Hostname: page-component-7c8c6479df-fqc5m Total loading time: 0 Render date: 2024-03-27T02:05:46.874Z Has data issue: false hasContentIssue false

The clinical context of copy number variation in the human genome

Published online by Cambridge University Press:  09 March 2010

Charles Lee
Affiliation:
Department of Pathology, Brigham and Women's Hospital and Harvard Medical School, Boston, MA, USA.
Stephen W. Scherer*
Affiliation:
The Centre for Applied Genomics and Program in Genetics & Genome Biology, Hospital for Sick Children, Toronto, Ontario, Canada. Department of Molecular Genetics, University of Toronto, Toronto, Ontario, Canada.
*
*Corresponding author: Stephen W. Scherer, The Hospital for Sick Children, MaRS Centre - East Tower, 101 College Street, Room 14-701, Toronto, Ontario, M5G 1L7, Canada. E-mail: stephen.scherer@sickkids.ca

Abstract

During the past five years, copy number variation (CNV) has emerged as a highly prevalent form of genomic variation, bridging the interval between long-recognised microscopic chromosomal alterations and single-nucleotide changes. These genomic segmental differences among humans reflect the dynamic nature of genomes, and account for both normal variations among us and variations that predispose to conditions of medical consequence. Here, we place CNVs into their historical and medical contexts, focusing on how these variations can be recognised, documented, characterised and interpreted in clinical diagnostics. We also discuss how they can cause disease or influence adaptation to an environment. Various clinical exemplars are drawn out to illustrate salient characteristics and residual enigmas of CNVs, particularly the complexity of the data and information associated with CNVs relative to that of single-nucleotide variation. The potential is immense for CNVs to explain and predict disorders and traits that have long resisted understanding. However, creative solutions are needed to manage the sudden and overwhelming burden of expectation for laboratories and clinicians to assay and interpret these complex genomic variations as awareness permeates medical practice. Challenges remain for understanding the relationship between genomic changes and the phenotypes that might be predicted and prevented by such knowledge.

Type
Review Article
Copyright
Copyright © Cambridge University Press 2010

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

References

1Lejeune, J., Gautier, M. and Turpin, R. (1959) [Study of somatic chromosomes from 9 mongoloid children.]. Comptes rendus hebdomadaires des séances de l'Académie des sciences 248, 1721-1722 [Article in French]Google Scholar
2Iafrate, A.J. et al. (2004) Detection of large-scale variation in the human genome. Nature Genetics 36, 949-951Google Scholar
3Sebat, J. et al. (2004) Large-scale copy number polymorphism in the human genome. Science 305, 525-528Google Scholar
4Ford, C.E. et al. (1959) A sex-chromosome anomaly in a case of gonadal dysgenesis (Turner's syndrome). Lancet 1, 711-713Google Scholar
5Jacobs, P.A. and Strong, J.A. (1959) A case of human intersexuality having a possible XXY sex-determining mechanism. Nature 183, 302-303Google Scholar
6Feuk, L., Carson, A.R. and Scherer, S.W. (2006) Structural variation in the human genome. Nature Reviews Genetics 7, 85-97Google Scholar
7Scherer, S.W. et al. (2007) Challenges and standards in integrating surveys of structural variation. Nature Genetics 39, S7-15Google Scholar
8Lee, C., Iafrate, A.J. and Brothman, A.R. (2007) Copy number variations and clinical cytogenetic diagnosis of constitutional disorders. Nature Genetics 39, S48-54Google Scholar
9Conrad, D.F. et al. (2006) A high-resolution survey of deletion polymorphism in the human genome. Nature Genetics 38, 75-81Google Scholar
10Khaja, R. et al. (2006) Genome assembly comparison identifies structural variants in the human genome. Nature Genetics 38, 1413-1418Google Scholar
11Beckmann, J.S., Estivill, X. and Antonarakis, S.E. (2007) Copy number variants and genetic traits: closer to the resolution of phenotypic to genotypic variability. Nature Reviews Genetics 8, 639-646Google Scholar
12Wain, L.V., Armour, J.A. and Tobin, M.D. (2009) Genomic copy number variation, human health, and disease. Lancet 374, 340-350Google Scholar
13Conrad, D.F. et al. (2009) Origins and functional impact of copy number variation in the human genome. Nature Oct 7; [Epub ahead of print]Google Scholar
14Buchanan, J.A. and Scherer, S.W. (2008) Contemplating effects of genomic structural variation. Genetics in Medicine 10, 639-647Google Scholar
15Varki, A., Geschwind, D.H. and Eichler, E.E. (2008) Explaining human uniqueness: genome interactions with environment, behaviour and culture. Nature Reviews Genetics 9, 749-763Google Scholar
16Redon, R. et al. (2006) Global variation in copy number in the human genome. Nature 444, 444-454Google Scholar
17Emanuel, B.S. and Shaikh, T.H. (2001) Segmental duplications: an ‘expanding’ role in genomic instability and disease. Nature Reviews Genetics 2, 791-800Google Scholar
18Scherer, S.W. et al. (2003) Human chromosome 7: DNA sequence and biology. Science 300, 767-772Google Scholar
19Fredman, D. et al. (2004) Complex SNP-related sequence variation in segmental genome duplications. Nature Genetics 36, 861-866Google Scholar
20Cheung, J. et al. (2003) Genome-wide detection of segmental duplications and potential assembly errors in the human genome sequence. Genome Biology 4, R25Google Scholar
21Kim, P.M. et al. (2008) Analysis of copy number variants and segmental duplications in the human genome: Evidence for a change in the process of formation in recent evolutionary history. Genome Research 18, 1865-1874Google Scholar
22Stankiewicz, P. and Lupski, J.R. (2002) Genome architecture, rearrangements and genomic disorders. Trends in Genetics 18, 74-82Google Scholar
23Lupski, J.R. et al. (1991) DNA duplication associated with Charcot-Marie-Tooth disease type 1A. Cell 66, 219-232Google Scholar
24Emanuel, B.S. and Saitta, S.C. (2007) From microscopes to microarrays: dissecting recurrent chromosomal rearrangements. Nature Reviews Genetics 8, 869-883Google Scholar
25Moore, J.K. and Haber, J.E. (1996) Cell cycle and genetic requirements of two pathways of nonhomologous end-joining repair of double-strand breaks in Saccharomyces cerevisiae. Molecular and Cellular Biology 16, 2164-2173Google Scholar
26Conrad, D.F. and Hurles, M.E. (2007) The population genetics of structural variation. Nature Genetics 39, S30-36Google Scholar
27Zhang, F. et al. (2009) The DNA replication FoSTeS/MMBIR mechanism can generate genomic, genic and exonic complex rearrangements in humans. Nature Genetics 41, 849-853Google Scholar
28Lee, J.A., Carvalho, C.M. and Lupski, J.R. (2007) A DNA replication mechanism for generating nonrecurrent rearrangements associated with genomic disorders. Cell 131, 1235-1247Google Scholar
29Carvalho, C.M. et al. (2009) Complex rearrangements in patients with duplications of MECP2 can occur by fork stalling and template switching. Human Molecular Genetics 18, 2188-2203Google Scholar
30Hastings, P.J. et al. (2009) Mechanisms of change in gene copy number. Nature Reviews Genetics 10, 551-564Google Scholar
31Hastings, P.J., Ira, G. and Lupski, J.R. (2009) A microhomology-mediated break-induced replication model for the origin of human copy number variation. PLoS Genetics 5, e1000327Google Scholar
32Koolen, D.A. et al. (2006) A new chromosome 17q21.31 microdeletion syndrome associated with a common inversion polymorphism. Nature Genetics 38, 999-1001Google Scholar
33Shaw-Smith, C. et al. (2006) Microdeletion encompassing MAPT at chromosome 17q21.3 is associated with developmental delay and learning disability. Nature Genetics 38, 1032-1037Google Scholar
34Stefansson, H. et al. (2005) A common inversion under selection in Europeans. Nature Genetics 37, 129-137Google Scholar
35Visser, R. et al. (2005) Identification of a 3.0-kb major recombination hotspot in patients with Sotos syndrome who carry a common 1.9-Mb microdeletion. American Journal of Human Genetics 76, 52-67Google Scholar
36Osborne, L.R. et al. (2001) A 1.5 million-base pair inversion polymorphism in families with Williams-Beuren syndrome. Nature Genetics 29, 321-325Google Scholar
37Scherer, S.W. and Osborne, L.R. (2006) Williams-Beuren syndrome. In Genomic Disorders: The Genomic Basis of Disease ( Lupski, J.R. and Stankiewicz, P., eds), pp. 221-236, Humana Press, Totowa, NJ, USAGoogle Scholar
38McCarroll, S.A. et al. (2008) Integrated detection and population-genetic analysis of SNPs and copy number variation. Nature Genetics 40, 1166-1174Google Scholar
39Zhang, J. et al. (2006) Development of bioinformatics resources for display and analysis of copy number and other structural variants in the human genome. Cytogenetic and Genome Research 115, 205-214Google Scholar
40Levy, S. et al. (2007) The diploid genome sequence of an individual human. PLoS Biology 5, e254Google Scholar
41Wheeler, D.A. et al. (2008) The complete genome of an individual by massively parallel DNA sequencing. Nature 452, 872-876Google Scholar
42Lander, E.S. et al. (2001) Initial sequencing and analysis of the human genome. Nature 409, 860-921Google Scholar
43International Human Genome Sequencing Consortium (2004) Finishing the euchromatic sequence of the human genome. Nature 431, 931-945Google Scholar
44Venter, J.C. et al. (2001) The sequence of the human genome. Science 291, 1304-1351Google Scholar
45Maher, B. (2008) Personal genomes: the case of the missing heritability. Nature 456, 18-21Google Scholar
46Manolio, T.A. et al. (2009) Finding the missing heritability of complex diseases. Nature 461, 747-753Google Scholar
47Cook, E.H. Jr, and Scherer, S.W. (2008) Copy-number variations associated with neuropsychiatric conditions. Nature 455, 919-923Google Scholar
48Carter, N.P. (2007) Methods and strategies for analyzing copy number variation using DNA microarrays. Nature Genetics 39, S16-21Google Scholar
49Wang, J. et al. (2008) The diploid genome sequence of an Asian individual. Nature 456, 60-65Google Scholar
50Bentley, D.R. et al. (2008) Accurate whole human genome sequencing using reversible terminator chemistry. Nature 456, 53-59Google Scholar
51Ahn, S.M. et al. (2009) The first Korean genome sequence and analysis: full genome sequencing for a socio-ethnic group. Genome Research 19, 1622-1629Google Scholar
52Kim, J.I. et al. (2009) A highly annotated whole-genome sequence of a Korean individual. Nature 460, 1011-1015Google Scholar
53McKernan, K.J. et al. (2009) Sequence and structural variation in a human genome uncovered by short-read, massively parallel ligation sequencing using two-base encoding. Genome Research 19, 1527-1541Google Scholar
54Drmanac, R. et al. Human genome sequencing using unchained base reads on self-assembling DNA nanoarrays. Science 327, 78-81Google Scholar
55Alkan, C. et al. (2009) Personalized copy number and segmental duplication maps using next-generation sequencing. Nature Genetics 41, 1061-1067Google Scholar
56Chiang, D.Y. et al. (2009) High-resolution mapping of copy-number alterations with massively parallel sequencing. Nature Methods 6, 99-103Google Scholar
57Lupski, J.R. (1998) Genomic disorders: structural features of the genome can lead to DNA rearrangements and human disease traits. Trends in Genetics 14, 417-422Google Scholar
58Hüffmeier, U. et al. (2009) Replication of LCE3C-LCE3B CNV as a risk factor for psoriasis and analysis of interaction with other genetic risk factors. Journal of Investigative Dermatology Dec 17; [Epub ahead of print]Google Scholar
59de Cid, R. et al. (2009) Deletion of the late cornified envelope LCE3B and LCE3C genes as a susceptibility factor for psoriasis. Nature Genetics 41, 211-215Google Scholar
60Shlien, A. et al. (2008) Excessive genomic DNA copy number variation in the Li-Fraumeni cancer predisposition syndrome. Proceedings of the National Academy of Sciences of the United States of America 105, 11264-11269Google Scholar
61Diskin, S.J. et al. (2009) Copy number variation at 1q21.1 associated with neuroblastoma. Nature 459, 987-991Google Scholar
62Lupski, J.R. (2007) Genomic rearrangements and sporadic disease. Nature Genetics 39, S43-47Google Scholar
63Cahan, P. et al. (2009) The impact of copy number variation on local gene expression in mouse hematopoietic stem and progenitor cells. Nature Genetics 41, 430-437Google Scholar
64Henrichsen, C.N., Chaignat, E. and Reymond, A. (2009) Copy number variants, diseases and gene expression. Human Molecular Genetics 18, R1-8Google Scholar
65Dathe, K. et al. (2009) Duplications involving a conserved regulatory element downstream of BMP2 are associated with brachydactyly type A2. American Journal of Human Genetics 84, 483-492Google Scholar
66Stranger, B.E. et al. (2007) Relative impact of nucleotide and copy number variation on gene expression phenotypes.Science 315, 848-853Google Scholar
67Merla, G. et al. (2006) Submicroscopic deletion in patients with Williams-Beuren syndrome influences expression levels of the nonhemizygous flanking genes. American Journal of Human Genetics 79, 332-341Google Scholar
68Firth, H.V. et al. (2009) DECIPHER: Database of Chromosomal Imbalance and Phenotype in Humans Using Ensembl Resources. American Journal of Human Genetics 84, 524-533Google Scholar
69Miller, D.T. et al. (2009) Microdeletion/duplication at 15q13.2q13.3 among individuals with features of autism and other neuropsychiatric disorders. Journal of Medical Genetics 46, 242-248Google Scholar
70Sharp, A.J. et al. (2008) A recurrent 15q13.3 microdeletion syndrome associated with mental retardation and seizures. Nature Genetics 40, 322-328Google Scholar
71O'Donovan, M.C., Kirov, G. and Owen, M.J. (2008) Phenotypic variations on the theme of CNVs. Nature Genetics 40, 1392-1393Google Scholar
72Brunetti-Pierri, N. et al. (2008) Recurrent reciprocal 1q21.1 deletions and duplications associated with microcephaly or macrocephaly and developmental and behavioral abnormalities. Nature Genetics 40, 1466-1471Google Scholar
73Mefford, H.C. et al. (2008) Recurrent rearrangements of chromosome 1q21.1 and variable pediatric phenotypes. New England Journal of Medicine 359, 1685-1699Google Scholar
74Carlson, C. et al. (1997) Molecular definition of 22q11 deletions in 151 velo-cardio-facial syndrome patients. American Journal of Human Genetics 61, 620-629Google Scholar
75Driscoll, D.A. et al. (1992) Deletions and microdeletions of 22q11.2 in velo-cardio-facial syndrome. American Journal of Medical Genetics 44, 261-268Google Scholar
76Coppinger, J. et al. (2009) Identification of familial and de novo microduplications of 22q11.21-q11.23 distal to the 22q11.21 microdeletion syndrome region. Human Molecular Genetics 18, 1377-1383Google Scholar
77Klopocki, E. et al. (2007) Complex inheritance pattern resembling autosomal recessive inheritance involving a microdeletion in thrombocytopenia-absent radius syndrome. American Journal of Human Genetics 80, 232-240Google Scholar
78Uhrig, S. et al. (2007) Impact of array comparative genomic hybridization-derived information on genetic counseling demonstrated by prenatal diagnosis of the TAR (thrombocytopenia-absent-radius) syndrome-associated microdeletion 1q21.1. American Journal of Human Genetics 81, 866-868Google Scholar
79Prior, T.W. (2007) Spinal muscular atrophy diagnostics. Journal of Child Neurology 22, 952-956Google Scholar
80Schonherr, N. et al. (2007) The centromeric 11p15 imprinting centre is also involved in Silver-Russell syndrome. Journal of Medical Genetics 44, 59-63Google Scholar
81Feuk, L. et al. (2006) Absence of a paternally inherited FOXP2 gene in developmental verbal dyspraxia. American Journal of Human Genetics 79, 965-972Google Scholar
82Bruder, C.E. et al. (2008) Phenotypically concordant and discordant monozygotic twins display different DNA copy-number-variation profiles. American Journal of Human Genetics 82, 763-771Google Scholar
83Piotrowski, A. et al. (2008) Somatic mosaicism for copy number variation in differentiated human tissues. Human Mutation 29, 1118-1124Google Scholar
84Gervasini, C. et al. (2007) High frequency of mosaic CREBBP deletions in Rubinstein-Taybi syndrome patients and mapping of somatic and germ-line breakpoints. Genomics 90, 567-573Google Scholar
85Roelfsema, J.H. and Peters, D.J. (2007) Rubinstein-Taybi syndrome: clinical and molecular overview. Expert Reviews in Molecular Medicine 9, 1-16Google Scholar
86Schorry, E.K. et al. (2008) Genotype-phenotype correlations in Rubinstein-Taybi syndrome. American Journal of Medical Genetics Part A 146A, 2512-2519Google Scholar
87Kozlowski, P. et al. (2007) Identification of 54 large deletions/duplications in TSC1 and TSC2 using MLPA, and genotype-phenotype correlations. Human Genetics 121, 389-400Google Scholar
88Robinson, D.O. et al. (2008) Genetic analysis of chromosome 11p13 and the PAX6 gene in a series of 125 cases referred with aniridia. American Journal of Medical Genetics Part A 146A, 558-569Google Scholar
89Kirov, G. et al. (2009) Support for the involvement of large copy number variants in the pathogenesis of schizophrenia. Human Molecular Genetics 18, 1497-1503Google Scholar
90Greenway, S.C. et al. (2009) De novo copy number variants identify new genes and loci in isolated sporadic tetralogy of Fallot. Nature Genetics 41, 931-935Google Scholar
91Need, A.C. et al. (2009) A genome-wide investigation of SNPs and CNVs in schizophrenia. PLoS Genetics 5, e1000373Google Scholar
92Lyle, R. et al. (2009) Genotype-phenotype correlations in Down syndrome identified by array CGH in 30 cases of partial trisomy and partial monosomy chromosome 21. European Journal of Human Genetics 17, 454-466Google Scholar
93Korbel, J.O. et al. (2009) The genetic architecture of Down syndrome phenotypes revealed by high-resolution analysis of human segmental trisomies. Proceedings of the National Academy of Sciences of the United States of America 106, 12031-12036Google Scholar
94Reisman, L.E. et al. (1966) Anti-mongolism. Studies in an infant with a partial monosomy of the 21 chromosome. Lancet 1, 394-397Google Scholar
95Ewart, A.K. et al. (1993) Hemizygosity at the elastin locus in a developmental disorder, Williams syndrome. Nature Genetics 5, 11-16Google Scholar
96Osborne, L.R. and Mervis, C.B. (2007) Rearrangements of the Williams-Beuren syndrome locus: molecular basis and implications for speech and language development. Expert Reviews in Molecular Medicine 9, 1-16Google Scholar
97Lupski, J.R. (2009) Genomic disorders ten years on. Genome Medicine 1, 42Google Scholar
98Somerville, M.J. et al. (2005) Severe expressive-language delay related to duplication of the Williams-Beuren locus. New England Journal of Medicine 353, 1694-1701Google Scholar
99Kriek, M. et al. (2006) Copy number variation in regions flanked (or unflanked) by duplicons among patients with developmental delay and/or congenital malformations; detection of reciprocal and partial Williams-Beuren duplications. European Journal of Human Genetics 14, 180-189Google Scholar
100Torniero, C. et al. (2008) Dysmorphic features, simplified gyral pattern and 7q11.23 duplication reciprocal to the Williams-Beuren deletion. European Journal of Human Genetics 16, 880-887Google Scholar
101Sharp, A.J. et al. (2006) Discovery of previously unidentified genomic disorders from the duplication architecture of the human genome. Nature Genetics 38, 1038-1042Google Scholar
102Christian, S.L. et al. (1999) Large genomic duplicons map to sites of instability in the Prader-Willi/Angelman syndrome chromosome region (15q11-q13). Human Molecular Genetics 8, 1025-1037Google Scholar
103Mignon-Ravix, C. et al. (2007) Recurrent rearrangements in the proximal 15q11-q14 region: a new breakpoint cluster specific to unbalanced translocations. European Journal of Human Genetics 15, 432-440Google Scholar
104Sahoo, T. et al. (2005) Array-based comparative genomic hybridization analysis of recurrent chromosome 15q rearrangements. American Journal of Medical Genetics A 139A, 106-113Google Scholar
105Sharp, A.J. et al. (2008) A recurrent 15q13.3 microdeletion syndrome associated with mental retardation and seizures. Nature Genetics 40, 322-328Google Scholar
106Stefansson, H. et al. (2008) Large recurrent microdeletions associated with schizophrenia. Nature 455, 232-236Google Scholar
107Pagnamenta, A.T. et al. (2009) A 15q13.3 microdeletion segregating with autism. European Journal of Human Genetics 17, 687-692Google Scholar
108Shinawi, M. et al. (2009) A small recurrent deletion within 15q13.3 is associated with a range of neurodevelopmental phenotypes. Nature Genetics 41, 1269-1271Google Scholar
109Helbig, I. et al. (2009) 15q13.3 microdeletions increase risk of idiopathic generalized epilepsy. Nature Genetics 41, 160-162Google Scholar
110Ben-Shachar, S. et al. (2009) Microdeletion 15q13.3: a locus with incomplete penetrance for autism, mental retardation, and psychiatric disorders. Journal of Medical Genetics 46, 382-388Google Scholar
111Pagnamenta, A.T. et al. (2009) A 15q13.3 microdeletion segregating with autism. European Journal of Human Genetics 17, 687-692Google Scholar
112van Bon, B.W. et al. (2009) Further delineation of the 15q13 microdeletion and duplication syndromes: a clinical spectrum varying from non-pathogenic to a severe outcome. Journal of Medical Genetics 46, 511-523Google Scholar
113Consortium, I.S. (2008) Rare chromosomal deletions and duplications increase risk of schizophrenia. Nature 455, 237-241Google Scholar
114Helbig, I. et al. (2009) 15q13.3 microdeletions increase risk of idiopathic generalized epilepsy. Nature Genetics 41, 160-162Google Scholar
115Buchanan, J.A. et al. (2009) The cycle of genome-directed medicine. Genome Medicine 1, 16Google Scholar
116Ali-Khan, S.E. et al. (2009) Whole genome scanning: resolving clinical diagnosis and management amidst complex data. Pediatric Research 66, 357-363Google Scholar
117Christiansen, J. et al. (2004) Chromosome 1q21.1 contiguous gene deletion is associated with congenital heart disease. Circulation Research 94, 1429-1435Google Scholar
118Szatmari, P. et al. (2007) Mapping autism risk loci using genetic linkage and chromosomal rearrangements. Nature Genetics 39, 319-328Google Scholar
119Mefford, H.C. et al. (2009) A method for rapid, targeted CNV genotyping identifies rare variants associated with neurocognitive disease. Genome Research 19, 1579-1585Google Scholar
120Pinto, D. et al. (2007) Copy-number variation in control population cohorts. Human Molecular Genetics 16 (Spec No. 2), R168-173Google Scholar
121Armengol, L., Rabionet, R. and Estivill, X. (2008) The emerging role of structural variations in common disorders: initial findings and discovery challenges. Cytogenetic and Genome Research 123, 108-117Google Scholar
122Schaschl, H., Aitman, T.J. and Vyse, T.J. (2009) Copy number variation in the human genome and its implication in autoimmunity. Clinical and Experimental Immunology 156, 12-16Google Scholar
123Ionita-Laza, I. et al. (2009) Genetic association analysis of copy-number variation (CNV) in human disease pathogenesis. Genomics 93, 22-26Google Scholar
124Hollox, E.J., Detering, J.C. and Dehnugara, T. (2009) An integrated approach for measuring copy number variation at the FCGR3 (CD16) locus. Human Mutation 30, 477-484Google Scholar
125Nuytten, H. et al. (2009) Accurate determination of copy number variations (CNVs): application to the alpha- and beta-defensin CNVs. Journal of Immunological Methods 344, 35-44Google Scholar
126McCarroll, S.A. et al. (2008) Deletion polymorphism upstream of IRGM associated with altered IRGM expression and Crohn's disease. Nature Genetics 40, 1107-1112Google Scholar
127Gonzalez, E. et al. (2005) The influence of CCL3L1 gene-containing segmental duplications on HIV-1/AIDS susceptibility. Science 307, 1434-1440Google Scholar
128Kulkarni, H. et al. (2008) CCL3L1-CCR5 genotype improves the assessment of AIDS Risk in HIV-1-infected individuals. PLoS One 3, e3165Google Scholar
129Shostakovich-Koretskaya, L. et al. (2009) Combinatorial content of CCL3L and CCL4L gene copy numbers influence HIV-AIDS susceptibility in Ukrainian children. AIDS 23, 679-688Google Scholar
130McKinney, C. et al. (2008) Evidence for an influence of chemokine ligand 3-like 1 (CCL3L1) gene copy number on susceptibility to rheumatoid arthritis. Annals of the Rheumatic Diseases 67, 409-413Google Scholar
131Colobran, R. et al. (2009) Copy number variation in the CCL4L gene is associated with susceptibility to acute rejection in lung transplantation. Genes and Immunity 10, 254-259Google Scholar
132McCarroll, S.A. et al. (2009) Donor-recipient mismatch for common gene deletion polymorphisms in graft-versus-host disease. Nature Genetics 41, 1341-1344Google Scholar
133Shlien, A. and Malkin, D. (2009) Copy number variations and cancer. Genome Medicine 1, 62Google Scholar
134Marcel, V. et al. (2009) TP53 PIN3 and MDM2 SNP309 polymorphisms as genetic modifiers in the Li-Fraumeni syndrome: impact on age at first diagnosis. Journal of Medical Genetics 46, 766-772Google Scholar
135Schwarzbraun, T. et al. (2009) Predictive diagnosis of the cancer prone Li-Fraumeni syndrome by accident: new challenges through whole genome array testing. Journal of Medical Genetics 46, 341-344Google Scholar
136Adam, M.P. et al. (2009) Clinical utility of array comparative genomic hybridization: uncovering tumor susceptibility in individuals with developmental delay. Journal of Pediatrics 154, 143-146Google Scholar
137Adams, S.A. et al. (2009) Impact of genotype-first diagnosis: the detection of microdeletion and microduplication syndromes with cancer predisposition by aCGH. Genetics in Medicine 11, 314-322Google Scholar
138Speicher, M.R. and Carter, N.P. (2005) The new cytogenetics: blurring the boundaries with molecular biology. Nature Reviews Genetics 6, 782-792Google Scholar
139Stuhrmann, M. et al. (2009) Testing the parents to confirm genotypes of CF patients is highly recommended: report of two cases. European Journal of Human Genetics 17, 417-419Google Scholar
140Girardet, A. et al. (2007) Negative genetic neonatal screening for cystic fibrosis caused by compound heterozygosity for two large CFTR rearrangements. Clinical Genetics 72, 374-377Google Scholar
141Tomaiuolo, R. et al. (2008) Epidemiology and a novel procedure for large scale analysis of CFTR rearrangements in classic and atypical CF patients: a multicentric Italian study. Journal of Cystic Fibrosis 7, 347-351Google Scholar
142McDevitt, T. and Barton, D. (2009) When good CF tests go bad. European Journal of Human Genetics 17, 403-405Google Scholar
143Vissers, L.E. et al. (2004) Mutations in a new member of the chromodomain gene family cause CHARGE syndrome. Nature Genetics 36, 955-957Google Scholar
144Fernandez, B.A. et al. (2009) Phenotypic spectrum associated with de novo and inherited deletions and duplications at 16p11.2 in individuals ascertained for diagnosis of autism spectrum disorder. Journal of Medical Genetics Sep 24; [Epub ahead of print]Google Scholar
145Kumar, R.A. et al. (2008) Recurrent 16p11.2 microdeletions in autism. Human Molecular Genetics 17, 628-638Google Scholar
146Marshall, C.R. et al. (2008) Structural variation of chromosomes in autism spectrum disorder. American Journal of Human Genetics 82, 477-488Google Scholar
147Weiss, L.A. et al. (2008) Association between microdeletion and microduplication at 16p11.2 and autism. New England Journal of Medicine 358, 667-675Google Scholar
148Shinawi, M. et al. (2009) Recurrent reciprocal 16p11.2 rearrangements associated with global developmental delay, behavioral problems, dysmorphism, epilepsy, and abnormal head size. Journal of Medical Genetics Nov 12; [Epub ahead of print]Google Scholar
149McCarthy, S.E. et al. (2009) Microduplications of 16p11.2 are associated with schizophrenia. Nature Genetics 41, 1223-1227Google Scholar
150Bochukova, E.G. et al. (2010) Large, rare chromosomal deletions associated with severe early-onset obesity. Nature 463, 666-670Google Scholar
151Perry, G.H. et al. (2008) The fine-scale and complex architecture of human copy-number variation. American Journal of Human Genetics 82, 685-695Google Scholar
152Zogopoulos, G. et al. (2007) Germ-line DNA copy number variation frequencies in a large North American population. Human Genetics 122, 345-353Google Scholar
153Jakobsson, M. et al. (2008) Genotype, haplotype and copy-number variation in worldwide human populations. Nature 451, 998-1003Google Scholar
154Armengol, L. et al. (2009) Identification of copy number variants defining genomic differences among major human groups. PLoS One 4, e7230Google Scholar
155Matsuzaki, H. et al. (2009) High resolution discovery and confirmation of copy number variants in 90 Yoruba Nigerians. Genome Biology 10, R125Google Scholar
156Yim, S.H. et al. (2010) Copy number variations in East-Asian population and their evolutionary and functional implications. Human Molecular Genetics Jan 15; [Epub ahead of print]Google Scholar
157Brookes, A.J. et al. (2009) Genomic variation in a global village: report of the 10th annual Human Genome Variation Meeting 2008. Human Mutation 30, 1134-1138Google Scholar
158Stankiewicz, P. and Lupski, J.R. (2010) Structural variation in the human genome and its role in disease. Annual Review of Medicine 61, 437-455Google Scholar
159Ilbery, P.L., Lee, C.W. and Winn, S.M. (1961) Incomplete trisomy in a mongoloid child exhibiting minimal stigmata. Medical Journal of Australia 48, 182-184Google Scholar
160Lejeune, J. et al. (1963) [3 cases of partial deletion of the short arm of a 5 chromosome.] Comptes rendus hebdomadaires des séances de l'Académie des sciences 257, 3098-3102 [Article in French]Google Scholar
161Caspersson, T. et al. (1969) Chemical differentiation with fluorescent alkylating agents in Vicia faba metaphase chromosomes. Experimental Cell Research 58, 128-140Google Scholar
162Orkin, S.H. (1978) The duplicated human alpha globin genes lie close together in cellular DNA. Proceedings of the National Academy of Sciences of the United States of America 75, 5950-5954Google Scholar
163Wyman, A.R. and White, R. (1980) A highly polymorphic locus in human DNA. Proceedings of the National Academy of Sciences of the United States of America 77, 6754-6758Google Scholar
164Bauman, J.G. et al. (1980) A new method for fluorescence microscopical localization of specific DNA sequences by in situ hybridization of fluorochromelabelled RNA. Experimental Cell Research 128, 485-490Google Scholar
165Van Prooijen-Knegt, A.C. et al. (1982) In situ hybridization of DNA sequences in human metaphase chromosomes visualized by an indirect fluorescent immunocytochemical procedure. Experimental Cell Research 141, 397-407Google Scholar
166Jeffreys, A.J., Wilson, V. and Thein, S.L. (1985) Individual-specific ‘fingerprints’ of human DNA. Nature 316, 76-79Google Scholar
167Monaco, A.P. et al. (1985) Detection of deletions spanning the Duchenne muscular dystrophy locus using a tightly linked DNA segment. Nature 316, 842-845Google Scholar
168Ray, P.N. et al. (1985) Cloning of the breakpoint of an X;21 translocation associated with Duchenne muscular dystrophy. Nature 318, 672-675Google Scholar
169Schmickel, R.D. (1986) Contiguous gene syndromes: a component of recognizable syndromes. Journal of Pediatrics 109, 231-241Google Scholar
170Kallioniemi, A. et al. (1992) Comparative genomic hybridization for molecular cytogenetic analysis of solid tumors. Science 258, 818-821Google Scholar
171 [No authors listed] (1996) A complete set of human telomeric probes and their clinical application. National Institutes of Health and Institute of Molecular Medicine collaboration. Nature Genetics 14, 86-89Google Scholar
172Pinkel, D. et al. (1998) High resolution analysis of DNA copy number variation using comparative genomic hybridization to microarrays. Nature Genetics 20, 207-211Google Scholar
173Stockley, T.L. et al. (2006) Strategy for comprehensive molecular testing for Duchenne and Becker muscular dystrophies. Genetic Testing 10, 229-243Google Scholar
174White, S.J. and den Dunnen, J.T. (2006) Copy number variation in the genome; the human DMD gene as an example. Cytogenetic and Genome Research 115, 240-246Google Scholar
175De Luca, A. et al. (2007) Deletions of NF1 gene and exons detected by multiplex ligation-dependent probe amplification. Journal of Medical Genetics 44, 800-808Google Scholar
176Raedt, T.D. et al. (2006) Conservation of hotspots for recombination in low-copy repeats associated with the NF1 microdeletion. Nature Genetics 38, 1419-1423Google Scholar
177Wimmer, K. et al. (2006) Spectrum of single- and multiexon NF1 copy number changes in a cohort of 1,100 unselected NF1 patients. Genes, Chromosomes & Cancer 45, 265-276Google Scholar
178Saugier-Veber, P. et al. (2007) Heterogeneity of NSD1 alterations in 116 patients with Sotos syndrome. Human Mutation 28, 1098-1107Google Scholar
179Fagali, C. et al. (2009) MLPA analysis in 30 Sotos syndrome patients revealed one total NSD1 deletion and two partial deletions not previously reported. European Journal of Medical Genetics 52, 333-336Google Scholar
180Woodward, K.J. (2008) The molecular and cellular defects underlying Pelizaeus-Merzbacher disease. Expert Reviews in Molecular Medicine 10, e14Google Scholar
181Brouwers, N. et al. (2006) Genetic risk and transcriptional variability of amyloid precursor protein in Alzheimer's disease. Brain 129, 2984-2991Google Scholar
182Rovelet-Lecrux, A. et al. (2006) APP locus duplication causes autosomal dominant early-onset al.zheimer disease with cerebral amyloid angiopathy. Nature Genetics 38, 24-26Google Scholar
183Sleegers, K. et al. (2006) APP duplication is sufficient to cause early onset al.zheimer's dementia with cerebral amyloid angiopathy. Brain 129, 2977-2983Google Scholar
184Durand, C.M. et al. (2007) Mutations in the gene encoding the synaptic scaffolding protein SHANK3 are associated with autism spectrum disorders. Nature Genetics 39, 25-27Google Scholar
185Moessner, R. et al. (2007) Contribution of SHANK3 mutations to autism spectrum disorder. American Journal of Human Genetics 81, 1289-1297Google Scholar
186Gauthier, J. et al. (2009) Novel de novo SHANK3 mutation in autistic patients. American Journal of Medical Genetics Part B, Neuropsychiatric Genetics 150B, 421-424Google Scholar
187Fantes, J. et al. (1995) Aniridia-associated cytogenetic rearrangements suggest that a position effect may cause the mutant phenotype. Human Molecular Genetics 4, 415-422Google Scholar
188Klopocki, E. et al. (2008) A microduplication of the long range SHH limb regulator (ZRS) is associated with triphalangeal thumb-polysyndactyly syndrome. Journal of Medical Genetics 45, 370-375Google Scholar
189Sun, M. et al. (2008) Triphalangeal thumb-polysyndactyly syndrome and syndactyly type IV are caused by genomic duplications involving the long range, limb-specific SHH enhancer. Journal of Medical Genetics 45, 589-595Google Scholar
190Fellermann, K. et al. (2006) A chromosome 8 gene-cluster polymorphism with low human beta-defensin 2 gene copy number predisposes to Crohn disease of the colon. American Journal of Human Genetics 79, 439-448Google Scholar
191Hollox, E.J. et al. (2008) Defensins and the dynamic genome: what we can learn from structural variation at human chromosome band 8p23.1. Genome Research 18, 1686-1697Google Scholar
192Aitman, T.J. et al. (2006) Copy number polymorphism in Fcgr3 predisposes to glomerulonephritis in rats and humans. Nature 439, 851-855Google Scholar
193Fanciulli, M. et al. (2007) FCGR3B copy number variation is associated with susceptibility to systemic, but not organ-specific, autoimmunity. Nature Genetics 39, 721-723Google Scholar
194Ibanez, P. et al. (2009) Alpha-synuclein gene rearrangements in dominantly inherited parkinsonism: frequency, phenotype, and mechanisms. Archives of Neurology 66, 102-108Google Scholar
195Burns, J.C. et al. (2005) Genetic variations in the receptor-ligand pair CCR5 and CCL3L1 are important determinants of susceptibility to Kawasaki disease. Journal of Infectious Diseases 192, 344-349Google Scholar
196Saitta, S.C. et al. (2004) Aberrant interchromosomal exchanges are the predominant cause of the 22q11.2 deletion. Human Molecular Genetics 13, 417-428Google Scholar
197Shaikh, T.H. et al. (2007) Low copy repeats mediate distal chromosome 22q11.2 deletions: sequence analysis predicts breakpoint mechanisms. Genome Research 17, 482-491Google Scholar
198Berg, J.S. et al. (2007) Speech delay and autism spectrum behaviors are frequently associated with duplication of the 7q11.23 Williams-Beuren syndrome region. Genetics in Medicine 9, 427-441Google Scholar
199Cusco, I. et al. (2008) Copy number variation at the 7q11.23 segmental duplications is a susceptibility factor for the Williams-Beuren syndrome deletion. Genome Research 18, 683-694Google Scholar
200Grisart, B. et al. (2009) 17q21.31 microduplication patients are characterised by behavioural problems and poor social interaction. Journal of Medical Genetics 46, 524-530Google Scholar
201Kirchhoff, M. et al. (2007) A 17q21.31 microduplication, reciprocal to the newly described 17q21.31 microdeletion, in a girl with severe psychomotor developmental delay and dysmorphic craniofacial features. European Journal of Medical Genetics 50, 256-263Google Scholar
202Shaffer, L.G. et al. (2007) The discovery of microdeletion syndromes in the post-genomic era: review of the methodology and characterization of a new 1q41q42 microdeletion syndrome. Genetics in Medicine 9, 607-616Google Scholar
203Ballif, B.C. et al. (2007) Discovery of a previously unrecognized microdeletion syndrome of 16p11.2-p12.2. Nature Genetics 39, 1071-1073Google Scholar
204Ghebranious, N. et al. (2007) A novel microdeletion at 16p11.2 harbors candidate genes for aortic valve development, seizure disorder, and mild mental retardation. American Journal of Medical Genetics Part A 143A, 1462-1471Google Scholar
205Ballif, B.C. et al. (2008) Expanding the clinical phenotype of the 3q29 microdeletion syndrome and characterization of the reciprocal microduplication. Molecular Cytogenetics 1, 8Google Scholar
206Goobie, S. et al. (2008) Molecular and clinical characterization of de novo and familial cases with microduplication 3q29: guidelines for copy number variation case reporting. Cytogenetic and Genome Research 123, 65-78Google Scholar
207Potocki, L. et al. (2007) Characterization of Potocki-Lupski syndrome (dup(17)(p11.2p11.2)) and delineation of a dosage-sensitive critical interval that can convey an autism phenotype. American Journal of Human Genetics 80, 633-649Google Scholar
208Tabor, H.K. and Cho, M.K. (2007) Ethical implications of array comparative genomic hybridization in complex phenotypes: points to consider in research. Genetics in Medicine 9, 626-631Google Scholar
209Ullmann, R. et al. (2007) Array CGH identifies reciprocal 16p13.1 duplications and deletions that predispose to autism and/or mental retardation. Human Mutation 28, 674-682Google Scholar
210Schaefer, G.B. and Mendelsohn, N.J. (2008) Genetics evaluation for the etiologic diagnosis of autism spectrum disorders. Genetics in Medicine 10, 4-12Google Scholar
211Glessner, J.T. et al. (2009) Autism genome-wide copy number variation reveals ubiquitin and neuronal genes. Nature 459, 569-573Google Scholar
212Abrahams, B.S. and Geschwind, D.H. (2008) Advances in autism genetics: on the threshold of a new neurobiology. Nature Reviews Genetics 9, 341-355Google Scholar
213Lachman, H.M. et al. (2007) Increase in GSK3beta gene copy number variation in bipolar disorder. American Journal of Medical Genetics Part B, Neuropsychiatric Genetics 144B, 259-265Google Scholar
214Burmeister, M., McInnis, M.G. and Zollner, S. (2008) Psychiatric genetics: progress amid controversy. Nature Reviews Genetics 9, 527-540Google Scholar
215Alaerts, M. and Del-Favero, J. (2009) Searching genetic risk factors for schizophrenia and bipolar disorder: learn from the past and back to the future. Human Mutation 30, 1139-1152Google Scholar
216Walsh, T. et al. (2008) Rare structural variants disrupt multiple genes in neurodevelopmental pathways in schizophrenia. Science 320, 539-543Google Scholar
217Xu, B. et al. (2008) Strong association of de novo copy number mutations with sporadic schizophrenia. Nature Genetics 40, 880-885Google Scholar
218Rujescu, D. et al. (2009) Disruption of the neurexin 1 gene is associated with schizophrenia. Human Molecular Genetics 18, 988-996Google Scholar
219Hughes, A.E. et al. (2006) A common CFH haplotype, with deletion of CFHR1 and CFHR3, is associated with lower risk of age-related macular degeneration. Nature Genetics 38, 1173-1177Google Scholar
220Maller, J. et al. (2006) Common variation in three genes, including a noncoding variant in CFH, strongly influences risk of age-related macular degeneration. Nature Genetics 38, 1055-1059Google Scholar
221Barber, J.C. et al. (2008) 8p23.1 duplication syndrome; a novel genomic condition with unexpected complexity revealed by array CGH. European Journal of Human Genetics 16, 18-27Google Scholar
222Hendrickson, B.C. et al. (2009) Differences in SMN1 allele frequencies among ethnic groups within North America. Journal of Medical Genetics 46, 641-644Google Scholar
223Alias, L. et al. (2009) Mutation update of spinal muscular atrophy in Spain: molecular characterization of 745 unrelated patients and identification of four novel mutations in the SMN1 gene. Human Genetics 125, 29-39Google Scholar
224Mantripragada, K.K. et al. (2009) Genome-wide high-resolution analysis of DNA copy number alterations in NF1-associated malignant peripheral nerve sheath tumors using 32K BAC array. Genes, Chromosomes & Cancer 48, 897-907Google Scholar

Further reading, resources and contacts

The Database of Chromosomal Imbalance and Phenotype in Humans using Ensembl Resources

(DECIPHER) provides tools that allow researchers to share information about copy number changes in patients: