Hostname: page-component-848d4c4894-wzw2p Total loading time: 0 Render date: 2024-04-30T18:39:49.045Z Has data issue: false hasContentIssue false

Genes and Politics: A New Explanation and Evaluation of Twin Study Results and Association Studies in Political Science

Published online by Cambridge University Press:  04 January 2017

Doron Shultziner*
Affiliation:
Department of Medical Education, Faculty of Medicine, Tel Aviv University, Tel Aviv, Israel e-mail: doron.shultziner@mail.huji.ac.il

Abstract

This article offers a new explanation for the results of twin studies in political science that supposedly disclose a genetic basis for political traits. I argue that identical twins tend to be more alike than nonidentical twins because the former are more similarly affected by the same environmental conditions, but the content of those greater trait similarities is nevertheless completely malleable and determined by particular environments. The twin studies method thus can neither prove nor refute the argument for a genetic basis of political traits such as liberal and conservative preferences or voting turnout. The meaning of heritability estimates results in twin studies are discussed, as well as the definition and function of the environment in the political science twin studies. The premature attempts to associate political traits with specific genes despite countertrends in genetics are also examined. I conclude by proposing that the alternative explanation of this article may explain certain puzzles in behavioral genetics, particularly why social and political traits have higher heritability estimates than common physical and medical traits. I map the main point of disagreements with the methodology and the interpretation of its results, and delineate the main operative implications for future research.

Type
Research Article
Copyright
Copyright © The Author 2013. Published by Oxford University Press on behalf of the Society for Political Methodology 

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

Alford, John, Funk, Carolyn, and Hibbing, John. 2005. Are political orientations genetically transmitted? American Political Science Review 99(2): 153–67.Google Scholar
Alford, John R., and Hibbing, John R. 2008. The new empirical biopolitics. Annual Review of Political Science 11: 183203.Google Scholar
Alford, John, and Hibbing, John. 2004. The origin of politics: An evolutionary theory of political behavior. Perspectives on Politics 2(4): 707–23.Google Scholar
Alford, John R., Funk, Carolyn L., and Hibbing, John R. 2008. Beyond liberals and conservatives to political genotypes and phenotypes. Perspectives on Politics 6(2): 321–8.Google Scholar
Bateson, Patrick. 1998. Genes, environment, and the development of behaviour. In The limits of reductionism in biology. Chichester, UK: Novartis Foundation.Google Scholar
Bateson, P., Barker, D., Clutton-Brock, T., Deb, D., D'Udine, B., Foley, R. A., Gluckman, P., Godfrey, K., Kirkwood, T., Lahr, M. M., McNamara, J., Metcalfe, N. B., Monaghan, P., Spencer, H. G., and Sultan, S. E. 2004. Developmental plasticity and human health. Nature 430(6998): 419–21.Google Scholar
Beckwith, Jon, and Morris, Corey A. 2008. Twin studies of political behavior: Untenable assumptions? Perspectives on Politics 6(4): 785–91.Google Scholar
Benjamin, Daniel. 2011. A white paper by Daniel Benjamin. In Genes, cognition, and social behavior: Next steps for foundations and researchers, ed. Lupia, A., 6677. Arlington, VA: National Science Foundation.Google Scholar
Biuso, Emily. 2008. Genopolitics. Times Magazine. December 12. http://www.nytimes.com/2008/12/14/magazine/14Ideas-Section2-B-t-007.html (accessed on November 6, 2012).Google Scholar
Bouchard, T. J., Lykken, D. T., Mcgue, M., Segal, N. L., and Tellegen, A. 1990. Sources of human psychological differences—the Minnesota study of twins reared apart. Science 250(4978): 223–8.Google Scholar
Bouchard, T. J., and 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., Tellegen, A., McGue, M., Keyes, M., and Krueger, R. 2003. Evidence for the construct validity and heritability of the Wilson-Patterson conservatism scale: A reared-apart twins study of social attitudes. Personality and Individual Differences 34(6): 959–69.Google Scholar
Brewer, M. B., and Pierce, K. P. 2005. Social identity complexity and outgroup tolerance. Personality and Social Psychology Bulletin 31(3): 428–37.Google Scholar
Brown, Donald E. 1991. Human universals. New York; London: McGraw-Hill.Google Scholar
Brown, Donald E. 1999. Human Universals. In The MIT encyclopedia of the cognitive sciences, eds. Wilson, R. A. and Keil, F. C. Cambridge, MA: MIT Press.Google Scholar
Brown, Donald E. 2004. Human universals, human nature, and human culture. Daedalus 4: 4754.Google Scholar
Caspi, A., McClay, J., Moffitt, T. E., Mill, J., Martin, J., Craig, I. W., Taylor, A., and Poulton, R. 2002. Role of genotype in the cycle of violence in maltreated children. Science 297(5582): 851–4.Google Scholar
Caspi, A., Sugden, K., Moffitt, T. E., Taylor, A., Craig, I. W., Harrington, H., McClay, J., Mill, J., Martin, J., Braithwaite, A., and Poulton, R. 2003. Influence of life stress on depression: Moderation by a polymorphism in the 5-HTT gene. Science 301(5631): 386–9.Google Scholar
Charney, Evan. 2008a. Genes and ideologies. Perspectives on Politics 6(2): 299319.Google Scholar
Charney, Evan. 2008b. Politics, genetics, and “greedy reductionism.” Perspectives on Politics 6(2): 337–43.Google Scholar
Charney, Evan. 2009. Physiology may not be (political) destiny. PsyCrit. February 25.Google Scholar
Charney, Evan, and English, William. 2012. Candidate genes and political behavior. American Political Science Review 106(1): 134.Google Scholar
Dawes, Christopher T., and Fowler, James H. 2009. Partisanship, voting, and the dopamine D2 receptor gene. Journal of Politics 71(3): 1157–71.Google Scholar
Easton, David. 1976. The relevance of biopolitics to political theory. In Biology and politics, ed. Somit, A. Paris: Mouton.Google Scholar
Falconer, Douglas S., and Mackay, Trudy F. C. 1996. Introduction to quantitative genetics, 4th ed. Essex, UK: Longmans Green, Harlow.Google Scholar
Feldman, M. W., and Otto, S. P. 1997. Twin studies, heritability, and intelligence. Science 278(5342): 1383–4.Google Scholar
Fergusson, D. M., Boden, J. M., Horwood, L. J., Miller, A. L., and Kennedy, M. A. 2011. MAOA, abuse exposure, and antisocial behavior: 30-year longitudinal study. British Journal of Psychiatry 198(6): 457–63.Google Scholar
Fowler, James H., Baker, Laura A., and Dawes, Christopher T. 2008. Genetic variation in political participation. American Political Science Review 102(2): 233–48.Google Scholar
Fowler, James H., and Dawes, Christopher T. 2008. Two genes predict voter turnout. Journal of Politics 70(3): 579–94.Google Scholar
Fowler, James H., and Schreiber, Darren. 2008. Biology, politics, and the emerging science of human nature. Science 322(5903): 912–4.Google Scholar
Gerber, A. S., Huber, G. A., Doherty, D., Dowling, C. M., and Ha, S. E. 2010. Personality and political attitudes: Relationships across issue domains and political contexts. American Political Science Review 104(1): 111–33.CrossRefGoogle Scholar
Gilbert, S. F. 2001. Ecological developmental biology: Developmental biology meets the real world. Developmental Biology 233(1): 112.Google Scholar
Gottlieb, G. 2003. On making behavioral genetics truly developmental. Human Development 46(6): 337–55.Google Scholar
Greenstein, Fred I. 1992. Can personality be studied systematically?. Political Psychology 13(1): 105–28.Google Scholar
Hatemi, Peter K., Byrne, Enda, and McDermott, Rose. Forthcoming. What is a “gene” and why does it matter for political science? Journal of Theoretical Politics.Google Scholar
Hatemi, P. K., Dawes, C. T., Frost-Keller, A., Settle, J. E., and Verhulst, B. 2011. Integrating social science and genetics: News from the political front. Biodemography and Social Biology 57(1): 6787.Google Scholar
Hatemi, P. K., Gillespie, N. A., Eaves, L. J., Maher, B. S., Webb, B. T., Heath, A. C., Medland, S. E., Smyth, D. C., Beeby, H. N., Gordon, S. D., Montgomery, G. W., Zhu, G., Byrne, E. M., and Martin, N. G. 2011. A genome-wide analysis of liberal and conservative political attitudes. Journal of Politics 73(1): 271–85.Google Scholar
Hatemi, P. K., Hibbing, J. R., Medland, S. E., Keller, M. C., Alford, J. R., Smith, K. B., Martin, N. G., and Eaves, L. J. 2010. Not by twins alone: Using the extended family design to investigate genetic influence on political beliefs. American Journal of Political Science 54(3): 798814.Google Scholar
Hatemi, P. K., Alford, J. R., Hibbing, J. R., Martin, N. G., and Eaves, L. J. 2009. Is there a “party” in your genes? Political Research Quarterly 62(3): 584600.Google Scholar
Hatemi, P. K., Funk, C. L., Medland, S. E., Maes, H. M., Silberg, J. L., Martin, N. G., and Eaves, L. J. 2009. Genetic and environmental transmission of political attitudes over a lifetime. Journal of Politics 71(3): 1141–56.Google Scholar
Hatemi, P. K., Medland, S. E., and Eaves, L. J. 2009. Do genes contribute to the “gender gap”? Journal of Politics 71(1): 262–76.Google Scholar
Hines, Samuel M. 1982. Politics and the evolution of inquiry in political science. Politics and the Life Sciences 1(1): 537.Google Scholar
Huang, S. Y., Lin, M. T., Lin, W. W., Huang, C. C., Shy, M. J., and Lu, R. B. 2009. Association of monoamine oxidase A (MAOA) polymorphisms and clinical subgroups of major depressive disorders in the Han Chinese population. World Journal of Biological Psychiatry 10(4): 544–51.Google Scholar
Johnson, W., Turkheimer, E., Gottesman, I. I., and Bouchard, T. J. 2009. Beyond heritability: Twin studies in behavioral research. Current Directions in Psychological Science 18(4): 217–20.Google Scholar
Joseph, Jay. 2010. The genetics of political attitudes and behavior: Claims and refutations. Ethical Human Psychology and Psychiatry 12(3): 200–17.Google Scholar
Keller, M. C., Medland, S. E., and Duncan, L. E. 2010. Are extended twin family designs worth the trouble? A comparison of the bias, precision, and accuracy of parameters estimated in four twin family models. Behavior Genetics 40(3): 377–93.Google Scholar
Kendler, Kenneth S. 2005. “A gene for …”: The nature of gene action in psychiatric disorders. American Journal of Psychiatry 162(7): 1243–52.Google Scholar
Lewontin, R. C. 1974. Analysis of variance and analysis of causes. American Journal of Human Genetics 26(3): 400–11.Google Scholar
Lieberman, M. D., Schreiber, D., and Ochsner, K. N. 2003. Is political cognition like riding a bicycle? How cognitive neuroscience can inform research on political thinking. Political Psychology 24(4): 681704.Google Scholar
Losco, Joseph. 1982. Biopolitics: On evolution in the inquiry into political science—a response to Hines 1982. Politics and the Life Sciences 1(1): 18–9.Google Scholar
Lupia, Arthur, ed. 2011. Genes, cognition, and social behavior: Next steps for foundations and researchers. In National Science Foundation's Political Science Program. Arlington, VA: National Science Foundation.Google Scholar
Lykken, D., and Tellegen, A. 1996. Happiness is a stochastic phenomenon. Psychological Science 7(3): 186–9.Google Scholar
Maher, B. 2008. Personal genomes: The case of the missing heritability. Nature 456(7218): 1821.Google Scholar
Maher, B. 2010. Hiding place for missing heritability uncovered. Nature. Published online 26 January, doi:10.1038/news.2010.33.Google Scholar
Mahoney, J. 2001. Beyond correlational analysis: Recent innovations in theory and method. Sociological Forum 16(3): 575–93.Google Scholar
Mameli, M., and Bateson, P. 2006. Innateness and the sciences. Biology & Philosophy 21(2): 155–88.Google Scholar
Manolio, T. A. 2010. Genome-wide association studies and assessment of the risk of disease. New England Journal of Medicine 363(2): 166–76.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. C., McCarroll, S. A., and Visscher, P. M. 2009. Finding the missing heritability of complex diseases. Nature 461(7265): 747–53.Google Scholar
Martin, N., Boomsma, D., and Machin, G. 1997. A twin-pronged attack on complex traits. Nature Genetics 17(4): 387–92.Google Scholar
Martin, N. G., Eaves, L. J., Heath, A. C., Jardine, R., Feingold, L. M., and Eysenck, H. J. 1986. Transmission of social attitudes. Proceedings of the National Academy of Sciences of the United States of America 83(12): 4364–8.Google Scholar
Masters, Roger D. 1990. Evolutionary biology and political theory. American Political Science Review 84(1): 195210.Google Scholar
McGue, M., and Lykken, D. T. 1992. Genetic influence on risk of divorce. Psychological Science 3(6): 368–73.Google Scholar
McGuffin, Peter, Riley, Brien, and Plomin, Robert. 2001. Genomics and behavior - toward behavioral genomics. Science 291(5507): 1232–3.CrossRefGoogle ScholarPubMed
Meaney, Michael J. 2001. Nature, nurture, and the disunity of knowledge. Annals of the New York Academy of Sciences 935: 5061.Google Scholar
Medland, S. E., and Hatemi, P. K. 2009. Political science, biometric theory, and twin studies: A methodological introduction. Political Analysis 17: 191214.Google Scholar
Molenaar, P. C. M., and Campbell, C. G. 2009. The new person-specific paradigm in psychology. Current Directions in Psychological Science 18(2): 112–7.Google Scholar
Mondak, J. J., Hibbing, M. V., Canache, D., Seligson, M. A., and Anderson, M. R. 2010. Personality and civic engagement: An integrative framework for the study of trait effects on political behavior. American Political Science Review 104(1): 85110.Google Scholar
Munafo, M. R., Durrant, C., Lewis, G., and Flint, J. 2009. Gene x environment interactions at the serotonin transporter locus. Biological Psychiatry 65(3): 211–9.Google Scholar
Munafo, M. R., and Flint, J. 2009. Replication and heterogeneity in gene x environment interaction studies. International Journal of Neuropsychopharmacology 12(6): 727–9.Google Scholar
Oxley, D. R., Smith, K. B., Alford, J. R., Hibbing, M. V., Miller, J. L., Scalora, M., Hatemi, P. K., and Hibbing, J. R. 2008. Political attitudes vary with physiological traits. Science 321(5896): 1667–70.Google Scholar
Pinker, Steven. 2003. The blank slate: The modern denial of human nature. London: Penguin.Google Scholar
Plomin, R. 1989. Environment and genes—determinants of behavior. American Psychologist 44(2): 105–11.Google Scholar
Plomin, R., Owen, M. J., and McGuffin, P. 1994. The genetic basis of complex human behaviors. Science 264(5166): 1733–9.Google Scholar
Price, T. D., Qvarnstrom, A., and Irwin, D. E. 2003. The role of phenotypic plasticity in driving genetic evolution. Proceedings of the Royal Society of London Series B—Biological Sciences 270(1523): 1433–40.Google Scholar
Prichard, Z., Mackinnon, A., Jorm, A. F., and Easteal, S. 2008. No evidence for interaction between MAOA and childhood adversity for antisocial behavior. American Journal of Medical Genetics Part B—Neuropsychiatric Genetics 147B (2): 228–32.Google Scholar
Risch, N., Herrell, R., Lehner, T., Liang, K. Y., Eaves, L., Hoh, J., Griem, A., Kovacs, M., Ott, J., and Merikangas, K. R. 2009. Interaction between the serotonin transporter gene (5-HTTLPR), stressful life events, and risk of depression: A meta-analysis. JAMA—Journal of the American Medical Association 301(23): 2462–71.Google Scholar
Rose, Steven P. R. 2006. Commentary: Heritability Estimates—long past their sell-by date. International Journal of Epidemiology 35(3): 525–7.Google Scholar
Rustow, Dankwark. A. 1970. Transitions to democracy—toward a dynamic model. Comparative Politics 2(3): 337–63.Google Scholar
Sarkar, Sahotra. 1999. From the reaktionsnorm to the adaptive norm: The norm of reaction, 1909–1960. Biology & Philosophy 14(2): 235–52.Google Scholar
Schonemann, P. H. 1997. On models and muddles of heritability. Genetica 99 (2–3): 97108.Google Scholar
Settle, J. E., Dawes, C. T., Christakis, N. A., and Fowler, J. H. 2010. Friendships moderate an association between a dopamine gene variant and political ideology. Journal of Politics 72(4): 1189–98.Google Scholar
Shultziner, Doron. 2010. Struggling for recognition: The psychological impetus for democratic progress. New York: Continuum Press.Google Scholar
Shultziner, Doron, Stevens, Thomas, Stevens, Martin, Stewart, Brian A., Hannagan, Rebbeca J., and Saltini-Semerari, Giulia. 2010. The causes and scope of political egalitarianism during the last glacial: A multi-disciplinary perspective. Biology & Philosophy 25(3): 319–46.Google Scholar
Smith, K. B., Alford, J. R., Hatemi, P. K., Eaves, L., Funk, C. L., and Hibbing, J. R. 2012. Biology, ideology, and epistemology: How do we know political attitudes are inherited and why should we care? American Journal of Political Science 56(1): 1733.Google Scholar
Somit, Albert. 1976. Biology and politics: Recent explorations. The Hague: Mouton.Google Scholar
Somit, Albert, and Peterson, Steven. 1998. Biopolitics after three decades—a balanced sheet. British Journal of Political Science 28(3): 559–71.Google Scholar
Suhay, Elizabeth, Kalmoe, Nathan, and McDermott, Christa. 2007. Why twin studies are problematic for the study of political ideology: Rethinking “are political orientations genetically transmitted?” Presented at the annual meeting of the International Society of Political Psychology, Portland, OR.Google Scholar
Sultan, S. E. 2003. Commentary: The promise of ecological developmental biology. Journal of Experimental Zoology 296(1): 17.Google Scholar
Tajfel, Henri, and Turner, John C. 1979. An integrative theory of intergroup conflict. In The social psychology of intergroup relations, eds. Austin, W. G. and Worchel, S. Monterey, CA: Brooks-Cole.Google Scholar
Todd, J. A. 2006. Statistical false positive or true disease pathway? Nature Genetics 38(7): 731–3.Google Scholar
Tooby, John, and Cosmides, Leda. 1992. The psychological foundations of culture. In The adapted mind: Evolutionary psychology and the generation of culture, eds. Barkow, J. H., Cosmides, L., and Tooby, J. New York: Oxford University Press.Google Scholar
Turkheimer, Eric. 2004. Spinach and ice cream: Why social science is so difficult. In Behavioral genetics principles: Perspectives in development, personality, and psychopathology, ed. DiLalla, L. Washington, DC: American Psychological Association.Google Scholar
Turkheimer, Eric. 2000. Three laws of behavior genetics and what they mean. Current Directions in Psychological Science 9(5): 160–4.Google Scholar
Turkheimer, E., Haley, A., Waldron, M., D'Onofrio, B., and Gottesman, I. I. 2003. Socioeconomic status modifies heritability of IQ in young children. Psychological Science 14(6): 623–8.Google Scholar
Turner, John C. 1982. Toward a cognitive definition of the social group. In Social identity and intergroup relations, ed. Tajfel, H. Cambridge, UK: Cambridge University Press.Google Scholar
Verhulst, Brad, Eaves, Lindon J., and Hatemi, Peter K. 2012. Correlation not causation: The relationship between personality traits and political ideologies. American Journal of Political Science 56(1): 3451.Google Scholar
Visscher, P. M., Hill, W. G., and Wray, N. R. 2008. Heritability in the genomics era—concepts and misconceptions. Nature Reviews Genetics 9(4): 255–66.Google Scholar
Visscher, P. M., Medland, S. E., Ferreira, M. A. R., Morley, K. I., Zhu, G., Cornes, B. K., Montgomery, G. W., and Martin, N. G. 2006. Assumption-free estimation of heritability from genome-wide identity-by-descent sharing between full siblings. Plos Genetics 2(3): 316–25.Google Scholar
Waller, N. G., Kojetin, B. A., Bouchard, T. J., Lykken, D. T., and Tellegen, A. 1990. Genetic and environmental influences on religious interests, attitudes, and values—a study of twins reared apart and together. Psychological Science 1(2): 138–42.Google Scholar
Weder, N., Yang, B. Z., Douglas-Palumberi, H., Massey, J., Krystal, J. H., Gelernter, J., and Kaufman, J. 2009. MAOA genotype, maltreatment, and aggressive behavior: The changing impact of genotype at varying levels of trauma. Biological Psychiatry 65(5): 417–24.Google Scholar
Winter, David. 2003. Personality and political behavior. In Oxford handbook of political psychology, ed. Sears, D. O., Huddy, L., and Jervis, R. New York: Oxford University Press.Google Scholar