Skip to main content
  • Review
  • Published:

Exploring the potential relevance of human-specific genes to complex disease

Abstract

Although human disease genes generally tend to be evolutionarily more ancient than non-disease genes, complex disease genes appear to be represented more frequently than Mendelian disease genes among genes of more recent evolutionary origin. It is therefore proposed that the analysis of human-specific genes might provide new insights into the genetics of complex disease. Cross-comparison with the Human Gene Mutation Database (http://www.hgmd.org) revealed a number of examples of disease-causing and disease-associated mutations in putatively human-specific genes. A sizeable proportion of these were missense polymorphisms associated with complex disease. Since both human-specific genes and genes associated with complex disease have often experienced particularly rapid rates of evolutionary change, either due to weaker purifying selection or positive selection, it is proposed that a significant number of human-specific genes may play a role in complex disease.

Human 'disease genes' have been known for some time to differ significantly from 'non-disease genes' in terms of their higher degree of evolutionary conservation [1–3]. Further, with respect to their evolutionary age, human disease genes appear not to be simply a random subset of all genes in the genome but are instead biased toward being of ancient (early metazoan) origin [4]. Concomitantly, a pronounced paucity of human lineage-specific genes is also evident among disease genes [4]. These initial findings were subsequently confirmed and elaborated upon by Cai et al.,[5] who determined the approximate age of evolutionary emergence of all human genes and then proceeded to compare disease genes with non-disease genes with respect to whether they were 'young', 'middle-aged' or 'old-aged'. For the purposes of their study, the origin of a given gene was determined by retracing its orthologues back to the species most distantly related to human. Genes that originated during the period since the adaptive radiation of the Laurasiatheria were described as 'young', the term 'middle-aged' was employed to describe those genes whose origin went back to the bony fish, and genes that emerged at some stage between yeast and Ciona (a tunicate) were ascribed the term 'old-aged'. Using these fairly crude descriptors of gene age, Cai et al.[5] confirmed that there was a tendency for Mendelian disease genes (ie those genes underlying single gene disorders) to be of more ancient evolutionary origin than non-disease genes. With Mendelian disease genes, the 'old-aged' genes were in the majority, closely followed by the 'middle-aged' genes. By contrast, most genes involved in the aetiology of complex disease were found to reside in the 'middle-aged' category. Although both Mendelian and complex disease genes were found to be under-represented in the 'young' category, the frequency of complex disease genes in this category was found to be more than twice that exhibited by the Mendelian disease genes [5]. Given this finding, we speculated that closer examination of the most recently acquired (ie human-specific) genes might well provide new insights into the genetics of complex disease.

Considerable efforts have been made to identify those genes that have been inactivated in the human lineage but which are still present in other higher primate species, including chimpanzee [6–9]. However, somewhat less attention has been paid so far either to human genes of relatively recent origin that are specific to the human lineage [10, 11] or to those genes that have been retained in the human genome despite having been lost in other primate species [12].

The first reported attempts to identify putative human-specific gene duplications were those of Fortna et al.[13] and Cheng et al.[14] The dataset of Cheng et al. comprised 88 complete gene duplications that were considered to have occurred since the divergence of human and chimpanzee. Of these, it was found that 13 have been reported in association with human inherited disease (see Human Gene Mutation Database (HGMD)[15]). These are the genes encoding Fc fragment of IgG, high affinity Ia, receptor (FCGR1A), DEAD box 1 protein (DDX11), cholinergic nicotinic acetylcholine receptor, alpha 7 subunit, exons 5-10 and family with sequence similarity 7A, exons A-E fusion (CHRFAM7A), aryl sulphotransferase (SULT1A1), CC chemokine ligand 4-like 1 (CCL4L1), killer cell immunoglobulin-like receptor, three domains, long cytoplasmic tail, 1 (KIR3DL1), mammalian STE20-like kinase 1 (MST1), dopamine receptor D5 (DRD5), succinate dehydrogenase complex (SDHA), survivor motor neurone 2 (SMN2), general transcription factor 2-I repeat domain-containing protein 2 (GTF2IRD2), neutrophil cytosolic factor 1 (NCF1), and aquaporin 7 (AQP7). Despite Cheng et al.[14] presenting expression data to support their claim that their 88 gene duplications involved functional duplicated gene copies, however, a question mark has remained over whether or not some of these genes may actually represent pseudogenes.

Probably, the most reliable dataset of human-specific gene duplications so far produced is that of Itan et al.[16] These workers identified 138 human-specific complete gene duplications that appear to have occurred since the divergence of human and chimpanzee. It was found that four of these human-specific genes are listed in the HGMD as having been reported in association with human inherited disease (Table 1). Indeed, these genes have been shown to harbour a number of different disease-causing mutations (DMs; including missense mutations and copy number variations) or disease-associated polymorphisms that may confer increased risk of a given disease state. In two of the four cases, the reported disease association was between a disease-associated polymorphism and a complex disease phenotype (ie susceptibility to infectious disease).

Table 1 Human-specific genes that have been associated with inherited disease

Intuitively, it might be supposed that those human genes that are present in more than one copy, as a consequence of a human lineage-specific increase in copy number, are better protected against the consequences of mutation in their functional parent genes by virtue of their newly acquired genetic redundancy [27, 28]. Using fairly stringent selection criteria, at least 27 human genes have so far been identified as having experienced a human lineage-specific increase in copy number [13, 29–32]. Contrary to expectation, nine of these genes (AQP7), cadherin 12, type 2 (CDH12), CHRFAM7A, DRD5, FCGR1A, GTF2IRD2, neuronal apoptosis inhibitory protein (NAIP), NCF1, and occludin (OCLN) are listed in the HGMD, although only five of them (FCGR1A, GTF2IRD2, NAIP, NCF1, OCLN) have been reported to harbour mutations that actually cause inherited disease (ie DMs; Table 2). It would be of considerable interest to ascertain (retrospectively) whether the particular patients in whom these mutations were described do indeed harbour extra functional copies of the relevant genes or whether perhaps these mutations have come to clinical attention precisely because the disease genes are effectively single copy in these particular individuals [27, 28]. An alternative explanation could, however, involve the loss of genetic redundancy through the post-duplication functional divergence of the gene copies leading to diversification through sub- or neo-functionalisation [42–45]. It is premature to speculate as to which of these postulates might provide an explanation for the observation that five of the nine genes listed in Table 2 can harbour clinically important mutations.

Table 2 Human genes identified as having experienced a human lineage-specific increase in copy number and which harbour mutations causing, or associated with, inherited disease

Another approach to this whole issue is to identify a set of human genes whose orthologues have been lost from the chimpanzee genome and then to ascertain how many of these genes are known to be involved in human inherited disease. The initial analysis of the chimpanzee genome (Chimpanzee Sequencing and Analysis Consortium, 2005) identified a total of 55 genes that are present in human but have been lost (or irrevocably disrupted) in chimpanzee. It should be noted that these genes may not necessarily be bona fide human-specific genes, since although they are absent from the chimpanzee genome, they may still be present in the genomes of other higher primates. This notwithstanding, cross-checking with the HGMD revealed that eight of these genes are known to be associated with either human inherited disease or disease susceptibility (Table 3). In seven of these eight cases, the reported disease association was between a disease-associated polymorphism and a complex disease phenotype (susceptibility to infectious disease or autoimmune disease). Although the numbers involved are clearly small, it would appear that there is at least a tendency for human-specific genes to harbour a greater-than-expected number of examples of polymorphisms associated with complex disease by comparison with the number of mutations causing Mendelian disease (bearing in mind that disease-associated polymorphisms constitute only a small minority of the lesions logged in the HGMD; see legend to Table 3).

Table 3 Human genes associated with an inherited disease that have been lost from the chimpanzee genome

Another intriguing finding emerges if Tables 1, 2, 3 are considered together: NAIP, OCLN and SMN2 are all located at 5q13.2; GTF2IRD2 and NCF1 are both located at 7q11.23, whereas butyrophilin-like 2 (BTNL2), HLA complex P5 (HCP5) and MHC class I polypeptide-related sequence A (MICA) are all located at 6p21.3. The most likely explanation for the clustering of the human-specific genes at 5q13.2 and 7q11.23 is that both of these chromosomal regions became duplicated specifically during the human lineage. On the other hand, the clustering of the BTNL2, HCP5 and MICA genes at 6p21.3 is likely to be due to the loss of this chromosomal region in chimpanzee. The clustering of human-specific genes at 5q13.2 has been noted previously [13]. The present data, however, suggest that at least two additional genomic regions (7q11.23 and 6p21.3) contain clusters of human-lineage specific genes.

Despite the above cited examples, there still appear to be relatively few 'young' (human-specific) genes among human disease genes. It may therefore be inferred that only a few young genes perform functions of sufficient importance to ensure that they have immediately come to clinical attention when mutated. This view is certainly supported by the observation that younger human, primate and mammalian genes tend to evolve more rapidly and are subject to weaker purifying selection than their more ancient counterparts [45, 60–62]. If it is assumed that the underlying mutation rate does not differ markedly between these categories of gene, there are essentially two potential explanations for very young genes experiencing a particularly rapid rate of evolutionary change: weaker purifying selection or an increase in positive selection. The available evidence suggests that both explanations are likely to pertain [62–65]. The majority of newly duplicated genes experience a period of relaxed purifying selection, while only a relatively small proportion exhibits a signature of positive selection consequent to the acquisition of new biological functions [66]. Presumably, in between the initial post-duplicational redundancy and eventual neofunctionalisation through genetic divergence, there is a 'half-way house' state in which a gene is relatively free to explore the acquisition of new functions while still being constrained to some extent by selection against the loss of those functions it has already acquired [67].

Whereas mutant alleles responsible for single gene disorders are usually under negative selection, alleles associated with complex disease appear to have been either under much less stringent purifying selection or may even have been subject to positive selection [3, 68, 69]. Genes that have experienced positive selection during the human lineage are likely to be characterised by human-specific functional adaptations. Since genes that have been subject to positive selection during human evolution have frequently also been implicated in disease,[70] we may surmise that the underlying pathological mutations may sometimes have interfered with these newly acquired functions. Consistent with this interpretation, Lappalainen et al.[71] have reported a statistically significant correlation between those regions of the human genome that have experienced recent positive selection in northern European populations and those regions that have been implicated in complex disease. This observation is also compatible with the tendency observed by the present authors for human-specific genes to harbour a disproportionate number of polymorphic variants associated with complex disease.

In terms of their expression profiles, there appears to be a tendency for rapidly evolving younger genes to be tissue-specific, whereas the more slowly evolving older genes are more broadly expressed [72–76]. It remains to be seen whether complex disease genes differ from Mendelian disease genes in terms of their expression characteristics.

More generally, it has been proposed that 'taxonomically restricted genes' may play a role in the generation of morphological diversity, thereby enabling organisms to adapt to changing environmental conditions [77]. If this is also true for the human lineage-specific genes discussed here, it follows that the acquisition of mutations in these genes may well reduce an individual's capacity to deal with a rapidly changing environment [78]. This chimes well with the (essentially unrelated) idea that some forms of common disease susceptibility may be a consequence of ancient human adaptations to a long-term stable environment ('thrifty alleles'). These ancestral alleles may now increase the risk of common disease in a changed environment consequent to the recent shift to a modern lifestyle [79]. The present authors therefore suggest that it would be a worthwhile exercise to construct a complete lexicon of human-specific genes, since these loci may well provide a happy hunting ground for those seeking to identify genes that play a key role in complex disease.

References

  1. Huang H, Winter EE, Wang H, Weinstock KG, et al: Evolutionary conservation and selection of human disease gene orthologs in the rat and mouse genomes. Genome Biol. 2004, 5: R47-10.1186/gb-2004-5-7-r47.

    Article  PubMed Central  PubMed  Google Scholar 

  2. Tu Z, Wang L, Xu M, Zhou X, et al: Further understanding human disease genes by comparing with housekeeping genes and other genes. BMC Genomics. 2006, 7: 31-10.1186/1471-2164-7-31.

    Article  PubMed Central  PubMed  Google Scholar 

  3. Blekhman R, Man O, Herrmann L, Boyko AR, et al: Natural selection on genes that underlie human disease susceptibility. Curr Biol. 2008, 18: 883-889. 10.1016/j.cub.2008.04.074.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  4. Domazet-Lošo T, Tautz D: An ancient evolutionary origin of genes associated with human genetic diseases. Mol Biol Evol. 2008, 25: 2699-2707. 10.1093/molbev/msn214.

    Article  PubMed Central  PubMed  Google Scholar 

  5. Cai JJ, Borenstein E, Chen R, Petrov DA: Similarly strong purifying selection acts on human disease genes of all evolutionary ages. Genome Biol Evol. 2009, 1: 131-144.

    Article  PubMed Central  PubMed  Google Scholar 

  6. Kehrer-Sawatzki H, Cooper DN: Understanding the recent evolution of the human genome: insights from human-chimpanzee genome comparisons. Hum Mutat. 2007, 28: 99-130. 10.1002/humu.20420.

    Article  CAS  PubMed  Google Scholar 

  7. Yu G: An exon-based comparative variant analysis pipeline to study the scale and role of frameshift and nonsense mutation in the human-chimpanzee divergence. Comp Funct Genomics. 2009, 406421-

    Google Scholar 

  8. Schrider DR, Costello JC, Hahn MW: All human-specific gene losses are present in the genome as pseudogenes. J Comput Biol. 2009, 16: 1419-1427. 10.1089/cmb.2009.0085.

    Article  CAS  PubMed  Google Scholar 

  9. Kim HL, Igawa T, Kawashima A, Satta Y, et al: Divergence, demography and gene loss along the human lineage. Philos Trans R Soc B. 2009, 365: 2451-2457.

    Article  Google Scholar 

  10. Nahon JL: Birth of "human-specific" genes during primate evolution. Genetica. 2003, 118: 193-208. 10.1023/A:1024157714736.

    Article  CAS  PubMed  Google Scholar 

  11. Stahl PD, Wainszelbaum MJ: Human-specific genes may offer a unique window into human cell signalling. Sci Signal. 2009, 2: e59-

    Article  Google Scholar 

  12. Knowles DG, McLysaght A: Recent de novo origin of human protein-coding genes. Genome Res. 2009, 19: 1752-1759. 10.1101/gr.095026.109.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  13. Fortna A, Kim Y, MacLaren E, Marshall K, et al: Lineage-specific gene duplication and loss in human and great ape evolution. PLoS Biol. 2004, 2: E207-10.1371/journal.pbio.0020207.

    Article  PubMed Central  PubMed  Google Scholar 

  14. Cheng Z, Ventura M, She X, Khaitovich P, et al: A genome-wide comparison of recent chimpanzee and human segmental duplications. Nature. 2005, 437: 88-93. 10.1038/nature04000.

    Article  CAS  PubMed  Google Scholar 

  15. Stenson PD, Mort M, Ball EV, Howells K, et al: The Human Gene Mutation Database: 2008 update. Genome Med. 2009, 1: 13-10.1186/gm13.

    Article  PubMed Central  PubMed  Google Scholar 

  16. Itan Y, Bryson K, Thomas MG: Detecting gene duplications in the human lineage. Ann Hum Genet. 2010, 74: 555-565. 10.1111/j.1469-1809.2010.00609.x.

    Article  CAS  PubMed  Google Scholar 

  17. Gonzalez E, Kulkarni H, Bolivar H, Mangano A, et al: The influence of CCL3L1 gene-containing segmental duplications on HIV-1/AIDS susceptibility. Science. 2005, 307: 1434-1440. 10.1126/science.1101160.

    Article  CAS  PubMed  Google Scholar 

  18. Mamtani M, Matsubara T, Shimizu C, Furukawa S, et al: Association of CCR2-CCR5 haplotypes and CCL3L1 copy number with Kawasaki disease, coronary artery lesions, and IVIG responses in Japanese children. PLoS One. 2010, 5: e11458-10.1371/journal.pone.0011458.

    Article  PubMed Central  PubMed  Google Scholar 

  19. McKinney C, Merriman ME, Chapman PT, Gow PJ, et al: Evidence for an influence of chemokine ligand 3-like 1 (CCL3L1) gene copy number on susceptibility to rheumatoid arthritis. Ann Rheum Dis. 2008, 67: 409-413.

    Article  CAS  PubMed  Google Scholar 

  20. Grünhage F, Nattermann J, Gressner OA, Wasmuth HE, et al: Lower numbers of the chemokine CCL3L1 gene in patients with chronic hepatitis C. J Hepatol. 2010, 52: 153-159. 10.1016/j.jhep.2009.11.001.

    Article  PubMed  Google Scholar 

  21. Minoretti P, Arra M, Emanuele U, Olivieri V, et al: A W148R mutation in the human FOXD4 gene segregating with dilated cardiomyopathy, obsessive-compulsive disorder, and suicidality. Int J Mol Med. 2007, 19: 369-372.

    CAS  PubMed  Google Scholar 

  22. Alter A, de Léséleuc L, Van Thuc N, Thai VH, et al: Genetic and functional analysis of common MRC1 exon 7 polymorphisms in leprosy susceptibility. Hum Genet. 2010, 127: 337-348. 10.1007/s00439-009-0775-x.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  23. Jedrzejowska M, Ryniewicz B, Kabzińska D, Drac H, et al: A patient with both Charcot-Marie-Tooth disease (CMT 1A) and mild spinal muscular atrophy (SMA 3). Neuromuscul Disord. 2008, 18: 339-341. 10.1016/j.nmd.2008.02.001.

    Article  PubMed  Google Scholar 

  24. Prior TW, Krainer AR, Hua Y, Swoboda KJ, et al: A positive modifier of spinal muscular atrophy in the SMN2 gene. Am J Hum Genet. 2009, 85: 408-413. 10.1016/j.ajhg.2009.08.002.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  25. Vezain M, Saugier-Veber P, Goina E, Touraine R, et al: A rare SMN2 variant in a previously unrecognized composite splicing regulatory element induces exon 7 inclusion and reduces the clinical severity of spinal muscular atrophy. Hum Mutat. 2010, 31: E1110-E1125. 10.1002/humu.21173.

    Article  PubMed  Google Scholar 

  26. Srivastava S, Mukherjee M, Panigrahi I, Shanker Pandey G, et al: SMN2-deletion in childhood-onset spinal muscular atrophy. Am J Med Genet. 2001, 101: 198-202. 10.1002/ajmg.1386.

    Article  CAS  PubMed  Google Scholar 

  27. Hsiao T-L, Vitkup D: Role of duplicate genes in robustness against deleterious human mutations. PLoS Genet. 2008, 4: e1000014-10.1371/journal.pgen.1000014.

    Article  PubMed Central  PubMed  Google Scholar 

  28. Ng PC, Levy S, Huang J, Stockwell TB, et al: Genetic variation in an individual human exome. PLoS Genet. 2008, 4: e1000160-10.1371/journal.pgen.1000160.

    Article  PubMed Central  PubMed  Google Scholar 

  29. Popesco MC, Maclaren EJ, Hopkins J, Dumas L, et al: Human lineage-specific amplification, selection, and neuronal expression of DUF1220 domains. Science. 2006, 313: 1304-1307. 10.1126/science.1127980.

    Article  CAS  PubMed  Google Scholar 

  30. Dumas L, Kim YH, Karimpour-Fard A, Cox M, et al: Gene copy number variation spanning 60 million years of human and primate evolution. Genome Res. 2007, 17: 1266-1277. 10.1101/gr.6557307.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  31. Armengol G, Knuutila S, Lozano JJ, Madrigal I, et al: Identification of human specific gene duplications relative to other primates by array CGH and quantitative PCR. Genomics. 2010, 95: 203-209. 10.1016/j.ygeno.2010.02.003.

    Article  CAS  PubMed  Google Scholar 

  32. Sudmant PH, Kitzman JO, Antonacci F, Alkan C, et al: Diversity of human copy number variation and multicopy genes. Science. 2010, 330: 641-646. 10.1126/science.1197005.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  33. Kondo H, Shimomura I, Kishida K, Kuriyama H, et al: Human aquaporin adipose (AQPap) gene. Genomic structure, promoter analysis and functional mutation. Eur J Biochem. 2002, 269: 1814-1826. 10.1046/j.1432-1033.2002.02821.x.

    Article  CAS  PubMed  Google Scholar 

  34. Rudd MF, Webb EL, Matakidou A, Sellick GS, et al: Variants in the GH-IGF axis confer susceptibility to lung cancer. Genome Res. 2006, 16: 693-701. 10.1101/gr.5120106.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  35. Raux G, Bonnet-Brilhault F, Louchart S, Houy E, et al: The -2 bp deletion in exon 6 of the "alpha 7-like" nicotinic receptor subunit gene is a risk factor for the P50 sensory gating deficit. Mol Psychiatry. 2002, 7: 1006-1011. 10.1038/sj.mp.4001140.

    Article  CAS  PubMed  Google Scholar 

  36. Cravchik A, Gejman PV: Functional analysis of the human D5 dopamine receptor missense and nonsense variants: differences in dopamine binding affinities. Pharmacogenetics. 1999, 9: 199-206.

    Article  CAS  PubMed  Google Scholar 

  37. van de Winkel JG, de Wit TP, Ernst LK, Capel PJ, et al: Molecular basis for a familial defect in phagocyte expression of IgG receptor I (CD64). J Immunol. 1995, 154: 2896-2903.

    CAS  PubMed  Google Scholar 

  38. Tipney HJ, Hinsley TA, Brass A, Metcalfe K, et al: Isolation and characterisation of GTF2IRD2, a novel fusion gene and member of the TFII-I family of transcription factors, deleted in Williams-Beuren syndrome. Eur J Hum Genet. 2004, 12: 551-560. 10.1038/sj.ejhg.5201174.

    Article  CAS  PubMed  Google Scholar 

  39. Roy N, Mahadevan MS, McLean M, Shutler G, et al: The gene for neuronal apoptosis inhibitory protein is partially deleted in individuals with spinal muscular atrophy. Cell. 1995, 80: 167-178. 10.1016/0092-8674(95)90461-1.

    Article  CAS  PubMed  Google Scholar 

  40. Roos D, de Boer M, Köker MY, Dekker J, et al: Chronic granulomatous disease caused by mutations other than the common GT deletion in NCF1, the gene encoding the pp47phox component of the phagocyte NADPH oxidase. Hum Mutat. 2006, 27: 1218-1229. 10.1002/humu.20413.

    Article  CAS  PubMed  Google Scholar 

  41. O'Driscoll MC, Daly SB, Urquhart JE, Black GC, et al: Recessive mutations in the gene encoding the tight junction protein occludin cause band-like calcification with simplified gyration and polymicrogyria. Am J Hum Genet. 2010, 87: 354-364. 10.1016/j.ajhg.2010.07.012.

    Article  PubMed Central  PubMed  Google Scholar 

  42. Lynch M, Force A: The probability of duplicate gene preservation by subfunctionalization. Genetics. 2009, 154: 459-473.

    Google Scholar 

  43. Shakhnovich BE, Koonin EV: Origins and impact of constraints in evolution of gene families. Genome Res. 2006, 16: 1529-1536. 10.1101/gr.5346206.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  44. Conrad B, Antonarakis SE: Gene duplication: A drive for phenotypic diversity and cause of human disease. Annu Rev Genomics Hum Genet. 2007, 8: 17-35. 10.1146/annurev.genom.8.021307.110233.

    Article  CAS  PubMed  Google Scholar 

  45. Han MV, Demuth JP, McGrath CL, Casola C, et al: Adaptive evolution of young gene duplicates in mammals. Genome Res. 2009, 19: 859-867. 10.1101/gr.085951.108.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  46. Chimpanzee Sequencing and Analysis Consortium: Initial sequence of the chimpanzee genome and comparison with the human genome. Nature. 2005, 437: 69-87. 10.1038/nature04072.

    Article  Google Scholar 

  47. Genovese G, Friedman DJ, Ross MD, Lecordier L, et al: Association of trypanolytic ApoL1 variants with kidney disease in African Americans. Science. 2010, 329: 841-845. 10.1126/science.1193032.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  48. Valentonyte R, Hampe J, Huse K, Rosenstiel P, et al: Sarcoidosis is associated with a truncating splice site mutation in BTNL2. Nat Genet. 2005, 37: 357-364. 10.1038/ng1519.

    Article  CAS  PubMed  Google Scholar 

  49. Zhou Q, Rammohan K, Lin S, Robinson N, et al: CD24 is a genetic modifier for risk and progression of multiple sclerosis. Proc Natl Acad Sci USA. 2003, 100: 15041-15046. 10.1073/pnas.2533866100.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  50. Sánchez E, Abelson AK, Sabio JM, González-Gay MA, et al: Association of a CD24 gene polymorphism with susceptibility to systemic lupus erythematosus. Arthritis Rheum. 2007, 56: 3080-3086. 10.1002/art.22871.

    Article  PubMed  Google Scholar 

  51. Li D, Zheng L, Jin L, Zhou Y, et al: CD24 polymorphisms affect risk and progression of chronic hepatitis B virus infection. Hepatology. 2009, 50: 735-742. 10.1002/hep.23047.

    Article  CAS  PubMed  Google Scholar 

  52. Fellay J, Shianna KV, Ge D, Colombo S, et al: A whole-genome association study of major determinants for host control of HIV-1. Science. 2007, 317: 944-947. 10.1126/science.1143767.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  53. Kirsten H, Petit-Teixeira E, Scholz M, Hasenclever D, et al: Association of MICA with rheumatoid arthritis independent of known HLA-DRB1 risk alleles in a family-based and a case control study. Arthritis Res Ther. 2009, 11: R60-10.1186/ar2683.

    Article  PubMed Central  PubMed  Google Scholar 

  54. Amroun H, Djoudi H, Busson M, Allat R, et al: Early-onset ankylosing spondylitis is associated with a functional MICA polymorphism. Hum Immunol. 2005, 66: 1057-1061. 10.1016/j.humimm.2005.09.004.

    Article  CAS  PubMed  Google Scholar 

  55. Castermans D, Wilquet V, Parthoens E, Huysmans C, et al: The neurobeachin gene is disrupted by a translocation in a patient with idiopathic autism. J Med Genet. 2003, 40: 352-356. 10.1136/jmg.40.5.352.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  56. Jönsson UB, Byström J, Stålenheim G, Venge P: Polymorphism of the eosinophil cationic protein-gene is related to the expression of allergic symptoms. Clin Exp Allergy. 2002, 32: 1092-1095. 10.1046/j.1365-2222.2002.01410.x.

    Article  PubMed  Google Scholar 

  57. Eriksson J, Reimert CM, Kabatereine NB, Kazibwe F, et al: The 434(G > C) polymorphism within the coding sequence of eosinophil cationic protein (ECP) correlates with the natural course of Schistosoma mansoni infection. Int J Parasitol. 2007, 37: 1359-1366. 10.1016/j.ijpara.2007.04.001.

    Article  CAS  PubMed  Google Scholar 

  58. Selman M, Lin HM, Montaño M, Jenkins AL, et al: Surfactant protein A and B genetic variants predispose to idiopathic pulmonary fibrosis. Hum Genet. 2003, 113: 542-550. 10.1007/s00439-003-1015-4.

    Article  CAS  PubMed  Google Scholar 

  59. Malik S, Greenwood CM, Eguale T, Kifle A, et al: Variants of the SFTPA1 and SFTPA2 genes and susceptibility to tuberculosis in Ethiopia. Hum Genet. 2006, 118: 752-759. 10.1007/s00439-005-0092-y.

    Article  CAS  PubMed  Google Scholar 

  60. Albà MA, Castresana J: Inverse relationship between evolutionary rate and age of mammalian genes. Mol Biol Evol. 2004, 22: 598-606. 10.1093/molbev/msi045.

    Article  PubMed  Google Scholar 

  61. Wolf YI, Novichkov PS, Karev GP, Koonin EV, et al: The universal distribution of evolutionary rates of genes and distinct characteristics of eukaryotic genes of different apparent ages. Proc Natl Acad Sci USA. 2009, 106: 7273-7280. 10.1073/pnas.0901808106.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  62. Cai JJ, Petrov DA: Relaxed purifying selection and possibly high rate of adaptation in primate lineage-specific genes. Genome Biol Evol. 2010, 2: 393-409. 10.1093/gbe/evq019.

    Article  PubMed Central  PubMed  Google Scholar 

  63. De S, Lopez-Bigas N, Teichmann SA: Patterns of evolutionary constraints on genes in humans. BMC Evol Biol. 2008, 8: 275-10.1186/1471-2148-8-275.

    Article  PubMed Central  PubMed  Google Scholar 

  64. Nguyen DQ, Webber C, Hehir-Kwa J, Pfundt R, et al: Reduced purifying selection prevails over positive selection in human copy number variant evolution. Genome Res. 2008, 18: 1711-1723. 10.1101/gr.077289.108.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  65. Kelley JL, Swanson WJ: Positive selection in the human genome: From genome scans to biological significance. Annu Rev Genomics Hum Genet. 2008, 9: 143-160. 10.1146/annurev.genom.9.081307.164411.

    Article  CAS  PubMed  Google Scholar 

  66. Lynch M, Conery JS: The evolutionary fate and consequences of duplicate genes. Science. 2000, 290: 1151-1155. 10.1126/science.290.5494.1151.

    Article  CAS  PubMed  Google Scholar 

  67. Chung WY, Albert R, Albert I, Nekrutenko A, et al: Rapid and asymmetric divergence of duplicate genes in the human gene coex-pression network. BMC Bioinformatics. 2006, 7: 46-10.1186/1471-2105-7-46.

    Article  PubMed Central  PubMed  Google Scholar 

  68. Nielsen R, Hubisz MJ, Hellmann I, Torgerson D, et al: Darwinian and demographic forces affecting human protein coding genes. Genome Res. 2009, 19: 838-849. 10.1101/gr.088336.108.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  69. Beekman M, Nederstigt C, Suchiman HED, Kremer D, et al: Genome-wide association study (GWAS)-identified disease risk alleles do not compromise human longevity. Proc Natl Acad Sci USA. 2010, 107: 18046-18049. 10.1073/pnas.1003540107.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  70. Vamathevan JJ, Hasan S, Emes RD, Amrine-Madsen H, et al: The role of positive selection in determining the molecular cause of species differences in disease. BMC Evol Biol. 2008, 8: 273-10.1186/1471-2148-8-273.

    Article  PubMed Central  PubMed  Google Scholar 

  71. Lappalainen T, Salmela E, Andersen PM, Dahlman-Wright K, et al: Genomic landscape of positive natural selection in northern European populations. Eur J Hum Genet. 2010, 18: 471-478. 10.1038/ejhg.2009.184.

    Article  PubMed Central  PubMed  Google Scholar 

  72. Winter EE, Goodstadt L, Ponting CP: Elevated rates of protein secretion, evolution, and disease among tissue-specific genes. Genome Res. 2004, 14: 54-61.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  73. Zhang L, Li W-H: Mammalian housekeeping genes evolve more slowly than tissue-specific genes. Mol Biol Evol. 2004, 21: 236-239.

    Article  PubMed  Google Scholar 

  74. Liao BY, Zhang J: Low rates of expression profile divergence in highly expressed genes and tissue-specific genes during mammalian evolution. Mol Biol Evol. 2006, 23: 1119-1128. 10.1093/molbev/msj119.

    Article  CAS  PubMed  Google Scholar 

  75. Hao L, Ge X, Wan H, Hu S, et al: Human functional genetic studies are biased against the medically most relevant primate-specific genes. BMC Evol Biol. 2010, 10: 316-10.1186/1471-2148-10-316.

    Article  PubMed Central  PubMed  Google Scholar 

  76. Park SG, Choi SS: Expression breadth and expression abundance behave differently in correlations with evolutionary rates. BMC Evol Biol. 2010, 10: 241-10.1186/1471-2148-10-241.

    Article  PubMed Central  PubMed  Google Scholar 

  77. Khalturin K, Hemmrich G, Fraune S, Augustin R, et al: More than just orphans: are taxonomically-restricted genes important in evolution?. Trends Genet. 2009, 25: 404-413. 10.1016/j.tig.2009.07.006.

    Article  CAS  PubMed  Google Scholar 

  78. Blekhman R, Oshlack A, Chabot AE, Smyth GK, et al: Gene regulation in primates evolves under tissue-specific selection pressures. PLoS Genet. 2008, 4: e1000271-10.1371/journal.pgen.1000271.

    Article  PubMed Central  PubMed  Google Scholar 

  79. Di Rienzo A, Hudson RR: An evolutionary framework for common diseases: the ancestral-susceptibility model. Trends Genet. 2005, 21: 596-601. 10.1016/j.tig.2005.08.007.

    Article  CAS  PubMed  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to David N Cooper.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Cooper, D.N., Kehrer-Sawatzki, H. Exploring the potential relevance of human-specific genes to complex disease. Hum Genomics 5, 99 (2011). https://doi.org/10.1186/1479-7364-5-2-99

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1479-7364-5-2-99

Keywords