Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-23T16:08:40.186Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 June 2012

Tessa M. Pollard
Affiliation:
University of Durham
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Western Diseases
An Evolutionary Perspective
, pp. 173 - 216
Publisher: Cambridge University Press
Print publication year: 2008

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

Abate, N., Carulli, L., Cabo-Chan, A., Chandalia, M., Snell, P. G. and Grundy, S. M. (2003). Genetic polymorphism PC-1 K121Q and ethnic susceptibility to insulin resistance. Journal of Clinical Endocrinology and Metabolism, 88, 5927–34.CrossRefGoogle ScholarPubMed
Adler, A. I., Stratton, I. M., Neil, A. W.et al. on behalf of the UK Prospective Diabetes Study Group (2000). Association of systolic blood pressure with macrovascular and microvascular complications of type 2 diabetes (UKPDS 36): prospective observational study. British Medical Journal, 321, 412–19.CrossRefGoogle ScholarPubMed
Adlerberth, I., Lindberg, E., Aberg, N.et al. (2005). Reduced enterobacterial and increased staphylococcal colonization of the infantile bowel: an effect of hygienic lifestyle?Pediatric Research, 59, 96–101.CrossRefGoogle ScholarPubMed
Adlercreutz, H. (2002). Phyto-estrogens and cancer. The Lancet Oncology, 3, 364–373.CrossRefGoogle ScholarPubMed
Agyemang, C. (2006). Rural and urban differences in blood pressure and hypertension in Ghana, West Africa. Public Health, 120, 525–33.CrossRefGoogle Scholar
Aiello, L. C. and Wells, J. C. K. (2002). Energetics and the evolution of the genus Homo. Annual Review of Anthropology, 31, 323–38.CrossRefGoogle Scholar
Aiello, L. C. and Wheeler, P. (1995). The expensive tissue hypothesis: the brain and digestive system in human and primate evolution. Current Anthropology, 36, 199–221.CrossRefGoogle Scholar
Akdis, M., Blaser, K. and Akdis, C. A. (2005). T regulatory cells in allergy: novel concepts in the pathogenesis, prevention, and treatment of allergic diseases. Journal of Allergy and Clinical Immunology, 116, 961–9.CrossRefGoogle Scholar
Alavanja, M. C. R., Hoppin, J. A. and Kamel, F. (2004). Health effects of chronic pesticide exposure: cancer and neurotoxicity. Annual Review of Public Health, 25, 155–97.CrossRefGoogle ScholarPubMed
Alberti, K. G. M. M., Zimmet, P. and Shaw, J. for the IDF Epidemiology Task Force Consensus Group (2005). The metabolic syndrome – a new worldwide definition. The Lancet, 366, 1059–62.CrossRefGoogle ScholarPubMed
Aligne, C. A., Auinger, P., Byrd, R. S. and Weitzman, M. (2000). Risk factors for pediatric asthma: contributions of poverty, race, and urban residence. American Journal of Respiratory and Critical Care Medicine, 162, 873–7.CrossRefGoogle ScholarPubMed
Allen, J. S. and Cheer, S. M. (1996). The non-thrifty genotype. Current Anthropology, 37, 831–42.CrossRefGoogle Scholar
Allen, N. B. and Badcock, P. B. T. (2003). The social risk hypothesis of depressed mood: evolutionary, psychosocial, and neurobiological processes. Psychological Bulletin, 129, 887–913.CrossRefGoogle Scholar
Allen, N. B. and Badcock, P. B. T. (2006). Darwinian models of depression: a review of evolutionary accounts of mood and mood disorders. Progress in Neuro-Psychopharmacology and Biological Psychiatry, 30, 815–26.CrossRefGoogle ScholarPubMed
Allen, N. E. and Key, T. J. (2000). The effects of diet on circulating sex hormone levels in men. Nutrition Research Reviews, 13, 159–84.CrossRefGoogle ScholarPubMed
Anand, S. S., Yusuf, S., Jacobs, R.et al. for the SHARE-AP Investigators (2001). Risk factors, atherosclerosis, and cardiovascular disease among Aboriginal people in Canada: the Study of Health Assessment and Risk Evaluation in Aboriginal Peoples (SHARE-AP). The Lancet, 358, 1147–53.CrossRefGoogle Scholar
Anderson, H. R. (1997). Air pollution and trends in asthma. In The Rising Trends in Asthma, ed. Chadwick, D. and Cardew, G.. New York: John Wiley and Sons, pp. 190–203.Google Scholar
Anderson, H. R., Ruggles, R., Strachan, D. P.et al. (2004). Trends in prevalence of symptoms of asthma, hay fever, and eczema in 12–14 year olds in the British Isles, 1995–2002: questionnaire survey. British Medical Journal, 328, 1052–3.CrossRefGoogle ScholarPubMed
Anderson, J. W., Johnstone, B. M. and Remley, D. T. (1999). Breast-feeding and cognitive development: a meta-analysis. American Journal of Clinical Nutrition, 70, 525–35.CrossRefGoogle ScholarPubMed
Anderson, M. (1988). Population Change in North-Western Europe, 1750–1850. Basingstoke, UK: Macmillan Education.CrossRefGoogle Scholar
Apter, D. and Vihko, R. (1990). Endocrine determinants of fertility: serum androgen concentrations during follow-up of adolescents into the third decade of life. Journal of Clinical Endocrinology and Metabolism, 71, 970–4.CrossRefGoogle ScholarPubMed
Apter, D., Reinila, M. and Vihko, R. (1989). Some endocrine characteristics of early menarche, a risk factor for breast cancer, are preserved into adulthood. International Journal of Cancer, 44, 783–7.CrossRefGoogle ScholarPubMed
Arnett, D. K., Xiong, B., McGovern, P. G., Blackburn, H. and Luepker, R. V. (2000). Secular trends in dietary macronutrient intake in Minneapolis-St. Paul, Minnesota, 1980–1992. American Journal of Epidemiology, 152, 868–73.CrossRefGoogle Scholar
Asher, M. I., Montefort, S., Björkstén, B.et al. and the ISAAC Phase Three Study Group (2006). Worldwide time trends in the prevalence of symptoms of asthma, allergic rhinoconjunctivitis, and eczema in childhood: ISAAC Phases One and Three repeat multicountry cross-sectional surveys. The Lancet, 368, 733–43.CrossRefGoogle ScholarPubMed
Asia–Pacific Cohort Studies Collaboration (2005). Smoking, quitting, and the risk of cardiovascular disease among women and men in the Asia-Pacific region. International Journal of Epidemiology, 34, 1036–45.CrossRefGoogle Scholar
Aspray, T. J., Mugusi, F., Rashid, S.et al. for the Essential Non-Communicable Disease Health Intervention Project (2000). Rural and urban differences in diabetes prevalence in Tanzania: the role of obesity, physical inactivity and urban living. Transactions of the Royal Society of Tropical Medicine and Hygiene, 94, 637–44.CrossRefGoogle Scholar
Astrup, A. and Finer, N. (2000). Redefining type 2 diabetes: ‘Diabesity’ or ‘Obesity dependent diabetes mellitus’?Obesity Reviews, 1, 57–9.CrossRefGoogle ScholarPubMed
Atwood, L. D., Heard-Costa, N. L., Cupples, L. A., Jaquish, C. E., Wilson, P. W. F. and D'Agostino, R. B. (2002). Genomewide linkage analysis of body mass index across 28 years of the Framingham Heart Study. American Journal of Human Genetics, 71, 1044–50.CrossRefGoogle ScholarPubMed
Baier, L. J. and Hanson, R. L. (2004). Genetic studies of the etiology of type 2 diabetes in Pima Indians. Diabetes, 53, 1181–6.CrossRefGoogle ScholarPubMed
Baker, P. T. (1984). Migration, genetics, and the degenerative diseases of South Pacific Islanders. In Migration and Mobility: Biosocial Aspects of Human Movement, ed. Boyce, A. J.. London: Taylor and Francis, pp. 209–39.Google Scholar
Baker, P. T. (1988). Infectious disease. In Human Biology: An Introduction to Human Evolution, Variation, Growth, and Adaptability, ed. Harrison, G. A., Tanner, J. M., Pilbeam, D. R. and Baker, P. T.. Oxford: Oxford University Press, pp. 508–28.Google Scholar
Baker, P. T., Hanna, J. M. and Baker, T. S., eds. (1986). The Changing Samoans: Behavior and Health in Transition. New York: Oxford University Press.Google Scholar
Balen, A. (1999). Pathogenesis of polycystic ovary syndrome – the enigma unravels?The Lancet, 354, 966–7.CrossRefGoogle ScholarPubMed
Ball, H. L. (2006). Parent–infant bed-sharing behavior: effects of feeding type and presence of father. Human Nature, 17, 301–18.CrossRefGoogle ScholarPubMed
Ball, H. L. and Klingaman, K. (2008). Breastfeeding and mother–infant sleep proximity: implications for infant care. In New Perspectives in Evolutionary Medicine, ed. Trevathan, W. R., Smith, E. O. and McKenna, J. J.. New York: Oxford University Press, pp. 226–241.Google Scholar
Barker, D. J. P. (1994). Mothers, Babies, and Disease in Later Life. London: BMJ Publishing Group.Google Scholar
Barker, D. J. P. and Osmond, C. (1986). Infant mortality, childhood nutrition and ischaemic heart disease in England and Wales. The Lancet, i, 1077–81.CrossRefGoogle Scholar
Barker, D. J. P., Osmond, C., Forsen, T. J., Kajantie, E. and Eriksson, J. G. (2005). Trajectories of growth among children who have coronary events as children. New England Journal of Medicine, 353, 1802–09.CrossRefGoogle ScholarPubMed
Barnes, K. C. (2006). Genetic epidemiology of health disparities in allergy and clinical immunology. Journal of Allergy and Clinical Immunology, 117, 243–54.CrossRefGoogle ScholarPubMed
Barnes, K. C., Armelagos, G. J. and Morreale, S. C. (1999). Darwinian medicine and the emergence of allergy. In Evolutionary Medicine, ed. Trevathan, W. R., Smith, E. O. and McKenna, J. J.. New York: Oxford University Press, pp. 209–43.Google Scholar
Barrett, R., Kuzawa, C. W., McDade, T. and Armelagos, G. J. (1998). Emerging and re-emerging infectious diseases: the third epidemiologic transition. Annual Review of Anthropology, 27, 247–71.CrossRefGoogle Scholar
Barroso, I. (2005). The genetics of type 2 diabetes. Diabetic Medicine, 22, 517–35.CrossRefGoogle ScholarPubMed
Baschetti, R. (1998). Diabetes epidemic in newly westernized populations: is it due to thrifty genes or to genetically unknown foods. Journal of the Royal Society of Medicine, 91, 622–25.CrossRefGoogle ScholarPubMed
Bassuk, S. S. and Manson, J. E. (2005). Epidemiological evidence for the role of physical activity in reducing risk of type 2 diabetes and cardiovascular disease. Journal of Applied Physiology, 99, 1193–204.CrossRefGoogle ScholarPubMed
Ben-Shlomo, Y. and Smith, Davey G. (1991). Deprivation in infancy or in adult life: which is more important for mortality risk?The Lancet, 337, 530–4.CrossRefGoogle ScholarPubMed
Bennett, K., Kabir, Z., Unal, B.et al. (2006). Explaining the recent decrease in coronary heart disease mortality rates in Ireland, 1985–2000. Journal of Epidemiology and Community Health, 60, 322–7.CrossRefGoogle Scholar
Bentley, G. R. (2000). Environmental pollutants and fertility. In Infertility in the Modern World: Present and Future Prospects, ed. Bentley, G. R. and Mascie-Taylor, C. G. N.. Cambridge: Cambridge University Press, pp. 85–152.CrossRefGoogle Scholar
Bentley, G. R., Jasieńska, G. and Goldberg, T. (1993). Is the fertility of agriculturalists higher than that of nonagriculturalists?Current Anthropology, 34, 778–85.CrossRefGoogle Scholar
Bentley, G. R., Harrigan, A. M. and Ellison, P. T. (1998). Dietary composition and ovarian function among Lese horticulturalist women of the Ituri Forest, Democratic Republic of Congo. European Journal of Clinical Nutrition, 52, 261–70.CrossRefGoogle ScholarPubMed
Bentley, G. R., Paine, R. R. and Boldsen, J. L. (2001). Fertility changes with the prehistoric transition to agriculture. In Reproductive Ecology and Human Evolution, ed. Ellison, P.. New York: Aldine de Gruyter, pp. 201–31.Google Scholar
Benyshek, D. C. and Watson, J. T. (2006). Exploring the thrifty genotype's food-shortage assumptions: a cross-cultural comparison of ethnographic accounts of food security among foraging and agricultural societies. American Journal of Physical Anthropology, 131, 120–6.CrossRefGoogle ScholarPubMed
Beral, V., Bull, D., Doll, R.et al. (1997). Breast cancer and hormone replacement therapy: collaborative reanalysis of data from 51 epidemiological studies of 52,705 women with breast cancer and 108,411 women without breast cancer. The Lancet, 350, 1047–59.Google Scholar
Beral, V., Bull, D., Doll, R.et al. (2002). Breast cancer and breastfeeding: collaborative reanalysis of individual data from 47 epidemiological studies in 30 countries, including 50,302 women with breast cancer and 969,973 women without the disease. The Lancet, 360, 187–95.Google Scholar
Beran, D. and Yudkin, J. S. (2006). Diabetes care in sub-Saharan Africa. The Lancet, 368, 1689–95.CrossRefGoogle ScholarPubMed
Bernstein, L. and Ross, R. K. (1993). Endogenous hormones and breast cancer risk. Epidemiologic Reviews, 15, 48–65.CrossRefGoogle ScholarPubMed
Bernstein, L., Yuan, J. M., Ross, R. K.et al. (1990). Serum hormone levels in pre-menopausal Chinese women in Shanghai and white women in Los Angeles: results from two breast cancer case-control studies. Cancer Causes and Control, 1, 51–8.CrossRefGoogle ScholarPubMed
Bernstein, L., Henderson, B. E., Hanisch, R., Sullivan-Halley, J. and Ross, R. K. (1994). Physical exercise and reduced risk of breast cancer in young women. Journal of the National Cancer Institute, 86, 1403–08.CrossRefGoogle ScholarPubMed
Berrington, A. (2004). Perpetual postponers? Women's, men's and couple's infertility intentions and subsequent fertility behaviour. Population Trends, 117, 9–19.Google Scholar
Biesele, M. and Howell, N. (1981). “The old people give you life”: aging among!Kung hunter–gatherers. In Other Ways of Growing Old, ed. Amoss, P. and Harrell, S.. Stanford: Stanford University Press, pp. 77–98.Google Scholar
Bilsborough, S. and Mann, N. (2006). A review of issues of dietary protein intake in humans. International Journal of Sport Nutrition and Exercise Metabolism, 16, 129–52.CrossRefGoogle ScholarPubMed
Bindon, J. R. and Baker, P. T. (1997). Bergmann's rule and the thrifty genotype. American Journal of Physical Anthropology, 104, 201–10.3.0.CO;2-0>CrossRefGoogle ScholarPubMed
Bjerve, K. S., Brubakk, A. M., Fougner, K. H., Johnsen, H., Midthjell, K. and Vik, T. (1993). Omega-3 fatty acids: essential fatty acids with important biological effects, and serum phospholipid fatty acids as markers of dietary omega-3 fatty acid intake. American Journal of Clinical Nutrition, 57 (Suppl.), 801S–6S.CrossRefGoogle ScholarPubMed
Björkstén, B., Naaber, P., Sepp, E. and Mikelsaar, M. (1999). The intestinal microbiota in allergic Estonian and Swedish 2-year-old children. Clinical and Experimental Allergy, 29, 342–6.CrossRefGoogle Scholar
Björntorp, P. and Rosmond, R. (2000). Obesity and cortisol. Nutrition, 16, 924–36.CrossRefGoogle ScholarPubMed
Blackley, R. (2001). The arrival of the Maoris in New Zealand, 1898. Emerging Infectious Diseases, 7, 914.CrossRefGoogle Scholar
Blair, P. S. and Ball, H. L. (2004). The prevalence and characteristics associated with parent–infant bed-sharing in England. Archives of Disease in Childhood, 89, 1106–10.CrossRefGoogle ScholarPubMed
Blumenthal, M. N., Langefeld, C. D., Beaty, T. H.et al. (2004). A genome-wide search for allergic response (atopy) genes in three ethnic groups: Collaborative Study on the Genetics of Asthma. Human Genetics, 114, 157–64.CrossRefGoogle Scholar
Jones, Blurton N. G., Hawkes, K. and O'Connell, J. F. (2002). Antiquity of postreproductive life: are there modern impacts on hunter–gatherer postreproductive life spans?American Journal of Human Biology, 14, 184–205.CrossRefGoogle Scholar
Bolling, K. (2006). Infant Feeding Survey 2005: Early Results. London: Information Centre, Government Statistical Service, p. 19.Google Scholar
Bonnet, M. H. and Arand, D. L. (1995). We are chronically sleep deprived. Sleep, 18, 908–11.CrossRefGoogle ScholarPubMed
Botha, J. L., Bray, F., Sankila, R. and Parkin, D. M. (2003). Breast cancer incidence and mortality trends in 16 European countries. European Journal of Cancer, 39, 1718–29.CrossRefGoogle ScholarPubMed
Bowlby, J. (1969). Attachment and Loss. New York: Basic Books.Google Scholar
Boyd, R. and Silk, J. B. (2006). How Humans Evolved. Los Angeles, University of California: University of California Press.Google Scholar
Boyden, S. V., ed. (1970). The Impact of Civilization on the Biology of Man. Canberra: Australian National University Press.Google Scholar
Boyden, S. V. (1987). Western Civilization in Biological Perspective: Patterns in Biohistory. Oxford: Oxford University Press.Google Scholar
Braman, S. S. (2006). The global burden of asthma. Chest, 130, 4S–12S.CrossRefGoogle Scholar
Braun, L. (2002). Race, ethnicity and health: can genetics explain disparities?Perspectives in Biology and Medicine, 45, 159–74.CrossRefGoogle ScholarPubMed
Bribiescas, R. G. (2001a). Reproductive ecology and life history of the human male. Yearbook of Physical Anthropology, 44, 148–76.CrossRefGoogle Scholar
Bribiescas, R. G. (2001b). Reproductive physiology of the human male: an evolutionary and life history perspective. In Reproductive Ecology and Human Evolution, ed. Ellison, P. T.. New York: Aldine de Gruyter, pp. 107–35.Google Scholar
Briefel, R. R. and Johnson, C. L. (2004). Secular trends in dietary intake in the United States. Annual Review of Nutrition, 24, 401–31.CrossRefGoogle ScholarPubMed
Brock, J. (1993). Native Plants of Northern Australia. New Holland: Frenchs Forest, Reed.Google Scholar
Brown, D. E. (2006). Measuring hormonal variation in the sympathetic nervous system: catecholamines. In Measuring Stress in Humans: A Practical Guide for the Field, ed. Ice, G. H. and James, G. D.. Cambridge: Cambridge University Press, pp. 94–121.CrossRefGoogle Scholar
Brown, E. S., Varghese, F. P. and McEwen, B. S. (2004). Association of depression with medical illness: does cortisol play a role?Biological Psychiatry, 55, 1–9.CrossRefGoogle ScholarPubMed
Brown, P. J. and Bentley-Condit, V. K. (1998). Culture, evolution, and obesity. In Handbook of Obesity, ed. Bray, G. A., Bouchard, C. and James, W. P. T.. New York: Marcel Dekker, pp. 143–55.Google Scholar
Brown, P. J. and Konner, M. (1987). An anthropological perspective on obesity. Annals of the New York Academy of Sciences, 499, 29–46.CrossRefGoogle ScholarPubMed
Brown, P. J. and Krick, S. V. (2001). The etiology of obesity: diet, television and the illusions of personal choice. In Obesity, Growth and Development, ed. Johnston, F. E. and Foster, G. D.. London: Smith-Gordon, pp. 111–28.Google Scholar
Brunner, E. and Marmot, M. (1999). Social organization, stress, and health. In Social Determinants of Health, ed. Marmot, M. and Wilkinson, R. G.. Oxford: Oxford University Press, pp. 17–43.Google Scholar
Burkitt, D. P. (1973). Some diseases characteristic of modern western civilization. British Medical Journal, 1, 274–8.CrossRefGoogle ScholarPubMed
Butte, N. F. (2001). The role of breastfeeding in obesity. Pediatric Clinics of North America, 48, 189–98.CrossRefGoogle ScholarPubMed
Cacioppo, J. T. and Hawkley, L. C. (2003). Social isolation and health, with an emphasis on underlying mechanisms. Perspectives in Biology and Medicine, 46, S39–52.CrossRefGoogle ScholarPubMed
Caldwell, J. C. (2001). Demographers and the study of mortality: scope, perspectives and theory. Annals of the New York Academy of Sciences, 954, 19–34.CrossRefGoogle Scholar
Calle, E. E. and Kaaks, R. (2004). Overweight, obesity and cancer: epidemiological evidence and proposed mechanisms. Nature Reviews Cancer, 4, 579–91.CrossRefGoogle ScholarPubMed
Calvani, M., Alessandri, C. and Bonci, E. (2002). Fever episodes in early life and the development of atopy in children with asthma. European Respiratory Journal, 20, 391–6.CrossRefGoogle ScholarPubMed
Camacho, T. C., Roberts, R. E., Lazarus, N. B., Kaplan, G. A. and Cohen, R. D. (1991). Physical activity and depression: evidence from the Alameda County Study. American Journal of Epidemiology, 134, 220–31.CrossRefGoogle ScholarPubMed
Carmichael, A. R. and Bates, T. (2004). Obesity and breast cancer: a review of the literature. The Breast, 13, 85–92.CrossRefGoogle ScholarPubMed
Carter, H. (1999). Urban origins: a review of theories. In The Pre-Industrial Cities and Technology Reader, ed. Chant, C.. London: Routledge and The Open University, pp. 7–14.Google Scholar
Carulli, L., Rondinella, S., Lombardini, S., Canedi, I., Loria, P. and Carulli, N. (2005). Diabetes, genetics and ethnicity. Alimentary Pharmacology and Therapeutics, 22 (Suppl. 2), 16–19.CrossRefGoogle ScholarPubMed
Catanese, D. M., Koetting O'Byrne, K. and Poston, W. S. C. (2001). The epidemiology of obesity in developed countries. In Obesity, Growth and Development, ed. Johnston, F. and Foster, G.. London: Smith-Gordon, pp. 69–90.Google Scholar
Cavalli-Sforza, L. L. and Feldman, M. W. (2003). The application of molecular genetic approaches to the study of human evolution. Nature Genetics, 33, 266–75.CrossRefGoogle Scholar
Cavallo, M. G., Fava, D., Monetini, L., Barone, F. and Pozzilli, P. (1996). Cell-mediated immune response to β casein in recent-onset insulin dependent diabetes: implications for disease pathogenesis. The Lancet, 348, 926–8.CrossRefGoogle ScholarPubMed
Chisholm, J. S. and Burbank, V. K. (2001). Evolution and inequality. International Journal of Epidemiology, 30, 206–11.CrossRefGoogle ScholarPubMed
Chlebowski, R. T., Blackburn, G. L., Thomson, C. A.et al. (2006). Dietary fat reduction and breast cancer outcome: interim efficacy results from the Women's Intervention Nutrition Study. Journal of the National Cancer Institute, 98, 1767–76.CrossRefGoogle ScholarPubMed
Chobanian, A. V. and Alexander, R. W. (1996). Exacerbation of atherosclerosis by hypertension. Archives of Internal Medicine, 156, 1952–6.CrossRefGoogle ScholarPubMed
Choi, Y. J., Cho, Y. M., Park, C. K.et al. (2006). Rapidly increasing diabetes-related mortality with socio-environmental changes in South Korea during the last two decades. Diabetes Research and Clinical Practice, 74, 295–300.CrossRefGoogle ScholarPubMed
Chow, C. K., Raju, P. K., Raju, R.et al. (2006). The prevalence and management of diabetes in rural India. Diabetes Care, 29, 1717–18.CrossRefGoogle ScholarPubMed
Cleave, T. L., Campbell, G. D. and Painter, N. S. (1969). Diabetes, Coronary Thrombosis, and the Saccharine Disease. Bristol: Wright.Google Scholar
Coe, K. and Steadman, L. B. (1995). The human breast and the ancestral reproductive cycle: a preliminary inquiry into breast cancer etiology. Human Nature, 6, 197–220.CrossRefGoogle ScholarPubMed
Cohen, M. N. (1989). Health and the Rise of Civilization. New Haven: Yale University Press.Google Scholar
Cohen, S. (2005). The Pittsburgh common cold studies: psychosocial predictors of susceptibility to respiratory infectious illness. International Journal of Behavioral Medicine, 12, 123–31.CrossRefGoogle ScholarPubMed
Cohen, S., Doyle, W. and Baum, A. (2006). Socioeconomic status is associated with stress hormones. Psychosomatic Medicine, 68, 414–20.CrossRefGoogle ScholarPubMed
Colagiuri, S., Colagiuri, R., Na'ati, S., Muimuiheata, S., Hussain, Z. and Palu, T. (2002). The prevalence of diabetes in the Kingdom of Tonga. Diabetes Care, 25, 1378–83.CrossRefGoogle ScholarPubMed
Cole, T. J., Bellizzi, M. C., Flegal, K. M. and Dietz, W. H. (2000). Establishing a standard definition for child overweight and obesity worldwide: international survey. British Medical Journal, 320, 1240–3.CrossRefGoogle ScholarPubMed
Colla, J., Buka, S., Harrington, D. and Murphy, J. M. (2006). Depression and modernization: a cross-cultural study of women. Social Psychiatry and Psychiatric Epidemiology, 41, 271–9.CrossRefGoogle Scholar
Collins, P. (2001). GPs and stress discourse: a preliminary report. Anthropology in Action, 8, 36–44.Google Scholar
Cook, D. C. (1984). Subsistence and health in the lower Illinois Valley: osteological evidence. In Palaeopathology at the Origins of Agriculture, ed. Cohen, M. N. and Armelagos, G. J.. Orlando, Florida: Academic Press, pp. 235–69.Google Scholar
Cooke, A., Zaccone, P., Raine, T., Phillips, J. M. and Dunne, D. W. (2004). Infection and autoimmunity: are we winning the war, only to lose the peace?Trends in Parasitology, 20, 316–21.CrossRefGoogle ScholarPubMed
Cookson, W. O. and Moffatt, M. F. (1997). Asthma: an epidemic in the absence of infection?Science, 275, 41–2.CrossRefGoogle Scholar
Cooper, P. J. (2004). Intestinal worms and human allergy. Parasite Immunology, 26, 455–67.CrossRefGoogle ScholarPubMed
Cooper, R. S., Rotimi, C., Ataman, S.et al. (1997). The prevalence of hypertension in seven populations of West African origin. American Journal of Public Health, 87, 160–8.CrossRefGoogle ScholarPubMed
Cooper, R. S., Amoah, A. G. and Mensah, G. A. (2003). High blood pressure: the foundation for epidemic cardiovascular disease in African populations. Ethnicity and Disease, 13 (Suppl. 2), S48–52.Google ScholarPubMed
Cordain, L., Gotshall, R. W. and Eaton, S. B. (1997). Evolutionary aspects of exercise. World Review of Nutrition and Dietetics, 81, 49–60.CrossRefGoogle Scholar
Cordain, L., Miller, Brand J., Eaton, S. B., Mann, N., Holt, S. H. A. and Speth, J. D. (2000). Plant–animal subsistence ratios and macronutrient energy estimations in worldwide hunter–gatherer diets. American Journal of Clinical Nutrition, 71, 682–92.CrossRefGoogle ScholarPubMed
Cordain, L., Eaton, S. B., Miller, Brand J., Mann, N. and Hill, K. (2002). The paradoxical nature of hunter–gatherer diets: meat-based, yet non-atherogenic. European Journal of Clinical Nutrition, 56 (Suppl. 1), S42–52.CrossRefGoogle ScholarPubMed
Cordain, L., Eades, M. R. and Eades, M. D. (2003). Hyperinsulinemic diseases of civilization: more than just Syndrome X. Comparative Biochemistry and Physiology Part A, 136, 95–112.CrossRefGoogle ScholarPubMed
Cordain, L., Eaton, S. B., Sebastian, A.et al. (2005). Origins and evolution of the Western diet: health implications for the 21st century. American Journal of Clinical Nutrition, 81, 341–54.CrossRefGoogle ScholarPubMed
Cosmides, L. and Tooby, J. (1997). Evolutionary Psychology: a Primer. University of California Santa Barbara: University of California Press.Google Scholar
Costa, D. L. (2000). Understanding the twentieth-century decline in chronic conditions among older men. Demography, 37, 53–72.CrossRefGoogle ScholarPubMed
Costa, D. L. (2004). The measure of man and older age mortality: evidence from the Gould sample. Journal of Economic History, 64, 1–23.CrossRefGoogle Scholar
Craig, P., Halavatau, V., Comino, E. and Caterson, I. (1999). Perception of body size in the Tongan community: differences from and similarities to an Australian sample. International Journal of Obesity, 23, 1288–94.CrossRefGoogle Scholar
Critser, G. (2003). Fat Land: How Americans became the Fattest People in the World. London: Penguin.Google Scholar
Crowther, R., Dinsdale, H., Rutter, H. and Kyffin, R. (2007). Analysis of the National Childhood Obesity Database 2005–06: South East Public Health Observatory.
Cuijpers, P., Straten, A. and Smit, F. (2005). Preventing the incidence of new cases of mental disorders – a meta-analytic review. Journal of Nervous and Mental Disease, 193, 119–25.CrossRefGoogle ScholarPubMed
Cunningham, A. S. (1995). Breastfeeding: adaptive behavior for child health and longevity. In Breastfeeding: Biocultural Perspectives, ed. Stuart-Macadam, P. and Dettwyler, K. A.. New York: Aldine de Gruyter, pp. 243–64.Google Scholar
Cunningham, G. R. (2006). Testosterone replacement therapy for late-onset hypogonadism. Nature Clinical Practice Urology, 3, 260–7.CrossRefGoogle ScholarPubMed
Daniel, M., Rowley, K. G., McDermott, R. and O'Dea, K. (2002). Diabetes and impaired glucose tolerance in Aboriginal Australians: prevalence and risk. Diabetes Research and Clinical Practice, 57, 23–33.CrossRefGoogle ScholarPubMed
Smith, Davey G. and Marmot, M. G. (1991). Trends in mortality in Britain: 1920–1986. Annals of Nutrition and Metabolism, 35 (Suppl. 1), 53–63.CrossRefGoogle ScholarPubMed
Davidson, K., Jonas, B., Dixon, K. and Markovitz, J. (2000). Do depression symptoms predict early hypertension incidence in young adults in the CARDIA study? Coronary artery risk development in young adults. Archives of Internal Medicine, 160, 1495–500.CrossRefGoogle ScholarPubMed
Davis, A. M., Kreutzer, R., Lipsett, M., King, G. and Shaikh, N. (2006). Asthma prevalence in Hispanic and Asian American ethnic subgroups: Results from the California Healthy Kids Survey. Pediatrics, 118, e363–70.CrossRefGoogle ScholarPubMed
Boever, E., Bacquer, D., Braeckman, L., Baele, L., Rosseneu, M. and Backer, G. (1995). Relation of fibrinogen to lifestyles and to cardiovascular risk factors in a working population. International Journal of Epidemiology, 24, 915–21.CrossRefGoogle Scholar
Jong, F. H., Oishi, K., Hayes, R. B.et al. (1991). Perpipheral hormone levels in controls and patients with prostatic cancer or benign prostatic hyperplasia – results from the Dutch-Japanese case-control study. Cancer Research, 51, 3445–50.Google ScholarPubMed
Laet, C. E. D. H. and Pols, H. A. P. (2000). Fractures in the elderly: epidemiology and demography. Best Practice and Research in Clinical Endocrinology and Metabolism, 14, 171–9.CrossRefGoogle ScholarPubMed
Delisle, H. F., Rivard, M. and Ekoe, J. M. (1995). Prevalence estimates of diabetes and other cardiovascular risk factors in the two largest Algonquin communities of Quebec. Diabetes Care, 18, 1255–9.CrossRefGoogle Scholar
DeMattia, L., Lemont, L. and Meurer, L. (2006). Do interventions to limit sedentary behaviours change behaviour and reduce childhood obesity? A critical review of the literature. Obesity Reviews, 8, 69–81.CrossRefGoogle Scholar
Der, G., Batty, D. and Deary, I. J. (2006). Effect of breast feeding on intelligence in children: prospective study, sibling pairs analysis, and meta-analysis. British Medical Journal, 333, 945–50.CrossRefGoogle ScholarPubMed
Despres, J. P. (2006). Is visceral obesity the cause of the metabolic syndrome?Annals of Medicine, 38, 52–63.CrossRefGoogle ScholarPubMed
Dettwyler, K. A. (1995a). Beauty and the breast: the cultural context of breastfeeding in the United States. In Breastfeeding: Biocultural Perspectives, ed. Stuart-Macadam, P. and Dettwyler, K. A.. New York: Aldine de Gruyter, pp. 167–215.Google Scholar
Dettwyler, K. A. (1995b). A time to wean: the hominid blueprint for the natural age of weaning in modern human populations. In Breastfeeding: Biocultural Perspectives, ed. Stuart-Macadam, P. and Dettwyler, K. A.. New York: Aldine de Gruyter, pp. 39–74.Google Scholar
Diamond, J. (2003). The double puzzle of diabetes. Nature, 423, 599–602.CrossRefGoogle ScholarPubMed
Dietz, W. H. (1996). The role of lifestyle in health: the epidemiology and consequences of inactivity. Proceedings of the Nutrition Society, 55, 829–40.CrossRefGoogle ScholarPubMed
Ding, E. L., Song, Y., Malik, V. and Liu, S. (2006). Sex differences of endogenous sex hormones and risk of type 2 diabetes: a systematic review and meta-analysis. Journal of the American Medical Association, 295, 1288–99.CrossRefGoogle ScholarPubMed
Dowse, G. K., Zimmet, P. Z. and King, H. (1991). Relationship between prevalence of impaired glucose tolerance and NIDDM in a population. Diabetes Care, 14, 968–74.CrossRefGoogle ScholarPubMed
Dowse, G. K., Gareeboo, H., Alberti, K. G. M. M.et al. (1995). Changes in population cholesterol concentrations and other cardiovascular risk factor levels after five years of the non-communicable disease intervention programme in Mauritius. British Medical Journal, 311, 1255–9.CrossRefGoogle ScholarPubMed
Drake, A. J. and Walker, B. R. (2004). The intergenerational effects of fetal programming: non-genomic mechanisms for the inheritance of low birth weight and cardiovascular risk. Journal of Endocrinology, 180, 1–16.CrossRefGoogle ScholarPubMed
Dressler, W. W. (1995). Modeling biocultural interactions: examples from studies of stress and cardiovascular disease. Yearbook of Physical Anthropology, 38, 27–56.CrossRefGoogle Scholar
Dressler, W. W., Balieiro, M. C., Ribeiro, R. P. and Santos, Dos J. E. (2005). Cultural consonance and arterial blood pressure in urban Brazil. Social Science and Medicine, 61, 527–40.CrossRefGoogle ScholarPubMed
Dudley, R. (2000). Evolutionary origins of human alcoholism in primate frugivory. Quarterly Review of Biology, 75, 3–15.CrossRefGoogle ScholarPubMed
Duggirala, R., Almasy, L., Blangero, J.et al. (2003). American Diabetes Association GENNID Study Group: Further evidence for a type 2 diabetes susceptibility locus on chromosome 11q. Genetic Epidemiology, 24, 240–2.CrossRefGoogle Scholar
Dunaif, A. (1997). Insulin resistance and the polycystic ovary syndrome: mechanism and implications for pathogenesis. Endocrine Reviews, 18, 774–800.Google ScholarPubMed
Dunn, J. E. (1975). Cancer epidemiology in populations of the United States – with an emphasis on Hawaii and California – and Japan. Cancer Research, 35, 3240–5.Google ScholarPubMed
Dunne, D. W. and Cooke, A. (2005). A worm's eye view of the immune system: consequences for evolution of human autoimmune disease. Nature Reviews Immunology, 5, 420–6.CrossRefGoogle ScholarPubMed
Durham, W. H. (1991). Coevolution: Genes, Culture, and Human Diversity. Stanford, California: Stanford University Press.Google Scholar
Eaton, S. B. and Eaton, S. B. (1999a). Breast cancer in evolutionary context. In Evolutionary Medicine, ed. Trevathan, W. R., Smith, E. O. and McKenna, J. J.. New York: Oxford University Press, pp. 429–42.Google Scholar
Eaton, S. B. and Eaton, S. B. (1999b). Hunter–gatherers and human health. In The Cambridge Encyclopedia of Hunters and Gatherers, ed. Lee, R. B. and Daly, R.. Cambridge: Cambridge University Press, pp. 449–56.Google Scholar
Eaton, S. B., Konner, M. and Shostak, M. (1988a). Stone agers in the fast lane: chronic degenerative diseases in evolutionary perspective. American Journal of Medicine, 84, 739–49.CrossRefGoogle Scholar
Eaton, S. B., Shostak, M. and Konner, M. (1988b). The Paleolithic Prescription: A Program of Diet and Exercise and a Design for Living. New York: Harper and Row.Google Scholar
Eaton, S. B., Pike, M. C., Short, R. V.et al. (1994). Women's reproductive biology in evolutionary context. Quarterly Review of Biology, 69, 353–67.CrossRefGoogle Scholar
Eaton, S. B., Eaton, S. B. and Konner, M. J. (1999). Paleolithic nutrition revisited. In Evolutionary Medicine, ed. Trevathan, W. R., Smith, E. O. and McKenna, J. J.. New York: Oxford University Press, pp. 313–32.Google Scholar
Eaton, S. B., Eaton, S. B. and Cordain, L. (2002). Evolution, diet, and health. In Human Diet: Its Origin and Evolution, ed. Ungar, P. S. and Teaford, M. F.. Westport, Connecticut: Bergin and Garvey, pp. 7–18.Google Scholar
Eckersley, R. (2005). Is modern Western culture a health hazard?International Journal of Epidemiology, 35, 252–8.CrossRefGoogle ScholarPubMed
Eisenmann, J. C. (2003). Secular trends in variables associated with the metabolic syndrome of North American children and adolescents: review and synthesis. American Journal of Human Biology, 15, 786–94.CrossRefGoogle ScholarPubMed
Ekbom, A. (2006). The developmental environment and the early origins of cancer. In Developmental Origins of Health and Disease, ed. Gluckman, P. and Hanson, M. A.. Cambridge: Cambridge University Press, pp. 415–25.CrossRefGoogle Scholar
Elliott, P., Stamler, J., Nichols, R.et al. (1996). Intersalt revisited: further analyses of 24-hour sodium excretion and blood pressure within and across populations. British Medical Journal, 312, 1249–53.CrossRefGoogle ScholarPubMed
Ellison, G. T. H. and Jones, Rees I. (2002). Social identities and the ‘new genetics’: scientific and social consequences. Critical Public Health, 12, 265–82.CrossRefGoogle Scholar
Ellison, P. T. (1990). Human ovarian function and reproductive ecology: new hypotheses. American Anthropologist, 92, 933–52.CrossRefGoogle Scholar
Ellison, P. T. (1996). Developmental influences on adult ovarian hormonal function. American Journal of Human Biology, 8, 725–34.3.0.CO;2-S>CrossRefGoogle ScholarPubMed
Ellison, P. T. (1999). Reproductive ecology and reproductive cancers. In Hormones, Health, and Behavior, ed. Panter-Brick, C. and Worthman, C. M.. Cambridge: Cambridge University Press, pp. 184–209.Google Scholar
Ellison, P. T. (2001). On Fertile Ground: A Natural History of Human Reproduction. Cambridge, Massachusetts: Harvard University Press.Google Scholar
Ellison, P. T. and Lager, C. (1986). Moderate recreational running is associated with lowered salivary progesterone profiles in women. American Journal of Obstetrics and Gynecology, 154, 1000–3.CrossRefGoogle ScholarPubMed
Ellison, P. T., Lipson, S. F., O'Rourke, M. T.et al. (1993a). Population variation in ovarian function. The Lancet, 342, 433–434.CrossRefGoogle Scholar
Ellison, P. T., Panter-Brick, C., Lipson, S. F. and O'Rourke, M. T. (1993b). The ecological context of human ovarian function. Human Reproduction, 8, 2248–58.CrossRefGoogle Scholar
Ellison, P. T., Bribiescas, R. G., Bentley, G. R.et al. (2002). Population variation in age-related decline in male salivary testosterone. Human Reproduction, 17, 3251–3.CrossRefGoogle ScholarPubMed
Ellison, P. T. and Jasiénska, G. (2007). Constraint, pathology, and adaptation: how can we tell them apart?American Journal of Human Biology, 19, 622–30.CrossRefGoogle ScholarPubMed
Elsom, D. M. (1992). Atmospheric Pollution: A Global Problem. Oxford: Blackwell.Google Scholar
Emanuel, M. B. (1988). Hay fever, a post industrial revolution epidemic: a history of its growth during the 19th century. Clinical Allergy, 18, 295–304.CrossRefGoogle ScholarPubMed
Endogenous Hormones and Breast Cancer Collaborative Group (2003). Body mass index, serum sex hormones, and breast cancer risk in postmenopausal women. Journal of the National Cancer Institute, 95, 1218–26.CrossRefGoogle Scholar
Erdal, D., Whiten, A., Boehm, C. and Knauft, B. (1994). On human egalitarianism: an evolutionary product of Machiavellian status escalation?Current Anthropology, 35, 175–83.CrossRefGoogle Scholar
Ernst, J. and Cacioppo, J. (1999). Lonely hearts: psychological perspectives on loneliness. Applied and Preventive Psychology, 8, 1–22.CrossRefGoogle Scholar
Eshed, V., Gopher, A. and Hershkovitz, I. (2006). Tooth wear and dental pathology at the advent of agriculture: new evidence from the Levant. American Journal of Physical Anthropology, 130, 145–59.CrossRefGoogle ScholarPubMed
Ewald, P. (2002). Plague Time: The New Germ Theory of Disease. New York: Anchor Books.Google Scholar
Ewing, R., Schmid, T., Killingsworth, R., Zlot, A. and Raudenbush, S. (2003). Relationship between urban sprawl and physical activity, obesity, and morbidity. American Journal of Health Promotion, 5, 47–57.CrossRefGoogle Scholar
Ezzati, M., Hoorn, Vander S., Lawes, C. M. M.et al. (2005). Rethinking the “diseases of affluence” paradigm: global patterns of nutritional risk in relation to economic development. PLoS Medicine, 2, e133.CrossRefGoogle ScholarPubMed
Falkner, B., Sherif, K., Sumner, A. and Kushner, H. (1999). Hyperinsulinism and sex hormones in young adult African Americans. Metabolism, 48, 107–12.CrossRefGoogle ScholarPubMed
Fall, C. H. D. (2001). Non-industrialised countries and affluence. British Medical Bulletin, 60, 33–50.CrossRefGoogle ScholarPubMed
Fanaro, S., Chierici, R., Guerrini, P. and Vigi, V. (2003). Intestinal microflora in early infancy: composition and development. Acta Paeditrica, 441 (Suppl.), 48–55.Google Scholar
FAO/WHO/UNU Expert Consultation (1985). Energy and Protein Requirements. Geneva: World Health Organization.Google Scholar
Fee, M. (2006). Racializing narratives: obesity, diabetes and the “Aboriginal” thrifty genotype. Social Science and Medicine, 62, 2988–97.CrossRefGoogle ScholarPubMed
Fein, S. B. and Roe, B. (1998). The effect of work status on initiation and duration of breast-feeding. American Journal of Public Health, 88, 1042–6.CrossRefGoogle ScholarPubMed
Feldman, H. A., Goldstein, I., Hatzichkristou, D. G., Krane, R. J. and McKinlay, J. B. (1994). Impotence and its medical and psychosocial correlates: results of the Massachusetts Male Aging Study. Journal of Urology, 151, 54–61.CrossRefGoogle ScholarPubMed
Fiennes, R. (1978). Zoonoses and the Origins and Ecology of Human Disease. London: Academic Press.Google Scholar
Fildes, V. (1995). The culture and biology of breastfeeding: An historical review of western Europe. In Breastfeeding: Biocultural Perspectives, ed. Stuart-Macadam, P. and Dettwyler, K. A.. New York: Aldine de Gruyter, pp. 101–26.Google Scholar
Finch, C. E. and Crimmins, E. M. (2004). Inflammatory exposure and historical changes in human life-spans. Science, 305, 1736–9.CrossRefGoogle ScholarPubMed
Flaherman, V. and Rutherford, G. W. (2006). A meta-analysis of the effect of high weight on asthma. Archives of Disease in Childhood, 91, 334–9.CrossRefGoogle ScholarPubMed
Flegal, K. M., Carroll, M. D., Kuczmarski, R. J. and Johnson, C. (1998). Overweight and obesity in the United States: prevalence and trends, 1960–1994. International Journal of Obesity, 22, 39–47.CrossRefGoogle Scholar
Fleming, J. O. and Cook, T. D. (2006). Multiple sclerosis and the hygiene hypothesis. Neurology, 67, 2085–6.CrossRefGoogle ScholarPubMed
Fletcher, E. S., Rugg-Gunn, A. J., Matthews, J. N. S.et al. (2004). Changes over 20 years in macronutrient intake and body mass index in 11- to 12-year-old adolescents living in Northumberland. British Journal of Nutrition, 92, 321–33.CrossRefGoogle ScholarPubMed
Fogel, R. W. and Costa, D. L. (1997). A theory of technophysio evolution, with some implications for forecasting population, health care costs, and pension costs. Demography, 34, 49–66.CrossRefGoogle ScholarPubMed
Foley, R. A. (1993). The influence of seasonality on hominid evolution. In Seasonality and Human Ecology, ed. Ulijaszek, S. J. and Strickland, S. S.. Cambridge: Cambridge University Press, pp. 17–37.CrossRefGoogle Scholar
Foley, R. A. (1996). The adaptive legacy of human evolution: A search for the environment of evolutionary adaptedness. Evolutionary Anthropology, 4, 194–203.CrossRefGoogle Scholar
Foley, R. A. and Lee, P. C. (1991). Ecology and energetics of encephalisation in hominid evolution. Philosophical Transactions of the Royal Society of London B, 334, 223–32.CrossRefGoogle Scholar
Ford, E. S., Giles, W. H. and Dietz, W. H. (2002). Prevalence of the metabolic syndrome among US adults: findings from the Third National Health and Nutrition Examination Survey. Journal of the American Medical Association, 287, 356–9.CrossRefGoogle ScholarPubMed
Forouhi, N. and Sattar, N. (2006). CVD risk factors and ethnicity – a homogeneous relationship?Atherosclerosis Supplements, 7, 11–19.CrossRefGoogle ScholarPubMed
Fox, C. S., Pencina, M. J., Meigs, J. B., Vasan, R. S., Levitzky, Y. S. and D'Agostino, R. B. (2006). Trends in the incidence of type 2 diabetes mellitus from the 1970s to the 1990s. Circulation, 113, 2914–18.CrossRefGoogle ScholarPubMed
Friedman, J. M. (2004). Modern science versus the stigma of obesity. Nature Medicine, 10, 563–9.CrossRefGoogle ScholarPubMed
Friedman, N. J. and Zeiger, R. S. (2005). The role of breast-feeding in the development of allergies and asthma. Journal of Allergy and Clinical Immunology, 115, 1238–48.CrossRefGoogle ScholarPubMed
Frisch, R. E., Wyshak, G., Albright, N. L.et al. (1985). Lower prevalence of breast cancer and cancers of the reproductive system among former college athletes compared to non-athletes. British Journal of Cancer, 52, 885–91.CrossRefGoogle ScholarPubMed
Frost, G. and Dornhorst, A. (2001). Starting the day the right way. The Lancet, 357, 736–7.CrossRefGoogle ScholarPubMed
Fullerton, S. M., Bartoszewicz, A., Ybazeta, G.et al. (2002). Geographic and haplotype structure of putative type 2 diabetes susceptibility variants at the Calpain-10 locus. American Journal of Human Genetics, 70, 1096–106.CrossRefGoogle Scholar
Furberg, A. S., Jasieńska, G., Bjurstam, N.et al. (2005). Metabolic and hormonal profiles: HDL cholesterol as a plausible biomarker of breast cancer risk. The Norwegian EBBA study. Cancer Epidemiology, Biomarkers and Prevention, 14, 33–40.Google ScholarPubMed
Fuster, V., Badimon, L., Badimon, J. J. and Chesebro, J. H. (1992a). The pathogenesis of coronary artery disease and the acute coronary syndromes: Part I. New England Journal of Medicine, 326, 242–50.Google Scholar
Fuster, V., Badimon, L., Badimon, J. J. and Chesebro, J. H. (1992b). The pathogenesis of coronary artery disease and the acute coronary syndromes: Part II. New England Journal of Medicine, 326, 310–18.Google Scholar
Gage, T. B. (2005). Are modern environments really bad for us?: revisiting the demographic and epidemiologic transitions. Yearbook of Physical Anthropology, 48, 96–117.CrossRefGoogle Scholar
Gale, E. A. M. (2002). A missing link in the hygiene hypothesis?Diabetologia, 45, 588–94.CrossRefGoogle ScholarPubMed
Gallou-Kabani, C. and Junien, C. (2005). Nutritional epigenomics of metabolic syndrome. Diabetes, 54, 1899–906.CrossRefGoogle ScholarPubMed
Galloway, A. (1997). The cost of reproduction and the evolution of postmenopausal osteoporosis. In The Evolving Female: A Life-History Perspective, ed. Morbeck, M. E., Galloway, A. and Zihlman, A. L.. Princeton: Princeton University Press, pp. 132–46.Google Scholar
Gandy, M. and Zumla, A. (2002). The resurgence of disease: social and historical perspectives on the ‘new’ tuberculosis. Social Science and Medicine, 55, 385–96.CrossRefGoogle ScholarPubMed
Gann, P. H., Hennekens, C. H., Ma, J., Longcope, C. and Stampfer, M. J. (1996). Prospective study of sex hormone levels and risk of prostate cancer. Journal of the National Cancer Institute, 88, 1118–26.CrossRefGoogle ScholarPubMed
Gapstur, S. M., Gann, P. H., Kopp, P., Colangelo, L., Longcope, C. and Liu, K. (2002). Serem androgen concentrations in young men: a longitudinal analysis of associations with age, obesity, and race. The CARDIA Male Hormone study. Cancer Epidemiology, Biomarkers and Prevention, 11, 1041–7.Google Scholar
Gardell, B. (1987). Efficiency and health hazards in mechanized work. In Work Stress, ed. Quick, J. C., Bhagat, R., Dalton, J. and Quick, J. D.. New York: Praeger, pp. 50–71.Google Scholar
Gaulin, S. J. C. (1980). Sexual dimorphism in the human post-reproductive life-span: possible causes. Journal of Human Evolution, 9, 227–32.CrossRefGoogle Scholar
Gibson, M. A. and Mace, R. (2005). Helpful grandmothers in rural Ethiopia: A study of the effect of kin on child survival and growth. Evolution and Human Behavior, 26, 469–82.CrossRefGoogle Scholar
Glassman, A. H., Helzer, J. E., Covey, L. S., Stetner, F., Tipp, J. E. and Johnson, J. (1990). Smoking, smoking cessation, and major depression. Journal of the American Medical Association, 264, 1546–9.CrossRefGoogle ScholarPubMed
Gleibermann, L. (1973). Blood pressure and dietary salt in human populations. Ecology of Food and Nutrition, 2, 143–56.CrossRefGoogle Scholar
Gluckman, P. and Hanson, M. (2005). The Fetal Matrix: Evolution, Development and Disease. Cambridge: Cambridge University Press.Google Scholar
Goldin, B. R., Adlercreutz, H., Gorbach, S. L.et al. (1986). The relationship between estrogen levels and diets of Caucasian American and Oriental immigrant women. American Journal of Clinical Nutrition, 44, 945–53.CrossRefGoogle ScholarPubMed
Gould, D. C., Petty, R. and Jacobs, H. S. (2000). The male menopause - does it exist?British Medical Journal, 320, 858–61.CrossRefGoogle ScholarPubMed
Gracey, M. and Spargo, R. M. (1987). The state of health of Aborigines in the Kimberley region. Medical Journal of Australia, 146, 200–4.Google ScholarPubMed
Gray, P. B., Kruger, A., Huisman, H. M., Wissing, M. P. and Vorster, H. H. (2006). Predictors of South African male testosterone levels: the THUSA study. American Journal of Human Biology, 18, 123–32.CrossRefGoogle ScholarPubMed
Gray-Donald, K., Jacobs-Starkey, L. and Johnson-Down, L. (2000). Food habits of Canadians: reduction in fat intake over a generation. Canadian Journal of Public Health, 91, 381–5.Google ScholarPubMed
Greenwood, D. C., Muir, K. R., Packham, C. J. and Madeley, R. J. (1996). Coronary heart disease: a review of the role of psychosocial stress and social support. Journal of Public Health Medicine, 18, 221–31.CrossRefGoogle ScholarPubMed
Griffiths, K., Denis, L., Turkes, A. and Morton, M. S. (1998). Phytoestrogens and diseases of the prostate gland. Baillière's Clinical Endocrinology and Metabolism, 12, 625–47.CrossRefGoogle ScholarPubMed
Gröland, M. M., Lehtonen, O. -P., Eerola, E. and Kero, P. (1999). Fecal microflora in healthy infants born by different methods of delivery: permanent changes in intestinal flora after Cesarian delivery. Journal of Pediatric Gastroenterology and Nutrition, 28, 19–25.CrossRefGoogle Scholar
Gross, L. S., Ford, E. S. and Liu, S. (2004). Increased consumption of refined carbohydrates and the epidemic of type 2 diabetes in the United States: an ecologic assessment. American Journal of Clinical Nutrition, 79, 774–9.CrossRefGoogle ScholarPubMed
Grossman, H., Bergmann, C. and Parker, S. (2006). Dementia: a brief review. Mount Sinai Journal of Medicine, 73, 985–92.Google ScholarPubMed
Grundy, J., Matthews, S., Bateman, B., Dean, T. and Arshad, S. H. (2002). Rising prevalence of allergy to peanuts in children: data from two sequential cohorts. Journal of Allergy and Clinical Immunology, 110, 784–9.CrossRefGoogle Scholar
Gu, K., Cowie, C. C. and Harris, M. I. (1998). Mortality in adults with and without diabetes in a national cohort of the US population, 1971–1993. Diabetes Care, 21, 1138–45.CrossRefGoogle Scholar
Guarner, F., Bourdet-Sicard, R., Brandtzaeg, P.et al. (2006). Mechanisms of disease: the hygiene hypothesis. Nature Clinical Practice Gastroenterology and Hepatology, 3, 275–84.CrossRefGoogle ScholarPubMed
Guest, C. S., O'Dea, K., Hopper, J. L. and Larkins, R. G. (1993). Hyperinsulinaemia and obesity in Aborigines of south-eastern Australia, with comparisons from rural and urban Europid populations. Diabetes Research and Clinical Practice, 20, 155–64.CrossRefGoogle Scholar
Guise, J. M., Austin, D. and Morris, C. D. (2005). Review of case-control studies related to breastfeeding and reduced risk of childhood leukemia. Pediatrics, 116, e724–31.CrossRefGoogle ScholarPubMed
Guyton, A. C. (1986). Textbook of Medical Physiology. Philadelphia: WB Saunders.Google Scholar
Haffner, S. M. (2003). Insulin resistance, inflammation, and the prediabetic state. American Journal of Cardiology, 92 (Suppl.), 18J–26J.CrossRefGoogle ScholarPubMed
Hagen, E. H. (1999). The functions of postpartum depression. Evolution and Human Behavior, 20, 325–59.CrossRefGoogle Scholar
Hajat, S., Haines, A. P., Atkinson, R. W., Bremner, S., Anderson, H. R. and Emberlin, J. (2001). Association between air pollution and daily consultations with General Practitioners for allergic rhinitis in London, United Kingdom. American Journal of Epidemiology, 153, 704–14.CrossRefGoogle ScholarPubMed
Hales, C. N. and Barker, D. J. P. (1992). Type 2 (non-insulin-dependent) diabetes mellitus: the thrifty phenotype hypothesis. Diabetologia, 35, 595–601.CrossRefGoogle ScholarPubMed
Hales, C. N. and Barker, D. J. P. (2001). The thrifty phenotype hypothesis. British Medical Bulletin, 60, 5–20.CrossRefGoogle ScholarPubMed
Hall, W. D., Clark, L. T., Wenger, N. H.et al. (2003). The metabolic syndrome in African Americans: a review. Ethnicity and Disease, 13, 414–28.Google ScholarPubMed
Hamilton, G. (2005). Filthy friends. New Scientist, 2495, 35–9.Google Scholar
Hankinson, S. E., Willett, W. C., Manson, J. E.et al. (1995). Alcohol, height, and adiposity in relation to estrogen and prolactin levels in postmenopausal women. Journal of the National Cancer Institute, 87, 1297–302.CrossRefGoogle ScholarPubMed
Hannaford, P. C., Selvaraj, S., Elliott, A. M., Angus, V., Iversen, L. and Lee, A. J. (2007). Cancer risk among users of oral contraceptives: cohort data from the Royal College of General Practioners' oral contraception study. British Medical Journal, 335, 651–4A.CrossRefGoogle Scholar
Hanson, R. L., Ehm, M. G., Pettitt, D. J.et al. (1998). An autosomal genomic scan for loci linked to type II diabetes mellitus and body-mass index in Pima Indians. American Journal of Human Genetics, 63, 1130–8.CrossRefGoogle ScholarPubMed
Harder, T., Bergmann, R., Kallischnigg, G. and Plagemann, A. (2005). Duration of breastfeeding and risk of overweight: A meta-analysis. American Journal of Epidemiology, 162, 397–403.CrossRefGoogle ScholarPubMed
Harder, T., Rodekamp, E., Schellong, K., Duderhausen, J. W., Plagemann, A. (2007). Birth weight and subsequent risk of type 2 diabetes: a meta-analysis. American Journal of Epidemiology, 165, 849–57.CrossRefGoogle ScholarPubMed
Hardy, R. and Kuh, D. (2002). Change in psychological and vasomotor symptom reporting during the menopause. Social Science and Medicine, 55, 1975–88.CrossRefGoogle ScholarPubMed
Harman, S. M., Metter, E. J., Tobin, J. D., Pearson, J. and Blackman, M. R. (2001). Longitudinal effects of aging on serum total and free testosterone levels in healthy men. Journal of Clinical Endocrinology and Metabolism, 86, 724–31.CrossRefGoogle ScholarPubMed
Harris, D. R. and Hillman, G. C., eds. (1989). Foraging and Farming: The Evolution of Plant Exploitation. London: Unwin Hyman.Google Scholar
Harris, M. I., Goldstein, D. E., Flegal, K. M.et al. (1998). Prevalence of diabetes, impaired fasting glucose, and impaired glucose tolerance in US adults. Diabetes Care, 21, 518–24.CrossRefGoogle Scholar
Harris, R., Tobias, M., Jeffreys, M., Waldegrave, K., Karlsen, S. and Nazroo, J. (2006). Effects of self-reported racial discrimination and deprivation on Māori health and inequalities in New Zealand: cross-sectional study. The Lancet, 367, 2005–9.CrossRefGoogle ScholarPubMed
Harris, S. B., Gittelsohn, J., Hanley, A. J. G.et al. (1997). The prevalence of NIDDM and associated risk factors in Native Canadians. Diabetes Care, 20, 185–7.CrossRefGoogle ScholarPubMed
Harrison, G. A. (1973). The effects of modern living. Journal of Biosocial Science, 5, 217–28.Google ScholarPubMed
Harrison, G. A., Tanner, J. A., Pilbeam, D. R. and Baker, P. T. (1988). Human Biology: An Introduction to Human Evolution, Variation, Growth, and Adaptability. Oxford: Oxford University Press.Google Scholar
Hart, R., Hickey, M. and Franks, S. (2004). Definitions, prevalence and symptoms of polycystic ovaries and polycystic ovary syndrome. Best Practice and Research in Clinical Obstetrics and Gynaecology, 18, 671–683.CrossRefGoogle ScholarPubMed
Harvie, M., Hooper, L. and Howell, A. H. (2003). Central obesity and breast cancer risk: a systematic review. Obesity Reviews, 4, 157–73.CrossRefGoogle ScholarPubMed
Hawkes, K. (2003). Grandmothers and the evolution of human longevity. American Journal of Human Biology, 15, 380–400.CrossRefGoogle ScholarPubMed
He, F. J., Nowson, C. A. and MacGregor, G. A. (2006). Fruit and vegetable consumption and stroke: meta-analysis of cohort studies. The Lancet, 367, 320–6.CrossRefGoogle ScholarPubMed
Hegele, R. A. and Bartlett, L. C. (2003). Genetics, environment and type 2 diabetes in the Oji-Cree population of northern Ontario. Canadian Journal of Diabetes, 27, 256–61.Google Scholar
Hegele, R. A., Cao, H., Hanley, A. J. G., Zinman, B., Harris, S. B. and Anderson, C. M. (2000). Clinical utility of HNF1A genotyping for diabetes in Aboriginal Canadians. Diabetes Care, 23, 775–8.CrossRefGoogle ScholarPubMed
Heim, C., Ehlert, U. and Hellhammer, D. H. (2000). The potential role of hypocortisolism in the pathophysiology of stress-related bodily disorders. Psychoneuroendocrinology, 25, 1–35.CrossRefGoogle ScholarPubMed
Heinig, M. J. (2001). Host defense benefits of breastfeeding for the infant. Pediatric Clinics of North America, 48, 105–23.CrossRefGoogle ScholarPubMed
Heinig, M. J. and Dewey, K. G. (1997). Health effects of breast feeding for mothers: a critical review. Nutrition Research Reviews, 10, 35–56.CrossRefGoogle ScholarPubMed
Helman, C. G. (1994). Culture, Health and Illness. Oxford: Butterworth-Heinemann.Google Scholar
Helmchen, L. A. and Henderson, R. M. (2004). Changes in the distribution of body mass index of white US men, 1890–2000. Annals of Human Biology, 31, 174–81.CrossRefGoogle ScholarPubMed
Hemingway, H. and Marmot, M. (1999). Psychosocial factors in the aetiology and prognosis of coronary heart disease: systematic review of prospective cohort studies. British Medical Journal, 318, 1460–7.CrossRefGoogle ScholarPubMed
Henderson, B. E., Ross, R. K., Pike, M. C. and Cassagrande, J. T. (1982). Endogenous hormones as a major factor in human cancer. Cancer Research, 42, 3232–9.Google ScholarPubMed
Hendrix, S. L., Wassertheil-Smoller, S., Johnson, K. C.et al. for the WHI Investigators (2006). Effects of conjugated equine estrogen on stroke in the Women's Health Initiative. Circulation, 113, 2425–34.CrossRefGoogle ScholarPubMed
Henry, C. J. K., Lightowler, H. J. and Al-Hourani, H. M. (2004). Physical activity and levels of inactivity in adolescent females aged 11–16 years in the United Arab Emirates. American Journal of Human Biology, 16, 346–53.CrossRefGoogle Scholar
Herbst, K. L. and Bhasin, S. (2004). Testosterone action on skeletal muscle: anabolic and catabolic signals. Current Opinion in Clinical Nutrition and Metabolic Care, 7, 271–7.CrossRefGoogle Scholar
Heuveline, P., Guillot, M. and Gwatkin, D. R. (2002). The uneven tides of the health transition. Social Science and Medicine, 55, 313–22.CrossRefGoogle ScholarPubMed
Hill, K. and Hurtado, A. M. (1991). The evolution of premature reproductive senescence and menopause in human females: an evaluation of the ‘grandmother hypothesis’. Human Nature, 2, 313–50.CrossRefGoogle Scholar
Hill, K. and Hurtado, A. M. (1996). Ache Life History. New York: Aldine de Gruyter.Google Scholar
Hill, K. and Kaplan, H. (1999). Life history traits in humans: theory and empirical studies. Annual Review of Anthropology, 28, 397–430.CrossRefGoogle Scholar
Hoffjan, S., Nicolae, D. and Ober, C. (2003). Association studies for asthma and atopic diseases: a comprehensive review of the literature. Respiratory Research, 4, 14.CrossRefGoogle ScholarPubMed
Holland, T. D. and O'Brien, M. J. (1997). Parasites, porotic hyperostosis, and the implications of changing perspectives. American Antiquity, 62, 183–93.CrossRefGoogle Scholar
Holman, R. C., Curns, A. T., Kaufman, S. F., Cheek, J. E., Pinner, R. W. and Schonberger, L. B. (2001). Trends in infectious disease hospitalizations among American Indians and Alaska Natives. American Journal of Public Health, 91, 425–31.Google ScholarPubMed
Horikawa, Y., Oda, N., Cox, N. J.et al. (2000). Genetic variation in the gene encoding calpain-10 is associated with type 2 diabetes mellitus. Nature Genetics, 26, 163–75.CrossRefGoogle ScholarPubMed
Howard, B. V., Lee, E. T., Cowan, L. D.et al. (1999). Rising tide of cardiovascular disease in American Indians. Circulation, 99, 2389–95.CrossRefGoogle ScholarPubMed
Hrdy, S. B. (2000). Mother Nature. London: Vintage.Google Scholar
Hsing, A. W. (1996). Hormones and prostate cancer: where do we go from here?Journal of the National Cancer Institute, 88, 1093–5.CrossRefGoogle Scholar
Hsing, A. W. and Devesa, S. S. (2001). Trends and patterns of prostate cancer: What do they suggest?Epidemologic Reviews, 23, 3–13.CrossRefGoogle ScholarPubMed
Hsing, A. W., Gao, Y. T., Chua, S., Deng, J. and Stanczyk, F. Z. (2003). Insulin resistance and prostate cancer risk. Journal of the National Cancer Institute, 95, 67–71.CrossRefGoogle ScholarPubMed
Hudson, J. I., Hiripi, E., Pope, H. G. and Kessler, R. C. (2007). The prevalence and correlates of eating disorders in the National Comorbidity Survey Replication. Biological Psychiatry, 61, 348–58.CrossRefGoogle ScholarPubMed
Hulley, S., Grady, D., Bush, T.et al. (1998). Randomized trial of estrogen plus progestin for secondary prevention of coronary heart disease in postmenopausal women. Journal of the American Medical Association, 280, 605–13.CrossRefGoogle ScholarPubMed
Hurtado, A. M., Arenas, I. and Hill, K. (1996). Evolutionary contexts of chronic allergic disease: the Hiwi of Venezuala. Human Nature, 8, 1–20.Google Scholar
Hussain, A., Rahim, M. A., Khan, Azad A. K., Ali, S. M. K. and Vaaler, S. (2005). Type 2 diabetes in rural and urban populations: diverse prevalence and associated risk factors in Bangladesh. Diabetic Medicine, 22, 931–6.CrossRefGoogle Scholar
Huxley, R., Neil, A. and Collins, R. (2002). Unravelling the fetal origins hypothesis: is there really an inverse association between birthweight and subsequent blood pressure?The Lancet, 360, 659–65.CrossRefGoogle ScholarPubMed
Huxley, R., Owens, J. F., Whincup, P. H., Cook, D. G., Colman, S. and Collins, R. (2004). Birth weight and subsequent cholesterol levels: exploration of the “fetal origins” hypothesis. Journal of the American Medical Association, 292, 2755–64.CrossRefGoogle ScholarPubMed
Huyghe, E., Matsuda, T. and Thonneau, P. (2003). Increasing incidence of testicular cancer worldwide: a review. Journal of Urology, 170, 5–11.CrossRefGoogle ScholarPubMed
Iglowstein, I., Jenni, O. G., Molinari, L. and Largo, R. H. (2003). Sleep duration from infancy to adolescence: reference values and generational trends. Pediatrics, 111, 302–7.CrossRefGoogle ScholarPubMed
Irons, W. (1998). Adaptively relevant environments versus the environment of evolutionary adaptedness. Evolutionary Anthropology, 6, 198–204.3.0.CO;2-B>CrossRefGoogle Scholar
ISAAC Steering Committee (1998). Worldwide variation in the prevalence of symptoms of asthma, allergic rhinoconjunctivitis, and atopic eczema: ISAAC. The Lancet, 351, 1225–32.CrossRef
Jackson, M. A., Kovi, J., Heshmat, M. Y.et al. (1980). Characterization of prostatic carcinoma among blacks: A comparison between a low-incidence area, Ibadan, Nigeria, and a high-incidence area, Washington, DC. Prostate, 1, 185–205.CrossRefGoogle Scholar
James, G. D., Jenner, D. A., Harrison, G. A. and Baker, P. T. (1985). Differences in catecholamine excretion rates, blood pressure and lifestyle among young Western Samoan men. Human Biology, 57, 635–47.Google ScholarPubMed
James, G. D., Crews, D. E. and Pearson, J. (1989). Catecholamines and stress. In Human Population Biology, ed. Little, M. A. and Haas, J. D.. Oxford: Oxford University Press.Google Scholar
James, G. D., Broege, P. A. and Schlussel, Y. R. (1996). Assessing cardiovascular risk and stress-related blood pressure variability in young women employed in waged jobs. American Journal of Human Biology, 8, 743–9.3.0.CO;2-U>CrossRefGoogle Scholar
James, W. P. T. (2002). Will feeding mothers prevent the Asian metabolic syndrome epidemic?Asia Pacific Journal of Clinical Nutrition, 11 (Suppl.), S516–23.CrossRefGoogle ScholarPubMed
James, W. P. T. (2005). Assessing obesity: are ethnic differences in body mass index and waist classification criteria justified?Obesity Reviews, 6, 179–81.CrossRefGoogle ScholarPubMed
Janes, C. R. (1990). Migration, changing gender roles and stress: the Samoan case. Medical Anthropology, 12, 145–67.CrossRefGoogle ScholarPubMed
Jasieńska, G. and Ellison, P. T. (1998). Physical work causes suppression of ovarian function in women. Proceedings of the Royal Society of London B, 265, 1847–51.CrossRefGoogle ScholarPubMed
Jasieńska, G. and Ellison, P. T. (2004). Energetic factors and seasonal changes in ovarian function in women from rural Poland. American Journal of Human Biology, 16, 563–80.CrossRefGoogle ScholarPubMed
Jasieńska, G. and Thune, I. (2001a). Lifestyle, hormones, and risk of breast cancer. British Medical Journal, 322, 586–7.CrossRefGoogle Scholar
Jasieńska, G. and Thune, I. (2001b). Lifestyle, progesterone, and risk of breast cancer – reply. British Medical Journal, 323, 1002.Google Scholar
Jasieńska, G., Thune, I. and Ellison, P. T. (2000). Energetic factors, ovarian steroids and the risk of breast cancer. European Journal of Cancer Prevention, 9, 231–9.CrossRefGoogle ScholarPubMed
Jasieńska, G., Thune, I. and Ellison, P. T. (2006a). Fatness at birth predicts adult susceptibility to ovarian suppression: an empirical test of the Predictive Adaptive Response hypothesis. Proceedings of the National Academy of Science, 103, 12759–62.CrossRefGoogle Scholar
Jasieńska, G., Ziomkiewicz, A., Thune, I., Lipson, S. F. and Ellison, P. T. (2006b). Habitual physical activity and estradiol levels in women of reproductive age. European Journal of Cancer Prevention, 15, 439–45.CrossRefGoogle Scholar
Jenike, M. R. (2001). Nutritional ecology: diet, physical activity and body size. In Hunter–Gatherers: An Interdisciplinary Perspective, ed. Panter-Brick, C., Layton, R. H. and Rowley-Conwy, P.. Cambridge: Cambridge University Press, pp. 205–39.Google Scholar
Jenner, D. A., Reynolds, V. and Harrison, G. A. (1980). Catecholamine excretion rates and occupation. Ergonomics, 23, 237–46.CrossRefGoogle ScholarPubMed
Jenner, D. A., Harrison, G. A. and Prior, I. A. M. (1987). Catecholamine excretion in Tokelauans living in three different environments. Human Biology, 59, 165–72.Google ScholarPubMed
Jin, B., Turner, L., Zhou, Z., Zhou, E. L. and Handelsman, D. J. (1999). Ethnicity and migration as determinants of human prostate size. Journal of Clinical Endocrinology and Metabolism, 84, 3613–19.Google ScholarPubMed
Jonas, B. S., Franks, P. and Ingram, D. D. (1997). Are symptoms of anxiety and depression risk factors for hypertension? Longitudinal evidence from the National Health and Nutrition Examination Survey I: epidemiological follow-up study. Archives of Family Medicine, 6, 43–9.CrossRefGoogle Scholar
Julian, D. G. and Cowan, J. C. (1992). Cardiology. London: Baillière Tindall.Google Scholar
Kaaks, R., Lukanova, A. and Kurzer, M. S. (2002). Obesity, endogenous hormones, and endometrial cancer risk: a synthetic review. Cancer Epidemiology, Biomarkers and Prevention, 11, 1531–43.Google ScholarPubMed
Kajantie, E. and Phillips, D. I. W. (2006). The effects of sex and hormonal status on the physiological response to acute psychosocial stress. Psychoneuroendocrinology, 31, 151–78.CrossRefGoogle ScholarPubMed
Kaliora, A. C., Dedoussis, G. V. Z. and Schmidt, H. (2006). Dietary antioxidants in preventing atherogenesis. Atherosclerosis, 187, 1–17.CrossRefGoogle ScholarPubMed
Kalliomäki, M., Salminen, S., Arvilommi, H., Kero, P., Koskinen, P. and Isolauri, E. (2001). Probiotics in primary prevention of atopic disease: a randomised placebo-controlled trial. The Lancet, 357, 1076–9.CrossRefGoogle ScholarPubMed
Kannel, W. B. (1987). New perspectives on cardiovascular risk factors. American Heart Journal, 114, 213–19.CrossRefGoogle ScholarPubMed
Kant, A. K. and Graubard, B. I. (2004). Eating out in America, 1987–2000: trends and nutritional correlates. Preventive Medicine, 38, 243–9.CrossRefGoogle ScholarPubMed
Karasek, R. A. (1979). Job demands, job decision latitude, and mental strain: implications for job redesign. Administrative Science Quarterly, 24, 285–308.CrossRefGoogle Scholar
Kaufman, J. S. and Hall, S. A. (2003). The slavery hypertension hypothesis: dissemination and appeal of a modern race theory. Epidemiology, 14, 111–18.CrossRefGoogle ScholarPubMed
Kaur, H., Choi, W. S., Mayo, M. S. and Harris, K. J. (2003). Duration of television watching is associated with increased body mass index. Journal of Pediatrics, 143, 506–11.CrossRefGoogle ScholarPubMed
Keighley, E. D., McGarvey, S. T., Quested, C., McCuddin, C., Viali, S. and Maga, U. A. (2007). Nutrition and health in modernizing Samoans: temporal trends and adaptive perspectives. In Health Changes in the Asia-Pacific Region: Biocultural and Epidemiological Approaches, ed. Ohtsuka, R. and Ulijaszek, S. J.. Cambridge: Cambridge University Press, pp. 147–91.CrossRefGoogle Scholar
Keil, J. E., Sutherland, S. E., Knapp, R. G. and Tyroler, H. A. (1992). Does equal socioeconomic status in black and white men mean equal risk of mortality?American Journal of Public Health, 82, 1133–6.CrossRefGoogle ScholarPubMed
Kemp, A. and Kakakios, A. (2004). Asthma prevention: breast is best?Journal of Paediatrics and Child Health, 40, 337–9.CrossRefGoogle ScholarPubMed
Kennedy, G. E. (2005). From the ape's dilemma to the weanling's dilemma: early weaning and its evolutionary context. Journal of Human Evolution, 48, 123–45.CrossRefGoogle ScholarPubMed
Kent, S. (1986). The influence of sedentism and aggregation on porotic hyperostosis and anemia: a case study. Man, 21, 605–36.CrossRefGoogle Scholar
Kero, J., Gissler, M., Hemminki, E. and Isolauri, E. (2001). Could T(h)1 and T(h)2 diseases coexist? Evaluation of asthma incidence in children with coeliac disease, type 1 diabetes, or rheumatoid arthritis: a register study. Journal of Allergy and Clinical Immunology, 108, 781–3.CrossRefGoogle ScholarPubMed
Key, T. J. A., Chen, J., Wang, D. Y., Pike, M. C. and Boreham, J. (1990). Sex hormones in women in rural China and in Britain. British Journal of Cancer, 62, 631–6.CrossRefGoogle ScholarPubMed
Kim, S. Y., Moon, S. and Popkin, B. M. (2000). The nutrition transition in South Korea. American Journal of Clinical Nutrition, 71, 44–53.CrossRefGoogle ScholarPubMed
King, H., Aubert, R. E. and Herman, W. H. (1998). Global burden of diabetes, 1995–2025. Diabetes Care, 21, 1414–31.CrossRefGoogle Scholar
Kleinman, A. and Good, B. (1985). Introduction: culture and depression. In Culture and Depression: Studies in the Anthropology and Cross-Cultural Psychiatry of Affect and Disorder, ed. Kleinman, A. and Good, B.. Berkeley: University of California Press, pp. 1–33.Google Scholar
Kleinman, A. M. (1977). Depression, somatization and the “new cross-cultural” psychiatry. Social Science and Medicine, 11, 3–10.CrossRefGoogle ScholarPubMed
Klerman, G. L. and Weissman, M. M. (1989). Increasing rates of depression. Journal of the American Medical Association, 261, 2229–35.CrossRefGoogle Scholar
Kliks, M. M. (1983). Paleoparasitology: on the origins and impact of human-helminth relationships. In Human Ecology and Infectious Disease, ed. Croll, N. A. and Cross, J. H.. New York: Academic Press.Google Scholar
Knowler, W. C. (1978). Diabetes incidence and prevalence in Pima Indians: a 19-fold greater incidence than in Rochester, Minnesota. American Journal of Epidemiology, 108, 497–505.CrossRefGoogle ScholarPubMed
Knowler, W. C., Pettit, D. J., Savage, P. J., et al. (1983). Diabetes mellitus in the Pima Indians: genetic and evolutionary considerations. American Journal of Physical Anthropology, 62, 107–14.CrossRefGoogle ScholarPubMed
Kohen-Avramoglu, R., Theriault, A. and Adeli, K. (2003). Emergence of the metabolic syndrome in childhood: an epidemiological overview and mechanistic link to dyslipidemia. Clinical Biochemistry, 36, 413–20.CrossRefGoogle ScholarPubMed
Kohler, H. P., Billari, F. C. and Ortega, J. A. (2002). The emergence of lowest-low fertility in Europe during the 1990s. Population and Development Review, 28, 641–86.CrossRefGoogle Scholar
Konner, M. and Worthman, C. M. (1980). Nursing frequency, gonadal function, and birth spacing among the!Kung hunter–gatherers. Science, 207, 788–91.CrossRefGoogle Scholar
Kraft, P., Pharoah, P., Chanock, S. J.et al. (2005). Genetic variation in the HSD17B1 gene and risk of prostate cancer. PLoS Genetics, 1, e68.CrossRefGoogle ScholarPubMed
Kramer, M. S. (1987). Determinants of low birth weight: methodological assessment and meta-analysis. Bulletin of the World Health Organization, 65, 663–737.Google ScholarPubMed
Kramer, M. S. and Kakuma, R. (2002). The Optimal Duration of Exclusive Breastfeeding: A Systematic Review. Geneva: World Health Organization.CrossRefGoogle Scholar
Krämer, U., Heinrich, J., Wjst, M. and Wichmann, H. -E. (1999). Age of entry to day nursery and allergy in later childhood. The Lancet, 352, 450–4.CrossRefGoogle Scholar
Krantz, D. S., Contrada, R. J., Hill, R. and Friedler, E. (1988). Environmental stress and biobehavioral antecedents of coronary heart disease. Journal of Consulting and Clinical Psychology, 56, 333–41.CrossRefGoogle ScholarPubMed
Krause, I. -B. (1989). Sinking heart: a Punjabi communication of distress. Social Science and Medicine, 29, 563–75.CrossRefGoogle ScholarPubMed
Kubzansky, L. D. and Kawachi, I. (2000). Going to the heart of the matter: do negative emotions cause coronary heart disease?Journal of Psychosomatic Research, 48, 323–37.CrossRefGoogle ScholarPubMed
Kuh, D. and Hardy, R., eds. (2002). A Life Course Approach to Women's Health. Oxford: Oxford University Press.CrossRefGoogle Scholar
Kuh, D., Ben-Shlomo, Y., Lynch, J., Hallqvist, J. and Power, C. (2003). Life course epidemiology. Journal of Epidemiology and Community Health, 57, 778–83.CrossRefGoogle ScholarPubMed
Kukkonen, K., Savilahti, E., Haahtela, T.et al. (2007). Probiotics and prebiotic galacto-oligosaccharides in the prevention of allergic diseases: a randomized, double-blind, placebo-controlled trial. Journal of Allergy and Clinical Immunology, 119, 192–8.CrossRefGoogle ScholarPubMed
Kuller, L. H., Meilahn, E. N., Cauley, J. A., Gutai, J. P. and Matthews, K. A. (1994). Epidemiologic studies of menopause: changes in risk factors and disease. Experimental Gerontology, 29, 495–509.CrossRefGoogle ScholarPubMed
Künzli, N., McConnel, R., Bates, D.et al. (2003). Breathless in Los Angeles: the exhausting search for clean air. American Journal of Public Health, 93, 1494–99.CrossRefGoogle ScholarPubMed
Kuzawa, C. W. (1998). Adipose tissue in human infancy and childhood: an evolutionary perspective. Yearbook of Physical Anthropology, 41, 177–210.3.0.CO;2-B>CrossRefGoogle Scholar
Kuzawa, C. W. (2005). Fetal origins of developmental plasticity: are fetal cues reliable predictors of future nutritional environments?American Journal of Human Biology, 17, 5–21.CrossRefGoogle ScholarPubMed
Lager, C. and Ellison, P. T. (1990). Effect of moderate weight loss on ovarian function assessed by salivary progesterone measurements. American Journal of Human Biology, 2, 303–12.CrossRefGoogle ScholarPubMed
Lake, A. and Townshend, T. (2006). Obesogenic environments: exploring the built and food environments. Journal of the Royal Society for the Promotion of Health, 126, 262–7.CrossRefGoogle ScholarPubMed
LaMonte, M. J., Barlow, C. E., Jurca, R., Kampert, J. B., Church, T. S. and Blair, S. N. (2005). Cardiorespiratory fitness is inversely associated with the incidence of metabolic syndrome. A prospective study of men and women. Circulation, 112, 505–12.CrossRefGoogle ScholarPubMed
Lancet Editorial (2006). A plea to abandon asthma as a disease concept. The Lancet, 368, 705.CrossRef
Lane, D., Beevers, D. G. and Lip, G. Y. H. (2002). Ethnic differences in blood pressure and the prevalence of hypertension in England. Journal of Human Hypertension, 16, 267–73.CrossRefGoogle ScholarPubMed
Larsen, C. S. (1995). Biological changes in human populations with agriculture. Annual Review of Anthropology, 24, 185–213.CrossRefGoogle Scholar
Larsen, C. S. (2002). Post-Pleistocene human evolution: bioarcheology of the agricultural transition. In Human Diet: Its Origin and Evolution, ed. Ungar, P. S. and Teaford, M. F.. Westport, Connecticut: Bergin and Garvey, pp. 19–36.Google Scholar
Larsen, C. S. (2003). Animal source foods and human health during evolution. Journal of Nutrition, 133, 3893S–7S.CrossRefGoogle ScholarPubMed
Lasker, G. (1969). Human biological adaptability: the ecological approach in physical anthropology. Science, 166, 1480–6.CrossRefGoogle ScholarPubMed
Vecchia, C. and Bosetti, C. (2004). Benefits and risks of oral contraceptives on cancer. European Journal of Cancer Prevention, 13, 467–70.CrossRefGoogle ScholarPubMed
Lawlor, D. A., Smith, Davey G., Leon, D. A., Sterne, J. A. C. and Ebrahim, S. (2002). Secular trends in mortality by stroke subtype in the 20th century: a retrospective analysis. The Lancet, 360, 1818–23.CrossRefGoogle ScholarPubMed
Souëf, P. N., Goldblatt, J. and Lynch, N. R. (2000). Evolutionary adaptation of inflammatory immune responses in human beings. The Lancet, 356, 242–4.CrossRefGoogle ScholarPubMed
Souëf, P. N., Candelaria, P. and Goldblatt, J. (2006). Evolution and respiratory genetics. European Respiratory Journal, 28, 1258–63.CrossRefGoogle ScholarPubMed
Lee, R. B. (1979). The!Kung San: Men, Women and Work in a Foraging Society. Cambridge: Cambridge University Press.Google Scholar
Leidy, L. E. (1999). Menopause in evolutionary perspective. In Evolutionary Medicine, ed. Trevathan, W. R., Smith, E. O. and McKenna, J. J.. New York: Oxford University Press, pp. 407–27.Google Scholar
Leidy Sievert, L. (2001). Aging and reproductive senescence. In Reproductive Ecology and Human Evolution, ed. Ellison, P. T.. New York: Aldine de Gruyter, pp. 267–92.Google Scholar
Sievert, Leidy L. (2006). Menopause: A Biocultural Perspective. New Brunswick, New Jersey: Rutgers University Press.Google Scholar
Leon, D. A., Chenet, L., Shkolnikov, V. M.et al. (1997). Huge variation in Russian mortality rates 1984–94: artefact, alcohol, or what?The Lancet, 350, 383–8.CrossRefGoogle ScholarPubMed
Leonard, W. R. (2000). Human nutritional evolution. In Human Biology: An Evolutionary and Biocultural Perspective, ed. Stinson, S., Bogin, B., Huss-Ashmore, R. and O'Rourke, D.. New York: Wiley-Liss, pp. 295–343.Google Scholar
Lett, H. S., Blumenthal, J. A., Babyak, M. A.et al. (2004). Depression as a risk factor for coronary artery disease: evidence, mechanisms and treatment. Psychosomatic Medicine, 66, 305–15.Google ScholarPubMed
Lewontin, R. C. (1982). Human Diversity. New York: Scientific American Library.Google Scholar
Li, C., Ford, E. S., Mokdad, A. H. and Cook, S. (2006). Recent trends in waist circumference and waist–height ratio among US children and adolescents. Pediatrics, 118, e1390–8.CrossRefGoogle ScholarPubMed
Lieberman, L. S. (2003). Dietary, evolutionary, and modernizing influences on the prevalence of type 2 diabetes. Annual Review of Nutrition, 23, 345–77.CrossRefGoogle ScholarPubMed
Linden, W., Stossel, C. and Maurice, J. (1996). Psychosocial interventions for patients with coronary artery disease: a meta-analysis. Archives of Internal Medicine, 156, 745–52.CrossRefGoogle ScholarPubMed
Lipson, S. F. (2001). Metabolism, maturation, and ovarian function. In Reproductive Ecology and Human Evolution, ed. Ellison, P. T.. New York: Aldine de Gruyter, pp. 235–48.Google Scholar
Lipson, S. F. and Ellison, P. T. (1996). Comparison of salivary steroid profiles in naturally occurring conception and non-conception cycles. Human Reproduction, 11, 2090–6.CrossRefGoogle ScholarPubMed
Liu, S., Willett, W. C., Manson, J. E., Hu, F. B., Rosner, B. and Colditz, G. (2003). Relation between changes in intakes of dietary fiber and grain products and changes in weight and development of obesity among middle-aged women. American Journal of Clinical Nutrition, 78, 920–7.CrossRefGoogle ScholarPubMed
Livingston, C. and Collison, M. (2002). Sex steroids and insulin resistance. Clinical Science, 102, 151–66.CrossRefGoogle Scholar
Lock, M. and Kaufert, P. (2001). Menopause, local biologies, and cultures of aging. American Journal of Human Biology, 13, 494–504.CrossRefGoogle Scholar
Ludwig, D. S. (2002). The glycemic index: physiological mechanisms relating to obesity, diabetes, and cardiovascular disease. Journal of the American Medical Association, 287, 2414–23.CrossRefGoogle Scholar
Ludwig, D. S., Peterson, K. E. and Gortmaker, S. L. (2001). Relation between consumption of sugar-sweetened drinks and childhood obesity: a prospective, observational analysis. The Lancet, 357, 505–8.CrossRefGoogle ScholarPubMed
Macleod, J. and Smith, Davey G. (2003). Psychosocial factors and public health: a suitable case for treatment. Journal of Epidemiology and Community Health, 57, 565–70.CrossRefGoogle Scholar
Maizels, R. M. (2005). Infections and allergy – helminths, hygiene and host immune regulation. Current Opinion in Immunology, 17, 656–61.CrossRefGoogle ScholarPubMed
Mann, J. (2004). Free sugars and human health: sufficient evidence for action?The Lancet, 363, 1068–70.CrossRefGoogle ScholarPubMed
Mann, J. I. (2002). Diet and risk of coronary heart disease and type 2 diabetes. The Lancet, 360, 783–9.CrossRefGoogle ScholarPubMed
Marmot, M., Shipley, M. and Rose, G. (1984). Inequalities in death – specific explanations of a general pattern?The Lancet, i, 1003–6.CrossRefGoogle Scholar
Marshall, C., Hitman, G. A., Partridge, C. J.et al. (2005). Evidence that an isoform of Calpain-10 is a regulator of exocytosis in pancreatic beta-cells. Molecular Endocrinology, 19, 213–24.CrossRefGoogle ScholarPubMed
Martin, R. M., Gunnell, D. and Smith, Davey G. (2005). Breastfeeding in infancy and blood pressure in later life: systematic review and meta-analysis. American Journal of Epidemiology, 161, 15–26.CrossRefGoogle ScholarPubMed
Martinez, F. D. (1994). Role of viral infections in the inception of asthma: could they be protective?Thorax, 49, 1189–91.CrossRefGoogle ScholarPubMed
Martorell, R., Khan, L. K., Hughes, M. L. and Grummer-Strawn, L. M. (1998). Obesity in Latin American women and children. Journal of Nutrition, 128, 1464–73.CrossRefGoogle ScholarPubMed
Mascie-Taylor, C. G. N. (1993). The biological anthropology of disease. In The Anthropology of Disease, ed. Mascie-Taylor, C. G. N.. Oxford: Oxford University Press, pp. 1–72.Google Scholar
Maskarinec, G., Franke, A., Williams, A.et al. (2004). Effects of a 2-year randomized soy intervention on sex hormone levels in premenopausal women. Cancer Epidemiology, Biomarkers and Prevention, 13, 1736–44.Google ScholarPubMed
Masoli, M., Fabian, D., Holt, S. and Beasley, R. for the Global Initiative for Asthma (GINA) Program (2004a). The global burden of asthma: executive summary of the GINA Dissemination Committee Report. Allergy, 59, 469–78.CrossRefGoogle Scholar
Masoli, M., Fabian, D., Holt, S. H. A. and Beasley, R. (2004b). Global Burden of Asthma: Global Initiative for Asthma.
Mathers, C. D. and Loncar, D. (2006). Projections of global mortality and burden of disease from 2002 to 2030. PLoS Medicine, 3, 2011–30.CrossRefGoogle ScholarPubMed
Matricardi, P. M., Bjorksten, B., Bonini, S.et al. for the EAACI Task Force 7 (2003). Microbial products in allergy prevention and therapy. Allergy, 58, 461–71.CrossRefGoogle ScholarPubMed
Matsuda, K., Nishi, Y., Okamatsu, Y., Kojima, M. and Matsuishi, T. (2006). Ghrelin and leptin: a link between obesity and allergy?Journal of Allergy and Clinical Immunology, 117, 705–6.CrossRefGoogle ScholarPubMed
Matthews, K. A. (2005). Psychological perspectives on the development of coronary heart disease. American Psychologist, 60, 783–96.CrossRefGoogle ScholarPubMed
Matthews, K. A., Raikkonen, K., Everson, S. A.et al. (2000). Do the daily experiences of healthy men and women vary according to occupational prestige and work strain?Psychosomatic Medicine, 62, 346–53.CrossRefGoogle ScholarPubMed
Mayor, S. (2005). 23% of babies in England are delivered by caesarean section. British Medical Journal, 330, 806.Google ScholarPubMed
McDaniel, C. N. and Gowdy, J. M. (2000). Paradise for Sale: A Parable of Nature. Berkeley, California: University of California Press.Google Scholar
McElroy, A. and Townsend, P. K. (2004). Medical Anthropology in Ecological Perspective. Boulder, Colorado: Westview Press.Google Scholar
McEwen, B. S. and Stellar, E. (1993). Stress and the individual – mechanisms leading to disease. Archives of Internal Medicine, 153, 2093–110.CrossRefGoogle ScholarPubMed
McEwen, B. S. and Wingfield, J. C. (2003). The concept of allostatis in biology and biomedicine. Hormones and Behavior, 43, 2–15.CrossRefGoogle Scholar
McGarvey, S. T., Bindon, J. R., Crews, D. E. and Schendel, D. E. (1989). Modernization and adiposity: causes and consequences. In Human Population Biology, ed. Little, M. A. and Haas, J. D.. Oxford: Oxford University Press, pp. 263–80.Google Scholar
McGee, R., Williams, S. and Elwood, M. (1994). Depression and the development of cancer – a meta-analysis. Social Science and Medicine, 1994, 187–92.CrossRefGoogle Scholar
McKeigue, P. M. (1996). Metabolic consequences of obesity and body fat pattern: lessons from migrant studies. In The Origins and Consequences of Obesity. CIBA Foundation Symposium 201, ed. Chadwick, D. J. and Cardew, G.. Chichester: John Wiley, pp. 54–63.CrossRefGoogle Scholar
McKeigue, P. M., Shah, B. and Marmot, M. G. (1991). Relation of central obesity and insulin resistance with high diabetes prevalence and cardiovascular risk in South Asians. The Lancet, 337, 382–6.CrossRefGoogle ScholarPubMed
McKenna, J. J., Mosko, S. and Richard, C. (1999). Breast-feeding and mother–infant cosleeping in relation to SIDS prevention. In Evolutionary Medicine, ed. Trevathan, W. R., Smith, E. O. and McKenna, J. J.. New York: Oxford University Press, pp. 53–74.Google Scholar
McMichael, A. J. (2001). Human Frontiers, Environments and Disease. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
McPherson, K., Steel, C. M. and Dixon, J. M. (2000). Breast cancer – epidemiology, risk factors, and genetics. British Medical Journal, 321, 624–8.CrossRefGoogle ScholarPubMed
Meek, J. Y. (2001). Breastfeeding in the workplace. Pediatric Clinics of North America, 48, 461–74.CrossRefGoogle ScholarPubMed
Menacker, F. (2005). Trends in Cesarean Rates for First Births and Repeat Cesarean Rates for Low Risk Women: United States, 1990–2000. National Vital Statistics Report 54: 1–12. Hyattsville, Maryland: National Center for Health Statistics.
Mendelsohn, M. E. and Karas, R. H. (1999). The protective effects of estrogen on the cardiovascular system. New England Journal of Medicine, 340, 1801–11.CrossRefGoogle ScholarPubMed
Meyer, V. F. (2001). The medicalization of the menopause: Critique and consequences. International Journal of Health Services, 31, 769–92.CrossRefGoogle ScholarPubMed
Midthjell, K., Kruger, O., Holmen, J.et al. (1999). Rapid changes in the prevalence of obesity and known diabetes in an adult Norwegian population – the Nord-Trondelag Health Surveys: 1984–1986 and 1995–1997. Diabetes Care, 22, 1813–20.CrossRefGoogle Scholar
Miller, W. C., Koceja, D. M. and Hamilton, E. J. (1997). A meta-analysis of the past 25 years of weight loss research using diet, exercise or diet plus exercise intervention. International Journal of Obesity and Related Metabolic Disorders, 21, 941–47.CrossRefGoogle ScholarPubMed
Misra, A. and Vikram, N. K. (2004). Insulin resistance syndrome (metabolic syndrome) and obesity in Asian Indians: evidence and implications. Nutrition, 20, 482–91.CrossRefGoogle ScholarPubMed
Mokdad, A. H., Serdula, M. K., Dietz, W. H., Bowman, B. A., Marks, J. S. and Koplan, J. P. (1999). The spread of the obesity epidemic in the United States, 1991–1998. Journal of the American Medical Association, 282, 1519–22.CrossRefGoogle ScholarPubMed
Moore, S. E., Halsall, I., Howarth, D., Poskitt, E. M. E. and Prentice, A. M. (2001). Glucose, insulin and lipid metabolism in rural Gambians exposed to early malnutrition. Diabetic Medicine, 18, 645–53.CrossRefGoogle ScholarPubMed
Muehlenbein, M. P. and Bribiescas, R. G. (2005). Testosterone-mediated immune functions and male life histories. American Journal of Human Biology, 17, 527–58.CrossRefGoogle ScholarPubMed
Mufunda, J., Chatora, R., Ndambukuwa, Y.et al. (2006). Emerging non-communicable disease epidemic in Africa: preventive measures from the WHO Regional Office for Africa. Ethnicity and Disease, 16, 521–6.Google ScholarPubMed
Murch, S. H. (2001). Toll of allergy reduced by probiotics. The Lancet, 357, 1057–9.CrossRefGoogle ScholarPubMed
Murray, C. J. L. and Lopez, A. D. (1997). Alternative projections of mortality and disability by cause 1990–2020: global burden of disease study. The Lancet, 349, 1498–504.CrossRefGoogle ScholarPubMed
Murray, C. J. L., Lauer, J. A., Hutubessy, R. C. W.et al. (2003). Effectiveness and costs of interventions to lower systolic blood pressure and cholesterol: a global and regional analysis on reduction of cardiovascular-disease risk. The Lancet, 361, 717–25.CrossRefGoogle ScholarPubMed
Muscat, J. E., Harris, R. E., Haley, N. J. and Wynder, E. L. (1991). Cigarette smoking and plasma cholesterol. American Heart Journal, 121, 141–7.CrossRefGoogle ScholarPubMed
Mussa, F. F., Chai, H., Wang, X., Yao, Q., Lumsden, A. B. and Chen, C. (2006). Chlamydia pneumoniae and vascular disease: an update. Journal of Vascular Surgery, 43, 1301–7.CrossRefGoogle ScholarPubMed
Must, A. and Colclough-Douglas, S. (2001). Adult health sequelae of pediatric obesity. In Obesity, Growth and Development, ed. Johnston, F. E. and Foster, G. D.. London: Smith-Gordon, pp. 185–98.Google Scholar
Must, A., Spadano, J., Coakley, E. H., Field, A. E., Colditz, G. and Dietz, W. H. (1999). The disease burden associated with overweight and obesity. Journal of the American Medical Association, 282, 1523–9.CrossRefGoogle ScholarPubMed
Muti, P. (2004). The role of endogenous hormones in the etiology and prevention of breast cancer: the epidemiological evidence. Annals of the New York Academy of Sciences, 1028, 273–182.CrossRefGoogle ScholarPubMed
National Sleep Foundation (2002). “Sleep in America” Poll. Washington, DC: National Sleep Foundation.
Nazroo, J. (2003). The structuring of ethnic inequalities in health: economic position, racial discrimination, and racism. American Journal of Public Health, 93, 277–84.CrossRefGoogle ScholarPubMed
Neel, J. V. (1962). Diabetes mellitus: a “thrifty” genotype rendered detrimental by “progress”?American Journal of Human Genetics, 14, 353–62.Google ScholarPubMed
Neel, J. V. (1982). The thrifty genotype revisited. In The Genetics of Diabetes Mellitus, ed. , J. Köbberling and , R. Tattersall. Amsterdam: Academic Press, pp. 137–47.Google Scholar
Neel, J. V., Weder, A. B. and Julius, S. (1998). Type II diabetes, essential hypertension, and obesity as ‘syndromes of impaired genetic homeostasis’: the ‘thrifty genotype’ hypothesis enters the 21st century. Perspectives in Biology and Medicine, 42, 44–74.CrossRefGoogle Scholar
Ness, R. B., Haggerty, C. L., Harger, G. and Ferrell, R. E. (2004). Differential distribution of allelic variants in cytokine genes among African Americans and white Americans. American Journal of Epidemiology, 160, 1033–8.CrossRefGoogle ScholarPubMed
Ness, R. B. (2000). Is depression an adaptation?Archives of General Psychiatry, 57, 14–20.CrossRefGoogle Scholar
Nesse, R. M. (2004). Natural selection and the elusiveness of happiness. Philosophical Transactions of the Royal Society of London B, 359, 1333–47.CrossRefGoogle ScholarPubMed
Nesse, R. M. and Williams, G. C. (1994). Evolution and Healing: The New Science of Darwinian Medicine. London: Phoenix.Google Scholar
Nesse, R. M., Stearns, S. C. and Omenn, G. S. (2006). Medicine needs evolution. Science, 311, 1071.CrossRefGoogle ScholarPubMed
Nettle, D. (2004). Evolutionary origins of depression: a review and reformulation. Journal of Affective Disorders, 81, 91–102.CrossRefGoogle ScholarPubMed
Netuveli, G., Hurwitz, B., Levy, M.et al. (2005). Ethnic variations in UK asthma frequency, morbidity, and health-service use: a systematic review and meta-analysis. The Lancet, 365, 312–17.CrossRefGoogle ScholarPubMed
Neuhausen, S. L. (1999). Ethnic differences in cancer risk resulting from genetic variation. Cancer, 86, 2575–82.3.0.CO;2-F>CrossRefGoogle ScholarPubMed
Newsome, C. A., Shiell, A. W., Fall, C. H. D., Phillips, D. I. W., Shier, R. and Law, C. M. (2003). Is birth weight related to later glucose and insulin metabolism? A systematic review. Diabetic Medicine, 20, 339–48.CrossRefGoogle ScholarPubMed
Nielsen, S. J. and Popkin, B. M. (2003). Patterns and trends in food portion sizes, 1977–1998. Journal of the American Medical Association, 289, 450–3.CrossRefGoogle ScholarPubMed
NIH State-of-the–Science Panel (2005). National Institutes of Health State-of-the–Science conference statement: management of menopause-related symptoms. Annals of Internal Medicine, 142, 1003–13.CrossRef
Núñez-de la Mora, A., Chatterton, R. T., Choudhury, O. A., Napolitano, D. A. and Bentley, G. R. (2007). Childhood conditions influence adult progesterone levels. PLoS Medicine, doi: 10.1371/journal. pmed. 0040167.CrossRef
O'Dea, K. (1984). Marked improvement in carbohydrate and lipid metabolism in diabetic Australian Aborigines after temporary reversion to traditional lifestyle. Diabetes, 33, 596–603.CrossRefGoogle ScholarPubMed
O'Dea, K. (1991). Traditional diet and food preferences of Australian Aboriginal hunter–gatherers. Philosophical Transactions of the Royal Society of London B, 334, 233–41.CrossRefGoogle ScholarPubMed
O'Dea, K. and Piers, L. (2002). Diabetes. In The Nutrition Transition: Diet and Disease in the Developing World, ed. Caballero, B. and Popkin, B. M.. Amsterdam: Elsevier, pp. 165–90.Google Scholar
O'Dea, K., Hopper, J., Patel, M., Traianedes, K. and Kubisch, D. (1993). Obesity, diabetes, and hyperlipidemia in a Central Australian Aboriginal community with a long history of acculturation. Diabetes Care, 16, 1004–10.CrossRefGoogle Scholar
O'Sullivan, A. J., Martin, A. and Brown, M. A. (2001). Efficient fat storage in premenopausal women and in early pregnancy: a role for estrogen. Journal of Clinical Endocrinology and Metabolism, 86, 4951–6.CrossRefGoogle ScholarPubMed
Oken, E. and Gillman, M. W. (2003). Fetal origins of obesity. Obesity Research, 11, 496–506.CrossRefGoogle ScholarPubMed
Omran, A. R. (1971). The epidemiologic transition: a theory of the epidemiology of population change. Milbank Memorial Fund Quarterly, 49, 509–38.CrossRefGoogle ScholarPubMed
Ong, K. K. and Dunger, D. B. (2004). Birth weight, infant growth and insulin resistance. European Journal of Endocrinology, 151, U131–U139.CrossRefGoogle ScholarPubMed
Ong, K. L. and Dunger, D. B. (2000). Thrifty genotypes and phenotypes in the pathogenesis of type 2 diabetes mellitus. Journal of Pediatric Endocrinology and Metabolism, 13, 1419–24.CrossRefGoogle ScholarPubMed
Ongphiphadhanakul, B., Rajatanavin, R., Chailurkit, L.et al. (1995). Serum testosterone and its relation to bone mineral density and body composition in normal males. Clinical Endocrinology, 43, 727–33.CrossRefGoogle ScholarPubMed
Orth-Gomer, K., Eriksson, I., Moser, V., Theorell, T. and Fredlund, P. (1994). Lipid lowering through work stress reduction. International Journal of Behavioral Medicine, 1, 204–14.CrossRefGoogle ScholarPubMed
Osei, K., Schuster, D. P., Owusu, S. K. and Amoah, A. G. B. (1997). Race and ethnicity determine serum insulin and C-peptide concentrations and hepatic insulin extraction and insulin clearance: comparative studies of three populations of west African ancestry and white Americans. Metabolism, 46, 53–8.CrossRefGoogle ScholarPubMed
Ostler, K., Thompson, C., Kinmonth, A. -L. K., Peveler, R. C. and Stevens, L. (2001). Influence of socio-economic deprivation on the prevalence and outcome of depression in primary care. British Journal of Psychiatry, 178, 12–17.CrossRefGoogle ScholarPubMed
Owen, C. G., Whincup, P. H., Odoki, K., Gilg, J. A. and Cook, D. G. (2002). Infant feeding and blood cholesterol: a study in adolescents and a systematic review. Pediatrics, 110, 597–608.CrossRefGoogle ScholarPubMed
Owens, J. F., Stoney, C. M. and Matthews, K. A. (1993). Menopausal status influences ambulatory blood pressure levels and blood pressure changes during mental stress. Circulation, 88, 2794–802.CrossRefGoogle ScholarPubMed
Palacios, S. (1999). Current perspectives on the benefits of HRT in menopausal women. Maturitas, 33, S1–S13.CrossRefGoogle ScholarPubMed
Pan, X. -R., Li, G. -W., Hu, Y. H.et al. (1997). Effects of diet and exercise in preventing NIDDM in people with impaired glucose tolerance. The Da Qing IGT and diabetes study. Diabetes Care, 20, 537–44.CrossRefGoogle Scholar
Paradies, Y. C., Montoya, M. J. and Fullerton, S. M. (2007). Racialised genetics and the study of complex diseases: the thrifty genotype revisited. Perspectives in Biology and Medicine, 50, 203–27.CrossRefGoogle Scholar
Parkin, D. M. and Fernández, L. M. G. (2006). Use of statistics to assess the global burden of breast cancer. Breast Journal, 12 (Suppl. 1), S70–80.CrossRefGoogle ScholarPubMed
Parkin, D. M., Bray, F., Ferlay, J. and Pisani, P. (2005). Global cancer statistics, 2002. CA: A Cancer Journal for Clinicians, 55, 74–108.Google ScholarPubMed
Parnia, S., Borwn, J. L. and Frew, A. J. (2002). The role of pollutants in allergic sensitization and the development of asthma. Allergy, 57, 1111–17.CrossRefGoogle Scholar
Pasquali, R. (2006). Obesity, fat distribution and infertility. Maturitas, 54, 363–71.CrossRefGoogle ScholarPubMed
Pasquali, R., Pelusi, C., Genghini, S., Cacciari, M. and Gambineri, A. (2003). Obesity and reproductive disorders in women. Human Reproduction Update, 9, 359–72.CrossRefGoogle ScholarPubMed
Patel, S. M. and Nestler, J. E. (2006). Fertility in polycystic ovary syndrome. Endocrinology and Metabolism Clinics of North America, 35, 137–55.CrossRefGoogle ScholarPubMed
Patz, J. A., Epstein, P. R., Burke, T. A. and Balbus, J. M. (1996). Global climate change and emerging infectious diseases. Journal of the American Medical Association, 275, 217–23.CrossRefGoogle ScholarPubMed
Pearce-Duvet, J. M. C. (2006). The origin of human pathogens: evaluating the role of agriculture and domestic animals in the evolution of human disease. Biological Reviews, 81, 369–82.CrossRefGoogle ScholarPubMed
Penders, J., Thijs, C., Vink, C.et al. (2006). Factors influencing the composition of the intestinal microbiota in early infancy. Pediatrics, 118, 511–21.CrossRefGoogle ScholarPubMed
Pérez-Perdomo, R., Pérez-Cardona, C., Disdier-Flores, O. and Cintrón, Y. (2003). Prevalence and correlates of asthma in the Puerto Rican population: Behavioral Risk Factor Surveillance System, 2000. Journal of Asthma, 40, 465–74.CrossRefGoogle ScholarPubMed
Pettitt, D. J., Aleck, K., Baird, H., Carraher, M., Bennett, P. and Knowler, W. C. (1988). Congenital susceptibility to NIDDM: role of intrauterine environment. Diabetes, 37, 622–8.CrossRefGoogle ScholarPubMed
Pettitt, D. J., Forman, M. R., Hanson, R. L., Knowler, W. C. and Bennett, P. H. (1997). Breastfeeding and incidence of non-insulin-dependent diabetes mellitus in Pima Indians. The Lancet, 350, 166–8.CrossRefGoogle ScholarPubMed
Pi-Sunyer, F. X. (2002). Glycemic index and disease. American Journal of Clinical Nutrition, 76 (Suppl.), 290S–8S.CrossRefGoogle ScholarPubMed
Pirart, J. (1978). Diabetes mellitus and its degenerative complications: a prospective study of 4,400 patients observed between 1947 and 1973 (Part 1). Diabetes Care, 1, 168–88.CrossRefGoogle Scholar
Platts-Mills, T. A. E., Vervloet, D., Thomas, W. R., Aalberse, R. C. and Chapman, M. D. (1997). Indoor allergens and asthma: report of the Third International Workshop. Journal of Allergy and Clinical Immunology, 100, S2–24.CrossRefGoogle ScholarPubMed
Platz, E. and Giovannucci, E. (2004). The epidemiology of sex steroid hormones and their signaling and metabolic pathways in the etiology of prostate cancer. Journal of Steroid Biochemistry and Molecular Biology, 92, 237–53.CrossRefGoogle ScholarPubMed
Pollard, I. (1994). A Guide to Reproduction: Social Issues and Human Concerns. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Pollard, T. M. (1997). Environmental change and cardiovascular disease: a new complexity. Yearbook of Physical Anthropology, 40, 1–24.3.0.CO;2-8>CrossRefGoogle Scholar
Pollard, T. M. (2000). Adrenaline. In Encyclopedia of Stress, ed. Fink, G.. San Diego: Academic Press, pp. 52–8.Google Scholar
Pollard, T. M. and Ice, G. H. (2006). Measuring hormonal variation in the hypothalamic pituitary adrenal axis: cortisol. In Measuring Stress in Humans: A Practical Guide for the Field, ed. Ice, G. H. and James, G. D.. Cambridge: Cambridge University Press, pp. 122–57.CrossRefGoogle Scholar
Pollard, T. M. and Unwin, N. C. (2007). Impaired reproductive function in western and ‘westernising’ populations: an evolutionary approach. In New Perspectives in Evolutionary Medicine, ed. Trevathan, W. R., Smith, E. O. and McKenna, J. J.. Oxford: Oxford University Press.Google Scholar
Pollard, T. M., Brush, G. and Harrison, G. A. (1991). Geographic distribution of within-population variability in blood pressure. Human Biology, 63, 643–61.Google Scholar
Pollard, T. M., Núñez-de la Mora, A. and Unwin, N. C. (in press). Evolutionary perspectives on type 2 diabetes in Asia. In Medicine and Evolution: Current Applications, Future Prospects, ed. Elton, S. and O'Higgins, P.. Boca Raton: Taylor and Francis.
Pollard, T. M., Ungpakorn, G., Harrison, G. A. and Parkes, K. R. (1996). Epinephrine and cortisol responses to work: a test of the models of Frankenhaeuser and Karasek. Annals of Behavioral Medicine, 18, 229–37.CrossRefGoogle ScholarPubMed
Pollock, K. (1988). On the nature of social stress: production of a modern mythology. Social Science and Medicine, 26, 381–92.CrossRefGoogle ScholarPubMed
Pond, C. M. (1998). The Fats of Life. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Popham, F. and Mitchell, R. (2006). Leisure time exercise and personal circumstances in the working age population: longitudinal analysis of the British household panel survey. Journal of Epidemiology and Community Health, 60, 270–4.CrossRefGoogle ScholarPubMed
Popkin, B. M. (1999). Urbanization, lifestyle changes and the nutrition transition. World Development, 27, 1905–16.CrossRefGoogle Scholar
Popkin, B. M. (2001). The nutrition transition and obesity in the developing world. Journal of Nutrition, 131, 871S–3S.CrossRefGoogle ScholarPubMed
Popkin, B. M. and Gordon-Larsen, P. (2004). The nutrition transition: worldwide obesity dynamics and their determinants. International Journal of Obesity, 28, S2–9.CrossRefGoogle ScholarPubMed
Poretsky, L., Cataldo, N. A., Rosenwaks, Z. and Giudice, L. C. (1999). The insulin-related ovarian regulatory system in health and disease. Endocrine Reviews, 20, 535–82.CrossRefGoogle ScholarPubMed
Portengen, L., Preller, L., Tielen, M., Doekes, G. and Heederik, D. (2005). Endotoxin exposure and atopic sensitization in adult pig farmers. Journal of Allergy and Clinical Immunology, 115, 797–802.CrossRefGoogle ScholarPubMed
Potter, L. B., Rogler, L. H. and Moscicki, E. K. (1995). Depression among Puerto Ricans in New York City – the Hispanic Health and Nutrition Examination Survey. Social Psychiatry and Psychiatric Epidemiology, 30, 185–93.CrossRefGoogle ScholarPubMed
Poulain, M., Doucet, M., Major, G. C.et al. (2006). The effect of obesity on chronic respiratory diseases: pathophysiology and therapeutic strategies. Canadian Medical Association Journal, 174, 1293–9.CrossRefGoogle ScholarPubMed
Prentice, A. M. (2001). Fires of life: the struggles of an ancient metabolism in a modern world. BNF Nutrition Bulletin, 26, 13–27.CrossRefGoogle Scholar
Prentice, A. M. and Jebb, S. A. (1995). Obesity in Britain: gluttony or sloth. British Medical Journal, 311, 437–9.CrossRefGoogle ScholarPubMed
Prentice, A. M., Rayco-Solon, P. and Moore, S. E. (2005). Insights from the developing world: thrifty genotypes and thrifty phenotypes. Proceedings of the Nutrition Society, 64, 153–61.CrossRefGoogle ScholarPubMed
Preston, S. H. (1976). Mortality Patterns in National Populations: With Special Reference to Recorded Causes of Death. New York: Academic Press.Google Scholar
Price, J. F. and Fowkes, G. R. (1997). Risk factors and the sex differential in coronary artery disease. Epidemiology, 8, 584–91.CrossRefGoogle ScholarPubMed
Prior, I. A. M. (1971). The price of civilization. Nutrition Today, 6, 2–11.CrossRefGoogle Scholar
Raben, N., Barbetti, F., Cama, A.et al. (1991). Normal coding sequence of insulin gene in Pima Indians and Nauruans, two groups with highest prevalence of type II diabetes. Diabetes, 40, 118–22.CrossRefGoogle ScholarPubMed
Rajkumar, L., Guzman, R. C., Yang, J., Thordarson, G., Talamantes, F. and Nandi, S. (2004). Prevention of mammary carcinogenesis by short-term estrogen and progestin treatments. Breast Cancer Research, 6, R31–7.CrossRefGoogle ScholarPubMed
Randolph, J. F., Sowers, M., Gold, E. B.et al. (2003). Reproductive hormones in the early menopausal transition: relationship to ethnicity, body size, and menopausal status. Journal of Clinical Endocrinology and Metabolism, 88, 1516–22.CrossRefGoogle ScholarPubMed
Ravelli, A. C. J., van der Meulen, J. H. P., Osmond, C., Barker, D. J. P. and Bleker, O. P. (2000). Infant feeding and adult glucose tolerance, lipid profile, blood pressure, and obesity. Archives of Disease in Childhood, 82, 248–52.CrossRefGoogle ScholarPubMed
Ray, C. and Stevens, J. R. (1995). Sacred Legends. Manotick, ON: Penumbra Press.Google Scholar
Reaven, G. M. (1988). Role of insulin resistance in human disease. Diabetes, 37, 1595–607.CrossRefGoogle ScholarPubMed
Rebuffé-Scrive, M., Enk, L., Crona, N.et al. (1985). Fat cell metabolism in different regions in women: effect of menstrual cycle, pregnancy, and lactation. Journal of Clinical Investigation, 75, 1973–6.CrossRefGoogle ScholarPubMed
Reilly, J. J., Jackson, D. M., Montgomery, C.et al. (2004). Total energy expenditure and physical activity in young Scottish children: mixed longitudinal study. The Lancet, 363, 211–12.CrossRefGoogle ScholarPubMed
Reilly, J. J., Armstrong, J., Dorosty, A. R.et al. for the Avon Longitudinal Study of Parents and Children Study Team (2005). Early life risk factors for obesity in childhood: cohort study. British Medical Journal, 330, 1357–9.CrossRefGoogle ScholarPubMed
Reinhard, K. J. (1988). Cultural ecology of prehistoric parasitism on the Colorado Plateau as evidenced by coprology. American Journal of Physical Anthropology, 77, 355–66.CrossRefGoogle ScholarPubMed
Relethford, J. H. (1994). Fundamental of Biological Anthropology. Mountain View, California: Mayfield Publishing Company.Google Scholar
Renehan, A. G., Zwahlen, M., Minder, C., O'Dwyer, S. T., Shalet, S. M. and Egger, M. (2004). Insulin-like growth factor (IGF)-1, IGF binding protein-3, and cancer risk: systematic review and meta-regression analysis. The Lancet, 363, 1346–53.CrossRefGoogle ScholarPubMed
Richards, M. P. (2002). A brief review of the archaeological evidence for Palaeolithic and Neolithic subsistence. European Journal of Clinical Nutrition, 56, 1–9.CrossRefGoogle ScholarPubMed
Ridker, P. M. (2002). On evolutionary biology, inflammation, infection and the causes of atherosclerosis. Circulation, 105, 2–4.Google ScholarPubMed
Riedler, J., Braun-Fahrlander, C., Eder, W.et al. and the ALEX study team (2001). Exposure to farming in early life and development of asthma and allergy: a cross-sectional survey. The Lancet, 358, 1129–33.CrossRefGoogle ScholarPubMed
Rissanen, A. M., Heliovaara, M., Knekt, P., Reunanen, A. and Aromaa, A. (1991). Determinants of weight gain and overweight in adult Finns. European Journal of Clinical Nutrition, 45, 419–30.Google ScholarPubMed
Ritenbaugh, C. and Goodby, C. S. (1989). Beyond the thrifty gene: metabolic implications of prehistoric migration into the new world. Medical Anthropology, 11, 227–36.CrossRefGoogle ScholarPubMed
Roberts, C. A. and Cox, M. (2003). Health and Disease in Britain. Stroud: Sutton Publishing.Google Scholar
Roberts, E. M. (2002). Racial and ethnic disparities in childhood asthma diagnosis: the role of clinical findings. Journal of the National Medical Association, 94, 215–23.Google ScholarPubMed
Robinson, T. N. (2001). Population-based obesity prevention for children and adolescents. In Obesity, Growth and Development, ed. Johnston, F. E. and Foster, G. D.. London: Smith–Gordon, pp. 129–41.Google Scholar
Rode, A. and Shephard, R. J. (1971). Cardiorespiratory fitness of an Arctic community. Journal of Applied Physiology, 31, 519–26.CrossRefGoogle ScholarPubMed
Rodriguez, M. A., Winkleby, M. A., Ahn, D., Sundquist, J. and Kraemer, H. C. (2002). Identification of population subgroups of children and adolescents with high asthma prevalence. Archives of Pediatric and Adolescent Medicine, 156, 269–75.CrossRefGoogle ScholarPubMed
Rook, G. A. W., Adams, V., Hunt, J., Palmer, R., Martinelli, R. and Brunet, Rosa L. (2004). Mycobacteria and other environmental organisms as immunomodulators for immunoregulatory disorders. Springer Seminars in Immunopathology, 25, 237–55.CrossRefGoogle ScholarPubMed
Roper, N. A., Bilous, R. W., Kelly, W. F., Unwin, N. C. and Connolly, V. M. (2001). Excess mortality in a population with diabetes and the impact of material deprivation: longitudinal, population based study. British Medical Journal, 322, 1389–93.CrossRefGoogle Scholar
Rose, D., Mannino, D. M. and Leaderer, B. P. (2006). Asthma prevalence among US adults, 1998–2000: role of Puerto Rican ethnicity and behavioral and geographic factors. American Journal of Public Health, 96, 880–8.CrossRefGoogle ScholarPubMed
Rosenblatt, K. A., Thomas, D. B. and The WHO Collaborative Study of Neoplasia and Steroid Contraceptives (1995). Prolonged lactation and endometrial cancer. International Journal of Epidemiology, 24, 499–503.CrossRefGoogle ScholarPubMed
Ross, R. K., Coetzee, G. A., Reichardt, J., Skinner, E. and Henderson, B. E. (1995). Does the racial-ethnic variation in prostate cancer risk have a hormonal basis?Cancer, 75, 1778–82.3.0.CO;2-J>CrossRefGoogle Scholar
Roumain, J., Charles, M. A., Courten, M. P.et al. (1998). The relationship of menstrual irregularity to Type 2 diabetes in Pima Indian women. Diabetes Care, 21, 346–9.CrossRefGoogle ScholarPubMed
Royal, C. D. M. and Dunston, G. M. (2004). Changing the paradigm from ‘race’ to human genome variation. Nature Genetics, 36, S5–7.CrossRefGoogle ScholarPubMed
Runciman, W. G. (2005). Stone age sociology. Journal of the Royal Anthropological Institute, 11, 129–42.CrossRefGoogle Scholar
Ryan, A. S., Wenjun, Z. and Acosta, A. (2002). Breastfeeding continues to increase into the new millenium. Pediatrics, 110, 1103–9.CrossRefGoogle Scholar
Sackett, R. D. (1996). Time, Energy, and the Indolent Savage. Unpublished PhD thesis. University of California, Los Angeles.
Saelens, B. E., Sallis, J. F., Black, J. B. and Chen, D. (2003). Neighbourhood-based differences in physical activity: an environment scale evaluation. American Journal of Public Health, 93, 1552–8.CrossRefGoogle Scholar
Sahlins, M. (1972). Stone Age Economics. New York: Aldine de Gruyter.Google Scholar
Sánchez-Castillo, C. P., Velásquez-Monroy, O., Lara-Esqueda, A.et al. (2005). Diabetes and hypertension increases in a society with abdominal obesity: results of the Mexican National Health Survey 2000. Public Health Nutrition, 8, 53–60.CrossRefGoogle Scholar
Sapolsky, R. M., Krey, L. C. and McEwen, B. S. (1986). The neuroendocrinology of stress and aging: the glucocorticoid cascade hypothesis. Endocrine Reviews, 7, 284–301.CrossRefGoogle ScholarPubMed
Sapolsky, R. M., Romero, L. M. and Munck, A. U. (2000). How do glucocorticoids influence stress responses? Integrating permissive, suppressive, stimulatory, and preparative actions. Endocrine Reviews, 21, 55–89.Google ScholarPubMed
Saxton, J. M. (2006). Diet, physical activity and energy balance and their impact on breast and prostate cancers. Nutrition Research Reviews, 19, 197–215.CrossRefGoogle ScholarPubMed
Schaub, B., Lauener, R. and Mutius, E. (2006). The many faces of the hygiene hypothesis. Journal of Allergy and Clinical Immunology, 117, 969–77.CrossRefGoogle ScholarPubMed
Schieffelin, E. L. (1985). The cultural analysis of depressive affect: an example from New Guinea. In Culture and Depression: Studies in the Anthropology and Cross-Cultural Psychiatry of Affect and Disorder, ed. Kleinman, A. and Good, B.. Berkeley: University of California Press, pp. 101–33.Google Scholar
Schofield, R. and Reher, D. (1991). The decline of mortality in Europe. In The Decline of Mortality in Europe, ed. Schofield, R., Reher, D. and Bideau, D.. Oxford: Clarendon, pp. 1–17.Google Scholar
Schröder, F. H. (1996). Impact of ethnic, nutritional and environmental factors on prostate cancer. In Pharmacology, Biology, and Clinical Applications of Androgens, ed. Bhasin, S., Gabelnick, H. L., Spieleretal, J. M.. New York: Wiley-Liss, pp. 121–36.Google Scholar
Schulz, L. O., Bennett, P. H., Ravussin, E.et al. (2006). Effects of traditional and western environments on prevalence of type 2 diabetes in Pima Indians in Mexico and the US. Diabetes Care, 29, 1866–71.CrossRefGoogle Scholar
Schulze, M. B., Manson, J. E., Ludwig, D. S.et al. (2004). Sugar-sweetened beverages, weight gain, and incidence of type 2 diabetes in young and middle-aged women. Journal of the American Medical Association, 292, 927–34.CrossRefGoogle Scholar
Schutz, Y., Weinsier, R. L. and Hunter, G. R. (2001). Assessment of free-living physical activity in humans: an overview of currently available and proposed new measures. Obesity Research, 9, 368–79.CrossRefGoogle ScholarPubMed
Scott, S. and Duncan, C. J. (2001). Biology of Plagues: Evidence from Historical Populations. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Scragg, R. K. R., Fraser, A. and Metcalf, P. A. (1996). Helicobacter pylori seropositivity and cardiovascular risk factors in a multicultural workforce. Journal of Epidemiology and Community Health, 50, 578–9.CrossRefGoogle Scholar
Scrivener, S., Yemaneberhan, H., Zebenigus, M.et al. (2001). Independent effects of intestinal parasite infection and domestic allergen exposure on risk of wheeze in Ethiopia: a nested case-control study. The Lancet, 358, 1493–9.CrossRefGoogle ScholarPubMed
Seale, C. (2000). Changing patterns of death and dying. Social Science and Medicine, 51, 917–30.CrossRefGoogle ScholarPubMed
Seidell, J. C. (1995). Obesity in Europe: scaling an epidemic. International Journal of Obesity, 19 (Suppl. 3), S1–4.Google ScholarPubMed
Seidell, J. C. (2000). Obesity, insulin resistance and diabetes – a worldwide epidemic. British Journal of Nutrition, 83 (Suppl. 1), S5–8.CrossRefGoogle ScholarPubMed
Sellen, D. W. and Smay, D. B. (2001). Relationship between subsistence and age at weaning in “preindustrial” societies. Human Nature, 12, 47–87.CrossRefGoogle ScholarPubMed
Sellers, E. A. C., Triggs-Raine, B., Rockman-Greenberg, C. and Dean, H. J. (2002). The prevalence of the HNF-1alpha G319S mutation in Canadian Aboriginal youth with type 2 diabetes. Diabetes Care, 25, 2202–6.CrossRefGoogle ScholarPubMed
Sephton, S. E., Sapolsky, R. M., Kraemer, H. C. and Spiegel, D. (2000). Diurnal cortisol rhythm as a predictor of breast cancer survival. Journal of the National Cancer Institute, 92, 994–1000.CrossRefGoogle ScholarPubMed
Shaneyfelt, T., Husein, R., Bubley, G. and Mantzoros, C. S. (2000). Hormonal predictors of prostate cancer: a meta-analysis. Journal of Clinical Oncology, 18, 847–53.CrossRefGoogle ScholarPubMed
Shanley, D. P. and Kirkwood, T. B. L. (2001). Evolution of the human menopause. Bioessays, 23, 282–7.3.0.CO;2-9>CrossRefGoogle ScholarPubMed
Sharpe, R. M. and Franks, S. (2002). Environment, lifestyle and infertility – an inter-generational issue. Nature Medicine, 8 (Suppl. 11), 33–40.CrossRefGoogle Scholar
Sharpe, R. M. and Skakkebaek, N. E. (1993). Are oestrogens involved in falling sperm counts and disorders of the male reproductive tract?The Lancet, 341, 1392–5.CrossRefGoogle ScholarPubMed
Shea, J. L. (2006). Parsing the ageing Asian woman: symptom results from the China Study of Midlife Women. Maturitas, 55, 36–50.CrossRefGoogle ScholarPubMed
Sherry, D. S. and Marlowe, F. W. (2007). Anthropometric data indicate nutritional homogeneity in Hadza Foragers of Tanzania. American Journal of Human Biology, 19, 107–18.CrossRefGoogle ScholarPubMed
Shetty, P. S., Henry, C. J. K., Black, A. E. and Prentice, A. M. (1996). Energy requirements of adults: an update on basal metabolic rates (BMRs) and physical activity levels (PALs). European Journal of Clinical Nutrition, 50 (Suppl. 1), S11–23.Google Scholar
Shimuzu, H., Ross, R. K., Bernstein, L., Pike, M. C. and Henderson, B. E. (1990). Serum oestrogen levels in postmenopausal women: comparison of American whites and Japanese in Japan. British Journal of Cancer, 62, 451–3.CrossRefGoogle Scholar
Shore, S. A. (2006). Obesity and asthma: cause for concern. Current Opinion in Pharmacology, 6, 230–6.CrossRefGoogle ScholarPubMed
Shusterman, D. J., Murphy, M. A. and Balmes, J. R. (1998). Subjects with seasonal allergic rhinitis and nonrhinitic subjects react differently to nasal provocation with chlorine gas. Journal of Allergy and Clinical Immunology, 101, 732–40.CrossRefGoogle Scholar
Sicherer, S. H., Muñoz-Furlong, A. and Sampson, H. A. (2003). Prevalence of peanut and tree nut allergy in the United States determined by means of a random digit dial telephone survey: a 5-year follow-up study. Journal of Allergy and Clinical Immunology, 112, 1203–7.CrossRefGoogle ScholarPubMed
Siegert, R. J. and Ward, T. (2002). Clinical psychology and evolutionary psychology: toward a dialogue. Review of General Psychology, 6, 235–59.CrossRefGoogle Scholar
Simhon, A., Douglas, J. R., Drasar, B. S. and Sotthill, J. F. (1982). Effect of feeding on infants' faecal flora. Archives of Disease in Childhood, 57, 54–8.Google ScholarPubMed
Simmons, R. (2005). Developmental origins of adult metabolic disease: concepts and controversies. TRENDS in Endocrinology and Metabolism, 16, 390–4.CrossRefGoogle ScholarPubMed
Sinclair, A. and O'Dea, K. (1993). The significance of arachidonic acid in hunter–gatherer diets: implications for the contemporary western diet. Journal of Food Lipids, 1, 143–57.CrossRefGoogle Scholar
Slattery, M. L. and Randall, D. E. (1988). Trends in coronary heart disease mortality and food consumption in the United States between 1909 and 1980. American Journal of Clinical Nutrition, 47, 1060–70.CrossRefGoogle ScholarPubMed
Smith, B. H. (1991). Dental development and the evolution of life history in Hominidae. American Journal of Physical Anthropology, 86, 157–74.CrossRefGoogle Scholar
Smith, C. J. (1994). Food habit and cultural changes among the Pima Indians. In Diabetes as a Disease of Civilization, ed. Joe, J. R. and Young, R. S.. New York: Mouton de Gruyter.CrossRefGoogle Scholar
Smith, C. J., Nelson, R. G., Hardy, S. A., Manahan, E. M., Bennett, P. H. and Knowler, W. C. (1996). Survey of the diet of Pima Indians using quantitative food frequency assessment and 24-hour recall. Journal of the American Dietetic Association, 96, 778–84.CrossRefGoogle ScholarPubMed
Smith, E. O. (2002). When Culture and Biology Collide: Why we are Stressed, Depressed, and Self-obsessed. New Brunswick, New Jersey: Rutgers University Press.Google Scholar
Smith, M. T. (1993). Genetic adaptation. In Human Adaptation, ed. Harrison, G. A.. Oxford: Oxford University Press, pp. 1–54.Google Scholar
Smyth, J., Ockenfels, M. C., Porter, L., Kirschbaum, C., Hellhammer, D. H. and Stone, A. A. (1998). Stressors and mood measured on a momentary basis are associated with salivary cortisol secretion. Psychoneuroendocrinology, 23, 353–70.CrossRefGoogle ScholarPubMed
Snijder, M. B., Dam, R. M., Visser, M. and Seidell, J. C. (2005). What aspects of body fat are particularly hazardous and how do we measure them?International Journal of Epidemiology, 35, 83–92.CrossRefGoogle Scholar
Solomon, C. G. (1999). The epidemiology of polycystic ovary syndrome. Endocrinology and Metabolism Clinics of North America, 28, 247–63.CrossRefGoogle ScholarPubMed
Sood, A., Ford, E. S. and Camargo, C. A. (2006). Association between leptin and asthma in adults. Thorax, 61, 300–5.CrossRefGoogle ScholarPubMed
Sowers, J. R. (2003). Obesity as a cardiovascular risk factor. American Journal of Medicine, 115, 37S–41S.CrossRefGoogle ScholarPubMed
Spiegel, K., Knutson, K., Leproult, R., Tasali, E. and Cauter, E. (2005). Sleep loss: a novel risk factor for insulin resistance and type 2 diabetes. Journal of Applied Physiology, 99, 2008–19.CrossRefGoogle ScholarPubMed
Spielman, R. S., Fajans, S. S., Neel, J. V., Pek, S., Floyd, J. C. and Oliver, W. J. (1982). Glucose tolerance in two unacculturated Indian tribes of Brazil. Diabetologia, 23, 90–3.CrossRefGoogle ScholarPubMed
Stamatakis, E., Primatesta, P., Chinn, S., Rona, R. and Falascheti, E. (2005). Overweight and obesity trends from 1974 to 2003 in English children: what is the role of socioeconomic factors?Archives of Disease in Childhood, 90, 999–1004.CrossRefGoogle ScholarPubMed
Stampfer, M. J. (2006). Cardiovascular disease and Alzheimer's disease: common links. Journal of Internal Medicine, 260, 211–23.CrossRefGoogle ScholarPubMed
Stampfer, M. J. and Rimm, E. B. (1995). Epidemiologic evidence for vitamin E in prevention of cardiovascular disease. American Journal of Clinical Nutrition, 62, S1365–9.CrossRefGoogle ScholarPubMed
Stanhope, J. M. and Prior, I. A. M. (1980). The Tokelau island migrant study: prevalence and incidence of diabetes mellitus. New Zealand Medical Journal, 92, 417–21.Google ScholarPubMed
Stearns, S. C. and Ebert, D. (2001). Evolution in health and disease: work in progress. Quarterly Review of Biology, 76, 417–32.CrossRefGoogle ScholarPubMed
Steckel, R. H. and Rose, J. C. (2002b). Conclusions. In The Backbone of History: Health and Nutrition in the Western Hemisphere, ed. Steckel, R. H. and Rose, J. C.. Cambridge: Cambridge University Press, pp. 583–9.CrossRefGoogle Scholar
Steckel, R. H. and Rose, J. C. (2002a). Patterns of health in the western hemisphere. In The Backbone of History: Health and Nutrition in the Western Hemisphere, ed. Steckel, R. H. and Rose, J. C.. Cambridge: Cambridge University Press, pp. 563–79.CrossRefGoogle Scholar
Steckel, R. H., Rose, J. C., Larsen, C. S. and Walker, P. L. (2002). Skeletal health in the Western Hemisphere from 4000BC to the present. Evolutionary Anthropology, 11, 142–55.CrossRefGoogle Scholar
Steffen, P. R., McNeilly, M., Anderson, N. and Sherwood, A. (2003). Effects of perceived racism and anger inhibition on ambulatory blood pressure in African Americans. Psychosomatic Medicine, 65, 746–50.CrossRefGoogle ScholarPubMed
Stene, L. C. and Nafstad, P. (2001). Relation between occurrence of type 1 diabetes and asthma. The Lancet, 357, 607–8.CrossRefGoogle ScholarPubMed
Steptoe, A. and Marmot, M. (2002). The role of psychobiological pathways in socio-economic inequalities in cardiovascular disease risk. European Heart Journal, 23, 13–25.CrossRefGoogle ScholarPubMed
Steptoe, A., Wardle, J., Lipsey, Z., Oliver, G., Pollard, T. M. and Davies, G. J. (1998). The effects of life stress on food choice. In The Nation's Diet: The Social Science of Food Choice, ed. Murcott, A.. Harlow: Addison Wesley Longman, pp. 29–42.Google Scholar
Stevens, R. G. (2006). Artificial lighting in the industrialized world: circadian disruption and breast cancer. Cancer Causes and Control, 17, 501–7.CrossRefGoogle ScholarPubMed
Stevens, R. G. and Rea, M. S. (2001). Light in the built environment: potential role of circadian disruption in endocrine disruption and breast cancer. Cancer Causes and Control, 12, 279–87.CrossRefGoogle ScholarPubMed
Stinson, S. (2002). Early childhood health in foragers. In Human Diet: Its Origin and Evolution, ed. Ungar, P. S. and Teaford, M. F.. Westport, Connecticut: Bergin and Garvey, pp. 37–48.Google Scholar
Storey, R. (1985). An estimate of mortality in a Pre–Columbian urban population. American Anthropologist, 83, 519–35.CrossRefGoogle Scholar
Strachan, D. P. (1989). Hay fever, hygiene and household size. British Medical Journal, 299, 1259–60.CrossRefGoogle ScholarPubMed
Strachan, D. P. (1997). Allergy and family size: a riddle worth solving. Clinical and Experimental Allergy, 27, 235–6.CrossRefGoogle ScholarPubMed
Strassmann, B. I. (1999). Menstrual cycling and breast cancer: an evolutionary perspective. Journal of Women's Health, 8, 193–202.CrossRefGoogle Scholar
Strassmann, B. I. and Dunbar, R. I. M. (1999). Human evolution and disease: putting the Stone Age in perspective. In Evolution in Health and Disease, ed. Stearns, S. C.. Oxford: Oxford University Press, pp. 91–101.Google Scholar
Strauss, R. S. and Pollack, H. A. (2001). Epidemic increase in childhood overweight, 1986–1998. Journal of the American Medical Association, 286, 2845–8.CrossRefGoogle ScholarPubMed
Stringer, C. (2002). Modern human origins: progress and prospects. Philosophical Transactions of the Royal Society of London B, 357, 563–79.CrossRefGoogle ScholarPubMed
Stringer, C. (2003). Out of Ethiopia. Nature, 423, 692–5.CrossRefGoogle ScholarPubMed
Stuart-Macadam, P. (1995). Breastfeeding in prehistory. In Breastfeeding: Biocultural Perspectives, ed. Stuart-Macadam, P. and Dettwyler, K. A.. New York: Aldine de Gruyter, pp. 75–99.Google Scholar
Stuart-Macadam, P. (1998). Iron deficiency anemia: exploring the difference. In Sex and Gender in Paleopathological Perspective, ed. Grauer, A. L. and Stuart-Macadam, P.. Cambridge: Cambridge University Press, pp. 45–63.Google Scholar
Sutton-Tyrrell, K., Wildman, R. P., Matthews, K. A.et al. (2005). Sex hormone-binding globulin and the free androgen index are related to cardiovascular risk factors in multiethnic premenopausal women and perimenopausal women enrolled in the study of women across the nation (SWAN). Circulation, 111, 1242–9.CrossRefGoogle Scholar
Swinburn, B. and Egger, G. (2004). The runaway weight gain train: too many accelerators, not enough brakes. British Medical Journal, 329, 736–9.CrossRefGoogle Scholar
Szathmáry, E. J. E. (1994). Non-insulin dependent diabetes mellitus among Aboriginal North Americans. Annual Review of Anthropology, 23, 457–82.CrossRefGoogle Scholar
Taheri, S. (2006). The link between short sleep duration and obesity: we should recommend more sleep to prevent obesity. Archives of Disease in Childhood, 91, 881–4.CrossRefGoogle ScholarPubMed
Taheri, S., Lin, L., Austin, D., Young, T. and Mignot, E. (2004). Short sleep duration is associated with reduced leptin, elevated ghrelin, and increased body mass index. PLoS Medicine, 1, 210–17.CrossRefGoogle ScholarPubMed
Talayero, J. M. P., Lizan-Garcia, M., Puime, A. O.et al. (2006). Full breastfeeding and hospitalization as a result of infections in the first year of life. Pediatrics, 118, E92–9.CrossRefGoogle Scholar
Tatz, C. (2005). Aboriginal Suicide is Different: Portrait of Life and Self-Destruction. Canberra: Aboriginal Studies Press.Google Scholar
Taylor, A. L., Dunstan, J. A. and Prescott, S. L. (2007). Probiotic supplementation for the first 6 months of life fails to reduce the risk of atopic dermatitis and increases the risk of allergen sensitization in high-risk children: a randomized controlled trial. Journal of Allergy and Clinical Immunology, 119, 184–91.CrossRefGoogle ScholarPubMed
Taylor, J. S., Kacmar, J. E., Nothnagle, M. and Lawrence, R. A. (2005). A systematic review of the literature associating breastfeeding with type 2 diabetes and gestational diabetes. Journal of the American College of Nutrition, 24, 320–6.CrossRefGoogle ScholarPubMed
Tesfaye, F., Nawi, N. G., Minh, H.et al. (2007). Association between body mass index and blood pressure across three populations in Africa and Asia. Journal of Human Hypertension, 21, 28–37.CrossRefGoogle ScholarPubMed
Thakore, J. H., Richards, P., Reznek, R. H., Martin, A. and Dinan, T. G. (1997). Increased intra-abdominal fat deposition in patients with major depressive illness as measured by computed tomography. Biological Psychiatry, 41, 1140–2.CrossRefGoogle ScholarPubMed
Thomas, R. B. (1998). The evolution of human adaptability paradigms: toward a biology of poverty. In Building a New Biocultural Synthesis, ed. Goodman, A. H. and Leatherman, T. L.. Ann Arbor: University of Michigan Press, pp. 43–74.Google Scholar
Thomson, N. J. (1991). Recent trends in aboriginal mortality. Medical Journal of Australia, 154, 235–9.Google ScholarPubMed
Thorburn, A. W. (2005). National prevalence of obesity – prevalence of obesity in Australia. Obesity Reviews, 6, 187–9.CrossRefGoogle Scholar
Tillin, T., Forouhi, N., Johnston, D. G., McKeigue, P., Chaturvedi, N. and Godsland, I. F. (2005). Metabolic syndrome and coronary heart disease in South Asians, African-Caribbeans and white Europeans: a UK population-based cross-sectional study. Diabetologia, 48, 649–56.CrossRefGoogle ScholarPubMed
Tishkoff, S. A. and Kidd, K. K. (2004). Implications of biogeography of human populations for ‘race’ and medicine. Nature Genetics, 36, S21–7.CrossRefGoogle Scholar
Tishkoff, S. A., Reed, F. A., Ranciaro, A.et al. (2007). Convergent adaptation of human lactase persistence in Europe. Nature Genetics, 39, 31–40.CrossRefGoogle Scholar
Tominaga, S. and Kuroishi, T. (1997). An ecological study on diet/nutrition and cancer in Japan. International Journal of Cancer, 71, S10, 2–6.3.0.CO;2-C>CrossRefGoogle Scholar
Tonetti, D. (2004). Prevention of breast cancer by recapitulation of pregnancy hormone levels. Breast Cancer Research, 6, E8.CrossRefGoogle ScholarPubMed
Travis, R. C., Allen, D. S., Fentiman, I. S. and Key, T. J. (2004). Melatonin and breast cancer: a prospective study. Journal of the National Cancer Institute, 96, 475–82.CrossRefGoogle ScholarPubMed
Treloar, A. E., Boynton, R. E., Behn, B. G. and Brown, B. W. (1967). Variation of the human menstrual cycle through reproductive life. International Journal of Fertility, 12, 77–126.Google ScholarPubMed
Tremblay, M. S., Katzmarzyk, P. T. and Willms, J. D. (2002). Temporal trends in overweight and obesity in Canada 1981–1996. International Journal of Obesity, 26, 538–43.CrossRefGoogle ScholarPubMed
Trevathan, W. R., Smith, E. O. and McKenna, J. J., eds. (1999). Evolutionary Medicine. New York: Oxford University Press.Google Scholar
Trevathan, W. R., Smith, E. O. and McKenna, J. J., eds. (2007). New Perspectives in Evolutionary Medicine. New York: Oxford University Press.Google Scholar
Trichopoulos, D., MacMahon, B. and Cole, P. (1972). Menopause and breast cancer risk. Journal of the National Cancer Institute, 48, 605–13.Google ScholarPubMed
Triggs-Raine, B. L., Kirkpatrick, R. D., Kelly, S. L.et al. (2002). HNF-1alpha G319S, a transactivation-deficient mutant, is associated with altered dynamics of diabetes onset in an Oji-Cree community. Proceedings of the National Academy of Science, 99, 4614–19.CrossRefGoogle Scholar
Troiano, R. P., Briefel, R. R., Carroll, M. D. and Bialostosky, K. (2000). Energy and fat intakes of children and adolescents in the United States: data from the National Health and Nutrition Examination Surveys. American Journal of Clinical Nutrition, 72, 1343S–53S.CrossRefGoogle ScholarPubMed
Trowell, H. C. and Burkitt, D. P. (1981a). Preface. In Western Diseases: Their Emergence and Prevention, ed. Trowell, H. C. and Burkitt, D. P.. Cambridge, Massachusetts: Harvard University Press, pp. xiii–xvi.Google Scholar
Trowell, H. C. and Burkitt, D. P., eds. (1981b). Western Diseases: Their Emergence and Prevention. Cambridge, Massachusetts: Harvard University Press.Google Scholar
Truswell, A. S. and Hansen, J. D. L. (1976). Medical research among the!Kung. In Kalahari Hunter–Gatherers: Studies of the!Kung San and their Neighbours, ed. Lee, R. B. and DeVore, I.. Cambridge, Massachusetts: Harvard University Press, pp. 166–94.CrossRefGoogle Scholar
Tunstall-Pedoe, H., Connaghan, J., Woodward, M., Tolonen, H. and Kuulasmaa, K. (2006). Pattern of declining blood pressure across replicate population surveys of the WHO MONICA project, mid-1980s to mid-1990s, and the role of medication. British Medical Journal, 332, 629–35.CrossRefGoogle ScholarPubMed
Ueshima, H., Okayama, A., Saitoh, S.et al. (2003). Differences in cardiovascular disease risk factors between Japanese in Japan and Japanese-Americans in Hawaii: the INTERLIPID study. Journal of Human Hypertension, 17, 631–9.CrossRefGoogle ScholarPubMed
Umetsu, D. T. and DeKruyff, R. H. (2006). The regulation of allergy and asthma. Immunological Reviews, 212, 238–55.CrossRefGoogle ScholarPubMed
Uusitalo, U., Feskens, E. J. M., Tuomilehto, J.et al. (1996). Fall in total cholesterol concentration over five years in association with changes in fatty acid composition of cooking oil in Mauritius: cross sectional survey. British Medical Journal, 313, 1044–6.CrossRefGoogle ScholarPubMed
Anders, S. M. and Watson, N. V. (2006). Menstrual cycle irregularities are associated with testosterone levels in healthy premenopausal women. American Journal of Human Biology, 18, 841–4.CrossRefGoogle ScholarPubMed
Blerkom, L. (2003). Role of viruses in human evolution. Yearbook of Physical Anthropology, 46, 14–46.CrossRefGoogle Scholar
Klink, J. J. L., Blonk, R. W. B., Schene, A. H. and van Dijk, F. J. H. (2001). The benefits of interventions for work-related stress. American Journal of Public Health, 91, 271–6.Google ScholarPubMed
Spuy, Z. M. and Dyer, S. J. (2004). The pathogenesis of infertility and early pregnancy loss in polycystic ovary syndrome. Best Practice and Research Clinical Obstetrics and Gynaecology, 18, 755–71.CrossRefGoogle ScholarPubMed
Eck, M. M. and Nicolson, N. A. (1994). Perceived stress and salivary cortisol in daily life. Annals of Behavioral Medicine, 16, 221–7.Google Scholar
Hooff, M. H. A., Voorhorst, F. J., Kaptein, M. B. H., Hirasing, R. A., Koppenaal, C. and Schoemaker, J. (2004). Predictive value of menstrual cycle pattern, body mass index, hormone levels and polycystic ovaries at age 15 years for oligo-amenorrhoea at age 18 years. Human Reproduction, 19, 383–92.CrossRefGoogle ScholarPubMed
Odijk, J., Kull, I., Borres, M. P.et al. (2003). Breastfeeding and allergic disease: a multidisciplinary review of the literature (1966–2001) on the mode of early feeding in infancy and its impact on later atopic manifestations. Allergy, 58, 833–43.CrossRefGoogle ScholarPubMed
Poppel, G., Kardinaal, A., Princen, H. and Kok, F. J. (1994). Antioxidants and coronary heart disease. Annals of Medicine, 26, 429–34.CrossRefGoogle ScholarPubMed
Vartiainen, E., Jousilahti, P., Alfthan, G., Sundvall, J., Pietinen, P. and Puska, P. (2000). Cardiovascular risk factors in Finland, 1972–1997. International Journal of Epidemiology, 29, 49–56.CrossRefGoogle Scholar
Verkasalo, P. K., Thomas, H. V., Appleby, P. N., Davey, G. K. and Key, T. J. (2001). Circulating levels of sex hormones and their relation to risk factors for breast cancer: a cross-sectional study in 1092 pre- and post-menopausal women (United Kingdom). Cancer Causes and Control, 12, 47–59.CrossRefGoogle Scholar
Villamor, E. and Cnattingius, S. (2006). Interpregnancy weight change and risk of adverse pregnancy outcomes: a population-based study. The Lancet, 368, 1164–70.CrossRefGoogle ScholarPubMed
Virtanen, S. M. and Knip, M. (2003). Nutritional risk predictors of β cell autoimmunity and type 1 diabetes at a young age. American Journal of Clinical Nutrition, 78, 1053–67.CrossRefGoogle Scholar
Vitzthum, V. J. (2001). Why not so great is still good enough. In Reproductive Ecology and Human Evolution, ed. Ellison, P. T.. New York: Aldine de Gruyter, pp. 179–202.Google Scholar
Vitzthum, V. J., Bentley, G. R., Spielvogel, H.et al. (2002). Salivary progesterone levels and rate of ovulation are significantly lower in poorer than in better-off urban-dwelling Bolivian women. Human Reproduction, 17, 1906–13.CrossRefGoogle ScholarPubMed
Vitzthum, V. J., Spielvogel, H. and Thornburg, J. (2004). Interpopulational differences in progesterone levels during conception and implantation in humans. Proceedings of the National Academy of Science, 101, 1443–8.CrossRefGoogle ScholarPubMed
Vogel, V. G., Costantino, J. P., Wickerham, D. L.et al. for the National Surgical Adjuvant Breast and Bowel Project (NSABP) (2006). Effects of tamoxifen vs raloxifene on the risk of developing invasive breast cancer and other disease outcomes: the NSABP study of tamoxifen and raloxifene (STAR) P-2 trial. Journal of the American Medical Association, 295, 2727–41.CrossRefGoogle ScholarPubMed
Ehrenstein, O. S., Mutius, E., Illi, S., Baumann, L., Bohm, O. and Kries, R. (2000). Reduced risk of hay fever and asthma among children of farmers. Clinical and Experimental Allergy, 30, 187–93.CrossRefGoogle Scholar
Hertzen, L. and Haahtela, T. (2005). Signs of reversing trends in prevalence of asthma. Allergy, 60, 283–92.CrossRefGoogle Scholar
Mutius, E., Martinez, F. D., Fritzsch, C., Nicolai, T., Reitmeir, P. and Thiemann, H. -H. (1994). Skin test reactivity and number of siblings. British Medical Journal, 308, 692–5.CrossRefGoogle Scholar
Vorona, R. D., Winn, M. P., Babineau, T. W., Eng, B. P., Feldman, H. R. and Ware, J. C. (2005). Overweight and obese patients in a primary care population report less sleep than patients with a normal body mass index. Archives of Internal Medicine, 165, 25–30.CrossRefGoogle Scholar
Waldron, I. (1991). Patterns and causes of gender differences in smoking. Social Science and Medicine, 32, 989–1005.CrossRefGoogle ScholarPubMed
Wallace, B. A. and Cumming, R. G. (2000). Systematic review of randomized trials of the effect of exercise on bone mass in pre- and postmenopausal women. Calcified Tissue International, 67, 10–18.CrossRefGoogle Scholar
Walters, V. (1993). Stress, anxiety and depression: women's accounts of their health problems. Social Science and Medicine, 36, 393–402.CrossRefGoogle ScholarPubMed
Wang, C., Catlin, D. H., Starcevic, B.et al. (2005). Low-fat high-fiber diet decreased serum and urine androgens in men. Journal of Clinical Endocrinology and Metabolism, 90, 3550–9.CrossRefGoogle ScholarPubMed
Wang, Y. Z., Mi, J., Shan, X. -Y., Wang, Q. J. and Ge, K. -Y. (2007). Is China facing an obesity epidemic and the consequences? The trends in obesity and chronic disease in China. International Journal of Obesity, 31, 177–88.CrossRefGoogle Scholar
Warren, M. (2004). A comparative review of the risks and benefits of hormone replacement therapy regimens. American Journal of Obstetrics and Gynecology, 190, 1141–67.CrossRefGoogle ScholarPubMed
Watts, J. (2006). Doctors blame air pollution for China's asthma increases. The Lancet, 368, 719–720.CrossRefGoogle ScholarPubMed
Weedon, M. N., Schwarz, P. E. H., Horikawa, Y.et al. (2003). Meta-analysis and a large association study confirm a role for Calpain-10 variation in type 2 diabetes susceptibility. American Journal of Human Genetics, 73, 1208–12.CrossRefGoogle Scholar
Weir, G., Laybutt, D., Kaneto, H., Bonner-Weir, S. and Sharma, A. (2001). Beta-cell adaptation and decompensation during the progression to diabetes. Diabetes, 50 (Suppl.1), 5154–9.CrossRefGoogle ScholarPubMed
Weiss, K. M. and Fullerton, S. M. (2005). Racing around, getting nowhere. Evolutionary Anthropology, 14, 165–9.CrossRefGoogle Scholar
Weiss, K. M., Ferrell, R. E. and Hanis, C. L. (1984). A New World Syndrome of metabolic diseases with a genetic and evolutionary basis. Yearbook of Physical Anthropology, 27, 153–78.CrossRefGoogle Scholar
Wells, J. C. K. (2006). The evolution of human fatness and susceptibility to obesity: an ethological approach. Biological Reviews, 81, 183–205.CrossRefGoogle ScholarPubMed
Wenzel, S. E. (2006). Asthma: defining of the persistent adult phenotypes. The Lancet, 368, 804–813.CrossRefGoogle ScholarPubMed
West, K. M. (1974). Diabetes in American Indians and other native populations of the New World. Diabetes, 23, 841–55.CrossRefGoogle ScholarPubMed
West, R. (2006). Tobacco control: present and future. British Medical Bulletin, 77–78, 123–36.CrossRefGoogle ScholarPubMed
Whitmer, R. A., Gunderson, E. P., Barrett-Connor, E., Quesenberry, C. P. and Yaffe, K. (2005). Obesity in middle age and future risk of dementia: a 27 year longitudinal population based study. British Medical Journal, 330, 1360–2.CrossRefGoogle ScholarPubMed
WHO Expert Consultation (2004). Appropriate body-mass index for Asian populations and its implications for policy and intervention strategies. The Lancet, 363, 157–63.CrossRefGoogle Scholar
WHO Global Infobase Team (2005). The SuRF Report 2. Surveillance of Chronic Disease Risk Factors: Country-Level Data and Comparable Estimates. Geneva: World Health Organization.
WHO International Consortium in Psychiatric Epidemiology (2000). Cross-national comparisons of the prevalences and correlates of mental disorders. Bulletin of the World Health Organization, 78, 413–26.Google Scholar
Wild, S., Roglic, G., Green, A., Sicree, R. and King, H. (2004). Global prevalence of diabetes: estimates for the year 2000 and projections for 2030. Diabetes Care, 27, 1047–53.CrossRefGoogle ScholarPubMed
Wild, S. H. and Byrne, C. D. (2006). Risk factors for diabetes and coronary heart disease. British Medical Journal, 333, 1009–11.CrossRefGoogle ScholarPubMed
Wilkinson, R. (1999). Health, hierarchy, and social anxiety. Annals of the New York Academy of Sciences, 896, 48–63.CrossRefGoogle ScholarPubMed
Williams, B. (1995). Westernised Asians and cardiovascular disease: nature or nurture. The Lancet, 345, 401–2.CrossRefGoogle ScholarPubMed
Williams, D. R. and Collins, C. (1995). US socioeconomic and racial differ-entials in health: patterns and explanations. Annual Review of Sociology, 21, 349–86.CrossRefGoogle Scholar
Williams, D. R., Neighbors, H. W. and Jackson, J. S. (2003). Racial/ethnic discrimination and health: findings from community studies. American Journal of Public Health, 93, 200–8.CrossRefGoogle ScholarPubMed
Williams, G. C. (1957). Pleiotropy, natural selection, and the evolution of senescence. Evolution, 11, 398–411.CrossRefGoogle Scholar
Williams, G. C. and Nesse, R. M. (1991). The dawn of Darwinian medicine. Quarterly Review of Biology, 66, 1–22.CrossRefGoogle ScholarPubMed
Williams, R. C., Long, J. C., Hanson, R. L., Sievers, M. L. and Knowler, W. C. (2000). Individual estimates of European genetic admixture associated with lower body-mass index, plasma glucose, and prevalence of type 2 diabetes in Pima Indians. American Journal of Human Genetics, 66, 527–38.CrossRefGoogle ScholarPubMed
Wilmoth, J. R. (2000). Demography of longevity: past, present, and future trends. Experimental Gerontology, 35, 1111–29.CrossRefGoogle ScholarPubMed
Wilson, B. D., Wilson, N. C. and Russell, D. G. (2001). Obesity and body fat distribution in the New Zealand population. New Zealand Medical Journal, 114, 127–30.Google ScholarPubMed
Wilson, T. W. and Grim, C. E. (1991). Biohistory of slavery and blood pressure differences in blacks today. A hypothesis. Hypertension, 17 (Suppl. 1), 1122–9.CrossRefGoogle ScholarPubMed
Wood, B. and Collard, M. (1999). The human genus. Science, 284, 65–71.CrossRefGoogle ScholarPubMed
World Health Organization (1999). World Health Report: Making a Difference. Geneva, World Health Organization.
World Health Organization (2003). Diet, Nutrition and the Prevention of Chronic Diseases. Geneva: World Health Organization.
Worthman, C. M. and Melby, M. K. (2002). Toward a comparative developmental ecology of human sleep. In Adolescent Sleep Patterns: Biological, Social and Psychological Influences, ed. Carskadon, M. A.. New York: Cambridge University Press, pp. 69–117.CrossRefGoogle Scholar
Wrigley, E. A. (1969). Population and History. London: Weidenfeld and Nicolson.
Writing Group for the Women's Health Initiative Investigators (2002). Risks and benefits of estrogen plus progestin in healthy postmenopausal women: principal results from the Women's Health Initiative randomized controlled trial. Journal of the American Medical Association, 288, 321–33.CrossRefGoogle Scholar
Wu, A. H., Pike, M. C. and Stram, D. O. (1999). Meta-analysis: dietary fat intake, serum estrogen levels, and the risk of breast cancer. Journal of the National Cancer Institute, 91, 529–34.CrossRefGoogle ScholarPubMed
Wu, Y. (2006). Overweight and obesity in China. British Medical Journal, 333, 362–3.CrossRefGoogle Scholar
Yajnik, C. (2000). Interactions of perturbations in intrauterine growth and growth during childhood on the risk of adult-onset disease. Proceedings of the Nutrition Society, 59, 257–65.CrossRefGoogle ScholarPubMed
Yajnik, C., Lubree, H., Rege, S.et al. (2002). Adiposity and hyperinsulinemia in Indians are present at birth. Journal of Clinical Endocrinology and Metabolism, 87, 5575–80.CrossRefGoogle ScholarPubMed
Yang, L., Parkin, D. M., Ferlay, J., Li, L. and Chen, Y. (2005). Estimates of cancer incidence in China for 2000 and projections for 2005. Cancer Epidemiology, Biomarkers and Prevention, 14, 243–50.Google ScholarPubMed
Yazdanbakhsh, M., Kremsner, P. G. and van Ree, R. (2002). Allergy, parasites, and the hygiene hypothesis. Science, 296, 490–4.CrossRefGoogle ScholarPubMed
Yemaneberhan, H., Bekele, Z., Venn, A., Lewis, S., Parry, E. and Britton, J. (1997). Prevalence of wheeze and asthma and relation to atopy in urban and rural Ethiopia. The Lancet, 350, 85–90.CrossRefGoogle Scholar
Yemaneberhan, H., Flohr, C., Lewis, S. A.et al. (2004). Prevalence and associated factors of atopic dermatitis symptoms in rural and urban Ethiopia. Clinical and Experimental Allergy, 34, 779–85.CrossRefGoogle ScholarPubMed
Yoon, K. -H., Lee, J. -H., Kim, J. -W.et al. (2006). Epidemic obesity and type 2 diabetes in Asia. The Lancet, 368, 1681–8.CrossRefGoogle Scholar
Young, D. B., Lin, H. and McCabe, R. D. (1995). Potassium's cardiovascular protective mechanisms. American Journal of Physiology: Regulatory, Integrative and Comparative Physiology, 37, R825–37.Google Scholar
Young, J. H., Chang, Y. P. C., Kim, J. D. O.et al. (2005). Differential susceptibility to hypertension is due to selection during the out-of-Africa expansion. PLoS Genetics, 1, 730–8.CrossRefGoogle ScholarPubMed
Young, T. K., Reading, J., Elias, B. and O'Neil, J. D. (2000). Type 2 diabetes mellitus in Canada's First Nations: status of an epidemic in progress. Canadian Medical Association Journal, 163, 561–6.Google ScholarPubMed
Yusuf, S., Reddy, S., Ounpuu, S. and Anand, S. (2001a). Global burden of cardiovascular diseases. Part I: General considerations, the epidemiologic transition, risk factors, and the impact of urbanization. Circulation, 104, 2746–53.CrossRefGoogle Scholar
Yusuf, S., Reddy, S., Ôunpuu, S. and Anand, S. (2001b). Global burden of cardiovascular diseases. Part II: Variations in cardiovascular disease by specific ethnic groups and geographic regions and prevention strategies. Circulation, 104, 2855–64.CrossRefGoogle Scholar
Zainudin, B. M. Z., Lai, C. K. W., Sporiano, J. B., Jia-Horng, W. and Guia, T. S. (2005). Asthma control in adults in Asia–Pacific. Respirology, 10, 579–86.CrossRefGoogle ScholarPubMed
Zhou, B. F., Stamler, J., Dennis, B.et al. (2003). Nutrient intakes of middle-aged men and women in China, Japan, United Kingdom, and United States in the late 1990s: the INTERMAP study. Journal of Human Hypertension, 17, 623–30.CrossRefGoogle ScholarPubMed
Ziegler, R. G., Hoover, R. N., Pike, M. C.et al. (1993). Migration patterns and breast cancer risk in Asian-American women. Journal of the National Cancer Institute, 85, 1819–27.CrossRefGoogle ScholarPubMed
Zimmet, P. (2000). Globalization, coca-colonization and the chronic disease epidemic: can the Doomsday scenario be averted?Journal of Internal Medicine, 247, 301–10.CrossRefGoogle ScholarPubMed
Zimmet, P., Taft, P., Guinea, A., Guthrie, W. and Thoma, K. (1977). The high prevalence of diabetes mellitus on a central Pacific island. Diabetologia, 13, 111–15.CrossRefGoogle ScholarPubMed
Zimmet, P., Faaiuso, S., Ainuu, J., Whitehouse, S., Milne, B. and DeBoer, W. (1981). The prevalence of diabetes in the rural and urban Polynesian population of Western Samoa. Diabetes, 30, 45–51.CrossRefGoogle Scholar
Zizza, C., Siega-Riz, A. M. and Popkin, B. M. (2001). Significant increases in young adults' snacking between 1977–1978 and 1994–1996 represents a cause of concern!Preventive Medicine, 32, 303–10.CrossRefGoogle Scholar
Zografos, G. C., Panou, M. and Panou, N. (2004). Common risk factors of breast and ovarian cancer: recent view. International Journal of Gynecogical Cancer, 14, 721–40.CrossRefGoogle ScholarPubMed
Zoratti, R. (1998). A review on ethnic differences in plasma triglyceride and high-density-lipoprotein cholesterol: is the lipid pattern the key factor for the low coronary heart disease rate in people of African origin?European Journal of Epidemiology, 14, 9–21.CrossRefGoogle ScholarPubMed
Abate, N., Carulli, L., Cabo-Chan, A., Chandalia, M., Snell, P. G. and Grundy, S. M. (2003). Genetic polymorphism PC-1 K121Q and ethnic susceptibility to insulin resistance. Journal of Clinical Endocrinology and Metabolism, 88, 5927–34.CrossRefGoogle ScholarPubMed
Adler, A. I., Stratton, I. M., Neil, A. W.et al. on behalf of the UK Prospective Diabetes Study Group (2000). Association of systolic blood pressure with macrovascular and microvascular complications of type 2 diabetes (UKPDS 36): prospective observational study. British Medical Journal, 321, 412–19.CrossRefGoogle ScholarPubMed
Adlerberth, I., Lindberg, E., Aberg, N.et al. (2005). Reduced enterobacterial and increased staphylococcal colonization of the infantile bowel: an effect of hygienic lifestyle?Pediatric Research, 59, 96–101.CrossRefGoogle ScholarPubMed
Adlercreutz, H. (2002). Phyto-estrogens and cancer. The Lancet Oncology, 3, 364–373.CrossRefGoogle ScholarPubMed
Agyemang, C. (2006). Rural and urban differences in blood pressure and hypertension in Ghana, West Africa. Public Health, 120, 525–33.CrossRefGoogle Scholar
Aiello, L. C. and Wells, J. C. K. (2002). Energetics and the evolution of the genus Homo. Annual Review of Anthropology, 31, 323–38.CrossRefGoogle Scholar
Aiello, L. C. and Wheeler, P. (1995). The expensive tissue hypothesis: the brain and digestive system in human and primate evolution. Current Anthropology, 36, 199–221.CrossRefGoogle Scholar
Akdis, M., Blaser, K. and Akdis, C. A. (2005). T regulatory cells in allergy: novel concepts in the pathogenesis, prevention, and treatment of allergic diseases. Journal of Allergy and Clinical Immunology, 116, 961–9.CrossRefGoogle Scholar
Alavanja, M. C. R., Hoppin, J. A. and Kamel, F. (2004). Health effects of chronic pesticide exposure: cancer and neurotoxicity. Annual Review of Public Health, 25, 155–97.CrossRefGoogle ScholarPubMed
Alberti, K. G. M. M., Zimmet, P. and Shaw, J. for the IDF Epidemiology Task Force Consensus Group (2005). The metabolic syndrome – a new worldwide definition. The Lancet, 366, 1059–62.CrossRefGoogle ScholarPubMed
Aligne, C. A., Auinger, P., Byrd, R. S. and Weitzman, M. (2000). Risk factors for pediatric asthma: contributions of poverty, race, and urban residence. American Journal of Respiratory and Critical Care Medicine, 162, 873–7.CrossRefGoogle ScholarPubMed
Allen, J. S. and Cheer, S. M. (1996). The non-thrifty genotype. Current Anthropology, 37, 831–42.CrossRefGoogle Scholar
Allen, N. B. and Badcock, P. B. T. (2003). The social risk hypothesis of depressed mood: evolutionary, psychosocial, and neurobiological processes. Psychological Bulletin, 129, 887–913.CrossRefGoogle Scholar
Allen, N. B. and Badcock, P. B. T. (2006). Darwinian models of depression: a review of evolutionary accounts of mood and mood disorders. Progress in Neuro-Psychopharmacology and Biological Psychiatry, 30, 815–26.CrossRefGoogle ScholarPubMed
Allen, N. E. and Key, T. J. (2000). The effects of diet on circulating sex hormone levels in men. Nutrition Research Reviews, 13, 159–84.CrossRefGoogle ScholarPubMed
Anand, S. S., Yusuf, S., Jacobs, R.et al. for the SHARE-AP Investigators (2001). Risk factors, atherosclerosis, and cardiovascular disease among Aboriginal people in Canada: the Study of Health Assessment and Risk Evaluation in Aboriginal Peoples (SHARE-AP). The Lancet, 358, 1147–53.CrossRefGoogle Scholar
Anderson, H. R. (1997). Air pollution and trends in asthma. In The Rising Trends in Asthma, ed. Chadwick, D. and Cardew, G.. New York: John Wiley and Sons, pp. 190–203.Google Scholar
Anderson, H. R., Ruggles, R., Strachan, D. P.et al. (2004). Trends in prevalence of symptoms of asthma, hay fever, and eczema in 12–14 year olds in the British Isles, 1995–2002: questionnaire survey. British Medical Journal, 328, 1052–3.CrossRefGoogle ScholarPubMed
Anderson, J. W., Johnstone, B. M. and Remley, D. T. (1999). Breast-feeding and cognitive development: a meta-analysis. American Journal of Clinical Nutrition, 70, 525–35.CrossRefGoogle ScholarPubMed
Anderson, M. (1988). Population Change in North-Western Europe, 1750–1850. Basingstoke, UK: Macmillan Education.CrossRefGoogle Scholar
Apter, D. and Vihko, R. (1990). Endocrine determinants of fertility: serum androgen concentrations during follow-up of adolescents into the third decade of life. Journal of Clinical Endocrinology and Metabolism, 71, 970–4.CrossRefGoogle ScholarPubMed
Apter, D., Reinila, M. and Vihko, R. (1989). Some endocrine characteristics of early menarche, a risk factor for breast cancer, are preserved into adulthood. International Journal of Cancer, 44, 783–7.CrossRefGoogle ScholarPubMed
Arnett, D. K., Xiong, B., McGovern, P. G., Blackburn, H. and Luepker, R. V. (2000). Secular trends in dietary macronutrient intake in Minneapolis-St. Paul, Minnesota, 1980–1992. American Journal of Epidemiology, 152, 868–73.CrossRefGoogle Scholar
Asher, M. I., Montefort, S., Björkstén, B.et al. and the ISAAC Phase Three Study Group (2006). Worldwide time trends in the prevalence of symptoms of asthma, allergic rhinoconjunctivitis, and eczema in childhood: ISAAC Phases One and Three repeat multicountry cross-sectional surveys. The Lancet, 368, 733–43.CrossRefGoogle ScholarPubMed
Asia–Pacific Cohort Studies Collaboration (2005). Smoking, quitting, and the risk of cardiovascular disease among women and men in the Asia-Pacific region. International Journal of Epidemiology, 34, 1036–45.CrossRefGoogle Scholar
Aspray, T. J., Mugusi, F., Rashid, S.et al. for the Essential Non-Communicable Disease Health Intervention Project (2000). Rural and urban differences in diabetes prevalence in Tanzania: the role of obesity, physical inactivity and urban living. Transactions of the Royal Society of Tropical Medicine and Hygiene, 94, 637–44.CrossRefGoogle Scholar
Astrup, A. and Finer, N. (2000). Redefining type 2 diabetes: ‘Diabesity’ or ‘Obesity dependent diabetes mellitus’?Obesity Reviews, 1, 57–9.CrossRefGoogle ScholarPubMed
Atwood, L. D., Heard-Costa, N. L., Cupples, L. A., Jaquish, C. E., Wilson, P. W. F. and D'Agostino, R. B. (2002). Genomewide linkage analysis of body mass index across 28 years of the Framingham Heart Study. American Journal of Human Genetics, 71, 1044–50.CrossRefGoogle ScholarPubMed
Baier, L. J. and Hanson, R. L. (2004). Genetic studies of the etiology of type 2 diabetes in Pima Indians. Diabetes, 53, 1181–6.CrossRefGoogle ScholarPubMed
Baker, P. T. (1984). Migration, genetics, and the degenerative diseases of South Pacific Islanders. In Migration and Mobility: Biosocial Aspects of Human Movement, ed. Boyce, A. J.. London: Taylor and Francis, pp. 209–39.Google Scholar
Baker, P. T. (1988). Infectious disease. In Human Biology: An Introduction to Human Evolution, Variation, Growth, and Adaptability, ed. Harrison, G. A., Tanner, J. M., Pilbeam, D. R. and Baker, P. T.. Oxford: Oxford University Press, pp. 508–28.Google Scholar
Baker, P. T., Hanna, J. M. and Baker, T. S., eds. (1986). The Changing Samoans: Behavior and Health in Transition. New York: Oxford University Press.Google Scholar
Balen, A. (1999). Pathogenesis of polycystic ovary syndrome – the enigma unravels?The Lancet, 354, 966–7.CrossRefGoogle ScholarPubMed
Ball, H. L. (2006). Parent–infant bed-sharing behavior: effects of feeding type and presence of father. Human Nature, 17, 301–18.CrossRefGoogle ScholarPubMed
Ball, H. L. and Klingaman, K. (2008). Breastfeeding and mother–infant sleep proximity: implications for infant care. In New Perspectives in Evolutionary Medicine, ed. Trevathan, W. R., Smith, E. O. and McKenna, J. J.. New York: Oxford University Press, pp. 226–241.Google Scholar
Barker, D. J. P. (1994). Mothers, Babies, and Disease in Later Life. London: BMJ Publishing Group.Google Scholar
Barker, D. J. P. and Osmond, C. (1986). Infant mortality, childhood nutrition and ischaemic heart disease in England and Wales. The Lancet, i, 1077–81.CrossRefGoogle Scholar
Barker, D. J. P., Osmond, C., Forsen, T. J., Kajantie, E. and Eriksson, J. G. (2005). Trajectories of growth among children who have coronary events as children. New England Journal of Medicine, 353, 1802–09.CrossRefGoogle ScholarPubMed
Barnes, K. C. (2006). Genetic epidemiology of health disparities in allergy and clinical immunology. Journal of Allergy and Clinical Immunology, 117, 243–54.CrossRefGoogle ScholarPubMed
Barnes, K. C., Armelagos, G. J. and Morreale, S. C. (1999). Darwinian medicine and the emergence of allergy. In Evolutionary Medicine, ed. Trevathan, W. R., Smith, E. O. and McKenna, J. J.. New York: Oxford University Press, pp. 209–43.Google Scholar
Barrett, R., Kuzawa, C. W., McDade, T. and Armelagos, G. J. (1998). Emerging and re-emerging infectious diseases: the third epidemiologic transition. Annual Review of Anthropology, 27, 247–71.CrossRefGoogle Scholar
Barroso, I. (2005). The genetics of type 2 diabetes. Diabetic Medicine, 22, 517–35.CrossRefGoogle ScholarPubMed
Baschetti, R. (1998). Diabetes epidemic in newly westernized populations: is it due to thrifty genes or to genetically unknown foods. Journal of the Royal Society of Medicine, 91, 622–25.CrossRefGoogle ScholarPubMed
Bassuk, S. S. and Manson, J. E. (2005). Epidemiological evidence for the role of physical activity in reducing risk of type 2 diabetes and cardiovascular disease. Journal of Applied Physiology, 99, 1193–204.CrossRefGoogle ScholarPubMed
Ben-Shlomo, Y. and Smith, Davey G. (1991). Deprivation in infancy or in adult life: which is more important for mortality risk?The Lancet, 337, 530–4.CrossRefGoogle ScholarPubMed
Bennett, K., Kabir, Z., Unal, B.et al. (2006). Explaining the recent decrease in coronary heart disease mortality rates in Ireland, 1985–2000. Journal of Epidemiology and Community Health, 60, 322–7.CrossRefGoogle Scholar
Bentley, G. R. (2000). Environmental pollutants and fertility. In Infertility in the Modern World: Present and Future Prospects, ed. Bentley, G. R. and Mascie-Taylor, C. G. N.. Cambridge: Cambridge University Press, pp. 85–152.CrossRefGoogle Scholar
Bentley, G. R., Jasieńska, G. and Goldberg, T. (1993). Is the fertility of agriculturalists higher than that of nonagriculturalists?Current Anthropology, 34, 778–85.CrossRefGoogle Scholar
Bentley, G. R., Harrigan, A. M. and Ellison, P. T. (1998). Dietary composition and ovarian function among Lese horticulturalist women of the Ituri Forest, Democratic Republic of Congo. European Journal of Clinical Nutrition, 52, 261–70.CrossRefGoogle ScholarPubMed
Bentley, G. R., Paine, R. R. and Boldsen, J. L. (2001). Fertility changes with the prehistoric transition to agriculture. In Reproductive Ecology and Human Evolution, ed. Ellison, P.. New York: Aldine de Gruyter, pp. 201–31.Google Scholar
Benyshek, D. C. and Watson, J. T. (2006). Exploring the thrifty genotype's food-shortage assumptions: a cross-cultural comparison of ethnographic accounts of food security among foraging and agricultural societies. American Journal of Physical Anthropology, 131, 120–6.CrossRefGoogle ScholarPubMed
Beral, V., Bull, D., Doll, R.et al. (1997). Breast cancer and hormone replacement therapy: collaborative reanalysis of data from 51 epidemiological studies of 52,705 women with breast cancer and 108,411 women without breast cancer. The Lancet, 350, 1047–59.Google Scholar
Beral, V., Bull, D., Doll, R.et al. (2002). Breast cancer and breastfeeding: collaborative reanalysis of individual data from 47 epidemiological studies in 30 countries, including 50,302 women with breast cancer and 969,973 women without the disease. The Lancet, 360, 187–95.Google Scholar
Beran, D. and Yudkin, J. S. (2006). Diabetes care in sub-Saharan Africa. The Lancet, 368, 1689–95.CrossRefGoogle ScholarPubMed
Bernstein, L. and Ross, R. K. (1993). Endogenous hormones and breast cancer risk. Epidemiologic Reviews, 15, 48–65.CrossRefGoogle ScholarPubMed
Bernstein, L., Yuan, J. M., Ross, R. K.et al. (1990). Serum hormone levels in pre-menopausal Chinese women in Shanghai and white women in Los Angeles: results from two breast cancer case-control studies. Cancer Causes and Control, 1, 51–8.CrossRefGoogle ScholarPubMed
Bernstein, L., Henderson, B. E., Hanisch, R., Sullivan-Halley, J. and Ross, R. K. (1994). Physical exercise and reduced risk of breast cancer in young women. Journal of the National Cancer Institute, 86, 1403–08.CrossRefGoogle ScholarPubMed
Berrington, A. (2004). Perpetual postponers? Women's, men's and couple's infertility intentions and subsequent fertility behaviour. Population Trends, 117, 9–19.Google Scholar
Biesele, M. and Howell, N. (1981). “The old people give you life”: aging among!Kung hunter–gatherers. In Other Ways of Growing Old, ed. Amoss, P. and Harrell, S.. Stanford: Stanford University Press, pp. 77–98.Google Scholar
Bilsborough, S. and Mann, N. (2006). A review of issues of dietary protein intake in humans. International Journal of Sport Nutrition and Exercise Metabolism, 16, 129–52.CrossRefGoogle ScholarPubMed
Bindon, J. R. and Baker, P. T. (1997). Bergmann's rule and the thrifty genotype. American Journal of Physical Anthropology, 104, 201–10.3.0.CO;2-0>CrossRefGoogle ScholarPubMed
Bjerve, K. S., Brubakk, A. M., Fougner, K. H., Johnsen, H., Midthjell, K. and Vik, T. (1993). Omega-3 fatty acids: essential fatty acids with important biological effects, and serum phospholipid fatty acids as markers of dietary omega-3 fatty acid intake. American Journal of Clinical Nutrition, 57 (Suppl.), 801S–6S.CrossRefGoogle ScholarPubMed
Björkstén, B., Naaber, P., Sepp, E. and Mikelsaar, M. (1999). The intestinal microbiota in allergic Estonian and Swedish 2-year-old children. Clinical and Experimental Allergy, 29, 342–6.CrossRefGoogle Scholar
Björntorp, P. and Rosmond, R. (2000). Obesity and cortisol. Nutrition, 16, 924–36.CrossRefGoogle ScholarPubMed
Blackley, R. (2001). The arrival of the Maoris in New Zealand, 1898. Emerging Infectious Diseases, 7, 914.CrossRefGoogle Scholar
Blair, P. S. and Ball, H. L. (2004). The prevalence and characteristics associated with parent–infant bed-sharing in England. Archives of Disease in Childhood, 89, 1106–10.CrossRefGoogle ScholarPubMed
Blumenthal, M. N., Langefeld, C. D., Beaty, T. H.et al. (2004). A genome-wide search for allergic response (atopy) genes in three ethnic groups: Collaborative Study on the Genetics of Asthma. Human Genetics, 114, 157–64.CrossRefGoogle Scholar
Jones, Blurton N. G., Hawkes, K. and O'Connell, J. F. (2002). Antiquity of postreproductive life: are there modern impacts on hunter–gatherer postreproductive life spans?American Journal of Human Biology, 14, 184–205.CrossRefGoogle Scholar
Bolling, K. (2006). Infant Feeding Survey 2005: Early Results. London: Information Centre, Government Statistical Service, p. 19.Google Scholar
Bonnet, M. H. and Arand, D. L. (1995). We are chronically sleep deprived. Sleep, 18, 908–11.CrossRefGoogle ScholarPubMed
Botha, J. L., Bray, F., Sankila, R. and Parkin, D. M. (2003). Breast cancer incidence and mortality trends in 16 European countries. European Journal of Cancer, 39, 1718–29.CrossRefGoogle ScholarPubMed
Bowlby, J. (1969). Attachment and Loss. New York: Basic Books.Google Scholar
Boyd, R. and Silk, J. B. (2006). How Humans Evolved. Los Angeles, University of California: University of California Press.Google Scholar
Boyden, S. V., ed. (1970). The Impact of Civilization on the Biology of Man. Canberra: Australian National University Press.Google Scholar
Boyden, S. V. (1987). Western Civilization in Biological Perspective: Patterns in Biohistory. Oxford: Oxford University Press.Google Scholar
Braman, S. S. (2006). The global burden of asthma. Chest, 130, 4S–12S.CrossRefGoogle Scholar
Braun, L. (2002). Race, ethnicity and health: can genetics explain disparities?Perspectives in Biology and Medicine, 45, 159–74.CrossRefGoogle ScholarPubMed
Bribiescas, R. G. (2001a). Reproductive ecology and life history of the human male. Yearbook of Physical Anthropology, 44, 148–76.CrossRefGoogle Scholar
Bribiescas, R. G. (2001b). Reproductive physiology of the human male: an evolutionary and life history perspective. In Reproductive Ecology and Human Evolution, ed. Ellison, P. T.. New York: Aldine de Gruyter, pp. 107–35.Google Scholar
Briefel, R. R. and Johnson, C. L. (2004). Secular trends in dietary intake in the United States. Annual Review of Nutrition, 24, 401–31.CrossRefGoogle ScholarPubMed
Brock, J. (1993). Native Plants of Northern Australia. New Holland: Frenchs Forest, Reed.Google Scholar
Brown, D. E. (2006). Measuring hormonal variation in the sympathetic nervous system: catecholamines. In Measuring Stress in Humans: A Practical Guide for the Field, ed. Ice, G. H. and James, G. D.. Cambridge: Cambridge University Press, pp. 94–121.CrossRefGoogle Scholar
Brown, E. S., Varghese, F. P. and McEwen, B. S. (2004). Association of depression with medical illness: does cortisol play a role?Biological Psychiatry, 55, 1–9.CrossRefGoogle ScholarPubMed
Brown, P. J. and Bentley-Condit, V. K. (1998). Culture, evolution, and obesity. In Handbook of Obesity, ed. Bray, G. A., Bouchard, C. and James, W. P. T.. New York: Marcel Dekker, pp. 143–55.Google Scholar
Brown, P. J. and Konner, M. (1987). An anthropological perspective on obesity. Annals of the New York Academy of Sciences, 499, 29–46.CrossRefGoogle ScholarPubMed
Brown, P. J. and Krick, S. V. (2001). The etiology of obesity: diet, television and the illusions of personal choice. In Obesity, Growth and Development, ed. Johnston, F. E. and Foster, G. D.. London: Smith-Gordon, pp. 111–28.Google Scholar
Brunner, E. and Marmot, M. (1999). Social organization, stress, and health. In Social Determinants of Health, ed. Marmot, M. and Wilkinson, R. G.. Oxford: Oxford University Press, pp. 17–43.Google Scholar
Burkitt, D. P. (1973). Some diseases characteristic of modern western civilization. British Medical Journal, 1, 274–8.CrossRefGoogle ScholarPubMed
Butte, N. F. (2001). The role of breastfeeding in obesity. Pediatric Clinics of North America, 48, 189–98.CrossRefGoogle ScholarPubMed
Cacioppo, J. T. and Hawkley, L. C. (2003). Social isolation and health, with an emphasis on underlying mechanisms. Perspectives in Biology and Medicine, 46, S39–52.CrossRefGoogle ScholarPubMed
Caldwell, J. C. (2001). Demographers and the study of mortality: scope, perspectives and theory. Annals of the New York Academy of Sciences, 954, 19–34.CrossRefGoogle Scholar
Calle, E. E. and Kaaks, R. (2004). Overweight, obesity and cancer: epidemiological evidence and proposed mechanisms. Nature Reviews Cancer, 4, 579–91.CrossRefGoogle ScholarPubMed
Calvani, M., Alessandri, C. and Bonci, E. (2002). Fever episodes in early life and the development of atopy in children with asthma. European Respiratory Journal, 20, 391–6.CrossRefGoogle ScholarPubMed
Camacho, T. C., Roberts, R. E., Lazarus, N. B., Kaplan, G. A. and Cohen, R. D. (1991). Physical activity and depression: evidence from the Alameda County Study. American Journal of Epidemiology, 134, 220–31.CrossRefGoogle ScholarPubMed
Carmichael, A. R. and Bates, T. (2004). Obesity and breast cancer: a review of the literature. The Breast, 13, 85–92.CrossRefGoogle ScholarPubMed
Carter, H. (1999). Urban origins: a review of theories. In The Pre-Industrial Cities and Technology Reader, ed. Chant, C.. London: Routledge and The Open University, pp. 7–14.Google Scholar
Carulli, L., Rondinella, S., Lombardini, S., Canedi, I., Loria, P. and Carulli, N. (2005). Diabetes, genetics and ethnicity. Alimentary Pharmacology and Therapeutics, 22 (Suppl. 2), 16–19.CrossRefGoogle ScholarPubMed
Catanese, D. M., Koetting O'Byrne, K. and Poston, W. S. C. (2001). The epidemiology of obesity in developed countries. In Obesity, Growth and Development, ed. Johnston, F. and Foster, G.. London: Smith-Gordon, pp. 69–90.Google Scholar
Cavalli-Sforza, L. L. and Feldman, M. W. (2003). The application of molecular genetic approaches to the study of human evolution. Nature Genetics, 33, 266–75.CrossRefGoogle Scholar
Cavallo, M. G., Fava, D., Monetini, L., Barone, F. and Pozzilli, P. (1996). Cell-mediated immune response to β casein in recent-onset insulin dependent diabetes: implications for disease pathogenesis. The Lancet, 348, 926–8.CrossRefGoogle ScholarPubMed
Chisholm, J. S. and Burbank, V. K. (2001). Evolution and inequality. International Journal of Epidemiology, 30, 206–11.CrossRefGoogle ScholarPubMed
Chlebowski, R. T., Blackburn, G. L., Thomson, C. A.et al. (2006). Dietary fat reduction and breast cancer outcome: interim efficacy results from the Women's Intervention Nutrition Study. Journal of the National Cancer Institute, 98, 1767–76.CrossRefGoogle ScholarPubMed
Chobanian, A. V. and Alexander, R. W. (1996). Exacerbation of atherosclerosis by hypertension. Archives of Internal Medicine, 156, 1952–6.CrossRefGoogle ScholarPubMed
Choi, Y. J., Cho, Y. M., Park, C. K.et al. (2006). Rapidly increasing diabetes-related mortality with socio-environmental changes in South Korea during the last two decades. Diabetes Research and Clinical Practice, 74, 295–300.CrossRefGoogle ScholarPubMed
Chow, C. K., Raju, P. K., Raju, R.et al. (2006). The prevalence and management of diabetes in rural India. Diabetes Care, 29, 1717–18.CrossRefGoogle ScholarPubMed
Cleave, T. L., Campbell, G. D. and Painter, N. S. (1969). Diabetes, Coronary Thrombosis, and the Saccharine Disease. Bristol: Wright.Google Scholar
Coe, K. and Steadman, L. B. (1995). The human breast and the ancestral reproductive cycle: a preliminary inquiry into breast cancer etiology. Human Nature, 6, 197–220.CrossRefGoogle ScholarPubMed
Cohen, M. N. (1989). Health and the Rise of Civilization. New Haven: Yale University Press.Google Scholar
Cohen, S. (2005). The Pittsburgh common cold studies: psychosocial predictors of susceptibility to respiratory infectious illness. International Journal of Behavioral Medicine, 12, 123–31.CrossRefGoogle ScholarPubMed
Cohen, S., Doyle, W. and Baum, A. (2006). Socioeconomic status is associated with stress hormones. Psychosomatic Medicine, 68, 414–20.CrossRefGoogle ScholarPubMed
Colagiuri, S., Colagiuri, R., Na'ati, S., Muimuiheata, S., Hussain, Z. and Palu, T. (2002). The prevalence of diabetes in the Kingdom of Tonga. Diabetes Care, 25, 1378–83.CrossRefGoogle ScholarPubMed
Cole, T. J., Bellizzi, M. C., Flegal, K. M. and Dietz, W. H. (2000). Establishing a standard definition for child overweight and obesity worldwide: international survey. British Medical Journal, 320, 1240–3.CrossRefGoogle ScholarPubMed
Colla, J., Buka, S., Harrington, D. and Murphy, J. M. (2006). Depression and modernization: a cross-cultural study of women. Social Psychiatry and Psychiatric Epidemiology, 41, 271–9.CrossRefGoogle Scholar
Collins, P. (2001). GPs and stress discourse: a preliminary report. Anthropology in Action, 8, 36–44.Google Scholar
Cook, D. C. (1984). Subsistence and health in the lower Illinois Valley: osteological evidence. In Palaeopathology at the Origins of Agriculture, ed. Cohen, M. N. and Armelagos, G. J.. Orlando, Florida: Academic Press, pp. 235–69.Google Scholar
Cooke, A., Zaccone, P., Raine, T., Phillips, J. M. and Dunne, D. W. (2004). Infection and autoimmunity: are we winning the war, only to lose the peace?Trends in Parasitology, 20, 316–21.CrossRefGoogle ScholarPubMed
Cookson, W. O. and Moffatt, M. F. (1997). Asthma: an epidemic in the absence of infection?Science, 275, 41–2.CrossRefGoogle Scholar
Cooper, P. J. (2004). Intestinal worms and human allergy. Parasite Immunology, 26, 455–67.CrossRefGoogle ScholarPubMed
Cooper, R. S., Rotimi, C., Ataman, S.et al. (1997). The prevalence of hypertension in seven populations of West African origin. American Journal of Public Health, 87, 160–8.CrossRefGoogle ScholarPubMed
Cooper, R. S., Amoah, A. G. and Mensah, G. A. (2003). High blood pressure: the foundation for epidemic cardiovascular disease in African populations. Ethnicity and Disease, 13 (Suppl. 2), S48–52.Google ScholarPubMed
Cordain, L., Gotshall, R. W. and Eaton, S. B. (1997). Evolutionary aspects of exercise. World Review of Nutrition and Dietetics, 81, 49–60.CrossRefGoogle Scholar
Cordain, L., Miller, Brand J., Eaton, S. B., Mann, N., Holt, S. H. A. and Speth, J. D. (2000). Plant–animal subsistence ratios and macronutrient energy estimations in worldwide hunter–gatherer diets. American Journal of Clinical Nutrition, 71, 682–92.CrossRefGoogle ScholarPubMed
Cordain, L., Eaton, S. B., Miller, Brand J., Mann, N. and Hill, K. (2002). The paradoxical nature of hunter–gatherer diets: meat-based, yet non-atherogenic. European Journal of Clinical Nutrition, 56 (Suppl. 1), S42–52.CrossRefGoogle ScholarPubMed
Cordain, L., Eades, M. R. and Eades, M. D. (2003). Hyperinsulinemic diseases of civilization: more than just Syndrome X. Comparative Biochemistry and Physiology Part A, 136, 95–112.CrossRefGoogle ScholarPubMed
Cordain, L., Eaton, S. B., Sebastian, A.et al. (2005). Origins and evolution of the Western diet: health implications for the 21st century. American Journal of Clinical Nutrition, 81, 341–54.CrossRefGoogle ScholarPubMed
Cosmides, L. and Tooby, J. (1997). Evolutionary Psychology: a Primer. University of California Santa Barbara: University of California Press.Google Scholar
Costa, D. L. (2000). Understanding the twentieth-century decline in chronic conditions among older men. Demography, 37, 53–72.CrossRefGoogle ScholarPubMed
Costa, D. L. (2004). The measure of man and older age mortality: evidence from the Gould sample. Journal of Economic History, 64, 1–23.CrossRefGoogle Scholar
Craig, P., Halavatau, V., Comino, E. and Caterson, I. (1999). Perception of body size in the Tongan community: differences from and similarities to an Australian sample. International Journal of Obesity, 23, 1288–94.CrossRefGoogle Scholar
Critser, G. (2003). Fat Land: How Americans became the Fattest People in the World. London: Penguin.Google Scholar
Crowther, R., Dinsdale, H., Rutter, H. and Kyffin, R. (2007). Analysis of the National Childhood Obesity Database 2005–06: South East Public Health Observatory.
Cuijpers, P., Straten, A. and Smit, F. (2005). Preventing the incidence of new cases of mental disorders – a meta-analytic review. Journal of Nervous and Mental Disease, 193, 119–25.CrossRefGoogle ScholarPubMed
Cunningham, A. S. (1995). Breastfeeding: adaptive behavior for child health and longevity. In Breastfeeding: Biocultural Perspectives, ed. Stuart-Macadam, P. and Dettwyler, K. A.. New York: Aldine de Gruyter, pp. 243–64.Google Scholar
Cunningham, G. R. (2006). Testosterone replacement therapy for late-onset hypogonadism. Nature Clinical Practice Urology, 3, 260–7.CrossRefGoogle ScholarPubMed
Daniel, M., Rowley, K. G., McDermott, R. and O'Dea, K. (2002). Diabetes and impaired glucose tolerance in Aboriginal Australians: prevalence and risk. Diabetes Research and Clinical Practice, 57, 23–33.CrossRefGoogle ScholarPubMed
Smith, Davey G. and Marmot, M. G. (1991). Trends in mortality in Britain: 1920–1986. Annals of Nutrition and Metabolism, 35 (Suppl. 1), 53–63.CrossRefGoogle ScholarPubMed
Davidson, K., Jonas, B., Dixon, K. and Markovitz, J. (2000). Do depression symptoms predict early hypertension incidence in young adults in the CARDIA study? Coronary artery risk development in young adults. Archives of Internal Medicine, 160, 1495–500.CrossRefGoogle ScholarPubMed
Davis, A. M., Kreutzer, R., Lipsett, M., King, G. and Shaikh, N. (2006). Asthma prevalence in Hispanic and Asian American ethnic subgroups: Results from the California Healthy Kids Survey. Pediatrics, 118, e363–70.CrossRefGoogle ScholarPubMed
Boever, E., Bacquer, D., Braeckman, L., Baele, L., Rosseneu, M. and Backer, G. (1995). Relation of fibrinogen to lifestyles and to cardiovascular risk factors in a working population. International Journal of Epidemiology, 24, 915–21.CrossRefGoogle Scholar
Jong, F. H., Oishi, K., Hayes, R. B.et al. (1991). Perpipheral hormone levels in controls and patients with prostatic cancer or benign prostatic hyperplasia – results from the Dutch-Japanese case-control study. Cancer Research, 51, 3445–50.Google ScholarPubMed
Laet, C. E. D. H. and Pols, H. A. P. (2000). Fractures in the elderly: epidemiology and demography. Best Practice and Research in Clinical Endocrinology and Metabolism, 14, 171–9.CrossRefGoogle ScholarPubMed
Delisle, H. F., Rivard, M. and Ekoe, J. M. (1995). Prevalence estimates of diabetes and other cardiovascular risk factors in the two largest Algonquin communities of Quebec. Diabetes Care, 18, 1255–9.CrossRefGoogle Scholar
DeMattia, L., Lemont, L. and Meurer, L. (2006). Do interventions to limit sedentary behaviours change behaviour and reduce childhood obesity? A critical review of the literature. Obesity Reviews, 8, 69–81.CrossRefGoogle Scholar
Der, G., Batty, D. and Deary, I. J. (2006). Effect of breast feeding on intelligence in children: prospective study, sibling pairs analysis, and meta-analysis. British Medical Journal, 333, 945–50.CrossRefGoogle ScholarPubMed
Despres, J. P. (2006). Is visceral obesity the cause of the metabolic syndrome?Annals of Medicine, 38, 52–63.CrossRefGoogle ScholarPubMed
Dettwyler, K. A. (1995a). Beauty and the breast: the cultural context of breastfeeding in the United States. In Breastfeeding: Biocultural Perspectives, ed. Stuart-Macadam, P. and Dettwyler, K. A.. New York: Aldine de Gruyter, pp. 167–215.Google Scholar
Dettwyler, K. A. (1995b). A time to wean: the hominid blueprint for the natural age of weaning in modern human populations. In Breastfeeding: Biocultural Perspectives, ed. Stuart-Macadam, P. and Dettwyler, K. A.. New York: Aldine de Gruyter, pp. 39–74.Google Scholar
Diamond, J. (2003). The double puzzle of diabetes. Nature, 423, 599–602.CrossRefGoogle ScholarPubMed
Dietz, W. H. (1996). The role of lifestyle in health: the epidemiology and consequences of inactivity. Proceedings of the Nutrition Society, 55, 829–40.CrossRefGoogle ScholarPubMed
Ding, E. L., Song, Y., Malik, V. and Liu, S. (2006). Sex differences of endogenous sex hormones and risk of type 2 diabetes: a systematic review and meta-analysis. Journal of the American Medical Association, 295, 1288–99.CrossRefGoogle ScholarPubMed
Dowse, G. K., Zimmet, P. Z. and King, H. (1991). Relationship between prevalence of impaired glucose tolerance and NIDDM in a population. Diabetes Care, 14, 968–74.CrossRefGoogle ScholarPubMed
Dowse, G. K., Gareeboo, H., Alberti, K. G. M. M.et al. (1995). Changes in population cholesterol concentrations and other cardiovascular risk factor levels after five years of the non-communicable disease intervention programme in Mauritius. British Medical Journal, 311, 1255–9.CrossRefGoogle ScholarPubMed
Drake, A. J. and Walker, B. R. (2004). The intergenerational effects of fetal programming: non-genomic mechanisms for the inheritance of low birth weight and cardiovascular risk. Journal of Endocrinology, 180, 1–16.CrossRefGoogle ScholarPubMed
Dressler, W. W. (1995). Modeling biocultural interactions: examples from studies of stress and cardiovascular disease. Yearbook of Physical Anthropology, 38, 27–56.CrossRefGoogle Scholar
Dressler, W. W., Balieiro, M. C., Ribeiro, R. P. and Santos, Dos J. E. (2005). Cultural consonance and arterial blood pressure in urban Brazil. Social Science and Medicine, 61, 527–40.CrossRefGoogle ScholarPubMed
Dudley, R. (2000). Evolutionary origins of human alcoholism in primate frugivory. Quarterly Review of Biology, 75, 3–15.CrossRefGoogle ScholarPubMed
Duggirala, R., Almasy, L., Blangero, J.et al. (2003). American Diabetes Association GENNID Study Group: Further evidence for a type 2 diabetes susceptibility locus on chromosome 11q. Genetic Epidemiology, 24, 240–2.CrossRefGoogle Scholar
Dunaif, A. (1997). Insulin resistance and the polycystic ovary syndrome: mechanism and implications for pathogenesis. Endocrine Reviews, 18, 774–800.Google ScholarPubMed
Dunn, J. E. (1975). Cancer epidemiology in populations of the United States – with an emphasis on Hawaii and California – and Japan. Cancer Research, 35, 3240–5.Google ScholarPubMed
Dunne, D. W. and Cooke, A. (2005). A worm's eye view of the immune system: consequences for evolution of human autoimmune disease. Nature Reviews Immunology, 5, 420–6.CrossRefGoogle ScholarPubMed
Durham, W. H. (1991). Coevolution: Genes, Culture, and Human Diversity. Stanford, California: Stanford University Press.Google Scholar
Eaton, S. B. and Eaton, S. B. (1999a). Breast cancer in evolutionary context. In Evolutionary Medicine, ed. Trevathan, W. R., Smith, E. O. and McKenna, J. J.. New York: Oxford University Press, pp. 429–42.Google Scholar
Eaton, S. B. and Eaton, S. B. (1999b). Hunter–gatherers and human health. In The Cambridge Encyclopedia of Hunters and Gatherers, ed. Lee, R. B. and Daly, R.. Cambridge: Cambridge University Press, pp. 449–56.Google Scholar
Eaton, S. B., Konner, M. and Shostak, M. (1988a). Stone agers in the fast lane: chronic degenerative diseases in evolutionary perspective. American Journal of Medicine, 84, 739–49.CrossRefGoogle Scholar
Eaton, S. B., Shostak, M. and Konner, M. (1988b). The Paleolithic Prescription: A Program of Diet and Exercise and a Design for Living. New York: Harper and Row.Google Scholar
Eaton, S. B., Pike, M. C., Short, R. V.et al. (1994). Women's reproductive biology in evolutionary context. Quarterly Review of Biology, 69, 353–67.CrossRefGoogle Scholar
Eaton, S. B., Eaton, S. B. and Konner, M. J. (1999). Paleolithic nutrition revisited. In Evolutionary Medicine, ed. Trevathan, W. R., Smith, E. O. and McKenna, J. J.. New York: Oxford University Press, pp. 313–32.Google Scholar
Eaton, S. B., Eaton, S. B. and Cordain, L. (2002). Evolution, diet, and health. In Human Diet: Its Origin and Evolution, ed. Ungar, P. S. and Teaford, M. F.. Westport, Connecticut: Bergin and Garvey, pp. 7–18.Google Scholar
Eckersley, R. (2005). Is modern Western culture a health hazard?International Journal of Epidemiology, 35, 252–8.CrossRefGoogle ScholarPubMed
Eisenmann, J. C. (2003). Secular trends in variables associated with the metabolic syndrome of North American children and adolescents: review and synthesis. American Journal of Human Biology, 15, 786–94.CrossRefGoogle ScholarPubMed
Ekbom, A. (2006). The developmental environment and the early origins of cancer. In Developmental Origins of Health and Disease, ed. Gluckman, P. and Hanson, M. A.. Cambridge: Cambridge University Press, pp. 415–25.CrossRefGoogle Scholar
Elliott, P., Stamler, J., Nichols, R.et al. (1996). Intersalt revisited: further analyses of 24-hour sodium excretion and blood pressure within and across populations. British Medical Journal, 312, 1249–53.CrossRefGoogle ScholarPubMed
Ellison, G. T. H. and Jones, Rees I. (2002). Social identities and the ‘new genetics’: scientific and social consequences. Critical Public Health, 12, 265–82.CrossRefGoogle Scholar
Ellison, P. T. (1990). Human ovarian function and reproductive ecology: new hypotheses. American Anthropologist, 92, 933–52.CrossRefGoogle Scholar
Ellison, P. T. (1996). Developmental influences on adult ovarian hormonal function. American Journal of Human Biology, 8, 725–34.3.0.CO;2-S>CrossRefGoogle ScholarPubMed
Ellison, P. T. (1999). Reproductive ecology and reproductive cancers. In Hormones, Health, and Behavior, ed. Panter-Brick, C. and Worthman, C. M.. Cambridge: Cambridge University Press, pp. 184–209.Google Scholar
Ellison, P. T. (2001). On Fertile Ground: A Natural History of Human Reproduction. Cambridge, Massachusetts: Harvard University Press.Google Scholar
Ellison, P. T. and Lager, C. (1986). Moderate recreational running is associated with lowered salivary progesterone profiles in women. American Journal of Obstetrics and Gynecology, 154, 1000–3.CrossRefGoogle ScholarPubMed
Ellison, P. T., Lipson, S. F., O'Rourke, M. T.et al. (1993a). Population variation in ovarian function. The Lancet, 342, 433–434.CrossRefGoogle Scholar
Ellison, P. T., Panter-Brick, C., Lipson, S. F. and O'Rourke, M. T. (1993b). The ecological context of human ovarian function. Human Reproduction, 8, 2248–58.CrossRefGoogle Scholar
Ellison, P. T., Bribiescas, R. G., Bentley, G. R.et al. (2002). Population variation in age-related decline in male salivary testosterone. Human Reproduction, 17, 3251–3.CrossRefGoogle ScholarPubMed
Ellison, P. T. and Jasiénska, G. (2007). Constraint, pathology, and adaptation: how can we tell them apart?American Journal of Human Biology, 19, 622–30.CrossRefGoogle ScholarPubMed
Elsom, D. M. (1992). Atmospheric Pollution: A Global Problem. Oxford: Blackwell.Google Scholar
Emanuel, M. B. (1988). Hay fever, a post industrial revolution epidemic: a history of its growth during the 19th century. Clinical Allergy, 18, 295–304.CrossRefGoogle ScholarPubMed
Endogenous Hormones and Breast Cancer Collaborative Group (2003). Body mass index, serum sex hormones, and breast cancer risk in postmenopausal women. Journal of the National Cancer Institute, 95, 1218–26.CrossRefGoogle Scholar
Erdal, D., Whiten, A., Boehm, C. and Knauft, B. (1994). On human egalitarianism: an evolutionary product of Machiavellian status escalation?Current Anthropology, 35, 175–83.CrossRefGoogle Scholar
Ernst, J. and Cacioppo, J. (1999). Lonely hearts: psychological perspectives on loneliness. Applied and Preventive Psychology, 8, 1–22.CrossRefGoogle Scholar
Eshed, V., Gopher, A. and Hershkovitz, I. (2006). Tooth wear and dental pathology at the advent of agriculture: new evidence from the Levant. American Journal of Physical Anthropology, 130, 145–59.CrossRefGoogle ScholarPubMed
Ewald, P. (2002). Plague Time: The New Germ Theory of Disease. New York: Anchor Books.Google Scholar
Ewing, R., Schmid, T., Killingsworth, R., Zlot, A. and Raudenbush, S. (2003). Relationship between urban sprawl and physical activity, obesity, and morbidity. American Journal of Health Promotion, 5, 47–57.CrossRefGoogle Scholar
Ezzati, M., Hoorn, Vander S., Lawes, C. M. M.et al. (2005). Rethinking the “diseases of affluence” paradigm: global patterns of nutritional risk in relation to economic development. PLoS Medicine, 2, e133.CrossRefGoogle ScholarPubMed
Falkner, B., Sherif, K., Sumner, A. and Kushner, H. (1999). Hyperinsulinism and sex hormones in young adult African Americans. Metabolism, 48, 107–12.CrossRefGoogle ScholarPubMed
Fall, C. H. D. (2001). Non-industrialised countries and affluence. British Medical Bulletin, 60, 33–50.CrossRefGoogle ScholarPubMed
Fanaro, S., Chierici, R., Guerrini, P. and Vigi, V. (2003). Intestinal microflora in early infancy: composition and development. Acta Paeditrica, 441 (Suppl.), 48–55.Google Scholar
FAO/WHO/UNU Expert Consultation (1985). Energy and Protein Requirements. Geneva: World Health Organization.Google Scholar
Fee, M. (2006). Racializing narratives: obesity, diabetes and the “Aboriginal” thrifty genotype. Social Science and Medicine, 62, 2988–97.CrossRefGoogle ScholarPubMed
Fein, S. B. and Roe, B. (1998). The effect of work status on initiation and duration of breast-feeding. American Journal of Public Health, 88, 1042–6.CrossRefGoogle ScholarPubMed
Feldman, H. A., Goldstein, I., Hatzichkristou, D. G., Krane, R. J. and McKinlay, J. B. (1994). Impotence and its medical and psychosocial correlates: results of the Massachusetts Male Aging Study. Journal of Urology, 151, 54–61.CrossRefGoogle ScholarPubMed
Fiennes, R. (1978). Zoonoses and the Origins and Ecology of Human Disease. London: Academic Press.Google Scholar
Fildes, V. (1995). The culture and biology of breastfeeding: An historical review of western Europe. In Breastfeeding: Biocultural Perspectives, ed. Stuart-Macadam, P. and Dettwyler, K. A.. New York: Aldine de Gruyter, pp. 101–26.Google Scholar
Finch, C. E. and Crimmins, E. M. (2004). Inflammatory exposure and historical changes in human life-spans. Science, 305, 1736–9.CrossRefGoogle ScholarPubMed
Flaherman, V. and Rutherford, G. W. (2006). A meta-analysis of the effect of high weight on asthma. Archives of Disease in Childhood, 91, 334–9.CrossRefGoogle ScholarPubMed
Flegal, K. M., Carroll, M. D., Kuczmarski, R. J. and Johnson, C. (1998). Overweight and obesity in the United States: prevalence and trends, 1960–1994. International Journal of Obesity, 22, 39–47.CrossRefGoogle Scholar
Fleming, J. O. and Cook, T. D. (2006). Multiple sclerosis and the hygiene hypothesis. Neurology, 67, 2085–6.CrossRefGoogle ScholarPubMed
Fletcher, E. S., Rugg-Gunn, A. J., Matthews, J. N. S.et al. (2004). Changes over 20 years in macronutrient intake and body mass index in 11- to 12-year-old adolescents living in Northumberland. British Journal of Nutrition, 92, 321–33.CrossRefGoogle ScholarPubMed
Fogel, R. W. and Costa, D. L. (1997). A theory of technophysio evolution, with some implications for forecasting population, health care costs, and pension costs. Demography, 34, 49–66.CrossRefGoogle ScholarPubMed
Foley, R. A. (1993). The influence of seasonality on hominid evolution. In Seasonality and Human Ecology, ed. Ulijaszek, S. J. and Strickland, S. S.. Cambridge: Cambridge University Press, pp. 17–37.CrossRefGoogle Scholar
Foley, R. A. (1996). The adaptive legacy of human evolution: A search for the environment of evolutionary adaptedness. Evolutionary Anthropology, 4, 194–203.CrossRefGoogle Scholar
Foley, R. A. and Lee, P. C. (1991). Ecology and energetics of encephalisation in hominid evolution. Philosophical Transactions of the Royal Society of London B, 334, 223–32.CrossRefGoogle Scholar
Ford, E. S., Giles, W. H. and Dietz, W. H. (2002). Prevalence of the metabolic syndrome among US adults: findings from the Third National Health and Nutrition Examination Survey. Journal of the American Medical Association, 287, 356–9.CrossRefGoogle ScholarPubMed
Forouhi, N. and Sattar, N. (2006). CVD risk factors and ethnicity – a homogeneous relationship?Atherosclerosis Supplements, 7, 11–19.CrossRefGoogle ScholarPubMed
Fox, C. S., Pencina, M. J., Meigs, J. B., Vasan, R. S., Levitzky, Y. S. and D'Agostino, R. B. (2006). Trends in the incidence of type 2 diabetes mellitus from the 1970s to the 1990s. Circulation, 113, 2914–18.CrossRefGoogle ScholarPubMed
Friedman, J. M. (2004). Modern science versus the stigma of obesity. Nature Medicine, 10, 563–9.CrossRefGoogle ScholarPubMed
Friedman, N. J. and Zeiger, R. S. (2005). The role of breast-feeding in the development of allergies and asthma. Journal of Allergy and Clinical Immunology, 115, 1238–48.CrossRefGoogle ScholarPubMed
Frisch, R. E., Wyshak, G., Albright, N. L.et al. (1985). Lower prevalence of breast cancer and cancers of the reproductive system among former college athletes compared to non-athletes. British Journal of Cancer, 52, 885–91.CrossRefGoogle ScholarPubMed
Frost, G. and Dornhorst, A. (2001). Starting the day the right way. The Lancet, 357, 736–7.CrossRefGoogle ScholarPubMed
Fullerton, S. M., Bartoszewicz, A., Ybazeta, G.et al. (2002). Geographic and haplotype structure of putative type 2 diabetes susceptibility variants at the Calpain-10 locus. American Journal of Human Genetics, 70, 1096–106.CrossRefGoogle Scholar
Furberg, A. S., Jasieńska, G., Bjurstam, N.et al. (2005). Metabolic and hormonal profiles: HDL cholesterol as a plausible biomarker of breast cancer risk. The Norwegian EBBA study. Cancer Epidemiology, Biomarkers and Prevention, 14, 33–40.Google ScholarPubMed
Fuster, V., Badimon, L., Badimon, J. J. and Chesebro, J. H. (1992a). The pathogenesis of coronary artery disease and the acute coronary syndromes: Part I. New England Journal of Medicine, 326, 242–50.Google Scholar
Fuster, V., Badimon, L., Badimon, J. J. and Chesebro, J. H. (1992b). The pathogenesis of coronary artery disease and the acute coronary syndromes: Part II. New England Journal of Medicine, 326, 310–18.Google Scholar
Gage, T. B. (2005). Are modern environments really bad for us?: revisiting the demographic and epidemiologic transitions. Yearbook of Physical Anthropology, 48, 96–117.CrossRefGoogle Scholar
Gale, E. A. M. (2002). A missing link in the hygiene hypothesis?Diabetologia, 45, 588–94.CrossRefGoogle ScholarPubMed
Gallou-Kabani, C. and Junien, C. (2005). Nutritional epigenomics of metabolic syndrome. Diabetes, 54, 1899–906.CrossRefGoogle ScholarPubMed
Galloway, A. (1997). The cost of reproduction and the evolution of postmenopausal osteoporosis. In The Evolving Female: A Life-History Perspective, ed. Morbeck, M. E., Galloway, A. and Zihlman, A. L.. Princeton: Princeton University Press, pp. 132–46.Google Scholar
Gandy, M. and Zumla, A. (2002). The resurgence of disease: social and historical perspectives on the ‘new’ tuberculosis. Social Science and Medicine, 55, 385–96.CrossRefGoogle ScholarPubMed
Gann, P. H., Hennekens, C. H., Ma, J., Longcope, C. and Stampfer, M. J. (1996). Prospective study of sex hormone levels and risk of prostate cancer. Journal of the National Cancer Institute, 88, 1118–26.CrossRefGoogle ScholarPubMed
Gapstur, S. M., Gann, P. H., Kopp, P., Colangelo, L., Longcope, C. and Liu, K. (2002). Serem androgen concentrations in young men: a longitudinal analysis of associations with age, obesity, and race. The CARDIA Male Hormone study. Cancer Epidemiology, Biomarkers and Prevention, 11, 1041–7.Google Scholar
Gardell, B. (1987). Efficiency and health hazards in mechanized work. In Work Stress, ed. Quick, J. C., Bhagat, R., Dalton, J. and Quick, J. D.. New York: Praeger, pp. 50–71.Google Scholar
Gaulin, S. J. C. (1980). Sexual dimorphism in the human post-reproductive life-span: possible causes. Journal of Human Evolution, 9, 227–32.CrossRefGoogle Scholar
Gibson, M. A. and Mace, R. (2005). Helpful grandmothers in rural Ethiopia: A study of the effect of kin on child survival and growth. Evolution and Human Behavior, 26, 469–82.CrossRefGoogle Scholar
Glassman, A. H., Helzer, J. E., Covey, L. S., Stetner, F., Tipp, J. E. and Johnson, J. (1990). Smoking, smoking cessation, and major depression. Journal of the American Medical Association, 264, 1546–9.CrossRefGoogle ScholarPubMed
Gleibermann, L. (1973). Blood pressure and dietary salt in human populations. Ecology of Food and Nutrition, 2, 143–56.CrossRefGoogle Scholar
Gluckman, P. and Hanson, M. (2005). The Fetal Matrix: Evolution, Development and Disease. Cambridge: Cambridge University Press.Google Scholar
Goldin, B. R., Adlercreutz, H., Gorbach, S. L.et al. (1986). The relationship between estrogen levels and diets of Caucasian American and Oriental immigrant women. American Journal of Clinical Nutrition, 44, 945–53.CrossRefGoogle ScholarPubMed
Gould, D. C., Petty, R. and Jacobs, H. S. (2000). The male menopause - does it exist?British Medical Journal, 320, 858–61.CrossRefGoogle ScholarPubMed
Gracey, M. and Spargo, R. M. (1987). The state of health of Aborigines in the Kimberley region. Medical Journal of Australia, 146, 200–4.Google ScholarPubMed
Gray, P. B., Kruger, A., Huisman, H. M., Wissing, M. P. and Vorster, H. H. (2006). Predictors of South African male testosterone levels: the THUSA study. American Journal of Human Biology, 18, 123–32.CrossRefGoogle ScholarPubMed
Gray-Donald, K., Jacobs-Starkey, L. and Johnson-Down, L. (2000). Food habits of Canadians: reduction in fat intake over a generation. Canadian Journal of Public Health, 91, 381–5.Google ScholarPubMed
Greenwood, D. C., Muir, K. R., Packham, C. J. and Madeley, R. J. (1996). Coronary heart disease: a review of the role of psychosocial stress and social support. Journal of Public Health Medicine, 18, 221–31.CrossRefGoogle ScholarPubMed
Griffiths, K., Denis, L., Turkes, A. and Morton, M. S. (1998). Phytoestrogens and diseases of the prostate gland. Baillière's Clinical Endocrinology and Metabolism, 12, 625–47.CrossRefGoogle ScholarPubMed
Gröland, M. M., Lehtonen, O. -P., Eerola, E. and Kero, P. (1999). Fecal microflora in healthy infants born by different methods of delivery: permanent changes in intestinal flora after Cesarian delivery. Journal of Pediatric Gastroenterology and Nutrition, 28, 19–25.CrossRefGoogle Scholar
Gross, L. S., Ford, E. S. and Liu, S. (2004). Increased consumption of refined carbohydrates and the epidemic of type 2 diabetes in the United States: an ecologic assessment. American Journal of Clinical Nutrition, 79, 774–9.CrossRefGoogle ScholarPubMed
Grossman, H., Bergmann, C. and Parker, S. (2006). Dementia: a brief review. Mount Sinai Journal of Medicine, 73, 985–92.Google ScholarPubMed
Grundy, J., Matthews, S., Bateman, B., Dean, T. and Arshad, S. H. (2002). Rising prevalence of allergy to peanuts in children: data from two sequential cohorts. Journal of Allergy and Clinical Immunology, 110, 784–9.CrossRefGoogle Scholar
Gu, K., Cowie, C. C. and Harris, M. I. (1998). Mortality in adults with and without diabetes in a national cohort of the US population, 1971–1993. Diabetes Care, 21, 1138–45.CrossRefGoogle Scholar
Guarner, F., Bourdet-Sicard, R., Brandtzaeg, P.et al. (2006). Mechanisms of disease: the hygiene hypothesis. Nature Clinical Practice Gastroenterology and Hepatology, 3, 275–84.CrossRefGoogle ScholarPubMed
Guest, C. S., O'Dea, K., Hopper, J. L. and Larkins, R. G. (1993). Hyperinsulinaemia and obesity in Aborigines of south-eastern Australia, with comparisons from rural and urban Europid populations. Diabetes Research and Clinical Practice, 20, 155–64.CrossRefGoogle Scholar
Guise, J. M., Austin, D. and Morris, C. D. (2005). Review of case-control studies related to breastfeeding and reduced risk of childhood leukemia. Pediatrics, 116, e724–31.CrossRefGoogle ScholarPubMed
Guyton, A. C. (1986). Textbook of Medical Physiology. Philadelphia: WB Saunders.Google Scholar
Haffner, S. M. (2003). Insulin resistance, inflammation, and the prediabetic state. American Journal of Cardiology, 92 (Suppl.), 18J–26J.CrossRefGoogle ScholarPubMed
Hagen, E. H. (1999). The functions of postpartum depression. Evolution and Human Behavior, 20, 325–59.CrossRefGoogle Scholar
Hajat, S., Haines, A. P., Atkinson, R. W., Bremner, S., Anderson, H. R. and Emberlin, J. (2001). Association between air pollution and daily consultations with General Practitioners for allergic rhinitis in London, United Kingdom. American Journal of Epidemiology, 153, 704–14.CrossRefGoogle ScholarPubMed
Hales, C. N. and Barker, D. J. P. (1992). Type 2 (non-insulin-dependent) diabetes mellitus: the thrifty phenotype hypothesis. Diabetologia, 35, 595–601.CrossRefGoogle ScholarPubMed
Hales, C. N. and Barker, D. J. P. (2001). The thrifty phenotype hypothesis. British Medical Bulletin, 60, 5–20.CrossRefGoogle ScholarPubMed
Hall, W. D., Clark, L. T., Wenger, N. H.et al. (2003). The metabolic syndrome in African Americans: a review. Ethnicity and Disease, 13, 414–28.Google ScholarPubMed
Hamilton, G. (2005). Filthy friends. New Scientist, 2495, 35–9.Google Scholar
Hankinson, S. E., Willett, W. C., Manson, J. E.et al. (1995). Alcohol, height, and adiposity in relation to estrogen and prolactin levels in postmenopausal women. Journal of the National Cancer Institute, 87, 1297–302.CrossRefGoogle ScholarPubMed
Hannaford, P. C., Selvaraj, S., Elliott, A. M., Angus, V., Iversen, L. and Lee, A. J. (2007). Cancer risk among users of oral contraceptives: cohort data from the Royal College of General Practioners' oral contraception study. British Medical Journal, 335, 651–4A.CrossRefGoogle Scholar
Hanson, R. L., Ehm, M. G., Pettitt, D. J.et al. (1998). An autosomal genomic scan for loci linked to type II diabetes mellitus and body-mass index in Pima Indians. American Journal of Human Genetics, 63, 1130–8.CrossRefGoogle ScholarPubMed
Harder, T., Bergmann, R., Kallischnigg, G. and Plagemann, A. (2005). Duration of breastfeeding and risk of overweight: A meta-analysis. American Journal of Epidemiology, 162, 397–403.CrossRefGoogle ScholarPubMed
Harder, T., Rodekamp, E., Schellong, K., Duderhausen, J. W., Plagemann, A. (2007). Birth weight and subsequent risk of type 2 diabetes: a meta-analysis. American Journal of Epidemiology, 165, 849–57.CrossRefGoogle ScholarPubMed
Hardy, R. and Kuh, D. (2002). Change in psychological and vasomotor symptom reporting during the menopause. Social Science and Medicine, 55, 1975–88.CrossRefGoogle ScholarPubMed
Harman, S. M., Metter, E. J., Tobin, J. D., Pearson, J. and Blackman, M. R. (2001). Longitudinal effects of aging on serum total and free testosterone levels in healthy men. Journal of Clinical Endocrinology and Metabolism, 86, 724–31.CrossRefGoogle ScholarPubMed
Harris, D. R. and Hillman, G. C., eds. (1989). Foraging and Farming: The Evolution of Plant Exploitation. London: Unwin Hyman.Google Scholar
Harris, M. I., Goldstein, D. E., Flegal, K. M.et al. (1998). Prevalence of diabetes, impaired fasting glucose, and impaired glucose tolerance in US adults. Diabetes Care, 21, 518–24.CrossRefGoogle Scholar
Harris, R., Tobias, M., Jeffreys, M., Waldegrave, K., Karlsen, S. and Nazroo, J. (2006). Effects of self-reported racial discrimination and deprivation on Māori health and inequalities in New Zealand: cross-sectional study. The Lancet, 367, 2005–9.CrossRefGoogle ScholarPubMed
Harris, S. B., Gittelsohn, J., Hanley, A. J. G.et al. (1997). The prevalence of NIDDM and associated risk factors in Native Canadians. Diabetes Care, 20, 185–7.CrossRefGoogle ScholarPubMed
Harrison, G. A. (1973). The effects of modern living. Journal of Biosocial Science, 5, 217–28.Google ScholarPubMed
Harrison, G. A., Tanner, J. A., Pilbeam, D. R. and Baker, P. T. (1988). Human Biology: An Introduction to Human Evolution, Variation, Growth, and Adaptability. Oxford: Oxford University Press.Google Scholar
Hart, R., Hickey, M. and Franks, S. (2004). Definitions, prevalence and symptoms of polycystic ovaries and polycystic ovary syndrome. Best Practice and Research in Clinical Obstetrics and Gynaecology, 18, 671–683.CrossRefGoogle ScholarPubMed
Harvie, M., Hooper, L. and Howell, A. H. (2003). Central obesity and breast cancer risk: a systematic review. Obesity Reviews, 4, 157–73.CrossRefGoogle ScholarPubMed
Hawkes, K. (2003). Grandmothers and the evolution of human longevity. American Journal of Human Biology, 15, 380–400.CrossRefGoogle ScholarPubMed
He, F. J., Nowson, C. A. and MacGregor, G. A. (2006). Fruit and vegetable consumption and stroke: meta-analysis of cohort studies. The Lancet, 367, 320–6.CrossRefGoogle ScholarPubMed
Hegele, R. A. and Bartlett, L. C. (2003). Genetics, environment and type 2 diabetes in the Oji-Cree population of northern Ontario. Canadian Journal of Diabetes, 27, 256–61.Google Scholar
Hegele, R. A., Cao, H., Hanley, A. J. G., Zinman, B., Harris, S. B. and Anderson, C. M. (2000). Clinical utility of HNF1A genotyping for diabetes in Aboriginal Canadians. Diabetes Care, 23, 775–8.CrossRefGoogle ScholarPubMed
Heim, C., Ehlert, U. and Hellhammer, D. H. (2000). The potential role of hypocortisolism in the pathophysiology of stress-related bodily disorders. Psychoneuroendocrinology, 25, 1–35.CrossRefGoogle ScholarPubMed
Heinig, M. J. (2001). Host defense benefits of breastfeeding for the infant. Pediatric Clinics of North America, 48, 105–23.CrossRefGoogle ScholarPubMed
Heinig, M. J. and Dewey, K. G. (1997). Health effects of breast feeding for mothers: a critical review. Nutrition Research Reviews, 10, 35–56.CrossRefGoogle ScholarPubMed
Helman, C. G. (1994). Culture, Health and Illness. Oxford: Butterworth-Heinemann.Google Scholar
Helmchen, L. A. and Henderson, R. M. (2004). Changes in the distribution of body mass index of white US men, 1890–2000. Annals of Human Biology, 31, 174–81.CrossRefGoogle ScholarPubMed
Hemingway, H. and Marmot, M. (1999). Psychosocial factors in the aetiology and prognosis of coronary heart disease: systematic review of prospective cohort studies. British Medical Journal, 318, 1460–7.CrossRefGoogle ScholarPubMed
Henderson, B. E., Ross, R. K., Pike, M. C. and Cassagrande, J. T. (1982). Endogenous hormones as a major factor in human cancer. Cancer Research, 42, 3232–9.Google ScholarPubMed
Hendrix, S. L., Wassertheil-Smoller, S., Johnson, K. C.et al. for the WHI Investigators (2006). Effects of conjugated equine estrogen on stroke in the Women's Health Initiative. Circulation, 113, 2425–34.CrossRefGoogle ScholarPubMed
Henry, C. J. K., Lightowler, H. J. and Al-Hourani, H. M. (2004). Physical activity and levels of inactivity in adolescent females aged 11–16 years in the United Arab Emirates. American Journal of Human Biology, 16, 346–53.CrossRefGoogle Scholar
Herbst, K. L. and Bhasin, S. (2004). Testosterone action on skeletal muscle: anabolic and catabolic signals. Current Opinion in Clinical Nutrition and Metabolic Care, 7, 271–7.CrossRefGoogle Scholar
Heuveline, P., Guillot, M. and Gwatkin, D. R. (2002). The uneven tides of the health transition. Social Science and Medicine, 55, 313–22.CrossRefGoogle ScholarPubMed
Hill, K. and Hurtado, A. M. (1991). The evolution of premature reproductive senescence and menopause in human females: an evaluation of the ‘grandmother hypothesis’. Human Nature, 2, 313–50.CrossRefGoogle Scholar
Hill, K. and Hurtado, A. M. (1996). Ache Life History. New York: Aldine de Gruyter.Google Scholar
Hill, K. and Kaplan, H. (1999). Life history traits in humans: theory and empirical studies. Annual Review of Anthropology, 28, 397–430.CrossRefGoogle Scholar
Hoffjan, S., Nicolae, D. and Ober, C. (2003). Association studies for asthma and atopic diseases: a comprehensive review of the literature. Respiratory Research, 4, 14.CrossRefGoogle ScholarPubMed
Holland, T. D. and O'Brien, M. J. (1997). Parasites, porotic hyperostosis, and the implications of changing perspectives. American Antiquity, 62, 183–93.CrossRefGoogle Scholar
Holman, R. C., Curns, A. T., Kaufman, S. F., Cheek, J. E., Pinner, R. W. and Schonberger, L. B. (2001). Trends in infectious disease hospitalizations among American Indians and Alaska Natives. American Journal of Public Health, 91, 425–31.Google ScholarPubMed
Horikawa, Y., Oda, N., Cox, N. J.et al. (2000). Genetic variation in the gene encoding calpain-10 is associated with type 2 diabetes mellitus. Nature Genetics, 26, 163–75.CrossRefGoogle ScholarPubMed
Howard, B. V., Lee, E. T., Cowan, L. D.et al. (1999). Rising tide of cardiovascular disease in American Indians. Circulation, 99, 2389–95.CrossRefGoogle ScholarPubMed
Hrdy, S. B. (2000). Mother Nature. London: Vintage.Google Scholar
Hsing, A. W. (1996). Hormones and prostate cancer: where do we go from here?Journal of the National Cancer Institute, 88, 1093–5.CrossRefGoogle Scholar
Hsing, A. W. and Devesa, S. S. (2001). Trends and patterns of prostate cancer: What do they suggest?Epidemologic Reviews, 23, 3–13.CrossRefGoogle ScholarPubMed
Hsing, A. W., Gao, Y. T., Chua, S., Deng, J. and Stanczyk, F. Z. (2003). Insulin resistance and prostate cancer risk. Journal of the National Cancer Institute, 95, 67–71.CrossRefGoogle ScholarPubMed
Hudson, J. I., Hiripi, E., Pope, H. G. and Kessler, R. C. (2007). The prevalence and correlates of eating disorders in the National Comorbidity Survey Replication. Biological Psychiatry, 61, 348–58.CrossRefGoogle ScholarPubMed
Hulley, S., Grady, D., Bush, T.et al. (1998). Randomized trial of estrogen plus progestin for secondary prevention of coronary heart disease in postmenopausal women. Journal of the American Medical Association, 280, 605–13.CrossRefGoogle ScholarPubMed
Hurtado, A. M., Arenas, I. and Hill, K. (1996). Evolutionary contexts of chronic allergic disease: the Hiwi of Venezuala. Human Nature, 8, 1–20.Google Scholar
Hussain, A., Rahim, M. A., Khan, Azad A. K., Ali, S. M. K. and Vaaler, S. (2005). Type 2 diabetes in rural and urban populations: diverse prevalence and associated risk factors in Bangladesh. Diabetic Medicine, 22, 931–6.CrossRefGoogle Scholar
Huxley, R., Neil, A. and Collins, R. (2002). Unravelling the fetal origins hypothesis: is there really an inverse association between birthweight and subsequent blood pressure?The Lancet, 360, 659–65.CrossRefGoogle ScholarPubMed
Huxley, R., Owens, J. F., Whincup, P. H., Cook, D. G., Colman, S. and Collins, R. (2004). Birth weight and subsequent cholesterol levels: exploration of the “fetal origins” hypothesis. Journal of the American Medical Association, 292, 2755–64.CrossRefGoogle ScholarPubMed
Huyghe, E., Matsuda, T. and Thonneau, P. (2003). Increasing incidence of testicular cancer worldwide: a review. Journal of Urology, 170, 5–11.CrossRefGoogle ScholarPubMed
Iglowstein, I., Jenni, O. G., Molinari, L. and Largo, R. H. (2003). Sleep duration from infancy to adolescence: reference values and generational trends. Pediatrics, 111, 302–7.CrossRefGoogle ScholarPubMed
Irons, W. (1998). Adaptively relevant environments versus the environment of evolutionary adaptedness. Evolutionary Anthropology, 6, 198–204.3.0.CO;2-B>CrossRefGoogle Scholar
ISAAC Steering Committee (1998). Worldwide variation in the prevalence of symptoms of asthma, allergic rhinoconjunctivitis, and atopic eczema: ISAAC. The Lancet, 351, 1225–32.CrossRef
Jackson, M. A., Kovi, J., Heshmat, M. Y.et al. (1980). Characterization of prostatic carcinoma among blacks: A comparison between a low-incidence area, Ibadan, Nigeria, and a high-incidence area, Washington, DC. Prostate, 1, 185–205.CrossRefGoogle Scholar
James, G. D., Jenner, D. A., Harrison, G. A. and Baker, P. T. (1985). Differences in catecholamine excretion rates, blood pressure and lifestyle among young Western Samoan men. Human Biology, 57, 635–47.Google ScholarPubMed
James, G. D., Crews, D. E. and Pearson, J. (1989). Catecholamines and stress. In Human Population Biology, ed. Little, M. A. and Haas, J. D.. Oxford: Oxford University Press.Google Scholar
James, G. D., Broege, P. A. and Schlussel, Y. R. (1996). Assessing cardiovascular risk and stress-related blood pressure variability in young women employed in waged jobs. American Journal of Human Biology, 8, 743–9.3.0.CO;2-U>CrossRefGoogle Scholar
James, W. P. T. (2002). Will feeding mothers prevent the Asian metabolic syndrome epidemic?Asia Pacific Journal of Clinical Nutrition, 11 (Suppl.), S516–23.CrossRefGoogle ScholarPubMed
James, W. P. T. (2005). Assessing obesity: are ethnic differences in body mass index and waist classification criteria justified?Obesity Reviews, 6, 179–81.CrossRefGoogle ScholarPubMed
Janes, C. R. (1990). Migration, changing gender roles and stress: the Samoan case. Medical Anthropology, 12, 145–67.CrossRefGoogle ScholarPubMed
Jasieńska, G. and Ellison, P. T. (1998). Physical work causes suppression of ovarian function in women. Proceedings of the Royal Society of London B, 265, 1847–51.CrossRefGoogle ScholarPubMed
Jasieńska, G. and Ellison, P. T. (2004). Energetic factors and seasonal changes in ovarian function in women from rural Poland. American Journal of Human Biology, 16, 563–80.CrossRefGoogle ScholarPubMed
Jasieńska, G. and Thune, I. (2001a). Lifestyle, hormones, and risk of breast cancer. British Medical Journal, 322, 586–7.CrossRefGoogle Scholar
Jasieńska, G. and Thune, I. (2001b). Lifestyle, progesterone, and risk of breast cancer – reply. British Medical Journal, 323, 1002.Google Scholar
Jasieńska, G., Thune, I. and Ellison, P. T. (2000). Energetic factors, ovarian steroids and the risk of breast cancer. European Journal of Cancer Prevention, 9, 231–9.CrossRefGoogle ScholarPubMed
Jasieńska, G., Thune, I. and Ellison, P. T. (2006a). Fatness at birth predicts adult susceptibility to ovarian suppression: an empirical test of the Predictive Adaptive Response hypothesis. Proceedings of the National Academy of Science, 103, 12759–62.CrossRefGoogle Scholar
Jasieńska, G., Ziomkiewicz, A., Thune, I., Lipson, S. F. and Ellison, P. T. (2006b). Habitual physical activity and estradiol levels in women of reproductive age. European Journal of Cancer Prevention, 15, 439–45.CrossRefGoogle Scholar
Jenike, M. R. (2001). Nutritional ecology: diet, physical activity and body size. In Hunter–Gatherers: An Interdisciplinary Perspective, ed. Panter-Brick, C., Layton, R. H. and Rowley-Conwy, P.. Cambridge: Cambridge University Press, pp. 205–39.Google Scholar
Jenner, D. A., Reynolds, V. and Harrison, G. A. (1980). Catecholamine excretion rates and occupation. Ergonomics, 23, 237–46.CrossRefGoogle ScholarPubMed
Jenner, D. A., Harrison, G. A. and Prior, I. A. M. (1987). Catecholamine excretion in Tokelauans living in three different environments. Human Biology, 59, 165–72.Google ScholarPubMed
Jin, B., Turner, L., Zhou, Z., Zhou, E. L. and Handelsman, D. J. (1999). Ethnicity and migration as determinants of human prostate size. Journal of Clinical Endocrinology and Metabolism, 84, 3613–19.Google ScholarPubMed
Jonas, B. S., Franks, P. and Ingram, D. D. (1997). Are symptoms of anxiety and depression risk factors for hypertension? Longitudinal evidence from the National Health and Nutrition Examination Survey I: epidemiological follow-up study. Archives of Family Medicine, 6, 43–9.CrossRefGoogle Scholar
Julian, D. G. and Cowan, J. C. (1992). Cardiology. London: Baillière Tindall.Google Scholar
Kaaks, R., Lukanova, A. and Kurzer, M. S. (2002). Obesity, endogenous hormones, and endometrial cancer risk: a synthetic review. Cancer Epidemiology, Biomarkers and Prevention, 11, 1531–43.Google ScholarPubMed
Kajantie, E. and Phillips, D. I. W. (2006). The effects of sex and hormonal status on the physiological response to acute psychosocial stress. Psychoneuroendocrinology, 31, 151–78.CrossRefGoogle ScholarPubMed
Kaliora, A. C., Dedoussis, G. V. Z. and Schmidt, H. (2006). Dietary antioxidants in preventing atherogenesis. Atherosclerosis, 187, 1–17.CrossRefGoogle ScholarPubMed
Kalliomäki, M., Salminen, S., Arvilommi, H., Kero, P., Koskinen, P. and Isolauri, E. (2001). Probiotics in primary prevention of atopic disease: a randomised placebo-controlled trial. The Lancet, 357, 1076–9.CrossRefGoogle ScholarPubMed
Kannel, W. B. (1987). New perspectives on cardiovascular risk factors. American Heart Journal, 114, 213–19.CrossRefGoogle ScholarPubMed
Kant, A. K. and Graubard, B. I. (2004). Eating out in America, 1987–2000: trends and nutritional correlates. Preventive Medicine, 38, 243–9.CrossRefGoogle ScholarPubMed
Karasek, R. A. (1979). Job demands, job decision latitude, and mental strain: implications for job redesign. Administrative Science Quarterly, 24, 285–308.CrossRefGoogle Scholar
Kaufman, J. S. and Hall, S. A. (2003). The slavery hypertension hypothesis: dissemination and appeal of a modern race theory. Epidemiology, 14, 111–18.CrossRefGoogle ScholarPubMed
Kaur, H., Choi, W. S., Mayo, M. S. and Harris, K. J. (2003). Duration of television watching is associated with increased body mass index. Journal of Pediatrics, 143, 506–11.CrossRefGoogle ScholarPubMed
Keighley, E. D., McGarvey, S. T., Quested, C., McCuddin, C., Viali, S. and Maga, U. A. (2007). Nutrition and health in modernizing Samoans: temporal trends and adaptive perspectives. In Health Changes in the Asia-Pacific Region: Biocultural and Epidemiological Approaches, ed. Ohtsuka, R. and Ulijaszek, S. J.. Cambridge: Cambridge University Press, pp. 147–91.CrossRefGoogle Scholar
Keil, J. E., Sutherland, S. E., Knapp, R. G. and Tyroler, H. A. (1992). Does equal socioeconomic status in black and white men mean equal risk of mortality?American Journal of Public Health, 82, 1133–6.CrossRefGoogle ScholarPubMed
Kemp, A. and Kakakios, A. (2004). Asthma prevention: breast is best?Journal of Paediatrics and Child Health, 40, 337–9.CrossRefGoogle ScholarPubMed
Kennedy, G. E. (2005). From the ape's dilemma to the weanling's dilemma: early weaning and its evolutionary context. Journal of Human Evolution, 48, 123–45.CrossRefGoogle ScholarPubMed
Kent, S. (1986). The influence of sedentism and aggregation on porotic hyperostosis and anemia: a case study. Man, 21, 605–36.CrossRefGoogle Scholar
Kero, J., Gissler, M., Hemminki, E. and Isolauri, E. (2001). Could T(h)1 and T(h)2 diseases coexist? Evaluation of asthma incidence in children with coeliac disease, type 1 diabetes, or rheumatoid arthritis: a register study. Journal of Allergy and Clinical Immunology, 108, 781–3.CrossRefGoogle ScholarPubMed
Key, T. J. A., Chen, J., Wang, D. Y., Pike, M. C. and Boreham, J. (1990). Sex hormones in women in rural China and in Britain. British Journal of Cancer, 62, 631–6.CrossRefGoogle ScholarPubMed
Kim, S. Y., Moon, S. and Popkin, B. M. (2000). The nutrition transition in South Korea. American Journal of Clinical Nutrition, 71, 44–53.CrossRefGoogle ScholarPubMed
King, H., Aubert, R. E. and Herman, W. H. (1998). Global burden of diabetes, 1995–2025. Diabetes Care, 21, 1414–31.CrossRefGoogle Scholar
Kleinman, A. and Good, B. (1985). Introduction: culture and depression. In Culture and Depression: Studies in the Anthropology and Cross-Cultural Psychiatry of Affect and Disorder, ed. Kleinman, A. and Good, B.. Berkeley: University of California Press, pp. 1–33.Google Scholar
Kleinman, A. M. (1977). Depression, somatization and the “new cross-cultural” psychiatry. Social Science and Medicine, 11, 3–10.CrossRefGoogle ScholarPubMed
Klerman, G. L. and Weissman, M. M. (1989). Increasing rates of depression. Journal of the American Medical Association, 261, 2229–35.CrossRefGoogle Scholar
Kliks, M. M. (1983). Paleoparasitology: on the origins and impact of human-helminth relationships. In Human Ecology and Infectious Disease, ed. Croll, N. A. and Cross, J. H.. New York: Academic Press.Google Scholar
Knowler, W. C. (1978). Diabetes incidence and prevalence in Pima Indians: a 19-fold greater incidence than in Rochester, Minnesota. American Journal of Epidemiology, 108, 497–505.CrossRefGoogle ScholarPubMed
Knowler, W. C., Pettit, D. J., Savage, P. J., et al. (1983). Diabetes mellitus in the Pima Indians: genetic and evolutionary considerations. American Journal of Physical Anthropology, 62, 107–14.CrossRefGoogle ScholarPubMed
Kohen-Avramoglu, R., Theriault, A. and Adeli, K. (2003). Emergence of the metabolic syndrome in childhood: an epidemiological overview and mechanistic link to dyslipidemia. Clinical Biochemistry, 36, 413–20.CrossRefGoogle ScholarPubMed
Kohler, H. P., Billari, F. C. and Ortega, J. A. (2002). The emergence of lowest-low fertility in Europe during the 1990s. Population and Development Review, 28, 641–86.CrossRefGoogle Scholar
Konner, M. and Worthman, C. M. (1980). Nursing frequency, gonadal function, and birth spacing among the!Kung hunter–gatherers. Science, 207, 788–91.CrossRefGoogle Scholar
Kraft, P., Pharoah, P., Chanock, S. J.et al. (2005). Genetic variation in the HSD17B1 gene and risk of prostate cancer. PLoS Genetics, 1, e68.CrossRefGoogle ScholarPubMed
Kramer, M. S. (1987). Determinants of low birth weight: methodological assessment and meta-analysis. Bulletin of the World Health Organization, 65, 663–737.Google ScholarPubMed
Kramer, M. S. and Kakuma, R. (2002). The Optimal Duration of Exclusive Breastfeeding: A Systematic Review. Geneva: World Health Organization.CrossRefGoogle Scholar
Krämer, U., Heinrich, J., Wjst, M. and Wichmann, H. -E. (1999). Age of entry to day nursery and allergy in later childhood. The Lancet, 352, 450–4.CrossRefGoogle Scholar
Krantz, D. S., Contrada, R. J., Hill, R. and Friedler, E. (1988). Environmental stress and biobehavioral antecedents of coronary heart disease. Journal of Consulting and Clinical Psychology, 56, 333–41.CrossRefGoogle ScholarPubMed
Krause, I. -B. (1989). Sinking heart: a Punjabi communication of distress. Social Science and Medicine, 29, 563–75.CrossRefGoogle ScholarPubMed
Kubzansky, L. D. and Kawachi, I. (2000). Going to the heart of the matter: do negative emotions cause coronary heart disease?Journal of Psychosomatic Research, 48, 323–37.CrossRefGoogle ScholarPubMed
Kuh, D. and Hardy, R., eds. (2002). A Life Course Approach to Women's Health. Oxford: Oxford University Press.CrossRefGoogle Scholar
Kuh, D., Ben-Shlomo, Y., Lynch, J., Hallqvist, J. and Power, C. (2003). Life course epidemiology. Journal of Epidemiology and Community Health, 57, 778–83.CrossRefGoogle ScholarPubMed
Kukkonen, K., Savilahti, E., Haahtela, T.et al. (2007). Probiotics and prebiotic galacto-oligosaccharides in the prevention of allergic diseases: a randomized, double-blind, placebo-controlled trial. Journal of Allergy and Clinical Immunology, 119, 192–8.CrossRefGoogle ScholarPubMed
Kuller, L. H., Meilahn, E. N., Cauley, J. A., Gutai, J. P. and Matthews, K. A. (1994). Epidemiologic studies of menopause: changes in risk factors and disease. Experimental Gerontology, 29, 495–509.CrossRefGoogle ScholarPubMed
Künzli, N., McConnel, R., Bates, D.et al. (2003). Breathless in Los Angeles: the exhausting search for clean air. American Journal of Public Health, 93, 1494–99.CrossRefGoogle ScholarPubMed
Kuzawa, C. W. (1998). Adipose tissue in human infancy and childhood: an evolutionary perspective. Yearbook of Physical Anthropology, 41, 177–210.3.0.CO;2-B>CrossRefGoogle Scholar
Kuzawa, C. W. (2005). Fetal origins of developmental plasticity: are fetal cues reliable predictors of future nutritional environments?American Journal of Human Biology, 17, 5–21.CrossRefGoogle ScholarPubMed
Lager, C. and Ellison, P. T. (1990). Effect of moderate weight loss on ovarian function assessed by salivary progesterone measurements. American Journal of Human Biology, 2, 303–12.CrossRefGoogle ScholarPubMed
Lake, A. and Townshend, T. (2006). Obesogenic environments: exploring the built and food environments. Journal of the Royal Society for the Promotion of Health, 126, 262–7.CrossRefGoogle ScholarPubMed
LaMonte, M. J., Barlow, C. E., Jurca, R., Kampert, J. B., Church, T. S. and Blair, S. N. (2005). Cardiorespiratory fitness is inversely associated with the incidence of metabolic syndrome. A prospective study of men and women. Circulation, 112, 505–12.CrossRefGoogle ScholarPubMed
Lancet Editorial (2006). A plea to abandon asthma as a disease concept. The Lancet, 368, 705.CrossRef
Lane, D., Beevers, D. G. and Lip, G. Y. H. (2002). Ethnic differences in blood pressure and the prevalence of hypertension in England. Journal of Human Hypertension, 16, 267–73.CrossRefGoogle ScholarPubMed
Larsen, C. S. (1995). Biological changes in human populations with agriculture. Annual Review of Anthropology, 24, 185–213.CrossRefGoogle Scholar
Larsen, C. S. (2002). Post-Pleistocene human evolution: bioarcheology of the agricultural transition. In Human Diet: Its Origin and Evolution, ed. Ungar, P. S. and Teaford, M. F.. Westport, Connecticut: Bergin and Garvey, pp. 19–36.Google Scholar
Larsen, C. S. (2003). Animal source foods and human health during evolution. Journal of Nutrition, 133, 3893S–7S.CrossRefGoogle ScholarPubMed
Lasker, G. (1969). Human biological adaptability: the ecological approach in physical anthropology. Science, 166, 1480–6.CrossRefGoogle ScholarPubMed
Vecchia, C. and Bosetti, C. (2004). Benefits and risks of oral contraceptives on cancer. European Journal of Cancer Prevention, 13, 467–70.CrossRefGoogle ScholarPubMed
Lawlor, D. A., Smith, Davey G., Leon, D. A., Sterne, J. A. C. and Ebrahim, S. (2002). Secular trends in mortality by stroke subtype in the 20th century: a retrospective analysis. The Lancet, 360, 1818–23.CrossRefGoogle ScholarPubMed
Souëf, P. N., Goldblatt, J. and Lynch, N. R. (2000). Evolutionary adaptation of inflammatory immune responses in human beings. The Lancet, 356, 242–4.CrossRefGoogle ScholarPubMed
Souëf, P. N., Candelaria, P. and Goldblatt, J. (2006). Evolution and respiratory genetics. European Respiratory Journal, 28, 1258–63.CrossRefGoogle ScholarPubMed
Lee, R. B. (1979). The!Kung San: Men, Women and Work in a Foraging Society. Cambridge: Cambridge University Press.Google Scholar
Leidy, L. E. (1999). Menopause in evolutionary perspective. In Evolutionary Medicine, ed. Trevathan, W. R., Smith, E. O. and McKenna, J. J.. New York: Oxford University Press, pp. 407–27.Google Scholar
Leidy Sievert, L. (2001). Aging and reproductive senescence. In Reproductive Ecology and Human Evolution, ed. Ellison, P. T.. New York: Aldine de Gruyter, pp. 267–92.Google Scholar
Sievert, Leidy L. (2006). Menopause: A Biocultural Perspective. New Brunswick, New Jersey: Rutgers University Press.Google Scholar
Leon, D. A., Chenet, L., Shkolnikov, V. M.et al. (1997). Huge variation in Russian mortality rates 1984–94: artefact, alcohol, or what?The Lancet, 350, 383–8.CrossRefGoogle ScholarPubMed
Leonard, W. R. (2000). Human nutritional evolution. In Human Biology: An Evolutionary and Biocultural Perspective, ed. Stinson, S., Bogin, B., Huss-Ashmore, R. and O'Rourke, D.. New York: Wiley-Liss, pp. 295–343.Google Scholar
Lett, H. S., Blumenthal, J. A., Babyak, M. A.et al. (2004). Depression as a risk factor for coronary artery disease: evidence, mechanisms and treatment. Psychosomatic Medicine, 66, 305–15.Google ScholarPubMed
Lewontin, R. C. (1982). Human Diversity. New York: Scientific American Library.Google Scholar
Li, C., Ford, E. S., Mokdad, A. H. and Cook, S. (2006). Recent trends in waist circumference and waist–height ratio among US children and adolescents. Pediatrics, 118, e1390–8.CrossRefGoogle ScholarPubMed
Lieberman, L. S. (2003). Dietary, evolutionary, and modernizing influences on the prevalence of type 2 diabetes. Annual Review of Nutrition, 23, 345–77.CrossRefGoogle ScholarPubMed
Linden, W., Stossel, C. and Maurice, J. (1996). Psychosocial interventions for patients with coronary artery disease: a meta-analysis. Archives of Internal Medicine, 156, 745–52.CrossRefGoogle ScholarPubMed
Lipson, S. F. (2001). Metabolism, maturation, and ovarian function. In Reproductive Ecology and Human Evolution, ed. Ellison, P. T.. New York: Aldine de Gruyter, pp. 235–48.Google Scholar
Lipson, S. F. and Ellison, P. T. (1996). Comparison of salivary steroid profiles in naturally occurring conception and non-conception cycles. Human Reproduction, 11, 2090–6.CrossRefGoogle ScholarPubMed
Liu, S., Willett, W. C., Manson, J. E., Hu, F. B., Rosner, B. and Colditz, G. (2003). Relation between changes in intakes of dietary fiber and grain products and changes in weight and development of obesity among middle-aged women. American Journal of Clinical Nutrition, 78, 920–7.CrossRefGoogle ScholarPubMed
Livingston, C. and Collison, M. (2002). Sex steroids and insulin resistance. Clinical Science, 102, 151–66.CrossRefGoogle Scholar
Lock, M. and Kaufert, P. (2001). Menopause, local biologies, and cultures of aging. American Journal of Human Biology, 13, 494–504.CrossRefGoogle Scholar
Ludwig, D. S. (2002). The glycemic index: physiological mechanisms relating to obesity, diabetes, and cardiovascular disease. Journal of the American Medical Association, 287, 2414–23.CrossRefGoogle Scholar
Ludwig, D. S., Peterson, K. E. and Gortmaker, S. L. (2001). Relation between consumption of sugar-sweetened drinks and childhood obesity: a prospective, observational analysis. The Lancet, 357, 505–8.CrossRefGoogle ScholarPubMed
Macleod, J. and Smith, Davey G. (2003). Psychosocial factors and public health: a suitable case for treatment. Journal of Epidemiology and Community Health, 57, 565–70.CrossRefGoogle Scholar
Maizels, R. M. (2005). Infections and allergy – helminths, hygiene and host immune regulation. Current Opinion in Immunology, 17, 656–61.CrossRefGoogle ScholarPubMed
Mann, J. (2004). Free sugars and human health: sufficient evidence for action?The Lancet, 363, 1068–70.CrossRefGoogle ScholarPubMed
Mann, J. I. (2002). Diet and risk of coronary heart disease and type 2 diabetes. The Lancet, 360, 783–9.CrossRefGoogle ScholarPubMed
Marmot, M., Shipley, M. and Rose, G. (1984). Inequalities in death – specific explanations of a general pattern?The Lancet, i, 1003–6.CrossRefGoogle Scholar
Marshall, C., Hitman, G. A., Partridge, C. J.et al. (2005). Evidence that an isoform of Calpain-10 is a regulator of exocytosis in pancreatic beta-cells. Molecular Endocrinology, 19, 213–24.CrossRefGoogle ScholarPubMed
Martin, R. M., Gunnell, D. and Smith, Davey G. (2005). Breastfeeding in infancy and blood pressure in later life: systematic review and meta-analysis. American Journal of Epidemiology, 161, 15–26.CrossRefGoogle ScholarPubMed
Martinez, F. D. (1994). Role of viral infections in the inception of asthma: could they be protective?Thorax, 49, 1189–91.CrossRefGoogle ScholarPubMed
Martorell, R., Khan, L. K., Hughes, M. L. and Grummer-Strawn, L. M. (1998). Obesity in Latin American women and children. Journal of Nutrition, 128, 1464–73.CrossRefGoogle ScholarPubMed
Mascie-Taylor, C. G. N. (1993). The biological anthropology of disease. In The Anthropology of Disease, ed. Mascie-Taylor, C. G. N.. Oxford: Oxford University Press, pp. 1–72.Google Scholar
Maskarinec, G., Franke, A., Williams, A.et al. (2004). Effects of a 2-year randomized soy intervention on sex hormone levels in premenopausal women. Cancer Epidemiology, Biomarkers and Prevention, 13, 1736–44.Google ScholarPubMed
Masoli, M., Fabian, D., Holt, S. and Beasley, R. for the Global Initiative for Asthma (GINA) Program (2004a). The global burden of asthma: executive summary of the GINA Dissemination Committee Report. Allergy, 59, 469–78.CrossRefGoogle Scholar
Masoli, M., Fabian, D., Holt, S. H. A. and Beasley, R. (2004b). Global Burden of Asthma: Global Initiative for Asthma.
Mathers, C. D. and Loncar, D. (2006). Projections of global mortality and burden of disease from 2002 to 2030. PLoS Medicine, 3, 2011–30.CrossRefGoogle ScholarPubMed
Matricardi, P. M., Bjorksten, B., Bonini, S.et al. for the EAACI Task Force 7 (2003). Microbial products in allergy prevention and therapy. Allergy, 58, 461–71.CrossRefGoogle ScholarPubMed
Matsuda, K., Nishi, Y., Okamatsu, Y., Kojima, M. and Matsuishi, T. (2006). Ghrelin and leptin: a link between obesity and allergy?Journal of Allergy and Clinical Immunology, 117, 705–6.CrossRefGoogle ScholarPubMed
Matthews, K. A. (2005). Psychological perspectives on the development of coronary heart disease. American Psychologist, 60, 783–96.CrossRefGoogle ScholarPubMed
Matthews, K. A., Raikkonen, K., Everson, S. A.et al. (2000). Do the daily experiences of healthy men and women vary according to occupational prestige and work strain?Psychosomatic Medicine, 62, 346–53.CrossRefGoogle ScholarPubMed
Mayor, S. (2005). 23% of babies in England are delivered by caesarean section. British Medical Journal, 330, 806.Google ScholarPubMed
McDaniel, C. N. and Gowdy, J. M. (2000). Paradise for Sale: A Parable of Nature. Berkeley, California: University of California Press.Google Scholar
McElroy, A. and Townsend, P. K. (2004). Medical Anthropology in Ecological Perspective. Boulder, Colorado: Westview Press.Google Scholar
McEwen, B. S. and Stellar, E. (1993). Stress and the individual – mechanisms leading to disease. Archives of Internal Medicine, 153, 2093–110.CrossRefGoogle ScholarPubMed
McEwen, B. S. and Wingfield, J. C. (2003). The concept of allostatis in biology and biomedicine. Hormones and Behavior, 43, 2–15.CrossRefGoogle Scholar
McGarvey, S. T., Bindon, J. R., Crews, D. E. and Schendel, D. E. (1989). Modernization and adiposity: causes and consequences. In Human Population Biology, ed. Little, M. A. and Haas, J. D.. Oxford: Oxford University Press, pp. 263–80.Google Scholar
McGee, R., Williams, S. and Elwood, M. (1994). Depression and the development of cancer – a meta-analysis. Social Science and Medicine, 1994, 187–92.CrossRefGoogle Scholar
McKeigue, P. M. (1996). Metabolic consequences of obesity and body fat pattern: lessons from migrant studies. In The Origins and Consequences of Obesity. CIBA Foundation Symposium 201, ed. Chadwick, D. J. and Cardew, G.. Chichester: John Wiley, pp. 54–63.CrossRefGoogle Scholar
McKeigue, P. M., Shah, B. and Marmot, M. G. (1991). Relation of central obesity and insulin resistance with high diabetes prevalence and cardiovascular risk in South Asians. The Lancet, 337, 382–6.CrossRefGoogle ScholarPubMed
McKenna, J. J., Mosko, S. and Richard, C. (1999). Breast-feeding and mother–infant cosleeping in relation to SIDS prevention. In Evolutionary Medicine, ed. Trevathan, W. R., Smith, E. O. and McKenna, J. J.. New York: Oxford University Press, pp. 53–74.Google Scholar
McMichael, A. J. (2001). Human Frontiers, Environments and Disease. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
McPherson, K., Steel, C. M. and Dixon, J. M. (2000). Breast cancer – epidemiology, risk factors, and genetics. British Medical Journal, 321, 624–8.CrossRefGoogle ScholarPubMed
Meek, J. Y. (2001). Breastfeeding in the workplace. Pediatric Clinics of North America, 48, 461–74.CrossRefGoogle ScholarPubMed
Menacker, F. (2005). Trends in Cesarean Rates for First Births and Repeat Cesarean Rates for Low Risk Women: United States, 1990–2000. National Vital Statistics Report 54: 1–12. Hyattsville, Maryland: National Center for Health Statistics.
Mendelsohn, M. E. and Karas, R. H. (1999). The protective effects of estrogen on the cardiovascular system. New England Journal of Medicine, 340, 1801–11.CrossRefGoogle ScholarPubMed
Meyer, V. F. (2001). The medicalization of the menopause: Critique and consequences. International Journal of Health Services, 31, 769–92.CrossRefGoogle ScholarPubMed
Midthjell, K., Kruger, O., Holmen, J.et al. (1999). Rapid changes in the prevalence of obesity and known diabetes in an adult Norwegian population – the Nord-Trondelag Health Surveys: 1984–1986 and 1995–1997. Diabetes Care, 22, 1813–20.CrossRefGoogle Scholar
Miller, W. C., Koceja, D. M. and Hamilton, E. J. (1997). A meta-analysis of the past 25 years of weight loss research using diet, exercise or diet plus exercise intervention. International Journal of Obesity and Related Metabolic Disorders, 21, 941–47.CrossRefGoogle ScholarPubMed
Misra, A. and Vikram, N. K. (2004). Insulin resistance syndrome (metabolic syndrome) and obesity in Asian Indians: evidence and implications. Nutrition, 20, 482–91.CrossRefGoogle ScholarPubMed
Mokdad, A. H., Serdula, M. K., Dietz, W. H., Bowman, B. A., Marks, J. S. and Koplan, J. P. (1999). The spread of the obesity epidemic in the United States, 1991–1998. Journal of the American Medical Association, 282, 1519–22.CrossRefGoogle ScholarPubMed
Moore, S. E., Halsall, I., Howarth, D., Poskitt, E. M. E. and Prentice, A. M. (2001). Glucose, insulin and lipid metabolism in rural Gambians exposed to early malnutrition. Diabetic Medicine, 18, 645–53.CrossRefGoogle ScholarPubMed
Muehlenbein, M. P. and Bribiescas, R. G. (2005). Testosterone-mediated immune functions and male life histories. American Journal of Human Biology, 17, 527–58.CrossRefGoogle ScholarPubMed
Mufunda, J., Chatora, R., Ndambukuwa, Y.et al. (2006). Emerging non-communicable disease epidemic in Africa: preventive measures from the WHO Regional Office for Africa. Ethnicity and Disease, 16, 521–6.Google ScholarPubMed
Murch, S. H. (2001). Toll of allergy reduced by probiotics. The Lancet, 357, 1057–9.CrossRefGoogle ScholarPubMed
Murray, C. J. L. and Lopez, A. D. (1997). Alternative projections of mortality and disability by cause 1990–2020: global burden of disease study. The Lancet, 349, 1498–504.CrossRefGoogle ScholarPubMed
Murray, C. J. L., Lauer, J. A., Hutubessy, R. C. W.et al. (2003). Effectiveness and costs of interventions to lower systolic blood pressure and cholesterol: a global and regional analysis on reduction of cardiovascular-disease risk. The Lancet, 361, 717–25.CrossRefGoogle ScholarPubMed
Muscat, J. E., Harris, R. E., Haley, N. J. and Wynder, E. L. (1991). Cigarette smoking and plasma cholesterol. American Heart Journal, 121, 141–7.CrossRefGoogle ScholarPubMed
Mussa, F. F., Chai, H., Wang, X., Yao, Q., Lumsden, A. B. and Chen, C. (2006). Chlamydia pneumoniae and vascular disease: an update. Journal of Vascular Surgery, 43, 1301–7.CrossRefGoogle ScholarPubMed
Must, A. and Colclough-Douglas, S. (2001). Adult health sequelae of pediatric obesity. In Obesity, Growth and Development, ed. Johnston, F. E. and Foster, G. D.. London: Smith-Gordon, pp. 185–98.Google Scholar
Must, A., Spadano, J., Coakley, E. H., Field, A. E., Colditz, G. and Dietz, W. H. (1999). The disease burden associated with overweight and obesity. Journal of the American Medical Association, 282, 1523–9.CrossRefGoogle ScholarPubMed
Muti, P. (2004). The role of endogenous hormones in the etiology and prevention of breast cancer: the epidemiological evidence. Annals of the New York Academy of Sciences, 1028, 273–182.CrossRefGoogle ScholarPubMed
National Sleep Foundation (2002). “Sleep in America” Poll. Washington, DC: National Sleep Foundation.
Nazroo, J. (2003). The structuring of ethnic inequalities in health: economic position, racial discrimination, and racism. American Journal of Public Health, 93, 277–84.CrossRefGoogle ScholarPubMed
Neel, J. V. (1962). Diabetes mellitus: a “thrifty” genotype rendered detrimental by “progress”?American Journal of Human Genetics, 14, 353–62.Google ScholarPubMed
Neel, J. V. (1982). The thrifty genotype revisited. In The Genetics of Diabetes Mellitus, ed. , J. Köbberling and , R. Tattersall. Amsterdam: Academic Press, pp. 137–47.Google Scholar
Neel, J. V., Weder, A. B. and Julius, S. (1998). Type II diabetes, essential hypertension, and obesity as ‘syndromes of impaired genetic homeostasis’: the ‘thrifty genotype’ hypothesis enters the 21st century. Perspectives in Biology and Medicine, 42, 44–74.CrossRefGoogle Scholar
Ness, R. B., Haggerty, C. L., Harger, G. and Ferrell, R. E. (2004). Differential distribution of allelic variants in cytokine genes among African Americans and white Americans. American Journal of Epidemiology, 160, 1033–8.CrossRefGoogle ScholarPubMed
Ness, R. B. (2000). Is depression an adaptation?Archives of General Psychiatry, 57, 14–20.CrossRefGoogle Scholar
Nesse, R. M. (2004). Natural selection and the elusiveness of happiness. Philosophical Transactions of the Royal Society of London B, 359, 1333–47.CrossRefGoogle ScholarPubMed
Nesse, R. M. and Williams, G. C. (1994). Evolution and Healing: The New Science of Darwinian Medicine. London: Phoenix.Google Scholar
Nesse, R. M., Stearns, S. C. and Omenn, G. S. (2006). Medicine needs evolution. Science, 311, 1071.CrossRefGoogle ScholarPubMed
Nettle, D. (2004). Evolutionary origins of depression: a review and reformulation. Journal of Affective Disorders, 81, 91–102.CrossRefGoogle ScholarPubMed
Netuveli, G., Hurwitz, B., Levy, M.et al. (2005). Ethnic variations in UK asthma frequency, morbidity, and health-service use: a systematic review and meta-analysis. The Lancet, 365, 312–17.CrossRefGoogle ScholarPubMed
Neuhausen, S. L. (1999). Ethnic differences in cancer risk resulting from genetic variation. Cancer, 86, 2575–82.3.0.CO;2-F>CrossRefGoogle ScholarPubMed
Newsome, C. A., Shiell, A. W., Fall, C. H. D., Phillips, D. I. W., Shier, R. and Law, C. M. (2003). Is birth weight related to later glucose and insulin metabolism? A systematic review. Diabetic Medicine, 20, 339–48.CrossRefGoogle ScholarPubMed
Nielsen, S. J. and Popkin, B. M. (2003). Patterns and trends in food portion sizes, 1977–1998. Journal of the American Medical Association, 289, 450–3.CrossRefGoogle ScholarPubMed
NIH State-of-the–Science Panel (2005). National Institutes of Health State-of-the–Science conference statement: management of menopause-related symptoms. Annals of Internal Medicine, 142, 1003–13.CrossRef
Núñez-de la Mora, A., Chatterton, R. T., Choudhury, O. A., Napolitano, D. A. and Bentley, G. R. (2007). Childhood conditions influence adult progesterone levels. PLoS Medicine, doi: 10.1371/journal. pmed. 0040167.CrossRef
O'Dea, K. (1984). Marked improvement in carbohydrate and lipid metabolism in diabetic Australian Aborigines after temporary reversion to traditional lifestyle. Diabetes, 33, 596–603.CrossRefGoogle ScholarPubMed
O'Dea, K. (1991). Traditional diet and food preferences of Australian Aboriginal hunter–gatherers. Philosophical Transactions of the Royal Society of London B, 334, 233–41.CrossRefGoogle ScholarPubMed
O'Dea, K. and Piers, L. (2002). Diabetes. In The Nutrition Transition: Diet and Disease in the Developing World, ed. Caballero, B. and Popkin, B. M.. Amsterdam: Elsevier, pp. 165–90.Google Scholar
O'Dea, K., Hopper, J., Patel, M., Traianedes, K. and Kubisch, D. (1993). Obesity, diabetes, and hyperlipidemia in a Central Australian Aboriginal community with a long history of acculturation. Diabetes Care, 16, 1004–10.CrossRefGoogle Scholar
O'Sullivan, A. J., Martin, A. and Brown, M. A. (2001). Efficient fat storage in premenopausal women and in early pregnancy: a role for estrogen. Journal of Clinical Endocrinology and Metabolism, 86, 4951–6.CrossRefGoogle ScholarPubMed
Oken, E. and Gillman, M. W. (2003). Fetal origins of obesity. Obesity Research, 11, 496–506.CrossRefGoogle ScholarPubMed
Omran, A. R. (1971). The epidemiologic transition: a theory of the epidemiology of population change. Milbank Memorial Fund Quarterly, 49, 509–38.CrossRefGoogle ScholarPubMed
Ong, K. K. and Dunger, D. B. (2004). Birth weight, infant growth and insulin resistance. European Journal of Endocrinology, 151, U131–U139.CrossRefGoogle ScholarPubMed
Ong, K. L. and Dunger, D. B. (2000). Thrifty genotypes and phenotypes in the pathogenesis of type 2 diabetes mellitus. Journal of Pediatric Endocrinology and Metabolism, 13, 1419–24.CrossRefGoogle ScholarPubMed
Ongphiphadhanakul, B., Rajatanavin, R., Chailurkit, L.et al. (1995). Serum testosterone and its relation to bone mineral density and body composition in normal males. Clinical Endocrinology, 43, 727–33.CrossRefGoogle ScholarPubMed
Orth-Gomer, K., Eriksson, I., Moser, V., Theorell, T. and Fredlund, P. (1994). Lipid lowering through work stress reduction. International Journal of Behavioral Medicine, 1, 204–14.CrossRefGoogle ScholarPubMed
Osei, K., Schuster, D. P., Owusu, S. K. and Amoah, A. G. B. (1997). Race and ethnicity determine serum insulin and C-peptide concentrations and hepatic insulin extraction and insulin clearance: comparative studies of three populations of west African ancestry and white Americans. Metabolism, 46, 53–8.CrossRefGoogle ScholarPubMed
Ostler, K., Thompson, C., Kinmonth, A. -L. K., Peveler, R. C. and Stevens, L. (2001). Influence of socio-economic deprivation on the prevalence and outcome of depression in primary care. British Journal of Psychiatry, 178, 12–17.CrossRefGoogle ScholarPubMed
Owen, C. G., Whincup, P. H., Odoki, K., Gilg, J. A. and Cook, D. G. (2002). Infant feeding and blood cholesterol: a study in adolescents and a systematic review. Pediatrics, 110, 597–608.CrossRefGoogle ScholarPubMed
Owens, J. F., Stoney, C. M. and Matthews, K. A. (1993). Menopausal status influences ambulatory blood pressure levels and blood pressure changes during mental stress. Circulation, 88, 2794–802.CrossRefGoogle ScholarPubMed
Palacios, S. (1999). Current perspectives on the benefits of HRT in menopausal women. Maturitas, 33, S1–S13.CrossRefGoogle ScholarPubMed
Pan, X. -R., Li, G. -W., Hu, Y. H.et al. (1997). Effects of diet and exercise in preventing NIDDM in people with impaired glucose tolerance. The Da Qing IGT and diabetes study. Diabetes Care, 20, 537–44.CrossRefGoogle Scholar
Paradies, Y. C., Montoya, M. J. and Fullerton, S. M. (2007). Racialised genetics and the study of complex diseases: the thrifty genotype revisited. Perspectives in Biology and Medicine, 50, 203–27.CrossRefGoogle Scholar
Parkin, D. M. and Fernández, L. M. G. (2006). Use of statistics to assess the global burden of breast cancer. Breast Journal, 12 (Suppl. 1), S70–80.CrossRefGoogle ScholarPubMed
Parkin, D. M., Bray, F., Ferlay, J. and Pisani, P. (2005). Global cancer statistics, 2002. CA: A Cancer Journal for Clinicians, 55, 74–108.Google ScholarPubMed
Parnia, S., Borwn, J. L. and Frew, A. J. (2002). The role of pollutants in allergic sensitization and the development of asthma. Allergy, 57, 1111–17.CrossRefGoogle Scholar
Pasquali, R. (2006). Obesity, fat distribution and infertility. Maturitas, 54, 363–71.CrossRefGoogle ScholarPubMed
Pasquali, R., Pelusi, C., Genghini, S., Cacciari, M. and Gambineri, A. (2003). Obesity and reproductive disorders in women. Human Reproduction Update, 9, 359–72.CrossRefGoogle ScholarPubMed
Patel, S. M. and Nestler, J. E. (2006). Fertility in polycystic ovary syndrome. Endocrinology and Metabolism Clinics of North America, 35, 137–55.CrossRefGoogle ScholarPubMed
Patz, J. A., Epstein, P. R., Burke, T. A. and Balbus, J. M. (1996). Global climate change and emerging infectious diseases. Journal of the American Medical Association, 275, 217–23.CrossRefGoogle ScholarPubMed
Pearce-Duvet, J. M. C. (2006). The origin of human pathogens: evaluating the role of agriculture and domestic animals in the evolution of human disease. Biological Reviews, 81, 369–82.CrossRefGoogle ScholarPubMed
Penders, J., Thijs, C., Vink, C.et al. (2006). Factors influencing the composition of the intestinal microbiota in early infancy. Pediatrics, 118, 511–21.CrossRefGoogle ScholarPubMed
Pérez-Perdomo, R., Pérez-Cardona, C., Disdier-Flores, O. and Cintrón, Y. (2003). Prevalence and correlates of asthma in the Puerto Rican population: Behavioral Risk Factor Surveillance System, 2000. Journal of Asthma, 40, 465–74.CrossRefGoogle ScholarPubMed
Pettitt, D. J., Aleck, K., Baird, H., Carraher, M., Bennett, P. and Knowler, W. C. (1988). Congenital susceptibility to NIDDM: role of intrauterine environment. Diabetes, 37, 622–8.CrossRefGoogle ScholarPubMed
Pettitt, D. J., Forman, M. R., Hanson, R. L., Knowler, W. C. and Bennett, P. H. (1997). Breastfeeding and incidence of non-insulin-dependent diabetes mellitus in Pima Indians. The Lancet, 350, 166–8.CrossRefGoogle ScholarPubMed
Pi-Sunyer, F. X. (2002). Glycemic index and disease. American Journal of Clinical Nutrition, 76 (Suppl.), 290S–8S.CrossRefGoogle ScholarPubMed
Pirart, J. (1978). Diabetes mellitus and its degenerative complications: a prospective study of 4,400 patients observed between 1947 and 1973 (Part 1). Diabetes Care, 1, 168–88.CrossRefGoogle Scholar
Platts-Mills, T. A. E., Vervloet, D., Thomas, W. R., Aalberse, R. C. and Chapman, M. D. (1997). Indoor allergens and asthma: report of the Third International Workshop. Journal of Allergy and Clinical Immunology, 100, S2–24.CrossRefGoogle ScholarPubMed
Platz, E. and Giovannucci, E. (2004). The epidemiology of sex steroid hormones and their signaling and metabolic pathways in the etiology of prostate cancer. Journal of Steroid Biochemistry and Molecular Biology, 92, 237–53.CrossRefGoogle ScholarPubMed
Pollard, I. (1994). A Guide to Reproduction: Social Issues and Human Concerns. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Pollard, T. M. (1997). Environmental change and cardiovascular disease: a new complexity. Yearbook of Physical Anthropology, 40, 1–24.3.0.CO;2-8>CrossRefGoogle Scholar
Pollard, T. M. (2000). Adrenaline. In Encyclopedia of Stress, ed. Fink, G.. San Diego: Academic Press, pp. 52–8.Google Scholar
Pollard, T. M. and Ice, G. H. (2006). Measuring hormonal variation in the hypothalamic pituitary adrenal axis: cortisol. In Measuring Stress in Humans: A Practical Guide for the Field, ed. Ice, G. H. and James, G. D.. Cambridge: Cambridge University Press, pp. 122–57.CrossRefGoogle Scholar
Pollard, T. M. and Unwin, N. C. (2007). Impaired reproductive function in western and ‘westernising’ populations: an evolutionary approach. In New Perspectives in Evolutionary Medicine, ed. Trevathan, W. R., Smith, E. O. and McKenna, J. J.. Oxford: Oxford University Press.Google Scholar
Pollard, T. M., Brush, G. and Harrison, G. A. (1991). Geographic distribution of within-population variability in blood pressure. Human Biology, 63, 643–61.Google Scholar
Pollard, T. M., Núñez-de la Mora, A. and Unwin, N. C. (in press). Evolutionary perspectives on type 2 diabetes in Asia. In Medicine and Evolution: Current Applications, Future Prospects, ed. Elton, S. and O'Higgins, P.. Boca Raton: Taylor and Francis.
Pollard, T. M., Ungpakorn, G., Harrison, G. A. and Parkes, K. R. (1996). Epinephrine and cortisol responses to work: a test of the models of Frankenhaeuser and Karasek. Annals of Behavioral Medicine, 18, 229–37.CrossRefGoogle ScholarPubMed
Pollock, K. (1988). On the nature of social stress: production of a modern mythology. Social Science and Medicine, 26, 381–92.CrossRefGoogle ScholarPubMed
Pond, C. M. (1998). The Fats of Life. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Popham, F. and Mitchell, R. (2006). Leisure time exercise and personal circumstances in the working age population: longitudinal analysis of the British household panel survey. Journal of Epidemiology and Community Health, 60, 270–4.CrossRefGoogle ScholarPubMed
Popkin, B. M. (1999). Urbanization, lifestyle changes and the nutrition transition. World Development, 27, 1905–16.CrossRefGoogle Scholar
Popkin, B. M. (2001). The nutrition transition and obesity in the developing world. Journal of Nutrition, 131, 871S–3S.CrossRefGoogle ScholarPubMed
Popkin, B. M. and Gordon-Larsen, P. (2004). The nutrition transition: worldwide obesity dynamics and their determinants. International Journal of Obesity, 28, S2–9.CrossRefGoogle ScholarPubMed
Poretsky, L., Cataldo, N. A., Rosenwaks, Z. and Giudice, L. C. (1999). The insulin-related ovarian regulatory system in health and disease. Endocrine Reviews, 20, 535–82.CrossRefGoogle ScholarPubMed
Portengen, L., Preller, L., Tielen, M., Doekes, G. and Heederik, D. (2005). Endotoxin exposure and atopic sensitization in adult pig farmers. Journal of Allergy and Clinical Immunology, 115, 797–802.CrossRefGoogle ScholarPubMed
Potter, L. B., Rogler, L. H. and Moscicki, E. K. (1995). Depression among Puerto Ricans in New York City – the Hispanic Health and Nutrition Examination Survey. Social Psychiatry and Psychiatric Epidemiology, 30, 185–93.CrossRefGoogle ScholarPubMed
Poulain, M., Doucet, M., Major, G. C.et al. (2006). The effect of obesity on chronic respiratory diseases: pathophysiology and therapeutic strategies. Canadian Medical Association Journal, 174, 1293–9.CrossRefGoogle ScholarPubMed
Prentice, A. M. (2001). Fires of life: the struggles of an ancient metabolism in a modern world. BNF Nutrition Bulletin, 26, 13–27.CrossRefGoogle Scholar
Prentice, A. M. and Jebb, S. A. (1995). Obesity in Britain: gluttony or sloth. British Medical Journal, 311, 437–9.CrossRefGoogle ScholarPubMed
Prentice, A. M., Rayco-Solon, P. and Moore, S. E. (2005). Insights from the developing world: thrifty genotypes and thrifty phenotypes. Proceedings of the Nutrition Society, 64, 153–61.CrossRefGoogle ScholarPubMed
Preston, S. H. (1976). Mortality Patterns in National Populations: With Special Reference to Recorded Causes of Death. New York: Academic Press.Google Scholar
Price, J. F. and Fowkes, G. R. (1997). Risk factors and the sex differential in coronary artery disease. Epidemiology, 8, 584–91.CrossRefGoogle ScholarPubMed
Prior, I. A. M. (1971). The price of civilization. Nutrition Today, 6, 2–11.CrossRefGoogle Scholar
Raben, N., Barbetti, F., Cama, A.et al. (1991). Normal coding sequence of insulin gene in Pima Indians and Nauruans, two groups with highest prevalence of type II diabetes. Diabetes, 40, 118–22.CrossRefGoogle ScholarPubMed
Rajkumar, L., Guzman, R. C., Yang, J., Thordarson, G., Talamantes, F. and Nandi, S. (2004). Prevention of mammary carcinogenesis by short-term estrogen and progestin treatments. Breast Cancer Research, 6, R31–7.CrossRefGoogle ScholarPubMed
Randolph, J. F., Sowers, M., Gold, E. B.et al. (2003). Reproductive hormones in the early menopausal transition: relationship to ethnicity, body size, and menopausal status. Journal of Clinical Endocrinology and Metabolism, 88, 1516–22.CrossRefGoogle ScholarPubMed
Ravelli, A. C. J., van der Meulen, J. H. P., Osmond, C., Barker, D. J. P. and Bleker, O. P. (2000). Infant feeding and adult glucose tolerance, lipid profile, blood pressure, and obesity. Archives of Disease in Childhood, 82, 248–52.CrossRefGoogle ScholarPubMed
Ray, C. and Stevens, J. R. (1995). Sacred Legends. Manotick, ON: Penumbra Press.Google Scholar
Reaven, G. M. (1988). Role of insulin resistance in human disease. Diabetes, 37, 1595–607.CrossRefGoogle ScholarPubMed
Rebuffé-Scrive, M., Enk, L., Crona, N.et al. (1985). Fat cell metabolism in different regions in women: effect of menstrual cycle, pregnancy, and lactation. Journal of Clinical Investigation, 75, 1973–6.CrossRefGoogle ScholarPubMed
Reilly, J. J., Jackson, D. M., Montgomery, C.et al. (2004). Total energy expenditure and physical activity in young Scottish children: mixed longitudinal study. The Lancet, 363, 211–12.CrossRefGoogle ScholarPubMed
Reilly, J. J., Armstrong, J., Dorosty, A. R.et al. for the Avon Longitudinal Study of Parents and Children Study Team (2005). Early life risk factors for obesity in childhood: cohort study. British Medical Journal, 330, 1357–9.CrossRefGoogle ScholarPubMed
Reinhard, K. J. (1988). Cultural ecology of prehistoric parasitism on the Colorado Plateau as evidenced by coprology. American Journal of Physical Anthropology, 77, 355–66.CrossRefGoogle ScholarPubMed
Relethford, J. H. (1994). Fundamental of Biological Anthropology. Mountain View, California: Mayfield Publishing Company.Google Scholar
Renehan, A. G., Zwahlen, M., Minder, C., O'Dwyer, S. T., Shalet, S. M. and Egger, M. (2004). Insulin-like growth factor (IGF)-1, IGF binding protein-3, and cancer risk: systematic review and meta-regression analysis. The Lancet, 363, 1346–53.CrossRefGoogle ScholarPubMed
Richards, M. P. (2002). A brief review of the archaeological evidence for Palaeolithic and Neolithic subsistence. European Journal of Clinical Nutrition, 56, 1–9.CrossRefGoogle ScholarPubMed
Ridker, P. M. (2002). On evolutionary biology, inflammation, infection and the causes of atherosclerosis. Circulation, 105, 2–4.Google ScholarPubMed
Riedler, J., Braun-Fahrlander, C., Eder, W.et al. and the ALEX study team (2001). Exposure to farming in early life and development of asthma and allergy: a cross-sectional survey. The Lancet, 358, 1129–33.CrossRefGoogle ScholarPubMed
Rissanen, A. M., Heliovaara, M., Knekt, P., Reunanen, A. and Aromaa, A. (1991). Determinants of weight gain and overweight in adult Finns. European Journal of Clinical Nutrition, 45, 419–30.Google ScholarPubMed
Ritenbaugh, C. and Goodby, C. S. (1989). Beyond the thrifty gene: metabolic implications of prehistoric migration into the new world. Medical Anthropology, 11, 227–36.CrossRefGoogle ScholarPubMed
Roberts, C. A. and Cox, M. (2003). Health and Disease in Britain. Stroud: Sutton Publishing.Google Scholar
Roberts, E. M. (2002). Racial and ethnic disparities in childhood asthma diagnosis: the role of clinical findings. Journal of the National Medical Association, 94, 215–23.Google ScholarPubMed
Robinson, T. N. (2001). Population-based obesity prevention for children and adolescents. In Obesity, Growth and Development, ed. Johnston, F. E. and Foster, G. D.. London: Smith–Gordon, pp. 129–41.Google Scholar
Rode, A. and Shephard, R. J. (1971). Cardiorespiratory fitness of an Arctic community. Journal of Applied Physiology, 31, 519–26.CrossRefGoogle ScholarPubMed
Rodriguez, M. A., Winkleby, M. A., Ahn, D., Sundquist, J. and Kraemer, H. C. (2002). Identification of population subgroups of children and adolescents with high asthma prevalence. Archives of Pediatric and Adolescent Medicine, 156, 269–75.CrossRefGoogle ScholarPubMed
Rook, G. A. W., Adams, V., Hunt, J., Palmer, R., Martinelli, R. and Brunet, Rosa L. (2004). Mycobacteria and other environmental organisms as immunomodulators for immunoregulatory disorders. Springer Seminars in Immunopathology, 25, 237–55.CrossRefGoogle ScholarPubMed
Roper, N. A., Bilous, R. W., Kelly, W. F., Unwin, N. C. and Connolly, V. M. (2001). Excess mortality in a population with diabetes and the impact of material deprivation: longitudinal, population based study. British Medical Journal, 322, 1389–93.CrossRefGoogle Scholar
Rose, D., Mannino, D. M. and Leaderer, B. P. (2006). Asthma prevalence among US adults, 1998–2000: role of Puerto Rican ethnicity and behavioral and geographic factors. American Journal of Public Health, 96, 880–8.CrossRefGoogle ScholarPubMed
Rosenblatt, K. A., Thomas, D. B. and The WHO Collaborative Study of Neoplasia and Steroid Contraceptives (1995). Prolonged lactation and endometrial cancer. International Journal of Epidemiology, 24, 499–503.CrossRefGoogle ScholarPubMed
Ross, R. K., Coetzee, G. A., Reichardt, J., Skinner, E. and Henderson, B. E. (1995). Does the racial-ethnic variation in prostate cancer risk have a hormonal basis?Cancer, 75, 1778–82.3.0.CO;2-J>CrossRefGoogle Scholar
Roumain, J., Charles, M. A., Courten, M. P.et al. (1998). The relationship of menstrual irregularity to Type 2 diabetes in Pima Indian women. Diabetes Care, 21, 346–9.CrossRefGoogle ScholarPubMed
Royal, C. D. M. and Dunston, G. M. (2004). Changing the paradigm from ‘race’ to human genome variation. Nature Genetics, 36, S5–7.CrossRefGoogle ScholarPubMed
Runciman, W. G. (2005). Stone age sociology. Journal of the Royal Anthropological Institute, 11, 129–42.CrossRefGoogle Scholar
Ryan, A. S., Wenjun, Z. and Acosta, A. (2002). Breastfeeding continues to increase into the new millenium. Pediatrics, 110, 1103–9.CrossRefGoogle Scholar
Sackett, R. D. (1996). Time, Energy, and the Indolent Savage. Unpublished PhD thesis. University of California, Los Angeles.
Saelens, B. E., Sallis, J. F., Black, J. B. and Chen, D. (2003). Neighbourhood-based differences in physical activity: an environment scale evaluation. American Journal of Public Health, 93, 1552–8.CrossRefGoogle Scholar
Sahlins, M. (1972). Stone Age Economics. New York: Aldine de Gruyter.Google Scholar
Sánchez-Castillo, C. P., Velásquez-Monroy, O., Lara-Esqueda, A.et al. (2005). Diabetes and hypertension increases in a society with abdominal obesity: results of the Mexican National Health Survey 2000. Public Health Nutrition, 8, 53–60.CrossRefGoogle Scholar
Sapolsky, R. M., Krey, L. C. and McEwen, B. S. (1986). The neuroendocrinology of stress and aging: the glucocorticoid cascade hypothesis. Endocrine Reviews, 7, 284–301.CrossRefGoogle ScholarPubMed
Sapolsky, R. M., Romero, L. M. and Munck, A. U. (2000). How do glucocorticoids influence stress responses? Integrating permissive, suppressive, stimulatory, and preparative actions. Endocrine Reviews, 21, 55–89.Google ScholarPubMed
Saxton, J. M. (2006). Diet, physical activity and energy balance and their impact on breast and prostate cancers. Nutrition Research Reviews, 19, 197–215.CrossRefGoogle ScholarPubMed
Schaub, B., Lauener, R. and Mutius, E. (2006). The many faces of the hygiene hypothesis. Journal of Allergy and Clinical Immunology, 117, 969–77.CrossRefGoogle ScholarPubMed
Schieffelin, E. L. (1985). The cultural analysis of depressive affect: an example from New Guinea. In Culture and Depression: Studies in the Anthropology and Cross-Cultural Psychiatry of Affect and Disorder, ed. Kleinman, A. and Good, B.. Berkeley: University of California Press, pp. 101–33.Google Scholar
Schofield, R. and Reher, D. (1991). The decline of mortality in Europe. In The Decline of Mortality in Europe, ed. Schofield, R., Reher, D. and Bideau, D.. Oxford: Clarendon, pp. 1–17.Google Scholar
Schröder, F. H. (1996). Impact of ethnic, nutritional and environmental factors on prostate cancer. In Pharmacology, Biology, and Clinical Applications of Androgens, ed. Bhasin, S., Gabelnick, H. L., Spieleretal, J. M.. New York: Wiley-Liss, pp. 121–36.Google Scholar
Schulz, L. O., Bennett, P. H., Ravussin, E.et al. (2006). Effects of traditional and western environments on prevalence of type 2 diabetes in Pima Indians in Mexico and the US. Diabetes Care, 29, 1866–71.CrossRefGoogle Scholar
Schulze, M. B., Manson, J. E., Ludwig, D. S.et al. (2004). Sugar-sweetened beverages, weight gain, and incidence of type 2 diabetes in young and middle-aged women. Journal of the American Medical Association, 292, 927–34.CrossRefGoogle Scholar
Schutz, Y., Weinsier, R. L. and Hunter, G. R. (2001). Assessment of free-living physical activity in humans: an overview of currently available and proposed new measures. Obesity Research, 9, 368–79.CrossRefGoogle ScholarPubMed
Scott, S. and Duncan, C. J. (2001). Biology of Plagues: Evidence from Historical Populations. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Scragg, R. K. R., Fraser, A. and Metcalf, P. A. (1996). Helicobacter pylori seropositivity and cardiovascular risk factors in a multicultural workforce. Journal of Epidemiology and Community Health, 50, 578–9.CrossRefGoogle Scholar
Scrivener, S., Yemaneberhan, H., Zebenigus, M.et al. (2001). Independent effects of intestinal parasite infection and domestic allergen exposure on risk of wheeze in Ethiopia: a nested case-control study. The Lancet, 358, 1493–9.CrossRefGoogle ScholarPubMed
Seale, C. (2000). Changing patterns of death and dying. Social Science and Medicine, 51, 917–30.CrossRefGoogle ScholarPubMed
Seidell, J. C. (1995). Obesity in Europe: scaling an epidemic. International Journal of Obesity, 19 (Suppl. 3), S1–4.Google ScholarPubMed
Seidell, J. C. (2000). Obesity, insulin resistance and diabetes – a worldwide epidemic. British Journal of Nutrition, 83 (Suppl. 1), S5–8.CrossRefGoogle ScholarPubMed
Sellen, D. W. and Smay, D. B. (2001). Relationship between subsistence and age at weaning in “preindustrial” societies. Human Nature, 12, 47–87.CrossRefGoogle ScholarPubMed
Sellers, E. A. C., Triggs-Raine, B., Rockman-Greenberg, C. and Dean, H. J. (2002). The prevalence of the HNF-1alpha G319S mutation in Canadian Aboriginal youth with type 2 diabetes. Diabetes Care, 25, 2202–6.CrossRefGoogle ScholarPubMed
Sephton, S. E., Sapolsky, R. M., Kraemer, H. C. and Spiegel, D. (2000). Diurnal cortisol rhythm as a predictor of breast cancer survival. Journal of the National Cancer Institute, 92, 994–1000.CrossRefGoogle ScholarPubMed
Shaneyfelt, T., Husein, R., Bubley, G. and Mantzoros, C. S. (2000). Hormonal predictors of prostate cancer: a meta-analysis. Journal of Clinical Oncology, 18, 847–53.CrossRefGoogle ScholarPubMed
Shanley, D. P. and Kirkwood, T. B. L. (2001). Evolution of the human menopause. Bioessays, 23, 282–7.3.0.CO;2-9>CrossRefGoogle ScholarPubMed
Sharpe, R. M. and Franks, S. (2002). Environment, lifestyle and infertility – an inter-generational issue. Nature Medicine, 8 (Suppl. 11), 33–40.CrossRefGoogle Scholar
Sharpe, R. M. and Skakkebaek, N. E. (1993). Are oestrogens involved in falling sperm counts and disorders of the male reproductive tract?The Lancet, 341, 1392–5.CrossRefGoogle ScholarPubMed
Shea, J. L. (2006). Parsing the ageing Asian woman: symptom results from the China Study of Midlife Women. Maturitas, 55, 36–50.CrossRefGoogle ScholarPubMed
Sherry, D. S. and Marlowe, F. W. (2007). Anthropometric data indicate nutritional homogeneity in Hadza Foragers of Tanzania. American Journal of Human Biology, 19, 107–18.CrossRefGoogle ScholarPubMed
Shetty, P. S., Henry, C. J. K., Black, A. E. and Prentice, A. M. (1996). Energy requirements of adults: an update on basal metabolic rates (BMRs) and physical activity levels (PALs). European Journal of Clinical Nutrition, 50 (Suppl. 1), S11–23.Google Scholar
Shimuzu, H., Ross, R. K., Bernstein, L., Pike, M. C. and Henderson, B. E. (1990). Serum oestrogen levels in postmenopausal women: comparison of American whites and Japanese in Japan. British Journal of Cancer, 62, 451–3.CrossRefGoogle Scholar
Shore, S. A. (2006). Obesity and asthma: cause for concern. Current Opinion in Pharmacology, 6, 230–6.CrossRefGoogle ScholarPubMed
Shusterman, D. J., Murphy, M. A. and Balmes, J. R. (1998). Subjects with seasonal allergic rhinitis and nonrhinitic subjects react differently to nasal provocation with chlorine gas. Journal of Allergy and Clinical Immunology, 101, 732–40.CrossRefGoogle Scholar
Sicherer, S. H., Muñoz-Furlong, A. and Sampson, H. A. (2003). Prevalence of peanut and tree nut allergy in the United States determined by means of a random digit dial telephone survey: a 5-year follow-up study. Journal of Allergy and Clinical Immunology, 112, 1203–7.CrossRefGoogle ScholarPubMed
Siegert, R. J. and Ward, T. (2002). Clinical psychology and evolutionary psychology: toward a dialogue. Review of General Psychology, 6, 235–59.CrossRefGoogle Scholar
Simhon, A., Douglas, J. R., Drasar, B. S. and Sotthill, J. F. (1982). Effect of feeding on infants' faecal flora. Archives of Disease in Childhood, 57, 54–8.Google ScholarPubMed
Simmons, R. (2005). Developmental origins of adult metabolic disease: concepts and controversies. TRENDS in Endocrinology and Metabolism, 16, 390–4.CrossRefGoogle ScholarPubMed
Sinclair, A. and O'Dea, K. (1993). The significance of arachidonic acid in hunter–gatherer diets: implications for the contemporary western diet. Journal of Food Lipids, 1, 143–57.CrossRefGoogle Scholar
Slattery, M. L. and Randall, D. E. (1988). Trends in coronary heart disease mortality and food consumption in the United States between 1909 and 1980. American Journal of Clinical Nutrition, 47, 1060–70.CrossRefGoogle ScholarPubMed
Smith, B. H. (1991). Dental development and the evolution of life history in Hominidae. American Journal of Physical Anthropology, 86, 157–74.CrossRefGoogle Scholar
Smith, C. J. (1994). Food habit and cultural changes among the Pima Indians. In Diabetes as a Disease of Civilization, ed. Joe, J. R. and Young, R. S.. New York: Mouton de Gruyter.CrossRefGoogle Scholar
Smith, C. J., Nelson, R. G., Hardy, S. A., Manahan, E. M., Bennett, P. H. and Knowler, W. C. (1996). Survey of the diet of Pima Indians using quantitative food frequency assessment and 24-hour recall. Journal of the American Dietetic Association, 96, 778–84.CrossRefGoogle ScholarPubMed
Smith, E. O. (2002). When Culture and Biology Collide: Why we are Stressed, Depressed, and Self-obsessed. New Brunswick, New Jersey: Rutgers University Press.Google Scholar
Smith, M. T. (1993). Genetic adaptation. In Human Adaptation, ed. Harrison, G. A.. Oxford: Oxford University Press, pp. 1–54.Google Scholar
Smyth, J., Ockenfels, M. C., Porter, L., Kirschbaum, C., Hellhammer, D. H. and Stone, A. A. (1998). Stressors and mood measured on a momentary basis are associated with salivary cortisol secretion. Psychoneuroendocrinology, 23, 353–70.CrossRefGoogle ScholarPubMed
Snijder, M. B., Dam, R. M., Visser, M. and Seidell, J. C. (2005). What aspects of body fat are particularly hazardous and how do we measure them?International Journal of Epidemiology, 35, 83–92.CrossRefGoogle Scholar
Solomon, C. G. (1999). The epidemiology of polycystic ovary syndrome. Endocrinology and Metabolism Clinics of North America, 28, 247–63.CrossRefGoogle ScholarPubMed
Sood, A., Ford, E. S. and Camargo, C. A. (2006). Association between leptin and asthma in adults. Thorax, 61, 300–5.CrossRefGoogle ScholarPubMed
Sowers, J. R. (2003). Obesity as a cardiovascular risk factor. American Journal of Medicine, 115, 37S–41S.CrossRefGoogle ScholarPubMed
Spiegel, K., Knutson, K., Leproult, R., Tasali, E. and Cauter, E. (2005). Sleep loss: a novel risk factor for insulin resistance and type 2 diabetes. Journal of Applied Physiology, 99, 2008–19.CrossRefGoogle ScholarPubMed
Spielman, R. S., Fajans, S. S., Neel, J. V., Pek, S., Floyd, J. C. and Oliver, W. J. (1982). Glucose tolerance in two unacculturated Indian tribes of Brazil. Diabetologia, 23, 90–3.CrossRefGoogle ScholarPubMed
Stamatakis, E., Primatesta, P., Chinn, S., Rona, R. and Falascheti, E. (2005). Overweight and obesity trends from 1974 to 2003 in English children: what is the role of socioeconomic factors?Archives of Disease in Childhood, 90, 999–1004.CrossRefGoogle ScholarPubMed
Stampfer, M. J. (2006). Cardiovascular disease and Alzheimer's disease: common links. Journal of Internal Medicine, 260, 211–23.CrossRefGoogle ScholarPubMed
Stampfer, M. J. and Rimm, E. B. (1995). Epidemiologic evidence for vitamin E in prevention of cardiovascular disease. American Journal of Clinical Nutrition, 62, S1365–9.CrossRefGoogle ScholarPubMed
Stanhope, J. M. and Prior, I. A. M. (1980). The Tokelau island migrant study: prevalence and incidence of diabetes mellitus. New Zealand Medical Journal, 92, 417–21.Google ScholarPubMed
Stearns, S. C. and Ebert, D. (2001). Evolution in health and disease: work in progress. Quarterly Review of Biology, 76, 417–32.CrossRefGoogle ScholarPubMed
Steckel, R. H. and Rose, J. C. (2002b). Conclusions. In The Backbone of History: Health and Nutrition in the Western Hemisphere, ed. Steckel, R. H. and Rose, J. C.. Cambridge: Cambridge University Press, pp. 583–9.CrossRefGoogle Scholar
Steckel, R. H. and Rose, J. C. (2002a). Patterns of health in the western hemisphere. In The Backbone of History: Health and Nutrition in the Western Hemisphere, ed. Steckel, R. H. and Rose, J. C.. Cambridge: Cambridge University Press, pp. 563–79.CrossRefGoogle Scholar
Steckel, R. H., Rose, J. C., Larsen, C. S. and Walker, P. L. (2002). Skeletal health in the Western Hemisphere from 4000BC to the present. Evolutionary Anthropology, 11, 142–55.CrossRefGoogle Scholar
Steffen, P. R., McNeilly, M., Anderson, N. and Sherwood, A. (2003). Effects of perceived racism and anger inhibition on ambulatory blood pressure in African Americans. Psychosomatic Medicine, 65, 746–50.CrossRefGoogle ScholarPubMed
Stene, L. C. and Nafstad, P. (2001). Relation between occurrence of type 1 diabetes and asthma. The Lancet, 357, 607–8.CrossRefGoogle ScholarPubMed
Steptoe, A. and Marmot, M. (2002). The role of psychobiological pathways in socio-economic inequalities in cardiovascular disease risk. European Heart Journal, 23, 13–25.CrossRefGoogle ScholarPubMed
Steptoe, A., Wardle, J., Lipsey, Z., Oliver, G., Pollard, T. M. and Davies, G. J. (1998). The effects of life stress on food choice. In The Nation's Diet: The Social Science of Food Choice, ed. Murcott, A.. Harlow: Addison Wesley Longman, pp. 29–42.Google Scholar
Stevens, R. G. (2006). Artificial lighting in the industrialized world: circadian disruption and breast cancer. Cancer Causes and Control, 17, 501–7.CrossRefGoogle ScholarPubMed
Stevens, R. G. and Rea, M. S. (2001). Light in the built environment: potential role of circadian disruption in endocrine disruption and breast cancer. Cancer Causes and Control, 12, 279–87.CrossRefGoogle ScholarPubMed
Stinson, S. (2002). Early childhood health in foragers. In Human Diet: Its Origin and Evolution, ed. Ungar, P. S. and Teaford, M. F.. Westport, Connecticut: Bergin and Garvey, pp. 37–48.Google Scholar
Storey, R. (1985). An estimate of mortality in a Pre–Columbian urban population. American Anthropologist, 83, 519–35.CrossRefGoogle Scholar
Strachan, D. P. (1989). Hay fever, hygiene and household size. British Medical Journal, 299, 1259–60.CrossRefGoogle ScholarPubMed
Strachan, D. P. (1997). Allergy and family size: a riddle worth solving. Clinical and Experimental Allergy, 27, 235–6.CrossRefGoogle ScholarPubMed
Strassmann, B. I. (1999). Menstrual cycling and breast cancer: an evolutionary perspective. Journal of Women's Health, 8, 193–202.CrossRefGoogle Scholar
Strassmann, B. I. and Dunbar, R. I. M. (1999). Human evolution and disease: putting the Stone Age in perspective. In Evolution in Health and Disease, ed. Stearns, S. C.. Oxford: Oxford University Press, pp. 91–101.Google Scholar
Strauss, R. S. and Pollack, H. A. (2001). Epidemic increase in childhood overweight, 1986–1998. Journal of the American Medical Association, 286, 2845–8.CrossRefGoogle ScholarPubMed
Stringer, C. (2002). Modern human origins: progress and prospects. Philosophical Transactions of the Royal Society of London B, 357, 563–79.CrossRefGoogle ScholarPubMed
Stringer, C. (2003). Out of Ethiopia. Nature, 423, 692–5.CrossRefGoogle ScholarPubMed
Stuart-Macadam, P. (1995). Breastfeeding in prehistory. In Breastfeeding: Biocultural Perspectives, ed. Stuart-Macadam, P. and Dettwyler, K. A.. New York: Aldine de Gruyter, pp. 75–99.Google Scholar
Stuart-Macadam, P. (1998). Iron deficiency anemia: exploring the difference. In Sex and Gender in Paleopathological Perspective, ed. Grauer, A. L. and Stuart-Macadam, P.. Cambridge: Cambridge University Press, pp. 45–63.Google Scholar
Sutton-Tyrrell, K., Wildman, R. P., Matthews, K. A.et al. (2005). Sex hormone-binding globulin and the free androgen index are related to cardiovascular risk factors in multiethnic premenopausal women and perimenopausal women enrolled in the study of women across the nation (SWAN). Circulation, 111, 1242–9.CrossRefGoogle Scholar
Swinburn, B. and Egger, G. (2004). The runaway weight gain train: too many accelerators, not enough brakes. British Medical Journal, 329, 736–9.CrossRefGoogle Scholar
Szathmáry, E. J. E. (1994). Non-insulin dependent diabetes mellitus among Aboriginal North Americans. Annual Review of Anthropology, 23, 457–82.CrossRefGoogle Scholar
Taheri, S. (2006). The link between short sleep duration and obesity: we should recommend more sleep to prevent obesity. Archives of Disease in Childhood, 91, 881–4.CrossRefGoogle ScholarPubMed
Taheri, S., Lin, L., Austin, D., Young, T. and Mignot, E. (2004). Short sleep duration is associated with reduced leptin, elevated ghrelin, and increased body mass index. PLoS Medicine, 1, 210–17.CrossRefGoogle ScholarPubMed
Talayero, J. M. P., Lizan-Garcia, M., Puime, A. O.et al. (2006). Full breastfeeding and hospitalization as a result of infections in the first year of life. Pediatrics, 118, E92–9.CrossRefGoogle Scholar
Tatz, C. (2005). Aboriginal Suicide is Different: Portrait of Life and Self-Destruction. Canberra: Aboriginal Studies Press.Google Scholar
Taylor, A. L., Dunstan, J. A. and Prescott, S. L. (2007). Probiotic supplementation for the first 6 months of life fails to reduce the risk of atopic dermatitis and increases the risk of allergen sensitization in high-risk children: a randomized controlled trial. Journal of Allergy and Clinical Immunology, 119, 184–91.CrossRefGoogle ScholarPubMed
Taylor, J. S., Kacmar, J. E., Nothnagle, M. and Lawrence, R. A. (2005). A systematic review of the literature associating breastfeeding with type 2 diabetes and gestational diabetes. Journal of the American College of Nutrition, 24, 320–6.CrossRefGoogle ScholarPubMed
Tesfaye, F., Nawi, N. G., Minh, H.et al. (2007). Association between body mass index and blood pressure across three populations in Africa and Asia. Journal of Human Hypertension, 21, 28–37.CrossRefGoogle ScholarPubMed
Thakore, J. H., Richards, P., Reznek, R. H., Martin, A. and Dinan, T. G. (1997). Increased intra-abdominal fat deposition in patients with major depressive illness as measured by computed tomography. Biological Psychiatry, 41, 1140–2.CrossRefGoogle ScholarPubMed
Thomas, R. B. (1998). The evolution of human adaptability paradigms: toward a biology of poverty. In Building a New Biocultural Synthesis, ed. Goodman, A. H. and Leatherman, T. L.. Ann Arbor: University of Michigan Press, pp. 43–74.Google Scholar
Thomson, N. J. (1991). Recent trends in aboriginal mortality. Medical Journal of Australia, 154, 235–9.Google ScholarPubMed
Thorburn, A. W. (2005). National prevalence of obesity – prevalence of obesity in Australia. Obesity Reviews, 6, 187–9.CrossRefGoogle Scholar
Tillin, T., Forouhi, N., Johnston, D. G., McKeigue, P., Chaturvedi, N. and Godsland, I. F. (2005). Metabolic syndrome and coronary heart disease in South Asians, African-Caribbeans and white Europeans: a UK population-based cross-sectional study. Diabetologia, 48, 649–56.CrossRefGoogle ScholarPubMed
Tishkoff, S. A. and Kidd, K. K. (2004). Implications of biogeography of human populations for ‘race’ and medicine. Nature Genetics, 36, S21–7.CrossRefGoogle Scholar
Tishkoff, S. A., Reed, F. A., Ranciaro, A.et al. (2007). Convergent adaptation of human lactase persistence in Europe. Nature Genetics, 39, 31–40.CrossRefGoogle Scholar
Tominaga, S. and Kuroishi, T. (1997). An ecological study on diet/nutrition and cancer in Japan. International Journal of Cancer, 71, S10, 2–6.3.0.CO;2-C>CrossRefGoogle Scholar
Tonetti, D. (2004). Prevention of breast cancer by recapitulation of pregnancy hormone levels. Breast Cancer Research, 6, E8.CrossRefGoogle ScholarPubMed
Travis, R. C., Allen, D. S., Fentiman, I. S. and Key, T. J. (2004). Melatonin and breast cancer: a prospective study. Journal of the National Cancer Institute, 96, 475–82.CrossRefGoogle ScholarPubMed
Treloar, A. E., Boynton, R. E., Behn, B. G. and Brown, B. W. (1967). Variation of the human menstrual cycle through reproductive life. International Journal of Fertility, 12, 77–126.Google ScholarPubMed
Tremblay, M. S., Katzmarzyk, P. T. and Willms, J. D. (2002). Temporal trends in overweight and obesity in Canada 1981–1996. International Journal of Obesity, 26, 538–43.CrossRefGoogle ScholarPubMed
Trevathan, W. R., Smith, E. O. and McKenna, J. J., eds. (1999). Evolutionary Medicine. New York: Oxford University Press.Google Scholar
Trevathan, W. R., Smith, E. O. and McKenna, J. J., eds. (2007). New Perspectives in Evolutionary Medicine. New York: Oxford University Press.Google Scholar
Trichopoulos, D., MacMahon, B. and Cole, P. (1972). Menopause and breast cancer risk. Journal of the National Cancer Institute, 48, 605–13.Google ScholarPubMed
Triggs-Raine, B. L., Kirkpatrick, R. D., Kelly, S. L.et al. (2002). HNF-1alpha G319S, a transactivation-deficient mutant, is associated with altered dynamics of diabetes onset in an Oji-Cree community. Proceedings of the National Academy of Science, 99, 4614–19.CrossRefGoogle Scholar
Troiano, R. P., Briefel, R. R., Carroll, M. D. and Bialostosky, K. (2000). Energy and fat intakes of children and adolescents in the United States: data from the National Health and Nutrition Examination Surveys. American Journal of Clinical Nutrition, 72, 1343S–53S.CrossRefGoogle ScholarPubMed
Trowell, H. C. and Burkitt, D. P. (1981a). Preface. In Western Diseases: Their Emergence and Prevention, ed. Trowell, H. C. and Burkitt, D. P.. Cambridge, Massachusetts: Harvard University Press, pp. xiii–xvi.Google Scholar
Trowell, H. C. and Burkitt, D. P., eds. (1981b). Western Diseases: Their Emergence and Prevention. Cambridge, Massachusetts: Harvard University Press.Google Scholar
Truswell, A. S. and Hansen, J. D. L. (1976). Medical research among the!Kung. In Kalahari Hunter–Gatherers: Studies of the!Kung San and their Neighbours, ed. Lee, R. B. and DeVore, I.. Cambridge, Massachusetts: Harvard University Press, pp. 166–94.CrossRefGoogle Scholar
Tunstall-Pedoe, H., Connaghan, J., Woodward, M., Tolonen, H. and Kuulasmaa, K. (2006). Pattern of declining blood pressure across replicate population surveys of the WHO MONICA project, mid-1980s to mid-1990s, and the role of medication. British Medical Journal, 332, 629–35.CrossRefGoogle ScholarPubMed
Ueshima, H., Okayama, A., Saitoh, S.et al. (2003). Differences in cardiovascular disease risk factors between Japanese in Japan and Japanese-Americans in Hawaii: the INTERLIPID study. Journal of Human Hypertension, 17, 631–9.CrossRefGoogle ScholarPubMed
Umetsu, D. T. and DeKruyff, R. H. (2006). The regulation of allergy and asthma. Immunological Reviews, 212, 238–55.CrossRefGoogle ScholarPubMed
Uusitalo, U., Feskens, E. J. M., Tuomilehto, J.et al. (1996). Fall in total cholesterol concentration over five years in association with changes in fatty acid composition of cooking oil in Mauritius: cross sectional survey. British Medical Journal, 313, 1044–6.CrossRefGoogle ScholarPubMed
Anders, S. M. and Watson, N. V. (2006). Menstrual cycle irregularities are associated with testosterone levels in healthy premenopausal women. American Journal of Human Biology, 18, 841–4.CrossRefGoogle ScholarPubMed
Blerkom, L. (2003). Role of viruses in human evolution. Yearbook of Physical Anthropology, 46, 14–46.CrossRefGoogle Scholar
Klink, J. J. L., Blonk, R. W. B., Schene, A. H. and van Dijk, F. J. H. (2001). The benefits of interventions for work-related stress. American Journal of Public Health, 91, 271–6.Google ScholarPubMed
Spuy, Z. M. and Dyer, S. J. (2004). The pathogenesis of infertility and early pregnancy loss in polycystic ovary syndrome. Best Practice and Research Clinical Obstetrics and Gynaecology, 18, 755–71.CrossRefGoogle ScholarPubMed
Eck, M. M. and Nicolson, N. A. (1994). Perceived stress and salivary cortisol in daily life. Annals of Behavioral Medicine, 16, 221–7.Google Scholar
Hooff, M. H. A., Voorhorst, F. J., Kaptein, M. B. H., Hirasing, R. A., Koppenaal, C. and Schoemaker, J. (2004). Predictive value of menstrual cycle pattern, body mass index, hormone levels and polycystic ovaries at age 15 years for oligo-amenorrhoea at age 18 years. Human Reproduction, 19, 383–92.CrossRefGoogle ScholarPubMed
Odijk, J., Kull, I., Borres, M. P.et al. (2003). Breastfeeding and allergic disease: a multidisciplinary review of the literature (1966–2001) on the mode of early feeding in infancy and its impact on later atopic manifestations. Allergy, 58, 833–43.CrossRefGoogle ScholarPubMed
Poppel, G., Kardinaal, A., Princen, H. and Kok, F. J. (1994). Antioxidants and coronary heart disease. Annals of Medicine, 26, 429–34.CrossRefGoogle ScholarPubMed
Vartiainen, E., Jousilahti, P., Alfthan, G., Sundvall, J., Pietinen, P. and Puska, P. (2000). Cardiovascular risk factors in Finland, 1972–1997. International Journal of Epidemiology, 29, 49–56.CrossRefGoogle Scholar
Verkasalo, P. K., Thomas, H. V., Appleby, P. N., Davey, G. K. and Key, T. J. (2001). Circulating levels of sex hormones and their relation to risk factors for breast cancer: a cross-sectional study in 1092 pre- and post-menopausal women (United Kingdom). Cancer Causes and Control, 12, 47–59.CrossRefGoogle Scholar
Villamor, E. and Cnattingius, S. (2006). Interpregnancy weight change and risk of adverse pregnancy outcomes: a population-based study. The Lancet, 368, 1164–70.CrossRefGoogle ScholarPubMed
Virtanen, S. M. and Knip, M. (2003). Nutritional risk predictors of β cell autoimmunity and type 1 diabetes at a young age. American Journal of Clinical Nutrition, 78, 1053–67.CrossRefGoogle Scholar
Vitzthum, V. J. (2001). Why not so great is still good enough. In Reproductive Ecology and Human Evolution, ed. Ellison, P. T.. New York: Aldine de Gruyter, pp. 179–202.Google Scholar
Vitzthum, V. J., Bentley, G. R., Spielvogel, H.et al. (2002). Salivary progesterone levels and rate of ovulation are significantly lower in poorer than in better-off urban-dwelling Bolivian women. Human Reproduction, 17, 1906–13.CrossRefGoogle ScholarPubMed
Vitzthum, V. J., Spielvogel, H. and Thornburg, J. (2004). Interpopulational differences in progesterone levels during conception and implantation in humans. Proceedings of the National Academy of Science, 101, 1443–8.CrossRefGoogle ScholarPubMed
Vogel, V. G., Costantino, J. P., Wickerham, D. L.et al. for the National Surgical Adjuvant Breast and Bowel Project (NSABP) (2006). Effects of tamoxifen vs raloxifene on the risk of developing invasive breast cancer and other disease outcomes: the NSABP study of tamoxifen and raloxifene (STAR) P-2 trial. Journal of the American Medical Association, 295, 2727–41.CrossRefGoogle ScholarPubMed
Ehrenstein, O. S., Mutius, E., Illi, S., Baumann, L., Bohm, O. and Kries, R. (2000). Reduced risk of hay fever and asthma among children of farmers. Clinical and Experimental Allergy, 30, 187–93.CrossRefGoogle Scholar
Hertzen, L. and Haahtela, T. (2005). Signs of reversing trends in prevalence of asthma. Allergy, 60, 283–92.CrossRefGoogle Scholar
Mutius, E., Martinez, F. D., Fritzsch, C., Nicolai, T., Reitmeir, P. and Thiemann, H. -H. (1994). Skin test reactivity and number of siblings. British Medical Journal, 308, 692–5.CrossRefGoogle Scholar
Vorona, R. D., Winn, M. P., Babineau, T. W., Eng, B. P., Feldman, H. R. and Ware, J. C. (2005). Overweight and obese patients in a primary care population report less sleep than patients with a normal body mass index. Archives of Internal Medicine, 165, 25–30.CrossRefGoogle Scholar
Waldron, I. (1991). Patterns and causes of gender differences in smoking. Social Science and Medicine, 32, 989–1005.CrossRefGoogle ScholarPubMed
Wallace, B. A. and Cumming, R. G. (2000). Systematic review of randomized trials of the effect of exercise on bone mass in pre- and postmenopausal women. Calcified Tissue International, 67, 10–18.CrossRefGoogle Scholar
Walters, V. (1993). Stress, anxiety and depression: women's accounts of their health problems. Social Science and Medicine, 36, 393–402.CrossRefGoogle ScholarPubMed
Wang, C., Catlin, D. H., Starcevic, B.et al. (2005). Low-fat high-fiber diet decreased serum and urine androgens in men. Journal of Clinical Endocrinology and Metabolism, 90, 3550–9.CrossRefGoogle ScholarPubMed
Wang, Y. Z., Mi, J., Shan, X. -Y., Wang, Q. J. and Ge, K. -Y. (2007). Is China facing an obesity epidemic and the consequences? The trends in obesity and chronic disease in China. International Journal of Obesity, 31, 177–88.CrossRefGoogle Scholar
Warren, M. (2004). A comparative review of the risks and benefits of hormone replacement therapy regimens. American Journal of Obstetrics and Gynecology, 190, 1141–67.CrossRefGoogle ScholarPubMed
Watts, J. (2006). Doctors blame air pollution for China's asthma increases. The Lancet, 368, 719–720.CrossRefGoogle ScholarPubMed
Weedon, M. N., Schwarz, P. E. H., Horikawa, Y.et al. (2003). Meta-analysis and a large association study confirm a role for Calpain-10 variation in type 2 diabetes susceptibility. American Journal of Human Genetics, 73, 1208–12.CrossRefGoogle Scholar
Weir, G., Laybutt, D., Kaneto, H., Bonner-Weir, S. and Sharma, A. (2001). Beta-cell adaptation and decompensation during the progression to diabetes. Diabetes, 50 (Suppl.1), 5154–9.CrossRefGoogle ScholarPubMed
Weiss, K. M. and Fullerton, S. M. (2005). Racing around, getting nowhere. Evolutionary Anthropology, 14, 165–9.CrossRefGoogle Scholar
Weiss, K. M., Ferrell, R. E. and Hanis, C. L. (1984). A New World Syndrome of metabolic diseases with a genetic and evolutionary basis. Yearbook of Physical Anthropology, 27, 153–78.CrossRefGoogle Scholar
Wells, J. C. K. (2006). The evolution of human fatness and susceptibility to obesity: an ethological approach. Biological Reviews, 81, 183–205.CrossRefGoogle ScholarPubMed
Wenzel, S. E. (2006). Asthma: defining of the persistent adult phenotypes. The Lancet, 368, 804–813.CrossRefGoogle ScholarPubMed
West, K. M. (1974). Diabetes in American Indians and other native populations of the New World. Diabetes, 23, 841–55.CrossRefGoogle ScholarPubMed
West, R. (2006). Tobacco control: present and future. British Medical Bulletin, 77–78, 123–36.CrossRefGoogle ScholarPubMed
Whitmer, R. A., Gunderson, E. P., Barrett-Connor, E., Quesenberry, C. P. and Yaffe, K. (2005). Obesity in middle age and future risk of dementia: a 27 year longitudinal population based study. British Medical Journal, 330, 1360–2.CrossRefGoogle ScholarPubMed
WHO Expert Consultation (2004). Appropriate body-mass index for Asian populations and its implications for policy and intervention strategies. The Lancet, 363, 157–63.CrossRefGoogle Scholar
WHO Global Infobase Team (2005). The SuRF Report 2. Surveillance of Chronic Disease Risk Factors: Country-Level Data and Comparable Estimates. Geneva: World Health Organization.
WHO International Consortium in Psychiatric Epidemiology (2000). Cross-national comparisons of the prevalences and correlates of mental disorders. Bulletin of the World Health Organization, 78, 413–26.Google Scholar
Wild, S., Roglic, G., Green, A., Sicree, R. and King, H. (2004). Global prevalence of diabetes: estimates for the year 2000 and projections for 2030. Diabetes Care, 27, 1047–53.CrossRefGoogle ScholarPubMed
Wild, S. H. and Byrne, C. D. (2006). Risk factors for diabetes and coronary heart disease. British Medical Journal, 333, 1009–11.CrossRefGoogle ScholarPubMed
Wilkinson, R. (1999). Health, hierarchy, and social anxiety. Annals of the New York Academy of Sciences, 896, 48–63.CrossRefGoogle ScholarPubMed
Williams, B. (1995). Westernised Asians and cardiovascular disease: nature or nurture. The Lancet, 345, 401–2.CrossRefGoogle ScholarPubMed
Williams, D. R. and Collins, C. (1995). US socioeconomic and racial differ-entials in health: patterns and explanations. Annual Review of Sociology, 21, 349–86.CrossRefGoogle Scholar
Williams, D. R., Neighbors, H. W. and Jackson, J. S. (2003). Racial/ethnic discrimination and health: findings from community studies. American Journal of Public Health, 93, 200–8.CrossRefGoogle ScholarPubMed
Williams, G. C. (1957). Pleiotropy, natural selection, and the evolution of senescence. Evolution, 11, 398–411.CrossRefGoogle Scholar
Williams, G. C. and Nesse, R. M. (1991). The dawn of Darwinian medicine. Quarterly Review of Biology, 66, 1–22.CrossRefGoogle ScholarPubMed
Williams, R. C., Long, J. C., Hanson, R. L., Sievers, M. L. and Knowler, W. C. (2000). Individual estimates of European genetic admixture associated with lower body-mass index, plasma glucose, and prevalence of type 2 diabetes in Pima Indians. American Journal of Human Genetics, 66, 527–38.CrossRefGoogle ScholarPubMed
Wilmoth, J. R. (2000). Demography of longevity: past, present, and future trends. Experimental Gerontology, 35, 1111–29.CrossRefGoogle ScholarPubMed
Wilson, B. D., Wilson, N. C. and Russell, D. G. (2001). Obesity and body fat distribution in the New Zealand population. New Zealand Medical Journal, 114, 127–30.Google ScholarPubMed
Wilson, T. W. and Grim, C. E. (1991). Biohistory of slavery and blood pressure differences in blacks today. A hypothesis. Hypertension, 17 (Suppl. 1), 1122–9.CrossRefGoogle ScholarPubMed
Wood, B. and Collard, M. (1999). The human genus. Science, 284, 65–71.CrossRefGoogle ScholarPubMed
World Health Organization (1999). World Health Report: Making a Difference. Geneva, World Health Organization.
World Health Organization (2003). Diet, Nutrition and the Prevention of Chronic Diseases. Geneva: World Health Organization.
Worthman, C. M. and Melby, M. K. (2002). Toward a comparative developmental ecology of human sleep. In Adolescent Sleep Patterns: Biological, Social and Psychological Influences, ed. Carskadon, M. A.. New York: Cambridge University Press, pp. 69–117.CrossRefGoogle Scholar
Wrigley, E. A. (1969). Population and History. London: Weidenfeld and Nicolson.
Writing Group for the Women's Health Initiative Investigators (2002). Risks and benefits of estrogen plus progestin in healthy postmenopausal women: principal results from the Women's Health Initiative randomized controlled trial. Journal of the American Medical Association, 288, 321–33.CrossRefGoogle Scholar
Wu, A. H., Pike, M. C. and Stram, D. O. (1999). Meta-analysis: dietary fat intake, serum estrogen levels, and the risk of breast cancer. Journal of the National Cancer Institute, 91, 529–34.CrossRefGoogle ScholarPubMed
Wu, Y. (2006). Overweight and obesity in China. British Medical Journal, 333, 362–3.CrossRefGoogle Scholar
Yajnik, C. (2000). Interactions of perturbations in intrauterine growth and growth during childhood on the risk of adult-onset disease. Proceedings of the Nutrition Society, 59, 257–65.CrossRefGoogle ScholarPubMed
Yajnik, C., Lubree, H., Rege, S.et al. (2002). Adiposity and hyperinsulinemia in Indians are present at birth. Journal of Clinical Endocrinology and Metabolism, 87, 5575–80.CrossRefGoogle ScholarPubMed
Yang, L., Parkin, D. M., Ferlay, J., Li, L. and Chen, Y. (2005). Estimates of cancer incidence in China for 2000 and projections for 2005. Cancer Epidemiology, Biomarkers and Prevention, 14, 243–50.Google ScholarPubMed
Yazdanbakhsh, M., Kremsner, P. G. and van Ree, R. (2002). Allergy, parasites, and the hygiene hypothesis. Science, 296, 490–4.CrossRefGoogle ScholarPubMed
Yemaneberhan, H., Bekele, Z., Venn, A., Lewis, S., Parry, E. and Britton, J. (1997). Prevalence of wheeze and asthma and relation to atopy in urban and rural Ethiopia. The Lancet, 350, 85–90.CrossRefGoogle Scholar
Yemaneberhan, H., Flohr, C., Lewis, S. A.et al. (2004). Prevalence and associated factors of atopic dermatitis symptoms in rural and urban Ethiopia. Clinical and Experimental Allergy, 34, 779–85.CrossRefGoogle ScholarPubMed
Yoon, K. -H., Lee, J. -H., Kim, J. -W.et al. (2006). Epidemic obesity and type 2 diabetes in Asia. The Lancet, 368, 1681–8.CrossRefGoogle Scholar
Young, D. B., Lin, H. and McCabe, R. D. (1995). Potassium's cardiovascular protective mechanisms. American Journal of Physiology: Regulatory, Integrative and Comparative Physiology, 37, R825–37.Google Scholar
Young, J. H., Chang, Y. P. C., Kim, J. D. O.et al. (2005). Differential susceptibility to hypertension is due to selection during the out-of-Africa expansion. PLoS Genetics, 1, 730–8.CrossRefGoogle ScholarPubMed
Young, T. K., Reading, J., Elias, B. and O'Neil, J. D. (2000). Type 2 diabetes mellitus in Canada's First Nations: status of an epidemic in progress. Canadian Medical Association Journal, 163, 561–6.Google ScholarPubMed
Yusuf, S., Reddy, S., Ounpuu, S. and Anand, S. (2001a). Global burden of cardiovascular diseases. Part I: General considerations, the epidemiologic transition, risk factors, and the impact of urbanization. Circulation, 104, 2746–53.CrossRefGoogle Scholar
Yusuf, S., Reddy, S., Ôunpuu, S. and Anand, S. (2001b). Global burden of cardiovascular diseases. Part II: Variations in cardiovascular disease by specific ethnic groups and geographic regions and prevention strategies. Circulation, 104, 2855–64.CrossRefGoogle Scholar
Zainudin, B. M. Z., Lai, C. K. W., Sporiano, J. B., Jia-Horng, W. and Guia, T. S. (2005). Asthma control in adults in Asia–Pacific. Respirology, 10, 579–86.CrossRefGoogle ScholarPubMed
Zhou, B. F., Stamler, J., Dennis, B.et al. (2003). Nutrient intakes of middle-aged men and women in China, Japan, United Kingdom, and United States in the late 1990s: the INTERMAP study. Journal of Human Hypertension, 17, 623–30.CrossRefGoogle ScholarPubMed
Ziegler, R. G., Hoover, R. N., Pike, M. C.et al. (1993). Migration patterns and breast cancer risk in Asian-American women. Journal of the National Cancer Institute, 85, 1819–27.CrossRefGoogle ScholarPubMed
Zimmet, P. (2000). Globalization, coca-colonization and the chronic disease epidemic: can the Doomsday scenario be averted?Journal of Internal Medicine, 247, 301–10.CrossRefGoogle ScholarPubMed
Zimmet, P., Taft, P., Guinea, A., Guthrie, W. and Thoma, K. (1977). The high prevalence of diabetes mellitus on a central Pacific island. Diabetologia, 13, 111–15.CrossRefGoogle ScholarPubMed
Zimmet, P., Faaiuso, S., Ainuu, J., Whitehouse, S., Milne, B. and DeBoer, W. (1981). The prevalence of diabetes in the rural and urban Polynesian population of Western Samoa. Diabetes, 30, 45–51.CrossRefGoogle Scholar
Zizza, C., Siega-Riz, A. M. and Popkin, B. M. (2001). Significant increases in young adults' snacking between 1977–1978 and 1994–1996 represents a cause of concern!Preventive Medicine, 32, 303–10.CrossRefGoogle Scholar
Zografos, G. C., Panou, M. and Panou, N. (2004). Common risk factors of breast and ovarian cancer: recent view. International Journal of Gynecogical Cancer, 14, 721–40.CrossRefGoogle ScholarPubMed
Zoratti, R. (1998). A review on ethnic differences in plasma triglyceride and high-density-lipoprotein cholesterol: is the lipid pattern the key factor for the low coronary heart disease rate in people of African origin?European Journal of Epidemiology, 14, 9–21.CrossRefGoogle ScholarPubMed

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

  • References
  • Tessa M. Pollard, University of Durham
  • Book: Western Diseases
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511841118.011
Available formats
×

Save book to Dropbox

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

  • References
  • Tessa M. Pollard, University of Durham
  • Book: Western Diseases
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511841118.011
Available formats
×

Save book to Google Drive

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

  • References
  • Tessa M. Pollard, University of Durham
  • Book: Western Diseases
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511841118.011
Available formats
×