DNA repair homepageWhat is DNA?  (intro / review)UV and skin cancerNucleotide excision repairBase excision repairDSB repairMismatch repair
 
 



NER in eukaryotic cells

 
      Although the general mechanisms of prokaryotic NER are reflected in the eukaryotic model, the processes are considerably more complex.  As in prokaryotes, eukaryotic NER can be subdivided into two overlapping pathways: a global genome repair (GGR) pathway which acts to remove damage from throughout the entire genome, regardless of transcription status, and a transcription-coupled repair (TCR) pathway which specifically targets transcription-blocking lesions and results in the preferential repair of the transcribed strand of active genes.  Interestingly, the relative contribution of these two pathways varies between different lesions. For example, (highly helix distorting) 6-4PPs are repaired so rapidly by GGR, that a TCR component of their repair can only be observed in GGR-defective cells [1].  In contrast, CPDs (which yield a relatively mild helical disruption) are repaired much more slowly by GGR in human cells and consequently a significant TCR component is detectable [2-4].  The contribution of TCR to repair of CPDs is even more pronounced in rodent cells, where GGR of these lesions is essentially non-existent [5].

 
 
  

 
 
 
  
Identification of human DNA repair genes
 

      Human genes (and the corresponding proteins) have been identified by their ability to correct the repair defects in cells from individuals with NER deficient syndromes [6, 7].  Other human repair genes were identified by the ability to correct repair deficient chinese hamster ovary (CHO) cell lines [8, 9], giving rise to the excision repair cross complementing (ERCC) group of genes.  Many of the ERCC genes were subsequently found to be identical to the XP and CS genes.  Some genes were also isolated by homology to previously identified rad group of yeast repair genes, which was facilitated by the high degree of conservation of eukaryotic NER genes (Table III).

 

 
  

 
 
  
Lesion recognition
 

      Of the identified human NER proteins, at least four have been implicated the process of damage recognition.  These proteins appear to segregate into two complexes: 

- The XPA [10, 11], XPE [12-14] and RPA [15-17] proteins can each bind damaged DNA independently, but interactions between them further increase this affinity [17-20]. 

- The second complex with a strong affinity for damaged DNA [21, 22] consists of the XPC protein in a tight association with a human homologue of the yeast rad 23 (HHR23B) protein.

      Like the prokaryotic model, it is likely that lesion recognition involves multiple steps, with the requirement for  identification of damage by one complex to be verified by the other before the repair complex is assembled.  However there is controversy regarding the order in which the NER proteins are assembled on UV photoproducts.   It is even possible that the order of assembly may vary depending on the specific nature of the lesion (see figure below). 

 The XPC protein has an affinity for both single-stranded [21, 23] and double-stranded DNA [21], and can bind a variety of DNA damage [22].  XPC exists in a tight complex with HHR23B [24] and is the first NER protein to recognise NA-AAF induced DNA damage [22].  XPC/HHR23B also associates with TFIIH [25].  Both XPC and TFIIH are required to form an open complex around a cis-diammine-dichloro-platinum (cis-platin) adduct allowing the subsequent addition of XP-A and RPA [26].  The formation of an open complex is presumably mediated by the helicase activities of the XPB and XPD components of TFIIH [25, 27-30], which then recruits XPA through a direct interaction [31].  Interestingly, the yeast homologue of HHR23B (Rad23) has been demonstrated to increase the interaction between TFIIH and Rad 14 (the yeast homologue of XPA) [32], although the effect of HHR23B on human NER remains controversial [21, 23].  The addition of XPA and RPA (through its interaction with XPA [17, 19, 33]) stabilises XPC/HHR23B and TFIIH on the damaged DNA to form pre-incision complex 1 (PIC1) [34, 35]. 

     Although the XPC/HHR23B complex appears to be the first to recognise several different forms of DNA damage [22, 26], it has been proposed that a second complex comprised of XPA, RPA, XPE and XPF/ERCC1 plays a significant role in the recognition of 6-4 PPs.  The XPA, XPE and RPA proteins all have a very strong, specific affinity for 6-4PPs [10, 12, 14, 15].  Whereas, neither XPA nor RPA nor a combination of the two is capable of discriminating between CPD containing DNA and undamaged DNA  [34].  A strong interaction between XPA and ERCC1 [36-39] further increases the affinity of XPA for UV damaged DNA [39], and cells from patients with mutations in the XPF gene exhibit deficient 6-4 PP repair but normal CPD repair [40-43] (see also Dual roles for ERCC1/XPF ?).  Both XPA [17, 19, 33] and XPE [18] directly interact with RPA, and these interactions further increase the binding affinity of these proteins for UV-damaged DNA [17-19].  These data suggest the existence of a large multi-subunit complex with a strong, specific affinity for 6-4PP.

     RPA is a ssDNA binding protein, which binds to the undamaged strand during NER [44] and can unwind short dsDNA substrates [45-47].  As 6-4 PPs strongly favour disruption of the DNA double helix [48], this distortion may be sufficient for RPA to further unwind the DNA allowing lesion identification by XPA (and/or the remainder of the 6-4 PP recognition complex).  Although XPA and TFIIH do not interact well in the absence of DNA damage, the binding of XPA to damaged DNA results in a conformational change which greatly increases its affinity for TFIIH [31, 49].  Recruitment of TFIIH is expected to also bring in XPC/HHR23, which appears to be necessary to promote the stable binding of XPA on UV irradiated DNA [34, 35, 50].  At the same time, it is likely that ERCC1/XPF disassociates from XPA, as it is not associated with the first stable protein complex on UV-damaged DNA [34, 35].  Consistent with this model, a kinetic analysis has shown that a complex of XPA and RPA binds a 6-4 PP adduct prior to XPC/HHR23B [51].  As XPE is not required for NER in vitro [52], its association with the stable recognition complex has not been examined.
 
 



 
  

 
  
Assembly of Pre-incision (PIC) complexes
 

     The first stable complex on damaged DNA (pre-incision complex 1 or PIC 1) consists of XPA, RPA, XPC/HHR23B and TFIIH [34, 35] (see figure above).  RPA binding to ssDNA protects ~30 nucleotides [53-55] and appears to define the size of the excision fragment in NER [44, 56]. 

     Addition of the 3' endonuclease [57], XPG, results in the concomitant loss of XPC from the complex and yields in PIC2 [35].  As RPA binds the undamaged DNA strand with a defined polarity [44], its interaction with XPG [17], specifically positions XPG at the 3' end of the fragment to be excised [44, 58].  However, an interaction between XPG and TFIIH [59], may also play a role in its recruitment to the repair complex.

     Finally, the XPF/ERCC1 complex is recruited to form PIC3 [35].  A direct interaction between RPA and XPF [60] positions XPF/ERCC1 at the 5' end of the fragment to be excised [44, 58].  The strong XPA-ERCC1 interaction does not appear to play a significant role in this positioning [58].  The interaction of RPA with XPG and XPF/ERCC1 stimulates the junction cutting activities of each of these structure specific endonucleases [44, 58, 60].  XPG incises the damaged strand at the 3' end of the open complex, followed closely by an incision at the 5' end of the complex mediated by XPF/ERCC1 [57, 61, 62].  Although there is some variability in the exact incision sites, the cuts are made at approximately 6 phosphodiester bonds 3' and 22 phosphodiester bonds 5' of the lesion, yielding an excision fragment of 27-29 nt [56].

     Unlike the prokaryotic model, in eukaryotic NER the dual incisions have been reported to result in excision and release of the damage containing oligonucleotide without the aid of additional repair factors [61], perhaps as a result of the helicase activities of XPB and XPD.  Resynthesis of the excised fragment is mediated by DNA pol d or e [63], and sealed by DNA ligase I [64].
 

 
 

 

Transcription-coupled repair in Eukaryotes
 

     Damage in the transcribed strand of active genes is believed to be recognised by its ability to block transcription complexes.  The CSB protein is a component of the RNA pol II elongation complex [65-68], and is believed responsible for recruiting TFIIH to stalled RNA pol II complexes [68-70].  The interaction between CSB and TFIIH is increased by the presence of CSA [68].  TFIIH-mediated unwinding of the DNA around the stalled RNA pol II complex may facilitate elongation past intrinsic pause sites, but is also likely to function in the recruitment of the NER complex when RNA pol II is blocked by DNA damage.  As eukaryotic genes are significantly larger than their prokaryotic counterparts, it is believed that the increased metabolic cost of aborting and reinitiating transcripts in eukaryotic cells has resulted in a mechanism which permits a stalled RNA pol complex to resume mRNA synthesis once the transcription blocking lesion has been removed.  Consistent with this model CSB, unlike the prokaryotic TRCF, does not dissociate the stalled RNA pol II complex from the DNA [71], furthermore a stalled RNA pol complex does not inhibit in vitro repair of the transcription blocking lesion [72]. 

     Although the excision repair patch lengths are similar for TCR and GGR [73], there are indications that the processes and/or complexes involved differ between the two processes.  In addition to the added requirement of the CS proteins, and absence of a role for XPC/HHR23B in TCR, the role of RPA also appears to differ between the two processes.  A yeast RPA mutant has been identified which, although it is completely GGR deficient, retains significant TCR activity [74].  Furthermore, although repair synthesis of GGR is performed by DNA polymerase d or e [75, 76], there is some suggestion that TCR is not [77].  Unfortunately, no in vitro system exists that supports TCR, in spite of the ability of cell extracts to simultaneously support both transcription and repair [78, 79].  Thus, characterisation of the TCR intermediates (and comparison to those of GGR) is not yet possible. 
 
 
 
 


 
DNA repair homepageWhat is DNA?  (intro / review)UV and skin cancerNucleotide excision repairBase excision repairDSB repairMismatch repair
  

  
References:

1. van Hoffen, A., et al., Transcription-coupled repair removes both cyclobutane pyrimidine dimers and 6-4 photoproducts with equal efficiency and in a sequential way from transcribed DNA in xeroderma pigmentosum group C fibroblasts. Embo J, 1995. 14(2): p. 360-7.

2. Mellon, I., et al., Preferential DNA repair of an active gene in human cells. Proc Natl Acad Sci U S A, 1986. 83(23): p. 8878-82.

3. Mellon, I., G. Spivak, and P.C. Hanawalt, Selective removal of transcription-blocking DNA damage from the transcribed strand of the mammalian DHFR gene. Cell, 1987. 51: p. 241-249.

4. Venema, J., et al., Xeroderma pigmentosum complementation group C cells remove pyrimidine dimers selectively from the transcribed strand of active genes. Molecular & Cellular Biology, 1991. 11(8): p. 4128-34.

5. Bohr, V.A., et al., DNA repair in an active gene: removal of pyrimidine dimers from the DHFR gene of CHO cells is much more efficient than in the genome overall. Cell, 1985. 40(2): p. 359-69.

6. Tanaka, K., et al., Analysis of a human DNA excision repair gene involved in group A xeroderma pigmentosum and containing a zinc-finger domain [see comments]. Nature, 1990. 348(6296): p. 73-6.

7. Legerski, R. and C. Peterson, Expression cloning of a human DNA repair gene involved in xeroderma pigmentosum group C. Nature, 1992. 359(6390): p. 70-3.

8. Westerveld, A., et al., Molecular cloning of a human DNA repair gene. Nature, 1984. 310(5976): p. 425-9.

9. Weber, C.A., et al., Molecular cloning and biological characterization of a human gene, ERCC2, that corrects the nucleotide excision repair defect in CHO UV5 cells. Mol Cell Biol, 1988. 8(3): p. 1137-46.

10. Jones, C.J. and R.D. Wood, Preferential binding of the xeroderma pigmentosum group A complementing protein to damaged DNA. Biochemistry, 1993. 32(45): p. 12096-104.

11. Tanaka, K., The Japan Society of Human Genetics Award Lecture. Molecular analysis of xeroderma pigmentosum group A gene. Japanese Journal of Human Genetics, 1993. 38(1): p. 1-14.

12. Abramic, M., A.S. Levine, and M. Protic, Purification of an ultraviolet-inducible, damage-specific DNA-binding protein from primate cells. Journal of Biological Chemistry, 1991. 266(33): p. 22493-500.

13. Hwang, B.J. and G. Chu, Purification and characterization of a human protein that binds to damaged DNA. Biochemistry, 1993. 32(6): p. 1657-66.

14. Reardon, J.T., et al., Comparative analysis of binding of human damaged DNA-binding protein (XPE) and Escherichia coli damage recognition protein (UvrA) to the major ultraviolet photoproducts: T[c,s]T, T[t,s]T, T[6-4]T, and T[Dewar]T. Journal of Biological Chemistry, 1993. 268(28): p. 21301-8.

15. Burns, J.L., et al., An affinity of human replication protein A for ultraviolet-damaged DNA. Journal of Biological Chemistry, 1996. 271(20): p. 11607-10.

16. Clugston, C.K., et al., Binding of human single-stranded DNA binding protein to DNA damaged by the anticancer drug cis-diamminedichloroplatinum (II). Cancer Res, 1992. 52(22): p. 6375-9.

17. He, Z., et al., RPA involvement in the damage-recognition and incision steps of nucleotide excision repair. Nature, 1995. 374(6522): p. 566-9.

18. Otrin, V.R., et al., Translocation of a UV-damaged DNA binding protein into a tight association with chromatin after treatment of mammalian cells with UV light. J Cell Sci, 1997. 110(Pt 10): p. 1159-68.

19. Li, L., et al., An interaction between the DNA repair factor XPA and replication protein A appears essential for nucleotide excision repair. Mol Cell Biol, 1995. 15(10): p. 5396-402.

20. Stigger, E., R. Drissi, and S.H. Lee, Functional analysis of human replication protein A in nucleotide excision repair. J Biol Chem, 1998. 273(15): p. 9337-43.

21. Reardon, J.T., D. Mu, and A. Sancar, Overproduction, purification, and characterization of the XPC subunit of the human DNA repair excision nuclease. Journal of Biological Chemistry, 1996. 271(32): p. 19451-6.

22. Sugasawa, K., et al., Xeroderma pigmentosum group C protein complex is the initiator of global genome nucleotide excision repair. Mol Cell, 1998. 2(2): p. 223-32.

23. Sugasawa, K., et al., HHR23B, a human Rad23 homolog, stimulates XPC protein in nucleotide excision repair in vitro. Molecular & Cellular Biology, 1996. 16(9): p. 4852-61.

24. Masutani, C., et al., Purification and cloning of a nucleotide excision repair complex involving the xeroderma pigmentosum group C protein and a human homologue of yeast RAD23. Embo J, 1994. 13(8): p. 1831-43.

25. Drapkin, R., et al., Dual role of TFIIH in DNA excision repair and in transcription by RNA polymerase II. Nature, 1994. 368(6473): p. 769-72.

26. Evans, E., et al., Mechanism of open complex and dual incision formation by human nucleotide excision repair factors. Embo J, 1997. 16(21): p. 6559-73.

27. Ma, L., et al., The xeroderma pigmentosum group B protein ERCC3 produced in the baculovirus system exhibits DNA helicase activity. Nucleic Acids Res, 1994. 22(20): p. 4095-102.

28. Hwang, J.R., et al., A 3' --> 5' XPB helicase defect in repair/transcription factor TFIIH of xeroderma pigmentosum group B affects both DNA repair and transcription. Journal of Biological Chemistry, 1996. 271(27): p. 15898-904.

29. Sung, P., et al., Human xeroderma pigmentosum group D gene encodes a DNA helicase. Nature, 1993. 365(6449): p. 852-5.

30. Schaeffer, L., et al., The ERCC2/DNA repair protein is associated with the class II BTF2/TFIIH transcription factor. EMBO Journal, 1994. 13(10): p. 2388-92.

31. Park, C.H., et al., The general transcription-repair factor TFIIH is recruited to the excision repair complex by the XPA protein independent of the TFIIE transcription factor. Journal of Biological Chemistry, 1995. 270(9): p. 4896-902.

32. Guzder, S.N., et al., Yeast DNA repair protein RAD23 promotes complex formation between transcription factor TFIIH and DNA damage recognition factor RAD14. J Biol Chem, 1995. 270(15): p. 8385-8.

33. Matsuda, T., et al., DNA repair protein XPA binds replication protein A (RPA). J Biol Chem, 1995. 270(8): p. 4152-7.

34. Mu, D., et al., Characterization of reaction intermediates of human excision repair nuclease. J Biol Chem, 1997. 272(46): p. 28971-9.

35. Wakasugi, M. and A. Sancar, Assembly, subunit composition, and footprint of human DNA repair excision nuclease. Proc Natl Acad Sci U S A, 1998. 95(12): p. 6669-74.

36. Li, L., et al., Specific association between the human DNA repair proteins XPA and ERCC1. Proc Natl Acad Sci U S A, 1994. 91(11): p. 5012-6.

37. Park, C.H. and A. Sancar, Formation of a ternary complex by human XPA, ERCC1, and ERCC4(XPF) excision repair proteins. Proceedings of the National Academy of Sciences of the United States of America, 1994. 91(11): p. 5017-21.

38. Li, L., et al., Mutations in XPA that prevent association with ERCC1 are defective in nucleotide excision repair. Mol Cell Biol, 1995. 15(4): p. 1993-8.

39. Nagai, A., et al., Enhancement of damage-specific DNA binding of XPA by interaction with the ERCC1 DNA repair protein. Biochemical & Biophysical Research Communications, 1995. 211(3): p. 960-6.

40. Arase, S., et al., A sixth complementation group in xeroderma pigmentosum. Mutat Res, 1979. 59(1): p. 143-6.

41. Zelle, B. and P.H. Lohman, Repair of UV-endonuclease-susceptible sites in the 7 complementation groups of xeroderma pigmentosum A through G. Mutat Res, 1979. 62(2): p. 363-8.

42. Zelle, B., F. Berends, and P.H. Lohman, Repair of damage by ultraviolet radiation in xeroderma pigmentosum cell strains of complementation groups E and F. Mutat Res, 1980. 73(1): p. 157-69.

43. Galloway, A.M., M. Liuzzi, and M.C. Paterson, Metabolic processing of cyclobutyl pyrimidine dimers and (6-4) photoproducts in UV-treated human cells. Evidence for distinct excision-repair pathways. J Biol Chem, 1994. 269(2): p. 974-80.

44. de Laat, W.L., et al., DNA-binding polarity of human replication protein A positions nucleases in nucleotide excision repair. Genes Dev, 1998. 12(16): p. 2598-609.

45. Georgaki, A., et al., DNA unwinding activity of replication protein A. FEBS Lett, 1992. 308(3): p. 240-4.

46. Georgaki, A. and U. Hubscher, DNA unwinding by replication protein A is a property of the 70 kDa subunit and is facilitated by phosphorylation of the 32 kDa subunit. Nucleic Acids Res, 1993. 21(16): p. 3659-65.

47. Treuner, K., U. Ramsperger, and R. Knippers, Replication protein A induces the unwinding of long double-stranded DNA regions. J Mol Biol, 1996. 259(1): p. 104-12.

48. Kim, J.K. and B.S. Choi, The solution structure of DNA duplex-decamer containing the (6-4) photoproduct of thymidylyl(3'-->5')thymidine by NMR and relaxation matrix refinement. Eur J Biochem, 1995. 228(3): p. 849-54.

49. Nocentini, S., et al., DNA damage recognition by XPA protein promotes efficient recruitment of transcription factor II H. J Biol Chem, 1997. 272(37): p. 22991-4.

50. Li, R.Y., et al., Interactions of the transcription/DNA repair factor TFIIH and XP repair proteins with DNA lesions in a cell-free repair assay. J Mol Biol, 1998. 281(2): p. 211-8.

51. Wakasugi, M. and A. Sancar, Order of assembly of human DNA repair excision nuclease. J Biol Chem, 1999. 274(26): p. 18759-68.

52. Mu, D., et al., Reconstitution of human DNA repair excision nuclease in a highly defined system. Journal of Biological Chemistry, 1995. 270(6): p. 2415-8.

53. Kim, C., R.O. Snyder, and M.S. Wold, Binding properties of replication protein A from human and yeast cells. Mol Cell Biol, 1992. 12(7): p. 3050-9.

54. Kim, C., B.F. Paulus, and M.S. Wold, Interactions of human replication protein A with oligonucleotides. Biochemistry, 1994. 33(47): p. 14197-206.

55. Blackwell, L.J., J.A. Borowiec, and I.A. Masrangelo, Single-stranded-DNA binding alters human replication protein A structure and facilitates interaction with DNA-dependent protein kinase. Mol Cell Biol, 1996. 16(9): p. 4798-807.

56. Huang, J.C., et al., Human nucleotide excision nuclease removes thymine dimers from DNA by incising the 22nd phosphodiester bond 5' and the 6th phosphodiester bond 3' to the photodimer. Proceedings of the National Academy of Sciences of the United States of America, 1992. 89(8): p. 3664-8.

57. O'Donovan, A., et al., XPG endonuclease makes the 3' incision in human DNA nucleotide excision repair. Nature, 1994. 371(6496): p. 432-5.

58. Matsunaga, T., et al., Replication protein A confers structure-specific endonuclease activities to the XPF-ERCC1 and XPG subunits of human DNA repair excision nuclease. Journal of Biological Chemistry, 1996. 271(19): p. 11047-50.

59. Iyer, N., et al., Interactions involving the human RNA polymerase II transcription/nucleotide excision repair complex TFIIH, the nucleotide excision repair protein XPG, and Cockayne syndrome group B (CSB) protein. Biochemistry, 1996. 35(7): p. 2157-67.

60. Bessho, T., et al., Reconstitution of human excision nuclease with recombinant XPF-ERCC1 complex. Journal of Biological Chemistry, 1997. 272(6): p. 3833-7.

61. Mu, D., D.S. Hsu, and A. Sancar, Reaction mechanism of human DNA repair excision nuclease. Journal of Biological Chemistry, 1996. 271(14): p. 8285-94.

62. Wakasugi, M., J.T. Reardon, and A. Sancar, The non-catalytic function of XPG protein during dual incision in human nucleotide excision repair. J Biol Chem, 1997. 272(25): p. 16030-4.

63. Wood, R.D. and M.K. Shivji, Which DNA polymerases are used for DNA-repair in eukaryotes? Carcinogenesis, 1997. 18(4): p. 605-10.

64. Nocentini, S., Rejoining kinetics of DNA single- and double-strand breaks in normal and DNA ligase-deficient cells after exposure to ultraviolet C and gamma radiation: an evaluation of ligating activities involved in different DNA repair processes. Radiat Res, 1999. 151(4): p. 423-32.

65. Selby, C.P. and A. Sancar, Cockayne syndrome group B protein enhances elongation by RNA polymerase II. Proc Natl Acad Sci U S A, 1997. 94(21): p. 11205-9.

66. Tantin, D., A. Kansal, and M. Carey, Recruitment of the putative transcription-repair coupling factor CSB/ERCC6 to RNA polymerase II elongation complexes. Mol Cell Biol, 1997. 17(12): p. 6803-14.

67 van Gool, A.J., et al., The Cockayne syndrome B protein, involved in transcription-coupled DNA repair, resides in an RNA polymerase II-containing complex. Embo J, 1997. 16(19): p. 5955-65.

68. Tantin, D., RNA polymerase II elongation complexes containing the Cockayne syndrome group B protein interact with a molecular complex containing the transcription factor IIH components xeroderma pigmentosum B and p62. J Biol Chem, 1998. 273(43): p. 27794-9.

69. Tu, Y., S. Bates, and G.P. Pfeifer, The transcription-repair coupling factor CSA is required for efficient repair only during the elongation stages of RNA polymerase II transcription. Mutat Res, 1998. 400(1-2): p. 143-51.

70. Tu, Y., S. Bates, and G.P. Pfeifer, Sequence-specific and Domain-specific DNA Repair in Xeroderma Pigmentosum and Cockayne Syndrome Cells. J. Biol. Chem., 1997. 272(33): p. 20747-20755.

71. Selby, C.P. and A. Sancar, Human transcription-repair coupling factor CSB/ERCC6 is a DNA-stimulated ATPase but is not a helicase and does not disrupt the ternary transcription complex of stalled RNA polymerase II. Journal of Biological Chemistry, 1997. 272(3): p. 1885-90.

72. Selby, C.P., et al., RNA polymerase II stalled at a thymine dimer: footprint and effect on excision repair. Nucleic Acids Research, 1997. 25(4): p. 787-93.

73. Bowman, K.K., C.A. Smith, and P.C. Hanawalt, Excision-repair patch lengths are similar for transcription-coupled repair and global genome repair in UV-irradiated human cells. Mutat Res, 1997. 385(2): p. 95-105.

74. Teng, Y., et al., Mutants with changes in different domains of yeast replication protein A exhibit differences in repairing the control region, the transcribed strand and the non-transcribed strand of the Saccharomyces cerevisiae MFA2 gene. J Mol Biol, 1998. 280(3): p. 355-63.

75. Wood, R.D. and M.K. Shivji, Which DNA polymerases are used for DNA-repair in eukaryotes? Carcinogenesis, 1997. 18(4): p. 605-10.

76. Coverley, D., et al., A role for the human single-stranded DNA binding protein HSSB/RPA in an early stage of nucleotide excision repair. Nucleic Acids Research, 1992. 20(15): p. 3873-80.

77. Tyrrell, R.M. and F. Amaudruz, Evidence for two independent pathways of biologically effective excision repair from its rate and extent in cells cultured from sun-sensitive humans. Cancer Res, 1987. 47(14): p. 3725-8.

78. Satoh, M.S. and P.C. Hanawalt, TFIIH-mediated nucleotide excision repair and initiation of mRNA transcription in an optimized cell-free DNA repair and RNA transcription assay. Nucleic Acids Research, 1996. 24(18): p. 3576-82.

79. Wang, Z., X. Wu, and E.C. Friedberg, A yeast whole cell extract supports nucleotide excision repair and RNA polymerase II transcription in vitro. Mutation Research, 1996. 364(1): p. 33-41.
 


  

 
 
DNA repair homepageWhat is DNA?  (intro / review)UV and skin cancerNucleotide excision repairBase excision repairDSB repairMismatch repair