Hostname: page-component-7c8c6479df-995ml Total loading time: 0 Render date: 2024-03-28T14:37:47.649Z Has data issue: false hasContentIssue false

Differential susceptibility to effects of maternal sensitivity? A study of candidate plasticity genes

Published online by Cambridge University Press:  15 September 2014

Jay Belsky*
Affiliation:
University of California, Davis
Daniel A. Newman
Affiliation:
University of Illinois at Urbana–Champaign
Keith F. Widaman
Affiliation:
University of California, Davis
Phil Rodkin
Affiliation:
University of Illinois at Urbana–Champaign
Michael Pluess
Affiliation:
Kings College London
R. Chris Fraley
Affiliation:
University of Illinois at Urbana–Champaign
Daniel Berry
Affiliation:
University of Illinois at Urbana–Champaign
Jonathan L. Helm
Affiliation:
University of California, Davis
Glenn I. Roisman
Affiliation:
University of Minnesota
*
Address correspondence and reprint requests to: Jay Belsky, Department of Human Ecology, University of California, Davis, 1331 Hart Hall, One Shields Avenue, Davis, CA 95616; E-mail: jbelsky@ucdavis.edu.

Abstract

Here we tested whether there was genetic moderation of effects of early maternal sensitivity on social–emotional and cognitive–linguistic development from early childhood onward and whether any detected Gene × Environment interaction effects proved consistent with differential-susceptibility or diathesis–stress models of Person × Environment interaction (N = 695). Two new approaches for evaluating models were employed with 12 candidate genes. Whereas maternal sensitivity proved to be a consistent predictor of child functioning across the primary-school years, candidate genes did not show many main effects, nor did they tend to interact with maternal sensitivity/insensitivity. These findings suggest that the developmental benefits of early sensitive mothering and the costs of insensitive mothering look more similar than different across genetically different children in the current sample. Although acknowledgement of this result is important, it is equally important that the generally null Gene × Environment results reported here not be overgeneralized to other samples, other predictors, other outcomes, and other candidate genes.

Type
Regular Articles
Copyright
Copyright © Cambridge University Press 2014 

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

Achenbach, T. M. (1991). Manual for the Child Behavior Checklist/4–18 and 1991 Profile. Burlington, VT: University of Vermont, Department of Psychiatry.Google Scholar
Achenbach, T. M., & Edelbrock, C. (1986). Manual for the Teacher's Report Form and Teacher Version of the Child Behavior Profile. Burlington, VT: University of Vermont, Department of Psychiatry.Google Scholar
Achenbach, T. M., Edelbrock, C., & Howell, C. (1987). Empirically-based assessment of the behavioral/ emotional problems of 2-3-year-old children. Journal of Abnormal Child Psychology, 15, 629650.CrossRefGoogle Scholar
Ainsworth, M. D. (1973). The development of infant–mother attachment. In Caldwell, B. M. & Ricciuti, H. N. (Eds.), Review of child development research (Vol. 3, pp. 194). Chicago: University of Chicago Press.Google Scholar
Bakermans-Kranenburg, M. J., & van IJzendoorn, M. H. (2011). Differential susceptibility to rearing environment depending on dopamine-related genes: New evidence and a meta-analysis. Development and Psychopathology, 23, 3952.CrossRefGoogle ScholarPubMed
Bakermans-Kranenburg, M. J., van IJzendoorn, M. H., & Juffer, F. (2003). Less is more: Meta-analysis of sensitivity and attachment interventions in early childhood. Psychological Bulletin, 129, 195215.CrossRefGoogle ScholarPubMed
Baumrind, D. (1967). Child-care practices anteceding three patterns of preschool behavior. Genetic Psychology Monographs, 75, 4388.Google ScholarPubMed
Baumrind, D. (1991). The influence of parenting style on adolescent competence and substance use. Journal of Early Adolescence, 11, 5695.CrossRefGoogle Scholar
Belsky, J. (1997). Variation in susceptibility to environmental influences: An evolutionary argument. Psychological Inquiry, 8, 182186.CrossRefGoogle Scholar
Belsky, J. (2005). Differential susceptibility to rearing influences: An evolutionary hypothesis and some evidence. In Ellis, B. & Bjorklund, D. (Eds.), Origins of the social mind: Evolutionary psychology and child development (pp. 139163). New York: Guilford Press.Google Scholar
Belsky, J., Bakermans-Kranenburg, M. J., & van IJzendoorn, M. H. (2007). For better and for worse: Differential susceptibility to environmental influence. Current Directions in Psychological Science, 16, 300304.CrossRefGoogle Scholar
Belsky, J., Jonassaint, C., Pluess, M., Stanton, M., Brummett, B., & Williams, R. (2009). Vulnerability genes or plasticity genes? Molecular Psychiatry, 14, 746754.CrossRefGoogle ScholarPubMed
Belsky, J., & Pluess, M. (2009). Beyond diathesis–stress: Differential susceptibility to environmental influences. Psychological Bulletin, 135, 885908.CrossRefGoogle ScholarPubMed
Belsky, J., & Pluess, M. (2012). Differential susceptibility to long-term effects of quality of child care on externalizing behavior in adolescence. International Journal of Behavioral Development, 36, 210.CrossRefGoogle Scholar
Belsky, J., & Pluess, M. (2013). Beyond risk, resilience and dysregulation: Phenotypic plasticity and human development. Development and Psychopathology, 25, 12431261.CrossRefGoogle ScholarPubMed
Belsky, J., Pluess, M., & Widaman, K. F. (2013). Confirmatory and competitive evaluation of alternative gene–environment interaction hypotheses. Journal of Child Psychology and Psychiatry, 54, 11351143.CrossRefGoogle ScholarPubMed
Belsky, J., Steinberg, L., & Draper, P. (1991). Childhood experience, interpersonal development and reproductive strategy: An evolutionary theory of socialization. Child Development, 62, 647670.CrossRefGoogle ScholarPubMed
Benjamini, Y., & Hochberg, Y. (1995). Controlling the false discovery rate: A practical and powerful approach to multiple testing. Journal of the Royal Statistical Society, Series B (Methodological), 57, 289300.CrossRefGoogle Scholar
Berry, D., Deater-Deckard, K., McCartney, K., Wang, Z., & Petrill, S. A. (2013). Gene–environment interaction between dopamine receptor D4 7-repeat polymorphism and early maternal sensitivity predicts inattention trajectories across middle childhood. Development and Psychopathology, 25, 291306.CrossRefGoogle ScholarPubMed
Brody, G. H., Beach, S. R. H., Chen, Y., Obasi, E., Philibert, R. A., Kogan, S. M., et al. (2011). Perceived discrimination, serotonin transporter linked polymorphic region status, and the development of conduct problems. Development and Psychopathology, 23, 617627.CrossRefGoogle ScholarPubMed
Carver, C. S., Johnson, S. L., Joormann, J., Kim, Y., & Nam, J. Y. (2011). Serotonin transporter polymorphism interacts with childhood adversity to predict aspects of impulsivity. Psychological Science, 22, 589595.CrossRefGoogle ScholarPubMed
Caspi, A., McClay, J., Moffitt, T., Mill, J., Martin, J., Craig, I., et al. (2002). Role of genotype in the cycle of violence in maltreted children. Science, 297, 851854.CrossRefGoogle ScholarPubMed
Chabris, C. F., Hebert, B. M., Benjamin, D. J., Beauchamp, J., Cesarini, D., van der Loos, M., et al. (2012). Most reported genetic associations with general intelligence are probably false positives. Psychological Science, 23, 13141323.CrossRefGoogle ScholarPubMed
Charney, E., & English, W. (2012). Candidate genes and political behavior. American Political Science Review, 106, 134.CrossRefGoogle Scholar
Clasen, P. C., Wells, T. T., Knopik, V. S., McGeary, J. E., & Beevers, C. G. (2011). 5-HTTLPR and BDNF Val66Met polymorphisms moderate effects of stress on rumination. Genes, Brain and Behavior, 10, 740746.CrossRefGoogle ScholarPubMed
Dearing, E., & Hamilton, L. C. (2006). Contemporary approaches and classic advice for analyzing mediating and moderating variables. Monographs of the Society for Research in Child Development, 71, 88104.Google Scholar
Deary, I. J. (2012). Intelligence. Annual Review of Psychology, 63, 453482.CrossRefGoogle ScholarPubMed
De Wolff, M., & van IJzendoorn, M. (1997). Sensitivity and attachment: A meta-analysis on parental antecedents of infant attachment. Child Development, 68, 571591.Google ScholarPubMed
DeYoung, C. G., Cicchetti, D., & Rogosch, F. A. (2011). Moderation of the association between childhood maltreatment and neuroticism by the corticotropin-releasing hormone receptor 1 gene. Journal of Child Psychology and Psychiatry, 52, 898906.CrossRefGoogle ScholarPubMed
Dressler, W. W., Balieiro, M. C., Ribeiro, R. P., & Dos Santos, J. E. (2009). Cultural consonance, a 5HT2A receptor polymorphism, and depressive symptoms: A longitudinal study of Gene × Culture interaction in urban Brazil. American Journal of Human Biology, 21, 9197.CrossRefGoogle ScholarPubMed
Drury, S. S., Gleason, M. M., Theall, K. P., Smyke, A. T., Nelson, C. A., Fox, N. A., et al. (2012). Genetic sensitivity to the caregiving context: The influence of 5HTTLPR and BDNF val66met on indiscriminate social behavior. Physiology & Behavior, 106, 728735.CrossRefGoogle Scholar
Ducci, F., Enoch, M. A., Hodgkinson, C., Xu, K., Catena, M., Robin, R. W., et al. (2008). Interaction between a functional MAOA locus and childhood sexual abuse predicts alcoholism and antisocial personality disorder in adult women. Molecular Psychiatry, 13, 334347.CrossRefGoogle ScholarPubMed
Duncan, L., & Keller, M. C. (2011). A critical review of the first 10 years of candidate gene-by-environment interaction research in psychiatry. American Journal of Psychiatry, 168, 10411049.CrossRefGoogle ScholarPubMed
Ellis, B. J., Boyce, W. T., Belsky, J., Bakermans-Kranenburg, M. J., & van IJzdenoorn, M. (2011). Differential susceptibility to the environment: An evolutionary–neurodevelopmental theory. Development and Psychopathology, 23, 15.CrossRefGoogle Scholar
Elovainio, M., Jokela, M., Kivimaki, M., Pulkki-Raback, L., Lehtimaki, T., Airla, N., et al. (2007). Genetic variants in the DRD2 gene moderate the relationship between stressful life events and depressive symptoms in adults: Cardiovascular risk in young Finns study. Psychosomatic Medicine, 69, 391395.CrossRefGoogle ScholarPubMed
Elzinger, B. M., Molendijk, M. L., Voshaar, R. C., Bus, B. A., Prickaerts, J., Spinhoven, P., et al. (2011). The impact of child abuse and recent stress on serum brain-derived neurotrophic factor and the moderating role of BDNF Val66Met. Psychopharmacology, 214, 319328.CrossRefGoogle Scholar
Enoch, M. A., Steer, C. D., Newman, T. K., Gibson, N., & Goldman, D. (2010). Early life stress, MAOA, and gene–environment interactions predict behavioral disinhibition in children. Genes, Brain and Behavior, 9, 6574.CrossRefGoogle ScholarPubMed
Fardo, D. W., Becker, K. D., Bertram, L., Tanzi, R. E., & Lange, C. (2009). Recovering unused information in genome-wide association studies: The benefit of analyzing SNPs out of Hardy–Weinberg equilibrium. European Journal of Human Genetics, 17, 16761682.CrossRefGoogle ScholarPubMed
Fortuna, K., van IJzendoorn, M. H., Mankuta, D., Kaitz, M., Avinun, R., Epstein, R. P., et al. (2011) Differential genetic susceptibility to child risk at birth in predicting observed maternal behavior. PLOS ONE, 6, e19765.CrossRefGoogle ScholarPubMed
Fraley, R. C., Roisman, G. I., & Haltigan, J. D. (2013). The legacy of early experiences in development: Formalizing alternative models of how early experiences are carried forward over time. Developmental Psychology, 49, 109126.CrossRefGoogle ScholarPubMed
Frazzetto, G., Di Lorenzo, G., Carola, V., Proietti, L., Sokolowska, E., Siracusano, A., et al. (2007). Early trauma and increased risk for physical aggression during adulthood: The moderating role of MAOA genotype. PLOS ONE, 2, e486.CrossRefGoogle ScholarPubMed
Gatt, J. M., Nemoroff, C. B., Dobson-Stone, C., Paul, R. H., Bryant, R. A., Schofield, P. R., et al. (2009). Interactions between BDNF Val66Met polymorphism and early life stress predict brain and arousal pathways to syndromal depression and anxiety. Molecular Psychiatry, 14, 681695.CrossRefGoogle ScholarPubMed
Gottesman, I. I., & Shields, J. (1967). A polygenic theory of schizophrenia. Proceedings of the National Academy of Sciences, 58, 199205.CrossRefGoogle ScholarPubMed
Gresham, F. M., & Elliott, S. N. (1990). The Social Skills Rating System. Circle Pines, MN: American Guidance Service.Google Scholar
Haltigan, J. D., Roisman, G. I., & Fraley, R. C. (2013). The predictive significance of early caregiving experiences for symptoms of psychopathology through midadolescence: Enduring or transient effects? Development and Psychopathology, 25, 209221.CrossRefGoogle ScholarPubMed
Hankin, B. L., Jenness, J., Abela, J. R., & Smolen, A. (2011). Interaction of 5-HTTLPR and idiographic stressors predicts prospective depressive symptoms specifically among youth in a multi-wave design. Journal of Clinical Child and Adolescent Psychology, 40, 572585.CrossRefGoogle Scholar
Hankin, B., Nederhof, E., Oppenheimer, C. W., Jenness, J., Young, J. F., Abela, J. R. Z., et al. (2011). Differential susceptibility in youth: Evidence that 5HTTLPR × Positive Parenting is associated with positive affect “for better and worse.Translational Psychiatry, 1, e44.CrossRefGoogle ScholarPubMed
Hartl, D. L., & Clark, A. G. (1989). Principles of population genetics (2nd ed.). Sunderland, MA: Sinauer Associates.Google Scholar
Hayden, E. P., Klein, D. N., Dougherty, L. R., Olino, T. M., Dyson, M. W., Durbin, C. E., et al. (in press). The role of brain-derived neurotrophic factor genotype, parental depression, and relationship discord in predicting early-emerging negative emotionality. Psychological Science.Google Scholar
Jaffari-Bimmel, N., Juffer, F., van IJzendoorn, M. H., Bakermans-Kranenburg, M. J., & Mooijaart, A. (2006). Social development from infancy to adolescence: Longitudinal and current factors in an adoption sample. Developmental Psychology, 42, 11431153.CrossRefGoogle Scholar
Jokela, M., Lehtimaki, T., & Keltikangas-Jarvinen, L. (2007a). The influence of urban/rural residency on depressive symptoms is moderated by the serotonin receptor 2A gene. American Journal of Medical Genetics: Part B, Neuropsychiatric Genetics, 144, 918922.Google Scholar
Jokela, M., Lehtimaki, T., & Keltikangas-Jarvinen, L. (2007b). The serotonin receptor 2A gene moderates the influence of parental socioeconomic status on adulthood harm avoidance. Behavior Genetics, 37, 567574.CrossRefGoogle ScholarPubMed
Juhasz, G., Dunham, J., McKie, S., Thomas, E., Downey, D., Chase, D., et al. (2011). The CREDB1–BDNF–NTRK2 pathway in depression. Biological Psychiatry, 69, 762771.CrossRefGoogle ScholarPubMed
Keltikangas-Jarvinen, L., Jokela, M., Hintsanen, M., Salo, J., Hintsa, T., Alatupa, S., et al. (2010). Does genetic background moderate the association between parental education and school achievement? Genes, Brain and Behavior, 9, 318324.CrossRefGoogle ScholarPubMed
Keltikangas-Jarvinen, L., Puttonen, S., Kivimaki, M., Elovainio, M., Rontu, R., & Lehtimaki, T. (2007). Tryptophan hydroxylase 1 gene haplotypes modify the effect of a hostile childhood environment on adulthood harm avoidance. Genes, Brain, and Behavior, 6, 305313.CrossRefGoogle ScholarPubMed
Knafo, A., Israel, S., & Ebstein, R. P. (2011). Heritability of children's prosocial bhavior and differential susceptibility to parenting by variation in the dopamine receptor D4 gene. Development and Psychopathology, 23, 5367.CrossRefGoogle Scholar
Kochanska, G. (2002). Mutually responsive orientation between mothers and their young children: A context for the early development of conscience. Current Directions in Psychological Science, 11, 191195.CrossRefGoogle Scholar
Kochanska, G., Kim, S., Barry, R. A., & Philibert, R. A. (2011). Children's genotypes interact with maternal responsive care in predicting children's competence: Diathesis–stress or differential susceptibility? Development and Psychopathology, 23, 605616.CrossRefGoogle ScholarPubMed
Kuepper, Y., Wielpuetz, C., Alexander, N., Mueller, E., Grant, P., & Hennig, J. (2012). 5-HTTLPR S-Allele: A genetic plasticity factor regarding the effects of life events on personality. Genes, Brain and Behavior, 34, 10701074.Google Scholar
Kuo, Z.-Y. (1967). The dynamics of behavior development: An epigenetic view. New York: Random House.Google Scholar
Laucht, M., Blomeyeer, D., Burchmann, A. F., Treutlein, J., Schmidt, M. H., Esser, G., et al. (2012). Catechol-O-methyltransferase Val158Met genotype, parenting practices and adolescent alcohol use: Testing the differential susceptibility hypothesis. Journal of Child Psychology and Psychiatry, 53, 351359.CrossRefGoogle Scholar
Mata, J., Thompson, R. J., & Gotlib, I. H. (2010). BDNF genotype moderates the relation between physical activity and depressive symptoms. Health Psychology, 29, 130133.CrossRefGoogle ScholarPubMed
Mills-Koonce, W. R., Propper, C. B., Gariepy, J. L., Blair, C., Garrett-Peters, P., & Cox, M. J. (2007). Bidirectional genetic and environmental influences on mother and child behavior: The family system as the unit of analyses. Development and Psychopathology, 19, 10731087.CrossRefGoogle Scholar
Neuman, R. J., Lobos, E., Reich, W., Henderson, C. A., Sun, L. W., & Todd, R. D. (2007). Prenatal smoking exposure and dopaminergic genotypes interact to cause a severe ADHD subtype. Biological Psychiatry, 61, 13201328.CrossRefGoogle Scholar
Newman, D. A. (2003). Longitudinal modeling with randomly and systematically missing data: A simulation of ad hoc, maximum likelihood, and multiple imputation techniques. Organizational Research Methods, 6, 328362.CrossRefGoogle Scholar
NICHD Early Child Care Research Network. (2001). Child-care and family predictors of preschool attachment and stability from infancy. Developmental Psychology, 37, 847862.CrossRefGoogle Scholar
NICHD Early Child Care Research Network. (2005). Child care and child development. New York: Guilford Press.Google Scholar
NICHD Early Child Care Research Network. (2006). Child-care effect sizes for the NICHD Study of Early Child Care and Youth Development. American Psychologist, 61, 99116.CrossRefGoogle Scholar
Nijmeijer, J. S., Hartman, C. A., Rommelse, N. N., Altink, M. E., Buschgens, C. J., Fliers, E. A., et al. (2010). Perinatal risk factors interacting with catechol-O-methyltransferase and the serotonin transporter gene predict ASD symptoms in children with ADHD. Journal of Child Psychology and Psychiatry, 51, 12421250.CrossRefGoogle ScholarPubMed
Obradović, J., & Boyce, W. T. (2009). Individual differences in behavioral, physiological, and genetic sensitivities to contexts: Implications for development and adaptation. Developmental Neuroscience, 31, 300308.CrossRefGoogle ScholarPubMed
Park, A., Sher, K. J., Todorov, A. A., & Heath, A. C. (2011). Interaction between DRD4 VNTR polymorphism and proximal and distal environments in alcohol dependence during emerging and young adulthood. Journal of Abnormal Psychology, 120, 585595.CrossRefGoogle ScholarPubMed
Patterson, G. R. (1982). Coercive family process. Eugene, OR: Castalia.Google Scholar
Patterson, G. R. (1986). Performance models for antisocial boys. American Psychologist, 41, 432444.CrossRefGoogle ScholarPubMed
Plomin, R. (2012). Child development and molecular genetics: 14 years later. Child Development. Advance online publication. doi: 10.1111/j.1467-8624.2012.01757.xCrossRefGoogle Scholar
Plomin, R., Haworth, C. M. A., Meaburn, E. L., Price, T. S., Wellcome Trust Case Control Consortium 2, & Davis, O. S. P. (2013). Common DNA markers can account for more than half of the genetic influence on cognitive abilities. Psychological Science. Advance online publication. doi:10.1177/0956797612457952CrossRefGoogle Scholar
Pluess, M., & Belsky, J. (2009). Differential susceptibility to rearing experience: The case of child-care. Journal of Child Psychology and Psychiatry and Allied Disciplines, 50, 396404.CrossRefGoogle Scholar
Pluess, M., & Belsky, J. (2010). Differential susceptibility to parenting and quality child care. Developmental Psychology, 45, 379390.CrossRefGoogle Scholar
Pluess, M., & Belsky, J. (2012). Vantage sensitivity: Individual differences in response to positive experiences. Psychological Bulletin. Advance online publication.Google Scholar
Propper, C., Mills-Koonce, W. R., Tucker Halpern, C., Hill-Soderlund, A. L., Carbone, M. A., Moore, G. A., et al. (2008). Gene–environment contributions to the development of infant vagal reactivity: The interaction of dopamine and maternal sensitivity. Child Development, 79, 13771394.CrossRefGoogle Scholar
Raymond, M., & Rousset, F. (1995). GENEPOP (Version 1.2): Population genetics software for exact tests and ecumenicism. Journal of Heredity, 86, 248249.CrossRefGoogle Scholar
Roisman, G. I., Booth-LaForce, C., Cauffman, E., Spieker, S., & the NICHD Early Child Care Research Network. (2009). The developmental significance of adolescent romantic relationships: Parent and peer predictors of quality and engagement at age 15. Journal of Youth and Adolescence, 38, 12941303.CrossRefGoogle ScholarPubMed
Roisman, G. I., & Fraley, R. C. (2012). A behavior-genetic study of the legacy of early caregiving experiences: Academic skills, social competence, and externalizing behavior in kindergarten. Child Development, 83, 728742.CrossRefGoogle ScholarPubMed
Roisman, G. I., Newman, D. A., Fraley, R. C., Haltigan, J. D., Groh, A. M., & Haydon, K. C. (2012). Distinguishing differential susceptibility from diathesis–stress: Recommendations for evaluating interaction effects. Development and Psychopathology, 24, 389409.CrossRefGoogle ScholarPubMed
Roisman, G. I., Susman, E., Barnett-Walker, K., Booth-LaForce, C., Owen, M. T., Belsky, J., et al. (2009). Early family and child-care antecedents of awakening cortisol levels in adolescence. Child Development, 80, 907920.CrossRefGoogle ScholarPubMed
Rothbart, M. K., & Bates, J. E. (2006). Temperament. In Eisenberg, N., Damon, W., & Lerner, R. M. (Eds.), Handbook of child psychology: Vol. 3. Social, emotional, and personality development (6th ed., pp. 99166). Hoboken, NJ: Wiley.Google Scholar
Rubin, D. B. (1987). Multiple imputation for nonresponse in surveys. New York: Wiley.CrossRefGoogle Scholar
Rutter, M. (2009). Understanding and testing risk mechanisms for mental disorders. Journal of Child Psychology and Psychiatry, 50, 4452.CrossRefGoogle ScholarPubMed
Salo, J., Jokela, M., Lehtimaki, T., & Keltikangas-Jarvinen, L. (2011). Serotonin receptor 2A gene moderates the effect of childhood maternal nurturance on adulthood social attachment. Genes, Brain and Behavior, 10, 702709.CrossRefGoogle ScholarPubMed
Sameroff, A. J. (1983). Developmental systems: Contexts and evolution. In Mussen, P. (Ed.), Handbook of child psychology (Vol. 1, pp. 237294). New York: Wiley.Google Scholar
Schafer, J. L. (1997). Analysis of incomplete multivariate data. London: Chapman & Hall.CrossRefGoogle Scholar
Schafer, J. L., & Graham, J. W. (2002). Missing data: Our view of the state of the art. Psychological Methods, 7, 147177.CrossRefGoogle ScholarPubMed
Sonuga-Barke, E. J., Oades, R. D., Psychogiou, L., Chen, W., Franke, B., Buitelaar, J., et al. (2009). Dopamine and serotonin transporter genotypes moderate sensitivity to maternal expressed emotion: The case of conduct and emotional problems in attention deficit/hyperactivity disorder. Journal of Child Psychology and Psychiatry and Allied Disciplines, 50, 10521063.CrossRefGoogle ScholarPubMed
Sroufe, L. A. (2000). Early relationships and the development of children. Infant Mental Health Journal, 21, 6774.3.0.CO;2-2>CrossRefGoogle Scholar
Sroufe, L. A., Egeland, B., Carlson, E. A., & Collins, W. A. (2005). The development of the person: The Minnesota study of risk and adaptation from birth to adulthood. New York: Guilford Press.Google Scholar
Suzuki, A., Matsumoto, Y., Shibuya, N., Sadahiro, R., Kamata, M., Goto, K., et al. (2011). The brain-derived neurotrophic factor Val66Met polymorphism modulates the effects of parental rearing on personality traits in healthy subjects. Genes, Brain and Behavior, 10, 385391.CrossRefGoogle ScholarPubMed
Troisi, A., Frazzetto, G., Carola, V., D'Amato, F. R., Moles, A., Siracusano, A., et al. (2010). Social hedonic capacity is associated with the A118G polymorphism of the μ-opioid receptor gene (OPRM1) in adult healthy volunteers and psychiatric patients. Social Neuroscience, 17, 110.Google Scholar
Troisi, A., Frazzetto, G., Carola, V., Di Lorenzo, G., Coviello, M., Siracusano, A., et al. (2012). Variationin the μ-opioid receptor gene (OPRM1) moderates the influence of early maternal care on fearful attachment. Social and Affective Nueroscience, 5, 542547.Google Scholar
Vandell, D. L., Belsky, J., Burchinal, M., Steinberg, L., Vandergrift, N., & the NICHD Early Child Care Research Network. (2010). Do effects of early child care extend to age 15 years? Results from the NICHD Study of Early Child Care and Youth Development. Child Development, 81, 737756.CrossRefGoogle ScholarPubMed
van IJzendoorn, M. H., Bakermans-Kranenburg, M. J., & Mesman, J. (2008). Dopamine system genes associated with parenting in the context of daily hassles. Genes, Brain, and Behavior, 7, 403410.CrossRefGoogle ScholarPubMed
Van Roekel, E., Goossens, L., Scholte, R. H., Engels, R. C., & Verhagen, M. (2011). The dopamine D2 receptor gene, perceived parental support, and adolescent loneliness: Longitudinal evidence for gene–environment interactions. Journal of Child Psychology and Psychiatry, 52, 10441051.CrossRefGoogle ScholarPubMed
Verschoor, E., & Markus, C. R. (2011). Affective and neuroendocrine stress reactivity to an academic examination: Influence of the 5-HTTLPR genotype and trait neuroticism. Biological Psychiatry, 87, 439449.CrossRefGoogle Scholar
Vinberg, M., Trajkovska, V., Bennike, B., Knorr, U., Knudsen, G. M., & Kessing, L. V. (2009). The BDNF Val66Met polymorphism: Relation to familiar risk of affective disorder, BDNF levels and salivary cortisol. Psychoneuroendrocrinlogy, 34, 13801389.CrossRefGoogle ScholarPubMed
Wacker, J., Mueller, E. M., Hennig, J., & Stemmler, G. (2012). How to consistently link extraversion and intelligence to the catechol-O-methyltransferase (COMT) gene: On defining and measuring psychological phenotypes in neurogenetic research. Journal of Personality and Social Psychology, 102, 427444.CrossRefGoogle Scholar
Wakschlag, L. S., Kistner, E. O., Pine, D. S., Biesecker, G., Pickett, K. E., Skol, A. D., et al. (2010). Interaction of prenatal exposure to cigatrettes and MAOA genotype in pathways to you antisocial behavior. Molecular Psychiatry, 15, 928937.CrossRefGoogle Scholar
Way, B. M., Taylor, S. E., & Eisenberger, N. I. (2009). Variation in the μ-opioid receptor gene (OPMR1) is associated with dispositional and neural sensitivity to social rejection. Proceedings of the National Academy of Sciences, 106, 1507915084.CrossRefGoogle Scholar
Widaman, K. F., Helm, J. L., Castro-Schilo, L., Pluess, M., Stallings, M. C., & Belsky, J. (2012). Distinguishing ordinal and disordinal interactions. Psychological Methods, 17, 615622.CrossRefGoogle ScholarPubMed
Widom, C. S., & Bruzustowicz, L. M. (2006). MAOA and the “cycle of violence”: Childhood abuse and neglect, MAOA genotype, and risk for violent and antisocial behavior. Biological Psychiatry, 60, 684689.CrossRefGoogle ScholarPubMed
Woodcock, R. W. (1990). Theoretical foundations of the WJ-R measures of cognitive ability. Journal of Psychoeducational Assessment, 8, 231258.CrossRefGoogle Scholar
Woodcock, R. W., & Johnson, M. B. (1989). Woodcock–Johnson Psycho-Educational Battery—Revised. Allen, TX: DLM.Google Scholar
Yong Zou, G., & Donner, A. (2006). The merits of testing Hardy–Weinberg equilibrium in the analysis of unmatched case-control data: A cautionary note. Annals of Human Genetics, 70, 923933.CrossRefGoogle Scholar
Zuckerman, M. (1999). Vulnerability to psychopathology: A biosocial model. Washington, DC: American Psychological Association.CrossRefGoogle Scholar
Supplementary material: File

Belsky Supplementary Material

Supplementary Material

Download Belsky Supplementary Material(File)
File 24.5 KB