Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-9pm4c Total loading time: 0 Render date: 2024-04-25T14:02:52.217Z Has data issue: false hasContentIssue false

2 - Human Emergence

Published online by Cambridge University Press:  05 August 2014

John L. Brooke
Affiliation:
Ohio State University
Get access

Summary

In the long run from 60 million years ago to 60,000 years ago, the global super-cycle moved from Greenhouse to Icehouse, from a warm, carbon-rich world in which ferns grew and dinosaurs grazed at the earth’s polar extremes to a cold, carbon-depleted world, where a new humanity stood on dry equatorial African shores on the edge of a global colonization leading to our very recent past. For decades, it has been a central tenet of evolutionary studies that the human condition has its origins in this inexorable march into Icehouse conditions. Prehuman and human history over the past 15 million years has a fundamental relationship to specific tectonic events and the regular oscillations of glacial climates: as Steven Stanley has put it, we are “children of the ice age.” More precisely, we are the children of the ice age tropics, if over time venturing into the grasslands and tundra bordering the ice itself. And it is always a shock to realize that the entirety of agriculturally based human civilization has unfolded in a brief 10,000-year interval embedded within this long sequence of glaciations, which in turn lies at the bottom of a 60-million-year descent into an Icehouse epoch. The chapters following this look with some care at this brief 10,000-year history. This chapter charts the global transition from Greenhouse to Icehouse, and then examines the emergence of humanity through the worst of the current Icehouse, between the divergence of bipedal human forbears from the arboreal great apes roughly 6 million years ago and the colonization of the entire world by modern human societies in the last phase of the Pleistocene ice ages.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2014

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Stanley, Steven M., Children of the Ice Age: How a Global Catastrophe Allowed Humans to Evolve (New York, 1996)
Butzer, Karl W., “Environment, Culture, and Human Evolution,” American Scientist 65 (1977), 572–84Google Scholar
Foley, Robert, Another Unique Species: Patterns in Human Evolutionary Ecology (Harlow, UK: 1987)
Mellars, Paul and Stringer, Chris, eds., The Human Revolution: Behavioural and Biological Perspectives on the Origins of Modern Humans (Edinburgh, 1992), 298–318
Stanley, Stephen M., “An Ecological Theory for the Origin of Homo,” Paleobiology 18 (1992), 237–57Google Scholar
Vrba, Elizabeth S. et al., eds. Paleoclimate and Evolution with Emphasis on Human Origins (New Haven, 1995)
deMenocal, Peter B., “Plio-Pleistocene African Climate,” Science 270 (1995), 53–9Google Scholar
Boaz, Noel Thomas, Eco Homo: How the Human Emerged from the Cataclysmic History of the Earth (New York, 1997)
Potts, Richard, Humanity’s Descent: The Consequences of Ecological Instability (New York, 1996)
Calvin, William H., A Brain for all Seasons: Human Evolution & Abrupt Climate Change (Chicago, 2002)
Fagan, Brian M., The Long Summer: How Climate Changed Civilization (New York, 2004)
Burroughs, William J., Climate Change in Prehistory: The End of the Reign of Chaos (New York, 2005)
Hetherington, Renée and Reid, Robert G. B., The Climate Connection: Climate Change and Modern Human Evolution (New York, 2010)
Hamilton, Robert M. et al., (NRC Committee on the Earth System Context of Hominin Evolution), Understanding Climate’s Influence on Human Evolution (Washington, DC, 2010)
deMenocal, Peter B., “Climate and Human Evolution,” Science 331 (2011), 540–2Google Scholar
Campbell, Bernard G. et al., Humankind Emerging, ninth edition (Boston, MA, 2006)
Larsen, Clark S., Our Origins: Discovering Physical Anthropology (New York, 2008)
Landis, G. P. et al., “Pele Hypothesis: Ancient Atmospheres and Geologic-Geochemical Controls of Evolution, Survival and Extinction,” in MacLeod, N. and Keller, G., eds., Cretaceous-Tertiary Mass Extinctions: Biotic and Environmental Changes (New York, 1996), 519–56
Royer, Dana L. et al., “CO2 as a Primary Driver of Phanerozoic Climate,” GSA Today 14/3 (March 2004), 6Google Scholar
Pearson, Paul N. and Palmer, Martin R., “Atmospheric Carbon Dioxide Concentrations over the Past 60 Million Years,” Nature 406 (2000), 695–9Google Scholar
Zachos, James et al., “Trends, Rhythms, and Aberrations in Global Climate 65 Ma to Present,” Science 292 (2001), 686–93Google Scholar
Decanto, Robert M. and Pollard, David, “Rapid Cenozoic Glaciation in Antarctica Induced by Declining Atmospheric CO2,” Nature 421 (2003), 245–9Google Scholar
Tripati, Aradna et al., “Eocene Bipolar Glaciation Associated with Global Carbon Cycle Changes,” Nature 436 (2005); 341–6Google Scholar
Pagani, Mark et al., “Marked Decline in Atmospheric Carbon Dioxide Concentrations during the Paleogene,” Science 309 (2005), 600–3Google Scholar
Hansen, James et al., “Target Atmospheric CO2: Where Should Humanity Aim?The Open Atmospheric Science Journal 2 (2008) 217–31Google Scholar
Ruddiman, William F., Earth’s Climate: Past and Future (New York, 2001), 147–71
van Andel, Tjeerd H., New Views on an Old Planet: A History of Global Change, second edition (New York, 1994), 175–233
Vrba, Elizabeth S. has developed her model of turnover pulses in a series of articles, most prominently “Ecological and Adaptive Changes Associated with Early Hominid Evolution,” in Delsen, E., ed., Ancestors: The Hard Evidence (New York, 1985), 63–71
“Late Pliocene Climate Events and Hominid Evolution,” in Grine, F. E., The Evolutionary History of the “Robust” Australopithecines (New York, 1988), 405–26
Turnover-Pulses, the Red Queen, and Related Topics,” AJS 293-a (1993), 418–52
The Pulse that Produced Us,” Natural History 102/5 (May, 1993), 47–51
Climate, Heterochrony, and Human Evolution,” JAnthRes 52 (1996), 1–28
Mass Turnover and Heterochrony Events in Response to Physical Change,” Paleobiology 31 (2005), 157–74
Hernandez, Manuel and Vrba, Elizabeth S., “Plio-Pleistocene Climatic Change in the Turkana Basin (East Africa): Evidence from Large Mammal Faunas,” JHumEv 50 (2006), 595–626Google Scholar
Frost, S. R., “African Pliocene and Pleistocene Cercopithecid Evolution and Global Climatic Change,” in Bobé, Rene et al., eds. Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence (Dordrect, 2007), 51–77
McKee, Jeffrey K., “Faunal Turnover Rates and Mammalian Biodiversity of the Late Pliocene and Pleistocene of Eastern Africa,” Paleobiology 27 (2001), 500–11Google Scholar
Behrensmeyer, Anna K., “Late Pliocene Faunal Turnover in the Turkana Basin, Kenya and Ethiopia,” Science 278 (1997), 1589–94Google Scholar
Prothero, D. R. and Heaton, T. H., “Faunal Stability during the Early Oligocene Climatic Crash,” PPP 127 (1996), 239–56Google Scholar
Harris, J. M., Kooba Fora Research Project, Vol. II: The Fossil Ungulates: Geology, Fossil Artiodactyls and Paleoenvironments (New York, 1991)
van Dam, Jan, “Long-Period Astronomical Forcing of Mammalian Turnover,” Nature 443 (2006), 687–91Google Scholar
Rohland, Nadin et al., “Genomic DNA Sequences from Mastodon and Woolly Mammoth Reveal Deep Speciation of Forest and Savannah Elephants,” PlosBiology 8 (2010), e1000564Google Scholar
Badgley, Catherine et al., “Ecological Changes in Miocene Mammalian Record Show Impact of Prolonged Climatic Forcing,” PNAS 105 (2008), 12145–9Google Scholar
Raia, P. et al., “Turnover Pulse or Red Queen? Evidence from the Large Mammal Communities during the Plio-Pleistocene of Italy,” PPP 221 (2005), 293–312Google Scholar
Meng, Jin and MacKenna, Malcolm C., “Faunal Turnovers of Palaeogene Mammals from the Mongolian Plateau,” Nature 394 (1998), 364–7Google Scholar
Potts, Richard, “Variability Selection in Hominid Evolution,” Evolutionary Anthropology 7 (1998), 81–96Google Scholar
–, “Environmental Hypotheses of Hominin Evolution,” YPA 41 (1998), 93–136
Behrensmeyer, Anna K., “Climate Change and Human Evolution,” Science 311 (2006), 476–8Google Scholar
Semino, Ornella et al., “The Genetic Legacy of Paleolithic Homo Sapiens sapiens in Extant Europeans: A Y Chromosome Perspective,” Science 290 (2000), 1155–9Google Scholar
Cote, Suzanne M., “Origins of African Hominoids: An Assessment of the Palaeobiogeographical Evidence,” Comptes Rendus Palevol 3 (2004), 323–40Google Scholar
Begun, David R., “Planet of the Apes,” SA 289/2 (August 2003), 74–83Google Scholar
Cameron, David W., Hominid Adaptations and Extinctions (Sydney, 2004), 162–79, 205–10
Zhisheng, An, “Evolution of Asian Monsoons and Phased Uplift of the Himalaya-Tibetan Plateau since Late Miocene Times,” Nature 441 (2001), 62–6Google Scholar
Guo, Z. T. et al., “Onset of Asian Desertification by 22 Myr Inferred from Loess Deposits in China,” Nature 416 (2002), 159–63Google Scholar
Edwards, Erika J. et al., “The Origins of C4 Grasslands: Integrating Evolutionary and Ecosystem Science,” Science 328 (2010), 587–91Google Scholar
de Menocal, Peter B., “African Climate Change and Faunal Evolution during the Pliocene-Pleistocene,” Earth and EPSL 220 (2004), 3–24 (8)Google Scholar
Rouchy, Jean-Marie et al., “The Messinian Salinity Crisis Revisited,” SedGeol 188/189 (2006), 1–8Google Scholar
Michels, Arne et al., “The Late Miocene Climate Response to a Modern Sahara Desert,” GPC 67 (2009), 193–204Google Scholar
Kumar, Sudhir et al., “Placing Confidence Limits on the Molecular Age of the Human-Chimpanzee Divergence,” PNAS 102 (2005), 18842–7Google Scholar
Jobling, M. A. et al., Human Evolutionary Genetics: Origins, Peoples, Diseases (New York, 2004), 215–17
Shi, Jinxiu et al., “Divergence of the Genes on Human Chromosome 21 between Human and other Hominoids and Variation of Substitution Rates among Transcription Units,” PNAS 100 (2003), 8331–6Google Scholar
Stauffer, R. L. et al., “Human and Ape Molecular Clocks and Constraints on Paleontological Hypotheses,” Journal of Heredity 92 (2001), 469–74Google Scholar
Coppen, Yves, “East Side Story: The Origin of Humankind,” SA 270/5 (May 1994). 88–95Google Scholar
Richmond, Brian G. and Junger, William L., “Orrin tugenensis Femoral Morphology and the Evolution of Human Bipedalism,” Science 319 (2008), 1662–5Google Scholar
White, Tim D., “Ardipithecus ramidus and the Peliobiology of Early Hominids,” Science 326 (2009), 75–86Google Scholar
Heile-Selassie, Yohannes et al., “A New Hominin Foot from Ethiopia Shows Multiple Pliocene Bipedal Adaptations,” Nature 483 (2012), 565–70Google Scholar
Patterson, Nick et al., “Genetic Evidence for Complex Speciation of Humans and Chimpanzees,” Nature 441 (2006), 1103–8Google Scholar
White, Tim D. et al., “Asa Issie, Aramis, and the Origin of Australopithecus,” Nature 330 (2006), 883–9Google Scholar
Bobé, René and Behrensmeyer, Anna K., “The Expansion of Grassland Ecosystems in Africa in Relation to Mammalian Evolution and the Origin of the Genus Homo,” PPP 207 (2004), 399–420Google Scholar
Bartoli, G., “Final Closure of Panama and the Onset of Northern Hemisphere Glaciation,” EPSL 237 (2005), 33–44Google Scholar
Cane, Mark A. and Molnar, Peter, “Closing of the Indonesian Seaway as a Precursor to East African Aridification around 3–4 million Years Ago,” Nature 411 (2001), 157–62Google Scholar
Bell, Martin and Walker, Michael J. C., Late Quaternary Environmental Change: Physical & Human Perspectives (Harlow and New York, 1992), 60–3
Hays, J. D. et al., “Variations in the Earth’s Orbit: Pacemaker of the Ice Ages,” Science 194 (1976), 1121–32Google Scholar
Zachos, James C. et al., “Climate Response to Orbital Forcing across the Oligocene-Miocene Boundary,” Science 292 (2001), 274–8Google Scholar
Westerhold, T. et al., “Middle to Late Oxygen Isotope Stratigraphy of ODP Site 1085 (SE Atlantic): New Constraints of Miocene Climate Variability and Sea-Level Fluctuations,” PPP 217 (2005), 205–22Google Scholar
Sun, Xiangjun and Wang, Pinxian, “How Old Is the Asian Monsoon System? – Palaeobotanical Records from China,” PPP 222 (2005), 181–222Google Scholar
Reed, Kaye E., “Paleoecological Patterns at the Hadar Hominin Site, Afar Regional State, Ethiopia,” JHumEv 54 (2008), 743–68Google Scholar
Collard, Mark, “Grades and Transitions in Human Evolution,” Proceedings of the British Academy 106 (2002), 61–100Google Scholar
Raymo, Maureen E., “The Initiation of Northern Hemispheric Glaciation,” Annual Reviews in Earth and Planetary Science 22 (1994), 353–83Google Scholar
The 41 kyr World: Milankovitch’s Other Unsolved Mystery,” Paleoceanography 18 (2003), 1011–16
Ruddiman, William F., “Orbital Insolation, Ice Volume, and Greenhouse Gases,” QSR 22 (2003), 1597–629, at 1624Google Scholar
Seki, Osamu et al., “Alkenone and Boron-Based Pliocene pCO2 Records,” EPSL 292 (2010), 201–11Google Scholar
Pagani, Mark et al., “High Earth-System Climate Sensitivity Determined from Pliocene Carbon Dioxide Concentrations,” NatGeosc 3 (2010), 27–30Google Scholar
Ravelo, Ana Christina, “Warmth and Glaciation,” NatGeosc 3 (2010), 672–4Google Scholar
Ruddiman, William F., “A Paleoclimatic Enigma?Science 328 (2010), 838–9Google Scholar
Wood, Bernard and Strait, David, “Patterns of Resource Use in Early Homo and Paranthropus,” JHumEv 46 (2004), 119–62Google Scholar
Feakins, Sarah J. et al., “Biomarker Records of Late Neogene Changes in Northeast African Vegetation,” Geology 33 (2005), 977–80Google Scholar
Wynn, Jonathan Guy, “Influence of Plio-Pleistocene Aridification on Human Evolution: Evidence from Paleosols of the Turkana Basin, Kenya,” AJPA 123 (2004), 106–18Google Scholar
Levin, Naomi E. et al., “Isotopic Evidence for Plio-Pleistocene Environmental Change at Gona, Ethiopia,” EPSL 219 (2004), 93–110Google Scholar
Cerling, Thure E., “Global Vegetation Change through the Miocene/Pliocene Boundary,” Nature 389 (1997), 153–8Google Scholar
Trauth, Martin H. et al., “Trends, Rhythms, and Events in Plio-Pleistocene African Climate,” QSR 28 (2009), 399–411Google Scholar
High- and Low-Latitude Forcing of Plio-Pleistocene East African Climate and Human Evolution,” JHumEv 53 (2007), 475–86
Kingston, John D. et al., “Astronomically Forced Climate Change in the Kenyan Rift Valley 2.7–2.55 Ma.: Implications for the Evolution of Early Hominin Ecosystems,” JHumEv 53 (2007), 487–503Google Scholar
Campisano, Christopher J. and Feibel, Craig S., “Connecting Local Environmental Sequences to Global Climate Patterns: Evidence from the Hominin-Bearing Hadar Formation, Ethiopia,” JHumEv 53 (2007), 515–27Google Scholar
Plummer, Thomas, “Flaked Stones and Old Bones: Biological and Cultural Evolution at the Dawn of Technology,” YPA 47 (2004), 118–64Google Scholar
Semaw, Sileshi et al., “2.6 Million-Year-Old Stone Tools and Associated Bones from OGS-6 and OGS-7, Gona, Afar, Ethiopia,” JHumEv 43 (2003), 169–77Google Scholar
Sussman, Randall L., “Hand Function and Tool Behavior in Early Hominids,” JHumEv 35 (1998), 23–46Google Scholar
Kimbel, W. H. et al., “Late Pliocene Homo and Oldowan Tools from the Hadar Formation (Kada Hadar Member), Ethiopia,” JHumEv 31 (1996), 549–61Google Scholar
Tattersall, Ian and Schwartz, Jeffrey H., “Evolution of the Genus Homo,” Annual Reviews: Earth and Planetary Science 37 (2009), 67–92Google Scholar
Berger, Lee R. et al., Australopithecus Sediba: A New Species of Homo-like Australopith from South Africa,” Science 328 (2010), 195–204Google Scholar
Stedman, Hansell H., “Myosin Gene Mutation Correlates with Anatomical Changes in the Human Lineage,” Nature 428 (2004), 415–18Google Scholar
Currie, Pete, “Muscling In on Hominid Evolution,” Nature 428 (2004), 373–4Google Scholar
Chou, Hsun-Hua et al., “Inactivation of CMP-N-Acetylneuraminic Acid Hydroxylase Occurred Prior to Brain Expansion during Human Evolution,” PNAS 99 (2002), 11636–41Google Scholar
Aiello, Leslie C. and Wheeler, Peter, “The Expensive-Tissue Hypothesis: The Brain and Digestive System in Human and Primate Evolution, CA (36 (1995), 199–221Google Scholar
Rose, Lisa and Marshall, Fiona, “Meat-Eating, Hominid Sociality, and Home Bases Revisited,” CA 37 (1996), 307–38Google Scholar
Blumenschine, Robert J. and Cavallo, John A., “Scavenging and Human Evolution,” SA 267/4 (Oct. 1992), 90–6Google Scholar
Falk, Dean, “Brain Evolution in Homo: The ‘Radiator’ Theory,” Behavioral and Brain Sciences 13 (1990), 333–81Google Scholar
Isaacs, Glynn “Aspects of Human Evolution,” in Bendall, D. S., ed. Evolution from Molecules to Men (New York, 1983), 509–43
Isaac, Glynn, “The Food-Sharing Behavior of Protohuman Hominids,” SA 238/4 (1978), 90–138Google Scholar
Gabunia, Leo et al., “Dmanisi and Dispersal,” Evolutionary Anthropology 10 (2001), 158–70Google Scholar
Lordkipanidze, David et al., “Postcranial Evidence from Early Homo from Dmanisi, Georgia,” Nature 449 (2007), 305–9Google Scholar
Spoor, F. et al., “Implications of New Early Homo Fossils from Ileret, East of Lake Turkana, Kenya,” Nature 448 (2007), 688–90Google Scholar
Lieberman, Daniel E., “Homing In on Early Homo,” Nature 339 (2007), 291–2Google Scholar
Lordkipanidze, David et al., “A Complete Skull from Demanisi, Georgia, and the Evolutionary Biology of Early Homo,” Nature 342 (2013), 326–31Google Scholar
Ashley, Gail M. et al., “Hominin Use of Springs and Wetlands: Paleoclimate and Archaeological Records from Olduvai Gorge (~1.79–1.74 Ma), PPP 272 (2009), 1–16Google Scholar
Owen, R. Bernart et al., “Diatomaceous Sediments and Environmental Change in the Pleistocene Olorgesailie Formation, Southern Kenyan Rift Valley,” PPP 269 (2008), 17–37Google Scholar
Ashley, Gail M., “Orbital Rhythms, Monsoons, and Playa Lake Response, Olduvai Basin, Equatorial East Africa (ca. 1.85–1.74 Ma),” Geology 35 (2007), 1091–4Google Scholar
Bobé, René et al., “Faunal Change, Environmental Variability and Late Pliocene Hominin Evolution,” JHumEv 42 (2002), 475–97Google Scholar
Bennett, Matthew R. et al., “Early Hominin Foot Morphology Based in 1.5-Million-Year-Old Footprints from Ileret, Kenya,” Science 323 (2009), 1197–201Google Scholar
Bramble, Dennis and Lieberman, Daniel E., “Endurance Running and the Evolution of Homo,” Nature 432 (2004), 345–52Google Scholar
Spoor, Fred and Wood, Bernard, “Implications of Early Hominid Labyrinthine Morphology for Evolution of Human Bipedal Locomotion,” Nature 369 (1994), 645–8Google Scholar
Anton, Susan C., “Natural History of Homo Erectus,” YPA 46 (2003), 126–70Google Scholar
Anton, Susan C. and Swisher, Carl C., III, “Early Dispersals of Homo from Africa,” Annual Reviews of Anthropology 33 (2004), 271–96Google Scholar
Goren-Inbar, Naama et al., “Pleistocene Milestones on the Out-of-Africa Corridor at Gesher Benot Ya’aqov, Israel,” Science 289 (2000), 944–7Google Scholar
Foley, Robert and Lahr, Marta Mirazón, “Mode 3 Technologies and the Evolution of Modern Humans,” CArchJ 7 (1997), 12Google Scholar
Rogers, Alan R. et al., “Genetic Variation at the MCIR Locus and the Time since Loss of Human Body Hair,” CA 45 (2004), 105–8Google Scholar
Jablonski, Nina G., “The Evolution of Human Skin and Skin Color,” Annual Reviews in Anthropology 3 (2004), 585–623Google Scholar
Jablonski, Nina G. and Chapin, George, “The Evolution of Human Skin Coloration,” JHumEv 39 (2000), 57–106Google Scholar
Wrangham, Richard, Catching Fire: How Cooking Made Us Human (New York, 2009)
Berna, Francesco et al., “Microstratigraphic Evidence of In Situ Fire in the Acheulean Strata of Wonderwerk Cave, Northern Cape Province, South Africa,” PNAS April 2, 2012Google Scholar
Goren-Inbar, Naama, “Evidence of Hominin Control of Fire at Gesher Benot Ya’aqov, Israel,” Science 304 (2004), 725–7Google Scholar
Alperson-Afil, Nira et al., “Spacial Organization of Hominin Activities at Gesher Benot Ya’aqov, Israel,” Science 326 (2009), 1677–80Google Scholar
Hawkes, Kristen, “Grandmothers and the Evolution of Human Longevity,” American Journal of Human Biology 15 (2003), 380–400Google Scholar
Brown, P. et al., “A New Small-Bodied Hominin from the Late Pleistocene of Flores, Indonesia,” Nature 431 (2004), 1055–61Google Scholar
Morwood, M. J. et al., “Archaeology and Age of a New Hominin from Flores in Eastern Indonesia,” Nature 431 (2004), 1087–91Google Scholar
Morwood, M. J. et al., “Further Evidence for Small-Bodied Hominins from the Late Pleistocene of Flores, Indonesia,” Nature 437 (2005), 1012–17Google Scholar
Wong, Kate, “The Littlest Human,” SA 292/2 (Feb. 2005), 56–65Google Scholar
Siddell, Mark et al., “Changes in Deep Pacific Temperature during the Mid-Pleistocene Transition and Quaternary,” QSR 29 (2010), 170–81Google Scholar
McClymont, Erin L. and Rosell-Melé, Antoni, “Links between the Onset of Modern Walker Circulation and the Mid-Pleistocene Climate Transition,” Geology 33 (2005), 389–92Google Scholar
Medina-Elizalde, Martin and Lea, David W., “The Mid-Pleistocene Transition in the Tropical Pacific,” Science 310 (2005), 1009–12Google Scholar
Raymo, M. E., “The Mid-Pleistocene Climate Transition: A Deep Sea Carbon Isotopic Perspective,” Paleoceanography 12 (1997), 546–59Google Scholar
Mudelsee, Manfred and Schultz, Michael, “The Mid-Pleistocene Climate Transition: Onset of 100 ka Cycle Lags Ice Volume Build-Up by 280 ka,” EPSL 151 (1997), 117–23Google Scholar
Shackleton, Nicholas J., “The 100,000-Year Ice-Age Cycle Identified and Found to Lag Temperature, Carbon Dioxide, and Orbital Eccentricity,” Science 289 (2000), 1897–902Google Scholar
Ruddiman, William F. and Raymo, Maureen E., “A Methane-Based Time Scale for Vostok Ice,” QSR 22 (2003), 141–55Google Scholar
Bergner, A. G. N. et al., “Tectonic and Climatic Control on Evolution of Rift Lakes in the Central Kenya Rift, East Africa,” QSR 28 (2009), 2804–16Google Scholar
Scholz, Christopher A. et al., “East African Megadroughts between 135 and 75 Thousand Years Ago and Bearing on Early-Modern Human Origins,” PNAS 104 (2007), 16416–21Google Scholar
Cohen, Andrew S., “Ecological Consequences of Early Late Pleistocene Megadroughts in Tropical Africa,” PNAS 104 (2007), 16422–7Google Scholar
Trauth, Martin H. et al., “East African Climate Change and Orbital Forcing during the Last 175 kyr BP,” EPSL 206 (2003), 297–313Google Scholar
Abbate, Ernesto, “A One-Million-Year-Old Homo Cranium from the Danakil (Afar) Depression of Eritrea,” Nature 393 (1998), 458–9Google Scholar
Asfaw, Berhane et al., “Remains of Homo Erectus from Bouri, Middle Awash, Ethiopia,” Nature 416 (2002), 317–20Google Scholar
Carbonell, Edward et al., “The First Hominin in Europe,” Nature 452 465–9
Manzi, Giorgio, “Human Evolution at the Matuyama-Brunhes Boundary,” Evolutionary Anthropology 13 (2004), 11–24Google Scholar
de Castro, J. M. Bermúdez, “The Atapuerca Sites and Their Contribution to the Knowledge of Human Evolution in Europe,” Evolutionary Anthropology 13 (2004), 25–41Google Scholar
Dennell, Robin, “Dispersal and Colonization, Long and Short Chronologies: How Continuous Is the Early Pleistocene Record for Hominids outside of East Africa,” JHumEv 45 (2003), 421–40Google Scholar
Rightmire, G. Philip, “‘Human Evolution in the Middle Pleistocene’: The Role of Homo Heidelbergensis,” Evolutionary Anthropology 6 (1998), 218–27Google Scholar
Finlayson, Clive, “Biogeography and Evolution of the Genus Homo,” Trends in Ecology and Evolution 20 (2005), 457–63Google Scholar
Rightmire, G. Philip, “Brain Size and Encephalization in Early to Mid-Pleistocene Homo,”AJPA 124 (2004), 109–23Google Scholar
Marwick, Ben, “Pleistocene Exchange Networks as Evidence for the Evolution of Language,” CArchJ 13 (2003), 67–81Google Scholar
Lahr, Marta Mirazón and Foley, Robert A., “Toward a Theory of Modern Human Origins: Geography, Demography, and Diversity in Recent Human Evolution,” YPA 41 (1998), 137–76Google Scholar
Parfitt, Simon A. et al., “Early Pleistocene Human Occupation at the Edges of the Boreal Zone in Northwest Europe,” Nature 466 (2010), 229–33Google Scholar
–, “The Earliest Record of Human Activity in Northern Europe,” Nature 438 (2005), 1008–12
Dennell, , “Dispersal and Colonization”; Finlayson, “Biogeography and Evolution.” Clive Gamble, in Timewalkers: The Prehistory of Global Colonization (Cambridge, MA, 1994), 135–6
Mounier, Aurélien et al., “Is Homo Heidelbergensis a Distinct People? New Insight on the Mauer Mandible,” JHumEv 56 (2009), 219–46Google Scholar
Fagan, Brian M., The Journey from Eden: The Peopling of Our World (London, 1990)
Calvin, William H., The Ascent of Mind: Ice Age Climates and the Evolution of Intelligence (New York, 1991)
Peer, Philip Van, “The Nile Corridor and the Out-of-Africa Model: An Examination of the Archaeological Record,” CA 39 (1998), S115–40Google Scholar
Green, Richard E. et al., “Analysis of One Million Base Pairs of Neanderthal DNA,” Nature 444 (2006), 330–6Google Scholar
Bischoff, James L. et al., “The Sima de los Huesos Hominids Date to beyond U/Th Equilibrium (>350 kyr) and Perhaps to 400–500 kyr,” JArchS 30 (2003), 275–80Google Scholar
Cann, R. L. et al., “Mitochondrial DNA and Human Evolution,” Nature 325 (1987), 31–6Google Scholar
Campbell, Michael C. and Tishkoff, Sarah A, “The Evolution of Human Genetic and Phenotypic Variation in Africa,” Current Biology 20 (2010), R166–R173Google Scholar
DeGiorgio, Michael et al., “Explaining Worldwide Patterns of Human Genetic Variation Using a Coalescent-Based Serial Founder Model F Migration Outward from Africa,” PNAS 106 (2009), 16057–62Google Scholar
Weaver, Timothy D. and Roseman, Charles C., “New Developments in the Genetic Evidence for Modern Human Origins,” Evolutionary Anthropology 17 (2008), 69–80Google Scholar
Relethford, J. H., “Genetic Evidence and the Modern Human Origins Debate,” Heredity 100 (2008), 555–63Google Scholar
Mellars, Paul, “The Impossible Coincidence: A Single-Species Model for the Origins of Modern Human Behavior in Europe,” Evolutionary Anthropology 14 (2005), 12–27Google Scholar
Pearson, Osbjorn M., “Has the Combination of Genetic and Fossil Evidence Solved the Riddle of Modern Human Origins?Evolutionary Anthropology 13 (2004), 145–59Google Scholar
Stringer, Chris, “Modern Human Origins: Progress and Prospects,” PTRS,LB 357 (2002), 563–79Google Scholar
Relethford, John H., Genetics and the Search for Modern Human Origins (New York, 2001)
Smith, Shelley L. and Harrold, Francis B., “A Paradigm’s Worth of Difference? Understanding the Impasse over Modern Human Origins,” YPA 40 (1997), 113–38Google Scholar
Grine, F. E., “Late Pleistocene Human Skull from Hofmeyr, South Africa, and Modern Human Origins,” Science 315 (2007), 226–9Google Scholar
Atkinson, Quentin D., “Phonemic Diversity Supports a Series Founder Effect Model of Language Expansion from Africa,” Science 332 (2011), 346–9Google Scholar
Klein, Richard G. with Edgar, Blake, The Dawn of Human Culture (New York, 2002)
Bar-Yosef, Ofer, “The Upper Paleolithic Revolution,” Annual Reviews in Anthropology 31 (2002), 363–93Google Scholar
Clark, Grahame, World Prehistory: A New Perspective (Cambridge, UK, 1977), 32–8
McBrearty, Sally and Brooks, Alison M., “The Revolution that Wasn’t: A New Interpretation of the Origin of Modern Human Behavior,” JHumEv 39 (2000), 494–7Google Scholar
Foley, and Lahr, Marta Mirazón, “On Stony Ground: Lithic Technology, Human Evolution, and the Emergence of Culture,” Evolutionary Anthropology 12 (2003), 109–22Google Scholar
Marean, Curtis W., “The Origin of Modern Human Behavior: Critique of the Models and Their Test Implications,” CA 44 (2003), 627–51Google Scholar
Barham, Lawrence and Robson-Brown, Kate, eds., Human Roots: Africa and Asia in the Middle Pleistocene (Bristol, 2001)
Clark, J. Desmond, in “The Middle Stone Age in East Africa and the Beginnings of Regional Identity,” JWP 2 (1998), 235–305Google Scholar
Ambrose, Stanley H., “Chronology of the Late Stone Age and Food Production in East Africa,” JArchS 25 (1998), 377–92Google Scholar
White, Tim et al., “Pleistocene Homo Sapiens from Middle Awash, Ethiopia,” Nature 423 (2003), 742–7 (Herto)Google Scholar
McDougall, Ian et al., “Stratigraphic Placement and Age of Modern Humans from Kibish, Ethiopia,” Nature 433 (2005), 733–6 (Omo)Google Scholar
Trinkaus, Erik, “Early Modern Humans,” Annual Reviews in Anthropology 34 (2005), 207–30Google Scholar
Karafet, T. M. et al., “Ancestral Asian Source(s) of New World Y-Chromosome Founder Haplotypes,” American Journal of Genetics 1999 (64), 817–31 (825)Google Scholar
Hammer, M. F. et al., “Out of Africa and Back Again: Nested Cladistic Analysis of Human Y Chromosome Variation,” Molecular Biology and Evolution 15 (1998), 427–41 (434)Google Scholar
Lieberman, Daniel E. et al., “The Evolution and Development of Cranial Form in Homo Sapiens,” PNAS 99 (2002), 1134–9Google Scholar
Lieberman, Philip, “On the Nature and Evolution of the Neural Bases of Human Language,” YPA 45 (2002), 36–62Google Scholar
Mithen, Stephen, Prehistory of the Mind: A Search for the Origins of Art, Religion and Science (London, 1996)
Bruner, Emiliano, “Morphological Differences in the Parietal Lobes within the Human Genus,” CA 51, Supplement 1 (2010), s77–s88Google Scholar
Dean, Christopher et al., “Growth Processes in Teeth Distinguish Modern Humans from Homo Erectus and Earlier Hominins,” Nature 414 (2001), 628–31Google Scholar
Coolidge, Frederick L. and Wynn, Thomas, “Executive Functions of the Frontal Lobes and the Evolutionary Ascendency of Homo Sapiens,” CArchJ 11 (2001), 255–60Google Scholar
Wyn, Thomas and Coolidge, Frederick L., “Beyond Symbolism and Language: Introduction to Supplement 1, Working Memory,” CA 51 Supplement 1 (2010), S5–S16Google Scholar
Rossano, Matt J., “How Our Ancestors Raised Children to Think as Modern Humans,” Biological Theory 5 (2010), 142–53Google Scholar
Crow, Timothy J., “The ‘Big Bang’ Theory of the Origin of Psychosis and the Faculty of Language,” Schizophrenia Research 102 (2008), 31–52Google Scholar
Crow, Timothy J., “Schizophrenia as the Price that Homo Sapiens Pay for Language : A Resolution of the Central Paradox in the Origin of the Species,” Brain Research Review 31(2000), 118–29Google Scholar
Crow, Timothy J., ed., The Speciation of Modern Homo Sapiens (Oxford, 2002)
Enard, Wolfgang et al., “Molecular Evolution of FOXP2, a Gene Involved in Speech and Language,” Nature 418 (2002), 869–72Google Scholar
Zhang, Jianshi et al., “Accelerated Protein Evolution and Origins of Human-Specific Features: FOXP2 as an Example,” Genetics 162 (2002), 1825–35Google Scholar
Kraus, Johannes et al., “The Derived FOXP2 Variant of Modern Humans was Shared with Neanderthals,” Current Biology 17 (2007), 1908–12Google Scholar
Coop, Graham et al., “The Timing of Selection at the Human FOXP2 Gene,” MBE 25 (2008), 1257–9Google Scholar
Maricic, Tomislav et al., “A Recent Evolutionary Change Affects a Regulatory Element in the the FOXP2 Gene,” MBE 30 (2013), 844–52Google Scholar
Corballis, Michael C., “Mirror Neurons and the Evolution of Language,” Brain & Language 112 (2010), 25–35Google Scholar
Corballis, Michael C., “The Evolution of Language,” Annals of the New York Academy of Sciences 1156 (2009), 19–43Google Scholar
Lahr, Marta Mirazón and Foley, Robert, “Multiple Dispersals and Modern Human Origins,” Evolutionary Anthropology 3 (1994), 48–60Google Scholar
Forster, Peter, “Ice Ages and the Mitochondrial DNA Chronology of Human Dispersals: A Review,” PTRS,LB 359 (2004), 255–64Google Scholar
Macaulay, Vincent et al., “Single, Rapid Coastal Settlement of Asia Revealed by Analysis of Complete Mitochondrial Genomes,” Science 308 (2005), 1034–6Google Scholar
Thangaraj, Kumarasamy, “Reconstructing the Origin of the Andaman Islanders,” Science 308 (2005), 996Google Scholar
Williams, Martin A. J. et al., “Environmental Impact of the 73 ka Toba Super-Eruption in South Asia,” PPP 284 (2009), 295–314Google Scholar
Ambrose, Stanley H., “Late Pleistocene Human Population Bottlenecks, Volcanic Winter, and Differentiation of Modern Humans,” JHumEv 34 (1998), 623–51Google Scholar
Rampino, Michael R. and Ambrose, Stanley H., “Volcanic Winter in the Garden of Eden: The Toba Superexplosion and the Late Pleistocene Human Population Crash,” in McCoy, F. W. and Heiken, G., eds., Volcanic Hazards and Disasters in Human Antiquity: GSA Special Paper 345 (Boulder, CO, 2000), 71–82
Oppenheimer, Clive, “Limited Global Change to Largest Known Quaternary Eruption, Toba ~ 74kyr BP?QSR 21 (2002), 1593–609Google Scholar
Gathorne-Hardy, F. J., Harcourt-Smith, W. E. H., and Ambrose, Stanley, JHumEv 45 (2003), 227–37
Mellars, Paul, “Going East: New Genetic and Archaeological Perspectives on the Modern Human Colonization of Eurasia,” Science 313 (2006), 796–800Google Scholar
Wells, Spencer, The Journey of Man: A Genetic Odyssey (Princeton, NJ, 2002), 178
Watson, Elizabeth et al., “Mitochondrial Footprints on Human Expansions in Africa,” AJHG 61 (1997), 691–704Google Scholar
Salas, Antonio et al., “The Making of the African mtDNA Landscape,” AJHG 71 (2002), 1082–111Google Scholar
Mishmar, Dan et al., “Natural Selection Shaped Regional mtDNA Variation in Humans,” PNAS 100 (2003), 171–6Google Scholar
Underhill, Peter A. et al., “Y Chromosome Sequence Variation and the History of Human Populations,” NatGen 26 (2000), 358–60Google Scholar
Hammer, Michael F. et al., “Hierarchical Patterns of Global Human Y-Chromosome Diversity,” MBE 18 (2001), 1189–203Google Scholar
Underhill, Peter A. et al., “The Phylogeography of Y Chromosome Binary Haplotypes and the Origins of Modern Human Populations,” Annals of Human Genetics 65 (2001), 43–62Google Scholar
Hammer, Michael F. and Zegura, Stephen L., “The Human Y Chromosome Haplogroup Tree: Nomenclature and Phylogeography of Its Major Divisions,” ARA 31 (2002), 303–21 (314)Google Scholar
Jobling, Mark A. and Tyler-Smith, Chris, “The Human Y Chromosome: An Evolutionary Marker Comes of Age,” NatGen 4 (2003), 598–610Google Scholar
Underhill, Peter A. and Kivisilk, Toomas, “Use of Y Chromosome and Mitochondrial DNA Population Structure in Tracing Human Migrations,” Annual Review of Genetics 41 (2007), 539–64Google Scholar
D’Errico, Francesco et al., “Nassarius Kraussianus Shell Beads from Blombos Cave: Evidence for Symbolic Behavior in the Middle Stone Age,” JHumEv 48 (2005), 3–24Google Scholar
Barham, Lawrence S., “Backed Tools in Middle Pleistocene Central Africa and Their Evolutionary Significance,” JHumEv 43 (2002), 585–603Google Scholar
Systematic Pigment Use in the Middle Pleistocene of South-Central Africa,” CA 43 (2002), 181–90
Yellen, John E., “Barbed Bone Points: Tradition and Continuity in Saharan and Sub-Saharan Africa,” African Archaeological Review 15 (1998), 173–98Google Scholar
Barton, R. N. E. et al., “OSL Dating of the Aterian Levels at Dar es-Soltan I (Rabat, Morocco) and Implications for the Dispersal of Modern Homo Sapiens,” QSR 28 (2009), 1914–31Google Scholar
Castaneda, Isla S. et al., “Wet Phases in the Sahara/Sahel Region and Human Migration Patterns in North Africa,” PNAS 106 (2009), 20159–63Google Scholar
Osborne, Anne H. et al., “A Humid Corridor across the Sahara for the Migration of Early Modern Humans out of Africa 120,000 Years Ago,” PNAS 105 (2008), 16444–7Google Scholar
Bouzouggar, Abdeljalil, “82,000-Year-Old Shell Beads from North Africa and Implications for the Origins of Modern Human Behavior,” PNAS 104 (2007), 9964–9Google Scholar
Garcea, Elena A. A., “Crossing Deserts and Avoiding Seas: Aterian North African-European Relations,” JAnthRes 60 (2004), 27–53Google Scholar
Walter, Robert C. et al., “Early Human Occupation of the Red Sea Coast of Eritrea during the Last Interglacial,” Nature 405 (2000), 65–9Google Scholar
Armitage, Simon J. et al., “The Southern Route ‘Out of Africa’: Evidence for an Early Expansion of Modern Human into Arabia,” Science 331 (2011), 453–6Google Scholar
Rose, Jeffrey I. et al., “The Nubian Complex of Dhofar, Oman: An African Middle Stone Age Industry in Southern Arabia,” Plos ONE 6 (2011), 28239Google Scholar
Rose, Jeffrey I., “The Question of Upper Pleistocene Connections between East Africa and South Arabia,” CA 45 (2004), 551–5Google Scholar
Bar-Yosef, Ofer, “The Role of Western Asia in Modern Human Origins,” PTRS,LB 337 (1992), 193–200Google Scholar
Holliday, Trenton W., “Evolution at the Crossroads: Modern Human Emergence in Western Asia,” AmAnth 102 (2000), 54–68Google Scholar
Strasser, Thomas F. et al., “Stone Age Seafaring in the Mediterranean: Evidence from the Plakias Region for Lower Palaeolithic and Mesolithic Habitation of Crete,” Hesperia 79 (2010), 145–90Google Scholar
Petraglia, Michael, “Middle Paleolithic Assemblages from the Indian Subcontinent before and after the Toba Super-Eruption, Science 317 (2007), 114–16Google Scholar
Haslam, Michael et al., “The 74 ka Toba Super-Eruption and Southern Indian Hominins: Archaeology, Lithic Technology, and Environments at Jwalapuram Locality 3,” JArchS 37 (2010), 3370–84Google Scholar
Petraglia, Michael D. et al., “Out of Africa: New Hypotheses and Evidence for the Dispersal of Homo Sapiens along the Indian Ocean Rim,” Annals of Human Biology 37 (2010), 288–311Google Scholar
Oppenheimer, Stephen, “The Great Arc of Dispersal of Modern Human: Africa to Australia,” QuatInt 202 (2009), 2–13Google Scholar
Field, Julia S. et al., “The Southern Dispersal Hypothesis and the South Asia Archaeological Record: Examinations of Dispersal Routes through GIS Analysis,” JAnthArch 26 (2007), 88–108Google Scholar
Goebel, Ted, “The Missing Years for Modern Humans,” Science 315 (2007), 194–6Google Scholar
James, Hannah V. A. and Petraglia, Michael D., “Modern Human Origins and the Evolution of Behavior in the Later Pleistocene Record of South Asia,” CA 46 (2005), S3–S27Google Scholar
Endicott, Phillip et al., “Genetic Evidence on Modern Human Dispersals in South Asia: Y Chromosome and Mitochondrial DNA Perspectives: The World through the Eye of Two Haploid Genomes,” in Petraglia, Michael D. and Allchin, Bridget, eds., The Evolution and History of Human Populations in South Asia (Dordrecht, 2007), 229–44
Rose, Jeffrey I., “New Light on Human Prehistory in the Arabo-Persian Gulf Oasis,” CA 51 (2010), 849–83Google Scholar
Beyin, Amanuel, “Upper Pliestocene Human Dispersal out of Africa,” A Review of the Current State of the Debate,” International Journal of Evolutionary Biology (2011), 615–94Google Scholar
Anikovich, M. V. et al., “Early Upper Paleolithic in Eastern Europe and Implications for the Dispersal of Modern Humans,” Science 317 (2007), 223–5Google Scholar
Olivieri, Anna et al., “The mtDNA Legacy of the Levantine Early Upper Paleolithic in Africa,” Science 314 (2006), 1767–70Google Scholar
Endicott, Phillip et al., “Evaluating the Mitochondrial Timescale of Human Evolution,” Cell: Trends in Ecology and Evolution 24 (2009), 515–21Google Scholar
Smith, Fred H. et al., “Modern Human Origins,” YPA 32 (1989), 35–68Google Scholar
Harding, Rosalind M., “Archaic African and Asian Lineages in the Genetic Ancestry of Modern Human,” AJHG 60 (1997), 770–89Google Scholar
Hawks, John et al., “Population Bottlenecks and Pleistocene Human Evolution,” MBE 17 (2000), 2–22Google Scholar
Wolpoff, Milford H., “Modern Human Ancestry at the Peripheries: A Test of the Replacement Theory,” Science 291 (2001), 293–7Google Scholar
Templeton, Alan, “Out of Africa Again and Again,” Nature 416 (2002), 45–51Google Scholar
Harding, Rosalind M. and McVean, Gil, “A Structured Ancestral Population for the Evolution of Modern Humans,” Current Opinion in Genetics & Development 14 (2004) 667–74CrossRefGoogle Scholar
Hammer, Michael F. et al., “Heterogeneous Patterns of Variation among Multiple Human X-Linked Loci: The Possible Role of Diversity-Reducing Selection in Non-Africans,” Genetics 167 (2004), 1841–53Google Scholar
Eswaran, Vinayak et al., “Genomics Refutes an Exclusively African Origin of Humans,” JHumEv 49 (2005), 1–18Google Scholar
Bräuer, Günter et al., “On the Reliability of Recent Tests…,” The Anatomical Record Part A 279A (2004), 701–7Google Scholar
Manica, Andrea et al., “The Effect of Ancient Population Bottlenecks on Human Phenotypic Variation,” Nature 448 (2007), 346–8Google Scholar
Laval, Guillaume et al., “Formulating a Historical and Demographic Model of Recent Human Evolution Based on Resequencing Data from Noncoding Regions, Plos ONE 5 (2010), e10284Google Scholar
Green, Richard E. et al., “A Draft Sequence of the Neanderthal Genome,” Science 328 (2010), 710–22Google Scholar
Caramelli, David et al., “28,000 Years Old Cro-Magnon mtDNA Sequence Differs from All Potentially Contaminating Modern Sequences,” Plos ONE 7 (2008), e2700Google Scholar
Chu, J. Y., “Genetic Relationships of Populations in China,” PNAS 95 (1998), 11763–8Google Scholar
Ke, Yuehai et al., “African Origin of Modern Humans in East Asia: A Tale of 12,000 Y Chromosomes,” Science 292 (2001), 1151–3Google Scholar
Reed, David L. et al., “Genetic Analysis of Lice Supports Direct Contact between Modern and Archaic Humans,” PLoS Biology 2 (2004), e340Google Scholar
Reich, David, “Genetic History of an Archaic Hominin Group from Denisova Cave in Siberia,” Nature 468 (2010), 1053–60Google Scholar
Abi-Rached, Laurent et al., “The Shaping of Modern Human Immune Systems by Multiregional Admixture with Archaic Humans,” Science 334 (2011), 89–94Google Scholar
Liu, Wu et al., “Human Remains from Zhirendong, South China, and Modern Human Emergence in East Asia,” PNAS 107 (2010), 19201–6Google Scholar
Harpending, Henry C. et al., “The Genetic Structure of Ancient Human Populations,” CA 34 (1993), 483–96Google Scholar
Sherry, Stephen T. et al., “Mismatch Distributions of mtDNA Reveal Recent Human Populations Expansions,” Human Biology 66 (1994), 761–75Google Scholar
Rogers, Alan R. and Jorde, Lynn B., “Genetic Evidence on Modern Human Origins,” Human Biology 67 (1995), 1–36Google Scholar
Harpending, Henry C. et al., “Genetic Traces of Ancient Demography,” PNAS 95 (1998), 1961–7Google Scholar
Eswaran, Vinayak, “A Diffusion Wave out of Africa: The Mechanism of the Modern Human Revolution?CA 43 (2002), 749–64Google Scholar
Marth, Gabor T. et al., “The Allele Frequency Spectrum in Genome-Wide Human Variation Data Reveals Signals of Differential Demographic History in Three Large World Populations,” Genetics 166 (2004), 351–72Google Scholar
Zhivotovsky, Lev A. et al., “Features of Evolution and Expansion of Modern Humans, Inferred from Genomewide Microsatellite Markers,” AJHG 72 (2003), 1171–86Google Scholar
Excoffier, Laurent, “Human Demographic History: Refining the Recent African Origin model,” Current Opinion in Genomics and Development 12 (2002), 675–82Google Scholar
Reconstructing the Demography of Prehistoric Human Populations from Molecular Data,” Evolutionary Anthropology 11/S1 (2002), 166–70
Eller, Elise, “Estimating Relative Population Sizes from Simulated Data Sets and the Question of Greater African Effective Size,” AJPA 116 (2001), 1–12Google Scholar
Eller, Elise, “Population Extinction and Recolonization in Human Demographic History,” Mathematical Biosciences 177 and 178 (2002), 1–10Google Scholar
Eller, Elise et al., “Local Extinction and Recolonization, Species Effective Population Size, and Modern Human Origins,” Human Biology 76 (2004), 689–709Google Scholar
Wakeley, John, “Metapopulation Models for Historical Inference,” Molecular Biology 13 (2004), 865–75Google Scholar
Tenesa, Albert et al., “Recent Human Effective Population Size Estimated from Linkage Disequilibrium,” Genomic Research 17 (2007), 520–6Google Scholar
Takahata, N., “A Genetic Perspective on the Origins and History of Humans,” Annual Review Ecol. Syst. 26 (1995), 343–72Google Scholar
Lee, Richard B. and Devore, Irven, eds., Man the Hunter (Chicago, IL, 1968)
Dumond, Don E., “The Limitation of Human Population: A Natural History,” Science 187 (1975), 713–21Google Scholar
Hassan, Fekri A., Demographic Archaeology (New York, 1981)
Harris, Marvin, Cannibals and Kings (New York, 1977), 11–25
Sahlins, Marshall, Stone Age Economics (New York, 1972)
Cohen, Mark Nathan and Armalegos, George J., Paleopathology at the Origins of Agriculture (New York, 1984)
Steckel, Richard and Rose, Jerome C., eds., The Backbone of History: Health and Nutrition in the Western Hemisphere (New York, 2002)
Keckler, Charles N. W., “Catastrophic Mortality in Simulations of Forager Age-at-Death: Where did all the Humans Go?” in Paine, Richard R., ed., Archaeological Demography: Multidisciplinary Approaches to Prehistoric Population (Carbondale, 1997), 205–28
Hill, Kim R. and Hurtado, A. Magdelena, Ache Life History: The Ecology and Demography of a Foraging People (New York, 1996)
Warrick, Gary, A Population History of the Huron-Petun, A.D. 500–1650 (New York, 2008), 46–51
Gignoux, Christopher R. et al., “Rapid, Global Demographic Expansions after the Origins of Agriculture,” PNAS 108 (2011), 6044–9Google Scholar
Ammerman, Albert J., “Late Pleistocene Population Dynamics: An Alternative View,” HumEcol 3 (1975), 219–33Google Scholar
Handwerker, W. Penn, “The First Demographic Transition: An Analysis of Subsistence Choices and Reproductive Consequences,” AmAnth 85 (1983), 5–27Google Scholar
Wood, James W., “A Theory of Preindustrial Population Dynamics: Demography, Economy, and Well-Being in Malthusian Systems,” CA 39 (1998), 121Google Scholar
Paine, Richard R., “If a Population Crashes in Prehistory, and There is no Paleodemographer There to Hear It, Does It Make a Sound?AJPA 112 (2000), 181–90Google Scholar
Shennan, Steven, “Population, Culture, and the Dynamics of Culture Change,” CA 41 (2000), 811–35Google Scholar
Shennan, Stephen, Genes, Memes, and Human History: Darwinian Archaeology and Cultural Evolution (London, 2002), 100–23
Boone, James L., “Subsistence Strategies and Early Human Population History: An Evolutionary Ecological Perspective,” WdArch 34 (2002), 6–25 (21)Google Scholar
Caldwell, John C. and Caldwell, Bruce K., “Pretransition Population Control and Equilibrium,” PopSt 57 (2003), 199–215Google Scholar
Crawford, Dorothy H., Deadly Companions: How Microbes Shaped Our History (Oxford, 2007), 46–53
Deacon, H. J. and Thackery, J. F., “Late Pleistocene Environmental Changes and Implications for the Archaeological Record in Southern Africa,” in Vogel, J. C., ed., Late Cainozoic Palaeoclimates of the Southern Hemisphere (Rotterdam, 1984), 375–90
Livi-Bacci, Massimo, Population and Nutrition: An Essay on European Demographic History (New York, 1991), 111–13
Rahmstorf, Stefan, “Ocean Circulation and Climate during the Past 120,000 Years,” Nature 419 (2002), 207–14Google Scholar
Clark, Peter et al., “The Role…”; Wallace S. Broecker, “The Chaotic Climate,” SA 273/5 (Nov. 1995), 62–8Google Scholar
Broecker, Wallace S. and Denton, George H., “What Drives Glacial Cycles?SA 262/1 (Jan. 1990), 48–56Google Scholar
Braun, Holger et al., “Possible Solar Origin of the 1,470 Glacial Climate Cycle Demonstrated in a Coupled Model,” Nature 438 (2005), 208–11Google Scholar
Staufer, B., “Atmospheric CO2 and Millennial-Scale Climate Changes during the Last Glacial Period,” Nature 392 (1998), 59–62Google Scholar
Alley, R. B., “Stochastic Resonance in the North Atlantic,” Paleoceanography 16 (2001), 190–8Google Scholar
Ganopolski, Andrey and Rahmstorf, Stefan, “Abrupt Climate Changes due to Stochastic Resonance,” Physical Review Letters 88/3 (2002), 038501Google Scholar
Fritz, S. C. et al., “Millennial-Scale Climate Variability during the Last Glacial Period in the Tropical Andes,” QSR 29 (2010), 1017–24Google Scholar
Tjallingii, Rik et al., “Coherent High- and Low-Latitude Control of the Northwest African Hydrological Balance,” NatGeosc 1 (2008), 670–5Google Scholar
Tierney, Jessica E. et al., “Northern Hemisphere Controls on Tropical African Climate during the Past 60,000 Years,” Science Express (Sept. 1, 2008)Google Scholar
Epica Community Members, “One-to-One Coupling of Glacial Climate Variability in Greenland and Antarctica,” Nature 444 (2006), 195–8Google Scholar
Wang, Xianfeng et al., “Wet Periods in Northeastern Brazil over the Past 210 kyr Linked to Distant Climate Anomalies,” Nature 432 (2004), 740–3Google Scholar
Altabet, Mark A. et al., “The Effect of Millennial-Scale Changes in Arabian Sea Denitrification on Atmospheric CO2,” Nature 415 (2002), 159–62Google Scholar
Stott, Lowell et al., “Super ENSO and Global Climate Oscillations at Millennial Time Scales,” Science 297 (2002), 222–6Google Scholar
Bond, Gerald et al., “A Pervasive Millennial-Scale Cycle in North Atlantic Holocene and Glacial Climate,” Science 278 (1997), 1257–66Google Scholar
Müller, Ulrich J. et al., “The Role of Climate in the Spread of Modern Humans into Europe,” QSR 30 (2011), 273–9Google Scholar
Shennan, Stephen, “Demography and Cultural Innovation: A Model and Its Implications for the Emergence of Modern Culture,” CArchJ 11 (2001), 15Google Scholar
Richerson, Peter J. and Boyd, Robert, “Institutional Evolution in the Holocene,” in Runciman, W. G., ed., The Origins of Social Institutions (Oxford, 2001), 197–234
Boserup, Ester’s classic statements were in The Condition for Agricultural Growth: The Economic of Agrarian Change under Population Pressure (Chicago, IL, 1965)
Lee, Ronald D., “Malthus and Boserup: A Dynamic Synthesis,” in Coleman, David and Schofield, Roger S., eds., The State of Population Theory: Forward from Malthus (Oxford, 1986), 96–103
Driscoll, Catherine, “Grandmothers, Hunters, and Human Life History,” Biology and Philosophy 24 (2009), 665–86Google Scholar
Caspari, Rachel and Lee, Sang-Hee, “Older Age Becomes Common Late in Human Evolution,” PNAS 101 (2004), 10895–900Google Scholar
Powell, Adam et al., “Late Pleistocene Demography and the Appearance of Modern Human Behavior,” Science 324 (2009), 1298–301Google Scholar
Zilhao, Joao, “The Emergence of Ornaments and Art: An Archaeological Perspective on the Origins of ‘Behavioral Modernity,’JArchRes 15 (2007), 1–54Google Scholar
Manning, Patrick, “Homo Sapiens Populates the Earth: A Provisional Synthesis, Privileging Linguistic Evidence,” JWH 17 (2006), 115–58Google Scholar
Driscoll, Catherine, “Grandmothers, Hunters and Human Life History,” BiolPlos 24 (2009), 665–86Google Scholar
Chase, Brian M., “South Africa Paleoenvironments during Marine Oxygen Isotope Stage 4: A Context for the Howiesons Poort and Still Bay Industries,” JArchS 37 (2010), 1359–66Google Scholar
Jacobs, Zenobia et al., “Ages for the Middle Stone Age of Southern Africa: Implications for Human Behavior and Dispersal,” Science 322 (2008), 733–5Google Scholar
Hovers, Erella and Belfer-Cohen, Anna, “’Now You See it, Now You Don’t’ – Modern Human Behavior in the Middle Paleolithic,” in Hovers, Erella and Kuhn, Steven L., eds., Transitions before the Transition: Evolution and Stability in the Middle Paleolithic and Middle Stone Age (New York, 2006), 295–304
Kuhn, S. L. et al., “The Early Upper Paleolithic and the Origins of Modern Human Behavior, in Brantingham, P. Jeffrey et al., eds., The Early Upper Paleolithic beyond Western Europe (Berkeley, 2004), 242–8
Diamond, Jared, Guns, Germs, and Steel: The Fate of Human Societies (New York, 1997), 111–12

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.

  • Human Emergence
  • John L. Brooke, Ohio State University
  • Book: Climate Change and the Course of Global History
  • Online publication: 05 August 2014
  • Chapter DOI: https://doi.org/10.1017/CBO9781139050814.004
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.

  • Human Emergence
  • John L. Brooke, Ohio State University
  • Book: Climate Change and the Course of Global History
  • Online publication: 05 August 2014
  • Chapter DOI: https://doi.org/10.1017/CBO9781139050814.004
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.

  • Human Emergence
  • John L. Brooke, Ohio State University
  • Book: Climate Change and the Course of Global History
  • Online publication: 05 August 2014
  • Chapter DOI: https://doi.org/10.1017/CBO9781139050814.004
Available formats
×