF and MEK Inhibitors Revealed Genetic and Non-Genetic Patient- and Drug-Specific Alterations and Remarkable Phenotypic Plasticity. Cells 2020, 9, 142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Pathria, G.; Garg, B.; Borgdorff, V.; Garg, K.; Wagner, C.; Superti-Furga, G.; Wagner, S.N. Overcoming MITF-conferred drug resistance through dual AURKA/MAPK targeting in human melanoma cells. Cell Death Dis. 2016, 7, e2135. [Google Scholar] [CrossRef] [Green Version]
  113. Arozarena, I.; Wellbrock, C. Overcoming resistance to BRAF inhibitors. Ann. Transl. Med. 2017, 5, 387. [Google Scholar] [CrossRef] [Green Version]
  114. Hertzman Johansson, C.; Azimi, A.; Frostvik Stolt, M.; Shojaee, S.; Wiberg, H.; Grafström, E.; Hansson, J.; Egyházi Brage, S. Association of MITF and other melanosome-related proteins with chemoresistance in melanoma tumors and cell lines. Melanoma Res. 2013, 23, 360–365. [Google Scholar] [CrossRef] [PubMed]
  115. Seoane, M.; Buhs, S.; Iglesias, P.; Strauss, J.; Puller, A.C.; Müller, J.; Gerull, H.; Feldhaus, S.; Alawi, M.; Brandner, J.M.; et al. Lineage-specific control of TFIIH by MITF determines transcriptional homeostasis and DNA repair. Oncogene 2019, 38, 3616–3635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Beuret, L.; Ohanna, M.; Strub, T.; Allegra, M.; Davidson, I.; Bertolotto, C.; Ballotti, R. BRCA1 is a new MITF target gene. Pigment Cell Melanoma Res. 2011, 24, 725–727. [Google Scholar] [CrossRef] [PubMed]
  117. Strub, T.; Giuliano, S.; Ye, T.; Bonet, C.; Keime, C.; Kobi, D.; Le Gras, S.; Cormont, M.; Ballotti, R.; Bertolotto, C.; et al. Essential role of microphthalmia transcription factor for DNA replication, mitosis and genomic stability in melanoma. Oncogene 2011, 30, 2319–2332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Krumm, A.; Barckhausen, C.; Kücük, P.; Tomaszowski, K.H.; Loquai, C.; Fahrer, J.; Krämer, O.H.; Kaina, B.; Roos, W.P. Enhanced Histone Deacetylase Activity in Malignant Melanoma Provokes RAD51 and FANCD2-Triggered Drug Resistance. Cancer Res. 2016, 76, 3067–3077. [Google Scholar] [CrossRef] [Green Version]
  119. Tuominen, R.; Jewell, R.; van den Oord, J.J.; Wolter, P.; Stierner, U.; Lindholm, C.; Hertzman Johansson, C.; Lindén, D.; Johansson, H.; Frostvik Stolt, M.; et al. MGMT promoter methylation is associated with temozolomide response and prolonged progression-free survival in disseminated cutaneous melanoma. Int. J. Cancer 2015, 136, 2844–2853. [Google Scholar] [CrossRef] [Green Version]
  120. Weller, M.; Stupp, R.; Reifenberger, G.; Brandes, A.A.; van den Bent, M.J.; Wick, W.; Hegi, M.E. MGMT promoter methylation in malignant gliomas: Ready for personalized medicine? Nat. Rev. Neurol. 2010, 6, 39–51. [Google Scholar] [CrossRef] [Green Version]
  121. Hoon, D.S.; Spugnardi, M.; Kuo, C.; Huang, S.K.; Morton, D.L.; Taback, B. Profiling epigenetic inactivation of tumor suppressor genes in tumors and plasma from cutaneous melanoma patients. Oncogene 2004, 23, 4014–4022. [Google Scholar] [CrossRef] [Green Version]
  122. Hassel, J.C.; Sucker, A.; Edler, L.; Kurzen, H.; Moll, I.; Stresemann, C.; Spieth, K.; Mauch, C.; Rass, K.; Dummer, R.; et al. MGMT gene promoter methylation correlates with tolerance of temozolomide treatment in melanoma but not with clinical outcome. Br. J. Cancer 2010, 103, 820–826. [Google Scholar] [CrossRef] [Green Version]
  123. Schraml, P.; von Teichman, A.; Mihic-Probst, D.; Simcock, M.; Ochsenbein, A.; Dummer, R.; Michielin, O.; Seifert, B.; Schläppi, M.; Moch, H.; et al. Predictive value of the MGMT promoter methylation status in metastatic melanoma patients receiving first-line temozolomide plus bevacizumab in the trial SAKK 50/07. Oncol. Rep. 2012, 28, 654–658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Robb, R.; Yang, L.; Shen, C.; Wolfe, A.R.; Webb, A.; Zhang, X.; Vedaie, M.; Saji, M.; Jhiang, S.; Ringel, M.D.; et al. Inhibiting BRAF Oncogene-Mediated Radioresistance Effectively Radiosensitizes BRAFV600E-Mutant Thyroid Cancer Cells by Constraining DNA Double-Strand Break Repair. Clin. Cancer Res. 2019, 25, 4749–4760. [Google Scholar] [CrossRef] [PubMed]
  125. Estrada-Bernal, A.; Chatterjee, M.; Haque, S.J.; Yang, L.; Morgan, M.A.; Kotian, S.; Morrell, D.; Chakravarti, A.; Williams, T.M. MEK inhibitor GSK1120212-mediated radiosensitization of pancreatic cancer cells involves inhibition of DNA double-strand break repair pathways. Cell Cycle 2015, 14, 3713–3724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Moriceau, G.; Hugo, W.; Hong, A.; Shi, H.; Kong, X.; Yu, C.C.; Koya, R.C.; Samatar, A.A.; Khanlou, N.; Braun, J.; et al. Tunable-combinatorial mechanisms of acquired resistance limit the efficacy of BRAF/MEK cotargeting but result in melanoma drug addiction. Cancer Cell 2015, 27, 240–256. [Google Scholar] [CrossRef] [Green Version]
  127. Das Thakur, M.; Salangsang, F.; Landman, A.S.; Sellers, W.R.; Pryer, N.K.; Levesque, M.P.; Dummer, R.; McMahon, M.; Stuart, D.D. Modelling vemurafenib resistance in melanoma reveals a strategy to forestall drug resistance. Nature 2013, 494, 251–255. [Google Scholar] [CrossRef] [PubMed]
  128. Nazarian, R.; Shi, H.; Wang, Q.; Kong, X.; Koya, R.C.; Lee, H.; Chen, Z.; Lee, M.K.; Attar, N.; Sazegar, H.; et al. Melanomas acquire resistance to B-RAF(V600E) inhibition by RTK or N-RAS upregulation. Nature 2010, 468, 973–977. [Google Scholar] [CrossRef] [Green Version]
  129. Hugo, W.; Shi, H.; Sun, L.; Piva, M.; Song, C.; Kong, X.; Moriceau, G.; Hong, A.; Dahlman, K.B.; Johnson, D.B.; et al. Non-genomic and Immune Evolution of Melanoma Acquiring MAPKi Resistance. Cell 2015, 162, 1271–1285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Bhattacharya, S.; Srinivasan, K.; Abdisalaam, S.; Su, F.; Raj, P.; Dozmorov, I.; Mishra, R.; Wakeland, E.K.; Ghose, S.; Mukherjee, S.; et al. RAD51 interconnects between DNA replication, DNA repair and immunity. Nucleic Acids Res. 2017, 45, 4590–4605. [Google Scholar] [CrossRef] [Green Version]
  131. Gopal, Y.N.; Deng, W.; Woodman, S.E.; Komurov, K.; Ram, P.; Smith, P.D.; Davies, M.A. Basal and treatment-induced activation of AKT mediates resistance to cell death by AZD6244 (ARRY-142886) in Braf-mutant human cutaneous melanoma cells. Cancer Res. 2010, 70, 8736–8747. [Google Scholar] [CrossRef] [Green Version]
  132. Deng, W.; Gopal, Y.N.; Scott, A.; Chen, G.; Woodman, S.E.; Davies, M.A. Role and therapeutic potential of PI3K-mTOR signaling in de novo resistance to BRAF inhibition. Pigment Cell Melanoma Res. 2012, 25, 248–258. [Google Scholar] [CrossRef]
  133. Shen, W.H.; Balajee, A.S.; Wang, J.; Wu, H.; Eng, C.; Pandolfi, P.P.; Yin, Y. Essential role for nuclear PTEN in maintaining chromosomal integrity. Cell 2007, 128, 157–170. [Google Scholar] [CrossRef] [Green Version]
  134. He, J.; Kang, X.; Yin, Y.; Chao, K.S.; Shen, W.H. PTEN regulates DNA replication progression and stalled fork recovery. Nat. Commun. 2015, 6, 7620. [Google Scholar] [CrossRef] [Green Version]
  135. Kang, X.; Song, C.; Du, X.; Zhang, C.; Liu, Y.; Liang, L.; He, J.; Lamb, K.; Shen, W.H.; Yin, Y. PTEN stabilizes TOP2A and regulates the DNA decatenation. Sci. Rep. 2015, 5, 17873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Sulkowski, P.L.; Scanlon, S.E.; Oeck, S.; Glazer, P.M. PTEN Regulates Nonhomologous End Joining By Epigenetic Induction of NHEJ1/XLF. Mol. Cancer Res. 2018, 16, 1241–1254. [Google Scholar] [CrossRef] [Green Version]
  137. Mendes-Pereira, A.M.; Martin, S.A.; Brough, R.; McCarthy, A.; Taylor, J.R.; Kim, J.S.; Waldman, T.; Lord, C.J.; Ashworth, A. Synthetic lethal targeting of PTEN mutant cells with PARP inhibitors. EMBO Mol. Med. 2009, 1, 315–322. [Google Scholar] [CrossRef] [PubMed]
  138. Turchick, A.; Liu, Y.; Zhao, W.; Cohen, I.; Glazer, P.M. Synthetic lethality of a cell-penetrating anti-RAD51 antibody in PTEN-deficient melanoma and glioma cells. Oncotarget 2019, 10, 1272–1283. [Google Scholar] [CrossRef]
  139. Aguissa-Touré, A.H.; Li, G. Genetic alterations of PTEN in human melanoma. Cell. Mol. Life Sci. 2012, 69, 1475–1491. [Google Scholar] [CrossRef]
  140. Zhou, X.P.; Gimm, O.; Hampel, H.; Niemann, T.; Walker, M.J.; Eng, C. Epigenetic PTEN silencing in malignant melanomas without PTEN mutation. Am. J. Pathol. 2000, 157, 1123–1128. [Google Scholar] [CrossRef] [Green Version]
  141. Mirmohammadsadegh, A.; Marini, A.; Nambiar, S.; Hassan, M.; Tannapfel, A.; Ruzicka, T.; Hengge, U.R. Epigenetic silencing of the PTEN gene in melanoma. Cancer Res. 2006, 66, 6546–6552. [Google Scholar] [CrossRef] [Green Version]
  142. Santamaría, P.G.; Floristán, A.; Fontanals-Cirera, B.; Vázquez-Naharro, A.; Santos, V.; Morales, S.; Yuste, L.; Peinado, H.; García-Gómez, A.; Portillo, F.; et al. Lysyl oxidase-like 3 is required for melanoma cell survival by maintaining genomic stability. Cell Death Differ. 2018, 25, 935–950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Mancuso, P.; Tricarico, R.; Bhattacharjee, V.; Cosentino, L.; Kadariya, Y.; Jelinek, J.; Nicolas, E.; Einarson, M.; Beeharry, N.; Devarajan, K.; et al. Thymine DNA glycosylase as a novel target for melanoma. Oncogene 2019, 38, 3710–3728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Burkel, F.; Jost, T.; Hecht, M.; Heinzerling, L.; Fietkau, R.; Distel, L. Dual mTOR/DNA-PK Inhibitor CC-115 Induces Cell Death in Melanoma Cells and Has Radiosensitizing Potential. Int. J. Mol. Sci. 2020, 21, 9321. [Google Scholar] [CrossRef]
  145. Seidel, J.A.; Otsuka, A.; Kabashima, K. Anti-PD-1 and Anti-CTLA-4 Therapies in Cancer: Mechanisms of Action, Efficacy, and Limitations. Front. Oncol. 2018, 8, 86. [Google Scholar] [CrossRef]
  146. Rowshanravan, B.; Halliday, N.; Sansom, D.M. CTLA-4: A moving target in immunotherapy. Blood 2018, 131, 58–67. [Google Scholar] [CrossRef] [PubMed]
  147. Chamoto, K.; Al-Habsi, M.; Honjo, T. Role of PD-1 in Immunity and Diseases. Curr. Top. Microbiol. Immunol. 2017, 410, 75–97. [Google Scholar] [CrossRef]
  148. Hodi, F.S.; O’Day, S.J.; McDermott, D.F.; Weber, R.W.; Sosman, J.A.; Haanen, J.B.; Gonzalez, R.; Robert, C.; Schadendorf, D.; Hassel, J.C.; et al. Improved survival with ipilimumab in patients with metastatic melanoma. N. Engl. J. Med. 2010, 363, 711–723. [Google Scholar] [CrossRef]
  149. Schachter, J.; Ribas, A.; Long, G.V.; Arance, A.; Grob, J.J.; Mortier, L.; Daud, A.; Carlino, M.S.; McNeil, C.; Lotem, M.; et al. Pembrolizumab versus ipilimumab for advanced melanoma: Final overall survival results of a multicentre, randomised, open-label phase 3 study (KEYNOTE-006). Lancet 2017, 390, 1853–1862. [Google Scholar] [CrossRef]
  150. Larkin, J.; Chiarion-Sileni, V.; Gonzalez, R.; Grob, J.J.; Cowey, C.L.; Lao, C.D.; Schadendorf, D.; Dummer, R.; Smylie, M.; Rutkowski, P.; et al. Combined Nivolumab and Ipilimumab or Monotherapy in Untreated Melanoma. N. Engl. J. Med. 2015, 373, 23–34. [Google Scholar] [CrossRef] [Green Version]
  151. Robert, C.; Schachter, J.; Long, G.V.; Arance, A.; Grob, J.J.; Mortier, L.; Daud, A.; Carlino, M.S.; McNeil, C.; Lotem, M.; et al. Pembrolizumab versus Ipilimumab in Advanced Melanoma. N. Engl. J. Med. 2015, 372, 2521–2532. [Google Scholar] [CrossRef]
  152. Hugo, W.; Zaretsky, J.M.; Sun, L.; Song, C.; Moreno, B.H.; Hu-Lieskovan, S.; Berent-Maoz, B.; Pang, J.; Chmielowski, B.; Cherry, G.; et al. Genomic and Transcriptomic Features of Response to Anti-PD-1 Therapy in Metastatic Melanoma. Cell 2016, 165, 35–44. [Google Scholar] [CrossRef] [Green Version]
  153. Wolchok, J.D.; Chiarion-Sileni, V.; Gonzalez, R.; Rutkowski, P.; Grob, J.J.; Cowey, C.L.; Lao, C.D.; Wagstaff, J.; Schadendorf, D.; Ferrucci, P.F.; et al. Overall Survival with Combined Nivolumab and Ipilimumab in Advanced Melanoma. N. Engl. J. Med. 2017, 377, 1345–1356. [Google Scholar] [CrossRef] [PubMed]
  154. Byrne, E.H.; Fisher, D.E. Immune and molecular correlates in melanoma treated with immune checkpoint blockade. Cancer 2017, 123, 2143–2153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Spranger, S.; Gajewski, T.F. Impact of oncogenic pathways on evasion of antitumour immune responses. Nat. Rev. Cancer 2018, 18, 139–147. [Google Scholar] [CrossRef] [PubMed]
  156. Paijens, S.T.; Vledder, A.; de Bruyn, M.; Nijman, H.W. Tumor-infiltrating lymphocytes in the immunotherapy era. Cell. Mol. Immunol. 2020. [Google Scholar] [CrossRef] [PubMed]
  157. Maibach, F.; Sadozai, H.; Seyed Jafari, S.M.; Hunger, R.E.; Schenk, M. Tumor-Infiltrating Lymphocytes and Their Prognostic Value in Cutaneous Melanoma. Front. Immunol. 2020, 11, 2105. [Google Scholar] [CrossRef]
  158. Tumeh, P.C.; Harview, C.L.; Yearley, J.H.; Shintaku, I.P.; Taylor, E.J.; Robert, L.; Chmielowski, B.; Spasic, M.; Henry, G.; Ciobanu, V.; et al. PD-1 blockade induces responses by inhibiting adaptive immune resistance. Nature 2014, 515, 568–571. [Google Scholar] [CrossRef] [PubMed]
  159. Galuppini, F.; Dal Pozzo, C.A.; Deckert, J.; Loupakis, F.; Fassan, M.; Baffa, R. Tumor mutation burden: From comprehensive mutational screening to the clinic. Cancer Cell Int. 2019, 19, 209. [Google Scholar] [CrossRef]
  160. Alexandrov, L.B.; Nik-Zainal, S.; Wedge, D.C.; Aparicio, S.A.; Behjati, S.; Biankin, A.V.; Bignell, G.R.; Bolli, N.; Borg, A.; Børresen-Dale, A.L.; et al. Signatures of mutational processes in human cancer. Nature. 2013, 500, 415–421. [Google Scholar] [CrossRef] [Green Version]
  161. Snyder, A.; Makarov, V.; Merghoub, T.; Yuan, J.; Zaretsky, J.M.; Desrichard, A.; Walsh, L.A.; Postow, M.A.; Wong, P.; Ho, T.S.; et al. Genetic basis for clinical response to CTLA-4 blockade in melanoma. N. Engl. J. Med. 2014, 371, 2189–2199. [Google Scholar] [CrossRef]
  162. Chalmers, Z.R.; Greenbowe, J.; Ali, S.M.; Balasubramanian, S.; Sun, J.X.; He, Y.; Frederick, D.T.; Puzanov, I.; Balko, J.M.; Cates, J.M.; et al. Targeted Next Generation Sequencing Identifies Markers of Response to PD-1 Blockade. Cancer Immunol. Res. 2016, 4, 959–967. [Google Scholar] [CrossRef] [Green Version]
  163. Van Allen, E.M.; Miao, D.; Schilling, B.; Shukla, S.A.; Blank, C.; Zimmer, L.; Sucker, A.; Hillen, U.; Foppen, M.H.G.; Goldinger, S.M.; et al. Genomic correlates of response to CTLA-4 blockade in metastatic melanoma. Science 2015, 350, 207–211. [Google Scholar] [CrossRef] [Green Version]
  164. Zaretsky, J.M.; Garcia-Diaz, A.; Shin, D.S.; Escuin-Ordinas, H.; Hugo, W.; Hu-Lieskovan, S.; Torrejon, D.Y.; Abril-Rodriguez, G.; Sandoval, S.; Barthly, L.; et al. Mutations Associated with Acquired Resistance to PD-1 Blockade in Melanoma. N. Engl. J. Med. 2016, 375, 819–829. [Google Scholar] [CrossRef]
  165. Lee, J.H.; Shklovskaya, E.; Lim, S.Y.; Carlino, M.S.; Menzies, A.M.; Stewart, A.; Pedersen, B.; Irvine, M.; Alavi, S.; Yang, J.; et al. Transcriptional downregulation of MHC class I and melanoma de- differentiation in resistance to PD-1 inhibition. Nat. Commun. 2020, 11, 1897. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Kümpers, C.; Jokic, M.; Haase, O.; Offermann, A.; Vogel, W.; Grätz, V.; Langan, E.A.; Perner, S.; Terheyden, P. Immune Cell Infiltration of the Primary Tumor, Not PD-L1 Status, Is Associated With Improved Response to Checkpoint Inhibition in Metastatic Melanoma. Front. Med. 2019, 6, 27. [Google Scholar] [CrossRef] [Green Version]
  167. Thomas, N.E.; Busam, K.J.; From, L.; Kricker, A.; Armstrong, B.K.; Anton-Culver, H.; Gruber, S.B.; Gallagher, R.P.; Zanetti, R.; Rosso, S.; et al. Tumor-infiltrating lymphocyte grade in primary melanomas is independently associated with melanoma-specific survival in the population-based genes, environment and melanoma study. J. Clin. Oncol. 2013, 31, 4252–4259. [Google Scholar] [CrossRef] [Green Version]
  168. Mandalà, M.; Merelli, B.; Massi, D. PD-L1 in melanoma: Facts and myths. Melanoma Manag. 2016, 3, 187–194. [Google Scholar] [CrossRef] [PubMed]
  169. Wang, Y.; Deng, W.; Li, N.; Neri, S.; Sharma, A.; Jiang, W.; Lin, S.H. Combining Immunotherapy and Radiotherapy for Cancer Treatment: Current Challenges and Future Directions. Front. Pharmacol. 2018, 9, 185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Biau, J.; Chautard, E.; Verrelle, P.; Dutreix, M. Altering DNA Repair to Improve Radiation Therapy: Specific and Multiple Pathway Targeting. Front. Oncol. 2019, 9, 1009. [Google Scholar] [CrossRef]
  171. Demaria, S.; Golden, E.B.; Formenti, S.C. Role of Local Radiation Therapy in Cancer Immunotherapy. JAMA Oncol. 2015, 1, 1325–1332. [Google Scholar] [CrossRef]
  172. Reits, E.A.; Hodge, J.W.; Herberts, C.A.; Groothuis, T.A.; Chakraborty, M.; Wansley, E.K.; Camphausen, K.; Luiten, R.M.; de Ru, A.H.; Neijssen, J.; et al. Radiation modulates the peptide repertoire, enhances MHC class I expression, and induces successful antitumor immunotherapy. J. Exp. Med. 2006, 203, 1259–1271. [Google Scholar] [CrossRef]
  173. Corso, C.D.; Ali, A.N.; Diaz, R. Radiation-induced tumor neoantigens: Imaging and therapeutic implications. Am. J. Cancer Res. 2011, 1, 390–412. [Google Scholar] [PubMed]
  174. Sharabi, A.B.; Lim, M.; DeWeese, T.L.; Drake, C.G. Radiation and checkpoint blockade immunotherapy: Radiosensitisation and potential mechanisms of synergy. Lancet Oncol. 2015, 16, e498–e509. [Google Scholar] [CrossRef]
  175. Durante, M.; Formenti, S.C. Radiation-Induced Chromosomal Aberrations and Immunotherapy: Micronuclei, Cytosolic DNA, and Interferon-Production Pathway. Front. Oncol. 2018, 8, 192. [Google Scholar] [CrossRef] [PubMed]
  176. Woo, S.R.; Fuertes, M.B.; Corrales, L.; Spranger, S.; Furdyna, M.J.; Leung, M.Y.; Duggan, R.; Wang, Y.; Barber, G.N.; Fitzgerald, K.A.; et al. STING-dependent cytosolic DNA sensing mediates innate immune recognition of immunogenic tumors. Immunity 2014, 41, 830–842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Hou, Y.; Liang, H.; Rao, E.; Zheng, W.; Huang, X.; Deng, L.; Zhang, Y.; Yu, X.; Xu, M.; Mauceri, H.; et al. Non-canonical NF-κB Antagonizes STING Sensor-Mediated DNA Sensing in Radiotherapy. Immunity 2018, 49, 490–503.e4. [Google Scholar] [CrossRef] [Green Version]
  178. Hopfner, K.P.; Hornung, V. Molecular mechanisms and cellular functions of cGAS-STING signalling. Nat. Rev. Mol. Cell. Biol. 2020, 21, 501–521. [Google Scholar] [CrossRef]
  179. Deng, L.; Liang, H.; Xu, M.; Yang, X.; Burnette, B.; Arina, A.; Li, X.D.; Mauceri, H.; Beckett, M.; Darga, T.; et al. STING-Dependent Cytosolic DNA Sensing Promotes Radiation-Induced Type I Interferon-Dependent Antitumor Immunity in Immunogenic Tumors. Immunity 2014, 41, 843–852. [Google Scholar] [CrossRef] [Green Version]
  180. Strojan, P. Role of radiotherapy in melanoma management. Radiol. Oncol. 2010, 44, 1–12. [Google Scholar] [CrossRef] [PubMed]
  181. Reijnen, C.; Küsters-Vandevelde, H.V.N.; Prinsen, C.F.; Massuger, L.F.A.G.; Snijders, M.P.M.L.; Kommoss, S.; Brucker, S.Y.; Kwon, J.S.; McAlpine, J.N.; Pijnenborg, J.M.A. Mismatch repair deficiency as a predictive marker for response to adjuvant radiotherapy in endometrial cancer. Gynecol. Oncol. 2019, 154, 124–130. [Google Scholar] [CrossRef]
  182. Dovedi, S.J.; Adlard, A.L.; Lipowska-Bhalla, G.; McKenna, C.; Jones, S.; Cheadle, E.J.; Stratford, I.J.; Poon, E.; Morrow, M.; Stewart, R.; et al. Acquired resistance to fractionated radiotherapy can be overcome by concurrent PD-L1 blockade. Cancer Res. 2014, 74, 5458–5468. [Google Scholar] [CrossRef] [Green Version]
  183. Derer, A.; Spiljar, M.; Bäumler, M.; Hecht, M.; Fietkau, R.; Frey, B.; Gaipl, U.S. Chemoradiation Increases PD-L1 Expression in Certain Melanoma and Glioblastoma Cells. Front. Immunol. 2016, 7, 610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Twyman-Saint Victor, C.; Rech, A.J.; Maity, A.; Rengan, R.; Pauken, K.E.; Stelekati, E.; Benci, J.L.; Xu, B.; Dada, H.; Odorizzi, P.M.; et al. Radiation and dual checkpoint blockade activate non-redundant immune mechanisms in cancer. Nature 2015, 520, 373–377. [Google Scholar] [CrossRef] [Green Version]
  185. Yilmaz, M.T.; Elmali, A.; Yazici, G. Abscopal Effect, From Myth to Reality: From Radiation Oncologists’ Perspective. Cureus 2019, 11, e3860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Harding, S.M.; Benci, J.L.; Irianto, J.; Discher, D.E.; Minn, A.J.; Greenberg, R.A. Mitotic progression following DNA damage enables pattern recognition within micronuclei. Nature 2017, 548, 466–470. [Google Scholar] [CrossRef] [Green Version]
  187. Stamell, E.F.; Wolchok, J.D.; Gnjatic, S.; Lee, N.Y.; Brownell, I. The abscopal effect associated with a systemic anti-melanoma immune response. Int. J. Radiat. Oncol. Biol. Phys. 2013, 85, 293–295. [Google Scholar] [CrossRef] [Green Version]
  188. Postow, M.A.; Callahan, M.K.; Barker, C.A.; Yamada, Y.; Yuan, J.; Kitano, S.; Mu, Z.; Rasalan, T.; Adamow, M.; Ritter, E.; et al. Immunologic correlates of the abscopal effect in a patient with melanoma. N. Engl. J. Med. 2012, 366, 925–931. [Google Scholar] [CrossRef] [Green Version]
  189. Park, S.S.; Dong, H.; Liu, X.; Harrington, S.M.; Krco, C.J.; Grams, M.P.; Mansfield, A.S.; Furutani, K.M.; Olivier, K.R.; Kwon, E.D. PD-1 Restrains Radiotherapy-Induced Abscopal Effect. Cancer Immunol. Res. 2015, 3, 610–619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Ribeiro Gomes, J.; Schmerling, R.A.; Haddad, C.K.; Racy, D.J.; Ferrigno, R.; Gil, E.; Zanuncio, P.; Buzaid, A.C. Analysis of the Abscopal Effect With Anti-PD1 Therapy in Patients With Metastatic Solid Tumors. J. Immunother. 2016, 39, 367–372. [Google Scholar] [CrossRef] [PubMed]
  191. Trommer, M.; Yeo, S.Y.; Persigehl, T.; Bunck, A.; Grüll, H.; Schlaak, M.; Theurich, S.; von Bergwelt-Baildon, M.; Morgenthaler, J.; Herter, J.M.; et al. Abscopal Effects in Radio-Immunotherapy-Response Analysis of Metastatic Cancer Patients With Progressive Disease Under Anti-PD-1 Immune Checkpoint. Front. Pharmacol. 2019, 10, 511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. ElJalby, M.; Pannullo, S.C.; Schwartz, T.H.; Parashar, B.; Wernicke, A.G. Optimal Timing and Sequence of Immunotherapy When Combined with Stereotactic Radiosurgery in the Treatment of Brain Metastases. World Neurosurg. 2019, 127, 397–404. [Google Scholar] [CrossRef] [PubMed]
  193. Qian, J.M.; Yu, J.B.; Kluger, H.M.; Chiang, V.L. Timing and type of immune checkpoint therapy affect the early radiographic response of melanoma brain metastases to stereotactic radiosurgery. Cancer 2016, 122, 3051–3058. [Google Scholar] [CrossRef] [Green Version]
  194. Patel, K.R.; Shoukat, S.; Oliver, D.E.; Chowdhary, M.; Rizzo, M.; Lawson, D.H.; Khosa, F.; Liu, Y.; Khan, M.K. Ipilimumab and Stereotactic Radiosurgery Versus Stereotactic Radiosurgery Alone for Newly Diagnosed Melanoma Brain Metastases. Am. J. Clin. Oncol. 2017, 40, 444–450. [Google Scholar] [CrossRef]
  195. Knispel, S.; Stang, A.; Zimmer, L.; Lax, H.; Gutzmer, R.; Heinzerling, L.; Weishaupt, C.; Pföhler, C.; Gesierich, A.; Herbst, R.; et al. Impact of a preceding radiotherapy on the outcome of immune checkpoint inhibition in metastatic melanoma: A multicenter retrospective cohort study of the DeCOG. J. Immunother. Cancer 2020, 8, e000395. [Google Scholar] [CrossRef] [PubMed]
  196. Mouw, K.W.; Goldberg, M.S.; Konstantinopoulos, P.A.; D’Andrea, A.D. DNA Damage and Repair Biomarkers of Immunotherapy Response. Cancer Discov. 2017, 7, 675–693. [Google Scholar] [CrossRef] [Green Version]
  197. Zhang, J.; Shih, D.J.H.; Lin, S.Y. Role of DNA repair defects in predicting immunotherapy response. Biomark. Res. 2020, 8, 23. [Google Scholar] [CrossRef]
  198. Le, D.T.; Uram, J.N.; Wang, H.; Bartlett, B.R.; Kemberling, H.; Eyring, A.D.; Skora, A.D.; Luber, B.S.; Azad, N.S.; Laheru, D.; et al. PD-1 Blockade in Tumors with Mismatch-Repair Deficiency. N. Engl. J. Med. 2015, 372, 2509–2520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Kelderman, S.; Schumacher, T.N.; Kvistborg, P. Mismatch Repair-Deficient Cancers Are Targets for Anti-PD-1 Therapy. Cancer Cell 2015, 28, 11–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Lemery, S.; Keegan, P.; Pazdur, R. First FDA Approval Agnostic of Cancer Site—When a Biomarker Defines the Indication. N. Engl. J. Med. 2017, 377, 1409–1412. [Google Scholar] [CrossRef]
  201. Conway, J.R.; Dietlein, F.; Taylor-Weiner, A.; AlDubayan, S.; Vokes, N.; Keenan, T.; Reardon, B.; He, M.X.; Margolis, C.A.; Weirather, J.L.; et al. Integrated molecular drivers coordinate biological and clinical states in melanoma. Nat. Genet. 2020, 52, 1373–1383. [Google Scholar] [CrossRef]
  202. Ponti, G.; Pellacani, G.; Tomasi, A.; Depenni, R.; Maccaferri, M.; Maiorana, A.; Orsi, G.; Giusti, F.; Cascinu, S.; Manfredini, M. Immunohistochemical mismatch repair proteins expression as a tool to predict the melanoma immunotherapy response. Mol. Clin. Oncol. 2020, 12, 3–8. [Google Scholar] [CrossRef] [Green Version]
  203. Mandal, R.; Samstein, R.M.; Lee, K.W.; Havel, J.J.; Wang, H.; Krishna, C.; Sabio, E.Y.; Makarov, V.; Kuo, F.; Blecua, P.; et al. Genetic diversity of tumors with mismatch repair deficiency influences anti-PD-1 immunotherapy response. Science 2019, 364, 485–491. [Google Scholar] [CrossRef]
  204. Biau, J.; Devun, F.; Jdey, W.; Kotula, E.; Quanz, M.; Chautard, E.; Sayarath, M.; Sun, J.S.; Verrelle, P.; Dutreix, M. A preclinical study combining the DNA repair inhibitor Dbait with radiotherapy for the treatment of melanoma. Neoplasia 2014, 16, 835–844. [Google Scholar] [CrossRef] [Green Version]
  205. Le Tourneau, C.; Dreno, B.; Kirova, Y.; Grob, J.J.; Jouary, T.; Dutriaux, C.; Thomas, L.; Lebbé, C.; Mortier, L.; Saiag, P.; et al. First-in-human phase I study of the DNA-repair inhibitor DT01 in combination with radiotherapy in patients with skin metastases from melanoma. Br. J. Cancer 2016, 114, 1199–1205. [Google Scholar] [CrossRef] [PubMed]
  206. Higuchi, T.; Flies, D.B.; Marjon, N.A.; Mantia-Smaldone, G.; Ronner, L.; Gimotty, P.A.; Adams, S.F. CTLA-4 Blockade Synergizes Therapeutically with PARP Inhibition in BRCA1-Deficient Ovarian Cancer. Cancer Immunol. Res. 2015, 3, 1257–1268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Jiao, S.; Xia, W.; Yamaguchi, H.; Wei, Y.; Chen, M.K.; Hsu, J.M.; Hsu, J.L.; Yu, W.H.; Du, Y.; Lee, H.H.; et al. PARP Inhibitor Upregulates PD-L1 Expression and Enhances Cancer-Associated Immunosuppression. Clin. Cancer Res. 2017, 23, 3711–3720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Jdey, W.; Thierry, S.; Russo, C.; Devun, F.; Al Abo, M.; Noguiez-Hellin, P.; Sun, J.S.; Barillot, E.; Zinovyev, A.; Kuperstein, I.; et al. Drug-Driven Synthetic Lethality: Bypassing Tumor Cell Genetics with a Combination of AsiDNA and PARP Inhibitors. Clin. Cancer Res. 2017, 23, 1001–1011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Clinical Trials. Available online: https://clinicaltrials.gov/ct2/home (accessed on 23 March 2021).
Figure 1. Melanoma cells treated with BRAF and MEK inhibitors depend on efficient DNA repair and genome stability for survival. In drug-naïve melanoma cells, inhibitors of BRAF (BRAFi) and MEK (MEKi) reduce the expression of DNA repair related genes, which can be enhanced by PARP and HDAC inhibitors (PARPi, HDACi) to increase melanoma cell death. Increased expression of NGFR, AXL and JARID1B protects the genome of melanoma cells intrinsically resistant to MAPK inhibitors (MAPKi). Such intrinsically drug-resistant melanoma cells are selected by MAPKi treatment and lead to disease progression. Occurring in parallel, development of acquired drug resistance through reactivation of ERK signaling may result in excessive ERK activity upon immediate drug withdrawal and DNA damage, which can be enhanced by DNA repair inhibitors such as PARP inhibitor (PARPi). Inhibitory effect is marked by red bar-headed arrows.
Figure 1. Melanoma cells treated with BRAF and MEK inhibitors depend on efficient DNA repair and genome stability for survival. In drug-naïve melanoma cells, inhibitors of BRAF (BRAFi) and MEK (MEKi) reduce the expression of DNA repair related genes, which can be enhanced by PARP and HDAC inhibitors (PARPi, HDACi) to increase melanoma cell death. Increased expression of NGFR, AXL and JARID1B protects the genome of melanoma cells intrinsically resistant to MAPK inhibitors (MAPKi). Such intrinsically drug-resistant melanoma cells are selected by MAPKi treatment and lead to disease progression. Occurring in parallel, development of acquired drug resistance through reactivation of ERK signaling may result in excessive ERK activity upon immediate drug withdrawal and DNA damage, which can be enhanced by DNA repair inhibitors such as PARP inhibitor (PARPi). Inhibitory effect is marked by red bar-headed arrows.
Ijms 22 03485 g001
Figure 2. Mechanisms of irradiation-induced stimulation of immune system. Ionizing radiation (IR) leads to DNA lesions that may result in mutant proteins and neoantigen presentation via MHC molecule for T cells to recognize. IR-induced DNA damage may cause cytosolic DNA or micronuclei (MN) accumulation and secretion, which is taken up by antigen presenting cells, and leads to cGAS-STING pathway-mediated type I interferon production and secretion. Finally, severe damage by IR can cause cell death and release of DAMP molecules, which are processed by antigen presenting cells and presented to T cells leading to T cell priming. Immunoinhibitory interaction of PD-L1 expressed by melanoma cells with PD-1 present on T cells can be targeted with nivolumab and pembrolizumab, while CD80/CD86 and CTLA-4 interaction between antigen presenting cell and T cell can be inhibited by ipilimumab. Inhibitors of IR-induced DNA repair are likely to augment the genome destabilizing immunostimulatory effect of IR. Inhibitory effect is marked by red bar-headed arrows.
Figure 2. Mechanisms of irradiation-induced stimulation of immune system. Ionizing radiation (IR) leads to DNA lesions that may result in mutant proteins and neoantigen presentation via MHC molecule for T cells to recognize. IR-induced DNA damage may cause cytosolic DNA or micronuclei (MN) accumulation and secretion, which is taken up by antigen presenting cells, and leads to cGAS-STING pathway-mediated type I interferon production and secretion. Finally, severe damage by IR can cause cell death and release of DAMP molecules, which are processed by antigen presenting cells and presented to T cells leading to T cell priming. Immunoinhibitory interaction of PD-L1 expressed by melanoma cells with PD-1 present on T cells can be targeted with nivolumab and pembrolizumab, while CD80/CD86 and CTLA-4 interaction between antigen presenting cell and T cell can be inhibited by ipilimumab. Inhibitors of IR-induced DNA repair are likely to augment the genome destabilizing immunostimulatory effect of IR. Inhibitory effect is marked by red bar-headed arrows.
Ijms 22 03485 g002
Table 1. Ongoing clinical trials in the treatment of melanoma that combine MAPK inhibitors or immune checkpoint inhibitors with chemotherapy, radiotherapy or drugs targeting proteins of relevance to genome stability.
Table 1. Ongoing clinical trials in the treatment of melanoma that combine MAPK inhibitors or immune checkpoint inhibitors with chemotherapy, radiotherapy or drugs targeting proteins of relevance to genome stability.
IdentifierPhaseEnrollmentPrimary Outcome MeasuresTreatment Regimen Description
NCT01676649 1230adverse eventsipilimumab + carboplatin + paclitaxel
NCT02097732 124LCRSRS + ipilimumab
NCT0239287 11/210adverse events and radiotherapy associated toxicitiesradiotherapy + dabrafenib + trametinib
NCT02617849 1230ORRcarboplatin + paclitaxel + pembrolizumab
NCT02718066 21/2118RP2DHBI-8000 (HDACi) + nivolumab
NCT0281602 2271ORRazacytidine(cytidine analog) + pembrolizumab
NCT02872259 21/292ORRBGB324 (AXL inhibitor) + pembrolizumab
BGB324 + dabrafenib + trametinib
NCT02974803 126intracranial ORSRS + dabrafenib + trametinib
NCT02978404 1226intracranial PFSnivolumab + radiosurgery
NCT02988817 21/2374DLTs, adverse eventsenapotamab vedotin (HuMax-AXL-ADC)
NCT03050060 22129ORRnelfinavir mesylate + pembrolizumab, nivolumab, or atezolizumab + hypofractionated radiation therapy
NCT03278665 21/240IAE4SC-202 (HDACi) + pembrolizumab
NCT03340129 22218intracranial response to immunotherapynivolumab + ipilimumab + SRS
NCT03425279 21/2120DLTs, MTD, ORRCAB-AXL-ADC (anti-AXL antibody drug conjugate)
NCT03430947 2232ORR in brainradiosurgery + vemurafenib + cobimetinib
NCT03448666 3253ORRelectrochemotherapy + pembrolizumab
NCT03474497 21/245ARRpembrolizumab + IL-2 + hypofractionated radiotherapy.
NCT03511391 1299PFSnivolumab or pembrolizumab or atezolizumab + SBRT
NCT03646617 2270number of adverse eventsipilumumab + nivolumab + HFRT
NCT03693014 2260ORRSBRT + ipilimumab, nivolumab, pembrolizumab or atezolizumab
NCT03765229 2214ORR, PFSentinostat (HDACi) + pembrolizumab
NCT03780608 1261ORRceralasertib (ATR inhibitor) + durvalumab (PD-1/PD-L1 inhibitor)
NCT03898908 2238intracranial ORRencorafenib + binimetinib + radiation
NCT03925350 2241ORRniraparib (PARPi)
NCT03958383 21/261IAE, MTD, MADradiation + nivolumab + ipilimumab + hu14.18-IL2
NCT04017897 2252ORRpembrolizumab or nivolumab + radiotherapy
NCT04042506 2215safety of SBRTSBRT + nivolumab
NCT04074096 32150intracranial PFSSRS + encorafenib + binimetinib
NCT04133948 21/245safety of patientsdomatinostat (HDACi) + nivolumab + ipilimumab
NCT04187833 2237best overall response (CR + PR)nivolumab + talazoparib (PARPi)
NCT04225390 2238CR, PR, SD or PDDTIC + re-exposure to immunotherapy
NCT04594187 23168time to regional nodal recurrencenodal radiation therapy + immunotherapy
NCT04620603 31/215tumor responselow dose rate brachytherapy + nivolumab
NCT04633902 3241ORRolaparib (PARPi) + pembrolizumab
NCT04793737 2N/A27ORRprecision radiation in patients on PD-1 inhibitor treatment that have tumor progression
1 active, not recruiting; 2 recruiting, 3 not yet recruiting; ACT, adoptive cell therapy; ARR, abscopal response rate; CR, complete response; DLTs, dosage-limiting toxicities; EFS, event-free survival; HFRT, hypofractionated radiotherapy; IAE, incidence of adverse effects; LCR, local control rate; MAD, maximum administered dose, MTD, maximum tolerated dose, N/A, not applicable; OR, objective/overall response; ORR, objective response rate; PD, progressive disease; PFS, progression-free survival; PR, partial response; RP2D, recommended phase II dose; SBRT, stereotactic body radiation therapy; SD, stable disease; SRS, stereotactic radiotherapy.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Osrodek, M.; Wozniak, M. Targeting Genome Stability in Melanoma—A New Approach to an Old Field. Int. J. Mol. Sci. 2021, 22, 3485. https://doi.org/10.3390/ijms22073485

AMA Style

Osrodek M, Wozniak M. Targeting Genome Stability in Melanoma—A New Approach to an Old Field. International Journal of Molecular Sciences. 2021; 22(7):3485. https://doi.org/10.3390/ijms22073485

Chicago/Turabian Style

Osrodek, Marta, and Michal Wozniak. 2021. "Targeting Genome Stability in Melanoma—A New Approach to an Old Field" International Journal of Molecular Sciences 22, no. 7: 3485. https://doi.org/10.3390/ijms22073485

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop