Genomic stability in normal cells is crucial to avoid oncogenesis. Accordingly, multiple components of the DNA damage response (DDR) operate as bona fide tumor suppressor proteins by preserving genomic stability, eliciting the demise of cells with unrepairable DNA lesions, and engaging cell-extrinsic oncosuppression via immunosurveillance. That said, DDR sig­naling can also favor tumor progression and resistance to therapy. Indeed, DDR signaling in cancer cells has been consistently linked to the inhibition of tumor-targeting immune responses. Here, we discuss the complex interactions between the DDR and inflammation in the context of oncogenesis, tumor progression, and response to therapy.

Significance:

Accumulating preclinical and clinical evidence indicates that DDR is intimately connected to the emission of immunomodulatory signals by normal and malignant cells, as part of a cell-extrinsic program to preserve organismal homeostasis. DDR-driven inflammation, however, can have diametrically opposed effects on tumor-targeting immunity. Understanding the links between the DDR and inflammation in normal and malignant cells may unlock novel immunotherapeutic paradigms to treat cancer.

According to widely accepted models, oncogenesis as the result of malignant transformation (i.e., the conversion of a healthy cell into a malignant cell precursor) followed by tumor progression (i.e., the expansion of malignant cells to generate clinically relevant neoplasms) involves two major components (1–3). On the one hand, genetic and/or epigenetic defects provide healthy cells with the ability to evade cell-intrinsic oncosuppressive mechanisms that would otherwise prevent transformation, such as regulated cell death (RCD; ref. 1). On the other hand, newly formed malignant cell precursors acquire additional genetic and/or epigenetic alterations that allow for the evasion of cell-extrinsic oncosuppression, notably anticancer immunosurveillance (2, 3). In line with this notion, numerous oncoproteins (e.g., KRAS, MYCN) not only provide cancer cell precursors with aggressive features linked to malignancy (e.g., accrued proliferative potential) but also favor immunoevasion and/or the establishment of an immunosuppressive microenvironment (4). Conversely, various oncosuppressor proteins (e.g., TP53, PTEN) not only support the preservation of homeostasis in healthy cells and/or favor the demise of cells damaged beyond repair (which are prone to undergo malignant transformation) but also mediate immunostimulatory effects that ultimately engage anticancer immunosurveillance (4). Cell-intrinsic and cell-extrinsic oncosuppression mechanisms are so efficient that only a few (if any) clinically manifest neoplasms develop in the lifespan of an individual despite abundant somatic mutagenesis in virtually every cell of the body (5, 6).

In healthy cells, the DNA damage response (DDR; Fig. 1) mediates robust oncosuppressive functions that involve both such cell-intrinsic and cell-extrinsic components. First, efficient DNA repair de facto prevents the accumulation of genetic defects that may initiate malignant transformation. Second, if DNA damage cannot be satisfactorily repaired, the DDR actively engages RCD programs that ensure the demise of cells at high risk for malignant transformation. Finally, some components of the DDR machinery directly regulate the emission of immunostimulatory signals that recruit or activate immune cells, hence supporting the elimination of potentially oncogenic cells (7). For instance, oncogene activation in healthy hepatocytes has been shown to stimulate tumor protein p53 (TP53, best known as p53)-dependent cellular senescence, coupled to the secretion of chemokines and cytokines that ultimately drive an innate and adaptive immune response against senescent cells (8, 9). Supporting the critical oncosuppressive role of the DDR, individuals with xeroderma pigmentosum (XP), an autosomal recessive disease driven by germline mutations affecting nucleotide excision repair (NER), exhibit a severe predisposition to ultraviolet (UV) light-driven tumors (10). Moreover, multiple bona fide tumor suppressors are intimately involved in DNA repair and/or the activation of RCD or cellular senescence downstream of failed DNA repair. These proteins, which are lost or mutated in a high percentage of human tumors, include (but are not limited to): p53 (11), ATM serine/threonine kinase (ATM; ref. 12), BRCA1 DNA repair associated (BRCA1; ref. 13), BRCA2 (13), and phosphatase and tensin homolog (PTEN; ref. 14). In line with this notion, virtually all established tumors exhibit at least some DDR defects (Table 1; ref. 15), which de facto provides the scientific rationale for using DNA-damaging agents including conventional chemotherapeutics and radiotherapy (RT) for clinical cancer management (16).

Figure 1.

Major DNA repair mechanisms and DDR signaling. Mammalian cells respond to specific DNA lesions by activating dedicated mechanisms that operate, at least initially, toward DNA repair and the preservation of cellular homeostasis. Such mechanisms include (but are not limited to) base excision repair (BER), mismatch repair (MMR), nonhomologous end-joining (NHEJ), homologous recombination (HR), nucleotide excision repair (NER), R-loop processing, the Fanconi anemia (FA) pathway, and replication fork processing. The DDR is also connected with the control of general processes including cell-cycle progression as part of a broader, cell-wide response to DNA damage. Importantly, the terminal effector of DDR signaling including p53 and p38MAPK can also promote the terminal proliferative inactivation (cellular senescence) or regulated demise of cells bearing irreparable DNA damage. Several DDR components are being explored as potential targets for the development of novel anticancer drugs. Only DDR factors mentioned in the text are depicted in the figure. DSB, DNA double-strand break; ssDNA, single-strand DNA.

Figure 1.

Major DNA repair mechanisms and DDR signaling. Mammalian cells respond to specific DNA lesions by activating dedicated mechanisms that operate, at least initially, toward DNA repair and the preservation of cellular homeostasis. Such mechanisms include (but are not limited to) base excision repair (BER), mismatch repair (MMR), nonhomologous end-joining (NHEJ), homologous recombination (HR), nucleotide excision repair (NER), R-loop processing, the Fanconi anemia (FA) pathway, and replication fork processing. The DDR is also connected with the control of general processes including cell-cycle progression as part of a broader, cell-wide response to DNA damage. Importantly, the terminal effector of DDR signaling including p53 and p38MAPK can also promote the terminal proliferative inactivation (cellular senescence) or regulated demise of cells bearing irreparable DNA damage. Several DDR components are being explored as potential targets for the development of novel anticancer drugs. Only DDR factors mentioned in the text are depicted in the figure. DSB, DNA double-strand break; ssDNA, single-strand DNA.

Close modal
Table 1.

Genetic DDR alterations in cancera

Cancer typeGenePrevalenceCohort sizeRef.
Bladder cancer ERCC2 12% 130 (288
Bladder cancer TP53 49% 130 (288
Breast cancerb BRCA1 1.36% 2,433 (289
Breast cancerb BRCA2 1.64% 2,433 (289
Breast cancerb CHEK2 2.22% 2,433 (289
Breast cancerb TP53 35.4% 2,433 (289
Colorectal cancer ATM 12% 145 (290
Colorectal cancer TP53 67% 145 (290
Endometrial cancer TP53 30% 91 (291
Hematologic malignancies TP53 10% 3,096 (292
Hepatocellular carcinoma TP53 31% 363 (293
Lung cancer TP53 46% 230 (294
Melanoma TP53 15% 318 (295
Pancreatic cancer ATM 5% 150 (296
Pancreatic cancer TP53 72% 150 (296
Prostate cancer ATM 6% 333 (297
Prostate cancer TP53 8% 333 (297
Prostate cancer BRCA2 3% 333 (297
Thyroid cancer ATM 1.2% 496 (298
Thyroid cancer CHEK2 1.2% 496 (298
Cancer typeGenePrevalenceCohort sizeRef.
Bladder cancer ERCC2 12% 130 (288
Bladder cancer TP53 49% 130 (288
Breast cancerb BRCA1 1.36% 2,433 (289
Breast cancerb BRCA2 1.64% 2,433 (289
Breast cancerb CHEK2 2.22% 2,433 (289
Breast cancerb TP53 35.4% 2,433 (289
Colorectal cancer ATM 12% 145 (290
Colorectal cancer TP53 67% 145 (290
Endometrial cancer TP53 30% 91 (291
Hematologic malignancies TP53 10% 3,096 (292
Hepatocellular carcinoma TP53 31% 363 (293
Lung cancer TP53 46% 230 (294
Melanoma TP53 15% 318 (295
Pancreatic cancer ATM 5% 150 (296
Pancreatic cancer TP53 72% 150 (296
Prostate cancer ATM 6% 333 (297
Prostate cancer TP53 8% 333 (297
Prostate cancer BRCA2 3% 333 (297
Thyroid cancer ATM 1.2% 496 (298
Thyroid cancer CHEK2 1.2% 496 (298

aLimited to most common tumor types.

bSomatic alterations only.

That said, DDR signaling in transformed cells can also favor tumor progression and resistance to therapy. For instance, some tumors that arise as a consequence of genetic DDR defects develop non-oncogene addiction to alternative/complementary DDR pathways as a means to limit genetic instability to degrees that are compatible with cell survival (17–19). This is the case of malignancies bearing BRCA1 or BRCA2 mutations, which require poly(ADP-ribose) polymerase 1 (PARP1) for survival (and hence are exquisitely sensitive to PARP inhibitors in a setting of synthetic lethality; ref. 20). Moreover, although cancer cells exhibit DDR alterations that enable the use of DNA-damaging agents in the clinic (16), numerous DDR inhibitors including (but not limited to) agents specific for ATM, ATR serine/threonine kinase (ATR), checkpoint kinase 1 (CHEK1), and WEE1 G2 checkpoint kinase (WEE1) are being investigated as chemo- or radiosensitizers, at least in part based on their ability to compromise DNA repair and thus promote therapy-driven RCD or cellular senescence (Fig. 1; refs. 20, 21). Finally, the hyperproliferative phenotype of malignant cells as well as the relatively adverse conditions of the tumor microenvironment (TME) is per se associated with high degrees of DNA damage, requiring at least some DDR activity not only to limit potentially lethal genetic drifts but also to contain inflammatory and immune reactions that are elicited by the unwarranted accumulation of nucleic acids in ectopic subcellular compartments and may result in the activation of anticancer immunosurveillance (22–24). In line with this notion, genetic signatures of poor tumor infiltration by CD8+ T cells (a class of lymphocytes that are critical for anticancer immunity) have recently been linked with an increased expression of DNA repair genes in various cohorts of patients with pediatric and adult tumors (25). Finally, XP-related tumors (which are driven by UV light in the context of genetic NER defects, see above) are susceptible to nonspecific immunostimulants including the Toll-like receptor 7 (TLR7) agonist imiquimod (26), suggesting the presence of an immunologically active TME at baseline.

That said, although a proficient DDR is strictly beneficial for healthy cells as it ensures oncosuppression, progressing neoplasms may obtain superior benefits from incomplete DNA repair and subcytotoxic DDR activation. Indeed, suboptimal DNA repair and DDR signaling enable the rapid accumulation of mutations that (when compatible with cell survival and unable to drive RCD) foster tumor heterogeneity and clonal diversification, generally endowing the tumor ecosystem as a whole with accrued capacities for adaptation to environmental and therapeutic challenges that operate as selective pressures according to Darwinian principles (27). This is particularly true for defects in DNA mismatch repair (MMR) that culminate with the establishment of microsatellite instability (MSI), a condition of genetic hypermutability that provides progressing tumors with a superior potential for adaptation (27).

Importantly, inflammatory and immune reactions (inclu­ding those elicited or inhibited by the DDR) may have opposed effects on tumor progression and response to therapy (22). On the one hand, efficient anticancer immunosurveillance (be it natural or elicited by therapy) relies on the initiation of acute and robust but ultimately resolving inflammatory reactions that involve activated immune effector cells (2, 3). Such effector cells include innate natural killer (NK) cells (which may be involved in early anticancer immunosurveillance as well as in the control of metastatic cancer dissemination; ref. 28), as well as adaptive, tumor-specific CD4+ and CD8+ T cells (which are critical for the optimal efficacy of numerous treatment regimens including conventional chemotherapy, RT and targeted anticancer agents; refs. 29–31). On the other hand, chronic, indolent and nonresolving inflammatory reactions have been shown to promote tumor progression and resistance to treatment (32). Indeed, nonresolving inflammation (as in nonhealing wounds) is generally associated with the establishment of a microenvironment that is dominated by immunosuppressive cytokines such as transforming growth factor beta 1 (TGFβ1) and interleukin 10 (IL10; ref. 33), as well as by regulatory cell populations, including so-called “M2-like” tumor-associated macrophages (TAM; ref. 34) and CD4+CD25+FOXP3+ regulatory T cells (Treg; ref. 35), coupled to the functional exhaustion of effector immune cells (36).

Here, we critically discuss the intricate links between the DDR and inflammation in the context of malignant transformation, tumor progression, and response to treatment, with emphasis on strategies harnessing DDR modulators to alter the immunological tumor configuration for therapeutic purposes. Mechanistic aspects of DNA repair and the DDR unrelated to inflammation have been recently reviewed elsewhere (23, 37) and hence will not be covered by this review.

Base excision repair (BER) is a DDR pathway activated by DNA lesions that do not considerably distort the tertiary structure of the double helix, notably bases that are oxidized by reactive oxygen species (ROS) or reactive nitrogen species (RNS), two major sources of DNA damage (38, 39). BER initiates with the removal of damaged/oxidized or misincorporated bases by DNA glycosidases, such as 8-oxoguanine DNA glycosylase (OGG1) and mutY DNA glycosylase (MUTYH), resulting in the generation of abasic sites that are detected and excised by apurinic/apyrimidinic endodeoxyribonuclease 1 (APEX1, best known as APE1). This culminates with the formation of single-strand breaks (SSB) that are sensed by PARP1 (see below), leading to the recruitment of additional BER and SSB repair factors to damaged sites (40). Finally, DNA synthesis by DNA polymerase beta (POLB) or delta (POLD), coupled to DNA ligation by DNA ligase 1 (LIG1) or LIG3 in complex with X-ray repair cross complementing 1 (XRCC1), can restore the original DNA sequence (Fig. 1). Although defects in BER have only sporadically been associated with human oncogenesis (40), multiple BER components have been shown to regulate inflammatory reactions in healthy or malignant cells.

For instance, OGG1 appears to mediate inflammatory functions in a variety of disease models, prompting the development of pharmacologic OGG1 inhibitors for use as nonsteroidal anti-inflammatory drugs (41). In line with this notion, the lack of Ogg1 limits inflammatory lesions induced in mice orally inoculated with the carcinogenic bacteria Helicobacter pylori, correlating with reduced serological signs of inflammation and blunted mutagenesis (42). However, Ogg1−/− mice are more susceptible to colorectal carcinogenesis as driven by azoxymethane (AOM) plus dextran sodium sulfate (DSS), which—similar to H. pylori-driven gastric carcinogenesis—involves a considerable inflammatory component, than their wild-type (WT) counterparts (43, 44). Similar findings have been obtained in mice lacking N-methylpurine DNA glycosylase (MPG), another glycosidase involved in BER (45). Conversely, Mutyh−/− mice appear to be protected from DSS/AOM-driven carcinogenesis (46). That said, the whole-body codeletion of Ogg1 and Mutyh considerably shortens the mouse lifespan, correlating with a considerable increase in the propensity for spontaneous oncogenesis, two effects that are further aggravated by the additional deletion of Msh2 (coding for an MMR component; ref. 47). Taken together, these findings exemplify the critical role of BER as a suppressor of oncogenesis driven by ROS and RNS. Accordingly, both OGG1 and MUTYH are mutated in a wide panel of human malignancies (48, 49).

APE1 is active in gastric epithelial cells infected by H. pylori, resulting in the activation of proinflammatory transcription factors such as AP-1 and NF-κB, interleukin 8 (IL8) production, and the establishment of an inflammatory response that supports not only gastric ulcer but also gastric transformation (50). Such a proinflammatory activity of APE1 also emerges in the context of tumor necrosis factor (TNF) signaling, which is key for organismal responses to chronic H. pylori infection and gastric carcinogenesis (51). APE1 also appears to be required for ROS generation in gastric epithelial cells exposed to H. pylori (52), knowing that (i) ROS and RNS are prominent sources of DNA damage and mutagens (39), (ii) ROS are potent activator of the NLR family pyrin domain containing 3 (NLRP3) inflammasome (53), and (iii) NLRP3 signaling coupled to IL1β and IL18 secretion has been linked to H. pylori pathogenicity (54). Thus, APE1-driven inflammatory reactions appear to contribute to gastric carcinogenesis in the context of chronic H. pylori infection.

Interestingly, APE1 drives tumor-promoting inflammation (TPI) largely because of its role in redox responses, rather than through its functions in BER. Moreover, it seems that the ability of APE1 to drive TPI may also involve its ability to regulate redox homeostasis in macrophages (55) and helper T cells (56). In line with this notion, POLB (which operates downstream of APE1 in BER) reportedly mediates robust anti-inflammatory, rather than proinflammatory, functions. Specifically, expression of a POLB variant that is associated with human gastric carcinoma (L22P) in mouse gastric epithelial cells not only favors chromosomal instability but also establishes TPI and promotes gastric carcinogenesis both as a standalone intervention (57) and in the context of H. pylori infection (58). Finally, depletion of POLB from human mammary carcinoma MCF7 cells drives potent type I interferon (IFN) synthesis downstream of faulty DNA repair, accrued cytosolic accumulation of double-strand DNA (dsDNA), and consequent activation of cyclic GMP-AMP synthase (CGAS), resulting in superior cancer cell sensitivity to elimination by NK cells (59). These findings suggest that proficient BER is key for preserving genetic integrity and inhibiting unwarranted inflammatory reactions. In line with this notion, low levels of XRCC1 correlated with increased tumor infiltration by CD8+ T cells, albeit in the context of high tumor grade and other features of aggressiveness, in two large cohorts of patients with breast cancer (60).

These observations exemplify the critical impact of BER components (via BER-dependent or independent mechanisms) on inflammatory responses in healthy and malignant tissues (Supplementary Table S1).

MMR resolves base mismatches and short insertions or deletions (61). Such alterations are detected by heterodimers consisting of mutS homolog 2 (MSH2) and MSH6 (known as MutSα), or MSH2 and MSH3 (known as MutSβ). In turn, MutSα (or MutSβ) provides a platform for mutL homolog 1 (MLH1) and PMS1 homolog 2, mismatch-repair system component (PMS2) heterodimers, an endonuclease complex stimulated by proliferating cell nuclear antigen (PCNA). This results in the incision of the nascent DNA strand, mismatch removal by exonuclease 1 (EXO1), and DNA repair synthesis (Fig. 1; ref. 61).

Heterozygous germline mutations in MMR components cause the so-called Lynch syndrome (LS), an inheritable predisposition for colorectal cancer (62). Moreover, somatic MMR defects provoke genetic hypermutability in the context of MSI, which has recently been approved as a tumor-agnostic marker for the use of immune-checkpoint inhibitors (ICI) targeting programmed cell death 1 (PDCD1, best known as PD-1) in patients with cancer (63). In line with this notion, a defective MMR correlates not only with an increased tumor mutational burden (TMB, which has been linked to ICI sensitivity in multiple cancer patient cohorts; ref. 64) but also with exacerbated inflammatory reactions and an inflamed TME at baseline (which has likewise been consistently associated with clinical ICI effectiveness; refs. 2, 65).

As compared with their WT counterparts, Msh2−/− mice are more sensitive to DSS/AOM-driven colorectal carcinogenesis (66), an effect that can be attenuated, at least in the distal colon, by the nonsteroidal anti-inflammatory drug sulindac (67). Along similar lines, the loss of Msh2 exacerbates the formation of colorectal lesions in tumor-prone APCMin/+ mice by a mechanism that does not originate from compromised repair of nitric oxide-induced DNA lesions (68) but rather involves the gut microbiome, both as a source of oncogenic metabolites (69) and as a driver of inflammation (70). Intriguingly, MSH2 is downregulated in response to TNF signaling in the liver (71), suggesting that, at least in some settings, TPI and DDR defects may be interconnected via a vicious feedforward loop (Fig. 2). Finally, although the combined loss of MSH2 and its binding partner MSH6 has been associated with a high TMB in a heterogeneous collection of MSI-high cancers (72), in a data set including 5 distinct cohorts of patients with lung adenocarcinoma, high MSH2 expression levels correlated positively not only with signatures of MMR proficiency but also with a high TMB, expression of CD274 (best known as PD-L1), and tumor infiltration by CD8+ T cells, pointing to an ICI-sensitive TME (73). Although the precise reasons underlying such an apparent discrepancy remain to be elucidated, tissue of origin may considerably influence the impact of MSH2 defects on the TME, as previously documented in prostate versus colorectal or endometrial tumors (74). On the other hand, several case reports confirm that LS-associated germline MSH6 mutations are linked to extraordinary sensitivity to ICIs not only in patients with colorectal cancer (75) but also in patients with uterine serous carcinoma (76) or endometrial tumors (77).

Figure 2.

The DDR-inflammation rheostat and cancer. Proficient DNA repair not only prevents the accumulation of potentially oncogenic mutations but also inhibits the secretion of multiple cytokines with immunostimulatory potential, thus preserving inflammatory homeostasis. Conversely, DDR alterations not only enable the accumulation of genetic defects that may drive oncogenesis but also derepress multiple proinflammatory pathways. Intriguingly, several immune mediators elicited by DDR defects negatively affect DDR proficiency, potentially engaging a feed-forward, self-amplificatory loop that compromises genomic stability and inflammatory homeostasis. Importantly, the net effect of such a loop on malignant transformation and tumor progression exhibits considerable degrees of context dependency, varying from accrued oncogenesis in the context of tumor-promoting inflammation to robust oncosuppression via cancer immunosurveillance.

Figure 2.

The DDR-inflammation rheostat and cancer. Proficient DNA repair not only prevents the accumulation of potentially oncogenic mutations but also inhibits the secretion of multiple cytokines with immunostimulatory potential, thus preserving inflammatory homeostasis. Conversely, DDR alterations not only enable the accumulation of genetic defects that may drive oncogenesis but also derepress multiple proinflammatory pathways. Intriguingly, several immune mediators elicited by DDR defects negatively affect DDR proficiency, potentially engaging a feed-forward, self-amplificatory loop that compromises genomic stability and inflammatory homeostasis. Importantly, the net effect of such a loop on malignant transformation and tumor progression exhibits considerable degrees of context dependency, varying from accrued oncogenesis in the context of tumor-promoting inflammation to robust oncosuppression via cancer immunosurveillance.

Close modal

Loss of MLH1 and its binding partner PMS2 has been linked to increased TMB in MSI-high solid tumors, although to lower levels than the combined loss of the MutSα components MSH2 and MSH6 (72). In endometrial carcinoma, loss of MLH1 and PMS2 has also been independently associated with PD-L1 positivity (78), again associating MSI with therapeutically actionable features of the TME. Accordingly, gastrointestinal tumors evolving in Mlh1−/− mice could be efficiently controlled by combining immunogenic chemotherapeutics such as gemcitabine or cyclophosphamide (29), a tumor-targeting vaccine (79, 80), and/or a PD-L1 blocker (81, 82).

Importantly, such a response pattern does not only reflect the abundant tumor neoantigen (TNA) load caused by the loss of MLH1 (which per se may induce some degree of anticancer immunity; ref. 83), but also the active secretion of proinflammatory factors, including type I IFN, C–C motif chemokine ligand 5 (CCL5) and C–X–C motif chemokine ligand 10 (CXCL10) via a CGAS- and stimulator of interferon response cGAMP interactor 1 (STING1)-dependent mechanism, by cells experiencing MMR defects (84–86). At least in some settings, such a response is initiated by EXO1-driven DNA hyperexcision and consequent accumulation of CGAS-activating DNA in the cytosol (87). Accordingly, Mlh1−/− mice are exquisitely sensitive to inflammation-driven colorectal carcinogenesis (88), a phenotype that can be further aggravated by mild inflammatory cues including low-dose radiation (89). Of note, MLH1 also resembles MSH2 in being (epigenetically) repressed by inflammatory cues (90) in the context of a vicious oncogenic cycle involving defective DDR and TPI (Fig. 2).

Taken together, these observations exemplify the major genoprotective and anti-inflammatory roles of a proficient MMR in both healthy and malignant cells (Supplementary Table S2).

Nonhomologous end-joining (NHEJ) is a potentially error-prone mechanism for the repair of DNA double-strand breaks (DSB) via direct ligation of DNA ends (91, 92). NHEJ is initiated by the binding of a heterodimer consisting of XRCC6 (best known as Ku70) and XRCC5 (best known as Ku80) to DSBs. The Ku70/Ku80 heterodimer promotes the recruitment and activation of protein kinase, DNA-activated, catalytic polypeptide (PRKDC, best known as DNA-PKcs). Processing of incompatible DSB ends by nucleases, like DNA cross-link repair 1C (DCLRE1C, best known as Artemis), and DNA polymerases, such as DNA polymerase mu (POLM) and lambda (POLL), facilitates DNA end-ligation by LIG4, XRCC4, and nonhomologous end-joining factor 1 (NHEJ1, best known XLF; refs. 91, 92; Fig. 1). Defects in NHEJ are fairly uncommon in human neoplasms, possibly with the exception of single-nucleotide polymorphisms in XRCC5, XRCC6, or their promoter regions, which have been associated with an increased risk for multiple solid tumors (93).

At odds with other DDR components (see above), both XRCC6 and XRCC5 appear to be upregulated in the context of proinflammatory signaling via TLR4 or prostaglandin-endoperoxide synthase 2 (PTGS2, best known as COX2) in healthy tissues including the liver (94), as well as in malignant cells (95). Diethylnitrosamine (DEN)-driven hepatic carcinogenesis was accelerated in male mice bearing a missense Tlr4 mutation, correlating with increased oxidative DNA damage, failed Ku70 and Ku80 upregulation early upon DEN-mediated hepatic injury, limited inflammatory signaling, and reduced numbers of intrahepatic macrophages (94). Importantly, most of these hepatic phenotypes could be reverted by adenovirus-driven Ku80 overexpression in hepatocytes, culminating with decelerated DEN-driven carcinogenesis (94). Of note, Ku80 levels negatively correlate with PD-L1 expression in post-RT biopsies from patients with a variety of tumors (96), suggesting that a proficient NHEJ or other NHEJ-unrelated Ku70 or Ku80 functions limit tumor-targeting immune responses culminating with interferon gamma (IFNG)-dependent PD-L1 upregulation in the TME.

Supporting the latter possibility, deletion of Xrcc6 (alone or in the context of mutant p53) causes intestinal alterations associated with a pronounced inflammatory response and increased propensity for carcinogenesis, a phenotype that the deletion of Lig4 cannot recapitulate (97). Of note, extranuclear Ku70 has been proposed to drive type III IFN secretion in response to the cytosolic accumulation of DNA, via a mechanism that involves the Ku70 DNA-binding domain and STING1 (98, 99). However, it seems unlikely that the loss of cytosolic DNA sensing by Ku70 may explain inflammation-associated colorectal carcinogenesis in Xrcc6−/− mice.

DNA-PKcs has also been proposed to operate as cytosolic DNA sensor in the context of viral infection, ultimately driving type I IFN production via a STING1- and interferon regulatory factor 3 (IRF3)-dependent mechanism (100). Such an extranuclear DNA-PKcs activity, which is antagonized by a variety of viral factors (101, 102), does not appear to involve physical interactions with Ku70 (103) and may not be operational in all cells. Indeed, pharmacologic inhibition (rather than activation) of DNA-PKcs with AZD7648 synergizes with RT to promote therapeutically relevant type I IFN responses in various syngeneic mouse tumor models (104). More­over, not only missense mutations of PRKDC are associated with autoimmune disorders (105), but DNA-PKcs has also been shown to catalyze the inactivating phosphorylation of CGAS, de facto suppressing antiviral immunity (106). Finally, a nuclear pool of STING1 has been proposed to support genomic stability by binding to DNA-PKcs and its interactors (107). Taken together, these latter observations suggest that the ability of DNA-PKcs to operate as a cytosolic DNA sensor should be revisited. Nonetheless, it appears that both DNA-PKcs and less so the DDR signal transducers ATR and mitogen-activated protein kinase 14 (MAPK14, best known as p38MAPK) are important for cancer cells experiencing DNA damage to emit adjuvant signals that actively engage tumor-specific immune responses irrespective of–or prior to–immunogenic cell death (ICD; refs. 108, 109). The precise molecular mechanisms underlying these latter observations, however, remain elusive.

Yet another NHEJ component, XRCC4, has been shown to affect inflammatory reactions (110). Specifically, XRCC4 physically interacts with RNA sensor RIG-I (RIGI, best known as RIG-I), a cytosolic pattern recognition receptor (PRR) that binds ectopic RNA molecules (111). Besides inhibiting NHEJ (and hence interfering with viral integration in the host genome), such an interaction promotes the oligomerization and ubiquitination of RIG-I in support of improved antiviral activity (110). However, whether the interaction between XRCC4 and RIG-I also modulates inflammatory responses in tumor models or patients with cancer has not yet been investigated.

Thus, NHEJ components operate as part of the DDR to repair DSBs in the nucleus and as regulators of inflammatory reactions in the cytosol (Supplementary Table S3).

Homologous recombination (HR) is a generally error-free pathway that operates in the S and G2 phases of the cell cycle to promote the repair of DSBs using homologous DNA templates (91, 92). HR is initiated by the 5′ to 3′ resection of DNA ends by the MRN complex, a heterotrimer consisting of MRE11 homolog, double-strand break repair nuclease (MRE11), RAD50 double-strand break repair protein (RAD50) and nibrin (NBN, best known as NBS1); and RB binding protein 8, endonuclease (RBBP8, best known as CTIP). This generates short stretches of single-strand DNA (ssDNA) that are elongated through the activities of nucleases, like EXO1 and DNA2, and helicases, such as BLM RecQ like helicase (BLM), resulting in long ssDNA overhangs that are coated and protected by replication protein A1 (RPA1, best known as RPA). BRCA1, BRCA2, and partner and localizer of BRCA2 (PALB2) favor the exchange of RPA with RAD51 recombinase (RAD51), which forms a nucleofilament that can invade the undamaged sister chromatid in search of homologous regions to utilize as template for repair. The invading strand initiates DNA synthesis, which can be followed by synthesis-dependent strand annealing, involving the release of the newly synthesized strand and its consequent reannealing to the second DSB end. Alternatively, double Holliday junctions can form, which can either undergo resolution by Holliday junction resolvases or dissolution by a complex containing BLM and DNA topoisomerase III (TOPOIII), resulting in the generation of crossover or noncrossover products (Fig. 1; refs. 91, 92). HR has been extensively investigated in the context of cancer, especially upon the identification of germline BRCA1 and BRCA2 mutations as risk factors for mammary and ovarian carcinogenesis (112), and the discovery that neoplasms bearing BRCA1 or BRCA2 mutations are exquisitely sensitive to PARP inhibition (20).

Each of the components of the MRN complex has been involved in the regulation of inflammatory reactions. For instance, an extranuclear pool of MRE11 has been shown to serve as a cytosolic DNA sensor via a mechanism that involves its partner RAD50 and culminates with STING1 activation followed by IRF3-dependent type I IFN synthesis (113). Moreover, in patients with Fanconi anemia (FA, see below), mitochondrial MRE11 actively degrades nascent mitochondrial DNA (mtDNA), resulting in mtDNA cytosolic spillage and CGAS activation (114), a process that is generally under negative regulation by autophagy (which degrades dysfunctional mitochondria; ref. 115) and caspases (which proteolytically inactivate CGAS and accelerate cell death; refs. 116, 117). Finally, the exonuclease activity of MRE11 can generate CGAS-activating DNA fragments and hence elicits innate immune signaling, at least in cells lacking RAD51, which would otherwise protect the newly synthesized DNA from MRE11 degradation (118). Extranuclear RAD50 also appears to interact with the PRR caspase recruitment domain family member 9 (CARD9), thus enabling IL1β production in response to cytosolic DNA accumulation in infected dendritic cells (DC) via an NF-κB–dependent mechanism (119). Whether this latter process is operational in noninfected or epithelial cells remains to be formally demonstrated.

Dissociation of the MRN complex as caused by adenoviral infections has also been suggested to promote NBS1 segregation in nuclear foci, coupled with ATM hyperactivation and consequent NF-κB signaling (120). Although this may result in genomic instability coupled with polyploidization (120), its impact on immunogenicity has not been clarified yet. Interestingly (at least in some circumstances), intracellular PD-L1 can act as an mRNA-binding protein to stabilize selected transcripts including NBS1, BRCA1, and other mRNAs encoding DDR components (121). This results in superior DDR proficiency, at least hypothetically coupled to the suppression of potentially oncosuppressive inflammatory reactions driven by suboptimal HR (121). Both NBS1 (122) and BRCA1 defects (see below) have indeed been associated with accrued inflammatory responses at baseline or in response to carcinogens.

EXO1 promotes HR by catalyzing long-range end-resection (123). EXO1 is also involved in other DDR pathways, including NER and MMR, largely reflecting its ability to remove DNA stretches containing lesions prior to the restoration of original DNA sequences (61, 124). EXO1 nuclease activity can generate ssDNA fragments that—unless degraded by the cytosolic exonuclease three prime repair exonuclease 1 (TREX1)—mediate interferogenic effects (125). A similar activity has also been documented for the end-resection factor BLM (125). Moreover, EXO1-dependent DNA degradation appears to underlie the ability of DSBs accumulating in response to RT to elicit PD-L1 exposure in cancer cells, a mechanism that is finely regulated by multiple HR components (126).

BRCA1 and BRCA2 promote HR by favoring DSB end-resection (BRCA1) and loading the recombinase RAD51 on resected DNA ends (BRCA1 and BRCA2; ref. 13). Besides generating non-oncogene addiction to PARP, loss of BRCA1 has been shown to promote oncogenesis and tumor progression in the context of immunosuppressive TME alterations that (at least in some settings) may nonetheless be actionable for therapeutic purposes. For instance, BRCA1 mutations have been associated with the recruitment of myeloid-derived suppressor cells (MDSC) to the mammary TME via CXCL12, resulting in therapeutic resistance to PD-1 blockers that could be abrogated with CXCL12-targeting strategies (127). Similarly, in preclinical models of ovarian cancer, loss of Brca1 results in an inflamed (but immunoresistant) TME at baseline as a consequence of chronic CGAS/STING1 activation and consequent secretion of vascular endothelial growth factor A (VEGFA; ref. 128). In this setting, sensitivity to dual PD-1 and cytotoxic T-lymphocyte associated protein 4 (CTLA4) blockage can be restored with VEGFA-neutralizing agents (or Sting1 deletion) optionally combined with PARP inhibition (128). Signatures of CGAS/STING1 and IFNG signaling have also been detected by comparing isogenic BRCA1-deficient versus BRCA1-proficient breast and ovarian cell lines, as well as by comparing patients with mutant vs. WT BRCA1 from The Cancer Genome Atlas (TCGA; ref. 129).

In triple-negative breast cancer (TNBC) models, Brca1 defects correlate with increased TMB and superior tumor infiltration by lymphocytes expressing PD-1 and CTLA4, offering a therapeutically actionable profile for dual immune-checkpoint blockage plus cisplatin-based immunotherapy (130). Disrupting the binding between BRCA1 and its physical interactor PALB2 facilitates age-driven hepatic oncogenesis, but results in the formation of tumors that respond to PD-1 blockage as a consequence of robust type I IFN secretion via CGAS (131). These findings suggest that the net effect of BRCA1 mutations on the TME and sensitivity to ICIs exhibits at least some degree of context dependency. In line with this notion, BRCA1 mutations have been correlated with increased tumor infiltration by a variety of immune cells including CD4+ and CD8+ T cells, in patients with hepatocellular carcinoma (HCC; ref. 132), breast cancer (133), or high-grade serous ovarian carcinoma (HGSOC; ref. 134). However, although in HCC this was linked to MSI and poor prognosis (132), a robust immune infiltrate and/or enhanced spatial interactions between malignant cells and T cells had positive prognostic value in patients with BRCA1-mutated breast cancer (133) and HGSOC (134).

Recent data argue against the assumption that BRCA1 and BRCA2 mutations would have similar immunological effects on developing tumors. Specifically, gene-expression programs related to both adaptive and innate immunity are enriched in BRCA2- over BRCA1-deficient tumors (135). Moreover, truncating mutations in BRCA2 are associated with superior response to ICIs as compared with their BRCA1 counterparts, not only as modeled in various mouse cancer cell lines but also as assessed in a patient cohort encompassing 95 patients with diverse malignancies (135). Along these lines, although the acute transcriptional response to experimental BRCA2 loss is largely dominated by genes involved in cell-cycle progression, DNA replication, and repair, late transcriptional alterations are mostly linked to IFN signaling, a process that is further magnified by PARP inhibition (136). Specifically, BRCA2 inactivation appears to promote the formation of micronuclei that recruit CGAS to initiate a potent interferogenic response that is accompanied by TNF secretion, and restores cell killing by exogenous TNF administration in a variety of otherwise TNF-resistant human cancer cell lines (137). In line with this notion, human cancer cells with homozygous or heterozygous deletions of BRCA2 have been attributed superior sensitivity to RCD driven by TNF-like death receptor ligands (138). These observations reveal unsuspected differences between the immune correlates of BRCA1 and BRCA2 defects that require additional investigation.

Similar to BRCA1 defects, RAD51 expression has been consistently associated with increased TMB and robust tumor infiltration by immune cells including CD4+ and CD8+ T cells in patients with HCC (139) and breast cancer (140, 141). However, in all these settings, elevated RAD51 levels correlated with advanced tumor stage, high grade, and dismal prognosis (139–141). These observations may reflect the establishment of chronic, indolent TPI with a detrimental impact on oncogenesis, and tumor progression. Supporting this notion, transcriptional signatures of type I and IFNγ signaling have consistently been associated with poor prognosis in patients with HCC and breast cancer (116, 142). Apparently at odds with these observations, however, depletion of RAD51 or its interacting partner RPA has been shown to cause robust accumulation of cytosolic ssDNA coupled to potent CGAS activation, especially in the context of defects in TREX1 (143). Conversely, CGAS appears to potently inhibit HR by compacting template DNA into a hypercondensed state that is less amenable to invasion by RAD51-coated ssDNA filaments (144), and by binding to and hence interfering with the activation of PARP1 (145). Such an effect, however, is independent from canonical CGAS/STING1 signaling.

In conclusion, HR signaling heavily impinges on the regulation of inflammatory reactions in cancer (Supplementary Table S4).

PARP1 and other members of the PARP protein family detect a variety of DNA lesions including SSBs resulting from the APE1-mediated processing of abasic sites during BER, ssDNA gaps formed during DNA replication, as well as DSBs in the context of HR and NHEJ, ultimately participating in the initiation of multiple DDR pathways (146). PARP inhibitors have been intensively investigated (and are currently approved for use) in patients with breast or ovarian neoplasms bearing BRCA1 or BRCA2 mutations (20). An expanding literature suggests that at least part of the therapeutic benefits mediated by PARP inhibitors in these patients emerges from immunologic mechanisms.

PARP inhibition reportedly mediates strong, STING1-dependent inflammatory responses culminating with type I IFN and/or CCL5 secretion in models of BRCA1-deficient and proficient ovarian cancer (147, 148), BRCA1-deficient (and less so BRCA1-proficient) TNBC (149, 150), BRCA1-competent non–small lung cell carcinoma (NSCLC) with defects in NER (149), BRCA2-deficient NSCLC and TNBC cells (136), as well as in renal cell carcinoma models lacking polybromo 1 (PBRM1), a chromatin remodeling enzyme (151), and models of small-cell lung carcinoma (SCLC; ref. 152).

Intriguingly, such an immunostimulatory activity has been linked to the ability of some PARP inhibitors to induce PARP1 trapping on DNA (153), resulting in accrued replication stress (see below), aggravated genomic instability, and reduced availability of cytosolic PARP1, which directly inhibits CGAS via PARylation (154). Conversely, CGAS accumulating in the nucleus as a consequence of DSBs interferes with PARP1 activation during HR (145). Such an effect, however, does not seem to involve canonical CGAS signaling but instead the dephosphorylation of CGAS on Y215, which is a target of BLK proto-oncogene, Src family tyrosine kinase (BLK; ref. 145).

Other members of the PARP family, including PARP7 (and most likely PARP2), share with PARP1 the ability to strongly suppress type I IFN signaling (155). In some models, genetic or pharmacologic PARP inhibition promoted the establishment of an inflamed TME in vivo (152, 156–158), correlating with the IFNG-driven upregulation of PD-L1 (149, 159) and synergism with ICIs (148, 152, 157–160) as well as immunotherapeutic agents stimulating STING1 signaling (161), targeting immunosuppressive macrophages (162), or promoting antibody-dependent cellular cytotoxicity by NK cells (163). Intriguingly, the ability of PARP inhibitors to drive PD-L1 expression in the TME may originate not only from the activation of cellular immunity coupled with IFNγ production but also from the ability of PARP1 to suppress the transcription of CD274 by PARylating signal transducer and activator of transcription 3 (STAT3; ref. 164). Along, similar lines, PARP1 has been shown to actively suppress the expression of various NK cell–activating ligands (NKAL) by acute myeloid leukemia cells, hence supporting their ability to evade NK cell–dependent immunosurveillance (which is particularly prominent in leukemia; refs. 165, 166).

Of note, Parp1−/− mice appear to be less (rather than more) sensitive to DSS/AOM-induced colorectal carcinogenesis, correlating with moderately increased DNA damage but decreased signs of inflammation in the intestinal epithelium (167). Intriguingly, such a chemopreventive effect appears to be maximal in heterozygous Parp1+/− hosts or in the context of incomplete pharmacologic inhibition of PARP, potentially reflecting the inhibition of TPI, notably the tumor-promoting activity of MDSCs, coupled to the reestablishment of robust anticancer immunosurveillance (156, 168).

Altogether, these data suggest that interfering with the kinetic whereby PARP normally interacts with damaged DNA, notably PARP trapping, results in robust inflammatory responses often linked to cytosolic STING1 activation (Supplementary Table S5).

The main DDR kinases are ATM, which is activated upon DSB formation (12), and ATR, which is activated in response to ssDNA stretches generated following DSB end-processing or replication fork stalling (169). The catalytic activity of ATM and ATR regulates (either directly or indirectly) various effector molecules that promote DNA repair in the context of a transient cell-cycle arrest, or actively drive cellular senescence or RCD when DNA is damaged beyond recovery. These effectors include CHEK2 (the prototypical ATM substrate) and CHEK1 (the prototypical ATR substrate) as well as p53, p38MAPK, and WEE1 (Fig. 1; refs. 12, 169). Somatic mutations or epigenetic mechanisms suppress (at least to some degree) DDR signaling in most human tumors (12, 170). This reflects the robust oncosuppressive function of proficient DDR signaling, both as a genoprotective mechanism in the context of mild DNA damage and as a driver of cellular senescence or RCD in cells bearing irreparable DNA lesions (171). That said, most cancer cells tend to become highly reliant on alternative DDR pathways for survival and proliferation, which has raised considerable interest in the development of DDR inhibitors (171). Accumulating evidence suggests that these agents may also evoke therapeutically relevant anticancer immune responses.

Atm−/− mice resemble their Msh2−/− counterparts as they exhibit increased sensitivity to colitis-driven colorectal carcinogenesis compared with WT mice (172). Along similar lines, the loss of ATM (which is frequent in human tumors) may generate non-oncogene addiction to PARP1 (at least in models of colorectal cancer and mantle cell lymphoma), which also occurs in the context of BRCA1 defects (173, 174). Moreover, ATM inhibition has been shown to derepress tumor-targeting immune responses by a variety of mechanisms. First, pharmacologic inhibition of ATM drives robust type I IFN secretion in preclinical tumor models, via both CGAS/STING1-independent (175) and -dependent processes (176, 177), representing an efficient therapeutic partner for ICIs (175, 176). Of note, at least in some of these settings, similar results could be obtained with inhibitors of CHEK2 (177). Second, ATM depletion by RNA interferences suppresses the expression of integrin αvβ3 by cancer cells experiencing DNA damage (178), knowing that integrin αvβ3 facilitates the uptake of viable (as compared with dying) cancer cells by antigen-presenting cells (APCs) and hence suppresses CD8+ T-cell cross-priming (178). Finally, pharmacologic ATM inhibition prevents PD-L1 expression as driven by chemotherapy in preclinical models of prostate cancer (179). In line with this notion, the loss of ATM as well as ATM mutations have been linked with a beneficial immune configuration of the TME and/or with increased sensitivity to ICIs in various cohorts of patients with cancer (177, 180, 181).

Although ATM signaling in neoplastic cells dampens inflammation, ATM activation in immune cells apparently supports (rather than inhibits) inflammatory responses. For instance, an extranuclear pool of ATM has been shown to underlie the ability of macrophages to acquire proinflammatory, antitumoral activities in response to the activation of NADPH oxidase 2 by radiotherapy (182). Similarly, ATM signaling in APCs exposed to low-dose chemotherapy drives an NF-κB–dependent signal transduction pathway that ultimately enhances antigen presentation by upregulating MHC class II surface levels (183). Whether the latter is associated with improved immune tumor control has not been explored. Of note, inflammatory signaling as driven by CGAS also regulates ATM functions. Specifically, the CGAS-dependent synthesis of cyclic GMP-AMP (cGAMP) in response to cytosolic dsDNA has been shown to promote ATM activation in STING1- and TANK binding kinase 1 (TBK1)-dependent (but type I IFN-independent) manner, resulting in cell-cycle arrest downstream of CHEK2 signaling but inhibited HR as a consequence of NAD+ depletion (184). These latter observations exemplify yet another scenario in which inflammatory cues and DDR inhibition may interact in a feedforward, self-amplifying mechanism (Fig. 2).

ATR activation as driven by Epstein–Barr virus (EBV) infection (which is responsible for most nasopharyngeal carcinoma cases) promotes the establishment of a strongly immunosuppressive microenvironment dominated by M2-like macrophages that is permissive for oncogenesis and tumor progression (185). Moreover, inhibited ATR signaling due to baculoviral IAP repeat containing 6 (Birc6) deletion in hepatocytes drives tumor-promoting inflammatory reactions that aggravate DEN-driven hepatic carcinogenesis (186). ATR underlies multiple immunosuppressive mechanisms including the upregulation of PD-L1 and CD47 (an antiphagocytic signal) also in cancer cells (187), de facto impairing the ability of RT to elicit therapeutically relevant anticancer immune responses as a consequence of defects in both immune priming (187) and immune execution (188). Thus, proficient ATR signaling mediates anti-inflammatory effects in both normal and malignant tissues. In stark contrast, however, Atr haploinsufficiency accelerates the formation of melanomas driven by the loss of Pten and activation of B-Raf proto-oncogene, serine/threonine kinase (BRAF) in mice, a phenotype that is accompanied by the establishment of an immunosuppressive TME with M2-like macrophages and PD-L1 expression (189). Whether such an apparently paradoxical effect originates from gene dosage changes remains to be clarified.

Irrespective of this unknown, pharmacologic ATR inhibitors reportedly mediate robust immunostimulatory effects, often dependent on cytosolic dsDNA accumulation and CGAS signaling, in preclinical models of prostate cancer (190), acute lymphocytic leukemia (ALL; ref. 191), head and neck squamous cell carcinoma (HNSCC; ref. 192), PBRM1-deficient renal cell carcinoma (151), and BRCA2-defective cervical carcinoma (193). Moreover, ATR inhibition synergizes with RT at driving CGAS-dependent immunostimulatory effects of therapeutic relevance that (at least in some settings) can be exacerbated by ICIs in models of hepatocellular carcinoma (194), human papillomavirus (HPV)+ lung carcinoma (195), and KRAS-mutant tumors (196). Intriguingly, a similar interferogenic response has been documented in nontransformed mammary epithelial MCF10A cells exposed to RT and ATR inhibition, although, in this context, it was mechanistically linked to sensing of cytosolic RNA, not DNA (197). ATR inhibitors have also been shown to synergize at the activation of therapeutically relevant (and sometimes ICI-actionable) tumor-targeting immune responses upon combination with WEE1 inhibitors (in models of ovarian cancer; ref. 198), oxaliplatin (in models of oxaliplatin-resistant colorectal cancer; ref. 199), and aurora kinase A (AURKA) blockers (in models of MYCN-amplified neuroblastoma; ref. 200). These latter observations reinforce the notion that inhibiting multiple components of the DDR may drive superior immunostimulatory effects, perhaps by abrogating (or at least interfering with the emergence of) compensatory non-oncogene addiction.

Immunostimulatory effects have also been observed upon inhibition of CHEK1. For instance, pharmacologic CHEK1 inhibition has been shown to synergize with the immunogenic chemotherapeutic gemcitabine and ICIs in preclinical models of SCLC, correlating with accrued tumor infiltration by CD8+ T cells, DCs, and proinflammatory M1-like macrophages at the expenses of anti-inflammatory M2-like macrophages and MDSCs (201). Similar to ATR blockers, CHEK1 inhibitors can also promote TNF secretion by ALL cells (191), although the immunologic consequences of this process remain unclear. Finally, pharmacologic CHEK1 inhibition has been shown to mediate potent, CGAS-dependent immunostimulatory effects and hence synergize with PD-L1 blockers in mouse models of SCLC (152).

That said, pharmacologic CHEK1 inhibition reportedly limits (rather than aggravates) colorectal carcinogenesis in the DSS/AOM model, apparently as a consequence of inhibited C–C motif chemokine ligand 2 (CCL2) expression and hence scarce recruitment of CCR2+ macrophages to the colon (202). Finally, it seems that multiple inflammatory cues converging on IRF1 activation (including, but not limited to, IFNG signaling) promote CHEK1 downregulation (203, 204). Such a circuitry, however, appears more plausible to facilitate the apoptotic demise of cancer cells experiencing DNA damage or replication stress than to inhibit DNA repair.

Activation of p53 in hepatocytes undergoing oncogene-driven senescence mediates robust cell-extrinsic oncosuppressive functions via CD4+ T cells and NK cells (8, 9). Intriguingly, such a homeostatic effect appears to require physiologic p53 signaling, as constitutive p53 activation resulting from the deletion of transformed mouse 3T3 cell double minute 2 (Mdm2, which encodes the main endogenous p53 inhibitor) instead paradoxically promotes KRAS-driven hepatic carcinogenesis upon the establishment of TPI (205). That said, p53 loss and cancer-associated TP53 mutations have been consistently linked to the establishment of TPI in support of tumor progression in the context of immunoevasion. For instance, a common p53 mutant (p53R249S) binds TBK1 and prevents it from interacting with STING1 and IRF3, de facto blocking type I IFN secretion downstream of CGAS signaling and hence limiting the infiltration of mouse TNBCs by NK cells in favor of increased amounts of immunosuppressive. M2-like macrophages, and accelerating tumor growth (206). Along similar lines, a gain-of-function mutation of p53 that is common in human neoplasms (R172H) alters the immune infiltrate of experimental KRAS-driven pancreatic tumors to promote the accumulation of neutrophils at the expenses of CD8+ and CD4+ T cells, resulting in accrued resistance to ICIs that can be ameliorated by neutrophil depletion (207). Furthermore, the loss of Trp53 (the mouse homolog of human TP53) specifically in KRAS-driven pancreatic cancer cells promotes the recruitment of immunosuppressive myeloid cells to the TME, culminating with accelerated tumor progression in the context of suppressed anticancer T cell–dependent immunity (208). In breast cancer, loss of p53 is also accompanied by the secretion of WNT ligands that stimulate local macrophages to produce IL1β, favoring a systemic inflammatory response that supports metastatic dissemination (209).

In line with these observations, p53 reactivation has been associated with restored cancer immunosurveillance and responsiveness to immunotherapy in numerous tumor models. For instance, pharmacologic MDM2 inhibition by nutlin-3a stimulates systemic antitumor immunity and tumor regression in highly infiltrated mouse lymphoma models (210) as well as in preclinical models of melanoma, especially when combined with an AURKA inhibitor (211). Moreover, pharmacologic MDM2 inhibition promotes interferogenic effects downstream of the derepression of endogenous retroviruses that are sensed as cytosolic double-strand RNA (dsRNA; ref. 212), and reportedly upregulates the costimulatory molecule CD80 on epithelial cancer cells (213). Along these lines, restoring p53 expression with nanoparticles reconfigures the immunologic TME in mouse models of HCC and oral squamous cell carcinoma as it synergizes with PD-1–targeting ICIs (214, 215). Similar immunostimulatory effects have been reported for nutlin-3a and a peptide-based MDM2-targeting vaccine in syngeneic models of HNSCC (216), for HDM201 (a potent and selective second-generation MDM2 inhibitor) combined with PD-1 or PD-L1 blockers in models of colorectal cancer and TNBC (217), as well as the for MDM2 antagonist APG-115 plus PD-1 inhibitors in models of colorectal cancer and HCC (218). Pharmacologic MDM2 antagonism has also been shown to increase the susceptibility of malignant cells to lysis by immune effector cells (219). Whether this effect mechanistically depends on p53, however, remains unclear. Further supporting the critical role of anticancer immunosurveillance in the efficacy of p53-restoring strategies, the orally active MDM2 antagonist DS-527 mediates optimal antileukemic effects only in immunocompetent hosts (220).

Intriguingly, MDM2 expression in T cells ensures STAT5 stability in support of tumor-targeting immunity, and disruption of the MDM2–p53 interaction with APG-115 exacerbates this effect by increasing the amount of p53-unbound MDM2 available for binding the STAT5 inhibitor Cbl proto-oncogene (CBL; ref. 221). Thus, disrupting the MDM2–p53 interaction also mediates immunostimulatory effects that originate from immune, rather than malignant, cells. Of note, mitochondrial antiviral signaling protein (MAVS), a signal transducer in cytosolic dsRNA sensing, has been proposed to interfere with the p53–MDM2 interaction and hence stabilize p53 in support of oncosuppression (222). Consistent with this notion, Mavs−/− mice are more sensitive to DSS/AOM-driven colorectal carcinogenesis than their WT counterparts (222). However, the relative contribution of cell-extrinsic inflammatory pathways versus cell-intrinsic genetic defects to accrued colorectal carcinogenesis in Mavs−/− mice remains to be clarified. Similar to p53, the p53 interactor tumor protein p53 binding protein 1 (TP53BP1, best known as 53BP1) has been shown to preserve inflammatory homeostasis in physiologic conditions, largely reflecting its ability to inhibit the formation of CGAS-activating cytoplasmic chromatin fragments (CCF; ref. 223). Mitochondrial dysfunction, however, initiates an oxidative pathway culminating with 53BP1 inhibition, CCF accumulation, and CGAS-dependent cellular senescence (223), which is coupled to the secretion of numerous proinflammatory factors. The ultimate effect of such a senescence-associated secretory phenotype on oncogenesis and tumor progression exhibits considerable context dependency (224).

Besides contributing to the cell-cycle arrest driven by DDR signaling (225, 226), the ATR substrate p38MAPK and its effector MAPK activated protein kinase 2 (MAPKAPK2, best known as MK2) appear to be linked to the regulation of inflammatory responses in a complex manner. p38MAPK has been involved in the immunogenicity of cancer cells injured by DNA-damaging agents (108). Moreover, the macrophage-specific deletion of Mapk14 limits DSS-driven colitis (227). Similar results have been obtained in Mapkapk2−/− mice (228), as well as in mice specifically lacking MK2 in the myeloid compartment (229, 230). Furthermore, MK2 has been suggested to promote replication stress as driven by DNA damage (231), which is also expected to promote inflammatory responses (see below). However, whole-body pharmacologic p38MAPK inhibition as well as Mapk14 deletion from intestinal epithelial cells (IEC) appears to aggravate, rather than limit, the inflammatory and oncogenic effects of DSS/AOM (232, 233). Thus, although p38MAPK/MK2 signaling in macrophages underlies DSS/AOM-driven oncogenesis, the same pathway may mediate oncosuppressive effects in IECs, potentially by preserving epithelial barrier function (233). Intriguingly, such a protective effect may no longer be relevant during tumor progression, as demonstrated by the fact that deleting Mapk14 in established DSS/AOM-driven carcinomas decelerates, rather than accelerates, tumor growth (233). Although this latter finding has been attributed to a cancer cell–intrinsic mitogenic role for p38MAPK signaling (233), the aforementioned observations exemplify the complexity of the links between the DDR and inflammation in different cell types.

WEE1, a kinase that inhibits cell-cycle progression by antagonizing the CHEK1 and CHEK2 substrate cell division cycle 25A (CDC25A) and is involved in DDR signaling, suppresses inflammatory responses in various preclinical cancer models. In line with this notion, pharmacologic WEE1 inhibition with AZD1775 potently activates STING1 in mouse SCLC, resulting in tumor infiltration by CD8+ T cells and restored sensitivity to PD-L1 blockers (234). Similar effects have been documented in mouse models of ovarian cancer cotreated with WEE1 and ATR inhibitors (198), as well as in models of gastric cancer lacking MUS81 structure-specific endonuclease subunit (MUS81; ref. 235). Moreover, WEE1 inhibitors appear to promote TNF secretion by ALL cells (191). Along these lines, WEE1 inhibition boosts anticancer immunity driven by RT in immunocompetent mice bearing syngeneic TNBC (236) and other tumors (237), at least in some setting synergizing with ICIs (237). Intriguingly, at least part of these effects originates from the ability of WEE1 inhibitors to reverse the G2—M block induced by RT (237), which has been associated with reduced sensitivity to immune effector molecules (238) similar to the epithelial–mesenchymal transition (another setting in which WEE1 inhibition restores sensitivity to immune effectors; ref. 239). Of note, WEE1 inhibition has also been associated with (i) the activation of ICD in mouse models of melanoma, a process that was exacerbated by concomitant AKT serine/threonine kinase 1 (AKT1) blockage, resulting in restored NK cell–dependent immunosurveillance and ICI sensitivity (240); (ii) the derepression of endogenous retroviruses (ERV), culminating with an interferogenic response driven by cytosolic dsRNA sensors independent of CGAS and STING1 (241); and (iii) the downregulation of baseline PD-L1 expression in pancreatic cancer cells, an effect that was magnified by concomitant ATM inhibition (242). Taken together, these observations exemplify the multipronged immunostimulatory potential of WEE1 inhibitors.

In summary, multiple DDR signal transducers are key regulators of inflammatory responses in both healthy and malignant cells (Supplementary Table S6).

Replication Stress Response

Replication stress occurs when the DNA replication machinery encounters obstacles, resulting in the stalling, collapse, or breakage of the Y-shaped DNA structures where DNA replication occurs, also known as replication forks (Fig. 1; ref. 243). Replication stress is common in cancer cells, owing not only to their hyperproliferative phenotype but also the accumulation of DNA lesions that interfere with replication fork progression (243). Of note, at least some DDR-targeting agents currently investigated for their anticancer activity de facto operate as inducers/aggravators of replication stress. Among others, these agents include ATR, CHEK1, and PARP1 inhibitors (244–246). Fork stalling leads to the accumulation of ssDNA stretches that activate ATR signaling to stabilize the replication fork, hence enabling the completion of DNA replication in support of preserved genomic integrity (247). Processing of stalled replication forks by nucleases such as MRE11 or DNA replication helicase/nuclease 2 (DNA2) can promote the generation of ssDNA fragments with interferogenic effects, especially in the context of SAM and HD domain containing deoxynucleoside triphosphate triphosphohydrolase 1 (SAMHD1) or FA complementation group D2 (FANCD2) defects (114, 248, 249).

PCNA, which is critical for physiologic DNA replication by acting as a DNA-sliding clamp that enhances the processivity of DNA polymerases (250), also prevents the MRE11-dependent production of potentially interferogenic ssDNA species at stalled replication forks (251). Specifically, PCNA phosphorylation on Y211 ensures optimal PCNA processivity in support of suppressed immunogenic signaling and limited NK cell–mediated immune tumor eradication (251). Interestingly, a plasma membrane pool of PCNA has also been proposed to operate as a coinhibitory ligand for the NK cell receptor natural cytotoxicity triggering receptor 2 (NCR2, best known as NKp44; ref. 252). The impact of PCNA phosphorylation on Y211 on such an immunosuppressive activity and the signals that regulate PCNA exposure on the plasma surface of malignant cells remain to be characterized.

Unresolved replication stress can also promote the formation of micronuclei (253), which are prone to rupture and hence can drive CGAS activation (254). In line with these observations, aggravating replication stress with a CHEK1 inhibitor plus low-dose hydroxyurea drives potent anticancer immunity in preclinical models of melanoma, with a major role for natural killer T (NKT) cells and limited synergy with ICIs (255). Similar observations have been made in preclinical models of squamous cell carcinoma and SCLC, where replication stress was induced by treatment with lysine demethylase 4A (KDM4A) or cyclin-dependent kinase 7 (CDK7) inhibitors, respectively, although in these settings anticancer immunity was primarily driven by CD8+ T cells and therapeutic effects could be improved by ICIs (256, 257).

The excessive accumulation of R-loops, which are three-stranded nucleic acid structures physiologically involved in transcription, has also been associated with replication stress and consequent accumulation of interferogenic nucleic acids including ssDNA fragments (258, 259) and RNA-DNA hybrids (260). This process, which is negatively regulated by the helicase DEAD-box helicase 41 (DDX41) and ERCC excision repair 1, endonuclease noncatalytic subunit (ERCC1, a regulatory component of NER), appears to be particularly relevant not only for age-related chronic inflammation, at least in the pancreas (259), but also for hematopoietic alterations potentially linked to leukemogenesis (258, 261). In line with this notion, germline DDX41 mutations are linked to an increased susceptibility to myeloid neoplasms in humans (262).

Taken together, these observations suggest that replication stress can be targeted to elicit therapeutically relevant anticancer immunity (Supplementary Table S7).

NER

NER is involved in the resolution of various DNA lesions, including intrastrand adducts and structures that distort the double helix. Damage sensing during NER is mediated by a heterotrimer comprising XPC complex subunit, DNA damage recognition and repair factor (XPC), RAD23 homolog B, nucleotide excision repair protein (RAD23B) and centrin 2 (CETN2). Next, the following factors are recruited: (i) the oligomeric transcription factor IIH (TFIIH), in support of DNA unwinding; (ii) XPA, DNA damage recognition and repair factor (XPA); (iii) RPA; (iv) ERCC excision repair 5, endonuclease (ERCC5, best known as XPG); and (v) an ERCC1/ERCC4 heterodimer (best known as XPF). Ultimately, this results in the incision of the damaged strand on both sides of the lesion by XPG and XPF, followed by DNA repair synthesis (Fig. 1; ref. 124). Only a few NER components have been implicated in the control of inflammatory reactions in cancer.

Defects in XPA have long been known to promote carcinogenesis in both humans and mice, but such an effect was largely attributed to genetic mechanisms (10). However, accelerated oncogenesis as driven by UV exposure and chemical carcinogens in Xpa−/− mice appears to involve an inflammatory component. Specifically, Xpa−/− mice exposed to UV light or topical dimethylbenz(a)anthracene (DMBA), a polycyclic aromatic hydrocarbon with pronounced oncogenic potential (263), develop a tumor-promoting dermal inflammatory response encompassing production of TNF, IL10, and prostaglandin E2 (PGE2) and accompanied by signs of systemic immunosuppression (264, 265). Along similar lines, Xpa−/− (but not WT) mice accumulate CXCL1 both dermally and systemically upon UVB exposure, and both CXCL1 neutralization with a specific antibody and systemic administration of an antioxidant limit skin oncogenesis in this model (266).

Finally, lung inflammation elicited by the intratracheal administration of lipopolysaccharide considerably impairs NER along with the downregulation of XPA and ERCC4 (267), providing yet another example of a vicious cycle linking TPI to DDR defects and vice versa (Fig. 2). Lending further support to this possibility, DSS/AOM-driven colitis is associated with a pronounced downregulation of XPF prior to overt oncogenesis (268). The relative contribution of XPF downregulation to inflammation-driven colorectal oncogenesis, however, remains to be formally defined.

In summary, components of the NER pathway regulate inflammatory responses that influence oncogenesis, tumor progression or response to treatment (Supplementary Table S7).

The FA Pathway

The FA pathway repairs DNA interstrand crosslinks (ICL), toxic lesions that interfere with DNA replication and transcription (269). Schematically, ICLs are detected by FA complementation group M (FANCM) and FA core complex associated protein 24 (FAAP24) resulting in the recruitment of a large multimeric ubiquitin ligase commonly known as the FA core complex to lesioned DNA. The core complex monoubiquitinates FANCD2 and FA complementation group I (FANCI), resulting in their recruitment to DNA repair foci. ICL repair is then carried out by a multitude of DDR factors, including BRCA1 interacting helicase 1 (BRIP1, best known as FANCJ), BRCA1, BRCA2, and PALB2 (Fig. 1; ref. 269). ­Germline mutations in any FA components (including BRCA1 and BRCA2) cause a rare human genetic disease characterized by bone marrow failure, genomic instability, and increased predisposition for acute myeloid leukemia and squamous cell carcinoma (269). Moreover, somatic mutations in FA genes are common in FA-independent tumors, especially squamous cell carcinoma as well as bladder and pancreatic cancer (269). Intriguingly, at least some tumors bearing somatic alterations in FA genes exhibit increased sensitivity to PARP inhibition (269). Whether this is related to the immunostimulatory activity of PARP inhibitors, though, remains to be elucidated.

Although the contribution of dysregulated inflammation to FA remains to be established, multiple FA proteins have been shown to interact with the control of inflammatory responses. For instance, FANCD2, FANCA, FANCC, FANCF, and FANCL (as well as BRCA1 and BRCA2) all appear to be required for the parkin RBR E3 ubiquitin protein ligase (PRKN)-dependent autophagic removal of dysfunctional mitochondria (mitophagy; ref. 270), which otherwise can activate a variety of inflammatory pathways (53). In line with this notion, the loss of FANCD2, BRCA1, or BRCA2 can promote CGAS activation coupled to TNF secretion and simultaneously restore sensitivity to exogenous TNF in multiple human cancer cell lines (137). Although such a proinflammatory effect was accompanied by the accumulation of CGAS-positive micronuclei (137), and at least in some models mechanistically involved the derepression of CGAS-activating type I long interspersed elements (LINE-1; ref. 271), the relative contribution of mitochondrial dysfunction to cytokine production driven by FANCD2 defects remains elusive.

Deletion of Fancd2 (or Fanca) from the mouse bone marrow results in the development of defective Tregs (272), which at least in part may contribute to the immune dysregulation of FA patients. Such a defect may originate from the hypersensitivity of FANCD2-deficient hematopoietic stem cells (HSC) to inflammatory cues, resulting in a rewiring of bioenergetic metabolism toward fatty acid oxidation and oxidative phosphorylation (273). A partially compromised Treg compartment may also account for the hypersensitivity of Fancd2−/− mice to carcinogen-driven skin oncogenesis (274). In this latter setting, FANCD2 has also been proposed to suppress oncogenesis by stabilizing TP63 (best known as p63; ref. 274). The consequences of this interaction on local inflammation, however, remain unexplored. Further supporting the notion that an intact FA pathway mediates anti-inflammatory effects, FANCA expression in HNSCC cells has been associated with radioresistance along with downregulation of IFN signaling coupled to an accrued secretion of immunomodulatory senescence-associated secretory phenotype (SASP) components (275). Conversely, high levels of FANCI have been linked to tumor infiltration by CD8+ and CD4+ T cells (but poor prognosis) in a cohort of patients with cervical cancer (276). The reasons underlying such an apparent discrepancy remain to be elucidated and may be linked to FA-unrelated functions of specific FA components.

Fancc−/− mice exhibit defects in IFNγ secretion by CD4+ T cells, presumably linked to the ability of FANCC to physically interact with (and support the function of) STAT1 in response to cytokine stimulation (277, 278). That said, FANCC has also been proposed to interact with heat-shock proteins, hence limiting the detrimental effects of inflammation on HSCs (279). In line with this notion, HSCs from FA patients with defects in FANCC are hypersensitive to IFNγ (280).

In summary, the FA pathway globally resembles other DDR cascades in its functions at the crossroad of genetic stability and inflammatory homeostasis (Supplementary Table S7).

The abundant literature discussed above delineates a leitmotif whereby the DDR, in most (if not all) of its variants, constitutively operates to preserve genetic and inflammatory homeostasis. Thus, DDR defects not only favor the accumulation of genetic lesions that may drive oncogenesis and tumor progression but also initiate inflammatory responses that can influence developing neoplasms in an ambiguous fashion. On the one hand, TPI may support disease progression and resistance to therapy. On the other hand, local inflammation may favor immunosurveillance and instead support sensitivity to (immuno)therapy. In line with this notion, several clinical trials are currently investigating the therapeutic profile of DDR inhibitors optionally combined with ICIs or other immunotherapeutic agents (Table 2). Preliminary findings from some of these studies are encouraging, generally pointing to an acceptable toxicity profile and at least some degree of clinical activity (281–283). Moreover, various ICIs are being used in multiple clinical indications (and extensively tested in clinical trials) together with agents that cause DNA damage along with inflammatory reactions, including several chemotherapeutics (29) as well as RT (30). Although the actual implication of DDR-driven inflammatory responses in the efficacy of these latter combinations remains to be formally demonstrated in patients, abundant preclinical data support this hypothesis. In this context, it is worth noting that cellular senescence driven by DDR signaling appears to be particularly immunogenic (284). Finally, the ability of DNA-damaging agents to elicit tumor-targeting immune responses has been shown to emerge from injured/dying (but hitherto alive) cancer cells (108, 109). Thus, it is tempting to speculate (but remains to be formally demonstrated) that the use of DNA-damaging agents at suboptimal doses (favoring cellular injury or senescence over rapid RCD) may result in superior synergism with ICIs or other immunotherapeutics. Additional work is required to clarify this possibility. At least theoretically, DDR inhibitors also stand out as promising combinatorial partners for exogenously administered PRR agonists (e.g., STING1 agonists, oncolytic viruses). Indeed, although DDR inhibition may per se drive robust PRR signaling in (at least some) malignant cells (Fig. 3), local PRR agonism may also engage nonmalignant components of the TME (which are refractory to DDR inhibitors). With the exception of two studies pointing to a positive interaction between ATM or ATR inhibitors and oncolytic viruses (not necessarily in the context of superior tumor-targeting immunity; refs. 285, 286), this possibility remains largely unexplored.

Table 2.

Clinical trials testing DDR-targeting agents plus immunotherapy in cancer patientsa

TargetAgentSettingPhaseStatusImmunotherapyBiomarkersNote(s)Trial ID
ATR AZD6738 Bile duct cancer Phase II Recruiting Durvalumab — After immunotherapy NCT04298008 
ATR AZD6738 Bile duct cancer Phase II Recruiting Durvalumab — After first-line chemotherapy NCT04298021 
ATR AZD6738 Gastric cancer Phase II Active, not recruiting Durvalumab — After second-line chemotherapy NCT03780608 
ATR AZD6738 HNSCC Phase I/II Recruiting Durvalumab — Advanced cancer NCT02264678 
  NSCLC       
ATR AZD6738 Melanoma Phase II Active, not recruiting Durvalumab — After PD-1/PD-L1 blockage NCT03780608 
ATR AZD6738 NSCLC Phase II Active, not recruiting Durvalumab KRAS mutations Biomarker-directed trial NCT02664935 
ATR AZD6738 NSCLC Phase II Recruiting Durvalumab — After PD-1/PD-L1 blockage NCT03334617 
ATR AZD6738 NSCLC Phase II Recruiting Durvalumab — After PD-1 blockage NCT03833440 
ATR AZD6738 SCLC Phase II Active, not recruiting Durvalumab — After second- or third-line therapy NCT04361825 
ATR AZD6738 SCLC Phase II Recruiting Durvalumab — Combination with cisplatin or carboplatin and etoposide NCT04699838 
ATR Berzosertib NSCLC Phase I/II Recruiting Pembrolizumab — Combination with gemcitabine and carboplatin NCT04216316 
ATR Berzosertib Solid tumors Phase I/II Recruiting Avelumab DDR defects Germline mutations NCT04266912 
ATR Elimusertib HNSCC Phase I Recruiting Pembrolizumab — Combination with SBRT NCT04576091 
ATR Elimusertib Solid tumors Phase I Active, not recruiting Pembrolizumab DDR defects Dose-escalation of elimusertib NCT04095273 
ATR M1774 Solid tumors Phase I Recruiting ICI (not specified) — Combination with additional DDR inhibitor (not specified) NCT05396833 
DNA-PKcs M3814 Hepatobiliary cancer Phase I/II Recruiting Avelumab — Combination with hypofractionated RT NCT04068194 
DNA-PKcs M3814 Prostate cancer Phase I/II Recruiting Avelumab — Combination with radium-223 dichloride NCT04071236 
MDM2 APG-115 Liposarcoma Phase II Recruiting Toripalimab Wild-type TP53 Biomarker-directed trial NCT04785196 
      MDM2 amplification   
MDM2 APG-115 Melanoma Phase I/II Recruiting Pembrolizumab — Dose-escalation of APG-11 NCT03611868 
MDM2 HDM201 AML Phase I Active, not recruiting MBG453 Wild-type TP53 Optionally combined with venetoclax NCT03940352 
  MDS       
MDM2 Navtemadlin Merkel cell carcinoma Phase I/II Recruiting Avelumab Wild-type TP53 No prior immunotherapy NCT03787602 
PARP Fluzoparib HNSCC Phase II Recruiting Camrelizumab — After first-line chemotherapy NCT04978012 
PARP Fluzoparib NSCLC Phase II Recruiting Tirelizumab — Fluzoparib as maintenance therapy NCT05392686 
PARP Fluzoparib SCLC n/a Not yet recruiting Camrelizumab — As consolidation treatment NCT04782089 
PARP Niraparib Breast cancer Phase II Recruiting HX008 DDR defects Definite pathogenic or suspected germline mutations NCT04508803 
PARP Niraparib Gynecologic tumors Phase II/III Recruiting Dostarlimab — After first-line chemotherapy NCT03651206 
PARP Niraparib Gynecologic tumors Phase III Active, not recruiting Atezolizumab — Combination with platinum-based chemotherapy NCT03598270 
PARP Niraparib HNSCC Phase II Recruiting Dostarlimab — HPV-negative tumors NCT04681469 
PARP Niraparib Mesothelioma Phase II Recruiting Dostarlimab — Platinum-sensitive tumors NCT03654833 
PARP Niraparib NSCLC Phase III Recruiting Pembrolizumab — As maintenance therapy NCT04475939 
PARP Niraparib Ovarian cancer Phase II Not yet recruiting Dostarlimab — Combination with bevacizumab NCT05065021 
PARP Niraparib Ovarian cancer Phase I/II Recruiting Dostarlimab — Combination with bevacizumab NCT03574779 
PARP Niraparib SCLC Phase II Active, not recruiting Dostarlimab — Including other high-grade neuroendocrine carcinomas NCT04701307 
PARP Niraparib TNBC Phase II Recruiting Dostarlimab — Combination with RT NCT04837209 
PARP Not specified SCLC n/a Not yet recruiting ICI (not specified) — Combination with RT and temozolomide NCT04790955 
PARP Olaparib Bladder cancer Phase I Active, not recruiting Durvalumab HR defects Biomarker-directed trial NCT02546661 
 Olaparib Breast cancer Phase II Recruiting Durvalumab BRCA1/2 mutations Germline or somatic mutations NCT03025035 
      HR deficiency   
PARP Olaparib Breast cancer Phase II Not yet recruiting Pembrolizumab DDR defects Deleterious germline mutations irrespective of HR status NCT05033756 
PARP Olaparib Breast cancer Phase I/II Recruiting Durvalumab — Optionally combined with cediranib NCT02484404 
  Colorectal cancer       
PARP Olaparib Breast cancer Phase II Active, not recruiting Pembrolizumab — Combination with carboplatin and gemcitabine NCT04191135 
PARP Olaparib Breast cancer Phase II Active, not recruiting Atezolizumab BRCA1/2 mutations Germline or somatic mutations NCT02849496 
PARP Olaparib Breast cancer Phase I/II Active, not recruiting Durvalumab BRCA1/2 mutations Germline mutations NCT02734004 
  Ovarian cancer       
PARP Olaparib Cervical cancer Phase II Recruiting Pembrolizumab — Competent or deficient FA repair NCT04483544 
PARP Olaparib Colorectal cancer Phase II Active, not recruiting Durvalumab — MMR proficient NCT03851614 
  Pancreatic cancer       
PARP Olaparib Endometrial cancer Phase II Active, not recruiting Durvalumab — — NCT03951415 
PARP Olaparib Gastric cancer Phase II Recruiting Pembrolizumab DDR defects Combination with SBRT NCT05379972 
PARP Olaparib Glioma Phase II Recruiting Durvalumab IDH1 mutation Biomarker-directed trial NCT03991832 
PARP Olaparib Gynecologic tumors Phase I/II Active, not recruiting Tremelimumab BRCA1/2 mutations Germline mutations NCT02571725 
PARP Olaparib Gynecologic tumors Phase II Active, not recruiting Durvalumab BRCA1/2 mutations Germline or somatic mutations NCT02953457 
PARP Olaparib Gynecologic tumors Phase II Recruiting Durvalumab BRCA1/2 mutations Combination with cediranib maleate NCT04739800 
PARP Olaparib Melanoma Phase II Recruiting Pembrolizumab HR defects Biomarker-directed trial NCT04633902 
PARP Olaparib NSCLC Phase III Recruiting Pembrolizumab — Combination with RT NCT04380636 
PARP Olaparib NSCLC Phase III Active, not recruiting Pembrolizumab — Combination with chemotherapy NCT03976362 
PARP Olaparib NSCLC Phase III Active, not recruiting Pembrolizumab — Combination with chemotherapy NCT03976323 
PARP Olaparib Ovarian cancer Phase II Active, not recruiting Tremelimumab — Platinum-sensitive tumors NCT04034927 
PARP Olaparib Pancreatic cancer Phase II Recruiting Pembrolizumab BRCA1/2 mutations Germline mutations NCT04548752 
PARP Olaparib Pancreatic cancer Phase II Recruiting Pembrolizumab HR defects Platinum-sensitive tumors NCT04666740 
PARP Olaparib Prostate cancer Phase III Active, not recruiting Pembrolizumab — HR-proficient lesions NCT03834519 
PARP Olaparib SCLC Phase I/II Recruiting Durvalumab — Combination with carboplatin, etoposide, and/or RT NCT04728230 
PARP Pamiparib Solid tumors Phase III Enrolling by invitation Tislelizumab — Dose escalation NCT04164199 
PARP Rucaparib Biliary tract cancer Phase II Active, not recruiting Nivolumab — After platinum-based chemotherapy NCT03639935 
PARP Rucaparib Ovarian cancer Phase III Active, not recruiting Nivolumab — After platinum-based chemotherapy NCT03522246 
PARP Rucaparib Solid tumors Phase II Recruiting Atezolizumab DDR defects Platinum-sensitive tumors NCT04276376 
PARP Talazoparib AML Phase I/II Not yet recruiting NK cells — After chemotherapy-based conditioning NCT05319249 
PARP Talazoparib Breast cancer Phase I/II Active, not recruiting Avelumab — As maintenance therapy NCT03964532 
PARP Talazoparib Solid tumors Phase I/II Active, not recruiting Avelumab BRCA1/2 mutations Dose escalation NCT03330405 
PARP Talazoparib Solid tumors Phase II Active, not recruiting Avelumab BRCA1/2 mutationsATM mutations Biomarker-directed trial NCT03565991 
      ATM mutations   
WEE1 Adavosertib Solid tumors Phase I Active, not recruiting Durvalumab — Dose-escalation of adavosertib NCT02617277 
TargetAgentSettingPhaseStatusImmunotherapyBiomarkersNote(s)Trial ID
ATR AZD6738 Bile duct cancer Phase II Recruiting Durvalumab — After immunotherapy NCT04298008 
ATR AZD6738 Bile duct cancer Phase II Recruiting Durvalumab — After first-line chemotherapy NCT04298021 
ATR AZD6738 Gastric cancer Phase II Active, not recruiting Durvalumab — After second-line chemotherapy NCT03780608 
ATR AZD6738 HNSCC Phase I/II Recruiting Durvalumab — Advanced cancer NCT02264678 
  NSCLC       
ATR AZD6738 Melanoma Phase II Active, not recruiting Durvalumab — After PD-1/PD-L1 blockage NCT03780608 
ATR AZD6738 NSCLC Phase II Active, not recruiting Durvalumab KRAS mutations Biomarker-directed trial NCT02664935 
ATR AZD6738 NSCLC Phase II Recruiting Durvalumab — After PD-1/PD-L1 blockage NCT03334617 
ATR AZD6738 NSCLC Phase II Recruiting Durvalumab — After PD-1 blockage NCT03833440 
ATR AZD6738 SCLC Phase II Active, not recruiting Durvalumab — After second- or third-line therapy NCT04361825 
ATR AZD6738 SCLC Phase II Recruiting Durvalumab — Combination with cisplatin or carboplatin and etoposide NCT04699838 
ATR Berzosertib NSCLC Phase I/II Recruiting Pembrolizumab — Combination with gemcitabine and carboplatin NCT04216316 
ATR Berzosertib Solid tumors Phase I/II Recruiting Avelumab DDR defects Germline mutations NCT04266912 
ATR Elimusertib HNSCC Phase I Recruiting Pembrolizumab — Combination with SBRT NCT04576091 
ATR Elimusertib Solid tumors Phase I Active, not recruiting Pembrolizumab DDR defects Dose-escalation of elimusertib NCT04095273 
ATR M1774 Solid tumors Phase I Recruiting ICI (not specified) — Combination with additional DDR inhibitor (not specified) NCT05396833 
DNA-PKcs M3814 Hepatobiliary cancer Phase I/II Recruiting Avelumab — Combination with hypofractionated RT NCT04068194 
DNA-PKcs M3814 Prostate cancer Phase I/II Recruiting Avelumab — Combination with radium-223 dichloride NCT04071236 
MDM2 APG-115 Liposarcoma Phase II Recruiting Toripalimab Wild-type TP53 Biomarker-directed trial NCT04785196 
      MDM2 amplification   
MDM2 APG-115 Melanoma Phase I/II Recruiting Pembrolizumab — Dose-escalation of APG-11 NCT03611868 
MDM2 HDM201 AML Phase I Active, not recruiting MBG453 Wild-type TP53 Optionally combined with venetoclax NCT03940352 
  MDS       
MDM2 Navtemadlin Merkel cell carcinoma Phase I/II Recruiting Avelumab Wild-type TP53 No prior immunotherapy NCT03787602 
PARP Fluzoparib HNSCC Phase II Recruiting Camrelizumab — After first-line chemotherapy NCT04978012 
PARP Fluzoparib NSCLC Phase II Recruiting Tirelizumab — Fluzoparib as maintenance therapy NCT05392686 
PARP Fluzoparib SCLC n/a Not yet recruiting Camrelizumab — As consolidation treatment NCT04782089 
PARP Niraparib Breast cancer Phase II Recruiting HX008 DDR defects Definite pathogenic or suspected germline mutations NCT04508803 
PARP Niraparib Gynecologic tumors Phase II/III Recruiting Dostarlimab — After first-line chemotherapy NCT03651206 
PARP Niraparib Gynecologic tumors Phase III Active, not recruiting Atezolizumab — Combination with platinum-based chemotherapy NCT03598270 
PARP Niraparib HNSCC Phase II Recruiting Dostarlimab — HPV-negative tumors NCT04681469 
PARP Niraparib Mesothelioma Phase II Recruiting Dostarlimab — Platinum-sensitive tumors NCT03654833 
PARP Niraparib NSCLC Phase III Recruiting Pembrolizumab — As maintenance therapy NCT04475939 
PARP Niraparib Ovarian cancer Phase II Not yet recruiting Dostarlimab — Combination with bevacizumab NCT05065021 
PARP Niraparib Ovarian cancer Phase I/II Recruiting Dostarlimab — Combination with bevacizumab NCT03574779 
PARP Niraparib SCLC Phase II Active, not recruiting Dostarlimab — Including other high-grade neuroendocrine carcinomas NCT04701307 
PARP Niraparib TNBC Phase II Recruiting Dostarlimab — Combination with RT NCT04837209 
PARP Not specified SCLC n/a Not yet recruiting ICI (not specified) — Combination with RT and temozolomide NCT04790955 
PARP Olaparib Bladder cancer Phase I Active, not recruiting Durvalumab HR defects Biomarker-directed trial NCT02546661 
 Olaparib Breast cancer Phase II Recruiting Durvalumab BRCA1/2 mutations Germline or somatic mutations NCT03025035 
      HR deficiency   
PARP Olaparib Breast cancer Phase II Not yet recruiting Pembrolizumab DDR defects Deleterious germline mutations irrespective of HR status NCT05033756 
PARP Olaparib Breast cancer Phase I/II Recruiting Durvalumab — Optionally combined with cediranib NCT02484404 
  Colorectal cancer       
PARP Olaparib Breast cancer Phase II Active, not recruiting Pembrolizumab — Combination with carboplatin and gemcitabine NCT04191135 
PARP Olaparib Breast cancer Phase II Active, not recruiting Atezolizumab BRCA1/2 mutations Germline or somatic mutations NCT02849496 
PARP Olaparib Breast cancer Phase I/II Active, not recruiting Durvalumab BRCA1/2 mutations Germline mutations NCT02734004 
  Ovarian cancer       
PARP Olaparib Cervical cancer Phase II Recruiting Pembrolizumab — Competent or deficient FA repair NCT04483544 
PARP Olaparib Colorectal cancer Phase II Active, not recruiting Durvalumab — MMR proficient NCT03851614 
  Pancreatic cancer       
PARP Olaparib Endometrial cancer Phase II Active, not recruiting Durvalumab — — NCT03951415 
PARP Olaparib Gastric cancer Phase II Recruiting Pembrolizumab DDR defects Combination with SBRT NCT05379972 
PARP Olaparib Glioma Phase II Recruiting Durvalumab IDH1 mutation Biomarker-directed trial NCT03991832 
PARP Olaparib Gynecologic tumors Phase I/II Active, not recruiting Tremelimumab BRCA1/2 mutations Germline mutations NCT02571725 
PARP Olaparib Gynecologic tumors Phase II Active, not recruiting Durvalumab BRCA1/2 mutations Germline or somatic mutations NCT02953457 
PARP Olaparib Gynecologic tumors Phase II Recruiting Durvalumab BRCA1/2 mutations Combination with cediranib maleate NCT04739800 
PARP Olaparib Melanoma Phase II Recruiting Pembrolizumab HR defects Biomarker-directed trial NCT04633902 
PARP Olaparib NSCLC Phase III Recruiting Pembrolizumab — Combination with RT NCT04380636 
PARP Olaparib NSCLC Phase III Active, not recruiting Pembrolizumab — Combination with chemotherapy NCT03976362 
PARP Olaparib NSCLC Phase III Active, not recruiting Pembrolizumab — Combination with chemotherapy NCT03976323 
PARP Olaparib Ovarian cancer Phase II Active, not recruiting Tremelimumab — Platinum-sensitive tumors NCT04034927 
PARP Olaparib Pancreatic cancer Phase II Recruiting Pembrolizumab BRCA1/2 mutations Germline mutations NCT04548752 
PARP Olaparib Pancreatic cancer Phase II Recruiting Pembrolizumab HR defects Platinum-sensitive tumors NCT04666740 
PARP Olaparib Prostate cancer Phase III Active, not recruiting Pembrolizumab — HR-proficient lesions NCT03834519 
PARP Olaparib SCLC Phase I/II Recruiting Durvalumab — Combination with carboplatin, etoposide, and/or RT NCT04728230 
PARP Pamiparib Solid tumors Phase III Enrolling by invitation Tislelizumab — Dose escalation NCT04164199 
PARP Rucaparib Biliary tract cancer Phase II Active, not recruiting Nivolumab — After platinum-based chemotherapy NCT03639935 
PARP Rucaparib Ovarian cancer Phase III Active, not recruiting Nivolumab — After platinum-based chemotherapy NCT03522246 
PARP Rucaparib Solid tumors Phase II Recruiting Atezolizumab DDR defects Platinum-sensitive tumors NCT04276376 
PARP Talazoparib AML Phase I/II Not yet recruiting NK cells — After chemotherapy-based conditioning NCT05319249 
PARP Talazoparib Breast cancer Phase I/II Active, not recruiting Avelumab — As maintenance therapy NCT03964532 
PARP Talazoparib Solid tumors Phase I/II Active, not recruiting Avelumab BRCA1/2 mutations Dose escalation NCT03330405 
PARP Talazoparib Solid tumors Phase II Active, not recruiting Avelumab BRCA1/2 mutationsATM mutations Biomarker-directed trial NCT03565991 
      ATM mutations   
WEE1 Adavosertib Solid tumors Phase I Active, not recruiting Durvalumab — Dose-escalation of adavosertib NCT02617277 

Abbreviations: AML, acute myeloid leukemia; MDS, myelodysplastic syndrome.

aAs per www.clinicaltrials.gov, limited to studies with “Not yet recruiting,” “Recruiting,” “Enrolling by invitation,” and “Active, not recruiting” status.

Figure 3.

Main interactions between DDR components and PRRs. Multiple components of the DDR interact with PRRs to control inflammatory responses. Such interactions can be either activating or inhibitory, and include (i) direct binding, (ii) indirect functional modulation, and (iii) altered availability of PRR agonists.

Figure 3.

Main interactions between DDR components and PRRs. Multiple components of the DDR interact with PRRs to control inflammatory responses. Such interactions can be either activating or inhibitory, and include (i) direct binding, (ii) indirect functional modulation, and (iii) altered availability of PRR agonists.

Close modal

One of the potential obstacles to the use of DDR inhibitors in cancer patients relates to a somehow nonnegligible potential for secondary, therapy-driven oncogenesis. This has already been documented for DNA-damaging therapeutics including RT (287), and may call for intratumoral (instead of systemic) drug administration to limit exposure for healthy tissues. From a biological perspective, one of the questions that requires immediate experimental attention involves the extent to which modulation of DDR components influences inflammation by (i) controlling the abundance of PRR agonists (as in the case of EXO1; ref. 125), (ii) engage DDR signal transducers (as in the case of p53 activation by CHEK2; ref. 8), or (iii) activate DDR-unrelated pathways (Fig. 4). Potential DDR-independent mechanisms that may be involved in the ability of some DDR components to regulate inflammation include control of redox homeostasis (as in the case APE1; ref. 52), modulation of the gut microbiome (as in the case of MSH2; refs. 69, 70), direct interactions between DDR components and PRRs or signal transducers thereof (as in the case of Ku70; refs. 98, 99), and an ectopic activity as PRR (as in the case of MRE11; Fig. 4; ref. 113). Precisely defining the signaling cascades that connect DDR components to the molecular machinery for inflammation may uncover novel targets that enable improved anticancer immune responses without weakening genetic homeostasis.

Figure 4.

Mechanisms through which the DDR can influence inflammation. DDR components can regulate inflammatory reactions by at least six different mechanisms, including (i) actively generating PRR agonists, (ii) initiating DDR-dependent transcriptional programs, (iii) modulating redox homeostasis, (iv) altering the gut microbiome, (v) physically interacting with PRR or signal transducers thereof, and (vi) directly operating as PRRs from ectopic subcellular localizations.

Figure 4.

Mechanisms through which the DDR can influence inflammation. DDR components can regulate inflammatory reactions by at least six different mechanisms, including (i) actively generating PRR agonists, (ii) initiating DDR-dependent transcriptional programs, (iii) modulating redox homeostasis, (iv) altering the gut microbiome, (v) physically interacting with PRR or signal transducers thereof, and (vi) directly operating as PRRs from ectopic subcellular localizations.

Close modal

Despite these and other unknowns, DDR-targeting agents stand out as promising partners for immunotherapeutic agents including (but not limited to) ICIs, partly owing to their ability to modulate inflammation.

G. Kroemer reports grants from Kaleido, Lytix Pharma, PharmaMar, Osasuna Therapeutics, Samsara Therapeutics, Sanofi, Tollys, and Vascage during the conduct of the study; personal fees from Reithera, Hevolution, and Longevity Vision Funds outside the submitted work; two patents for WO2014020041-A1 and WO2014020043-A1 licensed to Bayer, a patent for WO2008057863-A1 licensed to Bristol Myers Squibb, a patent for WO2019057742A1 licensed to Osasuna Therapeutics, a patent for WO2022049270A1 licensed to PharmaMar, a patent for WO2022048775-A1 licensed to PharmaMar, a patent for EP2664326-A1 licensed to Raptor Pharmaceuticals, a patent for GB202017553D0 licensed to Samsara Therapeutics, and a patent for EP3684471A1 licensed to Therafast Bio; is a scientific cofounder of EverImmune, Osasuna Therapeutics, Samsara Therapeutics, and Therafast Bio; is on the board of directors for the Bristol Myers Squibb Foundation France; and is on the nonremunerated scientific advisory boards for Institut Servier. G. Kroemer's wife, Laurence Zitvogel, has held research contracts with GSK, Incyte, Lytix, Kaleido, Innovate Pharma, Daiichi Sankyo, Pilege, Merus, Transgene, 9 Meters Biopharma, Tusk, and Roche, was on the board of directors for Transgene, is a cofounder of EverImmune, and holds patents covering the treatment of cancer and the therapeutic manipulation of the microbiota. G. Kroemer's brother, Romano Kroemer, was an employee of Sanofi and now consults for Boehringer Ingelheim. A. Ciccia reports grants from the NIH, the Pershing Square Sohn Foundation, the Basser Center, the Mary Kay Foundation, and the Irma T. Hirschl and Monique Weill-Caulier Research Foundation during the conduct of the study. L. Galluzzi reports grants from Lytix, Promontory, and Onxeo, and personal fees from AstraZeneca, OmniSEQ, Longevity Labs, Inzen, the Luke Heller TECPR2 Foundation, Sotio, Onxeo, Noxopharm, Imvax, EduCom, and Boehringer Ingelheim outside the submitted work. No disclosures were reported by the other authors.

V. Klapp is supported by a grant from the Luxembourg National Research Fund (FNR; PRIDE19/14254520/i2TRON). A. Ciccia is supported by two R01 grants (#R01CA197774; #R01CA227450) and one P01 grant from NIH (#P01CA174653), by a Pershing Square Sohn Cancer Research (PSSCR) Award, by a BRCA Award from the Basser Center, by a Cancer Research Grant from the Mary Kay Foundation, and by the Irma T. Hirschl and Monique Weill-Caulier Research Award. G. Kroemer is supported by the Ligue contre le Cancer (équipe labellisée); Agence National de la Recherche (ANR)—Projets blancs; AMMICa US23/CNRS UMS3655; Association pour la recherche sur le cancer (ARC); Cancéropôle Ile-de-France; Fondation pour la Recherche Médicale (FRM); a donation by Elior; Equipex Onco-Pheno-Screen; European Joint Programme on Rare Diseases (EJPRD); Gustave Roussy Odyssea, the European Union Horizon 2020 Projects Oncobiome and Crimson; Fondation Carrefour; Institut National du Cancer (INCa); Institut Universitaire de France; LabEx Immuno-Oncology (ANR-18-IDEX-0001); a Cancer Research ASPIRE Award from the Mark Foundation; the RHU Immunolife; Seerave Foundation; SIRIC Stratified Oncology Cell DNA Repair and Tumor Immune Elimination (SOCRATE); and SIRIC Cancer Research and Personalized Medicine (CARPEM). This study contributes to the IdEx Université de Paris ANR-18-IDEX-0001. The laboratory of L.G. (as a PI unless otherwise indicated) is or has been supported by two Breakthrough Level 2 grants from the US DoD BCRP (#BC180476P1; #BC210945), by a Transformative Breast Cancer Consortium Grant from the US DoD BCRP (#W81XWH2120034, PI: Formenti), by a U54 grant from NIH/NCI (#CA274291, PI: Deasy, Formenti, Weichselbaum), by a STARR Cancer Consortium grant (#I16-0064), by the 2019 Laura Ziskin Prize in Translational Research (#ZP-6177, PI: Formenti) from the Stand Up to Cancer (SU2C), by a Mantle Cell Lymphoma Research Initiative (MCL-RI, PI: Chen-Kiang) grant from the Leukemia and Lymphoma Society (LLS), by a Rapid Response Grant from the Functional Genomics Initiative (New York, US), by startup funds from the Dept. of Radiation Oncology at Weill Cornell Medicine (New York, US), by industrial collaborations with Lytix Biopharma (Oslo, Norway), Promontory (New York, US) and Onxeo (Paris, France), as well as by donations from Promontory (New York, US), the Luke Heller TECPR2 Foundation (Boston, US), Sotio a.s. (Prague, Czech Republic), Lytix Biopharma (Oslo, Norway), Onxeo (Paris, France), Ricerchiamo (Brescia, Italy), and Noxopharm (Chatswood, Australia).

Note: Supplementary data for this article are available at Cancer Discovery Online (http://cancerdiscovery.aacrjournals.org/).

1.
Hanahan
D
.
Hallmarks of cancer: new dimensions
.
Cancer Discov
2022
;
12
:
31
46
.
2.
Galluzzi
L
,
Chan
TA
,
Kroemer
G
,
Wolchok
JD
,
Lopez-Soto
A
.
The hallmarks of successful anticancer immunotherapy
.
Sci Transl Med
2018
;
10
:
eaat7807
.
3.
Dersh
D
,
Hollý
J
,
Yewdell
JW
.
A few good peptides: MHC class I-based cancer immunosurveillance and immunoevasion
.
Nat Rev Immunol
2021
;
21
:
116
28
.
4.
Spranger
S
,
Gajewski
TF
.
Impact of oncogenic pathways on evasion of antitumour immune responses
.
Nat Rev Cancer
2018
;
18
:
139
47
.
5.
Lee-Six
H
,
Olafsson
S
,
Ellis
P
,
Osborne
RJ
,
Sanders
MA
,
Moore
L
, et al
.
The landscape of somatic mutation in normal colorectal epithelial cells
.
Nature
2019
;
574
:
532
7
.
6.
Martincorena
I
,
Fowler
JC
,
Wabik
A
,
Lawson
ARJ
,
Abascal
F
,
Hall
MWJ
, et al
.
Somatic mutant clones colonize the human esophagus with age
.
Science
2018
;
362
:
911
7
.
7.
Galluzzi
L
,
Yamazaki
T
,
Kroemer
G
.
Linking cellular stress responses to systemic homeostasis
.
Nat Rev Mol Cell Biol
2018
;
19
:
731
45
.
8.
Kang
TW
,
Yevsa
T
,
Woller
N
,
Hoenicke
L
,
Wuestefeld
T
,
Dauch
D
, et al
.
Senescence surveillance of pre-malignant hepatocytes limits liver cancer development
.
Nature
2011
;
479
:
547
51
.
9.
Textor
S
,
Fiegler
N
,
Arnold
A
,
Porgador
A
,
Hofmann
TG
,
Cerwenka
A
.
Human NK cells are alerted to induction of p53 in cancer cells by upregulation of the NKG2D ligands ULBP1 and ULBP2
.
Cancer Res
2011
;
71
:
5998
6009
.
10.
Cleaver
JE
,
Lam
ET
,
Revet
I
.
Disorders of nucleotide excision repair: the genetic and molecular basis of heterogeneity
.
Nat Rev Genet
2009
;
10
:
756
68
.
11.
Levine
AJ
.
p53: 800 million years of evolution and 40 years of discovery
.
Nat Rev Cancer
2020
;
20
:
471
80
.
12.
Lee
JH
,
Paull
TT
.
Cellular functions of the protein kinase ATM and their relevance to human disease
.
Nat Rev Mol Cell Biol
2021
;
22
:
796
814
.
13.
Roy
R
,
Chun
J
,
Powell
SN
.
BRCA1 and BRCA2: different roles in a common pathway of genome protection
.
Nat Rev Cancer
2011
;
12
:
68
78
.
14.
Lee
YR
,
Chen
M
,
Pandolfi
PP
.
The functions and regulation of the PTEN tumour suppressor: new modes and prospects
.
Nat Rev Mol Cell Biol
2018
;
19
:
547
62
.
15.
Kennedy
RD
,
D'Andrea
AD
.
DNA repair pathways in clinical practice: lessons from pediatric cancer susceptibility syndromes
.
J Clin Oncol
2006
;
24
:
3799
808
.
16.
Hopkins
JL
,
Lan
L
,
Zou
L
.
DNA repair defects in cancer and therapeutic opportunities
.
Genes Dev
2022
;
36
:
278
93
.
17.
Li
H
,
Zimmerman
SE
,
Weyemi
U
.
Genomic instability and metabolism in cancer
.
Int Rev Cell Mol Biol
2021
;
364
:
241
65
.
18.
Petroni
G
,
Buqué
A
,
Coussens
LM
,
Galluzzi
L
.
Targeting oncogene and non-oncogene addiction to inflame the tumour microenvironment
.
Nat Rev Drug Discov
2022
;
21
:
440
62
.
19.
Huang
A
,
Garraway
LA
,
Ashworth
A
,
Weber
B
.
Synthetic lethality as an engine for cancer drug target discovery
.
Nat Rev Drug Discov
2020
;
19
:
23
38
.
20.
Huang
RX
,
Zhou
PK
.
DNA damage response signaling pathways and targets for radiotherapy sensitization in cancer
.
Signal Transduct Target Ther
2020
;
5
:
60
.
21.
Bouwman
P
,
Jonkers
J
.
The effects of deregulated DNA damage signalling on cancer chemotherapy response and resistance
.
Nat Rev Cancer
2012
;
12
:
587
98
.
22.
Vanpouille-Box
C
,
Demaria
S
,
Formenti
SC
,
Galluzzi
L
.
Cytosolic DNA sensing in organismal tumor control
.
Cancer Cell
2018
;
34
:
361
78
.
23.
Chabanon
RM
,
Rouanne
M
,
Lord
CJ
,
Soria
JC
,
Pasero
P
,
Postel-Vinay
S
.
Targeting the DNA damage response in immuno-oncology: developments and opportunities
.
Nat Rev Cancer
2021
;
21
:
701
17
.
24.
Ciccia
A
,
Elledge
SJ
.
The DNA damage response: making it safe to play with knives
.
Mol Cell
2010
;
40
:
179
204
.
25.
Higgs
EF
,
Bao
R
,
Hatogai
K
,
Gajewski
TF
.
Wilms tumor reveals DNA repair gene hyperexpression is linked to lack of tumor immune infiltration
.
J Immunother Cancer
2022
;
10
:
e004797
.
26.
Lambert
WC
,
Lambert
MW
.
Development of effective skin cancer treatment and prevention in xeroderma pigmentosum
.
Photochem Photobiol
2015
;
91
:
475
83
.
27.
Vitale
I
,
Shema
E
,
Loi
S
,
Galluzzi
L
.
Intratumoral heterogeneity in cancer progression and response to immunotherapy
.
Nat Med
2021
;
27
:
212
24
.
28.
López-Soto
A
,
Gonzalez
S
,
Smyth
MJ
,
Galluzzi
L
.
Control of metastasis by NK cells
.
Cancer Cell
2017
;
32
:
135
54
.
29.
Galluzzi
L
,
Humeau
J
,
Buqué
A
,
Zitvogel
L
,
Kroemer
G
.
Immunostimulation with chemotherapy in the era of immune checkpoint inhibitors
.
Nat Rev Clin Oncol
2020
;
17
:
725
41
.
30.
Rodriguez-Ruiz
ME
,
Vitale
I
,
Harrington
KJ
,
Melero
I
,
Galluzzi
L
.
Immunological impact of cell death signaling driven by radiation on the tumor microenvironment
.
Nat Immunol
2020
;
21
:
120
34
.
31.
Petroni
G
,
Buqué
A
,
Zitvogel
L
,
Kroemer
G
,
Galluzzi
L
.
Immunomodulation by targeted anticancer agents
.
Cancer Cell
2021
;
39
:
310
45
.
32.
Hou
J
,
Karin
M
,
Sun
B
.
Targeting cancer-promoting inflammation - have anti-inflammatory therapies come of age?
Nat Rev Clin Oncol
2021
;
18
:
261
79
.
33.
Propper
DJ
,
Balkwill
FR
.
Harnessing cytokines and chemokines for cancer therapy
.
Nat Rev Clin Oncol
2022
;
19
:
237
53
.
34.
Pittet
MJ
,
Michielin
O
,
Migliorini
D
.
Clinical relevance of tumour-associated macrophages
.
Nat Rev Clin Oncol
2022
;
19
:
402
21
.
35.
Togashi
Y
,
Shitara
K
,
Nishikawa
H
.
Regulatory T cells in cancer immunosuppression: implications for anticancer therapy
.
Nat Rev Clin Oncol
2019
;
16
:
356
71
.
36.
Philip
M
,
Schietinger
A
.
CD8(+) T cell differentiation and dysfunction in cancer
.
Nat Rev Immunol
2022
;
22
:
209
23
.
37.
Vitale
I
,
Manic
G
,
De Maria
R
,
Kroemer
G
,
Galluzzi
L
.
DNA damage in stem cells
.
Mol Cell
2017
;
66
:
306
19
.
38.
Zhao
S
,
Tadesse
S
,
Kidane
D
.
Significance of base excision repair to human health
.
Int Rev Cell Mol Biol
2021
;
364
:
163
93
.
39.
Renaudin
X
.
Reactive oxygen species and DNA damage response in cancer
.
Int Rev Cell Mol Biol
2021
;
364
:
139
61
.
40.
Krokan
HE
,
Bjoras
M
.
Base excision repair
.
Cold Spring Harb Perspect Biol
2013
;
5
:
a012583
.
41.
Visnes
T
,
Cazares-Korner
A
,
Hao
W
,
Wallner
O
,
Masuyer
G
,
Loseva
O
, et al
.
Small-molecule inhibitor of OGG1 suppresses proinflammatory gene expression and inflammation
.
Science
2018
;
362
:
834
9
.
42.
Touati
E
,
Michel
V
,
Thiberge
JM
,
Ave
P
,
Huerre
M
,
Bourgade
F
, et al
.
Deficiency in OGG1 protects against inflammation and mutagenic effects associated with H. pylori infection in mouse
.
Helicobacter
2006
;
11
:
494
505
.
43.
Liao
J
,
Seril
DN
,
Lu
GG
,
Zhang
M
,
Toyokuni
S
,
Yang
AL
, et al
.
Increased susceptibility of chronic ulcerative colitis-induced carcinoma development in DNA repair enzyme Ogg1 deficient mice
.
Mol Carcinog
2008
;
47
:
638
46
.
44.
Simon
H
,
Vartanian
V
,
Wong
MH
,
Nakabeppu
Y
,
Sharma
P
,
Lloyd
RS
, et al
.
OGG1 deficiency alters the intestinal microbiome and increases intestinal inflammation in a mouse model
.
PLoS One
2020
;
15
:
e0227501
.
45.
Meira
LB
,
Bugni
JM
,
Green
SL
,
Lee
CW
,
Pang
B
,
Borenshtein
D
, et al
.
DNA damage induced by chronic inflammation contributes to colon carcinogenesis in mice
.
J Clin Invest
2008
;
118
:
2516
25
.
46.
Casorelli
I
,
Pannellini
T
,
De Luca
G
,
Degan
P
,
Chiera
F
,
Iavarone
I
, et al
.
The Mutyh base excision repair gene influences the inflammatory response in a mouse model of ulcerative colitis
.
PLoS One
2010
;
5
:
e12070
.
47.
Xie
Y
,
Yang
H
,
Cunanan
C
,
Okamoto
K
,
Shibata
D
,
Pan
J
, et al
.
Deficiencies in mouse Myh and Ogg1 result in tumor predisposition and G to T mutations in codon 12 of the K-ras oncogene in lung tumors
.
Cancer Res
2004
;
64
:
3096
102
.
48.
Chevillard
S
,
Radicella
JP
,
Levalois
C
,
Lebeau
J
,
Poupon
MF
,
Oudard
S
, et al
.
Mutations in OGG1, a gene involved in the repair of oxidative DNA damage, are found in human lung and kidney tumours
.
Oncogene
1998
;
16
:
3083
6
.
49.
Sieber
OM
,
Lipton
L
,
Crabtree
M
,
Heinimann
K
,
Fidalgo
P
,
Phillips
RK
, et al
.
Multiple colorectal adenomas, classic adenomatous polyposis, and germ-line mutations in MYH
.
N Engl J Med
2003
;
348
:
791
9
.
50.
O'Hara
AM
,
Bhattacharyya
A
,
Mifflin
RC
,
Smith
MF
,
Ryan
KA
,
Scott
KG
, et al
.
Interleukin-8 induction by Helicobacter pylori in gastric epithelial cells is dependent on apurinic/apyrimidinic endonuclease-1/redox factor-1
.
J Immunol
2006
;
177
:
7990
9
.
51.
O'Hara
AM
,
Bhattacharyya
A
,
Bai
J
,
Mifflin
RC
,
Ernst
PB
,
Mitra
S
, et al
.
Tumor necrosis factor (TNF)-alpha-induced IL-8 expression in gastric epithelial cells: role of reactive oxygen species and AP endonuclease-1/redox factor (Ref)-1
.
Cytokine
2009
;
46
:
359
69
.
52.
den Hartog
G
,
Chattopadhyay
R
,
Ablack
A
,
Hall
EH
,
Butcher
LD
,
Bhattacharyya
A
, et al
.
Regulation of Rac1 and reactive oxygen species production in response to infection of gastrointestinal epithelia
.
PLoS Pathog
2016
;
12
:
e1005382
.
53.
Marchi
S
,
Guilbaud
E
,
Tait
SWG
,
Yamazaki
T
,
Galluzzi
L
.
Mitochondrial control of inflammation
.
Nat Rev Immunol
2022
:
1
15
.
54.
Ng
GZ
,
Menheniott
TR
,
Every
AL
,
Stent
A
,
Judd
LM
,
Chionh
YT
, et al
.
The MUC1 mucin protects against Helicobacter pylori pathogenesis in mice by regulation of the NLRP3 inflammasome
.
Gut
2016
;
65
:
1087
99
.
55.
Jedinak
A
,
Dudhgaonkar
S
,
Kelley
MR
,
Sliva
D
.
Apurinic/Apyrimidinic endonuclease 1 regulates inflammatory response in macrophages
.
Anticancer Res
2011
;
31
:
379
85
.
56.
Akhter
N
,
Takeda
Y
,
Nara
H
,
Araki
A
,
Ishii
N
,
Asao
N
, et al
.
Apurinic/apyrimidinic endonuclease 1/redox factor-1 (Ape1/Ref-1) modulates antigen presenting cell-mediated T helper cell type 1 responses
.
J Biol Chem
2016
;
291
:
23672
80
.
57.
Zhao
S
,
Klattenhoff
AW
,
Thakur
M
,
Sebastian
M
,
Kidane
D
.
Mutation in DNA polymerase beta causes spontaneous chromosomal instability and inflammation-associated carcinogenesis in mice
.
Cancers (Basel)
2019
;
11
:
1160
.
58.
Zhao
S
,
Thakur
M
,
Klattenhoff
AW
,
Kidane
D
.
Aberrant DNA polymerase beta enhances H. pylori infection induced genomic instability and gastric carcinogenesis in mice
.
Cancers (Basel)
2019
;
11
:
843
.
59.
Huang
M
,
Wu
T
,
Liu
R
,
Wang
M
,
Shi
M
,
Xin
J
, et al
.
Polβ modulates the expression of type I interferon via STING pathway
.
Biochem Biophys Res Commun
2022
;
621
:
137
43
.
60.
Green
AR
,
Aleskandarany
MA
,
Ali
R
,
Hodgson
EG
,
Atabani
S
,
De Souza
K
, et al
.
Clinical impact of tumor DNA repair expression and T-cell infiltration in breast cancers
.
Cancer Immunol Res
2017
;
5
:
292
9
.
61.
Kunkel
TA
,
Erie
DA
.
Eukaryotic mismatch repair in relation to DNA replication
.
Annu Rev Genet
2015
;
49
:
291
313
.
62.
Ricciardiello
L
,
Ahnen
DJ
,
Lynch
PM
.
Chemoprevention of hereditary colon cancers: time for new strategies
.
Nat Rev Gastroenterol Hepatol
2016
;
13
:
352
61
.
63.
Pestana
RC
,
Sen
S
,
Hobbs
BP
,
Hong
DS
.
Histology-agnostic drug development - considering issues beyond the tissue
.
Nat Rev Clin Oncol
2020
;
17
:
555
68
.
64.
Anagnostou
V
,
Bardelli
A
,
Chan
TA
,
Turajlic
S
.
The status of tumor mutational burden and immunotherapy
.
Nat Cancer
2022
;
3
:
652
6
.
65.
Morad
G
,
Helmink
BA
,
Sharma
P
,
Wargo
JA
.
Hallmarks of response, resistance, and toxicity to immune checkpoint blockade
.
Cell
2021
;
184
:
5309
37
.
66.
Kohonen-Corish
MR
,
Daniel
JJ
,
te Riele
H
,
Buffinton
GD
,
Dahlstrom
JE
.
Susceptibility of Msh2-deficient mice to inflammation-associated colorectal tumors
.
Cancer Res
2002
;
62
:
2092
7
.
67.
Mladenova
D
,
Daniel
JJ
,
Dahlstrom
JE
,
Bean
E
,
Gupta
R
,
Pickford
R
, et al
.
The NSAID sulindac is chemopreventive in the mouse distal colon but carcinogenic in the proximal colon
.
Gut
2011
;
60
:
350
60
.
68.
Belcheva
A
,
Green
B
,
Weiss
A
,
Streutker
C
,
Martin
A
.
Elevated incidence of polyp formation in APC(Min/+)Msh2−/− mice is independent of nitric oxide-induced DNA mutations
.
PLoS One
2013
;
8
:
e65204
.
69.
Belcheva
A
,
Irrazabal
T
,
Robertson
SJ
,
Streutker
C
,
Maughan
H
,
Rubino
S
, et al
.
Gut microbial metabolism drives transformation of MSH2-deficient colon epithelial cells
.
Cell
2014
;
158
:
288
99
.
70.
Tosti
E
,
Almeida
AS
,
Tran
TTT
,
Barbachan
ESM
,
Broin
PO
,
Dubin
R
, et al
.
Loss of MMR and TGFBR2 increases the susceptibility to microbiota-dependent inflammation-associated colon cancer
.
Cell Mol Gastroenterol Hepatol
2022
;
14
:
693
717
.
71.
Eso
Y
,
Takai
A
,
Matsumoto
T
,
Inuzuka
T
,
Horie
T
,
Ono
K
, et al
.
MSH2 dysregulation is triggered by proinflammatory cytokine stimulation and is associated with liver cancer development
.
Cancer Res
2016
;
76
:
4383
93
.
72.
Salem
ME
,
Bodor
JN
,
Puccini
A
,
Xiu
J
,
Goldberg
RM
,
Grothey
A
, et al
.
Relationship between MLH1, PMS2, MSH2 and MSH6 gene-specific alterations and tumor mutational burden in 1057 microsatellite instability-high solid tumors
.
Int J Cancer
2020
;
147
:
2948
56
.
73.
Jia
M
,
Yao
L
,
Yang
Q
,
Chi
T
.
Association of MSH2 expression with tumor mutational burden and the immune microenvironment in lung adenocarcinoma
.
Front Oncol
2020
;
10
:
168
.
74.
McCoy
P
,
Mangiola
S
,
Macintyre
G
,
Hutchinson
R
,
Tran
B
,
Pope
B
, et al
.
MSH2-deficient prostate tumours have a distinct immune response and clinical outcome compared to MSH2-deficient colorectal or endometrial cancer
.
Prostate Cancer Prostatic Dis
2021
;
24
:
1167
80
.
75.
Xie
T
,
Feng
Q
,
Li
Z
,
Lu
M
,
Li
J
,
Lizaso
A
, et al
.
Heterogeneous constitutional mismatch repair deficiency with MSH6 missense mutation clinically benefits from pembrolizumab and regorafenib combination therapy: a case report and literature review
.
Hered Cancer Clin Pract
2021
;
19
:
7
.
76.
Danley
KT
,
Schmitz
K
,
Ghai
R
,
Sclamberg
JS
,
Buckingham
LE
,
Burgess
K
, et al
.
A durable response to pembrolizumab in a patient with uterine serous carcinoma and lynch syndrome due to the MSH6 germline mutation
.
Oncologist
2021
;
26
:
811
7
.
77.
Santin
AD
,
Bellone
S
,
Buza
N
,
Choi
J
,
Schwartz
PE
,
Schlessinger
J
, et al
.
Regression of chemotherapy-resistant polymerase ε (POLE) ultra-mutated and MSH6 hyper-mutated endometrial tumors with nivolumab
.
Clin Cancer Res
2016
;
22
:
5682
7
.
78.
Kir
G
,
Olgun
ZC
,
Soylemez
T
,
Aydin
A
,
Demircan
B
,
Kaya
IA
, et al
.
PD-L1 expression in mismatch repair-deficient endometrial carcinoma and tumor-associated immune cells: differences between MLH1 methylated and nonmethylated subgroups
.
Int J Gynecol Pathol
2021
;
40
:
575
86
.
79.
Maletzki
C
,
Gladbach
YS
,
Hamed
M
,
Fuellen
G
,
Semmler
ML
,
Stenzel
J
, et al
.
Cellular vaccination of MLH1(-/-) mice - an immunotherapeutic proof of concept study
.
Oncoimmunology
2018
;
7
:
e1408748
.
80.
Maletzki
C
,
Wiegele
L
,
Nassar
I
,
Stenzel
J
,
Junghanss
C
.
Chemo-immunotherapy improves long-term survival in a preclinical model of MMR-D-related cancer
.
J Immunother Cancer
2019
;
7
:
8
.
81.
Salewski
I
,
Henne
J
,
Engster
L
,
Schneider
B
,
Lemcke
H
,
Skorska
A
, et al
.
Combined gemcitabine and immune-checkpoint inhibition conquers anti-PD-L1 resistance in low-immunogenic mismatch repair-deficient tumors
.
Int J Mol Sci
2021
;
22
:
5990
.
82.
Salewski
I
,
Kuntoff
S
,
Kuemmel
A
,
Feldtmann
R
,
Felix
SB
,
Henze
L
, et al
.
Combined vaccine-immune-checkpoint inhibition constitutes a promising strategy for treatment of dMMR tumors
.
Cancer Immunol Immunother
2021
;
70
:
3405
19
.
83.
Germano
G
,
Lamba
S
,
Rospo
G
,
Barault
L
,
Magrì
A
,
Maione
F
, et al
.
Inactivation of DNA repair triggers neoantigen generation and impairs tumour growth
.
Nature
2017
;
552
:
116
20
.
84.
Lu
C
,
Guan
J
,
Lu
S
,
Jin
Q
,
Rousseau
B
,
Lu
T
, et al
.
DNA sensing in mismatch repair-deficient tumor cells is essential for anti-tumor immunity
.
Cancer Cell
2021
;
39
:
96
108
.
e6
.
85.
Maletzki
C
,
Beyrich
F
,
Hühns
M
,
Klar
E
,
Linnebacher
M
.
The mutational profile and infiltration pattern of murine MLH1−/− tumors: concurrences, disparities and cell line establishment for functional analysis
.
Oncotarget
2016
;
7
:
53583
98
.
86.
Mowat
C
,
Mosley
SR
,
Namdar
A
,
Schiller
D
,
Baker
K
.
Anti-tumor immunity in mismatch repair-deficient colorectal cancers requires type I IFN-driven CCL5 and CXCL10
.
J Exp Med
2021
;
218
:
e20210108
.
87.
Guan
J
,
Lu
C
,
Jin
Q
,
Lu
H
,
Chen
X
,
Tian
L
, et al
.
MLH1 deficiency-triggered DNA hyperexcision by exonuclease 1 activates the cGAS-STING pathway
.
Cancer Cell
2021
;
39
:
109
21
.
88.
Taniguchi
K
,
Kakinuma
S
,
Tokairin
Y
,
Arai
M
,
Kohno
H
,
Wakabayashi
K
, et al
.
Mild inflammation accelerates colon carcinogenesis in Mlh1-deficient mice
.
Oncology
2006
;
71
:
124
30
.
89.
Morioka
T
,
Miyoshi-Imamura
T
,
Blyth
BJ
,
Kaminishi
M
,
Kokubo
T
,
Nishimura
M
, et al
.
Ionizing radiation, inflammation, and their interactions in colon carcinogenesis in Mlh1-deficient mice
.
Cancer Sci
2015
;
106
:
217
26
.
90.
Edwards
RA
,
Witherspoon
M
,
Wang
K
,
Afrasiabi
K
,
Pham
T
,
Birnbaumer
L
, et al
.
Epigenetic repression of DNA mismatch repair by inflammation and hypoxia in inflammatory bowel disease-associated colorectal cancer
.
Cancer Res
2009
;
69
:
6423
9
.
91.
Scully
R
,
Panday
A
,
Elango
R
,
Willis
NA
.
DNA double-strand break repair-pathway choice in somatic mammalian cells
.
Nat Rev Mol Cell Biol
2019
;
20
:
698
714
.
92.
Gillyard
T
,
Davis
J
.
DNA double-strand break repair in cancer: A path to achieving precision medicine
.
Int Rev Cell Mol Biol
2021
;
364
:
111
37
.
93.
Sishc
BJ
,
Davis
AJ
.
The role of the core non-homologous end joining factors in carcinogenesis and cancer
.
Cancers (Basel)
2017
;
9
:
81
.
94.
Wang
Z
,
Yan
J
,
Lin
H
,
Hua
F
,
Wang
X
,
Liu
H
, et al
.
Toll-like receptor 4 activity protects against hepatocellular tumorigenesis and progression by regulating expression of DNA repair protein Ku70 in mice
.
Hepatology
2013
;
57
:
1869
81
.
95.
Su
Y
,
Wang
J
.
JmjC domain-containing protein 8 (JMJD8) represses Ku70/Ku80 expression via attenuating AKT/NF-κB/COX-2 signaling
.
Biochim Biophys Acta Mol Cell Res
2019
;
1866
:
118541
.
96.
Kumazawa
T
,
Mori
Y
,
Sato
H
,
Permata
TBM
,
Uchihara
Y
,
Noda
SE
, et al
.
Expression of non-homologous end joining factor, Ku80, is negatively correlated with PD-L1 expression in cancer cells after X-ray irradiation
.
Oncol Lett
2022
;
23
:
29
.
97.
Puebla-Osorio
N
,
Kim
J
,
Ojeda
S
,
Zhang
H
,
Tavana
O
,
Li
S
, et al
.
A novel Ku70 function in colorectal homeostasis separate from nonhomologous end joining
.
Oncogene
2014
;
33
:
2748
57
.
98.
Sui
H
,
Zhou
M
,
Imamichi
H
,
Jiao
X
,
Sherman
BT
,
Lane
HC
, et al
.
STING is an essential mediator of the Ku70-mediated production of IFN-lambda1 in response to exogenous DNA
.
Sci Signal
2017
;
10
:
eaah5054
.
99.
Zhang
X
,
Brann
TW
,
Zhou
M
,
Yang
J
,
Oguariri
RM
,
Lidie
KB
, et al
.
Cutting edge: Ku70 is a novel cytosolic DNA sensor that induces type III rather than type I IFN
.
J Immunol
2011
;
186
:
4541
5
.
100.
Ferguson
BJ
,
Mansur
DS
,
Peters
NE
,
Ren
H
,
Smith
GL
.
DNA-PK is a DNA sensor for IRF-3-dependent innate immunity
.
Elife
2012
;
1
:
e00047
.
101.
Peters
NE
,
Ferguson
BJ
,
Mazzon
M
,
Fahy
AS
,
Krysztofinska
E
,
Arribas-Bosacoma
R
, et al
.
A mechanism for the inhibition of DNA-PK-mediated DNA sensing by a virus
.
PLoS Pathog
2013
;
9
:
e1003649
.
102.
Scutts
SR
,
Ember
SW
,
Ren
H
,
Ye
C
,
Lovejoy
CA
,
Mazzon
M
, et al
.
DNA-PK is targeted by multiple vaccinia virus proteins to inhibit DNA sensing
.
Cell Rep
2018
;
25
:
1953
65
.
103.
Ma
C
,
Spies
NP
,
Gong
T
,
Jones
CX
,
Chu
WM
.
Involvement of DNA-PKcs in the type I IFN response to CpG-ODNs in conventional dendritic cells in TLR9-dependent or -independent manners
.
PLoS One
2015
;
10
:
e0121371
.
104.
Nakamura
K
,
Karmokar
A
,
Farrington
PM
,
James
NH
,
Ramos-Montoya
A
,
Bickerton
SJ
, et al
.
Inhibition of DNA-PK with AZD7648 sensitizes tumor cells to radiotherapy and induces type I IFN-dependent durable tumor control
.
Clin Cancer Res
2021
;
27
:
4353
66
.
105.
Mathieu
AL
,
Verronese
E
,
Rice
GI
,
Fouyssac
F
,
Bertrand
Y
,
Picard
C
, et al
.
PRKDC mutations associated with immunodeficiency, granuloma, and autoimmune regulator-dependent autoimmunity
.
J Allergy Clin Immunol
2015
;
135
:
1578
88
.
106.
Sun
X
,
Liu
T
,
Zhao
J
,
Xia
H
,
Xie
J
,
Guo
Y
, et al
.
DNA-PK deficiency potentiates cGAS-mediated antiviral innate immunity
.
Nat Commun
2020
;
11
:
6182
.
107.
Cheradame
L
,
Guerrera
IC
,
Gaston
J
,
Schmitt
A
,
Jung
V
,
Goudin
N
, et al
.
STING protects breast cancer cells from intrinsic and genotoxic-induced DNA instability via a non-canonical, cell-autonomous pathway
.
Oncogene
2021
;
40
:
6627
40
.
108.
Sriram
G
,
Milling
LE
,
Chen
JK
,
Kong
YW
,
Joughin
BA
,
Abraham
W
, et al
.
The injury response to DNA damage in live tumor cells promotes antitumor immunity
.
Sci Signal
2021
;
14
:
eabc4764
.
109.
Kroemer
G
,
Galassi
C
,
Zitvogel
L
,
Galluzzi
L
.
Immunogenic cell stress and death
.
Nat Immunol
2022
;
23
:
487
500
.
110.
Guo
G
,
Gao
M
,
Gao
X
,
Zhu
B
,
Huang
J
,
Tu
X
, et al
.
Reciprocal regulation of RIG-I and XRCC4 connects DNA repair with RIG-I immune signaling
.
Nat Commun
2021
;
12
:
2187
.
111.
Vanpouille-Box
C
,
Hoffmann
JA
,
Galluzzi
L
.
Pharmacological modulation of nucleic acid sensors: therapeutic potential and persisting obstacles
.
Nat Rev Drug Discov
2019
;
18
:
845
67
.
112.
Fackenthal
JD
,
Olopade
OI
.
Breast cancer risk associated with BRCA1 and BRCA2 in diverse populations
.
Nat Rev Cancer
2007
;
7
:
937
48
.
113.
Kondo
T
,
Kobayashi
J
,
Saitoh
T
,
Maruyama
K
,
Ishii
KJ
,
Barber
GN
, et al
.
DNA damage sensor MRE11 recognizes cytosolic double-stranded DNA and induces type I interferon by regulating STING trafficking
.
Proc Natl Acad Sci U S A
2013
;
110
:
2969
74
.
114.
Luzwick
JW
,
Dombi
E
,
Boisvert
RA
,
Roy
S
,
Park
S
,
Kunnimalaiyaan
S
, et al
.
MRE11-dependent instability in mitochondrial DNA fork protection activates a cGAS immune signaling pathway
.
Sci Adv
2021
;
7
:
eabf9441
.
115.
Yamazaki
T
,
Kirchmair
A
,
Sato
A
,
Buqué
A
,
Rybstein
M
,
Petroni
G
, et al
.
Mitochondrial DNA drives abscopal responses to radiation that are inhibited by autophagy
.
Nat Immunol
2020
;
21
:
1160
71
.
116.
Rodriguez-Ruiz
ME
,
Buqué
A
,
Hensler
M
,
Chen
J
,
Bloy
N
,
Petroni
G
, et al
.
Apoptotic caspases inhibit abscopal responses to radiation and identify a new prognostic biomarker for breast cancer patients
.
Oncoimmunology
2019
;
8
:
e1655964
.
117.
Ning
X
,
Wang
Y
,
Jing
M
,
Sha
M
,
Lv
M
,
Gao
P
, et al
.
Apoptotic caspases suppress type I interferon production via the cleavage of cGAS, MAVS, and IRF3
.
Mol Cell
2019
;
74
:
19
31
.
118.
Bhattacharya
S
,
Srinivasan
K
,
Abdisalaam
S
,
Su
F
,
Raj
P
,
Dozmorov
I
, et al
.
RAD51 interconnects between DNA replication, DNA repair and immunity
.
Nucleic Acids Res
2017
;
45
:
4590
605
.
119.
Roth
S
,
Rottach
A
,
Lotz-Havla
AS
,
Laux
V
,
Muschaweckh
A
,
Gersting
SW
, et al
.
Rad50-CARD9 interactions link cytosolic DNA sensing to IL-1β production
.
Nat Immunol
2014
;
15
:
538
45
.
120.
Almuzaini
N
,
Moore
M
,
Robert-Guroff
M
,
Thomas
MA
.
Disruption of NBS1/MRN complex formation by E4orf3 supports NF-κB that licenses E1B55K-deleted adenovirus-infected cells to accumulate DNA>4n
.
Microbiol Spectr
2022
;
10
:
e0188121
.
121.
Tu
X
,
Qin
B
,
Zhang
Y
,
Zhang
C
,
Kahila
M
,
Nowsheen
S
, et al
.
PD-L1 (B7-H1) competes with the RNA exosome to regulate the DNA damage response and can be targeted to sensitize to radiation or chemotherapy
.
Mol Cell
2019
;
74
:
1215
26
.
122.
Pereira-Lopes
S
,
Tur
J
,
Calatayud-Subias
JA
,
Lloberas
J
,
Stracker
TH
,
Celada
A
.
NBS1 is required for macrophage homeostasis and functional activity in mice
.
Blood
2015
;
126
:
2502
10
.
123.
Cejka
P
,
Symington
LS
.
DNA End Resection: Mechanism and Control
.
Annu Rev Genet
2021
;
55
:
285
307
.
124.
Marteijn
JA
,
Lans
H
,
Vermeulen
W
,
Hoeijmakers
JH
.
Understanding nucleotide excision repair and its roles in cancer and ageing
.
Nat Rev Mol Cell Biol
2014
;
15
:
465
81
.
125.
Erdal
E
,
Haider
S
,
Rehwinkel
J
,
Harris
AL
,
McHugh
PJ
.
A prosurvival DNA damage-induced cytoplasmic interferon response is mediated by end resection factors and is limited by Trex1
.
Genes Dev
2017
;
31
:
353
69
.
126.
Sato
H
,
Niimi
A
,
Yasuhara
T
,
Permata
TBM
,
Hagiwara
Y
,
Isono
M
, et al
.
DNA double-strand break repair pathway regulates PD-L1 expression in cancer cells
.
Nat Commun
2017
;
8
:
1751
.
127.
Li
J
,
Shu
X
,
Xu
J
,
Su
SM
,
Chan
UI
,
Mo
L
, et al
.
S100A9-CXCL12 activation in BRCA1-mutant breast cancer promotes an immunosuppressive microenvironment associated with resistance to immunotherapy
.
Nat Commun
2022
;
13
:
1481
.
128.
Bruand
M
,
Barras
D
,
Mina
M
,
Ghisoni
E
,
Morotti
M
,
Lanitis
E
, et al
.
Cell-autonomous inflammation of BRCA1-deficient ovarian cancers drives both tumor-intrinsic immunoreactivity and immune resistance via STING
.
Cell Rep
2021
;
36
:
109412
.
129.
Cardenas
H
,
Jiang
G
,
Thomes Pepin
J
,
Parker
JB
,
Condello
S
,
Nephew
KP
, et al
.
Interferon-γ signaling is associated with BRCA1 loss-of-function mutations in high grade serous ovarian cancer
.
NPJ Precis Oncol
2019
;
3
:
32
.
130.
Nolan
E
,
Savas
P
,
Policheni
AN
,
Darcy
PK
,
Vaillant
F
,
Mintoff
CP
, et al
.
Combined immune checkpoint blockade as a therapeutic strategy for BRCA1-mutated breast cancer
.
Sci Transl Med
2017
;
9
:
eaal4922
.
131.
Ma
H
,
Kang
Z
,
Foo
TK
,
Shen
Z
,
Xia
B
.
Disrupted BRCA1-PALB2 interaction induces tumor immunosuppression and T-lymphocyte infiltration in HCC through cGAS-STING pathway
.
Hepatology
2023
;
77
:
33
47
.
132.
Mei
J
,
Wang
R
,
Xia
D
,
Yang
X
,
Zhou
W
,
Wang
H
, et al
.
BRCA1 is a novel prognostic indicator and associates with immune cell infiltration in hepatocellular carcinoma
.
DNA Cell Biol
2020
;
39
:
1838
49
.
133.
Jørgensen
N
,
Hviid
TVF
,
Nielsen
LB
,
Sønderstrup
IMH
,
Eriksen
JO
,
Ejlertsen
B
, et al
.
Tumour-infiltrating CD4-, CD8- and FOXP3-positive immune cells as predictive markers of mortality in BRCA1- and BRCA2-associated breast cancer
.
Br J Cancer
2021
;
125
:
1388
98
.
134.
Launonen
IM
,
Lyytikäinen
N
,
Casado
J
,
Anttila
EA
,
Szabó
A
,
Haltia
UM
, et al
.
Single-cell tumor-immune microenvironment of BRCA1/2 mutated high-grade serous ovarian cancer
.
Nat Commun
2022
;
13
:
835
.
135.
Samstein
RM
,
Krishna
C
,
Ma
X
,
Pei
X
,
Lee
KW
,
Makarov
V
, et al
.
Mutations in BRCA1 and BRCA2 differentially affect the tumor microenvironment and response to checkpoint blockade immunotherapy
.
Nat Cancer
2021
;
1
:
1188
203
.
136.
Reisländer
T
,
Lombardi
EP
,
Groelly
FJ
,
Miar
A
,
Porru
M
,
Di Vito
S
, et al
.
BRCA2 abrogation triggers innate immune responses potentiated by treatment with PARP inhibitors
.
Nat Commun
2019
;
10
:
3143
.
137.
Heijink
AM
,
Talens
F
,
Jae
LT
,
van Gijn
SE
,
Fehrmann
RSN
,
Brummelkamp
TR
, et al
.
BRCA2 deficiency instigates cGAS-mediated inflammatory signaling and confers sensitivity to tumor necrosis factor-alpha-mediated cytotoxicity
.
Nat Commun
2019
;
10
:
100
.
138.
De Toni
EN
,
Ziesch
A
,
Rizzani
A
,
Torok
HP
,
Hocke
S
,
Lu
S
, et al
.
Inactivation of BRCA2 in human cancer cells identifies a subset of tumors with enhanced sensitivity towards death receptor-mediated apoptosis
.
Oncotarget
2016
;
7
:
9477
90
.
139.
Xu
H
,
Xiong
C
,
Chen
Y
,
Zhang
C
,
Bai
D
.
Identification of Rad51 as a prognostic biomarker correlated with immune infiltration in hepatocellular carcinoma
.
Bioengineered
2021
;
12
:
2664
75
.
140.
Wu
R
,
Patel
A
,
Tokumaru
Y
,
Asaoka
M
,
Oshi
M
,
Yan
L
, et al
.
High RAD51 gene expression is associated with aggressive biology and with poor survival in breast cancer
.
Breast Cancer Res Treat
2022
;
193
:
49
63
.
141.
Li
F
,
Zhang
Y
,
Shi
Y
,
Liu
S
.
Comprehensive analysis of prognostic and immune infiltrates for RAD51 in human breast cancer
.
Crit Rev Eukaryot Gene Expr
2021
;
31
:
71
9
.
142.
Khodarev
N
,
Ahmad
R
,
Rajabi
H
,
Pitroda
S
,
Kufe
T
,
McClary
C
, et al
.
Cooperativity of the MUC1 oncoprotein and STAT1 pathway in poor prognosis human breast cancer
.
Oncogene
2010
;
29
:
920
9
.
143.
Wolf
C
,
Rapp
A
,
Berndt
N
,
Staroske
W
,
Schuster
M
,
Dobrick-Mattheuer
M
, et al
.
RPA and Rad51 constitute a cell intrinsic mechanism to protect the cytosol from self DNA
.
Nat Commun
2016
;
7
:
11752
.
144.
Jiang
H
,
Xue
X
,
Panda
S
,
Kawale
A
,
Hooy
RM
,
Liang
F
, et al
.
Chromatin-bound cGAS is an inhibitor of DNA repair and hence accelerates genome destabilization and cell death
.
Embo j
2019
;
38
:
e102718
.
145.
Liu
H
,
Zhang
H
,
Wu
X
,
Ma
D
,
Wu
J
,
Wang
L
, et al
.
Nuclear cGAS suppresses DNA repair and promotes tumorigenesis
.
Nature
2018
;
563
:
131
6
.
146.
da Costa
A
,
Chowdhury
D
,
Shapiro
GI
,
D'Andrea
AD
,
Konstantinopoulos
PA
.
Targeting replication stress in cancer therapy
.
Nat Rev Drug Discov
2023
;
22
:
38
58
.
147.
Ding
L
,
Kim
HJ
,
Wang
Q
,
Kearns
M
,
Jiang
T
,
Ohlson
CE
, et al
.
PARP inhibition elicits STING-dependent antitumor immunity in brca1-deficient ovarian cancer
.
Cell Rep
2018
;
25
:
2972
80
.
148.
Shen
J
,
Zhao
W
,
Ju
Z
,
Wang
L
,
Peng
Y
,
Labrie
M
, et al
.
PARPi triggers the STING-dependent immune response and enhances the therapeutic efficacy of immune checkpoint blockade independent of BRCAness
.
Cancer Res
2019
;
79
:
311
9
.
149.
Chabanon
RM
,
Muirhead
G
,
Krastev
DB
,
Adam
J
,
Morel
D
,
Garrido
M
, et al
.
PARP inhibition enhances tumor cell-intrinsic immunity inERCC1-deficient non-small cell lung cancer
.
J Clin Invest
2019
;
129
:
1211
28
.
150.
Pantelidou
C
,
Sonzogni
O
,
De Oliveria Taveira
M
,
Mehta
AK
,
Kothari
A
,
Wang
D
, et al
.
PARP inhibitor efficacy depends on CD8(+) T-cell recruitment via intratumoral STING pathway activation in BRCA-deficient models of triple-negative breast cancer
.
Cancer Discov
2019
;
9
:
722
37
.
151.
Chabanon
RM
,
Morel
D
,
Eychenne
T
,
Colmet-Daage
L
,
Bajrami
I
,
Dorvault
N
, et al
.
PBRM1 deficiency confers synthetic lethality to DNA repair inhibitors in cancer
.
Cancer Res
2021
;
81
:
2888
902
.
152.
Sen
T
,
Rodriguez
BL
,
Chen
L
,
Corte
CMD
,
Morikawa
N
,
Fujimoto
J
, et al
.
Targeting DNA damage response promotes antitumor immunity through STING-mediated T-cell activation in small cell lung cancer
.
Cancer Discov
2019
;
9
:
646
61
.
153.
Kim
C
,
Wang
XD
,
Yu
Y
.
PARP1 inhibitors trigger innate immunity via PARP1 trapping-induced DNA damage response
.
Elife
2020
;
9
:
e60637
.
154.
Wang
F
,
Zhao
M
,
Chang
B
,
Zhou
Y
,
Wu
X
,
Ma
M
, et al
.
Cytoplasmic PARP1 links the genome instability to the inhibition of antiviral immunity through PARylating cGAS
.
Mol Cell
2022
;
82
:
2032
49
.
155.
Gozgit
JM
,
Vasbinder
MM
,
Abo
RP
,
Kunii
K
,
Kuplast-Barr
KG
,
Gui
B
, et al
.
PARP7 negatively regulates the type I interferon response in cancer cells and its inhibition triggers antitumor immunity
.
Cancer Cell
2021
;
39
:
1214
26
.
156.
Juncheng
P
,
Joseph
A
,
Lafarge
A
,
Martins
I
,
Obrist
F
,
Pol
J
, et al
.
Cancer cell-autonomous overactivation of PARP1 compromises immunosurveillance in non-small cell lung cancer
.
J Immunother Cancer
2022
;
10
:
e004280
.
157.
Staniszewska
AD
,
Armenia
J
,
King
M
,
Michaloglou
C
,
Reddy
A
,
Singh
M
, et al
.
PARP inhibition is a modulator of anti-tumor immune response in BRCA-deficient tumors
.
Oncoimmunology
2022
;
11
:
2083755
.
158.
Wang
Z
,
Sun
K
,
Xiao
Y
,
Feng
B
,
Mikule
K
,
Ma
X
, et al
.
Niraparib activates interferon signaling and potentiates anti–PD-1 antibody efficacy in tumor models
.
Sci Rep
2019
;
9
:
1853
.
159.
Jiao
S
,
Xia
W
,
Yamaguchi
H
,
Wei
Y
,
Chen
MK
,
Hsu
JM
, et al
.
PARP inhibitor upregulates PD-L1 expression and enhances cancer-associated immunosuppression
.
Clin Cancer Res
2017
;
23
:
3711
20
.
160.
Higuchi
T
,
Flies
DB
,
Marjon
NA
,
Mantia-Smaldone
G
,
Ronner
L
,
Gimotty
PA
, et al
.
CTLA-4 blockade synergizes therapeutically with PARP inhibition in BRCA1-deficient ovarian cancer
.
Cancer Immunol Res
2015
;
3
:
1257
68
.
161.
Wang
Q
,
Bergholz
JS
,
Ding
L
,
Lin
Z
,
Kabraji
SK
,
Hughes
ME
, et al
.
STING agonism reprograms tumor-associated macrophages and overcomes resistance to PARP inhibition in BRCA1-deficient models of breast cancer
.
Nat Commun
2022
;
13
:
3022
.
162.
Mehta
AK
,
Cheney
EM
,
Hartl
CA
,
Pantelidou
C
,
Oliwa
M
,
Castrillon
JA
, et al
.
Targeting immunosuppressive macrophages overcomes PARP inhibitor resistance in BRCA1-associated triple-negative breast cancer
.
Nat Cancer
2021
;
2
:
66
82
.
163.
Fenerty
KE
,
Padget
M
,
Wolfson
B
,
Gameiro
SR
,
Su
Z
,
Lee
JH
, et al
.
Immunotherapy utilizing the combination of natural killer- and antibody dependent cellular cytotoxicity (ADCC)-mediating agents with poly (ADP-ribose) polymerase (PARP) inhibition
.
J Immunother Cancer
2018
;
6
:
133
.
164.
Ding
L
,
Chen
X
,
Xu
X
,
Qian
Y
,
Liang
G
,
Yao
F
, et al
.
PARP1 suppresses the transcription of PD-L1 by Poly(ADP-Ribosyl)ating STAT3
.
Cancer Immunol Res
2019
;
7
:
136
49
.
165.
Paczulla
AM
,
Rothfelder
K
,
Raffel
S
,
Konantz
M
,
Steinbacher
J
,
Wang
H
, et al
.
Absence of NKG2D ligands defines leukaemia stem cells and mediates their immune evasion
.
Nature
2019
;
572
:
254
9
.
166.
Rakova
J
,
Truxova
I
,
Holicek
P
,
Salek
C
,
Hensler
M
,
Kasikova
L
, et al
.
TIM-3 levels correlate with enhanced NK cell cytotoxicity and improved clinical outcome in AML patients
.
Oncoimmunology
2021
;
10
:
1889822
.
167.
Dörsam
B
,
Seiwert
N
,
Foersch
S
,
Stroh
S
,
Nagel
G
,
Begaliew
D
, et al
.
PARP-1 protects against colorectal tumor induction, but promotes inflammation-driven colorectal tumor progression
.
Proc Natl Acad Sci U S A
2018
;
115
:
E4061
e70
.
168.
Ghonim
MA
,
Ibba
SV
,
Tarhuni
AF
,
Errami
Y
,
Luu
HH
,
Dean
MJ
, et al
.
Targeting PARP-1 with metronomic therapy modulates MDSC suppressive function and enhances anti–PD-1 immunotherapy in colon cancer
.
J Immunother Cancer
2021
;
9
:
e001643
.
169.
Lecona
E
,
Fernandez-Capetillo
O
.
Targeting ATR in cancer
.
Nat Rev Cancer
2018
;
18
:
586
95
.
170.
McBride
KA
,
Ballinger
ML
,
Killick
E
,
Kirk
J
,
Tattersall
MH
,
Eeles
RA
, et al
.
Li-Fraumeni syndrome: cancer risk assessment and clinical management
.
Nat Rev Clin Oncol
2014
;
11
:
260
71
.
171.
Groelly
FJ
,
Fawkes
M
,
Dagg
RA
,
Blackford
AN
,
Tarsounas
M
.
Targeting DNA damage response pathways in cancer
.
Nat Rev Cancer
2023
;
23
:
78
94
.
172.
Westbrook
AM
,
Schiestl
RH
.
Atm-deficient mice exhibit increased sensitivity to dextran sulfate sodium-induced colitis characterized by elevated DNA damage and persistent immune activation
.
Cancer Res
2010
;
70
:
1875
84
.
173.
Wang
C
,
Jette
N
,
Moussienko
D
,
Bebb
DG
,
Lees-Miller
SP
.
ATM-deficient colorectal cancer cells are sensitive to the PARP inhibitor olaparib
.
Transl Oncol
2017
;
10
:
190
6
.
174.
Williamson
CT
,
Muzik
H
,
Turhan
AG
,
Zamò
A
,
O'Connor
MJ
,
Bebb
DG
, et al
.
ATM deficiency sensitizes mantle cell lymphoma cells to poly(ADP-ribose) polymerase-1 inhibitors
.
Mol Cancer Ther
2010
;
9
:
347
57
.
175.
Zhang
Q
,
Green
MD
,
Lang
X
,
Lazarus
J
,
Parsels
JD
,
Wei
S
, et al
.
Inhibition of ATM increases interferon signaling and sensitizes pancreatic cancer to immune checkpoint blockade therapy
.
Cancer Res
2019
;
79
:
3940
51
.
176.
Hu
M
,
Zhou
M
,
Bao
X
,
Pan
D
,
Jiao
M
,
Liu
X
, et al
.
ATM inhibition enhances cancer immunotherapy by promoting mtDNA leakage and cGAS/STING activation
.
J Clin Invest
2021
;
131
:
e139333
.
177.
Wang
L
,
Yang
L
,
Wang
C
,
Zhao
W
,
Ju
Z
,
Zhang
W
, et al
.
Inhibition of the ATM/Chk2 axis promotes cGAS/STING signaling in ARID1A-deficient tumors
.
J Clin Invest
2020
;
130
:
5951
66
.
178.
Jinushi
M
,
Chiba
S
,
Baghdadi
M
,
Kinoshita
I
,
Dosaka-Akita
H
,
Ito
K
, et al
.
ATM-mediated DNA damage signals mediate immune escape through integrin-αvβ3-dependent mechanisms
.
Cancer Res
2012
;
72
:
56
65
.
179.
Wang
Z
,
Zhang
X
,
Li
W
,
Su
Q
,
Huang
Z
,
Zhang
X
, et al
.
ATM/NEMO signaling modulates the expression of PD-L1 following docetaxel chemotherapy in prostate cancer
.
J Immunother Cancer
2021
;
9
:
e001758
.
180.
Yi
R
,
Lin
A
,
Cao
M
,
Xu
A
,
Luo
P
,
Zhang
J
.
ATM mutations benefit bladder cancer patients treated with immune checkpoint inhibitors by acting on the tumor immune microenvironment
.
Front Genet
2020
;
11
:
933
.
181.
Sun
L
,
Wang
RC
,
Zhang
Q
,
Guo
LL
.
ATM mutations as an independent prognostic factor and potential biomarker for immune checkpoint therapy in endometrial cancer
.
Pathol Res Pract
2020
;
216
:
153032
.
182.
Wu
Q
,
Allouch
A
,
Paoletti
A
,
Leteur
C
,
Mirjolet
C
,
Martins
I
, et al
.
NOX2-dependent ATM kinase activation dictates pro-inflammatory macrophage phenotype and improves effectiveness to radiation therapy
.
Cell Death Differ
2017
;
24
:
1632
44
.
183.
Oda
T
,
Nakamura
R
,
Kasamatsu
T
,
Gotoh
N
,
Okuda
K
,
Saitoh
T
, et al
.
DNA-double strand breaks enhance the expression of major histocompatibility complex class II through the ATM-NF-κB-IRF1-CIITA pathway
.
Cancer Gene Ther
2022
;
29
:
225
40
.
184.
Banerjee
D
,
Langberg
K
,
Abbas
S
,
Odermatt
E
,
Yerramothu
P
,
Volaric
M
, et al
.
A non-canonical, interferon-independent signaling activity of cGAMP triggers DNA damage response signaling
.
Nat Commun
2021
;
12
:
6207
.
185.
Zhang
B
,
Miao
T
,
Shen
X
,
Bao
L
,
Zhang
C
,
Yan
C
, et al
.
EB virus-induced ATR activation accelerates nasopharyngeal carcinoma growth via M2-type macrophages polarization
.
Cell Death Dis
2020
;
11
:
742
.
186.
Ge
C
,
Vilfranc
CL
,
Che
L
,
Pandita
RK
,
Hambarde
S
,
Andreassen
PR
, et al
.
The BRUCE-ATR signaling axis is required for accurate DNA replication and suppression of liver cancer development
.
Hepatology
2019
;
69
:
2608
22
.
187.
Hsieh
RC
,
Krishnan
S
,
Wu
RC
,
Boda
AR
,
Liu
A
,
Winkler
M
, et al
.
ATR-mediated CD47 and PD-L1 up-regulation restricts radiotherapy-induced immune priming and abscopal responses in colorectal cancer
.
Sci Immunol
2022
;
7
:
eabl9330
.
188.
Sun
LL
,
Yang
RY
,
Li
CW
,
Chen
MK
,
Shao
B
,
Hsu
JM
, et al
.
Inhibition of ATR downregulates PD-L1 and sensitizes tumor cells to T cell-mediated killing
.
Am J Cancer Res
2018
;
8
:
1307
16
.
189.
Chen
CF
,
Ruiz-Vega
R
,
Vasudeva
P
,
Espitia
F
,
Krasieva
TB
,
de Feraudy
S
, et al
.
ATR mutations promote the growth of melanoma tumors by modulating the immune microenvironment
.
Cell Rep
2017
;
18
:
2331
42
.
190.
Tang
Z
,
Pilié
PG
,
Geng
C
,
Manyam
GC
,
Yang
G
,
Park
S
, et al
.
ATR inhibition induces CDK1-SPOP signaling and enhances Anti-PD-L1 cytotoxicity in prostate cancer
.
Clin Cancer Res
2021
;
27
:
4898
909
.
191.
Ding
H
,
Vincelette
ND
,
McGehee
CD
,
Kohorst
MA
,
Koh
BD
,
Venkatachalam
A
, et al
.
CDK2-mediated upregulation of TNFα as a mechanism of selective cytotoxicity in acute leukemia
.
Cancer Res
2021
;
81
:
2666
78
.
192.
Chaudhary
R
,
Slebos
RJC
,
Song
F
,
McCleary-Sharpe
KP
,
Masannat
J
,
Tan
AC
, et al
.
Effects of checkpoint kinase 1 inhibition by prexasertib on the tumor immune microenvironment of head and neck squamous cell carcinoma
.
Mol Carcinog
2021
;
60
:
138
50
.
193.
Schoonen
PM
,
Kok
YP
,
Wierenga
E
,
Bakker
B
,
Foijer
F
,
Spierings
DCJ
, et al
.
Premature mitotic entry induced by ATR inhibition potentiates olaparib inhibition-mediated genomic instability, inflammatory signaling, and cytotoxicity in BRCA2-deficient cancer cells
.
Mol Oncol
2019
;
13
:
2422
40
.
194.
Sheng
H
,
Huang
Y
,
Xiao
Y
,
Zhu
Z
,
Shen
M
,
Zhou
P
, et al
.
ATR inhibitor AZD6738 enhances the antitumor activity of radiotherapy and immune checkpoint inhibitors by potentiating the tumor immune microenvironment in hepatocellular carcinoma
.
J Immunother Cancer
2020
;
8
:
e000340
.
195.
Dillon
MT
,
Bergerhoff
KF
,
Pedersen
M
,
Whittock
H
,
Crespo-Rodriguez
E
,
Patin
EC
, et al
.
ATR inhibition potentiates the radiation-induced inflammatory tumor microenvironment
.
Clin Cancer Res
2019
;
25
:
3392
403
.
196.
Vendetti
FP
,
Karukonda
P
,
Clump
DA
,
Teo
T
,
Lalonde
R
,
Nugent
K
, et al
.
ATR kinase inhibitor AZD6738 potentiates CD8+ T cell-dependent antitumor activity following radiation
.
J Clin Invest
2018
;
128
:
3926
40
.
197.
Feng
X
,
Tubbs
A
,
Zhang
C
,
Tang
M
,
Sridharan
S
,
Wang
C
, et al
.
ATR inhibition potentiates ionizing radiation-induced interferon response via cytosolic nucleic acid-sensing pathways
.
Embo J
2020
;
39
:
e104036
.
198.
Wu
X
,
Kang
X
,
Zhang
X
,
Xie
W
,
Su
Y
,
Liu
X
, et al
.
WEE1 inhibitor and ataxia telangiectasia and RAD3-related inhibitor trigger stimulator of interferon gene-dependent immune response and enhance tumor treatment efficacy through programmed death-ligand 1 blockade
.
Cancer Sci
2021
;
112
:
4444
56
.
199.
Combès
E
,
Andrade
AF
,
Tosi
D
,
Michaud
HA
,
Coquel
F
,
Garambois
V
, et al
.
Inhibition of ataxia-telangiectasia mutated and RAD3-related (ATR) overcomes oxaliplatin resistance and promotes antitumor immunity in colorectal cancer
.
Cancer Res
2019
;
79
:
2933
46
.
200.
Roeschert
I
,
Poon
E
,
Henssen
AG
,
Garcia
HD
,
Gatti
M
,
Giansanti
C
, et al
.
Combined inhibition of Aurora-A and ATR kinase results in regression of MYCN-amplified neuroblastoma
.
Nat Cancer
2021
;
2
:
312
26
.
201.
Sen
T
,
Della Corte
CM
,
Milutinovic
S
,
Cardnell
RJ
,
Diao
L
,
Ramkumar
K
, et al
.
Combination treatment of the oral CHK1 inhibitor, SRA737, and low-dose gemcitabine enhances the effect of programmed death ligand 1 blockade by modulating the immune microenvironment in SCLC
.
J Thorac Oncol
2019
;
14
:
2152
63
.
202.
Li
M
,
Huang
T
,
Li
X
,
Shi
Z
,
Sheng
Y
,
Hu
M
, et al
.
GDC-0575, a CHK1 inhibitor, impairs the development of colitis and colitis-associated cancer by inhibiting CCR2(+) macrophage infiltration in mice
.
Onco Targets Ther
2021
;
14
:
2661
72
.
203.
Yan
Y
,
Zheng
L
,
Du
Q
,
Cui
X
,
Dong
K
,
Guo
Y
, et al
.
Interferon regulatory factor 1 (IRF-1) downregulates checkpoint kinase 1 (CHK1) through miR-195 to upregulate apoptosis and PD-L1 expression in Hepatocellular carcinoma (HCC) cells
.
Br J Cancer
2021
;
125
:
101
11
.
204.
Kim
KS
,
Choi
KJ
,
Bae
S
.
Interferon-gamma enhances radiation-induced cell death via downregulation of Chk1
.
Cancer Biol Ther
2012
;
13
:
1018
25
.
205.
Makino
Y
,
Hikita
H
,
Fukumoto
K
,
Sung
JH
,
Sakano
Y
,
Murai
K
, et al
.
Constitutive activation of the tumor suppressor p53 in hepatocytes paradoxically promotes non-cell autonomous liver carcinogenesis
.
Cancer Res
2022
;
82
:
2860
73
.
206.
Ghosh
M
,
Saha
S
,
Bettke
J
,
Nagar
R
,
Parrales
A
,
Iwakuma
T
, et al
.
Mutant p53 suppresses innate immune signaling to promote tumorigenesis
.
Cancer Cell
2021
;
39
:
494
508
.
207.
Siolas
D
,
Vucic
E
,
Kurz
E
,
Hajdu
C
,
Bar-Sagi
D
.
Gain-of-function p53(R172H) mutation drives accumulation of neutrophils in pancreatic tumors, promoting resistance to immunotherapy
.
Cell Rep
2021
;
36
:
109578
.
208.
Blagih
J
,
Zani
F
,
Chakravarty
P
,
Hennequart
M
,
Pilley
S
,
Hobor
S
, et al
.
Cancer-specific loss of p53 leads to a modulation of myeloid and T cell responses
.
Cell Rep
2020
;
30
:
481
96
.
209.
Wellenstein
MD
,
Coffelt
SB
,
Duits
DEM
,
van Miltenburg
MH
,
Slagter
M
,
de Rink
I
, et al
.
Loss of p53 triggers WNT-dependent systemic inflammation to drive breast cancer metastasis
.
Nature
2019
;
572
:
538
42
.
210.
Guo
G
,
Yu
M
,
Xiao
W
,
Celis
E
,
Cui
Y
.
Local activation of p53 in the tumor microenvironment overcomes immune suppression and enhances antitumor immunity
.
Cancer Res
2017
;
77
:
2292
305
.
211.
Vilgelm
AE
,
Pawlikowski
JS
,
Liu
Y
,
Hawkins
OE
,
Davis
TA
,
Smith
J
, et al
.
Mdm2 and aurora kinase a inhibitors synergize to block melanoma growth by driving apoptosis and immune clearance of tumor cells
.
Cancer Res
2015
;
75
:
181
93
.
212.
Zhou
X
,
Singh
M
,
Sanz Santos
G
,
Guerlavais
V
,
Carvajal
LA
,
Aivado
M
, et al
.
Pharmacologic activation of p53 triggers viral mimicry response thereby abolishing tumor immune evasion and promoting antitumor immunity
.
Cancer Discov
2021
;
11
:
3090
105
.
213.
Scarpa
M
,
Marchiori
C
,
Scarpa
M
,
Castagliuolo
I
.
CD80 expression is upregulated by TP53 activation in human cancer epithelial cells
.
Oncoimmunology
2021
;
10
:
1907912
.
214.
Xiao
Y
,
Chen
J
,
Zhou
H
,
Zeng
X
,
Ruan
Z
,
Pu
Z
, et al
.
Combining p53 mRNA nanotherapy with immune checkpoint blockade reprograms the immune microenvironment for effective cancer therapy
.
Nat Commun
2022
;
13
:
758
.
215.
Moore
EC
,
Sun
L
,
Clavijo
PE
,
Friedman
J
,
Harford
JB
,
Saleh
AD
, et al
.
Nanocomplex-based TP53 gene therapy promotes anti-tumor immunity through TP53- and STING-dependent mechanisms
.
Oncoimmunology
2018
;
7
:
e1404216
.
216.
Kono
M
,
Kumai
T
,
Hayashi
R
,
Yamaki
H
,
Komatsuda
H
,
Wakisaka
R
, et al
.
Interruption of MDM2 signaling augments MDM2-targeted T cell-based antitumor immunotherapy through antigen-presenting machinery
.
Cancer Immunol Immunother
2021
;
70
:
3421
34
.
217.
Wang
HQ
,
Mulford
IJ
,
Sharp
F
,
Liang
J
,
Kurtulus
S
,
Trabucco
G
, et al
.
Inhibition of MDM2 promotes antitumor responses in p53 wild-type cancer cells through their interaction with the immune and stromal microenvironment
.
Cancer Res
2021
;
81
:
3079
91
.
218.
Fang
DD
,
Tang
Q
,
Kong
Y
,
Wang
Q
,
Gu
J
,
Fang
X
, et al
.
MDM2 inhibitor APG-115 synergizes with PD-1 blockade through enhancing antitumor immunity in the tumor microenvironment
.
J Immunother Cancer
2019
;
7
:
327
.
219.
Sahin
I
,
Zhang
S
,
Navaraj
A
,
Zhou
L
,
Dizon
D
,
Safran
H
, et al
.
AMG-232 sensitizes high MDM2-expressing tumor cells to T-cell-mediated killing
.
Cell Death Discov
2020
;
6
:
57
.
220.
Hayashi
Y
,
Goyama
S
,
Liu
X
,
Tamura
M
,
Asada
S
,
Tanaka
Y
, et al
.
Antitumor immunity augments the therapeutic effects of p53 activation on acute myeloid leukemia
.
Nat Commun
2019
;
10
:
4869
.
221.
Zhou
J
,
Kryczek
I
,
Li
S
,
Li
X
,
Aguilar
A
,
Wei
S
, et al
.
The ubiquitin ligase MDM2 sustains STAT5 stability to control T cell-mediated antitumor immunity
.
Nat Immunol
2021
;
22
:
460
70
.
222.
Zhang
W
,
Gong
J
,
Yang
H
,
Wan
L
,
Peng
Y
,
Wang
X
, et al
.
The mitochondrial protein MAVS stabilizes p53 to suppress tumorigenesis
.
Cell Rep
2020
;
30
:
725
38
.
223.
Vizioli
MG
,
Liu
T
,
Miller
KN
,
Robertson
NA
,
Gilroy
K
,
Lagnado
AB
, et al
.
Mitochondria-to-nucleus retrograde signaling drives formation of cytoplasmic chromatin and inflammation in senescence
.
Genes Dev
2020
;
34
:
428
45
.
224.
Faget
DV
,
Ren
Q
,
Stewart
SA
.
Unmasking senescence: context-dependent effects of SASP in cancer
.
Nat Rev Cancer
2019
;
19
:
439
53
.
225.
Manke
IA
,
Nguyen
A
,
Lim
D
,
Stewart
MQ
,
Elia
AE
,
Yaffe
MB
.
MAPKAP kinase-2 is a cell cycle checkpoint kinase that regulates the G2–M transition and S phase progression in response to UV irradiation
.
Mol Cell
2005
;
17
:
37
48
.
226.
Reinhardt
HC
,
Hasskamp
P
,
Schmedding
I
,
Morandell
S
,
van Vugt
MA
,
Wang
X
, et al
.
DNA damage activates a spatially distinct late cytoplasmic cell-cycle checkpoint network controlled by MK2-mediated RNA stabilization
.
Mol Cell
2010
;
40
:
34
49
.
227.
Chen
W
,
Liang
R
,
Yi
Y
,
Zhu
J
,
Zhang
J
.
P38alpha deficiency in macrophages ameliorates murine experimental colitis by regulating inflammation and immune process
.
Pathol Res Pract
2022
;
233
:
153881
.
228.
Ray
AL
,
Castillo
EF
,
Morris
KT
,
Nofchissey
RA
,
Weston
LL
,
Samedi
VG
, et al
.
Blockade of MK2 is protective in inflammation-associated colorectal cancer development
.
Int J Cancer
2016
;
138
:
770
5
.
229.
Suarez-Lopez
L
,
Sriram
G
,
Kong
YW
,
Morandell
S
,
Merrick
KA
,
Hernandez
Y
, et al
.
MK2 contributes to tumor progression by promoting M2 macrophage polarization and tumor angiogenesis
.
Proc Natl Acad Sci U S A
2018
;
115
:
E4236
E44
.
230.
Suarez-Lopez
L
,
Kong
YW
,
Sriram
G
,
Patterson
JC
,
Rosenberg
S
,
Morandell
S
, et al
.
MAPKAP kinase-2 drives expression of angiogenic factors by tumor-associated macrophages in a model of inflammation-induced colon cancer
.
Front Immunol
2020
;
11
:
607891
.
231.
Kopper
F
,
Bierwirth
C
,
Schon
M
,
Kunze
M
,
Elvers
I
,
Kranz
D
, et al
.
Damage-induced DNA replication stalling relies on MAPK-activated protein kinase 2 activity
.
Proc Natl Acad Sci U S A
2013
;
110
:
16856
61
.
232.
Nishimura
T
,
Andoh
A
,
Nishida
A
,
Shioya
M
,
Koizumi
Y
,
Tsujikawa
T
, et al
.
FR167653, a p38 mitogen-activated protein kinase inhibitor, aggravates experimental colitis in mice
.
World J Gastroenterol
2008
;
14
:
5851
6
.
233.
Gupta
J
,
del Barco Barrantes I, Igea
A
,
Sakellariou
S
,
Pateras
IS
,
Gorgoulis
VG
, et al
.
Dual function of p38alpha MAPK in colon cancer: suppression of colitis-associated tumor initiation but requirement for cancer cell survival
.
Cancer Cell
2014
;
25
:
484
500
.
234.
Taniguchi
H
,
Caeser
R
,
Chavan
SS
,
Zhan
YA
,
Chow
A
,
Manoj
P
, et al
.
WEE1 inhibition enhances the antitumor immune response to PD-L1 blockade by the concomitant activation of STING and STAT1 pathways in SCLC
.
Cell Rep
2022
;
39
:
110814
.
235.
Li
C
,
Shen
Q
,
Zhang
P
,
Wang
T
,
Liu
W
,
Li
R
, et al
.
Targeting MUS81 promotes the anticancer effect of WEE1 inhibitor and immune checkpoint blocking combination therapy via activating cGAS/STING signaling in gastric cancer cells
.
J Exp Clin Cancer Res
2021
;
40
:
315
.
236.
Wang
B
,
Sun
L
,
Yuan
Z
,
Tao
Z
.
Wee1 kinase inhibitor AZD1775 potentiates CD8+ T cell-dependent antitumour activity via dendritic cell activation following a single high dose of irradiation
.
Med Oncol
2020
;
37
:
66
.
237.
Patel
P
,
Sun
L
,
Robbins
Y
,
Clavijo
PE
,
Friedman
J
,
Silvin
C
, et al
.
Enhancing direct cytotoxicity and response to immune checkpoint blockade following ionizing radiation with Wee1 kinase inhibition
.
Oncoimmunology
2019
;
8
:
e1638207
.
238.
Sun
L
,
Moore
E
,
Berman
R
,
Clavijo
PE
,
Saleh
A
,
Chen
Z
, et al
.
WEE1 kinase inhibition reverses G2–M cell cycle checkpoint activation to sensitize cancer cells to immunotherapy
.
Oncoimmunology
2018
;
7
:
e1488359
.
239.
Hamilton
DH
,
Huang
B
,
Fernando
RI
,
Tsang
KY
,
Palena
C
.
WEE1 inhibition alleviates resistance to immune attack of tumor cells undergoing epithelial-mesenchymal transition
.
Cancer Res
2014
;
74
:
2510
9
.
240.
Dinavahi
SS
,
Chen
YC
,
Punnath
K
,
Berg
A
,
Herlyn
M
,
Foroutan
M
, et al
.
Targeting WEE1/AKT restores p53-dependent natural killer-cell activation to induce immune checkpoint blockade responses in “Cold” melanoma
.
Cancer Immunol Res
2022
;
10
:
757
69
.
241.
Guo
E
,
Xiao
R
,
Wu
Y
,
Lu
F
,
Liu
C
,
Yang
B
, et al
.
WEE1 inhibition induces anti-tumor immunity by activating ERV and the dsRNA pathway
.
J Exp Med
2022
;
219
:
e20210789
.
242.
Jin
MH
,
Nam
AR
,
Park
JE
,
Bang
JH
,
Bang
YJ
,
Oh
DY
.
Therapeutic Co-targeting of WEE1 and ATM downregulates PD-L1 expression in pancreatic cancer
.
Cancer Res Treat
2020
;
52
:
149
66
.
243.
Lin
YL
,
Pasero
P
.
Replication stress: from chromatin to immunity and beyond
.
Curr Opin Genet Dev
2021
;
71
:
136
42
.
244.
Cybulla
E
,
Vindigni
A
.
Leveraging the replication stress response to optimize cancer therapy
.
Nat Rev Cancer
2023
;
23
:
6
24
.
245.
Cong
K
,
Peng
M
,
Kousholt
AN
,
Lee
WTC
,
Lee
S
,
Nayak
S
, et al
.
Replication gaps are a key determinant of PARP inhibitor synthetic lethality with BRCA deficiency
.
Mol Cell
2021
;
81
:
3227
.
246.
Vaitsiankova
A
,
Burdova
K
,
Sobol
M
,
Gautam
A
,
Benada
O
,
Hanzlikova
H
, et al
.
PARP inhibition impedes the maturation of nascent DNA strands during DNA replication
.
Nat Struct Mol Biol
2022
;
29
:
329
38
.
247.
Saldivar
JC
,
Cortez
D
,
Cimprich
KA
.
The essential kinase ATR: ensuring faithful duplication of a challenging genome
.
Nat Rev Mol Cell Biol
2017
;
18
:
622
36
.
248.
Emam
A
,
Wu
X
,
Xu
S
,
Wang
L
,
Liu
S
,
Wang
B
.
Stalled replication fork protection limits cGAS-STING and P-body-dependent innate immune signalling
.
Nat Cell Biol
2022
;
24
:
1154
64
.
249.
Coquel
F
,
Silva
MJ
,
Techer
H
,
Zadorozhny
K
,
Sharma
S
,
Nieminuszczy
J
, et al
.
SAMHD1 acts at stalled replication forks to prevent interferon induction
.
Nature
2018
;
557
:
57
61
.
250.
Indiani
C
,
O'Donnell
M
.
The replication clamp-loading machine at work in the three domains of life
.
Nat Rev Mol Cell Biol
2006
;
7
:
751
61
.
251.
Wang
YL
,
Lee
CC
,
Shen
YC
,
Lin
PL
,
Wu
WR
,
Lin
YZ
, et al
.
Evading immune surveillance via tyrosine phosphorylation of nuclear PCNA
.
Cell Rep
2021
;
36
:
109537
.
252.
Kundu
K
,
Ghosh
S
,
Sarkar
R
,
Edri
A
,
Brusilovsky
M
,
Gershoni-Yahalom
O
, et al
.
Inhibition of the NKp44-PCNA immune checkpoint using a mAb to PCNA
.
Cancer Immunol Res
2019
;
7
:
1120
34
.
253.
Palmerola
KL
,
Amrane
S
,
De Los Angeles
A
,
Xu
S
,
Wang
N
,
de Pinho
J
, et al
.
Replication stress impairs chromosome segregation and preimplantation development in human embryos
.
Cell
2022
;
185
:
2988
3007
.
254.
Mackenzie
KJ
,
Carroll
P
,
Martin
CA
,
Murina
O
,
Fluteau
A
,
Simpson
DJ
, et al
.
cGAS surveillance of micronuclei links genome instability to innate immunity
.
Nature
2017
;
548
:
461
5
.
255.
Proctor
M
,
Gonzalez Cruz
JL
,
Daignault-Mill
SM
,
Veitch
M
,
Zeng
B
,
Ehmann
A
, et al
.
Targeting replication stress using CHK1 inhibitor promotes innate and NKT cell immune responses and tumour regression
.
Cancers (Basel)
2021
;
13
:
3733
.
256.
Zhang
W
,
Liu
W
,
Jia
L
,
Chen
D
,
Chang
I
,
Lake
M
, et al
.
Targeting KDM4A epigenetically activates tumor-cell-intrinsic immunity by inducing DNA replication stress
.
Mol Cell
2021
;
81
:
2148
65
.
257.
Zhang
H
,
Christensen
CL
,
Dries
R
,
Oser
MG
,
Deng
J
,
Diskin
B
, et al
.
CDK7 inhibition potentiates genome instability triggering anti-tumor immunity in small cell lung cancer
.
Cancer Cell
2020
;
37
:
37
54
.
258.
Weinreb
JT
,
Ghazale
N
,
Pradhan
K
,
Gupta
V
,
Potts
KS
,
Tricomi
B
, et al
.
Excessive R-loops trigger an inflammatory cascade leading to increased HSPC production
.
Dev Cell
2021
;
56
:
627
40
.
259.
Chatzidoukaki
O
,
Stratigi
K
,
Goulielmaki
E
,
Niotis
G
,
Akalestou-Clocher
A
,
Gkirtzimanaki
K
, et al
.
R-loops trigger the release of cytoplasmic ssDNAs leading to chronic inflammation upon DNA damage
.
Sci Adv
2021
;
7
:
eabj5769
.
260.
Crossley
MP
,
Song
C
,
Bocek
MJ
,
Choi
JH
,
Kousorous
J
,
Sathirachinda
A
, et al
.
R-loop-derived cytoplasmic RNA-DNA hybrids activate an immune response
.
Nature
2023
;
613
:
187
94
.
261.
Mosler
T
,
Conte
F
,
Longo
GMC
,
Mikicic
I
,
Kreim
N
,
Mockel
MM
, et al
.
R-loop proximity proteomics identifies a role of DDX41 in transcription-associated genomic instability
.
Nat Commun
2021
;
12
:
7314
.
262.
Makishima
H
,
Bowman
TV
,
Godley
LA
.
DDX41-associated susceptibility to myeloid neoplasms
.
Blood
2022
. doi .
263.
Buqué
A
,
Bloy
N
,
Perez-Lanzón
M
,
Iribarren
K
,
Humeau
J
,
Pol
JG
, et al
.
Immunoprophylactic and immunotherapeutic control of hormone receptor-positive breast cancer
.
Nat Commun
2020
;
11
:
3819
.
264.
Miyauchi-Hashimoto
H
,
Kuwamoto
K
,
Urade
Y
,
Tanaka
K
,
Horio
T
.
Carcinogen-induced inflammation and immunosuppression are enhanced in xeroderma pigmentosum group A model mice associated with hyperproduction of prostaglandin E2
.
J Immunol
2001
;
166
:
5782
91
.
265.
Miyauchi-Hashimoto
H
,
Tanaka
K
,
Horio
T
.
Enhanced inflammation and immunosuppression by ultraviolet radiation in xeroderma pigmentosum group A (XPA) model mice
.
J Invest Dermatol
1996
;
107
:
343
8
.
266.
Kunisada
M
,
Hosaka
C
,
Takemori
C
,
Nakano
E
,
Nishigori
C
.
CXCL1 inhibition regulates UVB-induced skin inflammation and tumorigenesis in Xpa-deficient mice
.
J Invest Dermatol
2017
;
137
:
1975
83
.
267.
Güngör
N
,
Haegens
A
,
Knaapen
AM
,
Godschalk
RW
,
Chiu
RK
,
Wouters
EF
, et al
.
Lung inflammation is associated with reduced pulmonary nucleotide excision repair in vivo
.
Mutagenesis
2010
;
25
:
77
82
.
268.
Sharp
SP
,
Malizia
RA
,
Walrath
T
,
D'Souza
SS
,
Booth
CJ
,
Kartchner
BJ
, et al
.
DNA damage response genes mark the early transition from colitis to neoplasia in colitis-associated colon cancer
.
Gene
2018
;
677
:
299
307
.
269.
Nalepa
G
,
Clapp
DW
.
Fanconi anaemia and cancer: an intricate relationship
.
Nat Rev Cancer
2018
;
18
:
168
85
.
270.
Sumpter
R
Jr.
,
Sirasanagandla
S
,
Fernández
ÁF
,
Wei
Y
,
Dong
X
,
Franco
L
, et al
.
Fanconi anemia proteins function in mitophagy and immunity
.
Cell
2016
;
165
:
867
81
.
271.
Brégnard
C
,
Guerra
J
,
Déjardin
S
,
Passalacqua
F
,
Benkirane
M
,
Laguette
N
.
Upregulated LINE-1 activity in the fanconi anemia cancer susceptibility syndrome leads to spontaneous pro-inflammatory cytokine production
.
EBioMedicine
2016
;
8
:
184
94
.
272.
Du
W
,
Erden
O
,
Wilson
A
,
Sipple
JM
,
Schick
J
,
Mehta
P
, et al
.
Deletion of Fanca or Fancd2 dysregulates Treg in mice
.
Blood
2014
;
123
:
1938
47
.
273.
Wu
L
,
Li
X
,
Lin
Q
,
Chowdhury
F
,
Mazumder
MH
,
Du
W
.
FANCD2 and HES1 suppress inflammation-induced PPARγ to prevent haematopoietic stem cell exhaustion
.
Br J Haematol
2021
;
192
:
652
63
.
274.
Park
E
,
Kim
H
,
Kim
JM
,
Primack
B
,
Vidal-Cardenas
S
,
Xu
Y
, et al
.
FANCD2 activates transcription of TAp63 and suppresses tumorigenesis
.
Mol Cell
2013
;
50
:
908
18
.
275.
Hess
J
,
Unger
K
,
Orth
M
,
Schötz
U
,
Schüttrumpf
L
,
Zangen
V
, et al
.
Genomic amplification of Fanconi anemia complementation group A (FancA) in head and neck squamous cell carcinoma (HNSCC): cellular mechanisms of radioresistance and clinical relevance
.
Cancer Lett
2017
;
386
:
87
99
.
276.
Liu
X
,
Liu
X
,
Han
X
.
FANCI may serve as a prognostic biomarker for cervical cancer: a systematic review and meta-analysis
.
Medicine (Baltimore)
2021
;
100
:
e27690
.
277.
Pang
Q
,
Fagerlie
S
,
Christianson
TA
,
Keeble
W
,
Faulkner
G
,
Diaz
J
, et al
.
The Fanconi anemia protein FANCC binds to and facilitates the activation of STAT1 by gamma interferon and hematopoietic growth factors
.
Mol Cell Biol
2000
;
20
:
4724
35
.
278.
Fagerlie
SR
,
Koretsky
T
,
Torok-Storb
B
,
Bagby
GC
.
Impaired type I IFN-induced Jak/STAT signaling in FA-C cells and abnormal CD4+ Th cell subsets in Fancc-/- mice
.
J Immunol
2004
;
173
:
3863
70
.
279.
Pang
Q
,
Keeble
W
,
Christianson
TA
,
Faulkner
GR
,
Bagby
GC
.
FANCC interacts with Hsp70 to protect hematopoietic cells from IFN-gamma/TNF-alpha-mediated cytotoxicity
.
Embo j
2001
;
20
:
4478
89
.
280.
Whitney
MA
,
Royle
G
,
Low
MJ
,
Kelly
MA
,
Axthelm
MK
,
Reifsteck
C
, et al
.
Germ cell defects and hematopoietic hypersensitivity to gamma-interferon in mice with a targeted disruption of the Fanconi anemia C gene
.
Blood
1996
;
88
:
49
58
.