Hostname: page-component-7c8c6479df-995ml Total loading time: 0 Render date: 2024-03-28T05:46:20.575Z Has data issue: false hasContentIssue false

Behavior genetics and postgenomics

Published online by Cambridge University Press:  24 October 2012

Evan Charney*
Affiliation:
Department of Public Policy and Political Science, Duke Institute for Brain Sciences, Duke Institute for Genome Sciences and Policy, Duke University, Durham, NC 90239. http://fds.duke.edu/db/Sanford/faculty/echar

Abstract

The science of genetics is undergoing a paradigm shift. Recent discoveries, including the activity of retrotransposons, the extent of copy number variations, somatic and chromosomal mosaicism, and the nature of the epigenome as a regulator of DNA expressivity, are challenging a series of dogmas concerning the nature of the genome and the relationship between genotype and phenotype. According to three widely held dogmas, DNA is the unchanging template of heredity, is identical in all the cells and tissues of the body, and is the sole agent of inheritance. Rather than being an unchanging template, DNA appears subject to a good deal of environmentally induced change. Instead of identical DNA in all the cells of the body, somatic mosaicism appears to be the normal human condition. And DNA can no longer be considered the sole agent of inheritance. We now know that the epigenome, which regulates gene expressivity, can be inherited via the germline. These developments are particularly significant for behavior genetics for at least three reasons: First, epigenetic regulation, DNA variability, and somatic mosaicism appear to be particularly prevalent in the human brain and probably are involved in much of human behavior; second, they have important implications for the validity of heritability and gene association studies, the methodologies that largely define the discipline of behavior genetics; and third, they appear to play a critical role in development during the perinatal period and, in particular, in enabling phenotypic plasticity in offspring. I examine one of the central claims to emerge from the use of heritability studies in the behavioral sciences, the principle of minimal shared maternal effects, in light of the growing awareness that the maternal perinatal environment is a critical venue for the exercise of adaptive phenotypic plasticity. This consideration has important implications for both developmental and evolutionary biology.

Type
Target Article
Copyright
Copyright © Cambridge University Press 2012 

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

Aarnoudse-Moens, C. S., Weisglas-Kuperus, N., van Goudoever, J. B., & Oosterlaan, J. (2009) Meta-analysis of neurobehavioral outcomes in very preterm and/or very low birth weight children. Pediatrics 124(2):717–28.CrossRefGoogle ScholarPubMed
Abbas, A. K., Lichtman, A. H. & Pillai, S. (2010) Cellular and molecular immunology, 6th ed. Saunders/Elsevier.Google Scholar
Abdolmaleky, H. M., Yaqubi, S., Papageorgis, P., Lambert, A. W., Ozturk, S., Sivaraman, V. & Thiagalingam, S. (2011) Epigenetic dysregulation of HTR2A in the brain of patients with schizophrenia and bipolar disorder. Schizophrenia Research 129(2):183–90.CrossRefGoogle ScholarPubMed
Acosta-Rojas, R., Becker, J., Munoz-Abellana, B., Ruiz, C., Carreras, E. & Gratacos, E. (2007) Twin chorionicity and the risk of adverse perinatal outcome. International Journal of Gynecology & Obstetrics 96(2):98102.CrossRefGoogle ScholarPubMed
Adegbite, A. L., Castille, S., Ward, S. & Bajoria, R. (2004) Neuromorbidity in preterm twins in relation to chorionicity and discordant birth weight. American Journal of Obstetrics & Gynecology 190(1):156–63.CrossRefGoogle ScholarPubMed
Adegbite, A. L., Castille, S., Ward, S. & Bajoria, R. (2005) Prevalence of cranial scan abnormalities in preterm twins in relation to chorionicity and discordant birth weight. European Journal of Obstetrics & Gynecology and Reproductive Biology 119(1):4755.Google Scholar
Adkinsregan, E., Ottinger, M. A. & Park, J. (1995) Maternal transfer of estradiol to egg-yolks alters sexual-differentiation of avian offspring. Journal of Experimental Zoology 271(6):466–70.CrossRefGoogle Scholar
Akbarian, S. (2010) Epigenetics of schizophrenia. Current Topics in Behavioral Neuroscience 4:611–28.CrossRefGoogle ScholarPubMed
Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K. & Walter, P. (2002) Molecular biology of the cell. Garland.Google Scholar
Alkan, C., Coe, B. P. & Eichler, E. E. (2011) Genome structural variation discovery and genotyping. Nature Reviews Genetics 12(5):363–76.CrossRefGoogle ScholarPubMed
Allegrucci, C., Thurston, A., Lucas, E. & Young, L. (2005) Epigenetics and the germline. Reproduction 129(2):137–49.CrossRefGoogle ScholarPubMed
Almahbobi, G., Williams, L. J. & Hall, P. F. (1992) Attachment of mitochondria to intermediate filaments in adrenal cells: Relevance to the regulation of steroid synthesis. Experimental Cell Research 200(2):361–69.CrossRefGoogle Scholar
Almal, S. H. & Padh, H. (2011) Implications of gene copy-number variation in health and diseases. Journal of Human Genetics 57(1):613.CrossRefGoogle ScholarPubMed
Almeida, S. S., Garcia, R. A. & de Oliveira, L. M. (1993) Effects of early protein malnutrition and repeated testing upon locomotor and exploratory behaviors in the elevated plus-maze. Physiology & Behavior 54(4):749–52.CrossRefGoogle ScholarPubMed
Anderson, P. & Doyle, L. W. (2003) Neurobehavioral outcomes of school-age children born extremely low birth weight or very preterm in the 1990s. JAMA 289(24):3264–72.Google Scholar
Anderson, P. J. & Doyle, L. W. (2004) Executive functioning in school-aged children who were born very preterm or with extremely low birth weight in the 1990s. Pediatrics 114(1):5057.Google Scholar
Andersson, A. M., Jorgensen, N., Main, K. M., Toppari, J., Rajpert-De Meyts, E., Leffers, H. & Skakkebek, N. J. (2008) Adverse trends in male reproductive health: We may have reached a crucial “tipping point.” International Journal of Andrology 31(2):7480.CrossRefGoogle ScholarPubMed
Andersson-Ellström, A., Forssman, L. & Milsom, I. (1996) Age of sexual debut related to life-style and reproductive health factors in a group of Swedish teenage girls. Acta Obstetricia et Gynecologica Scandinavica 75(5):484–89.CrossRefGoogle Scholar
Antonarakis, S. E., Lyle, R., Dermitzakis, E. T., Reymond, A. & Deutsch, S. (2004) Chromosome 21 and Down syndrome: From genomics to pathophysiology. Nature Reviews Genetics 5(10):725–38.CrossRefGoogle ScholarPubMed
Antonow-Schlorke, I., Schwab, M., Cox, L. A., Li, C., Stuchlik, K., Witte, O. W. & McDonald, T. J. (2011) Vulnerability of the fetal primate brain to moderate reduction in maternal global nutrient availability. Proceedings of the National Academy of Sciences 108(7):3011–16.Google Scholar
Anway, M. D., Rekow, S. S. & Skinner, M. K. (2008) Transgenerational epigenetic programming of the embryonic testis transcriptome. Genomics 91:3040.Google Scholar
Apanius, V., Penn, D., Slev, P. R., Ruff, L. R. & Potts, W. K. (1997) The nature of selection on the major histocompatibility complex. Critical Reviews in Immunology 17(2):179224.CrossRefGoogle ScholarPubMed
Atamna, H. (2004) Heme, iron & the mitochondrial decay of ageing. Ageing Research Reviews 3(3):303–18.Google Scholar
Badyaev, A. V. & Oh, K. P. (2008) Environmental induction and phenotypic retention of adaptive maternal effects. BMC Evolutionary Biology 8:3.CrossRefGoogle ScholarPubMed
Bailey, A., Le Couteur, A., Gottesman, I., Bolton, P., Simonoff, E., Yuzda, E. & Rutter, M. (1995) Autism as a strongly genetic disorder: Evidence from a British twin study. Psychological Medicine 25(1):6377.CrossRefGoogle ScholarPubMed
Bailey, J. A., Kidd, J. M. & Eichler, E. E. (2008) Human copy number polymorphic genes. Cytogenetic and Genome Research 123(1–4):234–43.CrossRefGoogle ScholarPubMed
Baillie, J. K., Barnett, M. W., Upton, K. R., Gerhardt, D. J., Richmond, T. A., De Sapio, F., Brennan, P. M., Rizzu, P., Smith, S., Fell, M., Talbot, R. T., Gustincich, S., Freeman, T. C., Mattick, J. S., Hume, D. A., Heutink, P., Carninci, P., Jeddeloh, J. A. & Faulkner, G. J. (2011) Somatic retrotransposition alters the genetic landscape of the human brain. Nature 479(7374): 534–37.CrossRefGoogle ScholarPubMed
Barha, C. K., Pawluski, J. L. & Galea, L. A. M. (2007) Maternal care affects male and female offspring working memory and stress reactivity. Physiology & Behavior 92(5):939–50.CrossRefGoogle ScholarPubMed
Batzer, M. A. & Deininger, P. L. (2002) Alu repeats and human genomic diversity. Nature Reviews Genetics 3(5):370–79.CrossRefGoogle ScholarPubMed
Beard, C. & Jaenisch, R. (1993) Role for DNA methylation in genomic imprinting. Nature 366:362–65.Google Scholar
Becques, A., Larose, C., Gouat, P. & Serra, J. (2010) Effects of pre- and postnatal olfactogustatory experience on early preferences at birth and dietary selection at weaning in kittens. Chemical Senses 35(1):4145.CrossRefGoogle ScholarPubMed
Bednarczyk, M. R., Aumont, A., Dècary, S., Bergeron, R. & Fernandes, K. J. L. (2009) Prolonged voluntary wheel-running stimulates neural precursors in the hippocampus and forebrain of adult CD1 mice. Hippocampus 19(10):913–27.CrossRefGoogle ScholarPubMed
Bellinger, L., Lilley, C. & Langley-Evans, S. C. (2004) Prenatal exposure to a maternal low-protein diet programmes a preference for high-fat foods in the young adult rat. British Journal of Nutrition 92(3):513–20.CrossRefGoogle ScholarPubMed
Bellinger, L., Sculley, D. V. & Langley-Evans, S. C. (2006) Exposure to undernutrition in fetal life determines fat distribution, locomotor activity and food intake in ageing rats. International Journal of Obesity 30:729–38.Google Scholar
Beraldi, R., Pittoggi, C., Sciamanna, I., Mattei, E. & Spadafora, C. (2006) Expression of LINE-1 retroposons is essential for murine preimplantation development. Molecular Reproduction and Development 73(3):279–87.Google Scholar
Bernardo, J. (1996) Maternal effects in animal ecology. American Zoology 36(2):83105.Google Scholar
Bernot, A. (2004) Genome, transcriptome & proteome. Wiley.Google Scholar
Bernstein, B. E., Meissner, A. & Lander, E. S. (2007) The mammalian epigenome. Cell 128(4):669–81.Google Scholar
Bettegowda, A., Lee, K. B. & Smith, G. W. (2008) Cytoplasmic and nuclear determinants of the maternal-to-embryonic transition. Reproduction, Fertility and Development 20(1):4553.CrossRefGoogle ScholarPubMed
Beutler, E. (2001) Discrepancies between genotype and phenotype in hematology: An important frontier. Blood 98(9):2597–602.CrossRefGoogle Scholar
Beversdorf, D. Q., Manning, S. E., Hillier, A., Anderson, S. L., Nordgren, R. E., Walters, S. E. & Bauman, M. L. (2005) Timing of prenatal stressors and autism. Journal of Autism and Developmental Disorders 35(4):471–78.CrossRefGoogle ScholarPubMed
Biemont, C. & Vieira, C. (2006) Genetics: Junk DNA as an evolutionary force. Nature 443(7111):521–24.Google Scholar
Bilkó, Á., Altbäcker, V. & Hudson, R. (1994) Transmission of food preference in the rabbit: The means of information transfer. Physiology & Behavior 56(5):907–12.CrossRefGoogle ScholarPubMed
Blickstein, I. (2006) Monochorionicity in perspective. Ultrasound Obstetrics and Gynecology 27(3):235–38.Google Scholar
Blickstein, I., Goldman, R. D. & Mazkereth, R. (2000) Adaptive growth restriction as a pattern of birth weight discordance in twin gestations. Obstetrics and Gynecology 96(6):986–90.Google ScholarPubMed
Bobadilla, J. L., Macek, M., Fine, J. P. & Farrell, P. M. (2002) Cystic fibrosis: A worldwide analysis of CFTR mutations-correlation with incidence data and application to screening. Human Mutation 19(6):575606.CrossRefGoogle ScholarPubMed
Bogdarina, I., Welham, S., King, P. J., Burns, S. P. & Clark, A. J. (2007) Epigenetic modification of the renin-angiotensin system in the fetal programming of hypertension. Circulation Research 100:520–26.CrossRefGoogle ScholarPubMed
Boklage, C. E. (2009) Traces of embryogenesis are the same in monozygotic and dizygotic twins: Not compatible with double ovulation. Human Reproduction 24(6):1255–66.Google Scholar
Boks, M. P., Derks, E. M., Weisenberger, D. J., Strengman, E., Janson, E., Sommer, I. E. & Ophoff, R. A. (2009) The relationship of DNA methylation with age, gender and genotype in twins and healthy controls. PLoS ONE 4(8):e6767.Google Scholar
Bollati, V. & Baccarelli, A. (2010) Environmental epigenetics. Heredity 105(1):105–12.Google Scholar
Borgaonkar, D. S. (1997) Chromosomal variation in man: A catalog of chromosomal variants and anomalies. Wiley-Liss.Google Scholar
Borght, K., Van der Kubor-Nyakas, D., Klauke, K., Eggen, B. J. L., Nyakas, C., Van der Zee, E. A. & Meerlo, P. (2009) Physical exercise leads to rapid adaptations in hippocampal vasculature: Temporal dynamics and relationship to cell proliferation and neurogenesis. Hippocampus 19(10):928–36.CrossRefGoogle ScholarPubMed
Bosker, F. J., Hartman, C. A., Nolte, I. M., Prins, B. P., Terpstra, P., Posthuma, D. & Nolen, W. A. (2011) Poor replication of candidate genes for major depressive disorder using genome-wide association data. Molecular Psychiatry 16(5):516–32.Google Scholar
Bouchard, T. J. (2004) Genetic influence on human psychological traits: A survey. Current Directions in Psychological Science 13:148–51.CrossRefGoogle Scholar
Bouchard, T. J. & McGue, M. (2003) Genetic and environmental influences on human psychological differences. Journal of Neurobiology 54(1):445.Google Scholar
Bouchard, T. J. & Segal, N. L. (1985) Environment and IQ. In: Handbook of intelligence: Theories, measurements & applications, ed. Wohlman, B. B., pp. 391464. Wiley.Google Scholar
Bourgeois, J. P., Goldman-Rakic, P. S. & Rakic, P. (2000) Formation, elimination and stabilization of synapses in the primate cerebral cortex. In: The new cognitive neurosciences, ed. Gazzaniga, M., pp. 4553. MIT Press.Google Scholar
Brook, J. S., Brook, D. W., Kats, N., Arencibia-Mireles, O. & Finch, S. J. (2009) Ecology and drug use: Universal predictors across time and place. Psychological Reports 104(3):9891006.CrossRefGoogle ScholarPubMed
Brosius, J. (1999) RNAs from all categories generate retrosequences that may be exapted as novel genes or regulatory elements. Gene 238(1):115–34.CrossRefGoogle ScholarPubMed
Brosius, J. (2003) The contribution of RNAs and retroposition to evolutionary novelties. Genetica 118(2–3):99116.Google Scholar
Brouha, B. (2003) Hot L1s account for the bulk of retrotransposition in the human population. Proceedings of the National Academy of Sciences USA 100:5280–85.Google Scholar
Brouwers, E. P. M., van Baar, A. L., & Pop, V. J. M. (2001) Maternal anxiety during pregnancy and subsequent infant development. Infant Behavior and Development 24(1):95106.CrossRefGoogle Scholar
Brown, A. S. (2002) Prenatal risk factors and schizophrenia. Expert Review of Neurotherapeutics 2(1):5360.Google Scholar
Brown, J. L. & Eklund, A. (1994) Kin recognition and the major histocompatibility complex: An integrative review. American Naturalist 143(3):435–61.CrossRefGoogle Scholar
Bruder, C. E. G., Piotrowski, A., Gijsbers, A. A. C. J., Andersson, R., Erickson, S., Diaz de Stahl, T., Menzel, U., Sandgren, J., von Tell, D., Poplawski, A., Crowley, M., Crasto, M. C., Partridge, E. C., Tiwari, H., Allison, D. B., Komorowski, J., van Ommen, G.-J. B., Boomsma, D. I., Pedersen, N. L., den Dunnen, J. T., Wirdefeldt, K. & Dumanski, J. P. (2008b) Phenotypically concordant and discordant monozygotic twins display different DNA copy-number-variation profiles. American Journal of Human Genetics 82:19.Google Scholar
Bruel-Jungerman, E., Laroche, S. & Rampon, C. (2005) New neurons in the dentate gyrus are involved in the expression of enhanced long-term memory following environmental enrichment. European Journal of Neuroscience 21(2):513–21.Google Scholar
Brunton, P. J. & Russell, J. A. (2011) Neuroendocrine control of maternal stress responses and fetal programming by stress in pregnancy. Progress in Neuropsychopharmacology & Biological Psychiatry 35(5):1178–91.Google Scholar
Brykczynska, U., Hisano, M., Erkek, S., Ramos, L., Oakeley, E. J., Roloff, T. C. & Peters, A. H. F. M. (2010) Repressive and active histone methylation mark distinct promoters in human and mouse spermatozoa. Nature Structural & Molecular Biology 17(6):679–87.Google Scholar
Buchanan, J. A. & Scherer, S. W. (2008) Contemplating effects of genomic structural variation. Genetics in Medicine 10:639–47.Google Scholar
Buizer-Voskamp, J. E., Muntjewerff, J.-W., Strengman, E., Sabatti, C., Stefansson, H., Vorstman, J. A. S. & Ophoff, R. A. (2011) Genome-wide analysis shows increased frequency of copy number variation deletions in Dutch schizophrenia patients. Biological Psychiatry 70(7):655–62.CrossRefGoogle ScholarPubMed
Burt, S. A. (2009) Rethinking environmental contributions to child and adolescent psychopathology: A meta-analysis of shared environmental influences. Psychological Bulletin 135(4):608–37.Google Scholar
Cai, Q., Zhu, Z., Li, H., Fan, X., Jia, N., Bai, Z. & Liu, J. (2007) Prenatal stress on the kinetic properties of Ca2+ and K+ channels in offspring hippocampal CA3 pyramidal neurons. Life Sciences 80(7):681–89.Google Scholar
Cajal, S. R. (1899) Comparative study of the sensory areas of the human cortex. Clark University.Google Scholar
Caldji, C., Diorio, J. & Meaney, M. J. (2003) Variations in maternal care alter GABAA receptor subunit expression in brain regions associated with fear. Neuropsychopharmacology 28(11):1950–59.Google Scholar
Callahan, H. S. & Pigliucci, M. (2002) Shade-induced plasticity and its ecological significance in wild populations of Arabidopsis thaliana . Ecology 83(7):1965–80.CrossRefGoogle Scholar
Callinan, P. A. & Batzer, M. A. (2006) Retrotransposable elements and human disease. In: Genome and disease, ed. Volff, J. N., pp. 104–15. S. Karger.Google Scholar
Cameron, N. M., Shahrokh, D., Del Corpo, A., Dhir, S. K., Szyf, M., Champagne, F. A. & Meaney, M. J. (2008) Epigenetic programming of phenotypic variations in reproductive strategies in the rat through maternal care. Journal of Neuroendocrinology 20(6):795801.Google Scholar
Cameron, N. M., Soehngen, E. & Meaney, M. J. (2011) Variation in maternal care influences ventromedial hypothalamus activation in the rat. Journal of Neuroendocrinology 23(5):393400.CrossRefGoogle ScholarPubMed
Cao, X. & Chen, Y. (2009) Mitochondria and calcium signaling in embryonic development. Seminars in Cell & Developmental Biology 20(3):337–45.Google Scholar
Cao, Y., Lu, H. M. & Liang, J. (2010) Probability landscape of heritable and robust epigenetic state of lysogeny in phage lambda. Proceedings of the National Academy of Sciences USA 107(43):18445–50.Google Scholar
Carroll, S. G. M., Tyfield, L., Reeve, L., Porter, H., Soothill, P. & Kyle, P. M. (2005) Is zygosity or chorionicity the main determinant of fetal outcome in twin pregnancies? American Journal of Obstetrics & Gynecology 193(3):757–61.CrossRefGoogle ScholarPubMed
Caspi, A., McClay, J., Moffitt, T. E., Mill, J., Martin, J., Craig, I. W. & Poulton, R. (2002) Role of genotype in the cycle of violence in maltreated children. Science 297(5582):851–54.Google Scholar
Cavalier-Smith, T. (2004) The membranome and membrane heredity in development and evolution. In: Organelles, genomes and eukaryote phylogeny: An evolutionary synthesis in the age of genomics, ed. Horner, D. S. & Hirt, R. P., pp. 335–51. CRC Press.Google Scholar
Chada, S. R. & Hollenbeck, P. J. (2004) Nerve growth factor signaling regulates motility and docking of axonal mitochondria. Current Biology 14(14):1272–76.Google Scholar
Chahrour, M. & Zoghbi, H. Y. (2007) The story of Rett syndrome: From clinic to neurobiology. Neuron 56(3):422–37.CrossRefGoogle ScholarPubMed
Champagne, F. A. (2008) Epigenetic mechanisms and the transgenerational effects of maternal care. Frontiers in Neuroendocrinology 29:386–97.Google Scholar
Champagne, F. A. (2010a) Early adversity and developmental outcomes. Perspectives on Psychological Science 5(5):564–74.CrossRefGoogle ScholarPubMed
Champagne, F. A. (2010b) Epigenetic influence of social experiences across the lifespan. Developmental Psychobiology 52(4):299311.Google Scholar
Champagne, F. A. & Curley, J. P. (2008) Maternal regulation of estrogen receptor alpha methylation. Current Opinion in Pharmacology 8(6):735–39.Google Scholar
Champagne, F. A. & Curley, J. P. (2009) The trans-generational fluence of maternal care on offspring gene expression and behavior in rodents. In: Maternal effects in mammals, ed. Maestripieri, D. & Mateo, J. M., pp. 182202. University of Chicago Press.Google Scholar
Champagne, F. A., Francis, D. D., Mar, A. & Meaney, M. J. (2003a) Variations in maternal care in the rat as a mediating influence for the effects of environment on development. Physiology & Behavior 79:359–71.CrossRefGoogle ScholarPubMed
Champagne, F. A., Weaver, I., Sharma, S. & Meaney, M. J. (2003b) Natural variations in maternal care are associated with estrogen receptor α expression and estrogen sensitivity in the medial preoptic area. Endocrinology 144(11):4720–24.Google Scholar
Chanock, S. J., Manolio, T., Boehnke, M., Boerwinkle, E., Hunter, D. J., Thomas, G. & Collins, F. S. (2007) Replicating genotype–phenotype associations. Nature 447(7145):655–60.Google Scholar
Charil, A., Laplante, D. P., Vaillancourt, C. & King, S. (2010) Prenatal stress and brain development. Brain Research Reviews 65(1):5679.Google Scholar
Charney, E. & English, W. (2012) Candidate genes and political behavior. American Political Science Review 106(1):134.Google Scholar
Chen, M. & Manley, J. L. (2009) Mechanisms of alternative splicing regulation: Insights from molecular and genomics approaches. Nature Reviews Molecular Cell Biology 10(11):741–54.CrossRefGoogle ScholarPubMed
Cheng, L.-C., Pastrana, E., Tavazoie, M. & Doetsch, F. (2009) miR-124 regulates adult neurogenesis in the subventricular zone stem cell niche. Nature Neuroscience 12(4):399408.Google Scholar
Chevrud, J. M. & Wolf, J. B. (2009) The genetics and evolutionary consequences of maternal effects. In: Maternal effects in mammals, ed. Maestripieri, D. & Mateo, J., pp. 1137. University of Chicago Press.CrossRefGoogle Scholar
Chisholm, J., Quinlivan, J., Petersen, R. & Coall, D. (2005) Early stress predicts age at menarche and first birth, adult attachment & expected lifespan. Human Nature 16(3):233–65.CrossRefGoogle ScholarPubMed
Chuang, J. C. & Jones, P. A. (2007) Epigenetics and MicroRNAs. Pediatric Research 61(5, Part 2):24R29R.CrossRefGoogle ScholarPubMed
Clark, A. G., Eisen, M. B., Smith, D. R., Bergman, C. M., Oliver, B., Markow, T. A. & MacCallum, I. (2007) Evolution of genes and genomes on the Drosophila phylogeny. Nature 450(7167):203–18.Google Scholar
Clarke, A. S. & Schneider, M. L. (1993) Prenatal stress has long-term effects on behavioral responses to stress in juvenile rhesus monkeys. Developmental Psychobiology 26(5):293304.Google Scholar
Clarke, A. S., Wittwer, D. J., Abbott, D. H. & Schneider, M. L. (1994) Long-term effects of prenatal stress on HPA axis activity in juvenile rhesus monkeys. Developmental Psychobiology 27(5):257–69.Google Scholar
Clay Montier, L. L., Deng, J. J. & Bai, Y. (2009) Number matters: Control of mammalian mitochondrial DNA copy number. Journal of Genetics and Genomics 36(3):125–31. Available at: http://www.sciencedirect.com/science/article/pii/S1673852708600995 Google Scholar
Clay, H. B., Sillivan, S. & Konradi, C. (2011) Mitochondrial dysfunction and pathology in bipolar disorder and schizophrenia. International Journal of Developmental Neuroscience 29(3):311–24.Google Scholar
Clement, T. M., Savenkova, M. I., Settles, M., Anway, M. D. & Skinner, M. K. (2010) Alterations in the developing testis transcriptome following embryonic vinclozolin exposure. Reproductive Toxicology 30(3):353–64.Google Scholar
Coe, C. L., Kramer, M., Czéh, B., Gould, E., Reeves, A. J., Kirschbaum, C. & Fuchs, E. (2003) Prenatal stress diminishes neurogenesis in the dentate gyrus of juvenile rhesus monkeys. Biological Psychiatry 54(10):1025–34.Google Scholar
Colhoun, H. M., McKeigue, P. M. & Davey Smith, G. (2003) Problems of reporting genetic associations with complex outcomes. Lancet 361(9360):865.Google Scholar
Collinge, J. (2001) Prion diseases of humans and animals: Their causes and molecular basis. Annual Review of Neuroscience 24:519–50.Google Scholar
Conrad, D. F., Keebler, J. E., Depristo, M. A., Lindsay, S. J., Zhang, Y., Casals, F. & Awadalla, P. (2011) Variation in genome-wide mutation rates within and between human families. Nature Genetics 43(7):712–14.Google Scholar
Conrad, D. F., Pinto, D., Redon, R., Feuk, L., Gokcumen, O., Zhang, Y. & Hurles, M. E. (2010) Origins and functional impact of copy number variation in the human genome. Nature 464(7289):704–12.Google Scholar
Conway, D., Lytton, H. & Pysh, F. (1980) Twin–singleton language differences. Canadian Journal of Behavioural Science/Revue Canadienne des Sciences du Comportement 12(3):264–71.Google Scholar
Cooke, L. & Fildes, A. (2011) The impact of flavour exposure in utero and during milk feeding on food acceptance at weaning and beyond. Appetite 27:27.Google Scholar
Coppieters, N. & Dragunow, M. (2011) Epigenetics in Alzheimer's disease: A focus on DNA modifications. Current Pharmaceutical Design 9:9.Google Scholar
Cordaux, R. & Batzer, M. A. (2009) The impact of retrotransposons on human genome evolution. Nature Reviews Genetics 10:691703.Google Scholar
Coskun, P. E., Flint Beal, M. & Wallace, D. C. (2004) Alzheimer's brains harbor somatic mtDNA control-region mutations that suppress mitochondrial transcription and replication. Proceedings of the National Academy of Sciences USA 101(29):10726–31.Google Scholar
Cotman, C. W. & Berchtold, N. C. (2002) Exercise: A behavioral intervention to enhance brain health and plasticity. Trends in Neurosciences 25(6):295301.Google Scholar
Coufal, N. G., Garcia-Perez, J. L., Peng, G. E., Yeo, G. W., Mu, Y., Lovci, M. T. & Gage, F. H. (2009) L1 retrotransposition in human neural progenitor cells. Nature 460(7259):1127–31.CrossRefGoogle ScholarPubMed
Coulon, A., Gandrillon, O. & Beslon, G. (2010) On the spontaneous stochastic dynamics of a single gene: Complexity of the molecular interplay at the promoter. BMC Systems Biology 4(1):2.Google Scholar
Coupé, B., Dutriez-Casteloot, I., Breton, C., Lefèvre, F., Mairesse, J., Dickes-Coopman, A. & Vieau, D. (2009) Perinatal undernutrition modifies cell proliferation and brain-derived neurotrophic factor levels during critical time-windows for hypothalamic and hippocampal development in the male rat. Journal of Neuroendocrinology 21(1):4048.CrossRefGoogle ScholarPubMed
Craig, N. L., Craigie, R., Gellert, M. & Lambowitz, A. M. (2002) Mobile DNA II. American Society for Microbiology.Google Scholar
Crameri, R., Schulz-Knappe, P. & Zucht, H.-D. (2005) The future of post-genomic biology at the proteomic level: An outlook. Combinatorial Chemistry & High Throughput Screening 8(8):807–10.Google Scholar
Crews, D. (2008) Epigenetics and its implications for behavioral neuroendocrinology. Frontiers in Neuroendocrinology 29:344–57.Google Scholar
Crews, D., Gore, A. C., Hsu, T. S., Dangleben, N. L., Spinetta, M., Schallert, T. & Skinner, M. K. (2007) Transgenerational epigenetic imprints on mate preference. Proceedings of the National Academy of Sciences 104(14):5942–46.Google Scholar
Crick, F. (1958) On protein synthesis. International union of biochemistry symposium series 33:109–28.Google Scholar
Crick, F. (1970) Central dogma of molecular biology. Nature 227:561–73.Google Scholar
Crow, J. & Kimura, M. (1970) An introduction to population genetics theory. Burgess.Google Scholar
Cuzin, F. & Rassoulzadegan, M. (2010) Non-Mendelian epigenetic heredity: Gametic RNAs as epigenetic regulators and transgenerational signals. Essays in Biochemistry 48(1):101106.Google ScholarPubMed
Cuzin, F., Grandjean, V. & Rassoulzadegan, M. (2008) Inherited variation at the epigenetic level: Paramutation from the plant to the mouse. Current Opinion in Genetics & Development 18(2):193–96.Google Scholar
Dadoune, J.-P. (2009) Spermatozoal RNAs: What about their functions? Microscopy Research and Technique 72(8):536–51.Google Scholar
Daoud, R., Da Penha Berzaghi, M., Siedler, F., Hubener, M. & Stamm, S. (1999) Activity-dependent regulation of alternative splicing patterns in the rat brain. European Journal of Neuroscience 11:788802.Google Scholar
Darnaudery, M. & Maccari, S. (2008) Epigenetic programming of the stress response in male and female rats by prenatal restraint stress. Brain Research Reviews 57(2):571–85.Google Scholar
Dawkins, R. (2006) The selfish gene. Oxford University Press.Google Scholar
Day, J. E. (1932) The development of language in twins: I. A comparison of twins and single children. Child Development 3(3):179–99.Google Scholar
Day, J. J. & Sweatt, J. D. (2011) Epigenetic mechanisms in cognition. Neuron 70(5):813–29. doi: 10.1016/j.neuron.2011.05.019. CrossRefGoogle ScholarPubMed
De Genna, N. M., Larkby, C. & Cornelius, M. D. (2011) Pubertal timing and early sexual intercourse in the offspring of teenage mothers. Journal of Youth and Adolescence 40(10):1315–28.CrossRefGoogle ScholarPubMed
Dear, P. H. (2009) Copy-number variation: The end of the human genome? Trends in Biotechnology 27(8):448–54.Google Scholar
Delhanty, J. D. (2011) Inherited aneuploidy: Germline mosaicism. Cytogenetic and Genome Research 133(2–4):136–40.Google Scholar
Desagher, S. & Martinou, J.-C. (2000) Mitochondria as the central control point of apoptosis. Trends in Cell Biology 10(9):369–77.Google Scholar
Dickinson, J. E., Duncombe, G. J., Evans, S. F., French, N. P. & Hagan, R. (2005) The long term neurologic outcome of children from pregnancies complicated by twin-to-twin transfusion syndrome. British Journal of Gynecology 112(1):6368.Google Scholar
Dimauro, S. & Davidzon, G. (2005) Mitochondrial DNA and disease. Annals of Medicine 37(3):222–32.Google Scholar
Dipple, K. M., Phelan, J. K. & McCabe, E. R. B. (2001) Consequences of complexity within biological networks: Robustness and health, or vulnerability and disease. Molecular Genetics and Metabolism 74(1–2):4550.Google Scholar
Dole, N., Savitz, D. A., Hertz-Picciotto, I., Siega-Riz, A. M., McMahon, M. J. & Buekens, P. (2003) Maternal stress and preterm birth. American Journal of Epidemiology 157(1):1424.Google Scholar
Dolinoy, D. C. (2008) The agouti mouse model: An epigenetic biosensor for nutritional and environmental alterations on the fetal epigenome. Nutrition Reviews 66(s1):S7S11.CrossRefGoogle Scholar
Dowsett, A. P. & Young, M. W. (1982) Differing levels of dispersed repetitive DNA among closely related species of Drosophila. Proceedings of the National Academy of Sciences 79(15):4570–74.Google Scholar
Dunkel Schetter, C. (2011) Psychological science on pregnancy: Stress processes, biopsychosocial models & emerging research issues. Annual Review of Psychology 62:531–58.Google Scholar
Edwards, A., Ayroles, J., Stone, E., Carbone, M., Lyman, R. & Mackay, T. (2009a) A transcriptional network associated with natural variation in Drosophila aggressive behavior. Genome Biology 10(7):R76.Google Scholar
Edwards, A., Rollmann, S. M., Morgan, T. J. & Mackay, T. F. (2006) Quantitative genomics of aggressive behavior in Drosophila melanogaster. PLoS Genetics 2:e154.Google Scholar
Edwards, A., Zwarts, L., Yamamoto, A., Callaerts, P. & Mackay, T. (2009b) Mutations in many genes affect aggressive behavior in Drosophila melanogaster. BMC Biology 7(1):29.Google Scholar
Elbaz, A., Nelson, L. M., Payami, H., Ioannidis, J. P., Fiske, B. K. & Annesi, G. (2006) Lack of replication of thirteen single-nucleotide polymorphisms implicated in Parkinson's disease: A large-scale international study. Lancet Neurology 5:917–23.CrossRefGoogle ScholarPubMed
Ellis, B. J. (2004) Timing of pubertal maturation in girls: An integrated life history approach. Psychological Bulletin 130(6):920–58.Google Scholar
Ellis, B. J. & Garber, J. (2000) Psychosocial antecedents of variation in girls' pubertal timing: Maternal depression, stepfather presence and marital and family stress. Child Development 71(2):485501.Google Scholar
Evsikov, A. V., Graber, J. H., Michael Brockman, J., Hampl, A., Holbrook, A. E., Singh, P. & Knowles, B. B. (2006) Cracking the egg: Molecular dynamics and evolutionary aspects of the transition from the fully grown oocyte to embryo. Genes & Development 20(19):2713–27.Google Scholar
Farber, S. L. (1981) Identical twins reared apart: A reanalysis. Basic Books.Google Scholar
Farthing, C. R., Ficz, G., Ng, R. K., Chan, C. F., Andrews, S., Dean, W. & Reik, W. (2008) Global mapping of DNA methylation in mouse promoters reveals epigenetic reprogramming of pluripotency genes. PLoS Genetics 4(6):e1000116.Google Scholar
Faulkner, G. J. (2011) Retrotransposons: Mobile and mutagenic from conception to death. FEBS Letters 585(11):1589–94.Google Scholar
Faulkner, G. J., Kimura, Y., Daub, C. O., Wani, S., Plessy, C., Irvine, K. M. & Carninci, P. (2009) The regulated retrotransposon transcriptome of mammalian cells. Nature Genetics 41(5):563–71.Google Scholar
Feinberg, A. P. & Irizarry, R. A. (2010) Stochastic epigenetic variation as a driving force of development, evolutionary adaptation & disease. Proceedings of the National Academy of Sciences 107(suppl. 1):1757–64.Google Scholar
Ferrell, J. E. Jr. (2002) Self-perpetuating states in signal transduction: Positive feedback, double-negative feedback and bistability. Current Opinion in Cell Biology 14(2):140–48.Google Scholar
Feuk, L., Carson, A. R. & Scherer, S. W. (2006) Structural variation in the human genome. Nature Reviews Genetics 7(2):8597.Google Scholar
Field, T., Diego, M., Dieter, J., Hernandez-Reif, M., Schanberg, S., Kuhn, C. & Bendell, D. (2004b) Prenatal depression effects on the fetus and the newborn. Infant Behavior and Development 27(2):216–29.Google Scholar
Field, T., Diego, M., Hernandez-Reif, M., Vera, Y., Gil, K., Schanberg, S. & Gonzalez-Garcia, A. (2004a) Prenatal maternal biochemistry predicts neonatal biochemistry. International Journal of Neuroscience 114(8):933–45.Google Scholar
Filonzi, L., Magnani, C., Lavezzi, A. M., Rindi, G., Parmigiani, S., Bevilacqua, G. & Marzano, F. N. (2009) Association of dopamine transporter and monoamine oxidase molecular polymorphisms with sudden infant death syndrome and stillbirth: New insights into the serotonin hypothesis. Neurogenetics 10(1):6572.Google Scholar
Fish, E., Shahrokh, D., Bagot, R., Caldji, C., Bredy, T., Szyf, M. & Meaney, M. J. (2004) Epigenetic programming of stress responses through variations in maternal care. Annals of the New York Academy of Sciences (The Uterus and Human Reproduction) 1036:167–80.CrossRefGoogle ScholarPubMed
Flanagan, J. M., Popendikyte, V., Pozdniakovaite, N., Sobolev, M., Assadzadeh, A., Schumacher, A. & Petronis, A. (2006) Intra- and interindividual epigenetic variation in human germ cells. American Journal of Human Genetics 79(1):6784.Google Scholar
Forestell, C. A. & Mennella, J. A. (2007) Early determinants of fruit and vegetable acceptance. Pediatrics 120(6):1247–54.CrossRefGoogle ScholarPubMed
Fouse, S. D., Shen, Y., Pellegrini, M., Cole, S., Meissner, A., Van Neste, L., & Fan, G. (2008) Promoter CpG methylation contributes to ES cell gene regulation in parallel with Oct4/Nanog, PcG complex, and histone H3 K4/K27 trimethylation. Cell Stem Cell 2(2):160–69.Google Scholar
Fraga, M. F., Ballesta, E., Paz, M. F., Ropero, S. & Setien, F. (2005) Epigenetic differences arise during the lifetime of monozygotic twins. Proceedings of the National Academy of Sciences USA 102:10604.Google Scholar
Francis, D. & Meaney, M. J. (1999) Maternal care and the development of stress responses. Current Opinion in Neurobiology 9:128.Google Scholar
Franke, B., Neale, B. & Faraone, S. (2009) Genome-wide association studies in ADHD. Human Genetics 126(1):1350.Google Scholar
Frydman, C., Camerer, C., Bossaerts, P. & Rangel, A. (2011) MAOA-L carriers are better at making optimal financial decisions under risk. Proceedings of the Royal Society B: Biological Sciences 278(1714):2053–59.Google Scholar
Fuchs, E. & Gould, E. (2000) Mini-review: In vivo neurogenesis in the adult brain: Regulation and functional implications. European Journal of Neuroscience 12:2211.Google Scholar
Fuemmeler, B. F., Agurs-Collins, T. D., McClernon, F. J., Kollins, S. H., Kail, M. E., Bergen, A. W. & Ashley-Koch, A. E. (2008) Genes implicated in serotonergic and dopaminergic functioning predict BMI categories. Obesity 16(2):348–55.Google Scholar
Fusco, G. & Minelli, A. (2010) Phenotypic plasticity in development and evolution: Facts and concepts. Introduction. Philosophical Transactions of the Royal Society of London B Biological Sciences 365(1540):547–56.Google Scholar
Galler, J. R., Ramsey, F., Solimano, G., Lowell, W. E. & Mason, E. (1983) The influence of early malnutrition on subsequent behavioral development: I. degree of impairment in intellectual performance. Journal of the American Academy of Child Psychiatry 22(1):815.Google Scholar
Galvan, A., Ioannidis, J. P. A. & Dragani, T. A. (2010) Beyond genome-wide association studies: Genetic heterogeneity and individual predisposition to cancer. Trends in Genetics 26(3):132–41.Google Scholar
Gangestad, S. W. & Simpson, J. A. (2000) The evolution of human mating: Trade-offs and strategic pluralism. Behavioral and Brain Sciences 23(4):573–87.Google Scholar
Gao, F.-B. (2008) Posttranscriptional control of neuronal development by microRNA networks. Trends in Neurosciences 31(1):2026.Google Scholar
Garland, T. Jr & Kelly, S. A. (2006) Phenotypic plasticity and experimental evolution. Journal of Experimental Biology 209(12):2344–61.Google Scholar
Gerrish, C. J. & Mennella, J. A. (2001) Flavor variety enhances food acceptance in formula-fed infants. American Journal of Clinical Nutrition 73(6):1080–85.Google Scholar
Gersen, S. L. & Keagle, M. B., eds. (2005) The principles of clinical cytogenetics. Humana Press.Google Scholar
Gibbs, R. A. (2003) The International HapMap Project. Nature 426(6968):789–96.Google Scholar
Gibson, J. & Cameron, A. (2008) Complications of monochorionic twins. Paediatrics and Child Health 18(12):568–73.Google Scholar
Gielen, M., van Beijsterveldt, C. E., Derom, C., Vlietinck, R., Nijhuis, J. G., Zeegers, M. P. & Boomsma, D. I. (2010) Secular trends in gestational age and birthweight in twins. Human Reproduction 25(9):2346–53.Google Scholar
Gilbert, N., Lutz, S., Morrish, T. A. & Moran, J. V. (2005) Multiple fates of L1 retrotransposition intermediates in cultured human cells. Molecular and Cellular Biology 25:7780–95.Google Scholar
Gilbert, S. F. & Epel, D. (2009) Ecological developmental biology: Integrating epigenetics, medicine & evolution. Sinauer Associates.Google Scholar
Gilgenkrantz, S. & Janot, C. (1983) Monozygotic twins discordant for trisomy 21 or chimeric dizygotic twins? American Journal of Medical Genetics 15(1):159–60.CrossRefGoogle ScholarPubMed
Gimpl, G. & Fahrenholz, F. (2001) The oxytocin receptor system: Structure, function & regulation. Physiological Reviews 81(2):629–83.Google Scholar
Gissis, S. B. & Jablonka, E., eds. (2011) Transformations of Lamarckism: From subtle fluids to molecular biology. Cambridge University Press.Google Scholar
Giulivi, C., Zhang, Y.-F., Omanska-Klusek, A., Ross-Inta, C., Wong, S., Hertz-Picciotto, I. & Pessah, I. N. (2010) Mitochondrial dysfunction in autism. JAMA 304(21):2389–96.Google Scholar
Glessner, J. T., Wang, K., Cai, G., Korvatska, O., Kim, C. E., Wood, S. & Hakonarson, H. (2009) Autism genome-wide copy number variation reveals ubiquitin and neuronal genes. Nature 459(7246):569–73.Google Scholar
Glinianaia, S. V., Obeysekera, M. A., Sturgiss, S. & Bell, R. (2011) Stillbirth and neonatal mortality in monochorionic and dichorionic twins: A population-based study. Human Reproduction 26(9):2549–57.Google Scholar
Glinianaia, S. V., Rankin, J. & Wright, C. (2008) Congenital anomalies in twins: A register-based study. Human Reproduction 23(6):1306–11.CrossRefGoogle ScholarPubMed
Gluckman, P. D. & Hanson, M. A. (2006) Developmental origins of health and disease. ed. Gluckman, P. D. & Hanson, M. A.. Cambridge University Press.Google Scholar
Gluckman, P. D., Hanson, M. A., Bateson, P., Beedle, A. S., Law, C. M., Bhutta, Z. A. & West-Eberhard, M. J. (2009) Towards a new developmental synthesis: Adaptive developmental plasticity and human disease. Lancet 373(9675):1654–57.Google Scholar
Godfrey, K. M., Sheppard, A., Gluckman, P. D., Lillycrop, K. A., Burdge, G. C., McLean, C. & Hanson, M. A. (2011) Epigenetic gene promoter methylation at birth is associated with child's later adiposity. Diabetes 60(5):1528–34.Google Scholar
Goodier, J. L. & Kazazian, H. H. (2008) Retrotransposons revisited: The restraint and rehabilitation of parasites. Cell 135:23.Google Scholar
Goyal, R., Goyal, D., Leitzke, A., Gheorghe, C. P. & Longo, L. D. (2010) Brain renin-angiotensin system: Fetal epigenetic programming by maternal protein restriction during pregnancy. Reproductive Sciences 17(3):227–38.Google Scholar
Graber, J. A., Brooks-Gunn, J. & Warren, M. P. (1995) The antecedents of menarcheal age: Heredity, family environment & stressful life events. Child Development 66(2):346–59.Google Scholar
Grabowski, P. (2011) Alternative splicing takes shape during neuronal development. Current Opinion in Genetics & Development 21(4):388–94.Google Scholar
Grafodatskaya, D., Chung, B., Szatmari, P. & Weksberg, R. (2010) Autism spectrum disorders and epigenetics. Journal of the American Academy of Child and Adolescent Psychiatry 49(8): 794809.Google Scholar
Grandjean, V., Gounon, P., Wagner, N., Martin, L., Wagner, K. D., Bernex, F. & Rassoulzadegan, M. (2009) The miR-124-Sox9 paramutation: RNA-mediated epigenetic control of embryonic and adult growth. Development 136(21):3647–55.Google Scholar
Gredilla, R., Bohr, V. A. & Stevnsner, T. (2010) Mitochondrial DNA repair and association with aging – An update. Experimental Gerontology 45(7–8):478–88.Google Scholar
Gressler, S. & Haslberger, A. G., eds. (2010) Epigenetics and human health: Linking hereditary, environmental, and nutritional aspects. Wiley-VCH.Google Scholar
Grimes, G. W. & Aufderheide, K. J. (1991) Cellular aspects of pattern formation: The problem of assembly. Karger.Google Scholar
Grizenko, N., Shayan, Y. R., Polotskaia, A., Ter-Stepanian, M. & Joober, R. (2008) Relation of maternal stress during pregnancy to symptom severity and response to treatment in children with ADHD. Journal of Psychiatry and Neuroscience 33(1):1016.Google Scholar
Groothuis, T. G. G. & Schwabl, H. (2008) Hormone-mediated maternal effects in birds: Mechanisms matter but what do we know of them? Philosophical Transactions of the Royal Society B: Biological Sciences 363(1497):1647–61.Google Scholar
Gruber, B. M. (2011) Epigenetics and etiology of neurodegenerative diseases. Postȩpy Higieny i Medycyny Doświadczalnej 65:542–51.Google Scholar
Guo, S. W. (2001) Does higher concordance in monozygotic twins than in dizygotic twins suggest a genetic component? Human Heredity 51(3):121–32.Google Scholar
Hack, K. E. A., Derks, J. B., Elias, S. G., Franx, A., Roos, C., Voerman, S. K. & Visser, G. H. A. (2008) Increased perinatal mortality and morbidity in monochorionic versus dichorionic twin pregnancies: Clinical implications of a large Dutch cohort study. BJOG: An International Journal of Obstetrics & Gynaecology 115(1):5867.Google Scholar
Hallmayer, J., Cleveland, S., Torres, A., Phillips, J., Cohen, B., Torigoe, T. & Risch, N. (2011) Genetic heritability and shared environmental factors among twin pairs with autism. Archives of General Psychiatry 68(11):1095–102.Google Scholar
Hancks, D. C. & Kazazian, H. H. Jr. (2010) SVA retrotransposons: Evolution and genetic instability. Seminars in Cancer Biology 20(4):234–45.Google Scholar
Härkönen, K. (2005) Pesticides and the induction of aneuploidy in human sperm. Cytogenetic and Genome Research 111(3–4):378–83.Google Scholar
Harris, A. & Seckl, J. (2011) Glucocorticoids, prenatal stress and the programming of disease. Hormones and Behavior 59(3):279–89.Google Scholar
Hassold, T. & Hunt, P. (2001) To err(meiotically) is human: The genesis of human aneuploidy. Nature Reviews Genetics 2:280–91.Google Scholar
Hastings, P. J., Lupski, J. R., Rosenberg, S. M. & Ira, G. (2009) Mechanisms of change in gene copy number. Nature Reviews Genetics 10(8):551–64.Google Scholar
Hausner, H., Nicklaus, S., Issanchou, S., Molgaard, C. & Moller, P. (2010) Breastfeeding facilitates acceptance of a novel dietary flavour compound. Clinical Nutrition 29(1):141–48.Google Scholar
Havlicek, J. & Craig Roberts, S. (2009) MHC-correlated mate choice in humans: A review. Psychoneuroendocrinology 34(4):497512.Google Scholar
Hay, D. A., Prior, M., Collett, S. & Williams, M. (1987) Speech and language development in preschool twins. Acta Geneticae Medicae et Gemellologiae 36(2):213–23.Google Scholar
Heijmans, B. T., Tobi, E. W., Stein, A. D., Putter, H., Blauw, G. J. & Susser, E. S. (2008) Persistent epigenetic differences associated with prenatal exposure to famine in humans. Proceedings of the National Academy of Science USA 105:17046–49.Google Scholar
Hepper, P. G. (1988) Adaptive fetal learning: Prenatal exposure to garlic affects postnatal preferences. Animal Behaviour 36:935–36.Google Scholar
Hepper, P. G. (1995) Human fetal “olfactory” learning. International Journal of Prenatal and Perinatal Psychology and Medicine 7(2):147–51.Google Scholar
Hilbricht, T., Varotto, S., Sgaramella, V., Bartels, D., Salamini, F. & Furini, A. (2008) Retrotransposons and siRNA have a role in the evolution of desiccation tolerance leading to resurrection of the plant Craterostigma plantagineum . New Phytologist 179(3):877–87.Google Scholar
Hirschhorn, J. N., Lohmueller, K., Byrne, E. & Hirschhorn, K. (2002) A comprehensive review of genetic association studies. Genetics in Medicine 4:4561.Google Scholar
Ho, D. H. & Burggren, W. W. (2010) Epigenetics and transgenerational transfer: A physiological perspective. Journal of Experimental Biology 213(1):316.Google Scholar
Horton, T. H. (2005) Fetal origins of developmental plasticity: Animal models of induced life history variation. American Journal of Human Biology 17(1):3443.Google Scholar
Horváthová, T., Nakagawa, S. & Uller, T. (2012) Strategic female reproductive investment in response to male attractiveness in birds. Proceedings of the Royal Society B: Biological Sciences 279(1726):163–70.Google Scholar
Horwitz, A. V., Videon, T. M., Schmitz, M. F. & Davis, D. (2003) Rethinking twins and environments: Possible social sources for assumed genetic influences in twin research. Journal of Health and Social Behavior 44(2):111–29.CrossRefGoogle ScholarPubMed
Hoskins, R., Smith, C., Carlson, J., Bernardo Carvalho, A., Halpern, A., Kaminker, J. & Karpen, G. (2002) Heterochromatic sequences in a Drosophila whole-genome shotgun assembly. Genome Biology 3(12):163–70.Google Scholar
Hroudová, J. & Fišar, Z. (2011) Connectivity between mitochondrial functions and psychiatric disorders. Psychiatry and Clinical Neurosciences 65(2):130–41.Google Scholar
Huang, C. R., Schneider, A. M., Lu, Y., Niranjan, T., Shen, P., Robinson, M. A. & Burns, K. H. (2010) Mobile interspersed repeats are major structural variants in the human genome. Cell 141(7):1171–82.Google Scholar
Huizink, A. C., de Medina, P. G., Mulder, E. J., Visser, G. H. & Buitelaar, J. K. (2002) Psychological measures of prenatal stress as predictors of infant temperament. Journal of the American Academy of Child & Adolescent Psychiatry 41(9):1078–85.Google Scholar
Hulten, M. A., Patel, S. D., Westgren, M., Papadogiannakis, N., Jonsson, A. M., Jonasson, J. & Iwarsson, E. (2010) On the paternal origin of trisomy 21 Down syndrome. Molecular Cytogenetics 3:4.Google Scholar
Hurles, M. E., Dermitzakis, E. T. & Tyler-Smith, C. (2008) The functional impact of structural variation in humans. Trends in Genetics 24:238–45.Google Scholar
Iafrate, A. J. (2004) Detection of large-scale variation in the human genome. Nature Genetics 36:949–51.Google Scholar
Ihara, Y. & Feldman, M. W. (2003) Evolution of disassortative and assortative mating preferences based on imprinting. Theoretical Population Biology 64(2):193200.Google Scholar
Ingram Cooke, R. W. (2010) Does neonatal and infant neurodevelopmental morbidity of multiples and singletons differ? Seminars in Fetal and Neonatal Medicine 15(6):362–66.Google Scholar
International Human Genome Sequencing Consortium (2004) Finishing the euchromatic sequence of the human genome. Nature 431(7011):931–45.Google Scholar
Ioannidis, J. P. A. (2006b) Common genetic variants for breast cancer: 32 largely refuted candidates and larger prospects. Journal of the National Cancer Institute 98(19):1350–53.Google Scholar
Ioannidis, J. P., Trikalinos, T. A. & Khoury, M. J. (2006) Implications of small effect sizes of individual genetic variants on the design and interpretation of genetic association studies of complex diseases. American Journal of Epidemiology 164(7):609–14.Google Scholar
Iourov, I. Y., Vorsanova, S. G. & Yurov, Y. B. (2006) Chromosomal variation in mammalian neuronal cells: Known facts and attractive hypotheses. International Review of Cytology 249:143–91.Google Scholar
Iourov, I. Y., Vorsanova, S. G., Liehr, T. & Yurov, Y. B. (2009) Aneuploidy in the normal, Alzheimer's disease and ataxia-telangiectasia brain: Differential expression and pathological meaning. Neurobiology of Disease 34:212–20.Google Scholar
Iskow, R. C., McCabe, M. T., Mills, R. E., Torene, S., Pittard, W. S., Neuwald, A. F. & Devine, S. E. (2010) Natural mutagenesis of human genomes by endogenous retrotransposons. Cell 141(7):1253–61.Google Scholar
Jablonka, E. (2004) Epigenetic epidemiology. International Journal of Epidemiology 33(5):929–35.Google Scholar
Jablonka, E. & Lamb, M. J. (2005) Evolution in four dimensions: Genetic, epigenetic, behavioral, and symbolic variation in the history of life. MIT Press.Google Scholar
Jablonka, E. & Raz, G. (2009) Transgenerational epigenetic inheritance: Prevalence, mechanisms and implications for the study of heredity and evolution. Quarterly Review of Biology 84:131–76.Google Scholar
Jackson, M. J., Beaudet, A. L. & O'Brien, W. E. (1986) Mammalian urea cycle enzymes. Annual Review of Genetics 20(1):431–64.Google Scholar
Jacobson, L. & Saplosky, R. (1991) The role of the hippocampus in feedback regulation of the hypothalamic–pituitary–adrenocortical axis. Endocrine Reviews 12(2):118–34.Google Scholar
Jia, R., Tai, F., An, S. & Zhang, X. (2011) Neonatal paternal deprivation or early deprivation reduces adult parental behavior and central estrogen receptor alpha expression in mandarin voles (Microtus mandarinus). Behavioural Brain Research 224(2):279–89.Google Scholar
Jirtle, R. L. & Skinner, M. K. (2007) Environmental epigenomics and disease susceptibility. Nature Reviews Genetics 8(4):253–62.Google Scholar
Joffe, M. (2010) What has happened to human fertility? Human Reproduction 25(2):295307.Google Scholar
Johannes, F. & Colome-Tatche, M. (2011) Quantitative epigenetics through epigenomic perturbation of isogenic lines. Genetics 188(1):215–27.Google Scholar
Johannes, F., Porcher, E., Teixeira, F. K., Saliba-Colombani, V., Simon, M., Agier, N. & Colot, V. (2009) Assessing the impact of transgenerational epigenetic variation on complex traits. PLoS Genetics 5(6):26.Google Scholar
Johnson, M. H. (2007) Essential reproduction. ed. Everitt, B. J.. Blackwell.Google Scholar
Joseph, J. (2004) The gene illusion. Algora.Google Scholar
Kamin, L. J. & Goldberger, A. S. (2002) Twin studies in behavioral research: A skeptical view. Theoretical Population Biology 61(1):8395.Google Scholar
Kaminsky, Z. A., Tang, T., Wang, S. C., Ptak, C., Oh, G. H., Wong, A. H., Feldcamp, L. A., Virtanen, C., Halfvarson, J., Tysk, C., McRae, A. F., Visscher, P. M., Montgomery, G. W., Gottesman, I. I., Martin, N. G. & Petronis, A. (2009) DNA methylation profiles in monozygotic and dizygotic twins. Nature Genetics 41(2):240–45.Google Scholar
Kaneda, M., Okano, M., Hata, K., Sado, T., Tsujimoto, N., Li, E. & Sasaki, H. (2004) Essential role for de novo DNA methyltransferase Dnmt3a in paternal and maternal imprinting. Nature 429(6994):900903.Google Scholar
Kano, H., Godoy, I., Courtney, C., Vetter, M. R., Gerton, G. L. & Ostertag, E. M. (2009) L1 retrotransposition occurs mainly in embryogenesis and creates somatic mosaicism. Genes & Development 23:1303–12.Google Scholar
Kato, M., Nakamura, M., Ichiba, M., Tomiyasu, A., Shimo, H., Higuchi, I. & Sano, A. (2011) Mitochondrial DNA deletion mutations in patients with neuropsychiatric symptoms. Neuroscience Research 69(4):331–36.Google Scholar
Kato, Y., Kaneda, M., Hata, K., Kumaki, K., Hisano, M. & Kohara, Y. (2007) Role of the Dnmt3 family in de novo methylation of imprinted and repetitive sequences during male germ cell development in the mouse. Human Molecular Genetics 16:2272–80.Google Scholar
Kaufmann, B. & van Oudenaarden, A. (2007) Stochastic gene expression: From single molecules to the proteome. Current Opinions in Genetics & Development 17:107–12.Google Scholar
Kaushal, D., Contos, J. J. A., Treuner, K., Yang, A. H., Kingsbury, M. A., Rehen, S. K. & Chun, J. (2003) Alteration of gene expression by chromosome loss in the postnatal mouse brain. Journal of Neuroscience 23(13):5599–606.Google Scholar
Kemme, K., Kaiser, S. & Sachser, N. (2007) Prenatal maternal programming determines testosterone response during social challenge. Hormones and Behavior 51(3):387–94.Google Scholar
Keverne, E. B. & Curley, J. P. (2008) Epigenetics, brain evolution and behaviour. Frontiers in Neuroendocrinology 29(3):398412.Google Scholar
Kim, J., Samaranayake, M. & Pradhan, S. (2009) Epigenetic mechanisms in mammals. Cellular and Molecular Life Sciences 66(4):596612.Google Scholar
Kingsbury, M. A. (2005) Aneuploid neurons are functionally active and integrated into brain circuitry. Proceedings of the National Academy of Science USA 102:6143–47.Google Scholar
Kinney, D. K., Munir, K. M., Crowley, D. J. & Miller, A. M. (2008) Prenatal stress and risk for autism. Neuroscience & Biobehavioral Reviews 32(8):1519–32.Google Scholar
Kirov, G., Grozeva, D., Norton, N., Ivanov, D., Mantripragada, K. K., Holmans, P. & O'Donovan, M. C. (2009) Support for the involvement of large copy number variants in the pathogenesis of schizophrenia. Human Molecular Genetics 18(8):1497–503.Google Scholar
Kitamura, T., Saitoh, Y., Takashima, N., Murayama, A., Niibori, Y., Ageta, H. & Inokuchi, K. (2009) Adult neurogenesis modulates the hippocampus-dependent period of associative fear memory. Cell 139(4):814–27.Google Scholar
Klein, J. (1986) Natural history of the histocompatibility complex. Wiley.Google Scholar
Knickmeyer, R. C., Kang, C., Woolson, S., Smith, J. K., Hamer, R. M., Lin, W. L. & Gilmore, J. H. (2011) Twin-singleton differences in neonatal brain structure. Twin Research and Human Genetics 14(3):268–76.CrossRefGoogle ScholarPubMed
Korbel, J. O. (2007) Paired-end mapping reveals extensive structural variation in the human genome. Science 318:420–26.Google Scholar
Kornblihtt, A. R., Pesce, C. G., Alonso, C. R., Cramer, P., Srebrow, A., Werbajh, S. & Muro, A. F. (1996) The fibronectin gene as a model for splicing and transcription studies. FASEB Journal 10(2):248–57.Google Scholar
Kovaleva, N. (2010) Germ-line transmission of trisomy 21: Data from 80 families suggest an implication of grandmaternal age and a high frequency of female-specific trisomy rescue. Molecular Cytogenetics 3(1):7.Google Scholar
Kovaleva, N. V. & Tahmasebi-Hesari, M. (2007) Gonadal mosaicism. Down Syndrome News 14(23).Google Scholar
Krichevsky, A. M. & Kosik, K. S. (2001) Neuronal RNA granules: A link between RNA localization and stimulation-dependent translation. Neuron 32(4):683–96.Google Scholar
Krichevsky, A. M., King, K. S., Donahue, C. P., Khrapko, K. & Kosik, K. S. (2003) A microRNA array reveals extensive regulation of microRNAs during brain development. RNA 9:1274–81.Google Scholar
Kuhn, H. G., Dickinson-Anson, H. & Gage, F. H. (1996) Neurogenesis in the dentate gyrus of the adult rat: Age-related decrease of neuronal progenitor proliferation. Journal of Neuroscience 16(6):2027–33.Google Scholar
Kuss, A. W. & Chen, W. (2008) MicroRNAs in brain function and disease. Current Neurology and Neuroscience Reports 8(3):190–97.Google Scholar
Kuzumaki, N., Ikegami, D., Tamura, R., Hareyama, N., Imai, S., Narita, M. & Narita, M. (2010) Hippocampal epigenetic modification at the brain-derived neurotrophic factor gene induced by an enriched environment. Hippocampus.Google Scholar
Lagos-Quintana, M., Rauhut, R., Yalcin, A., Meyer, J., Lendeckel, W. & Tuschl, T. (2002) Identification of tissue-specific microRNAs from mouse. Current Biology 12(9):735–39.Google Scholar
Lander, E. S. (2001) Initial sequencing and analysis of the human genome. Nature 409:860921.Google Scholar
Langley-Evans, S. C., Sherman, R. C., Welham, S. J., Nwagwu, M. O., Gardner, D. S. & Jackson, A. A. (1999) Intrauterine programming of hypertension: The role of the renin-angiotensin system. Biochemical Society Transactions 27(2):8893.Google Scholar
Lango, A., Hana Estrada, K., Lettre, G., Berndt, S. I., Weedon, M. N., Rivadeneira, F. & Hirschhorn, J. N. (2010) Hundreds of variants clustered in genomic loci and biological pathways affect human height. Nature 467(7317):832–38.Google Scholar
Laplante, P., Diorio, J. & Meaney, M. J. (2002) Serotonin regulates hippocampal glucocorticoid receptor expression via a 5-HT7 receptor. Developmental Brain Research 139(2):199203.Google Scholar
Larios-Del Toro, Y. E., Vasquez-Garibay, E. M., Gonzalez-Ojeda, A., Ramirez-Valdivia, J. M., Troyo-Sanroman, R. & Carmona-Flores, G. (2011) A longitudinal evaluation of growth outcomes at hospital discharge of very-low-birth-weight preterm infants. European Journal of Clinical Nutrition 16(10):191.Google Scholar
Lazinski, M. J., Shea, A. K. & Steiner, M. (2008) Effects of maternal prenatal stress on offspring development: A commentary. Archives of Women's Mental Health 11(5–6):363–75.Google Scholar
Lemke, H. A., Coutinho, A. & Lange, H. (2004) Lamarckian inheritance by somatically acquired maternal IgG phenotypes. Trends in Immunology 25(4):180–86.Google Scholar
Li, J., Olsen, J., Vestergaard, M. & Obel, C. (2010) Attention-deficit/hyperactivity disorder in the offspring following prenatal maternal bereavement: A nationwide follow-up study in Denmark. European Child & Adolescent Psychiatry 19(10):747–53.Google Scholar
Li, T., Spearow, J., Rubin, C. M. & Schmid, C. W. (1999) Physiological stresses increase mouse short interspersed element (SINE) RNA expression in vivo. Gene 239(2):367–72.Google Scholar
Li, Y., Zheng, H., Luo, R., Wu, H., Zhu, H., Li, R. & Wang, J. (2011) Structural variation in two human genomes mapped at single-nucleotide resolution by whole genome de novo assembly. Nature Biotechnology 29(8):723–30.Google Scholar
Liang, Q., Conte, N., Skarnes, W. C. & Bradley, A. (2008) Extensive genomic copy number variation in embryonic stem cells. Proceedings of the National Academy of Sciences 105(45):17453–56.Google Scholar
Lillycrop, K. A. & Burdge, G. C. (2011) Epigenetic changes in early life and future risk of obesity. International Journal of Obesity 35(1):7283.Google Scholar
Lillycrop, K. A., Phillips, E. S., Torrens, C., Hanson, M. A., Jackson, A. A. & Burdge, G. C. (2008) Feeding pregnant rats a protein-restricted diet persistently alters the methylation of specific cytosines in the hepatic PPAR alpha promoter of the offspring. British Journal of Nutrition 100:278–82.Google Scholar
Lim, L. P., Lau, N. C., Garrett-Engele, P., Grimson, A., Schelter, J. M., Castle, J. & Johnson, J. M. (2005) Microarray analysis shows that some microRNAs downregulate large numbers of target mRNAs. Nature 433(7027):769–73.Google Scholar
Lipniacki, T., Paszek, P., Marciniak-Czochra, A., Brasier, A. & Kimmel, M. (2006) Transcriptional stochasticity in gene expression. Journal of Theoretical Biology 238:348–67.Google Scholar
Lista, I. & Sorrentino, G. (2009) Biological mechanisms of physical activity in preventing cognitive decline. Cellular and Molecular Neurobiology.Google Scholar
Little, J., Higgins, J. P. T., Ioannidis, J. P. A., Moher, D., Gagnon, F., von Elm, E. & Birkett, N. (2009) Strengthening the reporting of genetic association studies (STREGA): An extension of the STROBE statement. Annals of Internal Medicine 150(3):206–15.Google Scholar
Liu, D., Diorio, J., Tannenbaum, B., Caldji, C. & Francis, D. (1997) Maternal care, hippocampal glucocorticoid receptors, and hypothalamic-pituitary-adrenal responses to stress. Science 277:1659.Google Scholar
Lucentini, J. (2004) Gene association studies typically wrong: Reproducible gene-disease associations are few and far between. Scientist 18(24):20(1).Google Scholar
Luco, R. F., Pan, Q., Tominaga, K., Blencowe, B. J., Pereira-Smith, O. M. & Misteli, T. (2010) Regulation of alternative splicing by histone modifications. Science 327(5968):9961000.Google Scholar
Lui, C. C., Wang, J.-Y., Tain, Y.-L., Chen, Y.-C., Chang, K.-A., Lai, M.-C. & Huang, L.-T. (2011) Prenatal stress in rat causes long-term spatial memory deficit and hippocampus MRI abnormality: Differential effects of postweaning enriched environment. Neurochemistry International 58(3):434–41.Google Scholar
Lynch, M., Koskella, B. & Schaack, S. (2006) Mutation pressure and the evolution of organelle genomic architecture. Science 311(5768):1727–30.Google Scholar
Maccari, S., Darnaudery, M., Morley-Fletcher, S., Zuena, A. R., Cinque, C. & Van Reeth, O. (2003) Prenatal stress and long-term consequences: Implications of glucocorticoid hormones. Neuroscience & Biobehavioral Reviews 27(1–2):119–27.Google Scholar
Machin, G. A. & Keith, L. G. (1999) An atlas of multiple pregnancy: Biology and pathology. Parthenon.Google Scholar
Maestripieri, D. & Mateo, J. M. (2009) Maternal effects in mammals. University of Chicago Press.Google Scholar
Maier, A. S., Chabanet, C., Schaal, B., Leathwood, P. D. & Issanchou, S. N. (2008) Breastfeeding and experience with variety early in weaning increase infants' acceptance of new foods for up to two months. Clinical Nutrition 27(6):849–57.Google Scholar
Maier, A., Chabanet, C., Schaal, B., Leathwood, P. & Issanchou, S. (2007) Food-related sensory experience from birth through weaning: Contrasted patterns in two nearby European regions. Appetite 49(2):429–40.Google Scholar
Makeyev, E. V., Zhang, J., Carrasco, M. A. & Maniatis, T. (2007) The MicroRNA miR-124 promotes neuronal differentiation by triggering brain-specific alternative pre-mRNA splicing. Molecular Cell 27(3):435–48.Google Scholar
Manolio, T. A., Collins, F. S., Cox, N. J., Goldstein, D. B., Hindorff, L. A., Hunter, D. J., McCarthy, M. I., Ramos, E. M., Cardon, L. R., Chakravarti, A., Cho, J. H., Guttmacher, A. E., Kong, A., Kruglyak, L., Mardis, E., Rotimi, C. N., Slatkin, M., Valle, D., Whittemore, A. S., Boehnke, M., Clark, A. G., Eichler, E. E., Gibson, G., Haines, J. L., Mackay, T. F., McCarroll, S. A. & Visscher, P. M. (2009) Finding the missing heritability of complex diseases. Nature 461(7265):747–53.Google Scholar
March, S. M., Abate, P., Spear, N. E. & Molina, J. C. (2009) Fetal exposure to moderate ethanol doses: Heightened operant responsiveness elicited by ethanol-related reinforcers. Alcoholism: Clinical & Experimental Research 33(11):1981–93.Google Scholar
Marchetto, M. C. N., Gage, F. H. & Muotri, A. R. (2010) Retrotransposition and neuronal diversity. In: Perspectives of stem cells: From tools for studying mechanisms of neuronal differentiation towards therapy, ed. Ulrich, H., pp. 8796. Springer.Google Scholar
Marshall, C. R., Noor, A., Vincent, J. B., Lionel, A. C., Feuk, L., Skaug, J. & Scherer, S. W. (2008) Structural variation of chromosomes in autism spectrum disorder. American Journal of Human Genetics 82(2):477–88.Google Scholar
Martienssen, R. A., Riggs, A. D. & Russo, V. E. A. (1996) Epigenetic mechanisms of gene regulation. Cold Spring Harbor Laboratory Press.Google Scholar
Martin, J. A. & Wang, Z. (2011) Next-generation transcriptome assembly. Nature Reviews Genetics 12(10):671–82.Google Scholar
Martin, S. L. (2009) Developmental biology: Jumping-gene roulette. Nature 460(7259):1087–88.Google Scholar
Mastripieri, D. & Mateo, J. M. (2009) The role of maternal effects in mammalian evolution and adaptation. In: Maternal effects in mammals, ed. Mastripieri, D. & Mateo, J. M., pp. 110. University of Chicago Press.Google Scholar
Mateo, J. M. (2009) Maternal influences on development, social relationships and survival behaviors. In: Maternal effects in mammals, ed. Maestripieri, D. & Mateo, J. M., pp. 133–58. University of Chicago Press.Google Scholar
Matlik, K., Redik, K. & Speek, M. (2006) L1 antisense promoter drives tissue-specific transcription of human genes. Journal of Biomedicine and Biotechnology 2006(1):71753.Google Scholar
Mattick, J. S. (2001) Non-coding RNAs: The architects of eukaryotic complexity. EMBO Reports 2(11):986–91.Google Scholar
Maulik, D. (2006) Fetal growth restriction: The etiology. Clinical Obstetrics and Gynecology 49(2):228–35.Google Scholar
MayPanloup, P., Chretien, M. F., Malthiery, Y. & Reynier, P. (2007) Mitochondrial DNA in the oocyte and the developing embryo. In: Current topics in developmental biology, ed. Justin, C. S. J., pp. 5183 Academic Press.Google Scholar
McBride, H. M., Neuspiel, M. & Wasiak, S. (2006) Mitochondria: More than just a powerhouse. Current Biology 16(14):R551–60.Google Scholar
McCabe, L. L. & McCabe, E. R. B. (2006) Complexity in genetic diseases: How patients inform the science by ignoring the dogma. American Journal of Medical Genetics Part A 140A(2):160–61.Google Scholar
McClellan, J. & King, M.-C. (2010) Genetic heterogeneity in human disease. Cell 141(2):210–17.Google Scholar
McEvoy, S. & Dodd, B. (1992) The communication abilities of 2- to 4-year-old twins. International Journal of Language & Communication Disorders 27(1):7387.Google Scholar
McGowan, P. O. & Szyf, M. (2010) The epigenetics of social adversity in early life: Implications for mental health outcomes. Neurobiology of Disease 39(1):6672.Google Scholar
McGowan, P. O., Meaney, M. J. & Szyf, M. (2008) Diet and the epigenetic(re)programming of phenotypic differences in behavior. Brain Research 1237:1224.Google Scholar
McGowan, P. O., Sasaki, A., D'Alessio, A. C., Dymov, S., Labonte, B., Szyf, M. & Meaney, M. J. (2009) Epigenetic regulation of the glucocorticoid receptor in human brain associates with childhood abuse. Nature Neuroscience 12(3):342–48.Google Scholar
McGue, M. & Bouchard, T. J. (1998) Genetic and environmental influences on human behavioral differences. Annual Review of Neuroscience 21(1):124.Google Scholar
McGue, M., Sharma, A. & Benson, P. (1996) The effect of common rearing on adolescent adjustment: Evidence from a U.S. adoption cohort. Developmental Psychology 32(4):604–13.Google Scholar
Meaney, M. J. & Aitken, D. H. (1985) The effects of early postnatal handling on hippocampal glucocorticoid receptor concentrations: Temporal parameters. Developmental Brain Research 22(2):301304.Google Scholar
Mehler, M. F. (2008) Epigenetic principles and mechanisms underlying nervous system functions in health and disease. Progress in Neurobiology 86:305–41.Google Scholar
Menard, J. L. & Hakvoort, R. M. (2007) Variations of maternal care alter offspring levels of behavioural defensiveness in adulthood: Evidence for a threshold model. Behavioural Brain Research 176(2):302–13.Google Scholar
Mennella, J. A. & Beauchamp, G. K. (1999) Experience with a flavor in mother's milk modifies the infant's acceptance of flavored cereal. Developmental Psychobiology 35(3):197203.Google Scholar
Mennella, J. A., Jagnow, C. P. & Beauchamp, G. K. (2001) Prenatal and postnatal flavor learning by human infants. Pediatrics 107(6):e88.Google Scholar
Mennella, J. A., Kennedy, J. M. & Beauchamp, G. K. (2006) Vegetable acceptance by infants: Effects of formula flavors. Early Human Development 82(7):463–68.Google Scholar
Mesquita, A. R., Wegerich, Y., Patchev, A. V., Oliveira, M., Leão, P., Sousa, N. & Almeida, O. F. X. (2009) Glucocorticoids and neuro- and behavioural development. Seminars in Fetal & Neonatal Medicine 14(3):130–35.Google Scholar
Meza-Sosa, K. F., Valle-García, D., Pedraza-Alva, G. & Pérez-Martínez, L. (2012) Role of microRNAs in central nervous system development and pathology. Journal of Neuroscience Research, 90(1):112.Google Scholar
Mills, R. E., Andrew Bennett, E., Iskow, R. C. & Devine, S. E. (2007) Which transposable elements are active in the human genome? Trends in Genetics 23(4):183–91.Google Scholar
Miltenberger, R. J., Mynatt, R. L., Wilkinson, J. E. & Woychik, R. P. (1997) The role of the agouti gene in the yellow obese syndrome. Journal of Nutrition 127(9):1902S.Google Scholar
Miura, A., Yonebayashi, S., Watanabe, K., Toyama, T., Shimada, H. & Kakutani, T. (2001) Mobilization of transposons by a mutation abolishing full DNA methylation in Arabidopsis. Nature 411(6834):212–14.Google Scholar
Moffitt, T. E., Caspi, A., Belsky, J. & Silva, P. A. (1992) Childhood experience and the onset of menarche: A test of a sociobiological model. Child Development 63(1):4758.Google Scholar
Molfese, D. L. (2011) Advancing neuroscience through epigenetics: Molecular mechanisms of learning and memory. Developmental Neuropsychology 36(7):810–27.Google Scholar
Morgan, C. P. & Bale, T. L. (2011) Early prenatal stress epigenetically programs dysmasculinization in second-generation offspring via the paternal lineage. Journal of Neuroscience 31(33):11748–55.Google Scholar
Morrow, E. M. (2010) Genomic copy number variation in disorders of cognitive development. Journal of the American Academy of Child & Adolescent Psychiatry 49(11):1091–104.Google Scholar
Mousseau, T. A. & Fox, C. W. (1998) The adaptive significance of maternal effects. Trends in Ecology and Evolution 13(10):403407.Google Scholar
Mueller, B. R. & Bale, T. L. (2008) Sex-specific programming of offspring emotionality after stress early in pregnancy. Journal of Neuroscience 28(36):9055–65.Google Scholar
Muotri, A. R., Marchetto, M. C. N., Coufal, N. G. & Gage, F. H. (2007) The necessary junk: New functions for transposable elements. Human Molecular Genetics 16(R2):R159–67.Google Scholar
Muotri, A. R., Marchetto, M. C. N., Coufal, N. G., Oefner, R., Yeo, G., Nakashima, K. & Gage, F. H. (2010) L1 retrotransposition in neurons is modulated by MeCP2. Nature 468(7322):443–46.Google Scholar
Muotri, A. R., Marchetto, M. C. N., Zhao, C. & Gage, F. H. (2009) Environmental influence on L1 retrotransposons in the adult hippocampus. Hippocampus 19(10):10021007.Google Scholar
Murgatroyd, C. & Spengler, D. (2011) Epigenetic programming of the HPA axis: Early life decides. Stress 19:19.Google Scholar
Nagaraj, N., Wisniewski, J. R., Geiger, T., Cox, J., Kircher, M., Kelso, J. & Mann, M. (2011) Deep proteome and transcriptome mapping of a human cancer cell line. Molecular Systems Biology 7(548):81.Google Scholar
Nagel, R. L. (2005) Epistasis and the genetics of human diseases. Comptes Rendus Biologies 328(7):606–15.Google Scholar
Naito, K., Cho, E., Yang, G., Campbell, M. A., Yano, K., Okumoto, Y. & Wessler, S. R. (2006) Dramatic amplification of a rice transposable element during recent domestication. Proceedings of the National Academy of Science USA 103(47):17620–25.Google Scholar
Naslund, K., Saetre, P., von Salome, J., Bergstrom, T. F., Jareborg, D. & Jazin, E. (2005) Genome-wide prediction of human VNTRs. Genomics 85(1):2435.Google Scholar
National Institute of Environmental Health (2010) Endocrine disruptors. Available at: http://www.niehs.nih.gov/health/topics/agents/endocrine/index.cfm.Google Scholar
Need, A. C. & Goldstein, D. B. (2010) Whole genome association studies in complex diseases: Where do we stand? Dialogues in Clinical Neuroscience 12(1):3746.Google Scholar
Need, A. C., Attix, D. K., McEvoy, J. M., Cirulli, E. T., Linney, K. L., Hunt, P. & Goldstein, D. B. (2009a) A genome-wide study of common SNPs and CNVs in cognitive performance in the CANTAB. Human Molecular Genetics 18(23):4650–61.Google Scholar
Need, A. C., Ge, D., Weale, M. E., Maia, J., Feng, S., Heinzen, E. L. & Goldstein, D. B. (2009b) A genome-wide investigation of SNPs and CNVs in schizophrenia. PLoS Genetics 5(2):e1000373.Google Scholar
Nelson, P. T., Kiriakidou, M., Mourelatos, Z., Tan, G. S., Jennings, M. H., Xie, K. & Wang, W.-X. (2010) High-throughput experimental studies to identify miRNA targets directly, with special focus on the mammalian brain. Brain Research 1338(0):122–30.Google Scholar
Nelson, P. T., Wang, W. X. & Rajeev, B. W. (2008) MicroRNAs (miRNAs) in Neurodegenerative Diseases. Brain Pathology 18(1):130–38.Google Scholar
Neves, G., Cooke, S. F. & Bliss, T. V. P. (2008) Synaptic plasticity, memory and the hippocampus: A neural network approach to causality. Nature Reviews Neuroscience 9(1):6575.Google Scholar
Nieuwint, A., Van Zalen-Sprock, R., Hummel, P., Pals, G., Van Vugt, J., Van Der Harten, H. & Madan, K. (1999) “Identical” twins with discordant karyotypes. Prenatal Diagnosis 19(1):7276.Google Scholar
Nilsen, T. W. & Graveley, B. R. (2010) Expansion of the eukaryotic proteome by alternative splicing. Nature 463(7280):457–63.Google Scholar
Nilsson, E. E., Anway, M. D., Stanfield, J. & Skinner, M. K. (2008) Transgenerational epigenetic effects of the endocrine disruptor vinclozolin on pregnancies and female adult onset disease. Reproduction 135:713–21.Google Scholar
Noble, D. (2010) Biophysics and systems biology. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 368(1914):1125–39.Google Scholar
Nobrega, M. A., Ovcharenko, I., Afzal, V. & Rubin, E. M. (2003) Scanning human gene deserts for long-range enhancers. Science 302(5644):413.Google Scholar
Notini, A. J., Craig, J. M. & White, S. J. (2008) Copy number variation and mosaicism. Cytogenetic and Genome Research 123(1–4):270–77.Google Scholar
Nussey, D. H., Wilson, A. J. & Brommer, J. E. (2007) The evolutionary ecology of individual phenotypic plasticity in wild populations. Journal of Evolutionary Biology 20(3):831–44.Google Scholar
O'Connor, T. G., Ben-Shlomo, Y., Heron, J., Golding, J., Adams, D. & Glover, V. (2005) Prenatal anxiety predicts individual differences in cortisol in pre-adolescent children. Biological Psychiatry 58(3):211–17.Google Scholar
Oberlander, T. F., Weinberg, J., Papsdorf, M., Grunau, R., Misri, S. & Devlin, A. M. (2008) Prenatal exposure to maternal depression, neonatal methylation of human glucocorticoid receptor gene (NR3C1) and infant cortisol stress responses. Epigenetics. 3(2):97106.Google Scholar
Ollikainen, M., Smith, K. R., Joo, E. J.-H., Kiat Ng, H., Andronikos, R., Novakovic, B., & Craig, J. M. (2010) DNA methylation analysis of multiple tissues from newborn twins reveals both genetic and intrauterine components to variation in the human neonatal epigenome. Human Molecular Genetics 19(21):4176–88.Google Scholar
Ooi, L. & Wood, I. C. (2008) Regulation of gene expression in the nervous system. Biochemical Journal 414:327–41.Google Scholar
Orchinik, L. J., Taylor, H. G., Espy, K. A., Minich, N., Klein, N., Sheffield, T. & Hack, M. (2011) Cognitive outcomes for extremely preterm/extremely low birth weight children in kindergarten. Journal of the International Neuropsychological Society 17(6):1067–79.Google Scholar
Orgel, L. E. & Crick, F. H. C. (1980) Selfish DNA: The ultimate parasite. Nature 284(5757):604607.Google Scholar
Ostertag, E. M. & Kazazian, H. H. Jr. (2001) Biology of mammalian L1 retrotransposons. Annual Review of Genetics 35:501–38.Google Scholar
Ostertag, E. M., Goodier, J. L., Zhang, Y. & Kazazian, H. H. (2003) SVA elements are nonautonomous retrotransposons that cause disease in humans. American Journal of Human Genetics 73:1444.Google Scholar
Paenke, I., Sendhoff, B. & Kawecki, T. J. (2007) Influence of plasticity and learning on evolution under directional selection. American Naturalist 170(2):5.Google Scholar
Painter, R. C., Roseboom, T. J. & Bleker, O. P. (2005) Prenatal exposure to the Dutch famine and disease in later life: An overview. Reproductive Toxicology 20:345–52.Google Scholar
Park, J.-H., Gail, M. H., Weinberg, C. R., Carroll, R. J., Chung, C. C., Wang, Z. & Chatterjee, N. (2011) Distribution of allele frequencies and effect sizes and their interrelationships for common genetic susceptibility variants. Proceedings of the National Academy of Sciences 108(44):18026–31.Google Scholar
Parnpiansil, P., Jutapakdeegul, N., Chentanez, T. & Kotchabhakdi, N. (2003) Exercise during pregnancy increases hippocampal brain-derived neurotrophic factor mRNA expression and spatial learning in neonatal rat pup. Neuroscience Letters 352(1):4548.Google Scholar
Paslakis, G., Bleich, S., Frieling, H. & Deuschle, M. (2011) Epigenetic mechanisms in major depression. Nervenarzt 26:126.Google Scholar
Paulsson, J. (2005) Models of stochastic gene expression. Physics of Life Reviews 2(2):157–75.Google Scholar
Pavelka, N., Rancati, G., Zhu, J., Bradford, W. D., Saraf, A., Florens, L. & Li, R. (2010) Aneuploidy confers quantitative proteome changes and phenotypic variation in budding yeast. Nature 468(7321):321–25.Google Scholar
Pedersen, P. E. & Blass, E. M. (1982) Prenatal and postnatal determinants of the 1st suckling episode in Albino-Rata. Developmental Psychobiology 15(4):349–55.Google Scholar
Pederson, D. R., Gleason, K. E., Moran, G. & Bento, S. (1998) Maternal attachment representations, maternal sensitivity and the infant-mother attachment relationship. Developmental Psychology 34(5):925–33.Google Scholar
Penn, D. & Potts, W. (1998) MHC-disassortative mating preferences reversed by cross-fostering. Proceedings of the Royal Society B Biological Sciences 265(1403):1299–306.Google Scholar
Pentinat, T., Ramon-Krauel, M., Cebria, J., Diaz, R. & Jimenez-Chillaron, J. C. (2010) Transgenerational inheritance of glucose intolerance in a mouse model of neonatal overnutrition. Endocrinology 151(12):5617–23.Google Scholar
Petronis, A. (2010) Epigenetics as a unifying principle in the aetiology of complex traits and diseases. Nature 465(7299):721–27.Google Scholar
Petterson, B., Nelson, K. B., Watson, L. & Stanley, F. (1993) Twins, triplets and cerebral palsy in births in Western Australia in the 1980s. British Medical Journal 307(6914):1239–43.Google Scholar
Pfeiffer, H., Lutz-Bonengel, S., Pollak, S., Fimmers, R., Baur, M. P. & Brinkmann, B. (2004) Mitochondrial DNA control region diversity in hairs and body fluids of monozygotic triplets. International Journal of Legal Medicine 118(2):7174.Google Scholar
Pigliucci, M. (2010) Phenotypic plasticity. In: Evolution: The extended synthesis, ed. Pigliucci, M. & Muller, G. B., pp. 355–78. MIT Press.Google Scholar
Pigliucci, M. & Murren, C. J. (2003) Perspective: Genetic assimilation and a possible evolutionary paradox: Can macroevolution sometimes be so fast as to pass us by? Evolution 57(7):1455–64.Google Scholar
Pigliucci, M., Murren, C. J. & Schlichting, C. D. (2006) Phenotypic plasticity and evolution by genetic assimilation. Journal of Experimental Biology 209(12):2362–67.Google Scholar
Pinto, D., Pagnamenta, A. T., Klei, L., Anney, R., Merico, D., Regan, R. & Betancur, C. (2010) Functional impact of global rare copy number variation in autism spectrum disorders. Nature 466(7304):368–72.Google Scholar
Piotrowski, A., Bruder, C. E., Andersson, R., Diaz de Stahl, T., Menzel, U., Sandgren, J., Poplawski, A., von Tell, D., Crasto, C., Bogdan, A., Bartoszewski, R., Bebok, Z., Krzyzanowski, M., Jankowski, Z., Partridge, E. C., Komorowski, J. & Dumanski, J. P. (2008) Somatic mosaicism for copy number variation in differentiated human tissues. Human Mutation 29:1118–24.Google Scholar
Plomin, R. (2004) Nature and nurture: An introduction to human behavioral genetics. Thomson.Google Scholar
Plomin, R. (2011) Why are children in the same family so different from one another? Non-shared environment three decades later. International Journal of Epidemiology 40(3):563–82.Google Scholar
Plomin, R. & Daniels, D. (1987) Why are children in the same family so different? Behavioral and Brain Sciences 10:116.Google Scholar
Plomin, R. & Davis, O. S. P. (2009) The future of genetics in psychology and psychiatry: Microarrays, genome-wide association & non-coding RNA. Journal of Child Psychology & Psychiatry 50(1/2):6371.Google Scholar
Plomin, R., DeFries, J. C. & Loehlin, J. C. (1977) Genotype-environment interaction and correlation in the analysis of human behavior. Psychological Bulletin 84(2):309–22.Google Scholar
Potts, W. K. & Wakeland, E. K. (1993) Evolution of MHC genetic diversity: A tale of incest, pestilence and sexual preference. Trends in Genetics 9(12):408–12.Google Scholar
Poulsen, P., Esteller, M., Vaag, A. & Frage, M. F. (2007) The epigenetic basis of twin discordance in age-related diseases. Pediatric Research 61(5, Part 2):38R42R.Google Scholar
Puri, D., Dhawan, J. & Mishra, R. K. (2010) The paternal hidden agenda: Epigenetic inheritance through sperm chromatin. Epigenetics 5:5.Google Scholar
Qvarnstrom, A. & Price, T. D. (2001) Maternal effects, paternal effects and sexual selection. Trends in Ecology & Evolution 16(2):95100.Google Scholar
Räikkönen, K., Seckl, J. R., Pesonen, A. K., Simons, A. & Van den Bergh, B. R. (2011) Stress, glucocorticoids and liquorice in human pregnancy: Programmers of the offspring brain. Stress 14(6):590603.Google Scholar
Raj, A. & van Oudenaarden, A. (2008) Nature, nurture, or chance: Stochastic gene expression and its consequences. Cell 135(2):216–26.Google Scholar
Rebollo, R., Horard, B., Hubert, B. & Vieira, C. (2010) Jumping genes and epigenetics: Towards new species. Gene 454(1–2):17.Google Scholar
Redden, D. T. & Allison, D. B. (2003) Nonreplication in genetic association studies of obesity and diabetes research. Journal of Nutrition 133(11):3323–26.Google Scholar
Redon, R., Ishikawa, S., Fitch, K. R., Feuk, L., Perry, G. H., Daniel Andrews, T., & Hurles, M. E. (2006) Global variation in copy number in the human genome. Nature 444(7118):444–54.Google Scholar
Rehen, S. K., Yung, Y. C., McCreight, M. P., Kaushal, D., Yang, A. H., Almeida, B. S. V., Kingsbury Cabral, K. M. S., McConnell, M. J., Anliker, B., Fontanoz, M., Chun, J. (2005) Constitutional aneuploidy in the normal human brain. Journal of Neuroscience 25:2176–80.Google Scholar
Reinders, J., Wulff, B. B., Mirouze, M., Mari-Ordonez, A., Dapp, M., Rozhon, W. & Paszkowski, J. (2009) Compromised stability of DNA methylation and transposon immobilization in mosaic Arabidopsis epigenomes. Genes & Development 23(8):939–50.Google Scholar
Reyes-Castro, L. A., Rodriguez, J. S., Rodriguez-Gonzalez, G. L., Wimmer, R. D., McDonald, T. J., Larrea, F. & Zambrano, E. (2011) Pre- and/or postnatal protein restriction in rats impairs learning and motivation in male offspring. International Journal of Developmental Neuroscience 29(2):177–82.Google Scholar
Reynolds, B. A. & Weiss, S. (1992) Generation of neurons and astrocytes from isolated cells of the adult mammalian central nervous system. Science 255(5052):1707–10.Google Scholar
Richards, E. J. (2009) Quantitative epigenetics: DNA sequence variation need not apply. Genes & Development 23(14):1601–605.Google Scholar
Richmond, J. (1990) Low-birth-weight infants. JAMA 263(22):3069–70.Google Scholar
Richmond, T. J. & Davey, C. A. (2003) The structure of DNA in the nucleosome core. Nature 423(6936):145–50.Google Scholar
Robbins, W. A., Elashoff, D. A., Xun, L., Jia, J., Li, N., Wu, G. & Wei, F. (2005) Effect of lifestyle exposures on sperm aneuploidy. Cytogenetic and Genome Research 111(3–4):371–77.Google Scholar
Rodríguez-Santiago, B., Malats, N., Rothman, N., Armengol, L., Garcia-Closas, M., Kogevinas, M. & Perez-Jurado, L. A. (2010) Mosaic uniparental disomies and aneuploidies as large structural variants of the human genome. American Journal of Human Genetics 87(1):129–38.Google Scholar
Rogers, J. G., Voullaire, L. & Gold, H. (1982) Monozygotic twins discordant for trisomy 21. American Journal of Medical Genetics 11(2):143–46.Google Scholar
Rollins, B., Martin, M. V., Sequeira, P. A., Moon, E. A., Morgan, L. Z., Watson, S. J. & Vawter, M. P. (2009) Mitochondrial variants in schizophrenia, bipolar disorder and major depressive disorder. PLoS ONE. 4(3):e4913. Epub 2009 Mar 17.Google Scholar
Roth, T. L., Roth, E. D. & Sweatt, J. D. (2010) Epigenetic regulation of genes in learning and memory. Essays in Biochemistry 48(1):263–74.Google Scholar
Rowold, D. J. & Herrera, R. J. (2000) Alu elements and the human genome. Genetica 108(1):5772.Google Scholar
Roy-Matton, N., Moutquin, J. M., Brown, C., Carrier, N. & Bell, L. (2011) The impact of perceived maternal stress and other psychosocial risk factors on pregnancy complications. Canadian Journal of Obstetrics and Gynaecology 33(4):344–52.Google Scholar
Rutter, M. & Mawhood, L. (1991) The long-term psychological sequelae of specific developmental disorders of speech and language. In: Biological risks for psychosocial disorders, ed. Rutter, M. & Caesar, P., pp. 233–59. Cambridge University Press.Google Scholar
Rutter, M., Thorpe, K., Greenwood, R., Northstone, K. & Golding, J. (2003) Twins as a natural experiment to study the causes of mild language delay: I: Design; twin-singleton differences in language and obstetric risks. Journal of Child Psychology and Psychiatry 44(3):326–41.Google Scholar
Ryan, B. C. & Vandenbergh, J. G. (2002) Intrauterine position effects. Neuroscience & Biobehavioral Reviews 26(6):665–78.Google Scholar
Sabol, S. Z., Hu, S. & Hamer, D. (1998) A functional polymorphism in the monoamine oxidase A gene promoter. Human Genetics 103(3):273–79.Google Scholar
Sakhai, S. A., Kriegsfeld, L. J. & Francis, D. D. (2011) Maternal programming of sexual attractivity in female Long Evans rats. Psychoneuroendocrinology 36(8):1217–25.Google Scholar
Salvioli, S., Bonafe, M., Capri, M., Monti, D. & Franceschi, C. (2001) Mitochondria, aging and longevity: A new perspective. FEBS Letters 492(1–2):913.Google Scholar
Sandman, C. A., Davis, E. P., Buss, C. & Glynn, L. M. (2011) Exposure to prenatal psychobiological stress exerts programming influences on the mother and her fetus. Neuroendocrinology 15:15.Google Scholar
Sapolsky, R. M., Michael Romero, L. & Munck, A. U. (2000) How do glucocorticoids influence stress responses? Integrating permissive, suppressive, stimulatory, and preparative actions. Endocrine Reviews 21(1):5589.Google Scholar
Saxbe, D. E. & Repetti, R. L. (2009) Brief report: Fathers' and mothers' marital relationship predicts daughters' pubertal development two years later. Journal of Adolescence 32(2):415–23.Google Scholar
Scarr, S. (1995) Commentary on Gottlieb's “Some conceptual deficiencies in ‘developmental’ behavioral genetics.” Human Development 38(3):154–58.Google Scholar
Scarr, S. (1997) Behavior–genetic and socialization theories of intelligence: Truce and reconciliation. In: Intelligence, heredity & environment, ed. Sternberg, E., pp. 341. Cambridge University Press.Google Scholar
Schaal, B., Marlier, L. & Soussignan, R. (2000) Human foetuses learn odours from their pregnant mother's diet. Chemical Senses 25(6):729–37.Google Scholar
Schaal, B., Orgeur, P. & Arnould, C. (1995) Olfactory preferences in newborn lambs: Possible influence of prenatal experience. Behaviour 132:351–65.Google Scholar
Scher, A. I., Petterson, B., Blair, E., Ellenberg, J. H., Grether, J. K., Haan, E. & Nelson, K. B. (2002) The risk of mortality or cerebral palsy in twins: A collaborative population-based study. Pediatric Research 52(5):671–81.Google Scholar
Schlichting, C. D. & Smith, H. (2002) Phenotypic plasticity linking molecular mechanisms with evolutionary outcomes. Evolutionary Ecology 16(3):189211.Google Scholar
Schnable, P. S., Ware, D., Fulton, R. S., Stein, J. C., Wei, F., Pasternak, S. & Wilson, R. K. (2009) The B73 maize genome: Complexity, diversity, and dynamics. Science 326(5956):1112–15.Google Scholar
Schneider, M. L. (1992) Prenatal stress exposure alters postnatal behavioral expression under conditions of novelty challenge in rhesus monkey infants. Developmental Psychobiology 25(7):529–40.Google Scholar
Schneider, M. L., Roughton, E. C., Koehler, A. J. & Lubach, G. R. (1999) Growth and development following prenatal stress exposure in primates: An examination of ontogenetic vulnerability. Child Development 70(2):263–74.Google Scholar
Sciamanna, I., Vitullo, P., Curatolo, A. & Spadafora, C. (2009) Retrotransposons, reverse transcriptase and the genesis of new genetic information. Gene 448(2):180–86.Google Scholar
Sebat, J., Lakshmi, B., Malhotra, D., Troge, J., Lese-Martin, C., Walsh, T. & Wigler, M. (2007) Strong association of de novo copy number mutations with autism. Science 316(5823):445–49.Google Scholar
Sebat, J., Lakshmi, B., Troge, J., Alexander, J. & Young, J. (2004) Large-scale copy number polymorphism in the human genome. Science 305:525.Google Scholar
Sebire, N. J., Snijders, R. J., Hughes, K., Sepulveda, W. & Nicolaides, K. H. (1997) The hidden mortality of monochorionic twin pregnancies. British Journal of Obstetrics and Gynaecology 104(10):1203–207.Google Scholar
Seckl, J. R. (1998) Physiologic programming of the fetus. Clinical Perinatology 25:939–64.Google Scholar
Seki, Y., Hayashi, K., Itoh, K., Mizugaki, M., Saitou, M. & Matsui, Y. (2005) Extensive and orderly reprogramming of genome-wide chromatin modifications associated with specification and early development of germ cells in mice. Developmental Biology 278(2):440–58.Google Scholar
Semke, E., Distel, H. & Hudson, R. (1995) Specific enhancement of olfactory receptor sensitivity associated with foetal learning of food odors in the rabbit. Naturwissenschaften 82(3):148–49.Google Scholar
Senoo, M., Okamura, K., Murotsuki, J., Yaegashi, N., Uehara, S. & Yajima, A. (2000) Growth pattern of twins of different chorionicity evaluated by sonographic biometry. Obstetrics and Gynecology 95(5):656–61.Google Scholar
Sgaramella, V. & Astolfi, P. A. (2010) Somatic genome variations interact with environment, genome and epigenome in the determination of the phenotype: A paradigm shift in genomics? DNA Repair 9(4):470–73.Google Scholar
Shahrezaei, V. & Swain, P. S. (2008) Analytical distributions for stochastic gene expression. Proceedings of the National Academy of Science USA 105(45):17256–61.Google Scholar
Shapiro, J. A. (2009) Revisiting the central dogma in the 21st century. Annals of the New York Academy of Sciences (Natural Genetic Engineering and Natural Genome Editing) 1178:628.Google Scholar
Shen, H., Liu, Y., Liu, P., Recker, R. R. & Deng, H.-W. (2005) Nonreplication in genetic studies of complex diseases—lessons learned from studies of osteoporosis and tentative remedies. Journal of Bone and Mineral Research 20(3):365–76.Google Scholar
Shi, Q., Ko, E., Barclay, L., Hoang, T., Rademaker, A. & Martin, R. (2001) Cigarette smoking and aneuploidy in human sperm. Molecular Reproduction and Development 59(4):417–21.Google Scholar
Shinohara, T. (2001) Mice containing a human chromosome 21 model behavioral impairment and cardiac anomalies of Down's syndrome. Human Molecular Genetics 10:1163–75.Google Scholar
Shur, N. (2009) The genetics of twinning: From splitting eggs to breaking paradigms. American Journal of Medical Genetics Part C: Seminars in Medical Genetics 151C(2):105109.Google Scholar
Sibai, B. M., Hauth, J., Caritis, S., Lindheimer, M. D., MacPherson, C., Klebanoff, M. & McNellis, D. (2000) Hypertensive disorders in twin versus singleton gestations. National Institute of Child Health and Human Development Network of Maternal-Fetal Medicine Units. American Journal of Obstetrics & Gynecology 182(4):938–42.Google Scholar
Siddiqui, F. & McEwan, A. (2007) Twins. Obstetrics, Gynecology, and Reproductive Medicine 17(10):289–95.Google Scholar
Simitzis, P. E., Deligeorgis, S. G., Bizelis, J. A. & Fegeros, K. (2008) Feeding preferences in lambs influenced by prenatal flavour exposure. Physiology & Behavior 93(3):529–36.Google Scholar
Singer, T., McConnell, M. J., Marchetto, M. C. N., Coufal, N. G. & Gage, F. H. (2010) LINE-1 retrotransposons: Mediators of somatic variation in neuronal genomes? Trends in Neurosciences. 33(8):345–54.Google Scholar
Skinner, M. K. (2008) What is an epigenetic transgenerational phenotype?: F3 or F2. Reproductive Toxicology 25(1):26.Google Scholar
Skinner, M. K., Anway, M. D., Savenkova, M. I., Gore, A. C. & Crews, D. (2008) Transgenerational epigenetic programming of the brain transcriptome and anxiety behavior. PLoS ONE 3:e3745.Google Scholar
Skinner, M. K., Manikkam, M. & Guerrero-Bosagna, C. (2010) Epigenetic transgenerational actions of environmental factors in disease etiology. Trends in Endocrinology & Metabolism 21(4):214–22.Google Scholar
Slatkin, M. (2009) Epigenetic inheritance and the missing heritability problem. Genetics 182(3):845–50.Google Scholar
Smith, N. A., Lyons, J. G. & McElrath, T. F. (2010) Protein:creatinine ratio in uncomplicated twin pregnancy. American Journal of Obstetrics & Gynecology 203(4):9.Google Scholar
Smotherman, W. P. (1982) Odor aversion learning by the rat fetus. Physiology & Behavior 29(5):769–71.Google Scholar
Smythe, J. W., Rowe, W. B. & Meaney, M. J. (1994) Neonatal handling alters serotonin(5-HT) turnover and 5-HT2 receptor binding in selected brain regions: Relationship to the handling effect on glucocorticoid receptor expression. Developmental Brain Research 80(1–2):183–89.Google Scholar
Song, L., Zheng, J., Li, H., Jia, N., Suo, Z., Cai, Q. & Zhu, Z. (2009) Prenatal stress causes oxidative damage to mitochondrial DNA in hippocampus of offspring rats. Neurochemical Research 34(4):739–45.Google Scholar
Srividhya, J., Li, Y. & Pomerening, J. R. (2011) Open cascades as simple solutions to providing ultrasensitivity and adaptation in cellular signaling. Physiology & Behavior 8(4):12.Google Scholar
Stefansson, H., Rujescu, D., Cichon, S., Pietilainen, O. P. H., Ingason, A., Steinberg, S. & Stefansson, K. (2008) Large recurrent microdeletions associated with schizophrenia. Nature 455(7210):232–36.Google Scholar
Storm, J. J. & Lima, S. L. (2010) Mothers forewarn offspring about predators: A transgenerational maternal effect on behavior. American Naturalist 175(3):382–90.Google Scholar
Stouder, C. & Paoloni-Giacobino, A. (2010) Transgenerational effects of the endocrine disruptor vinclozolin on the methylation pattern of imprinted genes in the mouse sperm. Reproduction 139:373–79.Google Scholar
Stranger, B. E. (2007) Relative impact of nucleotide and copy number variation on gene expression phenotypes. Science 315:848–53.Google Scholar
Sultan, C., Balaguer, P., Terouanne, B., Georget, V., Paris, F., Jeandel, C. & Nicolas, J.-C. (2001) Environmental xenoestrogens, antiandrogens and disorders of male sexual differentiation. Molecular and Cellular Endocrinology 178(1–2):99105.Google Scholar
Sutcliffe, A. G. & Derom, C. (2006) Follow-up of twins: Health, behaviour, speech, language outcomes and implications for parents. Early Human Development 82(6):379–86.Google Scholar
Suzuki, Y. & Nijhout, H. F. (2008) Genetic basis of adaptive evolution of a polyphenism by genetic accommodation. Journal of Evolutionary Biology 21(1):5766.Google Scholar
Szyf, M. (2009) The early life environment and the epigenome. Biochimica et Biophysica Acta(BBA) – General Subjects 1790(9):878–85.Google Scholar
Tabery, J. (2007) Biometric and developmental gene-environment interactions: Looking back, moving forward. Development and Psychopathology 19(4):961–76.Google Scholar
Tabery, J. & Griffiths, P. E. (2010) Historical and philosophical perspectives on behavioral genetics and developmental science. In: Handbook of developmental science, behavior and genetics, ed. Hood, K. E., Halpern, C. T., Greenberg, G. & Lerner, R. M., pp. 4160. Wiley-Blackwell.Google Scholar
Tamashiro, K. L. K. & Moran, T. H. (2010) Perinatal environment and its influences on metabolic programming of offspring. Physiology & Behavior 100(5):560–66.Google Scholar
Tanaka, Y., Yamashita, R., Suzuki, Y. & Nakai, K. (2010) Effects of Alu elements on global nucleosome positioning in the human genome. BMC Genomics 11:309.Google Scholar
Taylor, H. G., Espy, K. A. & Anderson, P. J. (2009) Mathematics deficiencies in children with very low birth weight or very preterm birth. Developmental Disabilities Research Reviews 15(1):5259.Google Scholar
Taylor, R. W. & Turnbull, D. M. (2005) Mitochondrial DNA mutations in human disease. Nature Reviews Genetics. 6(5):389402.Google Scholar
Temin, H. M. (1964) The participation of DNA in Rofs sarcoma virus production. Virology 23(4):486–94.Google Scholar
Tempest, H. G., Ko, E., Rademaker, A., Chan, P., Robaire, B. & Martin, R. H. (2009) Intra-individual and inter-individual variations in sperm aneuploidy frequencies in normal men. Fertility and Sterility 91(1):185–92.Google Scholar
Templado, C., Vidal, F. & Estop, A. (2011) Aneuploidy in human spermatozoa. Cytogenetic and Genome Research 133(2–4):9199.Google Scholar
Theise, N. D. & Harris, R. (2006) Postmodern biology: (Adult)( stem) cells are plastic, stochastic, complex, and uncertain. Handbook of Experimental Pharmacology 174:389408.Google Scholar
Thieffry, D. & Sanchez, L. (2002) Alternative epigenetic states understood in terms of specific regulatory structures. Annals of the New York Academy of Science 981:135–53.Google Scholar
Thompson, W. R. (1957) Influence of prenatal maternal anxiety on emotionality in young rats. Science 125:698.Google Scholar
Thorpe, K. (2006) Twin children's language development. Early Human Development 82(6):387–95.Google Scholar
Trzctńska, M., Tonkiss, J. & Galler, J. R. (1999) Influence of prenatal protein malnutrition on behavioral reactivity to stress in adult rats. Stress 3(1):7183.Google Scholar
Tsigos, C. & Chrousos, G. P. (2002) Hypothalamic-pituitary-adrenal axis, neuroendocrine factors and stress. Journal of Psychosomatic Research 53(4):865–71.Google Scholar
Tu, H.-P., Ko, A., Wang, S.-J., Lee, C.-H., Lea, R., Chiang, S.-L. & Ko, Y.-C. (2010) Monoamine oxidase A gene polymorphisms and enzyme activity associated with risk of gout in Taiwan aborigines. Human Genetics 127(2):223–29.Google Scholar
Uchida, I. A., deSa, D. J. & Whelan, D. T. (1983) 45,X/46,XX mosaicism in discordant monozygotic twins. Pediatrics 71(3):413–17.Google Scholar
Uller, T. (2008) Developmental plasticity and the evolution of parental effects. Trends in Ecology & Evolution 23(8):432–38.Google Scholar
Umur, A., Van Gemert, M. J. & Ross, M. G. (2001) Amniotic fluid and hemodynamic model in monochorionic twin pregnancies and twin-twin transfusion syndrome. American Journal of Physiology Regulatory, Integrative and Comparative Physiology 280(5):R1499–509.Google Scholar
Uzumcu, M., Suzuki, H. & Skinner, M. K. (2008) Effect of the anti-androgenic endocrine disruptor vinclozolin on embryonic testis cord formation and postnatal testis development and function. Reproductive Toxicology 18(6):765–74.Google Scholar
van Oudenaarden, A. (2009) Nature, nurture or just blind chance: Stochastic gene expression and its consequences. Biophysical Journal 96(3, Suppl. 1):15a.Google Scholar
von Engelhardt, N., Carere, C., Dijkstra, C. & Groothuis, T. G. G. (2006) Sex-specific effects of yolk testosterone on survival, begging and growth of zebra finches. Proceedings of the Royal Society B: Biological Sciences 273(1582):6570.Google Scholar
Vucetic, Z., Kimmel, J., Totoki, K., Hollenbeck, E. & Reyes, T. M. (2010) Maternal high-fat diet al.ters methylation and gene expression of dopamine and opioid-related genes. Endocrinology 151(10):4756–64.Google Scholar
Wade, M. J. (1998) The evolutionary genetics of maternal Effects. In: Maternal effects as adaptations, ed. Mousseau, T. A. & Fox, C. W., pp. 521. Oxford University Press.Google Scholar
Wadhwa, P. D., Sandman, C. A., Porto, M., Dunkel Schetter, C. & Garite, T. J. (1993) The association between prenatal stress and infant birth weight and gestational age at birth: A prospective investigation. American Journal of Obstetrics & Gynecology 169(4):858–65.Google Scholar
Walker, C. D. & McCormick, C. M. (2009) Development of the stress axis: Maternal and environmental influences. In: Hormones, brain and behavior, ed. Donald, W. P., Arthur, P. A., Susan, E. F., Anne, M. E. & Robert, T. R., pp. 1931–73. Academic Press.Google Scholar
Walker, D. M. & Gore, A. C. (2011) Transgenerational neuroendocrine disruption of reproduction. Nature Reviews Endocrinology 7(4):197207.Google Scholar
Walsh, T., McClellan, J. M., McCarthy, S. E., Addington, A. M., Pierce, S. B., Cooper, G. M. & Sebat, J. (2008) Rare structural variants disrupt multiple genes in neurodevelopmental pathways in schizophrenia. Science 320(5875):539–43.Google Scholar
Wang, H., Xing, J., Grover, D., Hedges, D. J., Han, K., Walker, J. A. & Batzer, M. A. (2005) SVA elements: A hominid-specific retroposon family. Journal of Molecular Biology 354(4):9941007.Google Scholar
Waterland, R. A. & Jirtle, R. L. (2003) Transposable elements: Targets for early nutritional effects on epigenetic gene regulation. Molecular and Cellular Biology 23(15):5293–300.Google Scholar
Waterland, R. A. & Michels, K. B. (2007) Epigenetic epidemiology of the developmental origins hypothesis. Annual Review of Nutrition 27(1):363–88.Google Scholar
Watkins, A. J., Wilkins, A., Cunningham, C., Perry, V. H., Seet, M. J., Osmond, C. & Fleming, T. P. (2008) Low protein diet fed exclusively during mouse oocyte maturation leads to behavioural and cardiovascular abnormalities in offspring. Journal of Physiology 586(8):2231–44.Google Scholar
Weatherall, D. J. (2000) Science, medicine & the future: Single gene disorders or complex traits: Lessons from the thalassaemias and other monogenic diseases. British Medical Journal 321(7269):1117–20.Google Scholar
Weaver, I. (2009) Shaping adult phenotypes through early life environments. Birth Defects Research Part C: Embryo Today: Reviews 87(4):314–26.Google Scholar
Weaver, I. C. G., D'Alessio, A. C., Brown, S. E., Hellstrom, I. C., Dymov, S., Sharma, S. & Meaney, M. J. (2007) The transcription factor nerve growth factor-inducible protein a mediates epigenetic programming: Altering epigenetic marks by immediate-early genes. Journal of Neuroscience 27(7):1756–68.Google Scholar
Weaver, I., Cervoni, N., Champagne, F. A., D'Alessio, A. C., Sharma, S. & Seckl, J. R. (2004) Epigenetic programming by maternal behavior. Nature Neuroscience 7:847–54.Google Scholar
Weaver, I., La Plante, P., Weaver, S., Parent, A., Sharma, S., Diorio, J. & Meaney, M. J. (2001) Early environmental regulation of hippocampal glucocorticoid receptor gene expression: Characterization of intracellular mediators and potential genomic target sites. Molecular and Cellular Endocrinology 185(1–2):205–18.Google Scholar
Weaver, I., Meaney, M. J. & Szyf, M. (2006) Maternal care effects on the hippocampal transcriptome and anxiety-mediated behaviors in the offspring that are reversible in adulthood. Proceedings of the National Academy of Science USA 103:3480.Google Scholar
Weber, J. L., David, D., Heil, J., Fan, Y., Zhao, C. F. & Marth, G. (2002) Human diallelic insertion/deletion polymorphisms. American Journal of Human Genetics 71(4):854–62.Google Scholar
Weber, M. A. & Sebire, N. J. (2010) Genetics and developmental pathology of twinning. Seminars in Fetal and Neonatal Medicine 15(6):313–18.Google Scholar
Wedekind, C. & Furi, S. (1997) Body odour preferences in men and women: Do they aim for specific MHC combinations or simply heterozygosity? Proceedings of the Royal Society B Biological Sciences 264(1387):1471–79.Google Scholar
Weinstock, M. (2008) The long-term behavioural consequences of prenatal stress. Neuroscience & Biobehavioral Reviews 32(6):1073–86.Google Scholar
Weiss, E., Loevy, H., Saunders, A., Pruzansky, S., Rosenthal, I. M. & Opitz, J. M. (1982) Monozygotic twins discordant for Ullrich-Turner syndrome. American Journal of Medical Genetics 13(4):389–99.Google Scholar
Welberg, L. A. M. & Seckl, J. R. (2001) Prenatal stress, glucocorticoids and the programming of the brain. Journal of Neuroendocrinology 13(2):113–28.Google Scholar
Wells, J. C. & Stock, J. T. (2011) Re-examining heritability: Genetics, life history and plasticity. Trends in Endocrinology & Metabolism 22(10):421–28.Google Scholar
West, M. J. & King, A. P. (1987) Settling nature and nurture into an ontogenetic niche. Developmental Psychobiology 20(5):549–62.Google Scholar
Westberry, J. M., Trout, A. L. & Wilson, M. E. (2010) Epigenetic regulation of estrogen receptor alpha gene expression in the mouse cortex during early postnatal development. Endocrinology 151:731–40.Google Scholar
West-Eberhard, M. J. (2003) Developmental plasticity and evolution. Oxford University Press.Google Scholar
Westra, J. W., Rivera, R. R., Bushman, D. M., Yung, Y. C., Peterson, S. E., Barral, S. & Chun, J. (2010) Neuronal DNA content variation (DCV) with regional and individual differences in the human brain. Journal of Comparative Neurology 518(19):39814000.Google Scholar
Wheeler, D. A., Srinivasan, M., Egholm, M., Shen, Y., Chen, L., McGuire, A. & Rothberg, J. M. (2008) The complete genome of an individual by massively parallel DNA sequencing. Nature 452(7189):872–76.Google Scholar
Whitelaw, N. C. & Whitelaw, E. (2006) How lifetimes shape epigenotype within and across generations. Human Molecular Genetics 15:R131.Google Scholar
Wiedmeier, J. E., Joss-Moore, L. A., Lane, R. H. & Neu, J. (2011) Early postnatal nutrition and programming of the preterm neonate. Nutrition Reviews 69(2):7682.Google Scholar
Wolff, G. L., Kodell, R. L., Moore, S. R. & Cooney, C. A. (1998) Maternal epigenetics and methyl supplements affect agouti gene expression in Avy/a mice. FASEB Journal 12(11):949–57.Google Scholar
Wong, C. C. Y., Caspi, A., Williams, B., Craig, I. W., Houts, R., Ambler, A., Moffit, T. E. & Mill, J. (2010) A longitudinal study of epigenetic variation in twins. Epigenetics 5(6):515–26.Google Scholar
Xing, J., Zhang, Y., Han, K., Salem, A. H., Sen, S. K., Huff, C. D. & Jorde, L. B. (2009) Mobile elements create structural variation: Analysis of a complete human genome. Genome Research.Google Scholar
Yang, J. A., Benyamin, B., McEvoy, B. P., Gordon, S., Henders, A. K., Nyholt, D. R. & Visscher, P. M. (2010) Common SNPs explain a large proportion of the heritability for human height. Nature Genetics 42(7):565–69.Google Scholar
Yeo, G., Holste, D., Kreiman, G. & Burge, C. B. (2004) Variation in alternative splicing across human tissues. Genome Biology 5:R74.Google Scholar
Yu, L. R., Conrads, T. P., Uo, T., Kinoshita, Y., Morrison, R. S., Lucas, D. A. & Veenstra, T. D. (2004) Global analysis of the cortical neuron proteome. Molecular Cell Proteomics 3(9):896907.Google Scholar
Yurov, Y. B., Iourov, I. Y., Vorsanova, S. G., Demidova, I. A., Kravetz, V. S., Beresheva, A. K. & Liehr, T. (2008) The schizophrenia brain exhibits low-level aneuploidy involving chromosome 1. Schizophrenia Research 98:137–47.Google Scholar
Zhang, F., Gu, W., Hurles, M. E. & Lupski, J. R. (2009) Copy number variation in human health, disease, and evolution. Annual Review of Genomics and Human Genetics 10(1):451–81.Google Scholar
Zhang, G.-H., Chen, M.-L., Liu, S.-S., Zhan, Y.-H., Quan, Y., Qin, Y.-M. & Deng, S.-P. (2011) Effects of mother's dietary exposure to acesulfame-k in pregnancy or lactation on the adult offspring's sweet preference. Chemical Senses 36(9):763–70.Google Scholar
Zhang, T. Y., Bagot, R., Parent, C., Nesbitt, C., Bredy, T. W., Caldji, C. & Meaney, M. J. (2006) Maternal programming of defensive responses through sustained effects on gene expression. Biological Psychology 73(1):7289.Google Scholar
Zhang, X. H. F. & Chasin, L. A. (2006) Comparison of multiple vertebrate genomes reveals the birth and evolution of human exons. Proceedings of the National Academy of Sciences USA 103(36):13427–32.Google Scholar
Zhao, C., Deng, W. & Gage, F. H. (2008) Mechanisms and functional implications of adult neurogenesis. Cell 132:645–60.Google Scholar
Zondervan, K. T. & Cardon, L. R. (2004) The complex interplay among factors that influence allelic association. Nature Reviews Genetics 5(2):89100.Google Scholar
Zuloaga, D. G., Puts, D. A., Jordan, C. L. & Breedlove, S. M. (2008) The role of androgen receptors in the masculinization of brain and behavior: What we've learned from the testicular feminization mutation. Hormones and Behavior 53(5):613–26.Google Scholar