Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

Mechanisms of change in gene copy number

Key Points

  • Copy number variants (CNVs) arise by homologous recombination (HR) between repeated sequences (recurrent CNVs) or by non-homologous recombination mechanisms that occur throughout the genome (non-recurrent CNVs).

  • Non-recurrent CNVs frequently show microhomology at their end-points, and can have a complex structure.

  • The locus-specific mutation frequencies for copy number variation and other structural changes are two to four orders of magnitude greater than for point mutations.

  • HR mechanisms generally achieve accurate repair of DNA damage.

  • Double-stranded breaks are repaired by HR or by end-joining mechanisms, which can be non-homologous.

  • Broken replication forks with single double-stranded ends are also repaired by HR.

  • There is evidence that repair of broken replication forks underlies some non-homologous recombination.

  • Repair of broken forks in stressed cells could cause non-homologous repair because of a downregulation of HR proteins induced by stress.

  • Models are presented for mechanisms by which stress might induce non-homologous events leading to copy number variation.

Abstract

Deletions and duplications of chromosomal segments (copy number variants, CNVs) are a major source of variation between individual humans and are an underlying factor in human evolution and in many diseases, including mental illness, developmental disorders and cancer. CNVs form at a faster rate than other types of mutation, and seem to do so by similar mechanisms in bacteria, yeast and humans. Here we review current models of the mechanisms that cause copy number variation. Non-homologous end-joining mechanisms are well known, but recent models focus on perturbation of DNA replication and replication of non-contiguous DNA segments. For example, cellular stress might induce repair of broken replication forks to switch from high-fidelity homologous recombination to non-homologous repair, thus promoting copy number change.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Complex structural variation.
Figure 2: Mechanisms of homologous recombination.
Figure 3: Change in copy number by homologous recombination.
Figure 4: The breakage–fusion–bridge cycle.
Figure 5: Replicative mechanisms for non-homologous structural change.

Similar content being viewed by others

References

  1. Iafrate, A. J. et al. Detection of large-scale variation in the human genome. Nature Genet. 36, 949–951 (2004).

    CAS  PubMed  Google Scholar 

  2. Kidd, J. M. et al. Mapping and sequencing of structural variation from eight human genomes. Nature 453, 56–64 (2008).

    CAS  PubMed Central  PubMed  Google Scholar 

  3. Korbel, J. O. et al. Paired-end mapping reveals extensive structural variation in the human genome. Science 318, 420–426 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Redon, R. et al. Global variation in copy number in the human genome. Nature 444, 444–454 (2006). A survey of 270 individuals from the human HapMap samples using SNP arrays and comparative genomic hybridization.

    CAS  PubMed Central  PubMed  Google Scholar 

  5. Sebat, J. et al. Large-scale copy number polymorphism in the human genome. Science 305, 525–528 (2004).

    CAS  PubMed  Google Scholar 

  6. Wong, K. K. et al. A comprehensive analysis of common copy-number variations in the human genome. Am. J. Hum. Genet. 80, 91–104 (2007).

    CAS  PubMed  Google Scholar 

  7. Bruder, C. E. G. et al. Phenotypically concordant and discordant monozygotic twins display different DNA copy-number-variation profiles. Am. J. Hum. Genet. 82, 1–9 (2008).

    Google Scholar 

  8. Piotrowski, A. et al. Somatic mosaicism for copy number variation in differentiated human tissues. Hum. Mutat. 29, 1118–1124 (2008).

    PubMed  Google Scholar 

  9. Beckmann, J. S., Estivill, X. & Antonarakis, S. E. Copy number variants and genetic traits: closer to the resolution of phenotypic to genotypic variability. Nature Rev. Genet. 8, 639–646 (2007).

    CAS  PubMed  Google Scholar 

  10. Dumas, L. et al. Gene copy number variation spanning 60 million years of human and primate evolution. Genome Res. 17, 1266–1277 (2007). This paper traces the history of copy number variation through the evolution of the primate lineage.

    CAS  PubMed Central  PubMed  Google Scholar 

  11. Nahon, J. L. Birth of 'human-specific' genes during primate evolution. Genetica 118, 193–208 (2003).

    CAS  PubMed  Google Scholar 

  12. Bailey, J. A. & Eichler, E. E. Primate segmental duplications: crucibles of evolution, diversity and disease. Nature Rev. Genet. 7, 552–564 (2006).

    CAS  PubMed  Google Scholar 

  13. Stankiewicz, P., Shaw, C. J., Withers, M., Inoue, K. & Lupski, J. R. Serial segmental duplications during primate evolution result in complex human genome architecture. Genome Res. 14, 2209–2220 (2004).

    CAS  PubMed Central  PubMed  Google Scholar 

  14. Marques-Bonet, T. et al. A burst of segmental duplications in the genome of the African great ape ancestor. Nature 457, 877–881 (2009).

    CAS  PubMed Central  PubMed  Google Scholar 

  15. Inoue, K. & Lupski, J. R. Molecular mechanisms for genomic disorders. Annu. Rev. Genomics Hum. Genet. 3, 199–242 (2002).

    CAS  PubMed  Google Scholar 

  16. Ohno, S. Evolution by Gene Duplication (Springer, Berlin, New York, 1970).

    Google Scholar 

  17. Rotger, M. et al. Partial deletion of CYP2B6 owing to unequal crossover with CYP2B7. Pharmacogenet. Genomics 17, 885–890 (2007).

    CAS  PubMed  Google Scholar 

  18. Zhang, F. et al. The DNA replication FoSTeS/MMBIR mechanism can generate human genomic, genic, and exon shuffling rearrangements. Nature Genet. 41, 849–853 (2009). Studies of disease associated rearrangements in proximal chromosome 17p reveal complex rearrangements of varying size and occurrence during mitosis with potential implications for genetic counselling regarding recurrence risk.

    CAS  PubMed  Google Scholar 

  19. Volik, S. et al. Decoding the fine-scale structure of a breast cancer genome and transcriptome. Genome Res. 16, 394–404 (2006).

    CAS  PubMed Central  PubMed  Google Scholar 

  20. Brodeur, G. M. & Hogarty, M. D. in The Genetic Basis of Human Cancer (eds. Vogelstein, B. & Kinzler, K. W.) 161–172 (McGraw-Hill, New York, 1998).

    Google Scholar 

  21. Frank, B. et al. Copy number variant in the candidate tumor suppressor gene MTUS1 and familial breast cancer risk. Carcinogenesis 28, 1442–1445 (2007).

    CAS  PubMed  Google Scholar 

  22. Lupski, J. R. Genomic disorders: structural features of the genome can lead to DNA rearrangement and human disease traits. Trends Genet. 14, 417–422 (1998).

    CAS  PubMed  Google Scholar 

  23. Stranger, B. E. et al. Relative impact of nucleotide and copy number variation on gene expression phenotypes. Science 315, 848–853 (2007).

    CAS  PubMed Central  PubMed  Google Scholar 

  24. Hastings, P. J., Ira, G. & Lupski, J. R. A microhomology-mediated break-induced replication model for the origin of human copy number variation. PLoS Genet. 5, e1000327 (2009). This review presents the MMBIR model for chromosomal rearrangement with detail of the evidence on which it is based.

    CAS  PubMed Central  PubMed  Google Scholar 

  25. Friedberg, E. C. et al. DNA Repair and Mutagenesis (ASM, Washington DC, 2005).

    Google Scholar 

  26. Lee, J. A., Carvalho, C. M. & Lupski, J. R. A DNA replication mechanism for generating nonrecurrent rearrangements associated with genomic disorders. Cell 131, 1235–1247 (2007). Description of the complex structure and microhomology of non-recurrent duplications seen in patients with a genomic disorder.

    CAS  PubMed  Google Scholar 

  27. Nobile, C. et al. Analysis of 22 deletion breakpoints in dystrophin intron 49. Hum. Genet. 110, 418–421 (2002).

    CAS  PubMed  Google Scholar 

  28. Carvalho, C. M. et al. Complex rearrangements in patients with duplications of MECP2 can occur by Fork Stalling and Template Switching. Hum. Mol. Genet. 18, 2188–2203 (2009).

    CAS  PubMed Central  PubMed  Google Scholar 

  29. Chen, J. M., Chuzhanova, N., Stenson, P. D., Férec, C. & Cooper, D. N. Intrachromosomal serial replication slippage in trans gives rise to diverse genomic rearrangements involving inversions. Hum. Mutat. 26, 362–373 (2005).

    PubMed  Google Scholar 

  30. Gajecka, M. et al. Unexpected complexity at breakpoint junctions in phenotypically normal individuals and mechanisms involved in generating balanced translocations t(1;22)(p36;q13). Genome Res. 18, 1733–1742 (2008).

    CAS  PubMed Central  PubMed  Google Scholar 

  31. Sheen, C. R. et al. Double complex mutations involving F8 and FUNDC2 caused by distinct break-induced replication. Hum. Mutat. 28, 1198–2006 (2007).

    CAS  PubMed  Google Scholar 

  32. Vissers, L. E. et al. Complex chromosome 17p rearrangements associated with low-copy repeats in two patients with congenital anomalies. Hum. Genet. 121, 697–709 (2007).

    CAS  PubMed Central  PubMed  Google Scholar 

  33. Stankiewicz, P. et al. Genome architecture catalyzes nonrecurrent chromosomal rearrangements. Am. J. Hum. Genet. 72, 1101–1116 (2003).

    CAS  PubMed Central  PubMed  Google Scholar 

  34. Lee, J. A. et al. Role of genomic architecture in PLP1 duplication causing Pelizaeus–Merzbacher disease. Hum. Mol. Genet. 15, 2250–2265 (2006).

    CAS  PubMed  Google Scholar 

  35. Lee, J. A. et al. Spastic paraplegia type 2 associated with axonal neuropathy and apparent PLP1 position effect. Ann. Neurol. 59, 398–403 (2006).

    CAS  PubMed  Google Scholar 

  36. Lovett, S. T., Hurley, R. L., Sutera, V. A. Jr, Aubuchon, R. H. & Lebedeva, M. A. Crossing over between regions of limited homology in Escherichia coli. RecA-dependent and RecA-independent pathways. Genetics 160, 851–859 (2002).

    CAS  PubMed Central  PubMed  Google Scholar 

  37. Liskay, R. M., Letsou, A. & Stachelek, J. L. Homology requirement for efficient gene conversion between duplicated chromosomal sequences in mammalian cells. Genetics 115, 161–167 (1987).

    CAS  PubMed Central  PubMed  Google Scholar 

  38. Reiter, L. T. et al. Human meiotic recombination products revealed by sequencing a hotspot for homologous strand exchange in multiple HNPP deletion patients. Am. J. Hum. Genet. 62, 1023–1033 (1998).

    CAS  PubMed Central  PubMed  Google Scholar 

  39. Stankiewicz, P. & Lupski, J. R. Genome architecture, rearrangements and genomic disorders. Trends Genet. 18, 74–82 (2002).

    CAS  PubMed  Google Scholar 

  40. Krogh, B. O. & Symington, L. S. Recombination proteins in yeast. Annu. Rev. Genet. 38, 233–271 (2004).

    CAS  PubMed  Google Scholar 

  41. Pâques, F. & Haber, J. E. Multiple pathways of recombination induced by double-strand breaks in Saccharomyces cerevisiae. Microbiol. Mol. Biol. Rev. 63, 349–404 (1999).

    PubMed Central  PubMed  Google Scholar 

  42. Esposito, M. S. Evidence that spontaneous mitotic recombination occurs at the two-strand stage. Proc. Natl Acad. Sci. USA 75, 4436–4440 (1978).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Stark, J. M. & Jasin, M. Extensive loss of heterozygosity is suppressed during homologous repair of chromosomal breaks. Mol. Cell. Biol. 23, 733–743 (2003).

    CAS  PubMed Central  PubMed  Google Scholar 

  44. Dupaigne, P. et al. The Srs2 helicase activity is stimulated by Rad51 filaments on dsDNA: implications for crossover incidence during mitotic recombination. Mol. Cell. 29, 243–254 (2008).

    CAS  PubMed  Google Scholar 

  45. Sun, W. et al. The FANCM ortholog Fml1 promotes recombination at stalled replication forks and limits crossing over during DNA double-strand break repair. Mol. Cell 32, 118–128 (2008).

    CAS  PubMed Central  PubMed  Google Scholar 

  46. Ira, G., Malkova, A., Liberi, G., Foiani, M. & Haber, J. E. Srs2 and Sgs1–Top3 suppress crossovers during double-strand break repair in yeast. Cell 115, 401–411 (2003).

    CAS  PubMed Central  PubMed  Google Scholar 

  47. Prakash, R. et al. Yeast Mph1 helicase dissociates Rad51-made D-loops: implications for crossover control in mitotic recombination. Genes Dev. 23, 67–79 (2009).

    CAS  PubMed Central  PubMed  Google Scholar 

  48. Wu, L. & Hickson, I. D. The Bloom's syndrome helicase suppresses crossing over during homologous recombination. Nature 426, 870–874 (2003).

    CAS  PubMed  Google Scholar 

  49. Prado, F. & Aguilera, A. Control of cross-over by single-strand DNA resection. Trends. Genet. 19, 428–431 (2003).

    CAS  PubMed  Google Scholar 

  50. Smith, C. E., Llorente, B. & Symington, L. S. Template switching during break-induced replication. Nature 447, 102–105 (2007). This paper shows experimental evidence from yeast on the nature of BIR. Specifically, template switching between homologous chromosomes (or sometimes non-homologous chromosomes, which causes translocation). It showed replication out to the telomere after switching.

    CAS  PubMed  Google Scholar 

  51. Bauters, M. et al. Nonrecurrent MECP2 duplications mediated by genomic architecture-driven DNA breaks and break-induced replication repair. Genome Res. 18, 847–858 (2008).

    CAS  PubMed Central  PubMed  Google Scholar 

  52. Deem, A. et al. Defective break-induced replication leads to half-crossovers in Saccharomyces cerevisiae. Genetics 179, 1845–1860 (2008).

    CAS  PubMed Central  PubMed  Google Scholar 

  53. Narayanan, V. & Lobachev, K. S. Intrachromosomal gene amplification triggered by hairpin-capped breaks requires homologous recombination and is independent of nonhomologous end-joining. Cell Cycle 6, 1814–1818 (2007).

    CAS  PubMed  Google Scholar 

  54. Payen, C., Koszul, R., Dujon, B. & Fischer, G. Segmental duplications arise from Pol32-dependent repair of broken forks through two alternative replication-based mechanisms. PLoS Genet. 4, e1000175 (2008). Evidence from yeast that LCRs arise by a replicative mechanism, specifically one involving BIR.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Schmidt, K. H., Wu, J. & Kolodner, R. D. Control of translocations between highly diverged genes by Sgs1, the Saccharomyces cerevisiae homolog of the Bloom's syndrome protein. Mol. Cell. Biol. 26, 5406–5420 (2006).

    CAS  PubMed Central  PubMed  Google Scholar 

  56. Lin, F. L., Sperle, K. & Sternberg, N. Model for homologous recombination during transfer of DNA into mouse L cells: role for DNA ends in the recombination process. Mol. Cell. Biol. 4, 1020–1034 (1984).

    CAS  PubMed Central  PubMed  Google Scholar 

  57. Sweigert, S. E. & Carroll, D. Repair and recombination of X-irradiated plasmids in Xenopus laevis oocytes. Mol. Cell. Biol. 10, 5849–5856 (1990).

    CAS  PubMed Central  PubMed  Google Scholar 

  58. Haber, J. E. Exploring the pathways of homologous recombination. Curr. Opin. Cell Biol. 4, 401–412 (1992).

    CAS  PubMed  Google Scholar 

  59. Elliott, B., Richardson, C. & Jasin, M. Chromosomal translocation mechanisms at intronic Alu elements in mammalian cells. Mol. Cell 17, 885–894 (2005).

    CAS  PubMed  Google Scholar 

  60. Rayssiguier, C., Thaler, D. S. & Radman, M. The barrier to recombination between Escherichia coli and Salmonella typhimurium is disrupted in mismatch repair mutants. Nature 342, 396–401 (1989).

    CAS  PubMed  Google Scholar 

  61. Unal, E. et al. DNA damage response pathway uses histone modification to assemble a double-strand break-specific cohesin domain. Mol. Cell 16, 991–1002 (2004).

    PubMed  Google Scholar 

  62. Strom, L., Lindroos, H. B., Shirahige, K. & Sjogren, C. Postreplicative recruitment of cohesin to double-strand breaks is required for DNA repair. Mol. Cell 16, 1003–1015 (2004).

    PubMed  Google Scholar 

  63. Kim, J. S., Krasieva, T. B., LaMorte, V., Taylor, A. M. & Yokomori, K. Specific recruitment of human cohesin to laser-induced DNA damage. J. Biol. Chem. 277, 45149–45153 (2002).

    CAS  PubMed  Google Scholar 

  64. Sjogren, C. & Nasmyth, K. Sister chromatid cohesion is required for postreplicative double-strand break repair in Saccharomyces cerevisiae. Curr. Biol. 11, 991–995 (2001).

    CAS  PubMed  Google Scholar 

  65. Kobayashi, T., Horiuchi, T., Tongaonkar, P., Vu, L. & Nomura, M. SIR2 regulates recombination between different rDNA repeats, but not recombination within individual rRNA genes in yeast. Cell 117, 441–453 (2004).

    CAS  PubMed  Google Scholar 

  66. Kobayashi, T. & Ganley, A. R. Recombination regulation by transcription-induced cohesin dissociation in rDNA repeats. Science 309, 1581–1584 (2005).

    CAS  PubMed  Google Scholar 

  67. Kaye, J. A. et al. DNA breaks promote genomic instability by impeding proper chromosome segregation. Curr. Biol. 14, 2096–2106 (2004).

    CAS  PubMed  Google Scholar 

  68. Soutoglou, E. et al. Positional stability of single double-strand breaks in mammalian cells. Nature Cell Biol. 9, 675–682 (2007).

    CAS  PubMed  Google Scholar 

  69. Oh, S. D. et al. BLM ortholog, Sgs1, prevents aberrant crossing-over by suppressing formation of multichromatid joint molecules. Cell 130, 259–272 (2007).

    CAS  PubMed Central  PubMed  Google Scholar 

  70. Jain, S. et al. A recombination execution checkpoint regulates the choice of homologous recombination pathway during DNA double-strand break repair. Genes Dev. 23, 291–303 (2009).

    CAS  PubMed Central  PubMed  Google Scholar 

  71. McVey, M. & Lee, S. E. MMEJ repair of double-strand breaks (director's cut): deleted sequences and alternative endings. Trends Genet. 24, 529–538 (2008).

    CAS  PubMed Central  PubMed  Google Scholar 

  72. Lieber, M. R. The mechanism of human nonhomologous DNA end joining. J. Biol. Chem. 283, 1–5 (2008).

    CAS  PubMed  Google Scholar 

  73. Daley, J. M., Palmbos, P. L., Wu, D. & Wilson, T. E. Nonhomologous end joining in yeast. Annu. Rev. Genet. 39, 431–451 (2005).

    CAS  PubMed  Google Scholar 

  74. Haviv-Chesner, A., Kobayashi, Y., Gabriel, A. & Kupiec, M. Capture of linear fragments at a double-strand break in yeast. Nucleic Acids Res. 35, 5192–5202 (2007).

    CAS  PubMed Central  PubMed  Google Scholar 

  75. Yu, X. & Gabriel, A. Ku-dependent and Ku-independent end-joining pathways lead to chromosomal rearrangements during double-strand break repair in Saccharomyces cerevisiae. Genetics 163, 843–856 (2003).

    CAS  PubMed Central  PubMed  Google Scholar 

  76. Nickoloff, J. A., De Haro, L. P., Wray, J. & Hromas, R. Mechanisms of leukemia translocations. Curr. Opin. Hematol. 15, 338–345 (2008).

    CAS  PubMed Central  PubMed  Google Scholar 

  77. McClintock, B. Chromosome organization and genic expression. Cold Spring Harb. Symp. Quant. Biol. 16, 13–47 (1951).

    CAS  PubMed  Google Scholar 

  78. Tanaka, H. & Yao, M. C. Palindromic gene amplification — an evolutionarily conserved role for DNA inverted repeats in the genome. Nature Rev. Cancer 9, 216–224 (2009).

    CAS  Google Scholar 

  79. Tanaka, H., Bergstrom, D. A., Yao, M. C. & Tapscott, S. J. Large DNA palindromes as a common form of structural chromosome aberrations in human cancers. Hum. Cell 19, 17–23 (2006).

    PubMed  Google Scholar 

  80. Coquelle, A., Pipiras, E., Toledo, F., Buttin, G. & Debatisse, M. Expression of fragile sites triggers intrachromosomal mammalian gene amplification and sets boundaries to early amplicons. Cell 89, 215–225 (1997).

    CAS  PubMed  Google Scholar 

  81. Shaw, C. J. & Lupski, J. R. Non-recurrent 17p11.2 deletions are generated by homologous and non-homologous mechanisms. Hum. Genet. 116, 1–7 (2005).

    CAS  PubMed  Google Scholar 

  82. Arlt, M. F. et al. Replication stress induces genome-wide copy number changes in human cells that resemble polymorphic and pathogenic variants. Am. J. Hum. Genet. 84, 339–350 (2009).

    CAS  PubMed Central  PubMed  Google Scholar 

  83. Durkin, S. G. et al. Replication stress induces tumor-like microdeletions in FHIT/FRA3B. Proc. Natl Acad. Sci. USA 105, 246–251 (2008).

    CAS  PubMed  Google Scholar 

  84. Kuo, M. T., Vyas, R. C., Jiang, L. X. & Hittelman, W. N. Chromosome breakage at a major fragile site associated with P-glycoprotein gene amplification in multidrug-resistant CHO cells. Mol. Cell Biol. 14, 5202–5211 (1994).

    CAS  PubMed Central  PubMed  Google Scholar 

  85. Coquelle, A., Rozier, L., Dutrillaux, B. & Debatisse, M. Induction of multiple double-strand breaks within an hsr by meganucleaseI-SceI expression or fragile site activation leads to formation of double minutes and other chromosomal rearrangements. Oncogene 21, 7671–7679 (2002).

    CAS  PubMed  Google Scholar 

  86. Michel, B., Ehrlich, S. D. & Uzest, M. DNA double-strand breaks caused by replication arrest. EMBO J. 16, 430–438 (1997).

    CAS  PubMed Central  PubMed  Google Scholar 

  87. Albertini, A. M., Hofer, M., Calos, M. P. & Miller, J. H. On the formation of spontaneous deletions: the importance of short sequence homologies in the generation of large deletions. Cell 29, 319–328 (1982).

    CAS  PubMed  Google Scholar 

  88. Farabaugh, P. J., Schmeissner, U., Hofer, M. & Miller, J. H. Genetic studies of the lac repressor. VII. On the molecular nature of spontaneous hotspots in the lacI gene of Escherichia coli. J. Mol. Biol. 126, 847–857 (1978).

    CAS  PubMed  Google Scholar 

  89. Ikeda, H., Shimizu, H., Ukita, T. & Kumagai, M. A novel assay for illegitimate recombination in Escherichia coli: stimulation of lambda bio transducing phage formation by ultra-violet light and its independence from RecA function. Adv. Biophys. 31, 197–208 (1995).

    CAS  PubMed  Google Scholar 

  90. Shimizu, H. et al. Short-homology-independent illegitimate recombination in Escherichia coli: distinct mechanism from short-homology-dependent illegitimate recombination. J. Mol. Biol. 266, 297–305 (1997).

    CAS  PubMed  Google Scholar 

  91. Bi, X. & Liu, L. F. recA-independent and recA-dependent intramolecular plasmid recombination: differential homology and requirement and distance effect. J. Mol. Biol. 235, 414–423 (1994).

    CAS  PubMed  Google Scholar 

  92. Mazin, A. V., Kuzminov, A. V., Dianov, G. L. & Salganik, R. I. Mechanisms of deletion formation in Escherichia coli plasmids. II. Deletions mediated by short direct repeats. Mol. Gen. Genet. 228, 209–214 (1991).

    CAS  PubMed  Google Scholar 

  93. Chedin, F., Dervyn, E., Dervyn, R., Ehrlich, S. D. & Noirot, P. Frequency of deletion formation decreases exponentially with distance between short direct repeats. Mol. Microbiol. 12, 561–569 (1994).

    CAS  PubMed  Google Scholar 

  94. Lovett, S. T., Gluckman, T. J., Simon, P. J., Sutera Jr, V. A. & Drapkin, P. T. Recombination between repeats in Escherichia coli by a recA-independent, proximity-sensitive mechanism. Mol. Gen. Genet. 254, 294–300 (1994).

    Google Scholar 

  95. Bierne, H., Vilette, D., Ehrlich, S. D. & Michel, B. Isolation of a dnaE mutation which enhances RecA-independent homologous recombination in the Escherichia coli chromosome. Mol. Microbiol. 24, 1225–1234 (1997).

    CAS  PubMed  Google Scholar 

  96. Saveson, C. J. & Lovett, S. T. Enhanced deletion formation by aberrant DNA replication in Escherichia coli. Genetics 146, 457–470 (1997).

    CAS  PubMed Central  PubMed  Google Scholar 

  97. Bzymek, M. & Lovett, S. T. Instability of repetitive DNA sequences: the role of replication in multiple mechanisms. Proc. Natl Acad. Sci. USA 98, 8319–8325 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  98. Lovett, S. T. & Feshenko, V. V. Stabilization of diverged tandem repeats by mismatch repair: evidence for deletion formation via a misaligned replication intermediate. Proc. Natl Acad. Sci. USA 93, 7120–7124 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  99. Cairns, J. & Foster, P. L. Adaptive reversion of a frameshift mutation in Escherichia coli. Genetics 128, 695–701 (1991).

    CAS  PubMed Central  PubMed  Google Scholar 

  100. Slack, A., Thornton, P. C., Magner, D. B., Rosenberg, S. M. & Hastings, P. J. On the mechanism of gene amplification induced under stress in Escherichia coli. PLoS Genet. 2, e48 (2006). Experimental evidence from E. coli suggesting that chromosomal structural change occurs by a replicative mechanism, and also revealing that the characteristics of amplification in E. coli are similar to those of non-recurrent changes seen in human genomic disorders. Proposes CNV formation by template switching between forks.

    PubMed Central  PubMed  Google Scholar 

  101. Kugelberg, E., Kofoid, E., Reams AB, Andersson, D. I. & Roth, J. R. Multiple pathways of selected gene amplification during adaptive mutation. Proc. Natl Acad. Sci. USA 103, 17319–17324 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  102. Allgood, N. D. & Silhavy, T. J. Escherichia coli xonA (sbcB) mutants enhance illegitimate recombination. Genetics 127, 671–680 (1991).

    CAS  PubMed Central  PubMed  Google Scholar 

  103. Bzymek, M., Saveson, C. J., Feschenko, V. V. & Lovett, S. T. Slipped misalignment mechanisms of deletion formation: in vivo susceptibility to nucleases. J. Bacteriol. 181, 477–482 (1999).

    CAS  PubMed Central  PubMed  Google Scholar 

  104. Lydeard, J. R., Jain, S., Yamaguchi, M. & Haber, J. E. Break-induced replication and telomerase-independent telomere maintenance require Pol32. Nature 448, 820–823 (2007).

    CAS  PubMed  Google Scholar 

  105. VanHulle, K. et al. Inverted DNA repeats channel repair of distant double-strand breaks into chromatid fusions and chromosomal rearrangements. Mol. Cell. Biol. 27, 2601–2614 (2007).

    CAS  PubMed Central  PubMed  Google Scholar 

  106. Davis, A. P. & Symington, L. S. RAD51-dependent break-induced replication in yeast. Mol. Cell. Biol. 24, 2344–2351 (2004).

    CAS  PubMed Central  PubMed  Google Scholar 

  107. McVey, M., Adams, M., Staeva-Vieira, E. & Sekelsky, J. J. Evidence for multiple cycles of strand invasion during repair of double-strand gaps in Drosophila. Genetics 167, 699–705 (2004).

    CAS  PubMed Central  PubMed  Google Scholar 

  108. Bindra, R. S., Crosby, M. E. & Glazer, P. M. Regulation of DNA repair in hypoxic cancer cells. Cancer Metastasis Rev. 26, 249–260 (2007).

    CAS  PubMed  Google Scholar 

  109. Huang, L. E., Bindra, R. S., Glazer, P. M. & Harris, A. L. Hypoxia-induced genetic instability — a calculated mechanism underlying tumor progression. J. Mol. Med. 85, 139–148 (2007).

    CAS  PubMed  Google Scholar 

  110. Bindra, R. S. & Glazer, P. M. Repression of RAD51 gene expression by E2F4/p130 complexes in hypoxia. Oncogene. 26, 2048–2057 (2007).

    CAS  PubMed  Google Scholar 

  111. Coquelle, A., Toledo, F., Stern, S., Bieth, A. & Debatisse, M. A new role for hypoxia in tumor progression: induction of fragile site triggering genomic rearrangements and formation of complex DMs and HSRs. Mol. Cell 2, 259–265 (1998).

    CAS  PubMed  Google Scholar 

  112. Hastings, P. J., Bull, H. J., Klump, J. R. & Rosenberg, S. M. Adaptive amplification: an inducible chromosomal instability mechanism. Cell 103, 723–731 (2000).

    CAS  PubMed  Google Scholar 

  113. Lombardo, M.-J., Aponyi, I. & Rosenberg, S. M. General stress response regulator RpoS in adaptive mutation and amplification in Escherichia coli. Genetics 166, 669–680 (2004).

    CAS  PubMed Central  PubMed  Google Scholar 

  114. Ponder, R. G., Fonville, N. C. & Rosenberg, S. M. A switch from high-fidelity to error-prone DNA double-strand break repair underlies stress-induced mutation. Mol. Cell 19, 791–804 (2005).

    CAS  PubMed  Google Scholar 

  115. Mullighan, C. G. et al. Genome-wide analysis of genetic alterations in acute lymphoblastic leukaemia. Nature 446, 758–764 (2007).

    CAS  PubMed  Google Scholar 

  116. Shao, L. et al. Identification of chromosome abnormalities in subtelomeric regions by microarray analysis: a study of 5,380 cases. Am. J. Med. Genet. A 146A, 2242–2251 (2008).

    PubMed Central  PubMed  Google Scholar 

  117. Yatsenko, S. A. et al. Molecular mechanisms for subtelomeric rearrangements associated with the 9q34.3 microdeletion syndrome. Hum. Mol. Genet. 18, 1924–1936 (2009).

    CAS  PubMed Central  PubMed  Google Scholar 

  118. Zhang, L., Lu, H. H., Chung, W. Y., Yang, J. & Li, W. H. Patterns of segmental duplication in the human genome. Mol. Biol. Evol. 22, 135–141 (2005).

    CAS  PubMed  Google Scholar 

  119. She, X. et al. The structure and evolution of centromeric transition regions within the human genome. Nature 430, 857–864 (2004).

    CAS  PubMed  Google Scholar 

  120. Nguyen, D. Q., Webber, C. & Ponting, C. P. Bias of selection on human copy-number variants. PLoS Genet. 2, e20 (2006).

    PubMed Central  PubMed  Google Scholar 

  121. Visser, R. et al. Identification of a 3.0-kb major recombination hotspot in patients with Sotos syndrome who carry a common 1.9-Mb microdeletion. Am. J. Hum. Genet. 76, 52–67 (2005).

    CAS  PubMed  Google Scholar 

  122. de Smith, A. J. et al. Small deletion variants have stable breakpoints commonly associated with Alu elements. PLoS ONE 3, e3104 (2008).

    PubMed Central  PubMed  Google Scholar 

  123. Bacolla, A. et al. Breakpoints of gross deletions coincide with non-B DNA conformations. Proc. Natl Acad. Sci. USA 101, 14162–14167 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  124. Bacolla, A., Wojciechowska, M., Kosmider, B., Larson, J. E. & Wells, R. D. The involvement of non-B DNA structures in gross chromosomal rearrangements. DNA Repair (Amst.) 5, 1161–1170 (2006).

    CAS  Google Scholar 

  125. Inagaki, H. et al. Chromosomal instability mediated by non-B DNA: cruciform conformation and not DNA sequence is responsible for recurrent translocation in humans. Genome Res. 19, 191–198 (2009).

    CAS  PubMed Central  PubMed  Google Scholar 

  126. Myers, S., Freeman, C., Auton, A., Donnelly, P. & McVean, G. A common sequence motif associated with recombination hot spots and genome instability in humans. Nature Genet. 40, 1124–1129 (2008).

    CAS  PubMed  Google Scholar 

  127. Shaw, C. J. & Lupski, J. R. Implications of human genome architecture for rearrangement-based disorders: the genomic basis of disease. Hum. Mol. Genet. 13, R57–R64 (2004).

    CAS  PubMed  Google Scholar 

  128. Bi, W. et al. Increased LIS1 expression affects human and mouse brain development. Nature Genet. 41, 168–177 (2009).

    CAS  PubMed  Google Scholar 

  129. Galhardo, R. S., Hastings, P. J. & Rosenberg, S. M. Mutation as a stress response and the regulation of evolvability. Crit. Rev. Biochem. Mol. Biol. 42, 399–435 (2007). A broad review of stress-induced mutation and its meaning for evolution.

    CAS  PubMed Central  PubMed  Google Scholar 

  130. Rosenberg, S. M. Evolving responsively: adaptive mutation. Nature Rev. Genet. 2, 504–515 (2001).

    CAS  PubMed  Google Scholar 

  131. Cirz, R. T. et al. Inhibition of mutation and combating the evolution of antibiotic resistance. PLoS Biol. 3, e176 (2005).

    PubMed Central  PubMed  Google Scholar 

  132. Riesenfeld, C., Everett, M., Piddock, L. J. & Hall, B. G. Adaptive mutations produce resistance to ciprofloxacin. Antimicrob. Agents Chemother. 41, 2059–2060 (1997).

    CAS  PubMed Central  PubMed  Google Scholar 

  133. Bindra, R. S. S. et al. Down-regulation of Rad51 and decreased homologous recombination in hypoxic cancer cells. Mol. Cell. Biol. 24, 8504–8518 (2004). This paper shows that HR enzymes are downregulated by stress in human cells, and that this is accompanied by reduction in HR.

    CAS  PubMed Central  PubMed  Google Scholar 

  134. Mihaylova, V. T. et al. Decreased expression of the DNA mismatch repair gene Mlh1 under hypoxic stress in mammalian cells. Mol. Cell. Biol. 23, 3265–3273 (2003).

    CAS  PubMed Central  PubMed  Google Scholar 

  135. Kolodner, R. D. et al. Germ-line msh6 mutations in colorectal cancer families. Cancer Res. 59, 5068–5074 (1999).

    CAS  PubMed  Google Scholar 

  136. Loeb, L. A. Mutator phenotype may be required for multistage carcinogenesis. Cancer Res. 51, 3075–3079 (1991).

    CAS  PubMed  Google Scholar 

  137. Modrich, P. Mismatch repair, genetic stability and tumour avoidance. Phil. Trans. R. Soc. Lond. B 347, 89–95 (1995).

    CAS  Google Scholar 

  138. Nguyen, D. Q. et al. Reduced purifying selection prevails over positive selection in human copy number variant evolution. Genome Res. 18, 1711–1723 (2008).

    CAS  PubMed Central  PubMed  Google Scholar 

  139. Perry, G. H. et al. Diet and the evolution of human amylase gene copy number variation. Nature Genet. 39, 1256–1260 (2007).

    CAS  PubMed  Google Scholar 

  140. Gonzalez, E. et al. The influence of CCL3L1 gene-containing segmental duplications on HIV-1/AIDS susceptibility. Science 307, 1434–1440 (2005).

    CAS  PubMed  Google Scholar 

  141. Higgs, D. R. et al. A review of the molecular genetics of the human alpha-globin gene cluster. Blood 73, 1081–1104 (1989).

    CAS  PubMed  Google Scholar 

  142. Nozawa, M., Kawahara, Y. & Nei, M. Genomic drift and copy number variation of sensory receptor genes in humans. Proc. Natl Acad. Sci. USA 104, 20421–20426 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  143. Jakobsson, M. et al. Genotype, haplotype and copy-number variation in worldwide human populations. Nature 451, 998–1003 (2008).

    CAS  PubMed  Google Scholar 

  144. Locke, D. P. et al. Linkage disequilibrium and heritability of copy-number polymorphisms within duplicated regions of the human genome. Am. J. Hum. Genet. 79, 275–290 (2006).

    CAS  PubMed Central  PubMed  Google Scholar 

  145. Sharp, A. J. et al. Segmental duplications and copy-number variation in the human genome. Am. J. Hum. Genet. 77, 78–88 (2005).

    CAS  PubMed Central  PubMed  Google Scholar 

  146. Fortna, A. et al. Lineage-specific gene duplication and loss in human and great ape evolution. PLoS Biol. 2, E207 (2004).

    PubMed Central  PubMed  Google Scholar 

  147. McCarroll, S. A. et al. Integrated detection and population-genetic analysis of SNPs and copy number variation. Nature Genet. 40, 1166–1174 (2008).

    CAS  PubMed  Google Scholar 

  148. Turner, D. J. et al. Germline rates of de novo meiotic deletions and duplications causing several genomic disorders. Nature Genet. 40, 90–95 (2008).

    CAS  PubMed  Google Scholar 

  149. Lam, K. W. & Jeffreys, A. J. Processes of copy-number change in human DNA: the dynamics of α-globin gene deletion. Proc. Natl Acad. Sci. USA 103, 8921–8927 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  150. Lam, K. W. & Jeffreys, A. J. Processes of de novo duplication of human α-globin genes. Proc. Natl Acad. Sci. USA 104, 10950–10955 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  151. Flores, M. et al. Recurrent DNA inversion rearrangements in the human genome. Proc. Natl Acad. Sci. USA 104, 6099–6106 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  152. Liang, Q., Conte, N., Skarnes, W. C. & Bradley, A. Extensive genomic copy number variation in embryonic stem cells. Proc. Natl Acad. Sci. USA 105, 17453–17456 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  153. Lupski, J. R. Genomic rearrangements and sporadic disease. Nature Genet. 39, S43–47 (2007).

    CAS  PubMed  Google Scholar 

  154. van Ommen G. J. Frequency of new copy number variation in humans. Nature Genet. 37, 333–334 (2005).

    CAS  PubMed  Google Scholar 

  155. Tuzun, E., Bailey, J. A. & Eichler, E. E. Recent segmental duplications in the working draft assembly of the brown Norway rat. Genome Res. 14, 493–506 (2004).

    CAS  PubMed Central  PubMed  Google Scholar 

  156. Graubert, T. A. et al. A high-resolution map of segmental DNA copy number variation in the mouse genome. PLoS Genet. 3, e3 (2007).

    PubMed Central  PubMed  Google Scholar 

  157. Lovett, S. T. Encoded errors: mutations and rearrangements mediated by misalignment at repetitive DNA sequences. Mol. Microbiol. 52, 1243–1253 (2004).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported by grants from the National Institutes of Health, R01 GM64022 to P.J.H., R01 NS59529 to J.R.L., R01 GM53158 and R01 CA85777 to S.M.R., and R01 GM80600 to G.I.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to P. J. Hastings.

Related links

Related links

FURTHER INFORMATION

Philip J. Hastings' homepage

James R. Lupski's homepage

Susan M. Rosenberg's homepage

Grzegorz Ira's homepage

Glossary

Array comparative genomic hybridization

A microarray-based technique to measure the relative amount of any DNA sequence.

Paired-end mapping

A technique whereby novel linkage relationships are detected by finding short sequences linked to other short sequences in DNA fragments of uniform size.

Non-allelic homologous recombination

Homologous recombination between lengths of homology in different genomic positions.

Double-stranded end

An end of dsDNA that is not protected by a telomere, which is the structure found at the ends of linear chromosomes.

Holliday junction

A point at which the strands of two dsDNA molecules exchange partners. This structure occurs as an intermediate in crossing over.

Gene conversion

A non-reciprocal transfer of an allelic difference from one chromosome to its homologue.

Crossover

A precisely reciprocal breakage of two DNA molecules followed by rejoining with exchanged partners.

Loss of heterozygosity

Loss of an allelic difference between two chromosomes in a diploid cell.

Helicase

An enzyme that separates the two nucleic acid strands of a double helix, resulting in the formation of regions of ssDNA or ssRNA.

Topoisomerase

An enzyme that can remove (or create) supercoiling and concatenation (interlocking) in duplex DNA by creating transitory breaks in one (type I topoisomerase) or both (type II topoisomerase) strands of the sugar–phosphate backbone.

Single-strand annealing

A double-stranded break repair mechanism that deletes sequence between repeats.

Alu

A family of short interspersed nuclear elements that are common in human and primate genomes.

Mismatch repair

A DNA repair system that corrects a mismatched base pair in duplex DNA by excision of a length of one strand followed by synthesis of the sequence complementary to the remaining strand.

Retrotransposon

A transposon (mobile element) that is copied from the host genome by transcription as RNA, and is later reverse-transcribed into DNA and reintegrated into the host genome.

Endonuclease

An enzyme that breaks the sugar–phosphate backbone of a DNA or RNA molecule where there is no free end.

Telomere

A structure at the ends of linear chromosomes that avoids shortening of chromosomes after replication, and that protects the end from homologous and non-homologous recombination.

Dicentric chromosome

A chromosome with two centromeres. These are pulled to opposite poles during mitosis but are unable to separate without chromosome breakage.

Amplification

Also called gene amplification. The formation of more than two repetitions of a chromosomal segment in tandem, dispersed or as autonomous circular molecules.

Fragile site

A position on a chromosome where spontaneous breaks occur frequently.

Okazaki fragment

The discontinuous length of DNA that is synthesized as one piece on the lagging strand template during DNA replication.

Amplicon

The repeat unit, or unit length of genome, that is amplified.

Exonuclease

An enzyme that degrades DNA or RNA from an end.

Heterochromatin

A highly condensed form of chromatin (the eukaryotic complex of DNA with proteins) that shows reduced gene expression and is replicated late in S phase.

LINE

Long interspersed nuclear elements. A class of transposable element lacking long terminal repeats.

SINE

Short interspersed nuclear elements. A class of short (<500 bp) transposable elements.

Evolvability

The capacity of an organism to evolve.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Hastings, P., Lupski, J., Rosenberg, S. et al. Mechanisms of change in gene copy number. Nat Rev Genet 10, 551–564 (2009). https://doi.org/10.1038/nrg2593

Download citation

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrg2593

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing