Skip to main content
Molecular Biomedicine logoLink to Molecular Biomedicine
. 2025 Dec 29;6:152. doi: 10.1186/s43556-025-00385-1

PARPs and PARP inhibitors: molecular mechanisms and clinical applications

Fei Wang 1,#, Zhuyi Guo 2,#, Michael J Carr 3,4, Weifeng Shi 1,2,5,
PMCID: PMC12748464  PMID: 41460301

Abstract

Poly (ADP-ribose) polymerases (PARPs) are a diverse family of enzymes that regulate genome stability, cell death, and stress responses through ADP-ribosylation. Among them, PARP1, PARP2, and PARP3 are central to cellular DNA repair, while tankyrases, and their isoforms, contribute to telomere maintenance, transcriptional regulation, immune signaling, and metabolism. Dysregulated PARP activity drives genomic instability, apoptosis, parthanatos, and tumor microenvironment remodeling, thereby linking PARPs to oncogenesis, immune escape, and therapy resistance. Clinically, PARP inhibitors (PARPi), such as olaparib, niraparib, rucaparib, and talazoparib, exploit synthetic lethality in homologous recombination–deficient tumors and are increasingly applied in ovarian, breast, prostate, and pancreatic cancers. Beyond oncology, preclinical studies demonstrate antiviral efficacy of PARPi against hepatitis B virus, human immunodeficiency virus, and coronaviruses, and also therapeutic potential in neurodegeneration, cardiovascular disease, fibrosis, and metabolic disorders. However, PARPi resistance arises through restoration of DNA repair, replication fork protection, epigenetic changes, and drug-target dynamics, while adverse events—including hematologic toxicity, gastrointestinal disturbance, and organ-specific effects—limit a broader use. Next-generation PARPi with improved isoform selectivity, PROteolysis-TArgeting Chimera (PROTAC) degraders, and rational combinations with ATR/CHK1 inhibitors, immune checkpoint blockade, or epigenetic modulators offer strategies to enhance efficacy and overcome resistance. Emerging biomarker-driven approaches, including liquid biopsies and functional assays, may further personalize therapy. By integrating canonical DNA repair roles with non-canonical signaling and host–virus interactions, PARPs represent pivotal regulators. Similarly, the versatile therapeutics of PARPi have implications that extend beyond oncology into a broader and diverse range of other human diseases.

Keywords: PARPs, PARPi, Cancer, Virus, Clinical applications

Introduction

Poly(ADP-ribose) polymerases (PARPs) are a family of enzymes that play essential roles in diverse cellular processes. Their initial discovery dates to 1963, when Chambon and colleagues identified a nuclear enzyme capable of synthesizing poly(ADP-ribose) (PAR) from nicotinamide adenine dinucleotide (NAD+), providing the first evidence of ADP-ribosylation as a novel post-translational modification [1]. Subsequent biochemical studies in the 1970s established PARP1 as the major nuclear isoform, which was rapidly recruited and activated in response to DNA double-strand breaks (DSBs) proving fundamental to DNA repair [2]. The cloning of PARP1 in the 1980s revealed its modular structure—including zinc-finger DNA-binding motifs, an automodification region, and a catalytic domain—which served to elucidate the mechanistic basis for its role in genome monitoring [3, 4]. A major conceptual advance emerged in the 1990s when genetic and pharmacologic studies demonstrated that PARP1 is indispensable for the base excision repair (BER) pathway, firmly linking PARP activity to both genome integrity and cellular survival [5]. Among the 17 PARP family members, PARP1 remains the most extensively studied enzyme and is central to cellular responses to DNA damage [6, 7]. The PARP family of enzymes, via poly-ADP ribosylation (PARylation), coordinate a diverse network of signaling pathways controlling genomic stability, transcriptional regulation, chromatin remodeling, and cell fate decisions [8].

The early 2000s marked a transformative era for the understanding of the PARP family of proteins with the emergence of the paradigm of synthetic lethality [9]. Pioneering studies in 2005 revealed that concurrent loss of both PARP1 activity and BRCA1/2-mediated homologous recombination (HR) resulted in catastrophic DNA damage and selective tumor cell death, which have fundamentally reshaped strategies for targeted cancer therapies. These insights directly enabled the development of the PARP inhibitors (PARPi), which block repair of single-strand breaks (SSBs), forcing their conversion into lethal DSBs in HR-deficient cancers [10, 11]. These mechanistic foundations supported the approval of multiple PARPi for an expanding range of malignancies, establishing PARPi as a new pillar of precision oncology [9]. Subsequent breakthroughs further accelerated the field, including the discovery of PARP-trapping as a distinct cytotoxic mechanism [12], and high-resolution structural studies of PARP1-DNA-inhibitor complexes via crystallography and cryo-electron microscopy, which revealed dynamic conformations essential for sensing DNA breaks [13, 14]. The subsequent identification of tankyrases (PARP5A/B) expanded PARP biology into telomere maintenance and Wnt signaling [15]. The discovery of catalytically inactive members, such as PARP9 and PARP13, together with mono-ADP-ribosylating enzymes, like PARP15, expanded the functional landscape of the PARP family into immune regulation and RNA metabolism, illustrating the plethora of cellular processes impinged upon and the evolving therapeutic scope for their clinical applications [16].

Despite these noteworthy advances, major challenges still remain, including tissue-specific toxicities, adaptive resistance, and an incomplete understanding of isoform-specific functions of PARPi. Today, the understanding of PARP biology has evolved from a single DNA repair enzyme to a multifunctional signaling network that governs genome maintenance, immunity, and metabolism [17, 18]. The development of next-generation PARPi, PROteolysis-TArgeting Chimera (PROTAC)-based PARP degraders, and isoform-specific inhibitors marks the latest phase in this trajectory, offering new opportunities to overcome resistance and extend therapeutic applications beyond oncology [1921]. 

This review aims to integrate the current knowledge of PARP biology with clinical experience regarding PARPi, highlighting their expanding roles not only in malignancy but also in infectious and other non-malignant diseases. We begin by outlining the classification, structure, and multifaceted biological functions of PARPs—covering their canonical roles in DNA repair, apoptosis, and parthanatos, as well as their non-canonical functions in immunity, transcription, and metabolism. We then explore the pathologic mechanisms of action of PARPs in cancer, viral infections [2225], and other diseases such as neurodegenerative, cardiovascular, and metabolic disorders [2630]. Next, we detail the mechanisms of action of PARPi in both normal and malignant cells, followed by a comprehensive discussion of their clinical applications in oncology, antiviral therapy, and other disease settings. Finally, we address emerging challenges including drug resistance and safety concerns, and conclude with future perspectives on next-generation inhibitors and novel therapeutic strategies. By systematically organizing these themes, this review provides an overarching and up-to-date resource for researchers and clinicians interested in the broadening therapeutic landscape of PARP targeting to treat a diverse array of human diseases.

Mechanisms and biological functions of PARPs

Classification and structure of PARPs

The PARP family of proteins comprise 17 enzymes that share a conserved catalytic core but differ markedly in their regulatory domains and biological functions (Fig. 1). Among these, PARP1, PARP2, and PARP3 are the most thoroughly characterized, largely due to their central roles in DNA repair and cellular stress responses [31]. PARP1, in particular, is rapidly activated upon sensing DSBs, catalyzing PARylation and coordinating the assembly of DNA repair factors necessary for safeguarding genome integrity [32, 33]. By contrast, PARP5A and PARP5B possess distinctive ankyrin-repeat clusters and sterile alpha motif (SAM) domains that mediate their functions in both telomere maintenance and the regulation of Wnt/β-catenin signaling [15, 34]. Other PARP family members—including PARP7, PARP12, and PARP13—participate primarily in innate immune responses, antiviral defenses, and transcriptional regulation [35]. Meanwhile, PARP9, PARP14, and PARP15, though lacking full enzymatic activity, operate as mono-ADP-ribosyltransferases or ADP-ribose-binding effectors that modulate diverse intracellular signaling pathways [16, 36].

Fig. 1.

Fig. 1

Domain organization and classification of human PARP family members. The figure is drawn based on the structures available from Uniprot. The 17 human poly(ADP-ribose) polymerases (PARPs) are grouped into five subfamilies based on their structural domains and functional features: DNA-dependent PARPs (PARP1, PARP2, and PARP3), Tankyrases (PARP5A, and PARP5B), CCCH-type PARPs (PARP7, PARP12, and PARP13), Macro domain-containing PARPs (PARP9, PARP14, and PARP15), and other PARPs with diverse structures (PARP4, PARP6, PARP8, PARP10, PARP11, and PARP16). ART (ADP-ribosyl transferase domain), essential for catalytic activity; Zn (Zn finger domain) and BRCT (BRCA C-terminal domain), involved in DNA binding and protein interactions; WGR (Trp-Gly-Arg domain), important for DNA-dependent activity; ARCs (ankyrin repeat cluster) and SAM (sterile alpha motif domain), characteristic of tankyrases for scaffolding functions; Macros domains, involved in ADP-ribose binding; HD (helical domain), important for enzymatic allosteric regulation and activation; VWFA (von Willebrand factor type A domain), mediating metal ion-dependent adhesion to partner proteins; WWE (Trp-Trp-Glu), VIT (vault protein interacting) domain and UIM (Ubiquitin Interacting Motif), mediating protein–protein interactions; RRM (RNA Recognition Motif), mediating the recognition and binding to damaged DNA

Despite their functional diversity, the vast majority of PARPs contain a C-terminal catalytic domain defined by the conserved HYE (His-Tyr-Glu) tripeptide motif (Fig. 1), which catalyzes the transfer of ADP-ribose from NAD⁺ to target proteins [37]. This reversible post-translational modification plays a critical role in modulating the function of targeted proteins, chromatin organization, and intracellular signaling dynamics [37, 38]. In contrast, the N-terminal and accessory domains vary substantially across the PARP family members and largely dictate their functional specialization. PARP1, for example, contains three zinc-finger motifs that allow rapid recognition of DNA lesions, an automodification region that regulates chromatin accessibility, and a WGR (Trp-Gly-Arg) domain that mediates DNA binding and intramolecular communication [14, 39]. PARP2 and PARP3 retain the WGR domain but lack zinc-finger modules, corresponding to their more focused involvement in SSB repair and chromatin association [14, 38]. The PARP5A/B enzymes feature extended ankyrin-repeat arrays that mediate protein interactions and drive substrate specificity, as well as SAM domains that promote polymerization and amplify downstream signaling events [40]. Catalytically attenuated members, such as PARP9, PARP13, and PARP15, instead employ macrodomains, RNA-binding motifs, or WWE (Trp-Trp-Glu) modules to regulate immune signaling and RNA metabolism [16, 41, 42]. Together, the conserved catalytic architecture paired with highly diversified regulatory domains underpins the diverse functional spectrum of the PARP family (Fig. 1).

PARPs in cellular homeostasis

The role of PARPs in DNA repair

PARPs are activated in response to SSBs or DSBs in DNA repair, particularly PARP1, PARP2, and PARP3, which are primary responders to cellular DNA damage. They recognize DNA SSBs or DSBs via their structural domains and rapidly recruit repair machinery by catalyzing the PARylation or mono(ADP-ribosyl)ation (MARylation) of substrate proteins. This activity initiates and coordinates various repair pathways, including BER, single-strand break repair (SSBR), and double-strand break repair (DSBR) [43, 44]. As the most extensively studied member, PARP1 is involved in HR and non-homologous end joining (NHEJ), highlighting its multifaceted role in maintaining genome integrity [4547]. The PAR strand synthesized by PARP1 also promotes chromatin remodeling, allowing better access of repair proteins to damaged DNA [47]. Furthermore, this critical function is the foundation for the use of PARPi in cancer therapy, which exploit "synthetic lethality" to selectively target tumors with HR deficiency, representing a major breakthrough in targeted oncology [48].

The role of PARPs in transcriptional regulation and chromatin remodeling

Recent studies have substantially broadened the conceptual framework of PARP biology, positioning PARP1 and related family members as active modulators of transcription rather than passive accessory factors. PARP1 is now recognized as a chromatin-associated regulator that fine-tunes promoter accessibility and shapes RNA polymerase II (Pol II) behavior during early transcriptional events. Through PARylation of histones and pause-associated factors, PARP1 can influence the transition from Pol II pausing to productive elongation, thereby enabling rapid transcriptional responses to stress and developmental cues [49]. This regulatory activity operates in a locus-specific manner: PARP1 stabilizes chromatin compaction at metabolically hyperactive regions while promoting accessibility at poised promoters, effectively redistributing transcriptional output according to cellular demand [50].

Beyond PARP1, recent work highlights the involvement of mono-ADP-ribosyltransferases (mono-ARTs), such as PARP7 and PARP14, in remodeling transcription factor networks. PARP7 has been shown to MARylate aryl hydrocarbon receptor (AHR), altering their DNA-binding behavior and downstream transcriptional programs [51, 52], whereas PARP14—often functioning with PARP9-DTX3L complexes—helps integrate cytokine signaling with chromatin-embedded transcriptional states [53, 54]. An emerging theme across these findings is that ADP-ribosylation acts as a tunable regulatory layer that intersects with chromatin structure and Pol II–proximal complexes to influence transcriptional outputs [55]. Importantly, recent evidence suggests that PARP1 also senses DNA:RNA hybrids (R-loops) formed during transcription and coordinates their resolution with RNase H2 and chromatin remodelers, positioning PARP activity at a critical interface between transcription and genome stability [56]. Notably, new findings further demonstrate that PARP1-dependent resolution of R-loops requires the coordinated action of EXD2 and CBP: when PARP1 is depleted or its catalytic activity is inhibited, EXD2 and CBP fail to localize to R-loop sites, leading to impaired hybrid removal [57]. Collectively, these discoveries support a shift in perspective: PARPs should be viewed not only as DNA-repair enzymes but as dynamic regulators of transcriptional circuitry whose activity influences both gene expression patterns and cellular homeostasis.

The role of PARPs in apoptosis

Beyond their role in DNA repair, PARPs are also involved in the regulation of apoptosis. In the early stages of apoptosis, caspase-3 is activated and cleaves PARP1, splitting it into two segments: a 24 kDa N-terminal and an 89 kDa C-terminal fragment [58]. This cleavage event is considered one of the quintessential hallmarks of apoptosis as it causes PARP1 to lose its DNA repair function and thereby promote cell death [59]. At the same time, excessive activation of PARP1 is also closely related to apoptosis in some viral infections. For example, certain DNA and RNA viruses trigger apoptosis by inducing DNA damage or by activating PARP1, thereby promoting virus release and transmission [60, 61]. In addition, PARP1 also triggers apoptosis by interacting with the stimulator of interferon genes (STING) protein [62]. Conversely, certain members like the tankyrases (PARP5A/B) [63], PARP9 [64], and PARP14 [65], exhibit anti-apoptotic functions, promoting cell survival. This dual and often opposing role in apoptosis and parthanatos highlights the complex and critical regulatory network governed by the PARP family in determining cell fate, whose dysregulation is intimately linked to tumorigenesis and therapy resistance.

The role of PARPs in parthanatos

Parthanatos is a distinct form of regulated cell death primarily mediated by PARP1. It is characterized by a series of biochemical events that lead to cell demise without the activation of caspases, which are the hallmark of apoptosis [66]. Upon sensing DNA damage, PARP1 becomes hyperactivated, leading to the rapid synthesis of PAR chains that modify various nuclear proteins. This process is crucial for the release and nuclear translocation of apoptosis-inducing factor (AIF), where it interacts with macrophage migration inhibitory factor (MIF) to enhance its nuclease activity, leading to extensive DNA fragmentation [67, 68]. Moreover, the depletion of NAD+ and ATP due to PARP1 activation also contributes to cellular energy failure and promotes parthanatos [69]. The interplay between PARP1 activity and cellular energy status is vital for determining cell fate, making PARP1 a promising therapeutic target in diseases where parthanatos is implicated. For example, inhibition of PARP1 has been shown to alleviate neuronal damage in models of neurodegenerative diseases and ischemic injury, highlighting the importance of this pathway in therapeutic interventions [7072].

Non-canonical functions of PARPs

Beyond canonical DNA repair and apoptosis, many PARP members play significant roles in a broad spectrum of cellular processes (Table 1). PARP1 participates in various biological processes such as ferroptosis [73], necroptosis [74, 75], pyroptosis [76], cellular metabolism [113], transcriptional regulation, and the cell stress response [114], thereby contributing to both immunomodulation and cellular invasiveness. PARP5A and PARP5B are involved in regulating cell proliferation through telomere maintenance and the Wnt/β-catenin signaling pathway [48, 63]. PARP7 [48], PARP13 [108], and PARP14 [54] act as critical regulators of innate immune responses, modulating immune microenvironments via interferon signaling. These non-canonical functions highlight the expansive influence of the PARP family proteins in cellular signaling networks and suggest novel perspectives for clinical trials involving combinatorial therapeutic strategies. It is important to note that not all PARP family members possess catalytic activity, e.g. PARP9 and PARP13, yet they still play key roles as scaffold proteins, molecular adaptors, or signal modulators via their protein-interaction domains (macrodomains and zinc fingers). For instance, PARP9 and PARP13 have been reported to facilitate the assembly of signaling complexes or function as RNA-binding proteins to regulate gene expression [115]. Even among the catalytically-active members, their protein products vary: PARP1, PARP2, PARP4, PARP5A, and PARP5B synthesize long-chain PAR polymers [77, 116], while most other members primarily catalyze MARylation [77]. This diversity in catalytic activity and molecular functions adds a significant layer of complexity to the regulatory repertoire of the PARP family.

Table 1.

Classification of the PARP family members with key domains and functions

PARP Member Functions Enzymatic Activity Reference
DNA Repair Apoptosis Parthanatos Others
PARP1 SSBR, DSBR Pro-apoptotic under severe DNA damage Primary inducer via PAR accumulation Ferroptosis, necroptosis, pyroptosis, mitosis, cell cycle, metastasis, transcription, chromatin structure modulation, membrane repair and cell stress response PAR, MAR [6, 48, 7383]
PARP2 SSBR, DSBR May suppress apoptosis / Mitosis, chromatin structure modulation, telomere maintenance PAR [6, 48, 82, 8489]
PARP3 SSBR, DSBR / / Mitosis, centrosome regulation, chromatin structure modulation, transcription PAR, MAR [6, 48, 77, 80, 82, 84, 86, 90, 91]
PARP4 DSBR Pro-apoptotic / Cell transportation and vault particle regulation PAR, MAR [48, 9294]
PARP5A (Tankyrase-1) DSBR Anti-apoptotic / Pyroptosis, cellular metabolism, telomere maintenance, mitosis, cell division, necroptosis PAR [6, 48, 63, 9597]
PARP5B (Tankyrase-2) DSBR Anti-apoptotic / Pyroptosis, mitosis, cellular metabolism, telomeric and centromeric regulation PAR [48, 63, 97]
PARP6 / / / Cell structure, adhesion, motility, spindle pole regulation and cell replication MAR [48, 98]
PARP7 / Anti-apoptotic / Transcription regulation, cell structure, adhesion and motility, innate immunity, and cell stress response MAR [48, 99101]
PARP8 / / / Membrane and nuclear envelope formation MAR [48]
PARP9 DSBR Anti-apoptotic / Lipid metabolism, innate immunity Catalytically Inactive [64, 77, 80, 102104]
PARP10 Nascent strand DNA gaps repair, translesion synthesis (TLS), DSBR Pro-apoptotic / Lipid metabolism, spindle pole regulation and cell replication MAR [36, 48, 102, 105, 106]
PARP11 / / / Spermatogenesis, membrane and nuclear envelope formation MAR [48]
PARP12 / Anti-apoptotic / Necroptosis, mouse oocyte meiotic maturation, cell stress response MAR [48, 107]
PARP13 / Pro-apoptotic in viral infection / Innate immunity Inactive [48, 77, 80, 108, 109]
PARP14 Homologous Recombination (HR), Replication Fork Protection, SSBR Anti-apoptotic / Lipid metabolism, transcription regulation, cell structure, adhesion, motility and innate immunity MAR [36, 48, 65, 102]
PARP15 / / / Post-translational modification of proteins and negative regulator of transcription MAR [48]
PARP16 / Pro-apoptotic / ER stress response, cell stress response, membrane and stabilization of amyloid precursor protein mRNA MAR [48, 110112]

BER base excision repair, SSR single-strand break, ER Endoplasmic reticulum, MAR MARylation, PAR PARylation

PARPs in disease pathogenesis

Mechanisms of PARPs in cancers

PARPs and genomic instability

PARPs collectively maintain genomic stability through distinct yet complementary mechanisms, and their dysregulation serves as a critical driver of tumorigenesis. PARP1 acts as a central sensor and catalytic effector by recognizing SSBs and catalyzing PAR chains to recruit DNA repair factors, primarily governing BER [117, 118]. PARP1 dysfunction leads to the accumulation of SSBs that can progress into lethal DSBs, directly promoting genomic instability. PARP2 shares overlapping functions with PARP1, playing essential roles in SSBR and DSBR, thereby contributing to genome integrity [119]. Compared with SSBR, PARP3 is more involved in DNA DSBR, influencing genomic stability through the regulation of HR and NHEJ [120]. Additionally, PARP5A and PARP5B contribute indirectly, though significantly, by regulating both telomere maintenance and modulating oncogenic signaling pathways. In cases of telomere dysfunction, DSBs can be triggered, affecting chromosome stability [48, 63]. Moreover, various other PARP family members are activated by distinct forms of DNA damage, and serve to orchestrate cellular responses to this damage, thereby modulating genomic instability (Table 1). Therefore, the dysfunction of PARPs will aggravate genomic instability from multiple levels (DNA repair, replication, and chromosome maintenance), thus driving the occurrence and subsequent development of tumors.

The role of PARPs in tumor cell death

PARPs play a dual role in tumorigenesis by regulating multiple cell death pathways: on the one hand, members such as PARP2, PARP5A/B, PARP7, and PARP14 inhibit apoptosis and, thus, serve to promote cancer development. PARP2 enhances nasopharyngeal carcinoma growth via SIRT1/AMPK signaling [88]. PARP5A/B promotes PTEN degradation via PARylation to activate PI3K-AKT signaling and can also modulate AXIN stability activating WNT/β-catenin pathways, thereby inhibiting apoptosis [63, 121]. PARP7 ADP-ribosylates FRA1 to suppress IRF1/IRF3-dependent apoptosis [99] and, finally, PARP14 inhibits the pro-apoptotic gene JNK1 to support metabolic growth in hepatocellular carcinoma [65]. On the other hand, however, PARP16 inhibits tumor cell growth and metastasis by activating the endoplasmic reticulum stress signaling pathway and inducing apoptosis [110]. PARP4, stimulated by interferon gamma (IFN-γ) and tumor necrosis factor-alpha (TNF-α), mediates apoptosis in human neuroblastoma cells [92]. PARP1 also exerts context-dependent effects: it promotes apoptosis to eliminate DNA-damaged cells early in tumorigenesis, but excessive activation depletes NAD+ and ATP, triggering parthanatos and inhibiting cancer progression [122]. During radiotherapy, radiation-induced PAR polymers, synthesized by PARP1, bind to STING and induce apoptosis [123]. Conversely, inhibiting PARP1 suppresses tumor growth through two mechanisms: inducing ferroptosis by repressing SLC7A11, or triggering pyroptosis via caspase-3-mediated gasdermin E cleavage [73, 76]. Additionally, PARP1 variably modulates autophagy via the MRPL21-PARP1 axis to promote chemoresistance in head and neck squamous cell carcinoma [124], while being activated by UHRF2 to promote autophagy and malignancy in hepatocellular carcinoma [125]. In the HT-29 human colon adenocarcinoma cell line, T lymphocytes can sensitize IFN-γ-induced PARP cleavage and promote tumor cell apoptosis [126]. In oesophageal carcinoma cells, IFN-λ1 induces apoptosis by inducing the cleavage of PARP, thus inhibiting the growth of cancer cells [127]. These multifaceted mechanisms underscore the therapeutic potential of targeting PARPs in combinatorial therapeutic strategies, involving chemotherapy, radiotherapy, and immunotherapy which requires further investigation in clinical trials to assess additive and synergistic effects.

The role of PARPs in the regulation of the tumor microenvironment

PARPs are also increasingly acknowledged for their involvement in the modulation of the tumor microenvironment (TME), which affects both tumor development and the anti-tumor immune responses [128130]. PARPs are actively engaged in the regulation of inflammatory responses within the TME, which may exacerbate tumor growth, contingent upon specific circumstances. For example, PARP7 promotes immune escape and tumor growth in the TME by negatively regulating the nucleic acid sensing-mediated type I interferon signaling pathway in tumor cells and inhibiting the inflammatory response [101]. PARP1 activates the IL-6-STAT3-cyclin D1 signaling axis to promote the progression of rectal cancer [131]. PARP14 can enhance the inflammatory response by supporting the polarization of macrophages to the tumor-promoting M2 type, and, thus, promote the occurrence and development of multiple tumors, such as breast cancer [132]. Conversely, PARP14 can also promote inflammatory cell infiltration by enhancing the expression of the IL-4-induced inflammatory cell chemokines CCL17 and CCL22, thus exacerbating the occurrence and development of follicular lymphoma [133]. Furthermore, PARPs regulate tumor cell migration and invasiveness by affecting microtubule stability. PARP7 is able to MARylate α-tubulin to promote microtubule instability, which may enhance the growth and motility of ovarian cancer cells [134]. The ability of PARPs to modulate the TME highlights their importance not only in tumor biology but also within the therapeutic arena, suggesting that targeting PARP activity may improve the efficacy of both chemo-, radio- and immunotherapies by reshaping the TME to support anti-tumor immunity [135137]. The dynamic interaction between PARPs, immune response, and microtubule stability offers a complex framework for cancer treatment, indicating that approaches aimed at modulating PARP activity could potentially serve to further improve therapeutic efficacies [138, 139].

The role of PARPs in replication stress tolerance and fork protection

Replication stress is increasingly recognized as a defining vulnerability of many cancers, and recent studies have clarified how multiple PARP family members safeguard replication fork integrity under such conditions. PARP1 is recruited to stalled replication forks, where it helps prevent Mre11-mediated nascent DNA degradation and supports fork stability and restart under replication stress [140]. Emerging evidence suggests that PARP2 may also contribute to fork stabilization in specific contexts [141]. More recent work shows that PARP1 also facilitates fork reversal, enabling template switching when forks encounter obstacles, although dysregulated PARP1 activity can lead to pathological hyper-reversal, fueling genomic instability and tumor evolution [47, 142]. In addition, PARP10 interacts with PCNA via its PIP-box motif and recruits RAD18 to promote PCNA mono-/poly-ubiquitination, thereby facilitating translesion DNA synthesis (TLS) and enabling cells to bypass DNA lesions under replicative stress conditions [143]. Collectively, these findings indicate that PARPs function as a hub linking replication-fork remodeling to DNA repair, enabling cancer cells to survive chronic replication stress. Therapeutically, disrupting fork-protective functions of PARPs may offer a strategy to selectively exploit replication stress–associated weaknesses in tumors with oncogene-driven hyperproliferation.

The role of PARPs in cancer epigenetics and transcriptional reprogramming

Increasing evidence demonstrates that the influence of PARPs on tumor biology extends into the regulation of chromatin architecture and transcriptional programs. PARP1 remodels cancer-associated chromatin by PARylating the nucleosomal core histones and chromatin-bound regulatory factors, thereby promoting chromatin accessibility at promoters and supporting c-MYC- and E2F-driven proliferative gene expression in tumor contexts [144]. Furthermore, PARP7 fine-tunes AHR-dependent transcriptional outputs in lung and liver cancers, directly regulating the expression of oncogenic networks [145]. PARP14, frequently overexpressed in hematologic malignancies, reinforces IL-4/STAT6 transcriptional circuits, promoting tumor cell survival and metabolic adaptation [133, 146]. Studies highlight that PARP-mediated ADP-ribosylation functions as a versatile epigenetic modification that integrates metabolic state, immune signaling, and cytokine inputs to reprogram oncogenic transcription [147, 148]. These mechanistic insights underscore the potential of PARPs as master regulators that couple chromatin dynamics to tumor-promoting transcriptional networks, providing a strong therapeutic rationale for combining PARP inhibition with transcription- and/or chromatin-targeted therapies.

Mechanisms of PARPs in virus infection

The role of PARPs in RNA virus infection

Crucially, PARPs can inhibit (Fig. 2a) or promote (Fig. 2b) the type I interferon (IFN) signaling pathway through multiple mechanisms during RNA virus infection, thereby affecting the host's antiviral immune responses. The influenza A virus (IAV) hemagglutinin (HA) protein can induce the degradation of the type I IFN receptor 1 (IFNAR1), and PARP1 plays a key role in this process. IAV infection induces cytoplasmic translocation of PARP1, and PARP1 interacts with HA to promote the ubiquitination and degradation of IFNAR1, thereby inhibiting the activation of the IFN signaling pathway (Fig. 2a) [24]. Upon vesicular stomatitis virus (VSV) infection, PARP11 is able to MARylate the ubiquitin E3 ligase β-TrCP, promoting its binding to IFNAR1, eventually resulting in ubiquitination and degradation of IFNAR1 (Fig. 2a). This process inhibits the activation of the IFN signaling pathway and attenuates host antiviral responses [149]. Meanwhile, PARP7 can inhibit TBK1 phosphorylation, thereby negatively regulating the production of IFN in VSV infection (Fig. 2a) [150].

Fig. 2.

Fig. 2

PARPs leverage the IFN signaling pathway to regulate host immunity upon RNA virus infection. a PARP1 and PARP11 promote IAV and VSV replication, respectively, by facilitating the ubiquitination of IFNAR. PARP7 promotes VSV replication by inhibiting the TBK1-IRF3 signaling pathway and reducing the production of IFNβ. b PARP9 can inhibit the replication of VSV, IAV, ReV and EMCV by promoting the production of IFN. In SARS-CoV-2 infection, IFN-γ-induced PARP14 can cooperate with PARP9 to regulate the production of IFN and inhibit viral replication

In contrast, PARP9 specifically recognizes the dsRNA genome of Reovirus (ReV), initiating a signaling cascade that activates interferon regulatory factor 3 (IRF3) and interferon regulatory factor 7 (IRF7) through the PI3K/AKT3 signaling pathway, promoting IFN production. Importantly, this process does not rely on the classical mitochondrial antiviral signaling protein (MAVS) innate immune adaptor signaling pathway, but, instead, operates by enhancing IFN signaling by direct phosphorylation of key residues of IRF3 and IRF7. Although studies have not clearly determined whether PARP9 directly binds to dsRNA produced by IAV and vesicular stomatitis virus (VSV), infection can activate the same signaling pathway induced by PARP9, which in turn promotes the production of type I interferon. This suggests that PARP9 may activate the antiviral signaling pathway by directly binding to the dsRNA of IAV and VSV, or by other indirect mechanisms, in response to infection with these viruses which warrants further studies (Fig. 2b) [151]. PARP9-DTX3L, as an E3 ubiquitin ligase complex, can interact with STAT1 and enhance the transcriptional activity of STAT1, thereby promoting the expression of interferon-stimulated genes (ISGs). The PARP9-DTX3L complex is also able to directly target the encephalomyocarditis virus (EMCV) 3C protease, degrading these viral proteins through the ubiquitin–proteasome system (Fig. 2b), inhibiting viral replication [152]. Similarly, in SARS-CoV-2 infection, IFN-γ induces ADP-ribosylation by activating the PARP9/DTX3L complex and PARP14. PARP9 and PARP14 can synergistically inhibit viral replication. PARP9-DTX3L can remove the ADP modification catalyzed by PARP14 to regulate macrophage activation. The Nsp3 of SARS-CoV-2 can also hydrolyze this modification, thereby inhibiting the host's antiviral responses (Fig. 2b) [54, 153].

In addition to modulating the IFN signaling pathway, PARPs can also regulate host immunity through pathways such as apoptosis, RNA degradation, viral transcription, interaction with host proteins, autophagy, and proteasome degradation (Fig. 3). IAV infection induces caspase-3 activation, which subsequently triggers PARP1 cleavage, culminating in apoptosis [154]. In IAV infection, IFN-γ-induced PARP12 regulates RIPK1 and RIPK3 through ADP-ribosylation and promotes RIPK1-RIPK3-mediated programmed cell necroptosis [155]. Furthermore, PARP1 can inhibit IAV replication following the activation of the PARP1 enzymatic activity. The inhibition of the PARP1 enzymatic activity can promote the RNA-dependent RNA polymerase (RdRp) activity of IAV and virus replication though the specific mechanisms are still unclear [156, 157]. Sirtuin 3 (Sirt3) decreases IAV replication and the associated inflammatory damage, oxidative stress, and mitochondrial dysfunction by inhibiting the PARP1 activity [158]. PARP7 can bind to Sindbis virus (SINV) RNA and exosome complex component 5 (EXOSC5) through its zinc finger domain and promote viral RNA degradation, thereby inhibiting SINV replication (Fig. 3) [159]. PARP1 promotes ELL2-super elongation complex (SEC) formation by increasing the level of ELL2 protein and inhibiting the expression of E3 ubiquitin ligase Siah1, which is crucial for the transcriptional activation of human immunodeficiency virus (HIV) [23].

Fig. 3.

Fig. 3

PARP modulates immune responses in response to RNA virus invasion through multiple pathways. PARP1 mediates apoptosis and regulates IAV replication. IFN-γ-induced PARP12 mediates necrosis and inhibits IAV replication. The role of PARP1 enzyme activity in IAV infection is complex and can be inhibited by RdRp to promote viral replication, and also Sirt3 inhibition can alleviate mitochondrial oxidative stress. PARP1 inhibits the expression of Siah1 and increases the level of the ELL2 protein, which in turn promotes the transcription of HIV. PARP1 activates transcription of PTEN and AKT, which in turn promotes autophagy to regulate JEV replication. PARP12 is able to ADP-ribosylate NS1 and NS3 proteins, resulting in their ubiquitination and proteasomal degradation, inhibiting ZIKV replication

PARPs play critical roles during the infection cycle of the important and genetically-related flaviviruses Japanese encephalitis virus (JEV) and Zika virus (ZIKV). During JEV infection, PARP1 can cause PARylation of AKT and affect its phosphorylation. At the same time, PTEN transcription can be upregulated, and then AKT is negatively regulated. PARP1-mediated AKT inactivation promotes autophagy and JEV pathogenesis (Fig. 3) [160]. Recently, studies have found that PARP12 is able to ADP-ribosylate the NS1 and NS3 proteins of ZIKV and can inhibit ZIKV replication (Fig. 3), leading to their ubiquitination and proteasomal degradation [161]. Collectively, these findings suggest that multiple PARPs respond differently to RNA virus infections with distinct antiviral strategies.

The role of PARPs in DNA virus infection

Some DNA viruses can take advantage of the enzymatic activity of PARP or promote the protein modification activities of PARP to facilitate viral replication (Fig. 4a). Herpes simplex virus type 1 (HSV-1) infection activates the DNA-dependent protein kinase (DNA-PK), promoting PARP1 phosphorylation and nuclear-to-cytoplasmic transport, which modifies the cyclic guanosine monophosphate (GMP)–adenosine monophosphate (AMP) synthase (cGAS) protein through PARylation, inhibiting its DNA binding ability and diminishing IFN responses [162, 163]. Interestingly, HSV-1 infection induces PARP5A phosphorylation via ERK, thereby promoting its expression and translocation to the nucleus by interacting with ICP0, thus enhancing HSV replication [164]. During another herpesvirus infection, human cytomegalovirus (HCMV; human herpesvirus 5, HHV-5), PARP1 is also activated and transferred from the nucleus to the cytoplasm, promoting viral replication. The UL76 protein of HCMV binds to the BRCT domain of PARP1, maintaining the host cell survival [165]. In Epstein-Barr virus (EBV; human herpesvirus 4, HHV-4) infection, PARP1 may modify EBNA1 through PAR, promoting the replication and maintenance of the EBV genome. Recent studies have shown that PARP1 can maintain the three-dimensional structure and latent state of the EBV genome by stabilizing the CTCF binding site (Fig. 4a) [166, 167].

Fig. 4.

Fig. 4

PARPs regulate immune responses to DNA viral infection via multiple cellular pathways. a PARP1 mediates the cGAS/STING signaling pathway through cytosolic translocation and promotes HSV-1 replication. PARP5 promotes HSV-1 replication through nuclear translocation. PARP1 binds to the UL76 protein of HCMV, inhibits IFN production, and promotes viral replication. PARP1 binds to EBNA1 and promotes replication and maintenance of the EBV genome. At the same time, PARP1 can also maintain the latent state of the EBV genome by stabilizing the CTCF binding site. b PARP1 modifies KSHV LANA and MHV68 RTA by PARylation to inhibit the replication of KSHV and MHV-68. PARP1 recruits NK cells through the CCL2-CCR2 axis, thereby inhibiting VV replication. PARP1 senses HBV DNA via the Ku70/80 complex, activates IRF1 and promotes secretion of CCL3 and CCL5 chemokines. PARP1 activates the ATM and ATR signaling pathways by sensing SSBs and DSBs, thereby inhibiting ADV replication. PARP5 binds to EBNA1 and inhibits EBV replication, and PARP1 is also able to bind to the BZLF1 promoter region of EBV, restricting EBV cleavage reactivation

The role of PARPs in DNA viruses is therefore multifaceted, and PARP can also inhibit viral replication through alteration of enzymatic activities, protein modification, and recruitment of immune cells (Fig. 4b). PARP1 can directly bind to the terminal repeat (TR) sequences of Kaposi’s sarcoma-associated herpesvirus (KSHV; human herpesvirus 8, HHV-8) and modify the latency-associated nuclear antigen (LANA) protein through PARylation to inhibit the replication of KSHV during the incubation period. The activity of PARP1 affects both the stability and replication efficiency of the viral genome, and is a key factor in the maintenance of the KSHV incubation period [168]. PARP1 inhibits replication of murine gammaherpesvirus 68 (MHV-68) by PARylation of the replication and transcription activator (RTA) protein, a key switch molecule in the lytic replication of MHV-68. The vPIP protein encoded by MHV-68 interacts with PARP1 to prevent the inhibition of RTA by PARP-1, thereby releasing the inhibition of RTA by the PARylation event (Fig. 4b) [169].

In a model of intraperitoneal vaccinia virus (VV) infection, PARP1 recruits NK cells to the site of infection, through the CCL2-CCR2 axis, thereby facilitating viral control (Fig. 4b) [7]. PARP1 senses hepatitis B virus (HBV) DNA in the cytoplasm through the DNA repair protein Ku70/80 complex, activates IRF1 and promotes the secretion of CCL3 and CCL5 chemokines (Fig. 4b), thereby recruiting immune cells and triggering hepatitis [22]. PARP1 activates the ATM and ATR signaling pathways by sensing intermediate products, such as SSBs and DSBs, produced during adenovirus (ADV) DNA replication, thereby inhibiting ADV replication. ADV can bind to PARP1 through its E4orf4 protein and inhibits its activity reducing PARP1-mediated DNA damage signaling and promoting viral replication [170]. In EBV infection, PARP5 binds to EBNA1 and inhibits the replication of EBV (Fig. 4b) [171]. PARP1 is also capable of binding to the BZLF1 promoter region of EBV (Fig. 4b), restricting EBV lytic reactivation [172]. In conclusion, during DNA infection, PARPs regulate the antiviral responses by multiple pathways.

Crosstalk mechanisms between cancer and viral infections

PARPs represent a critical node where cancer biology and viral infection intersect, as both contexts exploit DDR, metabolic reprogramming, and innate immune pathways. Mechanistically, cytoplasmic PARP1 can PARylate cGAS to curb its DNA binding, attenuate STING activation, and blunt type I interferon output—an immune-evasion strategy exploited during HSV-1 infection [163]. At the interface linking the DDR and innate immunity, PARP inhibition increases unresolved DNA and cytosolic DNA fragments, thereby activating cGAS–STING and enhancing antitumor immunity (and its synergy with checkpoint blockade), underscoring that the bidirectional control of immunity by PARPs is dependent on both context and timing [173, 174]. Metabolically, PARP activation consumes NAD+ and perturbs redox and energy flux. Live-cell imaging shows that PARP1-dependent NAD+ depletion drives a metabolic shift after DNA damage, illustrating how PARP signaling can reprogram bioenergetics in both stressed cancer and virus-infected cells [69]. Beyond PARP1, PARP14 directly promotes aerobic glycolysis (the Warburg effect) in hepatocellular carcinoma by modulating PKM2, linking ADP-ribosylation to pro-tumor glycolytic bias [175]. Viral infections likewise press on the same axis: coronaviruses induce MARylation of PARPs and deplete cellular NAD+, reconfiguring host metabolism and interferon responses in ways that can be countered by boosting NAD+ salvage [176]. Together, these data place PARPs at a shared mechanistic frontier of oncology and viral disease—coordinating DDR-derived innate sensing, immune evasion, and metabolic rewiring—and motivate the development of therapeutic strategies that co-target PARP/cGAS-STING and NAD+ metabolism in virus-associated cancers.

Mechanisms of PARPs in other diseases

PARPs in neurodegenerative diseases, cardiovascular diseases and fibrosis

Members of the PARPs family are, via distinct mechanistic effects, intimately involved in a variety of pathological processes. In neurodegenerative diseases, overactivation of PARP1 mediates neuroinflammation and triggers energy-exhausting cell death, and abnormal PARylation accelerates toxic protein aggregation [177]. The upregulation of PARP2/3 plays a dual role between DNA repair and apoptosis risk [178]. PARP14 affects the inflammatory response by regulating the M1/M2 polarization of microglia [179]. PARP14 inhibitors can promote the polarization of microglia to the anti-inflammatory M2 type, thereby reducing inflammation and improving cognitive function in Alzheimer's disease [179]. In cardiovascular diseases, PARP1 is a core regulatory factor, which drives atherosclerosis and vascular remodeling by regulating the phenotypic transformation of vascular smooth muscle cells, and leads to myocardial hypertrophy and fibrosis by promoting inflammatory response [180, 181]. In fibrotic diseases, PARP5A and PARP5B promote pathogenic fibroblast activation and tissue remodeling by overactivation of the Wnt/β-catenin signaling pathway [182, 183], while PARP1 also promotes fibrosis by regulating inflammatory signaling pathways, such as HMGB1-NF-κB [184].

PARPs in inflammatory and autoimmune disorders

PARPs also play critical and diverse roles in inflammatory and autoimmune disorders through regulation of key immune processes. PARP1 promotes the production of inflammatory factors by regulating signaling pathways, such as NF-κB, while the activation of the NLRP3 inflammasome directly leads to the maturation and release of inflammatory factors, which together aggravates intestinal inflammatory damage [185]. PARP14 promotes Th2 cell differentiation and IgE production in inflammation-related diseases, such as asthma and allergy, by regulating Stat6-dependent gene activation, thereby promoting disease progression [186]. The proinflammatory cytokines TNF-α and IFN-γ exacerbate Clostridioides difficile infection (CDI)-triggered apoptosis in enteric glial cells (EGCs) by increasing caspase-3/7/9 and PARP activation [187]. In autoimmune diseases, the balance between Th17 and regulatory T (Treg) cells is critical for disease pathogenesis. PARP1 and PARP14 play important roles in modulating this balance. Studies have demonstrated that PARP1 activation promotes Th17 cell differentiation while simultaneously suppressing Treg cell generation [188]. In contrast, PARP14 activation enhances Treg cell function and inhibits Th17 cell activity [28]. This bidirectional regulatory mechanism underscores the importance of PARP family members in immune modulation. Additionally, in experimental autoimmune encephalomyelitis (EAE), PARP2 regulates the development of neuroinflammation and neurological dysfunction by affecting the infiltration of proinflammatory Th1 and Th17 T helper lymphocytes in the central nervous system [189]. Consequently, therapeutic strategies targeting PARPs may offer promising possibilities for restoring Th17/Treg homeostasis, thereby providing potential novel interventions for the management of autoimmune disorders.

PARPs in metabolic diseases

PARPs play key roles in metabolic diseases and their mechanism of action involves a plethora of effects. For example, in fatty acid metabolism, PARPs affect the metabolic processing of fatty acids by regulating the expression of genes related to fatty acid synthesis and oxidation. Changes in the enzymatic activities of PARP1 and PARP2 can lead to increased fatty acid synthesis or decreased oxidation, which in turn affects fat accumulation in adipose tissue and the liver [190, 191]. In terms of cholesterol and lipoprotein metabolism, PARPs regulate key processes in cholesterol synthesis, transport, and lipoprotein metabolism. Deletion or inhibition of PARP2 can lead to increased cholesterol synthesis [192], while inhibition of PARP1 can improve cholesterol transport and lipoprotein metabolism [193, 194]. In addition, PARPs also affect the progression of metabolic diseases by regulating pathways such as the inflammatory response, oxidative stress, and cellular energy metabolism. For example, in non-alcoholic fatty liver disease (NAFLD) and alcoholic fatty liver disease (AFLD), activation of PARP1 and PARP2 leads to increased fat accumulation in hepatocytes [195, 196], while PARP inhibition attenuates these pathological changes [190, 195]. In obesity and type 2 diabetes, PARPs affect insulin sensitivity and energy metabolism by regulating metabolic processes in adipose tissue and muscle. Activation of PARP1 can inhibit lipolysis in adipose tissue and fatty acid oxidation in muscle, thus exacerbating insulin resistance [190, 191]. In atherosclerosis, the activation of PARP1 can promote the infiltration of inflammatory cells and the formation of foam cells, exacerbating the progression of atherosclerosis [194, 197].

Mechanisms of action of PARPi

Therapeutic targets and pharmacological properties of PARPi

PARPi are a class of small-molecule agents rationally designed to interfere with the catalytic function of PARP enzymes, particularly PARP1 and PARP2, which play central roles in SSB repair. Structurally, most clinically developed PARPi mimic nicotinamide, the natural substrate of PARP, and competitively bind to the conserved nicotinamide-binding pocket within the catalytic domain. This binding prevents the transfer of ADP-ribose units from NAD+ to target proteins, thereby blocking the formation of PAR chains that are essential for recruiting DNA repair factors [198]. In addition to catalytic inhibition, many PARPi also exert a so-called “PARP-trapping” effect, whereby the drug stabilizes PARP-DNA complexes, physically obstructing DNA replication and transcription and creating cytotoxic lesions [9]. This dual mechanism of action—catalytic inhibition plus PARP trapping—underpins their ability to induce synthetic lethality in tumors with HRR deficiencies, such as those harboring BRCA1/2 mutations [199].

Several PARPi, such as olaparib, rucaparib, niraparib, talazoparib, pamiparib, fluzoparib, and senaparib, have been developed and approved for clinical use. While all share the common strategy of targeting the catalytic NAD+ binding site, they differ in binding affinity, PARP isoform selectivity, trapping potency, and pharmacokinetic properties. For instance, olaparib, the first FDA-approved agent in this class, is indicated for BRCA-mutated ovarian and breast carcinomas [200]. Niraparib is employed as maintenance therapy in recurrent ovarian cancer, irrespective of the BRCA mutational status, while rucaparib is effective in tumors with HR deficiency [201]. Talazoparib demonstrates superior PARP-trapping activity, whereas veliparib shows favorable synergy with DNA-damaging chemotherapy [202].

Recent generations of PARPi, such as fluzoparib [203], pamiparib [204], and senaparib [205], exhibit enhanced PARP1 selectivity, stronger antitumor efficacy, and reduced off-target effects. These advances have not only expanded the therapeutic landscape for hard-to-treat cancers but also refined the paradigm of precision oncology by enabling more durable and tolerable treatment options for patients.

Mechanisms of PARPi in normal cells

Normal cells typically rely on HRR to repair DSBs. However, if the HRR function is impaired, PARPi may lead to genomic instability and cell death [206, 207]. Furthermore, PARPi can block excessive PARP activation, thereby preventing NAD+ depletion [208210], and mitigating inflammatory cytokine storms (Fig. 5) [209, 211, 212]. Additionally, PARPi interfere with a mitochondrial protective pathway. PARP can activate the nuclear ATM-NEMO complex, which then translocates to the cytoplasm and attaches to the outer mitochondrial membrane. Akt is activated and protects mitochondria from reactive oxygen species (ROS)-induced damage, allowing cells to survive (Fig. 5) [213, 214]. PARPi inhibit this process and mitochondria then continue to be induced and damaged by ROS. At the same time, PARPi leave ATF4 unPARylated and, thus, able to bind to the promoter of the MKP-1 encoding DNA region. The resulting transcription of MKP-1 mRNA followed by translation of the MKP-1 protein leads to the inactivation of JNK and p38 MAPKs (Fig. 5) [215]. Collectively, these mechanisms suggest that the role of PARPi in normal cells is a double-edged sword. Therefore, it is necessary to weigh their efficacy and side effects in specific clinical applications.

Fig. 5.

Fig. 5

Mechanisms of action of PARPi in normal cells. In normal cells, PARPi inhibit DSB repair, leading to cell death, and also inhibit the ATM-NEMO complex activation and translocation, whereby mitochondria will continue to be induced and damaged by ROS. At the same time, PARPi can block excessive PARP activation, NAD+ depletion, and inactivation of JNK and p38 MAPK

Mechanisms of PARPi in tumor cells

In tumor cells, when SSBs occur, PARP will bind to the DNA damage site and use NAD+ to generate PAR chains, recruiting DNA repair effector proteins, such as XRCC1, to complete the repair of SSBs (Fig. 6). PARPi will prevent the PARylation process by inhibiting the catalytic activity of PARPs, resulting in the incomplete repair of SSBs and DSBs and leading to cell death [216219]. Beyond catalytic inhibition, accumulating evidence demonstrates that PARP inhibitors also act through PARP trapping, whereby PARP1/2 are stably retained on DNA as toxic PARP–DNA complexes [119]. The formation of these stabilized complexes induces replication fork stalling and subsequent collapse, thereby aborting replication fork restart. This effect is particularly pronounced in HR-deficient backgrounds, where compromised fork protection renders stalled forks vulnerable to Mre11-mediated degradation of nascent DNA [140]. By concurrently obstructing DNA repair and exacerbating replication stress, PARP inhibitors achieve potent and tumor-selective lethality.

Fig. 6.

Fig. 6

Mechanisms of action of PARPi in tumor cells. In tumor cells, PARPi inhibit PARP to modify itself and other proteins through PARylation, inhibit its recruitment of repair factors (XRCC1 DNA ligase III and Pol-β), and inhibit DNA SSBs and DSBs repair. PARPi inhibit PARP trapping, where PARP1/2 remain stably bound to DNA as toxic PARP–DNA complexes. PARPi activate the cGAS/STING signaling pathway, promotes the transcription of IFNβ and the secretion of CXCL10 and CCL5, recruiting CD8+ T cells, and enhancing the immune response in the tumor microenvironment. Meanwhile, PARPi promote type I interferon signaling through TBK1–IRF3 activation, enhancing anti-tumor immunity. Potential impact of PARPi on endoplasmic reticulum function may involve functional interplay with H6PDH. In addition, both WEE1i and ATRi can cooperate with PARPi to enhance DNA damage effects and further inhibit cell cycle progression

In addition, PARPi lead to the accumulation of DNA damage, thereby activating intracellular cGAS, which catalyzes the production of cyclic GMP-AMP (cGAMP). cGAMP binds to and activates STING. The activation of STING further activates TANK-binding kinase 1 (TBK1), which then phosphorylates IRF3, allowing IRF3 to dimerize and transfer into the nucleus and promote the transcription of IFN-β (Fig. 6). The PARP7 inhibitor (RBN-2397) restores type I interferon signaling in tumor cells via the TBK1-IRF3 pathway, enhancing antitumor immunity and leading to tumor regression [101]. Meanwhile, PARPi activate the STING/TBK1/IRF3 pathway, resulting in increased expression of chemokines, such as the potent chemoattractants CXCL10 and CCL5, which target CD8+ T cells and enhance the immune response in the tumor microenvironment (Fig. 6) [220, 221]. In addition to directly inhibiting PARPs, PARPi may also affect cellular metabolism, thereby playing a broader role in cancer treatment [222]. For example, rucaparib is capable of binding to the NADP+-dependent hexose-6-phosphate dehydrogenase (H6PDH) in the endoplasmic reticulum, which is closely associated with redox regulation of this organelle and glucocorticoid metabolism [223]. Furthermore, PARPi can act synergistically with ATR and WEE1 inhibitors to suppress tumor progression. PARPi increase replication stress and promote the accumulation of DNA lesions, whereas ATR inhibition disables the S-phase cell cycle checkpoint and replication-fork protection, ultimately driving catastrophic fork collapse [224]. Critically, the synergy with WEE1 inhibition is a direct consequence of its location downstream of ATR. WEE1 blockade therefore bypasses the upstream ATR/CHK1 signal and forces a dysregulated cell cycle state [225]. This results in CDK1 hyperactivation, which drives excessive origin firing, nucleotide depletion, and premature mitotic entry. These events catastrophically amplify the replication stress and DNA damage initiated by PARPi [226]. These findings highlight that PARPi not only disrupt DNA repair and activate innate immune signaling but also modulate replication stress responses and cellular metabolism, thereby broadening their therapeutic potential in cancer treatment.

Clinical applications of PARPi

Clinical applications of PARPi as anti-tumor agents

In ovarian carcinoma, PARPi have redefined the standard of care by serving as maintenance therapy following platinum-based chemotherapy (Table 2). Olaparib was the first to demonstrate significant progression-free survival (PFS) benefit in BRCA-mutated patients [244]. For patients with newly-diagnosed advanced ovarian cancer (OC) and HR deficiency (HRD)-positive tumors, the combination of olaparib and bevacizumab has demonstrable efficacy [245, 246]. Niraparib has extended benefits in advanced ovarian cancer, HRD-positive and even HR-proficient populations due to its potent PARP trapping mechanism [119, 235, 236]. Rucaparib, with additional PARP3 inhibition, has also shown efficacy in platinum-sensitive recurrent ovarian cancer [247]. In breast cancer, olaparib and talazoparib are approved for germline BRCA-mutated, HER2-negative advanced disease, with talazoparib demonstrating superior PARP trapping and more robust tumor regression responses than olaparib [248]. Currently veliparib plus carboplatin is mainly employed to inhibit BRCA mutation-related breast cancer and triple-negative breast cancer [202].

Table 2.

Generational comparison of PARPi in anti-tumor therapy

Generation PARPi Mechanisms Approved/Investigated Indications Clinical trial number Common Adverse Effect References
First-generation Olaparib PARP1/2 inhibition, moderate PARP trapping, synthetic lethality in HR-deficient cells

Approved in Food and Drug Administration (FDA), European Medicines Agency (EMA) and China's National Medical Products Administration (NMPA): BRCA-mutated ovarian cancer, HER2 negative breast cancer, pancreatic, and prostate cancers, peritoneal cancer, endometrial cancer;

Phase III trials: Colorectal cancer, non-small cell lung cancer, small cell lung cancer, squamous cell cancer;

Phase II trials: Bladder cancer, cervical cancer, gastric cancer, glioblastoma, head and neck cancer, HER2 positive breast cancer, osteosarcoma

NCT03402841; NCT03286842; NCT02184195; NCT02032823, NCT03278717, NCT03106987,NCT02392676, NCT05171816, NCT03534453, NCT01924533, NCT01844986, NCT02000622, NCT06712472, NCT03976323, NCT03740165, NCT04456699,NCT04624204,NCT05432791, etc Nausea, fatigue, anemia [227230]
Rucaparib PARP1/2/3 inhibition, blocks DNA repair and induces cytotoxicity in HR-deficient cells

Approved in FDA, EMA and NMPA: BRCA-mutated ovarian cancer and BRCA-mutated metastatic castration-resistant prostate cancer;

Phase II trials: Breast cancer, cervical cancer, endometrial cancer, gastric cancer, malignant-mesothelioma, pancreatic cancer, small cell lung cancer, solid tumors, triple negative breast cancer, urogenital cancer;

Phase I/II trials: Non-small cell lung cancer

NCT02855944, NCT04227522, NCT01968213, NCT02975934, NCT03522246, NCT04455750, etc Fatigue, nausea, anemia, Alanine Aminotransferase (ALT)/Aspartate Aminotransferase (AST) elevation, dysgeusia [227, 230234]
Niraparib Potent PARP trapping, PARP1/2 inhibition, active in some HR-proficient tumors

Approved in FDA, EMA and NMPA: Advanced ovarian cancer, fallopian tube cancer, peritoneal cancer;

Phase III trials: Breast cancer, HER2 negative breast cancer, non-small cell lung cancer, prostate cancer, small cell lung cancer;

Phase II trials: Cervical cancer, cholangiocarcinoma, CNS cancer, endometrial cancer, glioblastoma; mesothelioma; neuroendocrine tumors, pancreatic cancer; renal cell carcinoma, solid tumors, squamous cell cancer, triple negative breast cancer, urogenital cancer, uveal melanoma

NCT01847274, NCT06388733, NCT04915755, NCT03651206, NCT02655016, NCT04475939, NCT01905592, NCT04497844, NCT03602859, NCT03981796, NCT03748641, NCT06915025, NCT05615818, NCT03516084, etc Thrombocytopenia, anemia, neutropenia, fatigue, hypertension [230, 235238]
Talazoparib Strongest PARP trapping, replication fork collapse and apoptosis

Approved in FDA, EMA and NMPA: BRCA-mutated breast cancer, HRR Gene–Altered metastatic castration-resistant prostate cancer (mCRPC);

Phase III trials: Ovarian cancer;

Phase II trials: Endometrial cancer, fallopian tube cancer, peritoneal cancer, solid tumors, triple negative breast cancer

NCT01945775, NCT03395197, NCT04821622, NCT03642132, etc Anemia, neutropenia, thrombocytopenia, alopecia [77, 227, 230, 239241]
Veliparib PARP1/2 inhibition with mininal PARP trapping, sensitizes cells to DNA-damaging agents

Phase III trials: Fallopian tube cancer, HER2 negative breast cancer, non-small cell lung cancer, ovarian cancer, peritoneal cancer, triple negative breast cancer, Combination with carboplatin in triple-negative breast cancer (TNBC), non-small cell lung cancer (NSCLC);

Phase II/III trials: Glioblastoma;

Phase II trials: Brain metastases, breast cancer, colorectal cancer, germ cell and embryonal neoplasms, germ cell cancer, malignant melanoma, pancreatic cancer, rectal cancer, solid tumors

NCT02163694, NCT02470585, NCT02106546, NCT02264990, NCT02032277, NCT02152982, etc Myelosuppression (notable with combination), nausea, fatigue [202, 230, 242]
Next-generation Pamiparib High PARP1/2 selectivity, blood–brain barrier penetration

Approved in NMPA: BRCA-mutated recurrent ovarian cancer, fallopian tube cancer, peritoneal cancer;

Phase II trials: Gastric cancer, HER2 negative breast cancer;

Phase I/II trials: Glioblastoma, glioma, solid tumors

NCT03519230, NCT04164199, etc Anemia, leukopenia, generally well tolerated in early studies [204, 230]
Fuzuloparib PARP1/2 selectivity

Approved in NMPA: Platinum-sensitive recurrent ovarian cancer, fallopian tube cancer, peritoneal cancer;

Phase III trials: Pancreatic cancer, prostate cancer; Phase II trials: Glioblastoma

NCT04691804, NCT06188455, NCT06533384, etc Anemia, leukopenia, thrombocytopenia [203, 230]
Senaparib Highly selective PARP1/2 inhibition with reduced off-target toxicity

Approved in NMPA: First-line maintenance in advanced ovarian cancer;

Phase II trials: Small cell lung cancer, solid tumors;

Phase I/II trials: Prostate cancer

NCT04822961, NCT07120451, NCT06617923, NCT04434482, NCT04089189, NCT05269316, etc Nausea, fatigue, anemia, leukopenia [205, 230, 243]

In prostate cancer, PARPi have demonstrated efficacy in metastatic, castration-resistant prostate cancer (mCRPC) harboring BRCA1/2 or other HR response gene alterations (Table 2). Olaparib gained approval following the PROfound trial [249], while combination strategies such as niraparib plus abiraterone have emerged to overcome resistance to androgen receptor signaling inhibitors [250]. In pancreatic cancer, a disease currently with limited treatment options, olaparib has shown benefit as maintenance therapy in germline BRCA-mutated patients [251], exploiting the same HR response vulnerabilities.

Recently developed agents, such as pamiparib, have gained approval for BRCA-mutated recurrent advanced ovarian, fallopian tube or primary peritoneal cancer [204]. Fuzuloparib has also been approved for platinum-sensitive ovarian cancers [203] (Table 2). Pamiparib offers high PARP1/2 selectivity and good blood–brain barrier permeability. Senaparib [205, 243], approved in 2025 as first-line maintenance for advanced ovarian cancer, represents a new generation of PARPi with reduced hematologic toxicity and improved tissue selectivity.

Collectively, PARPi have established themselves as a cornerstone of targeted therapy in precision oncology, driving the evolution of biomarker-driven, personalized treatment strategies and offering hope for durable disease control in genomically unstable tumors (Table 2).

Applications of PARPi in antiviral therapy

Beyond oncology, PARPi have also emerged as a promising therapeutic strategy for antiviral treatment, targeting both DNA and RNA viruses (Table 3), exemplified by HBV and coronaviruses [252, 253], respectively, and other clinically-significant viruses, such as high-risk human papillomavirus (HPV)-associated malignancies in the endocervix, head, and neck [263, 264] and pathogens with pandemic potential, including IAV [259]. Recent studies have demonstrated that the combination of PARPi with CRISPR/Cas9 technology can significantly enhance the antiviral efficacy against HBV by directly targeting covalently closed circular DNA (cccDNA). For instance, the inhibition of NHEJ pathways through PARPi has been shown to augment the antiviral effects of HBV-CRISPR, leading to marked reductions in both cccDNA and pregenomic RNA levels in infected hepatocytes [252]. Furthermore, stenoparib has demonstrated dose-dependent antiviral activity against SARS-CoV-2 and the seasonal human alphacoronavirus HCoV-NL63, through a dual mechanism of action by interfering with both viral entry and replication processes [253]. Additionally, the combination of PARPi with favipiravir for the treatment of SARS-CoV-2 has demonstrated potential synergistic effects that could enhance the overall treatment efficacy [254, 255]. Olaparib ameliorates IAV-induced pneumonia in mice by inhibiting the PARP1 activity, modulating inflammatory cytokines, reducing leukocyte infiltration, and inhibiting the NF-κB signaling pathway [259]. In HPV-associated tumors, such as cervical cancer and head and neck squamous cell carcinoma (HNSCC), persistent virus infection leads to damage to the host's DNA repair mechanisms, making tumor cells particularly sensitive to PARPi. Preclinical and early clinical studies have shown that PARPi exert antitumor effects also by enhancing the sensitivity of DNA damage treatments, such as radiotherapy and cisplatin, thereby improving treatment response rates and delaying tumor progression [264267]. The clinical application of PARPi in other viral infections, including KSHV [258], and HIV [256], is also underway, with preliminary data suggesting that they may inhibit viral replication effectively and potentially heralding the development of broad-spectrum antivirals.

Table 3.

Clinically significant viruses and PARPi applications

Virus PARPi Mechanism of Action Experimental Model/Stage Common Adverse Effect Reference
HBV Olaparib Inhibits NHEJ, enhancing CRISPR-mediated cccDNA cleavage HepG2-hNTCP-C4 cells, primary human hepatocytes (PHHs)/Preclinical No human safety data [252]
SARS-CoV-2/HCoV-NL63 Human Coronaviruses Stenoparib, CVL218 (mefuparib) Dual inhibition of viral entry and replication, synergistic with remdesivir or favipiravir Vero E6, Calu-3 cells/Preclinical No human safety data [253255]
HIV Talazoparib, 3-aminobenzamide, Olaparib, Talazoparib, 5-aminoisoquinolinone, PJ34 Enhances NK cell-mediated cytotoxicity, combination with the HDAC inhibitor vorinostat, inhibits HIV-1 LTR activation through NFκB suppression, disrupts actin cytoskeleton rearrangements via Rho GTPase inhibition J-Lat cell lines, human monocyte-derived macrophages (MDM)/Preclinical No human safety data [256, 257]
KSHV AZD2461 Induces DNA damage, reduces cell proliferation, combination with the CHK1 inhibitor UCN-01 Latency models (PEL cells)/Preclinical No human safety data [258]
IAV Olaparib Modulating inflammatory cytokines, reducing leukocyte infiltration, and suppressing the NF-κB signaling pathway Mice model No human safety data [259]
HPV/HBV related cancers

Cervical cancer: Veliparib, Olaparib;

HNSCC: Olaparib, Niraparib, Veliparib, Talazoparib, Rucaparib, CEP-9722

Cervical cancer: Enhance DNA damage, combination with the ATR inhibitor AZD6738;

HNSCC: Enhance DNA damage, boost immune responses by inducing PD-L1 expression, combination with cisplatin

Cervical cancer: Phase I/II trials;

HNSCC: Phase I-III trials

Anemia, neutropenia, thrombocytopenia, dyspnea, leukopenia [260262]

Notably, viruses from the class Bunyaviricetes, such as Crimean–Congo hemorrhagic fever virus (CCHFV) [268] and Rift Valley fever virus (RVFV) [269], have been shown to influence PARP cleavage during infection. While no direct studies have yet tested PARPi against bunyaviruses, these observations highlight the involvement of PARPs in bunyavirus-host interactions. Thus, bunyaviruses may serve as illustrative examples pointing to the broader possibility that PARPi could have therapeutic relevance in additional viral contexts beyond those currently reported.

Applications of PARPi in other diseases

It is increasingly evident that PARPi offer therapeutic potential beyond oncology and viral infections, particularly in non-malignant diseases driven by DNA damage, oxidative stress, and inflammatory signaling.

In neurodegenerative diseases, particularly Parkinson's disease, PJ34, a selective PARP1 inhibitor, has shown significant neuroprotective effects in experimental traumatic brain injury (TBI) by reducing neuronal death and microglial activation [26]. The PARP inhibitors 3-aminobenzamide (3-AB) or 5-aminoisoquinolinone (5-AIQ) reduce excessive PAR formation in the process of DNA damage repair by inhibiting the activity of PARP enzymes, thus alleviating the inflammatory response, apoptosis, and tissue damage, and, thus, have potential therapeutic effects in varied clinical entities, such as spinal cord injuries and stroke [27]. PARP14 inhibitors (such as RBN-3143) have shown therapeutic potential for atopic dermatitis (AD) in preclinical studies and can reduce the level of inflammatory factors, including Th17 cytokines [28]. In cardiovascular diseases, PARPi, such as 3-AB, PJ-34, INO-1001, BGP-15, and GPI 6150, reduce the consumption of NAD+ and ATP by inhibiting the overactivation of PARP1, thereby alleviating oxidative stress and inflammatory responses and protecting cells from necrosis. Considering heart failure and myocardial hypertrophy, PJ-34, INO-1001, and L-2286 play a role by improving contractile function, reducing myocardial cell death, and decreasing myocardial hypertrophy and fibrosis. In circulatory failure, 3-AB, PJ-34, and INO-1001 improve survival by promoting myocardial contractility and vascular function, reducing inflammation and tissue damage. In diseases such as hypertension, cardiovascular aging, atherosclerosis and vascular remodeling, diabetic cardiovascular complications, and angiogenesis, PJ-34, INO-1001, and other inhibitors have shown therapeutic potential by improving endothelial function, reducing plaque size and promoting plaque stability, reducing endothelial dysfunction, improving contraction and vascular function, and reducing VEGF or bFGF-induced angiogenesis [29]. Lastly, in organ fibrosis, PARPi, particularly HYDAMTIQ, inhibit TGF-β signaling, reduce fibroblast activation, and attenuate lung fibrosis progression in mice [30]. Olaparib inhibits the development of liver fibrosis by inhibiting PARP1 activity, reducing inflammatory responses, apoptosis, activation of TGF-β/Smad signaling pathway and oxidative stress [30].

In inflammatory and autoimmune diseases, olaparib effectively reduces the number of monocytes in the blood of mice and exerts anti-inflammatory effects by increasing IL-10 levels and inhibiting the concentrations of IL-1β and IL-6 in Crohn's disease models. 3-AB improves rectal bleeding, reduces blood glucose levels, lowers serum IL-1β levels, decreases weight loss, and improves histological scores of colon tissue sections in a mouse model of colitis-related diabetes [270]. In addition, RBN-3143, as a PARP14 inhibitor, is being used in clinical trials to treat atopic dermatitis, which primarily relieves symptoms in patients with atopic dermatitis by controlling inflammation [271].

In metabolic disorders, olaparib can prevent and alleviate NAFLD and non-alcoholic steatohepatitis (NASH), and diminish liver lipid accumulation by regulating liver lipid metabolism-related signaling pathways. PJ34 attenuates hepatocyte lipid accumulation in a mouse model of NAFLD induced by a high-fat and high-sucrose diet. TIQ-A reduces plasma cholesterol levels and reduces plaque formation in high-cholesterol diet-induced atherosclerotic mice. G-007LK reduces serum triglycerides, non-esterified fatty acids and total cholesterol levels, and improves obesity-related metabolic disorders in db/db mice. INO-1001 reduces LDL-C levels, inhibits the expression of inflammatory factors, and reduces the development of atherosclerosis in a murine model [102].

Although no PARPi are currently clinically approved outside oncology, the growing body of mechanistic and preclinical evidence underscores the promising scope of PARPi repurposing in a wide range of chronic and degenerative diseases.

PARPi mediated resistance to clinical therapy

PARPi are increasingly recognized as a dynamic process that evolves with disease progression and treatment duration [217, 272]. In ovarian cancer, continuous circulating tumor DNA (ctDNA) monitoring found that approximately 28% of the patients late in treatment developed acquired mutations in HRR-related genes, resulting in a significant reduction in PARPi sensitivity [273]. Meanwhile, drug resistance to PARPi can also be generated through other mechanisms, such as KAT6A-mediated liquid–liquid phase separation (LLPS), which reduces the retention of PARP1 at DNA damage sites and where the inhibition of KAT6A partially restores sensitivity [274]. Similarly, the efficiency of PARP1 trapping at DNA lesions, a key mechanism of PARPi cytotoxicity, is itself regulated by time-dependent, drug–target binding kinetics. Variations in the dissociation rate constant influences the persistence of PARP1–drug complexes and ultimately the depth of DNA repair inhibition [275]. It is worth noting that temporal changes in drug resistance are also seen in virus-associated cancers. In EBV-positive gastric cancer, EBNA1 enhances the sensitivity to olaparib by inhibiting the ATR kinase activity, however, may lead to drug resistance through the p38 MAPK pathway activation or ATR compensatory activation after long-term treatment [276]. Moreover, the heterogeneity within CDK12-mutated prostate cancers demonstrates that distinct mutational patterns drive differential therapeutic outcomes, with truncating mutations sustaining PARPi sensitivity and kinase-domain mutations leading to an intermediate reactive drug resistance [277]. Together, these studies reveal that genomic evolution, drug binding dynamics, and chromatin remodeling all contribute to temporally-regulated resistance, and highlight the stage specificity of drug resistance mechanisms, and the need for timely adjustment of treatment strategies by dynamic monitoring.

Given these temporally and contextually-regulated resistance mechanisms, monotherapy with PARPi is insufficient to ensure durable clinical responses [278]. Instead, combinatorial strategies appear essential to counteract both genomic adaptations and microenvironmental suppression [200, 279]. Preclinical studies have shown that STING agonists can reprogram M2-like TAMs toward antitumor M1 phenotypes and synergize with PARPi to restore immune activation [280], while immune checkpoint blockade targeting PD-1/PD-L1 can resensitize resistant tumors to T cell killing [281]. Rational drug combinations with DDR-targeting agents, such as ATR or WEE1 inhibitors, may further prevent the rapid outgrowth of resistant clones in genetically-heterogeneous tumors [282]. For moderately sensitive patients, such as those harboring CDK12 kinase domain mutations, the combination strategy of PARPi and other synergistic drugs can be considered in the early stage of treatment to avoid rapid screening of drug-resistant clones under the pressure of single agents. For sensitive subtypes, such as truncation mutations, CDK12 mutation status and replication stress-related biomarkers can be dynamically monitored to adjust treatment regimens early in disease progression [277].

In summary, these findings indicate that the most promising strategy to circumvent PARPi resistance is the application of stage-tailored and context-specific combination therapies, underpinned by dynamic biomarker monitoring.

Potential side effects and safety concerns of PARPi

Although PARPi are generally well tolerated, their toxicity profiles are not uniform and are strongly shaped by the cell- and tissue-specific expression of PARP enzymes, drug-specific pharmacodynamics, and organ-level pharmacokinetics. Hematologic toxicities, including anemia, neutropenia, and thrombocytopenia, represent the most common adverse events and reflect the high sensitivity of the bone marrow compartment [9, 283]. Mechanistically, PARP1 is broadly expressed, with particularly high abundance in lymphoid and hematopoietic cells, and experimental models demonstrate that combined PARP1/2 deficiency disrupts B cell development [284]. This underpins the clinical observation that hematological toxicity is a hallmark adverse effect of PARPi treatment. Furthermore, the severity of hematologic adverse events correlates with the extent of PARP trapping (rather than catalytic inhibition), with talazoparib exhibiting the strongest trapping activity and the highest hematologic toxicity [119, 248]. Clinical trials of niraparib additionally reveal a host factor–dependent pattern: patients with lower body weight or reduced baseline platelet counts experience a higher incidence of thrombocytopenia. As a result, individualized starting doses based on weight and platelet counts have become a standard practice to mitigate severe hematologic events [285, 286]. Rare, but clinically significant, late-onset complications include myelodysplastic syndrome (MDS) and acute myeloid leukemia (AML), which occur at a cumulative incidence of ~ 0.7%, typically after prolonged exposure, and are thought to represent organ-specific (hematopoietic) toxicity [287].

Beyond the bone marrow, other organ systems exhibit distinct toxicity signatures. In the gastrointestinal tract, PARP1 plays a role in epithelial injury responses, and preclinical data suggest that PARPi protected intestinal barrier integrity [288]. Clinically, nausea and vomiting are the most frequently reported gastrointestinal adverse events, whereas diarrhea and constipation are less common [289]. In the cardiovascular system, niraparib uniquely inhibits norepinephrine and dopamine transporters, leading to hypertension and tachycardia, which necessitates close blood pressure monitoring during the early treatment phase [290]. In the liver, rucaparib is frequently associated with transient elevations of alanine aminotransferase (ALT)/aspartate aminotransferase (AST), typically within the first few cycles and without concurrent bilirubin elevation, indicating a largely reversible hepatocellular reaction [291]. Renal toxicities also show organ-specific patterns: olaparib inhibits proximal tubular transporters (OCT2 and MATE1/2-K), resulting in a benign increase in serum creatinine without true glomerular filtration rate (GFR) impairment. In such cases, cystatin-C is recommended for more accurate assessment of renal function [292, 293].

Emerging strategies to reduce cell- and tissue-specific toxicity include the development of selective PARP1 inhibition (e.g., saruparib). Preliminary data from preclinical studies suggest that such selective inhibitors may have a reduced impact on hematological parameters [294].

In summary, the heterogeneity of PARP expression across cell types and organ systems plays a central role in shaping the adverse event spectrum of PARPi therapy. Clinical practice has increasingly shifted toward organ-specific monitoring and individualized dosing strategies, including weight- and platelet-adjusted initiation, early and frequent laboratory surveillance (liver function and renal biomarkers), and cardiovascular monitoring for agents, such as niraparib. Awareness of these cell- and tissue-specific determinants enhances the clinical relevance of PARPi safety management and informs the rational design of next-generation inhibitors with improved tolerability.

Conclusions and prospects

In this review, we have provided a comprehensive overview of PARP biology, PARPi, and existing and potential clinical applications. We have introduced the mechanisms and the known biological functions of PARPs, outlining their classification, structural characteristics, and canonical functions in DNA repair, parthanatos, and cellular homeostasis, alongside their non-canonical roles in transcriptional regulation, immunity, and metabolism. We have outlined the roles of PARPs in disease pathogenesis, including their contribution to genomic instability, tumor cell death regulation, tumor microenvironment remodeling, and viral infections, involving both DNA and RNA viruses. Beyond cancer and viral biology, we also highlighted their functions in neurodegenerative, cardiovascular, fibrotic, autoimmune, inflammatory, and metabolic diseases. Finally, we have reviewed the mechanisms of action of PARPi in normal and tumor cells, their clinical applications in oncology, antiviral therapy, and other disease contexts, as well as the challenges of drug resistance, toxicity, and safety management. Collectively, these insights position PARPs as central molecular regulators and PARPi as versatile therapeutic agents with implications far beyond oncology.

Next-generation PARPi and novel modalities

While current PARPi (e.g., olaparib, niraparib, rucaparib, and talazoparib) have transformed the treatment of HRR-deficient cancers, their broader application is limited by off-target toxicities and insufficient selectivity [295]. The next wave of innovation will involve highly selective PARPi [296], dual-target compounds, and novel modalities, such as PROTAC-based degraders [297] or allosteric inhibitors. These strategies may enhance efficacy, reduce adverse effects, and expand indications to patients without classical BRCA mutations. Furthermore, the potential of targeting other PARP family members (e.g., PARP7, PARP9, and PARP14) in immune and viral contexts remains underexplored and warrants systematic investigation; however, rigorous preclinical and translational studies are required to ensure that improved pharmacologic properties can be translated into sustained clinical benefit.

Overcoming resistance: novel combination strategies

A major barrier to PARPi efficacy is the emergence of resistance, frequently through restoration of HRR, stabilization of replication forks, or drug efflux mechanisms [298300]. Rational combination strategies—such as PARPi with ATR/CHK1 inhibitors [301], epigenetic modulators [302, 303], or immune checkpoint blockade [304]—offer promising avenues to resensitize tumors and potentiate antitumor immunity. The challenge lies in identifying optimal dosing regimens, scheduling, and biomarker-guided patient selection to balance efficacy with tolerability. Addressing these questions will be critical for sustaining long-term responses in the clinic.

Biomarker development and personalized therapies

The clinical translation of PARPi is still constrained by the lack of robust and dynamic biomarkers. Current reliance on BRCA1/2 and HRR gene mutations fails to capture the full spectrum of responsive patients [305]. Future work should focus on integrating liquid biopsy-based markers (ctDNA, and exosomal RNA), functional HRR assays, and single-cell/spatial multiomics platforms to monitor treatment response and resistance evolution in real time. Embedding such diagnostics into clinical decision-making frameworks will enable a new era of precision medicine approaches and extend PARPi benefits beyond narrowly-defined patient populations.

Beyond DNA repair: expanding therapeutic horizons

Although PARPs are traditionally recognized for their DNA repair roles, they also participate in immune signaling, transcriptional regulation, metabolism, and stress responses [306, 307]. These broader functions suggest therapeutic potential in non-oncologic contexts, including neurodegeneration, fibrosis, cardiovascular disorders, and autoimmune diseases [179, 186, 195, 196, 306308]. The challenge is to delineate the isoform-specific and context-dependent roles of PARPs to avoid unintended consequences of long-term inhibition [309]. Moreover, immunomodulatory effects of PARPi could be leveraged in viral infections or combined with immunotherapies, thereby broadening their clinical reach [307].

Overall, while PARPi have already redefined cancer therapy, their full potential remains unrealized. Future efforts must resolve key gaps: (1) clarifying isoform-specific functions across disease contexts; (2) designing next-generation, safer PARPi; (3) establishing robust biomarkers for personalized treatment; (4) optimizing rational combination strategies; and (5) elucidating the shared mechanisms of PARPs in cancer, viral infections, and other diseases. Addressing these research questions will not only enhance therapeutic precision but also determine whether PARPi can evolve from oncology-focused agents into broad-spectrum modulators of human disease.

Acknowledgements

We would like to acknowledge the researchers whose relevant works are cited in this review and all co-authors for their support. We acknowledge the use of Adobe Illustrator for the preparation of figures in this manuscript.

Authors’ contributions

F.W. and Z.G. performed the literature review. F.W. wrote the manuscript. M.J.C. proofread the manuscript. W.S. conceived this review, supervised the study, and proofread the manuscript. All authors have read and approved the final manuscript.

Funding

The authors are grateful to the financial support from the grant from the National Natural Science Foundation of China (No. 32325003 to W.S.), and the grants to F.W. from the China Postdoctoral Science Foundation (No. 2024M762020), Postdoctoral Fellowship Program of CPSF (No. GZB20240442), and Shanghai Sailing Program (No. 24YF2726900).

Data availability

No new data were generated or analyzed in this study. This work is a review study and does not rely on any external data.

Declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

All authors state that there is no conflict of interests.

Footnotes

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Fei Wang and Zhuyi Guo contributed equally to this study.

References

  • 1.Chambon P, Weill JD, Mandel P. Nicotinamide mononucleotide activation of a new DNA-dependent polyadenylic acid synthesizing nuclear enzyme. Biochem Biophys Res Commun. 1963;11:39–43. 10.1016/0006-291x(63)90024-x. [DOI] [PubMed] [Google Scholar]
  • 2.Sims JL, Berger SJ, Berger NA. Poly(ADP-ribose) polymerase inhibitors preserve oxidized nicotinamide adenine dinucleotide and adenosine 5’-triphosphate pools in DNA-damaged cells: mechanism of stimulation of unscheduled DNA synthesis. Biochemistry. 1983;22:5188–94. 10.1021/bi00291a019. [DOI] [PubMed] [Google Scholar]
  • 3.Ali AAE, Timinszky G, Arribas-Bosacoma R, Kozlowski M, Hassa PO, Hassler M, et al. The zinc-finger domains of PARP1 cooperate to recognize DNA strand breaks. Nat Struct Mol Biol. 2012;19:685–92. 10.1038/nsmb.2335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Kurosaki T, Ushiro H, Mitsuuchi Y, Suzuki S, Matsuda M, Matsuda Y, et al. Primary structure of human poly(ADP-ribose) synthetase as deduced from cDNA sequence. J Biol Chem. 1987;262:15990–7. [PubMed] [Google Scholar]
  • 5.Curtin NJ, Szabo C. Poly(ADP-ribose) polymerase inhibition: past, present and future. Nat Rev Drug Discov. 2020;19:711–36. 10.1038/s41573-020-0076-6. [DOI] [PubMed] [Google Scholar]
  • 6.Langelier MF, Eisemann T, Riccio AA, Pascal JM. PARP family enzymes: regulation and catalysis of the poly(ADP-ribose) posttranslational modification. Curr Opin Struct Biol. 2018;53:187–98. 10.1016/j.sbi.2018.11.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Shou Q, Fu H, Huang X, Yang Y. PARP-1 controls NK cell recruitment to the site of viral infection. JCI Insight. 2019;4:e121291. 10.1172/jci.insight.121291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Hottiger MO, Hassa PO, Luscher B, Schuler H, Koch-Nolte F. Toward a unified nomenclature for mammalian ADP-ribosyltransferases. Trends Biochem Sci. 2010;35:208–19. 10.1016/j.tibs.2009.12.003. [DOI] [PubMed] [Google Scholar]
  • 9.Lord CJ, Ashworth A. PARP inhibitors: synthetic lethality in the clinic. Science. 2017;355:1152–8. 10.1126/science.aam7344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Bryant HE, Schultz N, Thomas HD, Parker KM, Flower D, Lopez E, et al. Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase. Nature. 2005;434:913–7. 10.1038/nature03443. [DOI] [PubMed] [Google Scholar]
  • 11.Farmer H, McCabe N, Lord CJ, Tutt AN, Johnson DA, Richardson TB, et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature. 2005;434:917–21. 10.1038/nature03445. [DOI] [PubMed] [Google Scholar]
  • 12.Krastev DB, Wicks AJ, Lord CJ. PARP inhibitors – trapped in a toxic love affair. Cancer Res. 2021;81:5605–7. 10.1158/0008-5472.CAN-21-3201. [DOI] [PubMed] [Google Scholar]
  • 13.Langelier MF, Planck JL, Roy S, Pascal JM. Structural basis for DNA damage-dependent poly(ADP-ribosyl)ation by human PARP-1. Science. 2012;336:728–32. 10.1126/science.1216338. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Eustermann S, Wu WF, Langelier MF, Yang JC, Easton LE, Riccio AA, et al. Structural Basis of Detection and Signaling of DNA Single-Strand Breaks by Human PARP-1. Mol Cell. 2015; 60:742–754.10.1016/j.molcel.2015.10.032. [DOI] [PMC free article] [PubMed]
  • 15.Kim MK. Novel insight into the function of tankyrase. Oncol Lett. 2018; 16:6895–6902.10.3892/ol.2018.9551. [DOI] [PMC free article] [PubMed]
  • 16.Vyas S, Chesarone-Cataldo M, Todorova T, Huang YH, Chang P. A systematic analysis of the PARP protein family identifies new functions critical for cell physiology. Nat Commun. 2013;4:2240. 10.1038/ncomms3240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Lei P, Li W, Luo J, Xu N, Wang Y, Xie D, et al. PARP (Poly ADP-ribose Polymerase) Family in Health and Disease. MedComm (2020). 2025; 6:e70314.10.1002/mco2.70314. [DOI] [PMC free article] [PubMed]
  • 18.Bohi F, Hottiger MO. Expanding the Perspective on PARP1 and Its Inhibitors in Cancer Therapy: From DNA Damage Repair to Immunomodulation. Biomedicines. 2024; 12.10.3390/biomedicines12071617. [DOI] [PMC free article] [PubMed]
  • 19.Ray A, Opyrchal M. Targeting PARP1: a promising approach for next-generation poly (ADP-ribose) polymerase inhibitors. Curr Breast Cancer Rep. 2025;17:22. 10.1007/s12609-025-00582-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Chen L, Zou Y, Sun R, Huang M, Zhu X, Tang X, et al. Minimizing DNA trapping while maintaining activity inhibition via selective PARP1 degrader. Cell Death Dis. 2024;15:898. 10.1038/s41419-024-07277-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Wu Y, Wu M, Zheng X, Yu H, Mao X, Jin Y, et al. Discovery of a potent and selective PARP1 degrader promoting cell cycle arrest via intercepting CDC25C-CDK1 axis for treating triple-negative breast cancer. Bioorg Chem. 2024;142:106952. 10.1016/j.bioorg.2023.106952. [DOI] [PubMed] [Google Scholar]
  • 22.Li Y, Wu Y, Zheng X, Cong J, Liu Y, Li J, et al. Cytoplasm-translocated Ku70/80 complex sensing of HBV DNA induces hepatitis-associated chemokine secretion. Front Immunol. 2016;7:569. 10.3389/fimmu.2016.00569. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Yu D, Liu R, Yang G, Zhou Q. The PARP1-Siah1 axis controls HIV-1 transcription and expression of Siah1 substrates. Cell Rep. 2018;23(13):3741–9. 10.1016/j.celrep.2018.05.084. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Xia C, Wolf JJ, Sun C, Xu M, Studstill CJ, Chen J, et al. PARP1 Enhances Influenza A Virus Propagation by Facilitating Degradation of Host Type I Interferon Receptor. J Virol. 2020; 94.10.1128/JVI.01572-19. [DOI] [PMC free article] [PubMed]
  • 25.Papp H, Toth E, Bovari-Biri J, Banfai K, Juhasz P, Mahdi M, et al. The PARP inhibitor rucaparib blocks SARS-CoV-2 virus binding to cells and the immune reaction in models of COVID-19. Br J Pharmacol. 2024;181:4782–803. 10.1111/bph.17305. [DOI] [PubMed] [Google Scholar]
  • 26.Stoica BA, Loane DJ, Zhao Z, Kabadi SV, Hanscom M, Byrnes KR, et al. PARP-1 inhibition attenuates neuronal loss, microglia activation and neurological deficits after traumatic brain injury. J Neurotrauma. 2014;31:758–72. 10.1089/neu.2013.3194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Genovese T, Mazzon E, Muia C, Patel NS, Threadgill MD, Bramanti P, et al. Inhibitors of poly(ADP-ribose) polymerase modulate signal transduction pathways and secondary damage in experimental spinal cord trauma. J Pharmacol Exp Ther. 2005;312:449–57. 10.1124/jpet.104.076711. [DOI] [PubMed] [Google Scholar]
  • 28.Wu S, Zeng X, Liu J, Cong K, Lou S, Li Z, et al. Discovery and Optimization of Potent and Highly Selective PARP14 Inhibitors for the Treatment of Atopic Dermatitis. J Med Chem. 2025; 68:9755–9776.10.1021/acs.jmedchem.5c00564. [DOI] [PubMed]
  • 29.Pacher P, Szabo C. Role of poly(ADP-ribose) polymerase 1 (PARP-1) in cardiovascular diseases: the therapeutic potential of PARP inhibitors. Cardiovasc Drug Rev. 2007;25:235–60. 10.1111/j.1527-3466.2007.00018.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Rao PD, Sankrityayan H, Srivastava A, Kulkarni YA, Mulay SR, Gaikwad AB. ’PARP’ing fibrosis: repurposing poly (ADP ribose) polymerase (PARP) inhibitors. Drug Discov Today. 2020;25:1253–61. 10.1016/j.drudis.2020.04.019. [DOI] [PubMed] [Google Scholar]
  • 31.Zheng J, Deng Y, Fang C, Xiong S, Zhu X, Wu W, et al. Comprehensive dataset of interactors for the entire PARP family using TurboID proximity labeling. Sci Data. 2025;12:405. 10.1038/s41597-025-04722-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Pandey N, Black BE. Rapid detection and signaling of DNA damage by PARP-1. Trends Biochem Sci. 2021;46:744–57. 10.1016/j.tibs.2021.01.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Pascal JM. The comings and goings of PARP-1 in response to DNA damage. DNA Repair (Amst). 2018;71:177–82. 10.1016/j.dnarep.2018.08.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Mariotti L, Templeton CM, Ranes M, Paracuellos P, Cronin N, Beuron F, et al. Tankyrase Requires SAM Domain-Dependent Polymerization to Support Wnt-β-Catenin Signaling. Mol Cell. 2016;63:498–513. 10.1016/j.molcel.2016.06.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Du Q, Miao Y, He W, Zheng H. ADP-Ribosylation in Antiviral Innate Immune Response. Pathogens. 2023;12.10.3390/pathogens12020303. [DOI] [PMC free article] [PubMed]
  • 36.Dhoonmoon A, Nicolae CM. Mono-ADP-ribosylation by PARP10 and PARP14 in genome stability. NAR Cancer. 2023;5:zcad009. 10.1093/narcan/zcad009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Gupte R, Liu Z, Kraus WL. PARPs and ADP-ribosylation: recent advances linking molecular functions to biological outcomes. Genes Dev. 2017;31(2):101–26. 10.1101/gad.291518.116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.van Beek L, McClay E, Patel S, Schimpl M, Spagnolo L, Maia de Oliveira T. PARP Power: A Structural Perspective on PARP1, PARP2, and PARP3 in DNA Damage Repair and Nucleosome Remodelling. Int J Mol Sci. 2021; 22.10.3390/ijms22105112. [DOI] [PMC free article] [PubMed]
  • 39.Huang D, Su Z, Mei Y, Shao Z. The complex universe of inactive PARP1. Trends Genet. 2024;40:1074–85. 10.1016/j.tig.2024.08.009. [DOI] [PubMed] [Google Scholar]
  • 40.Pillay N, Mariotti L, Zaleska M, Inian O, Jessop M, Hibbs S, et al. Structural basis of tankyrase activation by polymerization. Nature. 2022;612:162–9. 10.1038/s41586-022-05449-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Jankevicius G, Hassler M, Golia B, Rybin V, Zacharias M, Timinszky G, et al. A family of macrodomain proteins reverses cellular mono-ADP-ribosylation. Nat Struct Mol Biol. 2013;20:508–14. 10.1038/nsmb.2523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Zhu H, Tang YD, Zhan G, Su C, Zheng C. The critical role of PARPs in regulating innate immune responses. Front Immunol. 2021;12:712556. 10.3389/fimmu.2021.712556. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Gibson BA, Kraus WL. New insights into the molecular and cellular functions of poly(ADP-ribose) and PARPs. Nat Rev Mol Cell Biol. 2012;13:411–24. 10.1038/nrm3376. [DOI] [PubMed] [Google Scholar]
  • 44.Schwarz SD, Xu J, Gunasekera K, Schurmann D, Vagbo CB, Ferrari E, et al. Covalent PARylation of DNA base excision repair proteins regulates DNA demethylation. Nat Commun. 2024; 15:184.10.1038/s41467-023-44209-8. [DOI] [PMC free article] [PubMed]
  • 45.Hochegger H, Dejsuphong D, Fukushima T, Morrison C, Sonoda E, Schreiber V, et al. Parp-1 protects homologous recombination from interference by Ku and ligase IV in vertebrate cells. EMBO J. 2006;25:1305–14. 10.1038/sj.emboj.7601015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Luijsterburg MS, de Krijger I, Wiegant WW, Shah RG, Smeenk G, de Groot AJL, et al. PARP1 Links CHD2-Mediated Chromatin Expansion and H3.3 Deposition to DNA Repair by Non-homologous End-Joining. Mol Cell. 2016; 61:547–562.10.1016/j.molcel.2016.01.019. [DOI] [PMC free article] [PubMed]
  • 47.Ray Chaudhuri A, Nussenzweig A. The multifaceted roles of PARP1 in DNA repair and chromatin remodelling. Nat Rev Mol Cell Biol. 2017; 18:610–621.10.1038/nrm.2017.53. [DOI] [PMC free article] [PubMed]
  • 48.Conceicao CJF, Moe E, Ribeiro PA, Raposo M. PARP1: a comprehensive review of its mechanisms, therapeutic implications and emerging cancer treatments. Biochimica et Biophysica Acta (BBA). 2025;1880:189282. 10.1016/j.bbcan.2025.189282. [DOI] [PubMed] [Google Scholar]
  • 49.Jubin T, Kadam A, Gani AR, Singh M, Dwivedi M, Begum R. Poly ADP-ribose polymerase-1: beyond transcription and towards differentiation. Semin Cell Dev Biol. 2017;63:167–79. 10.1016/j.semcdb.2016.07.027. [DOI] [PubMed] [Google Scholar]
  • 50.Zong W, Gong Y, Sun W, Li T, Wang ZQ. PARP1: Liaison of Chromatin Remodeling and Transcription. Cancers (Basel). 2022;14.10.3390/cancers14174162. [DOI] [PMC free article] [PubMed]
  • 51.Chen H, Gou X, Mao Y, O'Leary PC, Diolaiti ME, Ashworth A. PARP7 Inhibitors and AHR Agonists Act Synergistically across a Wide Range of Cancer Models. Mol Cancer Ther. 2025;24:56–68.10.1158/1535-7163.MCT-24-0211. [DOI] [PubMed]
  • 52.Popova K, Benedum J, Engl M, Lutgendorf-Caucig C, Fossati P, Widder J, et al. PARP7 as a new target for activating anti-tumor immunity in cancer. EMBO Mol Med. 2025;17:872–88. 10.1038/s44321-025-00214-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Li P, Lei Y, Qi J, Liu W, Yao K. Functional roles of ADP-ribosylation writers, readers and erasers. Front Cell Dev Biol. 2022;10:941356. 10.3389/fcell.2022.941356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Ribeiro VC, Russo LC, Hoch NC. PARP14 is regulated by the PARP9/DTX3L complex and promotes interferon γ-induced ADP-ribosylation. EMBO J. 2024;43(14):2908–28. 10.1038/s44318-024-00125-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Huang D, Kraus WL. The expanding universe of PARP1-mediated molecular and therapeutic mechanisms. Mol Cell. 2022;82:2315–34. 10.1016/j.molcel.2022.02.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Laspata N, Kaur P, Mersaoui SY, Muoio D, Liu ZS, Bannister MH, et al. PARP1 associates with R-loops to promote their resolution and genome stability. Nucleic Acids Res. 2023;51:2215–37. 10.1093/nar/gkad066. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Li Z, Liu Y, Liu Y, Zhang Y, Huen MSY, Lu H, et al. The PARP1-EXD2 axis orchestrates R-loop resolution to safeguard genome stability. Nat Chem Biol. 2025.10.1038/s41589-025-01952-x. [DOI] [PubMed]
  • 58.Kaufmann SH, Desnoyers S, Ottaviano Y, Davidson NE, Poirier GG. Specific proteolytic cleavage of poly(ADP-ribose) polymerase: an early marker of chemotherapy-induced apoptosis. Cancer Res. 1993;53:3976–85. [PubMed] [Google Scholar]
  • 59.Oliver FJ, de la Rubia G, Rolli V, Ruiz-Ruiz MC, de Murcia G, Murcia JM. Importance of poly(ADP-ribose) polymerase and its cleavage in apoptosis. Lesson from an uncleavable mutant. J Biol Chem. 1998;273:33533–9. 10.1074/jbc.273.50.33533. [DOI] [PubMed] [Google Scholar]
  • 60.Ryerson MR, Richards MM, Kvansakul M, Hawkins CJ, Shisler JL. Vaccinia Virus Encodes a Novel Inhibitor of Apoptosis That Associates with the Apoptosome. J Virol. 2017; 91:e01385–01317.10.1128/JVI.01385-17. [DOI] [PMC free article] [PubMed]
  • 61.Yeganeh B, Ghavami S, Rahim MN, Klonisch T, Halayko AJ, Coombs KM. Autophagy activation is required for influenza A virus-induced apoptosis and replication. Biochim Biophys Acta Mol Cell Res. 2018; 1865:364–378.10.1016/j.bbamcr.2017.10.014. [DOI] [PubMed]
  • 62.Sun Y, Aliyari SR, Parvatiyar K, Wang L, Zhen A, Sun W, et al. STING directly interacts with PAR to promote apoptosis upon acute ionizing radiation-mediated DNA damage. Cell Death Differ. 2025:1–13.10.1038/s41418-025-01457-z. [DOI] [PMC free article] [PubMed]
  • 63.Zimmerlin L, Zambidis ET. Pleiotropic roles of tankyrase/PARP proteins in the establishment and maintenance of human naive pluripotency. Exp Cell Res. 2020;390:111935. 10.1016/j.yexcr.2020.111935. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Camicia R, Bachmann SB, Winkler HC, Beer M, Tinguely M, Haralambieva E, et al. BAL1/ARTD9 represses the anti-proliferative and pro-apoptotic IFNgamma-STAT1-IRF1-p53 axis in diffuse large B-cell lymphoma. J Cell Sci. 2013; 126:1969–1980.10.1242/jcs.118174. [DOI] [PubMed]
  • 65.Iansante V, Choy PM, Fung SW, Liu Y, Chai J-G, Dyson J, et al. PARP14 promotes the Warburg effect in hepatocellular carcinoma by inhibiting JNK1-dependent PKM2 phosphorylation and activation. Nature Communications. 2015;6.10.1038/ncomms8882 [DOI] [PMC free article] [PubMed]
  • 66.David KK, Andrabi SA, Dawson TM, Dawson VL. Parthanatos, a messenger of death. Front Biosci (Landmark Ed). 2009; 14:1116–1128.10.2741/3297. [DOI] [PMC free article] [PubMed]
  • 67.Cande C, Cecconi F, Dessen P, Kroemer G. Apoptosis-inducing factor (AIF): key to the conserved caspase-independent pathways of cell death? J Cell Sci. 2002;115:4727–4734.10.1242/jcs.00210. [DOI] [PubMed]
  • 68.Yu SW, Wang H, Poitras MF, Coombs C, Bowers WJ, Federoff HJ, et al. Mediation of poly(ADP-ribose) polymerase-1-dependent cell death by apoptosis-inducing factor. Science. 2002;297:259–63. 10.1126/science.1072221. [DOI] [PubMed] [Google Scholar]
  • 69.Murata MM, Kong X, Moncada E, Chen Y, Imamura H, Wang P, et al. NAD+ consumption by PARP1 in response to DNA damage triggers metabolic shift critical for damaged cell survival. Mol Biol Cell. 2019;30:2584–97. 10.1091/mbc.E18-10-0650. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Liu L, Li J, Ke Y, Zeng X, Gao J, Ba X, et al. The key players of parthanatos: opportunities for targeting multiple levels in the therapy of parthanatos-based pathogenesis. Cell Mol Life Sci. 2022;79:60. 10.1007/s00018-021-04109-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Morris G, Walker AJ, Berk M, Maes M, Puri BK. Cell death pathways: a novel therapeutic approach for neuroscientists. Mol Neurobiol. 2018;55:5767–86. 10.1007/s12035-017-0793-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Park H, Kam TI, Peng H, Chou SC, Mehrabani-Tabari AA, Song JJ, et al. PAAN/MIF nuclease inhibition prevents neurodegeneration in Parkinson’s disease. Cell. 2022;185:1943–59. 10.1016/j.cell.2022.04.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Hong T, Lei G, Chen X, Li H, Zhang X, Wu N, et al. PARP inhibition promotes ferroptosis via repressing SLC7A11 and synergizes with ferroptosis inducers in BRCA-proficient ovarian cancer. Redox Biol. 2021;42:101928. 10.1016/j.redox.2021.101928. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Huang YL, Huang DY, Klochkov V, Chan CM, Chen YS, Lin WW. NLRX1 Inhibits LPS-Induced Microglial Death via Inducing p62-Dependent HO-1 Expression, Inhibiting MLKL and Activating PARP-1. Antioxidants (Basel). 2024; 13.10.3390/antiox13040481. [DOI] [PMC free article] [PubMed]
  • 75.Fujikawa DG. Programmed mechanisms of status epilepticus‐induced neuronal necrosis. Epilepsia Open. 2023;8(Suppl 1):S25–34. 10.1002/epi4.12593. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Kim C, Wang XD, Jang S, Yu Y. PARP1 inhibitors induce pyroptosis via caspase 3-mediated gasdermin E cleavage. Biochem Biophys Res Commun. 2023;646:78–85. 10.1016/j.bbrc.2023.01.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Richard IA, Burgess JT, O'Byrne KJ, Bolderson E. Beyond PARP1: The Potential of Other 1047 Members of the Poly (ADP-Ribose) Polymerase Family in DNA Repair and Cancer 1048 Therapeutics. Front Cell Dev Biol. 2021;10499:801200. 10.3389/fcell.2021.801200. [DOI] [PMC free article] [PubMed]
  • 78.Zhu H, Zheng C. When PARPs Meet Antiviral Innate Immunity. Trends Microbiol. 2021;29:776–8. 10.1016/j.tim.2021.01.002. [DOI] [PubMed]
  • 79. Suskiewicz MJ, Munnur D, Stromland O, Yang JC, Easton LE, Chatrin C, et al. Updated protein 1053 domain annotation of the PARP protein family sheds new light on biological function. 1054 Nucleic Acids Res. 2023;51:8217–36. 10.1093/nar/gkad514. [DOI] [PMC free article] [PubMed]
  • 80.Ke Y, Wang C, Zhang J, Zhong X, Wang R, Zeng X, et al. The Role of PARPs in Inflammation-1056 and Metabolic-Related Diseases: Molecular Mechanisms and Beyond. Cells. 2019;8:1057. 10.3390/cells8091047. [DOI] [PMC free article] [PubMed]
  • 81.Yang J, Wan SY, Song QY, Xie YH, Wan J, Zhou YH, et al. Angiopoietin-like protein 8 directs 1059 DNA damage responses towards apoptosis by stabilizing PARP1-DNA condensates. Cell 1060 Death Differ. 2025;32:672–88. 10.1038/s41418-024-01422-2. [DOI] [PMC free article] [PubMed]
  • 82.Slade D. Mitotic functions of poly(ADP-ribose) polymerases. Biochem Pharmacol. 2019;167:33–43. 10.1016/j.bcp.2019.03.028. [DOI] [PMC free article] [PubMed]
  • 83. Mashimo M, Kita M, Nobeyama A, Nomura A, Fujii T. PARP1 is activated by membrane 1064 damage and is involved in membrane repair through poly(ADP-ribosyl)ation. Genes Cells. 2022;27:305–12. 10.1111/gtc.12926. [DOI] [PubMed]
  • 84.Beck C, Robert I, Reina-San-Martin B, Schreiber V, Dantzer F. Poly(ADP-ribose) polymerases 1067 in double-strand break repair: focus on PARP1, PARP2 and PARP3. Exp Cell Res. 2014;329:18–25. 10.1016/j.yexcr.2014.07.003. [DOI] [PubMed]
  • 85.Bilkis R, Lake RJ, Cooper KL, Tomkinson A, Fan HY. The CSB chromatin remodeler regulates 1070 PARP1- and PARP2-mediated single-strand break repair at actively transcribed DNA 1071 regions. Nucleic Acids Res. 2023;51:7342–56. 10.1093/nar/gkad515. [DOI] [PMC free article] [PubMed]
  • 86.Kutuzov MM, Belousova EA, Ilina ES, Lavrik OI. Impact of PARP1, PARP2 & PARP3 on the 1073 Base Excision Repair of Nucleosomal DNA. Adv Exp Med Biol. 2020;1241:47–57. 10.1007/978-3-030-41283-8_4. [DOI] [PubMed]
  • 87.Muoio D, Laspata N, Dannenberg RL, Curry C, Darkoa-Larbi S, Hedglin M, et al. PARP2 1076 promotes Break Induced Replication-mediated telomere fragility in response to replication 1077 stress. Nat Commun. 2024;15:2857. 10.1038/s41467-024-47222-7. [DOI] [PMC free article] [PubMed]
  • 88.Zheng GD, Hu PJ, Chao YX, Zhou Y, Yang XJ, Chen BZ, et al. Nobiletin induces growth 1079 inhibition and apoptosis in human nasopharyngeal carcinoma C666-1 cells through regulating PARP-2/SIRT1/AMPK signaling pathway. Food Sci Nutr. 2019;7:1104–12. 10.1002/fsn3.953. [DOI] [PMC free article] [PubMed]
  • 89.Meng XW, Koh BD, Zhang JS, Flatten KS, Schneider PA, Billadeau DD, et al. Poly(ADP-1083 ribose) polymerase inhibitors sensitize cancer cells to death receptor-mediated apoptosis 1084 by enhancing death receptor expression. J Biol Chem. 2014;289:20543–58. 10.1074/jbc.M114.549220. [DOI] [PMC free article] [PubMed]
  • 90.Beck C, Rodriguez-Vargas JM, Boehler C, Robert I, Heyer V, Hanini N, et al. PARP3, a new 1087 therapeutic target to alter Rictor/mTORC2 signaling and tumor progression in BRCA1-1088 associated cancers. Cell Death Differ. 2019;26:1615–30. 10.1038/s41418-1089018-0233-1. [DOI] [PMC free article] [PubMed]
  • 91.Padella A, Ghelli Luserna Di Rora A, Marconi G, Ghetti M, Martinelli G, Simonetti G. 1091 Targeting PARP proteins in acute leukemia: DNA damage response inhibition and 1092 therapeutic strategies. J Hematol Oncol. 2022;15:10. 10.1186/s13045-022-109301228-0. [DOI] [PMC free article] [PubMed]
  • 92.Shelke GV, Jagtap JC, Kim DK, Shah RD, Das G, Shivayogi M, et al. TNF-alpha and IFN-1095 gamma Together Up-Regulates Par-4 Expression and Induce Apoptosis in Human 1096 Neuroblastomas. Biomedicines. 2017;6. 10.3390/biomedicines6010004. [DOI] [PMC free article] [PubMed]
  • 93. Li Y, Liu Y, Ma J, Yang Y, Yue Q, Zhu G, et al. PARP4 deficiency enhances sensitivity to ATM 1098 inhibitor by impairing DNA damage repair in melanoma. Cell Death Discov. 2025;11:35. 10.1038/s41420-025-02296-0. [DOI] [PMC free article] [PubMed]
  • 94. Frigon L, Pascal JM. Crystal structure and mutagenesis of a nucleic acid-binding BRCT domain 1101 in human PARP4. J Biol Chem. 2025;301:110277. 10.1016/j.jbc.2025.110277. [DOI] [PMC free article] [PubMed]
  • 95.DaRosa PA, Klevit RE, Xu W. Structural basis for tankyrase-RNF146 interaction reveals 1104 noncanonical tankyrase-binding motifs. Protein Sci. 2018;27:1057–105. 10.1002/pro.3413. [DOI] [PMC free article] [PubMed]
  • 96.Hou S, Zhang J, Jiang X, Yang Y, Shan B, Zhang M, et al. PARP5A and RNF146 phase 1107 separation restrains RIPK1-dependent necroptosis. Mol Cell. 2024;84:938–54.e938. 10.1016/j.molcel.2023.12.041. [DOI] [PubMed]
  • 97. Zhou B, Jiang ZH, Dai MR, Ai YL, Xiao L, Zhong CQ, et al. Full-length GSDME mediates 1110 pyroptosis independent from cleavage. Nat Cell Biol. 2024;26:1545–57. 10.1038/s41556-024-01463-2. [DOI] [PubMed]
  • 98.Huang JY, Wang K, Vermehren-Schmaedick A, Adelman JP, Cohen MS. PARP6 is a Regulator 1113 of Hippocampal Dendritic Morphogenesis. Sci Rep. 2016;6:18512. 10.1038/srep18512. [DOI] [PMC free article] [PubMed]
  • 99.Manetsch P, Bohi F, Nowak K, Leslie Pedrioli DM, Hottiger MO. PARP7-mediated ADP-1116 ribosylation of FRA1 promotes cancer cell growth by repressing IRF1- and IRF3-1117 dependent apoptosis. Proc Natl Acad Sci U S A. 2023;120:e2309047120. 10.1073/pnas.2309047120. [DOI] [PMC free article] [PubMed]
  • 100.Gu H, Yan W, Yang J, Liu B, Zhao X, Wang H, et al. Discovery of Highly Selective PARP7 1120 Inhibitors with a Novel Scaffold for Cancer Immunotherapy. J Med Chem. 2024;67:1932–48. 10.1021/acs.jmedchem.3c01764. [DOI] [PubMed]
  • 101. Gozgit JM, Vasbinder MM, Abo RP, Kunii K, Kuplast-Barr KG, Gui B, et al. PARP7 1123 negatively regulates the type I interferon response in cancer cells and its inhibition triggers 1124 antitumor immunity. Cancer Cell. 2021;39:1214–26.e1210. 10.1016/j.ccell.2021.06.018. [DOI] [PubMed]
  • 102. Szanto M, Gupte R, Kraus WL, Pacher P, Bai P. PARPs in lipid metabolism and related 1127 diseases. Prog Lipid Res. 2021;84:101117. 10.1016/j.plipres.2021.101117. [DOI] [PubMed]
  • 103.Bora NR, Kumar R, Kumar S. Development of an apoptotic lentogenic Newcastle disease 1129 virus strain by incorporating the p30 protein of African swine fever virus. Virology. 2025;1130 606. 10.1016/j.virol.2025.110477. [DOI] [PubMed]
  • 104. Robert I, Gaudot L, Yelamos J, Noll A, Wong HK, Dantzer F, et al. Robust immunoglobulin 1132 class switch recombination and end joining in Parp9-deficient mice. Eur J Immunol. 2017;47:665–76. 10.1002/eji.201646757. [DOI] [PubMed]
  • 105.Herzog N, Hartkamp JD, Verheugd P, Treude F, Forst AH, Feijs KL, et al. Caspase-dependent 1135 cleavage of the mono-ADP-ribosyltransferase ARTD10 interferes with its pro-apoptotic 1136 function. FEBS J. 2013;280:1330–43. 10.1111/febs.12124. [DOI] [PubMed]
  • 106. Khatib JB, Dhoonmoon A, Moldovan GL, Nicolae CM. PARP10 promotes the repair of 1138 nascent strand DNA gaps through RAD18 mediated translesion synthesis. Nat Commun. 2024;15:6197. 10.1038/s41467-024-50429-3. [DOI] [PMC free article] [PubMed]
  • 107.Cao G, Guo R, Chen M, Sun C, Wang J. PARP12 regulates mouse oocyte meiotic maturation. 1141 J Cell Physiol. 2023;238:1580–91. 10.1002/jcp.31037. [DOI] [PubMed]
  • 108.Todorova T, Bock FJ, Chang P. Poly(ADP-ribose) polymerase-13 and RNA regulation in 1036 immunity and cancer. Trends Mol Med. 2015;21:373–84. 10.1016/j.molmed.2015.03.002. [DOI] [PMC free article] [PubMed]
  • 109.Todorova T, Bock FJ, Chang P. PARP13 regulates cellular mRNA post-transcriptionally and 1143 functions as a pro-apoptotic factor by destabilizing TRAILR4 transcript. Nat Commun. 2014;5:5362. 10.1038/ncomms6362. [DOI] [PMC free article] [PubMed]
  • 110.Im JY, Kim SJ, Park JL, Han TH, Kim WI, Kim I, et al. CYB5R3 functions as a tumor 1146 suppressor by inducing ER stress-mediated apoptosis in lung cancer cells via the PERK-1147 ATF4 and IRE1alpha-JNK pathways. Exp Mol Med. 2024;56:235–49. 10.1038/s12276-024-01155-9. [DOI] [PMC free article] [PubMed]
  • 111. Jwa M, Chang P. PARP16 is a tail-anchored endoplasmic reticulum protein required for the 1150 PERK- and IRE1α-mediated unfolded protein response. Nature Cell Biology. 2012;14:1223–30. 10.1038/ncb2593. [DOI] [PMC free article] [PubMed]
  • 112.Wang J, Cheng Q, Zhang Y, Hong C, Liu J, Liu X, et al. PARP16-Mediated Stabilization of 1153 Amyloid Precursor Protein mRNA Exacerbates Alzheimer's Disease Pathogenesis. Aging 1154 Dis. 2023;14:1458–71. 10.14336/AD.2023.0119. [DOI] [PMC free article] [PubMed]
  • 113.Alemasova EE, Lavrik OI. Poly(ADP-ribosyl)ation by PARP1: reaction mechanism and 1030 regulatory proteins. Nucleic Acids Res. 2019;47:3811–27. 10.1093/nar/gkz120. [DOI] [PMC free article] [PubMed]
  • 114.Genois MM, Gagne JP, Yasuhara T, Jackson J, Saxena S, Langelier MF, et al. CARM1 regulates 1033 replication fork speed and stress response by stimulating PARP1. Mol Cell. 2021;81:784–800.e788. 10.1016/j.molcel.2020.12.010. [DOI] [PMC free article] [PubMed]
  • 115.Liao M, Long X, Chen Y, An J, Huang W, Xu X, et al. PARP9 exacerbates apoptosis and 1039 neuroinflammation via the PI3K pathway in the thalamus and hippocampus and cognitive 1040 decline after cortical infarction. J Neuroinflammation. 2025;22:43. 10.1186/s12974-025-03374-x. [DOI] [PMC free article] [PubMed]
  • 116. Sukhanova MV, Hamon L, Kutuzov MM, Joshi V, Abrakhi S, Dobra I, et al. A Single-Molecule 1043 Atomic Force Microscopy Study of PARP1 and PARP2 Recognition of Base Excision 1044 Repair DNA Intermediates. J Mol Biol. 2019;431:2655–73. 10.1016/j.jmb.2019.05.028. [DOI] [PubMed]
  • 117.Ogden TEH, Yang JC, Schimpl M, Easton LE, Underwood E, Rawlins PB, et al. Dynamics of the HD regulatory subdomain of PARP-1; substrate access and allostery in PARP activation and inhibition. Nucleic Acids Res. 2021;49:2266–88. 10.1093/nar/gkab020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Drew Y, Plummer R. The emerging potential of poly(ADP-ribose) polymerase inhibitors in the treatment of breast cancer. Curr Opin Obstet Gynecol. 2010;22:67–71. 10.1097/GCO.0b013e328334ff57. [DOI] [PubMed] [Google Scholar]
  • 119.Murai J, Huang SY, Das BB, Renaud A, Zhang Y, Doroshow JH, et al. Trapping of PARP1 and PARP2 by clinical PARP inhibitors. Cancer Res. 2012;72:5588–99. 10.1158/0008-5472.CAN-12-2753. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Beck C, Boehler C, Guirouilh Barbat J, Bonnet ME, Illuzzi G, Ronde P, et al. PARP3 affects the relative contribution of homologous recombination and nonhomologous end-joining pathways. Nucleic Acids Res. 2014;42:5616–32. 10.1093/nar/gku174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Li N, Zhang Y, Han X, Liang K, Wang J, Feng L, et al. Poly-ADP ribosylation of PTEN by tankyrases promotes PTEN degradation and tumor growth. Genes Dev. 2015;29:157–70. 10.1101/gad.251785.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Zhou Y, Liu L, Tao S, Yao Y, Wang Y, Wei Q, et al. Parthanatos and its associated components: promising therapeutic targets for cancer. Pharmacol Res. 2021;163:105299. 10.1016/j.phrs.2020.105299. [DOI] [PubMed] [Google Scholar]
  • 123.Sun Y, Aliyari SR, Parvatiyar K, Wang L, Zhen A, Sun W, et al. STING directly interacts with PAR to promote apoptosis upon acute ionizing radiation-mediated DNA damage. Cell Death Differ. 2025;32:1167–79. 10.1038/s41418-025-01457-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Guan R, Li C, Jiao R, Li J, Wei R, Feng C, et al. MRPL21-PARP1 axis promotes cisplatin resistance in head and neck squamous cell carcinoma by inhibiting autophagy through the PI3K/AKT/mTOR signaling pathway. J Exp Clin Cancer Res. 2025;44:221. 10.1186/s13046-025-03482-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Zhang Y, Wu K, Liu Y, Sun S, Shao Y, Li Q, et al. UHRF2 promotes the malignancy of hepatocellular carcinoma by PARP1 mediated autophagy. Cell Signal. 2023;109:110782. 10.1016/j.cellsig.2023.110782. [DOI] [PubMed] [Google Scholar]
  • 126.Zhang M, Guo R, Zhai Y, Yang D. LIGHT sensitizes IFNγ-mediated apoptosis of MDA-MB-231 breast cancer cells leading to down-regulation of anti-apoptosis Bcl-2 family members. Cancer Lett. 2003;195(2):201–10. 10.1016/s0304-3835(03)00148-4. [DOI] [PubMed] [Google Scholar]
  • 127.Li Q, Kawamura K, Ma G, Iwata F, Numasaki M, Suzuki N, et al. Interferon-lambda induces G1 phase arrest or apoptosis in oesophageal carcinoma cells and produces anti-tumour effects in combination with anti-cancer agents. Eur J Cancer. 2010; 46:180–190.10.1016/j.ejca.2009.10.002. [DOI] [PubMed]
  • 128.Brooks DM, Anand S, Cohen MS. Immunomodulatory roles of PARPs: shaping the tumor microenvironment, one ADP-ribose at a time. Curr Opin Chem Biol. 2023;77:102402. 10.1016/j.cbpa.2023.102402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Chan N, Pires IM, Bencokova Z, Coackley C, Luoto KR, Bhogal N, et al. Contextual synthetic lethality of cancer cell kill based on the tumor microenvironment. Cancer Res. 2010; 70:8045–8054.10.1158/0008-5472.CAN-10-2352. [DOI] [PMC free article] [PubMed]
  • 130.Yi XF, Gao RL, Sun L, Wu ZX, Zhang SL, Huang LT, et al. Dual antitumor immunomodulatory effects of PARP inhibitor on the tumor microenvironment: a counterbalance between anti-tumor and pro-tumor. Biomed Pharmacother. 2023;163:114770. 10.1016/j.biopha.2023.114770. [DOI] [PubMed] [Google Scholar]
  • 131.Dorsam 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:E4061–70. 10.1073/pnas.1712345115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Sturniolo I, Varoczy C, Regdon Z, Mazlo A, Muzsai S, Bacsi A, et al. PARP14 Contributes to the Development of the Tumor-Associated Macrophage Phenotype. Int J Mol Sci. 2024; 25.10.3390/ijms25073601. [DOI] [PMC free article] [PubMed]
  • 133.Mentz M, Keay W, Strobl CD, Antoniolli M, Adolph L, Heide M, et al. PARP14 is a novel target in STAT6 mutant follicular lymphoma. Leukemia. 2022;36:2281–92. 10.1038/s41375-022-01641-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Palavalli Parsons LH, Challa S, Gibson BA, Nandu T, Stokes MS, Huang D, et al. Identification of PARP-7 substrates reveals a role for MARylation in microtubule control in ovarian cancer cells. Elife. 2021; 10.10.7554/eLife.60481. [DOI] [PMC free article] [PubMed]
  • 135.Humer D, Klepsch V, Baier G. Potentiation of cancer immunogenicity by targeting PARP. J Immunother Cancer. 2025; 13.10.1136/jitc-2024-011056. [DOI] [PMC free article] [PubMed]
  • 136.Luo Y, Xia Y, Liu D, Li X, Li H, Liu J, et al. Neoadjuvant PARpi or chemotherapy in ovarian cancer informs targeting effector Treg cells for homologous-recombination-deficient tumors. Cell. 2024;187:4905. 10.1016/j.cell.2024.06.013. [DOI] [PubMed] [Google Scholar]
  • 137.Paul S, Sinha S, Kundu CN. Targeting cancer stem cells in the tumor microenvironment: an emerging role of PARP inhibitors. Pharmacol Res. 2022;184:106425. 10.1016/j.phrs.2022.106425. [DOI] [PubMed] [Google Scholar]
  • 138.Shen J, Xu F, Liu T, Ye Y, Xu S. NAD(+) metabolism-mediated SURF4-STING axis enhances T-cell anti-tumor effects in the ovarian cancer microenvironment. Cell Death Dis. 2025;16:640. 10.1038/s41419-025-07939-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Ran X, Wu BX, Vidhyasagar V, Song L, Zhang X, Ladak RJ, et al. PARP inhibitor radiosensitization enhances anti-PD-L1 immunotherapy through stabilizing chemokine mRNA in small cell lung cancer. Nat Commun. 2025;16:2166. 10.1038/s41467-025-57257-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Ying S, Hamdy FC, Helleday T. Mre11-dependent degradation of stalled DNA replication forks is prevented by BRCA2 and PARP1. Cancer Res. 2012;72:2814–21. 10.1158/0008-5472.CAN-11-3417. [DOI] [PubMed] [Google Scholar]
  • 141.Ronson GE, Piberger AL, Higgs MR, Olsen AL, Stewart GS, McHugh PJ, et al. PARP1 and PARP2 stabilise replication forks at base excision repair intermediates through Fbh1-dependent Rad51 regulation. Nat Commun. 2018;9:746. 10.1038/s41467-018-03159-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Wu CK, Shiu JL, Wu CL, Hung CF, Ho YC, Chen YT, et al. APLF facilitates interstrand DNA crosslink repair and replication fork protection to confer cisplatin resistance. Nucleic Acids Res. 2024;52:5676–97. 10.1093/nar/gkae211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Nicolae CM, Aho ER, Vlahos AH, Choe KN, De S, Karras GI, et al. The ADP-ribosyltransferase PARP10/ARTD10 interacts with proliferating cell nuclear antigen (PCNA) and is required for DNA damage tolerance. J Biol Chem. 2014; 289:13627–13637.10.1074/jbc.M114.556340. [DOI] [PMC free article] [PubMed]
  • 144.Demeny MA, Virag L. The PARP Enzyme Family and the Hallmarks of Cancer Part 1. Cell Intrinsic Hallmarks. Cancers (Basel). 2021; 13.10.3390/cancers13092042. [DOI] [PMC free article] [PubMed]
  • 145.Chen H, Diolaiti ME, O’Leary PC, Rojc A, Krogan NJ, Kim M, et al. A whole-genome CRISPR screen identifies AHR loss as a mechanism of resistance to a PARP7 inhibitor. Mol Cancer Ther. 2022;21(7):1076–89. 10.1158/1535-7163.MCT-21-0841. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Zhu Y, Liu Z, Wan Y, Zou L, Liu L, Ding S, et al. PARP14 promotes the growth and glycolysis of acute myeloid leukemia cells by regulating HIF-1α expression. Clin Immunol. 2022;242:109094. 10.1016/j.clim.2022.109094. [DOI] [PubMed] [Google Scholar]
  • 147.Li Z, Luo A, Xie B. The Complex Network of ADP-Ribosylation and DNA Repair: Emerging Insights and Implications for Cancer Therapy. Int J Mol Sci. 2023; 24.10.3390/ijms241915028. [DOI] [PMC free article] [PubMed]
  • 148.Sutcu HH, Matta E, Ishchenko AA. Role of PARP-catalyzed ADP-ribosylation in the crosstalk between DNA strand breaks and epigenetic regulation. J Mol Biol. 2020;432:1769–91. 10.1016/j.jmb.2019.12.019. [DOI] [PubMed] [Google Scholar]
  • 149.Guo T, Zuo Y, Qian L, Liu J, Yuan Y, Xu K, et al. ADP-ribosyltransferase PARP11 modulates the interferon antiviral response by mono-ADP-ribosylating the ubiquitin E3 ligase β-TrCP. Nat Microbiol. 2019;4(11):1872–84. 10.1038/s41564-019-0428-3. [DOI] [PubMed] [Google Scholar]
  • 150.Du X, Zhou J, Zhou Y, Chen Y, Kang Y, Zhao D, et al. PARP7i Clinical Candidate RBN‐2397 Exerts Antiviral Activity by Modulating Interferon‐β Associated Innate Immune Response in Macrophages. Drug Dev Res. 2024;85(7):e70013. 10.1002/ddr.70013. [DOI] [PubMed] [Google Scholar]
  • 151.Xing J, Zhang A, Du Y, Fang M, Minze LJ, Liu YJ, et al. Identification of poly(ADP-ribose) polymerase 9 (PARP9) as a noncanonical sensor for RNA virus in dendritic cells. Nat Commun. 2021;12(1):2681. 10.1038/s41467-021-23003-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Zhang Y, Mao D, Roswit WT, Jin X, Patel AC, Patel DA, et al. PARP9-DTX3L ubiquitin ligase targets host histone H2BJ and viral 3C protease to enhance interferon signaling and control viral infection. Nat Immunol. 2015;16:1215–27. 10.1038/ni.3279. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Kar P, Chatrin C, Dukic N, Suyari O, Schuller M, Zhu K, et al. PARP14 and PARP9/DTX3L regulate interferon-induced ADP-ribosylation. EMBO J. 2024; 43:2929–2953.10.1038/s44318-024-00126-0. [DOI] [PMC free article] [PubMed]
  • 154.Lin Z, Li Y, Gong G, Xia Y, Wang C, Chen Y, et al. Restriction of H1N1 influenza virus infection by selenium nanoparticles loaded with ribavirin via resisting caspase-3 apoptotic pathway. Int J Nanomedicine. 2018; 13:5787–5797.10.2147/IJN.S177658. [DOI] [PMC free article] [PubMed]
  • 155.Huang X, Li F, Liu L, Li Y, Zhang M, Ma G, et al. PARP12-mediated mono-ADP-ribosylation as a checkpoint for necroptosis and apoptosis. Proc Natl Acad Sci U S A. 2025;122:e2426660122. 10.1073/pnas.2426660122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Westera L, Jennings AM, Maamary J, Schwemmle M, Garcia-Sastre A, Bortz E. Poly-ADP Ribosyl Polymerase 1 (PARP1) Regulates Influenza A Virus Polymerase. Adv Virol. 2019; 2019:8512363.10.1155/2019/8512363. [DOI] [PMC free article] [PubMed]
  • 157.Piskunova TS, Yurova MN, Ovsyannikov AI, Semenchenko AV, Zabezhinski MA, Popovich IG, et al. Deficiency in poly(ADP-ribose) polymerase-1 (PARP-1) accelerates aging and spontaneous carcinogenesis in mice. Curr Gerontol Geriatr Res. 2008;2008:754190. 10.1155/2008/754190. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Yuan J, Li R, Song H, Liu X. Sirtuin3 Alleviated Influenza A Virus-Induced Mitochondrial Oxidative Stress and Inflammation in Lung Epithelial Cells via Regulating Poly (ADP-Ribose) Polymerase 1 Activity. Int Arch Allergy Immunol. 2023; 184:447–459.10.1159/000527535. [DOI] [PubMed]
  • 159.Kozaki T, Komano J, Kanbayashi D, Takahama M, Misawa T, Satoh T, et al. Mitochondrial damage elicits a TCDD-inducible poly(ADP-ribose) polymerase-mediated antiviral response. Proc Natl Acad Sci U S A. 2017;114:2681–6. 10.1073/pnas.1621508114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Desingu PA, Mishra S, Dindi L, Srinivasan S, Rajmani RS, Ravi V, et al. PARP1 inhibition protects mice against Japanese encephalitis virus infection. Cell Rep. 2023;42:113103. 10.1016/j.celrep.2023.113103. [DOI] [PubMed] [Google Scholar]
  • 161.Li L, Zhao H, Liu P, Li C, Quanquin N, Ji X, et al. PARP12 suppresses Zika virus infection through PARP-dependent degradation of NS1 and NS3 viral proteins. Sci Signal. 2018; 11:9332.10.1126/scisignal.aas9332. [DOI] [PMC free article] [PubMed]
  • 162.Wang F, Ma MT, Xu J, Liu H. Identification of phosphorylation site on PARP1 mediating its cytosolic translocation in virus-infected HeLa cells. STAR Protoc. 2022;3:101808. 10.1016/j.xpro.2022.101808. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.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; 82:2032–2049 e2037.10.1016/j.molcel.2022.03.034. [DOI] [PubMed]
  • 164.Li Z, Yamauchi Y, Kamakura M, Murayama T, Goshima F, Kimura H, et al. Herpes simplex virus requires poly(ADP-ribose) polymerase activity for efficient replication and induces extracellular signal-related kinase-dependent phosphorylation and ICP0-dependent nuclear localization of tankyrase 1. J Virol. 2012;86:492–503. 10.1128/JVI.05897-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Zhang W, Guo J, Chen Q. Role of PARP-1 in human cytomegalovirus infection and functional partners encoded by this virus. Viruses. 2022;14:2049. 10.3390/v14092049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Maestri D, Napoletani G, Kossenkov A, Preston-Alp S, Caruso LB, Tempera I. The three-dimensional structure of the EBV genome plays a crucial role in regulating viral gene expression in EBVaGC. Nucleic Acids Res. 2023;51(22):12092–110. 10.1093/nar/gkad936. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Morgan SM, Tanizawa H, Caruso LB, Hulse M, Kossenkov A, Madzo J, et al. The three-dimensional structure of Epstein-Barr virus genome varies by latency type and is regulated by PARP1 enzymatic activity. Nat Commun. 2022;13:187. 10.1038/s41467-021-27894-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Ohsaki E, Ueda K, Sakakibara S, Do E, Yada K, Yamanishi K. Poly(ADP-ribose) polymerase 1 binds to Kaposi's sarcoma-associated herpesvirus (KSHV) terminal repeat sequence and modulates KSHV replication in latency. J Virol. 2004; 78:9936–9946.10.1128/JVI.78.18.9936-9946.2004. [DOI] [PMC free article] [PubMed]
  • 169.Chung WC, Kim J, Kim BC, Kang HR, Son J, Ki H, et al. Structure-based mechanism of action of a viral poly(ADP-ribose) polymerase 1-interacting protein facilitating virus replication. IUCrJ. 2018;5:866–79. 10.1107/S2052252518013854. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Nebenzahl-Sharon K, Sharf R, Amer J, Shalata H, Khoury-Haddad H, Sohn SY, et al. An Interaction with PARP-1 and Inhibition of Parylation Contribute to Attenuation of DNA Damage Signaling by the Adenovirus E4orf4 Protein. J Virol. 2019; 93:e02253–02218.10.1128/JVI.02253-18. [DOI] [PMC free article] [PubMed]
  • 171.Deng Z, Atanasiu C, Zhao K, Marmorstein R, Sbodio JI, Chi NW, et al. Inhibition of Epstein-Barr virus OriP function by tankyrase, a telomere-associated poly-ADP ribose polymerase that binds and modifies EBNA1. J Virol. 2005; 79:4640–4650.10.1128/JVI.79.8.4640-4650.2005. [DOI] [PMC free article] [PubMed]
  • 172.Lupey-Green LN, Moquin SA, Martin KA, McDevitt SM, Hulse M, Caruso LB, et al. PARP1 restricts Epstein Barr Virus lytic reactivation by binding the BZLF1 promoter. Virology. 2017; 507:220–230.10.1016/j.virol.2017.04.006. [DOI] [PMC free article] [PubMed]
  • 173.Kim C, Wang XD, Yu Y. PARP1 inhibitors trigger innate immunity via PARP1 trapping-induced DNA damage response. Elife. 2020; 9.10.7554/eLife.60637. [DOI] [PMC free article] [PubMed]
  • 174.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. 10.1158/0008-5472.CAN-18-1003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Iansante V, Choy PM, Fung SW, Liu Y, Chai JG, Dyson J, et al. PARP14 promotes the Warburg effect in hepatocellular carcinoma by inhibiting JNK1-dependent PKM2 phosphorylation and activation. Nat Commun. 2015;6:7882. 10.1038/ncomms8882. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Heer CD, Sanderson DJ, Voth LS, Alhammad YMO, Schmidt MS, Trammell SAJ, et al. Coronavirus infection and PARP expression dysregulate the NAD metabolome: an actionable component of innate immunity. J Biol Chem. 2020;295:17986–96. 10.1074/jbc.RA120.015138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Thapa K, Khan H, Sharma U, Grewal AK, Singh TG. Poly (ADP-ribose) polymerase-1 as a promising drug target for neurodegenerative diseases. Life Sci. 2021;267:118975. 10.1016/j.lfs.2020.118975. [DOI] [PubMed] [Google Scholar]
  • 178.Motyl J, Wencel PL, Cieslik M, Strosznajder RP, Strosznajder JB. Alpha-synuclein alters differently gene expression of Sirts, PARPs and other stress response proteins: implications for neurodegenerative disorders. Mol Neurobiol. 2018; 55:727–740.10.1007/s12035-016-0317-1. [DOI] [PMC free article] [PubMed]
  • 179.Zhang Y, Chen JC, Zheng JH, Cheng YZ, Weng WP, Zhong RL, et al. Pterosin B improves cognitive dysfunction by promoting microglia M1/M2 polarization through inhibiting Klf5/Parp14 pathway. Phytomedicine. 2024;135:156152. 10.1016/j.phymed.2024.156152. [DOI] [PubMed] [Google Scholar]
  • 180.Xu W, Wu Y, Mao R, Jia Y, Jiang H, Zhang F, et al. Poly(ADP-ribose) polymerase 1 orchestrates vascular smooth muscle cell homeostasis in arterial disease. Exp Mol Med. 2025.10.1038/s12276-025-01501-5. [DOI] [PMC free article] [PubMed]
  • 181.Lin J, Wang X, Gu M, Chen Y, Xu J, VChau N, et al. Geniposide ameliorates atherosclerosis by restoring lipophagy via suppressing PARP1/PI3K/AKT signaling pathway. Phytomedicine. 2024;129:155617. 10.1016/j.phymed.2024.155617. [DOI] [PubMed] [Google Scholar]
  • 182.Distler A, Deloch L, Huang J, Dees C, Lin NY, Palumbo-Zerr K, et al. Inactivation of tankyrases reduces experimental fibrosis by inhibiting canonical Wnt signalling. Ann Rheum Dis. 2013;72:1575–80. 10.1136/annrheumdis-2012-202275. [DOI] [PubMed] [Google Scholar]
  • 183.Lakshmi TV, Bale S, Khurana A, Godugu C. Tankyrase as a novel molecular target in cancer and fibrotic diseases. Curr Drug Targets. 2017;18:1214–24. 10.2174/1389450117666160715152503. [DOI] [PubMed] [Google Scholar]
  • 184.Sunil VR, Vayas KN, Radbel J, Abramova E, Gow A, Laskin JD, et al. Impaired energy metabolism and altered functional activity of alveolar type II epithelial cells following exposure of rats to nitrogen mustard. Toxicol Appl Pharmacol. 2022;456:116257. 10.1016/j.taap.2022.116257. [DOI] [PubMed] [Google Scholar]
  • 185.Lu Q, Wang N, Wen D, Guo P, Liu Y, Fu S, et al. Baicalin attenuates lipopolysaccharide-induced intestinal inflammatory injury via suppressing PARP1-mediated NF-κB and NLRP3 signalling pathway. Toxicon. 2024;239:107612. 10.1016/j.toxicon.2024.107612. [DOI] [PubMed] [Google Scholar]
  • 186.Mehrotra P, Riley JP, Patel R, Li F, Voss L, Goenka S. PARP-14 functions as a transcriptional switch for Stat6-dependent gene activation. J Biol Chem. 2011;286:1767–76. 10.1074/jbc.M110.157768. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Fettucciari K, Marguerie F, Fruganti A, Marchegiani A, Spaterna A, Brancorsini S, et al. Clostridioides difficile toxin B alone and with pro-inflammatory cytokines induces apoptosis in enteric glial cells by activating three different signalling pathways mediated by caspases, calpains and cathepsin B. Cell Mol Life Sci. 2022;79:442. 10.1007/s00018-022-04459-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Rosado MM, Bennici E, Novelli F, Pioli C. Beyond DNA repair, the immunological role of PARP-1 and its siblings. Immunology. 2013;139:428–37. 10.1111/imm.12099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Kamboj A, Lu P, Cossoy MB, Stobart JL, Dolhun BA, Kauppinen TM, et al. Poly(ADP-ribose) polymerase 2 contributes to neuroinflammation and neurological dysfunction in mouse experimental autoimmune encephalomyelitis. J Neuroinflammation. 2013; 10:49.10.1186/1742-2094-10-49. [DOI] [PMC free article] [PubMed]
  • 190.Gariani K, Ryu D, Menzies KJ, Yi HS, Stein S, Zhang H, et al. Inhibiting poly ADP-ribosylation increases fatty acid oxidation and protects against fatty liver disease. J Hepatol. 2017;66:132–41. 10.1016/j.jhep.2016.08.024. [DOI] [PubMed] [Google Scholar]
  • 191.Bai P, Canto C, Brunyanszki A, Huber A, Szanto M, Cen Y, et al. PARP-2 regulates SIRT1 expression and whole-body energy expenditure. Cell Metab. 2011;13:450–60. 10.1016/j.cmet.2011.03.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Marton J, Peter M, Balogh G, Bodi B, Vida A, Szanto M, et al. Poly(ADP-ribose) polymerase-2 is a lipid-modulated modulator of muscular lipid homeostasis. Biochim Biophys Acta Mol Cell Biol Lipids. 2018; 1863:1399–1412.10.1016/j.bbalip.2018.07.013. [DOI] [PubMed]
  • 193.Hans CP, Feng Y, Naura AS, Zerfaoui M, Rezk BM, Xia H, et al. Protective effects of PARP-1 knockout on dyslipidemia-induced autonomic and vascular dysfunction in ApoE mice: effects on eNOS and oxidative stress. PLoS ONE. 2009;4:e7430. 10.1371/journal.pone.0007430. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Hans CP, Zerfaoui M, Naura AS, Troxclair D, Strong JP, Matrougui K, et al. Thieno[2,3-c]isoquinolin-5-one, a potent poly(ADP-ribose) polymerase inhibitor, promotes atherosclerotic plaque regression in high-fat diet-fed apolipoprotein E-deficient mice: effects on inflammatory markers and lipid content. J Pharmacol Exp Ther. 2009;329:150–8. 10.1124/jpet.108.145938. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Mukhopadhyay P, Horvath B, Rajesh M, Varga ZV, Gariani K, Ryu D, et al. PARP inhibition protects against alcoholic and non-alcoholic steatohepatitis. J Hepatol. 2017;66:589–600. 10.1016/j.jhep.2016.10.023. [DOI] [PubMed] [Google Scholar]
  • 196.Choi Y, Abdelmegeed MA, Song BJ. Diet high in fructose promotes liver steatosis and hepatocyte apoptosis in C57BL/6J female mice: role of disturbed lipid homeostasis and increased oxidative stress. Food Chem Toxicol. 2017;103:111–21. 10.1016/j.fct.2017.02.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Oumouna-Benachour K, Hans CP, Suzuki Y, Naura A, Datta R, Belmadani S, et al. Poly(ADP-ribose) polymerase inhibition reduces atherosclerotic plaque size and promotes factors of plaque stability in apolipoprotein E-deficient mice: effects on macrophage recruitment, nuclear factor-kappaB nuclear translocation, and foam cell death. Circulation. 2007; 115:2442–2450.10.1161/CIRCULATIONAHA.106.668756. [DOI] [PubMed]
  • 198.Curtin N. PARP inhibitors for anticancer therapy. Biochem Soc Trans. 2014;42:82–8. 10.1042/BST20130187. [DOI] [PubMed] [Google Scholar]
  • 199.Lin KY, Kraus WL. PARP inhibitors for cancer therapy. Cell. 2017;169:183. 10.1016/j.cell.2017.03.034. [DOI] [PubMed] [Google Scholar]
  • 200.Patel M, Nowsheen S, Maraboyina S, Xia F. The role of poly(ADP-ribose) polymerase inhibitors in the treatment of cancer and methods to overcome resistance: a review. Cell Biosci. 2020;10:35. 10.1186/s13578-020-00390-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Moore KN, Pothuri B, Monk B, Coleman RL. PARP inhibition as frontline therapy in ovarian cancer. Clin Adv Hematol Oncol. 2020;18:550–6. [PubMed] [Google Scholar]
  • 202.Loibl S, O'Shaughnessy J, Untch M, Sikov WM, Rugo HS, McKee MD, et al. Addition of the PARP inhibitor veliparib plus carboplatin or carboplatin alone to standard neoadjuvant chemotherapy in triple-negative breast cancer (BrighTNess): a randomised, phase 3 trial. Lancet Oncol. 2018; 19:497–509.10.1016/S1470-2045(18)30111-6. [DOI] [PubMed]
  • 203.Lee A. Fuzuloparib: first approval. Drugs. 2021;81(10):1221–6. 10.1007/s40265-021-01541-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Markham A. Pamiparib: first approval. Drugs. 2021;81:1343–8. 10.1007/s40265-021-01552-8. [DOI] [PubMed] [Google Scholar]
  • 205.Lee A. Senaparib: First Approval. Drugs. 2025.10.1007/s40265-025-02200-1. [DOI] [PubMed]
  • 206.De Vos M, Schreiber V, Dantzer F. The diverse roles and clinical relevance of PARPs in DNA damage repair: current state of the art. Biochem Pharmacol. 2012;84:137–46. 10.1016/j.bcp.2012.03.018. [DOI] [PubMed] [Google Scholar]
  • 207.Moynahan ME, Jasin M. Mitotic homologous recombination maintains genomic stability and suppresses tumorigenesis. Nat Rev Mol Cell Biol. 2010;11:196–207. 10.1038/nrm2851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Bai P, Canto C, Oudart H, Brunyanszki A, Cen Y, Thomas C, et al. PARP-1 inhibition increases mitochondrial metabolism through SIRT1 activation. Cell Metab. 2011;13:461–8. 10.1016/j.cmet.2011.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Hassa PO, Hottiger MO. The diverse biological roles of mammalian PARPS, a small but powerful family of poly-ADP-ribose polymerases. Front Biosci. 2008; 13:3046–3082.10.2741/2909. [DOI] [PubMed]
  • 210.Luo X, Kraus WL. On PAR with PARP: cellular stress signaling through poly(ADP-ribose) and PARP-1. Genes Dev. 2012;26:417–32. 10.1101/gad.183509.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Jagtap P, Szabo C. Poly(ADP-ribose) polymerase and the therapeutic effects of its inhibitors. Nat Rev Drug Discov. 2005;4:421–40. 10.1038/nrd1718. [DOI] [PubMed] [Google Scholar]
  • 212.Kauppinen TM, Swanson RA. Poly(ADP-ribose) polymerase-1 promotes microglial activation, proliferation, and matrix metalloproteinase-9-mediated neuron death. J Immunol. 2005;174:2288–96. 10.4049/jimmunol.174.4.2288. [DOI] [PubMed] [Google Scholar]
  • 213.Tapodi A, Bognar Z, Szabo C, Gallyas F, Sumegi B, Hocsak E. PARP inhibition induces Akt-mediated cytoprotective effects through the formation of a mitochondria-targeted phospho-ATM-NEMO-Akt-mTOR signalosome. Biochem Pharmacol. 2019;162:98–108. 10.1016/j.bcp.2018.10.005. [DOI] [PubMed] [Google Scholar]
  • 214.Miyamoto S, Murphy AN, Brown JH. Akt mediates mitochondrial protection in cardiomyocytes through phosphorylation of mitochondrial hexokinase-II. Cell Death Differ. 2008;15:521–9. 10.1038/sj.cdd.4402285. [DOI] [PubMed]
  • 215.Gallyas F, Jr., Sumegi B. Mitochondrial Protection by PARP Inhibition. Int J Mol Sci. 2020; 21:2767.10.3390/ijms21082767. [DOI] [PMC free article] [PubMed]
  • 216.D'Ambrosio C, Erriquez J, Capellero S, Cignetto S, Alvaro M, Ciamporcero E, et al. Cancer Cells Haploinsufficient for ATM Are Sensitized to PARP Inhibitors by MET Inhibition. Int J Mol Sci. 2022; 23:5770.10.3390/ijms23105770. [DOI] [PMC free article] [PubMed]
  • 217.Kim D, Nam HJ. PARP inhibitors: clinical limitations and recent attempts to overcome them. Int J Mol Sci. 2022;23:8412. 10.3390/ijms23158412. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Luo L, Keyomarsi K. PARP inhibitors as single agents and in combination therapy: the most promising treatment strategies in clinical trials for BRCA-mutant ovarian and triple-negative breast cancers. Expert Opin Investig Drugs. 2022; 31:607–631.10.1080/13543784.2022.2067527. [DOI] [PMC free article] [PubMed]
  • 219.Singh H, Singh AK, Kumar A, Kumar P. Patents on PARP-1 Inhibitors for the Management of Cancer from 2017–2023. Recent Pat Anticancer Drug Discov. 2024; 1:1–13.10.2174/0115748928315021240603073902. [DOI] [PubMed]
  • 220.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. 10.1158/2159-8290.CD-18-1218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.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. 10.1158/2159-8290.CD-18-1020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Doig CL, Lavery GG. PARP inhibitors: staying on target? Cell Chem Biol. 2016;23:1442–3. 10.1016/j.chembiol.2016.12.003. [DOI] [PubMed] [Google Scholar]
  • 223.Knezevic CE, Wright G, Rix LLR, Kim W, Kuenzi BM, Luo Y, et al. Proteome-wide Profiling of Clinical PARP Inhibitors Reveals Compound-Specific Secondary Targets. Cell Chem Biol. 2016; 23:1490–1503.10.1016/j.chembiol.2016.10.011. [DOI] [PMC free article] [PubMed]
  • 224.Kim H, Xu H, George E, Hallberg D, Kumar S, Jagannathan V, et al. Combining PARP with ATR inhibition overcomes PARP inhibitor and platinum resistance in ovarian cancer models. Nat Commun. 2020;11:3726. 10.1038/s41467-020-17127-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Mueller S, Haas-Kogan DA. WEE1 kinase as a target for cancer therapy. J Clin Oncol. 2015;33:3485–7. 10.1200/JCO.2015.62.2290. [DOI] [PubMed] [Google Scholar]
  • 226.Karnak D, Engelke CG, Parsels LA, Kausar T, Wei D, Robertson JR, et al. Combined inhibition of Wee1 and PARP1/2 for radiosensitization in pancreatic cancer. Clin Cancer Res. 2014;20:5085–96. 10.1158/1078-0432.CCR-14-1038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Mateo J, Lord CJ, Serra V, Tutt A, Balmana J, Castroviejo-Bermejo M, et al. A decade of clinical development of PARP inhibitors in perspective. Ann Oncol. 2019;30:1437–47. 10.1093/annonc/mdz192. [DOI] [PMC free article] [PubMed]
  • 228.Tutt A, Robson M, Garber JE, Domchek SM, Audeh MW, Weitzel JN, et al. Oral poly(ADP-1551 ribose) polymerase inhibitor olaparib in patients with BRCA1 or BRCA2 mutations and 1552 advanced breast cancer: a proof-of-concept trial. Lancet. 2010;376:235–44. 10.1016/S0140-6736(10)60892-6. [DOI] [PubMed]
  • 229.Domchek SM, Postel-Vinay S, Im SA, Park YH, Delord JP, Italiano A, et al. Olaparib and 1555 durvalumab in patients with germline BRCA-mutated metastatic breast cancer 1556 (MEDIOLA): an open-label, multicentre, phase 1/2, basket study. Lancet Oncol. 2020;21:1155–64. 10.1016/S1470-2045(20)30324-7. [DOI] [PubMed]
  • 230.Maiorano BA, Catalano M, Maiorano MFP, Signori A, Loizzi V, Cormio G, et al. 1559 Hematological toxicity of parp inhibitors in solid tumors: a systematic review and safety 1560 meta-analysis. Cancer Metastasis Rev. 2025;44:65. 10.1007/s10555-025-156110283-1. [DOI] [PMC free article] [PubMed]
  • 231.Moore DC, Ringley JT, Patel J. Rucaparib: A Poly(ADP-Ribose) Polymerase Inhibitor for 1563 BRCA-Mutated Relapsed Ovarian Cancer. J Pharm Pract. 2019;32:219–24. 10.1177/0897190017743131. [DOI] [PubMed]
  • 232.Cetin B, Wabl CA, Gumusay O. The DNA damaging revolution. Crit Rev Oncol Hematol. 2020;156:103117. 10.1016/j.critrevonc.2020.103117. [DOI] [PubMed]
  • 233. Kristeleit R, Leary A, Oaknin A, Redondo A, George A, Chui S, et al. PARP inhibition with 1568 rucaparib alone followed by combination with atezolizumab: Phase Ib COUPLET clinical 1569 study in advanced gynaecological and triple-negative breast cancers. Br J Cancer. 2024;131:820–31. 10.1038/s41416-024-02776-7. [DOI] [PMC free article] [PubMed]
  • 234. Loehr A, Hussain A, Patnaik A, Bryce AH, Castellano D, Font A, et al. Emergence of BRCA 1572 Reversion Mutations in Patients with Metastatic Castration-resistant Prostate Cancer After 1573 Treatment with Rucaparib. Eur Urol. 2023;83:200–9. 10.1016/j.eururo.2022.09.010. [DOI] [PMC free article] [PubMed]
  • 235. Wang L, Wang Q, Xu Y, Cui M, Han L. Advances in the Treatment of Ovarian Cancer Using 1519 PARP Inhibitors and the Underlying Mechanism of Resistance. Curr Drug Targets. 2020;21:167–78. 10.2174/1389450120666190925123507. [DOI] [PubMed]
  • 236.Li N, Zhu J, Yin R, Wang J, Pan L, Kong B, et al. Treatment With Niraparib Maintenance 1522 Therapy in Patients With Newly Diagnosed Advanced Ovarian Cancer: A Phase 3 1523 Randomized Clinical Trial. JAMA Oncol. 2023;9:1230–7. 10.1001/jamaoncol.2023.2283. [DOI] [PMC free article] [PubMed]
  • 237. Chi KN, Fleshner N, Chiuri VE, Van Bruwaene S, Hafron J, McNeel DG, et al. Niraparib with 1576 Abiraterone Acetate and Prednisone for Metastatic Castration-Resistant Prostate Cancer: 1577 Phase II QUEST Study Results. Oncologist. 2023;28:e309–12. 10.1093/oncolo/oyad008. [DOI] [PMC free article] [PubMed]
  • 238.Roberts HN, Maurice-Dror C, Chi KN. Combination niraparib and abiraterone for HRR-1580 altered metastatic castration-resistant prostate cancer. Future Oncol. 2025;21:201–11. 10.1080/14796694.2024.2442900. [DOI] [PMC free article] [PubMed]
  • 239.Hobbs EA, Litton JK, Yap TA. Development of the PARP inhibitor talazoparib for the 1583 treatment of advanced BRCA1 and BRCA2 mutated breast cancer. Expert Opin 1584 Pharmacother. 2021;22:1825–37. 10.1080/14656566.2021.1952181. [DOI] [PMC free article] [PubMed]
  • 240.Mittra A, Coyne G, Zlott J, Kummar S, Meehan R, Rubinstein L, et al. Pharmacodynamic 1586 effects of the PARP inhibitor talazoparib (MDV3800, BMN 673) in patients with BRCA-mutated advanced solid tumors. Cancer Chemother Pharmacol. 2024;93:177–89. 10.1007/s00280-023-04600-0. [DOI] [PMC free article] [PubMed]
  • 241.Mo W, Liu Q, Lin CC, Dai H, Peng Y, Liang Y, et al. mTOR Inhibitors Suppress Homologous 1590 Recombination Repair and Synergize with PARP Inhibitors via Regulating SUV39H1 in 1591 BRCA-Proficient Triple-Negative Breast Cancer. Clin Cancer Res. 2016;22:1699–712. 10.1158/1078-0432.CCR-15-1772. [DOI] [PMC free article] [PubMed]
  • 242.Thakur A, Chu YH, Rao NV, Mathew J, Grewal AS, Prabakaran P, et al. Leveraging a 1594 rationally designed veliparib-based anilide eliciting anti-leukemic effects for the design of 1595 pH-responsive polymer nanoformulation. Eur J Med Chem. 2024;273:116507. 10.1016/j.ejmech.2024.116507. [DOI] [PubMed]
  • 243. Cai SX, Ma N, Wang X, Guo M, Jiang Y, Tian YE. The Discovery of a Potent PARP1 Inhibitor 1545 Senaparib. Mol Cancer Ther. 2025;24:47–55. 10.1158/1535-7163.MCT-23-15460625. [DOI] [PubMed]
  • 244. Banerjee S, Moore KN, Colombo N, Scambia G, Kim BG, Oaknin A, et al. Maintenance olaparib for patients with newly diagnosed advanced ovarian cancer and a BRCA mutation 1510 (SOLO1/GOG 3004): 5-year follow-up of a randomised, double-blind, placebo-controlled, 1511 phase 3 trial. Lancet Oncol. 2021;22:1721–31. 10.1016/S1470-15122045(21)00531-3. [DOI] [PubMed]
  • 245.Cook SA, Tinker AV. PARP Inhibitors and the Evolving Landscape of Ovarian Cancer 1514 Management: A Review. BioDrugs. 2019;33:255–73. 10.1007/s40259-1515019-00347-4. [DOI] [PubMed]
  • 246.O'Malley DM, Krivak TC, Kabil N, Munley J, Moore KN. PARP Inhibitors in Ovarian Cancer: 1517 A Review. Target Oncol. 2023;18:471–503. 10.1007/s11523-023-00970-w. [DOI] [PMC free article] [PubMed]
  • 247.Monk BJ, Parkinson C, Lim MC, O'Malley DM, Oaknin A, Wilson MK, et al. A Randomized, 1526 Phase III Trial to Evaluate Rucaparib Monotherapy as Maintenance Treatment in Patients 1527 With Newly Diagnosed Ovarian Cancer (ATHENA-MONO/GOG-3020/ENGOT-ov45). J 1528 Clin Oncol. 2022;40:3952–64. 10.1200/JCO.22.01003. [DOI] [PMC free article] [PubMed]
  • 248.Sharma N, Bhati A, Aggarwal S, Shah K, Dewangan HK. PARP Pioneers: Using BRCA1/2 1530 Mutation-targeted Inhibition to Revolutionize Breast Cancer Treatment. Curr Pharm Des. 1531 2025;31:663–73. 10.2174/0113816128322894241004051814. [DOI] [PubMed]
  • 249.Saad F, Armstrong AJ, Shore N, George DJ, Oya M, Sugimoto M, et al. Olaparib Monotherapy 1533 or in Combination with Abiraterone for the Treatment of Patients with Metastatic 1534 Castration-Resistant Prostate Cancer (mCRPC) and a BRCA Mutation. Target Oncol. 2025;20:445–66. 10.1007/s11523-025-01146-4. [DOI] [PMC free article] [PubMed]
  • 250.Chi KN, Sandhu S, Smith MR, Attard G, Saad M, Olmos D, et al. Niraparib plus abiraterone 1537 acetate with prednisone in patients with metastatic castration-resistant prostate cancer and 1538 homologous recombination repair gene alterations: second interim analysis of the 1539 randomized phase III MAGNITUDE trial. Ann Oncol. 2023;34:772–82. 10.1016/j.annonc.2023.06.009. [DOI] [PMC free article] [PubMed]
  • 251.Golan T, Hammel P, Reni M, Van Cutsem E, Macarulla T, Hall MJ, et al. Maintenance Olaparib 1542 for Germline BRCA-Mutated Metastatic Pancreatic Cancer. N Engl J Med. 2019;381:317–27. 10.1056/NEJMoa1903387. [DOI] [PMC free article] [PubMed]
  • 252. Murai K, Kodama T, Hikita H, Shimoda A, Fukuoka M, Fukutomi K, et al. Inhibition of 1598 nonhomologous end joining-mediated DNA repair enhances anti-HBV CRISPR therapy. 1599 Hepatol Commun. 2022;6:2474–87. 10.1002/hep4.2014. [DOI] [PMC free article] [PubMed]
  • 253.Stone NE, Jaramillo SA, Jones AN, Vazquez AJ, Martz M, Versluis LM, et al. Stenoparib, an Inhibitor of Cellular Poly(ADP-Ribose) Polymerase, Blocks Replication of the SARS-CoV-2 and HCoV-NL63 Human Coronaviruses In Vitro. mBio. 2021; 12:e03495–03420.10.1128/mBio.03495-20. [DOI] [PMC free article] [PubMed]
  • 254. Ge Y, Tian T, Huang S, Wan F, Li J, Li S, et al. An integrative drug repositioning framework 1613 discovered a potential therapeutic agent targeting COVID-19. Signal Transduct Target Ther. 2021;6:165. 10.1038/s41392-021-00568-6. [DOI] [PMC free article] [PubMed]
  • 255.Zhao D, Xu W, Zhang X, Wang X, Ge Y, Yuan E, et al. Understanding the phase separation 1616 characteristics of nucleocapsid protein provides a new therapeutic opportunity against 1617 SARS-CoV-2. Protein Cell. 2021;12:734–40. 10.1007/s13238-021-00832-1618z. [DOI] [PMC free article] [PubMed]
  • 256. Tibebe H, Marquez D, McGraw A, Gagliardi S, Sullivan C, Hillmer G, et al. Targeting Latent 1635 HIV Reservoirs: Effectiveness of Combination Therapy with HDAC and PARP Inhibitors. 1636 Viruses. 2025;17. 10.3390/v17030400. [DOI] [PMC free article] [PubMed]
  • 257.Rom S, Reichenbach NL, Dykstra H, Persidsky Y. The dual action of poly(ADP-ribose) 1644 polymerase -1 (PARP-1) inhibition in HIV-1 infection: HIV-1 LTR inhibition and 1645 diminution in Rho GTPase activity. Front Microbiol. 2015;6:878. 10.3389/fmicb.2015.00878. [DOI] [PMC free article] [PubMed]
  • 258.Arena A, Gilardini Montani MS, Romeo MA, Benedetti R, Gaeta A, Cirone M. DNA damage 1631 triggers an interplay between wtp53 and c-Myc affecting lymphoma cell proliferation and 1632 Kaposi sarcoma herpesvirus replication. Biochim Biophys Acta Mol Cell Res. 2022;1869:119168. 10.1016/j.bbamcr.2021.119168. [DOI] [PubMed]
  • 259.Liu W, Ren X, Wang Q, Zhang Y, Du J. Pharmacological inhibition of poly (ADP-ribose) 1610 polymerase by olaparib ameliorates influenza-virus-induced pneumonia in mice. Eur J Clin 1611 Microbiol Infect Dis. 2021;40:159–67. 10.1007/s10096-020-04020-5. [DOI] [PMC free article] [PubMed]
  • 260.Tomao F, Santangelo G, Musacchio L, Di Donato V, Fischetti M, Giancotti A, et al. Targeting 1648 cervical cancer: Is there a role for poly (ADP-ribose) polymerase inhibition? J Cell Physiol. 2020;235:5050–8. 10.1002/jcp.29440. [DOI] [PubMed]
  • 261.Elayapillai SP, Dogra S, Lausen J, Parker M, Kennedy A, Benbrook DM, et al. ATR inhibition 1651 increases reliance on PARP-mediated DNA repair revealing an improved therapeutic 1652 strategy for cervical cancer. Gynecol Oncol. 2024;191:182–93. 10.1016/j.ygyno.2024.10.009. [DOI] [PMC free article] [PubMed]
  • 262.Moutafi M, Economopoulou P, Rimm D, Psyrri A. PARP inhibitors in head and neck cancer: 1655 Molecular mechanisms, preclinical and clinical data. Oral Oncol. 2021;117:105292. 10.1016/j.oraloncology.2021.105292. [DOI] [PubMed]
  • 263.Mann M, Singh VP, Kumar L. Cervical cancer: a tale from HPV infection to PARP inhibitors. 1605 Genes Dis. 2023;10:1445–56. 10.1016/j.gendis.2022.09.014. [DOI] [PMC free article] [PubMed]
  • 264.Citro S, Ghiani L, Doni M, Miccolo C, Tagliabue M, Ansarin M, et al. HPV-mediated PARP1 1607 regulation and drug sensitization in head and neck cancer. Oral Oncol. 2025;165:107307. 10.1016/j.oraloncology.2025.107307. [DOI] [PubMed]
  • 265.Dorna D, Kleszcz R, Paluszczak J. Triple Combinations of Histone Lysine Demethylase 1620 Inhibitors with PARP1 Inhibitor-Olaparib and Cisplatin Lead to Enhanced Cytotoxic 1621 Effects in Head and Neck Cancer Cells. Biomedicines. 2024;12. 10.3390/biomedicines12061359. [DOI] [PMC free article] [PubMed]
  • 266.M IJ, Mei X, Scutigliani EM, Rodermond HM, van Bochove GGW, Krawczyk PM, et al. 1624 Addition of PARP1-inhibition enhances chemoradiotherapy and thermoradiotherapy when 1625 treating cervical cancer in an in vivo mouse model. Int J Hyperthermia. 2025;42:2450514. 10.1080/02656736.2025.2450514. [DOI] [PubMed]
  • 267. M IJ, van Bochove GGW, Whitton D, Winiarczyk R, Honhoff C, Rodermond H, et al. PARP1-1628 Inhibition Sensitizes Cervical Cancer Cell Lines for Chemoradiation and Thermoradiation. 1629 Cancers (Basel). 2021;13. 10.3390/cancers13092092. [DOI] [PMC free article] [PubMed]
  • 268. Karlberg H, Tan YJ, Mirazimi A. Crimean-Congo haemorrhagic fever replication interplays 1638 with regulation mechanisms of apoptosis. J Gen Virol. 2015;96:538–46. 10.1099/jgv.0.000011. [DOI] [PubMed]
  • 269. Ahsan NA, Sampey GC, Lepene B, Akpamagbo Y, Barclay RA, Iordanskiy S, et al. Presence 1641 of Viral RNA and Proteins in Exosomes from Cellular Clones Resistant to Rift Valley Fever 1642 Virus Infection. Front Microbiol. 2016;7:139. 10.3389/fmicb.2016.00139. [DOI] [PMC free article] [PubMed]
  • 270.Singla S, Kumar V, Jena G. 3-aminobenzamide protects against colitis associated diabetes mellitus in male BALB/c mice: role of PARP-1, NLRP3, SIRT-1, AMPK. Biochimie. 2023;211:96–109. 10.1016/j.biochi.2023.03.009. [DOI] [PubMed] [Google Scholar]
  • 271.Santinelli-Pestana DV, Aikawa E, Singh SA, Aikawa M. PARPs and ADP-Ribosylation in Chronic Inflammation: A Focus on Macrophages. Pathogens. 2023; 12.10.3390/pathogens12070964. [DOI] [PMC free article] [PubMed]
  • 272.Zhou J, Du T, Wang X, Yao H, Deng J, Li Y, et al. Discovery of Quinazoline-2,4(1H,3H)-dione Derivatives Containing a Piperizinone Moiety as Potent PARP-1/2 Inhibitors horizontal line Design, Synthesis, In Vivo Antitumor Activity, and X-ray Crystal Structure Analysis. J Med Chem. 2023; 66:14095–14115.10.1021/acs.jmedchem.3c01152. [DOI] [PubMed]
  • 273.Kim YN, Shim Y, Seo J, Choi Z, Lee YJ, Shin S, et al. Investigation of PARP inhibitor resistance based on serially collected circulating tumor DNA in patients with BRCA-mutated ovarian cancer. Clin Cancer Res. 2023;29:2725–34. 10.1158/1078-0432.CCR-22-3715. [DOI] [PubMed] [Google Scholar]
  • 274.Zhan Z, Zhang J, Liang H, Wang C, Hong L, Liu W. Kat6a condensates impair PARP1 trapping of PARP inhibitors in ovarian cancer. Adv Sci. 2024;11:e2400140. 10.1002/advs.202400140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 275.Kanev PB, Varhoshkova S, Georgieva I, Lukarska M, Kirova D, Danovski G, et al. A unified mechanism for PARP inhibitor-induced PARP1 chromatin retention at DNA damage sites in living cells. Cell Rep. 2024;43:114234. 10.1016/j.celrep.2024.114234. [DOI] [PubMed] [Google Scholar]
  • 276.Moon SH, Park NS, Noh MH, Kim YS, Cheong SH, Hur DY. Olaparib-induced apoptosis through EBNA1-ATR-p38 MAPK signaling pathway in Epstein-Barr virus-positive gastric cancer cells. Anticancer Res. 2022;42:555–63. 10.21873/anticanres.15513. [DOI] [PubMed] [Google Scholar]
  • 277.Chou J, Robinson TM, Egusa EA, Lodha R, Zhang M, Badura M, et al. Synthetic lethal targeting of CDK12-deficient prostate cancer with PARP inhibitors. Clin Cancer Res. 2024;30:5445–58. 10.1158/1078-0432.CCR-23-3785. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 278.Rajawat J, Chandra A. Role of poly(ADP-ribose) polymerase (PARP1) in viral infection and its implication in SARS-CoV-2 pathogenesis. Curr Drug Targets. 2021;22:1477–84. 10.2174/1389450122666210120142746. [DOI] [PubMed] [Google Scholar]
  • 279.Gonzalez-Martin A, Pothuri B, Vergote I, DePont Christensen R, Graybill W, Mirza MR, et al. Niraparib in patients with newly diagnosed advanced ovarian cancer. N Engl J Med. 2019;381:2391–402. 10.1056/NEJMoa1910962. [DOI] [PubMed] [Google Scholar]
  • 280.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. 10.1038/s41467-022-30568-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 281.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; 23:3711–3720.10.1158/1078-0432.CCR-16-3215. [DOI] [PMC free article] [PubMed]
  • 282.Serra V, Wang AT, Castroviejo-Bermejo M, Polanska UM, Palafox M, Herencia-Ropero A, et al. Identification of a molecularly-defined subset of breast and ovarian cancer models that respond to WEE1 or ATR inhibition, overcoming PARP inhibitor resistance. Clin Cancer Res. 2022;28:4536–50. 10.1158/1078-0432.CCR-22-0568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 283.Litton JK, Rugo HS, Ettl J, Hurvitz SA, Goncalves A, Lee KH, et al. Talazoparib in patients with advanced breast cancer and a germline BRCA mutation. N Engl J Med. 2018;379:753–63. 10.1056/NEJMoa1802905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Menissier de Murcia J, Ricoul M, Tartier L, Niedergang C, Huber A, Dantzer F, et al. Functional interaction between PARP-1 and PARP-2 in chromosome stability and embryonic development in mouse. EMBO J. 2003; 22:2255–2263.10.1093/emboj/cdg206. [DOI] [PMC free article] [PubMed]
  • 285.Wu X, Zhu J, Yin R, Yang J, Liu J, Wang J, et al. Niraparib maintenance therapy using an individualised starting dose in patients with platinum-sensitive recurrent ovarian cancer (NORA): final overall survival analysis of a phase 3 randomised, placebo-controlled trial. EClinicalMedicine. 2024;72:102629. 10.1016/j.eclinm.2024.102629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 286.Vulsteke C, Chambers SK, Perez MJR, Chan JK, Raaschou-Jensen N, Zhuo Y, et al. Tolerability of the niraparib individualized starting dose in the PRIMA/ENGOT-OV26/GOG-3012 trial of niraparib first-line maintenance therapy. Eur J Cancer. 2024;208:114157. 10.1016/j.ejca.2024.114157. [DOI] [PubMed] [Google Scholar]
  • 287.Morice PM, Leary A, Dolladille C, Chretien B, Poulain L, Gonzalez-Martin A, et al. Myelodysplastic syndrome and acute myeloid leukaemia in patients treated with PARP inhibitors: a safety meta-analysis of randomised controlled trials and a retrospective study of the WHO pharmacovigilance database. Lancet Haematol. 2021;8:e122–34. 10.1016/S2352-3026(20)30360-4. [DOI] [PubMed] [Google Scholar]
  • 288.Kovacs D, Vantus VB, Vamos E, Kalman N, Schicho R, Gallyas F, et al. Olaparib: a clinically applied PARP inhibitor protects from experimental Crohn’s disease and maintains barrier integrity by improving bioenergetics through rescuing glycolysis in colonic epithelial cells. Oxid Med Cell Longev. 2021;2021:7308897. 10.1155/2021/7308897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 289.Jasim SA, Gupta J, Uthirapathy S, R R, Kazmi SW, Rizaev JA, et al. A comprehensive review of immunotherapy in gastrointestinal tumors with a focus on the role of combination therapy with PARP inhibitors. Naunyn Schmiedebergs Arch Pharmacol. 2025. 10.1007/s00210-025-04336-z. [DOI] [PubMed] [Google Scholar]
  • 290.Yang C, Song X, Sun H, Chen X, Liu C, Chen M. Cardiovascular adverse events associated with PARP inhibitors for ovarian cancer: a real world study (RWS) with Bayesian disproportional analysis based on the FDA adverse event reporting system (FAERS). Expert Opin Drug Saf. 2025;24:1039–46. 10.1080/14740338.2024.2390640. [DOI] [PubMed] [Google Scholar]
  • 291.Adrianto N, Mangkuliguna G, Tandiono EJ, Sibarani CNR. Efficacy and safety of rucaparib in patients with recurrent high-grade ovarian carcinoma: a systematic review and meta-analysis. Taiwan J Obstet Gynecol. 2024;63:601–9. 10.1016/j.tjog.2024.05.020. [DOI] [PubMed] [Google Scholar]
  • 292.Stanislawiak-Rudowicz J, Karbownik A, Szkutnik-Fiedler D, Otto F, Grabowski T, Wolc A, et al. Bidirectional pharmacokinetic drug interactions between olaparib and metformin. Cancer Chemother Pharmacol. 2024; 93:79–88.10.1007/s00280-023-04591-y. [DOI] [PMC free article] [PubMed]
  • 293.El-Gammacy TM, Shinkar DM, Mohamed NR, Al-Halag AR. Serum cystatin C as an early predictor of acute kidney injury in preterm neonates with respiratory distress syndrome. Scand J Clin Lab Invest. 2018;78:352–7. 10.1080/00365513.2018.1472803. [DOI] [PubMed] [Google Scholar]
  • 294.Illuzzi G, Staniszewska AD, Gill SJ, Pike A, McWilliams L, Critchlow SE, et al. Preclinical characterization of AZD5305, a next-generation, highly selective PARP1 inhibitor and trapper. Clin Cancer Res. 2022;28:4724–36. 10.1158/1078-0432.CCR-22-0301. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 295.Peng N, Chen D, Yang M, Zeng C, Li W. PARP inhibitors for HRR-deficient metastatic castration-resistant prostate cancer: mechanisms and clinical strategies. Eur J Pharm Sci. 2025;212:107193. 10.1016/j.ejps.2025.107193. [DOI] [PubMed] [Google Scholar]
  • 296.Thorsell AG, Ekblad T, Karlberg T, Low M, Pinto AF, Tresaugues L, et al. Structural Basis for Potency and Promiscuity in Poly(ADP-ribose) Polymerase (PARP) and Tankyrase Inhibitors. J Med Chem. 2017; 60:1262–1271.10.1021/acs.jmedchem.6b00990. [DOI] [PMC free article] [PubMed]
  • 297.Li G, Lin SS, Yu ZL, Wu XH, Liu JW, Tu GH, et al. A PARP1 PROTAC as a novel strategy against PARP inhibitor resistance via promotion of ferroptosis in p53-positive breast cancer. Biochem Pharmacol. 2022;206:115329. 10.1016/j.bcp.2022.115329. [DOI] [PubMed] [Google Scholar]
  • 298.McCann KE. Poly-ADP-ribosyl-polymerase inhibitor resistance mechanisms and their therapeutic implications. Curr Opin Obstet Gynecol. 2019;31:12–7. 10.1097/GCO.0000000000000517. [DOI] [PubMed] [Google Scholar]
  • 299.Francica P, Rottenberg S. Mechanisms of PARP inhibitor resistance in cancer and insights into the DNA damage response. Genome Med. 2018;10(1):101. 10.1186/s13073-018-0612-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 300.Jain A, Barge A, Parris CN. Combination strategies with PARP inhibitors in BRCA-mutated triple-negative breast cancer: overcoming resistance mechanisms. Oncogene. 2025;44:193–207. 10.1038/s41388-024-03227-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 301.Tang Q, Wang X, Wang H, Zhong L, Zou D. Advances in ATM, ATR, WEE1, and CHK1/2 inhibitors in the treatment of PARP inhibitor-resistant ovarian cancer. Cancer Biol Med. 2024;20:915–21. 10.20892/j.issn.2095-3941.2023.0260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 302.Baldan F, Mio C, Lavarone E, Di Loreto C, Puglisi F, Damante G, et al. Epigenetic bivalent marking is permissive to the synergy of HDAC and PARP inhibitors on TXNIP expression in breast cancer cells. Oncol Rep. 2015;33:2199–206. 10.3892/or.2015.3873. [DOI] [PubMed] [Google Scholar]
  • 303.Lovato A, Panasci L, Witcher M. Is there an epigenetic component underlying the resistance of triple-negative breast cancers to PARP inhibitors? Front Pharmacol. 2012;3:202. 10.3389/fphar.2012.00202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 304.Gupta T, Vinayak S, Telli M. Emerging strategies: PARP inhibitors in combination with immune checkpoint blockade in BRCA1 and BRCA2 mutation-associated and triple-negative breast cancer. Breast Cancer Res Treat. 2023;197:51–6. 10.1007/s10549-022-06780-4. [DOI] [PubMed] [Google Scholar]
  • 305.Ni J, Cheng X, Zhou R, Zhao Q, Xu X, Guo W, et al. Adverse events as a potential clinical marker of antitumor efficacy in ovarian cancer patients treated with poly ADP-ribose polymerase inhibitor. Front Oncol. 2021;11:724620. 10.3389/fonc.2021.724620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 306.Shen Y, Aoyagi-Scharber M, Wang B. Trapping poly(ADP-ribose) polymerase. J Pharmacol Exp Ther. 2015;353:446–57. 10.1124/jpet.114.222448. [DOI] [PubMed] [Google Scholar]
  • 307.Vida A, Marton J, Miko E, Bai P. Metabolic roles of poly(ADP-ribose) polymerases. Semin Cell Dev Biol. 2017;63:135–43. 10.1016/j.semcdb.2016.12.009. [DOI] [PubMed] [Google Scholar]
  • 308.Li N, Wang Y, Deng W, Lin SH. Poly (ADP-ribose) polymerases (PARPs) and PARP inhibitor-targeted therapeutics. Anticancer Agents Med Chem. 2019;19:206–12. 10.2174/1871520618666181109164645. [DOI] [PubMed] [Google Scholar]
  • 309.Woon EC, Threadgill MD. Poly(ADP-ribose)polymerase inhibition - where now? Curr Med Chem. 2005;12:2373–92. 10.2174/0929867054864778. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

No new data were generated or analyzed in this study. This work is a review study and does not rely on any external data.


Articles from Molecular Biomedicine are provided here courtesy of Springer

RESOURCES