Abstract
The poly(ADP-ribose) polymerase (PARP) superfamily of enzymes catalyses the ADP-ribosylation (ADPr) of target proteins by using nicotinamide adenine dinucleotide (NAD+) as a donor. ADPr reactions occur either in the form of attachment of a single ADP-ribose nucleotide unit on target proteins or in the form of ADP-ribose chains, with the latter called poly(ADP-ribosyl)ation. PARPs regulate many cellular processes, including the maintenance of genome stability and signal transduction. In this review, we focus on the PARP family members that possess the ability to modify proteins by poly(ADP-ribosyl)ation, namely PARP1, PARP2, Tankyrase-1, and Tankyrase-2. Here, we detail the cellular functions of PARP1 and PARP2 in the regulation of DNA damage response and describe the function of Tankyrases in Wnt-mediated signal transduction. Furthermore, we discuss how the understanding of these pathways has provided some major breakthroughs in the treatment of human cancer.
Keywords: adenosine diphosphate ribose, cancer, DNA damage response, PARPs, signalling
Introduction
ADP-ribosylation (ADPr) is a post-translational modification (PTM) conserved in bacteria, viruses, and in the majority of eukaryotes [1–3]. ADPr is mainly catalysed by the ADP-ribosyltransferase (ART) superfamily of proteins, which transfer a single or multiple ADP-ribose unit(s) from nicotinamide adenine dinucleotide (NAD+) onto target substrates [2–4]. Substrates of ADPr can be either proteins or nucleic acids [2–4]. ARTs are widespread in different organisms and regulate diverse cellular processes as the DNA damage response (DDR), transcription, RNA metabolism and antiviral response, cell division, unfolded protein response (UPR), stress granule formation, metabolism, and cell death to cite a few [5–18].
Similarly to other PTMs, ADPr operates by altering the function/localisation/stability of targets. Additionally, ADPr acts as a scaffold for the recruitment of effector proteins (‘readers’), which are able to recognise and bind the modification through specialised protein domains, such as the macrodomain, PBZ, and WWE domains [19–23].
ADPr is a ‘reversible’ PTM; ART activity is indeed counteracted by specific hydrolases (also called ‘erasers’). Two evolutionarily unrelated protein domains are known to reverse ART's activity; catalytic macrodomains (e.g. in PARG, MacroD1, MacroD2, and TARG1), and the ARH superfamily proteins (e.g. ARH1 and ARH3) [2,23–25]. Although with possible different substrate specificities, the existence of multiple erasers of ADPr in different cellular compartments (such as the nucleus, cytosol, and membranous organelles for instance mitochondria) ensures the capability of the cells to appropriately limit and fine-tune the ADPr signal when needed [26–32]. Additionally, three classes of enzymes have been described to cleave protein ADPr in a non-canonical manner; some members of NUDIX family (such as the mammal NUDT16) [33,34], members of the Ectonucleotide pyrophosphatase/phosphodiesterase family (such as the Phosphodiesterase I found in the poison glands of rattlesnakes and the vertebrate ENPP1) [35,36], and the Legionella pneumophila SdeA protein [37]. Those enzymes hydrolyse the ADPr phosphodiester bond into adenosine monophosphate (AMP) and a ribose-5′-phosphate moiety linked to the substrate molecule, a protein modification known as phosphoribosylation [2,33–37].
The best-studied ART family are the poly(ADP-ribose) polymerases (PARPs), also called diphtheria toxin-like ADP-ribosyltransferases (ARTDs) [38–40]. In humans, there are 18 genes encoding PARP catalytic domain (CAT) containing proteins [38]. A slightly different classification has been proposed by Vyas et al. [41] that does not include the diverged TpT1 member (see below).
Regarding the amino acid composition of the CAT domain, PARP1 (also called ARTD1), PARP2 (ARTD2), PARP3 (ARTD3), Tankyrase-1 (PARP5a/ARTD5), and Tankyrase-2 PARP5b/ARTD6 are characterised by a histidine–tyrosine–glutamate (HYE) triad in the catalytic pocket [41]. PARP6 (ARTD17), PARP7 (ARTD14), PARP8 (ARTD16), PARP10 (ARTD10), PARP11 (ARTD11), and PARP12 (ARTD12) are characterised by a histidine–tyrosine–isoleucine (HYI) triad [41]. PARP16 (ARTD16) is characterised by a histidine–tyrosine–tyrosine (HYI) motif [41]. PARP14 (ARTD8) and PARP15 (ARTD7) contain a histidine–tyrosine–leucine (HYL) triad [41]. PARP9 (ARTD9) holds a glutamine–tyrosine–threonine (QYT) [41]. The isoform 1 of PARP13 (ARTD13) is characterised by a tyrosine–tyrosine–valine (YYV) motif, while the CAT domain is absent in the isoform 2 of PARP13 (PARP13.2) [41]. Additionally, the highly divergent TpT1 (ARTD18) is also sometimes classified as a PARP-like protein. This protein contains a histidine–histidine–valine (HHV) triad in the CAT and in yeast catalyses a NAD+-dependent dephosphorylation of tRNA-splicing intermediates, generating ADPr-1-phopshate through a cyclic intermediate [2,38,42]. Most of the PARPs efficiently catalyse the transfer of ADP-ribose onto proteins, albeit with different specificities. PARPs usually transfer ADP-ribose onto aspartic/glutamic acid (via ester linkages; here named as Asp/Glu-ADPr) or serine (via O-glycosylation; here named as Ser-ADPr) residues on target molecules [3,43–45].
Members of the PARP family can be also classified based on their ability to perform mono(ADP-ribosyl)ation (MARylation) or poly(ADP-ribosyl)ation (PARylation) [41]; in the latter case, the ADP-ribose units are linked together through glycosidic ribose–ribose 1″ → 2′ bonds [46]. PARPs capable of synthesis of PARylation include the DNA damage-inducible PARP1 and PARP2, which are known for their ability to produce long (up to 200 ADP-ribose units) and heterogeneous chains of branched PAR [41,46,47], and Tankyrase-1 and Tankyrase-2 [41]. It should be noted that Tankyrases synthesise PAR polymers with an average chain length of 20 units and no detectable branching [41,48]. Tankyrases are involved in multiple cellular processes, such as telomere length maintenance, mitosis, and Wnt signalling regulation [3,4]. Although initially described as able to catalyse ADP-ribose polymers up to 15-mers [49], PARP3 is currently believed to be a MARylating enzyme [3,4,38,41].
In this review, we summarise the recent advances in the understanding of PARPs by focusing on those that carry out PARylation and have functions in genome stability and signal transduction. The targeting of both cellular processes has provided promising strategies for cancer therapy. Indeed, the understanding of the cellular processes regulated by PARPs owes much to the success of PARP inhibitors in preclinical and clinical trials.
Human PARPs
With the exception of TpT1 that appears to act as a RNA phosphotransferase, 17 human PARPs/ARTDs family members have been identified carrying a canonical ART domain [41]. PARPs are located in various cellular compartments and regulate major cellular functions, e.g. DNA damage response, transcription, chromatin structure regulation, UPR, metabolism, mitosis, telomere length maintenance, stress granule formation, antiviral response, and receptor-associated signalling [5–9,11,12,15,16,18].
The better understood PARylating PARPs/ARTDs are the DDR PARPs (PARP1 and PARP2) and Tankyrases 1 and 2 (see extensively below). Conversely, relatively little is known about the mono(ADP-ribosyl)ating (MARylating) PARPs/ARTDs and their physiological function. Recent efforts have sought to assign cellular functions to some MARylating PARPs. For instance, PARP3 is involved in the DDR and mitotic spindle assembly [50]; PARP4 (vPARP or ARTD4) has an unclear function at the mammalian vaults (ribonucleoprotein complexes) and it is possibly involved in antiviral response [51,52]; PARP6 has been proposed to have a role in cell cycle progression and has been associated with the development of colorectal cancer [53]. PARP9 possesses a unique MARylating activity specifically occurring on ubiquitin molecules and it has functions in DDR, transcription in lymphocytes, and antiviral response [52,54–57]. PARP10 is a binding protein and an inhibitor of MYC with inhibitory potential also on the NF-κB signalling pathway [58,59]. Moreover, PARP10 acts as a pro-apoptotic protein [60]. PARP11 is proposed to have a role in nuclear envelope biology [61,62], while PARP12 is a cytosolic protein but preferentially associates with the Golgi apparatus and regulates stress granule assembly, microRNA activity, and antiviral response [5,16,63]; PARP13 has been so far considered a catalytically inactive ART with functions in the assembly of stress granules [5] and in the regulation of microRNA [64], with implications in immunity and cancer [65]. PARP14 has been linked with multiple cellular functions such as survival of B-cells [66], cell migration [67], assembly of stress granules [5], transcription during inflammation processes [68], DDR [69], and antiviral response [52]. PARP15 is also involved in stress granule formation and antiviral response [5,52]. PARP16 is located at the endoplasmic reticulum and regulates UPR [8,70].
DNA repair PARPs and signalling
The main PARPs with a direct function in the DDR are the PARylating PARP1 and PARP2, and the MARylating PARP3 [71]. PARP1 is the founding member of the PARP/ARTD family and the best-studied one. At the N-terminal end, PARP1 contains three DNA-binding zinc finger domains (ZFI: amino acid (aa) 11–89; ZFII: aa 115–199; ZFIII: aa 233–373) [46,72,73]. The central domain of PARP1 contains the BRCA1 C-terminal (BRCT) domain and the tryptophan–glycine–arginine-rich (WGR) domain [74]. At the carboxyl-terminal region, PARP1 contains the PARP homology domain, comprising of the CAT responsible for ADP-ribosylation [75–77]. PARP1, PARP2, and PARP3 share ∼60% amino acid similarity within their catalytic and WGR domains but diverge at their N-termini. PARP2 and PARP3 lack the ZF and BRCT domains and instead possess shorter unstructured N-terminal regions with poorly understood functions. In the CAT domain, PARP1, PARP2, and PARP3 share a conserved structural feature known as helical domain (HD) [75]. The crystal structure of the essential domains of PARP1 in complex with DNA double-strand breaks revealed that the HD acts as an autoinhibitory domain by blocking the access to the NAD+-binding site [77,78]. However, in response to the binding of PARP1 with DNA, the HD rapidly unfolds, thus allowing PARP1 catalytic activity [79]. Outside of the CAT domain is the WGR domain, which is vital for the activation of all the three damage-dependent PARPs (PARP1, PARP2, and PARP3) [74,77]. Structural studies showed that the PARP1 WGR domain makes sequence-independent contacts with both DNA backbone, near the 5′-terminus, and the HD within the CAT domain [77]. As a result of the concomitant interaction with damaged DNA and HD, the WGR transfers the information of binding to DNA to the catalytic portion of PARP1. The intramolecular contact between these two domains triggered by damaged DNA structurally destabilises the HD, leading to the catalytic activation of PARP1 [74].
PARP1 and PARP2 have been recognised as central components of the base excision repair/single-strand break repair process (BER/SSBR) [71,80]. Moreover, PARP1 and PARP2 have been found to be activated by double-strand breaks (DSBs) and required for homologous recombination (HR) at stalled or collapsed replication forks as well as for non-homologous end joining (NHEJ) [71,80].
Upon binding to DNA breaks, the catalytic function of PARP1 is activated to generate extensive poly(ADP-ribose) chains (PAR chains) on proteins in the proximity of DNA damage, among which include DNA repair effectors and histones [73,78,81] (Figure 1A). It has also been proposed that PARP2 binds PARP1-neosynthetised PAR chains at DNA damage sites through its unstructured N-terminal region [82]. This interaction facilitates the subsequent activation of PARP2, which seems mainly responsible for the formation of branched chains of PAR [82]. Interestingly, PARP1, PARP2, as well as PARP3 have been recently reported to directly ADP-ribosylate the DNA breaks via free phosphate groups in cellular response to DNA damage; however, how this modification affects DDR is still unclear [83–85]. Overall, the absence or the inhibition of PARP1 or PARP2 in mice or human confers sensitivity to a variety of DNA damaging agents [6,86]. The double Parp1 and Parp2 knockout mice show embryonic lethality suggesting significant functional redundancies between the two proteins [87].
One of the main targets of PARP1 is the histone core proteins of nucleosomes (H2A, H2B, H3, and H4) as well as the linker histone H1 [43,88–90]. Although it is not completely clear as to the mechanistic role of the highly abundant histone modification, it has been proposed that it allows the opening of chromatin structure and a better accessibility for DDR factors, thus facilitating DNA repair [91,92] (Figure 1A). The most abundant ADPr histone sites are on the specific serine resides on their tails (Ser-ADPr) [90]. Ser-ADPr fully depends on a protein called histone PARylation factor 1 (HPF1), which was identified as a key protein controlling the DNA damage-inducible PARylation of histone proteins [89,90,93] (Figure 1A). HPF1 is a PARP1-binding protein that confers specificity for substrates to PARP1, allowing specific ADPr of serine residues in histones as well as many other factors involved in genome stability [89,90]. The main target residues of ADPr in histones are Ser6 of H2B and Ser10 of H3 [43,89,90]. Unbiased mass spectrometry studies have identified a common Ser-ADPr motif in response to DNA damage. This Ser-ADPr motif is characterised by a lysine (or less frequently an arginine) residue followed by serine, which acts as an acceptor site [89]. These recent results seem to somewhat contradict previous observations, mainly describing ADPr of histones and other DDR proteins on acidic residues [94–97]. While this could be partly due to different cell lines employed and type of stress used to activate the DNA damage, the main reasons lie in the methodology applied to prepare samples and detect the modified amino acids by mass spectrometry. For instance, the majority of studies describing Asp/Glu-ADPr sites in proteins in response to oxidative insults have been conducted by employing the hydroxylamine treatment, a method that enables the sole identification of Asp and Glu modified by ADPr, therefore excluding other types of modification [94].
Importantly, PARP1 PARylates itself (autoPARylation) on multiple acceptor sites that have been characterised by MS in vitro and in vivo [29,33–36,98–101]. AutoPARylation acts as an important scaffold for recruitment of DDR factors [80]. Excessive PARP1 automodification may also induce the release of PARP1 from DNA damage foci [102,103] (Figure 1A). Extensive automodification of PARP1 is suppressed by HPF1 [93]. HPF1 also changes automodification sites in PARP1 from Asp/Glu-ADPr to Ser-ADP, although the biochemical mechanism is still unknown [89,90]. PARP1- and HPF1-dependent Ser-ADPr also occurs on many proteins involved in the maintenance of genome stability [15,100,104], for instance, the high mobility group proteins [89]. Nevertheless, many other DDR proteins may be not regulated by the PARP1-2/HPF1 complex but instead controlled by distinct PARP complexes, perhaps leading to Asp/Glu-ADPr. Indeed, the absence of HPF1 leads to a reduced ADPr of histone proteins and DNA repair factors, and conversely to the enrichment of ADPr on Asp/Glu of PARP1 itself and many other proteins [90]. Further efforts will have to clarify what the functional importance/advantage of having Ser-ADPr instead of Asp/Glu-ADPr in response to DNA damage. Deficiency of HPF1 in cells leads to sensitivity to DNA damaging drugs and PARP inhibitors (PARPi; see below) [90,93].
As noted previously, PARP1 automodification as well as the PAR chains on histone core proteins serves as a scaffold for recruitment of critical DNA repair proteins, such as XRCC1 or different chromatin remodelling proteins, to the DNA break, thus facilitating DNA repair [19,20,105–107]. The turnover of longer chains of PAR after damage largely depends on PARG hydrolase function [108–110]. In cells, PARG is inefficient in cleaving short chains of PAR and is not capable of removing MARylation [24,110,111], which are instead a substrate for ARH3 hydrolase when ADP-ribose is linked to serine residues [90,110] (Figure 1B). Glutamate-linked MARylation is hydrolysed by macrodomain-containing proteins TARG1, MacroD1 and MacroD2 [2,4,29], but is a poor substrate for ARH3 [110]. In contrast, TARG1, MacroD1, and MacroD2 are unable to hydrolyse Ser-ADPr [110].
Modulation of DDR PAR signalling for cancer treatment
The enzymatic PARP1 inhibition by small-molecule analogues of NAD+ results in suppression of both SSB repair and BER and also in trapping PARP1 onto DNA lesions [112–114], which in turn causes the stalling and subsequent collapse of DNA replication forks, further resulting in replication-dependent DNA DSBs and possibly transcription conflicts [115,116]. DSBs are normally repaired by HR; however, in HR-deficient cells, such as BRCA1- and BRCA2-deficient backgrounds, lower-fidelity NHEJ occurs resulting in chromosomal aberration and ultimately cell death [117,118]. The catastrophic scenario induced by PARP inhibition could, therefore, be exploited to sensitise tumour cells to conventional treatments, such as chemotherapy or radiotherapy, which often cause DNA damage. Indeed, two groups in 2005 described the synthetic lethal (SL) interaction between PARP inhibition and BRCA1 or BRCA2 mutation suggesting a novel strategy for treating patients with BRCA mutant tumours [119–123]. Efforts in the last 20 years led to the development of many PARP inhibitors, including several that are already used in the clinics such as Veliparib (Abbvie), Rucaparib (Pfizer/Clovis), Olaparib (KuDOS/AstraZeneca), Niraparib (Merck/Tesaro) [124], Talazoparib (Lead/Biomarin/Medivation/Pfizer) [124,125], and Pamiparib (BeiGene/Merck Serono) [126,127], the latter with significant brain penetration.
All current clinically relevant PARPi are NAD+-mimetics and bind specifically to the PARP1 and PARP2 catalytic domain. PARPi are particularly effective (1000 times more sensitive) in the treatment of breast, ovarian, and other cancers that are BRCA1 and BRCA2 deficient [120,128–130].
Olaparib was the first PARPi approved by the U.S. Food and Drug Administration (FDA) for the treatment of patients carrying BRCA germline mutations (gBRCAm), whom have advanced ovarian cancer and already received previous lines of therapy [131,132]. The European Medicines Agency (EMA) also approved Olaparib as a maintenance treatment for BRCA mutant patients with platinum-sensitive gynecological cancers, such as high-grade serous epithelial ovarian, fallopian tube, or primary peritoneal cancers.
SL interactions between PARP inhibition and loss of function BRCA1 or BRCA2 can be expanded to sporadic tumours, a concept called ‘BRCAness’. Tumours that have not arisen from a germline gBRCAm are described to partially phenocopy the hereditary cancers in terms of HR defects [133,134]. Several mutations have been described to reproduce the phenotype of gBRCAm, such as somatically occurring mutations in either BRCA1 or BRCA2, somatic hypermethylation of BRCA1 promoter, or mutations in other genes that are involved in DSB repair and the stability of replication forks [133,134]. Beyond BRCA1 and BRCA2 mutations, BRCAness allowed to expand the SL approach to several tumours carrying deficiencies in many tumour suppressor genes involved in HR, such as ATM, ATR, PALB2, and the FANC gene family, which were shown to be sensitive to PARPi [134,135]. Notably, SL interactions between PARP inhibition and loss of function or amplification of other DDR factors, such as PALB2 and ATM, have been demonstrated by in vitro and xenograft studies [136–138]. Genome-wide sequencing projects revealed that somatic mutations in many other genes involved in HR occur in a wide spectrum of tumours [134], for instance, in high-grade serous ovarian cancer [139], advanced prostate cancer [140], and pancreatic cancer [141,142]. These and other cancers with HR mutations are therefore candidates for testing PARPi efficacy. Indeed, the use of Olaparib has also been approved for the treatment of metastatic, castration-resistant prostate cancer with genetic defects in DDR-encoding genes [140]. More studies will certainly expand the use of PARPi for the treatment of many other tumours, such as in acute myeloid leukaemia treatment [143]. Nevertheless, it is the obligation of the scientific community to improve the selectivity of targeted therapies. This can be achieved by an in-depth understanding of the molecular mechanisms regulated by PARPs and their cell type specificity as well as by a detailed comprehension of PARPi cytotoxic effects and by unveiling new functional interaction, such as with NEDD8/SCF and HDAC inhibitors [144,145].
A relevant topic for the clinic is the side effect of PARP inhibition. PARP1 and PARP2 have multiple important roles beyond the DDR, such as transcription, apoptosis, and immune function; the antitumour efficacy of PARPi might also reflect alterations in those functions [92]. Indeed, dose-limiting myelosuppression and central nervous system side effects were observed in some patients treated with Olaparib [146]. Being able to discriminate between the multiple cellular functions of PARP1 and PARP2 proteins, and specifically target the DDR pathway in a particular genetic background would be the next goal of the basic research-based translational research.
From a biochemical point of view, PARPi acts with different molecular mechanisms. For instance, some PARPi (such as Rucaparib, Olaparib, Niraparib, and Talazoparib) interfere with the catalytic cycle of PARP1. To a different extent, these drugs prevent PARP1 and PARP2 autoPARylation, thus trapping the enzymes on DNA lesions that they recognise [116–148] (Figure 1C); Talazoparib is ∼100 times more potent than Niraparib in trapping PARP1, which in turn traps PARP1 more potently than Olaparib and Rucaparib [147]. In contrast, Veliparib appears to have a limited ability to trap PARP1, despite its ability to inhibit PARylation [147]. These differences in trapping PARP1, rather than simply inhibiting PARylation, may be a better predictor of in vitro cytotoxicity in BRCA-deficient cells, with Talazoparib having the most profound cytotoxic effects [116,125,147].
Resistance to PARP inhibition is clinically relevant; multiple potential mechanisms of resistance to PARPi have been described. These include (i) silencing of DDR proteins, such as 53BP1 [149,150] or REV7 [151], which results in the reactivation of HR pathways; (ii) selective loss of PARG [152]; (iii) loss of function of proteins involved in destabilisation of the replication fork [153], such as EZH2 and MUS81 [154]; and (iv) loss of function of PARP1 itself [155]. Additionally, secondary ‘revertant’ mutations in BRCA1 or BRCA2 that restore sufficient HR function have been also described [156,157]; the same mutations are also responsible for resistance to platinum-based chemotherapy [134,156,158–160].
In the long term, the analysis of the gene expression of those DDR factors could be exploited to predict chemoresistance to PARPi. In addition, an in-depth understanding of cell type-specific PARP function is needed to address and improve patients' stratification for targeted therapies. Indeed, due to different expression of regulatory proteins and therefore with potentially different levels of regulation, the global ADPr in response to DNA damage may differ in a cell-specific manner [95]. For instance, some cell types may prefer Ser-ADPr to Asp/Glu-ADPr or vice versa, or follow alternative DDR pathways. Thus, knowing the origin of tumours and how the specific cell type respond to DNA damage induced by chemotherapy/radiotherapy may be useful to predict whether patients will be sensitive or not to PARPi.
A considerable effort has been made to identify additional SL interactions in ovarian cancers involving BRCA1 and BRCA2, which overcome PARPi resistance. The inhibition of the low-fidelity DNA polymerase-θ (Polθ) [118,161] and RAD52 [162] are synthetic lethal targets for the treatment of BRCA1-mutated cancers. These preclinical observations have led to ongoing efforts in the development of small molecule inhibiting Polθ and RAD52 [163].
In addition to PARP1 and PARP2, multiple players in ADPr regulation can be also targeted for cancer therapy, for instance, the hydrolases. The first study of breast cancer radiosensitisation by a cell-permeable specific inhibitor of PARG (PARGi; PDD00017273) [164] was reported by Gravells et al. [165]. PARGi sensitises cells to ionising radiation (IR) at the same magnitude of PARPi, although with different mechanisms of action. While PARPi radiosensitises to IR through replication-associated DNA breaks, which in turn are repaired at later times by NHEJ [121], inhibition of PARG leads to prolonged activation and persistence of PAR that results in faster DNA damage repair and rapid activation of NHEJ pathways [165]. Additionally, due to the pleiotropic functions of PARG, a severe mitotic phenotype induced by PARGi was observed, possibly due to the mitotic functions of PARG in reverting Tankyrase-mediated PARylation at the mitotic spindle (see below) [165].
The search for novel targets for cancer therapy may possibly open new research lines, perhaps looking for inhibitors of erasers and readers of ADPr (e.g. NUDT16, TARG1, ARH3, and ALC1). The advantage of targeting erasers of ADPr rather than PARPs may rely on the pleiotropic functions of catalytic enzymes, as demonstrated by the mitotic defects induced by PARGi in addition to the DNA repair ones [165]. Indeed, while each of the multiple PARPs shows a still not well-understood selectivity for certain substrates and regulates a restricted number of cellular functions, few erasers are required to act simultaneously on multiple processes and in different cellular compartments in order to revert any kind of ADPr event. Thus, compared with the inhibition of single PARPs, the inhibition of erasers of ADPr could potentially ensure the strongest impact on the overall physiology of cancer cells, thus affecting the cell survival and reducing the possibility of resistance to the treatment.
Tankyrases
Aside from the DDR, PARPs are recognised master regulators of multiple cellular functions, such as signal transduction, cell aging, and division. In this regard, Tankyrase-1 (also called PARP5a or TNKS1 or ARTD5) and Tankyrase-2 (also called PARP5b or TNKS2 or ARTD6) are poly(ADP-ribose) polymerases with functions in telomere length maintenance [166], HR-mediated DDR [167], mitosis [168–170], and Wnt and Notch-mediated signal transduction [171,172]. Notably, Tankyrases are novel and promising targets for cancer therapy [173].
Structurally, both Tankyrases are equipped with five Ankyrin repeats (ANK), a sterile alpha motif (SAM), and CAT domain [174]. Additionally, Tankyrase-1 contains a histidine, serine, and proline-rich (HPS) region [174]. While the SAM domain is required for multimerisation of Tankyrase molecules [175,176], the ANK domain serves as a binding platform for ADPr substrates [177]. In particular, the consensus hexapeptide motif RxxPDG (where ‘x’ refers to any amino acid) within the ANK domain was shown to be necessary and sufficient for interaction with protein targets [178]. Several targets of Tankyrases were identified, among them the telomeric-repeat binding factor-1 (TRF1) [168], the insulin-responsive amino-peptidase (IRAP) [178], the 182-kDa tankyrase-binding protein (TAB182) [178], the nuclear mitotic apparatus protein-1 (NuMA1) [178], AXIN1/2 [171], 3BP2 mutated in Cherubism disease [179,180], (Bhardwaj 2017) [172], and the CBP80/CBP20-dependent translation initiation factor (CTIF) [181]. Thus, the interaction with protein substrates can be considered the specificity determinant for Tankyrases. Compared with the knowledge achieved in the understanding of PARP1 activation mechanisms, it should be noted that very little is known about the processes of catalytic activation of Tankyrases, their amino acid specificity, and the modification sites in known Tankyrase substrates (AXIN above all). Recently, new details into the activation mechanism of Tankyrases have been provided. It has been shown that intermolecular SAM–SAM contacts are required to induce polymerisation of Tankyrase molecules, in both in vitro and in cellulo [175]. Tankyrase polymerisation seems to support the catalytic activity of the enzyme [175]. Formation of Tankyrase polymers also provides a nucleation point for the assembly of non-membranous structures, named signalosomes. The advantage of the formation of signalosomes is the local concentration of proteins, which can be both enzyme and substrates, for an efficient, transient, and spatially confined process [182,183]. Unfortunately, it is still unknown which signal triggers Tankyrase polymerisation and further activation of the catalytic activity.
Notably, Tankyrase-1 and Tankyrase-2 appear to have largely overlapping functions, as deletion of either gene leads to subtle phenotypes, while double knockout of both genes is embryonic lethal [184].
One of the best-characterised substrates of Tankyrases is AXIN, a key regulator of the canonical Wnt signalling pathway [173,185–188] (Figure 2). In the canonical pathway, Wnt regulates the level and subcellular localisation of the β-catenin transcription factor and is thus called a β-catenin-dependent Wnt pathway. In the absence of an activating Wnt signal, glycogen synthase kinase 3β (GSK3β) collaborates with the AXIN and APC (adenomatous polyposis coli) proteins and other factors to phosphorylate β-catenin [185]. The phosphorylated β-catenin is recognised and ubiquitinated by a complex containing a β-transducin repeat-containing protein (βTrCP), then degraded by the proteasome [185] (Figure 2). Wnt binds to a seven-pass transmembrane Frizzled receptor and its co-receptor, low-density lipoprotein receptor-related protein 6 (LRP6) on the cell surface. The resulting phosphorylation cascade inhibits the AXIN/GSK3β complex and stabilises the free pools of β-catenin, which can translocate into the nucleus [185] (Figure 2). In this context, Tankyrases are required to induce AXIN degradation and in turn β-catenin stabilisation. Tankyrases can indeed bind and PARylate AXIN. Tankyrase-mediated PARylation of AXIN acts as a scaffold for the recruitment for the WWE domain-equipped E3 ubiquitin ligase RNF146, which ubiquitinates AXIN leading to its proteasomal degradation [171] (Figure 2). In the nucleus, β-catenin binds to T-cell factor (TCF) transcriptional regulators along with other cofactors and modulates transcription of various genes [185]. Mutational mechanisms activating the WNT pathway and stabilising β-catenin have been found in human cancer, especially colorectal cancer (CRC) [185–188]. Cancer mutations include inactivation of APC or AXIN or activating mutations in β-catenin, all of which lead to constitutive transcription of β-catenin/TCF-regulated genes [185–188].
Modulation of Tankyrase signalling in cancer
The inhibition of both Tankyrases by the small NAD+ mimetic XAV939 was shown to inhibit growth of the APC-defective CRCs cell line by stabilising AXIN [171] (Figure 2). Tankyrase inhibitors would be presumably useful to target CRC cells that are characterised by constitutive activation of the Wnt pathway, such as those with upstream ligand-receptor defects or APC defects. In contrast, cancer cells expressing a mutant oncogenic β-catenin protein would presumably be resistant to AXIN stabilisation. Importantly, XAV939 inhibits Tankyrase-1, Tankyrase-2, PARP1, and PARP2 with comparable potency (e.g. IC50 values of 95, 5, 74, and 27 nM, respectively) [188]. More specific Tankyrase inhibitors were then developed with no measurable IC50 for PARP1 and PARP2, such as IWR-1 and IWR-2 [189–191]. Subsequently, structural studies have allowed structure-based drug design of more selective and potent Tankyrase inhibitors with cytotoxic potential demonstrated in CRC cell lines [184,192], for instance, JW74 [193], JW55 [194], WIKI4 [195,196], K-756 [197], G007-LK [198,199], and NVP-TNKS656 [200]. In vivo preclinical studies have demonstrated the antitumour activity of JW55 [194] and G007-LK [199] in xenograft and the Apc−/− mouse models. Major issues reported in preclinical studies are related to the intestinal toxicity of Tankyrase inhibitors per se [198,201].
As discussed for inhibitors of PARP1, the multiple functions of PARPs could limit the use of such molecules only in certain clinical conditions. As mentioned, Tankyrase proteins are evolutionarily conserved and function in telomere length regulation and sister telomere cohesion, GLUT4 vesicle translocation, and possibly also mitotic spindle pole regulation [174]. Thus, targeting of Tankyrases is expected to have varied effects on cells. Notably, inhibition of Tankyrase-1 resulted in synthetic lethal effects in cells with BRCA1 or BRCA2 defects, apparently due to exacerbation of the centrosome amplification phenotype associated with BRCA deficiency [202]. Of note, both BRCA1 and Tankyrases have connected functions in the regulation of bipolar mitotic spindle formation [203]. A better understanding of the molecular complexes and additional substrates of Tankyrases as well as knowing whether the expression/function of those proteins changes in tumours originating from specific cancer cells may help patients' stratification and prognosis.
Conclusions
In conclusion, ADPr is a central PTM regulating all the major cellular processes and thus affecting cell pathophysiology. Here, we have highlighted pertinent examples of how our understanding of ADPr signalling, gained from biochemical, structural, and cell biology studies, has been already exploited for the treatment of human cancer. Nevertheless, many other molecules linked to ADPr, thus not only ARTs, have or may have affect on cancer research; however, in many cases, their link with human pathology still needs to be uncovered.
Acknowledgements
The authors thank Dr Giovanna Grimaldi and Dr Daniela Corda (Institute of Protein Biochemistry, Naples, Italy), Dr Edward Bartlett (Sir William Dunn School of Pathology), and Dr Ian Gibbs-Seymour (Department of Biochemistry, University of Oxford, U.K.) for the helpful comments on the manuscript.
Abbreviations
- ADPr
ADP-ribosylation
- ANK
Ankyrin repeats
- APC
adenomatous polyposis coli
- ART
ADP-ribosyltransferase
- ARTD
ADP-ribosyltransferase diphtheria toxin-like
- Asp/Glu-ADPr
ADPr on aspartic/glutamic acids
- BER
base excision repair
- BRCT
BRCA1 C-terminal
- CAT
PARP catalytic domain
- CRC
colorectal cancer
- DDR
DNA damage response
- DSBs
double-strand breaks
- gBRCAm
BRCA germline mutations
- GSK3β
glycogen synthase kinase 3β
- HD
helical domain
- HPF1
histone PARylation factor 1
- HR
homologous recombination
- IR
ionising radiation
- MARylating
mono(ADP-ribosyl)ating
- MARylation
mono(ADP-ribosyl)ation
- NAD+
nicotinamide adenine dinucleotide
- NHEJ
non-homologous end joining
- PARPs
poly(ADP-ribose) polymerases
- PARylation
poly(ADP-ribosyl)ation
- PTM
post-translational modification
- SAM
sterile alpha motif
- Ser-ADPr
ADPr on serine residues
- SL
synthetic lethality
- SSBR
single-strand break repair
- TCF
T-cell factor
- UPR
unfolded protein response.
Author Contribution
L.P. and I.A. conceived and co-wrote the manuscript.
Funding
L.P. acknowledges funding from the Italian Foundation/Association for Cancer Research [FIRC/AIRC, Milan, Italy; grant 14895]. I.A. acknowledges funding from the Wellcome Trust [grant 101794], Cancer Research UK [grant C35050/A22284], and the European Research Council [grant 281739].
Competing Interests
The Authors declare that there are no competing interests associated with the manuscript.
References
- 1.Perina D., Mikoč A., Ahel J., C´etković H., Žaja R. and Ahel I. (2014) Distribution of protein poly(ADP-ribosyl)ation systems across all domains of life. DNA Repair 23, 4–16 10.1016/j.dnarep.2014.05.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Palazzo L., Mikoč A. and Ahel I. (2017) ADP-ribosylation: new facets of an ancient modification. FEBS J. 284, 2932–2946 10.1111/febs.14078 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Cohen M.S. and Chang P. (2018) Insights into the biogenesis, function, and regulation of ADP-ribosylation. Nat. Chem. Biol. 14, 236–243 10.1038/nchembio.2568 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Lüscher B., Bütepage M., Eckei L., Krieg S., Verheugd P. and Shilton B.H. (2018) ADP-ribosylation, a multifaceted posttranslational modification involved in the control of cell physiology in health and disease. Chem. Rev. 118, 1092–1136 10.1021/acs.chemrev.7b00122 [DOI] [PubMed] [Google Scholar]
- 5.Leung A.K., Vyas S., Rood J.E., Bhutkar A., Sharp P.A. and Chang P. (2011) Poly(ADP-ribose) regulates stress responses and microRNA activity in the cytoplasm. Mol. Cell 42, 489–499 10.1016/j.molcel.2011.04.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Gibson B.A. and Kraus W.L. (2012) New insights into the molecular and cellular functions of poly(ADP-ribose) and PARPs. Nat. Rev. Mol. Cell Biol. 13, 411–424 10.1038/nrm3376 [DOI] [PubMed] [Google Scholar]
- 7.Bai P. and Cantó C. (2012) The role of PARP-1 and PARP-2 enzymes in metabolic regulation and disease. Cell Metab. 16, 290–295 10.1016/j.cmet.2012.06.016 [DOI] [PubMed] [Google Scholar]
- 8.Jwa M. and Chang P. (2012) PARP16 is a tail-anchored endoplasmic reticulum protein required for the PERK- and IRE1α-mediated unfolded protein response. Nat. Cell Biol. 14, 1223–1230 10.1038/ncb2593 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Bock F.J., Todorova T.T. and Chang P. (2015) RNA regulation by poly(ADP-ribose) polymerases. Mol. Cell 58, 959–969 10.1016/j.molcel.2015.01.037 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Rack J.G., Morra R., Barkauskaite E., Kraehenbuehl R., Ariza A., Qu Y. et al. (2015) Identification of a class of protein ADP-ribosylating sirtuins in microbial pathogens. Mol. Cell 59, 309–320 10.1016/j.molcel.2015.06.013 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Bai P. (2015) Biology of poly(ADP-ribose) polymerases: the factotums of cell maintenance. Mol. Cell 58, 947–958 10.1016/j.molcel.2015.01.034 [DOI] [PubMed] [Google Scholar]
- 12.Bock F.J. and Chang P. (2016) New directions in poly(ADP-ribose) polymerase biology. FEBS J. 283, 4017–4031 10.1111/febs.13737 [DOI] [PubMed] [Google Scholar]
- 13.Lalić J., Posavec Marjanović M., Palazzo L., Perina D., Sabljić I., Žaja R. et al. (2016) Disruption of macrodomain protein SCO6735 increases antibiotic production in Streptomyces coelicolor. J. Biol. Chem. 291, 23175–23187 10.1074/jbc.M116.721894 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Jankevicius G., Ariza A., Ahel M. and Ahel I. (2016) The toxin-antitoxin system DarTG catalyzes reversible ADP-ribosylation of DNA. Mol. Cell 64, 1109–1116 10.1016/j.molcel.2016.11.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Gupte R., Liu Z. and Kraus W.L. (2017) PARPs and ADP-ribosylation: recent advances linking molecular functions to biological outcomes. Genes Dev. 31, 101–126 10.1101/gad.291518.116 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Catara G., Grimaldi G., Schembri L., Spano D., Turacchio G., Lo Monte M. et al. (2017) PARP1-produced poly-ADP-ribose causes the PARP12 translocation to stress granules and impairment of Golgi complex functions. Sci. Rep. 7, Article number: 14035 10.1038/s41598-017-14156-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Fehr A.R., Jankevicius G., Ahel I. and Perlman S. (2018) Viral macrodomains: unique mediators of viral replication and pathogenesis. Trends Microbiol. 26, 598–610 10.1016/j.tim.2017.11.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Di Girolamo, M. and Fabrizio, G. (2018) The ADP-ribosyl-transferases diphtheria toxin-like (ARTDs) family: an overview. Challenges 9, 24 10.3390/challe9010024 [DOI] [Google Scholar]
- 19.Ahel I., Ahel D., Matsusaka T., Clark A.J., Pines J., Boulton S.J. et al. (2008) Poly(ADP-ribose)-binding zinc finger motifs in DNA repair/checkpoint proteins. Nature 451, 81–85 10.1038/nature06420 [DOI] [PubMed] [Google Scholar]
- 20.Ahel D., Horejsí Z., Wiechens N., Polo S.E., Garcia-Wilson E., Ahel I. et al. (2009) Poly(ADP-ribose)-dependent regulation of DNA repair by the chromatin remodeling enzyme ALC1. Science 325, 1240–1243 10.1126/science.1177321 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Kalisch T., Amé J.-C., Dantzer F. and Schreiber V. (2012) New readers and interpretations of poly(ADP-ribosyl)ation. Trends Biochem. Sci. 37, 381–390 10.1016/j.tibs.2012.06.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Teloni F. and Altmeyer M. (2016) Readers of poly(ADP-ribose): designed to be fit for purpose. Nucleic Acids Res. 44, 993–1006 10.1093/nar/gkv1383 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Rack J.G., Perina D. and Ahel I. (2016) Macrodomains: structure, function, evolution, and catalytic activities. Annu. Rev. Biochem. 85, 431–454 10.1146/annurev-biochem-060815-014935 [DOI] [PubMed] [Google Scholar]
- 24.Slade D., Dunstan M.S., Barkauskaite E., Weston R., Lafite P., Dixon N. et al. (2011) The structure and catalytic mechanism of a poly(ADP-ribose) glycohydrolase. Nature 477, 616–620 10.1038/nature10404 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Mashimo M., Kato J. and Moss J. (2014) Structure and function of the ARH family of ADP-ribosyl-acceptor hydrolases. DNA Repair 23, 88–94 10.1016/j.dnarep.2014.03.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Winstall E., Affar E.B., Shah R., Bourassa S., Scovassi I.A. and Poirier G.G. (1999) Preferential perinuclear localization of poly(ADP-ribose) glycohydrolase. Exp. Cell Res. 251, 372–378 10.1006/excr.1999.4594 [DOI] [PubMed] [Google Scholar]
- 27.Meyer-Ficca M.L., Meyer R.G., Coyle D.L., Jacobson E.L. and Jacobson M.K. (2004) Human poly(ADP-ribose) glycohydrolase is expressed in alternative splice variants yielding isoforms that localize to different cell compartments. Exp. Cell Res. 297, 521–532 10.1016/j.yexcr.2004.03.050 [DOI] [PubMed] [Google Scholar]
- 28.Niere M., Mashimo M., Agledal L., Dölle C., Kasamatsu A., Kato J. et al. (2012) ADP-ribosylhydrolase 3 (ARH3), not poly(ADP-ribose) glycohydrolase (PARG) isoforms, is responsible for degradation of mitochondrial matrix-associated poly(ADP-ribose). J. Biol. Chem. 287, 16088–16102 10.1074/jbc.M112.349183 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Sharifi R., Morra R., Appel C.D., Tallis M., Chioza B., Jankevicius G. et al. (2013) Deficiency of terminal ADP-ribose protein glycohydrolase TARG1/C6orf130 in neurodegenerative disease. EMBO J. 32, 1225–1237 10.1038/emboj.2013.51 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Golia B., Moeller G.K., Jankevicius G., Schmidt A., Hegele A., Preißer J. et al. (2017) ATM induces MacroD2 nuclear export upon DNA damage. Nucleic Acids Res. 45, 244–254 10.1093/nar/gkw904 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Agnew T., Munnur D., Crawford K., Palazzo L., Mikoč A. and Ahel I. (2018) Macrod1 is a promiscuous ADP-ribosyl hydrolase localized to mitochondria. Front. Microbiol. 9, 20 10.3389/fmicb.2018.00020 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Bütepage M., Preisinger C., von Kriegsheim A., Scheufen A., Lausberg E., Li J. et al. (2018) Nucleolar-nucleoplasmic shuttling of TARG1 and its control by DNA damage-induced poly-ADP-ribosylation and by nucleolar transcription. Sci. Rep. 8, 6748 10.1038/s41598-018-25137-w [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Palazzo L., Thomas B., Jemth A.-S., Colby T., Leidecker O., Feijs K.L. et al. (2015) Processing of protein ADP-ribosylation by Nudix hydrolases. Biochem. J. 468, 293–301 10.1042/BJ20141554 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Daniels C.M., Thirawatananond P., Ong S.-E., Gabelli S.B. and Leung A.K. (2016) Nudix hydrolases degrade protein-conjugated ADP-ribose. Sci. Rep. 5, 18271 PMID: [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Chapman J.D., Gagné J.-P., Poirier G.G. and Goodlett D.R. (2013) Mapping PARP-1 auto-ADP-ribosylation sites by liquid chromatography-tandem mass spectrometry. J. Proteome Res. 12, 1868–1880 10.1021/pr301219h [DOI] [PubMed] [Google Scholar]
- 36.Palazzo L., Daniels C.M., Nettleship J.E., Rahman N., McPherson R.L., Ong S.-E. et al. (2016) ENPP1 processes protein ADP-ribosylation in vitro. FEBS J. 283, 3371–3388 10.1111/febs.13811 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Bhogaraju S., Kalayil S., Liu Y., Bonn F., Colby T., Matic I. et al. (2016) Phosphoribosylation of ubiquitin promotes serine ubiquitination and impairs conventional ubiquitination. Cell 167, 1636–1649.e13 10.1016/j.cell.2016.11.019 [DOI] [PubMed] [Google Scholar]
- 38.Hottiger M.O., Hassa P.O., Lüscher B., Schüler H. and Koch-Nolte F. (2010) Toward a unified nomenclature for mammalian ADP-ribosyltransferases. Trends Biochem. Sci. 35, 208–219 10.1016/j.tibs.2009.12.003 [DOI] [PubMed] [Google Scholar]
- 39.Pascal J.M. and Ellenberger T. (2015) The rise and fall of poly(ADP-ribose): An enzymatic perspective. DNA Repair 32, 10–16 10.1016/j.dnarep.2015.04.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Barkauskaite E., Jankevicius G. and Ahel I. (2015) Structures and mechanisms of enzymes employed in the synthesis and degradation of PARP-dependent protein ADP-ribosylation. Mol. Cell 58, 935–946 10.1016/j.molcel.2015.05.007 [DOI] [PubMed] [Google Scholar]
- 41.Vyas S., Matic I., Uchima L., Rood J., Zaja R., Hay R.T. et al. (2014) Family-wide analysis of poly(ADP-ribose) polymerase activity. Nat. Commun. 5, 4426 10.1038/ncomms5426 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Lopes R.R., Kessler A.C., Polycarpo C. and Alfonzo J.D. (2015) Cutting, dicing, healing and sealing: the molecular surgery of tRNA. Wiley Interdiscip. Rev. RNA 6, 337–349 10.1002/wrna.1279 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Leidecker O., Bonfiglio J.J., Colby T., Zhang Q., Atanassov I., Zaja R. et al. (2016) Serine is a new target residue for endogenous ADP-ribosylation on histones. Nat. Chem. Biol. 12, 998–1000 10.1038/nchembio.2180 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Crawford K., Bonfiglio J.J., Mikoč A., Matic I. and Ahel I. (2018) Specificity of reversible ADP-ribosylation and regulation of cellular processes. Crit. Rev. Biochem. Mol. Biol. 53, 64–82 10.1080/10409238.2017.1394265 [DOI] [PubMed] [Google Scholar]
- 45.Voorneveld J., Rack J.G.M., Ahel I., Overkleeft H.S., van der Marel G.A. and Filippov D.V. (2018) Synthetic α- and β-Ser-ADP-ribosylated peptides reveal α-Ser-ADPr as the native epimer. Org. Lett. 20, 4140–4143 10.1021/acs.orglett.8b01742 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.D'Amours D., Desnoyers S., D'Silva I. and Poirier G.G. (1999) Poly(ADP-ribosyl)ation reactions in the regulation of nuclear functions. Biochem. J. 342, 249–268 10.1042/bj3420249 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.de Murcia G. and Ménissier de Murcia J. (1994) Poly(ADP-ribose) polymerase: a molecular nick-sensor. Trends Biochem. Sci. 19, 172–176 10.1016/0968-0004(94)90280-1 [DOI] [PubMed] [Google Scholar]
- 48.Rippmann J.F., Damm K. and Schnapp A. (2002) Functional characterization of the poly(ADP-ribose) polymerase activity of tankyrase 1, a potential regulator of telomere length. J. Mol. Biol. 323, 217–224 10.1016/S0022-2836(02)00946-4 [DOI] [PubMed] [Google Scholar]
- 49.Rulten S.L., Fisher A.E., Robert I., Zuma M.C., Rouleau M., Ju L. et al. (2011) PARP-3 and APLF function together to accelerate nonhomologous end-joining. Mol. Cell 41, 33–45 10.1016/j.molcel.2010.12.006 [DOI] [PubMed] [Google Scholar]
- 50.Boehler C., Gauthier L.R., Mortusewicz O., Biard D.S., Saliou J.-M., Bresson A. et al. (2011) Poly(ADP-ribose) polymerase 3 (PARP3), a newcomer in cellular response to DNA damage and mitotic progression. Proc. Natl Acad. Sci. U.S.A. 108, 2783–2788 10.1073/pnas.1016574108 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Kickhoefer V.A., Siva A.C., Kedersha N.L., Inman E.M., Ruland C., Streuli M. et al. (1999) The 193-kD vault protein, VPARP, is a novel poly(ADP-ribose) polymerase. J. Cell. Biol. 146, 917–928 10.1083/jcb.146.5.917 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Daugherty M.D., Young J.M., Kerns J.A. and Malik H.S. (2014) Rapid evolution of PARP genes suggests a broad role for ADP-ribosylation in host-virus conflicts. PLoS Genet. 10, e1004403 10.1371/journal.pgen.1004403 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Tuncel H., Tanaka S., Oka S., Nakai S., Fukutomi R., Okamoto M. et al. (2012) PARP6, a mono(ADP-ribosyl) transferase and a negative regulator of cell proliferation, is involved in colorectal cancer development. Int. J. Oncol. 41, 2079–2086 10.3892/ijo.2012.1652 [DOI] [PubMed] [Google Scholar]
- 54.Yan Q., Xu R., Zhu L., Cheng X., Wang Z., Manis J. et al. (2013) BAL1 and its partner E3 ligase, BBAP, link poly(ADP-ribose) activation, ubiquitylation, and double-strand DNA repair independent of ATM, MDC1, and RNF8. Mol. Cell. Biol. 33, 845–857 10.1128/MCB.00990-12 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Yang C.-S., Jividen K., Spencer A., Dworak N., Ni L., Oostdyk L.T. et al. (2017) Ubiquitin modification by the E3 ligase/ADP-ribosyltransferase Dtx3L/Parp9. Mol. Cell 66, 503–516.e5 10.1016/j.molcel.2017.04.028 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Juszczynski P., Kutok J.L., Li C., Mitra J., Aguiar R.C. and Shipp M.A. (2006) BAL1 and BBAP are regulated by a gamma interferon-responsive bidirectional promoter and are overexpressed in diffuse large B-cell lymphomas with a prominent inflammatory infiltrate. Mol. Cell. Biol. 26, 5348–5359 10.1128/MCB.02351-05 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Camicia R., Bachmann S.B., Winkler H.C., Beer M., Tinguely M., Haralambieva E. et al. (2013) BAL1/ARTD9 represses the anti-proliferative and pro-apoptotic IFNγ-STAT1-IRF1-p53 axis in diffuse large B-cell lymphoma. J. Cell Sci. 126(Pt 9), 1969–1980 10.1242/jcs.118174 [DOI] [PubMed] [Google Scholar]
- 58.Yu M., Schreek S., Cerni C., Schamberger C., Lesniewicz K., Poreba E. et al. (2005) PARP-10, a novel Myc-interacting protein with poly(ADP-ribose) polymerase activity, inhibits transformation. Oncogene 24, 1982–1993 10.1038/sj.onc.1208410 [DOI] [PubMed] [Google Scholar]
- 59.Verheugd P., Forst A.H., Milke L., Herzog N., Feijs K.L., Kremmer E. et al. (2013) Regulation of NF-κB signalling by the mono-ADP-ribosyltransferase ARTD10. Nat. Commun. 4, 1683 10.1038/ncomms2672 [DOI] [PubMed] [Google Scholar]
- 60.Herzog N., Hartkamp J.D., Verheugd P., Treude F., Forst A.H., Feijs K.L. et al. (2013) Caspase-dependent cleavage of the mono-ADP-ribosyltransferase ARTD10 interferes with its pro-apoptotic function. FEBS J. 280, 1330–1343 10.1111/febs.12124 [DOI] [PubMed] [Google Scholar]
- 61.Meyer-Ficca M.L., Ihara M., Bader J.J., Leu N.A., Beneke S. and Meyer R.G. (2015) Spermatid head elongation with normal nuclear shaping requires ADP-ribosyltransferase PARP11 (ARTD11) in mice. Biol. Reprod. 92, 80 PMID: [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Carter-O'Connell I., Jin H., Morgan R.K., Zaja R., David L.L., Ahel I. et al. (2016) Identifying family-member-specific targets of mono-ARTDs by using a chemical genetics approach. Cell Rep. 14, 621–631 10.1016/j.celrep.2015.12.045 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Li L., Zhao H., Liu P., Li C., Quanquin N., Ji X. et al. (2018) PARP12 suppresses Zika virus infection through PARP-dependent degradation of NS1 and NS3 viral proteins. Sci. Signal. 11, pii: eaas9332 10.1126/scisignal.aas9332 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Seo G.J., Kincaid R.P., Phanaksri T., Burke J.M., Pare J.M., Cox J.E. et al. (2013) Reciprocal inhibition between intracellular antiviral signaling and the RNAi machinery in mammalian cells. Cell Host Microbe 14, 435–445 10.1016/j.chom.2013.09.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Todorova T., Bock F.J. and Chang P. (2015) Poly(ADP-ribose) polymerase-13 and RNA regulation in immunity and cancer. Trends Mol. Med. 21, 373–384 10.1016/j.molmed.2015.03.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Cho S.H., Goenka S., Henttinen T., Gudapati P., Reinikainen A., Eischen C.M. et al. (2009) PARP-14, a member of the B aggressive lymphoma family, transduces survival signals in primary B cells. Blood 113, 2416–2425 10.1182/blood-2008-03-144121 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Vyas S., Chesarone-Cataldo M., Todorova T., Huang Y.-H. and Chang P. (2013) A systematic analysis of the PARP protein family identifies new functions critical for cell physiology. Nat. Commun. 4, 2240 10.1038/ncomms3240 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Caprara G., Prosperini E., Piccolo V., Sigismondo G., Melacarne A., Cuomo A. et al. (2018) PARP14 controls the nuclear accumulation of a subset of type I IFN-inducible proteins. J. Immunol. 200, 2439–2454 10.4049/jimmunol.1701117 [DOI] [PubMed] [Google Scholar]
- 69.Schuller M., Riedel K., Gibbs-Seymour I., Uth K., Sieg C., Gehring A.P. et al. (2017) Discovery of a selective allosteric inhibitor targeting macrodomain 2 of polyadenosine-diphosphate-ribose polymerase 14. ACS Chem. Biol. 12, 2866–2874 10.1021/acschembio.7b00445 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Di Paola S., Micaroni M., Di Tullio G., Buccione R. and Di Girolamo M. (2012) PARP16/ARTD15 is a novel endoplasmic-reticulum-associated mono-ADP-ribosyltransferase that interacts with, and modifies karyopherin-ß1. PLoS ONE 7, e37352 10.1371/journal.pone.0037352 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Beck C., Robert I., Reina-San-Martin B., Schreiber V. and Dantzer F. (2014) Poly(ADP-ribose) polymerases in double-strand break repair: focus on PARP1, PARP2 and PARP3. Exp. Cell Res. 329, 18–25 10.1016/j.yexcr.2014.07.003 [DOI] [PubMed] [Google Scholar]
- 72.Tao Z., Gao P., Hoffman D.W. and Liu H.-W. (2008) Domain C of human poly(ADP-ribose) polymerase-1 is important for enzyme activity and contains a novel zinc-ribbon motif. Biochemistry 47, 5804–5813 10.1021/bi800018a [DOI] [PubMed] [Google Scholar]
- 73.Langelier M.-F., Servent K.M., Rogers E.E. and Pascal J.M. (2008) A third zinc-binding domain of human poly(ADP-ribose) polymerase-1 coordinates DNA-dependent enzyme activation. J. Biol. Chem. 283, 4105–4114 10.1074/jbc.M708558200 [DOI] [PubMed] [Google Scholar]
- 74.Langelier M.-F. and Pascal J.M. (2013) PARP-1 mechanism for coupling DNA damage detection to poly(ADP-ribose) synthesis. Curr. Opin. Struct. Biol. 23, 134–143 10.1016/j.sbi.2013.01.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Ruf A., Mennissier de Murcia J., de Murcia G. and Schulz G.E. (1996) Structure of the catalytic fragment of poly(AD-ribose) polymerase from chicken. Proc. Natl Acad. Sci. U.S.A. 93, 7481–7485 10.1073/pnas.93.15.7481 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Altmeyer M., Messner S., Hassa P.O., Fey M. and Hottiger M.O. (2009) Molecular mechanism of poly(ADP-ribosyl)ation by PARP1 and identification of lysine residues as ADP-ribose acceptor sites. Nucleic Acids Res. 37, 3723–3738 10.1093/nar/gkp229 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Langelier M.-F., Planck J.L., Roy S. and Pascal J.M. (2012) Structural basis for DNA damage-dependent poly(ADP-ribosyl)ation by human PARP-1. Science 336, 728–732 10.1126/science.1216338 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Dawicki-McKenna J.M., Langelier M.-F., DeNizio J.E., Riccio A.A., Cao C.D., Karch K.R. et al. (2015) PARP-1 activation requires local unfolding of an autoinhibitory domain. Mol. Cell 60, 755–768 10.1016/j.molcel.2015.10.013 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Langelier M.-F., Zandarashvili L., Aguiar P.M., Black B.E. and Pascal J.M. (2018) NAD+ analog reveals PARP-1 substrate-blocking mechanism and allosteric communication from catalytic center to DNA-binding domains. Nat. Commun. 9, 844 10.1038/s41467-018-03234-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Liu C., Vyas A., Kassab M.A., Singh A.K. and Yu X. (2017) The role of poly ADP-ribosylation in the first wave of DNA damage response. Nucleic Acids Res. 45, 8129–8141 10.1093/nar/gkx565 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Eustermann S., Wu W.-F., Langelier M.-F., Yang J.-C., Easton L.E., Riccio A.A. et al. (2015) Structural basis of detection and signaling of DNA single-strand breaks by human PARP-1. Mol. Cell 60, 742–754 10.1016/j.molcel.2015.10.032 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Chen Q., Kassab M.A., Dantzer F. and Yu X. (2018) PARP2 mediates branched poly ADP-ribosylation in response to DNA damage. Nat. Commun. 9, 3233 10.1038/s41467-018-05588-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Talhaoui I., Lebedeva N.A., Zarkovic G., Saint-Pierre C., Kutuzov M.M., Sukhanova M.V. et al. (2016) Poly(ADP-ribose) polymerases covalently modify strand break termini in DNA fragments in vitro. Nucleic Acids Res. 44, 9279–9295 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Munnur D. and Ahel I. (2017) Reversible mono-ADP-ribosylation of DNA breaks. FEBS J. 284, 4002–4016 10.1111/febs.14297 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Zarkovic G., Belousova E.A., Talhaoui I., Saint-Pierre C., Kutuzov M.M., Matkarimov B.T. et al. (2018) Characterization of DNA ADP-ribosyltransferase activities of PARP2 and PARP3: new insights into DNA ADP-ribosylation. Nucleic Acids Res. 46, 2417–2431 10.1093/nar/gkx1318 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Schreiber V., Dantzer F., Ame J.-C. and de Murcia G. (2006) Poly(ADP-ribose): novel functions for an old molecule. Nat. Rev. Mol. Cell. Biol. 7, 517–528 10.1038/nrm1963 [DOI] [PubMed] [Google Scholar]
- 87.Ménissier de Murcia J., Ricoul M., Tartier L., Niedergang C., Huber A., Dantzer F. et al. (2003) Functional interaction between PARP-1 and PARP-2 in chromosome stability and embryonic development in mouse. EMBO J. 22, 2255–2263 10.1093/emboj/cdg206 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Poirier G.G., de Murcia G., Jongstra-Bilen J., Niedergang C. and Mandel P. (1982) Poly(ADP-ribosyl)ation of polynucleosomes causes relaxation of chromatin structure. Proc. Natl Acad. Sci. U.S.A. 79, 3423–3427 10.1073/pnas.79.11.3423 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Bonfiglio J.J., Fontana P., Zhang Q., Colby T., Gibbs-Seymour I., Atanassov I. et al. (2017) Serine ADP-ribosylation depends on HPF1. Mol. Cell 65, 932–940.e6 10.1016/j.molcel.2017.01.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Palazzo L., Leidecker O., Prokhorova E., Dauben H., Matic I. and Ahel I. (2018) Serine is the major residue for ADP-ribosylation upon DNA damage. eLife 7, pii: e34334 10.7554/eLife.34334 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Satoh M.S. and Lindahl T. (1992) Role of poly(ADP-ribose) formation in DNA repair. Nature 356, 356–358 10.1038/356356a0 [DOI] [PubMed] [Google Scholar]
- 92.Krishnakumar R. and Kraus W.L. (2010) The PARP side of the nucleus: molecular actions, physiological outcomes, and clinical targets. Mol. Cell 39, 8–24 10.1016/j.molcel.2010.06.017 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Gibbs-Seymour I., Fontana P., Rack J.G.M. and Ahel I. (2016) HPF1/C4orf27 is a PARP-1-interacting protein that regulates PARP-1 ADP-ribosylation activity. Mol. Cell 62, 432–442 10.1016/j.molcel.2016.03.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Zhang Y., Wang J., Ding M. and Yu Y. (2013) Site-specific characterization of the Asp- and Glu-ADP-ribosylated proteome. Nat. Methods 10, 981–984 10.1038/nmeth.2603 [DOI] [PubMed] [Google Scholar]
- 95.Zhen Y., Zhang Y. and Yu Y. (2017) A cell-line-specific atlas of PARP-mediated protein Asp/Glu-ADP-ribosylation in breast cancer. Cell Rep. 21, 2326–2337 10.1016/j.celrep.2017.10.106 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Karch K.R., Langelier M.-F., Pascal J.M. and Garcia B.A. (2017) The nucleosomal surface is the main target of histone ADP-ribosylation in response to DNA damage. Mol. Biosyst. 13, 2660–2671 10.1039/C7MB00498B [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Gibson B.A., Conrad L.B., Huang D. and Kraus W.L. (2017) Generation and characterization of recombinant antibody-like ADP-ribose binding proteins. Biochemistry 56, 6305–6316 10.1021/acs.biochem.7b00670 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Daniels C.M., Ong S.-E. and Leung A.K. (2014) Phosphoproteomic approach to characterize protein mono- and poly(ADP-ribosyl)ation sites from cells. J. Proteome Res. 13, 3510–3522 10.1021/pr401032q [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Daniels C.M., Ong S.-E. and Leung A.K. (2015) The promise of proteomics for the study of ADP-ribosylation. Mol. Cell 58, 911–924 10.1016/j.molcel.2015.06.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Martello R., Leutert M., Jungmichel S., Bilan V., Larsen S.C., Young C. et al. (2016) Proteome-wide identification of the endogenous ADP-ribosylome of mammalian cells and tissue. Nat. Commun. 7, 12917 10.1038/ncomms12917 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Gagné J.-P., Langelier M.-F., Pascal J.M. and Poirier G.G. (2018) Hydrofluoric acid-based derivatization strategy to profile PARP-1 ADP-ribosylation by LC-MS/MS. J. Proteome Res. 17, 2542–2551 10.1021/acs.jproteome.8b00146 [DOI] [PubMed] [Google Scholar]
- 102.Ferro A.M. and Olivera B.M. (1982) Poly(ADP-ribosylation) in vitro: reaction parameters and enzyme mechanism. J. Biol. Chem. 257, 7808–7813 PMID: [PubMed] [Google Scholar]
- 103.Zahradka P. and Ebisuzaki K. (1982) A shuttle mechanism for DNA-protein interactions: the regulation of poly(ADP-ribose) polymerase. Eur. J. Biochem. 127, 579–585 10.1111/j.1432-1033.1982.tb06912.x [DOI] [PubMed] [Google Scholar]
- 104.Tallis M., Morra R., Barkauskaite E. and Ahel I. (2014) Poly(ADP-ribosyl)ation in regulation of chromatin structure and the DNA damage response. Chromosoma 123, 79–90 10.1007/s00412-013-0442-9 [DOI] [PubMed] [Google Scholar]
- 105.Caldecott K.W. (2014) DNA single-strand break repair. Exp. Cell Res. 329, 2–8 10.1016/j.yexcr.2014.08.027 [DOI] [PubMed] [Google Scholar]
- 106.Timinszky G., Till S., Hassa P.O., Hothorn M., Kustatscher G., Nijmeijer B. et al. (2009) A macrodomain-containing histone rearranges chromatin upon sensing PARP1 activation. Nat. Struct. Mol. Biol. 16, 923–929 10.1038/nsmb.1664 [DOI] [PubMed] [Google Scholar]
- 107.Li M. and Yu X. (2015) The role of poly(ADP-ribosyl)ation in DNA damage response and cancer chemotherapy. Oncogene 34, 3349–3356 10.1038/onc.2014.295 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Brochu G., Duchaine C., Thibeault L., Lagueux J., Shah G.M. and Poirier G.G. (1994) Mode of action of poly(ADP-ribose) glycohydrolase. Biochim. Biophys. Acta 1219, 342–350 10.1016/0167-4781(94)90058-2 [DOI] [PubMed] [Google Scholar]
- 109.Lambrecht M.J., Brichacek M., Barkauskaite E., Ariza A., Ahel I. and Hergenrother P.J. (2015) Synthesis of dimeric ADP-ribose and its structure with human poly(ADP-ribose) glycohydrolase. J. Am. Chem. Soc. 137, 3558–3564 10.1021/ja512528p [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Fontana P., Bonfiglio J.J., Palazzo L., Bartlett E., Matic I. and Ahel I. (2017) Serine ADP-ribosylation reversal by the hydrolase ARH3. eLife 6, pii: e28533 10.7554/eLife.28533 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Barkauskaite E., Brassington A., Tan E.S., Warwicker J., Dunstan M.S., Banos B. et al. (2013) Visualization of poly(ADP-ribose) bound to PARG reveals inherent balance between exo- and endo-glycohydrolase activities. Nat. Commun. 4, 2164 10.1038/ncomms3164 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Shall S. (1975) Seminar on poly(ADP-ribose) and ADP-ribosylation of proceedings: experimental manipulation of the specific activity of poly(ADP-ribose) polymerase. J. Biochem. 77, 2 10.1093/oxfordjournals.jbchem.a130859 [DOI] [PubMed] [Google Scholar]
- 113.Purnell M.R. and Whish W.J. (1980) Novel inhibitors of poly(ADP-ribose) synthetase. Biochem. J. 185, 775–777 10.1042/bj1850775 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Terada M., Fujiki H., Marks P.A. and Sugimura T. (1979) Induction of erythroid differentiation of murine erythroleukemia cells by nicotinamide and related compounds. Proc. Natl Acad. Sci. U.S.A. 76, 6411–6414 10.1073/pnas.76.12.6411 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Helleday T. (2011) The underlying mechanism for the PARP and BRCA synthetic lethality: clearing up the misunderstandings. Mol. Oncol. 5, 387–393 10.1016/j.molonc.2011.07.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Murai J., Huang S.-Y., Das B.B., Renaud A., Zhang Y., Doroshow J.H. et al. (2012) Trapping of PARP1 and PARP2 by clinical PARP inhibitors. Cancer Res. 72, 5588–5599 10.1158/0008-5472.CAN-12-2753 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Ceccaldi R., Rondinelli B. and D'Andrea A.D. (2016) Repair pathway choices and consequences at the double-strand break. Trends Cell Biol. 26, 52–64 10.1016/j.tcb.2015.07.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Mateos-Gomez P.A., Gong F., Nair N., Miller K.M., Lazzerini-Denchi E. and Sfeir A. (2015) Mammalian polymerase θ promotes alternative NHEJ and suppresses recombination. Nature 518, 254–257 10.1038/nature14157 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Ashworth A., Lord C.J. and Reis-Filho J.S. (2011) Genetic interactions in cancer progression and treatment. Cell 145, 30–38 10.1016/j.cell.2011.03.020 [DOI] [PubMed] [Google Scholar]
- 120.Farmer H., McCabe N., Lord C.J., Tutt A.N., Johnson D.A., Richardson T.B. et al. (2005) Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 434, 917–921 10.1038/nature03445 [DOI] [PubMed] [Google Scholar]
- 121.Bryant H.E., Schultz N., Thomas H.D., Parker K.M., Flower D., Lopez E. et al. (2005) Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase. Nature 434, 913–917 10.1038/nature03443 [DOI] [PubMed] [Google Scholar]
- 122.Bryant H.E., Petermann E., Schultz N., Jemth A.-S., Loseva O., Issaeva N. et al. (2009) PARP is activated at stalled forks to mediate Mre11-dependent replication restart and recombination. EMBO J. 28, 2601–2615 10.1038/emboj.2009.206 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Zaremba T. and Curtin N.J. (2007) PARP inhibitor development for systemic cancer targeting. Anticancer Agents Med. Chem. 7, 515–523 10.2174/187152007781668715 [DOI] [PubMed] [Google Scholar]
- 124.Lord C.J. and Ashworth A. (2017) PARP inhibitors: synthetic lethality in the clinic. Science 355, 1152–1158 10.1126/science.aam7344 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Shen Y., Rehman F.L., Feng Y., Boshuizen J., Bajrami I., Elliott R. et al. (2013) BMN 673, a novel and highly potent PARP1/2 inhibitor for the treatment of human cancers with DNA repair deficiency. Clin. Cancer Res. 19, 5003–5015 10.1158/1078-0432.CCR-13-1391 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Tang Z., Jiang B., Shi Z., Gong W., Liu Y., Wang X. et al. (2015) BGB-290, a novel PARP inhibitor with unique brain penetration ability, demonstrated strong synergism with temozolomide in subcutaneous and intracranial xenograft models. Cancer Res. 75, Abstract no. 1651 10.1158/1538-7445.AM2015-1651 [DOI] [Google Scholar]
- 127.Tang Z., Liu Y., Zhen Q., Ren B., Wang H., Shi Z. et al. (2015) BGB-290: A highly potent and specific PARP1/2 inhibitor potentiates anti-tumor activity of chemotherapeutics in patient biopsy derived SCLC models. Cancer Res. 75, Abstract no. 1653 10.1158/1538-7445.AM2015-1653 [DOI] [Google Scholar]
- 128.King M.-C. (2014) ‘The race’ to clone BRCA1. Science 343, 1462–1465 10.1126/science.1251900 [DOI] [PubMed] [Google Scholar]
- 129.Miki Y., Swensen J., Shattuck-Eidens D., Futreal P.A., Harshman K., Tavtigian S., et al. (1994) A strong candidate for the breast and ovarian cancer susceptibility gene BRCA1. Science 266, 66–71 10.1126/science.7545954 [DOI] [PubMed] [Google Scholar]
- 130.Wooster R., Bignell G., Lancaster J., Swift S., Seal S., Mangion J. et al. (1995) Identification of the breast cancer susceptibility gene BRCA2. Nature 378, 789–792 10.1038/378789a0 [DOI] [PubMed] [Google Scholar]
- 131.Kaufman B., Shapira-Frommer R., Schmutzler R.K., Audeh M.W., Friedlander M., Balmaña J. et al. (2015) Olaparib monotherapy in patients with advanced cancer and a germline BRCA1/2 mutation. J. Clin. Oncol. 33, 244–250 10.1200/JCO.2014.56.2728 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Kim G., Ison G., McKee A.E., Zhang H., Tang S., Gwise T. et al. (2015) FDA approval summary: olaparib monotherapy in patients with deleterious germline BRCA-mutated advanced ovarian cancer treated with three or more lines of chemotherapy. Clin. Cancer Res. 21, 4257–4261 10.1158/1078-0432.CCR-15-0887 [DOI] [PubMed] [Google Scholar]
- 133.Turner N., Tutt A. and Ashworth A. (2004) Hallmarks of ‘BRCAness’ in sporadic cancers. Nat. Rev. Cancer 4, 814–819 10.1038/nrc1457 [DOI] [PubMed] [Google Scholar]
- 134.Lord C.J. and Ashworth A. (2016) BRCAness revisited. Nat. Rev. Cancer 16, 110–120 10.1038/nrc.2015.21 [DOI] [PubMed] [Google Scholar]
- 135.McCabe N., Turner N.C., Lord C.J., Kluzek K., Białkowska A., Swift S. et al. (2006) Deficiency in the repair of DNA damage by homologous recombination and sensitivity to poly(ADP-ribose) polymerase inhibition. Cancer Res. 66, 8109–8115 10.1158/0008-5472.CAN-06-0140 [DOI] [PubMed] [Google Scholar]
- 136.Buisson R., Dion-Côté A.-M., Coulombe Y., Launay H., Cai H., Stasiak A.Z. et al. (2010) Cooperation of breast cancer proteins PALB2 and piccolo BRCA2 in stimulating homologous recombination. Nat. Struct. Mol. Biol. 17, 1247–1254 10.1038/nsmb.1915 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Williamson C.T., Muzik H., Turhan A.G., Zamò A., O'Connor M.J., Bebb D.G. et al. (2010) ATM deficiency sensitizes mantle cell lymphoma cells to poly(ADP-ribose) polymerase-1 inhibitors. Mol. Cancer Ther. 9, 347–357 10.1158/1535-7163.MCT-09-0872 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Weston V.J., Oldreive C.E., Skowronska A., Oscier D.G., Pratt G., Dyer M.J. et al. (2010) The PARP inhibitor olaparib induces significant killing of ATM-deficient lymphoid tumor cells in vitro and in vivo. Blood 116, 4578–4587 10.1182/blood-2010-01-265769 [DOI] [PubMed] [Google Scholar]
- 139.Bell D., Berchuck A., Birrer M., Chien J., Cramer D., Dao F., et al. (2011) Integrated genomic analyses of ovarian carcinoma. Nature 474, 609–615 10.1038/nature10166 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Mateo J., Carreira S., Sandhu S., Miranda S., Mossop H., Perez-Lopez R., et al. (2015) DNA-repair defects and olaparib in metastatic prostate cancer. N. Engl. J. Med. 373, 1697–1708 10.1056/NEJMoa1506859 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Waddell N., Pajic M., Patch A.-M., Chang D.K., Kassahn K.S., Bailey P., et al. (2015) Whole genomes redefine the mutational landscape of pancreatic cancer. Nature 518, 495–501 10.1038/nature14169 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Carnevale J. and Ashworth A. (2015) Assessing the significance of BRCA1 and BRCA2 mutations in pancreatic cancer. J. Clin. Oncol. 33, 3080–3081 10.1200/JCO.2015.61.6961 [DOI] [PubMed] [Google Scholar]
- 143.Zhao L. and So C.W. (2016) PARP-inhibitor-induced synthetic lethality for acute myeloid leukemia treatment. Exp. Hematol. 44, 902–907 10.1016/j.exphem.2016.07.007 [DOI] [PubMed] [Google Scholar]
- 144.Michelena J., Lezaja A., Teloni F., Schmid T., Imhof R. and Altmeyer M. (2018) Analysis of PARP inhibitor toxicity by multidimensional fluorescence microscopy reveals mechanisms of sensitivity and resistance. Nat. Commun. 9, 2678 10.1038/s41467-018-05031-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Liszczak G., Diehl K.L., Dann G.P. and Muir T.W. (2018) Acetylation blocks DNA damage-induced chromatin ADP-ribosylation. Nat. Chem. Biol. 14, 837–840 10.1038/s41589-018-0097-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Fong P.C., Boss D.S., Yap T.A., Tutt A., Wu P., Mergui-Roelvink M. et al. (2009) Inhibition of poly(ADP-ribose) polymerase in tumors from BRCA mutation carriers. N. Engl. J. Med. 361, 123–134 10.1056/NEJMoa0900212 [DOI] [PubMed] [Google Scholar]
- 147.Murai J., Huang S.-Y.N., Renaud A., Zhang Y., Ji J., Takeda S. et al. (2014) Stereospecific PARP trapping by BMN 673 and comparison with olaparib and rucaparib. Mol. Cancer Ther. 13, 433–443 10.1158/1535-7163.MCT-13-0803 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Pommier Y., O'Connor M.J. and de Bono J. (2016) Laying a trap to kill cancer cells: PARP inhibitors and their mechanisms of action. Sci. Transl. Med. 8, 362ps17 10.1126/scitranslmed.aaf9246 [DOI] [PubMed] [Google Scholar]
- 149.Bunting S.F., Callén E., Wong N., Chen H.-T., Polato F., Gunn A. et al. (2010) 53BP1 inhibits homologous recombination in Brca1-deficient cells by blocking resection of DNA breaks. Cell 141, 243–254 10.1016/j.cell.2010.03.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Jaspers J.E., Kersbergen A., Boon U., Sol W., van Deemter L., Zander S.A. et al. (2013) Loss of 53BP1 causes PARP inhibitor resistance in Brca1-mutated mouse mammary tumors. Cancer Discov. 3, 68–81 10.1158/2159-8290.CD-12-0049 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Xu G., Chapman J.R., Brandsma I., Yuan J., Mistrik M., Bouwman P. et al. (2015) REV7 counteracts DNA double-strand break resection and affects PARP inhibition. Nature 521, 541–544 10.1038/nature14328 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Gogola E., Duarte A.A., de Ruiter J.R., Wiegant W.W., Schmid J.A., de Bruijn R. et al. (2018) Selective loss of PARG restores PARylation and counteracts PARP inhibitor-Mediated synthetic lethality. Cancer Cell 33, 1078–1093.e12 10.1016/j.ccell.2018.05.008 [DOI] [PubMed] [Google Scholar]
- 153.Chaudhuri A.R., Callen E., Ding X., Gogola E., Duarte A.A., Lee J.-E. et al. (2016) Replication fork stability confers chemoresistance in BRCA-deficient cells. Nature 535, 382–387 10.1038/nature18325 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Rondinelli B., Gogola E., Yücel H., Duarte A.A., van de Ven M., van der Sluijs R. et al. (2017) EZH2 promotes degradation of stalled replication forks by recruiting MUS81 through histone H3 trimethylation. Nat. Cell Biol. 19, 1371–1378 10.1038/ncb3626 [DOI] [PubMed] [Google Scholar]
- 155.Pettitt S.J., Rehman F.L., Bajrami I., Brough R., Wallberg F., Kozarewa I. et al. (2013) A genetic screen using the PiggyBac transposon in haploid cells identifies Parp1 as a mediator of olaparib toxicity. PLoS ONE 8, e61520 10.1371/journal.pone.0061520 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Edwards S.L., Brough R., Lord C.J., Natrajan R., Vatcheva R., Levine D.A. et al. (2008) Resistance to therapy caused by intragenic deletion in BRCA2. Nature 451, 1111–1115 10.1038/nature06548 [DOI] [PubMed] [Google Scholar]
- 157.Barber L.J., Sandhu S., Chen L., Campbell J., Kozarewa I., Fenwick K. et al. (2013) Secondary mutations in BRCA2 associated with clinical resistance to a PARP inhibitor. J. Pathol. 229, 422–429 10.1002/path.4140 [DOI] [PubMed] [Google Scholar]
- 158.Sakai W., Swisher E.M., Karlan B.Y., Agarwal M.K., Higgins J., Friedman C. et al. (2008) Secondary mutations as a mechanism of cisplatin resistance in BRCA2-mutated cancers. Nature 451, 1116–1120 10.1038/nature06633 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Norquist B., Wurz K.A., Pennil C.C., Garcia R., Gross J., Sakai W. et al. (2011) Secondary somatic mutations restoring BRCA1/2 predict chemotherapy resistance in hereditary ovarian carcinomas. J. Clin. Oncol. 29, 3008–3015 10.1200/JCO.2010.34.2980 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Lord C.J. and Ashworth A. (2013) Mechanisms of resistance to therapies targeting BRCA-mutant cancers. Nat. Med. 19, 1381–1388 10.1038/nm.3369 [DOI] [PubMed] [Google Scholar]
- 161.Ceccaldi R., Liu J.C., Amunugama R., Hajdu I., Primack B. and Petalcorin M.I. (2015) O'connor KW, Konstantinopoulos PA, Elledge SJ, Boulton SJ, Yusufzai T, D'Andrea AD. Homologous-recombination-deficient tumours are dependent on Polθ-mediated repair. Nature 518, 258–262 10.1038/nature14184 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Murfuni I., Basile G., Subramanyam S., Malacaria E., Bignami M., Spies M. et al. (2013) Survival of the replication checkpoint deficient cells requires MUS81-RAD52 function. PLoS Genet. 9, e1003910 10.1371/journal.pgen.1003910 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163.Hengel S.R., Spies M.A. and Spies M. (2017) Small-molecule inhibitors targeting DNA repair and DNA repair deficiency in research and cancer therapy. Cell Chem. Biol. 24, 1101–1119 10.1016/j.chembiol.2017.08.027 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.James D.I., Smith K.M., Jordan A.M., Fairweather E.E., Griffiths L.A., Hamilton N.S. et al. (2016) First-in-class chemical probes against poly(ADP-ribose) glycohydrolase (PARG) inhibit DNA repair with differential pharmacology to olaparib. ACS Chem. Biol. 11, 3179–3190 10.1021/acschembio.6b00609 [DOI] [PubMed] [Google Scholar]
- 165.Gravells P., Neale J., Grant E., Nathubhai A., Smith K.M., James D.I. et al. (2018) Radiosensitization with an inhibitor of poly(ADP-ribose) glycohydrolase: a comparison with the PARP1/2/3 inhibitor olaparib. DNA Repair 61, 25–36 10.1016/j.dnarep.2017.11.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Smith S., Giriat I., Schmitt A. and de Lange T. (1998) Tankyrase, a poly(ADP-ribose) polymerase at human telomeres. Science 282, 1484–1487 10.1126/science.282.5393.1484 [DOI] [PubMed] [Google Scholar]
- 167.Nagy Z., Kalousi A., Furst A., Koch M., Fischer B. and Soutoglou E. (2016) Tankyrases promote homologous recombination and check point activation in response to DSBs. PLoS Genet. 12, e1005791 10.1371/journal.pgen.1005791 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Smith S. and de Lange T. (1999) Cell cycle dependent localization of the telomeric PARP, tankyrase, to nuclear pore complexes and centrosomes. J. Cell Sci. 112, 3649–3656 [DOI] [PubMed] [Google Scholar]
- 169.Dynek J.N. and Smith S. (2004) Resolution of sister telomere association is required for progression through mitosis. Science 304, 97–100 10.1126/science.1094754 [DOI] [PubMed] [Google Scholar]
- 170.Chang P., Coughlin M. and Mitchison T.J. (2005) Tankyrase-1 polymerization of poly(ADP-ribose) is required for spindle structure and function. Nat. Cell Biol. 7, 1133–1139 10.1038/ncb1322 [DOI] [PubMed] [Google Scholar]
- 171.Huang S.-M.A., Mishina Y.M., Liu S., Cheung A., Stegmeier F., Michaud G.A. et al. (2009) Tankyrase inhibition stabilizes axin and antagonizes Wnt signalling. Nature 461, 614–620 10.1038/nature08356 [DOI] [PubMed] [Google Scholar]
- 172.Bhardwaj A., Yang Y., Ueberheide B. and Smith S. (2017) Whole proteome analysis of human tankyrase knockout cells reveals targets of tankyrase-mediated degradation. Nat. Commun. 8, 2214 10.1038/s41467-017-02363-w [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Riffell J.L., Lord C.J. and Ashworth A. (2012) Tankyrase-targeted therapeutics: expanding opportunities in the PARP family. Nat. Rev. Drug Discov. 11, 923–936 10.1038/nrd3868 [DOI] [PubMed] [Google Scholar]
- 174.Hsiao S.J. and Smith S. (2008) Tankyrase function at telomeres, spindle poles, and beyond. Biochimie 90, 83–92 10.1016/j.biochi.2007.07.012 [DOI] [PubMed] [Google Scholar]
- 175.Mariotti L., Templeton C.M., Ranes M., Paracuellos P., Cronin N., Beuron F. et al. (2016) Tankyrase requires SAM domain-dependent polymerization to support Wnt-β-catenin signaling. Mol. Cell 63, 498–513 10.1016/j.molcel.2016.06.019 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Riccio A.A., McCauley M., Langelier M.-F. and Pascal J.M. (2016) Tankyrase sterile α motif domain polymerization Is required for Its role in Wnt signaling. Structure 24, 1573–1581 10.1016/j.str.2016.06.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Eisemann T., McCauley M., Langelier M.-F., Gupta K., Roy S., Van Duyne G.D. et al. (2016) Tankyrase-1 ankyrin repeats form an adaptable binding platform for targets of ADP-ribose modification. Structure 24, 1679–1692 10.1016/j.str.2016.07.014 [DOI] [PubMed] [Google Scholar]
- 178.Sbodio J.I. and Chi N.-W. (2002) Identification of a tankyrase-binding motif shared by IRAP, TAB182, and human TRF1 but not mouse TRF1. NuMA contains this RXXPDG motif and is a novel tankyrase partner. J. Biol. Chem. 277, 31887–31892 10.1074/jbc.M203916200 [DOI] [PubMed] [Google Scholar]
- 179.Guettler S., LaRose J., Petsalaki E., Gish G., Scotter A., Pawson T. et al. (2011) Structural basis and sequence rules for substrate recognition by Tankyrase explain the basis for cherubism disease. Cell 147, 1340–1354 10.1016/j.cell.2011.10.046 [DOI] [PubMed] [Google Scholar]
- 180.Levaot N., Voytyuk O., Dimitriou I., Sircoulomb F., Chandrakumar A., Deckert M. et al. (2011) Loss of Tankyrase-mediated destruction of 3BP2 is the underlying pathogenic mechanism of cherubism. Cell 147, 1324–1339 10.1016/j.cell.2011.10.045 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.Krastev D.B., Pettitt S.J., Campbell J., Song F., Tanos B.E., Stoynov S.S. et al. (2018) Coupling bimolecular PARylation biosensors with genetic screens to identify PARylation targets. Nat. Commun. 9, 2016 10.1038/s41467-018-04466-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182.Wu H. (2013) Higher-order assemblies in a new paradigm of signal transduction. Cell 153, 287–292 10.1016/j.cell.2013.03.013 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Bienz M. (2014) Signalosome assembly by domains undergoing dynamic head-to-tail polymerization. Trends Biochem. Sci. 39, 487–495 10.1016/j.tibs.2014.08.006 [DOI] [PubMed] [Google Scholar]
- 184.Lehtiö L., Chi N.-W. and Krauss S. (2013) Tankyrases as drug targets. FEBS J. 280, 3576–3593 10.1111/febs.12320 [DOI] [PubMed] [Google Scholar]
- 185.MacDonald B.T., Tamai K. and He X. (2009) Wnt/beta-catenin signaling: components, mechanisms, and diseases. Dev. Cell 17, 9–26 10.1016/j.devcel.2009.06.016 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186.Clevers H. and Nusse R. (2012) Wnt/β-catenin signaling and disease. Cell 149, 1192–1205 10.1016/j.cell.2012.05.012 [DOI] [PubMed] [Google Scholar]
- 187.Zhan T., Rindtorff N. and Boutros M. (2017) Wnt signaling in cancer. Oncogene 36, 1461–1473 10.1038/onc.2016.304 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Nusse R. and Clevers H. (2017) Wnt/β-catenin signaling, disease, and emerging therapeutic modalities. Cell 169, 985–999 10.1016/j.cell.2017.05.016 [DOI] [PubMed] [Google Scholar]
- 189.Thorsell A.-G., Ekblad T., Karlberg T., Löw M., Pinto A.F., Trésaugues L. et al. (2017) Structural basis for potency and promiscuity in poly(ADP-ribose) polymerase (PARP) and tankyrase inhibitors. J. Med. Chem. 60, 1262–1271 10.1021/acs.jmedchem.6b00990 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Chen B., Dodge M.E., Tang W., Lu J., Ma Z., Fan C.-W. et al. (2009) Small molecule-mediated disruption of Wnt-dependent signaling in tissue regeneration and cancer. Nat. Chem. Biol. 5, 100–107 10.1038/nchembio.137 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Gunaydin H., Gu Y. and Huang X. (2012) Novel binding mode of a potent and selective tankyrase inhibitor. PLoS ONE 7, e33740 10.1371/journal.pone.0033740 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Steffen J.D., Brody J.R., Armen R.S. and Pascal J.M. (2013) Structural implications for selective targeting of PARPs. Front. Oncol. 3, 301 10.3389/fonc.2013.00301 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Waaler J., Machon O., von Kries J.P., Wilson S.R., Lundenes E., Wedlich D. et al. (2011) Novel synthetic antagonists of canonical Wnt signaling inhibit colorectal cancer cell growth. Cancer Res. 71, 197–205 10.1158/0008-5472.CAN-10-1282 [DOI] [PubMed] [Google Scholar]
- 194.Waaler J., Machon O., Tumova L., Dinh H., Korinek V., Wilson S.R. et al. (2012) A novel tankyrase inhibitor decreases canonical Wnt signaling in colon carcinoma cells and reduces tumor growth in conditional APC mutant mice. Cancer Res. 72, 2822–2832 10.1158/0008-5472.CAN-11-3336 [DOI] [PubMed] [Google Scholar]
- 195.James R.G., Davidson K.C., Bosch K.A., Biechele T.L., Robin N.C., Taylor R.J. et al. (2012) WIKI4, a novel inhibitor of tankyrase and Wnt/β-catenin signaling. PLoS ONE 7, e50457 10.1371/journal.pone.0050457 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Haikarainen T., Venkannagari H., Narwal M., Obaji E., Lee H.-W., Nkizinkiko Y. et al. (2013) Structural basis and selectivity of tankyrase inhibition by a Wnt signaling inhibitor WIKI4. PLoS ONE 8, e65404 10.1371/journal.pone.0065404 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197.Okada-Iwasaki R., Takahashi Y., Watanabe Y., Ishida H., Saito J., Nakai R. et al. (2016) The discovery and characterization of K-756, a novel Wnt/β-Catenin pathway inhibitor targeting tankyrase. Mol. Cancer Ther. 15, 1525–1534 10.1158/1535-7163.MCT-15-0938 [DOI] [PubMed] [Google Scholar]
- 198.Lau T., Chan E., Callow M., Waaler J., Boggs J., Blake R.A. et al. (2013) A novel tankyrase small-molecule inhibitor suppresses APC mutation-driven colorectal tumor growth. Cancer Res. 73, 3132–3144 10.1158/0008-5472.CAN-12-4562 [DOI] [PubMed] [Google Scholar]
- 199.Voronkov A., Holsworth D.D., Waaler J., Wilson S.R., Ekblad B., Perdreau-Dahl H. et al. (2013) Structural basis and SAR for G007-LK, a lead stage 1,2,4-triazole based specific tankyrase 1/2 inhibitor. J. Med. Chem. 56, 3012–3023 10.1021/jm4000566 [DOI] [PubMed] [Google Scholar]
- 200.Shultz M.D., Cheung A.K., Kirby C.A., Firestone B., Fan J., Chen C.H. et al. (2013) Identification of NVP-TNKS656: the use of structure-efficiency relationships to generate a highly potent, selective, and orally active tankyrase inhibitor. J. Med. Chem. 56, 6495–6511 10.1021/jm400807n [DOI] [PubMed] [Google Scholar]
- 201.Zhong Y., Katavolos P., Nguyen T., Lau T., Boggs J., Sambrone A. et al. (2016) Tankyrase inhibition causes reversible intestinal toxicity in mice with a therapeutic index< 1. Toxicol. Pathol. 44, 267–278 10.1177/0192623315621192 [DOI] [PubMed] [Google Scholar]
- 202.McCabe N., Cerone M.A., Ohishi T., Seimiya H., Lord C.J. and Ashworth A. (2009) Targeting tankyrase 1 as a therapeutic strategy for BRCA-associated cancer. Oncogene 28, 1465–1470 10.1038/onc.2008.483 [DOI] [PubMed] [Google Scholar]
- 203.Palazzo L., Della Monica R., Visconti R., Costanzo V. and Grieco D. (2014) ATM controls proper mitotic spindle structure. Cell Cycle 13, 1091–1091 10.4161/cc.27945 [DOI] [PMC free article] [PubMed] [Google Scholar]