Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-27T01:06:57.964Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  11 February 2019

J. Tyler Faith
Affiliation:
University of Utah
R. Lee Lyman
Affiliation:
University of Missouri, Columbia
Get access
Type
Chapter
Information
Paleozoology and Paleoenvironments
Fundamentals, Assumptions, Techniques
, pp. 319 - 394
Publisher: Cambridge University Press
Print publication year: 2019

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

Abrams, Peter A., and Rowe, Locke. 1996. The Effects of Predation on the Age and Size of Maturity of Prey. Evolution 50:10521061.CrossRefGoogle ScholarPubMed
Agenbroad, Larry D. 2009. Mammuthus exilis from the California Channel Islands: Height, Mass, and Geologic Age. In Proceedings of the 7th California Islands Symposium, edited by Damiani, C. C and Garcelon, D. K, pp. 1519. Institute for Wildlife Studies, Arcata, CA.Google Scholar
Agenbroad, Larry D. 2012. Giants and Pygmies: Mammoths of Santa Rosa Island, California (USA). Quaternary International 255:28.Google Scholar
Ager, Derek V. 1979. Paleoecology. In Encyclopedia of Paleontology, edited by Fairbridge, Rhodes W. and Jablonski, David, pp. 530541. Dowden, Hutchinson and Ross, Stroudsburg, PA.CrossRefGoogle Scholar
Albarella, Umberto. 2002. “Size Matters”: How and Why Biometry is Still Important in Zooarchaeology. In Bones and the Man: Studies in Honour of Don Brothwell, edited by Dobney, Keith and O’Connor, Terry, pp. 5162. Oxbow Books, Oxford.Google Scholar
Alemseged, Zeresenay. 2003. An Integrated Approach to Taphonomy and Faunal Change in the Shungura Formation (Ethiopia) and its Implications for Hominid Evolution. Journal of Human Evolution 44:451478.Google Scholar
Alemseged, Z[eresenay], Bobe, R., and Geraads, D.. 2007. Comparability of Fossil Data and its Significance for the Interpretation of Hominin Environments: A Case Study in the Lower Omo Valley, Ethiopia. In Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence, edited by Bobe, R[ené, Alemseged, Z[eresenay, and Behrensmeyer, A[nna] K., pp. 159181. Springer, Dordrecht.Google Scholar
Alfimov, A. V., and Berman, D. I.. 2009. Possible Errors of the Mutual Climatic Range (MCR) Method in Reconstructing the Pleistocene Climate of Beringia. Entomological Review 89:487499.Google Scholar
Allan, Robert S. 1948. Geological Correlation and Paleoecology. Geological Society of America Bulletin 59:110.Google Scholar
Allègre, Claude J. 2008. Isotope Geochemistry. Cambridge University Press.Google Scholar
Allison, Peter A., and Bottjer, David J. (editors). 2011. Taphonomy: Process and Bias through Time. Topics in Geobiology 32. Springer, Dordrecht.CrossRefGoogle Scholar
Alroy, John. 2000. New Methods for Quantifying Macroevolutionary Patterns and Processes. Paleobiology 26:707733.Google Scholar
Alroy, John, Aberhan, Martin, Bottjer, David J. et al. 2008. Phanerozoic Trends in the Global Diversity of Marine Invertebrates. Science 321:97100.Google Scholar
Ambrose, Stanley H. 1998. Chronology of the Later Stone Age and Food Production in East Africa. Journal of Archaeological Science 25:377392.Google Scholar
Ambrose, Stanley H., and DeNiro, Michael J.. 1986. The Isotopic Ecology of East African Mammals. Oecologia 69:395406.Google Scholar
Ambrose, Stanley H., and Katzenberg, M. Anne (editors). 2000. Biogeochemical Approaches to Paleodietary Analysis. Kluwer Academic, New York.Google Scholar
Ambrose, Stanley H., and Krigbaum, John. 2003. Bone Chemistry and Bioarchaeology. Journal of Anthropological Archaeology 22:193199.Google Scholar
Ambrose, Stanley H., and Norr, Lynette. 1993. Experimental Evidence for the Relationship of the Carbon Isotope Ratios of Whole Diet and Dietary Protein to Those of Bone Collagen and Carbonate. In Prehistoric Human Bone: Archaeology at the Molecular Level, edited by Lambert, Joseph B. and Grupe, Gisela, pp. 137. Springer, Berlin.Google Scholar
Anderson, Alexander C. 1875. Further Remarks on the Reindeer; and on its Assumed Coexistence with the Hippopotamus. In Reliquiae Aquitanicae; Being Contributions to the Archaeology and Paleontology of Périgord and the Adjoining Provinces of Southern France, by Edourd Lartet and Henry Christy, edited by Jones, Thomas R., pp. 153160. Williams and Norgate, London.Google Scholar
Anderson, Elaine. 1968. Fauna of the Little Box Elder Cave, Converse County, Wyoming: The Carnivora. University of Colorado Studies, Series in Earth Sciences 6:159.Google Scholar
Andrews, Peter. 1990. Owls, Caves and Fossils: Predation, Preservation, and Accumulation of Small Mammal Bones in Caves, with an Analysis of the Pleistocene Cave Faunas from Westbury-sub-Mendip, Somerset, UK. University of Chicago Press.Google Scholar
Andrews, Peter. 1992. Community Evolution in Forest Habitats. Journal of Human Evolution 22:423438.Google Scholar
Andrews, Peter. 1995. Mammals as Palaeoecological Indicators. Acta Zoologica Cracoviensia 38:5972.Google Scholar
Andrews, Peter. 1996. Palaeoecology and Hominoid Palaeoenvironments. Biological Reviews 71:257300.Google Scholar
Andrews, Peter. 2006. Taphonomic Effects of Faunal Impoverishment and Faunal Mixing. Palaeogeography, Palaoeclimatology, Palaeoecology 241:572589.CrossRefGoogle Scholar
Andrews, Peter, and Hixson, Sylvia. 2014. Taxon-Free Methods of Palaeoecology. Annales Zoologici Fennici 51:269284.Google Scholar
Andrews, Peter, and Humphrey, Louise. 1999. African Miocene Environments and the Transition to Early Hominins. In African Biogeography, Climate Change, and Human Evolution, edited by Bromage, Timothy G. and Schrenk, Friedemann, pp. 282300. Oxford University Press.Google Scholar
Andrews, Peter, Lord, J. M., and Nesbit Evans, Elisabeth M.. 1979. Patterns of Ecological Diversity in Fossil and Modern Mammalian Faunas. Biological Journal of the Linnean Society 11:177205.Google Scholar
Anyonge, William. 1996. Microwear on Canines and Killing Behavior in Large Carnivores: Saber Function in Smilodon fatalis. Journal of Mammalogy 77:10591067.Google Scholar
Araújo, Miguel B., and Peterson, A. Townsend. 2012. Uses and Misuses of Bioclimatic Envelope Modeling. Ecology 93:15271539.Google Scholar
Araújo, Miguel B., Nogués-Bravo, David, Diniz-Filho, José Alexandre F. et al. 2007. Quaternary Climate Changes Explain Diversity among Reptiles and Amphibians. Ecography 31:815.CrossRefGoogle Scholar
Ashton, Kyle G. 2002. Patterns of Within-Species Body Size Variation of Birds: Strong Evidence for Bergmann’s Rule. Global Ecology and Biogeography 11:505523.CrossRefGoogle Scholar
Ashton, Kyle G., Tracy, Mark C., and de Queiroz, Alan. 2000. Is Bergmann’s Rule Valid for Mammals? American Naturalist 156:390415.Google Scholar
Assefa, Zelalem, Yirga, Solomon, and Reed, Kaye E.. 2008. The Large-Mammal Fauna from the Kibish Formation. Journal of Human Evolution 55:501512.Google Scholar
Atkinson, T. C., Briffa, K. R., Coope, G. R., Joachim, M. J., and Perry, D. W.. 1986. Climatic Calibration of Coleopteran Data. In Handbook of Holocene Palaeoecology and Palaeohydrology, edited by Bergland, B. E, pp. 851858. John Wiley and Sons, Chichester.Google Scholar
Atkinson, T. C., Briffa, K. R., and Coope, G. R.. 1987. Seasonal Temperatures in Britain during the Past 22,000 Years, Reconstructed Using Beetle Remains. Nature 325:587592.Google Scholar
Austin, Mike. 2007. Species Distribution Models and Ecological Theory: A Critical Assessment and Some Possible New Approaches. Ecological Modeling 200:119.Google Scholar
Avery, D. M. 1982. Micromammals as Palaeoenvironmental Indicators and an Interpretation of the Late Quaternary in the Southern Cape Province, South Africa. Annals of the South African Museum 85:183374.Google Scholar
Avery, D. M. 1988. Micromammals and Paleoenvironmental Interpretation in Southern Africa. Geoarchaeology 3:4152.Google Scholar
Avery, D. M. 1990. Holocene Climatic Change in Southern Africa: The Contribution of Micromammals to its Study. South African Journal of Science 86:407412.Google Scholar
Avery, D. M. 1992. The Environment of Early Modern Humans at Border Cave, South Africa: Micromammalian Evidence. Palaeogeography, Palaeoclimatology, Palaeoecology 91:7187.Google Scholar
Avery, D. M. 1999. A Preliminary Assessment of the Relationship between Trophic Variability in Southern African Barn Owls Tyto alba and Climate. Ostrich 70:179186.CrossRefGoogle Scholar
Avery, D. M. 2004. Size Variation in the Common Molerat Cryptomys hottentotus from Southern Africa and its Potential for Palaeoenvironmental Reconstruction. Journal of Archaeological Science 31:273282.Google Scholar
Avery, D. M. 2007. Micromammals as Palaeoenvironmental Indicators of the Southern African Quaternary. Transactions of the Royal Society of South Africa 62:1723.Google Scholar
Ayliffe, Linda, and Chivas, A. R.. 1990. Oxygen Isotope Composition of the Bone Phosphate of Australian Kangaroos: Potential as a Palaeoenvironmental Recorder. Geochimica et Cosmochimica Acta 54:26032609.Google Scholar
Badgley, Catherine. 1986. Counting Individuals in Mammalian Fossil Assemblages from Fluvial Environments. Palaios 1:328338.Google Scholar
Bailey, I. W., and Sinnott, E. W.. 1915. A Botanical Index of Cretaceous and Tertiary Climates. Science 41:831834.CrossRefGoogle ScholarPubMed
Bailey, I. W., and Sinnott, E. W.. 1916. The Climatic Distribution of Certain Types of Angiosperm Leaves. American Journal of Botany 3:2439.Google Scholar
Baker, G., Jones, L. H. P., and Wardrop, I. D.. 1959. Cause of Wear in Sheeps’ Teeth. Nature 184:15831584.Google Scholar
Balkwill, Darlene McCuaig, and Cumbaa, Stephen L.. 1992. A Guide to the Identification of Post-Cranial Bones of Bos taurus and Bison bison. Syllogeus 71. Canadian Museum of Nature, Ottawa.Google Scholar
Bañuls-Cardona, Sandra, López-García, J. M., Blain, Hugues-Alexandre, and Salomó, A. C.. 2012. Climate and Landscape during the Last Glacial Maximum in Southwestern Iberia: The Small-Vertebrate Association from the Sala De Las Chimeneas, Maltravieso, Extremadura. Comptes Rendus Palevol 11:3140.Google Scholar
Bañuls-Cardona, Sandra, López-García, J. M., Blain, Hugues-Alexandre, Lozano-Fernández, I., and Cuenca-Bescós, Gloria. 2014. The End of the Last Glacial Maximum in the Iberian Peninsula Characterized by the Small-Mammal Assemblages. Journal of Iberian Geology 40:1927.Google Scholar
Barnosky, Anthony D. 1998. What Causes “Disharmonious” Mammal Assemblages? In Quaternary Paleozoology in the Northern Hemisphere, edited by Saunders, Jeffrey J., Styles, Bonnie W., and Baryshnikov, Gennady F., pp. 173186. Illinois State Museum Scientific Papers 27. Illinois State University, Springfield.Google Scholar
Barnosky, Anthony D. 2001. Distinguishing the Effects of the Red Queen and Court Jester on Miocene Mammal Evolution in the Northern Rocky Mountains. Journal of Vertebrate Paleontology 21:172185.Google Scholar
Barnosky, Anthony D. (editor) 2004. Biodiversity Response to Climate Change in the Middle Pleistocene: The Porcupine Cave Fauna from Colorado. University of California Press, Berkeley.Google Scholar
Barnosky, Anthony D. 2009. Heatstroke: Nature in an Age of Global Warming. Island Press, Washington, DC.Google Scholar
Barnosky, Anthony D., and Carrasco, Marc A.. 2002. Effects of Oligo-Miocene Global Climate Changes on Mammalian Species Richness in the Northwestern Quarter of the USA. Evolutionary Ecology Research 4:811841.Google Scholar
Barnosky, Anthony D., and Hadly, Elizabeth A.. 2016. Tipping Points for Planet Earth. Thomas Dunne Books, St. Martin’s Press, New York.Google Scholar
Barnosky, Anthony D., Hadly, Elizabeth A., Gonzalez, Patrick et al. 2017. Merging Paleobiology with Conservation Biology to Guide the Future of Terrestrial Ecosystems. Science 355:eaah4787.Google Scholar
Barr, W. Andrew. 2014. Functional Morphology of the Bovid Astragalus in Relation to Habitat: Controlling Phylogenetic Signal in Ecomorphology. Journal of Morphology 275:12011216.Google Scholar
Barr, W. Andrew. 2015. Paleoenvironments of the Shungura Formation (Plio-Pleistocene: Ethiopia) Based on Ecomorphology of the Bovid Astragalus. Journal of Human Evolution 88:97107.Google Scholar
Barr, W. Andrew. 2017. Bovid Locomotor Functional Trait Distributions Reflect Land Cover and Annual Precipitation in Sub-Saharan Africa. Evolutionary Ecology Research 18:253269.Google Scholar
Barr, W. Andrew, and Scott, Robert S.. 2014. Phylogenetic Comparative Methods Complement Discriminant Function Analysis in Ecomorphology. American Journal of Physical Anthropology 153:663674.Google Scholar
Barry, John C., Morgan, Michèle, Flynn, Lawrence J. et al. 2002. Faunal and Environmental Change in the Late Miocene Siwaliks of Northern Pakistan. Paleobiology 28 (special issue 3):171.Google Scholar
Bartlein, P. J., Webb, T. III, and Fleri, E.. 1984. Holocene Climatic Change in the Northern Midwest: Pollen-Derived Estimates. Quaternary Research 22:361374.Google Scholar
Bate, D. M. A. 1937. Palaeontology: The Fossil Fauna of the Wady El-Mughara Caves. In The Stone Age of Mt. Carmel, vol. i: Excavations at the Wady El-Mughara, by Garrod, D. A. E. and Bate, D. M. A., pp. 135240. Oxford University Press.Google Scholar
Baxter, M. J. 2001. Methodological Issues in the Study of Assemblage Diversity. American Antiquity 66:715725.Google Scholar
Bedford, Jean N. 1978. A Technique for Sex Determination of Mature Bison Metapodials. Plains Anthropologist 14:4043.Google Scholar
Begon, Michael, Townsend, Colin R., and Harper, John L.. 2006. Ecology: From Individuals to Ecosystems, fourth edition. Blackwell, Malden, MA.Google Scholar
Behrensmeyer, Anna K. 1982. Time Resolution in Fluvial Vertebrate Assemblages. Paleobiology 8:211227.Google Scholar
Behrensmeyer, Anna K. 2006. Climate Change and Human Evolution. Science 311:476478.Google Scholar
Behrensmeyer, Anna K., and Chapman, Ralph E.. 1993. Models and Simulations of Time-Averaging in Terrestrial Veretebrate Accumulations. In Taphonomic Approaches to Time Resolution in Fossil Assemblages, edited by Kidwell, Susan M. and Behrensmeyer, Anna K., pp. 125149. Short Courses in Paleontology 6, The Paleontological Society, Knoxville, TN.Google Scholar
Behrensmeyer, Anna K., and Hook, Robert W.. 1992. Paleoenvironmental Contexts and Taphonomic Modes. In Terrestrial Ecosystems through Time: Evolutionary Paleoecology of Terrestrial Plants and Animals, edited by Behrensmeyer, Anna K., Damuth, John D., DiMichele, William A., Potts, Richard, Sues, Hans-Dieter, and Wing, Scott L., pp. 1536. University of Chicago Press.Google Scholar
Behrensmeyer, Anna K., Kidwell, Susan M., and Gastaldo, Robert A.. 2000. Taphonomy and Paleobiology. In Deep Time: Paleobiology’s Perspective, edited by Erwin, Douglas H. and Wing, Scott W., pp. 103147. Paleobiology (Supplement) 26.Google Scholar
Behrensmeyer, Anna K., Stayton, C. Tristan, and Chapman, Ralph E.. 2003. Taphonomy and Ecology of Modern Avifaunal Remains from Amboseli Park, Kenya. Paleobiology 29:5270.Google Scholar
Behrensmeyer, A[nna] K., Bobe, R[ené, and Alemseged, Z[eresenay. 2007a. Approaches to the Analysis of Faunal Change during the East African Pliocene. In Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence, edited by Bobe, R[ené, Alemseged, Z[eresenay, and Behrensmeyer, A[nna] K., pp. 124. Springer, Dordrecht.Google Scholar
Behrensmeyer, A[nna] K., Alemseged, Z[eresenay, and Bobe, R[ené. 2007b. Finale and Future: Investigating Faunal Evidence for Hominin Paleoecology in East Africa. In Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence, edited by Bobe, R[ené, Alemseged, Z[eresenay, and Behrensmeyer, A[nna] K., pp. 333345. Springer, Dordrecht.Google Scholar
Bell, Christopher J., Gauthier, Jacques A., and Bever, Gabe S.. 2010. Covert Biases, Circularity, and Apomorphies: A Critical Look at the North American Quaternary Herpetofaunal Stability Hypothesis. Quaternary International 217:3036.Google Scholar
Bell, Michael A., Travis, Matthew P., and Blouw, D. Max. 2005. Inferring Natural Selection in a Fossil Threespine Stickleback. Paleobiology 32:562577.Google Scholar
Belyea, Lisa R. 2007. Revealing the Emperor’s New Clothes: Niche-Based Palaeoenvironmental Reconstruction in the Light of Recent Ecological Theory. The Holocene 17:683688.Google Scholar
Berger, A[ndre], and Loutre, M. F.. 1991. Insolation Values for the Climate of the Last 10 Million Years. Quaternary Science Reviews 10:297317.Google Scholar
Bergerud, Arthur T. 2000. Caribou. In Ecology and Management of Large Mammals in North America, edited by Demarais, Stephen and Krausman, Paul R., pp. 658693. Prentice Hall, Upper Saddle River, NJ.Google Scholar
Bergmann, Carl. 1847. Ueber die Verhältnisse der Wärmeökonomie der Thiere zu Ihrer Grösse. Gottinger studien 3:595708.Google Scholar
Berke, Melissa A., Johnson, Thomas C., Werne, Josef P., Schouten, Stefan, and Sinninghe Damsté, Jaap S.. 2012. A Mid-Holocene Thermal Maximum at the End of the African Humid Period. Earth and Planetary Science Letters 351–352:95104.Google Scholar
Berner, Robert A., and Kothavala, Zavareth. 2001. Geocarb III: A Revised Model of Atmospheric CO2 over Phanerozoic Time. American Journal of Science 301:182204.Google Scholar
Berto, Claudio, Boscato, Paolo, Boschin, Francesco, Luzi, Elisa, and Ronchitelli, Annamaria. 2017. Paleoenvironmental and Paleoclimatic Context during the Upper Palaeolithic (Late Upper Pleistocene) in the Italian Peninsula: The Small Mammal Record from Grotta Paglicci (Rignano Garganico, Foggia, Southern Italy). Quaternary Science Reviews 168:3041.CrossRefGoogle Scholar
Bininda-Emonds, Olaf R. P., Cardillo, Marcel, Jones, Kate E. et al. 2007. The Delayed Rise of Present-Day Mammals. Nature 446:507512.Google Scholar
Birch, L. C. 1957. The Role of Weather in Determining the Distribution and Abundance of Animals. Cold Spring Harbor Symposium on Quantitative Biology 22:203218.Google Scholar
Birks, H. J[ohn] B. 1995. Quantitative Palaeoenvironmental Reconstructions. In Statistical Modelling of Quaternary Science Data, edited by Maddy, Darrel and Brew, John S., pp. 161254. Technical Guide 5. Quaternary Research Association, Cambridge, UK.Google Scholar
Birks, H. John, B., Hein, Oliver, Seppa, Heikki, and Bjune, Anne E.. 2010. Strengths and Weaknesses of Quantitative Climate Reconstructions Based on Late-Quaternary Biological Proxies. Open Ecology Journal 3:68110.Google Scholar
Blaauw, Maarten. 2012. Out of Tune: The Dangers of Aligning Proxy Archives. Quaternary Science Reviews 36:3849.Google Scholar
Blain, Hugues-Alexandre, Bailon, Salvador, Cuenca-Bescós, Gloria et al. 2009. Long-Term Climate Record Inferred from Early-Middle Pleistocene Amphibian and Squamate Reptile Assemblages at the Gran Dolina Cave, Atapuerca, Spain. Journal of Human Evolution 56:5565.Google Scholar
Blain, Hugues-Alexandre, Bailon, Salvador, Cuenca-Bescós, Gloria et al. 2010. Climate and Environment of the Earliest West European Hominins Inferred from the Amphibian and Squamate Reptile Assemblages: Sima del Elefante Lower Red Unit, Atapuerca, Spain. Quaternary Science Reviews 29:30343044.Google Scholar
Blain, Hugues-Alexandre, Lozano-Fernández, Iván, Ollé, Andreu, Rodríguez, Jesus, Santonja, Manuel, and Pérez-González, Alfredo. 2015. The Continental Record of Marine Isotope Stage 11 (Middle Pleistocene) on the Iberian Peninsula Characterized by the Herpetofaunal Assemblages. Journal of Quaternary Science 30:667678.Google Scholar
Blair, W. Frank. 1958. Distributional Patterns of Vertebrates in the Southern United States in Relation to Past and Present Environments. In Zoogeography, edited by Hubbs, Carl L., pp. 433468. Publication 51. American Association for the Advancement of Science, Washington, DC.Google Scholar
Blois, Jessica L., Feranec, Robert S., and Hadly, Elizabeth A.. 2008. Environmental Influences on Spatial and Temporal Patterns of Body-Size Variation in California Ground Squirrels (Spermophilus beecheyi). Journal of Biogeography 35:602613.Google Scholar
Blois, Jessica L., McGuire, Jenny L., and Hadly, Elizabeth A.. 2010. Small Mammal Diversity Loss in Response to Late-Pleistocene Climatic Change. Nature 465:771774.Google Scholar
Blois, Jessica L., Zametske, Phoebe L., Fitzpatrick, Matthew C., and Finnegan, Seth. 2013. Climate Change and the Past, Present, and Future of Biotic Interactions. Science 341:499504.Google Scholar
Blumenthal, Scott A., Levin, Naomi E., Brown, Francis H. et al. 2017. Aridity and Hominin Environments. Proceedings of the National Academy of Sciences USA 114:73317336.Google Scholar
Bobe, René. 2006. The Evolution of Arid Ecosystems in Eastern Africa. Journal of Arid Environments 66:564584.Google Scholar
Bobe, René, and Behrensmeyer, Anna K.. 2004. The Expansion of Grassland Ecosystems in Africa in Relation to Mammalian Evolution and the Origin of the Genus Homo. Palaeogeography, Palaeoclimatology, Palaeoecology 207:399420.Google Scholar
Bobe, René, and Eck, Gerald G.. 2001. Responses of African Bovids to Pliocene Climatic Change. Paleobiology 27 (special issue 2):148.Google Scholar
Bobe, René, and Leakey, Meave G.. 2009. Ecology of Plio-Pleistocene Mammals in the Omo-Turkana Basin and the Emergence of Homo. In The First Humans: Origin and Early Evolution of the Genus Homo, edited by Grine, Frederick E., Fleagle, John G., and Leakey, Richard E., pp. 173184. Springer, Dordrecht.Google Scholar
Bobe, René, Behrensmeyer, Anna K., and Chapman, Ralph E.. 2002. Faunal Change, Environmental Variability and Late Pliocene Hominin Evolution. Journal of Human Evolution 42:475497.Google Scholar
Bobe, R[ené, Alemseged, Z[eresenay, and Behrensmeyer, A[nna] K. (editors). 2007. Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. Springer, Dordrecht.Google Scholar
Boessneck, Joachim, and von den Driesch, Angela. 1978. The Significance of Measuring Animal Bones from Archaeological Sites. In Approaches to Faunal Analysis in the Middle East, edited by Meadow, Richard H. and Zeder, Melinda A., pp. 2539. Peabody Museum of Archaeology and Ethnology Bulletin 2. Harvard University, Cambridge, MA.Google Scholar
Boivin, Nicole L., Zeder, Melinda A., Fuller, Dorian Q. et al. 2016. Ecological Consequences of Human Niche Construction: Examining Long-Term Anthropogenic Shaping of Global Species Distributions. Proceedings of the National Academy of Sciences USA 113:63886396.Google Scholar
Bökönyi, Sándor. 1982. The Climatic Interpretation of Macrofaunal Assemblages in the Near East. In Palaeoclimates, Palaeoenvironments and Human Communities in the Eastern Mediterranean Region in Later Prehistory, edited by Bintliff, John L. and Van Zeist, Willem, pp. 149163. British Archaeological Reports, International Series 133. BAR, Oxford.Google Scholar
Bond, William J., and Midgley, Guy F.. 2000. A Proposed CO2–Controlled Mechanism of Woody Plant Invasion in Grasslands and Savannas. Global Change Biology 6:865869.Google Scholar
Bond, W[illiam] J., Midgley, G. F., and Woodward, F. I.. 2003. The Importance of Low Atmospheric CO2 and Fires in Promoting the Spread of Grasslands and Savannas. Global Change Biology 9:973982.Google Scholar
Boshoff, Andre F., and Kerley, Graham I. H.. 2001. Potential Distributions of the Medium- to Large-Sized Mammals in the Cape Floristic Region, Based on Historical Accounts and Habitat Requirements. African Zoology 36:245273.Google Scholar
Boshoff, Andre F., Kerley, Graham I. H., and Cowling, Richard M.. 2001. A Pragmatic Approach to Estimating the Distributions and Spatial Requirements of the Medium- to Large-Sized Mammals in the Cape Floristic Region, South Africa. Diversity and Distributions 7:2943.Google Scholar
Bottjer, David J., Cambell, Kathleen A., Schubert, Jennifer K., and Droser, Mary I.. 1995. Palaeoecological Models, Non-Uniformitarianism, and Tracking the Changing Ecology of the Past. In Marine Palaeoenvironmental Analysis from Fossils, edited by Bosence, Dan W. J and Allison, Peter A., pp. 726. Geological Society of America Special Paper 83. Geological Society of America, Boulder, CO.Google Scholar
Boudadi-Maligne, Myriam, and Escarguel, Gilles. 2014. A Biometric Re-Evaluation of Recent Claims for Early Upper Palaeolithic Wolf Domestication in Eurasia. Journal of Archaeological Science 45:8089.Google Scholar
Bowden, Joseph J., Eskildsen, Anne, Hansen, Rikke R. et al. 2015. High-Arctic Butterflies Become Smaller with Rising Temperature. Biology Letters 11:20150574.Google Scholar
Boyce, Mark S. 1978. Climatic Variability and Body Size Variation in the Muskrats (Ondatra zibethicus) of North America. Oecologia 36:119.Google Scholar
Boyce, Mark S. 1979. Seasonality and Patterns of Natural Selection for Life Histories. American Naturalist 114:569583.Google Scholar
Bradley, Raymond S. 1985. Quaternary Paleoclimatology: Methods of Paleoclimatic Reconstruction. Allen and Unwin, Boston.Google Scholar
Bradley, Raymond S. 2015. Paleoclimatology: Reconstructing Climates of the Quaternary, third edition. Elsevier, Amsterdam.Google Scholar
Brain, C. K. 1981. The Hunters or the Hunted? An Introduction to African Cave Taphonomy. University of Chicago Press.Google Scholar
Bray, P. J., Blockley, S. P. E., Coope, G. R. et al. 2006. Refining Mutual Climatic Range (MCR) Quantitative Estimates of Palaeotemperature Using Ubiquity Analysis. Quaternary Science Reviews 25:18651876.Google Scholar
Breslawski, Ryan P., and Byers, David A.. 2015. Assessing Measurement Error in Paleozoological Osteometrics with Bison Remains. Journal of Archaeological Science 53:235242.Google Scholar
Brink, James S. 1999. Preliminary Report on a Caprine from the Cape Mountains, South Africa. Archaeozoologia 10:1126.Google Scholar
Bronk Ramsey, Christopher. 2009. Bayesian Analysis of Radiocarbon Dates. Radiocarbon 51:337360.Google Scholar
Brook, Barry W., Ellis, Erle C., Perring, Michael P., Mackay, Anson W., and Blomqvist, Linus. 2013. Does the Terrestrial Biosphere Have Planetary Tipping Points? Trends in Ecology and Evolution 28:396401.Google Scholar
Brothwell, D., and Jones, R.. 1978. The Relevance of Small Mammal Studies to Archaeology. In Research Problems in Zooarchaeology, edited by Brothwell, D. R., Thomas, K. D, and Clutton-Brock, Juliet, pp. 4757. Occasional Publication 3. Institute of Archaeology, University of London.Google Scholar
Broughton, Jack M. 1997. Widening Diet Breadth, Declining Foraging Efficiency, and Prehistoric Harvest Pressure: Ichthyofaunal Evidence from the Emeryville Shellmound, California. Antiquity 71:845862.Google Scholar
Broughton, Jack M. 2000. Terminal Pleistocene Fish Remains from Homestead Cave, Utah, and Implications for Fish Biogeography in the Bonneville Basin. Copeia 2000:645656.Google Scholar
Broughton, Jack M., and Cannon, Michael D. (editors). 2010. Evolutionary Ecology and Archaeology: Applications to Problems in Human Evolution and Prehistory. University of Utah Press, Salt Lake City.Google Scholar
Broughton, Jack M., Madsen, David B., and Quade, Jay. 2000. Fish Remains from Homestead Cave and Lake Levels of the Past 13,000 Years in the Bonneville Basin. Quaternary Research 53:392401.Google Scholar
Broughton, Jack M., Cannon, Virginia I., Arnold, Shannon, Bogiatto, Raymond J., and Dalton, Kenya. 2006. The Taphonomy of Owl-Deposited Fish Remains and the Origin of the Homestead Cave Ichthyofauna. Journal of Taphonomy 4:6995.Google Scholar
Broughton, Jack M., Byers, David A., Bryson, Reid A., Eckerle, William, and Madsen, David B.. 2008. Did Climatic Seasonality Control Late Quaternary Artiodactyl Densities in Western North America? Quaternary Science Reviews 27:19161937.Google Scholar
Brown, Barnum. 1908. The Conrad Fissure, a Pleistocene Bone Deposit in Northern Arkansas: With Descriptions of Two New Genera and Twenty New Species of Mammals. American Museum of Natural History Memoir 9:155208.Google Scholar
Brown, James H. 1984. On the Relationship between Abundance and Distribution of Species. American Naturalist 124:255279.Google Scholar
Brown, James H., Gillooly, James F., Allen, Andrew P., Savage, Van M., and West, Geoffrey B.. 2004. Toward a Metabolic Theory of Ecology. Ecology 85:17711789.Google Scholar
Brown, P., Sutikna, T., Morwood, M. J. et al. 2004. A New Small-Bodied Hominin from the Late Pleistocene of Flores, Indonesia. Nature 431:10551061.Google Scholar
Brown, W. L., Jr., and Wilson, E. O.. 1956. Character Displacement. Systematic Zoology 5:4964.Google Scholar
Buckley, Michael. 2018. Zooarchaeology by Mass Spectrometry (ZooMS) Collagen Fingerprinting for the Species Identification of Archaeological Bone Fragments. In Zooarchaeology in Practice, edited by Giovas, Christina M. and LeFebvre, Michelle J., pp. 227247. Springer, Cham.Google Scholar
Buckley, Michael, Collins, Matthew, Thomas-Oates, Jane, and Wilson, Julie C.. 2009. Species Identification by Analysis of Bone Collagen Using Matrix-Assisted Laser Desorption/Ionisation Time-of-Flight Mass Spectrometry. Rapid Communications in Mass Spectrometry 23:38433854.Google Scholar
Bulinski, Katherine V. 2007. Analysis of Sample-Level Properties along a Paleoenvironmental Gradient: The Behavior of Evenness as a Function of Sample Size. Palaeogeography, Palaeoclimatology, Palaeoecology 253:490508.Google Scholar
Bunn, Henry T., Bartram, Laurence E., and Kroll, Ellen M.. 1988. Variability in Bone Assemblage Formation from Hadza Hunting, Scavenging, and Carcass Processing. Journal of Anthropological Archaeology 7:412457.Google Scholar
Burgman, Jenny H. E., Leichliter, Jennifer, Avenant, Nico L., and Ungar, Peter S.. 2016. Dental Microwear of Sympatric Rodent Species Sampled across Habitats in Southern Africa: Implications for Environmental Influence. Integrative Zoology 11:111127.Google Scholar
Burnham, Robyn J. 2008. Hide and Go Seek: What Does Presence Mean in the Fossil Record? Annals of the Missouri Botanical Garden 95:5171.Google Scholar
Burt, William H. 1958. The History and Affinities of the Recent Land Mammals of Western North America. In Zoogeography, edited by Hubbs, Carl L., pp. 131154. Publication 51. American Association for the Advancement of Science, Washington, DC.Google Scholar
Burt, William H., and Grossenheider, Richard P.. 1964. A Field Guide to the Mammals. Houghton Mifflin, Boston.Google Scholar
Burton, James H., and Price, T. Douglas. 1990. The Ratio of Barium to Strontium as a Paleodietary Indicator of Consumption of Marine Resources. Journal of Archaeological Science 17:547557.Google Scholar
Burton, James H., Price, T. Douglas, and Middleton, W. D.. 1999. Correlation of Bone Ba/Ca and Sr/Ca due to Biological Purification of Calcium. Journal of Archaeological Science 26:609616.Google Scholar
Butler, B. Robert. 1969. More Information on the Frozen Ground Features and Further Interpretation of the Small Mammal Sequence at the Wasden Site (Owl Cave), Bonneville County, Idaho. Tebiwa: Journal of the Idaho State University Museum 12(1):5863.Google Scholar
Butler, B. Robert. 1972a. The Holocene in the Desert West and its Cultural Significance. In Great Basin Cultural Ecology: A Symposium, edited by Fowler, Don D., pp. 512. Publications in the Social Sciences. 8. Desert Research Institute, Reno, NV.Google Scholar
Butler, B. Robert. 1972b. The Holocene or Postglacial Ecological Crisis on the Eastern Snake River Plain. Tebiwa: Journal of the Idaho State University Museum 15(1):4963.Google Scholar
Butler, B. Robert. 1976. The Evolution of the Modern Sagebrush–Grass Steppe Biome on the Eastern Snake River Plain. In Holocene Environmental Change in the Great Basin, edited by Elston, Robert, pp. 439. Research Paper 6. Nevada Archeological Survey, Reno.Google Scholar
Butler, B. Robert. 1978. A Guide to Understanding Idaho Archaeology: The Upper Snake and Salmon River Country. Idaho Museum of Natural History, Pocatello.Google Scholar
Butler, Kaylene, Louys, Julien, and Travouillon, Kenny. 2014. Extending Dental Mesowear Analyses to Australian Marsupials, with Applications to Six Plio-Pleistocene Kangaroos from Southeast Queensland. Palaeogeography, Palaeoclimatology, Palaeoecology 408:1125.Google Scholar
Butler, P. M. 1952. The Milk Molars of Perissodactyla, with Remarks on Molar Occlusion. Proceedings of the Zoological Society of London 121:777817.Google Scholar
Butler, Virginia L. 1988. Fish Feeding Behaviour and Fish Capture: The Case for Variation in Lapita Fishing Strategies. Archaeology in Oceania 29:8190.Google Scholar
Byers, John A. 1997. American Pronghorn: Social Adaptations and the Ghosts of Predators Past. University of Chicago Press.Google Scholar
Calaby, J. H. 1971. Man, Fauna, and Climate in Aboriginal Australia. In Aboriginal Man and Environment in Australia, edited by Mulvaney, D. J. and Golson, J., pp. 8093. Australian National University Press, Canberra.Google Scholar
Calandra, Ivan, and Merceron, Gildas. 2016. Dental Microwear Texture Analysis in Mammalian Ecology. Mammal Review 46:215228.Google Scholar
Campbell, B. D., and Smith, D. M.. 2000. A Synthesis of Recent Global Change Research on Pasture and Rangeland Production: Reduced Uncertainties and their Management Implications. Agriculture, Ecosystems and Environments 82:3955.Google Scholar
Campbell, Timothy L., Lewis, Patrick J., and Williams, Justin K.. 2011. Analysis of the Modern Distribution of South African Gerbilliscus (Rodentia: Gerbillinae) with Implications for Plio-Pleistocene Palaeoenvironmental Reconstruction. South African Journal of Science 107:Art. #497.Google Scholar
Campbell, Timothy L., Lewis, Patrick J., Thies, Monte L., and Williams, Justin K.. 2012. A Geographic Information Systems (GIS)-Based Analysis of Modern South African Rodent Distributions, Habitat Use, and Environmental Tolerances. Ecology and Evolution 2:28812894.Google Scholar
Cannon, Michael D. 1999. A Mathematical Model of the Effects of Screen Size on Zooarchaeological Relative Abundance Measures. Journal of Archaeological Science 26:205214.Google Scholar
Cannon, Michael D. 2003. A Model of Central Place Forager Prey Choice and an Application to Faunal Remains from Mimbres Valley, New Mexico. Journal of Anthropological Archaeology 22:125.Google Scholar
Cannon, Michael D. 2004. Geographic Variability in North American Mammal Community Richness during the Terminal Pleistocene. Quaternary Science Reviews 23:10991123.Google Scholar
Cannon, Michael D. 2013. NISP, Bone Fragmentation, and the Measurement of Taxonomic Abundance. Journal of Archaeological Method and Theory 20:397419.Google Scholar
Caporale, Salvatore S., and Ungar, Peter S.. 2016. Rodent Incisor Microwear as a Proxy for Ecological Reconstruction. Palaeogeography, Palaeoclimatology, Palaeoecology 446:225233.Google Scholar
Cartmill, Matt. 1967. The Early Pleistocene Mammalian Microfaunas of Sub-Saharan Africa and their Ecological Significance. Quaternaria 9:169198.Google Scholar
Caruso, Nicholas M., Sears, Michael D., Adams, Dean C., and Lips, Karen R.. 2015. Widespread Rapid Reductions in Body Size of Adult Salamanders in Response to Climate Change. Global Change Biology 20:17511759.Google Scholar
Casanovas-Vilar, Isaac, and Agusti, Jordi. 2007. Ecogeographical Stability and Climate Forcing in the Late Miocene (Vallesian) Rodent Record of Spain. Palaeogeography, Palaeoclimatology, Palaeoecology 248:169189.Google Scholar
Case, E. C. 1921. Criteria for the Determination of the Climatic Environment of Extinct Animals. Geological Society of America Bulletin 32:333338.Google Scholar
Case, E. C. 1936. Paleoecology of the Vertebrates. In Report of the Committee on Paleoecology, 1935–1936, edited by Twenhofel, W. H, pp. 1021. National Research Council, Division of Geology and Geography, Washington, DC.Google Scholar
Cassiliano, Michael L. 1994. Paleoecology and Taphonomy of Vertebrate Faunas from the Anza-Borrego Desert of California. Unpublished Ph.D. dissertation, Department of Geosciences, University of Arizona, Tucson.Google Scholar
Castaños, Jone, Castaños, Pedro, Murelaga, Xabier et al. 2014. Osteometric Analysis of the Scapula and Humerus of Rangifer tarandus and Cervus elaphus: A Contribution to the Discrimination of Late Pleistocene Cervids. Acta Palaeontologica Polonica 59:779786.Google Scholar
Cerling, Thure E., and Harris, John M.. 1999. Carbon Isotope Fractionation between Diet and Bioapatite in Ungulate Mammals and Implications for Ecological and Paleoecological Studies. Oecologia 120:347363.Google Scholar
Cerling, Thure E., Wang, Yang, and Quade, Jay. 1993. Expansion of C4 Ecosystems as an Indicator of Global Ecological Change in the Late Miocene. Nature 361:344345.Google Scholar
Cerling, Thure E., Harris, John M., MacFadden, Bruce J. et al. 1997. Global Vegetation Change through the Miocene–Pliocene Boundary. Nature 389:153158.Google Scholar
Cerling, Thure E., Harris, John M., and Passey, Benjamin H.. 2003. Diets of East African Bovidae Based on Stable Isotope Analysis. Journal of Mammalogy 84:456470.Google Scholar
Cerling, Thure E., Andanje, Samuel A., Blumenthal, Scott A. et al. 2015. Dietary Changes of Large Herbivores in the Turkana Basin, Kenya from 4 to 1 Ma. Proceedings of the National Academy of Sciences USA 112:1146711472.Google Scholar
Chaline, J. 1977. Rodents, Evolution, and Prehistory. Endeavor n.s. 1(2):4451.Google Scholar
Chang, Jie Christine, Shulmeister, James, and Woodward, Craig. 2015. A Chironomid Based Transfer Function for Reconstructing Summer Temperatures in Southeastern Australia. Palaeogeography, Palaeoclimatology, Palaeoecology 423:109121.Google Scholar
Chapin, F. Stuart, III, and Starfield, Anthony M.. 1997. Time Lags and Novel Ecosystems in Response to Transient Climatic Change in Alaska. Climate Change 35:449461.Google Scholar
Chase, Brian M., and Meadows, M. E.. 2007. Late Quaternary Dynamics of Southern Africa’s Winter Rainfall Zone. Earth-Science Reviews 84:103138.Google Scholar
Chase, Brian M., Faith, J. Tyler, Mackay, Alex et al. 2018. Climatic Controls on Later Stone Age Human Adaptation in Africa’s Southern Cape. Journal of Human Evolution 114:3544.Google Scholar
Chase, Jonathan M., and Leibold, Matthew A.. 2002. Spatial Scale Dictates the Productivity–Biodiversity Relationship. Nature 416:427430.Google Scholar
Cheatum, E. P., and Allen, Don. 1964. Limitations in Paleoecological Reconstruction Utilizing Data from Fossil Non-Marine Molluscs. In The Reconstruction of Past Environments: Proceedings, assembled by Hester, James J. and Schoenwetter, James, pp. 3133. Publication 3. Fort Burgwin Research Center, Taos, NM.Google Scholar
Chetboun, R., and Tchernov, Etian. 1983. Temporal and Spatial Morphological Variation in Meriones tristrami (Rodentia: Gerbillidae) from Israel. Israel Journal of Zoology 32:6390.Google Scholar
Churcher, C. S., and Wilson, M[ichael] C.. 1990. Methods in Quaternary Ecology #12: Vertebrates. Geoscience Canada 17:5978.Google Scholar
Clark, William C. 1985. Scales of Climate Impacts. Climatic Change 7:527.Google Scholar
Cleghorn, Naomi, and Marean, Curtis W.. 2004. Distinguishing Selective Transport and In Situ Attrition: A Critical Review of Analytical Approaches. Journal of Taphonomy 2:4367.Google Scholar
Cleghorn, Naomi, and Marean, Curtis W.. 2007. The Destruction of Skeletal Elements by Carnivores: The Growth of a General Model for Skeletal Element Destruction and Survival in Zooarchaeological Assemblages. In Breathing Life into Fossils: Taphonomic Studies in Honor of C. K. (Bob) Brain, edited by Pickering, Travis R., Schick, Kathy, and Toth, Nicholas, pp. 3766. Publication Series. 2. Stone Age Institute Press, Gosport, IN.Google Scholar
Cleland, Carol E. 2001. Historical Science, Experimental Science, and the Scientific Method. Geology 29:987990.Google Scholar
Cleland, Carol E. 2011. Prediction and Explanation in Historical Natural Science. British Journal for the Philosophy of Science 62:551582.Google Scholar
Cleland, Charles E. 1966. The Prehistoric Animal Ecology and Ethnozoology of the Upper Great Lakes Region. Anthropological Papers 29. Museum of Anthropology, University of Michigan, Ann Arbor.Google Scholar
Clements, Frederic E. 1936. Nature and the Structure of Climax. Journal of Ecology 24:252284.Google Scholar
Clements, Frederic E., and Shelford, Victor. 1939. Bio-Ecology. Wiley, New York.Google Scholar
Clyde, William C., and Gingerich, Philip D.. 1994. Rates of Evolution in the Dentition of Early Eocene Cantius: Comparison of Size and Shape. Paleobiology 20:506522.Google Scholar
Codding, Brian F., Bird, Douglas W., and Bird, Rebecca Bliege. 2010a. Interpreting Abundance Indices: Some Zooarchaeological Implications of Martu Foraging. Journal of Archaeological Science 37:32003210.Google Scholar
Codding, Brian F., Porcasi, Judith F., and Jones, Terry L.. 2010b. Explaining Prehistoric Variation in the Abundance of Large Prey: A Zooarchaeological Analysis of Deer and Rabbit Hunting along the Pecho Coast of Central California. Journal of Anthropological Archaeology 29:4761.Google Scholar
Cole, Kenneth L. 1995. Equable Climates, Mixed Assemblages, and the Regression Fallacy. In Late Quaternary Environments and Deep History: A Tribute to Paul S. Martin, edited by Steadman, David W. and Mead, Jim I., pp. 131138. Scientific Papers 3. Mammoth Site of Hot Springs, Hot Springs, SD.Google Scholar
Colwell, Robert K., and Rangel, Thiago F.. 2009. Hutchinson’s Duality: The Once and Future Niche. Proceedings of the National Academy of Sciences USA 106 (Supplement 2):1965119658.Google Scholar
Colwell, Robert K., Chao, Anne, Gotelli, Nicholas J. et al. 2012. Models and Estimators Linking Individual-Based and Sample-Based Rarefaction, Extrapolation and Comparison of Assemblages. Journal of Plant Ecology 5:321.Google Scholar
Coope, G. R. 1959. A Late Pleistocene Insect Fauna from Chelford, Cheshire. Proceedings of the Royal Society B 151:7086.Google Scholar
Coope, G. R. 1986. Coleoptera Analysis. In Handbook of Palaeoecology and Palaeohydrology, edited by Bergland, B. E, pp. 703713. John Wiley and Sons, Chichester.Google Scholar
Coppens, Yves, Howell, F. Clark, Isaac, Glynn Ll., and Leakey, Richard E. F (editors). 1976. Earliest Man and Environments in the Lake Rudolf Basin: Stratigraphy, Paleoecology, and Evolution. University of Chicago Press.Google Scholar
Cormie, A. B., Luz, B., and Schwarcz, H. P.. 1994. Relationship between the Hydrogen and Oxygen Isotopes of Deer Bone and their Use in the Estimation of Relative Humidity. Geochimica et Cosmochimica Acta 58:34393449.Google Scholar
Costeur, Loïc. 2004. Cenogram Analysis of the Rudabánya Mammalian Community: Palaeonvironmental Interpretations. Paleontographica Italica 90:303307.Google Scholar
Costeur, Loïc, and Legendre, Serge. 2008. Mammalian Communities Document a Latitudinal Environmental Gradient during the Miocene Climatic Optimum in Western Europe. Palaios 23:280288.Google Scholar
Craig, G. Y. 1961. Palaeozoological Evidence of Climate: (2) Invertebrates. In Descriptive Palaeoclimatology, edited by Nairen, A. E. M, pp. 207226. Interscience Publishers, New York.Google Scholar
Creighton, G. Ken. 1980. Static Allometry of Mammalian Teeth and the Correlation of Tooth Size and Body Size in Contemporary Mammals. Journal of Zoology (London) 191:435443.Google Scholar
Croft, Darin A. 2001. Cenozoic Environmental Change in South America as Indicated by Mammalian Body Size Distributions (Cenograms). Diversity and Distributions 7:271287.Google Scholar
Croft, Darin A., and Weinstein, Deborah. 2008. The First Application of the Mesowear Method to Endemic South American Ungulates (Notoungulata). Palaeogeography, Palaeoclimatology, Palaeoecology 269:103114.Google Scholar
Cruz-Uribe, Kathryn. 1983. The Mammalian Fauna from Redcliff Cave, Zimbabwe. South African Archaeological Bulletin 38:716.Google Scholar
Cruz-Uribe, Kathryn. 1988. The Use and Meaning of Species Diversity and Richness in Archaeological Faunas. Journal of Archaeological Science 15:179196.Google Scholar
Cuenca-Bescós, G[loria], Rofes, J., and Garcia-Pimienta, J.. 2005. Environmental Change across the Early–Middle Pleistocene Transition: Small Mammalian Evidence from the Trinchera Dolina Cave, Atapuerca, Spain. In Early–Middle Pleistocene Transitions: The Land–Ocean Evidence, edited by Head, M. J and Gibbard, P. L, pp. 277286. Special Publication 247. Geological Society of London.Google Scholar
Cuenca-Bescós, Gloria, Straus, Lawrence G., Morales, Manuel R. Gonzálex, and Pimienta, Juan C. García. 2009. The Reconstruction of Past Environments through Small Mammals: From the Mousterian to the Bronze Age in El Mirón Cave (Cantabria, Spain). Journal of Archaeological Science 36:947955.Google Scholar
Curran, Sabrina C. 2012. Expanding Ecomorphological Methods: Geometric Morphometric Analysis of Cervidae Post-Crania. Journal of Archaeological Science 39:11721182.Google Scholar
Cutler, Alan H., Behrensmeyer, Anna K., and Chapman, Ralph E.. 1999. Environmental Information in a Recent Bone Assemblage: Roles of Taphonomic Processes and Ecological Change. Palaeogeography, Palaeoclimatology, Palaeoecology 149:359372.Google Scholar
Cuvier, G. 1796. Notice sur le Squelette d’une Très-Grande Espèce de Quadrupède Inconnue jusqu’à Présent, Trouvé au Paraguay, et Déposé au Cabinet d’Histoire Naturelle de Madrid. Magasin Encyclopédique 1:303310.Google Scholar
Daams, Remmert, van der Meulen, Albert J., Peláez-Campomanes, Pablo, and Alvarez-Sierra, Maria A.. 1999. Trends in Rodent Assemblages from the Aragonian (Early–Middle Miocene) of the Calatayud-Daroca Basin, Aragón, Spain. In Hominoid Evolution and Climatic Change in Europe, vol. i: The Evolution of Neogene Terrestrial Ecosystems in Europe, edited by Agustí, Jordi, Rook, Lorenzo, and Andrews, Peter, pp. 127139. Cambridge University Press.Google Scholar
Dalquest, Walter W. 1965. New Pleistocene Formation and Local Fauna from Hardeman County, Texas. Journal of Paleontology 39:6379.Google Scholar
Daly, Patricia. 1969. Approaches to Faunal Analysis in Archaeology. American Antiquity 34:146153.Google Scholar
Damuth, John D. 1981. Population Density and Body Size in Mammals. Nature 290:699700.Google Scholar
Damuth, John D. 1982. Analysis of the Preservation of Community Structure in Assemblages of Fossil Mammals. Paleobiology 8:434446.Google Scholar
Damuth, John D. 1992. Taxon-Free Characterisation of Animal Communities. In Terrestrial Ecosystems through Time: Evolutionary Paleoecology of Terrestrial Plants and Animals, edited by Behrensmeyer, Anna K., Damuth, John D., DiMichele, William A., Potts, Richard, Sues, Hans-Dieter, and Wing, Scott L., pp. 183203. University of Chicago Press.Google Scholar
Damuth, John D., and Fortelius, Mikael. 2001. Reconstructing Mean Annual Precipitation, Based on Mammalian Dental Morphology and Local Species Richness. In Eeden Plenary Workshop on Late Miocene to Early Pliocene Environments and Ecosystems, edited by Agustí, J[ordi and Oms, O., pp. 2324. European Science Foundation, Sabadell, Spain.Google Scholar
Damuth, John D., and Janis, Christine M.. 2011. On the Relationship between Hypsodonty and Feeding Ecology in Ungulate Mammals, and its Utility in Palaeoecology. Biological Reviews 86:733758.Google Scholar
Damuth, John D., and MacFadden, Bruce J. (editors). 1990. Body Size in Mammalian Paleobiology: Estimation and Biological Implications. Cambridge University Press.Google Scholar
Damuth, John D., Fortelius, Mikael, Andrews, Peter et al. 2002. Reconstructing Mean Annual Precipitation Based on Mammalian Dental Morphology and Local Species Richness. Journal of Vertebrate Paleontology 22 (Supplement 3):48A.Google Scholar
Darlington, Philip J., Jr. 1957. Zoogeography: The Geographical Distribution of Animals. John Wiley and Sons, New York.Google Scholar
Darwin, Charles. 1859. On the Origin of Species by Means of Natural Selection, or the Preservation of Favoured Races in the Struggle for Life. John Murray, London.Google Scholar
Daubenmire, Rexford F. 1938. Merriam’s Life Zones of North America. Quarterly Review of Biology 13:317332.Google Scholar
Dauphin, Y., Kowalski, C., and Denys, C.. 1994. Assemblage Data and Bone and Teeth Modifications as an Aid to Paleoenvironmental Interpretations of the Open-Air Pleistocene Site of Tighenif (Algeria). Quaternary Research 42:340349.Google Scholar
Davies, Jeremy. 2016. The Birth of the Anthropocene. University of California Press, Berkeley.Google Scholar
Davies, T. Jonathan, Purvis, Andy, and Gittleman, John L.. 2009. Quaternary Climate Change and the Geographic Ranges of Mammals. American Naturalist 174:297307.Google Scholar
Davis, Edward Byrd, and Pyenson, Nicholas D.. 2007. Diversity Biases in Terrestrial Mammalian Assemblages and Quantifying the Differences between Museum Collections and Published Accounts: A Case Study from the Miocene of Nevada. Palaeogeography, Palaeoclimatology, Palaeoecology 250:139149.Google Scholar
Davis, Margaret B. 1976. Pleistocene Biogeography of Temperate Deciduous Forests. In Ecology of the Pleistocene: A Symposium, edited by West, R. C and Haag, W. G, pp. 1326. Geoscience and Man 13. Louisiana State University, Baton Rouge.Google Scholar
Davis, Margaret B. 1981. Quaternary History and the Stability of Forest Communities. In Forest Succession: Concepts and Application, edited by West, Darrell C., Shugart, Herman H., and Botkin, Daniel B., pp. 132153. Springer, New York.Google Scholar
Davis, Margaret B., and Shaw, Ruth G.. 2001. Range Shifts and Adaptive Responses to Quaternary Climate Change. Science 292:637679.Google Scholar
Davis, Margaret B., Shaw, Ruth G., and Etterson, Julie R.. 2005. Evolutionary Responses to Changing Climate. Ecology 86:17041714.Google Scholar
Davis, Matt, and Pineda-Munoz, Silvia. 2016. The Temporal Scale of Diet and Dietary Proxies. Ecology and Evolution 6:18831897.Google Scholar
Davis, Simon J. M. 1977. Size Variation of the Fox, Vulpes vulpes in the Palaearctic Region Today, and in Israel during the Late Quaternary. Journal of Zoology (London) 182:343351.Google Scholar
Davis, Simon J. M. 1981. The Effects of Temperature Change and Domestication on the Body Size of Late Pleistocene to Holocene Mammals of Israel. Paleobiology 7:101104.Google Scholar
Dawkins, W. Boyd. 1869. On the Distribution of British Postglacial Mammals. Quarterly Journal of the Geological Society of London 25:192217.Google Scholar
Dawkins, W. Boyd. 1871. On Pleistocene Climate and the Relation of the Pleistocene Mammalia to the Glacial Period. Popular Science Review 10:377397.Google Scholar
Dawkins, W. Boyd. 1874. Cave Hunting, Researches on the Evidence of Caves Respecting the Early Inhabitants of Europe. Macmillan, London.Google Scholar
Dawson, Terence P., Jackson, Stephen T., House, Joanna I., Prentice, Iain Colin, and Mace, Georgina M.. 2011. Beyond Predictions: Biodiversity Conservation in a Changing Climate. Science 332:5358.Google Scholar
Dayan, Tamar, Tchernov, Etian, Yom-Tov, Yoram and Simberloff, Daniel. 1989. Ecological Character Displacement in Saharo-Arabian Vulpes: Outfoxing Bergmann’s Rule. Oikos 55:263272.Google Scholar
Dayan, Tamar, Simberloff, Daniel, Tchernov, Etian, and Yom-Tov, Yoram. 1991. Calibrating the Paleothermometer: Climate, Communities, and the Evolution of Size. Paleobiology 17:189199.Google Scholar
Deacon, H. J. 1979. Excavations at Boomplaas Cave: A Sequence through the Upper Pleistocene and Holocene in South Africa. World Archaeology 10:241257.Google Scholar
Deacon, H. J., and Brooker, Mary. 1976. The Holocene and Upper Pleistocene Sequence in the Southern Cape. Annals of the South African Museum 71:203214.Google Scholar
Deacon, H. J., Deacon, Janette, and Brooker, Mary. 1976. Four Painted Stones from Boomplaas Cave, Oudtshoorn District. South African Archaeological Bulletin 31:141145.Google Scholar
Deacon, H. J., Deacon, Janette, Brooker, Mary, and Wilson, M. L.. 1978. The Evidence for Herding at Boomplaas Cave in the Southern Cape, South Africa. South African Archaeological Bulletin 33:3965.Google Scholar
Deacon, H. J., Scholtz, A., and Daitz, L. D.. 1983. Fossil Charcoals as a Source of Palaeoecological Information in the Fynbos Region. In Fynbos Palaeoecology: A Preliminary Synthesis, South African National Scientific Programmes Report No. 75, edited by Deacon, H. J, Hendey, Q. B. and Lambrechts, J. J. N, pp. 174182. Mills Litho, Cape Town.Google Scholar
Deacon, H. J., Deacon, Janette, Scholtz, A., Thackeray, J. F., and Brink, James S.. 1984. Correlation of Palaeoenvironmental Data from the Late Pleistocene and Holocene Deposits at Boomplaas Cave, Southern Cape. In Late Cainozoic Palaeoclimates of the Southern Hemisphere, edited by Vogel, J. C, pp. 339351. Balkema, Rotterdam.Google Scholar
Deacon, Hilary J. 1995. Two Late Pleistocene–Holocene Archaeological Depositories from the Southern Cape, South Africa. South African Archaeological Bulletin 50:121131.Google Scholar
Deacon, Janette. 1984. The Later Stone Age of Southernmost Africa. British Archaeological Reports, International Series 213. BAR, Oxford.Google Scholar
Deacon, Janette, and Lancaster, N.. 1988. Late Quaternary Palaeoenvironments of Southern Africa. Oxford University Press.Google Scholar
De Graaf, G. 1961. A Preliminary Investigation of the Mammalian Microfauna in Pleistocene Deposits of Caves in the Transvaal System. Palaeontologia Africana 7:59117.Google Scholar
DeGusta, David, and Vrba, Elisabeth S.. 2003. A Method for Inferring Paleohabitats from the Functional Morphology of Astragali. Journal of Archaeological Science 30:10091022.Google Scholar
DeGusta, David, and Vrba, Elisabeth S.. 2005a. Methods for Inferring Paleohabitats from Discrete Traits of the Bovid Postcranial Skeleton. Journal of Archaeological Science 32:11151123.Google Scholar
DeGusta, David, and Vrba, Elisabeth S.. 2005b. Methods for Inferring Paleohabitats from the Functional Morphology of Bovid Phalanges. Journal of Archaeological Science 32:10991113.Google Scholar
Delcourt, Hazel R., and Delcourt, Paul A.. 1988. Quaternary Landscape Ecology: Relevant Scales in Space and Time. Landscape Ecology 2:2344.Google Scholar
Delcourt, Hazel R., Delcourt, Paul A., and Webb, Thompson III. 1983. Dynamic Plant Ecology: The Spectrum of Vegetational Change in Space and Time. Quaternary Science Reviews 1:153175.Google Scholar
Demarais, Stephen, Miller, Karl V., and Jacobson, Harry A.. 2000. White-Tailed Deer. In Ecology and Management of Large Mammals in North America, edited by Demarais, Stephen and Krausman, Paul R., pp. 601628. Prentice Hall, Upper Saddle River, NJ.Google Scholar
DeMenocal, Peter, Ortiz, Joseph, Guilderson, Tom et al. 2000. Abrupt Onset and Termination of the African Humid Period: Rapid Climate Responses to Gradual Insolation Forcing. Quaternary Science Reviews 19:347361.Google Scholar
Demirel, Arzu, Andrews, Peter, Yalcinkay, Isin, and Ersoy, Ayhan. 2011. The Taphonomy and Palaeoenvironmental Implications of the Small Mammals from Karain Cave, Turkey. Journal of Archaeological Science 38:30483059.Google Scholar
Deng, Tao. 2009. Late Cenozoic Environmental Changes in the Linxia Basin (Gansu, China) as Indicated by Cenograms of Fossil Mammals. Vertebrata PalAsiatica 10:282298.Google Scholar
DeNiro, Michael J., and Epstein, Samuel. 1978. Influence of Diet on the Distribution of Carbon Isotopes in Animals. Geochimica et Cosmochimica Acta 42:495506.Google Scholar
DeNiro, Michael J., and Epstein, Samuel. 1981. Influence of Diet on the Distribution of Nitrogen Isotopes in Animals. Geochimica et Cosmochimica Acta 45:341351.Google Scholar
Denys, C., Dauphin, Y., Rzebik-Kowalski, B., and Kowalski, K.. 1996. Taphonomic Study of Algerian Owl Pellet Assemblages and Differential Preservation of Some Rodents: Palaeontological Implications. Acta Zoologica Cracoviensia 39:103116.Google Scholar
de Ruiter, Darryl J., Sponheimer, Matt, and Lee-Thorp, Julia A.. 2008. Indications of Habitat Association of Australopithecus robustus in the Bloubank Valley, South Africa. Journal of Human Evolution 55:10151030.Google Scholar
DeSantis, Larisa R. G. 2016. Dental Microwear Textures: Reconstructing Diets of Fossil Mammals. Surface Topography: Metrology and Properties 4:023002.Google Scholar
DeSantis, Larisa R. G., and Haupt, Ryan J.. 2014. Cougars’ Key to Survival through the Late Pleistocene Extinction: Insights from Dental Microwear Texture Analysis. Biology Letters 10:20140203.Google Scholar
DeSantis, Larisa R. G., Schubert, Blaine W., Scott, Jessica R., and Ungar, Peter S.. 2012a. Implications of Diet for the Extinction of Saber-Toothed Cats and American Lions. PLoS ONE 7:e52453.Google Scholar
DeSantis, Larisa R. G., Beavins Tracy, Rachel A., Koontz, Cassandra S., Roseberry, John C., and Velasco, Matthew C.. 2012b. Mammalian Niche Conservation through Deep Time. PLOS One 7:e35624.Google Scholar
Dietl, Gregory P., and Flessa, Karl W. (editors). 2009. Conservation Paleobiology: Using the Past to Manage for the Future. Paleontological Society Papers 15. Paleontological Society, Boulder, CO.Google Scholar
Dietl, Gregory P., and Flessa, Karl W. 2011. Conservation Paleobiology: Putting the Dead to Work. Trends in Ecology and Evolution 26:3037.Google Scholar
Dietl, Gregory P., Kidwell, Susan, Brenner, Mark et al. 2012. Conservation Paleobiology: Opportunities for the Earth Sciences. Paleontological Research Institute, Ithaca, NY.Google Scholar
Dietl, Gregory P., Kidwell, Susan M., Brenner, Mark et al. 2015. Conservation Paleobiology: Leveraging Knowledge of the Past to Inform Conservation and Restoration. Annual Review of Earth and Planetary Sciences 43:79103.Google Scholar
Diffenbaugh, Noah S., Ashfaq, Moetasim, Shuman, Bryan, Williams, John W., and Bartlein, Patrick J.. 2006. Summer Aridity in the United States: Response to Mid-Holocene Changes in Insolation and Sea Surface Temperature. Geophysical Research Letters 33:L22712.Google Scholar
DiMichele, W. A., Behrensmeyer, A. K., Olszewski, T. D. et al. 2004. Long-Term Stasis in Ecological Assemblages: Evidence from the Fossil Record. Annual Review of Ecology, Evolution, and Systematics 35:285322.Google Scholar
Dodd, J. Robert, and Stanton, Robert J., Jr. 1981. Paleoecology: Concepts and Applications. John Wiley and Sons, New York.Google Scholar
Dodd, J. Robert, and Stanton, Robert J., 1990. Paleoecology: Concepts and Applications, second edition. John Wiley and Sons, New York.Google Scholar
Dodson, Peter. 1973. The Significance of Small Bones in Paleoecological Interpretation. University of Wyoming Contributions to Geology 12:1519.Google Scholar
Domínguez-Rodrigo, Manuel. 2012. Critical Review of the MNI (Minimum Number of Individuals) as a Zooarchaeological Unit of Quantification. Archaeological and Anthropological Sciences 4:4759.Google Scholar
Domínguez-Rodrigo, Manuel, and Musiba, Charles M.. 2010. How Accurate are Paleoecological Reconstructions of Early Paleontological and Archaeological Sites? Evolutionary Biology 37:128140.Google Scholar
Donohue, Shelly L., DeSantis, Larisa R. G., Schubert, Blaine W., and Ungar, Peter S.. 2013. Was the Giant Short-Faced Bear a Hyper-Scavenger? A New Approach to the Dietary Study of Ursids Using Dental Microwear Textures. PLoS ONE 8:e77531.Google Scholar
Dorf, E. 1959. Climatic Changes of the Past and Present. Contributions from the Museum of Paleontology 13:181210. University of Michigan, Ann Arbor.Google Scholar
Dortch, Joe, and Wright, Richard. 2010. Identifying Palaeo-Environments and Changes in Aboriginal Subsistence from Dual-Patterned Faunal Assemblages, South-Western Australia. Journal of Archaeological Science 37:10531064.Google Scholar
Driesch, Angela von den. 1976. A Guide to the Measurement of Animal Bones from Archaeological Sites. Peabody Museum of Archaeology and Ethnology Bulletin 1. Harvard University, Cambridge, MA.Google Scholar
Driesch, Angela von den, and Deacon, Hilary J.. 1985. Sheep Remains from Boomplaas Cave, South Africa. South African Archaeological Bulletin 40:3944.Google Scholar
Driver, Jonathan C. 1992. Identification, Classification and Zooarchaeology. Circaea 9:3547.Google Scholar
Driver, Jonathan C. 2001. Paleoecological and Archaeological Implications of the Charlie Lake Cave Fauna, British Columbia. In People and Wildlife in Northern North America: Essays in Honor of R. Dale Guthrie, edited by Gerlach, S. Craig and Murray, Maribeth S., pp. 1322. British Archaeological Reports, International Series 944. BAR, Oxford.Google Scholar
Driver, Jonathan C. 2011. Identification, Classification and Zooarchaeology. Ethnobiology Letters 2:1939.Google Scholar
Driver, Jonathan C., and Woiderski, Joshua R.. 2008. Interpretation of the “Lagomorph Index” in the American Southwest. Quaternary International 185:311.Google Scholar
Duffield, Lathel F. 1973. Aging and Sexing the Post-Cranial Skeleton of Bison. Plains Anthropologist 18:132139.Google Scholar
Dulian, James J. 1975. Paleoecology of the Brayton Local Biota, Late Wisconsinan of Southwestern Iowa. Unpublished Master of Science thesis, University of Iowa, Iowa City.Google Scholar
Eastham, L. C., Feranec, R. S., and Begun, D. R.. 2016. Stable Isotopes Show Resource Partitioning among the Early Late Miocene Herbivore Community at Rudabánya II: Paleoenvironmental Implications for the Hominoid, Rudapithecus hungaricus. Palaeogeography, Palaeoclimatology, Palaeoecology 454:161174.Google Scholar
Eckert, C. G., Samis, K. E., and Lougheed, S. C.. 2008. Genetic Variation across Species’ Geographical Ranges: The Central–Marginal Hypothesis and Beyond. Molecular Ecology 17:11701188.Google Scholar
Efremov, J. A. 1940. Taphonomy: New Branch of Paleontology. Pan-American Geologist 74 :8193.Google Scholar
Ehleringer, James R., Sage, R. F., Flanagan, L. B., and Pearcy, R. W.. 1991. Climate Change and the Evolution of C4 Photosynthesis. Trends in Ecology and Evolution 6:9599.Google Scholar
Ehleringer, James R., Cerling, Thure E., and Helliker, Brent R.. 1997. C4 Photosynthesis, Atmospheric CO2, and Climate. Oecologia 112:285299.Google Scholar
Eisenmann, Véra. 1986. Comparative Osteology of Modern and Fossil Horses, Half-Asses, and Asses. In Equids in the Ancient World, edited by Meadow, Richard H. and Uerpmann, Hans-Peter, pp. 67116. Dr. Ludwig Reichert Verlag, Wiesbaden.Google Scholar
Elias, Robert W., Hirao, Yoshimitsu, and Patterson, Clair C.. 1982. The Circumvention of the Natural Biopurification of Calcium along Nutrient Pathways by Atmospheric Inputs of Industrial Lead. Geochimica et Cosmochimica Acta 46:25612580.Google Scholar
Elias, Scott A. 1997. The Mutual Climatic Range Method of Palaeoclimate Reconstruction Based on Insect Fossils: New Applications and Interhemispheric Comparisons. Quaternary Science Reviews 16:12171225.Google Scholar
Elias, Scott A. 2001. Mutual Climatic Range Reconstructions of Seasonal Temperatures Based on Late Pleistocene Fossil Assemblages in Eastern Beringia. Quaternary Science Reviews 20:7791.Google Scholar
Elias, Scott A., and Whitehouse, Nicki J.. 2014. G. Russell Coope: Papers Honouring his Life and Career. Quaternary International 341:15.Google Scholar
Elias, Scott A., Andrews, John T., and Anderson, Katherine H.. 1999. Insights on the Climatic Constraints on the Beetle Fauna of Coastal Alaska, USA, Derived from the Mutual Climatic Range Method of Palaeoclimate Reconstruction. Arctic and Alpine Research 31:9498.Google Scholar
Elith, Jane, and Leathwick, John R.. 2009. Species Distribution Models: Ecological Explanation and Prediction across Space and Time. Annual Review of Ecology, Evolution, and Systematics 40:677697.Google Scholar
Elton, Charles S. 1927. Animal Ecology. Sedgwick and Jackson, London.Google Scholar
Emerson, Thomas E. 1978. A New Method for Calculating the Live Weight of the Northern White-Tailed Deer from Osteoarchaeological Material. Mid-Continental Journal of Archaeology 3:3544.Google Scholar
Emery, Kitty F., and Thornton, Erin Kennedy. 2008. Zooarchaeological Habitat Analysis of Ancient Maya Landscape Changes. Journal of Ethnobiology 28:154178.Google Scholar
Emery, Kitty F., and Thornton, Erin Kennedy. 2012. Using Animal Remains to Reconstruct Ancient Landscapes and Climate in Central and Southern Maya Lowlands. In Proceedings of the General Session of the 11th International Council for Archaeozoology Conference, edited by Lefèvre, Christine, pp. 203225. British Archaeological Reports, International Series 2354. BAR, Oxford.Google Scholar
Emery, Kitty F., and Thornton, Erin Kennedy. 2014. Tracking Climate Change in the Ancient Maya World through Zooarchaeological Habitat Analysis. In The Great Maya Droughts in Cultural Context: Case Studies in Resilience and Vulnerability, edited by Iannone, Gyles, pp. 301331. University Press of Colorado, Boulder.Google Scholar
Emslie, Steven D. 1982. Osteological Identification of Long-Eared and Short-Eared Owls. American Antiquity 47:155157.Google Scholar
Endler, John A. 1982. Problems in Distinguishing Historical from Ecological Factors in Biogeography. American Zoologist 22:441452.Google Scholar
Erasmus, Barend F. N., Van Jaarswel, Albert S., Chown, Steven L., Kshatriya, Mrigesh, and Wessels, Konrad J.. 2002. Vulnerability of South African Animal Taxa to Climate Change. Global Change Biology 8:679693.Google Scholar
Eronen, Jussi T. 2006. Eurasian Neogene Large Herbivorous Mammals and Climate. Acta Zoologica Fennica 216:172.Google Scholar
Eronen, Jussi T., and Rook, Lorenzo. 2004. The Mio-Pliocene Primate Fossil Record: Dynamics and Habitat Tracking. Journal of Human Evolution 47:323341.Google Scholar
Eronen, Jussi T., Ataabadi, Majid M., Micheels, Arne et al. 2009. Distribution History and Climatic Controls of the Late Miocene Pikermian Chronofauna. Proceedings of the National Academy of Sciences USA 106:1186711871.Google Scholar
Eronen, Jussi T., David Polly, P., Fred, Marianne et al. 2010a. Ecometrics: The Traits that Bind the Past and Present Together. Integrative Zoology 5:88101.Google Scholar
Eronen, J[ussi] T., Puolamäki, K., Liu, L. et al. 2010b. Precipitation and Large Herbivorous Mammals I: Estimates from Present-Day Communities. Evolutionary Ecology Research 12:217233.Google Scholar
Eronen, J[ussi] T., Puolamäki, K., Liu, L. et al. 2010c. Precipitation and Large Herbivorous Mammals II: Application to Fossil Data. Evolutionary Ecology Research 12:235248.Google Scholar
Eronen, J[ussi] T., Janis, C., Chamberlain, C. P., and Mulch, A.. 2015. Mountain Uplift Explains Differences in Palaeogene Patterns of Mammalian Evolution and Extinction between North America and Europe. Proceedings of the Royal Society B 282:20150136.Google Scholar
Ervynck, Anton. 1999. Possibilities and Limitations of the Use of Archaeozoological Data in Biogeographical Analysis: A Review with Examples from the Benelux Region. Belgian Journal of Zoology 129:125138.Google Scholar
Escarguel, Gilles, Fara, Emmanuel, Brayard, Arnaud, and Legendre, Serge. 2011. Biodiversity is Not (and Never Has Been) a Bed of Roses! Comptes Rendus Biologies 334:351359.Google Scholar
Estes, James A. 1996. Predators and Ecosystem Management. Wildlife Society Bulletin 24:390396.Google Scholar
Estes, Richard, and Berberian, Paul. 1970. Paleoecology of a Late Cretaceous Vertebrate Community from Montana. Breviora 343:135.Google Scholar
Etheridge, Richard. 1958. Pleistocene Lizards of the Cragin Quarry Fauna of Meade County, Kansas. Copeia 1958:94101.Google Scholar
Etnier, Michael A. 2002. The Effects of Human Hunting on Northern Fur Seal (Callorhinus ursinus) Migration and Breeding Distributions in the Late Holocene. Unpublished Ph.D. dissertation, Department of Anthropology, University of Washington, Seattle.Google Scholar
Etnier, Michael A. 2004. Reevaluating Evidence of Density-Dependent Growth in Northern Fur Seals (Callorhinus ursinus) Based on Measurements of Archived Skeletal Specimens. Canadian Journal of Fisheries and Aquatic Sciences 61:16161626.Google Scholar
Evans, Alistair R. 2013. Shape Descriptors as Ecometrics in Dental Ecology. Hystrix 24:133140.Google Scholar
Evin, Allowen, Cucchi, Thomas, Escarguel, Gilles et al. 2014. Using Traditional Biometrical Data to Distinguish West Palearctic Wild Boar and Domestic Pigs in the Archaeological Record: New Methods and Standards. Journal of Archaeological Science 43:18.Google Scholar
Fagerstrom, J. A. 1964. Fossil Communities in Paleoecology: Their Recognition and Significance. Geological Society of America Bulletin 75:11971216.Google Scholar
Faith, J. Tyler. 2008. Eland, Buffalo, and Wild Pigs: Were Middle Stone Age Humans Ineffective Hunters? Journal of Human Evolution 55:2436.Google Scholar
Faith, J. Tyler. 2011a. Late Quaternary Dietary Shifts of the Cape Grysbok (Raphicerus melanotis) in Southern Africa. Quaternary Research 75:159165.Google Scholar
Faith, J. Tyler. 2011b. Ungulate Biogeography, Statistical Methods, and the Proficiency of Middle Stone Age Hunters. Journal of Human Evolution 60:315317.Google Scholar
Faith, J. Tyler. 2011c. Ungulate Community Richness, Grazer Extinctions, and Human Subsistence Behavior in Southern Africa’s Cape Floral Region. Palaeogeography, Palaeoclimatology, Palaeoecology 306:219227.Google Scholar
Faith, J. Tyler. 2012a. Conservation Implications of Fossil Roan Antelope (Hippotragus equinus) in Southern Africa’s Cape Floristic Region. In Paleontology in Ecology and Conservation, edited by Louys, Julien, pp. 239251. Springer, Heidelberg.Google Scholar
Faith, J. Tyler. 2012b. Palaeozoological Insights into Management Options for a Threatened Mammal: Southern Africa’s Cape Mountain Zebra (Equus zebra zebra). Diversity and Distributions 18:438447.Google Scholar
Faith, J. Tyler. 2013a. Taphonomic and Paleoecological Change in the Large Mammal Sequence from Boomplaas Cave, Western Cape, South Africa. Journal of Human Evolution 65: 715730.Google Scholar
Faith, J. Tyler. 2013b. Ungulate Diversity and Precipitation History since the Last Glacial Maximum in the Western Cape, South Africa. Quaternary Science Reviews 68:191199.Google Scholar
Faith, J. Tyler. 2014. Late Pleistocene and Holocene Mammal Extinctions on Continental Africa. Earth-Science Reviews 128:105121.Google Scholar
Faith, J. Tyler. 2018. Paleodietary Change and its Implications for Aridity Indices Derived from δ18O of Herbivore Tooth Enamel. Palaeogeography, Palaeoclimatology, Palaeoecology 490:571578.Google Scholar
Faith, J. Tyler, and Du, Andrew. 2018. The Measurement of Taxonomic Evenness in Zooarchaeology. Archaeological and Anthropological Sciences 10:1419–1428.Google Scholar
Faith, J. Tyler, and Behrensmeyer, Anna K.. 2013. Climate Change and Faunal Turnover: Testing the Mechanics of the Turnover-Pulse Hypothesis with South African Fossil Data. Paleobiology 39:609627.Google Scholar
Faith, J. Tyler, and Gordon, Adam D. 2007. Skeletal Element Abundances in Archaeofaunal Assemblages: Economic Utility, Sample Size, and Assessment of Carcass Transport Strategies. Journal of Archaeological Science 34:872882.Google Scholar
Faith, J. Tyler, and O’Connell, James F.. 2011. Revisiting the Late Pleistocene Mammal Extinction Record at Tight Entrance Cave, Southwestern Australia. Quaternary Research 76:397400.Google Scholar
Faith, J. Tyler, Domínguez-Rodrigo, Manuel, and Gordon, Adam D.. 2009. Long-Distance Carcass Transport at Olduvai Gorge? A Quantitative Examination of Bed I Skeletal Element Abundances. Journal of Human Evolution 56:247256.Google Scholar
Faith, J. Tyler, Choiniere, Jonah. N., Tryon, Christian A., Peppe, Daniel J., and Fox, David L.. 2011. Taxonomic Status and Paleoecology of Rusingoryx atopocranion (Mammalia, Artiodactyla), an Extinct Pleistocene Bovid from Rusinga Island, Kenya. Quaternary Research 75:697707.Google Scholar
Faith, J. Tyler, Potts, Richard, Plummer, Thomas W. et al. 2012. New Perspectives on Middle Pleistocene Change in the Large Mammal Faunas of East Africa: Damaliscus hypsodon sp. nov. (Mammalia, Artiodactyla) from Lainyamok, Kenya. Palaeogeography, Palaeoclimatology, Palaeoecology 361–362:8493.Google Scholar
Faith, J. Tyler, Tryon, Christian A., Peppe, Daniel J., and Fox, David L.. 2013. The Fossil History of Grévy’s Zebra (Equus grevyi) in Equatorial East Africa. Journal of Biogeography 40:359369.Google Scholar
Faith, J. Tyler, Tryon, Christian A., Peppe, Daniel J. et al. 2015. Paleoenvironmental Context of the Middle Stone Age Record from Karungu, Lake Victoria Basin, Kenya, and its Implications for Human and Faunal Dispersals in East Africa. Journal of Human Evolution 83:2845.Google Scholar
Faith, J. Tyler, Tryon, Christian A., and Peppe, Daniel J.. 2016a. Environmental Change, Ungulate Biogeography, and their Implications for Early Human Dispersals in Equatorial East Africa. In Africa from MIS 6–2: Population Dynamics and Paleoenvironments, edited by Jones, Sacha C. and Stewart, Brian A., pp. 233245. Springer, Dordrecht.Google Scholar
Faith, J. Tyler, Patterson, David B., Blegen, Nick et al. 2016b. Size Variation in Tachyoryctes splendens (East African Mole-Rat) and its Implications for Late Quaternary Temperature Change in Equatorial East Africa. Quaternary Science Reviews 140:3948.CrossRefGoogle Scholar
Faith, J. Tyler, Dortch, Joe, Jones, Chelsea, Shulmeister, James, and Travouillon, Kenny J.. 2017. Large Mammal Species Richness and Late Quaternary Precipitation Change in Southwestern Australia. Journal of Quaternary Science 32:760769.CrossRefGoogle Scholar
Falk, Carl R., and Semken, Holmes A., Jr. 1998. Taphonomy of Rodent and Insectivore Remains in North American Archaeological Sites: Selected Examples and Interpretations. In Quaternary Paleozoology in the Northern Hemisphere, edited by Saunders, Jeffrey J., Styles, Bonnie W., and Baryshnikov, Gennady F., pp. 285321. Illinois State Museum Scientific Papers 27. Illinois State Museum, Springfield.Google Scholar
Faure, Gunter, and Mensing, Teresa M.. 2004. Isotopes: Principles and Applications, third edition. Wiley, Hoboken, NJ.Google Scholar
Ferguson, Steven H. 2002. The Effects of Productivity and Seasonality on Life History: Comparing Age at Maturity among Moose (Alces alces) Populations. Global Ecology and Biogeography 11:303312.Google Scholar
Fernández-García, Mónica, and López-García, Juan Manuel. 2013. Palaeoecology and Biochronology Based on the Rodents Analysis from the Late Pleistocene/Holocene of Toll Cave (Moià, Barcelona). Spanish Journal of Paleontology 28:227238.CrossRefGoogle Scholar
Fernández-García, Mónica, López-García, Juan Manuel, and Lorenzo, Carlos. 2016. Palaeoecological Implications of Rodents as Proxies for the Late Pleistocene–Holocene Environmental and Climatic Changes in Northeastern Iberia. Comptes Rendus Palevol 15:707719.Google Scholar
Fernández-Jalvo, Yolanda, and Andrews, Peter. 2016. Atlas of Taphonomic Identifications: 1001+ Images of Fossil and Recent Mammal Bone Modification. Springer, Dordrecht.Google Scholar
Fernández-Jalvo, Yolanda, , Christiane Denys, Peter Andrews, Terry Williams, Yanicke Dauphin, , and Humphrey, Louise. 1998. Taphonomy and Palaeoecology of Olduvai Bed-I (Pleistocene, Tanzania). Journal of Human Evolution 34:137172.Google Scholar
Fernández-Jalvo, Y[olanda], Scott, L., and Andrews, P.. 2011. Taphonomy in Palaeoecological Interpretations. Quaternary Science Reviews 30:12961302.Google Scholar
Fichman, Martin. 1977. Wallace: Zoogeography and the Problem of Landbridges. Journal of the History of Biology 10:4563.Google Scholar
Fick, Stephen E., and Hijmans, Robert J.. 2017. WorldClim 2: New 1-km Spatial Resolution Climate Surfaces for Global Land Areas. International Journal of Climatology 37:43025315Google Scholar
Figueirido, Borja, Palmqvist, Paul, and Pérez-Claros, J. A.. 2009. Ecomorphological Correlates of Craniodental Variation in Bears and Paleobiological Implications for Extinct Taxa: An Approach Based on Geometric Morphometrics. Journal of Zoology (London) 277:7080.Google Scholar
Findley, James S. 1964. Paleoecological Reconstruction: Vertebrate Limitations. In The Reconstruction of Past Environments: Proceedings, assembled by Hester, James J. and Schoenwetter, James, pp. 2325. Publication 3. Fort Burgwin Research Center, Taos, NM.Google Scholar
Fisher, Jacob L. 2012. Shifting Prehistoric Abundances of Leporids at Five Finger Ridge, a Central Utah Archaeological Site. Western North American Naturalist 72:6068.Google Scholar
Fisher, Jacob L., and Valentine, Benjamin. 2013. Resource Depression, Climate Change, and Mountain Sheep in the Eastern Great Basin of Western North America. Archaeological and Anthropological Sciences 5:145157.Google Scholar
Flannery, Kent V. 1967. Vertebrate Fauna and Hunting Patterns. In The Prehistory of the Tehuacan Valley, vol. i: Environment and Subsistence, edited by Byers, Douglas S., pp. 132177. University of Texas Press, Austin.Google Scholar
Fleagle, John G. 1978. Size Distributions of Living and Fossil Primate Faunas. Paleobiology 4:6776.Google Scholar
Fleming, Theodore H. 1973. Number of Mammal Species in North and Central American Forest Communities. Ecology 54:555563.Google Scholar
Flynn, Lawrence J. 2003. Small Mammal Indicators of Forest Paleo-Environment in the Siwalik Deposits of the Potwar Plateau, Pakistan. In Distribution and Migration of Tertiary Mammals in Eurasia: A Volume in Honour of Hans de Bruun, edited by Reumer, Jelle W. F and Wessels, Wilma, pp. 183196. Deinsea 10. Natural History Museum, Rotterdam.Google Scholar
Fortelius, Mikael 1985. Ungulate Cheek Teeth: Developmental, Functional, and Evolutionary Interrelations. Acta Zoologica Fennica 180:176.Google Scholar
Fortelius, Mikael 2003. Evolution of Dental Capability in Western Eurasian Large Mammal Plant Eaters 22–2 Million Years Ago: A Case for Environmental Forcing Mediated by Biotic Processes. In The New Panorama of Animal Evolution, edited by Legakis, A., Sfenthourakis, S., Polymeni, R., and Thessalou-Legaki, M., pp. 6167. Pensoft, Sofia and Moscow.Google Scholar
Fortelius, Mikael, and Solounias, Nikos. 2000. Functional Characterization of Ungulate Molars Using the Abrasion–Attrition Wear Gradient: A New Method for Reconstructing Paleodiets. American Museum Novitates 3301:136.Google Scholar
Fortelius, Mikael, Eronen, Jussi T., Jernvall, Jukka et al. 2002. Fossil Mammals Resolve Regional Patterns of Eurasian Climate Change over 20 Million Years. Evolutionary Ecology Research 4:10051016.Google Scholar
Fortelius, Mikael, Eronen, Jussi T., Liu, Liping et al. 2003. Continental-Scale Hypsodonty Patterns, Climatic Palaeobiogeography, and Dispersal of Eurasian Neogene Large Mammal Herbivores. In Distribution and Migration of Tertiary Mammals in Eurasia, edited by Reumer, Jelle W. F. and Wessels, Wilma, pp. 111. Deinsea 10. Natural History Museum, Rotterdam.Google Scholar
Fortelius, Mikael, Eronen, Jussi T., Liu, Liping et al. 2006. Late Miocene and Pliocene Large Land Mammals and Climatic Changes in Eurasia. Palaeogeography, Palaeoclimatology, Palaeoecology 238:219227.Google Scholar
Fortelius, Mikael, Žliobaitė, Indrė, Kaya, Ferhat et al. 2016. An Ecometric Analysis of the Fossil Mammal Record of the Turkana Basin. Philosophical Transactions of the Royal Society B 371:20150232.Google Scholar
Foster, J. Bristol. 1964. Evolution of Mammals on Islands. Nature 202:234235.Google Scholar
Fowler, Melvin L., and Parmalee, Paul W.. 1959. Ecological Interpretation of Data on Archaeological Sites: The Modoc Rock Shelter. Transactions of the Illinois State Academy of Science 52(3–4):109119.Google Scholar
Francis, C. M. 2008. A Guide to the Mammals of Southeast Asia. Princeton University Press.Google Scholar
Franz-Odendaal, Tamara A., and Kaiser, Thomas M.. 2003. Differential Mesowear in the Maxillary and Mandibular Cheek Dentition of Some Ruminants (Artiodactyla). Annales Zoologici Fennici 40:395410.Google Scholar
Fraser, Danielle, and Theodor, Jessica M.. 2010. The Use of Gross Dental Wear in Dietary Studies of Extinct Lagomorphs. Journal of Paleontology 84:720729.Google Scholar
Frazier, Michael K. 1977. New Records of Neofiber leonardi (Rodentia: Cricetidae) and the Paleoecology of the Genus. Journal of Mammalogy 58:368373.Google Scholar
Fricke, Henry C., Clyde, William C., and O’Neill, James R.. 1998. Intra-Tooth Variations in δ18O (PO4) of Mammalian Tooth Enamel as a Record of Seasonal Variations in Continental Climate Variables. Geochimica et Cosmochimica Acta 62:18391850.Google Scholar
Fritts, Harold C., Blasing, Terence J., Hayden, Bruce P., and Kutzbach, John E.. 1971. Multivariate Techniques for Specifying Tree-Growth and Climate Relationships and for Reconstructing Anomalies in Paleoclimate. Journal of Applied Meteorology 10:845864.Google Scholar
Frost, S. R. 2007. African Pliocene and Pleistocene Cercopithecid Evolution and Global Climate Change. In Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence, edited by Bobe, R[ené, Alemseged, Z[eresenay and Behrensmeyer, A[nna] K., pp. 5176. Springer, Dordrecht.Google Scholar
Fukami, Tadashi. 2015. Historical Contingency in Community Assembly: Integrating Niches, Species Pools, and Priority Effects. Annual Review of Ecology, Evolution, and Systematics 46:123.Google Scholar
Gailer, Juan Pablo, Calandra, Ivan, Schulz-Kornas, Ellen, and Kaiser, Thomas M.. 2016. Morphology is Not Destiny: Discrepancy between Form, Function and Dietary Adaptation in Bovid Cheek Teeth. Journal of Mammalian Evolution 23:369383.CrossRefGoogle Scholar
Galster, S., Burgess, N. D., Fjeldså, J., Hansen, L. A., and Rahbek, C.. 2007. One Degree Resolution Databases of the Distribution of 1085 Mammals in Sub-Saharan Africa. Zoological Museum, University of Copenhagen.Google Scholar
Gamble, Clive. 1978. Optimising Information from Studies of Faunal Remains. In Sampling in Contemporary British Archaeology, edited by Cherry, John F., Gamble, Clive, and Shennan, Stephen, pp. 321353. British Archaeological Reports, British Series 50. BAR, Oxford.Google Scholar
Gandiwa, Edson. 2013. Top-Down and Bottom-Up Control of Large Herbivore Populations: A Review of Natural and Human-Induced Influences. Tropical Conservation Science 6:493505.Google Scholar
García-Alix, Antonio, Minwer-Barakat, Raef, Suárez, Elvira Martín, Freudenthal, Matthijs, and Martín, José M.. 2008. Late Miocene–Early Pliocene Climatic Evolution of the Granada Basin (Southern Spain) Deduced from the Paleoecology of the Micromammal Associations. Palaeogeography, Palaeoclimatology, Palaeoecology 265:214225.Google Scholar
García-Alix, Antonio, Minwer-Barakat, Raef, Suárez, Elvira Martín, and Freudenthal, Matthijs. 2009. Small Mammals from the Early Pleistocene of the Granada Basin, Southern Spain. Quaternary Research 72:265274.Google Scholar
García Yelo, B. A., Gómez Cano, A. R., Cantalapiedra, J. L. et al. 2014. Palaeoenvironmental Analysis of the Aragonian (Middle Miocene) Mammalian Faunas from the Madrid Basin Based on Body-Size Structure. Journal of Iberian Geology 40:129140.Google Scholar
Gardner, A. L. (editor). 2007. Mammals of South America, vol. i: Marsupials, Xenarthans, Shrews, and Bats. University of Chicago Press.Google Scholar
Gardner, Janet L., Peters, Anne, Kearney, Michael R., Joseph, Leo, and Heinsohn, Robert. 2011. Declining Body Size: A Third Universal Response to Warming? Trends in Ecology and Evolution 26:285291.Google Scholar
Garrard, Andrew N. 1982. The Environmental Implications of a Re-analysis of the Large Mammal Fauna from the Wadi El-Mughara Caves, Palestine. In Palaeoclimates, Palaeoenvironments and Human Communities in the Eastern Mediterranean Region in Later Prehistory, edited by Bintliff, John L. and Van Zeist, Willem, pp. 165187. British Archaeological Reports, International Series 133(i). BAR, Oxford.Google Scholar
Garrett, Nicole D., Fox, David L., McNulty, Kieran P. et al. 2015. Stable Isotope Paleoecology of Late Pleistocene Middle Stone Age Humans from Equatorial East Africa, Lake Victoria Basin, Kenya. Journal of Human Evolution 82:114.Google Scholar
Gasse, F., and Tekaia, F.. 1983. Transfer Functions for Estimating Paleoecological Conditions (Ph) from East African Diatoms. Hydrobiologia 103:8590.Google Scholar
Gasse, F., Juggins, S., and Khelifa, L. Ben. 1995. Diatom-Based Transfer Functions for Inferring Past Hydrochemical Characteristics of African Lakes. Palaeogeography, Palaeoclimatology, Palaeoecology 117:3154.Google Scholar
Gaston, Kevin J. 1996. Species Richness: Measure and Measurement. In Biodiversity: A Biology of Numbers and Difference, edited by Gaston, Kevin J., pp. 77113. Blackwell, Oxford.Google Scholar
Gaston, Kevin J. 2007. Latitudinal Gradient in Species Richness. Current Biology 17(15):R574.Google Scholar
Gauthreaux, Sidney A., Jr. 1980. The Influences of Long-Term and Short-Term Climatic Changes on the Dispersal and Migration of Organisms. In Animal Migration, Orientation and Navigation, edited by Gauthreaux, Sidney A., Jr., pp. 103174. Academic Press, New York.Google Scholar
Gavin, Daniel G., Oswald, W. Wyatt, Wahlf, Eugene R., and Williams, John W.. 2003. A Statistical Approach to Evaluating Distance Metrics and Analog Assignments for Pollen Records. Quaternary Research 60:356367.Google Scholar
Geist, Valerius. 1987. Bergmann’s Rule Is Invalid. Canadian Journal of Zoology 65:10351038.Google Scholar
Geist, Valerius. 1989. Environmentally Guided Phenotype Plasticity in Mammals and Some of its Consequences to Theoretical and Applied Biology. In Alternative Life-History Styles of Animals, edited by Bruton, Michael N., pp. 153176. Kluwer Academic, Dordrecht.Google Scholar
Geist, Valerius. 1992. Endangered Species and the Law. Nature 357:274276.Google Scholar
Geist, Valerius. 1998. Deer of the World: Their Evolution, Behavior, and Ecology. Stackpole Books, Mechanicsburg, PA.Google Scholar
George, Christian O. 2012. Alternative Approaches to the Identification and Reconstruction of Paleoecology of Quaternary Mammals. Unpublished Ph.D. dissertation, University of Texas, Austin.Google Scholar
George, T. Neville. 1958. The Ecology of Fossil Animals: I. Organism and Environment. Science Progress 46:677680.Google Scholar
Georgina, Dianna M. 2001. The Small Mammals of Lime Hills Cave I. In People and Wildlife in Northern North America: Essays in Honor of R. Dale Guthrie, edited by Gerlach, S. Craig and Murray, Maribeth S., pp. 2331. British Archaeological Reports, International Series 944. BAR, Oxford.Google Scholar
Gidley, James W., and Gazin, C. Lewis. 1938. The Pleistocene Vertebrate Fauna from Cumberland Cave, Maryland, United States Museum Bulletin 171. Smithsonian Institution, Washington, DC.Google Scholar
Gieke, James. 1872. On Changes of Climate during the Glacial Epoch. Fifth Paper. Geological Magazine 9:164170.Google Scholar
Gieke, James. 1881. Prehistoric Europe, a Geological Sketch. Edward Standford, London.Google Scholar
Gienapp, P., Teplitsky, C., Alho, J. S., Mills, J. A., and Merilä, J.. 2008. Climate Change and Evolution: Disentangling Environmental and Genetic Responses. Molecular Ecology 17:167178.Google Scholar
Gifford[-Gonzalez], Diane P. 1981. Taphonomy and Paleoecology: A Critical Review of Archaeology’s Sister Disciplines. In Advances in Archaeological Method and Theory, vol. iv, edited by Schiffer, Michael B., pp. 365438. Academic Press, New York.Google Scholar
Gifford-Gonzalez, Diane P. 1991a. Bones Are Not Enough: Analogues, Knowledge, and Interpretive Strategies in Zooarchaeology. Journal of Anthropological Archaeology 10:215254.Google Scholar
Gifford-Gonzalez, Diane [P.]. 1991b. Examining and Refining the Quadratic Crown Height Method of Age Estimation. In Human Predators and Prey Mortality, edited by Stiner, Mary C., pp. 4178. Westview Press, Boulder, CO.Google Scholar
Giglio, Louis, Randerson, James T., and der Werf, Guido van. 2013. Analysis of Daily, Monthly, and Annual Burned Area Using the Fourth-Generation Global Fire Emissions Database (GFED4). Journal of Geophysical Research 118:317328.Google Scholar
Gilmore, Raymond M. 1949. The Identification and Value of Mammal Bones from Archeological Excavations. Journal of Mammalogy 30:163169.Google Scholar
Gingerich, Philip D. 1989. New Earliest Wasatchian Mammalian Fauna from the Eocene of Northwestern Wyoming: Composition and Diversity in a Rarely Sampled High-Floodplain Assemblage. University of Michigan Papers in Paleontology 28:197.Google Scholar
Gingerich, Philip D. 1993. Quantification and Comparison of Evolutionary Rates. American Journal of Science 293 a:453478.Google Scholar
Gingerich, Philip D. 2006. Environment and Evolution through the Paleocene-Eocene Thermal Maximum. Trends in Ecology and Evolution 21:246253.Google Scholar
Giovas, Christina M. 2009. The Shell Game: Analytic Problems in Archaeological Mollusc Quantification. Journal of Archaeological Science 36:15571564.Google Scholar
Gleason, Henry A. 1926. The Individualistic Concept of the Plant Association. American Midland Naturalist 21:92110.Google Scholar
Gobalet, Kenneth W. 2001. A Critique of Faunal Analysis: Inconsistency among Experts in Blind Tests. Journal of Archaeological Science 28:377386.Google Scholar
Goillot, Cyrielle, Blondel, Cécile, and Peigné, Stéphane. 2009. Relationships between Dental Microwear and Diet in Carnivora (Mammalia): Implications for the Reconstruction of the Diet of Extinct Taxa. Palaeogeography, Palaeoclimatology, Palaeoecology 271: 1323.Google Scholar
Goldblatt, P., and Manning, John C.. 2002. Plant Diversity of the Cape Region of South Africa. Annals of the Missouri Botanical Garden 89:281302.Google Scholar
Gómez Cano, Ana Rosa, García Yelo, Blanca A., and Hernández-Fernández, Manuel. 2006. Cenogramas, Análisis Bioclimático y Muestreo en Faunas de Mamíferos: Implicaciones para la Aplicación de Métodos de Análisis Paleoecológico. Estudios Geológicos 62:135144.Google Scholar
Gómez Cano, , Rosa, Ana, Cantalapiedra, Juan L., Álvarez-Sierra, M. Ángels, and Fernández, Manuel Hernández. 2014. A Macroecological Glance at the Structure of Late Miocene Rodent Assemblages from Southwest Europe. Scientific Reports 4:6557.Google Scholar
Good, Stephen P., and Caylor, Kelly K.. 2011. Climatological Determinants of Woody Cover in Africa. Proceedings of the National Academy of Sciences USA 108:49024907.Google Scholar
Gordon, Elizabeth A. 1993. Screen Size and Differential Faunal Recovery: A Hawaiian Example. Journal of Field Archaeology 20:453460.Google Scholar
Gordon, K. D. 1988. A Review of Methodology and Quantification in Dental Microwear Analysis. Scanning Microscopy 2:11391147.Google Scholar
Gotelli, Nicholas J., and Colwell, Robert K.. 2001. Quantifying Biodiversity: Procedures and Pitfalls in the Measurements and Comparison of Species Richness. Ecology Letters 4:379391.Google Scholar
Gotelli, Nicholas J., and Graves, Gary R.. 1996. Null Models in Ecology. Smithsonian Institution Press, Washington, DC.Google Scholar
Gould, Stephen Jay. 1965. Is Uniformitarianism Necessary? American Journal of Science 263:223228.Google Scholar
Gould, Stephen Jay. 1966. Allometry and Size in Ontogeny and Phylogeny. Biological Reviews 41:587640.Google Scholar
Gould, Stephen Jay. 1970. Land Snail Communities and Pleistocene Climates in Bermuda: A Multivariate Analysis of Microgastropod Diversity. In Proceedings of North American Paleontological Convention, edited by Yochelson, Ellis L., pp. 486521. Allen Press, Lawrence, KS.Google Scholar
Gould, Stephen Jay. 1975. On the Scaling of Tooth Size in Mammals. American Zoologist 15:353362.Google Scholar
Gould, Stephen Jay. 1977. Eternal Metaphors in Palaeontology. In Patterns of Evolution as Illustrated by the Fossil Record, edited by Hallam, Anthony, pp. 126. Elsevier, Amsterdam.Google Scholar
Gould, Stephen Jay. 1987. Time’s Arrow, Time’s Cycle: Myth and Metaphor in the Discovery of Geological Time. Harvard University Press, Cambridge, MA.Google Scholar
Gould, Stephen Jay. 1990. The Golden Rule: A Proper Scale for our Environmental Crisis. Natural History 90(9):2430.Google Scholar
Gould, Stephen Jay, and Lewontin, Richard C.. 1979. The Spandrels of San Marco and the Panglossain Paradigm: A Critique of the Adaptationist Programme. Proceedings of the Royal Society B 205:581598.Google Scholar
Graham, Russell W. 1976. Late Wisconsin Mammalian Faunas and Environmental Gradients of the Eastern United States. Paleobiology 2:343350.Google Scholar
Graham, Russell W. 1979. Paleoclimates and Late Pleistocene Faunal Provinces in North America. In Pre-Llano Cultures of the Americas: Paradoxes and Possibilities, edited by Humphrey, Robert L. and Stanford, Dennis, pp. 4969. Anthropological Society of Washington, Washington, DC.Google Scholar
Graham, Russell W. 1981. Preliminary Report on Late Pleistocene Vertebrates from the Selby and Dutton Archeological/Paleontological Sites, Yuma County, Colorado. University of Wyoming Contributions to Geology 20:3356.Google Scholar
Graham, Russell W. 1984. Paleoenvironmental Implications of the Quaternary Distribution of the Eastern Chipmunk (Tamias striatus) in Central Texas. Quaternary Research 21:111114.Google Scholar
Graham, Russell W. 1985a. Diversity and Community Structure of the Late Pleistocene Mammal Fauna of North America. Acta Zoologica Fennica 170:181192.Google Scholar
Graham, Russell W. 1985b. Response of Mammalian Communities to Environmental Changes during the Late Quaternary. In Community Ecology, edited by Diamond, Jared and Case, Ted J., pp. 303313. Harper and Row, New York.Google Scholar
Graham, Russell W. 1991. Interpreting Fossils (Review of Peter Andrews’ Owls, Caves and Fossils). Science 253:213214.Google Scholar
Graham, Russell W. 2001. Comment on “Skeleton of Extinct North American Sea Mink (Mustela macrodon)” by Mead et al. Quaternary Research 56:419421.Google Scholar
Graham, Russell W. 2005. Quaternary Mammal Communities: Relevance of the Individualistic Response and Non-Analogue Faunas. In Paleobiogeography: Generating New Insights into the Coevolution of the Earth and its Biota, edited by Lieberman, Bruce S. and Stigall, Alycia L., pp. 141157. Paleontological Society Papers 11. Paleontological Society, Boulder, CO.Google Scholar
Graham, Russell W., and Lundelius, Ernest L., Jr. 1994. FAUNMAP: A Database Documenting Late Quaternary Distributions of Mammal Species in the United States, Illinois State Museum Scientific Papers 25. Illinois State Museum, Springfield.Google Scholar
Graham, Russell W., and Mead, Jim I.. 1987. Environmental Fluctuations and Evolution of Mammalian Faunas during the Last Deglaciation in North America. In North America and Adjacent Oceans during the Last Deglaciation, edited by Ruddiman, W. F. and Wright, H. E, Jr., pp. 371402. Geology of North America k-3. Geological Society of America, Boulder, CO.Google Scholar
Graham, Russell W., and Semken, Holmes A., Jr. 1987. Philosophy and Procedures for Paleoenvironmental Studies of Quaternary Mammalian Faunas. In Late Quaternary Mammalian Biogeography and Environments of the Great Plains and Prairies, edited by Graham, Russell W., Semken, Holmes A., Jr., and Graham, Mary A., pp. 117. Illinois State Museum Scientific Papers 22. Illinois State Museum, Springfield.Google Scholar
Graham, Russell W., Semken, Holmes A., Jr., and Graham, Mary A. (editors). 1987. Late Quaternary Mammalian Biogeography and Environments of the Great Plains and Prairies. Illinois State Museum Scientific Papers 22. Illinois State Museum, Springfield.Google Scholar
Grayson, Donald K. 1976. The Nightfire Island Avifauna and the Altithermal. In Holocene Environmental Change in the Great Basin, edited by Elston, Robert, pp. 74102. Research Paper 6. Nevada Archaeological Survey, Reno.Google Scholar
Grayson, Donald K. 1977. A Review of the Evidence for Early Holocene Turkeys in the Northern Great Basin. American Antiquity 42:110114.Google Scholar
Grayson, Donald K. 1978. Reconstructing Mammalian Communities: A Discussion of Shotwell’s Method of Paleoecological Analysis. Paleobiology 4:7781.Google Scholar
Grayson, Donald K. 1979a. Mount Mazama, Climatic Change, and Fort Rock Basin Archaeofaunas. In Volcanic Activity and Human Ecology, edited by Sheets, Payson D. and Grayson, Donald K., pp. 427457. Academic Press, New York.Google Scholar
Grayson, Donald K. 1979b. On the Quantification of Vertebrate Archaeofaunas. In Advances in Archaeological Method and Theory vol. ii, edited by Schiffer, Michael B., pp. 199237. Academic Press, New York.Google Scholar
Grayson, Donald K. 1980. Vicissitudes and Overkill: The Development of Explanations of Pleistocene Extinctions. In Advances in Archaeological Method and Theory vol. iii, edited by Schiffer, Michael B., pp. 357403. Academic Press, New York.Google Scholar
Grayson, Donald K. 1981. A Critical View of the Use of Archaeological Vertebrates in Paleoenvironmental Reconstruction. Journal of Ethnobiology 1:2838.Google Scholar
Grayson, Donald K. 1983a. The Establishment of Human Antiquity. Academic Press, New York.Google Scholar
Grayson, Donald K. 1983b. The Paleontology of Gatecliff Shelter. In The Archaeology of Monitor Valley ii: Gatecliff Shelter, by Thomas, David Hurst, pp. 99126. Anthropological Papers 59(1). American Museum of Natural History, New York.Google Scholar
Grayson, Donald K. 1984a. Nineteenth-Century Explanations of Pleistocene Extinctions: A Review and Analysis. In Quaternary Extinctions: A Prehistoric Revolution, edited by Martin, Paul S. and Klein, Richard G., pp. 539. University of Arizona Press, Tucson.Google Scholar
Grayson, Donald K. 1984b. Quantitative Zooarchaeology: Topics in the Analysis of Archaeological Faunas. Academic Press, Orlando, FL.Google Scholar
Grayson, Donald K. 1991a. Alpine Faunas from the White Mountains, California: Adaptive Change in the Prehistoric Great Basin? Journal of Archaeological Science 18:483506.Google Scholar
Grayson, Donald K. 1991b. The Small Mammals of Gatecliff Shelter: Did People Make a Difference? In Beamers, Bobwhites, and Blue-Points: Tributes to the Career of Paul W. Parmalee, edited by Purdue, James R., Klippel, Walter E., and Styles, Bonnie W., pp. 99109. Illinois State Museum Scientific Papers 23. Illinois State Museum, Springfield.Google Scholar
Grayson, Donald K. 1998. Moisture History and Small Mammal Community Richness during the Latest Pleistocene and Holocene, Northern Bonneville Basin, Utah. Quaternary Research 49:330334.Google Scholar
Grayson, Donald K. 2000a. The Homestead Cave Mammals. In Late Quaternary Paleoecology in the Bonneville Basin, edited by Madsen, David B., pp. 6789. Bulletin 130. Utah Geological Survey, Salt Lake City.Google Scholar
Grayson, Donald K. 2000b. Mammalian Responses to Middle Holocene Climatic Change in the Great Basin of the Western United States. Journal of Biogeography 27:181192.Google Scholar
Grayson, Donald K. 2006. The Late Quaternary Biogeographic Histories of Some Great Basin Mammals (Western USA). Quaternary Science Reviews 25:29642991.Google Scholar
Grayson, Donald K. 2011. The Great Basin: A Natural Prehistory. University of California Press, Berkeley.Google Scholar
Grayson, Donald K. 2016. Giant Sloths and Sabertooth Cats: Extinct Mammals and the Archaeology of the Ice Age Great Basin. University of Utah Press, Salt Lake City.Google Scholar
Grayson, Donald K, and Delpech, Françoise. 1998. Changing Diet Breadth in the Early Upper Paleolithic of Southwestern France. Journal of Archaeological Science 25:11191129.Google Scholar
Grayson, Donald K, 2003. Ungulates and the Middle-to-Upper Paleolithic Transition at Grotte XVI (Dordogne, France). Journal of Archaeological Science 30:16331648.Google Scholar
Grayson, Donald K, 2005. Pleistocene Reindeer and Global Warming. Conservation Biology 19:557562.Google Scholar
Grayson, Donald K, 2006. Was There Increasing Dietary Specialization across the Middle-to-Upper Paleolithic Transition in France? In When Neanderthals and Modern Humans Met, edited by Conard, Nicholas J., pp. 377417. Tübingen Publications in Prehistory. Kerns Verlag, Tübingen, Germany.Google Scholar
Grayson, Donald K, 2008. The Large Mammals of Roc de Combe (Lot, France): The Châtelperronian and Aurignacian Assemblages. Journal of Anthropological Archaeology 27:338362.Google Scholar
Grayson, Donald K, and Madsen, David B.. 2000. Biogeographic Implications of Recent Low-Elevation Recolonization by Neotoma cinerea in the Great Basin. Journal of Mammalogy 81:11001105.Google Scholar
Grayson, Donald K., Livingston, Stephanie D., Rickart, Eric, and Shaver, Monson W., III. 1996. Biogeographic Significance of Low-Elevation Records for Neotoma cinerea from the Northern Bonneville Basin, Utah. Great Basin Naturalist 56:191196.Google Scholar
Grayson, Donald K, Delpech, Françoise, Rigaud, Jean-Philippe, and Simek, Jan F.. 2001. Explaining the Development of Dietary Dominance by a Single Ungulate Taxon at Grotte XVI, Dordogne, France. Journal of Archaeological Science 28:115125.Google Scholar
Green, Jeremy L. 2009. Dental Microwear in the Orthodentine of the Xenarthra (Mammalia) and its Use in Reconstructing the Palaeodiet of Extinct Taxa: The Case Study of Nothrotheriops shastensis (Xenarthra, Tardigrada, Nothrotheriidae). Zoological Journal of the Linnean Society 156:201222.Google Scholar
Green, Jeremy L., and Kalthoff, Daniela C.. 2015. Xenarthran Dental Microstructure and Dental Microwear Analyses, with New Data for Megatherium americanum (Megatheriidae). Journal of Mammalogy 96:645657.Google Scholar
Green, Jeremy L., and Resar, Nicholas A.. 2012. The Link between Dental Microwear and Feeding Ecology in Tree Sloths and Armadillos (Mammalia: Xenarthra). Biological Journal of the Linnean Society 107:277294.Google Scholar
Greenacre, M. J., and Vrba, E. S.. 1984. Graphical Display and Interpretation of Antelope Census Data in African Wildlife Areas, Using Correspondence Analysis. Ecology 65:984997.Google Scholar
Grigson, Caroline. 1969. The Uses and Limitations of Differences in Absolute Size in the Distinction between the Bones of Aurochs (Bos primigenius) and Domestic Cattle (Bos taurus). In The Domestication and Exploitation of Plants and Animals, edited by Ucko, Peter J. and Dimbleby, G. W., pp. 277294. Aldine Atherton, Chicago.Google Scholar
Grigson, Caroline. 1989. Size and Sex: Evidence for the Domestication of Cattle in the Near East. In The Beginnings of Agriculture, edited by Milles, Annie, Williams, Diane, and Gardner, Neville, pp. 77109. British Archaeological Reports, International Series 496. BAR, Oxford.Google Scholar
Grimm, Guido W., and Potts, Alastair J.. 2016. Fallacies and Fantasies: The Theoretical Underpinnings of the Coexistence Approach for Palaeoclimate Reconstruction. Climate of the Past 12:611622.Google Scholar
Grine, Frederick E. 1986. Dental Evidence for Dietary Differences in Australopithecus and Paranthropus: A Quantitative Analysis of Permanent Molar Microwear. Journal of Human Evolution 15:783822.Google Scholar
Grine, Frederick E., and Kay, Richard F.. 1988. Early Hominid Diets from Quantitative Image Analysis of Dental Microwear. Nature 333:765768.Google Scholar
Grine, Frederick E., Ungar, Peter S., and Teaford, M. F.. 2002. Error Rates in Dental Microwear Quantification Using Scanning Electron Microscopy. Scanning 24:144153.Google Scholar
Grinnell, Joseph. 1914. Barriers to Distribution as Regards Birds and Mammals. American Naturalist 48:248254.Google Scholar
Grinnell, Joseph. 1917a. Field Tests of Theories Concerning Distributional Control. American Naturalist 51:115128.Google Scholar
Grinnell, Joseph. 1917b. The Niche-Relationships of the California Thrasher. Auk 34:427433.Google Scholar
Grinnell, Joseph. 1924. Geography and Evolution. Ecology 5:225229.Google Scholar
Grinnell, Joseph. 1928. Presence and Absence of Animals. University of California Chronicle 30:429450.Google Scholar
Groffman, Peter M., Baron, Jill S., Blett, Tamara et al. 2006. Ecological Thresholds: The Key to Successful Environmental Management or an Important Concept with No Practical Application? Ecosystems 9:113.Google Scholar
Guilday, John E. 1962. The Pleistocene Local Fauna of the Natural Chimneys, Augusta County, Virginia. Carnegie Museum of Natural History Annals 36:87122.Google Scholar
Guilday, John E. 1967. Differential Extinction during Late-Pleistocene and Recent Times. In Pleistocene Extinctions: The Search for a Cause, edited by Martin, P[aul] S. and Wright, H. E., Jr., pp. 121140. Yale University Press, New Haven, CT.Google Scholar
Guilday, John E. 1969. Small Mammal Remains from the Wasden Site (Owl Cave), Bonneville County, Idaho. Tebiwa, Journal of the Idaho State University Museum 12(1):4757.Google Scholar
Guilday, John E. 1971. The Pleistocene History of the Appalachian Mammal Fauna. In The Distributional History of the Biota of the Southern Appalachians, Part iii: Vertebrates, edited by Holt, Perry C., pp. 233262. Research Division Monograph 4. Virginia Polytechnic Institute and State University, Blacksburg.Google Scholar
Guilday, John E., and Adam, Eleanor K.. 1967. Small Mammal Remains from Jaguar Cave, Lemhi County, Idaho. Tebiwa, Journal of the Idaho State University Museum 10(1):2637.Google Scholar
Guilday, John E., andParmalee, Paul W.. 1972. Quaternary Periglacial Records of Voles of the Genus Phenacomys Merriam (Cricetidae: Rodentia). Quaternary Research 2:170175.Google Scholar
Guilday, John E., Martin, Paul S., and McCrady, A. D.. 1964. New Paris No. 4: A Pleistocene Cave Deposit in Bedford County, Pennsylvania. Bulletin of the National Speleological Society 26:121194.Google Scholar
Guilday, John E., Hamilton, Harold W., Anderson, Elaine, and Parmalee, Paul W.. 1978. The Baker Bluff Cave Deposit, Tennessee, and the Late Pleistocene Faunal Gradient. Bulletin of the Carnegie Museum of Natural History 11. Carnegie Museum of Natural History, Pittsburgh.Google Scholar
Guiot, J., De Beaulieu, Jean-Luc, Pons, A., and Reille, M.. 1989. A 140,000-Year Climatic Reconstruction from Two European Pollen Cores. Nature 338:309313.Google Scholar
Gunnell, Gregg F. 1994. Paleocene Mammals and Faunal Analysis of the Chappo Type Locality (Tiffanian), Green River Basin, Wyoming. Journal of Vertebrate Paleontology 14:81104.Google Scholar
Gunnell, Gregg F. 1997. Wasatchian–Bridgerian (Eocene) Paleoecology of the Western Interior of North America: Changing Paleoenvironments and Taxonomic Composition of Omomyid (Tarsiiformes) Primates. Journal of Human Evolution 32:105132.Google Scholar
Guo, Qinfeng. 2014. Central-Marginal Population Dynamics in Species Invasions. Frontiers in Ecology and Evolution 2: article 23.CrossRefGoogle Scholar
Guo, Qinfeng, Taper, Mark, Schoenberger, M., and Brandle, J.. 2005. Spatial-Temporal Population Dynamics across Species Range: From Centre to Margin. Oikos 108:4757.Google Scholar
Gustafson, Carl E. 1972. Faunal Remains from the Marmes Rockshelter and Related Archaeological Sites in the Columbia Basin. Unpublished Ph.D. dissertation, Department of Zoology, Washington State University, Pullman.Google Scholar
Guthrie, R. Dale. 1968a. Paleoecology of the Large-Mammal Community in Interior Alaska during the Late Pleistocene. American Midland Naturalist 79:346363.Google Scholar
Guthrie, R. Dale. 1968b. Paleoecology of a Late Pleistocene Small Mammal Community from Interior Alaska. Arctic 21:223244.Google Scholar
Guthrie, R. Dale. 1970. Bison Evolution and Zoogeography in North America during the Pleistocene. Quarterly Review of Biology 45:115.Google Scholar
Guthrie, R. Dale. 1982. Mammals of the Mammoth Steppe as Paleoenvironmental Indicators. In Paleoecology of Beringia, edited by Hopkins, David M., Matthews, John V., Jr., Schweger, Charles E., and Young, Steven B., pp. 307326. Academic Press, New York.Google Scholar
Guthrie, R. Dale. 1984a. Alaskan Megabucks, Megabulls, and Megarams: The Issue of Pleistocene Gigantism. In Contributions in Quaternary Vertebrate Paleontology: A Volume in Honor of John E. Guilday, edited by Genoways, Hugh H. and Dawson, Mary R., pp. 482510. Special Publication 8. Carnegie Museum of Natural History, Pittsburgh.Google Scholar
Guthrie, R. Dale. 1984b. Mosaics, Allochemics, and Nutrients: An Ecological Theory of Late Pleistocene Extinctions. In Quaternary Extinctions: A Prehistoric Revolution, edited by Martin, Paul S. and Klein, Richard G., pp. 259298. University of Arizona Press, Tucson.Google Scholar
Guthrie, R. Dale. 2003. Rapid Body Size Decline in Alaskan Pleistocene Horses before Extinction. Nature 426:169171.Google Scholar
Haber, F. 1959. Fossils and the Idea of a Process of Time in Natural History. In Forerunners of Darwin: 1745–1859, edited by Glass, B., Temkin, O., and Strauss, W. L, pp. 222261. Johns Hopkins University Press, Baltimore.Google Scholar
Hadly, Elisabeth A. 1997. Evolutionary and Ecological Response of Pocket Gophers (Thomomys talpoides) to Late-Holocene Climatic Change. Biological Journal of the Linnean Society 60:277296.Google Scholar
Hadly, Elisabeth A. 1999. Fidelity of Terrestrial Vertebrate Fossils to a Modern Ecosystem. Palaeogeography, Palaeoclimatology, Palaeoecology 149:389409.Google Scholar
Hadly, Elizabeth A., Kohn, Michael H., Leonard, Jennifer A., and Wayne, Robert K.. 1998. A Genetic Record of Population Isolation in Pocket Gophers during Holocene Climatic Change. Proceedings of the National Academy of Sciences USA 95:68936896.CrossRefGoogle ScholarPubMed
Hadly, Elizabeth A., Ramakrishnan, Uma, Chan, Yvonne L. et al. 2004. Genetic Response to Climatic Change: Insights from Ancient DNA and Phylochronology. PLoS Biology 2(10): e290.Google Scholar
Hafner, David J. 1993. North American Pika (Ochotona princeps) as a Late Quaternary Biogeographic Indicator Species. Quaternary Research 39:373380.Google Scholar
Hairston, Nelson G., Smith, Frederick E., and Slobodkin, Lawrence B.. 1960. Community Structure, Population Control, and Competition. American Naturalist 44:421425.Google Scholar
Hall, Brian K. 2002. Palaeontology and Evolutionary Developmental Biology: A Science of the Nineteenth and Twenty-First Centuries. Palaeontology 45:647669.Google Scholar
Hall, E. Raymond. 1981. The Mammals of North America, second edition. John Wiley and Sons, New York.Google Scholar
Hall, E. Raymond, and Kelson, Keith R.. 1959. The Mammals of North America. Ronald Press, New York.Google Scholar
Hammer, Øyvind, and Harper, David A. T.. 2006. Paleontological Data Analysis. Blackwell, Malden, MA.Google Scholar
Hammer, Øyvind, Harper, David A. T., and Ryan, Paul D.. 2001. Paleontological Statistics Software Package for Education and Data Analysis. Palaeontologia Electronica 4(1):9 pp.Google Scholar
Hampe, Arndt. 2004. Bioclimatic Envelope Models: What They Detect and What They Hide. Global Ecology and Biogeography 13:469471.Google Scholar
Hanley, Torrance C., and La Pierre, Kimberly J. (editors). 2015. Trophic Ecology: Bottom-Up or Top-Down Interactions across Aquatic and Terrestrial Systems. Cambridge University Press.Google Scholar
Hargrave, Lyndon L., and Emslie, Steven D.. 1979. Osteological Identification of Sandhill Crane versus Turkey. American Antiquity 44:295299.Google Scholar
Harris, Arthur H. 1963. Vertebrate Remains and Past Environmental Reconstruction in the Navajo Reservoir District. Papers in Anthropology 11. Museum of New Mexico, Santa Fe.Google Scholar
Harris, Arthur H. 1984. Neotoma in the Late Pleistocene of New Mexico and Chihuahua. In Contributions in Quaternary Vertebrate Paleontology: A Volume in Memorial to John E. Guilday, edited by Genoways, Hugh H. and Dawson, Mary R., pp. 164178. Special Publication 8. Carnegie Museum of Natural History, Pittsburgh.Google Scholar
Harris, Arthur H. 1985. Late Pleistocene Vertebrate Paleoecology of the West. University of Texas Press, Austin.Google Scholar
Harris, John M., and Cerling, Thure E.. 2002. Dietary Adaptations of Extant and Neogene African Suids. Journal of Zoology (London) 256:4554.Google Scholar
Harrison, J. L. 1962. The Distribution of Feeding Habits among Animals in a Tropical Rain Forest. Journal of Animal Ecology 31:5363.Google Scholar
Harrison, Sandy P., and Prentice, Colin I.. 2003. Climate and CO2 Controls on Global Vegetation Distribution at the Last Glacial Maximum: Analysis Based on Palaeovegetation Data, Biome Modelling and Palaeoclimate Simulations. Global Change Biology 9:9831004.Google Scholar
Haupt, Ryan J., DeSantis, Larisa R. G., Green, Jeremy L., and Ungar, Peter S.. 2013. Dental Microwear Texture as a Proxy for Diet in Xenarthrans. Journal of Mammalogy 94:856866.Google Scholar
Hawkins, Bradford A. 2001. Ecology’s Oldest Pattern? Trends in Ecology and Evolution 16:470.Google Scholar
Hawkins, Bradford A., and Porter, Eric E.. 2003. Relative Influence of Current and Historical Factors on Mammal and Bird Diversity Patterns in Deglaciated North America. Global Ecology and Biogeography 12:475481.Google Scholar
Hayek, Lee-Ann C., Bernor, Raymond L, Solounias, Nikos, and Steigerwald, Patricia. 1991. Preliminary Studies of Hipparionine Horse Diet as Measured by Tooth Microwear. Annales Zoologici Fennici 28:187200.Google Scholar
Hayward, M. W., Henschel, P., O’Brien, J. et al. 2006. Prey Preferences of the Leopard (Panthera pardus). Journal of Zoology (London) 270:298313.Google Scholar
Heaney, Lawrence R. 1978. Island Area and Body Size of Insular Mammals: Evidence from Tri-Colored Squirrel (Callosciurus prevosti) of Southeast Asia. Evolution 32:2944.Google Scholar
Hedges, R. E. M. 2002. Bone Diagenesis: An Overview of Processes. Archaeometry 44:319328.Google Scholar
Helgen, Kristofer M., Wells, Rod T., Kear, Benjamin P., Gerdtz, Wayne R., and Flannery, Timothy F.. 2006. Ecological and Evolutionary Significance of Sizes of Giant Extinct Kangaroos. Australian Journal of Zoology 54:293303.Google Scholar
Hedgpeth, Joel W., and Ladd, Harry S. (editors). 1957. Treatise on Marine Ecology and Paleoecology. Memoir 67 (vol. i: Ecology; vol. ii: Paleoecology). Geological Society of America, New York.Google Scholar
Heisler, Leanne M., Somers, Christopher M., and Poulin, Ray G.. 2014. Rodent Populations on the Northern Great Plains Respond to Weather Variation at a Landscape Level. Journal of Mammalogy 95:8290.Google Scholar
Hendey, Q. B. 1974. The Late Cenozoic Carnivora of the South-Western Cape Province. Annals of the South African Museum 63:1369.Google Scholar
Hengl, Tomslav, de Jesus, Jorge Mendes, MacMillan, Robert A. et al. 2014. Soilgrids1km – Global Soil Information Based on Automated Mapping. PLoS ONE 9: e105992.Google Scholar
Herm, D. 1972. Pitfalls in Paleoecologic Interpretation: An Integrated Approach to Avoid the Major Pits. In International Geographic Congress, 24th Session, Section 7: Paleontology, edited by Mamet, B. E and Westermann, G. E. G, pp. 8288. International Geographic Union, Montreal.Google Scholar
Hernández-Fernández, Manuel. 2001. Bioclimatic Discriminant Capacity of Terrestrial Mammal Faunas. Global Ecology and Biogeography 10:189204.Google Scholar
Hernández-Fernández, Manuel, and Peláez-Campomanes, Pablo. 2003. The Bioclimatic Model: A Method of Palaeoclimatic Qualitative Inference Based on Mammal Associations. Global Ecology and Biogeography 12:507517.Google Scholar
Hernández-Fernández, Manuel, and Peláez-Campomanes, Pablo. 2005. Quantitative Palaeoclimatic Inference Based on Terrestrial Mammal Faunas. Global Ecology and Biogeography 14:3956.Google Scholar
Hernández-Fernández, Manuel, Alberdi, M. T., Azanza, B. et al. 2006. Identification Problems of Arid Environments in the Neogene–Quaternary Mammal Record of Spain. Journal of Arid Environments 66:585608.Google Scholar
Hernández-Fernández, Manuel, Álvarez-Sierra, M. A., and Peláez-Campomanes, Pablo. 2007. Bioclimatic Analysis of Rodent Palaeofaunas Reveals Severe Climatic Changes in Southwestern Europe during the Plio-Pleistocene. Palaeogeography, Palaeoclimatology, Palaeoecology 251:500526.Google Scholar
Hester, James J. 1964. The Possibilities for Paleoecological Reconstruction – Archaeology. In The Reconstruction of Past Environments: Proceedings, assembled by Hester, James J. and Schoenwetter, James, pp. 1923. Publication 3. Fort Burgwin Research Center, Taos, NM.Google Scholar
Hester, James J., and Schoenwetter, James (assemblers). 1964. The Reconstruction of Past Environments. Publication 3. Fort Burgwin Research Center, Taos, NM.Google Scholar
Hibbard, Claude W. 1944. Stratigraphy and Vertebrate Paleontology of Pleistocene Deposits of Southwestern Kansas. Geological Society of America Bulletin 55:707754.Google Scholar
Hibbard, Claude W. 1949. Techniques of Collecting Microvertebrate Fossils. Contributions from the Museum of Paleontology 8:719. University of Michigan, Ann Arbor.Google Scholar
Hibbard, Claude W. 1955. The Jinglebob Interglacial (Sangamon?) Fauna from Kansas and its Climatic Significance. Contributions from the Museum of Paleontology 12:179228. University of Michigan, Ann Arbor.Google Scholar
Hibbard, Claude W. 1958. Summary of North American Pleistocene Mammalian Local Faunas. Papers of the Michigan Academy of Science, Arts and Letters 43:332.Google Scholar
Hibbard, Claude W. 1960. An Interpretation of Pliocene and Pleistocene Climates in North America. Annual Report of the Michigan Academy of Science, Arts and Letters 62:530.Google Scholar
Hibbard, Claude W. 1963. A Late Illinoian Fauna from Kansas and its Climatic Significance. Papers of the Michigan Academy of Science, Arts and Letters 48:187221.Google Scholar
Hibbard, C[laude] W., Ray, C. E., Savage, D. E., Taylor, D. W., and Guilday, J. E.. 1965. Quaternary Mammals of North America. In The Quaternary of the United States, edited by Wright, H. E, Jr., and Frey, David G., pp. 509525. Princeton University Press.Google Scholar
Higgs, E. S. 1967. Faunal Fluctuations and Climate in Libya. In Background to Evolution in Africa, edited by Bishop, Walter W. and Clark, J. Desmond, pp. 149163. University of Chicago Press.Google Scholar
Higham, C. F. 1969. The Metrical Attributes of Two Samples of Bovine Limb Bones. Journal of Zoology (London) 157:6374.Google Scholar
Hijmans, Robert J., and Graham, Catherine H.. 2006. The Ability of Climate Envelope Models to Predict the Effect of Climate Change on Species Distributions. Global Change Biology 12:22722281.Google Scholar
Hijmans, Robert J., Cameron, Susan E., Parra, Juan L., Jones, Peter G., and Jarvis, Andy. 2005. Very High Resolution Interpolated Climate Surfaces for Global Land Areas. International Journal of Climatology 25:19651978.Google Scholar
Hill, M. O. 1973. Diversity and Evenness: A Unifying Notation and its Consequences. Ecology 54:427432.Google Scholar
Hill, M. O., and Gauch, H. G., Jr. 1980. Detrended Correspondence Analysis: An Improved Ordination Technique. Vegetatio 42:4758.Google Scholar
Hill, Matthew E., Jr., Hill, Matthew G., and Widga, Christopher C.. 2008. Late Quaternary Bison Diminution on the Great Plains of North America: Evaluating the Role of Human Hunting Versus Climate Change. Quaternary Science Reviews 27:17521771.Google Scholar
Hobbs, Richard J., Higgs, Eric, and Harris, James A.. 2009. Novel Ecosystems: Implications for Conservation and Restoration. Trends in Ecology and Evolution 24:599605.Google Scholar
Hobbs, Richard J., Higgs, Eric, and Hall, Carol M. (editors). 2013. Novel Ecosystems: Intervening in the New Ecological World Order. Wiley-Blackwell, Oxford.Google Scholar
Hobbs, Richard J., Higgs, Eric, Hall, Carol M. et al. 2014. Managing the Whole Landscape: Historical, Hybrid, and Novel Ecosystems. Frontiers in Ecology and Environment 12:557564.Google Scholar
Hodgkins, Jamie, Marean, Curtis W., Turq, Alain et al. 2016. Climate-Mediated Shifts in Neandertal Subsistence Behaviors at Pech de l’Azé and Roc de Marsal (Dordogne Valley, France). Journal of Human Evolution 96:118.Google Scholar
Hoefs, Jochen. 2015. Stable Isotope Geochemistry, seventh edition. Springer, New York.Google Scholar
Hoffman, Antoni. 1979. Community Paleoecology as an Epiphenomenal Science. Paleobiology 5:357379.Google Scholar
Hoffmann, Robert S., and Jones, J. Knox, Jr. 1970. Influence of Late-Glacial and Post-Glacial Events on the Distribution of Recent Mammals on the Northern Great Plains. In Pleistocene and Recent Environments of the Central Great Plains, edited by Dort, Wakefield, Jr., and Jones, J. Knox, pp. 355394. Department of Geology, Special Publication 3. University of Kansas Press, Lawrence.Google Scholar
Hoffmeister, Donald F. 1964. Mammals and Life Zones: Time for Re-evaluation. Plateau 37:4655.Google Scholar
Hofmann, R. R. 1989. Evolutionary Steps of Ecophysiological Adaptation and Diversification of Ruminants: A Comparative View of their Digestive System. Oecologia 78:443457.Google Scholar
Hokr, Zdenek. 1951. A Method of the Quantitative Determination of the Climate in the Quaternary Period by Means of Mammal Associations. Sborník of the Geological Survey of Czechoslovakia 18:209219.Google Scholar
Holbrook, Sally J. 1975. Prehistoric Paleoecology of Northwestern New Mexico. Unpublished Ph.D. dissertation, University of California, Berkeley.Google Scholar
Holbrook, Sally J. 1977. Rodent Faunal Turnover and Prehistoric Community Stability in Northwestern New Mexico. American Naturalist 111: 11951208.Google Scholar
Holbrook, Sally J. 1980. Species Diversity Patterns in Some Present and Prehistoric Rodent Communities. Oecologia 44:355367.Google Scholar
Holbrook, Sally J. 1982a. Prehistoric Environmental Reconstruction by Mammalian Microfaunal Analysis, Grasshopper Pueblo. In Multidisciplinary Research at Grasshopper Pueblo, Arizona, edited by Longacre, William A., Holbrook, Sally J., and Graves, Michael W., pp. 7386. Anthropological Papers of the University of Arizona 40. University of Arizona, Tucson.Google Scholar
Holbrook, Sally J. 1982b. The Prehistoric Local Environment of Grasshopper Pueblo, Arizona. Journal of Field Archaeology 9:207215.Google Scholar
Holbrook, Sally J., and Mackey, James C.. 1976. Prehistoric Environmental Change in Northern New Mexico: Evidence from a Gallina Phase Archaeological Site. Kiva 41:309317.Google Scholar
Holdaway, Simon, and Wandsnider, LuAnn (editors). 2008. Time in Archaeology: Time Perspectivism Revisited. University of Utah Press, Salt Lake City.Google Scholar
Holdridge, L. R. 1947. Determination of World Plant Formations from Simple Climatic Data. Science 105:367368.Google Scholar
Holling, C. S. 1992. Cross-Scale Morphology, Geometry, and Dynamics of Ecosytems. Ecological Monographs 62:447502.Google Scholar
Holt, Ben G., Lessard, Jean-Philippe, Borregaard, Michael K. et al. 2013. An Update of Wallace’s Zoogeographic Regions of the World. Science 339:7478.Google Scholar
Holt, Robert D. 2003. On the Evolutionary Ecology of Species’ Ranges. Evolutionary Ecology Research 5:159178.Google Scholar
Holt, Robert D. 2009. Bringing the Hutchinsonian Niche into the 21st Century: Ecological and Evolutionary Perspectives. Proceedings of the National Academy of Sciences USA 106 (Supplement 2):1965919665.Google Scholar
Holt, Robert D., and Keitt, Timothy H.. 2005. Species’ Borders: A Unifying Theme in Ecology. Oikos 108:36.Google Scholar
Holtzman, Richard C. 1979. Maximum Likelihood Estimation of Fossil Assemblage Composition. Paleobiology 5:7789.Google Scholar
Hooijer, D. A. 1947. Pleistocene Remains of Panthera tigris (Linnaeus) Subspecies from Wansien, Szechwan, China, Compared with Fossil and Recent Tigers from Other Localities. American Museum Novitates 1346:117.Google Scholar
Hopkins, Samantha S. B. 2008. Reassessing the Mass of Exceptionally Large Rodents Using Toothrow Length and Area as Proxies for Body Mass. Journal of Mammalogy 89:232243.Google Scholar
Hopley, Philip J., and Maslin, Mark A.. 2010. Climate-Averaging of Terrestrial Faunas: An Example from the Plio-Pleistocene of South Africa. Paleobiology 36:3250.Google Scholar
Hopley, Philip J., Latham, Alf G., and Marshall, Jim D.. 2006. Palaeoenvironments and Palaeodiets of Mid-Pliocene Micromammals from Makapansgat Limeworks, South Africa: A Stable Isotope and Dental Microwear Approach. Palaeogeography, Palaeoclimatology, Palaeoecology 233:235251.Google Scholar
Hostetler, Mark E. 1997. Avian Body-Size Clumps and the Response of Birds to Scale-Dependent Landscape Structure in Suburban Habitats. Unpublished Ph.D. dissertation, University of Florida, Gainesville.Google Scholar
Howard, Hildegarde. 1930. A Census of the Pleistocene Birds of Rancho La Brea from the Collections of the Los Angeles Museum. Condor 32:8188.Google Scholar
Howell, F. Clark and Bourlière, François (editors). 1963. African Ecology and Human Evolution. Aldine Transaction, New Brunswick, NJ.Google Scholar
Hubbell, Stephen P. 2001. The Unified Neutral Theory of Biodiversity and Biogeography. Princeton University Press.Google Scholar
Hughes, Susan S. 2009. Noble Marten (Martes americana nobilis) Revisited: Its Adaptation and Extinction. Journal of Mammalogy 90:7492.Google Scholar
Hughes, Terry P., Carpenter, Stephen, Rockström, Johan, Scheffer, Marten, and Walker, Brian. 2013. Multiscale Regime Shifts and Planetary Boundaries. Trends in Ecology and Evolution 28:389395.Google Scholar
Hunt, Gene. 2006. Fitting and Comparing Models of Phyletic Evolution: Random Walks and Beyond. Paleobiology 32:578601.Google Scholar
Hunt, Gene. 2007. The Relative Importance of Directional Change, Random Walks, and Stasis in the Evolution of Fossil Lineages. Proceedings of the National Academy of Sciences USA 104:1840418408.Google Scholar
Hunt, Gene, Bell, Michael A., and Travis, Matthew P.. 2008. Evolution toward a New Adaptive Optimum: Phenotypic Evolution in a Fossil Stickleback Lineage. Evolution 62–63:700710.Google Scholar
Hunter, Mark D., and Price, Peter W.. 1992. Playing Chutes and Ladders: Heterogeneity and the Relative Roles of Bottom-Up and Top-Down Forces in Natural Communities. Ecology 73:724732.Google Scholar
Huntley, Brian. 2012. Reconstructing Palaeoclimates from Biological Proxies: Some often Overlooked Sources of Uncertainty. Quaternary Science Reviews 31:116.Google Scholar
Huppert, Amit, and Solow, Andrew R.. 2004. A Method for Reconstructing Climate from Fossil Beetle Assemblages. Proceedings of the Royal Society B 271:11251128.Google Scholar
Hurlbert, Stuart H. 1971. The Nonconcept of Species Diversity: A Critique and Alternative Parameters. Ecology 52:577586.Google Scholar
Huston, Michael A., and Wolverton, Steve. 2009. The Global Distribution of Net Primary Productivity: Resolving the Paradox. Ecological Monographs 79:343377.Google Scholar
Huston, Michael A., and Wolverton, Steve. 2011. Regulation of Animal Size by eNPP, Bergmann’s Rule, and Related Phenomena. Ecological Monographs 81:349405.Google Scholar
Hutchinson, G. Evelyn. 1957. Concluding Remarks. Cold Spring Harbor Symposium in Quantitative Biology 22:415427.Google Scholar
Hutchinson, G. Evelyn. 1978. An Introduction to Population Ecology. Yale University Press, New Haven, CT.Google Scholar
Hutson, William H. 1977. Transfer Functions under No-Analog Conditions: Experiments with Indian Ocean Planktonic Foraminifera. Quaternary Research 8:355367.Google Scholar
Imbrie, J., and Kipp, Nilva G.. 1971. A New Micropaleontological Method for Quantitative Paleoclimatology: Application to a Late Pleistocene Carribean Core. In The Late Cenozoic Glacial Ages, edited by Turekian, Karl K., pp. 71181. Yale University Press, New Haven, CT.Google Scholar
Imbrie, John, and Newell, Norman D. (editors). 1964. Approaches to Paleoecology. John Wiley and Sons, New York.Google Scholar
Inkpen, Robert John. 2008. Explaining the Past in the Geosciences. Philosophia 36:495507.Google Scholar
Inskeep, R. R. 1987. Nelson Bay Cave, Cape Province, South Africa: The Holocene Levels. British Archaeological Reports, International Series 357. BAR, Oxford.Google Scholar
Jablonski, David, and Sepkoski, J. John, Jr. 1996. Paleobiology, Community Ecology, and Scales of Ecological Pattern. Ecology 77:13671378.Google Scholar
Jackson, Donald A. 1997. Compositional Data in Community Ecology: The Paradigm or Peril of Proportions? Ecology 78:929940.Google Scholar
Jackson, Stephen T. 2009a. Alexander von Humboldt and the General Physics of the Earth. Science 324:596597.Google Scholar
Jackson, Stephen T. 2009b. Introduction: Humboldt, Ecology, and the Cosmos. In Essay on the Geography of Plants, by von Humboldt, Alexander, pp. 146. University of Chicago Press.Google Scholar
Jackson, Stephen T., and Overpeck, Jonathan T.. 2000. Responses of Plant Populations and Communities to Environmental Changes of the Late Quaternary. Paleobiology 26 (Supplement):194220.Google Scholar
Jackson, Stephen T, and Williams, John W.. 2004. Modern Analogs in Quaternary Paleoecology: Here Today, Gone Yesterday, Gone Tomorrow? Annual Review of Earth and Planetary Sciences 32:495537.Google Scholar
Jacobs, Zenobia, and Roberts, Richard. G.. 2017. Single-Grain OSL Chronologies for the Still Bay and Howieson’s Poort Industries and the Transition between Them. Journal of Human Evolution 107:113.Google Scholar
Jacobson, Jodi A. 2003. Identification of Mule Deer (Odocoileus hemionus) and White-Tailed Deer (Odocoileus virginianus) Postcranial Remains as a Means of Determining Human Subsistence Strategies. Plains Anthropologist 48:287297.Google Scholar
Jacobson, Jodi A. 2004. Determining Human Ecology on the Plains through the Identification of Mule Deer (Odocoileus hemionus) and White-Tailed Deer (Odocoileus virginianus) Postcranial Remains. Unpublished Ph.D. dissertation, Department of Anthropology, University of Tennessee, Knoxville.Google Scholar
Jaeger, J.-J., and Wesselman, H. B.. 1976. Fossil Remains of Micromammals from the Omo Group Deposits. In Earliest Man and Environments in the Lake Rudolf Basin: Stratigraphy, Paleoecology, and Evolution, edited by Coppen, Yves, Howell, F. Clark, Isaac, Glynn Ll., and Leakey, Richard E. F, pp. 351360. University of Chicago Press.Google Scholar
James, Frances C. 1970. Geographic Size Variation in Birds and its Relationship to Climate. Ecology 51:365390.Google Scholar
Jamniczky, Heather A., Brinkman, Donald B., and Russell, Anthony P.. 2003. Vertebrate Microsite Sampling: How Much is Enough? Journal of Vertebrate Paleontology 23:725734.Google Scholar
Jamniczky, Heather A., Brinkman, Donald B., and Russell, Anthony P.. 2008. How Much is Enough? A Repeatable, Efficient, and Controlled Sampling Protocol for Assessing Taxonomic Diversity and Abundance in Vertebrate Microfossil Assemblages. In Vertebrate Microfossil Assemblages: Their Role in Paleoecology and Paleobiogeography, edited by Sankey, Julia T. and Baszio, Sven, pp. 916. Indiana University Press, Bloomington.Google Scholar
Janis, Christine M. 1988. An Estimation of Tooth Volume and Hypsodonty Indices in Ungulate Mammals, and the Correlation of These Factors with Dietary Preference. In Teeth Revisited: Proceedings of the VIIth International Symposium on Dental Morphology, edited by Russell, D. E, Santoro, J. P, and Sigogneau-Russell, D., pp. 367387. Mémoires du Muséum National d’Historie Naturelle, Series C. Muséum National d’Historie Naturelle, Paris.Google Scholar
Janis, Christine M. 1995. Correlations between Craniodental Morphology and Feeding Behavior in Ungulates: Reciprocal Illuminations between Living and Fossil Taxa. In Functional Morphology in Vertebrate Paleontology, edited by Thomason, J. J, pp. 7698. Cambridge University Press.Google Scholar
Janis, Christine M., and Ehrhardt, D.. 1988. Correlation of Relative Muzzle Width and Relative Incisor Width with Dietary Preference in Ungulates. Zoological Journal of the Linnean Society 92:267284.Google Scholar
Janis, Christine M., and Fortelius, Mikael. 1988. On the Means Whereby Mammals Achieve Increased Functional Durability of their Dentitions, with Special Reference to Limiting Factors. Biological Reviews 63:197230.Google Scholar
Janis, Christine M., Damuth, John, and Theodor, Jessica M.. 2000. Miocene Ungulates and Terrestrial Primary Productivity: Where Have All the Browsers Gone? Proceedings of the National Academy of Sciences USA 97:78997904.Google Scholar
Janis, Christine M., Damuth, John, and Theodor, Jessica M.. 2004. The Species Richness of Miocene Browsers, and Implications for Habitat Type and Primary Productivity in the North American Grassland Biome. Palaeogeography, Palaeoclimatology, Palaeoecology 207:371398.Google Scholar
Jardine, Phillip E., Janis, Christine M., Sahney, Sarda, and Benton, Michael J.. 2012. Grit Not Grass: Concordant Patterns of Early Origin of Hypsodonty in Great Plains Ungulates and Glires. Palaeogeography, Palaeoclimatology, Palaeoecology 365–366:110.Google Scholar
Jass, Christopher N., Poteet, M. F., and Bell, C. J.. 2015. Response of Pocket Gophers (Geomys) to Late Quaternary Environmental Change on the Edwards Plateau of Central Texas. Historical Biology 27:192213.Google Scholar
Jehl, Joseph R., Jr. 1966. Subspecies of Recent and Fossil Birds. Auk 83:306307.Google Scholar
Jernvall, Jukka, Hunter, John P., and Fortelius, Mikael. 1996. Molar Tooth Diversity, Disparity, and Ecology in Cenozoic Ungulate Radiations. Science 274:14891492.Google Scholar
Johnson, Ralph Gordon. 1960. Environmental Interpretation of Pleistocene Marine Species. Journal of Geology 68:575576.Google Scholar
Jones, Emily Lena. 2015. Archaeofaunal Evidence of Human Adaptation to Climate Change in Upper Paleolithic Iberia. Journal of Archaeological Science: Reports 2:257263.Google Scholar
Jones, J. Knox, Jr., Armstrong, David M., and Choate, Jerry R.. 1985. Guide to Mammals of the Plains States. University of Nebraska Press, Lincoln.Google Scholar
Jones, Kate E., Biebly, Jon, Cardillo, Marcel et al. 2009. PanTHERIA: A Species-Level Database of Life History, Ecology, and Geography of Extant and Recently Extinct Mammals. Ecology 90:2648.Google Scholar
Jones, Terry L., Gobalet, Kenneth W., and Codding, Brian F.. 2016. The Archaeology of Fish and Fishing on the Central Coast of California: The Case for an Under-Exploited Resource. Journal of Anthropological Archaeology 41:88108.Google Scholar
Jouzel, J., Masson-Delmotte, V., Cattani, O. et al. 2007. Orbital and Millennial Antarctic Climate Variability over the Past 800,000 Years. Science 317:793796.Google Scholar
Jungers, William L., and German, Rebecca Z.. 1981. Ontogenetic and Interspecific Skeletal Allometry in Nonhuman Primates: Bivariate Versus Multivariate Analysis. American Journal of Physical Anthropology 55:195202.Google Scholar
Jungers, William L., Falsetti, Anthony B., and Wall, Christine E.. 1995. Shape, Relative Size, and Size-Adjustments in Morphometrics. American Journal of Physical Anthropology 38:137161.Google Scholar
Kaiser, Thomas M., and Fortelius, Mikael. 2003. Differential Mesowear in Occluding Upper and Lower Molars: Opening Mesowear Analysis for Lower Molars and Premolars in Hyposodont Horses. Journal of Morphology 258:6383.Google Scholar
Kaiser, Thomas M., and Solounias, Nikos. 2003. Extending the Tooth Mesowear Method to Extinct and Extant Equids. Geodiversitas 25:321345.Google Scholar
Kappelman, John. 1984. Plio-Pleistocene Environments of Bed I and Lower Bed II, Olduvai Gorge, Tanzania. Palaeogeography, Palaeoclimatology, Palaeoecology 48:171196.Google Scholar
Kappelman, John. 1988. Morphology and Locomotor Adaptations of the Bovid Femur in Relation to Habitat. Journal of Morphology 198:119130.Google Scholar
Kappelman, John. 1991. The Paleoenvironments of Kenyapithecus at Fort Ternan. Journal of Human Evolution 20:95129.Google Scholar
Kappelman, John, Plummer, Tom, Bishop, Laura, Duncan, Alex, and Appleton, Scott. 1997. Bovids as Indicators of Plio-Pleistocene Paleoenvironments in East Africa. Journal of Human Evolution 32:229256.Google Scholar
Karr, James R. 1971. Structure of Avian Communities in Selected Panama and Illinois Habitats. Ecological Monographs 41:207233.Google Scholar
Kaufman, Daniel. 1998. Measuring Archaeological Diversity: An Application of the Jackknife Technique. American Antiquity 63:7385.Google Scholar
Kay, Richard F. 1975. The Functional Adaptations of Primate Molar Teeth. American Journal of Physical Anthropology 43:195215.Google Scholar
Kay, Richard F. 1987. Analysis of Primate Dental Microwear Using Image Processing Techniques. Scanning Microscopy 1:657662.Google Scholar
Kay, Richard F., and Madden, Richard F.. 1997. Mammals and Rainfall: Paleoecology of the Middle Miocene at La Venta (Colombia, South America). Journal of Human Evolution 32:161199.Google Scholar
Kay, Richard F., Vizcaíno, Sergio F., and Bargo, M. Susana. 2012. A Review of the Paleoenvironment and Paleoecology of the Miocene Santa Cruz Formation. In Early Miocene Paleobiology in Patagonia: High-Latitude Paleocommunities of the Santa Cruz Formation, edited by Vizcaíno, Sergio F., Kay, Richard F., and Bargo, M. Susana, pp. 331365. Cambridge University Press.Google Scholar
Kaya, Ferhat, Kaymakçi, Nuretdin, Bibi, Faysal et al. 2016. Magnetostratigraphy and Paleoecology of the Hominid-Bearing Locality Çorakyerler, Tuglu Formation (Çankiri Basin, Central Anatolia). Journal of Vertebrate Paleontology 36:e1071710.Google Scholar
Kearney, Michael, and Porter, Warren. 2009. Mechanistic Niche Modeling: Combining Physiological and Spatial Data to Predict Species’ Ranges. Ecology Letters 12:334350.Google Scholar
Kenagy, G. J. 1973. Adaptations for Leaf Eating in the Great Basin Kangaroo Rat, Dipodomys microps. Oecologia 12:383412.Google Scholar
Kerley, Graham I. H., Pressey, Robert L., Cowling, Richard M., Boshoff, Andre F., and Sims-Castley, Rebecca. 2003. Options for the Conservation of Large and Medium-Sized Mammals in the Cape Floristic Region Hotspot, South Africa. Biological Conservation 112:169190.Google Scholar
Kidwell, Susan M. 2001. Preservation of Species Abundance in Marine Death Assemblages. Science 294:10911094.Google Scholar
Kidwell, Susan M., and Behrensmeyer, Anna K. (editors). 1993. Taphonomic Approaches to Time Resolution in Fossil Assemblages. Short Courses in Paleontology 6. Paleontological Society, Knoxville, TN.Google Scholar
Kidwell, Susan M., and Tomasovych, Adam. 2013. Implications of Time-Averaged Death Assemblages for Ecology and Conservation Biology. Annual Review of Ecology, Evolution and Systematics 44:539563.Google Scholar
King, Frances B., and Graham, Russell W.. 1981. Effects of Ecological and Paleoecological Patterns on Subsistence and Paleoenvironmental Reconstructions. American Antiquity 46:128142.Google Scholar
Kingdon, J., and Hoffman, M. (editors). 2013. Mammals of Africa, vol. vi: Pigs, Hippopotamuses, Chevrotain, Giraffes, Deer and Bovids. Bloomsbury, London.Google Scholar
Kingdon, J., Happold, D., Butynski, T., Hoffman, M. and Kalina, J. (editors). 2013. Mammals of Africa (vols. i–iv). Bloomsbury, London.Google Scholar
Kingston, John D. 2007. Shifting Adaptive Landscapes: Progress and Challenges in Reconstructing Early Hominid Environments. Yearbook of Physical Anthropology 50:2058.Google Scholar
Kipp, Nilva G. 1976. New Transfer Function for Estimating Past Sea-Surface Conditions from Sea-Bed Distribution of Planktonic Foraminiferal Assemblages in the North Atlantic. Geological Society of America Memoirs 145:342.Google Scholar
Klein, Richard G. 1972. The Late Quaternary Mammalian Fauna of Nelson Bay Cave (Cape Province, South Africa): Its Implications for Megafaunal Extinctions and Environmental and Cultural Change. Quaternary Research 2:135142.Google Scholar
Klein, Richard G. 1975. Paleoanthropological Implications of the Nonarcheological Bone Assemblage from Swartklip I, South-Western Cape Province, South Africa. Quaternary Research 5:275288.Google Scholar
Klein, Richard G. 1976a. The Fossil History of Raphicerus H. Smith, 1827 (Bovidae, Mammalia) in the Cape Biotic Zone. Annals of the South African Museum 71:169191.Google Scholar
Klein, Richard G. 1976b. The Mammalian Fauna of the Klasies River Mouth Sites, Southern Cape Province, South Africa. South African Archaeological Bulletin 31:7598.Google Scholar
Klein, Richard G. 1978. A Preliminary Report on the Larger Mammals from the Boomplaas Stone Age Cave Site, Cango Valley, Oudtshoorn District, South Africa. South African Archaeological Bulletin 33:6675.Google Scholar
Klein, Richard G. 1980. Environmental and Ecological Implications of Large Mammals from Upper Pleistocene and Holocene Sites in Southern Africa. Annals of the South African Museum 81:223283.Google Scholar
Klein, Richard G. 1983. Palaeoenvironmental Implications of Quaternary Large Mammals in the Fynbos Region. In Fynbos Palaeoecology: A Preliminary Synthesis, edited by Deacon, H. J, Hendey, Q. B, and Lambrechts, J. J. N, pp. 116138. South African National Scientific Programmes Report 75. Mills Litho, Cape Town.Google Scholar
Klein, Richard G. 1986. Carnivore Size and Quaternary Climatic Change in Southern Africa. Quaternary Research 25:153170.Google Scholar
Klein, Richard G. 1991. Size Variation in Cape Dune Molerat (Bathyergus suillus) and Late Quaternary Climatic Change in the Southwestern Cape Province, South Africa. Quaternary Research 36:243256.Google Scholar
Klein, Richard G. 1994.The Problem of Modern Human Origins. In Origins of Anatomically Modern Humans, edited by Nitecki, Matthew V. and Nitecki, Doris V., pp. 317. Plenum Press, New York.Google Scholar
Klein, Richard G. 1995. Anatomy, Behavior, and Modern Human Origins. Journal of World Prehistory 9:167198.Google Scholar
Klein, Richard G. 1998. Why Anatomically Modern People Did Not Disperse from Africa 100,000 Years Ago. In Neandertals and Modern Humans in Western Asia, edited by Akazawa, Takeru, Aoki, Kenichi, and Bar-Yosef, Ofer, pp. 509521. Plenum Press, New York.Google Scholar
Klein, Richard G. 2000. Archaeology and the Evolution of Human Behavior. Evolutionary Anthropology 9:1736.Google Scholar
Klein, Richard G., and Cruz-Uribe, Kathryn. 1983a. The Computation of Ungulate Age (Mortality) Profiles from Dental Crown Heights. Paleobiology 9:7078.Google Scholar
Klein, Richard G., and Cruz-Uribe, Kathryn. 1983b. Stone Age Population Numbers and Average Tortoise Size at Byneskranskop Cave 1 and Die Kelders Cave 1, Southern Cape Province, South Africa. South African Archaeological Bulletin 38:2630.Google Scholar
Klein, Richard G., and Cruz-Uribe, Kathryn. 1984. The Analysis of Animal Bones from Archeological Sites. University of Chicago Press.Google Scholar
Klein, Richard G., and Cruz-Uribe, Kathryn. 1987. Large Mammal and Tortoise Bones from Eland’s Bay Cave and Nearby Sites, Western Cape Province, South Africa. In Papers in the Prehistory of the Western Cape, South Africa, edited by Parkington, J. E and Hall, M., pp. 132163. British Archaeological Reports, International Series 322. BAR, Oxford.Google Scholar
Klein, Richard G., and Cruz-Uribe, Kathryn. 1996. Size Variation in the Rock Hyrax (Procavia capensis) and Late Quaternary Climatic Change in South Africa. Quaternary Research 46:193207.Google Scholar
Klein, Richard G., and Cruz-Uribe, Kathryn. 2000. Middle and Later Stone Age Large Mammal and Tortoise Remains from Die Kelders Cave 1, Western Cape Province, South Africa. Journal of Human Evolution 38:169195.Google Scholar
Klein, Richard G., and Scott, Katharine. 1989. Glacial/Interglacial Size Variation in Fossil Spotted Hyenas (Crocuta crocuta) from Britain. Quaternary Research 32:8895.Google Scholar
Klein, Richard G., Cruz-Uribe, Kathryn, and Beaumont, Peter B.. 1991. Environmental, Ecological, and Paleoanthropological Implications of the Late Pleistocene Mammalian Fauna from Equus Cave, Northern Cape Province, South Africa. Quaternary Research 36:94119.Google Scholar
Klein, Richard G., Cruz-Uribe, Kathryn, Halkett, David, Hart, Timothy, and Parkington, John E.. 1999. Paleoenvironmental and Human Behavioral Implications of the Boegoeberg 1 Late Pleistocene Hyena Den, Northern Cape Province, South Africa. Quaternary Research 52:393403.Google Scholar
Klein, Richard G., Avery, Graham, Cruz-Uribe, Kathryn et al. 2004. The Ysterfontein 1 Middle Stone Age Site, South Africa, and Early Human Exploitation of Coastal Resources. Proceedings of the National Academy of Sciences USA 101:57085715.Google Scholar
Klein, Richard G., Avery, Graham, Cruz-Uribe, Kathryn, and Steele, Teresa E.. 2007. The Mammalian Fauna Associated with an Archaic Hominin Skullcap and Later Acheulean Artifacts at Elandsfontein, Western Cape Province, South Africa. Journal of Human Evolution 52:164186.Google Scholar
Koch, Paul L. 1986. Clinal Geographic Variation in Mammals: Implications for the Study of Chronoclines. Paleobiology 12:269281.Google Scholar
Koch, Paul L., Tuross, Noreen, and Fogel, Marilyn L.. 1997. The Effects of Sample Treatment and Diagenesis on the Isotopic Integrity of Carbonate in Biogenic Hydroxylapatite. Journal of Archaeological Science 24:417429.Google Scholar
Koch, Paul L., Hoppe, Kathryn A., and Webb, David S.. 1998. The Isotopic Ecology of Late Pleistocene Mammals in North America. Part 1. Florida. Chemical Geology 152:119138.Google Scholar
Kohn, Matthew J. 1996. Predicting Animal δ18O: Accounting for Diet and Physiological Adaptation. Geochimica et Cosmochimica Acta 60:48114829.Google Scholar
Kolb, H. H. 1978. Variation in the Size of Foxes in Scotland. Biological Journal of the Linnean Society 10:291304.Google Scholar
Kooyman, Brian, and Sandgathe, Dennis. 2001. Sexually Dimorphic Size Variation in Holocene Bison as Revealed by Carpals and Tarsals. In People and Wildlife in Northern North America: Essays in Honor of R. Dale Guthrie, edited by Gerlach, S. Craig and Murray, Maribeth S., pp. 6778. British Archaeological Reports, International Series 944. BAR, Oxford.Google Scholar
Kovarovic, Kris, and Andrews, Peter. 2007. Bovid Postcranial Ecomorphological Survey of the Laetoli Paleoenvironment. Journal of Human Evolution 52:663680.Google Scholar
Kovarovic, Kris, Andrews, Peter, and Aiello, Leslie. 2002. The Palaeoecology of the Upper Ndolanya Beds at Laetoli, Tanzania. Journal of Human Evolution 43:395418.Google Scholar
Kovarovic, Kris, Aiello, Leslie C., Cardini, Andrea, and Lockwood, Charles A.. 2011. Discriminant Function Analyses in Archaeology: Are Classification Rates Too Good to Be True? Journal of Archaeological Science 38:30063018.Google Scholar
Kowalewski, Michal. 1996. Time-Averaging, Overcompleteness, and the Geological Record. Journal of Geology 104:317326.Google Scholar
Kowalewski, Michal, and Hoffmeister, Alan P.. 2003. Sieves and Fossils: Effects of Mesh Size on Paleontological Patterns. Palaios 18:460469.Google Scholar
Kubo, Mugino O., and Yamada, Eisuke. 2014. The Inter-Relationship between Dietary and Environmental Properties and Tooth Wear: Comparisons of Mesowear, Molar Wear Rate, and Hypsodonty Index of Extant Sika Deer Populations. PLoS ONE 9:e90745.Google Scholar
Kubo, Mugino O., Yamada, Eisuke, Fujita, Masaki, and Oshiro, Ituro. 2015. Paleoecological Reconstruction of Late Pleistocene Deer from the Ryukyu Islands, Japan: Combined Evidence of Mesowear and Stable Isotope Analyses. Palaeogeography, Palaeoclimatology, Palaeoecology 435:159166.Google Scholar
Kurtén, Björn. 1952. The Chinese Hipparion Fauna: A Quantitative Survey, with Comments on the Ecology of the Machairodonts and Hyaenids and the Taxonomy of Gazelles. Societas Scientiarum Fennica: Commentationes Biologicae 13(4).Google Scholar
Kurtén, Björn. 1957. The Bears and Hyenas of the Interglacials. Quaternaria 4:6981.Google Scholar
Kurtén, Björn. 1960. Chronology and Faunal Evolution of the Earlier European Glaciations. Societas Scientiarum Fennica Commentationes Biologicae 21:162.Google Scholar
Kurtén, Björn. 1965. The Carnivora of the Palestine Caves. Acta Zoologica Fennica 107:174.Google Scholar
Kutzbach, John E., and Guetter, Peter J.. 1986. The Influence of Changing Orbital Parameters and Surface Boundary Conditions on Climate Simulations for the Past 18000 Years. Journal of Atmospheric Sciences 43:17261759.Google Scholar
Kutzbach, J[ohn E.], Gallimore, R., Harrison, S., Behling, P., Selin, R., and Laarif, F.. 1998. Climate and Biome Simulations for the Past 21,000 Years. Quaternary Science Reviews 17:473506.Google Scholar
Ladd, Harry S. 1959. Ecology, Paleontology, and Stratigraphy. Science 129:6978.Google Scholar
Laland, Kevin N., and O’Brien, Michael J.. 2011. Cultural Niche Construction: An Introduction. Biological Theory 6:191202.Google Scholar
Lambert, W. David. 2006. Functional Convergence of Ecosystems: Evidence from Body Mass Distributions of North American Late Miocene Mammal Faunas. Ecosystems 9:97118.Google Scholar
Lambert, W. David, and Holling, Crawford S.. 1998. Causes of Ecosystem Transformation at the End of the Pleistocene: Evidence from Mammal Body-Mass Distributions. Ecosystems 1:157175.Google Scholar
Lambrides, A. B. J., and Weisler, M. I.. 2015. Assessing Protocols for Identifying Pacific Island Archaeological Fish Remains: The Contribution of Vertebrae. International Journal of Osteoarchaeology 25:838848.Google Scholar
Lande, Russell. 1996. Statistics and Partitioning of Species Diversity, and Similarity among Multiple Communities. Oikos 76:513.Google Scholar
Landres, Peter B. 1992. Temporal Scale Perspectives in Managing Biological Diversity. Transactions of the North American Wildlife and Natural Resources Conference 57:292307.Google Scholar
Lang, Richard W., and Harris, Arthur H.. 1984. The Faunal Remains from Arroyo Hondo Pueblo, New Mexico: A Study in Short-Term Subsistence Change. Arroyo Hondo Archaeological Series 5. School of American Research, Santa Fe, NM.Google Scholar
Larocque, I., Hall, R. I., and Grahn, E.. 2001. Chironomids as Indicators of Climate Change: A 100-Lake Training Set from a Subarctic Region of Northern Sweden (Lapland). Journal of Paleolimnology 26:307322.Google Scholar
Lartet, Edourd. 1867. Notes sur Deux Têtes de Carnassiers Fossiles (Ursus et Felis) et sur Quelques Débris de Rhinocéros Provenant des Découvertes Faites par M. Bourguignat dans les Cavernes du Midi de la France. Annales des Sciences Naturelles: Zoologie et Paleontologie 8:157194.Google Scholar
Lartet, Edourd. 1875. Notes on the Reindeer and Hippopotamus. In Reliquiae Aquitanicae; Being Contributions to the Archaeology and Paleontology of Périgord and the Adjoining Provinces of Southern France, by Edourd Lartet and Henry Christy, edited by Jones, Thomas R, pp. 147152. Williams and Norgate, London.Google Scholar
Lastrucci, Carlo L. 1963. The Scientific Approach: Basic Principles of the Scientific Method. Schenkman, Cambridge, MA.Google Scholar
Lavergne, Sébastien, Mouquet, Nicolas, Thuiller, Wilfried, and Ronce, Ophélie. 2010. Biodiversity and Climate Change: Integrating Evolutionary and Ecological Responses to Species and Communities. Annual Review of Ecology, Evolution, and Systematics 41:321350.Google Scholar
Lawing, A. Michelle, Head, Jason J., and Polly, P. David. 2012. The Ecology of Morphology: The Ecometrics of Locomotion and Macroenvironment in North American Snakes. In Paleontology in Ecology and Conservation, edited by Louys, Julien, pp. 117146. Springer, Heidelberg.Google Scholar
Lawrence, Barbara. 1973. Problems in the Inter-Site Comparisons of Faunal Remains. In Domestikationsforschung und Geschichte der Haustiere, edited by Matolcsi, J., pp. 397402. Akademiai Kiado, Budapest.Google Scholar
Lawrence, David R. 1968. Taphonomy and Information Losses in Fossil Communities. Geological Society of America Bulletin 79:13151330.Google Scholar
Lawrence, David R. 1971. The Nature and Structure of Paleoecology. Journal of Paleontology 45:593607.Google Scholar
Le, Jianning. 1992. Paleotemperature Estimation Methods: Sensitivity Test on Two Western Equatorial Pacific Cores. Quaternary Science Reviews 11:801820.Google Scholar
Le, Jianning, and Shackleton, Nick J.. 1994. Reconstructing Paleoenvironment by Transfer Function: Model Evaluation with Simulated Data. Marine Micropaleontology 24:187199.Google Scholar
Leach, Foss. 1986. A Method for the Analysis of Pacific Island Fishbone Assemblages and an Associated Database Management System. Journal of Archaeological Science 13:147159.Google Scholar
Lee-Thorp, Julia A. 2008. On Isotopes and Old Bones. Archaeometry 50:925950.Google Scholar
Lee-Thorp, Julia A., and Sponheimer, Matt. 2006. Contributions of Biogeochemistry to Understanding Hominin Dietary Ecology. Yearbook of Physical Anthropology 131:131148.Google Scholar
Lee-Thorp, Julia A., and Sponheimer, Matt. 2013. Hominin Ecology from Hard-Tissue Biogeochemistry. In Early Hominin Paleoecology, edited by Sponheimer, Matt, Lee-Thorp, Julia A., and Reed, Kaye E., pp. 281324. University Press of Colorado, Boulder.Google Scholar
Lee-Thorp, Julia A., and van der Merwe, Nikolaas J.. 1987. Carbon Isotope Analysis of Fossil Bone Apatite. South African Journal of Science 83:712715.Google Scholar
Lee-Thorp, Julia A., Sealy, Judith C., and van der Merwe, Nikolaas J.. 1989. Stable Carbon Isotope Ratio Differences between Bone Collagen and Bone Apatite, and their Relationship to Diet. Journal of Archaeological Science 16:585599.Google Scholar
Le Fur, Soizic, Fara, Emmanuel, and Vignaud, Patrick. 2011. Effect of Simulated Faunal Impoverishment and Mixture on the Ecological Structure of Modern Mammal Faunas: Implications for the Reconstruction of Mio-Pliocene African Palaeoenvironments. Palaeogeography, Palaeoclimatology, Palaeoecology 305:295309.Google Scholar
Legendre, L., and Gallagher, E. D.. 2001. Ecologically Meaningful Transformations for Ordination of Species Data. Oikos 129:271280.Google Scholar
Legendre, P., and Legendre, L.. 2012. Numerical Ecology. Elsevier, New York.Google Scholar
Legendre, Serge. 1986. Analysis of Mammalian Communities from the Late Eocene and Oligocene of Southern France. Paleovertebrata 16:191212.Google Scholar
Legendre, Serge. 1987. Les Communautés de Mammifères d’Europe Occidentale de L’Eocène Supérieur et Oligocène: Structures et Milieux. Münchner Geowissenschaftliche Abhandlungen A 10:301312.Google Scholar
Legendre, Serge. 1989. Les Communautés de Mammifères du Paléogène (Eocène Supérieur et Oligocène) d’Europe Occidentale: Structures, Milieux et Évolution. Münchner Geowissenschaftliche Abhandlungen A 16:1110.Google Scholar
Legendre, Serge, and Roth, Claudia. 1988. Correlation of Carnassial Tooth Size and Body Weight in Recent Carnivores (Mammalia). Historical Biology 1:8598.Google Scholar
Lenton, Timothy M., Held, Hermann, Kriegler, Elmar et al. 2008. Tipping Elements in the Earth’s Climate System. Proceedings of the National Academy of Sciences USA 105:17861793.Google Scholar
Leroux, Shawn J., and Loreau, Michel. 2015. Theoretical Perspectives on Bottom-Up and Top-Down Interactions across Ecosystems. In Trophic Ecology: Bottom-Up and Top-Down Interactions across Aquatic and Terrestrial Systems, edited by Hanley, Torrence C. and La Pierre, Kimberly J., pp. 127. Cambridge University Press.Google Scholar
Lesage, Louis, Crête, Michel, Huot, Jean, and Ouellet, Jean-Pierre. 2001. Evidence for a Trade-Off between Growth and Body Reserves in Northern White-Tailed Deer. Oecologia 126:3041.Google Scholar
Levin, Naomi E., Cerling, Thure E., Passey, Benjamin H., Harris, John M., and Ehleringer, James R.. 2006. A Stable Isotope Aridity Index for Terrestrial Environments. Proceedings of the National Academy of Sciences USA 103:1120111205.Google Scholar
Levin, Simon A. 1992. The Problem of Pattern and Scale in Ecology. Ecology 73:19431967.Google Scholar
Levinson, M. 1985. Are Fossil Rodents Useful in Palaeo-Ecological Interpretations? Annals of the Geological Survey of South Africa 19:5364.Google Scholar
Lewis, Patrick J., Buchanan, Briggs, and Johnson, Eileen. 2005. Sexing Bison Metapodials Using Principal Components Analysis. Plains Anthropologist 50:159172.Google Scholar
Linder, H. P. 2003. The Radiation of the Cape Flora, Southern Africa. Biological Reviews 78:597638.Google Scholar
Lindstedt, Stan L., and Boyce, Mark S.. 1985. Seasonality, Fasting Endurance, and Body Size in Mammals. American Naturalist 125:873878.Google Scholar
Lintulaakso, Kari, and Kovarovic, Kris. 2016. Diet and Locomotion, but Not Body Size, Differentiate Mammal Communities in Worldwide Tropical Ecosystems. Palaeogeography, Palaeoclimatology, Palaeoecology 454:2029.Google Scholar
Lister, A[drian] M. 1989. Rapid Dwarfing of Red Deer on Jersey in the Last Interglacial. Nature 342:539542.Google Scholar
Lister, A[drian] M. 1997. The Evolutionary Response of Vertebrates to Quaternary Environmental Change. In Past and Future Rapid Environmental Changes: The Spatial and Evolutionary Responses of Terrestrial Biota, edited by Huntley, B., Cramer, W., Prentice, A. V, and Allen, J. R. M, pp. 287302. NATO ASI Series 147. Springer, Berlin.Google Scholar
Liu, Liping, Eronen, Jussi T., and Fortelius, Mikael. 2009. Significant Mid-Latitude Aridity in the Middle Miocene of East Asia. Palaeogeography, Palaeoclimatology, Palaeoecology 279:201206.Google Scholar
Liu, Liping, Puolamäki, Kai, Eronen, Jussi T. et al. 2012. Dental Functional Traits of Mammals Resolve Productivity in Terrestrial Ecosystems Past and Present. Proceedings of the Royal Society B 279:27932799.Google Scholar
Livingston, Stephanie D. 1987. Prehistoric Biogeography of White-Tailed Deer in Washington and Oregon. Journal of Wildlife Management 51:649654.Google Scholar
Lomolino, Mark V. 1985. Body Size of Mammals on Islands: The Island Rule Reexamined. American Naturalist 125:310316.Google Scholar
Lomolino, Mark V. 2005. Body Size Evolution in Insular Vertebrates: Generality of the Island Rule. Journal of Biogeography 32:16381699.Google Scholar
Lomolino, Mark V., Riddle, Brett R., Whittaker, Robert J., and Brown, James H.. 2010. Biogeography, fourth edition. Sinauer Associates, Sunderland, MA.Google Scholar
Longinelli, Antonio. 1984. Oxygen Isotopes in Mammal Bone Phosphate: A New Tool for Paleohydrological and Paleoclimatological Research? Geochimica et Cosmochimica Acta 48:385390.Google Scholar
López-García, Juan Manuel, Blain, Hugues-Alexandre, Cuenca-Bescós, Gloria et al. 2010. Palaeoenvironmental and Palaeoclimatic Reconstruction of the Latest Pleistocene El Portalón Site, Sierra De Atapuerca, Northwestern Spain. Palaeogeography, Palaeoclimatology, Palaeoecology 292:453464.Google Scholar
López-García, Juan Manuel, Blain, Hugues-Alexandre, Bennàsar, Maria, and Fernández-Garcí, Mónica. 2014. Environmental and Climatic Context of Neanderthal Occupation in Southwestern Europe during mis3 Inferred from the Small-Vertebrate Assemblages. Quaternary International 326–327:319328.Google Scholar
López-García, Juan Manuel, Soler, Narcís, Maroto, Julià et al. 2015a. Palaeoenvironmental and Palaeoclimatic Reconstruction of the Latest Pleistocene of L’Arbreda Cave (Serinya, Girona, Northeastern Iberia) Inferred from the Small-Mammal (Insectivore and Rodent) Assemblages. Palaeogeography, Palaeoclimatology, Palaeoecology 435:244253.Google Scholar
López-García, Juan Manuel, Valle, Chiara dalla, Cremaschi, Mauro, and Peresani, Marco. 2015b. Reconstruction of the Neanderthal and Modern Human Landscape and Climate from the Fumane Cave Sequence (Verona, Italy) Using Small-Mammal Assemblages. Quaternary Science Reviews 128:113.Google Scholar
Lorimer, Jamie. 2015. Wildlife in the Anthropocene: Conservation after Nature. University of Minnesota Press, Minneapolis.Google Scholar
Losos, Jonathan B., and Miles, Donald B.. 1994. Adaptation, Constraint, and the Comparative Method: Phylogenetic Issues and Methods. In Ecological Morphology: Integrative Organismal Biology, edited by Wainwright, Peter C. and Reilly, Stephen M., pp. 6098. University of Chicago Press.Google Scholar
Louthan, Allison M., Doak, Daniel F., and Angert, Amy L.. 2015. Where and When Do Species Interactions Set Range Limits? Trends in Ecology and Evolution 30:780792.Google Scholar
Louys, Julien (editor). 2012. Paleontology in Ecology and Conservation. Springer, Heidelberg.Google Scholar
Louys, Julien, Meloro, Carlo, Elton, Sarah, Ditchfield, Peter, and Bishop, Laura C.. 2011a. Mammal Community Structure Correlates with Arboreal Heterogeneity in Faunally and Geographically Diverse Habitats: Implications for Community Convergence. Global Ecology and Biogeography 20:717729.Google Scholar
Louys, Julien, Meloro, Carlo, Elton, Sarah, Ditchfield, Peter, and Bishop, Laura C.. 2011b. Meoswear as a Means of Determining Diet in African Antelopes. Journal of Archaeological Science 38:14851495.Google Scholar
Louys, Julien, Montanari, Shaena, Plummer, Thomas, Hertel, Fritz, and Bishop, Laura C.. 2013. Evolutionary Divergence and Convergence in Shape and Size within African Antelope Proximal Phalanges. Journal of Mammalian Evolution 20:239248.Google Scholar
Louys, Julien, Meloro, Carlo, Elton, Sarah, Ditchfield, Peter, and Bishop, Laura C.. 2015. Analytical Framework for Reconstructing Heterogeneous Environmental Variables from Mammal Community Structure. Journal of Human Evolution 78:111.Google Scholar
Lowe, J. J., and Walker, M. J. C.. 1997. Reconstructing Quaternary Environments, second edition. Pearson, Prentice Hall, New York.Google Scholar
Lowe, J. J., and Walker, M. J. C.. 2015. Reconstructing Quaternary Environments, third edition. Routledge, London.Google Scholar
Lozek, V. 1986. Mollusca Analysis. In Handbook of Palaeoecology and Palaeohydrology, edited by Berglund, B. E, pp. 729740. John Wiley and Sons, Chichester.Google Scholar
Lubinski, Patrick M. 2000. A Comparison of Methods for Evaluating Ungulate Mortality Distributions. Archaeozoologia 11:121134.Google Scholar
Lucas, Peter W., Omar, Ridwaan, Al-Fadhalah, Khaled et al. 2013. Mechanisms and Causes of Wear in Tooth Enamel: Implications for Hominin Diets. Journal of the Royal Society: Interface 10:20120923.Google Scholar
Lundelius, Ernest L., Jr. 1960. Post Pleistocene Faunal Succession in Western Australia and its Climatic Interpretation. Proceedings of the International Geological Congress 21(4):142153.Google Scholar
Lundelius, Ernest L., 1964. The Use of Vertebrates in Paleoecological Reconstructions. In The Reconstruction of Past Environments: Proceedings, assembled by Hester, James J. and Schoenwetter, James, pp. 2631. Publication 3. Fort Burgwin Research Center, Taos, NM.Google Scholar
Lundelius, Ernest L., 1967. Late-Pleistocene and Holocene Faunal History of Central Texas. In Pleistocene Extinctions: The Search for a Cause, edited by Martin, P[aul] S. and Wright, H. E, Jr., pp. 287319. Yale University Press, New Haven, CT.Google Scholar
Lundelius, Ernest L., 1972. Vertebrate Remains from the Gray Sand. In Blackwater Locality No. 1: A Stratified, Early Man Site in Eastern New Mexico, edited by Hester, James J., pp. 148163. Publication 8. Fort Burgwin Research Center, Taos, NM.Google Scholar
Lundelius, Ernest L., 1974. The Last Fifteen Thousand Years of Faunal Change in North America. In History and Prehistory of the Lubbock Lake Site, edited by Black, Craig C., pp. 141160. The Museum Journal 15. Texas Tech University, Lubbock Lake.Google Scholar
Lundelius, Ernest L., 1976. Vertebrate Paleontology of the Pleistocene: An Overview. In Ecology of the Pleistocene: A Symposium, edited by West, R. C and Haag, W. G, pp. 4559. Geoscience and Man 13. Louisiana State University, Baton Rouge.Google Scholar
Lundelius, Ernest L., 1979. Post-Pleistocene Mammals from Pratt Cave and their Environmental Significance. In Biological Investigations in the Guadalupe Mountains National Park, Texas, edited by Genoways, Hugh H. and Baker, R. J, pp. 239258. Proceedings and Transactions Series 4. National Park Service, Washington, DC.Google Scholar
Lundelius, Ernest L., 1983. Climatic Implications of Late Pleistocene and Holocene Faunal Associations in Australia. Alcheringa 7:125149.Google Scholar
Lundelius, Ernest L., 1985. North American Pleistocene Mammals: Major Problems. Acta Zoologici Fennica 170:167171.Google Scholar
Lundelius, Ernest L., 1998. Development of Quaternary Vertebrate Paleontology in North America. In Quaternary Paleozoology in the Northern Hemisphere, edited by Saunders, Jeffrey J., Styles, Bonnie W., and Baryshnikov, Gennady F., pp. 235248. Illinois State Museum Scientific Papers 27. Illinois State Museum, Springfield.Google Scholar
Luo, Zhe-Xi, Chong-Xi, Yuan, Qing-Jin, Meng, and Ji, Qiang. 2010. A Jurassic Eutherian Mammal and Divergence of Marsupials and Placentals. Nature 476:442445.Google Scholar
Lupo, Karen D. 2007. Evolutionary Foraging Models in Zooarchaeological Analysis: Recent Applications and Future Challenges. Journal of Archaeological Research 15: 143189.Google Scholar
Lüthi, Dieter, Le Floch, Martine, Bereiter, Bernhard et al. 2008. High-Resolution Carbon Dioxide Concentration Record 650,000–800,000 Years before Present. Nature 453:379382.Google Scholar
Lyell, Charles. 1832. Principles of Geology, Being an Attempt to Explain Changes of the Earth’s Surface by Reference to Causes Now in Operation, vol. ii. John Murray, London.Google Scholar
Lyell, Charles. 1833. Principles of Geology, Being an Attempt to Explain Changes of the Earth’s Surface by Reference to Causes Now in Operation, vol. iii. John Murray, London.Google Scholar
Lyell, Charles. 1863. The Geological Evidences of the Antiquity of Man, with Remarks on Theories of the Origin of Species by Variation. G. W. Childs, Philadelphia.Google Scholar
Lyman, R. Lee. 1986. On the Analysis and Interpretation of Species List Data in Zooarchaeology. Journal of Ethnobiology 6:6781.Google Scholar
Lyman, R. Lee. 1987. On the Analysis of Vertebrate Mortality Profiles: Sample Size, Mortality Type, and Hunting Pressure. American Antiquity 52:125142.Google Scholar
Lyman, R. Lee. 1994. Vertebrate Taphonomy. Cambridge University Press.Google Scholar
Lyman, R. Lee. 1995. Determining When Rare (Zoo)archaeological Phenomena are Truly Absent. Journal of Archaeological Method and Theory 2:369424.Google Scholar
Lyman, R. Lee. 1996. Applied Zooarchaeology: The Relevance of Faunal Analysis to Wildlife Management. World Archaeology 28:110125.Google Scholar
Lyman, R. Lee. 2000. Building Cultural Chronology in Eastern Washington: The Influence of Geochronology, Index Fossils, and Radiocarbon Dating. Geoarchaeology 15:609648.Google Scholar
Lyman, R. Lee. 2002. Taxonomic Identification of Zooarchaeological Remains. Review of Archaeology 23:1320.Google Scholar
Lyman, R. Lee. 2003. The Influence of Time Averaging and Space Averaging on Application of Foraging Theory in Archaeology. Journal of Archaeological Science 30:595610.Google Scholar
Lyman, R. Lee. 2004. Late-Quaternary Diminution and Abundance of Prehistoric Bison (Bison sp.) in Eastern Washington State, USA. Quaternary Research 62:7685.Google Scholar
Lyman, R. Lee. 2006a. Identifying Bilateral Pairs of Deer (Odocoileus sp.) Bones: How Symmetrical is Symmetrical Enough? Journal of Archaeological Science 33:12371255.Google Scholar
Lyman, R. Lee. 2006b. Paleozoology in the Service of Conservation Biology. Evolutionary Anthropology 15:1119.Google Scholar
Lyman, R. Lee. 2008a. Estimating the Magnitude of Data Asymmetry in Paleozoological Biogeography. International Journal of Osteoarchaeology 18:8594.Google Scholar
Lyman, R. Lee. 2008b. Quantitative Paleozoology. Cambridge University Press.Google Scholar
Lyman, R. Lee. 2010a. Paleozoology’s Dependence on Natural History Collections. Journal of Ethnobiology 30:126136.Google Scholar
Lyman, R. Lee. 2010b. Prehistoric Anthropogenic Impacts to Local and Regional Faunas are Not Ubiquitous. In The Archaeology of Anthropogenic Environments, edited by Dean, Rebecca M., pp. 204224. Center for Archaeological Investigations Occasional Paper 37, Southern Illinois University Press, Carbondale.Google Scholar
Lyman, R. Lee. 2011. Paleoecological and Biogeographical Implications of Late Pleistocene Noble Marten (Martes americana nobilis) in Eastern Washington State, USA. Quaternary Research 75:176182.Google Scholar
Lyman, R. Lee. 2012a. Biodiversity, Paleozoology, and Conservation Biology. In Paleontology in Ecology and Conservation, edited by Louys, Julien, pp. 147169. Springer, Heidelberg.Google Scholar
Lyman, R. Lee. 2012b. Human-Behavioral and Paleoecological Implications of Terminal Pleistocene Fox Remains at the Marmes Site (45FR50), Eastern Washington State, USA. Quaternary Science Reviews 41:3948.Google Scholar
Lyman, R. Lee. 2012c. The Influence of Screen-Mesh Size, and Size and Shape of Rodent Teeth on Recovery. Journal of Archaeological Science 39:18541861.Google Scholar
Lyman, R. Lee. 2012d. Rodent-Prey Content in Long-Term Samples of Barn Owl (Tyto alba) Pellets from the Northwestern United States Reflects Local Agricultural Change. American Midland Naturalist 167:150163.Google Scholar
Lyman, R. Lee. 2012e. A Warrant for Applied Paleozoology. Biological Reviews 87:513525.Google Scholar
Lyman, R. Lee. 2013. Taxonomic Composition and Body-Mass Distribution in the Terminal Pleistocene Mammalian Fauna from the Marmes Site, Southeastern Washington State, USA. Paleobiology 39:345359.Google Scholar
Lyman, R. Lee. 2014a. Paleoenvironmental Implications of Two Relative Indicator Rodent Taxa during the Pleistocene to Holocene Transition in Southeastern Washington State, USA. Journal of Quaternary Science 29:691697.Google Scholar
Lyman, R. Lee. 2014b. Terminal Pleistocene Change in Mammal Communities in Southeastern Washington State, USA. Quaternary Research 81:295304.Google Scholar
Lyman, R. Lee. 2015a. The History of “Laundry Lists” in North American Zooarchaeology. Journal of Anthropological Archaeology 39:4250.Google Scholar
Lyman, R. Lee. 2015b. On the Variable Relationship between NISP and NTAXA in Bird Remains and in Mammal Remains. Journal of Archaeological Science 53:291296.Google Scholar
Lyman, R. Lee. 2016. The Mutual Climatic Range Technique is (Usually) Not the Area of Sympatry Technique when Reconstructing Paleoenvironments Based on Faunal Remains. Palaeogeography, Palaeoclimatology, Palaeoecology 454:7581.Google Scholar
Lyman, R. Lee. 2017. Paleoenvironmental Reconstruction from Faunal Remains: Ecological Basics and Analytical Assumptions. Journal of Archaeological Research 25:315371.Google Scholar
Lyman, R. Lee. 2018a. A Critical Review of Four Efforts to Resurrect MNI in Zooarchaeology. Journal of Archaeological Method and Theory, doi.org/10.1007/s10816-018-9365-3.Google Scholar
Lyman, R. Lee. 2018b. The History of MNI in North American Zooarchaeology. In Zooarchaeology in Practice: Case Studies in Methodology and Interpretation in Archaeofaunal Analysis, edited by Giovas, Christina M. and LeFebvre, Michelle J., pp. 1333. Springer, New York.Google Scholar
Lyman, R. Lee. 2018c. Observations on the History of Zooarchaeological Quantitative Units: Why NISP, Then MNI, Then NISP Again? Journal of Archaeological Science: Reports 18:4350.Google Scholar
Lyman, R. Lee, and Ames, Kenneth M.. 2004. Sampling to Redundancy in Zooarchaeology: Lessons from the Portland Basin, Northwestern Oregon and Southwestern Washington. Journal of Ethnobiology 24:329346.Google Scholar
Lyman, R. Lee, and Ames, Kenneth M.. 2007. On the Use of Species-Area Curves to Detect the Effects of Sample Size. Journal of Archaeological Science 34:19851990.Google Scholar
Lyman, R. Lee, and Cannon, Kenneth P. (editors). 2004. Zooarchaeology and Conservation Biology. University of Utah Press, Salt Lake City.Google Scholar
Lyman, R. Lee, and Lyman, R. Jay. 2003. Lessons from Temporal Variation in the Mammalian Faunas from Two Collections of Owl Pellets in Columbia County, Washington. International Journal of Osteoarchaeology 13:150156.Google Scholar
Lyman, R. Lee, and O’Brien, Michael J.. 2000. Chronometers and Units in Early Archaeology and Paleontology. American Antiquity 65:691707.Google Scholar
Lyman, R. Lee, and O’Brien, Michael J.. 2005. Within-Taxon Morphological Diversity in Late-Quaternary Neotoma as a Paleoenvironmental Indicator, Bonneville Basin, Northwestern Utah, USA. Quaternary Research 63:274282.Google Scholar
Lyman, R. Lee, and VanPool, Todd L.. 2009. Metric Data in Archaeology: A Study of Intra-Analyst and Inter-Analyst Variation. American Antiquity 74:485504.Google Scholar
Lynch, Michael, and Gabriel, Wilfried. 1987. Environmental Tolerance. American Naturalist 129:283303.Google Scholar
MacArthur, Robert H. 1964. Environmental Factors Affecting Bird Species Diversity. American Naturalist 98:387397.Google Scholar
McCain, Christy M., and King, Sarah R. B.. 2014. Body Size and Activity Times Mediate Mammalian Responses to Climate Change. Global Change Biology 20:17601769.Google Scholar
McCown, T. D. 1961. Animals, Climate and Palaeolithic Man. Kroeber Anthropological Society Papers 25:221230.Google Scholar
McDonald, H. Gregory, and Bryson, Reid A.. 2010. Modeling Pleistocene Local Climatic Parameters Using Macrophysical Climate Modeling and the Paleoecology of Pleistocene Megafauna. Quaternary International 217:131137.Google Scholar
McDonald, Jerry N. 1981. North American Bison: Their Classification and Evolution. University of California Press, Berkeley.Google Scholar
MacFadden, Bruce J., Solounias, Nikos, and Cerling, Thure E.. 1999. Ancient Diets, Ecology, and Extinction of 5-Million-Year-Old Horses from Florida. Science 283:824827.Google Scholar
McGill, Brian J., Etienne, Rampal S., Gray, John S. et al. 2007. Species Abundance Distributions: Moving Beyond Single Prediction Theories to Integration within an Ecological Framework. Ecology Letters 10:9951015.Google Scholar
McGuire, Jenny L. 2011. Identifying California Microtus Species Using Geometric Morphometrics Documents Quaternary Geographic Range Contractions. Journal of Mammalogy 92: 13831394.Google Scholar
McGuire, Jenny L., and Davis, Edward B.. 2014. Conservation Paleobiogeography: The Past, Present and Future of Species Distributions. Ecography 37: 10921094.Google Scholar
McIntosh, Robert P. 1975. H. A. Gleason – “Individualistic Ecologist” 1882–1975: His Contributions to Ecological Theory. Bulletin of the Torrey Botanical Club 102:253273.Google Scholar
McIntosh, Robert P. 1990. Henry Allan Gleason and the Individualistic Hypothesis: The Structure of a Botanist’s Career. Botanical Review 56:91161.Google Scholar
McIntosh, Robert P. 1995. H. A. Gleason’s “Individualistic Concept” and Theory of Animal Communities: A Continuing Controversy. Biological Reviews 70:317357.Google Scholar
McIntosh, Robert P. 1998. The Myth of Community as Organism. Perspectives in Biology and Medicine 41:426438.Google Scholar
McNab, Brian K. 1971. On the Ecological Significance of Bergmann’s Rule. Ecology 52:845854.Google Scholar
McNab, Brian K. 2010. Geographic and Temporal Correlations of Mammalian Size Reconsidered: A Resource Rule. Oecologia 164:1323.Google Scholar
Madsen, David B. (editor). 2000. Late Quaternary Paleoecology in the Bonneville Basin. Bulletin 130. Utah Geological Survey, Salt Lake City.Google Scholar
Madsen, David B., Rhode, David, Grayson, Donald K. et al. 2001. Late Quaternary Environmental Change in the Bonneville Basin, Western USA. Palaeogeography, Palaeoclimatology, Palaeoecology 167:243271.Google Scholar
Magurran, Anne E. 1988. Ecological Diversity and its Measurement. Princeton University Press.Google Scholar
Magurran, Anne E. 2004. Measuring Biological Diversity. Blackwell, Malden, MA.Google Scholar
Magurran, Anne E., and Henderson, Peter A.. 2003. Explaining the Excess of Rare Species in Natural Species Abundance Distributions. Nature 422:714716Google Scholar
Manley, Brian F. J. 2005. Multivariate Statistical Methods: A Primer. Chapman and Hall/CRC, Boca Raton, FL.Google Scholar
Mannion, Philip D., and Upchurch, Paul. 2010. Completeness Metrics and the Quality of the Sauropodomorph Fossil Record through Geological and Historical Time. Paleobiology 36:283302.Google Scholar
Marean, Curtis W. 1992. Implications of Late Quaternary Mammalian Fauna from Lukenya Hill (South-Central Kenya) for Paleoenvironmental Change and Faunal Extinctions. Quaternary Research 37:239255.Google Scholar
Marean, Curtis W., and Cleghorn, Naomi. 2003. Large Mammal Skeletal Element Transport: Applying Foraging Theory in a Complex Taphonomic System. Journal of Taphonomy 1:1542.Google Scholar
Marean, Curtis W., Mudida, Nina, and Reed, Kaye E.. 1994. Holocene Paleoeonvironmental Change in the Kenyan Central Rift as Indicated by Micromammals from Ekapune Ya Muto Rockshelter. Quaternary Research 41:376389.Google Scholar
Marean, Curtis W., Cawthra, Haley C., Cowling, Richard M. et al. 2014. Stone Age People in a Changing South African Greater Cape Floristic Region. In Fynbos: Ecology, Evolution, and Conservation of a Megadiverse Region, edited by Allsopp, Nicky, Colville, Jonathan F. and Verboom, G. Anthony, pp. 164199. Oxford University Press.Google Scholar
Mares, Michael A., and Willig, Michael R.. 1994. Inferring Biome Associations of Recent Mammals from Samples of Temperate and Tropical Faunas: Paleoecological Considerations. Historical Biology 8:3148.Google Scholar
Marshall, Charles R. 1990. Confidence Intervals on Stratigraphic Ranges. Paleobiology 16:110.Google Scholar
Marshall, Fiona, and Pilgram, Tom. 1993. NISP vs. MNI in Quantification of Body-Part Representation. American Antiquity 58:261269.Google Scholar
Martin, Paul S. 1958. Pleistocene Ecology and Biogeography of North America. In Zoogeography, edited by Hubbs, Carl L., pp. 375420. Publication 51. American Association for the Advancement of Science, Washington, DC.Google Scholar
Martin, Robert A. 1984. The Evolution of Cotton Rat Body Mass. In Contributions in Quaternary Vertebrate Paleontology: A Volume in Honor of John E. Guilday, edited by Genoways, Hugh H. and Dawson, Mary R., pp. 252266. Special Publication 8. Carnegie Museum of Natural History, Pittsburgh.Google Scholar
Martin, Robert A. 1990. Estimating Body Mass and Correlated Variables in Extinct Mammals: Travels in the Fourth Dimension. In Body Size in Mammalian Paleobiology: Estimation and Biological Implications, edited by Damuth, John and MacFadden, Bruce J., pp. 4968. Cambridge University Press.Google Scholar
Martínez-Meyer, Enrique, A, Peterson, Townsend, and Hargrove, William W.. 2004. Ecological Niches as Stable Distributional Constraints on Mammal Species, with Implications for Pleistocene Extinctions and Climate Change Projections for Biodiversity. Global Ecology and Biogeography 13:305314.Google Scholar
Maser, Chris, Hammer, E. Wayne, and Anderson, Stanley H.. 1970. Comparative Food Habits of Three Owl Species in Central Oregon. The Murrelet 51:2933.Google Scholar
Matthews, Thalassa, Denys, Christiane, and Parkington, J. E.. 2005. The Palaeoecology of the Micromammals from the Late Middle Pleistocene Site of Hoedjiespunt 1 (Cape Province, South Africa). Journal of Human Evolution 49:432451.Google Scholar
Matthews, Thalassa, Rector, Amy L., Jacobs, Zenobia, Herries, Andy I. R., and Marean, Curtis W.. 2011. Environmental Implications of Micromammals Accumulated Close to the MIS 6 to MIS 5 Transition at Pinnacle Point Cave 9 (Mossel Bay, Western Cape Province, South Africa). Palaeogeography, Palaeoclimatology, Palaeoecology 302:213229.Google Scholar
Mayr, Ernst 1954. Change of Genetic Environment and Evolution. In Evolution as a Process, edited by Huxley, Julian, Hardy, A. C and Ford, E. B, pp. 157180. George Allen and Unwin, London.Google Scholar
Mayr, Ernst 1956. Geographical Character Gradients and Climatic Adaptations. Evolution 10:105108.Google Scholar
Mayr, Ernst 1970. Populations, Species, and Evolution. Harvard University Press, Cambridge, MA.Google Scholar
Mead, Jim I. 1987. Quaternary Records of Pika, Ochotona, in North America. Boreas 16:165171.Google Scholar
Mead, Jim I., and Grady, Frederick. 1996. Ochotona (Lagomorpha) from Late Quaternary Cave Deposits in Eastern North America. Quaternary Research 45:93101.Google Scholar
Mead, Jim I., and Spaulding, W. Geoffrey. 1995. Pika (Ochotona) and Paleoecological Reconstructions of the Intermountain West, Nevada and Utah. In Late Quaternary Environments and Deep History: A Tribute to Paul S. Martin, edited by Steadman, David W. and Mead, Jim I., pp. 165186. Scientific Papers 3. Mammoth Site of Hot Springs, Hot Springs, SD.Google Scholar
Mead, Jim I., and Spiess, Arthur E.. 2001. Reply to Russell Graham about Mustela macrodon. Quaternary Research 56:422423.Google Scholar
Mead, Jim I., Bell, Christopher J., and Murray, Lyndon K.. 1992. Mictomys borealis (Northern Bog Lemming) and the Wisconsin Paleoecology of the East-Central Great Basin. Quaternary Research 37:229238.Google Scholar
Mead, Jim I., Spiess, Arthur E., and Sobolik, Kristin D.. 2000. Skeleton of Extinct North American Sea Mink (Mustela macrodon). Quaternary Research 53:247262.Google Scholar
Meadow, Richard H. 1999. The Use of Size Index Scaling Techniques for Research on Archaeozoological Collections from the Middle East. In Historia Animalium ex Ossibus: Beiträge zur Paläoanatomie, Archäologie, Ägyptologie, Ethnologie und Geschichte der Tiermedizin, Festrchrift für Angela von den Driesch, edited by Becker, Cornelia, Manhart, Henriette, Peters, Joris, and Schibler, Jörg, pp. 285300. Verlag Marie Leidorf, Rahden/Westf.Google Scholar
Meiri, Shai. 2011. Bergmann’s Rule – What’s in a Name? Global Ecology and Biogeography 29:203207.Google Scholar
Meiri, Shai, Cooper, Natalie, and Purvis, Andy. 2008. The Island Rule: Made to Be Broken? Proceedings of the Royal Society B 275:141148.Google Scholar
Meiri, Shai, and Dayan, Tamar. 2003. On the Validity of Bergmann’s Rule. Journal of Biogeography 30:331351.Google Scholar
Meiri, Shai, Dayan, Tamar, and Simberloff, Daniel. 2005. Area, Isolation and Body Size Evolution in Insular Carnivores. Ecology Letters 8:12111217.Google Scholar
Meloro, Carlo, and Kovarovic, Kris. 2013. Spatial and Ecometric Analyses of the Plio-Pleistocene Large Mammal Communities of the Italian Peninsula. Journal of Biogeography 40:14511462.Google Scholar
Meltzer, David J. 2006. Folsom: New Archaeological Investigations of a Classic Paleoindian Bison Kill. University of California Press, Berkeley.Google Scholar
Mendoza, Manuel, and Palmqvist, Paul. 2008. Hypsodonty in Ungulates: An Adaptation for Grass Consumption or for Foraging in Open Habitat? Journal of Zoology (London) 274:134142.Google Scholar
Mendoza, Manuel, Janis, Christine M., and Palmqvist, Paul. 2002. Characterizing Complex Craniodental Patterns Related to Feeding Behaviour in Ungulates: A Multivariate Approach. Journal of Zoology (London) 258:223246.Google Scholar
Mendoza, Manuel, Janis, Christine M., and Palmqvist, Paul. 2005. Ecological Patterns in the Trophic-Size Structure of Large Mammal Communities: A “Taxon-Free” Characterization. Evolutionary Ecology Research 7:505530.Google Scholar
Menge, Bruce A., and Sutherland, John P.. 1987. Community Regulation: Variation in Disturbance, Competition, and Predation in Relation to Environmental Stress and Recruitment. American Naturalist 130:730757.Google Scholar
Merceron, Gildas, and Ungar, Peter. 2005. Dental Microwear and Paleoecology of Bovids from the Early Pliocene of Langebaanweg, Western Cape Province, South Africa. South African Journal of Science 101:365370.Google Scholar
Merceron, Gildas, Blondel, Cécile, Brunet, Michel et al. 2004. The Late Miocene Paleoenvironment of Afghanistan as Inferred from Dental Microwear in Artiodactyls. Palaeogeography, Palaeoclimatology, Palaeoecology 207:143163.Google Scholar
Merceron, Gildas, de Bonis, Louis, Viriot, Laurent, and Blondel, Cécile. 2005a. Dental Microwear of Fossil Bovids from Northern Greece: Paleoenvironmental Conditions in the Eastern Mediterranean During the Messinian. Palaeogeography, Palaeoclimatology, Palaeoecology 217:173185.Google Scholar
Merceron, Gildas, Blondel, Cécile, de Bonis, Louis, Kuofos, Georges D., and Viriot, Laurent. 2005b. A New Method of Dental Microwear Analysis: Application to Extant Primates and Ouranopithecus macedoniensis (Late Miocene of Greece). Palaios 20:551561.Google Scholar
Merriam, C. Hart. 1890. Results of a Biological Survey of the San Francisco Mountain Region and Desert of the Little Colorado. North American Fauna 3. Government Printing Office, Washington, DC.Google Scholar
Merriam, C. Hart. 1892. The Geographical Distribution of Life in North America with Special Reference to the Mammalia. Proceedings of the Biological Society of Washington 7:164.Google Scholar
Merriam, C. Hart. 1894. Laws of Temperature Control of the Geographic Distribution of Terrestrial Animals and Plants. National Geographic Magazine 6:229238.Google Scholar
Merriam, C. Hart. 1895. The Geographic Distribution of Animals and Plants in North America. Yearbook of the US Department of Agriculture for 1894, pp. 203214.Google Scholar
Merriam, C. Hart. 1898. Life Zones and Crop Zones in the United States. Bulletin of the Division of the Biological Survey, USDA 10:179.Google Scholar
Mihlbachler, Matthew C., and Solounias, Nikos. 2006. Coevolution of Tooth Crown Height and Diet in Oreodonts (Myerycoidodontidae, Artiodactyla) Examined with Phylogenetically Independent Contrasts. Journal of Mammalian Evolution 13:1136.Google Scholar
Mihlbachler, Matthew C., Rivals, Florent, Solounias, Nikos, and Semprebon, Gina M.. 2011. Dietary Change and Evolution of Horses in North America. Science 331:11781181.Google Scholar
Millar, J. S., and Hickling, G. J.. 1990. Fasting Endurance and the Evolution of Mammalian Body Size. Functional Ecology 4:512.Google Scholar
Miller, Alden H. 1937. Biotic Associations and Life-Zones in Relation to the Pleistocene Birds of California. Condor 39:248252.Google Scholar
Miller, G. H., Beaumont, Peter B., Deacon, H. J. et al. 1999. Earliest Modern Humans in Southern Africa Dated by Isoleucine Epimerization in Ostrich Eggshell. Quaternary Science Reviews 18:15371548.Google Scholar
Miller, Joshua H. 2011. Ghosts of Yellowstone: Multi-Decadal Histories of Wildlife Populations Captured by Bones on a Modern Landscape. PLoS One 6(3):318057.Google Scholar
Miller, Joshua H., Behrensmeyer, Anna K., Du, Andrew et al. 2014. Ecological Fidelity of Functional Traits Based on Presence–Absence in a Modern Mammalian Bone Assemblage (Amboseli, Kenya). Paleobiology 40:560583.Google Scholar
Millien, Virginie, Lyons, S. Kathleen, Olson, Link et al. 2006. Ecotypic Variation in the Context of Global Climate Change: Revisiting the Rules. Ecology Letters 9:853869.Google Scholar
Mills, J. R. E. 1955. Ideal Dental Occlusion in the Primates. Dental Practice 6:4761.Google Scholar
Minagawa, Masao, and Wada, Eitaro. 1984. Stepwise Enrichment of 15N Along Food Chains: Further Evidence and the Relation between δ15N and Animal Age. Geochimica et Cosmochimica Acta 48:11351140.Google Scholar
Mittelbach, Gary G., Steiner, Christopher F., Scheiner, Samuel M. et al. 2001. What Is the Observed Relationship between Species Richness and Productivity? Ecology 82:23812396.Google Scholar
Mix, Alan C., Morey, Ann E., Pisias, Nicklas G., and Hostetler, Steven W.. 1999. Foraminiferal Faunal Estimates of Paleotemperature: Circumventing the No-Analog Problem Yields Cool Ice Age Tropics. Paleoceanography 14:350359.Google Scholar
Moffett, R. O., and Deacon, H. J.. 1977. The Flora and Vegetation in the Surrounds of Boomplaas Cave: Cango Valley. South African Archaeological Bulletin 32:127145.Google Scholar
Moine, Olivier, Rousseau, Denis-Didier, Jolly, Dominique, and Vianey-Liaud, Marc. 2002. Paleoclimatic Reconstruction Using Mutual Climatic Range on Terrestrial Mollusks. Quaternary Research 57:162172.Google Scholar
Monchot, Hervé, and Gendron, Daniel. 2010. Disentangling Long Bones of Foxes (Vulpes vulpes and Alopex lagopus) from Arctic Archaeological Sites. Journal of Archaeological Science 37:799806.Google Scholar
Montuire, Sophie, and Marcolini, Frederica. 2002. Palaeoenvironmental Significance of the Mammalian Faunas of Italy since the Pliocene. Journal of Quaternary Science 17:8796.Google Scholar
Moore, Jason R., Norman, David B., and Upchurch, Paul. 2007. Assessing Relative Abundances in Fossil Assemblages. Palaeogeography, Palaeoclimatology, Palaeoecology 253:317322.Google Scholar
Morales, Arturo, and Rosenlund, Knud. 1979. Fish Bone Measurements: An Attempt to Standardize the Measuring of Fish Bones from Archaeological Sites. Steenstrupia, Copenhagen.Google Scholar
Morin, Eugène. 2012. Reassessing Paleolithic Subsistence: The Neandertal and Modern Human Foragers of Saint-Césaire. Cambridge University Press.Google Scholar
Morin, Eugène, Ready, Elspeth, Boileau, Arianne, Beauval, Cédric, and Coumont, Marie-Pierre. 2017a. Problems of Identification and Quantification in Archaeozoological Analysis, Part I: Insights from a Blind Test. Journal of Archaeological Method and Theory 24:886937.Google Scholar
Morin, Eugène, Ready, Elspeth, Boileau, Arianne, Beauval, Cédric, and Coumont, Marie-Pierre. 2017b. Problems of Identification and Quantification in Archaeozoological Analysis, Part II: Presentation of an Alternative Counting Method. Journal of Archaeological Method and Theory 24:938973.Google Scholar
Morin, Xavier, and Lechowicz, Martin J.. 2008. Contemporary Perspectives on the Niche that Can Improve Models of Species Range Shifts under Climate Change. Biology Letters 4:573576.Google Scholar
Morlan, Richard E. 1984. Biostratigraphy and Biogeography of Quaternary Microtine Rodents from Northern Yukon Territory, Eastern Beringia. In Contributions in Quaternary Vertebrate Paleontology: A Volume in Memorial to John E. Guilday, edited by Genoways, Hugh H. and Dawson, Mary R., pp. 184199. Special Publication 8. Carnegie Museum of Natural History, Pittsburgh.Google Scholar
Morlan, Richard E. 1991. Bison Carpal and Tarsal Measurements: Bulls versus Cows and Calves. Plains Anthropologist 36:215227.Google Scholar
Morlot, A[dolphe von]. 1861. General Views on Archaeology. Annual Report of the Smithsonian Institution for 1860, pp. 284343.Google Scholar
Mosbrugger, Volker. 2009. Nearest-Living-Relative Method. In Encyclopedia of Paleoclimatology and Ancient Environments, edited by Gornitz, Vivien, pp. 607609. Encyclopedia of Earth Sciences Series. Springer, Dordrecht.Google Scholar
Mota-Vargas, Claudio, and Rojas-Soto, Octavio R.. 2016. Taxonomy and Ecological Niche Modeling: Implications for the Conservation of Wood Partridges (Genus Dendrortyx). Journal for Nature Conservation 29:113.Google Scholar
Munro, Natalie D. 2004. Zooarchaeological Measures of Hunting Pressure and Occupation Intensity in the Natufian: Implications for Agricultural Origins. Current Anthropology 45:S5–S33.Google Scholar
Murcia, Carolina, Aronson, James, Kattan, Gustavo H. et al. 2014. A Critique of the “Novel Ecosystem” Concept. Trends in Ecology and Evolution 29:548553.Google Scholar
Nagaoka, Lisa. 2005. Differential Recovery of Pacific Island Fish Remains. Journal of Archaeological Science 32:941955.Google Scholar
Nekola, Jeffrey C., and White, Peter S.. 1999. The Distance Decay of Similarity in Biogeography and Ecology. Journal of Biogeography 26:867878.Google Scholar
Nel, Thurid H., and Henshilwood, Christopher S.. 2016. The Small Mammal Sequence from the C. 76–72 ka Still Bay Levels at Blombos Cave, South Africa: Taphonomic and Palaeoecological Implications for Human Behaviour. PLoS ONE 11:e0159817.Google Scholar
Nelson, Bruce K., DeNiro, Michael J., Schoeninger, Margaret J., De Paolo, Donald J., and Hare, P. E.. 1986. Effects of Diagenesis on Strontium, Carbon, Nitrogen and Oxygen Concentration and Isotopic Composition of Bone. Geochimica et Cosmochimica Acta 50:19411949.Google Scholar
Nelson, Robert S., and Semken, Holmes A., Jr. 1970. Paleoecological and Stratigraphic Significance of the Muskrat in Pleistocene Deposits. Geological Society of America Bulletin 81:37333738.Google Scholar
Nesbit Evans, E. M., van Couvering, Judith H., and Andrews, Peter. 1981. Palaeoecology of Miocene Sites in Western Kenya. Journal of Human Evolution 10:3548.Google Scholar
Nikita, E. 2014. Estimation of the Original Number of Individuals Using Multiple Skeletal Elements. International Journal of Osteoarchaeology 24:660664.Google Scholar
Nogués-Bravo, David. 2009. Predicting the Past Distribution of Species Climatic Niches. Global Ecology and Biogeography 18:521531.Google Scholar
Nowak, Robert S., Nowak, Cheryl L., and Tusch, Robin J.. 2000. Probability that a Fossil Absent from a Sample Is Also Absent from the Paleolandscape. Quaternary Research 54:144154.Google Scholar
O’Connell, James F., Hawkes, Kristen, and Blurton-Jones, Nicholas. 1988. Hadza Hunting, Butchering, and Bone Transport and their Archaeological Implications. Journal of Anthropological Research 44:113161.Google Scholar
O’Connor, Anne. 2007. Finding Time for the Old Stone Age: A History of Palaeolithic Archaeology and Quaternary Geology in Britain, 1860–1960. Oxford University Press.Google Scholar
O’Connor, Terry (editor). 2005. Biosphere to Lithosphere: New Studies in Vertebrate Taphonomy. Oxbow Books, Oxford.Google Scholar
Odling-Smee, F. John, Laland, Kevin N., and Feldman, Marcus W.. 2003. Niche Construction: The Neglected Process in Evolution. Princeton University Press.Google Scholar
Odum, Eugene P. 1971. Fundamentals of Ecology, third edition. W. B. Saunders, Philadelphia.Google Scholar
Odum, Eugene P., and Barrett, Gary W.. 2005. Fundamentals of Ecology, fifth edition. Thomson Brooks/Cole, Belmont, CA.Google Scholar
O’Gara, Bart W., and Yoakum, Jim D. (editors). 2004. Pronghorn: Ecology and Management. University Press of Colorado, Boulder.Google Scholar
Olander, Heikki, Korhola, Atte, and Blom, Tom. 1997. Surface Sediment Chironomidae (Insecta: Diptera) Distributions along an Ecotonal Transect in Subarctic Fennoscandia: Developing a Tool for Palaeotemperature Reconstructions. Journal of Paleolimnology 18:4559.Google Scholar
Olcott, Susan P., and Barry, Ronald E.. 2000. Environmental Correlates of Geographic Variation in Body Size of the Eastern Cottontail (Sylvilagus floridanus). Journal of Mammalogy 81:986998.Google Scholar
Olff, Han, Ritchie, M. E., and Prins, Herbert H. T.. 2002. Global Environmental Controls of Diversity in Large Herbivores. Nature 415:901905.Google Scholar
Olsen, John W. 1982. Prehistoric Environmental Reconstruction by Vertebrate Faunal Analysis. In Multidisciplinary Research at Grasshopper Pueblo, Arizona, edited by Longacre, William A., Holbrook, Sally J., and Graves, Michael W., pp. 6372. Anthropological Papers 40. University of Arizona, Tucson.Google Scholar
Olson, D. M., Dinerstein, E., Wikramanayake, E. D. et al. 2001. Terrestrial Ecoregions of the World: A New Map of Life on Earth. BioScience 51:933938.Google Scholar
Olszewski, Thomas D. 1999. Taking Advantage of Time Averaging. Paleobiology 25:226238.Google Scholar
O’Regan, Hannah J., and Turner, Alan. 2004. The Interface between Conservation Biology, Palaeontology and Archaeozoology: Morphometrics and Population Viability Analysis. In The Future from the Past, edited by Lauwerier, Roel C. G. M and Plug, Ina, pp. 9096. Oxbow Books, Oxford.Google Scholar
Orians, Gordon H., and Milewski, A. V.. 2007. Ecology of Australia: The Effects of Nutrient-Poor Soils and Intense Fires. Biological Reviews 82:393423.Google Scholar
Orlando, Ludovic, and Cooper, Alan. 2014. Using Ancient DNA to Understand Evolutionary and Ecological Processes. Annual Review of Ecology, Evolution and Systematics 45:573598.Google Scholar
Ortiz, J. D., and Mix, A. C.. 1997. Comparison of Imbrie-Kipp Transfer Function and Modern Analog Temperature Estimates Using Sediment Trap and Core Top Foraminiferal Faunas. Paleoceanography 12:175190.Google Scholar
Orton, David C. 2014. Biometry in Zooarchaeology. In Encyclopedia of Global Archaeology, edited by Smith, Claire, pp. 902910. Springer, New York.Google Scholar
Overpeck, Jonathan T., Webb, Robert S., and Webb, Thompson III. 1992. Mapping Eastern North American Vegetation Change of the Past 18 Ka: No-Analogs and the Future. Geology 20:10711074.Google Scholar
Owen, Pamela R., Bell, Christopher J., and Mead, Emilee M.. 2000. Fossils, Diet, and Conservation of Black-Footed Ferrets (Mustela nigripes). Journal of Mammalogy 81:422433.Google Scholar
Owen-Smith, R. Norman. 1988. Megaherbivores: The Influence of Very Large Body Size on Ecology. Cambridge University Press.Google Scholar
Palmqvist, Paul, Gröcke, Darren R., Arribas, Alfonso, and Fariña, Richard A.. 2003. Paleoecological Reconstruction of a Lower Pleistocene Large Mammal Community Using Biogeochemical (δ13C, δ15N, δ18O, Sr:Zn) and Ecomorphological Approaches. Paleobiology 29:205229.Google Scholar
Parmesan, Camille. 2006. Ecological and Evolutionary Responses to Recent Climate Change. Annual Review of Ecology, Evolution, and Systematics 37:637669.Google Scholar
Pate, F. Donald. 1994. Bone Chemistry and Paleodiet. Journal of Archaeological Method and Theory 1:161209.Google Scholar
Paterson, James D. 1990. Comment – Bergmann’s Rule is Invalid: A Reply to V. Geist. Canadian Journal of Zoology 68:16101612.Google Scholar
Patton, J. L., Pardiñas, U. F. J, and D’Elía, G. (editors). 2015. Mammals of South America, vol. ii: Rodents. University of Chicago Press.Google Scholar
Patton, Thomas H. 1963. Fossil Vertebrates from Miller’s Cave, Llano County, Texas. Texas Memorial Museum Bulletin 7. University of Texas, Austin.Google Scholar
Payne, Sebastian. 1969. A Metrical Distinction between Sheep and Goat Metacarpals. In The Domestication and Exploitation of Plants and Animals, edited by Ucko, Peter J. and Dimbleby, G. W, pp. 295305. Aldine Atherton, Chicago.Google Scholar
Peel, M. C., Finlayson, B. L., and McMahon, T. A.. 2007. Updated World Map of the Köppen-Geiger Climate Classification. Hydrology and Earth System Sciences 11:16331644.Google Scholar
Peet, Robert K. 1974. The Measurement of Species Diversity. Annual Review of Ecology and Systematics 5:285307.Google Scholar
Peet, Robert K., Knox, Robert G., Case, J. Stephen, and Allen, R. B.. 1988. Putting Things in Order: The Advantages of Detrended Correspondence Analysis. American Naturalist 131:924934.Google Scholar
Peppe, Daniel J., Royer, Dana L., Cariglino, Bárbara et al. 2011. Sensitivity of Leaf Size and Shape to Climate: Global Patterns and Paleoclimatic Applications. New Phytologist 190:724739.Google Scholar
Pérez-Barbería, F. J., and Gordon, I. J.. 2001. Relationships between Oral Morphology and Feeding Style in the Ungulata: A Phylogenetically Controlled Evaluation. Proceedings of the Royal Society B 268:10211030.Google Scholar
Pérez-Crespo, Víctor A., Barrón-Ortiz, Christian R., Arroyo-Barales, Joaquín et al. 2016. Preliminary Data on the Diet and Habitat Preferences of Capromeryx mexicana (Mammalia: Antilocapridae) from the Late Pleistocene of Cedral, San Luis Potosí, Mexico. Southwestern Naturalist 61:152155.Google Scholar
Peters, J., and Brink, James S.. 1992. Comparative Postcranial Osteomorphology and Osteometry of Springbok, Antidorcas marsupialis (Zimmerman, 1780) and Grey Rhebok, Pelea capreolus (Forster, 1790) (Mammalia: Bovidae). Navorsinge van die Nasionale Museum Bloemfontein 8:161206.Google Scholar
Peters, Robert H. 1983. The Ecological Implications of Body Size. Cambridge University Press.Google Scholar
Peterson, A. Townsend. 2011. Ecological Niche Conservatism: A Time-Structured Review of Evidence. Journal of Biogeography 38:817827.Google Scholar
Peterson, A. Townsend, and Soberón, Jorge. 2012. Species Distribution Modeling and Ecological Niche Modeling: Getting the Concepts Right. Natureza & Conservação 10:16.Google Scholar
Peterson, Charles H. 1977. The Paleoecological Significance of Undetected Short-Term Variability. Journal of Paleontology 51:976981.Google Scholar
Petit, J. R., Jouzel, J., Raynaud, D. et al. 1999. Climate and Atmospheric History of the Past 420,000 Years from the Vostok Ice Core, Antarctica. Nature 399:429436.Google Scholar
Peyron, Odile, Guiot, Joël, Cheddadi, Rachid et al. 1998. Climatic Reconstruction in Europe for 18,000 Yr BP from Pollen Data. Quaternary Research 49:183196.Google Scholar
Phillips, Arthur M., III, House, D. A., and Phillips, B. G.. 1989. Expedition to the San Francisco Peaks: C. Hart Merriam and the Life Zone Concept. Plateau 60:1930.Google Scholar
Pianka, E. R. 1978. Evolutionary Ecology, second edition. Harper and Row, New York.Google Scholar
Pianka, E. R. 1988. Evolutionary Ecology, fourth edition. Harper and Row, New York.Google Scholar
Pickering, Travis, Schick, Kathy, and Toth, Nicholas (editors). 2007. Breathing Life into Fossils: Taphonomic Studies in Honor of C. K. (Bob) Brain. Stone Age Institute Publication Series no. 2. Stone Age Institute, Gosport, IN.Google Scholar
Pielou, E. C. 1966. The Measurement of Diversity in Different Types of Biological Collections. Journal of Theoretical Biology 13:131144.Google Scholar
Pierce, Becky M., Bleich, Vernon C., Monteith, Kevin L., and Bowyer, R. Terry. 2012. Top-Down Versus Bottom-Up Forcing: Evidence from Mountain Lions and Mule Deer. Journal of Mammalogy 93:977988.Google Scholar
Pinto, C. Miguel, Soto-Centeno, J. Angel et al. 2016. Archaeology, Biogeography, and Mammalogy Do Not Provide Evidence for Tarukas (Cervidae: Hippocamelus antisensis) in Ecuador. Journal of Mammalogy 97:4153.Google Scholar
Pinto-Llona, Ana C. 2013. Macrowear and Occlusal Microwear on Teeth of Cave Bears Ursus spelaeus and Brown Bears Ursus arctos: Inferences Concerning Diet. Palaeogeography, Palaeoclimatology, Palaeoecology 370:4150.Google Scholar
Plug, Ina. 2005. Osteomorphological Differences between Some Skeletal Elements of Labeobarbus kimberleyensis, Labeobarbus aeneus and Labeo capensis (Pisces: Cyprinidae). Annals of the Transvaal Museum 42:517.Google Scholar
Plummer, Thomas W., and Bishop, Laura C.. 1994. Hominid Paleoecology at Olduvai Gorge, Tanzania, as Indicated by Antelope Remains. Journal of Human Evolution 27:4775.Google Scholar
Plummer, Thomas W., Bishop, Laura C., and Hertel, Fritz. 2008. Habitat Preference of Extant African Bovids Based on Astragalus Morphology: Operationalizing Ecomophology for Palaeoenvironmental Reconstruction. Journal of Archaeological Science 35:30163027.Google Scholar
Plummer, Thomas W., Bishop, Laura C., Ditchfield, Peter et al. 2009. The Environmental Context of Oldowan Hominin Activities at Kanjera South, Kenya. In Interdisciplinary Approaches to the Oldowan, edited by Hovers, Erella and Braun, David R., pp. 149160. Springer, Dordrecht.Google Scholar
Plummer, Thomas W., Ferraro, Joseph V., Louys, Julien et al. 2015. Bovid Ecomorphology and Hominin Paleoenvironments of the Shungura Formation, Lower Omo River Valley, Ethiopia. Journal of Human Evolution 88:108126.Google Scholar
Pokines, James T. 1998. The Paleoecology of Lower Magdalenian Cantabrian Spain. British Archaeological Reports, International Series 713. BAR, Oxford.Google Scholar
Polley, H. Wayne. 1997. Implications of Rising Atmospheric Carbon Dioxide Concentrations for Rangeleands. Journal of Range Management 50:562577.Google Scholar
Polley, H. Wayne, Mayeux, Herman S., Johnson, Hyrum B., and Tischler, Charles R.. 1997. Viewpoint: Atmospheric CO2, Soil Water and Shrub/Grass Ratios on Rangelands. Journal of Range Management 50:278284.Google Scholar
Polly, P. David. 2010. Tiptoeing through the Trophics: Geographic Variation in Carnivoran Locomotor Ecomorphology in Relation to Environment. In Carnivoran Evolution: New Views on Phylogeny, Form, and Function, edited by Goswami, Anjali and Friscia, Anthony, pp. 374410. Cambridge University Press.Google Scholar
Polly, P. David, and Eronen, Jussi T.. 2011. Mammal Associations in the Pleistocene of Britain: Implications of Ecological Niche Modelling and a Method for Reconstructing Palaeoclimate. In The Ancient Human Occupation of Britain, edited by Ashton, N., Lewis, S., and Stringer, C., pp. 279304. Developments in Quaternary Science 14. Elsevier, Amsterdam.Google Scholar
Polly, P. David, and Sarwar, Sana. 2014. Extinction, Extirpation, and Exotics: Effects on the Correlation between Traits and Environment at the Continental Level. Annales Zoologici Fennici 51:209226.Google Scholar
Polly, P. David, Eronen, Jussi T., Fred, Marianne et al. 2011. History Matters: Ecometrics and Integrative Climate Change Biology. Proceedings of the Royal Society B 278:11311140.Google Scholar
Posadas, P., Crisci, J. V., and Katinas, L.. 2006. Historical Biogeography: A Review of its Basic Concepts and Critical Issues. Journal of Arid Environments 66:389403.Google Scholar
Potts, Richard. 1988. Early Hominid Activities at Olduvai. Aldine de Gruyter, New York.Google Scholar
Prentice, Colin I., Harrison, S. P., and Bartlein, P. J.. 2011. Global Vegetation and Terrestrial Carbon Cycle Changes after the Last Ice Age. New Phytologist 189:988998.Google Scholar
Preston, F. W. 1948. The Commonness, and Rarity, of Species. Ecology 29:254283.Google Scholar
Price, Gilbert J., and Sobbe, Ian H.. 2005. Pleistocene Palaeoecology and Environmental Change on the Darling Downs, Southeastern Queensland, Australia. Memoirs of the Queensland Museum 51:171201.Google Scholar
Price, Gilbert J., Zhao, Jian-xin, Feng, Yue-xing, and Hocknull, Scott A.. 2009. New Records of Plio-Pleistocene Koalas from Australia: Palaeoecological and Taxonomic Implications. Records of the Australian Museum 61:3948.Google Scholar
Price, T. Douglas, Blitz, Jennifer, Burton, James, and Ezzo, Joseph A.. 1992. Diagenesis in Prehistoric Bone: Problems and Solutions. Journal of Archaeological Science 19:513529.Google Scholar
Prideaux, Gavin J., Long, John A., Ayliffe, Linda K. et al. 2007. An Arid-Adapted Middle Pleistocene Vertebrate Fauna from South-Central Australia. Nature 445:422425.Google Scholar
PRISM Climate Group. 2004. PRISM Gridded Climate Data. Oregon State University, http://prism.oregonstate.edu.Google Scholar
Purdue, James R. 1980. Clinal Variation of Some Mammals during the Holocene in Missouri. Quaternary Research 13:242258.Google Scholar
Purdue, James R. 1986. The Size of White-Tailed Deer (Odocoileus virginianus) during the Archaic Period in Central Illinois. In Foraging, Collecting, and Harvesting: Archaic Period Subsistence and Settlement in the Eastern Woodlands, edited by Neusius, Sarah W., pp. 6595. Occasional Paper 6, Center for Archaeological Investigations, Southern Illinois University, Carbondale.Google Scholar
Purdue, James R. 1987. Estimation of Body Weight of White-Tailed Deer from Bone Size. Journal of Ethnobiology 7:112.Google Scholar
Purdue, James R. 1989. Changes during the Holocene in Size of White-Tailed Deer (Odocoileus virginianus) from Central Illinois. Quaternary Research 32:307316.Google Scholar
Qiao, Huijie, Soberón, Jorge, and Peterson, A. Townsend. 2015. No Silver Bullets in Correlative Ecological Niche Modeling: Insights from Testing among Many Potential Algorithms for Niche Estimation. Methods in Ecology and Evolution 6:11261136.Google Scholar
Quade, Jay, Cerling, Thure E., Barry, John C. et al. 1992. A 16–Ma Record of Paleodiet Using Carbon and Oxygen Isotopes in Fossil Teeth from Pakistan. Chemical Geology 94:183192.Google Scholar
Quirt-Booth, Tina, and Cruz-Uribe, Kathryn. 1997. Analysis of Leporid Remains from Prehistoric Singagua Sites, Northern Arizona. Journal of Archaeological Science 24:945960.Google Scholar
Rabenold, Diana, and Pearson, Osbjorn M.. 2014. Scratching the Surface: A Critique of Lucas et al. (2013)’s Conclusion that Phytoliths Do Not Abrade Enamel. Journal of Human Evolution 74:130133.Google Scholar
Raia, Pasquale, Carotenuto, Francesco, Meloro, Carlo, Piras, Paola, and Pushkina, Diana. 2010. The Shape of Contention: Adaptation, History, and Contingency in Ungulate Mandibles. Evolution 64:14891503.Google Scholar
Rainger, Ronald. 1981. The Continuation of the Morphological Tradition: American Paleontology, 1880–1910. Journal of the History of Biology 14:129158.Google Scholar
Rainger, Ronald. 1985. Paleontology and Philosophy: A Critique. Journal of the History of Biology 18:267287.Google Scholar
Rainger, Ronald. 1997. Everett C. Olson and the Development of Vertebrate Paleoecology and Taphonomy. Archives of Natural History 24:373396.Google Scholar
Raper, Diana J., and Zander, Holli. 2009. Paleoecology: An Untapped Resource for Teaching Environmental Change. International Journal of Environmental and Science Education 4:441447.Google Scholar
Raup, David M. 1977. Stochastic Models in Evolutionary Paleobiology. In Patterns of Evolution as Illustrated by the Fossil Record, edited by Hallam, Anthony, pp. 5978. Elsevier, Amsterdam.Google Scholar
Raup, David M., and Stanley, Steven M.. 1971. Principles of Paleontology. W. H. Freeman, San Francisco.Google Scholar
Rea, Amadeo M. 1986. Verification and Reverification: Problems in Archaeofaunal Studies. Journal of Ethnobiology 6:918.Google Scholar
Real, Leslie A., and Brown, James H. (editors). 1991. Foundations of Ecology: Classic Papers with Commentaries. University of Chicago Press.Google Scholar
Réale, Denis, McAdam, Andrew G., Boutin, Stan, and Berteaux, Dominique. 2003. Genetic and Plastic Responses of a Northern Mammal to Climatic Change. Proceedings of the Royal Society B 270:591596.Google Scholar
Rector, Amy L., and Reed, Kaye E.. 2010. Middle and Late Pleistocene Faunas of Pinnacle Point and their Paleoecological Implications. Journal of Human Evolution 59:340357.Google Scholar
Rector, Amy L., and Verrelli, Brian C.. 2010. Glacial Cycling, Large Mammal Community Composition, and Trophic Adaptations in the Western Cape, South Africa. Journal of Human Evolution 58:90102.Google Scholar
Redding, Richard W. 1978. Rodents and the Archaeological Paleoenvironment: Considerations, Problems, and the Future. In Approaches to Faunal Analysis in the Middle East, edited by Meadow, Richard H. and Zeder, Melinda A., pp. 6368. Peabody Museum of Archaeology and Ethnology Bulletin 2. Harvard University, Cambridge, MA.Google Scholar
Reed, Charles A. 1963. Osteo-archaeology. In Science in Archaeology, edited by Brothwell, D. and Higgs, E., pp. 204216. Basic Books, New York.Google Scholar
Reed, Charles A., and Braidwood, Robert J.. 1960. Toward the Reconstruction of the Environmental Sequence of Northeastern Iraq. In Prehistoric Investigations in Iraqi Kurdistan, edited by Braidwood, Robert J. and Howe, Bruce, pp. 163173. Studies in Ancient Oriental Civilization 31. University of Chicago Press.Google Scholar
Reed, D. N. 2007. Serengeti Micromammals and their Implications for Olduvai Paleoenvironments. In Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence, edited by Bobe, R[ené, Alemseged, Z[eresenay, and Behrensmeyer, A[nna] K., pp. 217255. Springer, Dordrecht.Google Scholar
Reed, Kaye E. 1997. Early Hominid Evolution and Ecological Change through the African Plio-Pleistocene. Journal of Human Evolution 32:289322.Google Scholar
Reed, Kaye E. 1998. Using Large Mammal Communities to Examine Ecological and Taxonomic Structure and Predict Vegetation in Extant and Extinct Assemblages. Paleobiology 24:384408.Google Scholar
Reed, Kaye E. 2008. Paleoecological Patterns at the Hadar Hominin Site, Afar Regional State, Ethiopia. Journal of Human Evolution 54:743768.Google Scholar
Reed, Kaye E. 2013. Multiproxy Paleoecology: Reconstructing Evolutionary Context in Paleoanthropology. In A Companion to Paleoanthropology, edited by Begun, David R., pp. 204225. Blackwell, Malden, MA.Google Scholar
Reed, Kaye E., Spencer, Lillian M., and Rector, Amy L.. 2013. Faunal Approaches to Early Hominin Paleoecology. In Early Hominin Paleoecology, edited by Sponheimer, Matt, Lee-Thorp, Julia A., Reed, Kaye E., and Ungar, Peter S., pp. 334. University Press of Colorado, Boulder.Google Scholar
Reimer, P. J., Bard, E., Bayliss, A. et al. 2013. IntCal 13 and Marine 13 Radiocarbon Age Calibration Curves 0–50,000 Years cal BP. Radiocarbon 55:18691887.Google Scholar
Reitz, Elizabeth J., and Ruff, Barbara. 1994. Morphometric Data for Cattle from North America and the Caribbean Prior to the 1850s. Journal of Archaeological Science 21:699713.Google Scholar
Reitz, Elizabeth J., and Wing, Elizabeth S.. 2008. Zooarchaeology, second edition. Cambridge University Press.Google Scholar
Rensberger, John M. 1978. Scanning Electron Microscopy of Wear and Occlusal Events in Some Small Herbivores. In Development, Function, and Evolution of Teeth, edited by Butler, Percy M. and Joysey, Kenneth A., pp. 415438. Academic Press, New York.Google Scholar
Rensch, Bernhard. 1938. Some Problems of Geographical Variation and Species Formation. Proceedings of the Linnean Society of London 150:275285.Google Scholar
Reynolds, Sally C. 2007. Mammalian Body Size Changes and Plio-Pleistocene Environmental Shifts: Implications for Understanding Hominin Evolution in Eastern and Southern Africa. Journal of Human Evolution 53:528548.Google Scholar
Rhoades, Robert E. 1978. Archaeological Use and Abuse of Ecological Concepts and Studies: The Ecotone Example. American Antiquity 43:608614.Google Scholar
Rhodes, R. Sanders, II. 1984. Paleoecology and Regional Paleoclimatic Implications of the Farmdalian Craigmile and Woodfordian Waubonsie Mammalian Local Faunas, Southwestern Iowa. Reports of Investigations 40. Illinois State Museum, Springfield.Google Scholar
Rhodes, R. Sanders, II, and Semken, Holmes A., Jr. 1986. Quaternary Biostratigraphy and Paleoecology of Fossil Mammals from the Loess Hills Region of Western Iowa. Proceedings of the Iowa Academy of Sciences 93(3):94139.Google Scholar
Richter, Kristine, Wilson, Julie, Jones, Andrew K. G. et al. 2011. Fish ’n Chips: ZooMS Peptide Mass Fingerprinting in a 96 Well Plate Format to Identify Fish Bone Fragments. Journal of Archaeological Science 38:15021510.Google Scholar
Ricklefs, Robert E., and Schluter, Dolph (editors). 1993. Species Diversity in Ecological Communities: Historical and Geographical Perspectives. University of Chicago Press.Google Scholar
Ries, Leslie, Fletcher, Robert J., Jr., Battin, James, and Sisk, Thomas D.. 2004. Ecological Responses to Habitat Edges: Mechanisms, Models, and Variability Explained. Annual Review of Ecology, Evolution and Systematics 35:491522.Google Scholar
Rivals, Florent, and Lister, Adrian M.. 2016. Dietary Flexibility and Niche Partitioning of Large Herbivores through the Pleistocene of Britain. Quaternary Science Reviews 146:116133.Google Scholar
Rivals, Florent, and Semprebon, Gina M.. 2006. A Comparison of the Dietary Habits of a Large Sample of the Pleistocene Pronghorn Stockoceros onusrosagris from the Papago Springs Cave in Arizona to the Modern Antilocapra americana. Journal of Vertebrate Paleontology 26:495500.Google Scholar
Rivals, Florent, Mihlbachler, Matthew C., and Solounias, Nikos. 2007a. Effect of Ontogenetic-Age Distribution in Fossil and Modern Samples on the Interpretation of Ungulate Paleodiets Using the Mesowear Method. Journal of Vertebrate Paleontology 27:763767.Google Scholar
Rivals, Florent, Solounias, Nikos, and Mihlbachler, Matthew C.. 2007b. Evidence for Geographic Variation in the Diets of Late Pleistocene and Early Holocene Bison in North America, and Differences from the Diets of Recent Bison. Quaternary Research 68:338346.Google Scholar
Rivals, Florent, Solounias, Nikos, and Schaller, George B.. 2011. Diet of Mongolian Gazelles and Tibetan Antelopes from Steppe Habitats Using Premaxillary Shape, Tooth Mesowear and Microwear Analyses. Mammalian Biology 76:358364.Google Scholar
Rivals, Florent, Semprebon, Gina M., and Lister, Adrian. 2012. An Examination of Dietary Diversity Patterns in Pleistocene Proboscideans (Mammuthus, Palaeoloxodon, and Mammut) from Europe and North America as Revealed by Dental Microwear. Quaternary International 255:188195.Google Scholar
Rivals, Florent, Julien, Marie-Anne, Kuitems, Margot et al. 2015. Investigation of Equid Paleodiet from Schöningen 13 II-4 through Dental Wear and Isotopic Analyses: Archaeological Implications. Journal of Human Evolution 89:129137.Google Scholar
Roberts, D. L., Bateman, M. D., Murray-Wallace, C. V., Carr, A. S., and Holmes, P. J.. 2009. West Coast Dune Plumes: Climate Driven Contrasts in Dunefield Morphogenesis along the Western and Southern South African Coasts. Palaeogeography, Palaeoclimatology, Palaeoecology 271:2438.Google Scholar
Roberts, Linda J. 1982. The Formulation and Application of a Technique, Based on Phalanges, for Discriminating the Sex of Plains Bison (Bison bison bison). Unpublished Master of Arts thesis, Department of Anthropology, University of Manitoba, Winnipeg.Google Scholar
Roberts, Michael F. 1970. Late Glacial and Postglacial Environments in Southeastern Wyoming. Palaeogeography, Palaeoclimatology, Palaeoecology 8:517.Google Scholar
Rodríguez, Jesús. 1999. Use of Cenograms in Mammalian Palaeoecology: A Critical Review. Lethaia 32:331347.Google Scholar
Rogers, Raymond R., Eberth, David A., and Fiorillo, Anthony R. (editors). 2007. Bonebeds: Genesis, Analysis, and Paleobiological Significance. University of Chicago Press.Google Scholar
Romano, Marco. 2015. Reviewing the Term Uniformitarianism in Modern Earth Sciences. Earth-Science Reviews 148:6576.Google Scholar
Romer, Alfred S. 1961. Palaeozoological Evidence of Climate: (I) Vertebrates. In Descriptive Paleoclimatology, edited by Nairn, A. E. M, pp. 183206. Wiley-Interscience, New York.Google Scholar
Rosenzweig, Michael L. 1968. The Strategy of Body Size in Mammalian Carnivores. American Midland Naturalist 80:299315.Google Scholar
Rosenzweig, Michael L. 1995. Species Diversity in Space and Time. Cambridge University Press.Google Scholar
Rosner, Hillary. 2015. Pine Beetle Epidemic: The Bug That’s Eating the Woods. National Geographic (April).Google Scholar
Rosvold, Jørgen, Andersen, Reidar, Linnell, John D. C., and Jufthammer, Anne Karin. 2013. Cervids in a Dynamic Northern Landscape: Holocene Changes in the Relative Abundance of Moose and Red Deer at the Limits of their Distributions. The Holocene 23:11431150.Google Scholar
Rowan, John, Faith, J. Tyler, Gebru, Y., and Fleagle, John G.. 2015. Taxonomy and Paleoecology of Fossil Bovidae (Mammalia, Artiodactyla) from the Kibish Formation, Southern Ethiopia: Implications for Dietary Change, Biogeography, and the Structure of Living Bovid Faunas of East Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 420:210222.Google Scholar
Rowe, Rebecca J., and Terry, Rebecca C.. 2014. Small Mammal Responses to Environmental Change: Integrating Past and Present Dynamics. Journal of Mammalogy 95:11571174.Google Scholar
Roy, Kaustuv, Valentine, James W., Jablonski, David, and Kidwell, Susan M.. 1996. Scales of Climatic Variability and Time Averaging in Pleistocene Biotas: Implications for Ecology and Evolution. Trends in Ecology and Evolution 11:458463.Google Scholar
Ruddiman, William F. 2013. The Anthropocene. Annual Review of Earth and Planetary Sciences 41:4568.Google Scholar
Rudwick, Martin J. S. 1971. Uniformity and Progression: Reflections on the Structure of Biological Theory in the Age of Lyell. In Perspectives in the History of Science and Technology, edited by Roller, D. H. D, pp. 209227. Oklahoma State University Press, Norman.Google Scholar
Rudwick, Martin J. S. 1978. Charles Lyell’s Dream of a Statistical Paleontology. Paleontology 21:225244.Google Scholar
Rudwick, Martin J. S. 1985. The Meaning of Fossils: Episodes in the History of Palaeontology, second edition. University of Chicago Press.Google Scholar
Rudwick, Martin J. S. 1996. Cuvier and Brongniart, William Smith, and the Reconstruction of Geohistory. Earth Sciences History 15:2536.Google Scholar
Rudwick, Martin J. S. 1997. Georges Cuvier, Fossil Bone, and Geological Catastrophes: New Translations and Interpretations of the Primary Texts. University of Chicago Press.Google Scholar
Ruff, Christopher B. 1991. Climate and Body Shape in Hominid Evolution. Journal of Human Evolution 21:81105.Google Scholar
Ruff, Christopher B. 1994. Morphological Adaptation to Climate in Modern and Fossil Hominids. Yearbook of Physical Anthropology 37:65107.Google Scholar
Rull, Valentí. 2012. Palaeobiodiversity and Taxonomic Resolution: Linking Past Trends with Present Patterns. Journal of Biogeography 39:10051006.Google Scholar
Running, Steven W., Ramakrishna, R. Nemani, Heinsch, Faith Ann et al. 2004. A Continuous Satellite-Derived Measure of Global Terrestrial Primary Production. BioScience 54:547560.Google Scholar
Ryan, Alan S. 1979. Wear Striation Direction on Primate Teeth: A Scanning Electron Microscope Examination. American Journal of Physical Anthropology 50:155167.Google Scholar
Ryder, M. L. 1992. What Are We Measuring? Circaea 10(2):8182.Google Scholar
Rymer, L. 1978. The Use of Uniformitarianism and Analogy in Palaeocology, Particularly Pollen Analysis. In Biology and Quaternary Environments, edited by Walker, D. and Guppy, J. C, pp. 245257. Australian Academy of Science, Canberra.Google Scholar
Sachs, Harvey Maurice, Webb, T[hompson III, and Clark, D. R.. 1977. Paleoecological Transfer Functions. Annual Review of Earth and Planetary Sciences 5:159178.Google Scholar
Sanchez, Julia L. 1996. A Re-evaluation of Mimbres Faunal Subsistence. Kiva 61:295307.Google Scholar
Sand, Håkan, Cederlund, Göran, and Danell, Kjell. 1995. Geographical and Latitudinal Variation in Growth Patterns and Adult Body Size of Swedish Moose (Alces alces). Oecologia 102:433442.Google Scholar
Sanders, Howard L. 1968. Marine Benthic Diversity: A Comparative Study. American Naturalist 102:243282.Google Scholar
Sandford, Mary K. (editor). 1993. Investigations of Ancient Human Tissue: Chemical Analyses in Anthropology. Gordon and Breach, Langhorne, PA.Google Scholar
Sandweiss, Daniel H., and Kelley, Alice R.. 2012. Archaeological Contributions to Climate Change Research: The Archaeological Record as a Paleoclimatic and Paleoenvironmental Archive. Annual Review of Anthropology 41:371391.Google Scholar
Sankaran, Mahesh, Hanan, Niall P., Scholes, Robert J. et al. 2005. Determinants of Woody Cover in African Savannas. Nature 438:846849.Google Scholar
Sauer, John, Niven, Daniel, Hines, James et al. 2017. The North American Breeding Bird Survey, Results and Analysis 1966–2015. Version 2.07.2017. USGS Patuxent Wildlife Research Center, Laurel, MD.Google Scholar
Sayre, Nathan F. 2005. Ecological and Geographical Scale: Parallels and Potential for Integration. Progress in Human Geography 29:276290.Google Scholar
Scheffer, Marten, Carpenter, Stephen R., Lenton, Timothy M. et al. 2012. Anticipating Critical Transitions. Science 338:344348.Google Scholar
Schmidt, Niels M., and Jensen, Per M.. 2003. Changes in Mammalian Body Length over 175 Years: Adaptations to a Fragmented Landscape? Conservation Ecology 7:6 (online).Google Scholar
Schmidt, Niels M., and Jensen, Per M.. 2005. Concomitant Patterns in Avian and Mammalian Body Length Changes in Denmark. Ecology and Society 10:5 (online).Google Scholar
Schmitt, Dave N., and Lupo, Karen D.. 2012. The Bonneville Estates Rockshelter Rodent Fauna and Changes in Late Pleistocene–Middle Holocene Climates and Biogeography in the Northern Bonneville Basin, USA. Quaternary Research 78:95102.Google Scholar
Schmitt, Dave N., Madsen, David B., and Lupo, Karen D.. 2002. Small-Mammal Data on Early and Middle Holocene Climates and Biotic Communities in the Bonneville Basin, USA. Quaternary Research 58:255260.Google Scholar
Schneider, David C. 2001. The Rise of the Concept of Scale in Ecology. BioScience 51:545553.Google Scholar
Schoeninger, Margaret J. 1995. Stable Isotope Studies in Human Evolution. Evolutionary Anthropology 4:8398.Google Scholar
Schoeninger, Margaret J., and DeNiro, Michael J.. 1984. Nitrogen and Carbon Isotopic Composition of Bone Collagen from Marine and Terrestrial Animals. Geochimica et Cosmochimica Acta 48:625639.Google Scholar
Scholtz, Anton. 1986. Palynological and Palaeobotanical Studies in the Southern Cape. Unpublished Master of Arts thesis, University of Stellenbosch, South Africa.Google Scholar
Schoonmaker, Peter K. 1998. Paleoecological Perspectives on Ecological Scale. In Ecological Scale: Theory and Applications, edited by Peterson, David L. and Parker, V. Thomas, pp. 79103. Columbia University Press, New York.Google Scholar
Schoville, Benjamin J., and Otárola-Castillo, Erik. 2014. A Model of Hunter-Gatherer Skeletal Element Transport: The Effect of Prey Body Size, Carriers, and Distance. Journal of Human Evolution 73:114.Google Scholar
Schubert, Blaine W., Ungar, Peter S., Sponheimer, Matt, and Reed, Kaye E.. 2006. Microwear Evidence for Plio-Pleistocene Bovid Diets from Makapansgat Limeworks Cave, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 241:301319.Google Scholar
Schultz, Gerald E. 1967. Four Superimposed Late-Pleistocene Vertebrate Faunas from Southwest Kansas. In Pleistocene Extinctions: The Search for a Cause, edited by Martin, P[aul] S. and Wright, H. E, Jr., pp. 321336. Yale University Press, New Haven, CT.Google Scholar
Schultz, Gerald E. 1969. Geology and Paleontology of a Late Pleistocene Basin in Southwest Kansas. Special Paper 105. Geological Society of America, Boulder, CO.Google Scholar
Schultz, Gerald E. 2010. Pleistocene (Irvingtonian, Cudahyan) Vertebrates from the Texas Panhandle, and their Geographic and Paleoecologic Significance. Quaternary International 217: 195224.Google Scholar
Schweitzer, Franz R., and Wilson, M. L.. 1982. Byneskranskop 1, a Late Quaternary Living Site in the Southern Cape Province, South Africa. Annals of the South African Museum 88:1203.Google Scholar
Scott, G. H. 1963. Uniformitarianism, the Uniformity of Nature, and Paleoecology. New Zealand Journal of Geology and Geophysics 6:510527.Google Scholar
Scott, Jessica R. 2012a. Dental Microwear Texture Analysis of Extant African Bovidae. Mammalia 76:157174.Google Scholar
Scott, Jessica R. 2012b. Dental Microwear Texture Analysis of Pliocene Bovids from Four Early Hominin Fossil Sites in Eastern Africa: Implications for Paleoenvironmental Dynamics and Human Evolution. Unpublished Ph.D. thesis, Department of Anthropology, University of Arkansas, Fayettefille.Google Scholar
Scott, Robert S., and Barr, W. Andrew. 2014. Ecomorphology and Phylogenetic Risk: Implications for Habitat Reconstruction Using Fossil Bovids. Journal of Human Evolution 73:4757.Google Scholar
Scott, Robert S., Kappelman, John, and Kelley, Jay. 1999. The Paleoenvironment of Sivapithecus parvada. Journal of Human Evolution 36:245274.Google Scholar
Scott, Robert S., Ungar, Peter S., Bergstrom, Torbjorn S. et al. 2005. Dental Microwear Texture Analysis Shows Within-Species Diet Variability in Fossil Hominins. Nature 436:693695.Google Scholar
Scott, Robert S., Ungar, Peter S., Bergstrom, Torbjorn S. et al. 2006. Dental Microwear Texture Analysis: Technical Considerations. Journal of Human Evolution 51:339349.Google Scholar
Scott, Robert S., Teaford, Mark F., and Ungar, Peter S.. 2012. Dental Microwear Texture and Anthropoid Diets. American Journal of Physical Anthropology 147:551579.Google Scholar
Sealy, Judith, Lee-Thorp, Julia, Loftus, Emma, Faith, J. Tyler, and Marean, Curtis W.. 2016. Late Quaternary Environmental Change in the Southern Cape, South Africa, from Stable Carbon and Oxygen Isotopes in Faunal Tooth Enamel from Boomplaas Cave. Journal of Quaternary Science 31:919927.Google Scholar
Secord, Ross, Bloch, Jonathan I., Chester, Stephen G. B. et al. 2012. Evolution of the Earliest Horses Driven by Climate Change in the Paleocene–Eocene Thermal Maximum. Science 335:959962.Google Scholar
Seddon, Alistair W. R., Mackay, Anson W., Baker, Ambroise G. et al. 2014. Looking Forward through the Past: Identification of 50 Priority Research Questions in Palaeoecology. Journal of Ecology 102:256267.Google Scholar
Semken, Holmes A., Jr. 1966. Stratigraphy and Paleontology of the McPherson Equus Beds (Sandahl Local Fauna), McPherson County, Kansas. Contributions from the Museum of Paleontology 20:121178. University of Michigan, Ann Arbor.Google Scholar
Semken, Holmes A., 1980. Holocene Climatic Reconstructions Derived from the Three Micromammal Bearing Cultural Horizons of the Cherokee Sewer Site, Northwestern Iowa. In The Cherokee Excavations, edited by Anderson, Duane C. and Semken, Holmes A., Jr., pp. 6799. Academic Press, New York.Google Scholar
Semken, Holmes A., 1983. Holocene Mammalian Biogeography and Climatic Change in the Eastern and Central United States. In Late-Quaternary Environments of the United States, vol. ii: The Holocene, edited by Wright, H. E, Jr., pp. 182207. University of Minnesota Press, Minneapolis.Google Scholar
Semken, Holmes A., 1988. Environmental Interpretations of the “Disharmonous” Late Wisconsinan Biome of Southeastern North America. In Late Pleistocene and Early Holocene Paleoecology and Archeology of the Eastern Great Lakes Region, edited by Laub, Richard S., Miller, Norton G., and Steadman, David W., pp. 185194. Bulletin 33. Buffalo Society of Natural Sciences, Buffalo, NY.Google Scholar
Semken, Holmes A., Jr., and Graham, Russell W.. 1987. Summary: Environmental Analysis and Plains Archaeology. In Late Quaternary Mammalian Biogeography and Environments of the Great Plains and Prairies, edited by Graham, Russell W., Semken, Holmes A., Jr., and Graham, Mary Ann, pp. 474480. Scientific Papers 22. Illinois State Museum, Springfield.Google Scholar
Semken, Holmes A., Jr., and Graham, Russell W.. 1996. Paleoecologic and Taphonomic Patterns Derived from Correspondence Analysis of Zooarchaeological and Paleontological Faunal Samples, a Case Study from the North American Prairie/Forest Ecotone. Acta Zoologica Cracoviensia 39:477490.Google Scholar
Semken, Holmes A., Jr., and Wallace, Steven C.. 2002. Key to Arvicoline (“Microtine” Rodents) and Arvicoline-Like Lower First Molars Recovered from Late Wisconsinan and Holocene Archaeological and Palaeontological Sites in Eastern North America. Journal of Archaeological Science 29:2331.Google Scholar
Semken, Holmes A., Jr., Graham, Russell W., and Stafford, Thomas W., Jr. 2010. AMS 14C Analysis of Late Pleistocene Non-Analog Faunal Components from 21 Cave Deposits in Southeastern North America. Quaternary International 217:240255.Google Scholar
Semprebon, Gina M., Rivals, Florent, Solounias, Nikos, and Hulbert, Richard C., Jr. 2016a. Paleodietary Reconstruction of Fossil Horses from the Eocene through Pleistocene of North America. Palaeogeography, Palaeoclimatology, Palaeoecology 442:110127.Google Scholar
Semprebon, Gina M., Tao, Deng, Hasjanova, Jelena, and Solounias, Nikos. 2016b. An Examination of the Dietary Habits of Platybelodon grangeri from the Linxia Basin of China: Evidence from Dental Microwear of Molar Teeth and Tusks. Palaeogeography, Palaeoclimatology, Palaeoecology 457:109116.Google Scholar
Sénégas, F., and Thackeray, J. F.. 2008. Temperature Indices Based on Relative Abundances of Rodent Taxa in South Africa Plio-Pleistocene Assemblages. Annals of the Transvaal Museum 45:143144.Google Scholar
Sepkoski, David. 2012. Rereading the Fossil Record: The Growth of Paleobiology as an Evolutionary Discipline. University of Chicago Press.Google Scholar
Sepkoski, J. John, Jr. 1988. Alpha, Beta, or Gamma: Where Does All the Diversity Go? Paleobiology 14:221234.Google Scholar
Sesé Benito, Carmen. 1994. Paleoclimatical Interpretation of the Quaternary Small Mammals of Spain. Geobios 27:753767.Google Scholar
Sexton, Jason P., McIntyre, Patrick J., Angert, Amy L., and Rice, Kevin J.. 2009. Evolution and Ecology of Range Limits. Annual Review of Ecology, Evolution and Systematics 40:415436.Google Scholar
Shabel, Alan B., Barnosky, Anthony D., Van Leuvan, Tonya, Bibi, Faysal, and Kaplane, Matthew H.. 2004. Irvingtonian Mammals from the Badger Room in Porcupine Cave: Age, Taphonomy, Climate, and Ecology. In Biodiversity Response to Climate Change in the Middle Pleistocene: The Porcupine Cave Fauna from Colorado, edited by Barnosky, Anthony D., pp. 295317. University of California Press, Berkeley.Google Scholar
Shaffer, Brian S., and Sanchez, Julia L. J.. 1994. Comparison of 1/8’’- and 1/4’’- Mesh Recovery of Controlled Samples of Small-to-Medium-Sized Mammals. American Antiquity 59:525530.Google Scholar
Shaffer, Brian S., and Schick, Christopher P.. 1995. Environment and Animal Procurement by the Mogollon of the Southwest. North American Archaeologist 16:117132.Google Scholar
Shapiro, Amy E., Venkataraman, Virek V., Nguyen, Nga, and Fashing, Peter J.. 2016. Dietary Ecology of Fossil Theropithecus: Inferences from Dental Microwear Textures of Extant Geladas from Ecologically Diverse Sites. Journal of Human Evolution 99:19.Google Scholar
Sharp, Zachary. 2007. Principles of Stable Isotope Geochemistry. Pearson Prentice Hall, Upper Saddle River, NJ.Google Scholar
Sheets, H. David., and Mitchell, Charles E.. 2001. Why the Null Matters: Statistical Tests, Random Walks, and Evolution. Genetica 112–113:105125.Google Scholar
Shelford, Victor E. 1913. Animal Communities in Temperate America. University of Chicago Press.Google Scholar
Shelford, Victor E. 1931. Some Concepts of Bioecology. Ecology 12:455467.Google Scholar
Sheridan, Jennifer A., and Bickford, David. 2011. Shrinking Body Size as an Ecological Response to Climate Change. Nature Climate Change 1:401406.Google Scholar
Shi, Guang R. 1993. Multivariate Data Analysis in Palaeoecology and Palaeobiogeography: A Review. Palaeogeography, Palaeoclimatology, Palaeoecology 105:199234.Google Scholar
Shipman, Pat, and Harris, J. M.. 1988. Habitat Preference and Paleoecology of Australopithecus boisei in Eastern Africa. In Evolutionary History of the “Robust” Australopithecines, edited by Grine, Frederick E., pp. 343381. Aldine De Gruyter, New York.Google Scholar
Sholander, P. F. 1955. Evolution of Climatic Adaptation in Homeotherms. Evolution 9:1526.Google Scholar
Shotwell, J. Arnold. 1955. An Approach to the Paleoecology of Mammals. Ecology 36:327337.Google Scholar
Shotwell, J. Arnold. 1958. Inter-Community Relationships in Hemphillian (Mid-Pliocene) Mammals. Ecology 39:271282.Google Scholar
Shotwell, J. Arnold. 1963. The Juntura Basin: Studies in Earth History and Paleoecology. Transactions of the American Philosophical Society 53:177.Google Scholar
Shurin, Jonathan B., and Allen, Emily G.. 2001. Effects of Competition, Predation, and Dispersal on Species Richness at Local and Regional Scales. American Naturalist 158:624637.Google Scholar
Simberloff, Daniel. 1998. Flagships, Umbrellas, and Keystones: Is Single-Species Management Passé in the Landscape Era? Biological Conservation 83:247257.Google Scholar
Simpson, George G. 1936. Data on the Relationships of Local and Continental Mammalian Faunas. Journal of Paleontology 10:410414.Google Scholar
Simpson, George G. 1937. The Fort Union of the Crazy Mountain Field, Montana, and its Mammalian Faunas. United States National Museum Bulletin 169. Smithsonian Institution, Washington, DC.Google Scholar
Simpson, George G. 1940. Mammals and Land Bridges. Journal of the Washington Academy of Science 30:137163.Google Scholar
Simpson, George G. 1942. The Beginnings of Vertebrate Paleontology in North America. Proceedings of the American Philosophical Society 81:130188.Google Scholar
Simpson, George G. 1943a. Criteria for Genera, Species, and Subspecies in Zoology and Paleozoology. Annals of the New York Academy of Sciences 44:145178.Google Scholar
Simpson, George G. 1943b. Mammals and the Nature of Continents. American Journal of Science 241:131.Google Scholar
Simpson, George G. 1947. Holarctic Mammalian Faunas and Continental Relationship during the Cenozoic. Geological Society of America Bulletin 58:613688.Google Scholar
Simpson, George G. 1952. Probabilities of Dispersal in Geologic Time. Bulletin of the American Museum of Natural History 99:163176.Google Scholar
Simpson, George G. 1953. Life of the Past: An Introduction to Paleontology. Yale University Press, New Haven, CT.Google Scholar
Simpson, George G. 1964. Species Density of North American Recent Mammals. Systematic Zoology 13:5773.Google Scholar
Simpson, George G. 1970. Uniformitarianism: An Inquiry into Principle, Theory, and Method in Geohistory and Biohistory. In Essays in Evolution and Genetics, edited by Hecht, M. K and Steere, W. C, pp. 4396. Appleton-Century-Crofts, New York.Google Scholar
Simpson, George G., Roe, Anna, and Lewontin, Richard C.. 1960. Quantitative Zoology, revised edition. Harcourt, Brace, New York.Google Scholar
Sinninghe Damsté, Jaap S., Verschuren, Dirk, Osssebaar, Jort et al. 2011. A 25,000–Year Record of Climate-Induced Changes in Lowland Vegetation of Eastern Equatorial Africa Revealed by Stable Carbon-Isotopic Composition of Fossil Plant Leaf Waxes. Earth and Planetary Science Letters 302:236246.Google Scholar
Skead, C. J. 2011. Historical Incidence of the Large Land Mammals in the Broader Western and Northern Cape. Centre for African Conservation Ecology, Nelson Mandela Metropolitan University, Port Elizabeth.Google Scholar
Skinner, J. D., and Chimimba, Christian T.. 2005. The Mammals of the Southern African Subregion. Cambridge University Press.Google Scholar
Slaughter, Bob H. 1966. The Moore Pit Local Fauna: Pleistocene of Texas. Journal of Paleontology 40:7891.Google Scholar
Slaughter, Bob H. 1967. Animal Ranges as a Clue to Late-Pleistocene Extinction. In Pleistocene Extinctions: The Search for a Cause, edited by Martin, P[aul] S. and Wright, H. E, Jr., pp. 155167. Yale University Press, New Haven, CT.Google Scholar
Smith, Benjamin, and Wilson, J. Bastow. 1996. A Consumer’s Guide to Evenness Indices. Oikos 76:7082.Google Scholar
Smith, C. Lavett. 1954. Pleistocene Fishes of the Berends Fauna of Beaver County, Oklahoma. Copeia 1954:282289.Google Scholar
Smith, Douglas W., Peterson, Rolf O., and Houston, Douglas B.. 2003. Yellowstone after Wolves. BioScience 53:330340.Google Scholar
Smith, Felisa A., and Betancourt, Julio L.. 1998. Response of Bushy-Tailed Woodrats (Neotoma cinerea) to Late Quaternary Climatic Change in the Colorado Plateau. Quaternary Research 50:111.Google Scholar
Smith, Felisa A., and Betancourt, Julio L.. 2003. The Effect of Holocene Temperature Fluctuations on the Evolution and Ecology of Neotoma (Woodrats) in Idaho and Northwestern Utah. Quaternary Research 59:160171.Google Scholar
Smith, Felisa A., and Lyons, S. Kathleen. 2011. How Big Should a Mammal Be? A Macroecological Look at Mammalian Body Size over Space and Time. Philosophical Transactions of the Royal Society B 366:23642378.Google Scholar
Smith, Felisa A., Betancourt, Julio L., and Brown, James H.. 1995. Evolution of Body Size in the Woodrat over the Past 25,000 Years of Climate Change. Science 270:20122014.Google Scholar
Smith, Felisa A., Browning, Hillary, and Sheperd, Ursula L.. 1998. The Influence of Climate Change on the Body Mass of Woodrats Neotoma in an Arid Region of New Mexico, USA. Ecography 21:140148.Google Scholar
Smith, Felisa A., Crawford, Dolly L., Harding, Larisa E. et al. 2009. A Tale of Two Species: Extirpation and Range Expansion during the Late Quaternary in an Extreme Environment. Global and Planetary Change 65:122133.Google Scholar
Smith, Felisa A., Gittleman, John L., and Brown, James H. (editors). 2014. Foundations of Macroecology: Classic Papers with Commentaries. University of Chicago Press.Google Scholar
Smith, James P. 1919. Climatic Relations of the Tertiary and Quaternary Faunas of the California Region. Proceedings of the California Academy of Sciences (fourth series) 9(4):123173.Google Scholar
Smith, Philip W. 1957. An Analysis of Post-Wisconsin Biogeography of the Prairie Peninsula Region Based on Distributional Phenomena among Terrestrial Vertebrate Populations. Ecology 38:205218.Google Scholar
Soberón, Jorge. 2007. Grinnellian and Eltonian Niches and Geographic Distributions of Species. Ecology Letters 10:11151123.Google Scholar
Soberón, Jorge, and Arroyo-Peña, B.. 2017. Are Fundamental Niches Larger than the Realized? Testing a 50–Year-Old Prediction by Hutchinson. PLoS One 12(4): e0175138.Google Scholar
Soberón, Jorge, and Nakamura, Miguel. 2009. Niches and Distributional Areas: Concepts, Methods, and Assumptions. Proceedings of the National Academy of Sciences USA 106 (Supplement 2):1964419650.Google Scholar
Soberón, Jorge, and Peterson, A. Townsend. 2005. Interpretation of Models of Fundamental Ecological Niches and Species’ Distributional Areas. Biodiversity Informatics 2:110.Google Scholar
Socha, Pawel. 2014. Rodent Palaeofaunas from Bisnick Cave (Kraków–Czestochowa Upland, Poland): Palaeoecological, Palaeoclimatic and Biostratigraphic Reconstruction. Quaternary International 326–327:6481.Google Scholar
Sokal, Robert R., and Rohlf, F. James. 1995. Biometry, third edition. W. H. Freeman, New York.Google Scholar
Soligo, Christophe. 2002. Primatology, Paleoecology, and a New Method for Assessing Taphonomic Bias in Fossil Assemblages. Evolutionary Anthropology (Supplement) 1:2427.Google Scholar
Soligo, Christophe and Andrews, Peter. 2005. Taphonomic, Taxonomic and Historical Bias of Faunal Structure in Early Hominin Localities. Journal of Human Evolution 49:206229.Google Scholar
Solounias, Nikos, and Moelleken, Sonja M. C.. 1992a. Dietary Adaptations of Two Goat Ancestors and Evolutionary Considerations. Geobios 25:797809.Google Scholar
Solounias, Nikos, and Moelleken, Sonja M. C.. 1992b. Tooth Microwear Analysis of Eotragus sansaniensis (Mammalia: Ruminantia), One of the Oldest Known Bovids. Journal of Vertebrate Paleontology 12:113121.Google Scholar
Solounias, Nikos, Tariq, Muhammad, Hou, Sukuan, Danowitz, Melinda, and Harrison, Mary. 2014. A New Method of Tooth Mesowear and a Test of It on Domestic Goats. Annales Zoologici Fennici 51:111118.Google Scholar
Southwood, T. R. E. 1987. The Concept and Nature of the Community. In Organization of Communities: Past and Present, edited by Gee, James H. R and Giller, Paul S., pp. 327. Blackwell Scientific, Oxford.Google Scholar
Sparks, B. W. 1961. The Ecological Interpretation of Quaternary Non-Marine Mollusca. Proceedings of the Linnean Society of London 172:7180.Google Scholar
Spellerberg, Ian F., and Fedor, Peter J.. 2003. A Tribute to Claude Shannon (1916–2001) and a Plea for More Rigorous Use of Species Richness, Species Diversity and the “Shannon-Wiener” Index. Global Ecology and Biogeography 12:177179.Google Scholar
Spencer, Lillian M. 1995. Morphological Correlates of Dietary Resource Partitioning in the African Bovidae. Journal of Mammalogy 76:448471.Google Scholar
Spencer, Lillian M. 1997. Dietary Adaptations of Plio-Pleistocene Bovidae: Implications for Hominid Habitat Use. Journal of Human Evolution 32:201228.Google Scholar
Sponheimer, Matt, and Lee-Thorp, Julia A.. 2001. The Oxygen Isotope Composition of Mammalian Enamel Carbonate from Morea Estate, South Africa. Oecologia 126:153157.Google Scholar
Sponheimer, Matt, and Lee-Thorp, Julia A.. 2003. Using Carbon Isotope Data of Fossil Bovid Communities for Palaeoenvironmental Reconstruction. South African Journal of Science 99:273275.Google Scholar
Sponheimer, Matt, and Lee-Thorp, Julia A.. 2006. Enamel Diagenesis at South African Australopith Sites: Implications for Paleoecological Reconstruction with Trace Elements. Geochimica et Cosmochimica Acta 70:16441654.Google Scholar
Sponheimer, Matt, Reed, Kaye, and Lee-Thorp, Julia A.. 1999. Combining Isotopic and Ecomorphological Data to Refine Bovid Paleodietary Reconstruction: A Case Study from the Makapansgat Limeworks Hominin Locality. Journal of Human Evolution 36:705718.Google Scholar
Sponheimer, Matt, Lee-Thorp, Julia A., DeRuiter, Darryl J. et al. 2003. Diets of Southern African Bovidae: Stable Isotope Evidence. Journal of Mammalogy 84:471479.Google Scholar
Staff, George M., Powell, Eric N., Stanton, Robert J., and Cummins, Hays. 1985. Biomass – Is It a Useful Tool in Paleocommunity Reconstruction? Lethaia 18:209232.Google Scholar
Stafford, Thomas W., Semken, Holmes A., Jr., Graham, Russell W. et al. 1999. First Accelerator Mass Spectrometry 14C Dates Documenting Contemporaneity of Nonanalog Species in Late Pleistocene Mammal Communities. Geology 27:903906.Google Scholar
Stahl, Peter W. 2005. An Exploratory Osteological Study of the Muscovy Duck (Cairina moschata) (Aves: Anatidae) with Implications for Neotropical Archaeology. Journal of Archaeological Science 32:915929.Google Scholar
Staver, A. Carla, Archibald, Sally, and Levin, Simon A.. 2011. The Global Extent and Determinants of Savanna and Forest as Alternative Biome States. Science 334:230232.Google Scholar
Steele, Teresa E. 2003. Using Mortality Profiles to Infer Behavior in the Fossil Record. Journal of Mammalogy 84:418430.Google Scholar
Steele, Teresa E. 2005. Comparing Methods for Analyzing Mortality Profiles in Zooarchaeological and Paleontological Samples. International Journal of Osteoarchaeology 15:404420.Google Scholar
Stegner, Mary Allison. 2015. The Mescal Cave Fauna (San Bernadino County, California) and Testing Assumptions of Habitat Fidelity in the Quaternary Fossil Record. Quaternary Research 83:582587.Google Scholar
Steinhauser, F. 1979. Climatic Atlas of North and Central America: I, Maps of Mean Temperature and Precipitation. World Meteorological Organization, UNESCO, and Cartographia, Geneva.Google Scholar
Stephens, John J. 1960. Stratigraphy and Paleontology of a Late Pleistocene Basin, Harper County, Oklahoma. Geological Society of America Bulletin 71:15751702.Google Scholar
Sterling, K. B. 1989. Builders of the US Biological Survey, 1885–1930. Journal of Forest History 33:180187.Google Scholar
Stewart, John R. 2005. The Use of Modern Geographical Ranges in the Identification of Archaeological Bird Remains. Documenta Archaeobiologiae 3:4354.Google Scholar
Stewart, John R. 2009. The Quaternary Fossil Record as a Source of Data for Evidence-Based Conservation: Is the Past the Key to the Future? In Holocene Extinctions, edited by Turvey, Samuel T., pp. 249261. Oxford University Press.Google Scholar
Stiner, Mary C. 1990. The Use of Mortality Patterns in Archaeological Studies of Hominid Predatory Adaptations. Journal of Anthropological Archaeology 9:305351.Google Scholar
Stiner, Mary C., Munro, Natalie D., and Surovell, Todd A.. 2000. The Tortoise and the Hare: Small-Game Use, the Broad Spectrum Revolution, and Paleolithic Demography. Current Anthropology 41:3973.Google Scholar
Stock, Chester. 1929. A Census of the Pleistocene Mammals of Rancho La Brea, Based on the Collections of the Los Angeles Museum. Journal of Mammalogy 10:281289.Google Scholar
Stonehouse, Bernard. 1997. Animal Responses to Climate. In Applied Climatology: Principles and Practice, edited by Thompson, Russell D. and Parry, Allen, pp. 141152. Routledge, London.Google Scholar
Storer, John E. 2003. Environments of Pleistocene Beringia: Analysis of Faunal Composition Using Cenograms. In Advances in Mammoth Research, edited byReumer, Jelle W. F, De Vos, John, and Mol, Dick, pp. 405414. Deinsea 9. Natural History Museum, Rotterdam.Google Scholar
Street-Perrott, F. Alayne, and Perrott, R. A.. 1993. Holocene Vegetation, Lake Levels and Climate of Africa. In Global Climates since the Last Glacial Maximum, edited by Wright, H. E Jr., Kutzbach, J. E, Webb, T. III, Ruddimann, W. F, Street-Perrott, F. A, and Bartlein, P. J, pp. 318356. University of Minnesota Press, Minneapolis.Google Scholar
Stynder, Deano D. 2009. The Diets of Ungulates from the Hominid Fossil-Bearing Site of Elandsfontein, Western Cape, South Africa. Quaternary Research 71:6270.Google Scholar
Su, Denise F., and Harrison, Terry. 2007. The Paleoecology of the Upper Laetolil Beds at Laetoli: A Reconsideration of the Large Mammal Evidence. In Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence, edited by Bobe, R[ené, Alemseged, Z[eresenay, and Behrensmeyer, A[nna] K., pp. 279313. Springer, Dordrecht.Google Scholar
Sukselainen, Leena, Fortelius, Mikael, and Harrison, Terry. 2015. Co-occurrence of Pliopithecoid and Hominoid Primates in the Fossil Record: An Ecometric Analysis. Journal of Human Evolution 84:2541.Google Scholar
Svenning, Jens-Christian, Eiserhardt, Wolf L., Normand, Signe, Ordonez, Alejandro, and Sandel, Brody. 2015. The Influence of Paleoclimate on Present-Day Patterns in Biodiversity and Ecosystems. Annual Review of Ecology, Evolution, and Systematics 46:551572.Google Scholar
Swihart, Robert K., Gehring, Thomas M., Kolozsvary, Mary Beth, and Nupp, Thomas E.. 2003. Responses of “Resistant” Vertebrates to Habitat Loss and Fragmentation: The Importance of Niche Breadth and Range Boundary. Diversity and Distributions 9:118.Google Scholar
Taber, Richard D., Raedeke, Kenneth, and McCaughran, Donald A.. 1982. Population Characteristics. In Elk of North America: Ecology and Management, edited by Thomas, Jack W. and Toweill, Dale E., pp. 279298. Stackpole Books, Harrisburg, PA.Google Scholar
Talma, A. S., and Vogel, John C.. 1992. Late Quaternary Paleotemperatures Derived from a Speleothem from Cango Caves, Cape Province, South Africa. Quaternary Research 37:203213Google Scholar
Taylor, Dwight W. 1965. The Study of Pleistocene Nonmarine Mollusks in North America. In The Quaternary of the United States, edited by Wright, H. E, Jr., and Frey, D. G, pp. 597611. Princeton University Press.Google Scholar
Taylor, Lucy A., Kaiser, Thomas M., Schwitzer, Christoph et al. 2013. Detecting Inter-Cusp and Inter-Tooth Wear Patterns in Rhinocerotids. PLoS ONE 8:e80921.Google Scholar
Tchernov, Eitan. 1968. Succession of Rodent Faunas during the Upper Pleistocene of Israel. Paul Parey, Hamburg.Google Scholar
Tchernov, Eitan. 1975. Rodent Faunas and Environmental Changes in the Pleistocene of Israel. In Rodents in Desert Environments, edited by Prakash, I. and Gosh, P. K., pp. 331362. Junk, The Hague.Google Scholar
Tchernov, Eitan. 1979. Polymorphism, Size Trends and Pleistocene Paleoclimatic Response of the Subgenus Sylvaemus (Mammalia: Rodentia) in Israel. Israel Journal of Zoology 28:131159.Google Scholar
Tchernov, Eitan. 1982. Faunal Responses to Environmental Changes in the Eastern Mediterranean during the Last 20,000 Years. In Palaeoclimates, Palaeoenvironments and Human Communities in the Eastern Mediterranean Region in Later Prehistory, edited by Bintliff, John L. and Van Zeist, Willem, pp. 105129. British Archaeological Reports, International Series 133. BAR, Oxford.Google Scholar
Tchernov, Eitan, and Horwitz, Liora Kolska. 1991. Body Size Diminution under Domestication: Unconscious Selection in Primeval Domesticates. Journal of Anthropological Archaeology 10:5475.Google Scholar
Teaford, Mark F. 1991. Dental Microwear: What Can It Tell Us about Diet and Dental Function? In Advances in Dental Anthropology, edited by Kelley, M. A and Larsen, C. S, pp. 341356. Wiley-Liss, New York.Google Scholar
Teaford, Mark F. 1994. Dental Microwear and Dental Function. Evolutionary Anthropology 3:1730.Google Scholar
Teaford, Mark F. 2006. What Do We Know and Not Know about Dental Microwear and Diet? In Evolution of the Human Diet: The Known, the Unknown, and the Unknowable, edited by Ungar, Peter, pp. 106132. Oxford University Press.Google Scholar
Teaford, Mark F., and Walker, Alan. 1984. Quantitative Differences in Dental Microwear between Primate Species with Different Diets and a Comment on the Presumed Diet of Sivapithecus. American Journal of Physical Anthropology 64:191200.Google Scholar
Teaford, Mark F., and Oyen, Ordean J.. 1989. In Vivo and In Vitro Turnover in Dental Microwear. American Journal of Physical Anthropology 80:447460.Google Scholar
Teaford, Mark F., Ungar, Peter, and Grine, Frederic E.. 2013. Dental Microwear and Paleoecology. In Early Hominin Paleoecology, edited by Sponheimer, Matt, Lee-Thorp, Julia A., Reed, Kaye E., and Ungar, Peter, pp. 251279. University Press of Colorado, Boulder.Google Scholar
Tedford, Richard H. 1970. Principles and Practices of Mammalian Geochronology in North America. In Proceedings of the North American Paleontological Convention, edited by Yochelson, Ellis L., pp. 666703. Allen Press, Lawrence, KS.Google Scholar
Telford, R. J., Andersson, C., Birks, H. J. B., and Juggins, S.. 2004. Biases in the Estimation of Transfer Function Prediction Errors. Paleoceanography 19:PA4014.Google Scholar
Telford, R. J., and Birks, H. J. B.. 2009. Evaluation of Transfer Functions in Spatially Structured Environments. Quaternary Science Reviews 28:13091316.Google Scholar
Teplitsky, Céline, Mills, James A., Alho, Jussi S., Yarall, John W., and Merilä, Juha. 2008. Bergmann’s Rule and Climate Change Revisited: Disentangling Environmental and Genetic Responses in a Wild Bird Population. Proceedings of the National Academy of Sciences USA 105:1349213496.Google Scholar
ter Braak, C. J. F. 1987. Ordination. In Data Analysis in Community and Landscape Ecology, edited by Jongman, R. H, ter Braak, C. J. F, and van Tongeren, O. F. R, pp. 91173. Cambridge University Press.Google Scholar
ter Braak, C. J. F., Juggins, S., Birks, H. J. B., and Van der Voet, H.. 1993. Weighted Averaging Partial Least Squares Regression (Wa-Pls): Definition and Comparison with Other Methods for Species–Environment Calibration. In Multivariate Environmental Statistics, edited by Patil, G. P and Rao, C. R, pp. 525560. Elsevier, Amsterdam.Google Scholar
Terry, Rebecca C. 2007. Inferring Predator Identity from Skeletal Damage of Small-Mammal Prey Remains. Evolutionary Ecology Research 9:199219.Google Scholar
Terry, Rebecca C. 2008. Modeling the Effects of Predation, Prey Cycling, and Time Averaging on Relative Abundance in Raptor-Generated Small Mammal Death Assemblages. Palaios 23:402410.Google Scholar
Terry, Rebecca C. 2009. Paleoecology: Methods. Encyclopedia of Life Sciences, a0003274. John Wiley and Sons, Chichester.Google Scholar
Terry, Rebecca C. 2010a. The Dead Do Not Lie: Using Skeletal Remains for Rapid Assessment of Historical Small-Mammal Community Baselines. Proceedings of the Royal Society B 277:11931201.Google Scholar
Terry, Rebecca C. 2010b. On Raptors and Rodents: Testing the Ecological Fidelity and Spatiotemporal Resolution of Cave Death Assemblages. Paleobiology 36:137160.Google Scholar
Terry, Rebecca C., Li, Cheng (Lily), and Hadly, Elizabeth A.. 2011. Predicting Small-Mammal Responses to Climatic Warming: Autecology, Geographic Range, and the Holocene Fossil Record. Global Change Biology 17:30193034.Google Scholar
Terry, Rebecca C., and Novak, Mark. 2015. Where Does the Time Go?: Mixing and the Depth-Dependent Distribution of Fossil Ages. Geology 43:487490.Google Scholar
Terry, Rebecca C., and Rowe, Rebecca J.. 2015. Energy Flow and Functional Compensation in Great Basin Small Mammals under Natural and Anthropogenic Environmental Change. Proceedings of the National Academy of Sciences USA 112:96569661.Google Scholar
Thackeray, J. Francis. 1987. Late Quaternary Environmental Changes Inferred from Small Mammalian Fauna, Southern Africa. Climatic Change 10:285305.Google Scholar
Thackeray, J. Francis. 1990. Temperature Indices from Late Quaternary Sequences in South Africa: Comparisons with the Vostok Core. South African Geographic Journal 72:4749.Google Scholar
Thackeray, J. Francis. 1992. Chronology of Late Pleistocene Deposits Associated with Homo sapiens at Klasies River Mouth, South Africa. Palaeoecology of Africa and the Surrounding Islands 23:177191.Google Scholar
Thackeray, J. Francis. 2002. Palaeoenvironmental Change and Re-assessment of the Age of Late Pleistocene Deposits at Die Kelders Cave, South Africa. Journal of Human Evolution 43:749753.Google Scholar
Thackeray, J. F[rancis], and Avery, D. M.. 1990. A Comparison between Temperature Indices for Late Pleistocene Sequences at Klasies River and Border Cave, South Africa. Palaeoecology of Africa 21:311316.Google Scholar
Thomas, Kenneth D., and Mannino, Marcello A.. 2017. Making Numbers Count: Beyond Minimum Numbers of Individuals (MNI) for the Quantification of Mollusc Assemblages from Shell Matrix Sites. Quaternary International 427(a):4758.Google Scholar
Thuiller, Wilfried, Lavorel, Sandra, and Araújo, Miguel B.. 2005. Niche Properties and Geographic Extent as Predictors of Species Sensitivity to Climate Change. Global Ecology and Biogeography 14:347357.Google Scholar
Tiffney, Bruce H. 2008. Phylogeography, Fossils, and Northern Hemisphere Biogeography: The Role of Physiological Uniformitarianism. Annals of the Missouri Botanical Garden 95:135143.Google Scholar
Tilman, David, and Pacala, Stephen. 1993. The Maintenance of Species Richness in Plant Communities. In Species Diversity in Ecological Communities, edited by Ricklefs, Robert E. and Schluter, Dolph, pp. 1325. University of Chicago Press.Google Scholar
Tipper, John C. 1979. Rarefaction and Rarefiction: The Use and Abuse of a Method in Paleoecology. Paleobiology 5:423434.Google Scholar
Todd, Lawrence C. 1986. Determination of Sex of Bison Upper Forelimb Bones: The Humerus and Radius. Wyoming Archaeologist 29(1–3):109124.Google Scholar
Todd, Lawrence C. 1987. Taphonomy of the Horner II Bone Bed. In The Horner Site: The Type Site of the Cody Cultural Complex, edited by Frison, George C. and Todd, Lawrence C., pp. 107198. Academic Press, Orlando, FL.Google Scholar
Travouillon, K[enny] J., and Legendre, S.. 2009. Using Cenograms to Investigate Gaps in Mammalian Body Mass Distributions in Australian Mammals. Palaeogeography, Palaeoclimatology, Palaeoecology 272:6984.Google Scholar
Travouillon, Kenny J., Archer, Michael, Legendre, Serge, and Hand, Suzanne J.. 2007. Finding the Minimum Sample Richness (MSR) for Multivariate Analyses: Implications for Palaeoecology. Historical Biology 19:315320.Google Scholar
Travouillon, Kenny J., Escarguel, Gilles, Legendre, Serge, Archer, Michael, and Hand, Suzanne J.. 2011. The Use of MSR (Minimum Sample Richness) for Sample Assemblage Comparisons. Paleobiology 37:696709.Google Scholar
Trinkaus, Erik. 1981. Neanderthal Limb Proportions and Cold Adaptation. In Aspects of Human Evolution, edited by Stringer, C. B, pp. 187224. Taylor and Francis, London.Google Scholar
Tryon, Christian A., and Tyler Faith, J.. 2016. A Demographic Perspective on the Middle to Later Stone Age Transition from Nasera Rockshelter, Tanzania. Philosophical Transactions of the Royal Society B 371:20150238.Google Scholar
Tryon, Christian A., Faith, J. Tyler, Peppe, Daniel J. et al. 2010. The Pleistocene Archaeology and Environments of the Wasiriya Beds, Rusinga Island, Kenya. Journal of Human Evolution 59:657671.Google Scholar
Tsubamoto, Takehisa, Egi, Naoko, Takai, Masanaru, Sein, Chit, and Maung, Maung. 2005. Middle Eocene Ungulate Mammals from Myanmar: A Review with Description of New Specimens. Acta Palaeontologica Polonica 50:117138.Google Scholar
Turvey, Samuel T., and Blackburn, Tim M.. 2011. Determinants of Species Abundance in the Quaternary Vertebrate Fossil Record. Paleobiology 37:537546.Google Scholar
Turvey, Samuel T., and Cooper, Joanne H.. 2009. The Past is Another Country: Is Evidence for Prehistoric, Historical, and Present-Day Extinction Really Comparable? In Holocene Extinctions, edited by Turvey, Samuel T., pp. 193212. Oxford University Press.Google Scholar
Ulbricht, Arlett, Maul, Lutz C., and Schulz, Ellen. 2015. Can Mesowear Analysis Be Applied to Small Mammals? A Pilot-Study on Leporines and Murines. Mammalian Biology 80:1420.Google Scholar
Ungar, Peter S. 2015. Mammalian Dental Function and Wear: A Review. Biosurface and Biotribology 1:2541.Google Scholar
Ungar, Peter S., Brown, Christopher A., Bergstrom, Torbjorn S., and Walker, Alan. 2003. Quantification of Dental Microwear by Tandem Scanning Confocal Microscopy and Scale-Sensitive Fractal Analyses. Scanning 25:183193.Google Scholar
Ungar, Peter S., Merceron, Gildas, and Scott, Robert S.. 2007. Dental Microwear Texture Analysis of Varswater Bovids and Early Pliocene Paleoenvironments of Langebaanweg, Western Cape Province, South Africa. Journal of Mammalian Evolution 14:163181.Google Scholar
Ungar, Peter S., Scott, Robert S., Scott, Jessica R., and Teaford, Mark. 2008. Dental Microwear Analysis: Historical Perspectives and New Approaches. In Technique and Application in Dental Anthropology, edited by Irish, Joel D. and Nelson, Greg C., pp. 389425. Cambridge University Press.Google Scholar
Utescher, T., Bruch, A. A., Erdei, B. et al. 2014. The Coexistence Approach: Theoretical Background and Practical Considerations of Using Plant Fossils for Climate Quantification. Palaeogeography, Palaeoclimatology, Palaeoecology 410:5873.Google Scholar
Valverde, J. A. 1964. Remarques sur la Structure et l’Évolution des Communautés de Vertébrés Terrestres. I. Structure d’une Commumauté. II. Rapports entre Prédateurs et Proies. Revue d’Écologie: La Terre et la Vie 111:121154.Google Scholar
Van Couvering, Judith A. 1980. Community Evolution in East Africa. In Fossils in the Making: Vertebrate Taphonomy and Paleoecology, edited by Behrensmeyer, Anna K. and Hill, Andrew P., pp. 272298. University of Chicago Press.Google Scholar
van der Klaauw, Cornelius J. 1948. Ecological Studies and Reviews: IV. Ecological Morphology. Bibliotheca Biotheoretica 4:27111.Google Scholar
van der Meulen, Albert J., and Daams, Remmert. 1992. Evolution of Early-Middle Miocene Rodent Faunas in Relation to Long-Term Palaeoenvironmental Changes. Palaeogeography, Palaeoclimatology, Palaeoecology 93:227253.Google Scholar
VanPool, Todd L., and Leonard, Robert D.. 2009. Quantitative Archaeology. Blackwell, Cambridge, MA.Google Scholar
Van Riper, A. Bowdoin. 1993. Men among the Mammoths: Victorian Science and the Discovery of Human Prehistory. University of Chicago Press.Google Scholar
Van Valen, L. 1973. Pattern and the Balance of Nature. Evolutionary Theory 1:3149.Google Scholar
Van Valkenburgh, Blaire. 1987. Skeletal Indicators of Locomotor Behavior in Living and Extinct Carnivores. Journal of Vertebrate Paleontology 7:162182.Google Scholar
Van Valkenburgh, Blaire. 1988. Trophic Diversity in Past and Present Guilds of Large Predatory Mammals. Paleobiology 14:155173.Google Scholar
Van Valkenburgh, Blaire, Teaford, Mark F., and Walker, Alan. 1990. Molar Microwear and Diet in Large Carnivores: Inferences Concerning Diet in the Sabretooth Cat, Smilodon fatalis. Journal of Zoology (London) 222:319340.Google Scholar
Varela, Sara, Lobo, Jorge M., and Hortal, Joaquín. 2011. Using Species Distribution Models in Paleobiogeography: A Matter of Data, Predictors and Concepts. Palaeogeography, Palaeoclimatology, Palaeoecology 310:451463.Google Scholar
Vartanyan, S. L., Garutt, V. E., and Sher, A. V.. 1993. Holocene Dwarf Mammoths from Wrangel Island in the Siberian Arctic. Nature 362:337–340.Google Scholar
Verts, B. J., and Carraway, Leslie N.. 1998. Land Mammals of Oregon. University of California Press, Berkeley.Google Scholar
Visher, S. S. 1954. Climatic Atlas of the United States. Harvard University Press, Cambridge, MA.Google Scholar
Vlok, Jan, and Schutte-Vlok, A. L.. 2010. Plants of the Klein Karoo. Umdaus Press, Hatfield, South Africa.Google Scholar
Vogel, John C. 2001. Radiometric Dates for the Middle Stone Age in South Africa. In Humanity from African Naissance to Coming Millennia: Colloquia in Human Biology and Paleoanthropology, edited by Tobias, Phillip V., Raath, Michael A., Maggi-Cecchi, Jacopo, and Doyle, Gerald A., pp. 261268. Florence University Press.Google Scholar
von Humboldt, Alexander. 1850. Views of Nature: Or Contemplations on the Sublime Phenomena of Creation; with Scientific Illustrations. Translated by E. C. Otté and H. G. Bohn. H. G. Bohn, London.Google Scholar
Voorhies, Michael R. 1970. Sampling Difficulties in Reconstructing Late Tertiary Mammalian Communities. In Proceedings of the North American Paleontological Convention, edited by Yochelson, Ellis L., pp. 454468. Allen Press, Lawrence, KS.Google Scholar
Vrba, E[lisabeth] S. 1974. Chronological and Ecological Implications of the Fossil Bovidae at the Sterkfontein Australopithecine Site. Nature 250:1923.Google Scholar
Vrba, E[lisabeth] S. 1975. Some Evidence of Chronology and Palaeoecology of Sterkfontein, Swartkrans and Kromdraai from the Fossil Bovidae. Nature 254:301304.Google Scholar
Vrba, E[lisabeth] S. 1980. The Significance of Bovid Remains as Indicators of Environment and Predation Patterns. In Fossils in the Making: Vertebrate Taphonomy and Paleoecology, edited by Behrensmeyer, Anna K. and Hill, Andrew P., pp. 247271. University of Chicago Press.Google Scholar
Vrba, E[lisabeth] S. 1985. Environment and Evolution: Alternative Causes of the Temporal Distribution of Evolutionary Events. South African Journal of Science 81:229236.Google Scholar
Vrba, E[lisabeth] S. 1992. Mammals as a Key to Evolutionary Theory. Journal of Mammalogy 73:128.Google Scholar
Vrba, E[lisabeth] S. 1995. The Fossil Record of African Antelopes (Mammalia, Bovidae) in Relation to Human Evolution and Paleoclimate. In Paleoclimate and Evolution with Emphasis on Human Origins, edited by Vrba, Elisabeth S., Denton, George H., Partridge, Timothy C., and Burckle, Lloyd H., pp. 385424. Yale University Press, New Haven, CT.Google Scholar
Waide, R. B., Willig, M. R., Steiner, C. F. et al. 1999. The Relationship between Productivity and Species Richness. Annual Review of Ecology and Systematics 30:257300.Google Scholar
Wainwright, Peter C., and Reilly, Stephen M. (editors). 1994. Ecological Morphology: Integrative Organismal Biology. University of Chicago.Google Scholar
Wake, David B., Hadly, Elizabeth A., and Ackerly, David D.. 2009. Biogeography, Changing Climates, and Niche Evolution. Proceedings of the National Academy of Sciences USA 106:1963119636.Google Scholar
Walde, Dale. 2004. Distinguishing Sex of Bison bison bison Using Discriminant Function Analysis. Canadian Journal of Archaeology 28:100116.Google Scholar
Walker, Alan, Hoeck, H. N., and Perez, L.. 1978. Microwear of Mammalian Teeth as an Indicator of Diet. Science 201:908910.Google Scholar
Walker, D. 1978. Envoi. In Biology and Quaternary Environments, edited by Walker, D. and Guppy, J. C, pp. 259264. Australian Academy of Science, Canberra.Google Scholar
Walker, D., and Guppy, J. C (editors). 1978. Biology and Quaternary Environments. Australian Academy of Science, Canberra.Google Scholar
Walker, Danny N. 1982. Early Holocene Vertebrate Fauna. In The Agate Basin Site: A Record of the Paleoindian Occupation of the Northwestern High Plains, edited by Frison, George C. and Stanford, Dennis J., pp. 274308. Academic Press, New York.Google Scholar
Walker, Danny N. 2007. Vertebrate Fauna. In Medicine Lodge Creek: Holocene Archaeology of the Eastern Big Horn Basin, Wyoming, vol. i, edited by Frison, George C. and Walker, Danny N., pp. 177208. Clovis Press, Avondale, CO.Google Scholar
Walker, Danny N., and Frison, George C.. 1982. Studies on Amerindian Dogs, 3: Prehistoric Wolf/Dog Hybrids from the Northwestern Plains. Journal of Archaeological Science 9:125172.Google Scholar
Walter, H. 1970. Vegetationszonen und Klima. Eugen Ulmer, Stuttgart.Google Scholar
Wang, Yang, and Cerling, Thure E.. 1994. A Model of Fossil Tooth and Bone Diagenesis: Implications for Paleodiet Reconstructions from Stable Isotopes. Palaeogeography, Palaeoclimatology, Palaeoecology 107:281289.Google Scholar
Wartenburg, Daniel, Ferson, Scott, and Rohlf, F. James. 1987. Putting Things in Order: A Critique of Detrended Correspondence Analysis. American Naturalist 129:434448.Google Scholar
Wasserman, David, and Nash, Donald J.. 1979. Variation in Body Size, Hair Length, and Hair Density in the Deer Mouse Peromyscus maniculatus along an Altitudinal Gradient. Holarctic Ecology 2:115118.Google Scholar
Watt, Cortney, Mitchell, Sean, and Salewski, Volker. 2010. Bergmann’s Rule: A Concept Cluster? Oikos 119:89100.Google Scholar
Weaver, Timothy D., Steele, Teresa E., and Klein, Richard G.. 2011. The Abundance of Eland, Buffalo, and Wild Pigs in Middle and Later Stone Age Sites. Journal of Human Evolution 60:309314.Google Scholar
Webb, Thompson, III, and Bryson, Reid A.. 1972. Late- and Postglacial Climatic Change in the Northern Midwest, USA: Quantitative Estimates Derived from Fossil Pollen Spectra by Multivariate Statistical Analysis. Quaternary Research 2:70115.Google Scholar
Webb, Thompson, III, and Clark, D. R.. 1977. Calibrating Micropaleontological Data in Climatic Terms: A Critical Review. Annals of the New York Academy of Sciences 288:93118.Google Scholar
Weinstock, Jaco. 1997. The Relationship between Body Size and Environment: The Case of Late Pleistocene Reindeer (Rangifer tarandus). Archaeofauna 6:123135.Google Scholar
Weinstock, Jaco. 2002. Reindeer Hunting in the Upper Paleolithic: Sex Ratios as a Reflection of Different Procurement Strategies. Journal of Archaeological Science 29:365377.Google Scholar
Weisler, M. I. 1993. The Importance of Fish Otoliths in Pacific Island Archaeofaunal Analysis. New Zealand Journal of Archaeology 15:131159.Google Scholar
Weissbrod, Lior, Malkinson, Dan, Cucchi, Thomas et al. 2014. Ancient Urban Ecology Reconstructed from Archaeozoological Remains of Small Mammals in the Near East. Plos One 9: e91795.Google Scholar
Wells, R. T. 1978. Fossil Mammals in the Reconstruction of Quaternary Environments with Examples from the Australian Fauna. In Biology and Quaternary Environments, edited by Walker, D. and Guppy, J. C, pp. 103124. Australian Academy of Science, Canberra.Google Scholar
West, Geoffrey B., Brown, James H., and Enquis, Brian J.. 1997. A General Model for the Origin of Allometric Scaling Laws in Biology. Science 276:122126.Google Scholar
Western, David, and Behrensmeyer, Anna K.. 2009. Bone Assemblages Track Animal Community Structure over 40 Years in an African Savanna Ecosystem. Science 324:10611064.Google Scholar
White, Ethan P., Morgan Ernest, S. K., Kerkhoff, Andrew J., and Enquist, Brian J.. 2007. Relationships between Body Size and Abundance in Ecology. Trends in Ecology and Evolution 22:323330.Google Scholar
White, T. C. R. 2008. The Role of Food, Weather and Climate in Limiting the Abundance of Animals. Biological Reviews 83:227248.Google Scholar
White, Theodore E. 1953. Studying Osteological Material. Plains Archaeological Conference Newsletter 6(1):5867.Google Scholar
White, Theodore E. 1954. Preliminary Analysis of the Fossil Vertebrates of the Canyon Ferry Reservoir Area. Proceedings of the United States National Museum 103:395438.Google Scholar
White, Theodore E. 1956. The Study of Osteological Material from the Plains. American Antiquity 21:401404.Google Scholar
Whittaker, Robert J. 1975. Communities and Ecosystems, second edition. Macmillan, New York.Google Scholar
Whittaker, Robert J., Willis, Katherine J., and Field, Richard. 2001. Scale and Species Richness: Towards a General Hierarchical Theory of Species Diversity. Journal of Biogeography 28:453570.Google Scholar
Whittaker, Robert J., Riddle, Brett R., Hawkins, Bradford A., and Ladle, Richard J.. 2013. The Geographical Distribution of Life and the Problem of Regionalization: 100 Years after Alfred Russell Wallace. Journal of Biogeography 40:22092214.Google Scholar
Whittington, H. B. 1964. Taxonomic Basis of Paleoecology. In Approaches to Paleoecology, edited by Imbrie, John and Newell, Norman D., pp. 1927. John Wiley and Sons, New York.Google Scholar
Widga, Chris. 2013. Evolution of the High Plains Paleoindian Landscape: The Paleoecology of Great Plains Faunal Assemblages. In Paleoindian Lifeways of the Cody Complex, edited by Knell, Edward J. and Muñiz, Mark P., pp. 6992. University of Utah Press, Salt Lake City.Google Scholar
Wiens, John A. 1989. Spatial Scaling in Ecology. Functional Ecology 3:385397.Google Scholar
Wiens, John J., and Donoghue, Michael J.. 2004. Historical Biogeography, Ecology and Species Richness. Trends in Ecology and Evolution 19:639644.Google Scholar
Wiens, John J., Ackerly, David D., Allen, Andrew P. et al. 2010. Niche Conservatism as an Emerging Principle in Ecology and Conservation Biology. Ecology Letters 13:13101324.Google Scholar
Wiens, John A., Hayward, Gregory D., Safford, Hugh D., and Giffen, Catherine M. (editors). 2012a. Historical Environmental Variation in Conservation and Natural Resource Management. Wiley-Blackwell, Chichester.Google Scholar
Wiens, John A., Safford, Hugh D., McGarigal, Kevin, Romme, William H., and Manning, Mary. 2012b. What is the Scope of “History” in Historical Ecology? Issues of Scale in Management and Conservation. In Historical Environmental Variation in Conservation and Natural Resource Management, edited by Wiens, John A., Hayward, Greogry D., Safford, Hugh D., and Giffen, Catherine M., pp. 6375. Wiley-Blackwell, Chichester.Google Scholar
Wiggington, John D., and Dobson, F. Stephen. 1999. Environmental Influences on Geographic Variation in Body Size of Western Bobcats. Canadian Journal of Zoology 77:802813.Google Scholar
Williams, John W., and Jackson, Stephen T.. 2007. Novel Climates, No-Analog Communities, and Ecological Surprises. Frontiers in Ecology and the Environment 5:475482.Google Scholar
Williams, John W., Shuman, Bryan N., and Webb, Thompson III. 2001. Dissimilarity Analyses of Late-Quaternary Vegetation and Climate in Eastern North America. Ecology 82:33463362.Google Scholar
Williams, John W., Shuman, Bryan, Bartlein, Patrick J., Diffenbaugh, Noah S., and Webb, Thompson III. 2010. Rapid, Time-Transgressive, and Variable Responses to Early Holocene Midcontinental Drying in North America. Geology 38:135138.Google Scholar
Williams, John W., Blois, Jessica L., Gill, Jacquelyn L. et al. 2013. Model Systems for a No-Analog Future: Species Associations and Climates during the Last Deglaciation. Annals of the New York Academy of Sciences 1297:2943.Google Scholar
Williams, Susan H., and Kay, Richard F.. 2001. A Comparative Test of Adaptive Explanations for Hypsodonty in Ungulates and Rodents. Journal of Mammalian Evolution 8:207229.Google Scholar
Willig, M. R., Kaufman, D. M., and Stevens, R. D.. 2003. Latitudinal Gradients of Biodiversity: Pattern, Process, Scale, and Synthesis. Annual Review of Ecology, Evolution, and Systematics 34:273309.Google Scholar
Wilson, Don E., and Reeder, DeeAnn M. (editors) 2005. Mammal Species of the World: A Taxonomic and Geographic Reference, third edition. Johns Hopkins University Press, Baltimore.Google Scholar
Wilson, Michael [C]. 1973. The Early Historic Fauna of Southern Alberta: Some Steps to Interpretation. In Historical Archaeology in Northwestern North America, edited by Getty, Ronald M. and Fladmark, Knut R., pp. 213248. University of Calgary Archaeological Association, Calgary, Alberta.Google Scholar
Wilson, Michael [C]. 1978. Archaeological Kill Site Populations and the Holocene Evolution of the Genus Bison. In Bison Procurement and Utilization: A Symposium, edited by Davis, Leslie B. and Wilson, Michael, pp. 922. Plains Anthropologist 23(82).Google Scholar
Wilson, M. V. H. 2001. Fossils as Environmental Indicators: Taphonomic Evidence. In Paleobiology II, edited by Briggs, Derek E. G and Crowther, Peter R., pp. 467470. Blackwell Science, Oxford.Google Scholar
Wilson, R. L. 1968. Systematics and Faunal Analysis of a Lower Pliocene Vertebrate Assemblage from Trego County, Kansas. Contributions from the Museum of Paleontology 22:75126. University of Michigan, Ann Arbor.Google Scholar
Wing, Scott L., Sues, Hans-Dieter, Potts, Richard, DiMichele, William A., and Behrensmeyer, Anna K.. 1992. Evolutionary Paleoecology. In Terrestrial Ecosystems through Time: Evolutionary Paleoecology of Terrestrial Plants and Animals, edited by Behrensmeyer, Anna K., Damuth, John D., DiMichele, William A., Potts, Richard, Sues, Hans-Dieter, and Wing, Scott L., pp. 113. University of Chicago Press.Google Scholar
Winkler, Alisa J., and Gose, Wulf. 2003. Mammalian Fauna and Paleomagnetics of the Middle Irvingtonian (Early Pleistocene) Fyllan Cave and Kitchen Door Localities, Travis County, Texas. In Ice Age Cave Faunas of North America, edited by Schubert, Blaine W., Mead, Jim I., and Graham, Russell W., pp. 215261. Indiana University Press, Bloomington.Google Scholar
Wisz, M. S., Pottier, J., Kissling, W. D. et al. 2013. The Role of Biotic Interactions in Shaping Distributions and Realised Assemblages of Species: Implications for Species Distribution Modeling. Biological Reviews 88:1530.Google Scholar
Withnell, Charles B., and Ungar, Peter S.. 2014. A Preliminary Analysis of Dental Microwear as a Proxy for Diet and Habitat in Shrews. Mammalia 78:409415.Google Scholar
Wolff, Ronald G. 1973. Hydrodynamic Sorting and Ecology of a Pleistocene Mammalian Assemblage from California (USA). Palaeogeography, Palaeoclimatology, Palaeoecology 13:91101.Google Scholar
Wolff, Ronald G. 1975. Sampling and Sample Size in Ecological Analyses of Fossil Mammals. Paleobiology 1:195204.Google Scholar
Wolverton, Steve. 2005. The Effects of the Hypsithermal on Prehistoric Foraging Efficiency in Missouri. American Antiquity 70:91106.Google Scholar
Wolverton, Steve. 2008. Harvest Pressure and Environmental Carrying Capacity: An Ordinal-Scale Model of Effects on Unulate Prey. American Antiquity 73:179199.Google Scholar
Wolverton, Steve. 2013. Data Quality in Zooarchaeological Faunal Identification. Journal of Archaeological Method and Theory 20:381396.Google Scholar
Wolverton, Steve, and Lyman, R. Lee (editors). 2012. Conservation Biology and Applied Zooarchaeology. University of Arizona Press, Tucson.Google Scholar
Wolverton, Steve, Kennedy, James H., and Cornelius, John D.. 2007. A Paleozoological Perspective on White-Tailed Deer (Odocoileus virginianus texana) Population Density and Body Size in Central Texas. Environmental Management 39:545552.Google Scholar
Wolverton, Steve, Huston, Michael A., Kennedy, James H., Cagle, Kevin, and Cornelius, John D.. 2009. Conformation to Bergmann’s Rule in White-Tailed Deer Can Be Explained by Food Availability. American Midland Naturalist 162:403417.Google Scholar
Wolverton, Steve, Nagaoka, Lisa, and Rick, Torben C.. 2016. Applied Zooarchaeology: Five Case Studies. Eliot Werner, Clinton Corners, NY.Google Scholar
Wood, Bernard A. 1979. An Analysis of Tooth and Body Size Relationships in Five Primate Taxa. Folia Primatologica 31:187211.Google Scholar
Wood, David L., and Barnosky, Anthony D.. 1994. Middle Pleistocene Climate Change in the Colorado Rocky Mountains Indicated by Fossil Mammals from Porcupine Cave. Quaternary Research 41:366375.Google Scholar
Wood, David M., and del Moral, Roger. 1987. Mechanisms of Early Primary Succession in Subalpine Habitats on Mount St. Helens. Ecology 68:780790Google Scholar
Woodcock, D. W. 1992. Climate Reconstruction Based on Biological Indicators. Quarterly Review of Biology 67:457477.Google Scholar
Woodring, W. P. 1951. Basic Assumption Underlying Paleoecology. Science 113:482483.Google Scholar
Woodward, C. A., and Shulmeister, J.. 2006. New Zealand Chironomids as Proxies for Human-Induced and Natural Environmental Change: Transfer Functions for Temperature and Lake Production (Chlorophyll a). Journal of Paleolimnology 36:407429.Google Scholar
Wu, Jianguo. 2007. Scale and Scaling: A Cross-Disciplinary Perspective. In Key Topics in Landscape Ecology, edited by Wu, Jianguo and Hobbs, Richard, pp. 115142. Cambridge University Press.Google Scholar
Yackulic, Charles B., Sanderson, Eric W., and Uriarte, María. 2011. Anthropogenic and Environmental Drivers of Modern Range Loss in Large Mammals. Proceedings of the National Academy of Sciences USA 108:40244029.Google Scholar
Yalden, D. W. 2001. Mammals as Climatic Indicators. In Handbook of Archaeological Sciences, edited by Brothwell, D. R and Pollard, A. M, pp. 147154. John Wiley and Sons, Chichester.Google Scholar
Yamada, Eisuke, Hasumi, Erl, Miyazato, Nao, Akahoshi, Megumi, Watabe, Mahito, and Nakaya, Hideo. 2016. Mesowear Analyses of Sympatric Ungulates from the Late Miocene Maragheh, Iran. Palaeobiodiversity and Palaeoenvironments 96:445452.Google Scholar
Yann, Lindsey T., DeSantis, Larisa R. G., Haupt, Ryan J. et al. 2013. The Application of an Oxygen Isotope Aridity Index to Terrestrial Paleoenvironmental Reconstructions in Pleistocene North America. Paleobiology 39:576590.Google Scholar
Yom-Tov, Yoram, and Geffen, Eli. 2006. Geographic Variation in Body Size: The Effects of Ambient Temperature and Precipitation. Oecologia 148:213218.Google Scholar
Yom-Tov, Yoram, and Nix, Henry. 1986. Climatological Correlates for Body Size of Five Species of Australian Mammals. Biological Journal of the Linnean Society 29:245262.Google Scholar
Yom-Tov, Yoram, and Yom-Tov, Jonathan. 2005. Global Warming, Bergmann’s Rule and Body Size in the Masked Shrew Sorex cinereus Kerr in Alaska. Journal of Animal Ecology 74:803808.Google Scholar
Yom-Tov, Yoram, Yom-Tov, Shlomith, Wright, Jonathan, Thorne, Chris J. R., and Feu, Richard Du. 2006. Recent Changes in Body Weight and Wing Length among Some British Passerine Birds. Oikos 112:91101.Google Scholar
Yom-Tov, Yoram, Yom-Tov, Shlomith, and Jarrell, Gordon. 2008. Recent Increase in Body Size of the American Marten Martes americana in Alaska. Biological Journal of the Linnean Society 93:701707.Google Scholar
Yom-Tov, Yoram, Leader, Noam, Yom-Tov, Shlomith, and Baagøe, Hans J.. 2010. Temperature Trends and Recent Decline in Body Size of the Stone Marten Martes foina in Denmark. Mammalian Biology 75:146150.Google Scholar
Young, Kenneth R. 2014. Biogeography of the Anthropocene: Novel Species Assemblages. Progress in Physical Geography 38:664673.Google Scholar
Zeder, Melinda A. 2001. A Metrical Analysis of Modern Goats (Capra hircus aegagrus and C. h. hircus) from Iran and Iraq: Implications for the Study of Caprine Domestication. Journal of Archaeological Science 28:6179.Google Scholar
Zeder, Melinda A., and Hesse, Brian. 2000. The Initial Domestication of Goats (Capra hircus) in the Zagros Mountains 10,000 Years Ago. Science 287:22542257.Google Scholar
Zeuner, F. E. 1936. Climatic Research Based on the Association of Species in Fossil Faunas and Floras. In Problems in Paleontology, vol. i, edited by Hartmann-Weinberg, A., pp. 200216. Publications of the Laboratory of Paleontology, Moscow University.Google Scholar
Zeuner, F. E. 1961. Faunal Evidence for Pleistocene Climates. Annals of the New York Academy of Science 95(1):502507.Google Scholar
Zhang, Hanwen, Wang, Yuan, Janis, Christine M., Goodall, Robert H., and Purnell, Mark A.. 2017. An Examination of Feeding Ecology in Pleistocene Proboscideans from Southern China (Sinomastodon, Stegodon, Elephas), by Means of Dental Microwear Texture Analysis. Quaternary International 445:6070.Google Scholar
Zhang, Yi Ge, Pagani, Mark, Liu, Zhongui, Bohaty, Steven M., and DeConto, Robert. 2013. A 40-Million-Year History of Atmospheric CO2. Philosophical Transactions of the Royal Society A 371:20130096.Google Scholar
Ziegler, Alan C. 1973. Inference from Prehistoric Faunal Remains. Addison-Wesley Module in Anthropology 43. Addison-Wesley, Reading, MA.Google Scholar
Zohar, Irit, and Belmaker, Miriam. 2005. Size Does Matter: Methodological Comments on Sieve Size and Species Richness in Fishbone Assemblages. Journal of Archaeological Science 32:635641.Google Scholar

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.

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.

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.

Available formats
×