Neuroscience 145 (2007) 1260 –1266
DNA SINGLE-STRAND BREAK REPAIR AND SPINOCEREBELLAR ATAXIA WITH AXONAL NEUROPATHY-1 S. F. EL-KHAMISYa,b AND K. W. CALDECOTTa*
Tyrosyl DNA phosphodiesterase 1 (TDP1) and spinocerebellar ataxia with axonal neuropathy (SCAN1)
a
Genome Damage and Stability Centre, University of Sussex, Science Park Road, Falmer, Brighton BN1 9RQ, UK
b
Biochemistry Department, Faculty of Pharmacy, Ain Shams University, P.O.B. 11566, Cairo, Egypt
Disorders in DNA repair typically result in multiple pathological phenotypes, depending on the nature of the repair defect (de Boer and Hoeijmakers, 2000; O’Driscoll and Jeggo, 2006). For example, in addition to spinocerebellar ataxia, the DNA double-strand break (DSB) response-defective disorders ataxia telangiectasia (AT) and ataxia telangiectasia like disorder (ATLD) exhibit radiosensitivity, immunodeficiency, and chromosomal instability (Stewart et al., 1999; Taylor et al., 2004; O’Driscoll and Jeggo, 2006). In contrast, Takashima et al. (2002) recently identified an autosomal recessive ataxia that they denoted SCAN1, the phenotype of which lacks chromosomal instability and cancer predisposition and appears to be restricted to the nervous system. SCAN1 is associated with a mutation in TDP1, a protein that is primarily involved in the repair of DNA strand breaks created by topoisomerase 1 (Top1). Top1 relaxes superhelical tension in DNA and during its normal catalytic cycle generates a reversible and transient intermediate known as the Top1 cleavage complex, in which Top1 is covalently attached via a tyrosyl residue (human Tyr723) to the 3=-terminus of a single-stranded nick (Wang, 2002). Following release of torsional stress, Top1 reseals the nick and restores the integrity of the double helix (Pommier et al., 2003). The rate of religation is normally much faster than the rate of cleavage and thus the steady state concentration of these intermediates is very low (Pommier et al., 2003; Rideout et al., 2004). However, under certain circumstances, Top1 cleavage complexes may become irreversible, most likely due to misalignment of the 5=-hydroxyl that is no longer able to act as a nucleophile in the religation reaction (Pommier et al., 2003). Consequently, this irreversible “abortive” Top1-associated DNA break requires a DNA repair process for its removal. For example, Top1 cleavage complexes can be converted into abortive Top1-associated DNA DSBs by collision with a replication fork (Avemann et al., 1988; Holm et al., 1989; Hsiang et al., 1989; Ryan et al., 1991; Strumberg et al., 2000). In addition, collision with the transcription machinery converts Top1 cleavage complexes into abortive Top1-associated DNA SSBs, and occasionally abortive Top1-associated DSBs if two Top1 cleavage complexes are closely spaced on opposite strands of DNA (Kroeger and Rowe, 1989; Bendixen et al., 1990; Wu and Liu, 1997). Finally, some mutagenic DNA adducts (Pourquier et al., 1998, 2001; Pommier et al., 2000a) and en-
Abstract—DNA single-strand breaks (SSBs) are the commonest DNA lesions arising spontaneously in cells, and if not repaired may block transcription or may be converted into potentially lethal/clastogenic DNA double-strand breaks (DSBs). Recently, evidence has emerged that defects in the rapid repair of SSBs preferentially impact the nervous system. In particular, spinocerebellar ataxia with axonal neuropathy (SCAN1) is a human disease that is associated with mutation of TDP1 (tyrosyl DNA phosphodiesterase 1) protein and with a defect in repairing certain types of SSBs. Although SCAN1 is a rare neurodegenerative disorder, understanding the molecular basis of this disease will lead to better understanding of neurodegenerative processes. Here we review recent progress in our understanding of TDP1, single-strand break repair (SSBR), and neurodegenerative disease. © 2006 IBRO. Published by Elsevier Ltd. All rights reserved.
DNA is under continuous threat and thousands of DNA single-strand breaks (SSBs) arise in each cell per day (Lindahl, 1993). SSBs can arise either directly (e.g. from attack of deoxyribose by reactive oxygen species) or indirectly (e.g. via enzymatic cleavage of the phosphodiester backbone during DNA base excision repair). Cells have employed different mechanisms to repair DNA SSBs, which are collectively termed DNA single-strand break repair (SSBR) (Thompson and West, 2000; Caldecott, 2004). Defects in the repair of, or response to, DNA damage have been associated with many human disorders such as cancer, immunodeficiency, and neurodegeneration. Recent data suggest that defects in the repair or response to DNA SSBs may have particular impact on the nervous system, and are associated with spinocerebellar ataxia (Date et al., 2001; Moreira et al., 2001; Takashima et al., 2002).
*Corresponding author. Tel: ⫹44-01273-877511; fax: ⫹44-01273678121. E-mail address:
[email protected] (K. Caldecott). Abbreviations: CPT, camptothecin; DSB, double-strand break; DSBR, double-strand break repair; HR, homologous recombination; LCL, lymphoblastoid cell line; SCAN1, spinocerebellar ataxia with axonal neuropathy; SSB, single-strand break; SSBR, single-strand break repair; TDP1, tyrosyl DNA phosphodiesterase 1; Top1, topoisomerase 1.
0306-4522/07$30.00⫹0.00 © 2006 IBRO. Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.neuroscience.2006.08.048
1260
S. F. El-Khamisy and K. W. Caldecott / Neuroscience 145 (2007) 1260 –1266
dogenous DNA lesions such as abasic sites, nicks, gaps, and DNA secondary structures can convert Top1 cleavage complexes into abortive Top1-associated DNA SSBs (Fig. 1) (Lanza et al., 1996; Pourquier et al., 1997a,b; Christiansen and Westergaard, 1999; Pommier et al., 2000b). In 1996, Nash and colleagues (Yang et al., 1996) employed a single-stranded DNA oligonucleotide linked to a tyrosine residue via a 3=-phosphoryl group, mimicking an abortive Top1-associated DNA strand break, to identify an activity in Saccharomyces cerevisiae extract that was able to remove Top1 from the 3=-terminus of DNA. They named the activity Tdp1, and the gene encoding this activity was subsequently cloned (Pouliot et al., 1999). Tdp1 is highly conserved in eukaryotes, with identifiable orthologs in a wide variety of animals, including humans (Pouliot et al., 1999; Cheng et al., 2002; Pommier et al., 2003; Connelly and Leach, 2004; Davies et al., 2004). Sequence analysis of S. cerevisiae Tdp1 (Pouliot et al., 1999; Interthal et al., 2001) and the crystal structure of human TDP1 (Davies et al., 2002, 2003) reveal it
1261
to be a member of the phospholipase D (PLD) superfamily. These enzymes contain two copies of a highly conserved sequence (HXK(X)4D(X)6GSXN), known as HKD motif (Interthal et al., 2001) and catalyze phosphoryl transfer reactions. Despite lacking the aspartate conserved in other HKD motifs, the crystal structure of human TDP1 has shown that both histidine and lysine residues of the HKD motifs are clustered in the active centre of the enzyme (Davies et al., 2002, 2003, 2004). Our initial understanding of the role for TDP1 in DNA repair came from studies conducted in budding yeast. Yeast Tdp1 provides one of several potentially redundant mechanisms for removing Top1 peptide from a DSB created by collision of Top1 cleavage complexes with DNA replication forks (Pouliot et al., 2001; Vance and Wilson, 2002; Deng et al., 2005). Once Top1 peptide has been removed, the DSB is then channeled into the homologous recombination (HR) pathway for double-strand break repair (DSBR) (Pouliot et al., 2001; Vance and Wilson,
Fig. 1. Model for the generation of DNA-strand breaks from Top1 cleavage complexes. (A) Top1 relaxes torsional stress in DNA by introducing a reversible and transient DNA nick in which Top1 is covalently attached to the 3=-end. CPT reversibly inhibits the ligation step of Top1, thereby increasing the half-life of Top1 cleavage complexes. Note that some endogenous DNA lesions can also stabilize Top1 cleavage complexes (e.g. base mismatches) or induce their formation (e.g. 8-oxoguanine). (B) Top1 cleavage complexes can be converted into irreversible DNA DSBs by collision with a replication fork, or DNA SSBs by collision with the transcriptional machinery or by proximity to some types of DNA lesions.
1262
S. F. El-Khamisy and K. W. Caldecott / Neuroscience 145 (2007) 1260 –1266
2002). Tdp1 in yeast appears to operate in conjunction with HR to repair Top1-associated DSBs, because mutation of Tdp1 is epistatic to mutations in critical components of the HR pathway, such as Rad52 (Pouliot et al., 2001; Vance and Wilson, 2002). As in yeast, it is likely that human TDP1 plays a role during DNA replication. The cytotoxic effect of the Top1 specific inhibitor camptothecin (CPT) is S-phase specific in most wild-type mammalian cell lines examined (Horwitz and Horwitz, 1973; D’Arpa et al., 1990; Ryan et al., 1994; Poot et al., 1999), and is ablated by co-incubation with aphidicolin, an inhibitor of DNA replication (Holm et al., 1989; Ryan et al., 1991). Thus, ongoing replication forks appear to be required to convert Top1 cleavage complexes into cytotoxic DSBs, in wild type proliferating cells at least (Shao et al., 1999; Arnaudeau et al., 2001; Culmsee et al., 2001; Godthelp et al., 2002; Wu et al., 2002; Furuta et al., 2003; Hinz et al., 2003; van Waardenburg et al., 2004). It is therefore plausible that TDP1 inhibitors may enhance the anti-cancer activity of Top1 inhibitors and act as anti-proliferative agents (Liao et al., 2006). Consistent with a possible role for human TDP1 during replication, lymphoblastoid cell lines (LCLs) from affected SCAN1 individuals appear to show more late S-phase arrest than normal cells following CPT treatment (Interthal et al., 2005b). This could reflect a direct role for TDP1 during replication or an accumulation of un-repaired SSBs that are consequently converted into DSBs by collision with replication forks. Despite its pivotal role in the repair of certain types of replication-dependent DSBs it seems unlikely that this can account for the pathology of SCAN1, which appears to be restricted to post-mitotic neurons and lacks any obvious sign of genetic instability. It is possible that SCAN1 results from a defect in the repair of replication-independent DSBs, which might arise from overlapping Top1-SSBs for example. However, since (like other SSBs) the occurrence of Top1-SSBs is most likely several orders of magnitude more frequent than Top1-DSBs, and since SSBs can block transcription (Kroeger and Rowe, 1989; Bendixen et al., 1990; Zhou and Doetsch, 1994; Wu and Liu, 1997), it seems plausible that a defect in the repair of SSBs might account for the neurodegenerative pathology in SCAN1. Consistent with this idea, whereas we have failed to observe a measurable defect in the repair of replicationindependent DSBs in CPT-treated SCAN1 LCLs, we have observed a profound defect in their ability to repair replication-independent SSBs. Intriguingly, SCAN1 cells were also less able than normal cells to repair oxidative DNA SSBs induced by H2O2, though the defect was not as pronounced as for CPT (El-Khamisy et al., 2005). This could reflect a direct requirement for TDP1 in processing oxidative DNA breaks, since it has been shown that TDP1 can remove glycolate from 3=-phosphoglycolate termini of strand breaks (Inamdar et al., 2002) and a variety of other 3=-adducts from DNA (Interthal et al., 2005a). In addition, yeast expressing an active site mutant of Tdp1 (His182Ala) is hypersensitive to bleomycin, a radiomimetic agent which generates DNA breaks that are almost exclusively terminated with 3=-phosphoglycolate (Liu et al., 2004). More-
over, in vitro assays have recently shown that SCAN1 cell extracts lack any detectable processing of 3=-phosphoglycolate at single-stranded oligomers or at 3=-overhangs of DSBs (Zhou et al., 2005). However, similar assays with substrates that mimic oxidative SSBs suggest that APE1 is the major 3=-phosphoglycolate-processing activity in human cell extracts (Parsons et al., 2004). This suggests that TDP1 may be a major contributor in removing 3=-phosphoglycolate at sites of DSBs, but not at SSBs. An alternative explanation for the role of TDP1-dependent SSBR at oxidative SSBs would be an increased formation of Top1-associated SSBs following treatment with H2O2. Indeed, the presence of nicks (Pourquier et al., 1997a), abasic sites (Pourquier et al., 1997b), modified bases (Pourquier et al., 1999; Lesher et al., 2002), and modified sugars (Chrencik et al., 2003) has been shown to stabilize and/or induce Top1 cleavage complexes. Consistent with this, Top1-deficient P388/CPT45 murine leukemia cells are more resistant to H2O2 than Top1-proficient parental cells (Daroui et al., 2004). In addition, over-expression of Top1 in yeast confers sensitivity to H2O2 and top1-mutant strains are more resistant to H2O2 than their isogenic wild-type strains (Daroui et al., 2004). Finally, studies employing HeLa cells have confirmed that H2O2 does indeed induce Top1-DNA lesions (Daroui et al., 2004). Yeast two-hybrid and coimmunoprecipitation experiments have provided evidence for a direct interaction between TDP1 and DNA ligase III␣, a component of the DNA SSBR machinery (El-Khamisy et al., 2005). In agreement with this, it has been shown that TDP1 co-immunoprecipitates with XRCC1 from rodent cell extract (Plo et al., 2003). The interaction between TDP1 and DNA ligase III␣ is mediated by the N-terminal domain of TDP1, a region that is poorly conserved or absent from Tdp1 of lower eukaryotes (Interthal et al., 2001; El-Khamisy et al., 2005). Budding yeasts lack orthologs of various components of SSBR machinery found in mammals, including XRCC1 and DNA ligase III␣. Therefore it seems likely that the N-terminal domain of human TDP1 may have evolved to couple this enzyme to the SSBR machinery, through an interaction with DNA ligase III␣. Current models of SSBR suggest sequential recruitment of the enzymes such that DNA repair-intermediates are passed from one enzyme to the other in a molecular relay (Mol et al., 2000; Wilson and Kunkel, 2000). However, at least some of the XRCC1 appears to be present in pre-formed multi-protein complexes (Vidal et al., 2001; Whitehouse et al., 2001; Caldecott, 2003; Clements et al., 2004), though it is unclear whether these interactions are occurring in response to “endogenous” levels of DNA damage or are truly constitutive. It should be noted that although we have not yet detected a measurable defect in the repair of replicationindependent DSBs in SCAN1 cells following CPT or oxidative stress, it is possible and indeed perhaps likely that one exists. Nevertheless, the observations that SSBs can be potent blocks to transcription and that they arise several orders of magnitude more frequently than DSBs renders
S. F. El-Khamisy and K. W. Caldecott / Neuroscience 145 (2007) 1260 –1266
these lesions a likely threat to cellular function, if they are not repaired rapidly. SCAN1 and neurodegeneration It is currently unclear why loss of TDP1 does not result in increased genetic instability and cancer. However, a likely contributing factor is the ability of alternative end processing proteins such as ERCC1/XPF, in conjunction with HR repair, to compensate for defective TDP1-SSBR by removing un-repaired SSBs during DNA replication. Consistent with this, HR is selectively elevated in SCAN1 cells as indicated by an increased frequency of “spontaneous” and “CPT-induced” sister chromatid exchanges (El-Khamisy et al., 2005). Thus, it seems possible that the concerted activity of structure specific nucleases and the HR machinery can fulfill a backup role in the absence of TDP1-
1263
dependent SSBR, thereby allowing proliferating SCAN1 cells to tolerate un-repaired SSBs, at physiological levels of DNA damage at least. This scheme is also applicable to other types of un-repaired SSBs, since structure specific nucleases may enable removal of a wide variety of damaged 5= and 3=-termini (Fig. 2). It is also currently unclear why the nervous system might be so sensitive to loss of TDP1, compared with other largely non-cycling tissues for example. One likely factor is that the nervous system consumes ⬃20% of inhaled oxygen and possesses low levels of antioxidant enzymes (Barzilai et al., 2002), and so encounters high levels of oxidative stress. Consequently, since TDP1 may be required for the repair of SSBs and DSBs arising either directly or indirectly through oxidative damage, the nervous system may be particularly dependent on TDP1. In
Fig. 2. Model for the impact of SSBs on proliferating and non-proliferating cells in the absence of SSBR. In both non-cycling and cycling cells SSBs are normally rapidly repaired by SSBR. However, in SSBR-defective non-cycling cells (left), SSBs persist and block transcription leading to cell death. Compared with other largely non-cycling tissues, we propose that the elevated oxidative stress, increased transcriptional demand, and limited cellular regenerative capacity in the nervous system render this tissue particularly dependent upon SSBR. In cycling cells lacking SSBR (right) un-repaired SSBs are converted into DSBs by collision with DNA replication forks (red arrow). Structure-specific 5= and 3=-nucleases can then process the damaged 5=-and 3=-termini (yellow ovals) enabling the DSB to be faithfully repaired by HR. For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.
1264
S. F. El-Khamisy and K. W. Caldecott / Neuroscience 145 (2007) 1260 –1266
addition, there is a high transcriptional demand in postmitotic neurons. For example, cortical neurons incorporate ⬃5.5-fold more 3H-uridine than astrocytes (Morris and Geller, 1996) and possess twice the number of transcription initiation sites as compared with glial cells (Flangas and Bowman, 1970; Sarkander and Uthoff, 1976; Sarkander and Dulce, 1978), and un-repaired SSBs can be potent blocks to RNA polymerases, which may subsequently trigger cell death. The limited regenerative capacity of the nervous system, compared with other predominantly non-cycling tissues, may also be a factor, and may render this tissue particularly susceptible to cell loss. In contrast, cell loss from tissues with greater regenerative capacity may be better tolerated. For example, myocytes and adipocytes are also terminally differentiated, but can be replaced by precursor cells (Vierck et al., 2000; Yan, 2000; Nouspikel and Hanawalt, 2002).
CONCLUSIONS Whereas great insights into the relationship between TDP1, DNA strand break repair, and neurodegeneration have emerged, there is still much that is unresolved. For example, what are the physiologically relevant substrates fro Tdp1 in vivo. In addition to 3=-termini created by Top1 and oxidative damage, recent data suggest that Tdp1 may also operate on certain types of 5=-termini (Nitiss et al., 2006), increasing the types of lesion that may impact on neurological integrity in SCAN1 individuals. Also, why does loss of Tdp1 selectively affect certain neuronal tissue such as the cerebellum, and what is the relative contribution of SSBs versus DSBs in neurodegeneration? The development of animal models with specific defects in SSBR and DSBR will be an important tool to address these questions. Acknowledgments—We apologize to colleagues whose primary research papers may not have been cited due to space constraints. This work was supported by an MRC programme grant to K.W.C. and an ORS to S.F.E-K.
REFERENCES Arnaudeau C, Lundin C, Helleday T (2001) DNA double-strand breaks associated with replication forks are predominantly repaired by homologous recombination involving an exchange mechanism in mammalian cells. J Mol Biol 307:1235–1245. Avemann K, Knippers R, Koller T, Sogo JM (1988) Camptothecin, a specific inhibitor of type I DNA topoisomerase, induces DNA breakage at replication forks. Mol Cell Biol 8:3026 –3034. Barzilai A, Rotman G, Shiloh Y (2002) ATM deficiency and oxidative stress: a new dimension of defective response to DNA damage. DNA Repair (Amst) 1:3–25. Bendixen C, Thomsen B, Alsner J, Westergaard O (1990) Camptothecin-stabilized topoisomerase I-DNA adducts cause premature termination of transcription. Biochemistry 29:5613–5619. Caldecott KW (2003) Protein-protein interactions during mammalian DNA single-strand break repair. Biochem Soc Trans 31:247–251. Caldecott KW (2004) DNA single-strand breaks and neurodegeneration. DNA Repair (Amst) 3:875– 882. Cheng TJ, Rey PG, Poon T, Kan CC (2002) Kinetic studies of human tyrosyl-DNA phosphodiesterase, an enzyme in the topoisomerase I DNA repair pathway. Eur J Biochem 269:3697–3704.
Chrencik JE, Burgin AB, Pommier Y, Stewart L, Redinbo MR (2003) Structural impact of the leukemia drug 1-beta-D-arabinofuranosylcytosine (Ara-C) on the covalent human topoisomerase I-DNA complex. J Biol Chem 278:12461–12466. Christiansen K, Westergaard O (1999) Mapping of eukaryotic DNA topoisomerase I catalyzed cleavage without concomitant religation in the vicinity of DNA structural anomalies. Biochim Biophys Acta 1489:249 –262. Clements PM, Breslin C, Deeks ED, Byrd PJ, Ju L, Bieganowski P, Brenner C, Moreira MC, Taylor AM, Caldecott KW (2004) The ataxia-oculomotor apraxia 1 gene product has a role distinct from ATM and interacts with the DNA strand break repair proteins XRCC1 and XRCC4. DNA Repair (Amst) 3:1493–1502. Connelly JC, Leach DR (2004) Repair of DNA covalently linked to protein. Mol Cell 13:307–316. Culmsee C, Bondada S, Mattson MP (2001) Hippocampal neurons of mice deficient in DNA-dependent protein kinase exhibit increased vulnerability to DNA damage, oxidative stress and excitotoxicity. Brain Res Mol Brain Res 87:257–262. D’Arpa P, Beardmore C, Liu LF (1990) Involvement of nucleic acid synthesis in cell killing mechanisms of topoisomerase poisons. Cancer Res 50:6919 – 6924. Daroui P, Desai SD, Li TK, Liu AA, Liu LF (2004) Hydrogen peroxide induces topoisomerase I-mediated DNA damage and cell death. J Biol Chem 279:14587–14594. Date H, Onodera O, Tanaka H, Iwabuchi K, Uekawa K, Igarashi S, Koike R, Hiroi T, Yuasa T, Awaya Y, Sakai T, Takahashi T, Nagatomo H, Sekijima Y, Kawachi I, Takiyama Y, Nishizawa M, Fukuhara N, Saito K, Sugano S, Tsuji S (2001) Early-onset ataxia with ocular motor apraxia and hypoalbuminemia is caused by mutations in a new HIT superfamily gene. Nat Genet 29:184 –188. Davies DR, Interthal H, Champoux JJ, Hol WG (2002) The crystal structure of human tyrosyl-DNA phosphodiesterase, Tdp1. Structure (Camb) 10:237–248. Davies DR, Interthal H, Champoux JJ, Hol WG (2003) Crystal structure of a transition state mimic for Tdp1 assembled from vanadate, DNA, and a topoisomerase I-derived peptide. Chem Biol 10:139–147. Davies DR, Interthal H, Champoux JJ, Hol WG (2004) Explorations of peptide and oligonucleotide binding sites of tyrosyl-DNA phosphodiesterase using vanadate complexes. J Med Chem 47:829 – 837. de Boer J, Hoeijmakers JH (2000) Nucleotide excision repair and human syndromes. Carcinogenesis 21:453– 460. Deng C, Brown JA, You D, Brown JM (2005) Multiple endonucleases function to repair covalent topoisomerase I complexes in Saccharomyces cerevisiae. Genetics 170:591– 600. El-Khamisy SF, Saifi GM, Weinfeld M, Johansson F, Helleday T, Lupski JR, Caldecott KW (2005) Defective DNA single-strand break repair in spinocerebellar ataxia with axonal neuropathy-1. Nature 434:108 –113. Flangas AL, Bowman RE (1970) Differential metabolism of RNA in neuronal-enriched and glial-enriched fractions of rat cerebrum. J Neurochem 17:1237–1245. Furuta T, Takemura H, Liao ZY, Aune GJ, Redon C, Sedelnikova OA, Pilch DR, Rogakou EP, Celeste A, Chen HT, Nussenzweig A, Aladjem MI, Bonner WM, Pommier Y (2003) Phosphorylation of histone H2AX and activation of Mre11, Rad50, and Nbs1 in response to replication-dependent DNA double-strand breaks induced by mammalian DNA topoisomerase I cleavage complexes. J Biol Chem 278:20303–20312. Godthelp BC, Wiegant WW, van Duijn-Goedhart A, Scharer OD, van Buul PP, Kanaar R, Zdzienicka MZ (2002) Mammalian Rad51C contributes to DNA cross-link resistance, sister chromatid cohesion and genomic stability. Nucleic Acids Res 30:2172–2182. Hinz JM, Helleday T, Meuth M (2003) Reduced apoptotic response to camptothecin in CHO cells deficient in XRCC3. Carcinogenesis 24:249 –253. Holm C, Covey JM, Kerrigan D, Pommier Y (1989) Differential requirement of DNA replication for the cytotoxicity of DNA topoisomerase
S. F. El-Khamisy and K. W. Caldecott / Neuroscience 145 (2007) 1260 –1266 I and II inhibitors in Chinese hamster DC3F cells. Cancer Res 49:6365– 6368. Horwitz SB, Horwitz MS (1973) Effects of camptothecin on the breakage and repair of DNA during the cell cycle. Cancer Res 33:2834–2836. Hsiang YH, Lihou MG, Liu LF (1989) Arrest of replication forks by drugstabilized topoisomerase I-DNA cleavable complexes as a mechanism of cell killing by camptothecin. Cancer Res 49:5077–5082. Inamdar KV, Pouliot JJ, Zhou T, Lees-Miller SP, Rasouli-Nia A, Povirk LF (2002) Conversion of phosphoglycolate to phosphate termini on 3= overhangs of DNA double strand breaks by the human tyrosylDNA phosphodiesterase hTdp1. J Biol Chem 277:27162–27168. Interthal H, Pouliot JJ, Champoux JJ (2001) The tyrosyl-DNA phosphodiesterase Tdp1 is a member of the phospholipase D superfamily. Proc Natl Acad Sci U S A 98:12009 –12014. Interthal H, Chen HJ, Champoux JJ (2005a) Human Tdp1 cleaves a broad spectrum of substrates, including phosphoamide linkages. J Biol Chem 280:36518 –36528. Interthal H, Chen HJ, Kehl-Fie TE, Zotzmann J, Leppard JB, Champoux JJ (2005b) SCAN1 mutant Tdp1 accumulates the enzymeDNA intermediate and causes camptothecin hypersensitivity. EMBO J 24:2224 –2233. Kroeger PE, Rowe TC (1989) Interaction of topoisomerase 1 with the transcribed region of the Drosophila HSP 70 heat shock gene. Nucleic Acids Res 17:8495– 8509. Lanza A, Tornaletti S, Rodolfo C, Scanavini MC, Pedrini AM (1996) Human DNA topoisomerase I-mediated cleavages stimulated by ultraviolet light-induced DNA damage. J Biol Chem 271:6978–6986. Lesher DT, Pommier Y, Stewart L, Redinbo MR (2002) 8-Oxoguanine rearranges the active site of human topoisomerase I. Proc Natl Acad Sci U S A 99:12102–12107. Liao Z, Thibaut L, Jobson A, Pommier Y (2006) Inhibition of human tyrosyl-DNA phosphodiesterase (Tdp1) by aminoglycoside antibiotics and ribosome inhibitors. Mol Pharmacol 70:366 –372. Lindahl T (1993) Instability and decay of the primary structure of DNA. Nature 362:709 –715. Liu C, Pouliot JJ, Nash HA (2004) The role of TDP1 from budding yeast in the repair of DNA damage. DNA Repair (Amst) 3:593–601. Mol CD, Hosfield DJ, Tainer JA (2000) Abasic site recognition by two apurinic/apyrimidinic endonuclease families in DNA base excision repair: the 3= ends justify the means. Mutat Res 460:211–229. Moreira MC, Barbot C, Tachi N, Kozuka N, Mendonca P, Barros J, Coutinho P, Sequeiros J, Koenig M (2001) Homozygosity mapping of Portuguese and Japanese forms of ataxia-oculomotor apraxia to 9p13, and evidence for genetic heterogeneity. Am J Hum Genet 68:501–508. Morris EJ, Geller HM (1996) Induction of neuronal apoptosis by camptothecin, an inhibitor of DNA topoisomerase-I: evidence for cell cycle-independent toxicity. J Cell Biol 134:757–770. Nitiss KC, Malik M, He X, White SW, Nitiss JH. (2006) Proc Natl Acad Sci U S A 103:8953– 8958. Nouspikel T, Hanawalt PC (2002) DNA repair in terminally differentiated cells. DNA Repair (Amst) 1:59 –75. O’Driscoll M, Jeggo PA (2006) The role of double-strand break repair: insights from human genetics. Nat Rev Genet 7:45–54. Parsons JL, Dianova II, Dianov GL (2004) APE1 is the major 3=phosphoglycolate activity in human cell extracts. Nucleic Acids Res 32:3531–3536. Plo I, Liao ZY, Barcelo JM, Kohlhagen G, Caldecott KW, Weinfeld M, Pommier Y (2003) Association of XRCC1 and tyrosyl DNA phosphodiesterase (Tdp1) for the repair of topoisomerase I-mediated DNA lesions. DNA Repair (Amst) 2:1087–1100. Pommier Y, Kohlhagen G, Pourquier P, Sayer JM, Kroth H, Jerina DM (2000a) Benzo[a]pyrene diol epoxide adducts in DNA are potent suppressors of a normal topoisomerase I cleavage site and powerful inducers of other topoisomerase I cleavages. Proc Natl Acad Sci U S A 97:2040 –2045. Pommier Y, Laco GS, Kohlhagen G, Sayer JM, Kroth H, Jerina DM (2000b) Position-specific trapping of topoisomerase I-DNA cleav-
1265
age complexes by intercalated benzo[a]-pyrene diol epoxide adducts at the 6-amino group of adenine. Proc Natl Acad Sci U S A 97:10739–10744. Pommier Y, Redon C, Rao VA, Seiler JA, Sordet O, Takemura H, Antony S, Meng L, Liao Z, Kohlhagen G, Zhang H, Kohn KW (2003) Repair of and checkpoint response to topoisomerase I-mediated DNA damage. Mutat Res 532:173–203. Poot M, Gollahon KA, Rabinovitch PS (1999) Werner syndrome lymphoblastoid cells are sensitive to camptothecin-induced apoptosis in S-phase. Hum Genet 104:10 –14. Pouliot JJ, Robertson CA, Nash HA (2001) Pathways for repair of topoisomerase I covalent complexes in Saccharomyces cerevisiae. Genes Cells 6:677– 687. Pouliot JJ, Yao KC, Robertson CA, Nash HA (1999) Yeast gene for a Tyr-DNA phosphodiesterase that repairs topoisomerase I complexes. Science 286:552–555. Pourquier P, Bjornsti MA, Pommier Y (1998) Induction of topoisomerase I cleavage complexes by the vinyl chloride adduct 1,N6-ethenoadenine. J Biol Chem 273:27245–27249. Pourquier P, Pilon AA, Kohlhagen G, Mazumder A, Sharma A, Pommier Y (1997a) Trapping of mammalian topoisomerase I and recombinations induced by damaged DNA containing nicks or gaps. Importance of DNA end phosphorylation and camptothecin effects. J Biol Chem 272:26441–26447. Pourquier P, Ueng LM, Kohlhagen G, Mazumder A, Gupta M, Kohn KW, Pommier Y (1997b) Effects of uracil incorporation, DNA mismatches, and abasic sites on cleavage and religation activities of mammalian topoisomerase I. J Biol Chem 272:7792–7796. Pourquier P, Waltman JL, Urasaki Y, Loktionova NA, Pegg AE, Nitiss JL, Pommier Y (2001) Topoisomerase I-mediated cytotoxicity of N-methyl-N=-nitro-N-nitrosoguanidine: trapping of topoisomerase I by the O6-methylguanine. Cancer Res 61:53–58. Pourquier P, Ueng LM, Fertala J, Wang D, Park HJ, Essigmann JM, Bjornsti MA, Pommier Y (1999) Induction of reversible complexes between eukaryotic DNA topoisomerase I and DNA-containing oxidative base damages 7,8-dihydro-8-oxoguanine and 5-hydroxycytosine. J Biol Chem 274:8516 – 8523. Rideout MC, Raymond AC, Burgin AB Jr (2004) Design and synthesis of fluorescent substrates for human tyrosyl-DNA phosphodiesterase I. Nucleic Acids Res 32:4657– 4664. Ryan AJ, Squires S, Strutt HL, Johnson RT (1991) Camptothecin cytotoxicity in mammalian cells is associated with the induction of persistent double strand breaks in replicating DNA. Nucleic Acids Res 19:3295–3300. Ryan AJ, Squires S, Strutt HL, Evans A, Johnson RT (1994) Different fates of camptothecin-induced replication fork-associated doublestrand DNA breaks in mammalian cells. Carcinogenesis 15:823–828. Sarkander HI, Uthoff CG (1976) Comparison of the number of RNA initiation sites in rat brain fractions enriched in neuronal or glial nuclei. FEBS Lett 72:53–56. Sarkander HI, Dulce HJ (1978) Studies on the regulation of RNA synthesis in neuronal and glial nuclei isolated from rat brain. Exp Brain Res 31:317–327. Shao RG, Cao CX, Zhang H, Kohn KW, Wold MS, Pommier Y (1999) Replication-mediated DNA damage by camptothecin induces phosphorylation of RPA by DNA-dependent protein kinase and dissociates RPA:DNA-PK complexes. EMBO J 18:1397–1406. Stewart GS, Maser RS, Stankovic T, Bressan DA, Kaplan MI, Jaspers NG, Raams A, Byrd PJ, Petrini JH, Taylor AM (1999) The DNA double-strand break repair gene hMRE11 is mutated in individuals with an ataxia-telangiectasia-like disorder. Cell 99:577–587. Strumberg D, Pilon AA, Smith M, Hickey R, Malkas L, Pommier Y (2000) Conversion of topoisomerase I cleavage complexes on the leading strand of ribosomal DNA into 5=-phosphorylated DNA double-strand breaks by replication runoff. Mol Cell Biol 20:3977–3987. Takashima H, Boerkoel CF, John J, Saifi GM, Salih MA, Armstrong D, Mao Y, Quiocho FA, Roa BB, Nakagawa M, Stockton DW, Lupski JR (2002) Mutation of TDP1, encoding a topoisomerase I-depen-
1266
S. F. El-Khamisy and K. W. Caldecott / Neuroscience 145 (2007) 1260 –1266
dent DNA damage repair enzyme, in spinocerebellar ataxia with axonal neuropathy. Nat Genet 32:267–272. Taylor AM, Groom A, Byrd PJ (2004) Ataxia-telangiectasia-like disorder (ATLD): its clinical presentation and molecular basis. DNA Repair (Amst) 3:1219 –1225. Thompson LH, West MG (2000) XRCC1 keeps DNA from getting stranded. Mutat Res 459:1–18. van Waardenburg RC, de Jong LA, van Delft F, van Eijndhoven MA, Bohlander M, Bjornsti MA, Brouwer J, Schellens JH (2004) Homologous recombination is a highly conserved determinant of the synergistic cytotoxicity between cisplatin and DNA topoisomerase I poisons. Mol Cancer Ther 3:393– 402. Vance JR, Wilson TE (2002) Yeast Tdp1 and Rad1-Rad10 function as redundant pathways for repairing Top1 replicative damage. Proc Natl Acad Sci U S A 99:13669 –13674. Vidal AE, Boiteux S, Hickson ID, Radicella JP (2001) XRCC1 coordinates the initial and late stages of DNA abasic site repair through protein-protein interactions. EMBO J 20:6530 – 6539. Vierck J, O’Reilly B, Hossner K, Antonio J, Byrne K, Bucci L, Dodson M (2000) Satellite cell regulation following myotrauma caused by resistance exercise. Cell Biol Int 24:263–272. Wang JC (2002) Cellular roles of DNA topoisomerases: a molecular perspective. Nat Rev Mol Cell Biol 3:430 – 440. Whitehouse CJ, Taylor RM, Thistlethwaite A, Zhang H, Karimi-Busheri F, Lasko DD, Weinfeld M, Caldecott KW (2001) XRCC1 stimulates
human polynucleotide kinase activity at damaged DNA termini and accelerates DNA single-strand break repair. Cell 104:107–117. Wilson SH, Kunkel TA (2000) Passing the baton in base excision repair. Nat Struct Biol 7:176 –178. Wu J, Liu LF (1997) Processing of topoisomerase I cleavable complexes into DNA damage by transcription. Nucleic Acids Res 25:4181–4186. Wu J, Yin MB, Hapke G, Toth K, Rustum YM (2002) Induction of biphasic DNA double strand breaks and activation of multiple repair protein complexes by DNA topoisomerase I drug 7-ethyl-10hydroxy-camptothecin. Mol Pharmacol 61:742–748. Yan Z (2000) Skeletal muscle adaptation and cell cycle regulation. Exerc Sport Sci Rev 28:24 –26. Yang SW, Burgin AB Jr, Huizenga BN, Robertson CA, Yao KC, Nash HA (1996) A eukaryotic enzyme that can disjoin dead-end covalent complexes between DNA and type I topoisomerases. Proc Natl Acad Sci U S A 93:11534 –11539. Zhou T, Lee JW, Tatavarthi H, Lupski JR, Valerie K, Povirk LF (2005) Deficiency in 3=-phosphoglycolate processing in human cells with a hereditary mutation in tyrosyl-DNA phosphodiesterase (TDP1). Nucleic Acids Res 33:289 –297. Zhou W, Doetsch PW (1994) Transcription bypass or blockage at single-strand breaks on the DNA template strand: effect of different 3= and 5= flanking groups on the T7 RNA polymerase elongation complex. Biochemistry 33:14926 –14934.
(Accepted 7 August 2006) (Available online 11 October 2006)