ADP-ribosylation has primarily been known as post-translational modification of proteins. As signalling strategy conserved in all domains of life, it modulates substrate activity, localisation, stability or interactions, thereby regulating a variety of cellular processes and microbial pathogenicity. Yet over the last years, there is increasing evidence of non-canonical forms of ADP-ribosylation that are catalysed by certain members of the ADP-ribosyltransferase family and go beyond traditional protein ADP-ribosylation signalling. New macromolecular targets such as nucleic acids and new ADP-ribose derivatives have been established, notably extending the repertoire of ADP-ribosylation signalling. Based on the physiological relevance known so far, non-canonical ADP-ribosylation deserves its recognition next to the traditional protein ADP-ribosylation modification and which we therefore review in the following.

ADP-ribosylation is a multifaceted modification of macromolecules that regulates a variety of cellular processes ranging from DNA damage repair, chromatin and telomere-related dynamics, RNA biogenesis, to stress and immune responses including antiviral defence as well as microbial metabolism, pathogenicity and nitrogen fixation [1–6]. The diverse enzyme superfamily of ADP-ribosyltransferases (ARTs) catalyses the ADP-ribosylation reaction which is characterised by the transfer of ADP-ribose from nicotinamide adenine dinucleotide (NAD+) onto target substrates via N-, O-, or S-glycosidic linkages with concomitant release of nicotinamide. While the majority of ARTs transfer only single ADP-ribose units onto their targets [7], resulting in substrate mono-ADP-ribosylation (MARylation), a subset of eukaryotic ARTs is also capable of repeatedly attaching an ADP-ribose unit to the respective preceding one. The latter process is known as poly-ADP-ribosylation (PARylation) forming long poly-ADP-ribose (PAR) chains reaching up to 200 units with occasional branching [8,9]. Whether ADP-ribose is attached as monomer or as linear or branched oligomers, influences the outcome for following downstream processes [10–12]. The structural heterogeneity and characteristic negative charge of PAR can additionally alter the properties of the substrate, thus providing another means of target modification, and affects the biophysical properties of local subcellular environments, e.g. by regulating phase separating processes during DNA repair and stress granule formation [13,14].

ADP-ribosylation has traditionally been considered as a post-translational modification (PTM) of proteins. All started almost 60 years ago with the identification of a polymer of ADP-ribose by Pierre Chambon and colleagues which was initially mistaken for a poly(A) reaction product while studying the RNA synthesis by RNA polymerase [15,16]. Following the observation of enzymes being present in mammalian cell extracts that can generate ADP-ribose polymers from NAD+ [17], bacterial toxins were then identified to cause pathogenicity through their function as ARTs. Diphtheria toxin, produced by Corynebacterium diphtheriae, was among the first to be characterised and found to ADP-ribosylate the eukaryotic elongation factor 2 (eEF2), which results in inhibition of protein biosynthesis and consequently toxicity [18,19]. Studies on cholera toxin isolated from Vibrio cholerae followed [20] which was shown to ADP-ribosylate a specific arginine in the regulatory subunit (Gsα) of heterotrimeric G-proteins that controls adenylate cyclase function, thus leading to unregulated production of cAMP. The elevated cAMP levels activate protein kinase A which opens normally gated channels in the plasma membrane, which results in the profuse watery diarrhea characteristic for cholera pathogenesis [21–24]. These beginnings of the ART research field defined the ADP-ribosylation of proteins as the canonical reaction of ARTs to this day. Furthermore, from those first discoveries research on ARTs and the ADP-ribosylation system expanded immensely over the next decades. ARTs were studied by phylogenetic analyses, which identified these enzymes in all domains of life and some viruses [25], and by functional and structural analyses, revealing enzymatic mechanistic details, target substrates and specificities with identification of exact modification sites. The latter benefitted greatly from proteomics approaches which confirmed the early known glutamate, aspartate, arginine and diphthamide (a modified histidine) residues as ADP-ribosylation acceptors but also identified as such serine, histidine, threonine, tyrosine, lysine and cysteine [7,26–28]. Serine residues were thereby found to be the most common ADP-ribosylation amino acid targets in human cells [29–32].

ADP-ribosylation signalling is regulated by a dynamic interplay between ARTs and ADP-ribosylhydrolase enzymes which are able to reverse the reaction of the transferases by removing the ADP-ribose modifications [1,33–35]. The ART superfamily consists of more than 20 families and its members can be grouped (based on homology of their catalytic domain to the first characterised bacterial ART toxins) into the diphtheria toxin-like ADP-ribosyltransferases (ARTDs) and cholera toxin-like ADP-ribosyltransferases (ARTCs). Four ARTCs are expressed in humans which are extracellular and partly membrane-associated or secreted proteins [27,36]. Most of the eukaryotic members of the ARTD family are referred to as PARPs which form the largest subfamily of ARTs with seventeen members in humans. PARP1, the enzyme involved in DNA repair and chromatin regulation, is the founding member of PARPs, the best characterised so far as well as the also most ubiquitous expressed and abundant PARP [37,38]. Amongst the PARPs, PARP1, 2 and the tankyrases (TNKS1 and 2) show PARylation activity, while all others, with the exception of the catalytic inactive PARP13, catalyse MARylation [27]. Furthermore, the tRNA phosphotransferase TRPT1 (KptA in bacteria/Tpt1 in yeast) was identified as a highly divergent PARP family member that is conserved over all domains of life and catalyses tRNA dephosphorylation [39,40]. Finally, the sirtuin family, that is evolutionary unrelated to ARTs, contains also members capable of catalysing an ADP-ribose transfer reaction, by acting usually as protein deacetylases of lysines producing an O-acetyl-ADP-ribose (OAADPr) molecule [41] or less frequently by ADP-ribosylating proteins [42].

The ART fold is structurally highly conserved and binds NAD+ in a bent and constrained confirmation. Characteristic for the ART fold are mainly two known conserved three amino acid motifs that are critical for NAD+ binding and catalysis [43]. The identity of the motif, either [H-Y-E] present in ARTDs or [R-S-E] in ARTCs (including variants of both motifs), allows classification of ARTs to those respective subfamilies [36,44]. Both motifs have a glutamate in common which is a key residue for catalysing the ADP-ribosylation reaction and ART variants lacking this acidic residue itself substitute it from their substrate through a process termed substrate-assisted catalysis [45]. The ADP-ribosylation reaction generally proceeds via an SN1 mechanism which is characterised by NAD+ cleavage, generating a reactive oxocarbenium ion as an intermediate for nucleophilic attack of the targeted acceptor. This is accompanied by anomeric inversion of the carbon C1 of the adenine-distal ribose and leaving of NAM as a reaction by-product (Figure 1) [44,46,47].

Reversal of ADP-ribosylation is provided by enzymes capable of cleaving the ADP-ribose modifications from the respective targets. Two evolutionary unrelated protein families are characterised to catalyse the reversal of the canonical protein ADP-ribosylation, i.e. macrodomain-containing enzymes (e.g. PARG, MacroD1/2, TARG1) and the ADP-ribosylhydrolases (e.g. ARH1, ARH3), which show differences regarding target residue specificities and activities [35,48]. PARG efficiently degrades PAR chains, however, is unable to remove the terminal protein-linked ADP-ribose unit [49,50]. This is taken over by MARylation-reversing hydrolases, whereby TARG1, MacroD1/2 show strong activity on acidic residues [51,52], ARH1 reverses arginine ADP-ribosylation [53,54] and ARH3 specifically removes serine-linked ADP-ribose modifications [55,56]. Furthermore, removal of ADP-ribosylation was seen to be catalysed by the NUDIX family member NUDT16 as well as ENPP1 which both cleave the ADP-ribose pyrophosphate bond, thus remove AMP and consequently leave a phosphoribosyl moiety on the target protein [57,58].

The concerted action and interplay of ARTs and hydrolases on a specific target residues is mechanistically best understood for serine-linked ADP-ribosylation in the context of DNA damage response [59]. For this, PARP1/2 are essential but not sufficient as it requires HPF1 as accessory factor to direct the ADP-ribosylation of serine and the synthesis of longer or shorter PAR chains [32,60,61]. By forming a composite active site with HPF1, PARP1/2 substrate specificity is switched from acidic residues towards serine residues on target substrates [62], whereby the PARP2/HPF1 complex was shown to bridge two DSBs in a conformation compatible with DNA ligation after double-strand breaks [63]. Reversal of serine ADP-ribosylation is controlled by the PAR-degrading PARG [50,64] followed by removal of the remaining ADP-ribose unit by ARH3 [55,56,60]. This way, serine ADP-ribosylation signalling is regulated in a tight and timely manner and with it the numerous associated processes utilising it for preserving genome stability [32,65].

Research on ADP-ribosylation signalling with its players is moving fast and expanding the last years owing to new methodologies and tools which are able to uncover new aspects including substrate targets, catalytic mechanisms of ARTs and hydrolases, as well as the cross-talk and combined signalling of ADP-ribosylation with other modifications such as ubiquitination. With in particular increasingly more ADP-ribosylation reactions and products uncovered that are seen as ‘atypical' or ‘special cases', it yet becomes clear that the ADP-ribosylation mark is far more than a protein PTM in the classical sense. This prompts to rethink the traditional view on ADP-ribosylation and with this review, we use the opportunity to provide an overview of those non-canonical ADP-ribosylation reactions (Figure 1).

ADP-ribosylation of guanosines

One of the first non-canonical ADP-ribosylation of guanosines reaction was described with the discovery of the toxin pierisin-1 isolated from cabbage butterfly larvae, Pieris rapae, that showed mono-ADP-ribosylation activity specifically on nucleic acids instead on protein targets [66]. More pierisin family members (pierisin-1–5, ScARP, Scabin) and related proteins including CARP-1 from shellfish species were identified which all were characterised to belong to the ARTC subclass and to attach ADP-ribose to the N2 amino group of guanosine bases in either dsDNA, ssDNA or guanine-derived nucleosides (Figure 2, top left panel) [67–70]. Thus, although the same guanine specificity is shared, the pierisin-like enzymes show differences in substrate preference and their reaction has also a relaxed specificity regarding the target DNA motif. The mechanism of the enzymatic activity and the substrate recognition of pierisin-like ARTs has been best understood on the models of pierisin-1, ScARP and Scabin [68,70,71]. Studies on piersin-1 revealed that the activity of its N-terminal catalytic domain is controlled (in contrast with ScARP and Scabin) by an autoinhibitory linker that connects to the C-terminal ricin B-like domains. The latter enable binding to surface glycosphingolipid receptors for internalisation into cells. Following its incorporation into lysozymes, pierisin-1 is cleaved which releases the catalytic cleft-occupying linker region and consequently activates the enzyme. In this form, pierisin-1 is released into the cytosol and migrates into the nucleus to target genomic DNA for ADP-ribosylation [71]. Furthermore, a complex structure of ScARP with NADH and GDP provided details regarding the target recognition by the ART-conserved substrate binding region, the ADP-ribosylating turn-turn (ARTT) [68]. And finally, complementing kinetic studies of the DNA ADP-ribosylation reaction were performed with Scabin, a toxin secreted from the plant pathogen Streptomyces scabies, that shows preference for modifying dsDNA with a single-base overhang on either terminus, i.e. nicked dsDNA substrates [70].

So far, the ADP-ribosylation reaction catalysed by pierisin-like ART members is understood for being irreversible which induces strong genotoxicity and cytotoxicity including in several human cancer cell lines, making such toxins likely to be utilised as protective agents against microbes, viruses or (in the case of butterflies) parasitic wasps [72,73]. Yet apart from the natural purpose of these toxins, in an application-oriented study silkworms were bioengineered to express a less toxic variant of piersin-1 which instead of inducing apoptosis resulted in the dysfunction of their silk glands. The modified silkworms consequently produced cocoons without the silk protein fibroin and solely with the glue-like glycoprotein sericin which could be used as new biomedical material e.g. for tissue engineering [74].

ADP-ribosylation of thymidines

The ARTC-class pierisin family members are not the only ARTs found to be capable of ADP-ribosylation of DNA bases. In 2016, a bacterial ARTD family member phylogenetically related to the eukaryotic PARPs was identified that catalyses the mono-ADP-ribosylation of thymidines (Figure 2, bottom left panel) and therefore named based on its general enzymatic activity DNA-ADP-ribosyltransferase, DarT [75]. In contrast with pierisins, the ADP-ribosylation reaction by DarT is characterised through target specificity and reversibility. Thus, DarT specifically modifies thymidine bases present in ssDNA, thereby showing no activity on other macro-biomolecules such as dsDNA, RNA or proteins. Its sequence specificity somewhat varies among different bacterial species, yet a four-base motif can be generalised for all DarTs with flexibility in all nucleotide positions apart from the third as being the thymidine that is targeted for ADP-ribosylation. As example, DarT of Thermus aquaticus and enteropathogenic E. coli (EPEC), which were among the first biochemically characterised variants, show a motif preference for TNTC and TTT/TCT, respectively (the underlined T is ADP-ribosylated) [75,76]. Only recently, mechanistic studies on Thermus sp. 2.9 revealed that DarT links ADP-ribose with the anomeric carbon of the adenine-distal ribose to the in-ring nitrogen of the thymidine base. The reaction requires an additional catalytic arginine residue in the active site that is assumed to be particularly essential for proton abstraction from the thymine nitrogen and that extends the canonical set of so far known ART catalytic residues [77]. It is indicative for the evolving diversity of ARTs with their ability to adapt to versatile specialised functions unrelated to the classical protein-targeting activity. The DarT-catalysed reaction can be reversed by the macrodomain-containing hydrolase DarG (DNA ADP-ribosylglycohydrolase) acting, compared with DarT, in a motif sequence-independent manner, thus is also able to reverse DarT reactions from non-cognate species [75,77]. By being downstream-coded in the same operon, DarG is genetically linked to DarT and together, DarT and DarG (DarTG) form a toxin-antitoxin (TA) pair. The DarTG TA system was the first to be discovered utilising ADP-ribosylation and is found in many bacterial including pathogenic species such as Mycobacterium tuberculosis, Pseudomonas aeruginosa, Acinetobacter baumannii and EPEC [75]. As the toxin of the system, DarT induces strong bacteriostatic effects and activates the SOS-response since the thymidine-linked ADP-ribosylation modifications are perceived as severe DNA damage requiring two consecutive DNA-repair pathways (HR and NER) to be resolved in a DarG-independent manner [75,76]. DarT expression was also shown to exert highly toxic effects in human cells which can however be protected by the endogenous human TARG1 enzymatic activity [78]. TARG1 displays close structural homology to DarG sharing the same catalytic lysine residue in the active site [52] and was found, like DarG, to be able to reverse thymidine-linked DNA ADP-ribosylation. Consequently, the expression of bacterial DarT toxin in TARG1-deficient human cells causes extreme replication stress impacting replication fork progression and activating the DNA-damage response at DNA replication sites [78]. For bacteria, DarG — that is functioning as the antitoxin — is an essential gene for survival and relevant for enabling smooth replication processes in DarT-expressing cells by providing DarT control and regulation via the hydrolytic activity of its macrodomain as well as physical sequestration through complex formation [75,76]. The relevance of DarG for bacterial survival and growth has been started to be understood in mycobacteria in which DarTG is among the three (out of ∼80 putative) TA systems that encode an essential antitoxin as demonstrated by transposon mutagenesis studies [79]. Mechanistically, mycobacterial DarT was shown to specifically ADP-ribosylate TTTW sequences that are abundant at the origin of chromosome replication (OriC). These OriC ADP-ribose modifications are assumed to impair the loading and DNA unwinding activity of the main replicative helicase DnaB (which is furthermore transcriptionally linked to DarTG), resulting in the phenotypic control of bacterial growth [77]. It is noteworthy, that carefully controlled, slow and nonreplicating growth states are key for M. tuberculosis, resulting in persistent, potentially life-long infection and antibiotic tolerance. DarT activity could therefore be a strategy employed by bacteria to induce persistence, i.e. a dormancy-like state that has been involved in gaining antibiotic resistance [75,80]. The latter was also seen as a result of increased mutability of mycobacterial strains upon experimental depletion of DarG [81]. Thus, it seems conclusive that DNA base ADP-ribosylation is used not only for targeted DNA damage to induce host-protective toxicity but instead is a far more complex signalling strategy for regulating bacterial physiology and pathogenicity. The observation that in pathogenic E. coli and cyanobacteria DarTG is often found to be inserted in an operon structure containing a type I restriction modification system suggested a link of DarTG to antiphage response and immunity [75,76]. Indeed, recently a role of DarT in antiphage defence was uncovered, where DarT is thought to modify invading viral DNA, thereby preventing its replication and consequently the production of virions, as a mechanism to control phage infections [82].

Phosphate-linked DNA/RNA ADP-ribosylation

While the ADP-ribosylation of DNA bases requires specific recognition of the nucleotide by the ART, a more general way for DNA ADP-ribosylation signalling is the attachment of ADP-ribose to the nucleic acid backbone — a reaction that is catalysed by different PARP family members (Figure 2, top right panel) [83]. Corresponding to their observed ADP-ribosylation activity on protein substrates, PARP1 and PARP2 were shown to directly PARylate the terminal phosphate ends of DNA whereas PARP3 was characterised for MARylation of the latter [84]. The enzymes thereby show slight differences regarding their substrate preferences: For PARP1, the 3′-terminal phosphate at dsDNA break (blunt) ends on gapped DNA duplexes were found as major acceptor sites [85], while PARP2/3 preferentially target the 5′-terminal phosphate at blunt ends of nicked dsDNA [84,86]. The 5′-phosphorylated DNA breaks, which are recognised via DNA-binding domains, are thought to function as allosteric activators of the ADP-ribosylation activity of PARP1-3 by inducing structural alterations in the catalytic domain and thus relieving the autoinhibitory state [87]. The tRNA 2′-phosphotransferase 1 TRPT1/Tpt1/KptA also targets the 5′-phosphorylated ends for ADP-ribosylation, yet of ssDNA, which was shown as conserved Tpt1 activity among pro-/eukaryotic and archaeal species [88,89]. The DNA mono-ADP-ribosylation reactions catalysed by PARPs and TRPT1/Tpt1/KptA were all shown to be reversible by hydrolases including PARG, MacroD2, TARG1 and ARH3 [84]. Although these phosphate-linked ADP-ribosylation products were so far only observed in vitro, there is strong indication for their in vivo existence and thus physiological relevance [86]. The substrates used for testing PARP DNA ADP-ribosylation activity represent different types of DNA damage, hence allow to speculate about potential roles of this modification in the respective DNA repair pathways. For example, ADP-ribosylation at DNA ends could facilitate recruitment of DNA repair factors, protect DNA ends from unregulated nuclease activity [84] or serve as substrate for DNA ligases facilitating dsDNA ligation [90]. Finally, it is also hypothesised that this type of DNA ADP-ribosylation may be an erroneous activity of PARPs leaving ADP-ribose-DNA adducts similar to DNA adenylates formed upon abortive DNA ligations which are then repaired by PARG in a non-canonical DNA repair reaction and similar to aprataxin [84,91].

Similarly to DNA, the 5′-phosphate termini of RNA ends are targeted for MARylation by the human PARP10, 11, 15 and TRPT1 [88], as well as by the TRPT1/Tpt1/KptA homologues from lower organisms (Figure 2, bottom right panel) [88,89]. RNA ADP-ribosylation is also a reversible process through hydrolytic activity of PARG, TARG1, MacroD1/2 and ARH3 [88]. Interestingly, phosphate-linked RNA MARylation catalysed by PARP10 can also be reversed by viral macrodomains including the ones encoded by coronaviruses [88] and directly links this modification to potential functions in an immunity-related context. PARP10 is known to be induced by interferons, to regulate NF-κB signalling and to have inhibitory effects on viral replication including the positive-sense RNA alphavirus VEEV [92–94]. The RNA-recognizing motif (RRM) domains in PARP10 may be responsible to differentiate host from invading viral RNA on which the ADP-ribose modification is then terminally attached by the PARP10 catalytic domain. The phosphate-linked ADP-ribosylation is hypothesised to act as a non-canonical RNA cap preventing viral RNA translation or triggering signal transduction [88] while the viral macrodomain evolved to counteract this antiviral response. More generally, terminal RNA ADP-ribosylation may be seen as non-canonical RNA capping providing RNA stability against nuclease-mediated degradation [88] or involved in RNA signal transduction [95].

In conclusion, over the last years ADP-ribosylation of nucleic acids was established as a function of well-known and newly identified ARTs and will find its relevance in different physiological contexts next to protein ADP-ribosylation. Genomic data provides evidence that DNA ADP-ribosylating enzyme systems are conserved in a variety of organisms of all domains of life and the evolution of specific enzymatic proteins to guarantee the reversibility of the reaction (i.e. transferases/hydrolases) further underlines that nucleic acid ADP-ribosylation likely presents a wide-spread form of ADP-ribosylation signalling. The current technical challenges regarding its detection and tracing are about to be overcome which will shed light on this so far largely unexplored facet of ART research.

Several classes of enzymes are known to produce ADP-ribosylated small chemical molecules as by-products of their activities (Figure 3). Among these are the sirtuins that produce OAADPr by catalysing the deacetylation of lysine residues from proteins including histones for the regulation of processes in all kingdoms of life (Figure 3, top left panel) [96–98]. In contrast with histone deacetylases (HDACs) and HDAC-related enzymes that utilise an active-site metal ion (Zn2+ or Fe2+) to direct a water-mediated attack hydrolysing the acetyl-lysine residue [99], the deacetylation activity of sirtuins is an NAD+-dependent catalysis, releasing nicotinamide, deacetylated lysine and OAADPr in a multi-step reaction [100,101]. OAADPr may moreover act as an important signalling molecule itself with functions implicated in TRPM2 cation channel gating, modulation of the cellular redox status as well as gene silencing by facilitating the loading of the Sir2–4 silencing complex onto nucleosomes [102,103]. Its binding to the macrodomain of the histone variant macroH2A1.1 might furthermore be relevant for inducing macroH2A1.1-dependent chromatin changes [104]. Other macrodomain-containing proteins, i.e. MacroD1/2 and TARG1, along with ARH3 are also able to recognise this NAD+ metabolite, and hydrolyse it to ADP-ribose and acetate [105–107]. The relevance of MacroD protein in OAADPr hydrolysis was confirmed in vivo using the filamentous fungus Neurospora crassa as model organism, in which deletion of MacroD led to a notable increase in OAADPr levels [105].

The highly conserved Tpt1 family presents another class of enzymes that produces ADP-ribosylated signalling molecules by transferring a single ADP-ribose unit to terminal 2′-phosphates of tRNA which is an intermediate step in the tRNA splicing process in plants, fungi and yeasts [39]. The ADP-ribosylation reaction is followed by non-enzymatic generation of ADP-ribose-cyclic phosphate, releasing the mature tRNA (Figure 3, top right panel) [108]. Since prokaryotes and other eukaryotes seem to lack intron-containing tRNAs requiring splicing, these Tpt1 homologues likely exert their functions as phosphotransferases on alternative targets, including through their ADP-ribosylation activity on 5′-phosphorylated RNA ends as mentioned above. Although it remains to be clarified whether ADP-ribose-cyclic phosphate has itself a signalling function, it is known to be hydrolysed to ADP-ribose by yeast Poa1P protein and other homologous macrodomain proteins including MacroD1/2 [109,110].

Recently, a novel type II TA system, referred to as MbcTA, was identified that is found in multiple mycobacterial species and that encodes a toxin, MbcT, showing structural similarity to ARTs as well as NADases [111]. MbcT was characterised as a NAD+ phosphorylase generating ADP-ribose-1″-phosphate by catalysing the transfer of ADP-ribose from NAD+ onto inorganic phosphates, whereby the latter functions as stimulator for the reaction itself. In the absence of the antitoxin MbcA, MbcT activity was found to result in depletion of intracellular NAD+ which triggered rapid cell death in M. tuberculosis, consequently prolonging the survival of infected mice (Figure 3, bottom left panel) [111].

Finally, some ARTs are known to have evolved in bacteria as a defence mechanism against viruses, other bacterial species and also including antimicrobial molecules [44]. Among these is the Arr-ms family, named after the rifampin ADP-ribosyltransferases (Arr), whose members belong to the ARTD-class. They were identified in mycobacterial species as well as gram-negative pathogenic bacteria, that catalyse the ADP-ribosylation of rifamycin antibiotics as a strategy to confer antibiotic resistance for the bacterium [112,113]. The antibacterial action of rifamycins is based on their inhibition of the DNA-dependent RNA polymerase through binding with high affinity [114]. Arr enzymes attach the ADP-ribose unit to the hydroxyl moiety on carbon 23 which interacts with the amide backbone of the RNA polymerase. ADP-ribosylation thus interrupts this interaction and leads to inactivation of the antibiotic (Figure 3, bottom right panel) [112,114]. The observations by Baysarowich et al. [112] that Arr enzymes show broad substrate specificity with similar kinetic constants and that specific interactions between Arr and rifampin are absent from in the co-crystal structure suggests that these ARTs likely have other cellular functions including possibly the ADP-ribosylation of other small molecules.

Besides the different substrates that can be targeted and modified by ARTs, ADP-ribosylation of proteins itself can be subjected to variations which are beyond its signalling forms as canonical target MARylation or PARylation (Figure 4). The ART-catalysed ADP-ribose modification can be derivatised on the target protein, providing a possibility to escape the reversibility of the reaction through common hydrolases and making it a more persistent signal [115]. Furthermore, ADP-ribosylation activity and substrate specificity can be additionally regulated by being dependent on prior target modification by other PTMs [42]. Some of the latter are known to cross-talk with ADP-ribosylation signalling which in case of ubiquitin as protein PTM includes the possibility to be a target for ADP-ribosylation itself and thus, to combine PTM signals [116,117]. Moreover, ADP-ribosylation can function as a linker enabling the attachment of macromolecules to the target substrate [118–122].

ADP-riboxanation

The human pathogen Shigella flexneri is among the few bacterial species living freely in the host cytosol, thus inevitably exposing its lipopolysaccharides (LPS) to the inflammasome and inflammatory caspases [123]. It evolved therefore a virulence mechanism that prevents LPS-induced pyroptosis that is mediated by caspase-11 or caspase-4 using a type III secretion system (T3SS) effector, OspC3 [115]. OspC3 targets caspase-conserved arginine residues in caspase 4/11 (Arg314/Arg310) by ADP-ribosylation. Yet instead of transferring the ADP-ribose unit in a canonical manner to an arginine Nω nitrogen, OspC3 first ADP-ribosylates the Nδ nitrogen which is followed by a (non-enzymatic) internal deamination initiated by the 2″-hydroxyl group of the ADP-ribose to remove one Nω nitrogen, thereby forming an oxazolidine ring. The OspC3-catalysed reaction was therefore termed ‘ADP-riboxanation’ (Figure 4, top left panel). Due to its non-canonical arginine linkage and ribose modification compared with arginine ADP-ribosylation, Li et al. [115] found that ADP-riboxanation cannot be hydrolysed by the ADP-ribosylhydrolase ARH1, which catalyses canonical ADP-ribosyl-arginine linkages, or other host hydrolases including ARH3, TARG1 and MacroD1/2. This makes ADP-riboxanation of caspase-4/11 more pathogenically advantageous which was shown to block caspase activation and cleavage of their substrate, the pore-forming protein GSDMD, due to structural interference with the GSDMD-binding exosite [115].

Lipoylation-dependent ADP-ribosylation

A distinct class of sirtuins referred to as ‘SirTMs’ and predominantly identified in bacterial and fungal pathogens was shown to lack the characteristic protein deacetylase activity of sirtuins and instead reliably catalyses the ADP-ribosylation of proteins [42]. In Staphylococcus aureus and Streptococcus pyogenes, this SirTM activity was furthermore dependent on prior lipoylation of its specific target, the lipoyl-carrier protein GcvH-L, by the lipoate-protein ligase A, LplA2. GcvH-L and LplA2 are encoded together with SirTM within the same operon which also includes a macrodomain protein, termed MacroD, that is capable of reversing GcvH-L ADP-ribosylation — in contrast with its homologues in humans (MacroD1) and E. coli (YmdB) (Figure 4, top middle panel) [42,124]. These operon-specific activities showing a cross-talk between lipoylation (probably acting as scavenger of reactive oxygen species) and MARylation (thought for guiding protein interactions) were suggested to be implicated in regulating oxidative stress response in these pathogens, thus providing a host defence mechanism [42].

ADP-ribosylation of ubiquitin

Over the last years, there have been growing examples of the interplay between ADP-ribosylation and ubiquitination. Ubiquitination is typically catalysed by the three-enzyme cascade involving ubiquitin-activating enzymes (E1s), ubiquitin-conjugating enzymes (E2s), and ubiquitin ligase enzymes (E3s), which attach the carboxy terminus of ubiquitin to, in most cases, the ε-amino group of a substrate lysine via an isopeptide bond [125]. However, heterodimerization of the E3 ligase DTX3L with its co-expressed binding partner PARP9 was shown to instead result in ADP-ribosylation of the ubiquitin carboxy terminus; the attachment of ADP-ribose to the C-terminal glycine (Gly76) of ubiquitin not only interferes with its conjugation to substrates but also with its activation and transfer along the E1–E2–E3 cascade [116,126]. The ubiquitin ADP-ribosylation activity of the DTX3L/PARP9 complex is stimulated by PAR binding of the macrodomains of PARP9. This is analogous to the allosteric activation of the E3 ligase RNF146 through PAR binding of its WWE domain [127]. PARP9 was suggested to be a catalytically active ART, restraining the E3 function of DTX3L in the context of DNA repair [116], yet the exact mechanistic role of PARP9 in the ubiquitin ADP-ribosylation process is still unclear. Further mechanistic studies moreover revealed the conserved RING-DTC (‘Deltex carboxyl-terminal’) domains from DTX3L and other human Deltex family members (DTX1–4) as being able to catalyse the linkage between ubiquitin and ADP-ribose [117]. While the RING domain recruits the ubiquitin-loaded E2, the DTC domain binds the NAD+ substrate, whereby the linker between the RING and DTC domain facilitates the juxtaposition of both domains for the ADP-ribose transfer onto the carboxylate group of ubiquitin's glycine terminus (Figure 4, top right panel). Nonspecific deubiquitinases are able to recognise and reverse the ADP-ribosylation modification on ubiquitin, indicating a dynamic nature of this signal [117]. For DTX2, it was shown that it is predominantly associated with the DNA damage response and PARP1 through binding of PARylated DNA repair proteins [128].

Phosphoribosyl-linked ubiquitination of proteins

A non-canonical type of protein ADP-ribosylation is utilised by bacterial effectors belonging to the SidE family (SdeA, SdeB, SdeC, and SidE) that are produced by Legionella pneumophila, the pathogen causing pneumonia infections known as Legionnaires’ disease [120,129,130]. SidE-type enzymes are characterised by the linkage of a phosphodiesterase (PDE) domain to an ART domain [122,131]. The combination of both active domains allows the enzymes to catalyse the conjugation of ubiquitin via a phosphoribosyl moiety to serine residues of host substrates (Figure 4, bottom left panel). Mechanistically, the ART domain (a member of the ARTC-class) first transfers ADP-ribose from NAD+ to the side chain of Arg42 on ubiquitin (Ub) to generate ADPr-Ub. This MARylation product is then recognised by the PDE domain which cleaves the ADP-ribose pyrophosphate bond resulting in phosphoribosylated ubiquitin (PR-Ub) that is then conjugated to serine residues in substrate proteins [118,121,122]. Hence, ADP-ribosylation is employed by SidE-type enzymes as an intermediate step to link ubiquitin (independently of E1 and E2 enzymes or ATP consumption) to their respective targets. The latter are several endoplasmic reticulum (ER)-associated human Rab GTPases and the ER protein reticulon 4 (RTN4) to control the dynamics of tubular ER for replication processes [129,132]. Reversal and thus regulation of the phosphoribosyl serine ubiquitination is achieved by DUPs (‘deubiquitinases for PR’), DupA and DupB, with specifically bind and cleave PR-Ub from the modified substrate serine [133]. PR-Ub itself has also a pathogenic function by inhibiting the host's conventional ubiquitination cascades and thereby impairing numerous cellular processes including mitophagy, TNF signalling, and proteasomal degradation [120].

RNAylation

A subset of regulatory RNAs is found abundantly as NAD-RNA in E. coli [134]. Most recently, an ARTC-class member produced by the T4 bacteriophage, ModB, was characterised to catalyse the attachment of RNA chains to host acceptor proteins via a diphosphoriboside linkage by utilising NAD+-capped RNAs as substrate [119]. Thus in this case, ADP-ribosylation also functions as a linker between RNA and target protein, with the modification termed ‘RNAylation’ (Figure 4, bottom right panel). ModB was shown to RNAylate specific arginine residues of its targets, including the ribosomal protein S1 (rS1) and RNase E [119,135]. The RNAylation modification could be reversed by human ARH1 in vitro [119]. The role of these RNA-protein conjugates is so far unclear but was suggested to promote recruitment of phage mRNAs for ribosomal biosynthesis. Furthermore, RNAylation destabilises rS1 which could contribute to the shut-down of host mRNA translation during T4 infection and thus promote bacterial lysis [119,135,136].

In summary, ADP-ribosylation displays a remarkable versatility. The molecular structure of the ADP-ribose unit provides several possibilities of linkages that allows ADP-ribosylation in its appearances as either monomeric or linear and branched polymeric form but also to be used for linking different biomolecules. Thereby, ARTs mainly direct — in some cases together with accessory factors — target specificity regarding the substrate type (proteins, nucleic acids, small molecules), the modification site and linkage to protein residue or nucleotide and the general form of ADP-ribosylation signal (MAR or PAR). So far, protein ADP-ribosylation is best-characterised and understood regarding its physiological relevance in various cellular processes. However, over the recent years, the ever more examples of non-canonical ADP-ribosylation reactions discovered extended the repertoire of this signalling strategy. These include new target substrates such as nucleic acids but also variations of the familiar protein ADP-ribosylation modification regarding cross-talk to other PTMs and inventive ways to modulate host-microbial interaction. Considering that ADP-ribosylation is a very ancient type of target-modifying signal with conservation among all domains of life, its versatility may have allowed to evolve many more non-canonical ADP-ribosylation-based modifications only waiting to be explored. The evolutionary spread may also indicate that these ADP-riboslyation reactions are more than just ‘special cases' or small ‘side reactions' which questions our description as being ‘non-canonical'. The heavy focus on ADP-ribosylation as protein modification may have just overshadowed the richness of ADP-ribosylation as general (protein-unrelated) signalling event. Its study will not only uncover exciting facets of life and evolution but also comes with great potential for the development of new antimicrobials and anticancer agents as well as biotechnological tools [73–75,77,78]

The authors declare that there are no competing interests associated with the manuscript.

Wellcome Trust (grant numbers 101794 and 210634), BBSRC (BB/ R007195/1), Ovarian Cancer Research Alliance (Collaborative Research Development Grant #813369) and Cancer Research UK (C35050/ A22284).

Open access for this article was enabled by the participation of University of Oxford in an all-inclusive Read & Publish pilot with Portland Press and the Biochemical Society under a transformative agreement with JISC.

We are grateful to Johannes G. M. Rack and Andreja Mikoč for the critical reading of the manuscript. Work in the Ivan Ahel laboratory is funded by the Wellcome Trust (grant numbers 101794 and 210634), BBSRC (BB/ R007195/1), Ovarian Cancer Research Alliance (Collaborative Research Development Grant #813369) and Cancer Research UK (C35050/ A22284).

ADPR

ADP-ribose

ART

ADP-ribosyltransferase

ARTDs

ADP-ribosyltransferases

LPS

lipopolysaccharides

NAD+

nicotinamide adenine dinucleotide

OAADPr

O-acetyl-ADP-ribose

PAR

poly-ADP-ribose

PDE

phosphodiesterase

PTM

post-translational modification

TA

toxin-antitoxin

1
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
2
Challa
,
S.
,
Stokes
,
M.S.
and
Kraus
,
W.L.
(
2021
)
Marts and marylation in the cytosol: biological functions, mechanisms of action, and therapeutic potential
.
Cells
10
,
313
3
Palazzo
,
L.
,
Mikolčević
,
P.
,
Mikoč
,
A.
and
Ahel
,
I.
(
2019
)
ADP-ribosylation signalling and human disease
.
Open Biol.
9
,
190041
4
Mikolčević
,
P.
,
Hloušek-Kasun
,
A.
,
Ahel
,
I.
and
Mikoč
,
A.
(
2021
)
ADP-ribosylation systems in bacteria and viruses
.
Comput. Struct. Biotechnol. J.
19
,
2366
2383
5
Crawford
,
K.
,
Oliver
,
P.L.
,
Agnew
,
T.
,
Hunn
,
B.H.M.
and
Ahel
,
I.
(
2021
)
Behavioural characterisation of macrod1 and macrod2 knockout mice
.
Cells
10
,
1
24
6
Eleazer
,
R.
and
Fondufe-Mittendorf
,
Y.N.
(
2021
)
The multifaceted role of PARP1 in RNA biogenesis
.
Wiley Interdiscip. Rev. RNA
12
,
1
21
7
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
,
1
13
8
Alvarez-gonzalez
,
R.
and
Jacobson
,
M.K.
(
1987
)
Characterization of polymers of adenosine diphosphate ribose generated in vitro and in vivo
.
Biochemistry
26
,
3218
3224
9
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
Bonfiglio
,
J.J.
,
Leidecker
,
O.
,
Dauben
,
H.
,
Longarini
,
E.J.
,
Colby
,
T.
,
San Segundo-Acosta
,
P.
et al (
2020
)
An HPF1/PARP1-based chemical biology strategy for exploring ADP-ribosylation
.
Cell
183
,
1086
1102.e23
11
Prokhorova
,
E.
,
Agnew
,
T.
,
Wondisford
,
A.R.
,
Tellier
,
M.
,
Kaminski
,
N.
,
Beijer
,
D.
et al (
2021
)
Unrestrained poly-ADP-ribosylation provides insights into chromatin regulation and human disease
.
Mol. Cell.
81
,
2640
2655.e8
12
Aberle
,
L.
,
Krüger
,
A.
,
Reber
,
J.M.
,
Lippmann
,
M.
,
Hufnagel
,
M.
,
Schmalz
,
M.
et al (
2020
)
PARP1 catalytic variants reveal branching and chain length-specific functions of poly(ADP-ribose) in cellular physiology and stress response
.
Nucleic Acids Res.
48
,
10015
10033
13
Leung
,
A.K.L.
(
2020
)
Poly(ADP-ribose): a dynamic trigger for biomolecular condensate formation
.
Trends Cell Biol.
30
,
370
383
14
Reber
,
J.M.
and
Mangerich
,
A.
(
2021
)
Why structure and chain length matter: on the biological significance underlying the structural heterogeneity of poly(ADP-ribose)
.
Nucleic Acids Res.
49
,
8432
8448
15
Chambon
,
P.
,
Weill
,
J.D.
and
Mandel
,
P.
(
1963
)
Nicotinamide mononucleotide activation of a new DNA-dependent polyadenylic acid synthesizing nuclear enzyme
.
Biochem. Biophys. Res. Commun.
11
,
39
43
16
Chambon
,
P.
,
Weill
,
J.D.
,
Doly
,
J.
,
Strosser
,
M.T.
and
Mandel
,
P.
(
1966
)
On the formation of a novel adenylic compound by enzymatic extracts of liver nuclei
.
Biochem. Biophys. Res. Commun.
25
,
638
643
17
Nishizuka
,
Y.
,
Ueda
,
K.
,
Nakazawa
,
K.
and
Hayaishi
,
O.
(
1967
)
Studies on the polymer of adenosine diphosphate ribose. I. Enzymic formation from nicotinamide adenine dinuclotide in mammalian nuclei
.
J. Biol. Chem.
242
,
3164
3171
18
Gill
,
D.M.
,
Pappenheimer
,
A.M.
,
Brown
,
R.
and
Kurnick
,
J.T.
(
1969
)
Studies on the mode of action of diphtheria toxin. VII. toxin-stimulated hydrolysis of nicotinamide adenine dinucleotide in mammalian cell extracts
.
J. Exp. Med.
129
,
1
21
19
Honjo
,
T.
,
Nishizuka
,
Y.
and
Hayaishi
,
O.
(
1968
)
Diphtheria toxin-dependent adenosine diphosphate ribosylation of aminoacyl transferase II and inhibition of protein synthesis
.
J. Biol. Chem.
243
,
3553
3555
20
Finkelstein
,
R.A.
and
LoSpalluto
,
J.J.
(
1969
)
Pathogenesis of experimental cholera: preparation and isolation of choleragen and choleragenoid
.
J. Exp. Med.
130
,
185
202
21
Cassel
,
D.
and
Pfeuffer
,
T.
(
1978
)
Mechanism of cholera toxin action: covalent modification of the guanyl nucleotide-binding protein of the adenylate cyclase system
.
Proc. Natl Acad. Sci. U.S.A.
75
,
2669
2673
22
Moss
,
J.
and
Vaughan
,
M.
(
1977
)
Mechanism of action of choleragen. Evidence for ADP ribosyltransferase activity with arginine as an acceptor
.
J. Biol. Chem.
252
,
2455
2457
23
Gill
,
D.M.
and
Meren
,
R.
(
1978
)
ADP-ribosylation of membrane proteins catalyzed by cholera toxin: basis of the activation of adenylate cyclase
.
Proc. Natl Acad. Sci. U.S.A.
75
,
3050
3054
24
DiRita
,
V.J
. (
2001
) Molecular Basis of Vibrio cholerae Pathogenesis. In
Principles of Bacterial Pathogenesis
(Eduardo A.G., ed.), pp.
457
508
,
Academic Press, San Diego, California
25
Perina
,
D.
,
Mikoč
,
A.
,
Ahel
,
J.
,
Ćetković
,
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
26
Buch-Larsen
,
S.C.
,
Hendriks
,
I.A.
,
Lodge
,
J.M.
,
Rykær
,
M.
,
Furtwängler
,
B.
,
Shishkova
,
E.
et al (
2020
)
Mapping physiological ADP-ribosylation using activated ion electron transfer dissociation
.
Cell Rep.
32
,
108176
27
Lüscher
,
B.
,
Ahel
,
I.
,
Altmeyer
,
M.
,
Ashworth
,
A.
,
Bai
,
P.
,
Chang
,
P.
et al (
2021
)
ADP-ribosyltransferases, an update on function and nomenclature
.
FEBS J.
28
Jørgensen
,
R.
,
Merrill
,
A.R.
,
Yates
,
S.P.
,
Marquez
,
V.E.
,
Schwan
,
A.L.
,
Boesen
,
T.
et al (
2005
)
Exotoxin A-eEF2 complex structure indicates ADP ribosylation by ribosome mimicry
.
Nature
436
,
979
984
29
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
,
e34334
30
Larsen
,
S.C.
,
Hendriks
,
I.A.
,
Lyon
,
D.
,
Jensen
,
L.J.
and
Nielsen
,
M.L.
(
2018
)
Systems-wide analysis of serine ADP-ribosylation reveals widespread occurrence and site-specific overlap with phosphorylation
.
Cell Rep.
24
,
2493
2505.e4
31
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
32
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
33
Pascal
,
J.M.
and
Ellenberger
,
T.
(
2015
)
The rise and fall of poly(ADP-ribose): an enzymatic perspective
.
DNA Repair
32
,
10
16
34
O'Sullivan
,
J.
,
Tedim Ferreira
,
M.
,
Gagné
,
J.P.
,
Sharma
,
A.K.
,
Hendzel
,
M.J.
,
Masson
,
J.Y.
et al (
2019
)
Emerging roles of eraser enzymes in the dynamic control of protein ADP-ribosylation
.
Nat. Commun.
10
,
1
14
35
Rack
,
J.G.M.
,
Palazzo
,
L.
and
Ahel
,
I.
(
2020
)
(ADP-ribosyl)hydrolases: structure, function, and biology
.
Genes Dev.
34
,
263
284
36
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
37
Luo
,
X.
and
Lee Kraus
,
W.
(
2012
)
On par with PARP: cellular stress signaling through poly(ADP-ribose) and PARP-1
.
Genes Dev.
26
,
417
432
38
Yamanaka
,
H.
,
Penning
,
C.A.
,
Willis
,
E.H.
,
Wasson
,
D.B.
and
Carson
,
D.A.
(
1988
)
Characterization of human poly(ADP-ribose) polymerase with autoantibodies
.
J. Biol. Chem.
263
,
3879
3883
39
Lopes
,
R.R.S.
,
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
40
Palazzo
,
L.
,
Mikoč
,
A.
and
Ahel
,
I.
(
2017
)
ADP-ribosylation: new facets of an ancient modification
.
FEBS J.
284
,
2932
2946
41
Smith
,
J.S.
,
Avalos
,
J.
,
Celic
,
I.
,
Muhammad
,
S.
,
Wolberger
,
C.
and
Boeke
,
J.D.
(
2002
)
SIR2 family of NAD+-dependent protein deacetylases
.
Methods Enzymol.
353
,
282
300
42
Rack
,
J.G.M.
,
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
43
Domenighini
,
M.
and
Rappuoli
,
R.
(
1996
)
Three conserved consensus sequences identify the NAD-binding site of ADP-ribosylating enzymes, expressed by eukaryotes, bacteria and T-even bacteriophages
.
Mol. Microbiol.
21
,
667
674
44
Aravind
,
L.
,
Zhang
,
D.
,
de Souza
,
R.F.
,
Anand
,
S.
and
Iyer
,
L.M
. (
2015
) The Natural History of ADP-Ribosyltransferases and the ADP-Ribosylation System. In
Endogenous ADP-Ribosylation
(
Koch-Nolte
,
F.
, ed.), pp.
3
32
,
Springer International Publishing
,
Cham
45
Kleine
,
H.
,
Poreba
,
E.
,
Lesniewicz
,
K.
,
Hassa
,
P.O.
,
Hottiger
,
M.O.
,
Litchfield
,
D.W.
et al (
2008
)
Substrate-assisted catalysis by PARP10 limits its activity to mono-ADP-ribosylation
.
Mol. Cell
32
,
57
69
46
Jørgensen
,
R.
,
Wang
,
Y.
,
Visschedyk
,
D.
and
Merrill
,
A.R.
(
2008
)
The nature and character of the transition state for the ADP-ribosyltransferase reaction
.
EMBO Rep.
9
,
802
809
47
Ueda
,
K.
(
1985
)
ADP-ribosylation
.
Annu. Rev. Biochem.
54
,
73
100
48
Rack
,
J.G.M.
,
Perina
,
D.
and
Ahel
,
I.
(
2016
)
Macrodomains: structure, function, evolution, and catalytic activities
.
Annu. Rev. Biochem.
85
,
431
454
49
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
,
1
8
50
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
51
Jankevicius
,
G.
,
Hassler
,
M.
,
Golia
,
B.
,
Rybin
,
V.
,
Zacharias
,
M.
,
Timinszky
,
G.
et al (
2013
)
A family of macrodomain proteins reverses cellular mono-ADP-ribosylation
.
Nat. Struct. Mol. Biol.
20
,
508
514
52
Sharifi
,
R.
,
Morra
,
R.
,
Denise
,
C.
,
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
53
Moss
,
J.
,
Oppenheimer
,
N.J.
,
West
,
R.E.
and
Stanley
,
S.J.
(
1986
)
Amino acid specific ADP-ribosylation: substrate specificity of an ADP-ribosylarginine hydrolase from Turkey erythrocytes
.
Biochemistry
25
,
5408
5414
54
Rack
,
J.G.M.
,
Ariza
,
A.
,
Drown
,
B.S.
,
Henfrey
,
C.
,
Bartlett
,
E.
,
Shirai
,
T.
et al (
2018
)
(ADP-ribosyl)hydrolases: structural basis for differential substrate recognition and inhibition
.
Cell Chem. Biol.
25
,
1533
1546.e12
55
Rack
,
J.G.M.
,
Liu
,
Q.
,
Zorzini
,
V.
,
Voorneveld
,
J.
,
Ariza
,
A.
,
Honarmand Ebrahimi
,
K.
et al (
2021
)
Mechanistic insights into the three steps of poly(ADP-ribosylation) reversal
.
Nat. Commun.
12
,
1
14
56
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
,
e28533
57
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
58
Palazzo
,
L.
,
Thomas
,
B.
,
Jemth
,
A.S.
,
Colby
,
T.
,
Leidecker
,
O.
,
Feijs
,
K.L.H.
et al (
2015
)
Processing of protein ADP-ribosylation by nudix hydrolases
.
Biochem. J.
468
,
293
301
59
Palazzo
,
L.
,
Suskiewicz
,
M.J.
and
Ahel
,
I.
(
2021
)
Serine ADP-ribosylation in DNA-damage response regulation
.
Curr. Opin. Genet. Dev.
71
,
106
113
60
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
61
Suskiewicz
,
M.J.
,
Palazzo
,
L.
,
Hughes
,
R.
and
Ahel
,
I.
(
2021
)
Progress and outlook in studying the substrate specificities of PARPs and related enzymes
.
FEBS J.
288
,
2131
2142
62
Suskiewicz
,
M.J.
,
Zobel
,
F.
,
Ogden
,
T.E.H.
,
Fontana
,
P.
,
Ariza
,
A.
,
Yang
,
J.C.
et al (
2020
)
HPF1 completes the PARP active site for DNA damage-induced ADP-ribosylation
.
Nature
579
,
598
602
63
Bilokapic
,
S.
,
Suskiewicz
,
M.J.
,
Ahel
,
I.
and
Halic
,
M.
(
2020
)
Bridging of DNA breaks activates PARP2–HPF1 to modify chromatin
.
Nature
585
,
609
613
64
Lin
,
W.
,
Amé
,
J.C.
,
Aboul-Ela
,
N.
,
Jacobson
,
E.L.
and
Jacobson
,
M.K.
(
1997
)
Isolation and characterization of the cDNA encoding bovine poly(ADP- ribose) glycohydrolase
.
J. Biol. Chem.
272
,
11895
11901
65
Hendriks
,
I.A.
,
Buch-Larsen
,
S.C.
,
Prokhorova
,
E.
,
Elsborg
,
J.D.
,
Rebak
,
A.K.L.F.S.
,
Zhu
,
K.
et al (
2021
)
The regulatory landscape of the human HPF1- and ARH3-dependent ADP-ribosylome
.
Nat. Commun.
12
,
1
16
66
Watanabe
,
M.
,
Enomoto
,
S.
,
Takamura-Enya
,
T.
,
Nakano
,
T.
,
Koyama
,
K.
,
Sugimura
,
T.
et al (
2004
)
Enzymatic properties of pierisin-1 and its N-terminal domain, a guanine-specific ADP-ribosyltransferase from the cabbage butterfly
.
J. Biochem.
135
,
471
477
67
Takamura-Enya
,
T.
,
Watanabe
,
M.
,
Totsuka
,
Y.
,
Kanazawa
,
T.
,
Matsushima-Hibiya
,
Y.
,
Koyama
,
K.
et al (
2001
)
Mono(ADP-ribosyl)ation of 2′-deoxyguanosine residue in DNA by an apoptosis-inducing protein pierisin-1 from cabbage butterfly
.
Proc. Natl Acad. Sci. U.S.A.
98
,
12414
12419
68
Yoshida
,
T.
and
Tsuge
,
H.
(
2018
)
Substrate N2 atom recognition mechanism in pierisin family DNA-targeting, guanine-specific ADP-ribosyltransferase ScARP
.
J. Biol. Chem.
293
,
13768
13774
69
Nakano
,
T.
,
Matsushima-Hibiya
,
Y.
,
Yamamoto
,
M.
,
Enomoto
,
S.
,
Matsumoto
,
Y.
,
Totsuka
,
Y.
et al (
2006
)
Purification and molecular cloning of a DNA ADP-ribosylating protein, CARP-1, from the edible clam Meretrix lamarckii
.
Proc. Natl Acad. Sci. U.S.A.
103
,
13652
13657
70
Lyons
,
B.
,
Lugo
,
M.R.
,
Carlin
,
S.
,
Lidster
,
T.
and
Rod Merrill
,
A.
(
2018
)
Characterization of the catalytic signature of Scabin toxin, a DNA-targeting ADP-ribosyltransferase
.
Biochem. J.
475
,
225
245
71
Oda
,
T.
,
Hirabayashi
,
H.
,
Shikauchi
,
G.
,
Takamura
,
R.
,
Hiraga
,
K.
,
Minami
,
H.
et al (
2017
)
Structural basis of autoinhibition and activation of the DNA-targeting ADP-ribosyltransferase pierisin-1
.
J. Biol. Chem.
292
,
15445
15455
72
Nakano
,
T.
,
Takahashi-Nakaguchi
,
A.
,
Yamamoto
,
M.
and
Watanabe
,
M
. (
2015
) Pierisins and CARP-1: ADP-Ribosylation of DNA by ARTCs in Butterflies and Shellfish. In
Endogenous ADP-Ribosylation
(
Koch-Nolte
,
F.
, ed.), pp.
127
149
,
Springer International Publishing
,
Cham
73
Kono
,
T.
,
Watanabe
,
M.
,
Koyama
,
K.
,
Kishimoto
,
T.
,
Fukushima
,
S.
,
Sugimura
,
T.
et al (
1999
)
Cytotoxic activity of pierisin, from the cabbage butterfly, Pieris rapae, in various human cancer cell lines
.
Cancer Lett.
137
,
75
81
74
Otsuki
,
R.
,
Yamamoto
,
M.
,
Matsumoto
,
E.
,
Iwamoto
,
S.I.
,
Sezutsu
,
H.
,
Suzui
,
M.
et al (
2017
)
Bioengineered silkworms with butterfly cytotoxinmodified silk glands produce sericin cocoons with a utility for a new biomaterial
.
Proc. Natl Acad. Sci. U.S.A.
114
,
6740
6745
75
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
76
Lawarée
,
E.
,
Jankevicius
,
G.
,
Cooper
,
C.
,
Ahel
,
I.
,
Uphoff
,
S.
and
Tang
,
C.M.
(
2020
)
DNA ADP-ribosylation stalls replication and is reversed by recF-mediated homologous recombination and nucleotide excision repair
.
Cell Rep.
30
,
1373
1384
77
Schuller
,
M.
,
Butler
,
R.E.
,
Ariza
,
A.
,
Tromans-Coia
,
C.
,
Jankevicius
,
G.
,
Claridge
,
T.D.W.
et al (
2021
)
Molecular basis for darT ADP-ribosylation of a DNA base
.
Nature
596
,
597
602
78
Tromans-Coia
,
C.
,
Sanchi
,
A.
,
Moeller
,
G.K.
,
Timinszky
,
G.
,
Lopes
,
M.
and
Ahel
,
I.
(
2021
)
TARG1 protects against toxic DNA ADP-ribosylation
.
Nucleic Acids Res.
49
,
10477
10492
79
Dejesus
,
M.A.
,
Gerrick
,
E.R.
,
Xu
,
W.
,
Park
,
S.W.
,
Long
,
J.E.
,
Boutte
,
C.C.
et al (
2017
)
Comprehensive essentiality analysis of the Mycobacterium tuberculosis genome via saturating transposon mutagenesis
.
MBio
8
,
e01233-16
80
Gerdes
,
K.
and
Maisonneuve
,
E.
(
2012
)
Bacterial persistence and toxin-antitoxin loci
.
Annu. Rev. Microbiol.
66
,
103
123
81
Zaveri
,
A.
,
Wang
,
R.
,
Botella
,
L.
,
Sharma
,
R.
,
Zhu
,
L.
,
Wallach
,
J.B.
et al (
2020
)
Depletion of the DarG antitoxin in Mycobacterium tuberculosis triggers the DNA-damage response and leads to cell death
.
Mol. Microbiol.
114
,
641
652
82
LeRoux
,
M.
,
Srikant
,
S.
,
Littlehale
,
M.L.
,
Teodoro
,
G.
,
Doron
,
S.
,
Badiee
,
M.
et al)
The DarTG toxin-antitoxin system provides phage defense by ADP-ribosylating viral DNA
.
bioRxiv
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
84
Munnur
,
D.
and
Ahel
,
I.
(
2017
)
Reversible mono-ADP-ribosylation of DNA breaks
.
FEBS J.
284
,
4002
4016
85
Matta
,
E.
,
Kiribayeva
,
A.
,
Khassenov
,
B.
,
Matkarimov
,
B.T.
and
Ishchenko
,
A.A.
(
2020
)
Insight into DNA substrate specificity of PARP1-catalysed DNA poly(ADP-ribosyl)ation
.
Sci. Rep.
10
,
1
11
86
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
87
Langelier
,
M.F.
,
Riccio
,
A.A.
and
Pascal
,
J.M.
(
2014
)
PARP-2 and PARP-3 are selectively activated by 5’ phosphorylated DNA breaks through an allosteric regulatory mechanism shared with PARP-1
.
Nucleic Acids Res.
42
,
7762
7775
88
Munnur
,
D.
,
Bartlett
,
E.
,
Mikolčević
,
P.
,
Kirby
,
I.T.
,
Rack
,
J.G.M.
,
Mikoč
,
A.
et al (
2019
)
Reversible ADP-ribosylation of RNA
.
Nucleic Acids Res.
47
,
5658
5669
89
Munir
,
A.
,
Banerjee
,
A.
and
Shuman
,
S.
(
2018
)
NAD+-dependent synthesis of a 5-phospho-ADP-ribosylated RNA/DNA cap by RNA 2-phosphotransferase Tpt1
.
Nucleic Acids Res.
46
,
9617
9624
90
Belousova
,
E.A.
,
Ishchenko
,
A.A.
and
Lavrik
,
O.I.
(
2018
)
DNA is a New target of Parp3
.
Sci. Rep.
8
,
4176
91
Ahel
,
I.
,
Rass
,
U.
,
El-Khamisy
,
S.F.
,
Katyal
,
S.
,
Clements
,
P.M.
,
McKinnon
,
P.J.
et al (
2006
)
The neurodegenerative disease protein aprataxin resolves abortive DNA ligation intermediates
.
Nature
443
,
713
716
92
Verheugd
,
P.
,
Forst
,
A.H.
,
Milke
,
L.
,
Herzog
,
N.
,
Feijs
,
K.L.H.
,
Kremmer
,
E.
et al (
2013
)
Regulation of NF-κB signalling by the mono-ADP-ribosyltransferase ARTD10
.
Nat. Commun.
4
,
1683
93
Atasheva
,
S.
,
Frolova
,
E.I.
and
Frolov
,
I.
(
2014
)
Interferon-stimulated poly(ADP-Ribose) polymerases are potent inhibitors of cellular translation and virus replication
.
J. Virol.
88
,
2116
2130
94
Atasheva
,
S.
,
Akhrymuk
,
M.
,
Frolova
,
E.I.
and
Frolov
,
I.
(
2012
)
New PARP gene with an anti-alphavirus function
.
J. Virol.
86
,
8147
8160
95
Groslambert
,
J.
,
Prokhorova
,
E.
and
Ahel
,
I.
(
2021
)
ADP-ribosylation of DNA and RNA
.
DNA Repair
105
,
103144
96
Jackson
,
M.D.
and
Denu
,
J.M.
(
2002
)
Structural identification of 2′- and 3′-O-acetyl-ADP-ribose as novel metabolites derived from the Sir2 family β-NAD+ -dependent histone/protein deacetylases
.
J. Biol. Chem.
277
,
18535
18544
97
Liou
,
G.G.
,
Tanny
,
J.C.
,
Kruger
,
R.G.
,
Walz
,
T.
and
Moazed
,
D.
(
2005
)
Assembly of the SIR complex and its regulation by O-acetyl-ADP-ribose, a product of NAD-dependent histone deacetylation
.
Cell
121
,
515
527
98
Landry
,
J.
,
Sutton
,
A.
,
Tafrov
,
S.T.
,
Heller
,
R.C.
,
Stebbins
,
J.
,
Pillus
,
L.
et al (
2000
)
The silencing protein SIR2 and its homologs are NAD-dependent protein deacetylases
.
Proc. Natl Acad. Sci. U.S.A.
97
,
5807
5811
99
Lombardi
,
P.M.
,
Cole
,
K.E.
,
Dowling
,
D.P.
and
Christianson
,
D.W.
(
2011
)
Structure, mechanism, and inhibition of histone deacetylases and related metalloenzymes
.
Curr. Opin. Struct. Biol.
21
,
735
743
100
Smith
,
B.C.
,
Hallows
,
W.C.
and
Denu
,
J.M.
(
2008
)
Mechanisms and molecular probes of sirtuins
.
Chem. Biol.
15
,
1002
1013
101
Sauve
,
A.A.
,
Celic
,
I.
,
Avalos
,
J.
,
Deng
,
H.
,
Boeke
,
J.D.
and
Schramm
,
V.L.
(
2001
)
Chemistry of gene silencing: the mechanism of NAD+-dependent deacetylation reactions
.
Biochemistry
40
,
15456
15463
102
Hoff
,
K.G.
and
Wolberger
,
C.
(
2005
)
Getting a grip on O-acetyl-ADP-ribose
.
Nat. Struct. Mol. Biol.
12
,
560
561
103
Tong
,
L.
and
Denu
,
J.M.
(
2010
)
Function and metabolism of sirtuin metabolite O-acetyl-ADP-ribose
.
Biochim. Biophys. Acta Proteins Proteomics
1804
,
1617
1625
104
Kustatscher
,
G.
,
Hothorn
,
M.
,
Pugieux
,
C.
,
Scheffzek
,
K.
and
Ladurner
,
A.G.
(
2005
)
Splicing regulates NAD metabolite binding to histone macroH2A
.
Nat. Struct. Mol. Biol.
12
,
624
625
105
Chen
,
D.
,
Vollmar
,
M.
,
Rossi
,
M.N.
,
Phillips
,
C.
,
Kraehenbuehl
,
R.
,
Slade
,
D.
et al (
2011
)
Identification of macrodomain proteins as novel O-acetyl-ADP-ribose deacetylases
.
J. Biol. Chem.
286
,
13261
13271
106
Peterson
,
F.C.
,
Chen
,
D.
,
Lytle
,
B.L.
,
Rossi
,
M.N.
,
Ahel
,
I.
,
Denu
,
J.M.
et al (
2011
)
Orphan macrodomain protein (Human C6orf130) is an O-Acyl-ADP-ribose deacylase
.
J. Biol. Chem.
286
,
35955
35965
107
Ono
,
T.
,
Kasamatsu
,
A.
,
Oka
,
S.
and
Moss
,
J.
(
2006
)
The 39-kDa poly(ADP-ribose) glycohydrolase ARH3 hydrolyzes O-acetyl-ADP-ribose, a product of the Sir2 family of acetyl-histone deacetylases
.
Proc. Natl Acad. Sci. U.S.A.
103
,
16687
16691
108
Sawaya
,
R.
,
Schwer
,
B.
and
Shuman
,
S.
(
2005
)
Structure: function analysis of the yeast NAD+-dependent
.
RNA
11
,
107
113
109
Shull
,
N.P.
,
Spinelli
,
S.L.
and
Phizicky
,
E.M.
(
2005
)
A highly specific phosphatase that acts on ADP-ribose 1″-phosphate, a metabolite of tRNA splicing in saccharomyces cerevisiae
.
Nucleic Acids Res.
33
,
650
660
110
Neuvonen
,
M.
and
Ahola
,
T.
(
2009
)
Differential activities of cellular and viral macro domain proteins in binding of ADP-ribose metabolites
.
J. Mol. Biol.
385
,
212
225
111
Freire
,
D.M.
,
Gutierrez
,
C.
,
Garza-Garcia
,
A.
,
Grabowska
,
A.D.
,
Sala
,
A.J.
,
Ariyachaokun
,
K.
et al (
2019
)
An NAD+ phosphorylase toxin triggers Mycobacterium tuberculosis cell death
.
Mol. Cell
73
,
1282
1291.e8
112
Baysarowich
,
J.
,
Koteva
,
K.
,
Hughes
,
D.W.
,
Ejim
,
L.
,
Griffiths
,
E.
,
Zhang
,
K.
et al (
2008
)
Rifamycin antibiotic resistance by ADP-ribosylation: structure and diversity of Arr
.
Proc. Natl Acad. Sci. U.S.A.
105
,
4886
4891
113
Dabbs
,
E.R.
,
Yazawa
,
K.
,
Mikami
,
Y.
,
Miyaji
,
M.
,
Morisaki
,
N.
,
Iwasaki
,
S.
et al (
1995
)
Ribosylation by mycobacterial strains as a new mechanism of rifampin inactivation
.
Antimicrob. Agents Chemother.
39
,
1007
1009
114
Floss
,
H.G.
and
Yu
,
T.W.
(
2005
)
Rifamycin - mode of action, resistance, and biosynthesis
.
Chem. Rev.
105
,
621
632
115
Li
,
Z.
,
Liu
,
W.
,
Fu
,
J.
,
Cheng
,
S.
,
Xu
,
Y.
,
Wang
,
Z.
et al (
2021
)
Shigella evades pyroptosis by arginine ADP-riboxanation of caspase-11
.
Nature
599
,
290
295
116
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
117
Chatrin
,
C.
,
Gabrielsen
,
M.
,
Buetow
,
L.
,
Nakasone
,
M.A.
,
Ahmed
,
S.F.
,
Sumpton
,
D.
et al (
2020
)
Structural insights into ADP-ribosylation of ubiquitin by deltex family E3 ubiquitin ligases
.
Sci. Adv.
6
,
eabc0418
118
Akturk
,
A.
,
Wasilko
,
D.J.
,
Wu
,
X.
,
Liu
,
Y.
,
Zhang
,
Y.
,
Qiu
,
J.
et al (
2018
)
Mechanism of phosphoribosyl-ubiquitination mediated by a single Legionella effector
.
Nature
557
,
729
733
119
Höfer
,
K.
,
Schauerte
,
M.
,
Grawenhoff
,
J.
,
Wulf
,
A.
,
Welp
,
L.M.
,
Billau
,
F.A.
et al (
2021
)
Viral ADP-ribosyltransferases attach RNA chains to host proteins
.
bioRxiv
120
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
121
Dong
,
Y.
,
Mu
,
Y.
,
Xie
,
Y.
,
Zhang
,
Y.
,
Han
,
Y.
,
Zhou
,
Y.
et al (
2018
)
Structural basis of ubiquitin modification by the Legionella effector SdeA
.
Nature
557
,
674
678
122
Kalayil
,
S.
,
Bhogaraju
,
S.
,
Bonn
,
F.
,
Shin
,
D.
,
Liu
,
Y.
,
Gan
,
N.
et al (
2018
)
Insights into catalysis and function of phosphoribosyl-linked serine ubiquitination
.
Nature
557
,
734
738
123
Ray
,
K.
,
Marteyn
,
B.
,
Sansonetti
,
P.J.
and
Tang
,
C.M.
(
2009
)
Life on the inside: the intracellular lifestyle of cytosolic bacteria
.
Nat. Rev. Microbiol.
7
,
333
340
124
Appel
,
C.D.
,
Feld
,
G.K.
,
Wallace
,
B.D.
and
Williams
,
R.S.
(
2016
)
Structure of the sirtuin-linked macrodomain SAV0325 from Staphylococcus aureus
.
Protein Sci.
25
,
1682
1691
125
Komander
,
D.
and
Rape
,
M.
(
2012
)
The ubiquitin code
.
Annu. Rev. Biochem.
81
,
203
229
126
Dye
,
B.T.
and
Schulman
,
B.A.
(
2007
)
Structural mechanisms underlying posttranslational modification by ubiquitin-like proteins
.
Annu. Rev. Biophys. Biomol. Struct.
36
,
131
150
127
DaRosa
,
P.A.
,
Wang
,
Z.
,
Jiang
,
X.
,
Pruneda
,
J.N.
,
Cong
,
F.
,
Klevit
,
R.E.
et al (
2014
)
Allosteric activation of the RNF146 ubiquitin ligase by a poly(ADP-ribosyl)ation signal
.
Nature
517
,
223
226
128
Ahmed
,
S.F.
,
Buetow
,
L.
,
Gabrielsen
,
M.
,
Lilla
,
S.
,
Chatrin
,
C.
,
Sibbet
,
G.J.
et al (
2020
)
DELTEX2 C-terminal domain recognizes and recruits ADP-ribosylated proteins for ubiquitination
.
Sci. Adv.
6
,
1
15
129
Qiu
,
J.
,
Sheedlo
,
M.J.
,
Yu
,
K.
,
Tan
,
Y.
,
Nakayasu
,
E.S.
,
Das
,
C.
et al (
2016
)
Ubiquitination independent of E1 and E2 enzymes by bacterial effectors
.
Nature
533
,
120
124
130
Bardill
,
J.P.
,
Miller
,
J.L.
and
Vogel
,
J.P.
(
2005
)
IcmS-dependent translocation of SdeA into macrophages by the Legionella pneumophila type IV secretion system
.
Mol. Microbiol.
56
,
90
103
131
Wang
,
Y.
,
Shi
,
M.
,
Feng
,
H.
,
Zhu
,
Y.
,
Liu
,
S.
,
Gao
,
A.
et al (
2018
)
Structural insights into non-canonical ubiquitination catalyzed by SidE
.
Cell
173
,
1231
1243.e16
132
Kotewicz
,
K.M.
,
Ramabhadran
,
V.
,
Sjoblom
,
N.
,
Vogel
,
J.P.
,
Haenssler
,
E.
,
Zhang
,
M.
et al (
2017
)
A single legionella effector catalyzes a multistep ubiquitination pathway to rearrange tubular endoplasmic reticulum for replication
.
Cell Host Microbe
21
,
169
181
133
Shin
,
D.
,
Mukherjee
,
R.
,
Liu
,
Y.
,
Gonzalez
,
A.
,
Bonn
,
F.
,
Liu
,
Y.
et al (
2020
)
Regulation of phosphoribosyl-linked serine ubiquitination by deubiquitinases DupA and DupB
.
Mol. Cell
77
,
164
179.e6
134
Cahová
,
H.
,
Winz
,
M.L.
,
Höfer
,
K.
,
Nübel
,
G.
and
Jäschke
,
A.
(
2015
)
NAD captureseq indicates NAD as a bacterial cap for a subset of regulatory RNAs
.
Nature
519
,
374
377
135
Depping
,
R.
,
Lohaus
,
C.
,
Meyer
,
H.E.
and
Rüger
,
W.
(
2005
)
The mono-ADP-ribosyltransferases Alt and modB of bacteriophage T4: target proteins identified
.
Biochem. Biophys. Res. Commun.
335
,
1217
1223
136
Uzan
,
M.
and
Miller
,
E.S.
(
2010
)
Post-transcriptional control by bacteriophage T4: MRNA decay and inhibition of translation initiation
.
Virol. J.
7
,
1
22
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).