Skip to main content Accessibility help
×
Hostname: page-component-5c6d5d7d68-sv6ng Total loading time: 0 Render date: 2024-08-19T06:27:08.863Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 June 2014

Graeme Finlay
Affiliation:
University of Auckland
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Human Evolution
Genes, Genealogies and Phylogenies
, pp. 284 - 350
Publisher: Cambridge University Press
Print publication year: 2013

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

Waller, J, Fabulous Science: Fact and Fiction in the History of Scientific Discovery (Oxford: Oxford University Press, 2002), 181–4Google Scholar
Padian, K (2008). Darwin’s enduring legacy. Nature 451, 632–4; Bowler, BJ (2009). Darwin’s originality. Science 323, 223–6Google Scholar
Vollbrecht, E and Sigmon, B (2005). Amazing grass: developmental genetics of maize domestication. Biochemical Society Transactions 33, 1502–6; Tian, F, Stevens, NM and Buckler, ES IV (2009). Tracking footprints of maize domestication and evidence for a massive selective sweep on chromosome 10. Proceedings of the National Academy of Sciences of the USA 106 (Suppl. 1), 9979–86Google Scholar
National Academy of Sciences, Science, Evolution, and Creationism (Washington DC: National Academies Press, 2008), available at Google Scholar
Waller, , Fabulous Science, 199–202
Polkinghorne, J, Science and Creation (London: SPCK, 1988), 54Google Scholar
Jeeves, M, quoted in Henry, CF (ed.), Horizons of Science: Christian Scholars Speak Out (San Francisco: Harper & Row, 1978), 29Google Scholar
MacKay, DM, The Clockwork Image: A Christian Perspective on Science (London: Intervarsity Press, 1974), 54Google Scholar
Spencer, N, God and Darwin (London: SPCK, 2009), 48, 78–9, 82–4, 124Google Scholar
See ; see also Editorial (2005). Dealing with design. Nature 434, 1053; Brumfiel, G (2005). Intelligent design: who has designs on your students’ minds?Nature 434, 1062; Leshner, AI (2005). Redefining science. Science 309, 221Google Scholar
Brooke, JH (2008). Charles Darwin on religion: a statement prepared for the International Society for Science and Religion, available at ; Shapin, S (2010). The Darwin Show. London Review of Books 32, 3, available at Google Scholar
Russel, CA, Cross-Currents: Interactions between Science and Faith (Leicester: Intervarsity Press, 1985), 150Google Scholar
Livingstone, DN, Darwin’s Forgotten Defenders (Grand Rapids, MI: Eerdmans, 1987), 118Google Scholar
Including John Stek, Bruce Waltke and John Walton. See for example Van Till, HJ, Snow, RE, Stek, JH and Young, DA, Portraits of Creation (Grand Rapids, MI: Eerdmans, 1990)Google Scholar
Wenham, GJ, Word Biblical Commentary: Genesis 1–15 (Nashville: Thomas Nelson, 1987); Waltke, BK and Fredricks, CJ, Genesis: A Commentary (Grand Rapids, MI: Zondervan, 2001); Wenham, GJ, Exploring the Old Testament: Volume 1 The Pentateuch (London: SPCK, 2003); Walton, JH, The Lost World of Genesis One (Downers Grove, IL: Intervarsity Press, 2009)Google Scholar
Stott, JRW, Understanding the Bible (London: Scripture Union, 1984), 48–9; Livingstone, DN and Noll, MA (2000). B.B. Warfield (1851–1921): a biblical inerrantist as evolutionist. Isis 91, 283–304; Wright, NT, ; Bauckham, R, Bible and Ecology (London: Darton, Longman & Todd, 2010), 158–60Google Scholar
Livingstone, , Defenders, 102–5; Lindberg, DC and Numbers, RL, God and Nature: Historical Essays on the Encounter between Christianity and Science (Berkeley, CA: University of California Press, 1986), 376–7Google Scholar
Lindberg, and Numbers, , God and Nature, 402–3
Hunter, GW, A Civic Biology: Presented in Problems. (New York: American Book Company, 1914), 196, 261–3Google Scholar
Spencer, , God and Darwin, 87–98
For example, Wang, HH, Isaacs, FJ, Carr, PA et al. (2009). Programming cells by multiplex genome engineering and accelerated evolution. Nature 460, 894–8; Worsdorfer, B, Woycechowsky, KJ and Hilvert, D (2011). Directed evolution of a protein container. Science 331, 589–92Google Scholar
Haught, J, in Conway Morris, S (ed.), The Deep Structure of Biology: Is Convergence Sufficiently Ubiquitous to Give a Directional Signal? (West Conshohocken, PA: Templeton Foundation Press, 2008), 218Google Scholar
Polkinghorne, , Science and Creation, 47–50; Burge, T (2002). What else does physics tell us about God?Science and Christian Belief 14, 79–80; Gingerich, O, God’s Universe (Cambridge, MA: Harvard University Press, 2006), 118–20; Morris, Conway, Deep Structure, 228–30; Ewart, P (2009). The necessity of chance: randomness, purpose and the sovereignty of God. Science and Christian Belief 21, 111–31Google Scholar
Vermeij, GJ (2006). Historical contingency and the purported uniqueness of evolutionary innovations. Proceedings of the National Academy of Sciences of the USA 103, 1804–9CrossRefGoogle Scholar
Conway Morris, S, Life’s Solution: Inevitable Humans in a Lonely Universe (Cambridge: Cambridge University Press, 2003); Morris, Conway, Deep Structure, 46CrossRefGoogle Scholar
Ohlrich, C, The Suffering God (London: Triangle, 1982); Polkinghorne, J, The Way the World Is (London: SPCK, 1983), 69–77; Rolston, H III, Genes, Genesis and God (Cambridge: Cambridge University Press, 1999), 303–7Google Scholar
See Wright, , ref. 16
For exceptions, see Dingli, D and Nowak, MA (2006). Infectious tumour cells. Nature 443, 35–6; Belov, K (2011). The role of the major histocompatibility complex in the spread of contagious cancers. Mammalian Genome 22, 83–90Google Scholar
Katzir, N, Rechavi, G, Cohen, GB et al. (1985). ‘Retroposon’ insertion into the cellular oncogene c-myc in canine transmissible venereal tumor. Proceedings of the National Academy of Sciences of the USA 82, 1054–8; Murgia, C, Pritchard, JK, Kim, SY et al. (2006). Clonal origin and evolution of a transmissible cancer. Cell 126, 477–87; Rebbeck, CA, Thomas, R, Breen, M et al. (2009). Origins and evolution of a transmissible cancer. Evolution 63, 2340–9; Rebbeck, CA, Leroi, AM and Burt, A (2011). Mitochondrial capture by a transmissible cancer. Science 331, 303Google Scholar
Pearse, A-M and Swift, K (2006). Transmission of devil facial-tumour disease. Nature 439, 549; Murchison, EP, Tovar, C, Hsu, A et al. (2010). The Tasmanian devil transcriptome reveals Schwann cell origins of a clonally transmissible cancer. Science 327, 84–7; Murchison, EP, Schulz-Trieglaff, OB, Ning, Z et al. (2012). Genome sequencing and analysis of the Tasmanian devil and its transmissible cancer. Cell 148, 780–91; for overviews, see Jones, ME and McCallum, H (2011). The devil’s cancer. Scientific American 304(6), 72–7; Belov, K (2012). Contagious cancer: lessons from the devil and the dog. BioEssays 34, 285–92Google Scholar
Gill, P, Ivanov, PL, Kimpton, C et al. (1994). Identification of the remains of the Romanov family by DNA analysis. Nature Genetics 6, 130–5; Ivanov, PL, Wadhams, MJ, Roby, RK et al. (1996). Mitochondrial DNA sequence heteroplasmy in the Grand Duke of Russia Georgij Romanov establishes the authenticity of the remains of Tsar Nicholas II. Nature Genetics 12, 417–20Google Scholar
Coble, MD, Loreille, OM, Wadhams, MJ et al. (2009). Mystery solved: the identification of the two missing Romanov children using DNA analysis. PLoS ONE 4, e4838; Rogaev, EI, Grigorenko, AP, Moliaka, YK et al. (2009). Genomic identification in the historical case of the Nicholas II royal family. Proceedings of the National Academy of Sciences of the USA 106, 5258–63Google Scholar
Rogaev, EI, Grigorenko, AP, Faskhutdinova, G et al. (2009). Genotype analysis identifies the cause of the ‘royal disease’. Science 326, 817; Lannoy, N and Hermans, C (2010). The ‘royal disease’ – haemophilia A or B? A haematological mystery is finally solved. Haemophilia 16, 843–7CrossRefGoogle Scholar
Coble, MD (2011). The identification of the Romanovs: can we (finally) put the controversies to rest?Investigative Genetics 2, 20Google Scholar
Avise, JC, Inside the Human Genome: A Case for Non-Intelligent Design (Oxford: Oxford University Press, 2010); particularly recommended is Fairbanks, DJ, Relics of Eden: The Powerful Evidence of Evolution in Human DNA (Amherst, MA, and New York: Prometheus Books, 2007); from a more theological perspective, see Alexander, DR, Creation and Evolution: Do We Have to Choose? (Oxford: Monarch, 2008); Pattemore, P, Am I My Keepers Brother?: Human Origins from a Christian and a Scientific Perspective (see , 2011)CrossRefGoogle Scholar
Finlay, GJ (2003). Homo divinus: the ape that bears God’s image. Science and Christian Belief 15, 17–40; Finlay, GJ (2008). Evolution as created history. Science and Christian Belief 20, 67–89; Finlay, GJ (2008). Human evolution: how random process fulfils divine purpose. Perspectives on Science and Christian Faith 60(2), 103–14; Finlay, GJ. The emergence of human distinctiveness: the genetic story, in Jeeves, M (ed.), Rethinking Human Nature: a Multidisciplinary Approach (Grand Rapids, MI: Eerdmans, 2011)Google Scholar
Weinberg, RA, The Biology of Cancer (New York: Garland Science, 2007)Google Scholar
Boccardo, E and Villa, LL (2007). Viral origins of human cancer. Current Medicinal Chemistry 14, 2526–39; Dayaram, T and Marriott, SJ (2008). Effect of transforming viruses on molecular mechanisms associated with cancer. Journal of Cellular Physiology 216, 309–14; Moore, MS and Chang, Y (2010). Why do viruses cause cancer? Highlights of the first century of human tumour virology. Nature Reviews Cancer 10, 878–89Google Scholar
Rubin, H (2011). The early history of tumor virology: Rous, RIF, and RAV. Proceedings of the National Academy of Sciences of the USA 108, 14389–96CrossRefGoogle Scholar
Mortreux, F, Gabet, A-S and Wattel, E (2003). Molecular and cellular aspects of HTLV-1, associated leukemogenesis in vivo. Leukemia 17, 26–38; Karpas, A (2004). Human retroviruses in leukaemia and AIDS: reflections on their discovery, biology and epidemiology. Biological Reviews of the Cambridge Philosophical Society 79, 911–33; Verdonck, K, Gonzalez, E, Van Dooren, S et al. (2007). Human T-lymphotropic virus 1: recent knowledge about an ancient infection. Lancet Infectious Diseases 7, 266–81; Boxus, M and Willems, L (2009). Mechanisms of HTLV-1 persistence and transformation. British Journal of Cancer 101, 1497–501; Goncalves, DU, Proietti, FA, Ribas, JGR et al. (2010). Epidemiology, treatment, and prevention of human T-cell leukemia virus type 1-associated diseases. Clinical Microbiological Reviews 23, 577–89Google Scholar
Leclercq, I, Mortreux, F, Cavrois, M et al. (2000). Host sequences flanking the human T-cell leukemia virus type 1 provirus in vivo. Journal of Virology 74, 2305–12CrossRefGoogle Scholar
Chou, KS, Okayama, A, Su, I-J et al. (1996). Preferred nucleotide sequence at the integration target site of human T-cell leukemia virus type I from patients with adult T-cell leukemia. International Journal of Cancer 65, 20–4Google Scholar
Seiki, M, Hattori, S, Hirayama, Y and Yoshida, M (1983). Human adult T-cell leukemia virus: complete nucleotide sequence of the provirus genome integrated into leukemia cell DNA. Proceedings of the National Academy of Sciences of the USA 80, 3618–22CrossRefGoogle Scholar
Hacein-Bey-Abina, S, Von Kalle, C, Schmidt, M et al. (2003). LMO2-associated clonal T cell proliferation in two patients after gene therapy for SCID-X1. Science 302, 415–19; Deichmann, A, Hasein-Bey-Abina, S, Schmidt, M et al. (2007). Vector integration is nonrandom and clustered and influences the fate of lymphopoiesis in SCID-X1 gene therapy. Journal of Clinical Investigation 117, 2225–32; Kaiser, J (2009). β-Thalassemia treatment succeeds, with a caveat. Science 326, 1468–9Google Scholar
Hayward, WS, Neel, BG and Atrin, SM (1981). Activation of a cellular onc gene by promoter insertion in ALV-induced lymphoid leukosis. Nature 290, 475–80; Peters, G, Lee, AE and Dickson, C (1984). Activation of cellular gene by mouse mammary tumour virus may occur early in mammary tumour development. Nature 309, 273–5; Nusse, R, Van Ooyen, A, Rijsewijk, F et al. (1985). Retroviral insertional mutagenesis in murine mammary cancer. Proceedings of the Royal Society of London Series B 226, 3–13; Moules, V, Pomier, C, Sibon, D et al. (2005). Fate of premalignant clones during the asymptomatic phase preceding lymphoid malignancy. Cancer Research 65, 1234–43Google Scholar
Westaway, D, Payne, G and Varmus, HE (1984). Proviral deletions and oncogene base-substitutions in insertionally mutagenized c-myc alleles may contribute to the progression of avian bursal tumors. Proceedings of the National Academy of Sciences of the USA 81, 843–7CrossRefGoogle Scholar
Vinokurova, S, Wentzensen, N, Einenkel, J et al. (2005). Clonal history of papillomavirus-induced dysplasia in the female lower genital tract. Journal of the National Cancer Institute 97, 1816–21; see also Luft, F, Klaes, R, Nees, M et al. (2001). Detection of integrated papillomavirus sequences by ligation-mediated PCR (DIPS-PCR) and molecular characterization in cervical cancer cells. International Journal of Cancer 92, 9–17CrossRefGoogle Scholar
Ng, IO-L, Guan, X-Y, Poon, R T-P et al. (2003). Determination of the molecular relationship between multiple tumour nodules in hepatocellular carcinoma differentiates multicentric origin from intrahepatic metastases. Journal of Pathology 199, 345–53; Saigo, K, Yoshida, K, Ikeda, R et al. (2008). Integration of hepatitis B virus DNA into the myeloid/lymphoid or mixed lineage leukaemia (MLL4) gene and rearrangements of MLL4 in human hepatocellular carcinoma. Human Mutation 29, 703–8Google Scholar
Feng, H, Shuda, M, Chang, Y and Moore, PS (2008). Clonal integration of a polyomavirus in human Merkel cell carcinoma. Science 319, 1096–100; Gandhi, RK, Rosenberg, AS and Somach, SC (2009). Merkel cell polyomavirus: an update. Journal of Cutaneous Pathology 36, 1327–9CrossRefGoogle ScholarPubMed
Weiss, RA (2006). The discovery of endogenous retroviruses. Retrovirology 3, 67; Stoye, JP (2012). Studies of endogenous retroviruses reveal a continuing evolutionary saga. Nature Reviews Microbiology 10, 395–406; Feschotte, C and Gilbert, C (2012). Endogenous viruses: insights into viral evolution and impact on host biology. Nature Reviews Genetics 13, 283–96Google Scholar
Lower, R, Lower, J and Kurth, R (1996). The viruses in all of us: characteristics and biological significance of human endogenous retrovirus sequences. Proceedings of the National Academy of Sciences of the USA 93, 5177–84; Bromham, L (2002). The human zoo: endogenous retroviruses in the human genome. Trends in Ecology and Evolution 17, 91–7; Bannert, N and Kurth, R (2004). Retroelements and the human genome: new perspectives on an old relation. Proceedings of the National Academy of Sciences of the USA 101 (Suppl. 2), 14572–9; Kurth, R and Bannert, N (2010). Beneficial and detrimental effects of human endogenous retroviruses. International Journal of Cancer 126, 306–14Google Scholar
Smit, AFA (1999). Interspersed repeats and other mementos of transposable elements in mammalian genomes. Current Opinion in Genetics and Development 9, 657–63; Gifford, R and Tristem, M (2003). The evolution, distribution and diversity of endogenous retroviruses. Virus Genes 26, 291; Mayer, J and Meese, E (2005). Human endogenous retroviruses in the primate lineage and their influence on host genomes. Cytogenetic and Genome Research 110, 448–56; Blikstad, V, Benachenhou, F, Sperber, GO and Blomberg, J (2008). Evolution of human endogenous retroviral sequences: a conceptual account. Cellular and Molecular Life Sciences 65, 3348–65; Blomberg, J, Benachenhou, F, Blikstad, V et al. (2009). Classification and nomenclature of endogenous retroviral sequences (ERVs): problems and recommendations. Gene 448, 115–23Google Scholar
Bonner, TI, O’Connell, C and Cohen, M (1982). Cloned endogenous retroviral sequences from human DNA. Proceedings of the National Academy of Sciences of the USA 79, 4709–13; see also Mariani-Constantini, R, Horn, TM and Callahan, R (1989). Ancestry of a human endogenous retrovirus family. Journal of Virology 63, 4982–5Google Scholar
Johnson, WE and Coffin, JM (1999). Constructing primate phylogenies from ancient retrovirus sequences. Proceedings of the National Academy of Sciences of the USA 96, 10254–60CrossRefGoogle Scholar
Lindeskog, M, Mager, DL and Blomberg, J (1999). Isolation of a human endogenous retroviral HERV-H element with an open env reading frame. Virology 258, 441–50; de Parsival, N, Casella, J-F, Gressin, L and Heidmann, T (2001). Characterisation of the three HERV-H proviruses with an open envelope reading frame encompassing the immunosuppressive domain and evolutionary history in primates. Virology 279, 558–69; Benit, L, Calteau, A and Heidmann, T (2003). Characterization of the low-copy HERV-Fc family: evidence for recent integration in primates of elements with coding envelope genes. Virology 312, 159–68Google Scholar
Barbulescu, M, Turner, G, Seaman, MI et al. (1999). Many human endogenous retrovirus K (HERV-K) proviruses are unique to humans. Current Biology 9, 861–8CrossRefGoogle Scholar
Turner, G, Barbulescu, M, Su, M et al. (2001). Insertional polymorphisms of full-length endogenous retroviruses in humans. Current Biology 11, 1531–5; Jha, AR, Pillai, SK, York, VA et al. (2009). Cross-sectional dating of novel haplotypes of HERV-K 113 and HERV-K 115 indicate these proviruses originated in Africa before Homo sapiens. Molecular Biology and Evolution 26, 2617–26Google Scholar
Jha, AR, Nixon, DF, Rosenberg, MG et al. (2011). Human endogenous retrovirus K106 (HERV-K106) was infectious after the emergence of anatomically modern humans. PLoS ONE 6, e20234Google Scholar
Agoni, L, Golden, A, Guha, C and Lenz, J (2012). Neandertal and Denisovan retroviruses. Current Biology 22, R437–8CrossRefGoogle Scholar
Subramanian, RP, Wildschutte, JH, Russo, C and Coffin, JM (2011). Identification, characterization, and comparative genomic distribution of the HERV-K (HML-2) group of human endogenous retroviruses. Retrovirology 8, 90Google Scholar
Lebedev, YB, Belonovitch, OS, Zybrova, NV et al. (2000). Differences in HERV-K LTR insertions in orthologous loci of humans and the great apes. Gene 247, 265–77; Sverdlov, ED (2000). Retroviruses and primate evolution. BioEssays 22, 161–71; Kurdyukov, SG, Lebedev, YB, Artamonova, et al. (2001). Full-sized HERV-K (HML-2) human endogenous retroviral LTR sequences on human chromosome 21: map locations and evolutionary history. Gene 273, 51–61; Nadezhdin, EV, Lebedev, YB, Glazkova, DV et al. (2001). Identification of paralogous HERV-K LTRs on human chromosomes 3, 4, 7 and 11 containing clusters of olfactory receptor genes. Molecular Genetics and Genomics 265, 820–5CrossRefGoogle ScholarPubMed
Sin, H-S, Koh, E, Taya, M et al. (2011). A novel Y chromosome microdeletion with the loss of an endogenous retrovirus-related, testis-specific transcript in the AZFb region. Journal of Urology 186, 1545–52CrossRefGoogle Scholar
Subramanian, et al. (2011), ref. 24; see also Hughes, JF and Coffin, JM (2004). Human endogenous retrovirus K solo-LTR formation and insertional polymorphisms: implications for human and viral evolution. Proceedings of the National Academy of Sciences of the USA 101, 1668–72; Belshaw, R, Watson, J, Katzourakis, A et al. (2007). Rate of recombinational deletion among human endogenous retroviruses. Journal of Virology 81, 9437–42Google Scholar
Huh, J-W, Kim, D-S, Ha, H-S et al. (2007). Formation of a new solo-LTR of the human endogenous retrovirus H family in human chromosome 21. Molecules and Cells 22, 360–3Google Scholar
Choi, J, Koh, E, Matsui, F et al. (2008). Study of azoospermia factor-a deletion caused by homologous recombination between the human endogenous retroviral elements and population-specific alleles in Japanese infertile males. Fertility and Sterility 89, 1177–82CrossRefGoogle ScholarPubMed
Hughes, JF and Coffin, JM (2001). Evidence for genomic rearrangements mediated by human endogenous retroviruses during primate evolution. Nature Genetics 29, 487–9; Hughes, JF and Coffin, JM (2005). Human endogenous retroviral elements as indicators of ectopic recombination events in the primate genome. Genetics 171, 1183–94CrossRefGoogle ScholarPubMed
Lindeskog, M, Medstrand, P, Cunningham, AA and Blomberg, J (1999). Coamplification and dispersion of adjacent human endogenous retroviral HERV-H and HERV-E elements: presence of spliced hybrid transcripts in normal leukocytes. Virology 244, 219–29CrossRefGoogle Scholar
Jamain, S, Girondot, M, Leroy, P et al. (2001). Transduction of the human gene FAM8A1 by endogenous retrovirus during primate evolution. Genomics 78, 38–49CrossRefGoogle Scholar
International Human Genome Sequencing Consortium (2001). Initial sequencing and analysis of the human genome. Nature 409, 860–921; Li, W-H, Gu, Z, Wang, H and Nekrutenko, A (2001). Evolutionary analyses of the human genome. Nature 409, 847–9CrossRefGoogle Scholar
Chimpanzee Sequencing and Analysis Consortium (2005). Initial sequence of the chimpanzee genome and comparison with the human genome. Nature 437, 69–87CrossRefGoogle Scholar
Prufer, K, Munch, K, Hellmann, I et al. (2012). The bonobo genome compared with the human and chimpanzee genomes. Nature 486, 527–31 (for ERVs and TEs, see Supplementary Information)CrossRefGoogle Scholar
Ventura, M, Catacchio, CR, Alkan, C et al. (2011). Gorilla genome structural variation reveals evolutionary parallelisms with chimpanzee. Genome Research 21, 1640–9; for a genome analysis (without reference to ERVs), see Scally, A, Dutheil, JY, Hillier, LW et al. (2012). Insights into hominid evolution from the gorilla genome sequence. Nature 483, 169–75CrossRefGoogle ScholarPubMed
Locke, DP, Hillier, LW, Warren, WC et al. (2011). Comparative and demographic analysis of orang-utan genomes. Nature 469, 529–33 (for ERVs, see Supplementary Information)CrossRefGoogle ScholarPubMed
Rhesus Macaque Genome Sequencing and Analysis Consortium (2007). Evolutionary and biomedical insights from the rhesus macaque genome. Science 316, 222–34; Han, K, Konkel, MK, Xing, J et al. (2007). Mobile DNA in Old World monkeys: a glimpse through the rhesus macaque genome. Science 316, 238–40Google Scholar
Meyer, M, Kircher, M, Gansauge, M-T et al. (2012). A high-coverage genome sequence from an archaic Denisovan individual. Science 338, 222–6CrossRefGoogle ScholarPubMed
Green, RE, Krause, J, Briggs, AW et al. (2012). A draft sequence of the Neandertal genome. Science 328, 710–22CrossRefGoogle Scholar
Kijima, TE and Innan, H (2010). On the estimation of the insertion time of retrotransposable elements. Molecular Biology and Evolution 27, 896–904CrossRefGoogle ScholarPubMed
de Parsival, N and Heidmann, T (2005). Human endogenous retroviruses: from infectious elements to human genes. Cytogenetic and Genome Research 110, 318–32CrossRefGoogle Scholar
Herve, CA, Forrest, G, Lower, R et al. (2004). Conservation and loss of the ERV3 open reading frame in primates. Genomics 83, 940–3CrossRefGoogle Scholar
Aagaard, L, Villesen, P, Kjeldbjerg, AL and Pedersen, FS (2005). The approximately 30-million-year-old ERVPb1 envelope gene is evolutionarily conserved among hominoids and Old World monkeys. Genomics 86, 685–91; Kjeldbjerg, AL, Villesen, P, Aagaard, L and Pedersen, FS (2008). Gene conversion and purifying selection of a placenta-specific ERV-V envelope gene during simian evolution. BMC Evolutionary Biology 8, 266; for ERV-Pb1 envelope fusogenic capacity, see Aagaard, L, Bjerregaard, B, Kjeldbjerg, AL et al. (2012). Silencing of endogenous envelope genes in human choriocarcinoma cells shows that envPb1 is involved in heterotypic cell fusions. Journal of General Virology 93, 1696–9CrossRefGoogle ScholarPubMed
Bonnaud, B, Beliaeff, J, Bouton, O et al. (2005). Natural history of the ERVWE1 endogenous retroviral locus. Retrovirology 2, 57CrossRefGoogle ScholarPubMed
Mi, S, Lee, X, Li, X et al. (2000). Syncytin is a captive retroviral envelope protein involved in placental morphogenesis. Nature 403, 785–9CrossRefGoogle ScholarPubMed
Tolosa, JM, Schjenken, JE, Clifton, VL et al. (2012). The endogenous retroviral envelope protein syncytin-1 inhibits LPS/PHA-stimulated cytokine responses in human blood and is sorted into placental exosomes. Placenta 33, 933–41; Fahlbusch, FB, Ruebner, M, Volkert, G et al. (2012). Corticotropin-releasing hormone stimulates expression of leptin, 11beta-HSD2 and syncytin-1 in primary human trophoblasts. Reproductive Biology and Endocrinology 10, 80CrossRefGoogle ScholarPubMed
Prudhomme, S, Oriol, G and Mallet, F (2004). A retroviral promoter and a cellular enhancer define a bipartite element which controls env ERVWE1 placental expression. Journal of Virology 78, 12157–68; Matsuura, K, Jigami, T, Taniue, K et al. (2011). Identification of a link between Wnt/β-catenin signalling and the cell fusion pathway. Nature Communications 2, 548; Ruebner, M, Langbein, M, Strissel, PL et al. (2012). Regulation of the human endogenous retroviral syncytin-1 and cell–cell fusion by the nuclear hormone receptors PPARγ/RXRα in placentogenesis. Journal of Cellular Biochemistry 113, 2383–96Google Scholar
Muir, A, Lever, AML and Moffett, A (2006). Human endogenous retrovirus-W envelope (syncytin) is expressed in both villous and extravillous trophoblast populations. Journal of General Virology 87, 2067–71; Hayward, MD, Potgens, AJG, Drewlo, S et al. (2007). Distribution of human endogenous retrovirus type W receptor in normal human villous placenta. Pathology 39, 406–12CrossRefGoogle ScholarPubMed
Mi et al. (2000), ref. 46; Mallet, F, Bouton, O, Prudhomme, S et al. (2004). The endogenous locus ERVWE1 is a bona fide gene involved in hominoid placental physiology. Proceedings of the National Academy of Sciences of the USA 101, 1731–6Google Scholar
Caceres, M, NISC Comparative Sequencing Program and Thomas, JW (2006). The gene of retroviral origin syncytin 1 is specific to hominoids and is inactive in Old World monkeys. Journal of Heredity 97, 100–6; see also Bonnaud, B, Bouton, O, Oriol, G et al. (2004). Evidence of selection on the domesticated ERVWE1 env retroviral element involved in placentation. Molecular Biology and Evolution 21, 1895–901CrossRefGoogle ScholarPubMed
Søe, K, Andersen, TL, Hobolt-Pederson, A-S et al. (2011). Involvement of human endogenous retroviral syncytin-1 in human osteoclast fusion. Bone 48, 837–46CrossRefGoogle ScholarPubMed
Malassine, A, Blaise, S, Handschuh, K et al. (2007). Expression of the fusogenic HERV-FRD Env glycoprotein (syncytin 2) in human placenta is restricted to villous cytotrophoblastic cells. Placenta 28, 185–91; Esnault, C, Priet, S, Ribet, D et al. (2008). A placenta-specific receptor for the fusogenic, endogenous retrovirus-derived, human syncytin-2. Proceedings of the National Academy of Sciences of the USA 105, 17532–7CrossRefGoogle ScholarPubMed
Blaise, S, de Parseval, N, Benit, L and Heidmann, T (2003). Genome-wide screening for fusogenic human endogenous retrovirus envelopes identifies syncytin 2, a gene conserved on primate evolution. Proceedings of the National Academy of Sciences of the USA 100, 13013–18; Mangeney, M, Renard, M, Schlecht-Louf, G et al. (2007). Placental syncytins: genetic disjunction between the fusogenic and immunosuppressive activity of retroviral envelope proteins. Proceedings of the National Academy of Sciences of the USA 104, 20534–9CrossRefGoogle Scholar
Tolosa, et al. (2012), ref. 47; see also Kammerer, U, Germeyer, A, Stengel, S et al. (2011). Human endogenous retrovirus K (HERV-K) is expressed in villous and extravillous cytotrophoblast cells of the human placenta. Journal of Reproductive Immunology 91, 1–8Google Scholar
Heidmann, O, Vernochet, C, Dupressoir, A and Heidmann, T (2009). Identification of an endogenous retroviral envelope gene with fusogenic activity and placenta-specific expression in the rabbit: a new ‘syncytin’ in a third order of mammals. Retrovirology 6, 107; Vernochet, C, Heidmann, O, Dupressoir, A et al. (2011). A syncytin-like endogenous retroviral envelope gene of the guinea pig specifically expressed in the placenta junctional zone and conserved in Caviomorpha. Placenta 32, 885–92; Cornelis, G, Heidmann, O, Bernard-Stoecklin, S et al. (2012). Ancestral capture of syncytin-Car1, a fusogenic endogenous retroviral envelope gene involved in placentation and conserved in Carnivora. Proceedings of the National Academy of Sciences of the USA 109, E432–41CrossRefGoogle Scholar
Dunlap, KA, Palmarini, M, Varela, M et al. (2006). Endogenous retroviruses regulate periimplantation placental growth and differentiation. Proceedings of the National Academy of Sciences of the USA 103, 14390–5; Baba, K, Nakaya, Y, Shojima, T et al. (2011). Identification of novel endogenous betaretroviruses which are transcribed in bovine placenta. Journal of Virology 85, 1237–45CrossRefGoogle ScholarPubMed
Dupressoir, A, Vernochet, C, Bawa, O et al. (2009). Syncytin-A knockout mice demonstrate the critical role in placentation of a fusogenic, endogenous retrovirus-derived, envelope gene. Proceedings of the National Academy of Sciences of the USA 106, 12127–32; Dupressoir, A, Vernochet, C, Harper, F et al. (2011). A pair of co-opted retroviral envelope syncytin genes is required for formation of the two-layered murine placental syncytiotrophoblast. Proceedings of the National Academy of Sciences of the USA 108, E1164–73CrossRefGoogle ScholarPubMed
Dunlap, et al. (2006), ref. 57; Black, SG, Arnaud, F, Palmarini, M and Spencer, TE (2010). Endogenous retroviruses in endogenous differentiation and placental development. American Journal of Reproductive Immunology 64, 255–64Google Scholar
Kudaka, W, Oda, T, Jinno, Y et al. (2008). Cellular localization of placenta-specific human endogenous retrovirus (HERV) transcripts and their possible implication in pregnancy-induced hypertension. Placenta 29, 282–9; Langbein, M, Strick, R, Strissel, PL et al. (2008). Impaired cytotrophoblast cell–cell fusion is associated with reduced syncytin and increased apoptosis in patients with placental dysfunction. Molecular Reproduction and Development 75, 175–83; Ruebner, M, Strissel, PL, Langbein, M et al. (2010). Impaired cell fusion and differentiation in placentae from patients with intrauterine growth restriction correlate with reduced levels of HERV envelope genes. Journal of Molecular Medicine 88, 1143–56; Vargas, A, Toufaily, C, LeBellego, F et al. (2011). Reduced expression of both syncytin 1 and syncytin 2 correlates with severity of preeclampsia. Reproductive Sciences 18, 1085–91CrossRefGoogle ScholarPubMed
Mallasine, A, Frendo, JL, Blaise, S et al. (2008). Human endogenous retrovirus-FRD envelope protein (syncytin 2) expression in normal and trisomy 21-affected placenta. Retrovirology 5, 6CrossRefGoogle Scholar
Lynch, C and Tristem, M (2003). A co-opted gypsy-type LTR-retrotransposon is conserved in the genomes of humans, sheep, mice, and rats. Current Biology 13, 1518–23; Brandt, J, Veith, AM and Volff, J-N (2005). A family of neofunctionalized Ty3/gypsy retrotransposon genes in mammalian genomes. Cytogenetic and Genome Research 110, 307–17; Volff, J-N (2009). Cellular genes derived from Gypsy/Ty3 retrotransposons in mammalian genomes. Annals of the New York Academy of Sciences 1178, 233–43CrossRefGoogle ScholarPubMed
Clark, MB, Janicke, M, Gottesbuhren, U et al. (2007). Mammalian gene PEG10 expresses two reading frames by high efficiency –1 frameshifting in embryonic-associated tissues. Journal of Biological Chemistry 282, 37359–69CrossRefGoogle ScholarPubMed
Ono, R, Nakamura, K, Inoue, K et al. (2006). Deletion of Peg10, an imprinted gene acquired from a retrotransposon, causes early embryonic lethality. Nature Genetics 38, 101–6; Lim, AL, Ng, S, Leow, SCP et al. (2012). Epigenetic state and expression of imprinted genes in umbilical cord correlates with growth parameters in human pregnancy. Journal of Medical Genetics 49, 689–97CrossRefGoogle ScholarPubMed
Suzuki, S, Ono, R, Narita, T et al. (2007). Retrotransposon silencing by DNA methylation can drive mammalian genomic imprinting. PLoS Genetics 3, e55CrossRefGoogle ScholarPubMed
Sekita, Y, Wagatsuma, H, Nakamura, K et al. (2008). Role of transposon-derived imprinted gene, RTL1, in the feto-maternal interface of mouse placenta. Nature Genetics 40, 243–8CrossRefGoogle Scholar
Edwards, CA, Mungall, AJ, Matthews, L et al. (2008). The evolution of the DLK1-DIO3 imprinted domain in mammals. PLoS Biology 6, e135CrossRefGoogle ScholarPubMed
Knox, K and Baker, JC (2008). Genomic evolution of the placenta using co-option and duplication and divergence. Genome Research 8, 695–705; Rawn, SM and Cross, JC (2008). The evolution, regulation, and function of placental-specific genes. Annual Review of Cell and Developmental Biology 24, 159–81CrossRefGoogle Scholar
Sugimoto, J and Schust, DJ (2009). Human endogenous retroviruses and the placenta. Reproductive Sciences 16, 1023–33; Kaneko-Ishino, T and Ishino, F (2010). Retrotransposon silencing by DNA methylation contributed to the evolution of placentation and genomic imprinting in mammals. Development, Growth and Differentiation 52, 533–43; Dupressoir, A, Lavialle, C and Heidmann, T(2012). From ancestral infectious retroviruses to bona fide cellular genes: role of the captured syncytins in placentation. Placenta 33, 663–71CrossRefGoogle ScholarPubMed
Piriyapongsa, J, Polavarapu, N, Borodovsky, M and McDonald, J (2007). Exonization of the LTR transposable elements in human genome. BMC Genomics 8, 291CrossRefGoogle ScholarPubMed
Huh, JW, Kim, TH, Yi, JM et al. (2006). Molecular evolution of the periphilin gene in relation to human endogenous retrovirus M element. Journal of Molecular Evolution 62, 730–7CrossRefGoogle ScholarPubMed
Ling, J, Pi, W, Bollag, R et al. (2002). The solitary long terminal repeats of ERV-9 endogenous retroviruses are conserved during primate evolution and possess enhancer activities in embryonic and hematopoietic cells. Journal of Virology 76, 2410–23; Pi, W, Zhu, X, Wu, M et al. (2010). Long-range function of an intergenic retrotransposon. Proceedings of the National Academy of Sciences of the USA 107, 12992–7CrossRefGoogle ScholarPubMed
Dunn, CA, van de Lagemaat, LN, Baillie, GJ and Mager, DL (2005). Endogenous retrovirus long terminal repeats as ready-to-use mobile promoters: the case of primate β3GAL- T5. Gene 364, 2–12CrossRefGoogle ScholarPubMed
Schulte, AM and Wellstein, A (1998). Structure and phylogenetic analysis of an endogenous retrovirus inserted into the human growth factor gene pleiotrophin. Journal of Virology 72, 6065–72; Ball, M, Carmody, M, Wynne, F et al. (2009). Expression of pleiotrophin and its receptors in human placenta suggests roles in trophoblast life cycle and angiogenesis. Placenta 30, 649–53Google ScholarPubMed
Bieche, I, Laurent, A, Laurendeau, I et al. (2003). Placenta-specific INSL4 expression is mediated by a human endogenous retroviral element. Biology of Reproduction 68, 1422–9CrossRefGoogle Scholar
Landry, J-R and Mager, DL (2003). Functional analysis of the endogenous retroviral promoter of the human endothelin B receptor gene. Journal of Virology 77, 7459–66CrossRefGoogle ScholarPubMed
Cohen, CJ, Rebollo, R, Babovic, S et al. (2011). Placenta-specific expression of the IL-2 receptor B subunit from an endogenous retroviral promoter. Journal of Biological Chemistry 286, 35543–52CrossRefGoogle Scholar
Emera, D, Casola, C, Lynch, VJ et al. (2012). Convergent evolution of endometrial prolactin expression in primates, mice, and elephants through the independent recruitment of transposable elements. Molecular Biology and Evolution 29, 39–47; Emera, D and Wagner, DP (2012). Transformation of a transposon into a derived prolactin promoter with function during human pregnancy. Proceedings of the National Academy of Sciences of the USA 109, 11246–51CrossRefGoogle ScholarPubMed
Huh, J-W, Ha, H-S, Kim, D-S and Kim, H-S (2008). Placenta-restricted expression of LTR-derived NOS3. Placenta 29, 602–8CrossRefGoogle ScholarPubMed
Wang, T, Zeng, J, Lowe, CB et al. (2007). Species-specific endogenous retroviruses shape the transcriptional network of the human tumor suppressor protein p53. Proceedings of the National Academy of Sciences of the USA 104, 18613–18CrossRefGoogle ScholarPubMed
Beyer, U, Moll-Rocek, J, Moll, UM and Dobbelstein, M (2011). Endogenous retrovirus drives hitherto unknown proapoptotic p63 isoforms in the male germ line of humans and great apes. Proceedings of the National Academy of Sciences of the USA 108, 3624–9CrossRefGoogle ScholarPubMed
Conley, AB, Piriyapongsa, J and Jordan, IK (2008). Retroviral promoters in the human genome. Bioinformatics 24, 1563–7CrossRefGoogle ScholarPubMed
van de Lagemaat, LN, Landry, J-R, Mager, DL and Medstrand, P (2003). Transposable elements in mammals promote regulatory variation and diversification of genes with specialized functions. Trends in Genetics 19, 530–6; Cohen, CJ, Lock, WM and Mager, DL (2009). Endogenous retroviral LTRs as promoters for human genes: a critical assessment. Gene 448, 105–14CrossRefGoogle ScholarPubMed
Barbulescu, M, Turner, G, Su, M et al. (2001). A HERV-K provirus in chimpanzees, bonobos and gorillas, but not humans. Current Biology 11, 779–83CrossRefGoogle Scholar
Scally, et al. (2012), ref. 36
Maksakova, IA, Romanish, MT, Gagnier, L et al. (2006). Retroviral elements and their hosts: insertional mutagenesis in the mouse germ-line. PLoS Genetics 2, e2CrossRefGoogle ScholarPubMed
Theodorou, V, Kimm, MA, Boer, M et al. (2007). MMTV insertional mutagenesis identifies genes, gene families, and pathways involved in mammary cancer. Nature Genetics 39, 759–69; Ishiguro, T, Avila, H, Lin, SY et al. (2010). Gene trapping identifies chloride channel 4 as a novel inducer of colon cancer cell migration, invasion and metastasis. British Journal of Cancer 102, 774–82; Dail, M, Li, Q, McDaniel, A et al. (2010). Mutant Ikzf1, Kras, and Notch1 cooperate in T lineage leukemogenesis and modulate responses to targeted agents. Proceedings of the National Academy of Sciences of the USA 107, 5106–11CrossRefGoogle ScholarPubMed
Deichmann, et al. (2007), ref. 8
Barbulescu, et al. (1999), ref. 20
Hughes, and Coffin, (2001), ref. 30
Lee, YN and Bieniasz, PD (2007). Reconstitution of an infectious endogenous retrovirus. PLoS Pathogens 3, e10CrossRefGoogle Scholar
Barbulescu, et al. (2001), ref. 84
Dewannieux, M, Harper, F, Richaud, A et al. (2006). Identification of an infectious progenitor for the multiple copy HERV-K human endogenous retroelements. Genome Research 16, 1548–56CrossRefGoogle ScholarPubMed
Turner, et al. (2001), ref. 21
Agoni, et al. (2012), ref. 23
Lapuk, AV, Khil, PP, Lavrentieva, IV et al. (1999). A human endogenous retrovirus-like (HERV) LTR formed more than 10 million years ago due to an insertion of HERV-H LTR into the 5′ LTR of HERV-K is situated on human chromosomes 10, 19, and Y. Journal of General Virology 80, 835–9CrossRefGoogle Scholar
Young, GR, Eksmond, U, Salcedo, R et al. (2012). Resurrection of endogenous retrovirus in antibody-deficient mice. Nature 491, 774–8CrossRefGoogle ScholarPubMed
Stoye, (2012), ref. 14
Dewannieux, et al. (2006), ref. 93; Lee, and Bieniasz, (2007), ref. 91; Brady, T, Lee, YN, Ronen, K et al. (2009). Integration target site selection by a resurrected human endogenous retrovirus. Genes and Development 23, 633–42Google Scholar
Hanke, K, Kramer, P, Seeher, S et al. (2009). Reconstitution of the ancestral glycoprotein of a human endogenous retrovirus-K and modulation of its functional activity by truncation of the cytoplasmic domain. Journal of Virology 83, 12790–800CrossRefGoogle ScholarPubMed
Kraus, B, Boller, K, Reuter, A and Schnierle, BS (2011). Characterization of the human endogenous retrovirus K gag protein: identification of protease cleavage sites. Retrovirology 8, 21CrossRefGoogle ScholarPubMed
For ERV-Kcon sequence, see Lee, and Bieniasz, (2007), ref. 91 (Supporting Information); for K106 and K110 sequences, see Barbulescu, et al. (1999), ref. 20, available at and ; for regulatory elements, see Jha, et al. (2011), ref 22; Fuchs, NV, Kraft, M, Tondera, C et al. (2011). Expression of the human endogenous retrovirus (HERV) group HML-2/HERV-K does not depend on canonical promoter elements but is regulated by transcription factors Sp1 and Sp3. Journal of Virology 85, 3436–48Google Scholar
Horie, M, Honda, T, Suzuki, Y et al. (2010). Endogenous non-retroviral RNA elements in mammalian genomes. Nature 463, 84; Katzourakis, A and Gifford, RJ (2010). Endogenous viral elements in animal genomes. PLoS Genetics 6, e1001191; Feschotte, and Gilbert, (2012), ref. 14CrossRefGoogle ScholarPubMed
Liu, H, Fu, Y, Xie, J et al. (2011). Widespread endogenization of densoviruses and parvoviruses in animal and human genomes. Journal of Virology 85, 9863–76CrossRefGoogle ScholarPubMed
Arbuckle, JH, Medveczky, MM, Luka, J et al. (2010). The latent human herpesvirus-6A genome specifically integrates in telomeres of human chromosomes in vivo and in vitro. Proceedings of the National Academy of Sciences of the USA 107, 5563–8CrossRefGoogle ScholarPubMed
Kapitonov, VV, Pavlicek, A and Jurka, J. Anthology of repetitive DNA, in Meyers, RA (ed.), Encyclopedia of Molecular Cell Biology and Molecular Medicine Volume 1 (Weinheim: Wiley-VCH Verlag, 2004), 251–305; Kazazian, HH (2004). Mobile elements: drivers of genome evolution. Science 303, 1626–32; Jurka, J, Kapitonov, VV, Kohany, O and Jurka, MV (2007). Repetitive sequences in complex genomes: structure and evolution. Annual Review of Genomics and Human Genetics 8, 241–59; Wicker, T, Sabot, F, Hua-Van, A et al. (2007). A unified classification system for eukaryotic transposable elements. Nature Reviews Genetics 8, 973–82Google Scholar
Kramerov, DA and Vassetzky, NS (2011). Origin and evolution of SINEs in eukaryotic genomes. Heredity 107, 487–95CrossRefGoogle ScholarPubMed
Giordano, J, Ge, Y, Gelfand, Y et al. (2007). Evolutionary history of mammalian transposons determined by genome-wide defragmentation. PLoS Computational Biology 3, e137CrossRefGoogle ScholarPubMed
Deininger, PL and Batzer, MA (2002). Mammalian retroelements. Genome Research 12, 1455–65; Kazazian, HH (2007). Progress in understanding the biology of the human mutagen LINE-1. Human Mutation 28, 527–39; Goodier, JL and Kazazian, HH (2008). Retrotransposons revisited: the restraint and rehabilitation of parasites. Cell 135, 23–35; Cordaux, R and Batzer, MA (2009). The impact of retrotransposons on human genome evolution. Nature Reviews Genetics 10, 691–703CrossRefGoogle ScholarPubMed
Khan, H, Smit, A and Boissinot, S (2006). Molecular evolution and tempo of amplification of human LINE-1 retrotransposons since the origin of primates. Genome Research 16, 78–87; Kordis, D, Lovšin, N and Gubenšek, F (2006). Phylogenomic analysis of the L1 retrotransposons in Deuterostomia. Systematic Biology 55, 886–901; Waters, PD, Dobigny, G, Waddell, PJ and Robinson, TJ (2007). Evolutionary history of LINE-1 in the major clades of placental mammals. PLoS ONE 1, e258CrossRefGoogle ScholarPubMed
Giordano, et al. (2007), ref. 3
Eickbush, TH and Jumburuthugoda, VK (2008). The diversity of retrotransposons and the properties of their reverse transcriptases. Virus Research 134, 221–34CrossRefGoogle ScholarPubMed
Cordaux, R, Hedges, DJ and Batzer, MA (2004). Retrotransposition of Alu elements: how many sources?Trends in Genetics 20, 464–7; Bennett, EA, Keller, H, Mills, RE et al. (2008). Active Alu retrotransposons in the human genome. Genome Research 18, 1875–83CrossRefGoogle Scholar
Price, AL, Eskin, E and Pevzner, PA (2004). Whole-genome analysis of Alu repeat elements reveals complex evolutionary history. Genome Research 14, 2245–52; Churakov, G, Grundmann, N, Kuritzin, A et al. (2010). A novel web-based TinT application and the chronology of the primate Alu retroposon activity. BMC Evolutionary Biology 10, 376CrossRefGoogle ScholarPubMed
Wang, HH, Xing, J, Grover, D et al. (2005). SVA elements: a hominid-specific retroposon family. Journal of Molecular Biology 354, 994–1007; Hancks, DC and Kazazian, HH (2010). SVA retrotransposons: evolution and genetic instability. Seminars in Cancer Biology 20, 234–45CrossRefGoogle ScholarPubMed
Baillie, JK, Barnett, MW, Upton, KR et al. (2011). Somatic retrotransposition alters the genetic landscape of the human brain. Nature 479, 534–7; Tyekucheva, S, Yolken, RH, McCombie, WR et al. (2011). Establishing the baseline level of repetitive element expression in the human cortex. BMC Genomics 12, 495CrossRefGoogle ScholarPubMed
Kaneko, H, Dridi, S, Tarallo, V et al. (2011). DICER1 deficit induces Alu toxicity in age-related macular degeneration. Nature 471, 325–30; Tarrallo, V, Hirano, Y, Gelfand, BD et al. (2012). DICER1 loss and Alu RNA induce age-related macular degeneration via the NLRP3 inflammasome and MyD88. Cell 149, 847–59CrossRefGoogle ScholarPubMed
Wallace, NA, Belancio, VP and Deininger, PL (2008). L1 mobile element expression causes multiple types of toxicity. Gene 419, 75–81; Belancio, VP, Roy-Engel, AM, Pochampally, RR and Deininger, P (2010). Somatic expression of LINE-1 elements in human tissues. Nucleic Acids Research 38, 3909–22CrossRefGoogle ScholarPubMed
Han, JS and Boeke, JD (2004). A highly active synthetic mammalian retrotransposon. Nature 429, 314–18; van den Hurk, JAJM, Meij, IC, del Carmen Seleme, M et al. (2007). L1 retrotransposition can occur early in human embryonic development. Human Molecular Genetics 16, 1587–92; see also Bogerd, HP, Wiegand, HL, Hulme, AE et al. (2006). Cellular inhibitors of long interspersed element 1 and Alu retrotransposition. Proceedings of the National Academy of Sciences of the USA 103, 8780–5; Chiu, Y-L, Witkowska, HE, Hall, SC et al. (2006). High-molecular-mass APOBEC3G complexes restrict Alu retrotransposition. Proceedings of the National Academy of Sciences of the USA 103, 15588–93; Bennett, et al. (2008), ref. 8; Hancks, DC, Goodier, JL, Mandal, PK et al. (2011). Retrotransposition of marked SVA elements by human L1s in cultured cells. Human Molecular Genetics 20, 3386–400CrossRefGoogle ScholarPubMed
Muotri, AR, Chu, VT, Marchetto, MCN et al. (2005). Somatic mosaicism in neuronal precursor cells mediated by L1 retrotransposition. Nature 435, 903–10; An, W, Han, JS, Wheelan, SJ et al. (2006). Active retrotransposition by a synthetic L1 element in mice. Proceedings of the National Academy of Sciences of the USA 103, 18662–7; Okudaira, N, Goto, M, Yanobu-Takanashi, R et al. (2011). Involvement of retrotransposition of long interspersed nucleotide element-1 in skin carcinogenesis induced by 7, 12-dimethylbenz[a]anthracene and 12-O-tetradecanylphorbol-13-acetate. Cancer Science 102, 2000–6CrossRefGoogle ScholarPubMed
Miki, Y, Nishisho, I, Horii, A et al. (1991). Disruption of the APC gene by a retrotransposal insertion of L1 sequence in a colon cancer. Cancer Research 52, 643–5Google Scholar
Lee, E, Iskow, R, Yang, L et al. (2012). Landscape of somatic retrotransposition in human cancers. Science 337, 967–71CrossRefGoogle ScholarPubMed
Cordaux, R, Hedges, DJ, Herke, SW and Batzer, M (2006). Estimating the retrotransposition rate of human Alu elements. Gene 373, 134–7; Xing, J, Zhang, Y, Han, K et al. (2009). Mobile elements create structural variation: analysis of a complete human genome. Genome Research 19, 1516–26; Huang, CRL, Schneider, AM, Lu, Y et al. (2010). Mobile interspersed repeats are major structural variants in the human genome. Cell 141, 1171–82; Rouchka, E, Montoya-Durango, DE, Stribinskis, V et al. (2010). Assessment of genetic variation for the LINE-1 retrotransposon from next generation sequence data. BMC Bioinformatics 11 (Suppl 9), S12; Ewing, AD and Kazazian, HH (2011). Whole-genome resequencing allows detection of many rare LINE-1 insertion alleles in humans. Genome Research 21, 985–90CrossRefGoogle ScholarPubMed
Iskow, RC, McCabe, MT, Mills, RE et al. (2010). Natural mutagenesis of human genomes by endogenous retrotransposons. Cell 141, 1253–61; Hormozdiari, F, Alkan, C, Ventura, M et al. (2011). Alu repeat discovery and characterization within human genomes. Genome Research 21, 840–9CrossRefGoogle ScholarPubMed
Chen, J-M, Stenson, PD, Cooper, DN and Ferec, C (2005). A systematic study of LINE-1 endonuclease-dependent retrotranspositional events causing human genetic disease. Human Genetics 117, 411–27; Callinan, PA and Batzer, MA (2006). Retrotransposable elements and human disease. Genome Dynamics 1, 104–15; Chen, J-M, Ferec, C and Cooper, DN (2006). LINE-1 endonuclease-dependent retrotranspositional events causing human genetic disease: mutation detection bias and multiple mechanisms of target gene disruption. Journal of Biomedicine and Biotechnology 2006(1), Article ID 56182; Belancio, VP, Hedges, DJ and Deininger, P (2008). Mammalian non-LTR retrotransposons: for better or worse, in sickness and in health. Genome Research 18, 343–58; Belancio, VP, Deininger, PL and Roy-Engel, AM (2009). LINE dancing in the human genome: transposable elements and disease. Genome Medicine 1, 97; Burns, KH and Boeke, JD (2012). Human transposon tectonics. Cell 149, 740–52CrossRefGoogle ScholarPubMed
Wallace, MR, Andersen, LB, Saulino, AM et al. (1991). A de novo Alu insertion results in neurofibromatosis type 1. Nature 353, 864–6CrossRefGoogle Scholar
Gallus, GN, Cardaioli, E, Rufa, A et al. (2010). Alu-element insertion in an OPA1 intron sequence associated with autosomal dominant optic atrophy. Molecular Vision 16, 178–83Google Scholar
Watanabe, M, Kobayashi, K, Jin, F et al. (2005). Founder SVA retrotransposal insertion in Fukuyama-type congenital muscular dystrophy and its origin in Japanese and Northeast Asian populations. American Journal of Medical Genetics Part A 138, 344–8; Taniguchi-Ikeda, M, Kobayashi, K, Kanagawa, M et al. (2011). Pathogenic exon-trapping by SVA retrotransposon and rescue in Fukuyama muscular dystrophy. Nature 478, 127–31CrossRefGoogle ScholarPubMed
Takasu, M, Hayashi, R, Maruya, E et al. (2007). Deletion of entire HLA-A gene accompanied by an insertion of a retrotransposon. Tissue Antigens 70, 144–50CrossRefGoogle ScholarPubMed
Machado, PM, Brandao, RD, Cavaco, BM et al. (2007). Screening for a BRCA2 rearrangement in high-risk breast/ovarian cancer families: evidence for a founder effect and analysis of the associated phenotypes. Journal of Clinical Oncology 25, 2027–34CrossRefGoogle ScholarPubMed
Bouchet, C, Vuillaumier-Barrot, S, Gonzales, M et al. (2007). Detection of an Alu insertion in the POMT1 gene from three French Walker–Warburg syndrome families. Molecular Genetics and Metabolism 90, 93–6CrossRefGoogle Scholar
Konkel, MK and Batzer, MA (2010). A mobile threat to genome stability: the impact of non-LTR retrotransposons upon the human genome. Seminars in Cancer Biology 20, 211–21CrossRefGoogle ScholarPubMed
Katzir, et al. (1985), see Prologue, ref. 29
Ovchinnikov, I, Rubin, A and Swergold, GD (2002). Tracing the LINEs of human evolution. Proceedings of the National Academy of Sciences of the USA 99, 10522–7; Mathews, L, Chi, SY, Greenberg, N et al. (2003). Large differences between LINE-1 amplification rates in the human and chimpanzee lineages. American Journal of Human Genetics 72, 739–48CrossRefGoogle ScholarPubMed
The data are from the whole-genome studies, see Chapter 1, refs. 33–8. The detailed data for the LINE-1, LINE-2 and Alu elements were from the dataset of Prufer, et al. (2012), see Chapter 1, ref. 35; see also Mills, RE, Bennett, EA, Iskow, RC et al. (2006). Recently mobilized transposons in the human and chimpanzee genomes. American Journal of Human Genetics 78, 671–9; Mills, RE, Bennett, EA, Iskow, RC and Devine, SE (2007). Which transposable elements are active in the human genome?Trends in Genetics 23, 183–91; Lee, J, Cordaux, R, Han, K et al. (2007). Different evolutionary fates of recently integrated human and chimpanzee LINE-1 retrotransposons. Gene 390, 18–27Google Scholar
Buzdin, A, Ustyugova, S, Gogvadze, E et al. (2002). A new family of chimeric retrotranscripts formed by a full copy of U6 small nuclear RNA fused to the 3′ terminus of L1. Genomics 80, 402–6; Buzdin, A, Gogvadze, E, Kovalskaya, E et al. (2003). The human genome contains many types of chimeric retrogenes generated through in vivo RNA recombination. Nucleic Acids Research 31, 4385–90; Gogvadze, EV, Buzdin, AA and Sverdlov, ED (2005). Multiple template switches on LINE-directed reverse transcription: the most probable formation mechanism for the double and triple chimeric retroelements in mammals. Russian Journal of Bioorganic Chemistry 31, 74; Buzdin, A, Gogvadze, E and Lebrun, M-H (2007). Chimeric retrogenes suggest a role for the nucleolus in LINE amplification. FEBS Letters 581, 2877–82CrossRefGoogle ScholarPubMed
Hasnaoui, M, Doucet, AJ, Meziane, O and Gilbert, N (2009). Ancient repeat sequence derived from U6 snRNA in primate genomes. Gene 448, 139–44CrossRefGoogle ScholarPubMed
Ling, J, Zhang, L, Jin, H et al. (2004). Dynamic retrotransposition of ERV-9 LTR and L1 in the β-globin gene locus during primate evolution. Molecular Phylogenetics and Evolution 30, 867–71CrossRefGoogle ScholarPubMed
Hamdi, H, Nishio, H, Zielinski, R and Dugaiczyk, A (1999). Origin and phylogenetic distribution of Alu DNA repeats: irreversible events in the evolution of primates. Journal of Molecular Biology 289, 861–71CrossRefGoogle ScholarPubMed
Ajala, AR, Almeida, SS, Rangel, M et al. (2012). Association of ACE gene insertion/deletion polymorphism with birth weight, blood pressure levels, and ACE activity in healthy children. American Journal of Hypertension 7, 827–32CrossRefGoogle Scholar
Martinez, J, Dugaiczyk, LJ, Zielinski, R and Dugaiczyk, A (2001). Human genetic disorders: a phylogenetic perspective. Journal of Molecular Biology 308, 587–96CrossRefGoogle ScholarPubMed
Gibbons, R and Dugaiczyk, A (2005). Phylogenetic roots of Alu-mediated rearrangements leading to cancer. Genome 48, 160–7CrossRefGoogle Scholar
Hamdi, HK, Nishio, H, Travis, J et al. (2000). Alu-mediated phylogenetic novelties in gene regulation and development. Journal of Molecular Biology 299, 931–9CrossRefGoogle ScholarPubMed
Salem, A-H, Ray, DA, Xing, J et al. (2003). Alu elements and hominid phylogenetics. Proceedings of the National Academy of Sciences of the USA 100, 12787–91; Xing, J, Salem, A-H, Hedges, DJ et al. (2003). Comprehensive analysis of two Alu Yd subfamilies. Journal of Molecular Evolution 57 (Suppl 1), S76–89; Salem, A-H, Ray, DA, Hedges, DJ et al. (2005). Analysis of the human Alu Ye lineage. BMC Evolutionary Biology 5, 18CrossRefGoogle ScholarPubMed
Hedges, DJ, Callinan, PA, Cordaux, R et al. (2004). Differential Alu mobilization and polymorphism among the human and chimpanzee lineages. Genome Research 14, 1068–75CrossRefGoogle ScholarPubMed
Sheikh, TH and Deininger, PL (1996). The role and amplification of the HS Alu subfamily founder gene. Journal of Molecular Evolution 42, 15–21; Han, K, Xing, J, Wang, H et al. (2005). Under the genomic radar: the stealth model of Alu amplification. Genome Research 15, 655–64; Walker, JA, Konkel, MK, Ullmer, B et al. (2012). Orangutan Alu quiescence reveals possible source element: support for ancient backseat drivers. Mobile DNA 3, 8CrossRefGoogle Scholar
See Chapter 1, refs. 33–8
Lee, J, Ha, J, Son, S-Y and Han, K (2012). Human genome deletions generated by SVA-associated events. Comparative and Functional Genomics 2012, Article ID 807270; see also Chapter 1, ref. 38CrossRefGoogle ScholarPubMed
See Chapter 1, refs. 33–8
Kouprina, N, Pavlicek, A, Mochida, GH et al. (2004). Accelerated evolution of the ASPM gene controlling brain size begins prior to human brain expansion. PLoS Biology 2, e126; Pavlicek, A, Noskov, VN, Kouprina, N et al. (2004). Evolution of the tumor suppressor BRCA1 locus in primates: implications for cancer predisposition. Human Molecular Genetics 13, 2737–51CrossRefGoogle ScholarPubMed
Crouau-Roy, B and Clisson, I (2000). Evolution of an Alu DNA element of type Sx in the lineage of primates and the origin of an associated tetranucleotide microsatellite. Genome 43, 642–8CrossRefGoogle ScholarPubMed
Schmitz, J, Ohme, M and Zischler, H (2001). SINE insertions in cladistic analyses and the phylogenetic affiliations of Tarsius bancanus to other primates. Genetics 157, 777–84; Schmitz, J, Roos, C and Zischler, H (2005). Primate phylogeny: molecular evidence from retroposons. Cytogenetic and Genome Research 108, 26–37; for striking confirmation of simian-tarsier monophylicity, see Hartig G, Churakov G, Warren WC et al. (2013). Retrophylogenomics place tarsiers on the evolutionary branch of anthropoids. Scientific Report 3, 1756Google Scholar
Kuryshev, VY, Skryabin, BV, Kremerskothen, J et al. (2001). Birth of a gene: locus of neuronal BC200 snmRNA in three prosimians and human BC200 pseudogenes as archives of change in the Anthropoidea lineage. Journal of Molecular Biology 309, 1049–66CrossRefGoogle Scholar
Schmitz, J and Zischler, H (2003). A novel family of tRNA-related SINEs in the colugo and two new retrotransposable markers separating dermopterans from primates. Molecular Phylogenetics and Evolution 28, 341–9CrossRefGoogle Scholar
Liu, GE, Alkan, C, Jiang, L et al. (2009). Comparative analysis of Alu repeats in primate genomes. Genome Research 19, 876–85CrossRefGoogle ScholarPubMed
Kojima, KK (2010). Alu monomer revisited: recent generation of Alu monomers. Molecular Biology and Evolution 28, 13–15CrossRefGoogle ScholarPubMed
Park, S-J, Huh, J-W, Kim, Y-H et al. (2012). Intron retention and TE exonization events in ZRANB2. Comparative and Functional Genomics 2012, 170208CrossRefGoogle ScholarPubMed
Schmitz, J, Ohme, M, Suryobroto, B and Zischler, H (2002). The colugo (Cynocephalus variegatus, Dermoptera): the primates’ gliding sister?Molecular Biology and Evolution 19, 2308–12CrossRefGoogle ScholarPubMed
Levin, HL and Moran, JV (2011). Dynamic interactions between transposable elements and their hosts. Nature Reviews Genetics 12, 615–27CrossRefGoogle ScholarPubMed
Levy, A, Schwartz, S and Ast, G (2010). Large-scale discovery of insertion hotspots and preferential integration sites of human transposed elements. Nucleic Acids Research 38, 1515–30CrossRefGoogle ScholarPubMed
Wimmer, K, Callens, T, Wernstedt, A and Messiaen, L (2011). The NF1 gene contains hotspots for L1 endonuclease-dependent de novo insertion. PLoS Genetics 11, e1002371Google Scholar
Cantrell, MA, Filanoski, BJ, Ingermann, AR et al. (2001). An ancient retrovirus-like element contains hot spot for SINE insertion. Genetics 158, 769–77; for preferred integration sites in human disease, see Levin, and Moran, (2011), ref. 54Google ScholarPubMed
Vincent, BJ, Myers, JS, Ho, HJ et al. (2003). Following the LINEs: an analysis of primate genomic variation at human-specific LINE-1 insertion sites. Molecular Biology and Evolution 20, 1338–48; Ho, HJ, Ray, DA, Salem, A-H et al. (2005). Straightening out the LINEs: LINE-1 orthologous loci. Genomics 85, 201–7CrossRefGoogle ScholarPubMed
Roy-Engel, AM, Carroll, ML, El-Sawy, M et al. (2002). Non-traditional Alu evolution and primate genomic diversity. Journal of Molecular Biology 316, 1033–40; Salem, A-H, Kilroy, GE, Watkins, WS et al. (2003). Recently integrated Alu elements and human genomic diversity. Molecular Biology and Evolution 20, 1349–61CrossRefGoogle ScholarPubMed
Salem, et al. (2003), ref. 39
van de Lagemaat, L, Gagnier, L, Medstrand, P and Mager, DL (2005). Genomic deletions and precise removal of transposable elements mediated by short identical DNA segments in primates. Genome Research 15, 1243–9CrossRefGoogle ScholarPubMed
Ray, DA, Xing, J, Salem, A-H and Batzer, MA (2006). SINEs of a nearly perfect character. Systematic Biology 55, 928–35; Ray, DA (2007). SINEs of progress: mobile elements applications to molecular ecology. Molecular Ecology 16, 19–33CrossRefGoogle Scholar
Meyer, TJ, McLain, AT, Oldenburg, JM et al. (2012). An Alu-based phylogeny of gibbons (Hylobatidae). Molecular Biology and Evolution 29, 3441–50CrossRefGoogle Scholar
Xing, J, Wang, H, Han, K et al. (2005). A mobile element-based phylogeny of Old World monkeys. Molecular Phylogenetics and Evolution 37, 872–80; Xing, J, Wang, H, Zhang, Y et al. (2007). A mobile element-based evolutionary history of guenone (tribe Cercopithecini). BMC Biology 5, 5; Li, J, Han, K, Xing, J et al. (2009). Phylogeny of the macaques (Cercopithecidae: Macaca) based on Alu elements. Gene 448, 242–9CrossRefGoogle ScholarPubMed
Singer, SS, Schmitz, J, Schwiegk, C and Zischler, H (2002). Molecular cladistic markers in New World monkey phylogeny (Platyrrhini, Primates). Molecular Phylogenetics and Evolution 26, 490–501; Ray, DA, Xing, J, Hedges, DJ et al. (2005). Alu insertion loci and platyrrhine primate phylogeny. Molecular Phylogenetics and Evolution 35, 117–26CrossRefGoogle Scholar
Roos, C, Schmitz, J and Zischler, H (2004). Primate jumping genes elucidate strepsirrhine phylogeny. Proceedings of the National Academy of Sciences of the USA 101, 10650–4; McLain, AT, Meyer, TJ, Faulk, C et al. (2012). An Alu-based phylogeny of lemurs (Infraorder: Lemuriformes). PLoS ONE 7, e44035CrossRefGoogle ScholarPubMed
Xing, J, Witherspoon, DJ, Ray, DA et al. (2007). Mobile DNA elements in primate and human evolution. American Journal of Physical Anthropology 134 (Suppl 45), 2–19CrossRefGoogle Scholar
Herke, SW, Xing, J, Ray, DA et al. (2006). A SINE-based dichotomous key for primate identification. Gene 390, 39–51CrossRefGoogle ScholarPubMed
Murphy, WJ, Eizirik, E, O’Brien, SJ et al. (2001). Resolution of the early placental mammal radiation using Bayesian phylogenetics. Science 294, 2348–51; Binina-Emonds, ORP, Cardillo, M, Jones, KE et al. (2007). The delayed rise of present-day mammals. Nature 446, 507–12; Springer, MS and Murphy, WJ (2007). Mammalian evolution and biomedicine: new views from phylogeny. Biological Reviews of the Cambridge Philosophical Society 82, 375–92; Wildman, DE, Uddin, M, Opazo, JC et al. (2007). Genomics, biogeography, and the diversification of placental mammals. Proceedings of the National Academy of Sciences of the USA 104, 14395–400; Prasad, AB, Allard, MW, NISC Comparative Sequencing Program and Green, ED (2008). Confirming the phylogeny of mammals by use of large comparative sequence data sets. Molecular Biology and Evolution 25, 1795–808; Meredith, RW, Janecka, JE, Gatesy, J et al. (2011). Impacts of the Cretaceous terrestrial revolution and the KPg extinction on mammal diversification. Science 334, 521–4; Stringer, MS, Meredith, RW, Janecka, JE and Murphy, WJ (2011). The historical biogeography of Mammalia. Philosophical Transactions of the Royal Society of London Series B, Biological Sciences 366, 2478–502CrossRefGoogle ScholarPubMed
Thomas, JW, Touchman, JW, Blakesley, RW et al. (2003). Comparative analyses of multi-species sequences from targeted genomic regions. Nature 424, 788–93 (Phylogenetic Analyses, Supplementary Information 5)CrossRefGoogle ScholarPubMed
Hughes, DC (2000). MIRs as agents of mammalian gene evolution. Trends in Genetics 16, 60–2CrossRefGoogle ScholarPubMed
Lin, L, Jiang, P, Shen, S et al. (2009). Large-scale analysis of exonised mammalian-wide interspersed repeats in primate genomes. Human Molecular Genetics 18, 2204–14CrossRefGoogle Scholar
Silva, JC, Shabalina, SA, Harris, DG et al. (2003). Conserved fragments of transposable elements in intergenic regions: evidence for widespread recruitment of MIR- and L2-derived sequences within the mouse and human genomes. Genetical Research 82, 1–18; Zhu, L, Swergold, GD and Seldin, MF (2003). Examination of sequence homology between human chromosome 20 and the mouse genome: intense conservation of many genomic elements. Human Genetics 113, 60–70CrossRefGoogle ScholarPubMed
Kriegs, JO, Churakov, G, Jurka, J et al. (2007). Evolutionary history of 7SL RNA-derived SINEs in Supraprimates. Trends in Genetics 23, 158–61CrossRefGoogle ScholarPubMed
Armiger, V, Erlandsson, R, Pielberg, G et al. (2003). Comparative sequence analysis of the PRKAG3 region between human and pig: evolution of repetitive sequences and potential new exons. Cytogenetic and Genome Research 102, 163–72Google Scholar
Lowe, CB, Bejerano, G and Haussler, D (2007). Thousands of human mobile element fragments undergo strong purifying selection near developmental genes. Proceedings of the National Academy of Sciences of the USA 104, 8005–10CrossRefGoogle ScholarPubMed
Lindblad-Toh, K, Garber, M, Zuk, O et al. (2011). A high-resolution map of human evolutionary constraint using 29 mammals. Nature 478, 476–82CrossRefGoogle ScholarPubMed
Kriegs, JO, Churakov, G, Kiefmann, M et al. (2006). Retroposed elements as archives for the evolutionary history of placental mammals. PLoS Biology 4, e91CrossRefGoogle ScholarPubMed
Nishihara, H, Hasegawa, M and Okada, N (2006). Pegasoferae, an unexpected mammalian clade revealed by tracking ancient retroposon insertions. Proceedings of the National Academy of Sciences of the USA 103, 9929–34 (see Supporting Information)CrossRefGoogle ScholarPubMed
Yu, L and Zhang, Y-P (2005). Evolutionary implications of multiple SINE insertions in an intronic region from diverse mammals. Mammalian Genome 16, 651–60CrossRefGoogle Scholar
Smith, AM, Sanchez, M-J, Follows, GA et al. (2008). A novel mode of enhancer evolution: the TAL1 stem cell enhancer recruited a MIR element to specifically boost its activity. Genome Research 18, 1422–32CrossRefGoogle Scholar
Kriegs, et al. (2006), ref. 78
Blanchette, M, Green, ED, Miller, W and Haussler, D (2004). Reconstructing large regions of an ancestral mammalian genome in silico. Genome Research 14, 2412–23CrossRefGoogle ScholarPubMed
Murphy, WJ, Pringle, TH, Crider, TA et al. (2007). Using genomic data to unravel the root of the placental mammal tree. Genome Research 17, 413–21CrossRefGoogle Scholar
Churakov, G, Kriegs, JO, Baertsch, R et al. (2009). Mosaic retroposon insertion patterns in placental mammals. Genome Research 19, 868–75CrossRefGoogle ScholarPubMed
Nishihara, H, Maruyama, S and Okada, N (2009). Retroposon analysis and recent geological data suggest near-simultaneous divergence of the three superorders of mammals. Proceedings of the National Academy of Sciences of the USA 106, 5235–40CrossRefGoogle ScholarPubMed
Data from Kriegs, et al. (2006), ref. 78; Nishihara, et al. (2006), ref. 79; Kriegs, et al. (2007), ref. 74; Churakov, et al. (2009), ref. 85; Nishihara, et al. (2009), ref. 86; data in dark grey boxes are from refs. 88–96
Krull, M, Petrusma, M, Makalowski, W et al. (2007). Functional persistence of exonised mammalian-wide interspersed repeat elements (MIRs). Genome Research 17, 1139–45CrossRefGoogle Scholar
Santangelo, AM, de Souza, FSJ, Franchini, LF et al. (2007). Ancient exaptation of a CORE-SINE retroposon into a highly conserved mammalian neuronal enhancer of the proopiomelanocortin gene. PLoS Genetics 3, e166; Alfoldi, J, De Palma, F, Grabherr, M et al. (2011). The genome of the green anole lizard and a comparative analysis with birds and mammals. Nature 477, 587–91 (Supplementary Table 18); Franchini, LF, Lopez-Leal, R, Nasif, S et al. (2011). Convergent evolution of two mammalian neuronal enhancers by sequential exaptation of unrelated retroposons. Proceedings of the National Academy of Sciences of the USA 108, 15270–5CrossRefGoogle ScholarPubMed
Kamal, M, Xie, X and Lander, ES (2006). A large family of ancient repeat elements in the human genome is under strong selection. Proceedings of the National Academy of Sciences of the USA 103, 2740–5; see also Lowe, et al. (2007), ref. 76CrossRefGoogle ScholarPubMed
Bejerano, G, Lowe, CB, Ahituv, N et al. (2006). A distal enhancer and an ultraconserved exon are derived from a novel retroposon. Nature 441, 87–90CrossRefGoogle ScholarPubMed
Lowe, CB, Bejerano, G, Salama, SR and Haussler, D (2010). Endangered species hold clues to human evolution. Journal of Heredity 101, 437–47CrossRefGoogle ScholarPubMed
Nishihara, H, Smit, AFA and Okada, N (2006). Functional sequences derived from SINEs in the mammalian genome. Genome Research 16, 864–74; Xie, X, Kamal, M and Lander, ES (2006). A family of conserved noncoding elements derived from an ancient transposable element. Proceedings of the National Academy of Sciences of the USA 103, 11659–64; Hirakawa, M, Nishihara, H, Kanehisa, M and Okada, N (2009). Characterization and evolutionary landscape of AmnSINE1 in Amniota genomes. Gene 441, 100–10CrossRefGoogle ScholarPubMed
Sasaki, T, Nishihara, H, Hirakawa, M et al. (2008). Possible involvement of SINEs in mammalian-specific brain formation. Proceedings of the National Academy of Sciences of the USA 105, 4220–5; Okada, N, Sasaki, T, Shimogori, T and Nishihara, H (2010). Emergence of mammals by emergency: exaptation. Genes to Cells 15, 801–12CrossRefGoogle ScholarPubMed
Gentles, AJ, Wakefield, MJ, Kohany, O et al. (2007). Evolutionary dynamics of transposable elements in the short-tailed opossum Monodelphis domestica. Genome Research 17, 992–1004CrossRefGoogle ScholarPubMed
Elisaphenko, EA, Kolesnikov, NN, Shevchenko, AI et al. (2008). A dual origin of the XIST gene from a protein-coding gene and a set of transposable elements. PLoS ONE 3, e2521CrossRefGoogle ScholarPubMed
Jurka, J, Bao, W, Kojima, KK et al. (2012). Different groups of repetitive elements preserved in mammals correspond to different periods of regulatory innovation in vertebrates. Biology Direct 7, 36CrossRefGoogle Scholar
Alfoldi, et al. (2011), ref. 89
Churakov, G, Sadasivuni, MK, Rosenbloom, KR et al. (2010). Rodent evolution: back to the root. Molecular Biology and Evolution 27, 1315–26; Kriegs, JO, Zemann, A, Churakov, G et al. (2010). Retroposon insertions provide insights into deep lagomorph evolution. Molecular Biology and Evolution 27, 2678–81CrossRefGoogle ScholarPubMed
Hallstrom, BM, Schneider, A, Zoller, S and Janke, A (2011). A genomic approach to examine the complex evolution of laurasiatherian mammals. PLoS ONE 6, e28199CrossRefGoogle ScholarPubMed
Shimamura, M, Yasue, H, Ohshima, K et al. (1997). Molecular evidence from retroposons that whales form a clade within even-toed ungulates. Nature 388, 666–70; Shedlock, AM and Okada, N (2000). SINE insertions: powerful tools for molecular systematics. BioEssays 22, 148–60; Wong, K (2002). The mammals that conquered the seas. Scientific American 286(5), 70–9CrossRefGoogle Scholar
Nikaido, M, Matsuno, F, Hamilton, H et al. (2001). Retroposon analysis of major cetacean lineages: the monophyly of toothed whales and the paraphyly of river dolphins. Proceedings of the National Academy of Sciences of the USA 98, 7384–9; Sasaki, T, Nikaido, M, Wada, S et al. (2006). Balaenoptera omurai is a newly discovered baleen whale that represents an ancient evolutionary lineage. Molecular Phylogenetics and Evolution 41, 40–52; Chen, Z, Xu, S, Zhou, K and Yang, G (2011). Whale phylogeny and rapid radiation events revealed using novel retrotransposed elements and their flanking sequences. BMC Evolutionary Biology 11, 314; Meredith, RW, Gatesy, J, Cheng, J and Springer, MS (2011). Pseudogenisation of the tooth gene enamelysin (MMP20) in the common ancestor of extant baleen whales. Proceedings, Biological Sciences/The Royal Society 278, 993–1002CrossRefGoogle ScholarPubMed
Nijman, IJ, van Tessel, P and Lenstra, JA (2002). SINE retrotransposon during the evolution of the pecoran ruminants. Journal of Molecular Evolution 54, 9–16; Nilsson, M, Klassert, D, Bertelsen, M et al. (2012). Activity of ancient RTE retroposons during the evolution of cows, spiral-horned antelopes and nilgais (Bovinae). Molecular Biology and Evolution 29, 2885–8CrossRefGoogle ScholarPubMed
Nishihara, H, Satta, Y, Nikaido, M et al. (2005). A retroposon analysis of Afrotherian phylogeny. Molecular Biology and Evolution 22, 1823–33; Moller-Krull, M, Delsuc, F, Churakov, G et al. (2007). Retroposed elements and their flanking regions resolve the evolutionary history of xenarthran mammals (armadillos, anteaters, and sloths). Molecular Biology and Evolution 24, 2573–82CrossRefGoogle ScholarPubMed
Nilsson, MA, Churakov, G, Sommer, M et al. (2010). Tracking marsupial evolution using archaic genomic retroposon insertions. PLoS Biology 8, e1000436CrossRefGoogle ScholarPubMed
Watanabe, M, Nikaido, M, Tsuda, TT et al. (2006). The rise and fall of the CR1 subfamily in the lineage leading to penguins. Gene 365, 57–66; Kaiser, VB, van Tuinen, M and Ellegren, H (2007). Insertion events of CR1 retrotransposable elements elucidate the phylogenetic branching order in galliform birds. Molecular Biology and Evolution 24, 338–47CrossRefGoogle ScholarPubMed
Suh, A, Paus, M, Kiefmann, M et al. (2011). Mesozoic retroposons reveal parrots as the closest living relatives of passerine birds. Nature Communications 2, 443CrossRefGoogle ScholarPubMed
Han, K-L, Braun, EL, Kimball, RT et al. (2011). Are transposable element insertions homoplasy free? An examination using the avian tree of life. Systematic Biology 60, 375–86CrossRefGoogle ScholarPubMed
Shedlock, AM, Takahashi, K and Okada, N (2004), SINEs of speciation: tracking lineages with retroposons. Trends in Ecology and Evolution 19, 545–53CrossRefGoogle ScholarPubMed
Kuryshev, et al. (2001), ref. 48; see also Martignetti, JA and Brosius, J (1993). BC200 RNA: a neural RNA polymerase III product encoded by a monomer Alu element. Proceedings of the National Academy of Sciences of the USA 90, 11563–7; Duning, K, Buck, F, Barnekow, A and Kremerskothen, J (2008). SYNCRIP, a component of dendritically localised mRNPs, binds to the translation regulator BC200 RNA. Journal of Neurochemistry 105, 351–9Google Scholar
Parrott, AM, Tsai, M, Batchu, P et al. (2011). The evolution and expression of the snaR family of small non-coding RNAs. Nucleic Acids Research 39, 1485–500; Parrott, AM and Mathews, MB (2011). The evolution and consequences of snaR family transposition in primates. Mobile Genetic Elements 1, 291–5CrossRefGoogle ScholarPubMed
Courseaux, A and Nahon, J-L (2001). Birth of two chimeric genes in the Hominidae lineage. Science 291, 1293–7; Schmieder, S, Darre-Toulemonde, F, Arguel, M-J et al. (2008). Primate-specific spliced PMCHL RNAs are non-protein coding in human and macaque tissues. BMC Evolutionary Biology 8, 330CrossRefGoogle ScholarPubMed
Li, C-Y, Zhang, Y, Wang, Z et al. (2010). A human-specific de novo protein-coding gene associated with human brain functions. PLoS Computational Biology 6, e1000734CrossRefGoogle ScholarPubMed
Wu, M, Li, L and Sun, Z (2007). Transposable element fragments in protein-coding regions and their contributions to human functional proteins. Gene 401, 165–71; Schmitz, J and Brosius, J (2011). Exonization of transposed elements: a challenge and opportunity for evolution. Biochimie 93, 1928–34CrossRefGoogle ScholarPubMed
Damert, A, Lower, J and Lower, R (2004). Leptin receptor isoform 219.1: an example of protein evolution by LINE-1-mediated human-specific retrotransposition of a coding SVA element. Molecular Biology and Evolution 21, 647–51CrossRefGoogle ScholarPubMed
Mola, G, Vela, V, Fernandez-Figueras, MI et al. (2007). Exonization of Alu-generated splice variants in the survivin gene of human and non-human primates. Journal of Molecular Biology 366, 1055–63CrossRefGoogle ScholarPubMed
Krull, M, Brosius, J and Schmitz, J (2005). Alu-SINE exonization: en route to protein-coding function. Molecular Biology and Evolution 22, 1702–11CrossRefGoogle ScholarPubMed
Huh, J-W, Kim, Y-H, Lee, S-R et al. (2010). Four different ways of alternative transcripts generation mechanism in ADRA1A gene. Genes and Genetic Systems 85, 65–73CrossRefGoogle ScholarPubMed
Moller-Krull, M, Zemann, A, Roos, C et al. (2008). Beyond DNA: RNA editing and steps toward Alu exonization in primates. Journal of Molecular Biology 382, 601–9CrossRefGoogle ScholarPubMed
Singer, SS, Mannel, DN, Hehlgans, T et al. (2004). From ‘junk’ to gene: curriculum vitae of a primate receptor isoform gene. Journal of Molecular Biology 341, 883–6CrossRefGoogle ScholarPubMed
Lee, J-R, Huh, J-W, Kim, D-S et al. (2009). Lineage-specific evolutionary events on SFTPB gene: Alu recombination-mediated deletion (ARMD), exonization, and alternative splicing events. Gene 435, 29–35CrossRefGoogle Scholar
Park, et al. (2012), ref. 52
Lin, L, Shen, S, Tye, A et al. (2008). Diverse splicing patterns of exonised Alu elements in human tissues. PLoS Genetics 4, e1000225CrossRefGoogle Scholar
Hasler, J and Strub, K (2006). Alu elements as regulators of gene expression. Nucleic Acids Research 34, 5491–7; Feschotte, C (2008). Transposable elements and the evolution of regulatory networks. Nature Reviews Genetics 9, 397–405CrossRefGoogle ScholarPubMed
Pheasant, M and Mattick, JS (2007). Raising the estimate of functional human sequences. Genome Research 17, 245–53; Petherick, A (2008). The production line. Nature 454, 1042–5; Mattick, JS (2011). The central role of RNA in human development and cognition. FEBS Letters 585, 1600–16CrossRefGoogle ScholarPubMed
Mikkelsen, TS, Wakefield, MJ, Aken, B et al. (2007). Genome of the marsupial Monodelphis domestica reveals innovation in non-coding sequences. Nature 447, 167–77CrossRefGoogle ScholarPubMed
Lowe, CB and Haussler, D (2012). 29 mammalian genomes reveal novel exaptations of mobile elements for likely regulatory functions in the human genome. PLoS ONE 7, e43128CrossRefGoogle ScholarPubMed
Thurman, RE, Rynes, E, Humbert, R et al. (2012). The accessible chromatin landscape of the human genome. Nature 489, 75–82 (Supplementary Material, Figures 4 and 5, and Table 3)CrossRefGoogle ScholarPubMed
Conley, AB, Miller, WJ and Jordan, IK (2008). Human cis natural antisense transcripts initiated by transposable elements. Trends in Genetics 24, 53–6; Faulkner, GL, Kimura, Y, Daub, CO et al. (2009). The regulated retrotransposon transcriptome of mammalian cells. Nature Genetics 41, 563CrossRefGoogle ScholarPubMed
Bourque, G, Leong, B, Vega, VB et al. (2008). Evolution of the mammalian transcription factor binding repertoire via transposable elements. Genome Research 18, 1752–62; Polavarapu, N, Marino-Ramirez, L, Landsman, D et al. (2008). Evolutionary rates and patterns for human transcription factor binding sites derived from repetitive DNA. BMC Genomics 9, 226CrossRefGoogle ScholarPubMed
Polak, P and Domany, E (2006). Alu elements contain many binding sites for transcription factors and may play a role in regulation of developmental processes. BMC Genomics 7, 133CrossRefGoogle ScholarPubMed
Gombart, AF, Saito, T and Koeffler, HP (2009). Exaptation of an ancient Alu short interspersed element provides a highly conserved vitamin D-mediated innate immune response in humans and primates. BMC Genomics 10, 321CrossRefGoogle ScholarPubMed
Laperriere, D, Wang, T-T, White, JH and Mader, S (2007). Widespread Alu repeat-driven expansion of consensus DR2 retinoic acid response elements during primate evolution. BMC Genomics 8, 23; Mason, CE, Shu, F-J, Wang, C et al. (2010). Location analysis for the estrogen receptor-α reveals binding to diverse ERE sequences and widespread binding within repetitive DNA elements. Nucleic Acids Research 38, 2355–68; Bolotin, E, Chellappa, K, Huang-Verslues, W et al. (2011). Nuclear receptor HNF4alpha binding sequences are widespread in Alu repeats. BMC Genomics 12, 560CrossRefGoogle ScholarPubMed
Zemojtel, T, Kielbasa, SM, Arndt, PF et al. (2009). Methylation and deamination of CpGs generate p53-binding sites on a genomic scale. Trends in Genetics 25, 63–6CrossRefGoogle ScholarPubMed
Zemojtel, T, Kielbasa, SM, Arndt, PF et al. (2011). CpG deamination creates transcription factor binding sites with high efficiency. Genome Biology and Evolution 3, 1304–11CrossRefGoogle ScholarPubMed
Antonaki, A, Demetriades, C, Polyzos, A et al. (2011). Genomic analysis reveals a novel NK-κB binding site in Alu repetitive elements. Journal of Biological Chemistry 286, 38768–82CrossRefGoogle Scholar
Pandey, R, Mandal, AK, Jha, V and Mukerji, M (2011). HSF binding in Alu repeats expands its involvement in stress through an antisense mechanism. Genome Biology 12, R117CrossRefGoogle ScholarPubMed
Lynch, VJ, Leclerc, RD, May, G and Wagner, GP (2011). Transposon-mediated re-wiring of gene regulatory networks contributed to the evolution of pregnancy in mammals. Nature Genetics 43, 1154–9CrossRefGoogle Scholar
Johnson, R, Gamblin, RJ, OOi, L et al. (2006). Identification of the REST regulon reveals extensive transposable element-mediated binding site duplication. Nucleic Acids Research 34, 3862–77CrossRefGoogle ScholarPubMed
Schmidt, D, Schwalie, PC, Wilson, MD et al. (2012). Waves of retrotransposon expansion remodel genome organization and CTCF binding in multiple mammalian lineages. Cell 148, 335–48CrossRefGoogle ScholarPubMed
Reviewed in ref. 20
Han, K, Lee, J, Meyer, TJ et al. (2008). L1 recombination-associated deletions generate human genomic variation. Proceedings of the National Academy of Sciences of the USA 105, 19366–71CrossRefGoogle ScholarPubMed
Sen, SK, Han, K, Wang, J et al. (2006). Human genomic deletions mediated by recombination between Alu elements. American Journal of Human Genetics 79, 41–53CrossRefGoogle ScholarPubMed
Han, K, Lee, J, Meyer, TJ et al. (2007). Alu recombination-mediated structural deletions in the chimpanzee genome. PLoS Genetics 3, 1939–49CrossRefGoogle ScholarPubMed
Wang, X, Mitra, N, Secundino, I et al. (2012). Specific inactivation of two immunomodulatory SIGLEC genes during human evolution. Proceedings of the National Academy of Sciences of the USA 109, 9935–40CrossRefGoogle ScholarPubMed
Zhang, R, Wang, Y-Q and Su, B (2008). Molecular evolution of a primate-specific microRNA family. Molecular Biology and Evolution 25, 1493–502; Lehnert, S, Van Loo, P, Thilakarathne, PJ et al. (2009). Evidence for co-evolution between human microRNAs and Alu-repeats. PLoS ONE 4, e4456CrossRefGoogle ScholarPubMed
Lee, J, Han, K, Meyer, TJ et al. (2008). Chromosomal inversions between human and chimpanzee lineages caused by retrotransposons. PLoS ONE 3, e4047CrossRefGoogle ScholarPubMed
Kelkar, YD, Eckert, KA, Chiaromonte, F and Makova, KD (2011). A matter of life or death: how microsatellites emerge in and vanish from the human genome. Genome Research 21, 2038–48; Ahmed, M and Liang, P (2012). Transposable elements are a significant contributor to tandem repeats in the human genome. Comparative and Functional Genomics 2012, Article ID 947089CrossRefGoogle ScholarPubMed
Kurosaki, T, Ueda, S, Ishida, T et al. (2012). The unstable CCTG repeat responsible for myotonic dystrophy type 2 originates from an AluSx element insertion into an early primate genome. PLoS ONE 7, e38379 (Supporting Information, Figure S4)CrossRefGoogle Scholar
Thompson, R, Zoppis, S and McCord, B (2012). An overview of DNA typing methods for human identification: past, present, and future. Methods in Molecular Biology 830, 3–16CrossRefGoogle ScholarPubMed
Huda, A and Jordan, IK (2009). Epigenetic regulation of mammalian genomes by transposable elements. Annals of the New York Academy of Sciences 1178, 276–84CrossRefGoogle ScholarPubMed
Oliver, KR and Greene, WK (2009). Transposable elements: powerful facilitators of evolution. BioEssays 31, 703–14; Oliver, KR and Green, WK (2011). Mobile DNA and the TE-thrust hypothesis: supporting evidence from the primates. Mobile DNA 2, 8CrossRefGoogle Scholar
Hagan, CR, Sheffield, RF and Rudin, CM (2003). Human Alu element retrotransposition induced by genotoxic stress. Nature Genetics 3, 219–20; Stribinskis, V and Ramos, KS (2006). Activation of human long interspersed nuclear element 1 retrotransposition by benzo(a)pyrene, an ubiquitous environmental carcinogen. Cancer Research 66, 2616–20; Teneng, I, Stribinskis, V and Ramos, KS (2007). Context-specific regulation of LINE-1. Genes to Cells 12, 1101–10; Okudaira, N, Iijima, K, Koyama, T et al. (2010). Induction of long interspersed nucleotide element-1 (L1) retrotransposition by 6-formylindolo[3,2-b]carbazole (FICZ), a tryptophan photoproduct. Proceedings of the National Academy of Sciences of the USA 107, 18487–92; Giorgi, G, Marcantonio, P and Del Re, B (2011). LINE-1 retrotransposition in human neuroblastoma cells is affected by oxidative stress. Cell and Tissue Research 346, 383–91; Okudaira, et al. (2011), ref. 15; Teneng, I, Montoya-Durango, DE, Quertermous, JL et al. (2011). Reactivation of L1 retrotransposon by benzo(a)pyrene involves complex genetic and epigenetic regulation. Epigenetics 6, 355–67CrossRefGoogle Scholar
Ishizaka, Y, Okudaira, N, Tamura, M et al. (2012). Modes of retrotransposition by long interspersed element-1 by environmental factors. Frontiers in Microbiology 3, Article 191CrossRefGoogle ScholarPubMed
Zeh, DW, Zeh, JA and Ishida, Y (2009). Transposable elements and an epigenetic basis for punctuated equilibria. BioEssays 31, 715–26; Rebollo, R, Horard, B, Hubert, B and Vieira, C (2010). Jumping genes and epigenetics: towards new species. Gene 454, 1–7CrossRefGoogle ScholarPubMed
Britten, RJ (2010). Transposable element insertions have strongly affected human evolution. Proceedings of the National Academy of Sciences of the USA 107, 19945–8CrossRefGoogle ScholarPubMed
Mattick, (2011), ref. 125
de Koning, AP, Gu, W, Castoe, TA et al. (2011). Repetitive elements may comprise over two-thirds of the human genome. PLoS Genetics 7, e1002384CrossRefGoogle ScholarPubMed
Werren, JH (2011). Selfish genetic elements, genetic conflict, and evolutionary innovation. Proceedings of the National Academy of Sciences of the USA 108 (Suppl. 2), 10863–70CrossRefGoogle ScholarPubMed
Interview (1998). Dr. Charles Dinarello elected to the U.S. National Academy of Sciences. International Cytokine Society Newsletter 6, 1
Drayna, D (2005). Founder mutations. Scientific American 294(4), 78–85CrossRefGoogle Scholar
Eiberg, H, Troelsen, J, Nielsen, M et al. (2008). Blue eye colour in humans may be caused by a perfectly associated founder mutation in a regulatory element located within the HERC2 gene inhibiting OCA2 expression. Human Genetics 123, 177–87; Sturm, RA, Duffy, DL, Zhao, ZZ et al. (2008). A single SNP in an evolutionary conserved region within intron 86 of the HERC2 gene determines human blue–brown eye colour. American Journal of Human Genetics 82, 424–31CrossRefGoogle Scholar
Dalbagni, DG, Zhi-Ping, R, Herr, H et al. (2001). Genetic alterations in TP53 in recurrent urothelial cancer: a longitudinal study. Clinical Cancer Research 7, 2797–801; Denzinger, S, Mohren, K, Knuechel, R et al. (2006). Improved clonality analysis of multifocal bladder tumors by combination of histopathologic organ mapping, loss of heterozygosity, fluorescence in situ hybridization, and p53 analysis. Human Pathology 37, 143–51Google ScholarPubMed
Hafner, C, Knuechel, R, Stoehr, R and Hartman, A (2002). Clonality of multifocal urothelial carcinomas: 10 years of molecular genetic studies. International Journal of Cancer 101, 1–6CrossRefGoogle ScholarPubMed
Gerlinger, M, Rowan, AJ, Horswell, S et al. (2012). Tumour heterogeneity and branched evolution revealed by multiregion sequencing. New England Journal of Medicine 366, 883–92CrossRefGoogle ScholarPubMed
Greaves, LC, Preston, SL, Tadrous, PJ et al. (2006). Mitochondrial DNA mutations are established in human colonic stem cells, and mutated clones expand by crypt fission. Proceedings of the National Academy of Sciences of the USA 103, 714–19CrossRefGoogle ScholarPubMed
Leedham, SJ, Graham, TA, Oukrif, D et al. (2009). Clonality, founder mutations, and field cancerization in human ulcerative colitis-associated neoplasia. Gastroenterology 136, 542–50; Salk, JJ, Salipante, SJ, Risques, RA et al. (2009). Clonal expansions in ulcerative colitis identify patients with neoplasia. Proceedings of the National Academy of Sciences of the USA 106, 20871–6; Hayashi, H, Miyagi, Y, Sekiyama, A et al. (2012). Colorectal small cell carcinoma in ulcerative colitis with identical rare p53 gene mutation to associated adenocarcinoma and dysplasia. Journal of Crohn’s and Colitis 6, 112–15CrossRefGoogle ScholarPubMed
Jones, S, Chen, W-D, Parmigiani, G et al. (2008). Comparative lesion sequencing provides insights into tumor evolution. Proceedings of the National Academy of Sciences of the USA 105, 4283–8CrossRefGoogle ScholarPubMed
Campbell, PJ, Pleasance, ED, Stephens, PJ et al. (2008). Subclonal phylogenetic structures in cancer revealed by ultra-deep sequencing. Proceedings of the National Academy of Sciences of the USA 105, 13081–6; Shah, SP, Morin, RD, Khattra, J et al. (2009). Mutational evolution in a lobular breast tumour, profiled at single nucleotide resolution. Nature 461, 809–13; Campbell, PJ, Yachida, S, Mudie, LJ et al. (2010). The patterns and dynamics of genetic instability in metastatic pancreatic cancer. Nature 467, 1109–13; Yachida, S, Jones, S, Bozic, I et al. (2010). Distant metastasis occurs late during the genetic evolution of pancreatic cancer. Nature 467, 1114–17; Tao, Y, Ruan, J, Yeh, A-H et al. (2011). Rapid growth of a hepatocellular carcinoma and the driving mutations revealed by cell-population genetic analysis of whole-genome data. Proceedings of the National Academy of Sciences of the USA 108, 12042–7CrossRefGoogle ScholarPubMed
Navin, N, Kendal, J, Troge, J et al. (2011). Tumour evolution inferred by single-cell sequencing. Nature 472, 90–4CrossRefGoogle ScholarPubMed
Greaves, M (2010). Cancer stem cells? Back to Darwin. Seminars in Cancer Biology 20, 65–70; Stratton, MR (2011). Exploring the genomes of cancer cells: progress and promise. Science 331, 1553–8; Greaves, M and Maley, CC (2012). Clonal evolution in cancer. Nature 481, 306–13CrossRefGoogle ScholarPubMed
Lieber, MR (2008). The mechanism of human nonhomologous end joining. Journal of Biological Chemistry 283, 1–5CrossRefGoogle Scholar
Pace, JK, Sen, SK, Batzer, MA and Feschotte, C (2009). Repair-mediated duplication by capture of proximal chromosome DNA has shaped vertebrate genome evolution. PLoS Genetics 5, e1000469 (Supplementary Information, Figure S1)CrossRefGoogle Scholar
Thomas, EE, Srebro, N, Sebat, J et al. (2004). Distribution of short paired duplications in primate genomes. Proceedings of the National Academy of Sciences of the USA 101, 10349–54CrossRefGoogle Scholar
Morrish, TA, Gilbert, N, Myers, JS et al. (2002). DNA repair mediated by endonuclease-independent LINE-1 retrotransposition. Nature Genetics 31, 159–65CrossRefGoogle ScholarPubMed
Sen, SK, Huang, CT, Han, K and Batzer, M (2007). Endonuclease-independent insertion provides an alternative pathway for L1 retrotransposition in the human genome. Nucleic Acids Research 35, 3741–51CrossRefGoogle ScholarPubMed
Srikanta, D, Sen, SK, Huang, CT et al. (2009). An alternative pathway for Alu retrotransposition suggests a role in DNA double-strand break repair. Genomics 93, 205–12CrossRefGoogle Scholar
Willett-Brozick, JE, Savul, SA, Richey, LE and Baysal, BE (2001). Germ line insertion of mtDNA at the breakpoint junction of a reciprocal constitutional translocation. Human Genetics 109, 216–23CrossRefGoogle ScholarPubMed
Millar, DS, Tysoe, C, Lazarou, LP et al. (2010). An isolated case of lissencephaly caused by the insertion of a mitochondrial genome-derived DNA sequence into the 5′ untranslated region of the PAFAH1B1 (L1S1) gene. Human Genomics 4, 384–93CrossRefGoogle Scholar
Bensasson, D, Zhang, D-X, Hartl, DL and Hewitt, GM (2001). Mitochondrial pseudogenes: evolution’s misplaced witness. Trends in Ecology and Evolution 16, 314–21; Hazkani-Covo, E, Zeller, RM and Martin, W (2010). Molecular poltergeists: mitochondrial DNA copies (numts) in sequenced nuclear genomes. PLoS Genetics 6, e1000834CrossRefGoogle Scholar
Tourmen, Y, Baris, O, Dessen, P et al. (2002). Structure and chromosomal distribution of human mitochondrial pseudogenes. Genomics 80, 71–7; Woischnik, M and Moraes, CT (2002). Pattern of organization of human mitochondrial pseudogenes in the nuclear genome. Genome Research 12, 885–93; Ricchetti, M, Tekaia, F and Dujon, B (2004). Continued colonisation of the human genome by mitochondrial DNA. PLoS Biology 2, e273; Gherman, A, Chen, PE, Teslovich, TM et al. (2007). Population bottlenecks as a potential major shaping force of human genome architecture. PLoS Genetics 3, e119; Simone, D, Calabrese, FM, Lang, M et al. (2011). The reference human nuclear mitochondrial sequences compilation validated and implemented on the UCSC genome browser. BMC Genomics 12, 517CrossRefGoogle ScholarPubMed
Zischler, H, Geisert, H and Castresana, J (1998). A hominoid-specific nuclear insertion of the mitochondrial D-loop: implications for reconstructing ancestral mitochondrial sequences. Molecular Biology and Evolution 15, 463–9; Zischler, H (2000). Nuclear integrations of mitochondrial DNA in primates; inference of associated mutational events. Electrophoresis 21, 531–6; Schmitz, J, Ohme, M and Zischler, H (2002). The complete mitochondrial sequence of Tarsius bancanus: evidence for an extensive nucleotide compositional plasticity of primate mitochondrial DNA. Molecular Biology and Evolution 19, 544–53; Ovchinnikov, IV and Kholina, OI (2010). Genome digging: insight into the mitochondrial genome of Homo. PLoS ONE 5, e14278CrossRefGoogle ScholarPubMed
Ricchetti, et al. (2004), ref. 22; Hazkani-Covo, E and Graur, D (2007). A comparative analysis of numt evolution in human and chimpanzee. Molecular Biology and Evolution 24, 13–8; Hazkani-Covo, E and Covo, S (2008). Numt-mediated double-strand break repair mitigates deletions during primate genome evolution. PLoS Genetics 4, e1000237; Lang, M, Sazzini, M, Calabrese, FM et al. (2012). Polymorphic numts trace human population relationships. Human Genetics 131, 757–71; Calabrese, FM, Simone, D and Attimonelli, M (2012). Primates and mouse numts in the UCSC genome browser. BMC Bioinformatics 13 (Suppl. 4), S15 (figures kindly supplied by Dr F Calabrese, personal communication).Google Scholar
Jensen-Seaman, MI, Wildschutte, JH, Soto-Calderon, ID and Anthony, NM (2009). A comparative approach shows differences in patterns of numt insertion during hominoid evolution. Journal of Molecular Evolution 68, 688–99; Hazkani-Covo, E (2009). Mitochondrial insertions into primate nuclear genomes suggest the use of numts as a tool for phylogeny. Molecular Biology and Evolution 26, 2175–9; Soto-Calderon, ID, Lee, EJ, Jensen-Seaman, MI and Anthony, NM (2012). Factors affecting the relative abundance of nuclear copies of mitochondrial DNA (numts) in hominoids. Journal of Molecular Evolution 75, 102–11; Tsuji, J, Frith, MC, Tomii, K and Horton, P (2012). Mammalian numt insertion is non-random. Nucleic Acids Research 40, 9073–88 (see Supplementary Material)CrossRefGoogle ScholarPubMed
Yunis, JJ and Prakash, O (1986). The origin of man: a chromosomal pictorial legacy. Science 215, 1525–30CrossRefGoogle Scholar
Kehrer-Sawatzki, H and Cooper, DN (2007). Structural differences between the human and chimpanzee genomes. Human Genetics 120, 759–78CrossRefGoogle Scholar
IJdo, JW, Baldini, A, Ward, DC et al. (1991). Origin of human chromosome 2: an ancestral telomere–telomere fusion. Proceedings of the National Academy of Sciences of the USA 88, 9051–5CrossRefGoogle Scholar
Fan, Y, Linardopoulou, E, Friedman, C et al. (2002). Genomic structure and evolution of the ancestral chromosome fusion site in 2q13–2q14.1 and paralogous regions on other human chromosomes. Genome Research 12, 1651–62; Hillier, LW, Graves, TA, Fulton, RS et al. (2005). Generation and annotation of the DNA sequences of human chromosomes 2 and 4. Nature 434, 724–31CrossRefGoogle ScholarPubMed
Azzalin, CM, Nergadze, SG and Giulotto, E (2001). Human intrachromosomal telomeric repeats: sequence organization and mechanisms of origin. Chromosoma 110, 75–82; Nergadze, SG, Rocchi, M, Azzalin, CM et al. (2004). Insertion of telomeric repeats at intrachromosomal break sites during primate evolution. Genome Research 14, 1704–10; Nergadze, SG, Santagostino, MA, Salzano, A et al. (2007). Contribution of telomerase RNA retrotranscription to DNA double-strand break repair during mammalian genome evolution. Genome Biology 8, R260CrossRefGoogle Scholar
Nagy, R, Sweet, K and Eng, C (2004). Highly penetrant cancer syndromes. Oncogene 23, 6445–70CrossRefGoogle ScholarPubMed
Harbour, JW (1998). Overview of RB gene mutations in patients with retinoblastoma: implications for clinical genetic screening. Ophthalmology 105, 1442–7; Balogh, K, Patócs, A, Majnik, J et al. (2004). Genetic screening methods for the detection of mutations responsible for multiple endocrine neoplasia type 1. Molecular Genetics and Metabolism 83, 74–81; Ferla, R, Calo, V, Cascio, S et al. (2007). Founder mutations in BRCA1 and BRCA2 genes. Annals of Oncology 18 (Suppl. 6), vi93–8; Walsh, T, Casadei, S, Coats, KH et al. (2006). Spectrum of mutations in BRCA1, BRCA2, CHEK2, and TP53 in families at high risk of breast cancer. Journal of the American Medical Association 295, 1379–88; Lindstrom, E, Shimokawa, T, Toftgard, R and Zaphiropoulos, PG (2006). PTCH mutations: distribution and analysis. Human Mutation 27, 215–9; Nordstrom-O’Brien, M, van der Luijt, RB and van Rooijen, E (2010). Genetic analysis of von Hippel–Lindau disease. Human Mutation 31, 521–37CrossRefGoogle ScholarPubMed
Garritano, S, Gemignani, F, Palmeri, EI et al. (2009). Detailed haplotype analysis at the TP53 locus in p.R337H mutation carriers in the population of Southern Brazil: evidence for a founder effect. Human Mutation 31, 143–50Google Scholar
Zheng, D and Gerstein, MB (2007). The ambiguous boundary between genes and pseudogenes: the dead rise up, or do they?Trends in Genetics 23, 219–24; Harrison, PM, Hegyi, H, Balasubramian, S et al. (2002). Molecular fossils in the human genome: identification and analysis of the pseudogenes in chromosomes 21 and 22. Genome Research 12, 272–80; Zheng, D, Frankish, A, Baertsch, R et al. (2007). Pseudogenes in the ENCODE regions: consensus annotation, analysis of transcription, and evolution. Genome Research 17, 839–51CrossRefGoogle ScholarPubMed
Baguley, BC and Finlay, GJ (1988). Derivatives of amsacrine: determinants required for high activity against the Lewis lung carcinoma. Journal of the National Cancer Institute 80, 195–9; Baguley, BC and Finlay, GJ (1988). Relationship between the structure of analogues of amsacrine and their degree of cross-resistance to adriamycin-resistant P388 leukaemia cells. European Journal of Cancer and Clinical Oncology 24, 205–10CrossRefGoogle ScholarPubMed
Neff, MW, Robertson, KR, Wong, AK et al. (2004). Breed distribution and history of canine mdr1–1Δ, a pharmacogenetic mutation that marks the emergence of breeds from the collie lineage. Proceedings of the National Academy of Sciences of the USA 101, 11725–30CrossRefGoogle ScholarPubMed
Balasubramanian, S, Habegger, L, Frankish, A et al. (2011). Gene inactivation and its implications for annotation in the era of personal genomics. Genes and Development 25, 1–10; MacArthur, DG, Balasubramanian, S, Frankish, A et al. (2012). A systematic survey of loss-of-function variants in human protein-coding genes. Science 335, 823–8CrossRefGoogle ScholarPubMed
Rompler, H, Schulz, A, Pitra, C et al. (2005). The rise and fall of the chemoattractant receptor GPR33. Journal of Biological Chemistry 280, 31068–75; Bohnekamp, J, Boselt, I, Saalbach, A et al. (2010). Involvement of the chemokine-like receptor GPR33 in innate immunity. Biochemical and Biophysical Research Communications 396, 272–7CrossRefGoogle ScholarPubMed
Galvani, AP and Novembre, J (2005). The evolutionary history of the CCR5-Δ32 HIV-resistance mutation. Microbes and Infection 7, 302–9CrossRefGoogle Scholar
Mercereau-Puijalon, O and Menard, D (2010). Plasmodium vivax and the Duffy antigen: a paradigm revisited. Transfusion clinique et biologique 17, 176–83CrossRefGoogle Scholar
Xue, Y, Daly, A, Yngvadottir, B et al. (2006). Spread of an inactive form of caspase-12 in humans is due to recent positive selection. American Journal of Human Genetics 78, 659–70CrossRefGoogle ScholarPubMed
Hermel, E and Klapstein, KD (2011). A possible mechanism for maintenance of the deleterious allele of human CASPASE-12. Medical Hypotheses 77, 803–6; Hervella, M, Plantinga, TS, Alonso, S et al. (2012). The loss of functional caspase-12 in Europe is a pre-Neolithic event. PLoS ONE 7, e37022CrossRefGoogle ScholarPubMed
Wang, et al. (2012), Chapter 2, ref. 145
Chou, H-H, Takematsu, H, Diaz, S et al. (1998). A mutation in human CMP–sialic acid hydroxylase occurred after the Homo–Pan divergence. Proceedings of the National Academy of Sciences of the USA 95, 11751–6; Hayakawa, T, Satta, Y, Gagneux, P et al. (2001). Alu-mediated inactivation of the human CMP-N-acetylneuraminic acid hydroxylase gene. Proceedings of the National Academy of Sciences of the USA 98, 11399–404CrossRefGoogle ScholarPubMed
Hedlund, M, Padler-Karavani, V, Varki, NM and Varki, A (2008). Evidence for a human-specific mechanism for diet and antibody-mediated inflammation in carcinoma progression. Proceedings of the National Academy of Sciences of the USA 105, 18936–41; Taylor, RE, Gregg, CJ, Padler-Karavani, V et al. (2010). Novel mechanism for the generation of human xeno-autoantibodies against the nonhuman sialic acid N-glycolyneuraminic acid. Journal of Experimental Medicine 207, 1637–46; Ghaderi, D, Taylor, RE, Padler-Karavani, V et al. (2010). Implications of the presence of N-glycolylneuraminic acid in recombinant therapeutic glycoproteins. Nature Biotechnology 28, 863–7; Varki, A (2010). Uniquely human evolution of sialic acid genetics and biology. Proceedings of the National Academy of Sciences of the USA 107 (Suppl 2), 8939–46CrossRefGoogle ScholarPubMed
Winter, H, Langbein, L, Krawczak, M et al. (2001). Human type I hair keratin pseudogene ψhHaA has functional orthologs in the chimpanzee and gorilla: evidence for the inactivation of the human gene after the Pan–Homo divergence. Human Genetics 108, 37–42CrossRefGoogle Scholar
Kazantseva, A, Goltsov, A, Zinchenko, R et al. (2006). Human hair growth deficiency is linked to a genetic defect in the phospholipase gene LIPH. Science 314, 982–5CrossRefGoogle ScholarPubMed
Stedman, HH, Kozyak, BW, Nelson, A et al. (2004). Myosin gene mutation correlates with anatomical changes in the human lineage. Nature 428, 415–18CrossRefGoogle ScholarPubMed
Kim, HL, Igawa, T, Kawashima, A et al. (2010). Divergence, demography and gene loss along the human lineage. Philosophical Transactions of the Royal Society Series B: Biological Sciences 365, 2541–7CrossRefGoogle ScholarPubMed
Zhu, J, Sanborn, JZ, Diekhans, M et al. (2007). Comparative genomics search for losses of long-established genes on the human lineage. PLoS Computational Biology 3, e247CrossRefGoogle ScholarPubMed
Annilo, T and Dean, M (2004). Degeneration of an ATP-binding cassette transporter gene, ABCC13, in different mammalian lineages. Genomics 84, 34–46CrossRefGoogle ScholarPubMed
Martinez-Arias, R, Calafell, F, Mateu, E et al. (2001). Sequence variability of a human pseudogene. Genome Research 11, 1071–85; Wafaei, JR and Choy, FYM(2005). Glucocerebrosidase recombinant allele: molecular evolution of the glucocerebrosidase gene and pseudogene in primates. Blood Cells, Molecules and Diseases 35, 277–85CrossRefGoogle ScholarPubMed
Apoil, P-A, Roubinet, F, Despiau, S et al. (2000). Evolution of α2-fucosyltransferase genes in primates: relation between an intronic Alu-Y element and red cell expression of ABH antigens. Molecular Biology and Evolution 17, 337–51; Saunier, K, Barreaud, J-P, Eggen, A et al. (2001). Organization of the bovine α2-fucosyltransferase gene cluster suggests that the Sec1 gene might have been shaped through a nonautonomous L1-retrotransposition event within the same locus. Molecular Biology and Evolution 18, 2083–91CrossRefGoogle Scholar
Oda, M, Satta, Y, Takenaka, O and Takahata, N (2002). Loss of urate oxidase activity in hominoids and its evolutionary implications. Molecular Biology and Evolution 19, 640–53CrossRefGoogle ScholarPubMed
Keebaugh, AC and Thomas, JW (2010). The evolutionary fate of the genes encoding the purine catabolic enzymes in hominoids, birds and reptiles. Molecular Biology and Evolution 27, 1359–69CrossRefGoogle ScholarPubMed
Ivell, R, Pusch, W, Balvers, M et al. (2000). Progressive inactivation of the haploid gene for the sperm-specific endozepine-like peptide (ELP) through primate evolution. Gene 255, 335–45CrossRefGoogle ScholarPubMed
Goldberg, A, Wildman, DE, Schmidt, TR et al. (2003). Adaptive evolution of cytochrome c oxidase subunit VIII in anthropoid primates. Proceedings of the National Academy of Sciences of the USA 100, 5873–8CrossRefGoogle ScholarPubMed
Koike, C, Fung, JJ, Geller, DA et al. (2002). Molecular basis of evolutionary loss of the α1,3-galactosyltransferase gene in higher primates. Journal of Biological Chemistry 277, 10114–20; Koike, C, Uddin, M, Wildman, DE et al. (2007). Functionally important glycosyltransferase gain and loss during catarrhine primate emergence. Proceedings of the National Academy of Sciences of the USA 104, 559–64 (Supplementary Information, Data Set 1)CrossRefGoogle ScholarPubMed
Wigglesworth, KM, Racki, WJ, Mishra, R et al. (2011). Rapid recruitment and activation of macrophages by anti-Gal/αGal liposome interaction accelerates wound healing. Journal of Immunology 186, 4422–32CrossRefGoogle Scholar
Morisset, M, Richard, C, Astier, C et al. (2012). Anaphylaxis to pork kidney is related to IgE antibodies specific for galactose-alpha-1,3-galactose. Allergy 67, 699–704; Wolver, SE, Sun, DR, Commins, SP and Schwartz, LB (2013). A peculiar case of anaphylaxis: no more steak? The journey to discovery of a newly recognized allergy to galactose-alpha-1,3-galactose found in mammalian meat. Journal of General Internal Medicine 28, 322–5CrossRefGoogle ScholarPubMed
Cai, X and Patel, S (2010). Degeneration of an intracellular ion channel in the primate lineage by relaxation of selective constraints. Molecular Biology and Evolution 27, 2352–9CrossRefGoogle ScholarPubMed
Ohta, Y and Nishikimi, M (1999). Random nucleotide substitutions in primate non-functional gene for L-gulono-γ-lactone oxidase, the missing enzyme in L-ascorbic acid biosynthesis. Biochimica et Biophysica Acta 1472, 408–11; Inai, Y, Ohta, Y and Nishikimi, M (2003). The whole structure of the human nonfunctional L-gulono-γ-lactone oxidase gene – the gene responsible for scurvy – and the evolution of repetitive sequences thereon. Journal of Nutritional Science and Vitaminology 49, 315–19CrossRefGoogle Scholar
Zhu, et al. (2007), ref. 50
Lachapelle, MY and Drouin, G (2011). Inactivation dates of the human and guinea pig vitamin C genes. Genetica 139, 199–207 (Supplementary Figure 1)CrossRefGoogle ScholarPubMed
Cui, J, Yuan, X, Wang, L et al. (2011). Recent loss of vitamin C biosynthesis ability in bats. PLoS ONE 6, e27114CrossRefGoogle ScholarPubMed
Drouin, G, Godin, JR and Page, B (2011). The genetics of vitamin C loss in vertebrates. Current Genomics 12, 371–8CrossRefGoogle ScholarPubMed
Johnson, RJ, Gaucher, EA, Sautin, YY et al. (2008). The planetary biology of ascorbate and uric acid and their relationship with the epidemic of obesity and cardiovascular disease. Medical Hypotheses 71, 22–31; Johnson, RJ, Andrews, P, Benner, SA and Oliver, W (2010). The evolution of obesity: insights from the mid-Miocene. Transactions of the American Clinical and Climatological Association 121, 295–305CrossRefGoogle ScholarPubMed
Fernandez, E, Torrents, D, Zorzano, A et al. (2005). Identification and functional characterization of a novel low-affinity aromatic-preferring amino acid transporter (arpAT). Journal of Biological Chemistry 280, 19364–72; Casals, F, Ferrer-Admetlla, A, Chillaron, J et al. (2008). Is there selection for the pace of successive inactivation of the arpAT gene in primates?Journal of Molecular Evolution 67, 23–8CrossRefGoogle Scholar
Naidu, S, Peterson, ML and Spear, BT (2010). Alpha-fetoprotein-related gene (ARG): a member of the albumin gene family that is no longer functional in primates. Gene 449, 95–102CrossRefGoogle ScholarPubMed
Zhang, Z and Gerstein, M (2003). The human genome has 49 cytochrome c pseudogenes, including a relic of a primordial gene that still functions in mouse. Gene 312, 61–72; sequences are from ; Pierron, D, Opazo, JC, Heiske, M et al. (2011). Silencing, positive selection and parallel evolution: busy history of primate cytochromes c. PLoS ONE 6, e26269 (Supporting Information, Figure S3)CrossRefGoogle Scholar
Zhang, ZD, Frankish, A, Hunt, T et al. (2010). Identification and analysis of unitary pseudogenes: historic and contemporary gene losses in humans and other primates. Genome Biology 11, R26CrossRefGoogle ScholarPubMed
Liman, ER (2006). Use it or lose it: molecular evolution of sensory signalling in primates. Pflugers Archiv: European Journal of Physiology 453, 125–31; Niimura, Y and Nei, M (2006). Evolutionary dynamics of olfactory and other chemosensory receptor genes in vertebrates. Journal of Human Genetics 51, 505–17; Rouquier, S and Giorgi, D (2007). Olfactory receptor gene repertoires in mammals. Mutation Research 616, 95–102; Nei, M, Niimura, Y and Nozawa, M (2008). The evolution of animal chemosensory receptor repertoires: roles of chance and necessity. Nature Reviews Genetics 9, 951–63CrossRefGoogle Scholar
Mundy, NI (2006). Genetic basis of olfactory communication in primates. American Journal of Primatology 68, 559–67CrossRefGoogle ScholarPubMed
Liman, ER and Innan, H (2003). Released selective pressure on an essential component of pheromone transduction in primate evolution. Proceedings of the National Academy of Sciences of the USA 100, 3328–32; Zhang, J and Webb, DM (2003). Evolutionary deterioration of the vomeronasal pheromone transduction pathway in catarrhine primates. Proceedings of the National Academy of Sciences of the USA 100, 8337–41CrossRefGoogle Scholar
Yu, L, Jin, W, Wang, J-X et al. (2010). Characterization of TRPC2, an essential genetic component of VNS chemoreception provides insights into the evolution of pheromonal olfaction in secondary-adapted marine mammals. Molecular Biology and Evolution 27, 1467–77CrossRefGoogle ScholarPubMed
Zhang, and Webb, (2003), ref. 74; Mundy, NI and Cook, S (2003). Positive selection during the diversification of class I vomeronasal receptor-like (V1RL) genes, putative pheromone receptor genes, in human and primate evolution. Molecular Biology and Evolution 20, 1805–10Google Scholar
Young, JM, Kambere, M, Trask, B and Lane, RP (2005). Divergent V1R repertoires in five species: amplification in rodents, decimation in primates, and a surprisingly small repertoire in dogs. Genome Research 15, 231–40; Young, JM, Massa, HF, Hsu, L and Trask, BJ (2010). Extreme variability among mammalian V1R receptors. Genome Biology 20, 10–18CrossRefGoogle Scholar
Young, JM and Trask B, (2007). V2R gene families degenerated in primates, dog and cow, but expanded in opossum. Trends in Genetics 23, 212–5; Suarez, R, Fernandez-Aburto, P, Manger, PR and Mpodozis, J (2011). Deterioration of the Gαo vomeronasal pathway in sexually dimorphic mammals. PLoS ONE 6, e23436CrossRefGoogle ScholarPubMed
Go, Y and Niimura, Y (2008). Similar numbers but different repertoires of olfactory receptor genes in humans and chimpanzees. Molecular Biology and Evolution 25, 1897–907; Matsui, A, Go, Y and Niimura, Y (2010). Degeneration of olfactory receptors gene repertories in primates: no direct link to full trichromatic vision. Molecular Biology and Evolution 27, 1192–200CrossRefGoogle ScholarPubMed
Gilad, Y, Man, O, Paabo, S and Lancet, D (2003). Human-specific loss of olfactory receptor genes. Proceedings of the National Academy of Sciences of the USA 100, 3324–7CrossRefGoogle ScholarPubMed
Gilad, Y, Man, O and Glusman, G (2005). A comparison of the human and chimpanzee olfactory receptor gene repertoires. Genome Research 15, 224–30CrossRefGoogle ScholarPubMed
Olender, T, Waszak, SM, Viavant, M et al. (2012). Personal receptor repertoires: olfaction as a model. BMC Genomics 13, 414; Pierron, D, Cortés, NG, Letellier, T and Grossman, LI (2013). Current relaxation of selection on the human genome: tolerance of deleterious mutations on olfactory receptors. Molecular Phylogenetics and Evolution 66, 558–64CrossRefGoogle Scholar
Freitag, J, Ludwig, G, Andreini, I et al. (1998). Olfactory receptors in aquatic and terrestrial vertebrates. Journal of Comparative Physiology A 183, 635–50; Kishida, T, Kubota, S, Shirayama, Y and Fukami, H (2007). The olfactory receptor gene repertoires in secondary-adapted marine vertebrates: evidence for reduction of the functional proportions in cetaceans. Biology Letters 3, 428–30; Niimura, Y and Nei, M (2007). Extensive gains and losses of olfactory receptor genes in mammalian evolution. PLoS ONE 2, e708; Hayden, S, Bekaert, M, Crider, TA et al. (2010). Ecological adaptation determines functional mammalian olfactory subgenomes. Genome Research 20, 1–9CrossRefGoogle ScholarPubMed
Young, JM, Waters, H, Dong, C et al. (2007). Degeneration of the olfactory guanylyl cyclase D gene during primate evolution. PLoS ONE 2 e884CrossRefGoogle ScholarPubMed
Wooding, S, Bufe, B, Gassi, C et al. (2006). Independent evolution of bitter-taste sensitivity in humans and chimpanzees. Nature 440, 930–4; Lalueza-Fox, C, Gigli, E, de la Rasilla, M et al. (2009). Bitter taste perception in Neanderthals through the analysis of the TAS2R38 gene. Biology Letters 5, 809–11CrossRefGoogle ScholarPubMed
Fischer, A, Gilad, Y, Man, O and Paabo, S (2004). Evolution of bitter taste receptors in humans and apes. Molecular Biology and Evolution 22, 432–6; Perry, CM, Erkner, A and le Coutre, J (2004). Divergence of T2R chemosensory receptor families in humans, bonobos, and chimpanzees. Proceedings of the National Academy of Sciences of the USA 101, 14830–4CrossRefGoogle ScholarPubMed
Li, X, Li, W, Wang, H et al. (2005). Pseudogenization of a sweet-receptor gene accounts for cats’ indifference towards sugar. PLoS Genetics 1, 27–35; Li, X, Li, W, Wang, H et al. (2006). Cats lack a sweet taste receptor. Journal of Nutrition 136, 1932S–4S; Jiang, P, Josue, J, Li, X et al. (2012). Major taste loss in carnivorous mammals. Proceedings of the National Academy of Sciences of the USA 109, 4956–61CrossRefGoogle Scholar
Zhao, H, Zhou, Y, Pinto, CM et al. (2010). Evolution of the sweet taste receptor gene Tas1R2 in bats. Molecular Biology and Evolution 27, 2642–50CrossRefGoogle ScholarPubMed
Zhao, H, Yang, J-R, Xu, H and Zhang, J (2010). Pseudogenization of the umami taste receptor gene Tas1r1 in the giant panda coincided with its dietary switch to bamboo. Molecular Biology and Evolution 27, 2669–73; Jiang, et al. (2012), ref. 87; Sato, JJ and Wolsan, M (2012). Loss or major reduction of umami taste sensation in pinnipeds. Naturwissenschaften 99, 655–9CrossRefGoogle ScholarPubMed
Renard, C, Chardon, P and Vaiman, N (2003). The phylogenetic history of the MHC class I gene families in pigs, including a fossil gene predating mammalian radiation. Journal of Molecular Evolution 57, 420–34CrossRefGoogle ScholarPubMed
Brawand, D, Wahli, W and Kaessmann, H (2008). Loss of egg yolk genes in mammals and the origin of lactation and placentation. PLoS Biology 6, e63CrossRefGoogle ScholarPubMed
Davit-Béal, T, Tucker, AS and Sire, J-Y (2009). Loss of teeth and enamel in tetrapods: fossil record, genetic data and morphological adaptations. Journal of Anatomy 214, 477–501CrossRefGoogle ScholarPubMed
Deméré, TA, McGowen, MR, Berta, A and Gatesy, J (2008). Morphological and molecular evidence for a stepwise evolutionary transition from teeth to baleen in mysticete whales. Systematic Biology 57, 15–37; Meredith, RW, Gatesy, J, Murphy, WJ et al. (2009). Molecular decay of the tooth gene enamelin (ENAM) mirrors the loss of enamel in the fossil record of placental mammals. PLoS Genetics 5, e1000634CrossRefGoogle ScholarPubMed
Meredith, et al. (2011), Chapter 2, ref. 102
Louchart, A and Viriot, L (2011). From snout to beak: the loss of teeth in birds. Trends in Ecology and Evolution 26, 663–73CrossRefGoogle ScholarPubMed
Harris, MP, Hasso, SM, Ferguson, MWJ and Fallon, JF (2006). The development of archosaurian first-generation teeth in a chicken mutant. Current Biology 16, 371–7CrossRefGoogle Scholar
Al-Hashimi, N, Lafont, A-G, Delgado, S et al. (2010). The enamelin genes in lizard, crocodile and frog, and the pseudogene in the chicken provide new insights on enamelin evolution in tetrapods. Molecular Biology and Evolution 27, 2078–94CrossRefGoogle ScholarPubMed
Sire, J-Y, Delgado, SC and Girondot, M (2008). Hen’s teeth with enamel cap: from dream to impossibility. BMC Evolutionary Biology 8, 246; McKnight, DA and Fisher, LW (2009). Molecular evolution of dentin phosphoprotein among toothed and toothless animals. BMC Evolutionary Biology 9, 299; Meredith RW, Gatesy J and Springer S (2013). Molecular decay of enamel matrix protein genes in turtles and other edentulous amniotes. BMC Evolutionary Biology 13, 20CrossRefGoogle ScholarPubMed
Kawasaki, K, Lafont, A-G and Sire, J-Y (2011). The evolution of milk casein genes from tooth genes before the origin of mammals. Molecular Biology and Evolution 28, 2053–61CrossRefGoogle ScholarPubMed
Esnault, C, Maestre, J and Heidmann, T (2000). Human LINE retrotransposons generate processed pseudogenes. Nature Genetics 24, 363–7CrossRefGoogle ScholarPubMed
Awano, H, Malueka, RG, Yagi, M et al. (2010). Contemporary retrotransposition of a novel non-coding gene induces exon-skipping in dystrophin mRNA. Journal of Human Genetics 55, 785–90; see also Tabata, A, Sheng, J-S, Ushikai, M et al. (2008). Identification of 13 novel mutations including a retrotransposal insertion in SLC25A13 gene and frequency of 30 mutations found in patients with citrin deficiency. Journal of Human Genetics 53, 534–45CrossRefGoogle ScholarPubMed
Zhang, Z, Carriero, N and Gerstein, M (2004). Comparative analysis of processed pseudogenes in the mouse and human genomes. Trends in Genetics 20, 62–7; Zhang, Z and Gerstein, M (2004). Large-scale analysis of pseudogenes in the human genome. Current Opinion in Genetics and Development 14, 328–35; Pavlicek, A, Gentles, AJPaces, J et al. (2006). Retrotransposition of processed pseudogenes: the impact of RNA stability and translational control. Trends in Genetics 22, 69–73CrossRefGoogle ScholarPubMed
Nachman, MW and Crowell, SL (2000). Estimate of the mutations rate per nucleotide in humans. Genetics 156, 297–304; Booth, HAF and Holland, PWH (2004). Eleven daughters of NANOG. Genomics 84, 229–38; Fairbanks, DJ and Maughan, PJ (2006). Evolution of the NANOG pseudogene family in the human and chimpanzee genomes. BMC Evolutionary Biology 6, 12Google ScholarPubMed
Fairbanks, DJ, Fairbanks, AD, Ogden, TH et al. (2012). NANOGP8: evolution of a human-specific retro-oncogene. G3: Genes, Genomes, Genetics 2, 1447–57Google Scholar
Ejima, Y and Yang, L (2003). Trans mobilization of genomic DNA as a mechanism for retrotransposon-mediated exon shuffling. Human Molecular Genetics 12, 1321–8CrossRefGoogle ScholarPubMed
Devor, EJ, Dill-Devor, RM, Magee, HJ and Waziri, R (1998). Serine hydroxymethyltransferase pseudogene SHMT-ps1: a unique genetic marker of the order Primates. Journal of Experimental Zoology 282, 150–6; Friedberg, F and Rhoads, AR (2000). Calculation and verification of the ages of retroprocessed pseudogenes. Molecular Phylogenetics and Evolution 16, 127–30; Devor, EJ (2001). Use of molecular beacons to verify that the serine hydroxymethyltransferase pseudogene SHMT-ps1 is unique to the order Primates. Genome Biology 2, research0006.1; Gotter, AL and Reppert, SM (2001). Analysis of human Per4. Molecular Brain Research 92, 19–26; Devor, EJ and Moffat-Wilson, K (2004). An ancient RNASE H1 splice junction mutant preserved in a 19-million-year-old genetic fossil in ape genomes. Journal of Heredity 95, 257–613.0.CO;2-Y>CrossRefGoogle ScholarPubMed
Devor, EJ (2006). Primate microRNAs miR-220 and miR-492 lie within processed pseudogenes. Journal of Heredity 97, 186CrossRefGoogle ScholarPubMed
Locke, et al. (2011) (Supplemental Section S10), Chapter 1, ref. 37
Schmitz, J, Churakov, G, Zischler, H and Brosius, J (2004). A novel class of mammalian-specific tailless retropseudogenes. Genome Research 14, 1911–15CrossRefGoogle ScholarPubMed
Perreault, J, Noel, J-F, Briere, F et al. (2005). Retropseudogenes derived from the human Ro/SS-A autoantigen-associated hY RNAs. Nucleic Acids Research 33, 2032–41CrossRefGoogle ScholarPubMed
Janecka, JE, Miller, W, Pringle, TH et al. (2007). Molecular and genomic data identify the closest living relatives of primates. Science 318, 792–4CrossRefGoogle ScholarPubMed
Poux, C, van Rheede, T, Madsen, O and de Jong, WW (2002). Sequence gaps join mice and men: phylogenetic evidence from deletions in two proteins. Molecular Biology and Evolution 19, 2035–7; de Jong, WW, van Dijk, MAM, Poux, C et al. (2003). Indels in protein-coding sequences of Euarchontoglires constrain the rooting of the eutherian tree. Molecular Phylogenetics and Evolution 28, 328–40; Springer, MS, Stanhope, MJ, Madsen, MJ and de Jong, WW (2004). Molecules consolidate the placental mammal tree. Trends in Ecology and Evolution 19, 430–8CrossRefGoogle ScholarPubMed
Murphy, et al. (2007), Chapter 2, ref. 84
van Rheede, T, Bastiaans, T, Boone, DN et al. (2006). The platypus is in its place: nuclear genes and indels confirm the sister group relation of Monotremes and Therians. Molecular Biology and Evolution 23, 587–97CrossRefGoogle ScholarPubMed
Wetterbom, A, Sevov, M, Cavelier, L and Bergstrom, TF (2006). Comparative genomic analysis of human and chimpanzee indicates a key role for indels in primate evolution. Journal of Molecular Evolution 63, 682–90CrossRefGoogle ScholarPubMed
Chaisson, MJ, Raphael, BJ and Pevzner, PA (2006). Microinversions in mammalian evolution. Proceedings of the National Academy of Sciences of the USA 103, 19824–9CrossRefGoogle ScholarPubMed
Perelman, P, Johnson, WE, Roos, C et al. (2011). A molecular phylogeny of living primates. PLoS Genetics 7, e1001342 (Supporting Information, Table S5)CrossRefGoogle ScholarPubMed
McLean, CY, Reno, PL, Pollen, AA et al. (2011). Human-specific loss of regulatory DNA and the evolution of human-specific traits. Nature 471, 216–19CrossRefGoogle ScholarPubMed
See ref. 32
Bekpen, C, Marques-Bonet, T, Alkan, C et al. (2009). Death and resurrection of the human IRGM gene. PLoS Genetics 5, e1000403; Bekpen, C, Xavier, RJ and Eichler, EE (2010). Human IRGM gene ‘to be or not to be’. Seminars in Immunopathology 32, 437–44CrossRefGoogle ScholarPubMed
Poliseno, L, Salmena, L, Zhang, J et al. (2010). A coding-independent function of gene and pseudogene mRNAs regulates tumour biology. Nature 465, 1033–8; Han, YJ, Ma, SF, Yourek, G et al. (2011). A transcribed pseudogene of MYLK promotes cell proliferation. FASEB Journal 25, 2305–12; Pink, RC, Wicks, K, Caley, DP et al. (2011). Pseudogenes: pseudo-functional or key regulators in health and disease?RNA 17, 792–8CrossRefGoogle ScholarPubMed
Rebollo, R, Romanish, MT and Mager, DL (2012). Transposable elements: an abundant and natural source of regulatory sequences for host cells. Annual Review of Genetics 46, 21–42CrossRefGoogle Scholar
Vogt, N, Lefevre, S-H, Apoiu, F et al. (2004). Molecular structure of double-minute chromosomes bearing amplified copies of the epidermal growth factor receptor gene in gliomas. Proceedings of the National Academy of Sciences of the USA 101, 11368–73; Storlazzi, CT, Lonoce, A, Guastadisegni, AC et al. (2010). Gene amplification as double minutes or homogeneously staining regions in solid tumors: origin and structure. Genome Research 20, 1198–206CrossRefGoogle ScholarPubMed
Isoda, T, Ford, AM, Tomizawa, D et al. (2009). Immunologically silent cancer clone transmission from mother to offspring. Proceedings of the National Academy of Sciences of the USA 106, 17882–5CrossRefGoogle ScholarPubMed
Tannock, IF and Hill, RP, The Basic Science of Oncology, 3rd edn (New York: McGraw-Hill, 1998), 156Google Scholar
Rabes, HM (2001). Gene rearrangements in radiation-induced thyroid carcinogenesis. Medical and Pediatric Oncology 36, 574–82; Zhu, Z, Ciampi, R, Nikiforova, MN et al. (2006). Prevalence of RET/PTC rearrangements in thyroid papillary carcinomas: effects of the detection methods and genetic heterogeneity. Journal of Clinical Endocrinology and Metabolism 91, 3603–10; Kaye, FJ (2009). Mutation-associated fusion cancer genes in solid tumours. Molecular Cancer Therapeutics 8, 1399–408; Clark, JP and Cooper, CS (2010). ETS gene fusions in prostate cancer. Nature Reviews Urology 6, 429–39; Palanisamy, N, Ateeq, B, Kalyana-Sundaram, S et al. (2010). Rearrangements of the RAF kinase pathway in prostate cancer, gastric cancer, and melanoma. Nature Medicine 16, 793–8; Sasaki, T, Rodig, SJ, Chirieac, LR and Janne, PA (2010). The biology and treatment of EML4-ALK non-small cell lung cancer. European Journal of Cancer 46, 1773–80; Zitzelsberger, H, Bauer, V, Thomas, G and Unger, K (2010). Molecular rearrangements in papillary thyroid carcinomas. Clinica Chimica Acta 411, 301–8CrossRefGoogle ScholarPubMed
Iafrate, AJ, Feuk, L, Rivera, MN et al. (2004). Detection of large-scale variation in the human genome. Nature Genetics 36, 949–51; Sebat, J, Lakshmi, B, Troge, J et al. (2004). Large-scale copy number polymorphism in the human genome. Science 305, 525–8CrossRefGoogle ScholarPubMed
Conrad, DF, Pinto, D, Redon, R et al. (2010). Origins and functional impact of copy number variation in the human genome. Nature 464, 704–12; Sudmant, PH, Kitzman, JO, Antonacci, F et al. (2010). Diversity of human copy number variation and multicopy genes. Science 330, 641–6CrossRefGoogle ScholarPubMed
Nozowa, M, Kawahara, Y and Nei, M (2007). Genomic drift and copy number variation of sensory receptor genes. Proceedings of the National Academy of Sciences of the USA 104, 20421–6; Iskow, RC, Gokcumen, O and Lee, C (2012). Exploring the role of copy number variants in human adaptation. Trends in Genetics 28, 245–57CrossRefGoogle Scholar
Stults, DM, Killen, MW, Pierce, HH et al. (2008). Genomic architecture and inheritance of ribosomal RNA gene clusters. Genome Research 18, 13–18; Stults, DM, Killen, MW, Williamson, EP et al. (2009). Human rRNA gene clusters are recombinational hotspots in cancer. Cancer Research 69, 9096–104CrossRefGoogle ScholarPubMed
Perry, GH, Dominy, NJ, Claw, KG et al. (2007). Diet and the evolution of human amylase gene copy number variation. Nature Genetics 39, 1256–60; Mandel, AL, Peyrot des Gachons, C, Plank, KL et al. (2010). Individual differences in AMY1 copy number, salivary α-amylase levels, and the perception of oral starch. PLoS ONE 5, e13352; Mandel, AL and Breslin, PAS (2012). High endogenous salivary amylase activity is associated with improved glycemic homeostasis following starch ingestion in adults. Journal of Nutrition 142, 853–8CrossRefGoogle ScholarPubMed
Bailey, JA and Eichler, EE (2006). Primate segmental duplications: crucibles of evolution, diversity and disease. Nature Reviews Genetics 7, 552–64; Marques-Bonet, T, Girirajan, S and Eichler, EE (2009). The origins and impact of primate segmental duplications. Trends in Genetics 25, 443–54; Carvalho, CMV, Zhang, F and Lupski, JR (2010). Genomic disorders: a window into human gene and genome evolution. Proceedings of the National Academy of Sciences of the USA 107 (Suppl. 1), 1765–71CrossRefGoogle ScholarPubMed
Wimmer, R, Kirsch, S, Rappold, GA and Schempp, W (2002). Direct evidence for the Homo–Pan clade. Chromosome Research 10, 55–61CrossRefGoogle ScholarPubMed
Keller, MP, Seifried, BA and Chance, PF (1999). Molecular evolution of the CMT1A region: a human- and chimpanzee-specific repeat. Molecular Biology and Evolution 16, 1019–26CrossRefGoogle ScholarPubMed
Delabre, C, Nakauchi, H, Bontrop, R et al. (1993). Duplication of the CD8 β-chain gene as a marker of the human–gorilla–chimpanzee clade. Proceedings of the National Academy of Sciences of the USA 90, 7049–53CrossRefGoogle Scholar
Saglio, G, Storlazzi, CT, Giugliano, E et al. (2002). A 76-kb duplication maps close to the BCR gene on chromosome 22 and the ABL gene on chromosome 9: possible involvement in the genesis of the Philadelphia chromosome translocation. Proceedings of the National Academy of Sciences of the USA 99, 9882–7; Albano, F, Anelli, L, Zagaria, A et al. (2010). Genomic segmental duplications on the basis of the t(9;22) rearrangement in chronic myeloid leukemia. Oncogene 29, 2509–16CrossRefGoogle ScholarPubMed
Munch, C, Kirsch, S, Fernandez, AMG and Schempp, W (2008). Evolutionary analysis of the highly dynamic CHEK2 duplicon in anthropoids. BMC Evolutionary Biology 8, 269CrossRefGoogle ScholarPubMed
Cheng, Z, Ventura, M, She, X et al. (2005). A genome-wide comparison of recent chimpanzee and human segmental duplications. Nature 437, 88–93; Linardopoulou, EV, Williams, EM, Fan, Y et al. (2005). Human subtelomeres are hot spots of interchromosomal recombination and segmental duplication. Nature 437, 94–100CrossRefGoogle ScholarPubMed
Perry, GH, Tchinda, J, McGrath, SD et al. (2006). Hotspots for copy number variation in chimpanzees and humans. Proceedings of the National Academy of Sciences of the USA 103, 8006–11; Sudmant, et al. (2010), ref. 6CrossRefGoogle ScholarPubMed
Johnson, ME, NISC Comparative Sequencing Program, Cheng, Z et al. (2006). Recurrent duplication-driven transposition of DNA during hominoid evolution. Proceedings of the National Academy of Sciences of the USA 103, 17626–31; She, X, Liu, G, Ventura, M et al. (2006). A preliminary comparative analysis of primate segmental duplications shows elevated substitution rates and a great-ape expansion of intrachromosomal duplications. Genome Research 16, 576–83CrossRefGoogle Scholar
Marques-Bonet, T, Kidd, JM, Ventura, M et al. (2009). A burst of segmental duplications in the genome of the African great ape ancestor. Nature 457, 877–81CrossRefGoogle ScholarPubMed
Ji, X and Zhao, S (2008). DA and Xiao – two giant and composite LTR-retrotransposon-like elements identified in the human genome. Genomics 91, 249–58; Li, X, Slife, J, Patel, N and Zhao, S (2009). Stepwise evolution of two giant composite LTR-retrotransposon-like elements DA and Xiao. BMC Evolutionary Biology 9, 128CrossRefGoogle ScholarPubMed
Itoh, T, Toyoda, A, Taylor, TD et al. (2005). Identification of large ancient duplications associated with human gene deserts. Nature Genetics 37, 1041–3CrossRefGoogle ScholarPubMed
Conrad, B and Antonarakis, SE (2007). Gene duplication: a drive for phenotypic diversity and cause of genetic disease. Annual Review of Genomics and Human Genetics 8, 17–35; Kaessmann, H (2010). Origins, evolution, and phenotypic impact of new genes. Genome Research 20, 1313–26CrossRefGoogle Scholar
Fortna, A, Kim, Y, MacLaren, E et al. (2004). Lineage-specific gene duplication and loss in human and great ape evolution. PLoS Biology 2, e207; Dumas, L, Kim, YH, Karimpour-Fard, A et al. (2007). Gene copy number variation spanning 60 million years of human and primate evolution. Genome Research 17, 1266–72; Armengol, G, Knuutila, S, Lozano, J-J et al. (2010). Identification of human-specific gene duplications relative to other primates by array CGH and quantitative PCR. Genomics 95, 203–9CrossRefGoogle ScholarPubMed
Johnson, ME, Viggiano, L, Bailey, JA et al. (2001). Positive selection of a gene family during the emergence of humans and African apes. Nature 413, 514–19CrossRefGoogle ScholarPubMed
Giannuzzi, G, Siswara, P, Malig, M et al. (2012). Evolutionary dynamism of the LRRC37 gene family. Genome Research 23, 46–59CrossRefGoogle ScholarPubMed
Vandepoele, K, van Roy, N, Staes, K et al. (2005). A novel gene family NBPF: intricate structure generated by gene duplications during primate evolution. Molecular Biology and Evolution 22, 2265–74; Popesco, MC, MacLaren, EJ, Hopkins, J et al. (2006). Human lineage-specific amplification, selection, and neuronal expression of DUF1220 domains. Science 313, 1304–7CrossRefGoogle ScholarPubMed
Hayakawa, T, Angata, T, Lewis, AL et al. (2005). A human-specific gene in microglia. Science 309, 1693; Wang, X, Chow, R, Deng, L et al. (2011). Expression of Siglec-11 by human and chimpanzee ovarian stromal cells, with uniquely human ligands: implications for human ovarian physiology and pathology. Glycobiology 21, 1038–48; Wang, X, Mitra, N, Cruz, P et al. (2012). Evolution of Siglec-11 and Siglec-16 genes in hominins. Molecular Biology and Evolution 29, 2073–86Google ScholarPubMed
Suh, A, Kriegs, JO, Brosius, J and Schmitz, J (2011). Retroposon insertions and the chronology of avian sex chromosome evolution. Molecular Biology and Evolution 28, 2993–7CrossRefGoogle ScholarPubMed
Bellott, DW, Skaletsky, H, Pyntikova, T et al. (2010). Convergent evolution of chicken Z and human X chromosomes by expansion and gene acquisition. Nature 466, 612–16CrossRefGoogle Scholar
Rozen, S, Skaletsky, H, Marszalek, JD et al. (2003). Abundant gene conversion between arms of palindromes in human and ape Y chromosomes. Nature 423, 873–6; Skaletsky, H, Kuroda-Kawaguchi, T, Minx, PJ et al. (2003). The male-specific region of the human Y chromosome is a mosaic of discrete sequence classes. Nature 423, 825–37CrossRefGoogle ScholarPubMed
Sin, H-S, Kim, D-S, Murayama, M et al. (2010). Human endogenous retrovirus K14C drove genomic diversification of the Y chromosome during primate evolution. Journal of Human Genetics 55, 717–25CrossRefGoogle Scholar
Warburton, PE, Giordano, J, Cheung, F et al. (2004). Inverted repeat structure of the human genome: the X-chromosome contain a preponderance of large, highly homologous inverted repeats that contain testes genes. Genome Research 14, 1861–9 (Supplemental Research, Data Supplement 3)CrossRefGoogle ScholarPubMed
Bhowmick, BK, Satta, Y and Takahata, N (2007). The origin and evolution of human ampliconic gene families and ampliconic structure. Genome Research 17, 441–50CrossRefGoogle ScholarPubMed
Yu, Y-H, Lin, Y-W, Yu, J-F et al. (2008). Evolution of the DAZ gene and the AZFc region on primate Y chromosomes. BMC Evolutionary Biology 8, 96; Hughes, JF, Skaletsky, H and Page, DC (2012). Sequencing of rhesus macaque Y chromosome clarifies origins and evolution of the DAZ (Deleted in AZoospermia) genes. BioEssays 34, 1035–44CrossRefGoogle ScholarPubMed
Vitek, WS, Pagidas, K, Gu, G et al. (2012). Xq;autosome translocation in POF: Xq27.2 deletion resulting in haploinsufficiency for SPANX. Journal of Assisted Reproduction and Genetics 29, 63–6CrossRefGoogle ScholarPubMed
Kouprina, N, Mullokandov, M, Rogozin, IM et al. (2004). The SPANX gene family of cancer testis-specific antigens: rapid evolution and amplification in African great apes and hominids. Proceedings of the National Academy of Sciences of the USA 101, 3077–82; Kouprina, N, Noskov, VN, Pavlicek, A et al. (2007). Evolutionary diversification of SPANX-N sperm protein gene structure and expression. PLoS ONE 4, e359CrossRefGoogle ScholarPubMed
Kouprina, N, Pavlicek, A, Noskov, VN et al. (2005). Dynamic structure of the SPANX gene cluster mapped to the prostate cancer susceptibility locus HPCX at Xq27. Genome Research 15, 1477–86CrossRefGoogle Scholar
Hansen, MA, Nielsen, JE, Retelska, D et al. (2008). A shared promoter region suggests a common ancestor for the human VCX/Y, SPANX, and CSAG gene families and the murine CYPT family. Molecular Reproduction and Development 75, 219–29CrossRefGoogle Scholar
Ruault, M, Ventura, M, Galtier, N et al. (2003). BAGE genes generated by juxtacentromeric reshuffling in the Hominidae lineage are under selective pressure. Genomics 81, 391–9CrossRefGoogle Scholar
Artamonova, and Gelfand, MS (2004). Evolution of the exon–intron structure and alternative splicing of the MAGE-A family of cancer/testis antigens. Journal of Molecular Evolution 59, 620–31; Katsura, Y and Satta, Y (2011). Evolutionary history of the cancer immunity antigen MAGE gene family. PLoS ONE 6, e20365; Zhao, Q, Caballero, OL, Simpson, AJG and Strausberg, RL (2012). Differential evolution of MAGE genes based on expression pattern and selection pressure. PLoS ONE 7, e48240CrossRefGoogle ScholarPubMed
Liu, Y, Zhu, Q and Zhu, N (2008). Recent duplication and positive selection of the GAGE gene family. Genetica 133, 31–5; Killen, MW, Taylor, TL, Stults, DM et al. (2011). Configuration and rearrangement of the human GAGE gene clusters. American Journal of Translational Research 3, 234–42CrossRefGoogle ScholarPubMed
Paulding, CA, Ruvolo, M and Haber, DA (2003). The Tre2 (USP6) oncogene is a hominoid-specific gene. Proceedings of the National Academy of Sciences of the USA 100, 2507–11; information on the junction sequence was provided by Dr Paulding, personal communicationGoogle Scholar
Kuryshev, VY, Vorobyov, E, Zink, D et al. (2006). An anthropoid-specific segmental duplication on human chromosome 1q22. Genomics 88, 143–51CrossRefGoogle ScholarPubMed
Perez-Maya, AA, Rodriguez-Sanchez, IP, de Jong, P et al. (2012). The chimpanzee GH locus: composition, organisation, and evolution. Mammalian Genome 23, 387–98CrossRefGoogle Scholar
Scally, et al. (2012), Chapter 1, ref. 36
Maston, GA and Ruvolo, M (2002). Chorionic gonadotropin has a recent origin within primates and an evolutionary history of selection. Molecular Biology and Evolution 19, 320–35; Hallast, P, Rull, K and Laan, M (2007). The evolution and genomic landscape of CGB1 and CGB2 genes. Molecular and Cellular Endocrinology 260–2, 2–11; Hallast, P, Saarela, J, Palotie, A and Laan, M (2008). High divergence in primate-specific duplicated regions: human and chimpanzee chorionic gonadotropin beta genes. BMC Evolutionary Biology 8, 195; Nagirnaja, L, Rull, K, Uuskula, L et al. (2010). Genomics and genetics of gonadotropin beta-subunit genes: unique FSHB and duplicated LHB/CGB loci. Molecular and Cellular Endocrinology 329, 4–16; Parrott, AM, Sriram, G, Liu, Y and Mathews, MB (2011). Expression of type II chorionic gonadotropin genes supports a role in the male reproductive system. Molecular and Cellular Biology 31, 287–99CrossRefGoogle Scholar
Bo, M and Boime, I (1992). Identification of the transcriptionally active genes of the chorionic gonadotropin β gene cluster in vivo. Journal of Biological Chemistry 267, 3179–84; Hallast, et al. (2007), Parrott, et al. (2011), ref. 46Google Scholar
Parrott, et al. (2011), Parrott, and Mathews, (2011), Chapter 2, ref. 111
Samuelson, LC, Phillips, RS and Swanberg, LJ (1996). Amylase gene structures in primates: retroposon insertions and promoter evolution. Molecular Biology and Evolution 13, 767–79CrossRefGoogle ScholarPubMed
Irwin, DM, Biegel, JM and Stewart, C-B (2011). Evolution of the mammalian lysozyme gene family. BMC Evolutionary Biology 11, 166CrossRefGoogle ScholarPubMed
Sorrentino, S (2010). The eight human ‘canonical’ ribonucleases: molecular diversity, catalytic properties, and special biological actions of the enzyme proteins. FEBS Letters 584, 2194–200CrossRefGoogle ScholarPubMed
Narita, Y, Oda, S, Takenaka, O and Kageyama, T (2010). Lineage-specific duplication and loss of pepsinogen genes in hominoid evolution. Journal of Molecular Evolution 70, 313–24CrossRefGoogle ScholarPubMed
Lawrence, MG, Stephens, CR, Need, EF et al. (2012). Long terminal repeats act as androgen-responsive enhancers for the PSA-kallikrein locus. Endocrinology 153, 3199–210; Marques, PI, Bernardino, R, Fernandes, T et al. (2012). Birth-and-death of KLK3 and KLK2 in primates: evolution driven by reproductive biology. Genome Biology and Evolution 4, 1331–8CrossRefGoogle ScholarPubMed
Zhang, YE, Landback, P, Vibranovsky, MD and Long, M (2011). Accelerated recruitment of new brain development genes into the human genome. PLoS Biology 9, e1001179CrossRefGoogle ScholarPubMed
Vandepoele, et al. (2005), ref. 26
Vandepoele, K and van Roy, F (2007). Insertion of a HERV(K) LTR in the intron of NBPF3 is not required for its transcriptional activity. Virology 362, 1–5; Abrarova, N, Simonova, L, Vinogradova, T and Sverdlov, E (2011). Different transcription activity of HERV-K LTR-containing and LTR-lacking genes of the KIAA1245/NBPF gene subfamily. Genetica 139, 733–41CrossRefGoogle Scholar
O’Bleness, MS, Dickens, CM, Dumas, LJ et al. (2012). Evolutionary history and domain organisation of DUF1220 protein domains. G3: Genes, Genomes, Genetics 2, 977–86CrossRefGoogle ScholarPubMed
Dumas, LJ, O’Bleness, MS, Davis, MJ et al. (2012). DUF1220-domain copy number implicated in human brain-size pathology and evolution. American Journal of Human Genetics 91, 444–54CrossRefGoogle ScholarPubMed
Dennis, MY, Nuttle, X, Sudmant, PH et al. (2012). Evolution of human-specific neural SRGAP2 genes by incomplete segmental duplication. Cell 149, 912–22CrossRefGoogle ScholarPubMed
Charrier, C, Joshi, K, Coutinho-Budd, J et al. (2012). Inhibition of SRGAP2 function by its human-specific paralogs induces neoteny during spine maturation. Cell 149, 923–35CrossRefGoogle ScholarPubMed
Deeb, SS, Jorgensen, AL, Battisti, L et al. (1994). Sequence divergence of the red and green visual pigments in great apes and humans. Proceedings of the National Academy of Sciences of the USA 91, 7262–6CrossRefGoogle ScholarPubMed
Dulai, KS, von Dornum, M, Mollon, JD and Hunt, DM (1999). The evolution of trichromatic color vision by opsin gene duplication in New World and Old World primates. Genome Research 9, 629–38; for a review, see Jacobs, GH (2008). Primate colour vision: a comparative perspective. Visual Neuroscience 25, 619–33Google ScholarPubMed
Ueyama, H, Torii, R, Tanabe, S et al. (2004). An insertion/deletion TEX28 polymorphism and its application to analysis of red/green visual pigment arrays. Journal of Human Genetics 49, 548–57CrossRefGoogle Scholar
Goodman, M (1999). The genomic record of humankind’s evolutionary roots. American Journal of Human Genetics 64, 31–9CrossRefGoogle ScholarPubMed
Fitch, DHA, Bailey, WJ, Tagle, DA et al. (1991). Duplication of the γ-globin gene mediated by L1 long interspersed repetitive elements in an early ancestor of simian primates. Proceedings of the National Academy of Sciences of the USA 88, 7396–400CrossRefGoogle Scholar
Opazo, JC, Hoffmann, FG and Storz, JF (2008). Differential loss of embryonic globin genes during the radiation of placental mammals. Proceedings of the National Academy of Sciences of the USA 105, 12950–5; Neumann, R, Lawson, VE and Jeffreys, AJ (2010). Dynamics and processes of copy number instability in human γ-globin genes. Proceedings of the National Academy of Sciences of the USA 107, 8304–9CrossRefGoogle ScholarPubMed
Wolf, R, Ruzicka, T and Yuspa, SH (2011). Novel S100A7 (psoriasin)/S100A15 (koebnerisin) subfamily: highly homologous but distinct in regulation and function. Amino Acids 41, 789–96CrossRefGoogle ScholarPubMed
Kulski, JK, Lim, CP, Dunn, DS and Bellgard, M (2003). Genomic and phylogenetic analysis of the S100A7 (psoriasin) gene duplications within the region of the S100 gene cluster on human chromosome 1q21. Journal of Molecular Evolution 56, 397–406CrossRefGoogle ScholarPubMed
Lehrer, RI and Lu, W (2012). α-Defensins in human innate immunity. Immunological Reviews 245, 84–112CrossRefGoogle ScholarPubMed
Semple, CA, Gautier, P, Taylor, K and Dorin, JR (2006). The changing of the guard: molecular diversity and rapid evolution of β-defensins. Molecular Diversity 10, 575–84; Hollox, EJ, Barber, JCK, Brookes, AJ and Armour, JAL (2008). Defensins and the dynamic genome: what we can learn from structural variation at human chromosome band 8p23.1. Genome Research 18, 1686–97; Fode, P, Jespersgaard, C, Hardwick, RJ et al. (2011). Determination of beta-defensin genomic copy number in different populations: a comparison of three methods. PLoS ONE 6, e16768CrossRefGoogle ScholarPubMed
Whittington, CM, Papenfuss, AT, Bansal, P et al. (2008). Defensins and the convergent evolution of the platypus and reptile venom genes. Genome Research 18, 986–94CrossRefGoogle ScholarPubMed
Lehrer, and Lu, (2012), ref. 69; Das, S, Nikolaidis, N, Goto, H et al. (2010). Comparative genomics and evolution of the alpha-defensin multigene family in primates. Molecular Biology and Evolution 27, 2333–43Google Scholar
Nguyen, TX, Cole, AM and Lehrer, RI (2003). Evolution of primate θ-defensins: a serpentine path to a sweet tooth. Peptides 24, 1647–54CrossRefGoogle ScholarPubMed
Penberthy, WT, Chari, S, Cole, AL and Cole, AM (2011). Retrocyclins and their activity against HIV-1. Cellular and Molecular Life Sciences 68, 2231–42; Doss, M, Ruchala, P, Tecle, T et al. (2012). Hapivirins and diprovirins: novel θ-defensin analogs with potent activity against influenza A virus. Journal of Immunology 188, 2759–68; for a review see Lehrer, RI, Cole, AM and Selsted, ME (2012). θ-Defensins: cyclic peptides with endless potential. Journal of Biological Chemistry 287, 27014–19CrossRefGoogle ScholarPubMed
Venkataraman, N, Cole, AL, Ruchala, P et al. (2009). Reawakening retrocyclins: ancestral human defensins active against HIV-1. PLoS Biology 7, e95CrossRefGoogle ScholarPubMed
Parham, P, Abi-Rached, L, Matevosyan, L et al. (2010). Primate-specific regulation of natural killer cells. Journal of Medical Primatology 39, 194–212; Han, K, Lou, DI and Sawyer, SL (2011). Identification of a genomic reservoir for new TRIM genes in primate genomes. PLoS Genetics 7, e1002388; Munk, C, Willemsen, A and Bravo, IG (2012). An ancient history of gene duplications, fusions and losses in the evolution of APOBEC3 mutators in mammals. BMC Evolutionary Biology 12, 71CrossRefGoogle ScholarPubMed
Shiina, T, Tamiya, G, Oka, A et al. (1999). Molecular dynamics of MHC genesis unraveled by sequence analysis of the 1,796,938-bp HLA class 1 region. Proceedings of the National Academy of Sciences of the USA 96, 13282–7; Anzai, T, Shiina, T, Kimura, N et al. (2003). Comparative sequencing of human and chimpanzee MHC class I regions unveils insertions/deletions as the major path to genomic divergence. Proceedings of the National Academy of Sciences of the USA 100, 7708–13CrossRefGoogle Scholar
Fukami-Kobayashi, K, Shiina, T, Anzai, T et al. (2005). Genomic evolution of MHC class I region in primates. Proceedings of the National Academy of Sciences of the USA 102, 9230–4CrossRefGoogle ScholarPubMed
Kulski, JK, Anzai, T and Inoko, H (2005). ERVK9, transposons and the evolution of MHC class I duplicons within the alpha-block of the human and chimpanzee. Cytogenetic and Genome Research 110, 181–92CrossRefGoogle ScholarPubMed
Kulski, JK, Anzai, T, Shiina, T and Inoko, H (2004). Rhesus macaque class I duplicon structures, organization, and evolution within the alpha block of the major histocompatibility complex. Molecular Biology and Evolution 21, 2079–91; Sawai, H, Kawamoto, Y, Takahata, N and Satta, Y (2004). Evolutionary relationships of major histocompatibility complex class I genes in simian primates. Genetics 166, 1897–907; Kulski, et al. (2005), ref. 79CrossRefGoogle Scholar
Doxiadis, GG, Hoof, I, de Groot, N and Bontrop, RE (2012). Evolution of HLA-DRB genes. Molecular Biology and Evolution 29, 3843–53CrossRefGoogle ScholarPubMed
Das, S, Nozawa, M, Klein, J and Nei, M (2008). Evolutionary dynamics of the immunoglobulin heavy chain variable region genes in vertebrates. Immunogenetics 60, 47–55; Pramanik, S, Cui, X, Wang, H-Y et al. (2011). Segmental duplication as one of the driving forces underlying the diversity of the human immunoglobulin heavy chain variable gene region. BMC Genomics 12, 78CrossRefGoogle ScholarPubMed
Das, S (2009). Evolutionary origin and genomic organisation of micro-RNA genes in immunoglobulin lambda variable region gene family. Molecular Biology and Evolution 26, 1179–89; Than, NG, Romero, R, Goodman, M et al. (2009). A primate subfamily of galectins expressed at the maternal–fetal interface that promote immune cell death. Proceedings of the National Academy of Sciences of the USA 106, 9731–6CrossRefGoogle ScholarPubMed
Nowick, K, Hamilton, AT, Zhang, H and Stubbs, L (2010). Rapid sequence and expression divergence suggest selection for novel function in primate-specific KRAB-ZNF genes. Molecular Biology and Evolution 27, 2606–17; Nowick, K, Fields, C, Gernat, T et al. (2011). Gain, loss and divergence in primate zinc finger genes: a rich resource for evolution of gene regulatory differences between species. PLoS ONE 6, e21553CrossRefGoogle ScholarPubMed
Shen, S, Lin, L, Cai, JJ et al. (2011). Widespread establishment and regulatory impact of Alu exons in human genes. Proceedings of the National Academy of Sciences of the USA 108, 2837–42; Thomas, JH and Schneider, SE (2011). Coevolution of retroelements and tandem zinc finger genes. Genome Research 21, 1800–12CrossRefGoogle ScholarPubMed
Bosch, N, Caceres, M, Cardone, MF et al. (2007). Characterisation and evolution of the novel gene family FAM90A in primates originated by multiple duplication and rearrangement events. Human Molecular Genetics 16, 2572–82; Bosch, N, Escaramis, G, Mercader, JM et al. (2008). Analysis of the multi-copy gene family FAM90A as a copy number variant in different ethnic backgrounds. Gene 420, 113–17CrossRefGoogle Scholar
Matsunami, M, Sumiyama, K and Saitou, N (2010). Evolution of conserved, non-coding sequences within the vertebrate Hox clusters through the two-round whole-genome duplications revealed by phylogenetic footprinting analysis. Journal of Molecular Evolution 71, 427–36CrossRefGoogle ScholarPubMed
Zhong, Y-F and Holland, PWH (2011). The dynamics of vertebrate homeobox gene evolution: gains and losses of genes in mouse and human lineages. BMC Evolutionary Biology 11, 169; Zhong, Y-F and Holland, PWH (2011). Correction: The dynamics of vertebrate homeobox gene evolution: gain and loss of genes in mouse and human lineages. BMC Evolutionary Biology 11, 204CrossRefGoogle Scholar
Niu, A-L, Wang, Y-Q, Zhang, H et al. (2011). Rapid evolution and copy number variation of primate RHOXF2, an X-linked homeobox gene, involved in male reproduction and possibly brain function. BMC Evolutionary Biology 11, 298CrossRefGoogle ScholarPubMed
Leidenroth, A, Clapp, J, Mitchell, LM et al. (2012). Evolution of DUX gene macrosatellites in placental mammals. Chromosoma 121, 489–97CrossRefGoogle ScholarPubMed
Ding, W, Lin, L, Chen, B and Dai, J (2006). L1 elements, processed pseudogenes and retrogenes in mammalian genomes. IUBMB Life 58, 677–85; Kaessmann, H, Vinckenbosch, N and Long, M (2009). RNA-based gene duplication: mechanistic and evolutionary insights. Nature Reviews Genetics 10, 19–31CrossRefGoogle ScholarPubMed
Parker, HG, VonHoldt, BM, Quignon, P et al. (2009). An expressed Fgf4 retrogene is associated with breed-defining chondrodysplasia in domestic dogs. Science 325, 995–8CrossRefGoogle ScholarPubMed
Marques, AC, Dupanloup, I, Vinckenbosch, N et al. (2005). Emergence of young human genes after a burst of retroposition in primates. PLoS Biology 3, e357; Vinckenbosch, N, Dupanloup, I and Kaessmann, H (2006). Evolutionary fate of retroposed gene copies in the human genome. Proceedings of the National Academy of Sciences of the USA 103, 3220–5; Fablet, M, Bueno, M, Potrzebowski, L and Kaessmann, H (2010). Evolutionary origin and functions of retrogene introns. Molecular Biology and Evolution 26, 2147–56CrossRefGoogle ScholarPubMed
Baertsch, R, Diekhans, M, Kent, WJ et al. (2008). Retrocopy contribution to the evolution of the human genome. BMC Genomics 9, 466CrossRefGoogle Scholar
Burki, F and Kaessmann, H (2004). Birth and adaptive evolution of a hominoid gene that supports high neurotransmitter flux. Nature Genetics 36, 1061–3CrossRefGoogle ScholarPubMed
Rosso, L, Marques, AC, Reidhert, AS and Kaessmann, H (2008). Mitochondrial targeting adaption of the hominoid-specific glutamate dehydrogenase driven by positive Darwinian selection. PLoS Genetics 4, e1000150; Kotzamani, D and Plaitakis, A (2012). Alpha helical structures in the leader sequence of human GLUD2 glutamate dehydrogenase responsible for mitochondrial import. Neurochemistry International 61, 463–9CrossRefGoogle Scholar
Spanaki, C, Zaganas, I, Kleopa, KA and Plaitakis, A (2010). Human GLUD2 glutamate dehydrogenase is expressed in neural and testicular supporting cells. Journal of Biological Chemistry 285, 16748–56; Zaganas, I, Spanaki, C and Plaitakis, A (2012). Expression of human GLUD2 glutamate dehydrogenase in human tissues: functional implications. Neurochemistry International 61, 455–62CrossRefGoogle ScholarPubMed
Kanavouras, K, Mastorodemos, V, Borompokas, M et al. (2007). Properties and molecular evolution of human GLUD2 (neural and testicular tissue-specific) glutamate dehydrogenase. Journal of Neuroscience Research 85, 3398–406; Spanaki, S, Zaganas, I, Kounoupa, Z and Plaitakis, A (2012). The complex regulation of human glud1 and glud2 glutamate dehydrogenases and its implications in nerve tissue biology. Neurochemistry International 61, 470–81CrossRefGoogle ScholarPubMed
Plaitakis, A, Latsoudis, H, Kanavouras, K et al. (2010). Gain-of-function variant in GLUD2 glutamate dehydrogenase modifies Parkinson’s disease onset. European Journal of Human Genetics 18, 336–41CrossRefGoogle ScholarPubMed
Rosso, L, Marques, AC, Weier, M et al (2008). Birth and rapid subcellular adaptation of a hominoid-specific CDC14 protein. PLoS Biology 6, e140CrossRefGoogle ScholarPubMed
Babushok, DV, Ohshima, K, Ostertag, EM et al. (2007). A novel testis ubiquitin-binding protein gene arose by exon shuffling in hominoids. Genome Research 17, 1129–38; Ohshima, K and Igarashi, K (2010). Inference for the initial stage of domain shuffling: tracing the evolutionary fate of the PIPSL retrogene in hominoids. Molecular Biology and Evolution 27, 2522–33CrossRefGoogle ScholarPubMed
Zingler, N, Willhoeft, U, Brose, HP et al. (2005). Analysis of 5′ junctions of human LINE-1 and Alu retrotransposons suggests an alternative model for 5′-end attachment requiring microhomology-mediated end joining. Genome Research 15, 780–9; Kojima, KK (2010). Different integration site structures between L1 protein-mediated retrotransposition in cis and retrotransposition in trans. Mobile DNA 1, 17CrossRefGoogle ScholarPubMed
Kojima, KK and Okada, N (2009). mRNA retrotransposition coupled with 5′ inversion as a possible source of new genes. Molecular Biology and Evolution 26, 1405–20CrossRefGoogle ScholarPubMed
Lee, Y, Ise, T, Ha, D et al. (2006). Evolution and expression of chimeric POTE-actin genes in the human genome. Proceedings of the National Academy of Sciences of the USA 103, 17885–90CrossRefGoogle ScholarPubMed
Liu, XF, Bera, TK, Liu, LJ and Pastan, I (2009). A primate-specific POTE-actin fusion protein plays a role in apoptosis. Apoptosis 14, 1237–44; Bera, TK, Walker, DA, Sherins, RJ and Pastan, I (2012). POTE protein, a cancer-testis antigen, is highly expressed in spermatids in human testis and is associated with apoptotic cells. Biochemical and Biophysical Research Communications 417, 1271–4CrossRefGoogle Scholar
Vinckenbosch, et al. (2006), ref. 93 (Supporting Data Set 1); adapted by Wood, AJ, Roberts, RG, Monk, D et al. (2007). A screen for retrotransposed imprinted genes reveals an association between X chromosome homology and maternal germ-line methylation. PLoS Genetics 3, 192; see also Svensson, O, Arvestad, L and Lagergren, J (2006). Genome-wide survey for biologically functional pseudogenes. PLoS Computational Biology 2, e46Google Scholar
Luo, C, Lu, X, Stubbs, L and Kim, J (2006). Rapid evolution of recently retroposed transcription factor YY2 in mammalian genomes. Genomics 87, 348–55; Kim, JD, Faulk, C and Kim, J (2007). Retroposition and evolution of the DNA-binding motifs of YY1, YY2 and REX1. Nucleic Acids Research 35, 3442–52CrossRefGoogle ScholarPubMed
Chen, L, Tioda, T, Coser, KR et al. (2011). Genome-wide analysis of YY2 versus YY1 target genes. Nucleic Acids Research 38, 4011–26CrossRefGoogle Scholar
Guallar, D, Perez-Palacios, R, Climent, M et al. (2012). Expression of endogenous retroviruses is negatively regulated by the pluipotency marker REX1/Zpf42. Nucleic Acids Research 40; 8993–9007CrossRefGoogle Scholar
Wood, et al. (2007), ref. 106; McCole, RB, Loughran, NB, Chahal, M et al. (2011). A case-by-case evolutionary analysis of four imprinted retrogenes. Evolution 65, 1413–27Google Scholar
Bradley, J, Baltus, A, Skaletsky, H et al. (2004). An X-to-autosome retrogene is required for spermatogenesis in mice. Nature Genetics 36, 872–6; Rohozinski, J, Lamb, DL and Bishop, CE (2006). UTP14c is a recently acquired retrogene associated with spermatogenesis and fertility in man. Biology of Reproduction 74, 644–51CrossRefGoogle ScholarPubMed
Yang, F, Skaletsky, H and Wang, PJ (2007). Ubl4b, an X-derived retrogene, is specifically expressed in post-meiotic germ cells in mammals. Gene Expression Patterns 7, 131–6CrossRefGoogle Scholar
McLysaght, A (2008). Evolutionary steps of sex chromosomes are reflected in retrogenes. Trends in Genetics 24, 478–81; Potrzebowski, L, Vinckenbosch, N, Marques, AC et al. (2008). Chromosomal gene movements reflect the recent origin and biology of therian sex chromosomes. PLoS Biology 6, e80CrossRefGoogle ScholarPubMed
Potrzebowski, L, Vinckenbosch, N and Kaessmann, H (2010). The emergence of new genes on the young therian X. Trends in Genetics 26, 1–4CrossRefGoogle Scholar
Ciomborowska, J, Rosikiewicz, W, Szklarczyk, D et al. (2013). ‘Orphan’ retrogenes in the human genome. Molecular Biology and Evolution 30, 384–96CrossRefGoogle ScholarPubMed
Weber, MJ (2006). Mammalian small nucleolar RNAs are mobile genetic elements. PLoS Genetics 2, e205; Luo, Y and Li, S (2007). Genome-wide analyses of retrogenes derived from the human box H/ACA snoRNAs. Nucleic Acids Research 35, 559–71CrossRefGoogle ScholarPubMed
Kapitonov, et al. (2004), Chapter 2, ref. 1
Collier, LS, Carlson, CM, Ravimohan, S et al. (2005). Cancer gene discovery in solid tumours using transposon-based somatic mutagenesis in the mouse. Nature 436, 272–6; Dupuy, AJ, Akagi, K, Largaespada, DA et al. (2005). Mammalian mutagenesis using a highly mobile somatic Sleeping Beauty transposon system. Nature 436, 221–6; Rad, R, Rad, L, Wang, W et al. (2010). PiggyBac transposon mutagenesis: a tool for cancer gene discovery in mice. Science 330, 1104–7; Starr, TK, Scott, PM, Marsh, BM et al. (2011). A Sleeping Beauty transposon-mediated screen identifies murine susceptibility genes for adenomatous polyposis coli (Apc)-dependent intestinal tumorigenesis. Proceedings of the National Academy of Sciences of the USA 108, 5765–70CrossRefGoogle ScholarPubMed
Pace, JK, II and Feschotte, C (2007). The evolutionary history of human DNA transposons: evidence for intense activity in the primate lineage. Genome Research 17, 422–32CrossRefGoogle ScholarPubMed
Pritham, EJ and Feschotte, C (2007). Massive amplification of rolling-circle transposons in the lineage of the bat Myotis lucifugus. Proceedings of the National Academy of Sciences of the USA 104, 1895–900; Ray, DA, Pagan, HJT, Thompson, ML and Stevens, RD (2007). Bats with hATs: evidence for DNA transposon activity in genus Myotis. Molecular Biology and Evolution 24, 632–9CrossRefGoogle ScholarPubMed
Volff, J-N (2006). Turning junk into gold: domestication of transposable elements and the creation of new genes in eukaryotes. BioEssays 28, 913–22CrossRefGoogle ScholarPubMed
Piriyapongsa, J and Jordan, IK (2007). A family of microRNA genes from miniature inverted-repeat transposable elements. PLoS ONE 2, e203; Yuan, Z, Sun, X, Jiang, D et al. (2010). Origin and evolution of a placental-specific microRNA family in the human genome. BMC Evolutionary Biology 10, 346CrossRefGoogle ScholarPubMed
Piriyapongsa, J, Marino-Ramirez, L and Jordan, IK (2007). Origin and evolution of human microRNAs from transposable elements. Genetics 176, 1323–37; Borchert, GM, Holton, NW, Williams, JD et al. (2011). Comprehensive analysis of microRNA genomic loci identifies pervasive repetitive-element origins. Mobile Genetic Elements 1, 8–17CrossRefGoogle ScholarPubMed
Cordaux, R, Udit, S, Batzer, MA and Feschotte, C (2006). Birth of a chimeric gene by capture of the transposase gene from a mobile element. Proceedings of the National Academy of Sciences of the USA 103, 8101–6CrossRefGoogle ScholarPubMed
Shaleen, M, Williamson, E, Nickoloff, J et al. (2010). Metnase/SETMAR: a domesticated primate transposase that enhances DNA repair, replication and decatenation. Genetica 138, 559–66; Wray, J, Williamson, EA, Chester, S et al. (2010). The transposase domain protein Metnase/SETMAR suppresses chromosomal translocations. Cancer Genetics and Cytogenetics 200, 184–90Google Scholar
Hromas, R, Williamson, EA, Fnu, S et al. (2012). Chk1 phosphorylation of metnase enhances DNA repair but inhibits replication fork restart. Oncogene 31, 4245–54CrossRefGoogle ScholarPubMed
Almeida, LM, Silva, IT, Silva WA, Jr et al. (2007). The contribution of transposable elements to Bos taurus gene structure. Gene 390, 180–9CrossRefGoogle Scholar
Markljung, E, Jiang, L, Jaffe, JD et al. (2009). ZBED6, a novel transcription factor derived from a domesticated DNA transposon regulates IGF2 expression and muscle growth. PLoS Biology 7, e1000256; Andersson, L, Andersson, G, Hjalm, G et al. (2010). ZBED6: the birth of a new transcription factor in the common ancestor of placental mammals. Transcription 1, 144–8CrossRefGoogle ScholarPubMed
Casola, C, Hucks, D and Feschotte, F (2008). Convergent domestication of pogo-like transposases into centromere-binding proteins in fission yeast and mammals. Molecular Biology and Evolution 25, 29–41CrossRefGoogle ScholarPubMed
Kojima, KK and Jurka, J (2011). Crypton transposons: identification of new diverse families and ancient domestication events. Mobile DNA 2, 12CrossRefGoogle ScholarPubMed
Sinzelle, L, Kapitonov, VV, Grzela, DP et al. (2008). Transposition of a reconstructed Harbinger element in human cells and functional homology with two transposon-derived cellular genes. Proceedings of the National Academy of Sciences of the USA 105, 4715–20; Smith, JJ, Sumiyama, J and Amemiya, CT (2012). A living fossil in the genome of a living fossil: Harbinger transposons in the coelacanth genome. Molecular Biology and Evolution 29, 985–93CrossRefGoogle ScholarPubMed
Fugmann, SD (2010). The origins of the Rag genes: from transposition to V(D)J recombination. Seminars in Immunology 22, 10–16CrossRefGoogle ScholarPubMed
Tautz, D and Domazet-Loso, T (2011). The evolutionary origin of orphan genes. Nature Reviews Genetics 12, 692–702CrossRefGoogle ScholarPubMed
Li, Y, Qian, Y-P, Yu, X-J et al. (2004). Recent origin of a hominid-specific splice form of neuropsin, a gene involved in learning and memory. Molecular Biology and Evolution 21, 2111–15CrossRefGoogle Scholar
Lu, Z-X, Peng, J and Su, B (2007). A human-specific mutation leads to the origin of a novel splice form of neuropsin (KLK8), a gene involved in learning and memory. Human Mutation 28, 978–84CrossRefGoogle Scholar
Gianfrancesco, F, Esposito, T, Casu, G et al. (2004). Emergence of talanin protein associated with human uric acid nephrolithiasis in the Hominidae lineage. Gene 339, 131–8CrossRefGoogle ScholarPubMed
Knowles, DG and McLysaght, A (2009). Recent de novo origin of human protein-coding genes. Genome Research 19, 1752–9CrossRefGoogle ScholarPubMed
Wu, DD, Irwin, DM and Zhang, Y-P (2011). De novo origin of human protein-coding genes. PLoS Genetics 7, e1002379CrossRefGoogle ScholarPubMed
Tay, S-K, Blythe, J and Lipovich, L (2009). Global discovery of primate-specific genes in the human genome. Proceedings of the National Academy of Sciences of the USA 106, 12019–24; Toll-Riera, M, Bosch, N, Bellora, N et al. (2009). Origin of primate orphan genes: a comparative genomics approach. Molecular Biology and Evolution 26, 603–12 (Supplementary Information)CrossRefGoogle ScholarPubMed
Hu, HY, He, L, Forminykh, K et al. (2012). Evolution of the human-specific microRNA miR-941. Nature Communications 3, 114; Iwama, H, Kato, K, Imachi, A et al. (2012). Human microRNAs originated from two periods at accelerated rates in mammalian evolution. Molecular Biology and Evolution 30, 613–26CrossRefGoogle ScholarPubMed
Duret, L, Chureau, C, Samain, S et al. (2006). The Xist gene evolved in eutherians by pseudogenization of a protein-coding gene. Science 312, 1653–5; Yen, ZC, Meyer, IM, Karalic, S and Brown, CJ (2007). A cross-species comparison of X-chromosome inactivation in Eutheria. Genomics 90, 453–63 (Supplementary Data, Figure 5)CrossRefGoogle Scholar
Elisaphenko, EA, Kolesnikov, NN, Shevchenko, AI et al. (2008). A dual origin of the Xist gene from a protein-coding gene and a set of transposable elements. PLoS ONE 3, e2521CrossRefGoogle Scholar
Romito, A and Rougeulle, C (2011). Origin and evolution of the long non-coding genes in the X-inactivation centre. Biochimie 93, 1935–42CrossRefGoogle Scholar
Gontan, C, Achame, EM, Demmers, J et al. (2012). RNF-12 initiates X-chromosome inactivation by targeting REX1 for degradation. Nature 485, 386–90CrossRefGoogle ScholarPubMed
He, S, Liu, S and Zhu, H (2011). The sequence, structure and evolutionary features of HOTAIR in mammals. BMC Evolutionary Biology 11, 102; see also Cartault, F, Munier, P, Benko, E et al. (2012). Mutation in a primate-conserved retrotransposon reveals a non-coding RNA as a mediator of infantile encephalopathy. Proceedings of the National Academy of Sciences of the USA 109, 4980–5; Novikova, IV, Hennelly, SP and Sanbonmatsu, KY (2012). Structural architecture of the long non-coding RNA, steroid receptor RNA activator. Nucleic Acids Research 40, 5034–51CrossRefGoogle ScholarPubMed
Kelley, DR and Rinn, JL (2012). Transposable elements reveal a stem cell-specific class of long noncoding RNAs. Genome Biology 13, R107CrossRefGoogle ScholarPubMed
Xie, C, Zhang, YE, Chen, J-Y et al. (2012). Hominoid-specific de novo protein-coding genes originating from long non-coding RNAs. PLoS Genetics 8, e1002942Google Scholar
Matsunami, et al. (2010), ref. 87
Lek, M, Quinlan, KGR and North, KN (2009). The evolution of skeletal muscle performance: gene duplication and divergence of human sarcomeric α-actinins. BioEssays 32, 17–25CrossRefGoogle Scholar
Huminiecki, L and Heldin, C-H (2010). 2R and the modeling of vertebrate signal transduction engine. BMC Biology 8, 146; Manning, G and Scheeff, E (2010). How the vertebrates were made: selective pruning of a double-duplicated genome. BMC Biology 8, 144CrossRefGoogle Scholar
Hoffmann, FG and Opazo, JC (2011). Evolution of the relaxin/insulin-like gene family in placental mammals: implications for its early evolution. Journal of Molecular Evolution 72, 72–9; Arroyo, JI, Hoffmann, FG and Opazo, JC (2012). Gene turnover and differential retention in the relaxin/insulin-like gene family in primates. Molecular Phylogenetics and Evolution 63, 768–76CrossRefGoogle ScholarPubMed
Hellstein, U, Aspden, JL, Rio, DC and Rokhsar, DS (2011). A segmental duplication generates a functional intron. Nature Communications 2, 454CrossRefGoogle Scholar
Kordis, D (2011). Extensive intron gain in the ancestor of placental mammals. Biology Direct 6, 59; Kordis, D and Kokosar, J (2012). What can domesticated genes tell us about the intron gain in mammals?International Journal of Evolutionary Biology 2012, Article ID 278981CrossRefGoogle ScholarPubMed
Wu, DD, Irwin, DM and Zhang, YP (2008). Molecular evolution of the keratin- associated protein gene family in mammals, role in the evolution of human hair. BMC Evolutionary Biology 8, 241; similarly, Vandeburgh, W and Bossuyt, F (2012). Radiation and functional diversification of alpha keratins during early vertebrate evolution. Molecular Biology and Evolution 29, 995–1004CrossRefGoogle Scholar
Check Hayden, E (2009). Darwin 200: The other strand. Nature 457, 776–9; Varki, A, Geschwind, DH and Eichler, EE (2010). Explaining human uniqueness: genome interactions with environment, behaviour, and culture. Nature Reviews Genetics 9, 749–63CrossRefGoogle ScholarPubMed
Lai, Y, Di Nardo, A, Nakatsuji, T et al. (2009). Commensal bacteria regulate toll-like receptor 3-dependent inflammation after skin injury. Nature Medicine 15, 1377–82CrossRefGoogle ScholarPubMed
Fukata, M, Chen, A, Klepper, A et al. (2006). Cox-2 is regulated by Toll-like receptor-4 (TLR4) signaling: role in proliferation and apoptosis in the intestine. Gastroenterology 131, 862–77CrossRefGoogle ScholarPubMed
Abreu, MT (2010). Toll-like receptor signalling in the intestinal epithelium: how bacterial recognition shapes intestinal function. Nature Reviews Immunology 10, 131–44; van Baarlen, P, Troost, F, van der Meer, C et al. (2011). Human mucosal in vivo transcriptome responses to three lactobacilli indicate how probiotics may modulate cellular pathways. Proceedings of the National Academy of Sciences of the USA 108 (Suppl. 1), 4562–9; Wells, JM, Rossi, O, Meijerink, M and van Baarlen, P (2011). Epithelial crosstalk at the microbiota–mucosal interface. Proceedings of the National Academy of Sciences of the USA 108 (Suppl. 1), 4607–14CrossRefGoogle ScholarPubMed
Bouskra, D, Brezillon, C, Berard, M et al. (2008). Lymphoid tissue genesis by commensals through NOD1 regulates intestinal homeostasis. Nature 456, 507–10CrossRefGoogle ScholarPubMed
Guarner, F, Bourdet-Sicard, R, Brandtzaeg, P et al. (2006). Mechanisms of disease: the hygiene hypothesis revisited. Nature Clinical Practice Gastroenterology and Hepatology 3, 275–84; Elliott, DE, Summers, RW and Weinstock, JV (2007). Helminths as governors of immune-mediated inflammation. International Journal for Parasitology 37, 457–64; Liu, Z, Liu, Q, Bleich, D et al. (2010). Regulation of type 1 diabetes, tuberculosis, and asthma by parasites. Journal of Molecular Medicine 88, 27–38; Scher, JU and Abramson, SB (2011). The microbiome and rheumatoid arthritis. Nature Reviews Rheumatology 7, 569–78CrossRefGoogle ScholarPubMed
Maslowski, KM and Mackay, CR (2011). Diet, gut microbiota and immune responses. Nature Immunology 12, 5–9; Pennisi, E (2011). Newsfocus: girth and the gut. Science 332, 32–3; Rooks, MG and Garrett, WS (2011). Sharing the bounty. Scientist 25(8), 38CrossRefGoogle ScholarPubMed
Mazmanian, SK, Round, JL and Kasper, DL (2008). A microbial symbiosis factor prevents intestinal inflammatory disease. Nature 453, 620–5; Mazmanian, S (2009). The microbial health factor. Scientist 23(8), 34; Round, JL and Mazmanian, SK (2009). The gut microbiota shapes intestinal immune responses during health and disease. Nature Reviews Immunology 9, 313–23CrossRefGoogle ScholarPubMed
Atarashi, K, Tanoue, T, Shima, T et al. (2011). Induction of colonic regulatory T cells by indigenous Clostridium species. Science 331, 337–41CrossRefGoogle Scholar
Wen, L, Ley, RE, Volchkov, PY et al. (2008). Innate immunity and intestinal microbiota in the development of type 1 diabetes. Nature 455, 1109–13; Cook, A (2009). Review series on helminths, immune modulation and the hygiene hypothesis: how might infection modulate the onset of type 1 diabetes?Immunology 126, 12–17; Hubner, MP, Stocker, JT and Mitre, E (2009). Inhibition of type 1 diabetes in filaria-infected non-obese diabetic mice is associated with a T helper type 2 shift and induction of FoxP3+ regulatory T cells. Immunology 127, 512–22; Zaccone, P, Burton, OT, Gibbs, S et al. (2010). Immune modulation by Schistosoma mansoni antigens in NOD mice: effects on both innate and adaptive immune systems. Journal of Biomedicine and Biotechnology 2010, 795210CrossRefGoogle ScholarPubMed
Cho, I, Yamanishi, S, Cox, L et al. (2012). Antibiotics in early life alter the murine colonic microbiome and adiposity. Nature 488, 621–6CrossRefGoogle ScholarPubMed
Reddy, A and Fried, B (2009). An update on the use of helminths to treat Crohn’s and other autoimmune diseases. Parasitology Research 104, 217–21; Jouvin, M-H and Kinet, J-P (2012). Trichuris suis ova: testing a helminth-based therapy as an extension of the hygiene hypothesis. Journal of Allergy and Clinical Immunology 130, 3–10CrossRefGoogle Scholar
Palmer, R (2011). Fecal matters. Nature Medicine 17, 150–2; Borody, TJ and Khoruts, A (2012). Fecal microbiota transplantation and emerging applications. Nature Reviews Gastroenterology and Hepatology 9, 88–96; Damman, CJ, Miller, SI, Surawicz, CM and Zisman, TL (2012). The microbiome and inflammatory bowel disease: is there a therapeutic role for fecal microbiota transplantation?American Journal of Gastroenterology 107, 1452–9CrossRefGoogle ScholarPubMed
Correale, J and Farez, M (2007). Association between parasite infection and immune responses in multiple sclerosis. Annals of Neurology 61, 97–108; Correale, J and Farez, MF (2012). Does helminth activation of toll-like receptors modulate immune responses in multiple sclerosis patients?Frontiers in Cellular and Infection Microbiology 2, 112CrossRefGoogle ScholarPubMed
von Mutius, E and Vercelli, D (2010). Farm living: effects on childhood asthma and allergy. Nature Reviews Immunology 10, 861–8CrossRefGoogle ScholarPubMed
Wickens, K, Black, PN, Stanley, TV et al. (2008). A differential effect of 2 probiotics in the prevention of eczema and atopy: a double-blind, randomized, placebo-controlled trial. Journal of Allergy and Clinical Immunology 122, 788–94; Pelucchi, C, Chatenoud, L, Turati, F et al. (2012). Probiotics supplementation during pregnancy or infancy for the prevention of atopic dermatitis: a meta-analysis. Epidemiology 23, 402–14CrossRefGoogle ScholarPubMed
Heijtz, RD, Wang, S, Anuar, F et al. (2011). Normal gut microbiota modulates brain development and behaviour. Proceedings of the National Academy of Sciences of the USA 108, 3047–52; Collins, SM, Surette, M and Bercik, B (2012). The interplay between the intestinal microbiota and the brain. Nature Reviews Microbiology 10, 735–42CrossRefGoogle Scholar
Rook, GAW and Lowry, CA (2008). The hygiene hypothesis and psychiatric disorders. Trends in Immunology 29, 150–8; Blaser, MJ and Falkow, S (2009). What are the consequences of the disappearing human microbiota?Nature Reviews Microbiology 7, 887–94; Rook, GAW (2009). Review series on helminths, immune modulation and the hygiene hypothesis: the broader implications of the hygiene hypothesis. Immunology 126, 3–11CrossRefGoogle ScholarPubMed
Claesson, MJ, Jeffery, IB, Conde, S et al. (2012). Gut microbiota composition correlates with diet and health in the elderly. Nature 488, 178–84; Hughes, V (2012). Microbiome: cultural differences. Nature 492, S14–15CrossRefGoogle ScholarPubMed
Khaitovich, P, Enard, W, Lachmann, M and Paabo, S (2006). Evolution of primate gene expression. Nature Reviews Genetics 7, 693–702CrossRefGoogle ScholarPubMed
Somel, M, Franz, H, Yan, Z et al. (2009). Transcriptional neoteny in the human brain. Proceedings of the National Academy of Sciences of the USA 106, 5743–8CrossRefGoogle ScholarPubMed
Maguire, EA, Gadian, DG, Johnsrude, IS et al. (2000). Navigation-related structural change in the hippocampi of taxi drivers. Proceedings of the National Academy of Sciences of the USA 97, 4398–403CrossRefGoogle ScholarPubMed
Munte, TF, Altenmuller, E and Jancke, L (2002). The musician’s brain as a model of neuroplasticity. Nature Reviews Neuroscience 3, 473–8; Draganski, B, Gaser, C, Busch, V et al. (2004). Changes in grey matter induced by training. Nature 427, 311–12; Draganski, B, Gaser, C, Kempermann, G et al. (2006). Temporal and spatial dynamics of brain structure changes during extensive learning. Journal of Neuroscience 26, 6314–17; Jancke, L, Koeneke, S, Hoppe, A et al. (2009). The architecture of the golfer’s brain. PLoS ONE 4, e4785; Zatorre, RJ, Fields, RD and Johansen-Berg, H (2011). Plasticity in grey and white: neuroimaging in brain structure during learning. Nature Neuroscience 15, 528–36CrossRefGoogle ScholarPubMed
Sameroff, A (2010). A unified theory of development: a dialectic integration of nature and nurture. Child Development 81, 6–22CrossRefGoogle ScholarPubMed
Stiles, J (2011). Brain development and the nature versus nurture debate. Progress in Brain Research 189, 3–22CrossRefGoogle ScholarPubMed
Morishita, H and Hensch, TK (2008). Critical period revisited: impact on vision. Current Opinion in Neurobiology 18, 101–7CrossRefGoogle ScholarPubMed
Bedny, M, Konkle, T, Pelphrey, K et al. (2010). Sensitive period for a multimodal response in human visual motion area MT/MST. Current Biology 20, 1900–6CrossRefGoogle ScholarPubMed
Leppanen, JM and Nelson, CA (2009). Tuning the developing brain to social signals of emotions. Nature Reviews Neuroscience 10, 37–47CrossRefGoogle ScholarPubMed
de Villiers-Sidani, E, Simpson, KL, Lu, Y-F et al. (2008). Manipulating critical period closure across different sectors of the primary auditory cortex. Nature Neuroscience 11, 957–65CrossRefGoogle Scholar
Kuhl, PL (2010). Brain mechanisms in early language acquisition. Neuron 67, 713–27; Perani, D, Saccuman, MC, Scifo, P et al. (2012). Neural networks at birth. Proceedings of the National Academy of Sciences of the USA 108, 16056–61CrossRefGoogle ScholarPubMed
Baumeister, RF and Masicampo, EJ (2011). Conscious thought is for facilitating social and cultural interactions: how mental simulations serve the animal–culture interface. Psychological Review 117, 945–71CrossRefGoogle Scholar
Dehaene, S, Pegado, F, Braga, F et al. (2010). How learning to read changes the cortical networks for vision and language. Science 330, 1359–64CrossRefGoogle Scholar
Maggi, S, Irwin, LJ, Siddiqi, A and Hertzman, C (2010). The social determinants of early child development: an overview. Journal of Paediatrics and Child Health 46, 627–35CrossRefGoogle ScholarPubMed
Bradshaw, GA, Schore, AN, Brown, JL et al. (2005). Elephant breakdown. Nature 433, 807CrossRefGoogle ScholarPubMed
Bock, J and Braun, K (2011). The impact of perinatal stress on the functional maturation of prefronto-cortical synaptic circuits: implications for the pathophysiology of ADHD?Progress in Brain Research 189, 155–69; Mogi, K, Nagasawa, M and Kikusui, T (2011). Developmental consequences and biological significance of mother–infant bonding. Progress in Neuro-Psychopharmacology and Biological Psychiatry 35, 1232–41; Kolb, B, Mychasiuk, R, Muhammad, A et al. (2012). Experience and the developing prefrontal cortex. Proceedings of the National Academy of Sciences of the USA 109 (Suppl 2), 17186–93CrossRefGoogle ScholarPubMed
Makinodan, M, Rosen, KM, Ito, S and Corfas, G (2012). A critical period for social experience-dependent oligodendrocyte maturation and myelination. Science 337, 1357–60CrossRefGoogle ScholarPubMed
Perry, BD (2002). Childhood experience and the expression of genetic potential: what childhood neglect tells us about nature and nurture. Brain and Mind 3, 79–100; McGowan, PO, Sasaki, A, D’Alessio, AC et al. (2009). Epigenetic regulation of the glucocorticoid receptor in human brain associates with childhood abuse. Nature Neuroscience 12, 342–8; Neigh, GN, Gillespie, CF and Nemeroff, CB (2009). The neurobiological toll of child abuse and neglect. Trauma, Violence and Neglect 10, 389–410; Heim, C, Shugart, M, Craighead, WE and Nemeroff, CB (2010). Neurobiological and psychiatric consequences of child abuse and neglect. Developmental Psychobiology 52, 671–90; Teicher, MH, Anderson, CM and Polcari, A (2012). Childhood maltreatment is associated with reduced volume in the hippocampal subfields CA3, dentate gyrus, and subiculum. Proceedings of the National Academy of Sciences of the USA 109, E563–72CrossRefGoogle Scholar
Pollak, SD, Nelson, CA, Schlaak, MF et al. (2010). Neurodevelopmental effects of early deprivation in postinstitutionalized children. Child Development 81, 224–36; see also Majer, M, Nater, UM, Lin, J-MS et al. (2010). Association of childhood trauma with cognitive function in healthy adults: a pilot study. BMC Neurology 10, 61CrossRefGoogle ScholarPubMed
Chugani, HT, Behen, ME, Muzik, O et al. (2001). Local brain functional activity following early deprivation: a study of postinstitutional Romanian orphans. NeuroImage 14, 1290–301; see also Maheu, FS, Dozier, M, Guyer, AE et al. (2010). A preliminary study of medial temporal lobe function in youths with a history of caregiver deprivation and emotional neglect. Cognitive, Affective and Behavioral Neuroscience 10, 34–49CrossRefGoogle Scholar
Eluvathingal, TJ, Chugani, HT, Behen, ME et al. (2006). Abnormal brain connectivity in children after early severe socioemotional deprivation: a diffusion tensor imaging study. Pediatrics 117, 2093–100; Sheridan, MA, Fox, NA, Zeanah, CH et al. (2012). Variation in neural development as a result of exposure to institutionalization early in childhood. Proceedings of the National Academy of Sciences of the USA 109, 12927–32CrossRefGoogle ScholarPubMed
Teicher, MH, Dumont, NL, Ito, Y et al. (2004). Childhood neglect is associated with reduced corpus callosum area. Biological Psychiatry 56, 80–5CrossRefGoogle ScholarPubMed
Benetti, S, McCrory, E, Arulanantham, S et al. (2010). Attachment style, affective loss and grey matter volume: a voxel-based morphometry study. Human Brain Mapping 31, 1482–9CrossRefGoogle ScholarPubMed
Frodl, T, Reinhold, E, Koutsouleris, N et al. (2010). Interaction of childhood stress with hippocampus and prefrontal cortex volume reduction in major depression. Journal of Psychiatric Research 44, 799–807CrossRefGoogle ScholarPubMed
Spinelli, S, Chefer, S, Suomi, SJ et al. (2009). Early-life stress induces long-term morphologic changes in primate brain. Archives of General Psychiatry 66, 658–65; Arabadzisz, D, Diaz-Heijtz, R, Knuesel, I et al. (2010). Primate early-life stress leads to long-term mild hippocampal decreases in glucocorticoid receptor expression. Biological Psychiatry 11, 1106–9; Feng, X, Wang, L, Yang, S et al. (2011). Maternal separation produces lasting changes in cortisol and behaviour in rhesus monkeys. Proceedings of the National Academy of Sciences of the USA 108, 14312–17; Cole, SW, Conti, G, Arevalo, JMG et al. (2012). Transcriptional modulation of the developing immune system by early-life social adversity. Proceedings of the National Academy of Sciences of the USA 109, 20578–83CrossRefGoogle ScholarPubMed
Sallet, J, Mars, RB, Noonan, MP et al. (2011). Social network size affects neural circuits in macaques. Science 334, 697–700CrossRefGoogle ScholarPubMed
Hackman, DA, Farah, MJ and Meaney, MJ (2010). Socioeconomic status and the brain: mechanistic insights from human and animal research. Nature Reviews Neuroscience 11, 651–9; Maggi, et al. (2010), ref. 33CrossRefGoogle ScholarPubMed
Perry, (2002), ref. 37; Venderwert, RE, Marshall, PJ, Nelson, CE, III et al. (2010). Timing of intervention affects brain electrical activity in children exposed to severe psychosocial neglect. PLoS ONE 5, e11415Google Scholar
Messer, N, Selfish Genes and Christian Ethics (London: SCM Press, 2007), 113, 128Google Scholar
Fonagy, P and Luyten, P (2009). A developmental, mentalization-based approach to the understanding and treatment of borderline personality disorder. Development and Psychopathology 21, 1355–81CrossRefGoogle ScholarPubMed
Colvert, E, Rutter, M, Kreppner, JH et al. (2008). Do theory of mind and executive function deficits underlie the adverse outcomes associated with profound early deprivation? Findings from the English and Romanian adoptees study. Journal of Abnormal Child Psychology 36, 1057–68; Lewis-Morrarty, E, Dozier, M, Bernard, K et al. (2012). Cognitive flexibility and theory of mind outcomes among foster children: preschool follow-up results of a randomized clinical trial. Journal of Adolescent Health 51 (Suppl. 2), S17–22CrossRefGoogle ScholarPubMed
Pyers, JE and Senghas, A (2009). Language promotes false-belief understanding. Psychological Science 20, 805–12CrossRefGoogle ScholarPubMed
Kobayashi, C, Glover, GH and Temple, E (2007). Cultural and linguistic effects on neural bases of ‘theory of mind’ in American and Japanese children. Brain Research 1164, 95–107; Kobayashi, C, Glover, GH and Temple, E (2008). Switching language switches mind: linguistic effects on developmental neural bases of ‘theory of mind’. Social Cognitive and Affective Neuroscience 3, 62–70CrossRefGoogle ScholarPubMed
Kapogiannis, D, Barbey, AK, Su, M et al. (2009). Cognitive and neural foundations of religious belief. Proceedings of the National Academy of Sciences of the USA 106, 4876–81CrossRefGoogle ScholarPubMed
Jeeves, M, in Jeeves, (ed.), Rethinking Human Nature (2011), see Prologue, ref. 36
Rochat, P (2007). Intentional action arises from early reciprocal exchanges. Acta Psychologica 124, 8–25CrossRefGoogle ScholarPubMed
Steeves, HP (2003). Humans and animals at the divide: the case of feral children. Between the Species III (an online journal for the study of philosophy and animals, )Google Scholar
Villa-Vicencio, C and De Gruchy, J, Doing Ethics in Context: South African Perspectives (Claremont: David Philip and Maryknoll, NY: Orbis Books, 1994), 196; De Gruchy, J, Christianity and Democracy (Cambridge: Cambridge University Press, 1995), 188–92Google Scholar
Boyd, R, Richerson, PJ and Henrich, J (2011). The cultural niche: why social learning is essential for human adaptation. Proceedings of the National Academy of Sciences of the USA 108 (Suppl. 2), 10918–25CrossRefGoogle ScholarPubMed
Perry, JC, Sigal, JJ, Boucher, S and Pare, N (2006). Seven institutionalized children and their adaptation in late adulthood: the children of Duplessis (les enfants de Duplessis). Psychiatry 69, 283–301CrossRefGoogle Scholar
Cozza, CJ (2006). Case studies of the orphans of Duplessis: the power of stories. Psychiatry 69, 325–7CrossRefGoogle ScholarPubMed
Broom, BC (2010). A reappraisal of the role of ‘mindbody’ factors in chronic urticaria. Postgraduate Medical Journal 86, 365–70CrossRefGoogle ScholarPubMed
Baumeister, and Masicampo, (2011), ref. 31
McAdams, DP and Olson, BD (2010). Personality development: continuity and change over the life course. Annual Review of Psychology 61, 517–42CrossRefGoogle ScholarPubMed
McLean, KC, Pasupathi, M and Pals, JL (2007). Selves creating stories creating selves: a process model of self-development. Personality and Social Psychology Review 11, 262–78; see also Fivush, R, Habermas, T, Waters, TEA and Zaman, W (2011). The making of autobiographical memory: intersections of culture, narratives and identity. International Journal of Psychology 46, 321–45CrossRefGoogle ScholarPubMed
Mauron, A (2001). Is the genome the secular equivalent of the soul?Science 291, 831–2CrossRefGoogle ScholarPubMed
Wright, NT, The New Testament and the People of God (London: SPCK, 1992), 32, 38, 40–5, 116Google Scholar
Birch, BC and Rasmussen, LL, Bible and Ethics in the Christian Life (Minneapolis: Augsburg, 1989), 106–7Google Scholar
Hauerwas, S, Truthfulness and Tragedy (Notre Dame: University of Notre Dame Press, 1977), 71–8, 223–5; Vision and Virtue (Notre Dame: University of Notre Dame Press, 1981), 68–77Google Scholar
Lk 6:27–8; Mt 5:44
Lk 10:25–37
Rolston, H, III, Genes, Genesis and God (Cambridge: Cambridge University Press, 1999), 248CrossRefGoogle Scholar
Mk 10:43–4; also 9:35; Mt 23:11
Paabo, S (2001). The human genome and our view of ourselves. Science 291, 1219–20CrossRefGoogle ScholarPubMed
Ayala, FJ, Darwin’s Gift to Science and Religion (Washington, DC: Joseph Henry Press, 2007), 177–8Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • Graeme Finlay, University of Auckland
  • Book: Human Evolution
  • Online publication: 05 June 2014
  • Chapter DOI: https://doi.org/10.1017/CBO9781139627092.008
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • Graeme Finlay, University of Auckland
  • Book: Human Evolution
  • Online publication: 05 June 2014
  • Chapter DOI: https://doi.org/10.1017/CBO9781139627092.008
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • Graeme Finlay, University of Auckland
  • Book: Human Evolution
  • Online publication: 05 June 2014
  • Chapter DOI: https://doi.org/10.1017/CBO9781139627092.008
Available formats
×