In this review, Azarm et al. provide an overview of PARPs in DNA repair, highlighting recent studies on their role during DNA replication and in phase separation. The authors also discuss the role of tankyrases in telomere maintenance and chromosome resolution.
Keywords: PARP, tankyrase, genome integrity
Abstract
Effective maintenance and stability of our genomes is essential for normal cell division, tissue homeostasis, and cellular and organismal fitness. The processes of chromosome replication and segregation require continual surveillance to insure fidelity. Accurate and efficient repair of DNA damage preserves genome integrity, which if lost can lead to multiple diseases, including cancer. Poly(ADP-ribose) a dynamic and reversible posttranslational modification and the enzymes that catalyze it (PARP1, PARP2, tankyrase 1, and tankyrase 2) function to maintain genome stability through diverse mechanisms. Here we review the role of these enzymes and the modification in genome repair, replication, and resolution in human cells.
Background: PARP1 and PARP2 and tankyrase 1 and tankyrase 2
Enzymes that catalyze PARylation
ADP-ribosylation (ADPr) is a posttranslational modification that regulates multiple cellular processes (Gupte et al. 2017). ADPr is catalyzed by the ADP-ribosyltransferase (ART) superfamily of enzymes (Palazzo et al. 2017). The best-studied family of this group is the poly(ADP-ribose) polymerases (PARPs). The PARP family is comprised of 17 members that share an ART diphtheria toxin-like (ARTD) domain and, hence, is also referred to as the ARTD family (Amé et al. 2004; Hottiger et al. 2010). PARPs transfer a single or multiple ADP-ribose unit(s) from nicotinamide adenine dinucleotide (NAD+) to protein acceptors. PARP family members that perform poly(ADP-ribosyl)ation (PARylation) include PARP1 and PARP2, which synthesize long branched chains of up to 200 units, and tankyrase 1 and 2, which synthesize shorter unbranched chains of up to 20 units (de Murcia and Menissier de Murcia 1994; D'Amours et al. 1999; Rippmann et al. 2002; Vyas et al. 2014). The remaining members of the PARP family are not as well studied, but most catalyze mono(ADP-ribosyl)ation (MARylation) and have a range of functions throughout the cell (Lüscher et al. 2018). Here we focus on the four family members that catalyze PARylation (PARP1 and PARP2 and tankyrase 1 and tankyrase 2) and how they function in the nucleus to promote genome stability. We first compare and contrast PARP versus tankyrase and discuss the role of the poly(ADP-ribose) (PAR) modification. We provide an overview of PARPs in DNA repair, highlighting recent studies on their role during DNA replication and in phase separation. Finally, we focus on the role of tankyrases in telomere maintenance and chromosome resolution.
Compare and contrast: PARP1 and PARP2 versus tankyrase 1 and tankyrase 2
Structure, mode of action, and function
Structure
PARP1, the founding member of the PARP family, is the most well studied and known for its role in DNA damage repair (DDR) (Martin-Hernandez et al. 2017; Ray Chaudhuri and Nussenzweig 2017). The primary structure of PARP1 comprises an N-terminal three-zinc-finger domain (Zn1, Zn2, and Zn3), an internal BRCA C terminus (BRCT) domain followed by a Trp–Gly–Arg (WGR) domain, and a C-terminal catalytic domain comprised of a helical domain (HD) in close association with the ART domain (Fig. 1A; Langelier et al. 2012, 2018a). One of the earliest events in DNA repair is recruitment of PARP1 to DNA. PARP1 binds to a range of DNA structures like cruciforms (Brázda et al. 2011) and DNA damage structures, including nicked and gapped single-strand breaks (SSBs) and overhang and blunt-ended double-strand breaks (DSBs) (Beck et al. 2014b). In the absence of genotoxic stress PARP1 is catalytically inactive. The HD regulates catalytic activity by blocking NAD+ binding. Upon detection of DNA damage, the Zn fingers and WGR domain organize around the DNA break, inducing an allosteric destabilization of the HD that enables NAD+ to access the catalytic active site (Dawicki-McKenna et al. 2015; Langelier et al. 2018b). PARP2 and the closely related PARP3 [a mono(ADP-ribosyl) transferase] lack the Zn fingers and BRCT domain of PARP1, but share a common WGR domain that interacts with DNA (albeit with different substrate specificities) and regulates their catalytic activity in response to DNA damage (Langelier et al. 2014; Grundy et al. 2016; Obaji et al. 2016, 2018; Riccio et al. 2016a).
Figure 1.

Primary structure and mode of action for PARP1 and PARP2 and tankyrase 1 and tankyrase 2. (A) Schematic representation of PARP1 and PARP2. (Zinc fingers) Zn1, Zn2, and Zn3; (BRCT) BRCA C terminus; (WGR) Trp–Gly–Arg domain; (HD) helical domain; (ARDT) ADP-ribosyltransferase. (B) PARP1 or PARP2 is regulated through its catalytic activity. Inactive PARP1 binds to breaks and is activated to undergo autoPARylation. PAR serves to recruit DNA repair proteins. (C) Schematic representation of tankyrase 1 and tankyrase 2. (HPS) His–Pro–Ser domain; (SAM) sterile α module; (ART) ADP-ribosyltransferase. (D) Tankyrase 1 or tankyrase 2 is regulated through its selection of binding partners. Constitutively active tankyrase binds to RxxxxG-containing proteins and localizes throughout the cell for multiple functions. Tankyrase may (or may not) PARylate its partner.
Binding to breaks stimulates rapid and extensive synthesis of PAR onto PARP1 itself (autoPARylation) (Satoh and Lindahl 1992). The PAR then may prevent immediate inappropriate processing of the damage by blocking access to nucleases (Caron et al. 2019) while simultaneously acting as a recruitment platform for nucleases (Wang et al. 2019) and other proteins that bind PAR to repair the damage (Fig. 1B). Excessive PARylation can induce release of PARP1 from the DNA. The turnover of the PAR chains is also regulated rapidly through the glycohydrolase PARG (Lin et al. 1997; Slade et al. 2011; O'Sullivan et al. 2019).
Tankyrase 1 and Tankyrase 2 are closely related proteins with diverse cellular functions ranging from chromosome resolution to Wnt/β-catenin signaling (Hsiao and Smith 2008; Haikarainen et al. 2014). They have a similar primary structure comprised of a C-terminal catalytic PARP domain, a SAM (sterile α module) domain, an ankyrin repeat domain, and an N-terminal HPS (His, Pro, and Ser) domain of unknown function, unique to tankyrase 1 (Fig. 1C; Smith et al. 1998). The SAM domain of tankyrases can promote homo- and hetero-oligomerization of tankyrase 1 and tankyrase 2; polymerization is required for full catalytic activity (De Rycker and Price 2004; Mariotti et al. 2016; Riccio et al. 2016b; Fan et al. 2018). A distinguishing feature of tankyrases is their ability to interact (through the ankyrin repeat domain) with a broad range of binding partners (Fig. 1D). The ankyrin domain is organized into five ankyrin repeat clusters (ARCs), which serve as a basic unit for recognizing an eight amino acid segment (with a strict requirement for Arg at position 1 and Gly at position 6 [RxxxxG]) in its binding partners (Sbodio and Chi 2002; Seimiya and Smith 2002; De Rycker et al. 2003; Guettler et al. 2011; Eisemann et al. 2016). Four of the five ARCs (1, 2, 4, and 5) are each capable of recognizing a peptide motif. Proteomic and in silico screens have identified hundreds of potential tankyrase interacting proteins (Guettler et al. 2011; Bhardwaj et al. 2017; Li et al. 2017). Over 40 human tankyrase-binding partners have been validated by coimmunoprecipitation; they localize throughout the cell and most contain a consensus peptide that (where it has been tested) binds to both tankyrase 1 and tankyrase 2 (Fig. 2). This review will focus on those partners that reside in the nucleus and have been shown to influence genome stability.
Figure 2.

List of validated tankyrase-binding partners and their localizations throughout the cell. (A) A chronological list of human tankyrase-binding partners that have been validated by protein coimmunoprecipitation. For each protein, the list includes its subcellular location(s), its “RxxxxG” consensus tankyrase-binding site (“——” if not found in the amino acid sequence), and a reference for the publication identifying the interaction. Protein location(s) were curated through a literature search and the UniProt protein database. (N.D.) Not determined. (a) The protein contains additional binding motifs in its sequence. (b) The binding site is noncanonical. (c) The binding site is conserved among protein family members. (B) Schematic representation of the localization of tankyrase binding proteins (listed in A) to the indicated organelle(s). Circle size is based on the number of tankyrase-binding proteins localized to the organelle(s). The overlap between an organelle(s) circle with the “nucleus” circle represents the number proteins found in both.
Mode of action
Despite the fact that tankyrases and PARP1 and PARP2 share the ability to catalyze synthesis of ADP-ribose polymers, tankyrases differ from PARP1 and PARP2 in multiple ways. Tankyrases do not bind directly to DNA and they are not induced by DNA breaks to promote rapid synthesis of long PAR chains (Cook et al. 2002). The tankyrase 1 and tankyrase 2 catalytic ART domains lack the regulatory HD found in PARPs 1 and 2, rendering them accessible to NAD+ binding. Analysis of the PARylation status in cells shows tankyrase 1 to be constitutively autoPARylated with relatively short PAR chains (Smith and de Lange 2000; Cook et al. 2002; Chang et al. 2005). So far, there does not appear to be a mechanism for dramatic induction of tankyrase PARP activity along the lines of the massive stimulation of PARP1 by DNA damage. Studies suggest that SAM domain-mediated polymerization can regulate tankyrase catalytic activity (Mariotti et al. 2016; Riccio et al. 2016b; Fan et al. 2018), although how this would be regulated in vivo to stimulate or induce tankyrase-mediated PARylation remains to be determined. Another distinguishing feature between PARP1 and PARP2 and tankyrases is that while PARP1 and PARP2 are predominantly nuclear, tankyrases (in addition to their nuclear localization) are distributed throughout the cell at many sites, consistent with their broad array of binding partners (Fig. 2). Thus, while the regulatory mechanism for PARP1 and PARP2 occurs through repression and activation of their catalytic ART domains, the regulatory mechanism for tankyrases occurs through its ankyrin repeat selection of RxxxxG-containing binding partners. Tankyrase-binding partners can recruit tankyrases to different subcellular sites (Hsiao and Smith 2008). Once tankyrase binds a partner, it may or may not PARylate it (Smith et al. 1998; Bisht et al. 2012; Eisemann et al. 2019). Subsequently, tankyrase-mediated PARylation can influence the stability, localization, or function of the target.
Consequences of deletion
The effect of knockout of these PARPs in a whole organism, such as mice, indicates a redundant and essential role for each class. Mice lacking PARP1 or PARP2 are viable, but exhibit sensitivity to genotoxic agents, consistent with a role for each protein in DNA repair (de Murcia et al. 1997; Wang et al. 1997; Masutani et al. 2000; Forsyth et al. 2002). Double deletion of PARP1 and PARP2 leads to early embryonic lethality at Day 8.0 in mice, indicating redundancy between the two proteins and suggesting a role in early development (Menissier de Murcia et al. 2003). The requirements for PARP1 and PARP2 at the stage during gastrulation when cellular proliferation increases dramatically, suggests a fundamental role for PARP in DNA replication (see below).
Mice lacking tankyrase 2 are viable and fertile, but have a short stature phenotype (Chiang et al. 2006; Hsiao et al. 2006). Mice deficient in tankyrase 1 appear to develop normally and have no defects in body size (Chiang et al. 2008), but suffer from a metabolic disorder (Yeh et al. 2009; Zhong et al. 2016). The double knockout is embryonic lethal, indicating functional redundancy and suggesting a role in development (Chiang et al. 2008). The requirement for tankyrases in development may be due to their interactions with signaling proteins such as Axin and Notch that play essential roles in development (Huang et al. 2009; Bhardwaj et al. 2017).
In contrast to the effect of deletion on mouse development, double deletion of PARP1 and PARP2 or tankyrase 1 and tankyrase 2 is not lethal in human cells grown in culture. Double-knockout PARP1/PARP2 cell lines were generated in normal human (hTERT RPE-1) and cancer (U2OS) cells and double-knockout TNKS1/TNKS2 cell lines were generated in cancer (HEK293T) cells (Bhardwaj et al. 2017; Hanzlikova et al. 2017; Ronson et al. 2018). The observation that these cell lines are viable indicate that (at least in the absence of exogenous genotoxic stress or other deficiencies) some human cell lines can survive without PARP1 and PARP2 or without tankyrase 1 and tankyrase 2.
PAR: a docking site for interacting proteins
The amino acid target sites of PARylation
Significant progress has been made on the proteome-wide determination of PARylated amino acids in human cells. Proteomic analyses have identified a diverse group of amino acids including Asp, Glu, Lys, Arg, and Cys (Daniels et al. 2015; Martello et al. 2016). PARP1 was previously found to be automodified on Asp, Glu, and Lys. However, recent analysis revealed Ser as the major target (Leidecker et al. 2016; Bilan et al. 2017) and further that the histone PARylation factor (HPF1) associates with PARP1 and PARP2 to target specifically Ser-ADPr (Gibbs-Seymour et al. 2016; Bonfiglio et al. 2017). Accurate analysis of ADPr site localization is inherently challenging. Recent advances in proteomic techniques will likely lead to revisions of previous analyses and open up future discovery. Indeed, very recent work has expanded the ADP-ribosylome, identifying Tyr as an ADPr acceptor site (Leslie Pedrioli et al. 2018) and revealing crosstalk between Ser-ADPr and Ser phosphorylation (Larsen et al. 2018). Much less is known about tankyrase target sites. However, recent analysis for tankyrase 1 in vitro shows Glu and Asp as primary targets and also that HPF1 does not alter tankyrase 1 substrate specificity to target Ser for modification (Eisemann et al. 2019).
PAR-binding modules
The PAR modification can serve as a docking site for PAR-binding proteins (Teloni and Altmeyer 2016; Wei and Yu 2016). Proteins can bind to PAR through a range of interacting modules including: PAR-binding motifs (PBMs), short sequences (∼20 amino acids) with a loosely defined consensus (Pleschke et al. 2000; Gagné et al. 2008); macrodomains, larger (∼130- to 190-amino-acid) globular domains that bind to ADP-ribose monomer or the terminal ADP-ribose moiety in a PAR chain (Karras et al. 2005; Feijs et al. 2013); PAR-binding zinc fingers (PBZ), short modules (∼30 amino acids) that bind to ADP-ribose monomers or to the ADP-ribose–ADP-ribose junctions of PAR chains (Ahel et al. 2008; Eustermann et al. 2010; Li et al. 2010; Oberoi et al. 2010); and WWE domains, comprised of conserved tryptophan (W) and glutamate (E) residues that bind to the iso-ADP-ribose moiety, the internal unit of the PAR polymer (Wang et al. 2012). Additional domains that can recognize PAR include: the forkhead-associated (FHA) domains of aprataxin (APTX) and polynucleotidekinase-3 phosphatase (PNKP), which interact with iso-ADP-ribose (like WWE domains) and the BRCA1 C-terminal (BRCT) domains of DNA ligase IV, XRCC1, and NBS1, which recognize the ADP-ribosyl moiety within PAR (Li and Yu 2013; Li et al. 2013; Breslin et al. 2015).
In addition to those described above, a number of protein motifs that bind to RNA and DNA have been found to also recognize PAR, thereby broadening the potential influence of PAR in nuclear functions. These motifs include the RNA recognition motif (RRM) found in hnRNP proteins (Gagné et al. 2008; Ji and Tulin 2009); serine/arginine repeats (SR) found in splicing factors (Malanga et al. 2008); the OB-fold found in the single-stranded binding proteins SSB1 and BRCA2 (Zhang et al. 2014, 2015a); the N terminus of the PilT protein (PIN) domain found in EXO1, GEN1, and SMG5 (Zhang et al. 2015b); and regions rich in arginines and glycines (RG/RGG), also termed glycine arginine-rich (GAR) domains, found in proteins involved in DNA damage signaling, transcription, and RNA processing (Thandapani et al. 2013).
PAR-binding modules in chromatin remodeling and DNA repair
Proteins containing PAR-binding modules can be recruited in a PAR-dependent manner by PARP1 to sites of DNA breaks to modify chromatin structure and facilitate DNA repair. For example, the chromatin remodeler amplified in liver cancer 1(ALC1), also known as chromodomain–helicase–DNA-binding protein 1-like (CHD1L), contains a PAR-binding macrodomain that is activated in a PAR-dependent manner to enable nucleosome remodeling and DNA repair (Ahel et al. 2009; Gottschalk et al. 2009, 2012). Two DNA damage repair proteins, checkpoint with forkhead and ring finger domains (CHFR) and aprataxin polynucleotide kinase-like factor (APLF) contain, respectively, single and tandem PBZ motifs (Ahel et al. 2008). APFL is recruited to sites of DNA damage through interactions with ADP-ribose through its PAR-binding PBZ motif to promote DNA repair by nonhomologous end joining (NHEJ) (Rulten et al. 2008, 2011). CHFR is required for the antephase checkpoint, dependent on its PBZ domain and PAR synthesis (Ahel et al. 2008). The chromatin remodeler CHD2 is recruited to DSBs by PAR binding through an uncharacterized motif, where it promotes chromatin decondensation to facilitate DNA repair by NHEJ (Luijsterburg et al. 2016). X-ray repair cross-complementing protein 1 (XRCC1) is recruited in a PARP1 dependent manner through a PBM, to act as a scaffold in assembly and activation of the DNA base excision repair (BER) machinery (Masson et al. 1998; Okano et al. 2003).
PAR-binding modules in protein degradation
Most examples of PAR-binding module recruitment that have been described thus far relate to PARP1. This may be due to the greater abundance of PARP1-induced PAR chains in cells compared with tankyrase-induced PAR chains. However, one PAR interactor, the E3 ligase RN146, associates specifically with tankyrase 1 and tankyrase 2. RNF146 interacts with PARylated substrates through its internal WWE domain (Callow et al. 2011; Kang et al. 2011; Zhang et al. 2011; Zhou et al. 2011; Wang et al. 2012). The PAR ligand allosterically activates the E3 ligase through a conformational change in the N-terminal RING domain to promote K48-ubiquitination of its targets (DaRosa et al. 2015). RNF146 can also bind to the ARC domains in tankyrase 1 or tankyrase 2 directly through motifs in its C terminus, thereby favoring PARylated tankyrases and their bound PARylated targets for ubiquitination and degradation (DaRosa et al. 2015, 2018). In this way, tankyrases regulate the cellular levels of several disease-related cytoplasmic proteins including Axin, 3BP2, PTEN, and the angiomotins, which are key regulators of the Wnt/β-catenin, SRC, AKT, and Hippo signaling pathways, respectively (Huang et al. 2009; Levaot et al. 2011; Li et al. 2015; Wang et al. 2015). Whether this mechanism is used to target degradation of tankyrase targets in the nucleus to influence genome integrity has not been determined.
Chromosome repair
PARP1 and PARP2 are central components in the single-strand break (SSB) repair pathway. They are also activated at double-strand breaks (DSBs) and at stalled replication forks, where they promote homologous recombination (HR) and NHEJ. In contrast, tankyrase 1 and tankyrase 2 do not have a canonical role in DNA repair, although several recent studies suggest an indirect role through their binding partners. Finally, recent studies show that PARP1-dependent PAR can facilitate DNA repair through phase separation and transient compartmentalization of DNA damage sites.
PARP1 and PARP2 in DNA repair
Single-strand break repair
SSBs can arise directly by oxidative damage to the sugar backbone, indirectly during DNA BER, or as a result of the abortive activity of topoisomerase 1 (Ray Chaudhuri and Nussenzweig 2017). Spontaneous single-strand breaks are rapidly detected and bound by PARP1. DNA-binding activates PARP1 to undergo autoPARylation (Satoh and Lindahl 1992). PAR recruits XRCC1, a core factor in SSBR that serves as a scaffold for SSB proteins (Caldecott et al. 1994). XRCC1 has an internal BRCT domain that binds directly to PARylated PARP1 or PARP2, thereby directing its recruitment to activated PARP1 or PARP2 (Caldecott 2019). XRCC1 then acts as a scaffold for protein-binding partners necessary to process damaged termini, including PNKP, APTX, and tyrosyl-DNA phosphodiesterase 1 (TDP1). XRCC1 recruits Pol β through its N-terminal domain to replace the single missing nucleotide at the SSBs and LIG3 through a C-terminal BRCT domain to ligate the nick (Ray Chaudhuri and Nussenzweig 2017; Caldecott 2019).
Double-strand break repair
DSBs are produced following exposure to DNA-damaging agents such as γ-irradiation. DSBs are repaired by either HR or NHEJ (Martin-Hernandez et al. 2017; Ray Chaudhuri and Nussenzweig 2017). PARP1 detects and binds DS breaks, activating its autoPARylation. PARP1 may promote recruitment of the MRE11 nuclease (through its putative PAR-binding domain), which could contribute to DNA end processing and channel the pathway choice to HR (Haince et al. 2008). PARP1 also contributes to early recruitment of BRCA1 (through its PAR-binding domain), which subsequently loads RAD51 onto DNA, an essential step for HR (Li and Yu 2013). In G1, when the sister chromatid is not available for HR, NHEJ is the preferred mechanism for DSBR. DSBs are bound by KU70–KU80 dimers, which recruit DNA-PK catalytic subunit (DNA-PKcs). PARP1 binds and PARylates DNA-PKcs, stimulating its kinase activity, and may additionally help to recruit the chromatin remodeler CHD2 to promote LIG4-dependent classical NHEJ (cNHEJ) (Ruscetti et al. 1998; Spagnolo et al. 2012; Luijsterburg et al. 2016). Alternative NHEJ (aNHEJ) facilitates ligation independently of KU and LIG4, instead relying on MRE11 (Truong et al. 2013). PARP1 can compete with KU for access to DNA ends. In the absence of KU, PARP1 may promote recruitment of MRE11 to process the ends and channel repair to aNHEJ (Haince et al. 2008; Cheng et al. 2011). In a highly mutagenic process, the resected ends are then joined through sequence microhomology, and the gaps are filled in by POLQ and ligated by LIG3 (Ceccaldi et al. 2015; Mateos-Gomez et al. 2015).
Stalled replication forks
An obstruction in the DNA template (such as an SSB in the leading strand template) can lead to a stalled fork, which can either undergo fork reversal where the nascent strands anneal in a “chicken foot” structure or fork collapse due to replication run-off of the leading strand (Cortez 2015). Either scenario results in a one-ended DSB that can be detected by PARP1 or PARP2 (Ray Chaudhuri and Nussenzweig 2017; Hanzlikova and Caldecott 2019). PARP-binding suppresses KU binding to prevent NHEJ, which can be toxic at one-ended DSBs (Hochegger et al. 2006; Sugimura et al. 2008). At the reversed fork PARP1 also inhibits the RECQ1 helicase to prevent premature fork reset and restart (Popuri et al. 2012; Berti et al. 2013). PARP activity can then promote the HR pathway. PARP1 or PARP2 physically recruits MRE11 (potentially through PAR binding) to promote DNA end-resection creating a 3′ single-stranded tail for BRCA2-assisted RAD51 loading (Haince et al. 2008; Li and Yu 2013). RAD51 facilitates DNA stand exchange/template switching to promote reannealing of the two strands of the reversed or collapsed fork with their complimentary sisters for DNA replication restart (Pasero and Vindigni 2017). In the case of reversed forks, this can also be achieved by RECQ1 once the lesion is repaired and PARP1 is released.
MARylating PARPs in DNA repair
Several PARP family members that catalyze mono(ADP-ribosylation) also function in DNA repair and genome integrity. PARP3, which is structurally related to PARP1 and PARP2, contributes to the cellular response to DSBs (Boehler et al. 2011; Rulten et al. 2011). PARP3 cooperates with KU to drive pathway choice to cNHEJ (Beck et al. 2014a) and facilitates association of APFL to damaged DNA, which promotes retention of XRCC4/DNA ligase IV to accelerate ligation during NHEJ (Rulten et al. 2011; Fenton et al. 2013). Two other MARylating PARPS, PARP10 and PARP14, which are structurally and functionally related to each other, play distinct roles in alleviating replication stress. PARP10 recognizes ubiquitinated PCNA, which is required to recruit translesion DNA polymerases for restart of stalled replication forks (Nicolae et al. 2014; Shahrour et al. 2016; Zafar and Eoff 2017). PARP10 and its interaction with PCNA is required for efficient translesion synthesis and resistance to replication fork stalling (Nicolae et al. 2014). PARP14 interacts with RAD51 to promote HR for relief of replication stress (Nicolae et al. 2015).
Tankyrase 1 and tankyrase 2 in DNA repair
Unlike PARP1 and PARP2, tankyrase 1 and tankyrase 2 do not bind directly to DNA and there is little evidence to suggest a direct role for tankyrases in DNA damage-induced repair. Depletion of tankyrase 1 does lead to an increase in DNA damage foci, but this may be indirect due to defects in telomere and rDNA resolution (see below) (Hsiao and Smith 2009; Daniloski et al. 2019). However, several recent studies suggest that tankyrases may play a role in DNA repair, indirectly through its binding partners: mediator of DNA damage checkpoint protein 1 (MDC1), MERIT40, or TRF1. MDC1 is essential for spreading of the DDR signaling on chromatin surrounding DSBs and plays a role in DNA repair (Jungmichel and Stucki 2010). Tankyrase 1 and tankyrase 2 bind to MDC1 through tankyrase binding sites 948-RGEPEG and 1993-RRLLEG (Nagy et al. 2016). Tankyrase 1 or tankyrase 2 (overexpressed in the nucleus) localizes to sites of DNA damage (laser stripes or I-SCEI-induced DS breaks)-dependent on MDC1 and promotes DSB repair by HR (Nagy et al. 2016).
MERIT40, a component of the BRCA1-A and BRISC complexes was shown to bind to tankyrase 1 or tankyrase 2 through tankyrase binding sites 28-RSNPEG and 48-RSEGEG (Guettler et al. 2011). MERIT40 recruits tankyrase 1 (overexpressed in the nucleus) to γH2AX foci in X-ray irradiated cells, and the tankyrase-MERIT40 interaction is necessary for viability, suggesting a role for tankyrase in regulation of the DDR (Okamoto et al. 2018). It should be noted that in the above two examples tankyrase is artificially localized to the nucleus through a nuclear localization signal (NLS). This may amplify tankyrase function in the nucleus, since normally only a small fraction of endogenous tankyrase is nuclear.
Lastly, tankyrase 1 has been implicated in DNA repair through its interaction with the telomere-binding protein TRF1, which binds to tankyrase 1 and tankyrase 2 through a 13-RGCADG tankyrase-binding site (Smith et al. 1998). Tankyrase 1 can be recruited to sites of induced telomere oxidative damage where it PARylates TRF1 and leads to recruitment of XRCC1. Tankyrase 1 inhibition or expression of a TRF1 tankyrase-binding site mutant sensitizes cells to induced telomere oxidative damage, suggesting a role for tankyrase 1 in facilitating SSBR at damaged telomeres through PARylation (Yang et al. 2017).
PAR-induced liquid demixing in DNA repair
Liquid–liquid demixing or phase separation can dynamically organize soluble intracellular components into confined compartments. Phase separation can be initiated by intrinsically disordered proteins (IDPs) that contain low-complexity domains (LCDs): unstructured repetitive sequences that can phase separate into liquid droplets (Chong and Forman-Kay 2016; Banani et al. 2017). Some LCD-containing RNA-binding proteins can self-assemble into structures, and in some cases assembly can be nucleated by RNA itself (Chong et al. 2018). One example is FUS, an RNA-binding protein containing an N-terminal LCD (comprised of a prion-like domain; PLD) and multiple RGG domains (Thandapani et al. 2013). Studies showed that RNA binding could nucleate formation of higher order FUS structures, dependent on the PLD and RGG domains (Schwartz et al. 2013). PAR polymers (like RNA) could serve as seeding platforms to recruit proteins that multimerize through their LCDs to form cellular compartments (Leung 2014). Indeed, FUS binds directly to PAR in vitro through its RGG domains and is recruited to sites of DNA damage, dependent on PARP1 and PARylation activity (Mastrocola et al. 2013; Rulten et al. 2014).
Analysis of FUS compartments in vivo reveals all the hallmarks of liquid droplets: They undergo rapid internal rearrangement, are spherical, and two droplets can fuse and relax into one sphere (Patel et al. 2015). Further analysis shows that FUS, as well as two other members of the TET family of proteins, EWS and TAF15, are recruited to laser micro-irradiated sites of DNA damage in a PAR-dependent manner with the RGG domains serving as PAR sensors (Altmeyer et al. 2015). The N-terminal PLDs assemble into spherical structures in cells by liquid demixing. Live-cell microscopy shows that PAR seeds the liquid demixing at sites of DNA damage and further that the activity is enhanced by depletion of PARG and inhibited by PARP inhibitors (PARPi). The purified TET proteins or a model protein comprised of a PLD and RGG form aggregates in vitro that are consistently larger in the presence of PAR, demonstrating the intrinsic ability of PAR to nucleate IDP-aggregation.
Recently, this model was elaborated in vitro at the single molecule level using atomic force microscopy to dissect the steps (Singatulina et al. 2019). FUS is recruited to PAR synthesized by PARP1 on damaged DNA, and FUS triggers formation of large compartments in which damaged DNA is enriched. Consistent with the in vivo studies described above, both the PLD and RGG motifs are required for formation of the compartments. These compartments can reversibly dissociate by hydrolysis of PAR by PARG. Together these data indicate that PAR-mediated recruitment of FUS can facilitate DNA repair through transient phase-separated compartmentalization of DNA damage sites. Whether tankyrase-dependent PAR can promote liquid demixing remains to be determined.
Chromosome replication
As described above, PARP1 and PARP2 are activated by SSBs, DSBs, and stalled replication forks. PARP1 and PARP2 bind to the damaged DNA structure, become catalytically active, and PARylate themselves (and other proteins) to promote DNA repair. However, what about during a normal S phase in proliferating cells in the absence of exogenous damage? Is there a role for PARP1 during DNA replication? Historically it has been difficult to identify sites of PAR in unperturbed cells, although several studies hint at a role for PARP1 and PAR in DNA replication. Recent studies suggest a role for PARP1 in normal proliferating cells during S phase in Okazaki fragment processing (Hanzlikova et al. 2018) and regulation of replication fork speed (Maya-Mendoza et al. 2018). To date a role has not been described for tankyrase 1 and tankyrase 2 in replication fork progression; however, they do play a role in replication of chromosome ends by telomerase.
PARP1 and Okazaki fragment processing
During DNA replication PARP1 can bind directly to nicks of the lagging strands. Early studies showed increased PARP activity upon initiation of DNA replication and in newly replicated chromatin (Lehmann et al. 1974; Anachkova et al. 1989). Depletion of PARG, the enzyme that cleaves PAR, using siRNA led to slowing of replication fork progression and to accumulation of defective replication intermediates (Ray Chaudhuri et al. 2015). PARG interacts directly with PCNA and colocalizes to PCNA-containing replication foci (Mosavi et al. 2004; Kaufmann et al. 2017). Recent advances have led to new, cell-active in vitro chemical probes, which offer potent and selective inhibition of PARG in the cellular context (James et al. 2016). Armed with this new class of PARG inhibitor, Caldecott and colleagues set out to detect endogenous PAR by treating cells briefly (15–60 min) with inhibitor (Hanzlikova et al. 2018). PAR was detected specifically in S phase of the cell cycle and at sites of DNA replication (based on proximity to PCNA), dependent on PARP1 activity. The S phase PAR was not due to DNA damage or replication stress, rather it resulted from unligated Okazaki fragments. Okazaki fragments are first processed by FEN1, and then ligated together by LIG1. Treatment of FEN1-inhibited or LIG1-depleted cells with PARG inhibitor led to a >10-fold increase in S-phase PCNA-associated PAR. Moreover, treatment with emetine, an inhibitor of DNA replication that prevents formation of Okazaki fragments, completely blocked the appearance of S phase PAR. The authors postulated that a subfraction of the 30 to 50 million Okazaki fragments that are synthesized during replication of the human genome might escape canonical processing by FEN1 and LIG1, and instead, through PAR synthesis recruit the SSBR machinery to complete ligation (Hanzlikova et al. 2018; Hanzlikova and Caldecott 2019). Indeed, it has been demonstrated that the SSBR scaffold protein XRCC1 is recruited to these sites of incomplete replication, dependent on PARP1 and PARP2 (Breslin et al. 2015). Together these data indicate the PARP-dependent SSBR machinery as a “backup” pathway for processing unligated Okazaki fragments.
PARP1 and replication fork speed
Another way in which PARP1 may impact chromosome replication is through regulation of replication fork speed. A recent study identified a new regulatory network involving PARylation and p21 as suppressors of DNA replication fork speed (Maya-Mendoza et al. 2018). Treatment of cells with PARP inhibitors for 24 h resulted in a 60% increase in fork speed measured by DNA fiber analysis (from 1.0 to 1.6 kb/min), in contrast to the accepted model that PARP inhibitors induce fork stalling (Bryant et al. 2005). The increased fork speed was accompanied by a DDR in cells. Depletion of LIG1 and FEN1 led to accelerated fork speed (1.2 and 1.3 kb/min, respectively), but it was less than that induced by PARPi and did not induce a DDR. Indeed, when PARPi was titrated down in concentration or duration to induce acceleration <40%, a DDR was not induced, suggesting that increases >40% lead to a DDR. Knockdown of PARP1 protein led to increased fork speed (1.2 kb/min), but it was less than the effect of PARPi. The PARP knockdown cells did not show reduced PAR levels at the fork, indicating that other PARPs may be involved in fork-speed control. PARPi treatment of PARP1 knockdown cells did not further accelerate fork speed or a DDR, indicating a requirement for PARP1 protein.
To gain insight to control of the pathway, a connection between replication fork speed and PARP1, PARylation, and p53–p21 was explored. p53 activates p21 and is itself PARylated, and PARP1 binds p21 and controls its expression (el-Deiry et al. 1993; Waga et al. 1994; Frouin et al. 2003; Madison and Lundblad 2010; Lee et al. 2012). p21 levels were found to increase in PARP1 knockdown cells, but not in PARPi-treated cells. Depletion of p21 did not affect PARylation levels, but did lead to increased fork speed (1.7 kb/min) that was additive with PARPi (2.3 kb/min). A fork speed regulatory network was proposed with p21 and PARP1/PARylation as interacting arms that are connected through PARylation-mediated regulation of p53 activity, as well as PARP1-mediated inhibition of p21 expression. Reduction of p21 or PARylation would lead to an imbalance, resulting in increased fork speed. Thus, at steady state during normal DNA replication, the network would limit fork speed to ensure genome stability (Maya-Mendoza et al. 2018; Quinet and Vindigni 2018).
Tankyrase 1 and tankyrase 2 and telomere length maintenance
Human telomeres are comprised of TTAGGG repeats and shelterin, a six-subunit complex that includes the double-stranded telomere DNA-binding proteins TRF1 and TRF2 (de Lange 2005). Telomeres rely on shelterin and shelterin-binding proteins to mediate the specialized mechanisms required for their replication (Gilson and Géli 2007; Stewart et al. 2012) and protection (Palm and de Lange 2008; de Lange 2018). In normal human cells after multiple rounds of cell division, shortened telomeres are unable to recruit sufficient shelterin to protect chromosome ends, resulting in a persistent DNA damage response that signals replicative senescence (d'Adda di Fagagna et al. 2003; Takai et al. 2003). Shortening can be counteracted by telomerase a reverse transcriptase (Greider and Blackburn 1985, 1987; Lingner et al. 1997) that is repressed in the human soma (Wright et al. 1996), but is up-regulated in most human cancers (Kim et al. 1994) and is required for tumor cell growth (Hahn et al. 1999). Telomere length is regulated by TRF1; overexpression leads to telomere shortening and depletion to telomere lengthening by telomerase (van Steensel and de Lange 1997).
Tankyrase 1 was initially identified as a binding partner for TRF1 (Smith et al. 1998). Tankyrase PARylates TRF1, and this modification releases TRF1 from DNA, much in the way PARyation of PARP1 releases it from DNA breaks. Tankyrase is recruited to telomeres by TRF1, dependent on its tankyrase-binding site (Hsiao and Smith 2008). Overexpression of tankyrase 1 or tankyrase 2 in the nucleus evicts TRF1 from telomeres, resulting in its degradation and allowing telomerase-mediated recruitment and telomere lengthening (Smith and de Lange 2000; Cook et al. 2002; Chang et al. 2003). In these experiments (as described above for tankyrase in DNA repair) artificial expression of tankyrase in the nucleus may amplify its nuclear function. CRISPR-generated knockout of tankyrase 1 or tankyrase 2 in human HEK293T cancer cells has no effect on telomere length, but the double knockout induces telomere shortening, which can be rescued by reintroduction of tankyrase 1 or tankyrase 2 (Bhardwaj et al. 2017). Thus, tankyrase is required to maintain telomere length, but either tankyrase 1 or tankyrase 2 is sufficient.
Chromosome resolution
Resolution of sister chromatids in mitosis is essential for proper distribution of genetic material to the daughter cells. Sister chromatids are held together from the time of their replication in S phase until their separation at mitosis by protein complexes termed cohesins (Nasmyth and Haering 2009; Nishiyama 2019). In human cells, cohesins are removed in two stages: first in prophase from chromosome arms, and second in metaphase from centromeres (Waizenegger et al. 2000; Peters and Nishiyama 2012). Human repetitive chromosomal regions (telomeres and ribosomal DNA) rely on additional specialized tankyrase-dependent mechanisms for their resolution. Here we focus on the role of tankyrase in chromosome resolution and genome integrity. The mitotic functions of tankyrase and PARPs were recently reviewed elsewhere (Slade 2019).
Telomere resolution requires tankyrase 1 and tankyrase 2
Cohesion between telomeres is normally resolved at the same time as chromosome arms in G2/M (Ofir et al. 2002; Yalon et al. 2004). However, when cells are treated with tankyrase siRNA chromosome arms and centromeres resolve, but telomeres remain cohered in mitosis (Fig. 3; Dynek and Smith 2004). Cells undergo a prolonged anaphase, but ultimately exit mitosis; although some cell types (like HeLa) exhibit a prolonged mitotic arrest (Dynek and Smith 2004; Kim and Smith 2014). Cells also exhibit an increase in DNA damage foci and in sister telomere fusions (Hsiao and Smith 2009). Rescue of persistent cohesion requires a PARP-active tankyrase 1 (Bisht et al. 2013b). The persistent cohesion phenotype is recapitulated in CRISPR-generated HEK293T TNKS1 knockout cell lines (Bhardwaj et al. 2017). Surprisingly, TNKS2 knockout cells show the same phenotype as TNKS1 knockout, and the double-knockout cells show an even greater level of persistent cohesion than the single knockouts. Thus, in contrast to telomere length maintenance where either tankyrase 1 or tankyrase 2 is sufficient, resolution of telomere cohesion requires both tankyrase 1 and tankyrase 2 (Bhardwaj et al. 2017). CRISPR-generated mutation of the tankyrase-binding site in TRF1 (from 13-RGCADG to RGDADP) leads to persistent telomere cohesion, indicating that TRF1-mediated recruitment of tankyrase is required for resolution (Azarm et al. 2020).
Figure 3.

Tankyrase is required for resolution of telomeres and rDNA. (Top panel) In control cells, arm and telomere cohesion is released in G2/prophase, followed by resolution of centromere and rDNA at the metaphase to anaphase transition. (Bottom panel) In the absence of tankyrase 1, arm cohesion is released in G2/prophase, followed by release of centromere cohesion at the metaphase to anaphase transition. Resolution of telomeres and rDNA is delayed until anaphase leading to anaphase delay, rDNA damage, and aneuploidy.
Telomere resolution requires RNF8-mediated stabilization of tankyrase 1
Tankyrase 1 is recruited to telomeres in late S/G2 phase, aligned with the timing of sister telomere resolution (Bisht et al. 2012, 2013a). What regulates tankyrase to act specifically at telomeres within this window? The answer lies in cell cycle-regulated ubiquitination and deubiquitination of tankyrase 1. Tankyrase is normally turned over rapidly due to RNF146-mediated K48-linked ubiquitination and degradation by the proteasome (Callow et al. 2011; Zhang et al. 2011). However, in late S/G2, the damage-responsive E3 ubiquitin ligase RNF8 promotes K63-linked ubiquitination on tankyrase 1 (Tripathi and Smith 2017). The damage activating RNF8 originates from the endogenous ATM-mediated signaling at newly replicated telomeres (Verdun et al. 2005). The K63-linked ubiquitin promotes transient stabilization of tankyrase 1 to permit timely resolution of cohesion in late S/G2. RNF8 depletion (like tankyrase 1 depletion) leads to persistent telomere cohesion and sister telomere fusions (Hsiao and Smith 2009; Tripathi and Smith 2017).
Following resolution of cohesion, upon nuclear envelope breakdown, tankyrase 1 is exposed to the cytoplasmic K63-deubiquitinating enzyme complex BRISC (Cooper et al. 2010; Feng et al. 2010; Hu et al. 2011). Tankyrase 1 binds to a 333-RPQAVG tankyrase binding site in the BRISC scaffold subunit ABRO1 (Tripathi and Smith 2017). ABRO1 depletion leads to increased K63-ubiquitinated tankyrase 1, premature resolution of telomere cohesion, and sister telomere fragility/loss, demonstrating that limiting tankyrase 1 stabilization and activity to the S/G2 window of the cell cycle is vital for telomere integrity. These phenotypes can be rescued by introduction of wild-type ABRO1, but not a tankyrase binding site mutant (from 333-RPQAVG to RPQAVR). This study suggests that K63-ubiquitination can shunt tankyrase 1 out of the RNF146-mediated K48-ubiquitination degradation cycle into a transiently stabilized state to function in resolution of telomere cohesion (Tripathi and Smith 2017).
Telomere resolution in aging cells
As human cells age and their telomeres shorten they exhibit a natural block in resolution of telomere cohesion. In presenescent cells aged in culture and early passage cells from aged individuals telomeres are cohered in mitosis, whereas other chromosomal regions are resolved (Ofir et al. 2002; Yalon et al. 2004). Introduction of telomerase into presenescent cells rescues the persistent telomere cohesion, suggesting that it is coordinated with telomere shortening (Yalon et al. 2004). Tankyrase levels are not reduced in aging cells, yet overexpression of PARP-active tankyrase 1 rescues the persistent telomere cohesion, hinting at a mechanism that prevents endogenous tankyrase from acting at telomeres in these aged cells (Kim and Smith 2014). Indeed, a recent study shows that shortened telomeres recruit insufficient TRF1 and as a consequence inadequate tankyrase 1 to resolve sister telomere cohesion (Azarm et al. 2020). Overexpression of wild-type TRF1, but not a tankyrase-binding site mutant (from 13-RGCADG to RGDAAA), promotes recruitment of tankyrase 1 to aging cell telomeres and forces resolution of cohesion. Surprisingly, persistent cohesion is beneficial to aging cells; forcing resolution results in deleterious phenotypes: excessive interchromosomal subtelomere recombination, DNA damage, and premature activation of checkpoint-mediated senescence (Azarm et al. 2020). The limited recruitment of tankyrase at shortened telomeres promotes persistent telomere cohesion, protecting chromosome ends from engaging in damage-inducing subtelomere recombination that signals premature cell cycle arrest. Thus, in aging cells the gradual loss of telomere repeats and the accompanying limited recruitment of tankyrase 1 ensures an integrated and measured onset of replicative senescence.
Telomere resolution in ALT cancer cells
Persistent telomere cohesion in mitosis has also been observed to occur naturally in cancer cells that use alternative lengthening of telomeres (ALT), a recombination-based mechanism of telomere maintenance (Bryan et al. 1997; Ramamoorthy and Smith 2015; Sobinoff and Pickett 2017). As with aging cells, tankyrase 1 levels are not reduced in ALT cells, yet overexpression of PARP-active tankyrase 1 can rescue the persistent telomere cohesion (Ramamoorthy and Smith 2015). ALT cell telomeres range from exceptionally long to critically short (Henson et al. 2002). Could the population of shortened telomeres with insufficient TRF1 contribute to persistent telomere cohesion as in aging cells? Indeed, introduction of telomerase into ALT cells rescues the persistent cohesion, linking the phenotype to shortened telomeres (Azarm et al. 2020). Overexpression of wild-type TRF1, but not a tankyrase-binding-site mutant, recruits tankyrase 1 to telomeres, forces resolution of cohesion, and leads to similar phenotypes in ALT as in aging cells: subtelomere recombination, DNA damage, and (in checkpoint-proficient U2OS ALT cells) senescence (Azarm et al. 2020).
A noncanonical role for the PAR-binding macroH2A1.1
In both aging normal cells and ALT cancer cells, reduced TRF1 at shortened telomeres limits tankyrase 1 recruitment (Ramamoorthy and Smith 2015; Azarm et al. 2020). There appears to be an additional mechanism (at least in ALT cells) that acts directly through tankyrase 1. A common feature of ALT cancer cell lines is loss of the SWI/SNF-like ATPase ATRX (Heaphy et al. 2011; Lovejoy et al. 2012). Along with its roles in chromatin remodeling and histone deposition, ATRX sequesters the soluble pool of macroH2A (Ratnakumar et al. 2012), a histone variant comprised of an N-terminal H2A-like domain and a C-terminal PAR-binding macrodomain (Gamble and Kraus 2010; Cantariño et al. 2013). MacroH2A exists as three isoforms (macroH2A1.1, macroH2A1.2, and macroH2A2), but only one macroH2A1.1 (by virtue of an alternative splice) binds PAR (Kustatscher et al. 2005; Timinszky et al. 2009). Loss of ATRX in ALT cells frees the soluble pool of macroH2A1.1 to bind and sequester PARylated tankyrase 1 away from telomeres (Ramamoorthy and Smith 2015). Overexpression of ATRX (or depletion of macroH2A1.1) forces resolution of telomere cohesion, phenocopying TRF1 or tankyrase 1 overexpression. Whether tankyrase 1 is regulated through a similar mechanism in aging cells remains to be determined, but the possibility is supported by studies showing that macroH2A1.1 is highly expressed in cells undergoing senescence (Sporn et al. 2009; Chen et al. 2015).
Ribosomal DNA resolution
To accommodate the high demand for ribosomal RNA, the genes encoding the ribosomal RNA (rDNA) exist in many copies (Gonzalez and Sylvester 1995). In humans the rDNA is in repetitive clusters on the short arms of the five acrocentric human chromosomes adjacent to the telomeres (Henderson et al. 1972; McStay 2016). While it was known that the bulk of the human genome resolves by prophase, except for the centromeres, which resolve at the metaphase-to-anaphase transition (Waizenegger et al. 2000; Nagasaka et al. 2016), the timing of rDNA resolution was unknown. Two recent studies show that rDNA is not resolved until late in anaphase after the rest of the genome (Daniloski et al. 2019; Potapova et al. 2019) and also that tankyrase 1 is required. In tankyrase 1-depleted cells the rDNA loci of sister chromatids remain connected and stretch across the segregating DNA mass in anaphase (Fig. 3; Daniloski et al. 2019). The stretching results from the inaction of condensin II. Two condensin complexes (condensin I and condensin II) with complementing activities exist in human cells (Ono et al. 2003, 2004; Hirota et al. 2004). They each have the common SMC2 and four subunits plus three additional related, but unique, subunits. An in-silico search for tankyrase-binding sites across all human condensin subunits revealed them in only one, CAP-D3 of condensin II. Depletion of CAP-D3 (but not its condensin I counterpart, CAP-D2) led to the same rDNA stretching as tankyrase 1 depletion. Tankyrase 1 binds to CAP-D3 through its 519-RSEPSG tankyrase-binding site, and wild-type CAP-D3 (but not a tankyrase-binding site mutant, RSEPSA) rescues the rDNA stretching and facilitates topoisomerase IIα-mediated segregation of rDNA.
Analysis of the fate of the rDNA in tankyrase 1-depleted cells reveals dramatic consequences for genome integrity: rDNA-specific damage in mitosis, rDNA-containing micronuclei, and nondisjunction of rDNA-containing acrocentric chromosomes (Daniloski et al. 2019). How does defective rDNA resolution lead to damage? The rDNA clusters, like subtelomeres, are “hot spots” for recombination (Killen et al. 2009; Stults et al. 2009; Salim and Gerton 2019). Defective rDNA resolution may lead to aberrant recombination that induces damage, similar to the subtelomere recombination-driven damage in replicative senescence described above.
Telomere and rDNA resolution—a connection through tankyrase 1
Two types of human repetitive sequences (telomeres and the rDNA) require tankyrase 1 for their resolution in mitosis. Surprisingly, despite the common role for tankyrase, resolution of telomeres and rDNA occurs at different stages in mitosis: prophase and anaphase, respectively (Fig. 3). Is their resolution connected? Tankyrase 1 is recruited to telomeres through TRF1, but what directs it to the rDNA? Considering the timing of resolution (telomeres in G2/prophase, and rDNA in anaphase), perhaps tankyrase is initially recruited to telomeres (including those of acrocentric chromosomes) through TRF1, and then relocalizes to the adjacent rDNA locus. The studies described above show timely rDNA resolution occurs in normal cells and in telomerase-positive cancer cells. It remains to be determined whether the process is defective (like resolution of telomere cohesion) in aging and ALT cancer cells. Loss of ATRX in ALT cancer cells has been connected to a reduction in rDNA copy number (Udugama et al. 2018). As described above, loss of ATRX in ALT cells sequesters tankyrase away from telomeres, preventing resolution of telomere cohesion (Ramamoorthy and Smith 2015). Perhaps it also keeps tankyrase 1 away from rDNA, preventing rDNA resolution. The observed loss of rDNA in ALT could be a consequence of defective rDNA resolution.
Perspectives
In this review, we discussed the role of PARP1, PARP2, tankyrase 1, and tankyrase 2 in chromosome repair, replication, and resolution in human cells. Though these four proteins share the ability to PARylate themselves and their binding partners, they differ in their contributions to genome integrity. PARP1 and PARP2 auto-PARylation provides a harbor at sites of DNA damage, disrupted replication forks, or unprocessed Okazaki fragments, for PAR-binding proteins to dock and resolve these aberrations. The harboring of proteins to preserve genome stability is taken further through evidence of compartmentalization through PAR-mediated liquid-liquid demixing at sites of DNA damage. Auto-PARylation of tankyrases, on the other hand, does not appear to motivate function. Rather, it is the ability to interact with an array of diverse binding partners at specific locations and times, as evidenced by the example of tankyrase 1 localization to telomeres through its binding partner TRF1 in G2/M to promote timely resolution of sister telomere cohesion. Together, the combined activities of these four PARPs contribute significantly to genome maintenance and stability. Against the backdrop of expanding strategies for PARP inhibitors in the clinic, future research will continue to identify new targets and provide new insights.
Acknowledgments
This review was supported by the National Institutes of Health under award numbers R01GM129780 and R01CA200751 to S.S. and F31CA221162 to K.A.
Footnotes
Article published online ahead of print. Article and publication date are online at http://www.genesdev.org/cgi/doi/10.1101/gad.334730.119.
References
- Ahel I, Ahel D, Matsusaka T, Clark AJ, Pines J, Boulton SJ, West SC. 2008. Poly(ADP-ribose)-binding zinc finger motifs in DNA repair/checkpoint proteins. Nature 451: 81–85. 10.1038/nature06420 [DOI] [PubMed] [Google Scholar]
- Ahel D, Horejsi Z, Wiechens N, Polo SE, Garcia-Wilson E, Ahel I, Flynn H, Skehel M, West SC, Jackson SP, 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]
- Altmeyer M, Neelsen KJ, Teloni F, Pozdnyakova I, Pellegrino S, Grøfte M, Rask MD, Streicher W, Jungmichel S, Nielsen ML, et al. 2015. Liquid demixing of intrinsically disordered proteins is seeded by poly(ADP-ribose). Nat Commun 6: 8088 10.1038/ncomms9088 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Amé JC, Spenlehauer C, de Murcia G. 2004. The PARP superfamily. Bioessays 26: 882–893. 10.1002/bies.20085 [DOI] [PubMed] [Google Scholar]
- Anachkova B, Russev G, Poirier GG. 1989. DNA replication and poly(ADP-ribosyl)ation of chromatin. Cytobios 58: 19–28. [PubMed] [Google Scholar]
- Azarm A, Bhardwaj A, Kim E, Smith S. 2020. Persistent telomere cohesion protects aged cells from premature senescence. bioRxiv doi:10.1101/2020.02.03.932145. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Banani SF, Lee HO, Hyman AA, Rosen MK. 2017. Biomolecular condensates: organizers of cellular biochemistry. Nat Rev Mol Cell Biol 18: 285–298. 10.1038/nrm.2017.7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Beck C, Boehler C, Guirouilh Barbat J, Bonnet ME, Illuzzi G, Ronde P, Gauthier LR, Magroun N, Rajendran A, Lopez BS, et al. 2014a. PARP3 affects the relative contribution of homologous recombination and nonhomologous end-joining pathways. Nucleic Acids Res 42: 5616–5632. 10.1093/nar/gku174 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Beck C, Robert I, Reina-San-Martin B, Schreiber V, Dantzer F. 2014b. 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]
- Berti M, Ray Chaudhuri A, Thangavel S, Gomathinayagam S, Kenig S, Vujanovic M, Odreman F, Glatter T, Graziano S, Mendoza-Maldonado R, et al. 2013. Human RECQ1 promotes restart of replication forks reversed by DNA topoisomerase I inhibition. Nat Struct Mol Biol 20: 347–354. 10.1038/nsmb.2501 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bhardwaj A, Yang Y, Ueberheide B, 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]
- Bilan V, Leutert M, Nanni P, Panse C, Hottiger MO. 2017. Combining higher-energy collision dissociation and electron-transfer/higher-energy collision dissociation fragmentation in a product-dependent manner confidently assigns proteomewide ADP-ribose acceptor sites. Anal Chem 89: 1523–1530. 10.1021/acs.analchem.6b03365 [DOI] [PubMed] [Google Scholar]
- Bisht KK, Dudognon C, Chang WG, Sokol ES, Ramirez A, Smith S. 2012. GDP-mannose-4,6-dehydratase is a cytosolic partner of tankyrase 1 that inhibits its poly(ADP-ribose) polymerase activity. Mol Cell Biol 32: 3044–3053. 10.1128/MCB.00258-12 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bisht KK, Daniloski Z, Smith S. 2013a. SA1 binds directly to DNA through its unique AT-hook to promote sister chromatid cohesion at telomeres. J Cell Sci 126: 3493–3503. 10.1242/jcs.130872 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bisht KK, Daniloski Z, Smith S. 2013b. SA1 binds directly to DNA through its unique AT-hook to promote sister chromatid cohesion at telomeres. J Cell Sci 126: 3493–3503. 10.1242/jcs.130872 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boehler C, Gauthier LR, Mortusewicz O, Biard DS, Saliou JM, Bresson A, Sanglier-Cianferani S, Smith S, Schreiber V, Boussin F, et al. 2011. Poly(ADP-ribose) polymerase 3 (PARP3), a newcomer in cellular response to DNA damage and mitotic progression. Proc Natl Acad Sci 108: 2783–2788. 10.1073/pnas.1016574108 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bonfiglio JJ, Fontana P, Zhang Q, Colby T, Gibbs-Seymour I, Atanassov I, Bartlett E, Zaja R, Ahel I, Matic I. 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]
- Brázda V, Laister RC, Jagelská EB, Arrowsmith C. 2011. Cruciform structures are a common DNA feature important for regulating biological processes. BMC Mol Biol 12: 33 10.1186/1471-2199-12-33 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Breslin C, Hornyak P, Ridley A, Rulten SL, Hanzlikova H, Oliver AW, Caldecott KW. 2015. The XRCC1 phosphate-binding pocket binds poly (ADP-ribose) and is required for XRCC1 function. Nucleic Acids Res 43: 6934–6944. 10.1093/nar/gkv623 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bryan TM, Englezou A, Dalla-Pozza L, Dunham MA, Reddel RR. 1997. Evidence for an alternative mechanism for maintaining telomere length in human tumors and tumor-derived cell lines. Nat Med 3: 1271–1274. 10.1038/nm1197-1271 [DOI] [PubMed] [Google Scholar]
- Bryant HE, Schultz N, Thomas HD, Parker KM, Flower D, Lopez E, Kyle S, Meuth M, Curtin NJ, Helleday T. 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]
- Caldecott KW. 2019. XRCC1 protein; Form and function. DNA Repair (Amst) 81: 102664 10.1016/j.dnarep.2019.102664 [DOI] [PubMed] [Google Scholar]
- Caldecott KW, McKeown CK, Tucker JD, Ljungquist S, Thompson LH. 1994. An interaction between the mammalian DNA repair protein XRCC1 and DNA ligase III. Mol Cell Biol 14: 68–76. 10.1128/MCB.14.1.68 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Callow MG, Tran H, Phu L, Lau T, Lee J, Sandoval WN, Liu PS, Bheddah S, Tao J, Lill JR, et al. 2011. Ubiquitin ligase RNF146 regulates tankyrase and Axin to promote Wnt signaling. PLoS One 6: e22595 10.1371/journal.pone.0022595 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cantariño N, Douet J, Buschbeck M. 2013. MacroH2A–an epigenetic regulator of cancer. Cancer Lett 336: 247–252. 10.1016/j.canlet.2013.03.022 [DOI] [PubMed] [Google Scholar]
- Caron MC, Sharma AK, O'Sullivan J, Myler LR, Ferreira MT, Rodrigue A, Coulombe Y, Ethier C, Gagné JP, Langelier MF, et al. 2019. Poly(ADP-ribose) polymerase-1 antagonizes DNA resection at double-strand breaks. Nat Commun 10: 2954 10.1038/s41467-019-10741-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ceccaldi R, Liu JC, Amunugama R, Hajdu I, Primack B, Petalcorin MI, O'Connor KW, Konstantinopoulos PA, Elledge SJ, Boulton SJ, et al. 2015. Homologous-recombination-deficient tumours are dependent on Polθ-mediated repair. Nature 518: 258–262. 10.1038/nature14184 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chang W, Dynek JN, Smith S. 2003. TRF1 is degraded by ubiquitin-mediated proteolysis after release from telomeres. Genes Dev 17: 1328–1333. 10.1101/gad.1077103 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chang W, Dynek JN, Smith S. 2005. NuMA is a major acceptor of poly(ADP-ribosyl)ation by tankyrase 1 in mitosis. Biochem J 391: 177–184. 10.1042/BJ20050885 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen H, Ruiz PD, McKimpson WM, Novikov L, Kitsis RN, Gamble MJ. 2015. MacroH2A1 and ATM play opposing roles in paracrine senescence and the senescence-associated secretory phenotype. Mol Cell 59: 719–731. 10.1016/j.molcel.2015.07.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cheng Q, Barboule N, Frit P, Gomez D, Bombarde O, Couderc B, Ren GS, Salles B, Calsou P. 2011. Ku counteracts mobilization of PARP1 and MRN in chromatin damaged with DNA double-strand breaks. Nucleic Acids Res 39: 9605–9619. 10.1093/nar/gkr656 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chiang YJ, Nguyen ML, Gurunathan S, Kaminker P, Tessarollo L, Campisi J, Hodes RJ. 2006. Generation and characterization of telomere length maintenance in tankyrase 2-deficient mice. Mol Cell Biol 26: 2037–2043. 10.1128/MCB.26.6.2037-2043.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chiang YJ, Hsiao SJ, Yver D, Cushman SW, Tessarollo L, Smith S, Hodes RJ. 2008. Tankyrase 1 and tankyrase 2 are essential but redundant for mouse embryonic development. PLoS One 3: e2639 10.1371/journal.pone.0002639 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chong PA, Forman-Kay JD. 2016. Liquid-liquid phase separation in cellular signaling systems. Curr Opin Struct Biol 41: 180–186. 10.1016/j.sbi.2016.08.001 [DOI] [PubMed] [Google Scholar]
- Chong PA, Vernon RM, Forman-Kay JD. 2018. RGG/RG motif regions in RNA binding and phase separation. J Mol Biol 430: 4650–4665. 10.1016/j.jmb.2018.06.014 [DOI] [PubMed] [Google Scholar]
- Cook BD, Dynek JN, Chang W, Shostak G, Smith S. 2002. Role for the related poly(ADP-Ribose) polymerases tankyrase 1 and 2 at human telomeres. Mol Cell Biol 22: 332–342. 10.1128/MCB.22.1.332-342.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cooper EM, Boeke JD, Cohen RE. 2010. Specificity of the BRISC deubiquitinating enzyme is not due to selective binding to Lys63-linked polyubiquitin. J Biol Chem 285: 10344–10352. 10.1074/jbc.M109.059667 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cortez D. 2015. Preventing replication fork collapse to maintain genome integrity. DNA Repair (Amst) 32: 149–157. 10.1016/j.dnarep.2015.04.026 [DOI] [PMC free article] [PubMed] [Google Scholar]
- d'Adda di Fagagna F, Reaper PM, Clay-Farrace L, Fiegler H, Carr P, Von Zglinicki T, Saretzki G, Carter NP, Jackson SP. 2003. A DNA damage checkpoint response in telomere-initiated senescence. Nature 426: 194–198. 10.1038/nature02118 [DOI] [PubMed] [Google Scholar]
- D'Amours D, Desnoyers S, D'Silva I, Poirier GG. 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]
- Daniels CM, Ong SE, Leung AK. 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]
- Daniloski Z, Bisht KK, McStay B, Smith S. 2019. Resolution of human ribosomal DNA occurs in anaphase, dependent on tankyrase 1, condensin II, and topoisomerase IIα. Genes Dev 33: 276–281. 10.1101/gad.321836.118 [DOI] [PMC free article] [PubMed] [Google Scholar]
- DaRosa PA, Wang Z, Jiang X, Pruneda JN, Cong F, Klevit RE, Xu W. 2015. Allosteric activation of the RNF146 ubiquitin ligase by a poly(ADP-ribosyl)ation signal. Nature 517: 223–226. 10.1038/nature13826 [DOI] [PMC free article] [PubMed] [Google Scholar]
- DaRosa PA, Klevit RE, Xu W. 2018. Structural basis for tankyrase-RNF146 interaction reveals noncanonical tankyrase-binding motifs. Protein Sci 27: 1057–1067. 10.1002/pro.3413 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dawicki-McKenna JM, Langelier MF, DeNizio JE, Riccio AA, Cao CD, Karch KR, McCauley M, Steffen JD, Black BE, Pascal JM. 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]
- de Lange T. 2005. Shelterin: the protein complex that shapes and safeguards human telomeres. Genes Dev 19: 2100–2110. 10.1101/gad.1346005 [DOI] [PubMed] [Google Scholar]
- de Lange T. 2018. Shelterin-mediated telomere protection. Annu Rev Genet 52: 223–247. 10.1146/annurev-genet-032918-021921 [DOI] [PubMed] [Google Scholar]
- de Murcia G, Menissier 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]
- de Murcia JM, Niedergang C, Trucco C, Ricoul M, Dutrillaux B, Mark M, Oliver FJ, Masson M, Dierich A, LeMeur M, et al. 1997. Requirement of poly(ADP-ribose) polymerase in recovery from DNA damage in mice and in cells. Proc Natl Acad Sci 94: 7303–7307. 10.1073/pnas.94.14.7303 [DOI] [PMC free article] [PubMed] [Google Scholar]
- De Rycker M, Price CM. 2004. Tankyrase polymerization is controlled by its sterile α motif and poly(ADP-ribose) polymerase domains. Mol Cell Biol 24: 9802–9812. 10.1128/MCB.24.22.9802-9812.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- De Rycker M, Venkatesan RN, Wei C, Price CM. 2003. Vertebrate tankyrase domain structure and sterile α motif (SAM)-mediated multimerization. Biochem J 372: 87–96. 10.1042/bj20021450 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dynek JN, 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]
- Eisemann T, McCauley M, Langelier MF, Gupta K, Roy S, Van Duyne GD, Pascal JM. 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]
- Eisemann T, Langelier MF, Pascal JM. 2019. Structural and functional analysis of parameters governing tankyrase-1 interaction with telomeric repeat-binding factor 1 and GDP-mannose 4,6-dehydratase. J Biol Chem 294: 14574–14590. 10.1074/jbc.RA119.009200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- el-Deiry WS, Tokino T, Velculescu VE, Levy DB, Parsons R, Trent JM, Lin D, Mercer WE, Kinzler KW, Vogelstein B. 1993. WAF1, a potential mediator of p53 tumor suppression. Cell 75: 817–825. 10.1016/0092-8674(93)90500-P [DOI] [PubMed] [Google Scholar]
- Eustermann S, Brockmann C, Mehrotra PV, Yang JC, Loakes D, West SC, Ahel I, Neuhaus D. 2010. Solution structures of the two PBZ domains from human APLF and their interaction with poly(ADP-ribose). Nat Struct Mol Biol 17: 241–243. 10.1038/nsmb.1747 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fan C, Yarravarapu N, Chen H, Kulak O, Dasari P, Herbert J, Yamaguchi K, Lum L, Zhang X. 2018. Regulation of tankyrase activity by a catalytic domain dimer interface. Biochem Biophys Res Commun 503: 1780–1785. 10.1016/j.bbrc.2018.07.113 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feijs KL, Forst AH, Verheugd P, Lüscher B. 2013. Macrodomain-containing proteins: regulating new intracellular functions of mono(ADP-ribosyl)ation. Nat Rev Mol Cell Biol 14: 443–451. 10.1038/nrm3601 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feng L, Wang J, Chen J. 2010. The Lys63-specific deubiquitinating enzyme BRCC36 is regulated by two scaffold proteins localizing in different subcellular compartments. J Biol Chem 285: 30982–30988. 10.1074/jbc.M110.135392 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fenton AL, Shirodkar P, Macrae CJ, Meng L, Koch CA. 2013. The PARP3- and ATM-dependent phosphorylation of APLF facilitates DNA double-strand break repair. Nucleic Acids Res 41: 4080–4092. 10.1093/nar/gkt134 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Forsyth NR, Wright WE, Shay JW. 2002. Telomerase and differentiation in multicellular organisms: turn it off, turn it on, and turn it off again. Differentiation 69: 188–197. 10.1046/j.1432-0436.2002.690412.x [DOI] [PubMed] [Google Scholar]
- Frouin I, Maga G, Denegri M, Riva F, Savio M, Spadari S, Prosperi E, Scovassi AI. 2003. Human proliferating cell nuclear antigen, poly(ADP-ribose) polymerase-1, and p21waf1/cip1. A dynamic exchange of partners. J Biol Chem 278: 39265–39268. 10.1074/jbc.C300098200 [DOI] [PubMed] [Google Scholar]
- Gagné JP, Isabelle M, Lo KS, Bourassa S, Hendzel MJ, Dawson VL, Dawson TM, Poirier GG. 2008. Proteome-wide identification of poly(ADP-ribose) binding proteins and poly(ADP-ribose)-associated protein complexes. Nucleic Acids Res 36: 6959–6976. 10.1093/nar/gkn771 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gamble MJ, Kraus WL. 2010. Multiple facets of the unique histone variant macroH2A: from genomics to cell biology. Cell Cycle 9: 2568–2574. 10.4161/cc.9.13.12144 [DOI] [PubMed] [Google Scholar]
- Gibbs-Seymour I, Fontana P, Rack JGM, 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]
- Gilson E, Géli V. 2007. How telomeres are replicated. Nat Rev Mol Cell Biol 8: 825–838. 10.1038/nrm2259 [DOI] [PubMed] [Google Scholar]
- Gonzalez IL, Sylvester JE. 1995. Complete sequence of the 43-kb human ribosomal DNA repeat: analysis of the intergenic spacer. Genomics 27: 320–328. 10.1006/geno.1995.1049 [DOI] [PubMed] [Google Scholar]
- Gottschalk AJ, Timinszky G, Kong SE, Jin J, Cai Y, Swanson SK, Washburn MP, Florens L, Ladurner AG, Conaway JW, et al. 2009. Poly(ADP-ribosyl)ation directs recruitment and activation of an ATP-dependent chromatin remodeler. Proc Natl Acad Sci 106: 13770–13774. 10.1073/pnas.0906920106 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gottschalk AJ, Trivedi RD, Conaway JW, Conaway RC. 2012. Activation of the SNF2 family ATPase ALC1 by poly(ADP-ribose) in a stable ALC1·PARP1·nucleosome intermediate. J Biol Chem 287: 43527–43532. 10.1074/jbc.M112.401141 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Greider CW, Blackburn EH. 1985. Identification of a specific telomere terminal transferase activity in Tetrahymena extracts. Cell 43: 405–413. 10.1016/0092-8674(85)90170-9 [DOI] [PubMed] [Google Scholar]
- Greider CW, Blackburn EH. 1987. The telomere terminal transferase of Tetrahymena is a ribonucleoprotein enzyme with two kinds of primer specificity. Cell 51: 887–898. 10.1016/0092-8674(87)90576-9 [DOI] [PubMed] [Google Scholar]
- Grundy GJ, Polo LM, Zeng Z, Rulten SL, Hoch NC, Paomephan P, Xu Y, Sweet SM, Thorne AW, Oliver AW, et al. 2016. PARP3 is a sensor of nicked nucleosomes and monoribosylates histone H2BGlu2. Nat Commun 7: 12404 10.1038/ncomms12404 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guettler S, LaRose J, Petsalaki E, Gish G, Scotter A, Pawson T, Rottapel R, Sicheri F. 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]
- Gupte R, Liu Z, Kraus WL. 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]
- Hahn WC, Stewart SA, Brooks MW, York SG, Eaton E, Kurachi A, Beijersbergen RL, Knoll JH, Meyerson M, Weinberg RA. 1999. Inhibition of telomerase limits the growth of human cancer cells. Nat Med 5: 1164–1170. 10.1038/13495 [DOI] [PubMed] [Google Scholar]
- Haikarainen T, Krauss S, Lehtio L. 2014. Tankyrases: structure, function and therapeutic implications in cancer. Curr Pharm Des 20: 6472–6488. 10.2174/1381612820666140630101525 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Haince JF, McDonald D, Rodrigue A, Déry U, Masson JY, Hendzel MJ, Poirier GG. 2008. PARP1-dependent kinetics of recruitment of MRE11 and NBS1 proteins to multiple DNA damage sites. J Biol Chem 283: 1197–1208. 10.1074/jbc.M706734200 [DOI] [PubMed] [Google Scholar]
- Hanzlikova H, Caldecott KW. 2019. Perspectives on PARPs in S phase. Trends Genet 35: 412–422. 10.1016/j.tig.2019.03.008 [DOI] [PubMed] [Google Scholar]
- Hanzlikova H, Gittens W, Krejcikova K, Zeng Z, Caldecott KW. 2017. Overlapping roles for PARP1 and PARP2 in the recruitment of endogenous XRCC1 and PNKP into oxidized chromatin. Nucleic Acids Res 45: 2546–2557. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hanzlikova H, Kalasova I, Demin AA, Pennicott LE, Cihlarova Z, Caldecott KW. 2018. The importance of Poly(ADP-Ribose) polymerase as a sensor of unligated Okazaki fragments during DNA r. Mol Cell 71: 319–331.e3. 10.1016/j.molcel.2018.06.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heaphy CM, Subhawong AP, Hong SM, Goggins MG, Montgomery EA, Gabrielson E, Netto GJ, Epstein JI, Lotan TL, Westra WH, et al. 2011. Prevalence of the alternative lengthening of telomeres telomere maintenance mechanism in human cancer subtypes. Am J Pathol 179: 1608–1615. 10.1016/j.ajpath.2011.06.018 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Henderson AS, Warburton D, Atwood KC. 1972. Location of ribosomal DNA in the human chromosome complement. Proc Natl Acad Sci 69: 3394–3398. 10.1073/pnas.69.11.3394 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Henson JD, Neumann AA, Yeager TR, Reddel RR. 2002. Alternative lengthening of telomeres in mammalian cells. Oncogene 21: 598–610. 10.1038/sj.onc.1205058 [DOI] [PubMed] [Google Scholar]
- Hirota T, Gerlich D, Koch B, Ellenberg J, Peters JM. 2004. Distinct functions of condensin I and II in mitotic chromosome assembly. J Cell Sci 117: 6435–6445. 10.1242/jcs.01604 [DOI] [PubMed] [Google Scholar]
- Hochegger H, Dejsuphong D, Fukushima T, Morrison C, Sonoda E, Schreiber V, Zhao GY, Saberi A, Masutani M, Adachi N, et al. 2006. Parp-1 protects homologous recombination from interference by Ku and Ligase IV in vertebrate cells. EMBO J 25: 1305–1314. 10.1038/sj.emboj.7601015 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hottiger MO, Hassa PO, Lüscher B, Schüler H, 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]
- Hsiao SJ, 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]
- Hsiao SJ, Smith S. 2009. Sister telomeres rendered dysfunctional by persistent cohesion are fused by NHEJ. J Cell Biol 184: 515–526. 10.1083/jcb.200810132 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hsiao SJ, Poitras MF, Cook BD, Liu Y, Smith S. 2006. Tankyrase 2 poly(ADP-ribose) polymerase domain-deleted mice exhibit growth defects but have normal telomere length and capping. Mol Cell Biol 26: 2044–2054. 10.1128/MCB.26.6.2044-2054.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hu X, Kim JA, Castillo A, Huang M, Liu J, Wang B. 2011. NBA1/MERIT40 and BRE interaction is required for the integrity of two distinct deubiquitinating enzyme BRCC36-containing complexes. J Biol Chem 286: 11734–11745. 10.1074/jbc.M110.200857 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang SM, Mishina YM, Liu S, Cheung A, Stegmeier F, Michaud GA, Charlat O, Wiellette E, Zhang Y, Wiessner S, et al. 2009. Tankyrase inhibition stabilizes axin and antagonizes Wnt signalling. Nature 461: 614–620. 10.1038/nature08356 [DOI] [PubMed] [Google Scholar]
- James DI, Smith KM, Jordan AM, Fairweather EE, Griffiths LA, Hamilton NS, Hitchin JR, Hutton CP, Jones S, Kelly P, 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]
- Ji Y, Tulin AV. 2009. Poly(ADP-ribosyl)ation of heterogeneous nuclear ribonucleoproteins modulates splicing. Nucleic Acids Res 37: 3501–3513. 10.1093/nar/gkp218 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jungmichel S, Stucki M. 2010. MDC1: the art of keeping things in focus. Chromosoma 119: 337–349. 10.1007/s00412-010-0266-9 [DOI] [PubMed] [Google Scholar]
- Kang HC, Lee YI, Shin JH, Andrabi SA, Chi Z, Gagne JP, Lee Y, Ko HS, Lee BD, Poirier GG, et al. 2011. Iduna is a poly(ADP-ribose) (PAR)-dependent E3 ubiquitin ligase that regulates DNA damage. Proc Natl Acad Sci 108: 14103–14108. 10.1073/pnas.1108799108 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Karras GI, Kustatscher G, Buhecha HR, Allen MD, Pugieux C, Sait F, Bycroft M, Ladurner AG. 2005. The macro domain is an ADP-ribose binding module. EMBO J 24: 1911–1920. 10.1038/sj.emboj.7600664 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaufmann T, Grishkovskaya I, Polyansky AA, Kostrhon S, Kukolj E, Olek KM, Herbert S, Beltzung E, Mechtler K, Peterbauer T, et al. 2017. A novel non-canonical PIP-box mediates PARG interaction with PCNA. Nucleic Acids Res 45: 9741–9759. 10.1093/nar/gkx604 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Killen MW, Stults DM, Adachi N, Hanakahi L, Pierce AJ. 2009. Loss of Bloom syndrome protein destabilizes human gene cluster architecture. Hum Mol Genet 18: 3417–3428. 10.1093/hmg/ddp282 [DOI] [PubMed] [Google Scholar]
- Kim MK, Smith S. 2014. Persistent telomere cohesion triggers a prolonged anaphase. Mol Biol Cell 25: 30–40. 10.1091/mbc.e13-08-0479 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim NW, Piatyszek MA, Prowse KR, Harley CB, West MD, Ho PL, Coviello GM, Wright WE, Weinrich SL, Shay JW. 1994. Specific association of human telomerase activity with immortal cells and cancer. Science 266: 2011–2015. 10.1126/science.7605428 [DOI] [PubMed] [Google Scholar]
- Kustatscher G, Hothorn M, Pugieux C, Scheffzek K, Ladurner AG. 2005. Splicing regulates NAD metabolite binding to histone macroH2A. Nat Struct Mol Biol 12: 624–625. 10.1038/nsmb956 [DOI] [PubMed] [Google Scholar]
- Langelier MF, Planck JL, Roy S, Pascal JM. 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]
- Langelier MF, Riccio AA, Pascal JM. 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. 10.1093/nar/gku474 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Langelier MF, Eisemann T, Riccio AA, Pascal JM. 2018a. PARP family enzymes: regulation and catalysis of the poly(ADP-ribose) posttranslational modification. Curr Opin Struct Biol 53: 187–198. 10.1016/j.sbi.2018.11.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Langelier MF, Zandarashvili L, Aguiar PM, Black BE, Pascal JM. 2018b. 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]
- Larsen SC, Hendriks IA, Lyon D, Jensen LJ, Nielsen ML. 2018. Systems-wide analysis of serine ADP-ribosylation reveals widespread occurrence and site-specific overlap with phosphorylation. Cell Rep 24: 2493–2505.e4. 10.1016/j.celrep.2018.07.083 [DOI] [PubMed] [Google Scholar]
- Lee MH, Na H, Kim EJ, Lee HW, Lee MO. 2012. Poly(ADP-ribosyl)ation of p53 induces gene-specific transcriptional repression of MTA1. Oncogene 31: 5099–5107. 10.1038/onc.2012.2 [DOI] [PubMed] [Google Scholar]
- Lehmann AR, Kirk-Bell S, Shall S, Whish WJ. 1974. The relationship between cell growth, macromolecular synthesis and poly ADP-ribose polymerase in lymphoid cells. Exp Cell Res 83: 63–72. 10.1016/0014-4827(74)90688-0 [DOI] [PubMed] [Google Scholar]
- Leidecker O, Bonfiglio JJ, Colby T, Zhang Q, Atanassov I, Zaja R, Palazzo L, Stockum A, Ahel I, Matic I. 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]
- Leslie Pedrioli DM, Leutert M, Bilan V, Nowak K, Gunasekera K, Ferrari E, Imhof R, Malmström L, Hottiger MO. 2018. Comprehensive ADP-ribosylome analysis identifies tyrosine as an ADP-ribose acceptor site. EMBO Rep 19 10.15252/embr.201745310 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leung AK. 2014. Poly(ADP-ribose): an organizer of cellular architecture. J Cell Biol 205: 613–619. 10.1083/jcb.201402114 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Levaot N, Voytyuk O, Dimitriou I, Sircoulomb F, Chandrakumar A, Deckert M, Krzyzanowski PM, Scotter A, Gu S, Janmohamed S, 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]
- Li M, Yu X. 2013. Function of BRCA1 in the DNA damage response is mediated by ADP-ribosylation. Cancer Cell 23: 693–704. 10.1016/j.ccr.2013.03.025 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li GY, McCulloch RD, Fenton AL, Cheung M, Meng L, Ikura M, Koch CA. 2010. Structure and identification of ADP-ribose recognition motifs of APLF and role in the DNA damage response. Proc Natl Acad Sci 107: 9129–9134. 10.1073/pnas.1000556107 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li M, Lu LY, Yang CY, Wang S, Yu X. 2013. The FHA and BRCT domains recognize ADP-ribosylation during DNA damage response. Genes Dev 27: 1752–1768. 10.1101/gad.226357.113 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li N, Zhang Y, Han X, Liang K, Wang J, Feng L, Wang W, Songyang Z, Lin C, Yang L, et al. 2015. Poly-ADP ribosylation of PTEN by tankyrases promotes PTEN degradation and tumor growth. Genes Dev 29: 157–170. 10.1101/gad.251785.114 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li X, Han H, Zhou MT, Yang B, Ta AP, Li N, Chen J, Wang W. 2017. Proteomic analysis of the human tankyrase protein interaction network reveals its role in pexophagy. Cell Rep 20: 737–749. 10.1016/j.celrep.2017.06.077 [DOI] [PubMed] [Google Scholar]
- Lin W, Amé JC, Aboul-Ela N, Jacobson EL, Jacobson MK. 1997. Isolation and characterization of the cDNA encoding bovine poly(ADP-ribose) glycohydrolase. J Biol Chem 272: 11895–11901. 10.1074/jbc.272.18.11895 [DOI] [PubMed] [Google Scholar]
- Lingner J, Hughes TR, Shevchenko A, Mann M, Lundblad V, Cech TR. 1997. Reverse transcriptase motifs in the catalytic subunit of telomerase. Science 276: 561–567. 10.1126/science.276.5312.561 [DOI] [PubMed] [Google Scholar]
- Lovejoy CA, Li W, Reisenweber S, Thongthip S, Bruno J, de Lange T, De S, Petrini JH, Sung PA, Jasin M, et al. 2012. Loss of ATRX, genome instability, and an altered DNA damage response are hallmarks of the alternative lengthening of telomeres pathway. PLoS Genet 8: e1002772 10.1371/journal.pgen.1002772 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luijsterburg MS, de Krijger I, Wiegant WW, Shah RG, Smeenk G, de Groot AJL, Pines A, Vertegaal ACO, Jacobs JJL, Shah GM, et al. 2016. PARP1 links CHD2-mediated chromatin expansion and H3.3 deposition to DNA repair by non-homologous end-joining. Mol Cell 61: 547–562. 10.1016/j.molcel.2016.01.019 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lüscher B, Bütepage M, Eckei L, Krieg S, Verheugd P, Shilton BH. 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]
- Madison DL, Lundblad JR. 2010. C-terminal binding protein and poly(ADP)ribose polymerase 1 contribute to repression of the p21waf1/cip1 promoter. Oncogene 29: 6027–6039. 10.1038/onc.2010.338 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Malanga M, Czubaty A, Girstun A, Staron K, Althaus FR. 2008. Poly(ADP-ribose) binds to the splicing factor ASF/SF2 and regulates its phosphorylation by DNA topoisomerase I. J Biol Chem 283: 19991–19998. 10.1074/jbc.M709495200 [DOI] [PubMed] [Google Scholar]
- Mariotti L, Templeton CM, Ranes M, Paracuellos P, Cronin N, Beuron F, Morris E, Guettler S. 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]
- Martello R, Leutert M, Jungmichel S, Bilan V, Larsen SC, Young C, Hottiger MO, Nielsen ML. 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]
- Martin-Hernandez K, Rodriguez-Vargas JM, Schreiber V, Dantzer F. 2017. Expanding functions of ADP-ribosylation in the maintenance of genome integrity. Semin Cell Dev Biol 63: 92–101. 10.1016/j.semcdb.2016.09.009 [DOI] [PubMed] [Google Scholar]
- Masson M, Niedergang C, Schreiber V, Muller S, Menissier-de Murcia J, de Murcia G. 1998. XRCC1 is specifically associated with poly(ADP-ribose) polymerase and negatively regulates its activity following DNA damage. Mol Cell Biol 18: 3563–3571. 10.1128/MCB.18.6.3563 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mastrocola AS, Kim SH, Trinh AT, Rodenkirch LA, Tibbetts RS. 2013. The RNA-binding protein fused in sarcoma (FUS) functions downstream of poly(ADP-ribose) polymerase (PARP) in response to DNA damage. J Biol Chem 288: 24731–24741. 10.1074/jbc.M113.497974 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Masutani M, Nozaki T, Nakamoto K, Nakagama H, Suzuki H, Kusuoka O, Tsutsumi M, Sugimura T. 2000. The response of Parp knockout mice against DNA damaging agents. Mutat Res 462: 159–166. 10.1016/S1383-5742(00)00033-8 [DOI] [PubMed] [Google Scholar]
- Mateos-Gomez PA, Gong F, Nair N, Miller KM, Lazzerini-Denchi E, 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]
- Maya-Mendoza A, Moudry P, Merchut-Maya JM, Lee M, Strauss R, Bartek J. 2018. High speed of fork progression induces DNA replication stress and genomic instability. Nature 559: 279–284. 10.1038/s41586-018-0261-5 [DOI] [PubMed] [Google Scholar]
- McStay B. 2016. Nucleolar organizer regions: genomic ‘dark matter’ requiring illumination. Genes Dev 30: 1598–1610. 10.1101/gad.283838.116 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Menissier de Murcia J, Ricoul M, Tartier L, Niedergang C, Huber A, Dantzer F, Schreiber V, Amé JC, Dierich A, LeMeur M, 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]
- Mosavi LK, Cammett TJ, Desrosiers DC, Peng ZY. 2004. The ankyrin repeat as molecular architecture for protein recognition. Protein Sci 13: 1435–1448. 10.1110/ps.03554604 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nagasaka K, Hossain MJ, Roberti MJ, Ellenberg J, Hirota T. 2016. Sister chromatid resolution is an intrinsic part of chromosome organization in prophase. Nat Cell Biol 18: 692–699. 10.1038/ncb3353 [DOI] [PubMed] [Google Scholar]
- Nagy Z, Kalousi A, Furst A, Koch M, Fischer B, 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]
- Nasmyth K, Haering CH. 2009. Cohesin: its roles and mechanisms. Annu Rev Genet 43: 525–558. 10.1146/annurev-genet-102108-134233 [DOI] [PubMed] [Google Scholar]
- Nicolae CM, Aho ER, Vlahos AH, Choe KN, De S, Karras GI, Moldovan GL. 2014. The ADP-ribosyltransferase PARP10/ARTD10 interacts with proliferating cell nuclear antigen (PCNA) and is required for DNA damage tolerance. J Biol Chem 289: 13627–13637. 10.1074/jbc.M114.556340 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nicolae CM, Aho ER, Choe KN, Constantin D, Hu HJ, Lee D, Myung K, Moldovan GL. 2015. A novel role for the mono-ADP-ribosyltransferase PARP14/ARTD8 in promoting homologous recombination and protecting against replication stress. Nucleic Acids Res 43: 3143–3153. 10.1093/nar/gkv147 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nishiyama T. 2019. Cohesion and cohesin-dependent chromatin organization. Curr Opin Cell Biol 58: 8–14. 10.1016/j.ceb.2018.11.006 [DOI] [PubMed] [Google Scholar]
- Obaji E, Haikarainen T, Lehtiö L. 2016. Characterization of the DNA dependent activation of human ARTD2/PARP2. Sci Rep 6: 34487 10.1038/srep34487 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Obaji E, Haikarainen T, Lehtiö L. 2018. Structural basis for DNA break recognition by ARTD2/PARP2. Nucleic Acids Res 46: 12154–12165. 10.1093/nar/gky927 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oberoi J, Richards MW, Crumpler S, Brown N, Blagg J, Bayliss R. 2010. Structural basis of poly(ADP-ribose) recognition by the multizinc binding domain of checkpoint with forkhead-associated and RING domains (CHFR). J Biol Chem 285: 39348–39358. 10.1074/jbc.M110.159855 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ofir R, Yalon-Hacohen M, Segev Y, Schultz A, Skorecki KL, Selig S. 2002. Replication and/or separation of some human telomeres is delayed beyond S-phase in pre-senescent cells. Chromosoma 111: 147–155. 10.1007/s00412-002-0199-z [DOI] [PubMed] [Google Scholar]
- Okamoto K, Ohishi T, Kuroiwa M, Iemura SI, Natsume T, Seimiya H. 2018. MERIT40-dependent recruitment of tankyrase to damaged DNA and its implication for cell sensitivity to DNA-damaging anticancer drugs. Oncotarget 9: 35844–35855. 10.18632/oncotarget.26312 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Okano S, Lan L, Caldecott KW, Mori T, Yasui A. 2003. Spatial and temporal cellular responses to single-strand breaks in human cells. Mol Cell Biol 23: 3974–3981. 10.1128/MCB.23.11.3974-3981.2003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ono T, Losada A, Hirano M, Myers MP, Neuwald AF, Hirano T. 2003. Differential contributions of condensin I and condensin II to mitotic chromosome architecture in vertebrate cells. Cell 115: 109–121. 10.1016/S0092-8674(03)00724-4 [DOI] [PubMed] [Google Scholar]
- Ono T, Fang Y, Spector DL, Hirano T. 2004. Spatial and temporal regulation of Condensins I and II in mitotic chromosome assembly in human cells. Mol Biol Cell 15: 3296–3308. 10.1091/mbc.e04-03-0242 [DOI] [PMC free article] [PubMed] [Google Scholar]
- O'Sullivan J, Tedim Ferreira M, Gagné JP, Sharma AK, Hendzel MJ, Masson JY, Poirier GG. 2019. Emerging roles of eraser enzymes in the dynamic control of protein ADP-ribosylation. Nat Commun 10: 1182 10.1038/s41467-019-08859-x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Palazzo L, Mikoč A, 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]
- Palm W, de Lange T. 2008. How shelterin protects mammalian telomeres. Annu Rev Genet 42: 301–334. 10.1146/annurev.genet.41.110306.130350 [DOI] [PubMed] [Google Scholar]
- Pasero P, Vindigni A. 2017. Nucleases acting at stalled forks: how to reboot the replication program with a few shortcuts. Annu Rev Genet 51: 477–499. 10.1146/annurev-genet-120116-024745 [DOI] [PubMed] [Google Scholar]
- Patel A, Lee HO, Jawerth L, Maharana S, Jahnel M, Hein MY, Stoynov S, Mahamid J, Saha S, Franzmann TM, et al. 2015. A liquid-to-solid phase transition of the ALS protein FUS accelerated by disease mutation. Cell 162: 1066–1077. 10.1016/j.cell.2015.07.047 [DOI] [PubMed] [Google Scholar]
- Peters JM, Nishiyama T. 2012. Sister chromatid cohesion. Cold Spring Harb Perspect Biol 4: a011130 10.1101/cshperspect.a011130 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pleschke JM, Kleczkowska HE, Strohm M, Althaus FR. 2000. Poly(ADP-ribose) binds to specific domains in DNA damage checkpoint proteins. J Biol Chem 275: 40974–40980. 10.1074/jbc.M006520200 [DOI] [PubMed] [Google Scholar]
- Popuri V, Croteau DL, Brosh RM Jr, Bohr VA. 2012. RECQ1 is required for cellular resistance to replication stress and catalyzes strand exchange on stalled replication fork structures. Cell Cycle 11: 4252–4265. 10.4161/cc.22581 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Potapova TA, Unruh JR, Yu Z, Rancati G, Li H, Stampfer MR, Gerton JL. 2019. Superresolution microscopy reveals linkages between ribosomal DNA on heterologous chromosomes. J Cell Biol 218: 2492–2513. 10.1083/jcb.201810166 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Quinet A, Vindigni A. 2018. Superfast DNA replication causes damage in cancer cells. Nature 559: 186–187. 10.1038/d41586-018-05501-6 [DOI] [PubMed] [Google Scholar]
- Ramamoorthy M, Smith S. 2015. Loss of ATRX suppresses resolution of telomere cohesion to control recombination in ALT cancer cells. Cancer Cell 28: 357–369. 10.1016/j.ccell.2015.08.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ratnakumar K, Duarte LF, LeRoy G, Hasson D, Smeets D, Vardabasso C, Bonisch C, Zeng T, Xiang B, Zhang DY, et al. 2012. ATRX-mediated chromatin association of histone variant macroH2A1 regulates α-globin expression. Genes Dev 26: 433–438. 10.1101/gad.179416.111 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ray Chaudhuri A, Nussenzweig A. 2017. The multifaceted roles of PARP1 in DNA repair and chromatin remodelling. Nat Rev Mol Cell Biol 18: 610–621. 10.1038/nrm.2017.53 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ray Chaudhuri A, Ahuja AK, Herrador R, Lopes M. 2015. Poly(ADP-ribosyl) glycohydrolase prevents the accumulation of unusual replication structures during unperturbed S phase. Mol Cell Biol 35: 856–865. 10.1128/MCB.01077-14 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Riccio AA, Cingolani G, Pascal JM. 2016a. PARP-2 domain requirements for DNA damage-dependent activation and localization to sites of DNA damage. Nucleic Acids Res 44: 1691–1702. 10.1093/nar/gkv1376 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Riccio AA, McCauley M, Langelier MF, Pascal JM. 2016b. 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]
- Rippmann JF, Damm K, 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]
- Ronson GE, Piberger AL, Higgs MR, Olsen AL, Stewart GS, McHugh PJ, Petermann E, Lakin ND. 2018. PARP1 and PARP2 stabilise replication forks at base excision repair intermediates through Fbh1-dependent Rad51 regulation. Nat Commun 9: 746 10.1038/s41467-018-03159-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rulten SL, Cortes-Ledesma F, Guo L, Iles NJ, Caldecott KW. 2008. APLF (C2orf13) is a novel component of poly(ADP-ribose) signaling in mammalian cells. Mol Cell Biol 28: 4620–4628. 10.1128/MCB.02243-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rulten SL, Fisher AE, Robert I, Zuma MC, Rouleau M, Ju L, Poirier G, Reina-San-Martin B, Caldecott KW. 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]
- Rulten SL, Rotheray A, Green RL, Grundy GJ, Moore DA, Gómez-Herreros F, Hafezparast M, Caldecott KW. 2014. PARP-1 dependent recruitment of the amyotrophic lateral sclerosis-associated protein FUS/TLS to sites of oxidative DNA damage. Nucleic Acids Res 42: 307–314. 10.1093/nar/gkt835 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ruscetti T, Lehnert BE, Halbrook J, Le Trong H, Hoekstra MF, Chen DJ, Peterson SR. 1998. Stimulation of the DNA-dependent protein kinase by poly(ADP-ribose) polymerase. J Biol Chem 273: 14461–14467. 10.1074/jbc.273.23.14461 [DOI] [PubMed] [Google Scholar]
- Salim D, Gerton JL. 2019. Ribosomal DNA instability and genome adaptability. Chromosome Res 27: 73–87. 10.1007/s10577-018-9599-7 [DOI] [PubMed] [Google Scholar]
- Satoh MS, Lindahl T. 1992. Role of poly(ADP-ribose) formation in DNA repair. Nature 356: 356–358. 10.1038/356356a0 [DOI] [PubMed] [Google Scholar]
- Sbodio JI, Chi NW. 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]
- Schwartz JC, Wang X, Podell ER, Cech TR. 2013. RNA seeds higher-order assembly of FUS protein. Cell Rep 5: 918–925. 10.1016/j.celrep.2013.11.017 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Seimiya H, Smith S. 2002. The telomeric poly(ADP-ribose) polymerase, tankyrase 1, contains multiple binding sites for telomeric repeat binding factor 1 (TRF1) and a novel acceptor, 182-kDa tankyrase-binding protein (TAB182). J Biol Chem 277: 14116–14126. 10.1074/jbc.M112266200 [DOI] [PubMed] [Google Scholar]
- Shahrour MA, Nicolae CM, Edvardson S, Ashhab M, Galvan AM, Constantin D, Abu-Libdeh B, Moldovan GL, Elpeleg O. 2016. PARP10 deficiency manifests by severe developmental delay and DNA repair defect. Neurogenetics 17: 227–232. 10.1007/s10048-016-0493-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Singatulina AS, Hamon L, Sukhanova MV, Desforges B, Joshi V, Bouhss A, Lavrik OI, Pastré D. 2019. PARP-1 activation directs FUS to DNA damage sites to form PARG-reversible compartments enriched in damaged DNA. Cell Rep 27: 1809–1821.e5. 10.1016/j.celrep.2019.04.031 [DOI] [PubMed] [Google Scholar]
- Slade D. 2019. Mitotic functions of poly(ADP-ribose) polymerases. Biochem Pharmacol 167: 33–43. 10.1016/j.bcp.2019.03.028 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Slade D, Dunstan MS, Barkauskaite E, Weston R, Lafite P, Dixon N, Ahel M, Leys D, Ahel I. 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]
- Smith S, de Lange T. 2000. Tankyrase promotes telomere elongation in human cells. Curr Biol 10: 1299–1302. 10.1016/S0960-9822(00)00752-1 [DOI] [PubMed] [Google Scholar]
- Smith S, Giriat I, Schmitt A, 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]
- Sobinoff AP, Pickett HA. 2017. Alternative lengthening of telomeres: DNA repair pathways converge. Trends Genet 33: 921–932. 10.1016/j.tig.2017.09.003 [DOI] [PubMed] [Google Scholar]
- Spagnolo L, Barbeau J, Curtin NJ, Morris EP, Pearl LH. 2012. Visualization of a DNA-PK/PARP1 complex. Nucleic Acids Res 40: 4168–4177. 10.1093/nar/gkr1231 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sporn JC, Kustatscher G, Hothorn T, Collado M, Serrano M, Muley T, Schnabel P, Ladurner AG. 2009. Histone macroH2A isoforms predict the risk of lung cancer recurrence. Oncogene 28: 3423–3428. 10.1038/onc.2009.26 [DOI] [PubMed] [Google Scholar]
- Stewart JA, Chaiken MF, Wang F, Price CM. 2012. Maintaining the end: roles of telomere proteins in end-protection, telomere replication and length regulation. Mutat Res 730: 12–19. 10.1016/j.mrfmmm.2011.08.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stults DM, Killen MW, Williamson EP, Hourigan JS, Vargas HD, Arnold SM, Moscow JA, Pierce AJ. 2009. Human rRNA gene clusters are recombinational hotspots in cancer. Cancer Res 69: 9096–9104. 10.1158/0008-5472.CAN-09-2680 [DOI] [PubMed] [Google Scholar]
- Sugimura K, Takebayashi S, Taguchi H, Takeda S, Okumura K. 2008. PARP-1 ensures regulation of replication fork progression by homologous recombination on damaged DNA. J Cell Biol 183: 1203–1212. 10.1083/jcb.200806068 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takai H, Smogorzewska A, de lange T. 2003. DNA damage foci at dysfunctional telomeres. Mol Cell 13: 1549-1556. 10.1016/S0960-9822(03)00542-6 [DOI] [PubMed] [Google Scholar]
- Teloni F, 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]
- Thandapani P, O'Connor TR, Bailey TL, Richard S. 2013. Defining the RGG/RG motif. Mol Cell 50: 613–623. 10.1016/j.molcel.2013.05.021 [DOI] [PubMed] [Google Scholar]
- Timinszky G, Till S, Hassa PO, Hothorn M, Kustatscher G, Nijmeijer B, Colombelli J, Altmeyer M, Stelzer EH, Scheffzek K, 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]
- Tripathi E, Smith S. 2017. Cell cycle-regulated ubiquitination of tankyrase 1 by RNF8 and ABRO1/BRCC36 controls the timing of sister telomere resolution. EMBO J 36: 503–519. 10.15252/embj.201695135 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Truong LN, Li Y, Shi LZ, Hwang PY, He J, Wang H, Razavian N, Berns MW, Wu X. 2013. Microhomology-mediated end joining and homologous recombination share the initial end resection step to repair DNA double-strand breaks in mammalian cells. Proc Natl Acad Sci 110: 7720–7725. 10.1073/pnas.1213431110 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Udugama M, Sanij E, Voon HPJ, Son J, Hii L, Henson JD, Chan FL, Chang FTM, Liu Y, Pearson RB, et al. 2018. Ribosomal DNA copy loss and repeat instability in ATRX-mutated cancers. Proc Natl Acad Sci 115: 4737–4742. 10.1073/pnas.1720391115 [DOI] [PMC free article] [PubMed] [Google Scholar]
- van Steensel B, de Lange T. 1997. Control of telomere length by the human telomeric protein TRF1. Nature 385: 740–743. 10.1038/385740a0 [DOI] [PubMed] [Google Scholar]
- Verdun RE, Crabbe L, Haggblom C, Karlseder J. 2005. Functional human telomeres are recognized as DNA damage in G2 of the cell cycle. Mol Cell 20: 551–561. 10.1016/j.molcel.2005.09.024 [DOI] [PubMed] [Google Scholar]
- Vyas S, Matic I, Uchima L, Rood J, Zaja R, Hay RT, Ahel I, Chang P. 2014. Family-wide analysis of poly(ADP-ribose) polymerase activity. Nat Commun 5: 4426 10.1038/ncomms5426 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Waga S, Hannon GJ, Beach D, Stillman B. 1994. The p21 inhibitor of cyclin-dependent kinases controls DNA replication by interaction with PCNA. Nature 369: 574–578. 10.1038/369574a0 [DOI] [PubMed] [Google Scholar]
- Waizenegger IC, Hauf S, Meinke A, Peters JM. 2000. Two distinct pathways remove mammalian cohesin from chromosome arms in prophase and from centromeres in anaphase. Cell 103: 399–410. 10.1016/S0092-8674(00)00132-X [DOI] [PubMed] [Google Scholar]
- Wang ZQ, Stingl L, Morrison C, Jantsch M, Los M, Schulze-Osthoff K, Wagner EF. 1997. PARP is important for genomic stability but dispensable in apoptosis. Genes Dev 11: 2347–2358. 10.1101/gad.11.18.2347 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Z, Michaud GA, Cheng Z, Zhang Y, Hinds TR, Fan E, Cong F, Xu W. 2012. Recognition of the iso-ADP-ribose moiety in poly(ADP-ribose) by WWE domains suggests a general mechanism for poly(ADP-ribosyl)ation-dependent ubiquitination. Genes Dev 26: 235–240. 10.1101/gad.182618.111 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang W, Li N, Li X, Tran MK, Han X, Chen J. 2015. Tankyrase inhibitors target YAP by stabilizing angiomotin family proteins. Cell Rep 13: 524–532. 10.1016/j.celrep.2015.09.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y, Luo W, Wang Y. 2019. PARP-1 and its associated nucleases in DNA damage response. DNA Repair (Amst) 81: 102651 10.1016/j.dnarep.2019.102651 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wei H, Yu X. 2016. Functions of PARylation in DNA damage repair pathways. Genomics Proteomics Bioinformatics 14: 131–139. 10.1016/j.gpb.2016.05.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wright WE, Piatyszek MA, Rainey WR, Byrd W, Shay JW. 1996. Telomerase activity in human germline and embryonic tissues and cells. Dev Genetics 18: 173–179. [DOI] [PubMed] [Google Scholar]
- Yalon M, Gal S, Segev Y, Selig S, Skorecki KL. 2004. Sister chromatid separation at human telomeric regions. J Cell Sci 117: 1961–1970. 10.1242/jcs.01032 [DOI] [PubMed] [Google Scholar]
- Yang L, Sun L, Teng Y, Chen H, Gao Y, Levine AS, Nakajima S, Lan L. 2017. Tankyrase1-mediated poly(ADP-ribosyl)ation of TRF1 maintains cell survival after telomeric DNA damage. Nucleic Acids Res 45: 3906–3921. 10.1093/nar/gkx083 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yeh TY, Beiswenger KK, Li P, Bolin KE, Lee RM, Tsao TS, Murphy AN, Hevener AL, Chi NW. 2009. Hypermetabolism, hyperphagia, and reduced adiposity in tankyrase-deficient mice. Diabetes 58: 2476–2485. 10.2337/db08-1781 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zafar MK, Eoff RL. 2017. Translesion DNA synthesis in cancer: molecular mechanisms and therapeutic opportunities. Chem Res Toxicol 30: 1942–1955. 10.1021/acs.chemrestox.7b00157 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Y, Liu S, Mickanin C, Feng Y, Charlat O, Michaud GA, Schirle M, Shi X, Hild M, Bauer A, et al. 2011. RNF146 is a poly(ADP-ribose)-directed E3 ligase that regulates axin degradation and Wnt signalling. Nat Cell Biol 13: 623–629. 10.1038/ncb2222 [DOI] [PubMed] [Google Scholar]
- Zhang F, Chen Y, Li M, Yu X. 2014. The oligonucleotide/oligosaccharide-binding fold motif is a poly(ADP-ribose)-binding domain that mediates DNA damage response. Proc Natl Acad Sci 111: 7278–7283. 10.1073/pnas.1318367111 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang F, Shi J, Bian C, Yu X. 2015a. Poly(ADP-Ribose) Mediates the BRCA2-dependent early DNA damage response. Cell Rep 13: 678–689. 10.1016/j.celrep.2015.09.040 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang F, Shi J, Chen SH, Bian C, Yu X. 2015b. The PIN domain of EXO1 recognizes poly(ADP-ribose) in DNA damage response. Nucleic Acids Res 43: 10782–10794. 10.1093/nar/gkv939 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhong L, Ding Y, Bandyopadhyay G, Waaler J, Börgeson E, Smith S, Zhang M, Phillips SA, Mahooti S, Mahata SK, et al. 2016. The PARsylation activity of tankyrase in adipose tissue modulates systemic glucose metabolism in mice. Diabetologia 59: 582–591. 10.1007/s00125-015-3815-1 [DOI] [PubMed] [Google Scholar]
- Zhou ZD, Chan CH, Xiao ZC, Tan EK. 2011. Ring finger protein 146/Iduna is a poly(ADP-ribose) polymer binding and PARsylation dependent E3 ubiquitin ligase. Cell Adh Migr 5: 463–471. 10.4161/cam.5.6.18356 [DOI] [PMC free article] [PubMed] [Google Scholar]
