Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-wzw2p Total loading time: 0 Render date: 2024-04-30T12:59:19.597Z Has data issue: false hasContentIssue false

17 - Tropical and temperate seasonal influences on human evolution

Published online by Cambridge University Press:  10 August 2009

Kaye E. Reed
Affiliation:
Department of Anthropology/Institute of Human Origins, Arizona State University, Tempe AZ 85287 USA
Jennifer L. Fish
Affiliation:
Max Planck Institute of Molecular Cell Biology and Genetics, Pfotenhauerstrasse 108 01307 Dresden Germany
Diane K. Brockman
Affiliation:
University of North Carolina, Charlotte
Carel P. van Schaik
Affiliation:
Universität Zürich
Get access

Summary

Introduction

Climatic and subsequent habitat change has often been invoked as a driving force of evolutionary change in hominins, mammals, and other taxa (Vrba 1988a, 1988b, 1992, 1995; Bromage & Schrenk 1999; Potts 1998; Bobe & Eck 2001; Janis et al. 2002). Global climatic change in the mid Pliocene Epoch has been suggested as a cause for hominin speciation events (Vrba 1995) and is correlated with changes in dentition and jaw morphology of two hominin lineages (Teaford & Ungar 2000). Climatic change also influences seasonality, such that drying trends, for example, likely instigate short wet seasons, while the reverse is also true. Although Foley (1987) indicated that seasonal differences were likely important in determining hominin foraging effort, and Blumenschine (1987) posited a dry-season scavenging niche for Pleistocene hominids, little attention has been given to how seasonal changes over time might contribute to differences among hominin behavioral adaptations. Seasonal changes refer to changes in the lengths of regular four-season patterns in temperate climates or wet and dry seasonal differences in the tropics over geological time.

Evolutionary changes in fossil hominins are detected through changes in morphology that represent different behavioral adaptations. Fossil hominin diets are inferred from comparisons with extant primates in features such as tooth size (Hylander 1975; Kay 1984; Ungar & Grine 1991), molar shearing crests (Kay 1984), dental microwear (Grine 1981; Teaford 1988; Ungar 1998), biomechanics (Hylander 1988; Daegling & Grine 1991), and isotopic signatures (Sponheimer & Lee-Thorp 1999; van der Merwe et al. 2003).

Type
Chapter
Information
Seasonality in Primates
Studies of Living and Extinct Human and Non-Human Primates
, pp. 489 - 518
Publisher: Cambridge University Press
Print publication year: 2005

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

Aiello, L. & Wheeler, P. (1995). The expensive-tissue hypothesis. Current Anthropology, 36, 199–221.CrossRefGoogle Scholar
Andrews, P. (1989). Palaeoecology of Laetoli. Journal of Human Evolution, 18, 173–81.CrossRefGoogle Scholar
Andrews, P. & Humphrey, L. (1999). African Miocene environments and the transition to early hominins. In Paleoclimate and Evolution with Emphasis on Human Origins, ed. Vrba, E. S., Denton, G. H., Partridge, T. C., & Burckle, L. C.. New Haven, CT: Yale University Press, pp. 282–300.Google Scholar
Archibold, O. W. (1995). Ecology of Wild Vegetation. London: Chapman & Hall.CrossRefGoogle Scholar
Bar-Yosef, O. & Meignen, L. (1992). Insights into Middle Paleolithic cultural variability. In The Middle Paleolithic: Adaptation, Behavior, and Variability, ed. Dibble, H. & Mellars, P.. Philadelphia: The University Museum, University of Pennsylvania, pp. 163–82.Google Scholar
Behrensmeyer, A. K., Todd, N., Potts, R., & McBrinn, G. E. (1997). Late Pliocene faunal turnover in the Turkana Basin, Kenya and Ethiopia. Science, 278, 1589–94.CrossRefGoogle ScholarPubMed
Blumenschine, R. J. (1987). Characteristics of an early hominid scavenging niche. Current Anthropology, 28, 383–407.CrossRefGoogle Scholar
Bobe, R. & Eck, G. G. (2001). Responses of African bovids to Pliocene climatic change. Paleobiology, 27 (Suppl), 1–47.2.0.CO;2>CrossRefGoogle Scholar
Bocherens, H., Billiou, D., Mariotti, A., et al. (1999). Palae-environmental and Palaeodietary implications of isotopic biogeochemistry of last interglacial Neanderthal and mammal bones in Scladina Cave (Belgium). Journal of Archaeological Science, 26, 599–607.CrossRefGoogle Scholar
Bonnefille, R. (1995). A reassessment of the Plio-Pleistocene pollen record of East Africa. In Paleoclimate and Evolution with Emphasis on Human Origins, ed. Vrba, E. S., Denton, G. H., Partridge, T. C., & Burckle, L. C.. New Haven, CT: Yale University Press, pp. 299–310.Google Scholar
Brain, C. K. & Shipman, P. (1993). The Swartkrans bone tools. In Swartkrans: A Cave's Chronicle of Early Man, ed. Brain, C. K.. Pretoria, South Africa: Transvaal Museum Monograph, pp. 195–215.Google Scholar
Bromage, T. G. & Schrenk, F. (1999). African Biogeography, Climate Change & Human Evolution. New York: Oxford University Press.Google Scholar
Cachel, S. & Harris, J. W. K. (1998). The lifeways of Homo erectus inferred from archaeology and evolutionary ecology: a perspective from east Africa. In Early Human Behaviour in Global Context: The Rise and Diversity of the Lower Palaeolithic Record, ed. Petraglia, M. D. & Korisettar, R.. London: Routledge, pp. 280–303.Google Scholar
Cerling, T. (1992). Development of grasslands and savannas in East Africa during the Neogene. Palaeogeography, Palaeoclimatology, Palaeoecology, 97, 241–7.CrossRefGoogle Scholar
Cole, M. M. (1986). The Savannas: Biogeography and Geobotany. London: Academic Press.Google Scholar
Connell, J. H. & Orias, E. (1964). The ecological regulation of species diversity. American Naturalist, 98, 399–414.CrossRefGoogle Scholar
Crowley, & North, (1991) Paleoclimatology. Oxford: Oxford University Press.Google Scholar
Daegling, D. & Grine, F. E. (1991). Compact-bone distribution and biomechanics of early hominid mandibles. American Journal of Physical Anthropology, 86, 321–39.CrossRefGoogle ScholarPubMed
DeMenocal, P. B., & Bloemendal, J. (1995). Plio-Pleistocene subtropical African climate variability and the paleoenvironment of hominid evolution: a combined data-model approach, In Paleoclimate and Evolution with Emphasis on Human Origins, ed. Vrba, E. S., Denton, G. H., Partridge, T. C., & Burckle, L. C.. New Haven, CT: Yale University Press, pp. 262–88.Google Scholar
D'Errico, F., Backwell, L. R., & Berger, L. R. (2001). Bone tool use in termite foraging by early hominids and its impact on our understanding of early hominid behaviour. South African Journal of Science, 3–4: 71–5.Google Scholar
Fish, J. L. & Lockwood, C. A. (2003). Dietary constraints on encephalization in Primates. American Journal of Physical Anthropology, 120, 171–81.CrossRefGoogle ScholarPubMed
Fizet, M., Mariotti, A., & Bocherens, H. (1995) Effect of diet, physiology, and climate on carbon and nitrogen stable isotopes of collagen in a Late Pleistocene anthropic palaeoecosystem: Marillac, Charente, France. Journal of Archaeological Science, 22, 67–79.CrossRefGoogle Scholar
Fleagle, J. G. (1999). Primate Adaptation and Evolution. New York: Academic Press.Google Scholar
Foley, R. (1987). Another Unique Species. Harlow, UK: Longman Scientific & Technical.Google Scholar
Gabunia, L., Vekua, A., Lordkipanidze, D., et al. (2000a). Earliest Pleistocene hominid cranial remains from Dmanisi, Republic of Georgia: taxonomy, geological setting, and age. Science, 288, 1019–25.CrossRefGoogle Scholar
Gabunia, L., Vekua, A., Lordkipanidze, D., et al. (2000b). The environmental context of early human occupation in Georgia (Transcaucasia). Journal of Human Evolution, 34, 785–802.CrossRefGoogle Scholar
Gamble, C. S. (1986). The Palaeolithic Settlement of Europe. Cambridge: Cambridge University Press.Google Scholar
Gaudzinski, S. (1996). On bovid assemblages and their consequences for the knowledge of subsistence patterns in the Middle Paleolithic. Proceedings of the Prehistoric Society, 62, 19–39.CrossRefGoogle Scholar
Gaudzinski, S. and Roebroeks, W. (2000). Adults only: reindeer hunting at the Middle Paleolithic site Salzgitter Lebenstedt, Northern Germany. Journal of Human Evolution, 38, 497–521.CrossRefGoogle Scholar
Geraads, D. (1997). Le grande faune associee aux derniers Neandertaliens de Zafarraya (Andalousie, Espagne): systematique et essai d'interpretation. CRAS, Paris, Sciences de la terre et des planetes, 325, 725–31.Google Scholar
Golovanova, L., Hoffecker, J., Kharitonov, V., & Romanova, G. (1999). Mezmaiskaya Cave: a Neanderthal occupation in the Northern Caucasus. Current Anthropology, 40, 77–86.CrossRefGoogle Scholar
Grine, F. E. (1981). Trophic differences between gracile and robust australopithecines: a scanning electron-microscope analysis of occlusal events. South African Journal of Science 77, 203–30.Google Scholar
Grine, F., Klein, R., and Volman, T. (1991). Dating, archaeology and human fossils from the Middle Stone Age levels of Die Kelders, South Africa. Journal of Human Evolution, 21, 363–95.CrossRefGoogle Scholar
Hailemichael, M. (1999). The Pliocene environment of Hadar, Ethiopia: a comparative isotopic study of paleosol carbonates and lacustrine mollusk shells of the Hadar Formation and of modern analog. Ph. D. thesis, Case Western Reserve University.
Haile-Selassie, Y. (2001). Late Miocene hominids from the Middle Awash, Ethiopia. Nature, 412, 178–81.CrossRefGoogle ScholarPubMed
Hatley, T. & Kappelman, J. (1980). Bears, pigs, and Plio-Pleistocene hominids: a case for the exploitation of below ground food resources. Human Ecology, 8, 371–87.CrossRefGoogle Scholar
Hylander, W. (1975). Human mandible: lever or link. American Journal of Physical Anthropology, 43, 227–42.CrossRefGoogle ScholarPubMed
Hylander, W.(1988). Implications of in vivo experiments for interpreting the functional significance of “robust” Australopithecine jaws. In Evolutionary History of the “Robust” Australopithecines, ed. Grine, F. E.. New York: Aldine de Gruyter, pp. 55–8.Google Scholar
Jacobs, B. F. (1999a). The use of leaf form to estimate Miocene rainfall variables in tropical Africa. XVI International Botanical Congress, Abstract #4570.
Jacobs, B. F.(1999b). Estimation of rainfall variables from leaf characters in tropical Africa. Palaeogeography, Palaeoclimatology, Palaeoecology, 145, 231–50.CrossRefGoogle Scholar
Jacobs, B. F.(2002). Estimation of low latitude paleoclimates using fossil angiosperm leaves: examples from the Miocene Tugen Hills, Kenya. Paleobiology, 28, 399–421.2.0.CO;2>CrossRefGoogle Scholar
Janis, C. 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, Paris, 1986, ed. D. E. Russell, J. P. Santoro, & D. Sigogneau-Russell. Memoires du Museum Nationale d' Histoire Naturelle, Paris, Serie C, 53, 367–87.
Janis, C. M., Damuth, J. M., and Theodor, J. (2002). The origins and evolution of the North American grassland biome: the story from hoofed mammals. Palaeogeography, Palaeclimatology, Palaeoecology, 177, 183–98.CrossRefGoogle Scholar
Kay, R. F. (1984). On the use of anatomical features to infer foraging behavior in extinct primates. In Adaptations for Foraging in Nonhuman Primates: Contributions to an Organismal Biology of Prosimians, Monkeys and Apes, ed. Rodman, P. S. & Cant, J. G. H.. New York: Columbia University Press, pp. 21–53.Google Scholar
Kimbel, W. H. (1995). Hominid speciation and Pliocene climate change. In Paleoclimate and Evolution with Emphasis on Human Origins, ed. Vrba, E. S., Denton, G. H., Partridge, T. C., & Burckle, L. C.. New Haven, CT: Yale University Press, pp. 425–37.Google Scholar
Klein, R. (1976). The mammalian fauna of the Klasies River Mouth sites, Southern Cape Province, South Africa. South African Archaeological Bulletin, 31, 75–98.CrossRefGoogle Scholar
Klein, R.(1999) The Human Career: Human Biological and Cultural Origins. Chicago: Chicago University Press.Google Scholar
Lalueza Fox, C. & Perez-Perez, A. (1993). The diet of the Neanderthal child Gibraltar 2 (Devil's Tower) through the study of the vestibular striation pattern. Journal of Human Evolution, 24, 29–41.Google Scholar
Lalueza, C., Perez-Perez, A., & Turbon, D. (1996) Dietary inferences through buccal microwear analysis of Middle and Upper Pleistocene human fossils. American Journal of Physical Anthropology, 100, 367–87.3.0.CO;2-R>CrossRefGoogle ScholarPubMed
Leakey, M. G., Spoor, F., Brown, F. H., et al. (2001). New hominin genus from eastern Africa shows diverse middle Pliocene lineages. Nature, 410, 433–40.CrossRefGoogle ScholarPubMed
Leiberman, D. (1993). The rise and fall of seasonal mobility among hunter-gatherers. Current Anthropology, 35, 569–98.Google Scholar
Leiberman, D., & Shea, J. (1994). Behavioral differences between archaic and modern humans in the Levantine Mousterian. American Anthropologist, 96, 300–332.CrossRefGoogle Scholar
Lind, E. M. and Morrison, M. E. S. (1974). East African Vegetation. Bristol, UK: Longman.Google Scholar
Marean, C. W. (1989). Sabertooth cats and their relevance for early hominid diet and evolution. Journal of Human Evolution, 18, 559–82.CrossRefGoogle Scholar
Marean, C. W.(1997). Hunter-gatherer foraging strategies in tropical grasslands: model building and testing in the East African Middle and Later Stone Age. Journal of Anthropological Archaeology, 16, 189–225.CrossRefGoogle Scholar
Marean, C. W. and Kim, S. Y. (1998). Mousterian large mammal remains from Kobeh Cave (Zagros Mountains, Iran): behavioral implications for Neandertals and early modern humans. Current Anthropology, 39, S79–114.CrossRefGoogle Scholar
Marlow, J. R., Lange, C. B., Wefer, G., & Rosell-Mele, A. (2000). Upwelling intensification as part of the Pliocene–Pleistocene climate transition. Science, 290, 288–91.Google ScholarPubMed
McDougal, I., Braoen, F. H., & Fleagle, J. G. (2005). Stratigraphic placement and age of modern humans from Kibish, Ethiopia. Nature, 433, 733–6.CrossRefGoogle Scholar
McHenry, H. (1992). How big were early hominids?Evolutionary Anthropology, 1, 15–20.CrossRefGoogle Scholar
Monahan, C. M. (1996). New zooarchaeological data from Bed II, Olduvai Gorge, Tanzania: implications for hominid behavior in the Early Pleistocene. Journal of Human Evolution, 31, 93–128.CrossRefGoogle Scholar
O'Brien, E. (1998). Water-energy dynamics, climate, and prediction of woody plant species richness: an interim general model. Journal of Biogeography, 25, 379–98.CrossRefGoogle Scholar
O'Connell, J., Hawkes, K., & Blurton Jones, N. (1999). Grandmothering and the evolution of Homo erectus. Journal of Human Evolution, 36, 461–85.CrossRefGoogle ScholarPubMed
Pearson, O. M. (2000). Activity, climate, and postcranial robusticity. Current Anthropology, 41, 569–607.Google ScholarPubMed
Potts, R. (1998). Environmental hypotheses of hominin evolution. Yearbook of Physical Anthropology, 41, 93–136.3.0.CO;2-X>CrossRefGoogle Scholar
Pratt, D. J. and Gwynne, M. D. (1977). Rangeland Management and Ecology in East Africa. London: Hodder and Stoughton.Google Scholar
Rabinovich, R. & Tchernov, E. (1995). Chronological, paleoecological and taphonomical aspects of the Middle Paleolithic Site of Qafzeh, Israel. In Archaeozoology of the Near East II: Proceedings of the Second International Symposium on the Archaeozoology of Southwestern Asia and Adjacent Areas, ed. Buitenhuis, H. & Uerpman, H.-P.Leiden: Backhuys Publishers, pp. 5–44.Google Scholar
Reed, K. E. (1997). Early hominid evolution and ecological change through the African Plio-Pleistocene. Journal of Human Evolution, 32, 289–322.CrossRefGoogle ScholarPubMed
Reed, K. E.(1998). Using large mammal communities to examine ecological and taxonomic organization and predict vegetation in extant and extinct assemblages. Paleobiology, 24, 384–408.Google Scholar
Reed, K. E.(2002). The use of paleocommunity and taphonomic studies in reconstructing primate behavior. In Reconstructing Primate Behavior in the Fossil Record, ed. Plavcan, M. J., Kay, R., Schaik, C., & Jungers, W. L.. New York: Kluwer Academic/Plenum Press, pp. 217–59.CrossRefGoogle Scholar
Reed, K. E.(in preparation) Paleoecology of the Hadar Formation, Ethiopia.
Richards, M., Pettitt, P., Trinkaus, E., Smith, F., Paunovic, M., & Karavanic, I. (2000). Neanderthal diet at Vindija and Neanderthal predation: the evidence from stable isotopes. Proceedings of the National Academy of Sciences, USA, 97, 7663–66.CrossRefGoogle ScholarPubMed
Rightmire, P. (1996). The human cranium from Bodo, Ethiopia: evidence for speciation in the Middle Pleistocene?Journal of Human Evolution, 31, 21–39.CrossRefGoogle Scholar
Ritchie, M. E. & Olff, H. (1999). Spatial scaling laws yield a synthetic theory of biodiversity. Nature, 400, 557–60.CrossRefGoogle ScholarPubMed
Roebroeks, W. (2001). Hominid behaviour and the earliest occupation of Europe: an exploration. Journal of Human Evolution, 41, 437–61.CrossRefGoogle Scholar
Roebroeks, W., Conrad, N., & Kolfschoten, T. (1992). Dense forests, cold steppes, and the Paleolithic settlement of Northern Europe. Current Anthropology, 33, 551–86.CrossRefGoogle Scholar
Rosenberger, A. L. & Kinzey, W. G. (1976). Functional patterns of molar occlusion in platyrrhine primates. American Journal of Physical Anthropology, 45, 281–98.CrossRefGoogle ScholarPubMed
Ruff, C., Trinkaus, E., & Holliday, T. (1997). Body mass and encephalization in Pleistocene Homo. Nature, 387, 173–6.CrossRefGoogle ScholarPubMed
Ryan, A. C. & Johanson, D. C. (1989). Anterior dental microwear in Australopithecus afarensis: comparisons with human and nonhuman primates. Journal of Human Evolution, 18, 235–68.CrossRefGoogle Scholar
Schefuss, E., Pancost, R. D., Jansen, J. H. F., & Damste, J. S. S. (2000). The mid-Pleistocene climate transition: insight from organic geochemical records from the tropical Atlantic. Journal of Conference Abstracts, 5, 886.Google Scholar
Semaw, S., Renne, P., Harris, J. W. K., et al. (1997). 2.5-million-year-old stone tools from Gona, Ethiopia. Nature, 385, 333–6.CrossRefGoogle ScholarPubMed
Shea, J. (2003). Neandertals, competition, and the origin of modern human behavior in the Levant. Evolutionary Anthropology, 12, 173–87.CrossRefGoogle Scholar
Spencer, L. M. (1997). Dietary adaptations of Plio-Pleistocene Bovidae: implications for hominid habitat use. Journal of Human Evolution, 32, 201–28.CrossRefGoogle ScholarPubMed
Speth, J. & Tchernov, E. (1998). The role of hunting and scavenging in Neandertal procurement strategies: new evidence from Kebara cave (Israel). In Neandertals and Modern Humans in Western Asia, ed. Akazawa, T., Aoki, K., & Bar-Yosef, B.. New York: Plenum Press, pp. 223–39.Google Scholar
Sponheimer, M. & Lee-Thorp, J. (1999). Isotopic evidence for the diet of an early hominid, Australopithecus africanus. Science, 283, 368–70.CrossRefGoogle ScholarPubMed
Suzuki, H. & Takai, F. (1970). The Amud Man and His Cave Site. Tokyo: Academic Press of Japan.Google Scholar
Teaford, M. (1988). Scanning electron-microscope diagnosis of wear patterns versus artifacts on fossil teeth. Scanning Microscopy, 2, 1149–66.Google ScholarPubMed
Teaford, M. & Ungar, P. (2000). Diet and the evolution of the earliest human ancestors. Proceedings of the National Academy of Sciences, USA, 97, 13506–11.CrossRefGoogle ScholarPubMed
Thieme, H. (1997). Lower Palaeolithic hunting spears from Germany. Nature, 385, 807–10.CrossRefGoogle ScholarPubMed
Trinkaus, E. & Zimmerman, M. R. (1982). Trauma among the Shanidar Neanderthals. American Journal of Physical Anthropology, 57, 61–76.CrossRefGoogle Scholar
Ungar, P. (1998). Dental allometry, morphology and wear as evidence for diet in fossil primates. Evolutionary Anthropology, 6, 205–17.3.0.CO;2-9>CrossRefGoogle Scholar
Ungar, P. & Grine, F. (1991). Incisor size and wear in Australopithecus africanus and Paranthropus robustus. Journal of Human Evolution, 20, 313–40.CrossRefGoogle Scholar
Andel, T. H. & Tzedakis, P. C. (1996). Paleolithic landscapes of Europe and environs, 150,000–25,000 years ago: an overview. Quaternary Science Review, 15, 481–500.CrossRefGoogle Scholar
Merwe, N. J., Thackeray, J. F., Lee-Thorp, J. A., & Luyt, J. (2003). The carbon isotope ecology and diet of Australopithecus africanus at Sterkfontein, South Africa. Journal of Human Evolution, 44, 581–97.CrossRefGoogle ScholarPubMed
Vekua, A., Lordkipanidze, D., Rightmire, G. P., et al. (2002). A new skull of early Homo from Dmanisi, Georgia. Science, 297, 85–9.CrossRefGoogle ScholarPubMed
Vrba, E. S. (1988a). Late Pliocene climatic events and hominid evolution. In Evolutionary History of the “Robust” Australopithecines, ed. Grine, F. E.. New York: Aldine de Gruyter, pp. 405–26.Google Scholar
Vrba, E. S.(1988b). Evolution, species and fossils: how does life evolve?South African Journal of Science, 76, 61–84.Google Scholar
Vrba, E. S.(1992). Mammals as a key to evolutionary theory. Journal of Mammalogy, 73, 1–28.CrossRefGoogle Scholar
Vrba, E.(1995). On the connections between paleoclimate and evolution. In Paleoclimate and Evolution with Emphasis on Human Origins, ed. Vrba, E. S., Denton, G. H., Partridge, T. C., & Burckle, L. C.. New Haven, CT: Yale University Press, pp. 385–424.Google Scholar
White, F (1983).The Vegetation of Africa: A Descriptive Memoir to Accompany UNESCO/AETFAT/UNSO Vegetation Maps of Africa. Paris: United Nations Educational, Scientific and Cultural Organisation.
White, T. D., Asfar, B., Dectusta, D., et al. (2003). Pleistocene Homo sapiens from Awash, Ethiopia. Nature, 423, 742–7.CrossRefGoogle ScholarPubMed
Woldegabriel, G.White, T. D., Suwa, G., et al. (1994). Ecological and temporal placement of early Pliocene hominids at Aramis, Ethiopia. Nature, 371, 330–33.CrossRefGoogle ScholarPubMed
Wrangham, R. W., Jones, J. H., Laden, G., Pilbeam, D., & Conklin-Brittain, N. (1999). The raw and the stolen: cooking and the ecology of human origins. Current Anthropology, 40, 567–94.Google ScholarPubMed
Wynn, J. (2000). Paleosols, stable carbon isotopes, and paleoenvironmental interpretation of Kanapoi, Northern Kenya. Journal of Human Evolution, 39, 411–32.CrossRefGoogle ScholarPubMed

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

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
×