Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-26T11:01:48.876Z Has data issue: false hasContentIssue false

Part II - Approaches: Concepts and Methods

Published online by Cambridge University Press:  10 November 2017

Carole L. Crumley
Affiliation:
University of North Carolina, Chapel Hill
Tommy Lennartsson
Affiliation:
Swedish Biodiversity Centre, Uppsala
Anna Westin
Affiliation:
Swedish Biodiversity Centre, Uppsala
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Issues and Concepts in Historical Ecology
The Past and Future of Landscapes and Regions
, pp. 143 - 272
Publisher: Cambridge University Press
Print publication year: 2017

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

References

Anderson, K. (1997). A walk on the wild side: a critical geography of domestication. Progress in Human Geography, 21, 463–85.Google Scholar
Archibald, S. (2008). African grazing lawns – how fire, rainfall, and grazer numbers interact to affect grass community status. Journal of Wildlife Management, 72, 492501.CrossRefGoogle Scholar
Augustine, D. J. (2003). Long-term, livestock mediated redistribution of nitrogen and phosphorus in an East African savanna. Journal of Applied Ecology, 40, 137–49.Google Scholar
Augustine, D. J. & McNaughton, S. J. (2004). Regulation of shrub dynamics by native browsing ungulates on East African rangeland. Journal of Applied Ecology, 41, 4558.CrossRefGoogle Scholar
Atlestam, P-O. (1942). Bohusläns ljunghedar, en geografisk studie. Gothenburg: Meddelanden från Göteborgs högskolas Geografiska institution 30.Google Scholar
Balée, W. (2006). The research program of historical ecology. Annual Review of Anthropology, 35, 7598.Google Scholar
Barad, K. (2007). Meeting the Universe Halfway: Quantum Physics and the Entanglement of Matter and Meaning. Durham, NC: Duke University Press.Google Scholar
Barrow, E. C. G. (1990). Usufruct rights to trees: the role of Ekwar in dryland central Turkana, Kenya. Human Ecology, 18, 163–76.CrossRefGoogle Scholar
Bennet, J. (2010). Vibrant Matter: A Political Ecology of Things. Durham, NC: Duke University Press.Google Scholar
Berglund, B. E., ed. (1991). The cultural landscape during 6000 years in southern Sweden. Ecological Bulletins, 41, 1495.Google Scholar
Blench, R. (2007). The intertwined history of silk-cotton and baobab. In Cappers, R. J., ed., Fields of Change: Progress in African Archaeobotany. Groningen Archaeological Studies 5. Groningen: Barkhuis Publishing & Groningen University Library, pp. 119.Google Scholar
Boles, J. C. & Lane, P. J. (2016). The green, green grass of home: an archaeo-ecological approach to pastoralist settlement in central Kenya. Azania: Archaeological Research in Africa, 51(4), 507–30. Doi: 10.1080/0067270X.2016.1249587.Google Scholar
Bourdieu, P. (1977). Outline of a Theory of Practice. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Boyd, R., Richerson, P. J., & Henrich, J. (2011). The cultural niche: why social learning is essential for human adaptation. Proceedings of the National Academy of Sciences USA, 108, 10918–25.CrossRefGoogle ScholarPubMed
Broughton, J. M., Cannon, M. D., & Bartelink, E. J. (2010). Evolutionary ecology, resource depression, and niche construction theory: applications to Central California hunter-gatherers and Mimbres-Mogollon agriculturalists. Journal of Archaeological Method and Theory, 17, 371421.CrossRefGoogle Scholar
Butzer, K. W. (1982). Archaeology as Human Ecology: Method and Theory for a Contextual Approach. Cambridge: Cambridge University Press.Google Scholar
Butzer, K. W. (1996). Ecology in the long view: settlement histories, agrosystemic strategies, and ecological performance. Journal of Field Archaeology, 23, 141–50.Google Scholar
Butzer, K. W. (2005). Environmental history in the Mediterranean world: cross-disciplinary investigation of cause-and-effect for degradation and soil erosion. Journal of Archaeological Science, 32, 17731800.Google Scholar
Campbell, Å. (1948). Från Vildmark till bygd. En Etnologisk undersökning av nybyggarkulturen I Lappland före industrialismens genombrott. Landsmåls- och Folkminnesarkivet, Uppsala, B:5.Google Scholar
Callon, M. & Law, J. (1995). Agency and the hybrid collectif. South Atlantic Quaterly, 94, 481507.Google Scholar
Causey, M. (2008). Delineating Pastoralist Behaviour and Long-Term Environmental Change: A GIS Landscape Approach on the Laikipia Plateau, Kenya, PhD Dissertation, School of Archaeology, University of Oxford.Google Scholar
Cech, P. G., Kuster, T., Edwards, P. J., & Venterink, H. O. (2008). Effects of herbivory, fire and N2-fixation on nutrient limitation in a humid African savannah. Ecosystems, 11, 9911004.Google Scholar
Chase, J. M. & Leibold, J. M. (2003). Ecological Niches: Linking Classical and Contemporary Approaches. Chicago: University of Chicago Press.Google Scholar
Childe, V. G. (1928). The Most Ancient East: The Oriental Prelude to European Prehistory. London: Kegan Paul.Google Scholar
Clement, R. C., Denevan, W. M., Heckenberger, M. J., Junqueira, A. B., Neves, E. G., Teixeira, W. G., & Woods, W. I. (2015). The domestication of Amazonia before European conquest. Proceeding of the Royal Society B, 282, 20150813.Google ScholarPubMed
Clements, F. (1916). Plant Succession. Washington, DC: Carnegie Institution.Google Scholar
Cousins, S. A. O. & Eriksson, O. (2002). The influence of management history and habitat on plant species richness in a rural hemiboreal landscape, Sweden. Landscape Ecology, 17, 517–29.Google Scholar
Cowles, H. C. (1899). The ecological relations of the vegetation on the sand dunes of Lake Michigan. Botanical Gazette, 27, 95117, 167–202, 281–308, 361–91.Google Scholar
Crumley, C. (1987). Celtic settlement before the Conquest: the dialectics of landscape and power. In Crumley, C. L. & Marquardt, W H., eds., Regional Dynamics: Burgundian Landscapes in Historical Perspective. San Diego, CA: Academic Press, pp. 403–29.Google Scholar
Crumley, C. L. (ed.). (1994). Historical Ecology: Cultural Knowledge and Changing Landscapes. Santa Fe, NM: School of American Research Press.Google Scholar
Crumley, C. L. (2007). Historical ecology: integrated thinking at multiple temporal and spatial scales. In Hornborg, A. & Crumley, C., eds., The World System and the Earth System: Global Socioenvironmental Change and Sustainability since the Neolithic. Walnut Creek, CA: Left Coast Press, pp. 1528.Google Scholar
Danchin, E. (2013). Avatars of information: towards an inclusive evolutionary synthesis. Trends in Ecology and Evolution, 28, 351–8.Google Scholar
Darwin, C. R. (1859). On the Origin of Species by means of Natural Selection, or the Preservation of Favoured Races in the Struggle for Life. London: John Murray.CrossRefGoogle Scholar
Davies, M. I. J. (2013). Environment in North American and European archaeology. In Davies, M. I. J. & M’Mbogori, F. N., eds., Humans and the Environment: New Archaeological Perspectives for the 21st Century. Oxford: Oxford University Press, pp. 325.Google Scholar
Deleuze, G. & Guattari, F. [1980] (2010). A Thousand Plateaus. London: Bloomsbury Academic.Google Scholar
Del Mármol, C. & Vaccaro, I. (2015). Changing ruralities: between abandonment and redefinition in the Catalan Pyrenees. Anthropological Forum, 25, 2141.CrossRefGoogle Scholar
Denevan, W. M. (2011). The ‘pristine myth’ revisited. Geographical Review, 101, 576–91.Google Scholar
Dieter, J. (2012). Evolution and the paleolithic. Notae Praehistoricae, 32, 257–87.Google Scholar
Dobney, K. & Larson, G. (2006). Genetics and animal domestication: new windows on an elusive process. Journal of Zoology, 269, 261–71.Google Scholar
Dunnell, R. C. (1971). Systematics in Prehistory. New York: Free Press.Google Scholar
Dupouey, J.-L., Dambrine, E., Lafitte, J. D., & Moares, C. (2002). Irreversible impact of past land use on forest soils and biodiversity. Ecology, 83, 2978–84.Google Scholar
Duvall, C. S. (2007). Human settlement and baobab distribution in south-western Mali. Journal of Biogeography, 34, 1947–61.Google Scholar
Earle, T. & Spriggs, M. (2015). Political economy in prehistory: a Marxist approach to Pacific sequences. Current Anthropology, 56, 515–44.Google Scholar
Egerton, F. N. (1973). Changing concepts of the balance of nature. Quarterly Review of Biology, 48, 322–50.Google Scholar
Ellen, R. (1982). Environment, Subsistence and System: The Ecology of Small-Scale Social Formations. Cambridge: Cambridge University Press.Google Scholar
Ellis, E. C., Kaplan, J. O., Fuller, D. Q., Vavrus, S., Goldewijk, K. K., & Verburg, P. H. (2013). Used planet: a global history. Proceedings of the National Academy of Sciences USA, 110, 7978–85.Google Scholar
Ellis, E. C. & Ramankutty, N. (2008). Putting people in the map: anthropogenic biomes of the world. Frontiers in Ecology and the Environment, 6, 439–47.Google Scholar
Emanuelsson, U. (2009). The Rural Landscape of Europe: How Man Has Shaped European Nature. Stockholm: The Swedish Research Council Formas.Google Scholar
Erickson, C. L. (2006). The domesticated landscapes of the Bolivian Amazon. In Balée, W. & Erickson, C. L., eds., Time and Complexity in Historical Ecology. New York: Columbia University Press, pp. 235–78.Google Scholar
Erickson, C. L. (2008). Amazonia: the historical ecology of a domesticated landscape. In Silverman, H. & Isbell, W. H., eds., Handbook of South American Archaeology. New York: Springer, pp. 157–83.Google Scholar
Erickson, C. L. (2010). The transformation of environment into landscape: the historical ecology of monumental earthwork construction in the Bolivian Amazon. Diversity, 2, 618–52.Google Scholar
Eriksson, O. (2013). Species pools in cultural landscapes – niche construction, ecological opportunity and niche shifts. Ecography, 36, 403–13.Google Scholar
Eriksson, O. & Arnell, M. (2017). Niche construction, entanglement and landscape domestication in Scandinavian infield systems. Landscape Research, 42(1), 7888. Doi: 10.1080/01426397.2016.1255316.Google Scholar
Eriksson, O. & Cousins, S. A. O. (2014). Historical landscape perspectives on grasslands in Sweden and the Baltic region. Land, 3, 300–21.CrossRefGoogle Scholar
Escobar, A. (1996). Construction nature: elements for a post structuralist political ecology. Futures, 28, 325–43.Google Scholar
Fairhead, J. & Leach, M. (1996). Misreading the African Landscape: Society and Ecology in a Forest-Savanna Mosaic. Cambridge: Cambridge University Press.Google Scholar
Forman, R. T. T. & Godron, M. (1986). Landscape Ecology. New York: Wiley.Google Scholar
Foster, D. R. (2002). Thoreau’s country: a historical-ecological perspective on conservation in a New England landscape. Journal of Biogeography, 29, 1537–55.Google Scholar
Foucault, M. (1970). The Order of Things. New York: Random House.Google Scholar
Fraser, J. A., Frausin, V., & Jarvis, A. (2015). An intergenerational transmission of sustainability? Ancestral habitus and food production in a traditional agro-ecosystem of the Upper Guinea Forest, West Africa. Global Environmental Change, 31, 226–38.Google Scholar
Frausin, V., Fraser, J. A., Narmah, W., Lahai, M. K., Winnebah, T. R., Fairhead, J., & Leach, M. (2014). ‘God made the soil, but we made it fertile’: gender, knowledge, and practice in the formation and use of African Dark Earths in Liberia and Sierra Leone. Human Ecology, 42, 695710.Google Scholar
Fried, M. (1967). The Evolution of Political Society: An Essay in Political Economy. New York: Random House.Google Scholar
Friedman, J. & Rowlands, M. (1978). Notes toward an epigenetic model of the evolution of ‘civilization’. In Friedman, J. & Rowlands, M., eds., The Evolution of Social Systems. London: Duckworth, pp. 201–76.Google Scholar
Gadd, C. J. (2011). The agricultural revolution in Sweden 1700–1870. In Myrdal, J. & Morell, M., eds., The Agrarian History of Sweden from 4000 BC to AD 2000. Lund: Nordic Academic Press, pp. 118–64.Google Scholar
Gero, J. & Conkey, W., eds. (1991). Engendering Archaeology. Oxford: Blackwell.Google Scholar
Giddens, A. (1984). The Constitution of Society: Outline of the Theory of Structuration. Cambridge: Polity Press.Google Scholar
Gillson, L. (2004). Testing non-equilibrium theories in savannas: 1400 years of vegetation change in Tsavo National Park, Kenya. Ecological Complexity, 1, 281–98.Google Scholar
Gilman, A. (1981). The development of social stratification in Bronze Age Europe. Current Anthropology, 22, 124.Google Scholar
Gleason, H. A. (1926). The individualistic concept of the plant association. Bulletin of the Torrey Botanical Club, 53, 726.Google Scholar
Godsoe, W. (2010). I can’t define the niche but I know it when I see it: a formal link between statistical theory and the ecological niche. Oikos, 119, 5360.Google Scholar
Gosden, C. (2005). What do objects want? Journal of Archaeological Method and Theory, 12, 193211.Google Scholar
Gustavsson, E., Dahlström, A., Emanuelsson, , Wissman, M., , J., & Lennartsson, T. (2011). Combining historical and ecological knowledge to optimise biodiversity conservation in semi-natural grasslands. In Pujol, J. L., ed., The Importance of Biological Interactions in the Study of Biodiversity. New York: InTech Publishers, pp. 173–96.Google Scholar
Haraway, D. (1991). Simians, Cyborgs, and Women: The Reinvention of Nature. London: Free Association Books.Google Scholar
Harris, D. (1996). Domesticatory relationships of people, plants and animals. In Ellen, R. & Fukui, K., eds., Redefining Nature: Ecology, Culture and Domestication. Oxford: Berg, pp. 437–63.Google Scholar
Harris, S. E. (2012). Cyprus as a degraded landscape or resilient environment in the wake of colonial intrusion. Proceedings of the National Academy of Sciences USA, 109, 3670–5.Google Scholar
Hermy, M. & Verheyen, K. (2007). Legacies of the past in the present-day forest biodiversity: a review of past land-use effects on forest plant species composition and diversity. Ecological Research, 22, 361–71.Google Scholar
Higgs, E., Falk, D. A., Guerrini, A., Hall, M., Harris, J., Hobbs, R. J., Jackson, S. T., Rhemtulla, J. M., & Throop, W. (2014). The changing role of history in restoration ecology. Frontiers in Ecology and the Environment, 12, 499506.Google Scholar
Hinchcliffe, S. (2007). Geographies of Nature: Societies, Environments, Ecologies. Los Angeles: Sage Publications.Google Scholar
Hodder, I. (1982). Theoretical archaeology: a reactionary view. In Hodder, I., ed., Symbolic and Structural Archaeology. Cambridge: Cambridge University Press, pp. 116.Google Scholar
Hodder, I. (2012). Entangled: An Archaeology of the Relationships between Humans and Things. Oxford: Blackwell Wiley.Google Scholar
Hubbell, S. P. & Foster, R. S. (1986). Biology, chance, and history and the structure of tropical rainforest tree communities. In Diamond, J. & Case, J. D., eds., Community Ecology. New York: Harper and Row, pp. 314–43.Google Scholar
Hutchinson, G. E. (1957). Concluding remarks. Cold Spring Harbor Symposium of Quantitative Biology, 22, 415–27.Google Scholar
Jarman, M. R. & Wilkinson, P. F. (1972). Criteria of animal domestication. In Higgs, E. S., ed., Papers in Economic Prehistory. Cambridge: Cambridge University Press, pp. 8396.Google Scholar
Johansson, L. (1947). Bebyggelse och folkliv i det gamla Frostviken. Landsmåls- och Folkminnesarkivet i Uppsala B:3. Uppsala.Google Scholar
Josefsson, T., Gunnarsson, B., Liedgren, L., Bergman, I., & Östlund, L. (2010). Historical human influence on forest composition and structure in boreal Fennoscandia. Canadian Journal of Forest Research, 40, 872–84.Google Scholar
Kareiva, P., Watts, S., McDonald, R., & Boucher, T. (2007). Domesticated nature: shaping landscapes and ecosystems for human welfare. Science, 316, 1866–9.Google Scholar
Kendal, J., Tehrani, J. J., & Odling-Smee, J. (2011). Human niche construction in interdisciplinary focus. Philosophical Transactions of the Royal Society B, 366, 785–92.Google Scholar
Kirch, P. V. (1984). The Evolution of Polynesian Chiefdoms. Cambridge: Cambridge University Press.Google Scholar
Kjellström, R. (2012). Nybyggarliv i Vilhelmina. 1. Träd och växter som resurs. Acta Academiae Regiae Gustavi Adolphi 119, CBM Publications 65. Uppsala.Google Scholar
Kricher, J. (2009). The Balance of Nature: Ecology’s Enduring Myth. Princeton, NJ: Princeton University Press.Google Scholar
Lagerås, P. (2007). The Ecology of Expansion and Abandonment: Medieval and Post-medieval Land-Use and Settlement Dynamics in a Landscape Perspective. Stockholm: National Heritage Board, Sweden.Google Scholar
Laland, K. N. & O’Brien, M. J. (2011). Cultural niche construction: an introduction. Biological Theory, 6, 191202.Google Scholar
Lamprey, H. F., Halevy, G., & Makacha, S. (1974). Interactions between Acacia bruchid seed beetles and large herbivores. East African Wildlife Journal, 12, 81–5.Google Scholar
Lane, P. J. (2011). An outline of the later Holocene archaeology and precolonial history of the Ewaso Basin, Kenya. Smithsonian Contributions to Zoology, 632, 1130.Google Scholar
Lane, P. J. (2016). Entangled banks and the domestication of East African pastoralist landscapes. In Fernandini, F. & Der, L., eds., Archaeology of Entanglement. Walnut Creek, CA: Left Coast Press, pp. 127–50.Google Scholar
Latour, B. (1999). Pandora’s Hope: Essays in the Reality of Science Studies. Cambridge, MA: Harvard University Press.Google Scholar
Latour, B. (2000). When things strike back – a possible contribution of science studies to the social sciences. British Journal of Sociology, 51, 107–23.Google Scholar
Latour, B. (2005). Reassembling the Social: An Introduction to Actor-Network-Theory. Oxford: Oxford University Press.CrossRefGoogle Scholar
Law, J. & Mol, A. (2002). Complexities: Social Studies of Knowledge Practices. Durham, NC: Duke University Press.Google Scholar
Leach, H. (2003). Human domestication reconsidered. Current Anthropology, 44, 349–68.Google Scholar
Maley, J. & Brenac, P. (1998). Vegetation dynamics, palaeoenvironments and climatic changes in the forests of western Cameroon during the last 28,000 years B.P. Review of Palaeobotany and Palynology, 99, 157–87.Google Scholar
Maranz, S. (2009). Tree mortality in the African Sahel indicates an anthropogenic ecosystem displaced by climate change. Journal of Biogeography, 36, 1181–93.Google Scholar
Marcus, G. A. & Fischer, M. F., eds. (1986). Anthropology as Cultural Critique: An Experimental Moment in the Human Sciences. Chicago: University of Chicago Press.Google Scholar
Marquardt, W. H. (1992). Dialectical archaeology. In Schiffer, M. B., ed., Archaeological Method and Theory, Volume 4. Tucson: University of Arizona Press, pp. 101–40.Google Scholar
Mather, C. (2003). Shrines and the domestication of landscape. Journal of Anthropological Research, 59, 2345.Google Scholar
McCann, J. (1999). Green Land, Brown Land, Black Land: An Environmental History of Africa, 1800–1990. London: James Currey.Google Scholar
Mesoudi, A. (2011). Cultural Evolution: How Darwinian Theory Can Explain Human Culture and Synthesize the Social Sciences. Chicago: University of Chicago Press.Google Scholar
Meyer, W. J. & Crumley, C. L. (2011). Historical ecology: using what works to cross the divide. Atlantic Europe in the first millennium BC: crossing the divide. In Moore, T. & Armada, L., eds., Atlantic Europe in the First Millennium BC: Crossing the Divide. Oxford: Oxford University Press, pp. 109–34.Google Scholar
Morris, D. L., Western, D., & Maitumo, D. (2009). Pastoralist’s livestock and settlements influence game bird diversity and abundance in a savanna ecosystem of southern Kenya. African Journal of Ecology, 47, 4855.Google Scholar
Muchiru, A. N., Western, D. J., & Reid, R. S. (2008). The role of abandoned pastoral settlements in the dynamics of African large herbivore communities. Journal of Arid Environments, 72, 940–52.Google Scholar
Myrdal, J. (2011). Farming and feudalism. In Myrdal, J. & Morell, M., eds., The Agrarian History of Sweden from 4000 BC to AD 2000. Lund: Nordic Academic Press, pp. 72117.Google Scholar
Nyerges, A. E. ed. (1997). The Ecology of Practice: Studies of Food Crop Production in Sub-Saharan West Africa. Amsterdam: Gordon & Breach.Google Scholar
O’Brien, M. J. & Laland, K. N. (2012). Genes, culture, and agriculture: an example of human niche construction. Current Anthropology, 53, 434–70.Google Scholar
O’Connor, T. P. (1997). Working at relationships: another look at animal domestication. Antiquity, 71, 149–56.Google Scholar
Odling-Smee, J., Erwin, D., Palkovacs, E. P., Feldman, M. W., & Laland, K. N. (2013). Niche construction theory: a practical guide for ecologists. Quarterly Review of Biology, 88, 328.Google Scholar
Odling-Smee, F. J., Laland, K. N., & Feldman, M. W. (2003). Niche Construction: The Neglected Process in Evolution. Princeton, NJ: Princeton University Press.Google Scholar
Odum, H. T. (1983). Systems Ecology: An Introduction. New York: Wiley.Google Scholar
O’Neill, R. V., DeAngelis, D. L., Waide, J. B., & Allen, T. F. H. (1986). A Hierarchical Concept of Ecosystems. Princeton, NJ: Princeton University Press.Google Scholar
Pearman, P. B., Guisan, A., Broennimann, O., & Randin, C. F. (2008). Niche dynamics in space and time. Trends in Ecology and Evolution, 23, 149–58.Google Scholar
Pedersen, E. A. & Widgren, M. (2011). Agriculture in Sweden, 800 BC–AD 1000. In Myrdal, J. & Morell, M., eds., The Agrarian History of Sweden from 4000 BC to AD 2000. Lund: Nordic Academic Press, pp. 4671.Google Scholar
Pickett, S. T. A. & White, P. S. (1985). Patch dynamics: a synthesis. In Pickett, S. T. A. & White, P. S., eds., The Ecology of Natural Disturbance and Patch Dynamics. San Diego, CA: Academic Press, pp. 371–84.Google Scholar
Pinker, S. (2010). The cognitive niche: coevolution of intelligence, sociality, and language. Proceedings of the National Academy of Sciences USA, 107, 8993–9.Google Scholar
Post, D. M. & Palkovacs, E. P. (2009). Eco-evolutionary feedbacks in community and ecosystem ecology: interactions between the ecological theatre and the evolutionary play. Philosophical Transactions of the Royal Society B, 364, 1629–40.Google Scholar
Rackham, O. (1996). Ecology and pseudo-ecology: the example of ancient Greece. In Shipley, G., ed., Human Landscapes in Classical Antiquity: Environment and Culture. London and New York: Routledge, pp. 1643.Google Scholar
Rappaport, R. A. (1968). Pigs for the Ancestors: Ritual in the Ecology of a New Guinea People. New Haven, CT: Yale University Press.Google Scholar
Reid, R. S. & Ellis, J. E. (1995). Impacts of pastoralists on woodlands in South Turkana, Kenya: livestock-mediated tree recruitment. Ecological Applications, 5, 978–92.Google Scholar
Renfrew, C. (1972). The Emergence of Civilization: The Cyclades and the Aegean in the Third Millennium B.C. London: Methuen.Google Scholar
Renfrew, C., Rowlands, M. J., & Segraves, B. A. eds. (1982). Theory and Explanation in Archaeology: The Southampton Conference. London: Academic Press.Google Scholar
Rindos, D. (1984). The Origins of Agriculture: An Evolutionary Perspective. London: Academic Press.Google Scholar
Robbins, P. (2012). Political Ecology, 2nd ed. Oxford: Wiley-Blackwell.Google Scholar
Rosman, A. & Rubel, P. G. (1989). Stalking the wild pig: hunting and horticulture in Papua New Guinea. In Kent, S., ed., Farmers as Hunters. Cambridge: Cambridge University Press, pp. 2736.Google Scholar
Russell, N. (2002). The wild side of animal domestication. Society & Animals, 10, 285302.Google Scholar
Sahlins, M. (1985). Islands of History. Chicago: University of Chicago Press.Google Scholar
Seddon, A. W. R., Machay, A. W, Baker, A. G., & Birks, H. G. (2014). Looking forward through the past: identification of 50 priority research questions in palaeoecology. Journal of Ecology, 102, 256–67.Google Scholar
Shanks, M. & Tilley, C. (1987). Re-constructing Archaeology: Theory and Practice. Cambridge: Cambridge University Press.Google Scholar
Smith, B. D. (2007). Niche construction and the behavioral context of plant and animal domestication. Evolutionary Anthropology, 16, 188–99.Google Scholar
Smith, B. D. (2011a). A cultural niche construction theory of initial domestication. Biological Theory, 6, 260–71.Google Scholar
Smith, B. D. (2011b). General patterns of niche construction and the management of ‘wild’ plant and animal resources by small-scale pre-industrial societies. Philosophical Transactions of the Royal Society B, 366, 836–48.Google Scholar
Soberón, J. (2007). Grinnellian and Eltonian niches and geographic distributions of species. Ecology Letters, 10, 1115–23.Google Scholar
Söderström, B. & Reid, R. S. (2010). Abandoned pastoral settlements provide concentrations of resources for savannah birds. Acta Oecologica, 36, 184–96.Google Scholar
Spector, J. D. (1993). What this Awl Means: Feminist Archaeology at a Wahpeton Dakota Village. Minneapolis: Minnesota Historical Society Press.Google Scholar
Spriggs, M., ed. (1984). Marxist Perspectives in Archaeology. Cambridge: Cambridge University Press.Google Scholar
Straight, B., Lane, P. J., Hilton, C., & Letua, M. (2016). ‘Dust People’: Samburu perspectives on disaster, identity, and landscape. Journal of Eastern African Studies, 10, 168–88.Google Scholar
Szabó, P. (2010). Why history matters in ecology: an interdisciplinary perspective. Environmental Conservation, 37, 380–7.Google Scholar
Szabó, P., & Hedl, R. (2011). Advancing the integration of history and ecology for conservation. Conservation Biology, 25, 680–7.Google Scholar
Terrell, J. E. & Hart, J. O. (2006). Domesticated landscapes. In David, B. & Thomas, J., eds., Handbook of Landscape Archaeology. Walnut Creek, CA: Left Coast Press, pp. 328–32.Google Scholar
Thompson, J. N. (1994). The Coevolutionary Process. Chicago: University of Chicago Press.Google Scholar
Thrift, N. (2008). Non-representational Theory: Space, Politics, Affect. London and New York: Routledge.Google Scholar
Tooby, J. & DeVore, I. (1987). The reconstruction of hominid behavioral evolution through strategic modeling. In Kinzey, W. G., ed., The Evolution of Human Behavior: Primate Models. Albany: State University of New York Press, pp. 183237.Google Scholar
Vayda, A. P., ed. (1969). Environment and Cultural Behavior. New York: The Natural History Press.Google Scholar
Vellend, M., Brown, C. D., Kharouba, H. M., McCune, J. L., & Myers-Smith, I. H. (2013). Historical ecology: using unconventional data sources to test for effects of global environmental change. American Journal of Botany, 100, 12941305.Google Scholar
Vera, F. W. M. (2000). Grazing Ecology and Forest History. Wallingford: CABI Publishing.Google Scholar
Verheyen, K., Bossuyt, B., Hermy, M., & Tack, G. (1999). The land use history (1278–1990) of a mixed hardwood forest in western Belgium and its relationship with chemical soil characteristics. Journal of Biogeography, 26, 1115–28.Google Scholar
Vestbö-Franzén, A. (2005). Råg och rön: om mat, människor och landskapsförändringar i norra Småland, ca 1550–1700. Dept. of Cultural Geography, Stockholm University 132.Google Scholar
Warming, E. (1895). Plantesamfund. Kjøbenhavn: P.G. Philipsens Forlag. (English edition: Oecology of Plants. 1909. Oxford: Clarendon Press).Google Scholar
Watson, P. J., LeBlanc, S. A., & Redman, C. L. (1971). Explanation in Archeology: An Explicitly Scientific Approach. New York: Columbia University Press.Google Scholar
Watt, A. S. (1947). Pattern and process in the plant community. Journal of Ecology, 35, 122.Google Scholar
Welinder, S. (2011). Early farming households. In Myrdal, J. & Morell, M., eds., The Agrarian History of Sweden from 4000 BC to AD 2000. Lund: Nordic Academic Press, pp. 1845.CrossRefGoogle Scholar
Whatmore, S. (2002). Hybrid Geographies: Natures, Culture, Spaces. London: Sage Publications.Google Scholar
Whiten, A. & Erdal, D. (2012). The human socio-cognitive niche and its evolutionary origins. Philosophical Transactions of the Royal Society B, 367, 2119–29.Google Scholar
Whittaker, R. H. (1975). Communities and Ecosystems. New York: Macmillan.Google Scholar
Widgren, M. (2012). Landscape research in a world of domesticated landscapes: the role of values, theory, and concepts. Quaternary International, 251, 117–24.Google Scholar
Wiegand, K., Ward, D., & Saltz, D. (2005). Multi-scale patterns and bush encroachment in an arid savanna with a shallow soil layer. Journal of Vegetation Science, 16, 311–20.Google Scholar
Willis, K. J., Araújo, M. B., Bennett, , Figueroa-Rangel, K. D., Froyd, B., , C. A., & Myers, N. (2007). How can a knowledge of the past help to conserve the future? Biodiversity conservation and the relevance of long-term ecological studies. Philosophical Transactions of the Royal Society B, 362, 175–86.Google Scholar
Willis, K. J., Gillson, L., & Brncic, T. M. (2004). How ‘virgin’ is virgin rainforest? Science, 304, 402–3.Google Scholar
Wilson, D. S. (2005). Evolutionary social constructivism. In Gottschall, J. & Wilson, D. S., eds., The Literary Animal: Evolution and the Nature of Narrative. Evanston, IL: Northwestern University Press, pp. 2037.Google Scholar
Wolch, J. R. & Emel, J., eds. (1998). Animal Geographies: Place, Politics, and Identity in the Nature-Culture Borderlands. London: Verso.Google Scholar
Wolf, E. R. (1982). Europe and the People without History. Berkeley: University of California Press.Google Scholar
Worster, D. (1994). Natures Economy, a History of Ecological Ideas, 2nd ed. Cambridge: Cambridge University Press.Google Scholar
Wu, J. & Loucks, O. (1995). From balance of nature to hierarchical patch dynamics: a paradigm shift in ecology. Quarterly Review of Biology, 70, 439–66.Google Scholar
Wylie, A. (1985). Putting Shakertown back together: critical theory in archaeology. Journal of Anthropological Archaeology, 4, 133–47.Google Scholar
Young, T. P., Partridge, N., & Macrae, A. (1995). Long-term glades in Acacia bushland and their edge effects in Laikipia, Kenya. Ecological Applications, 5, 97108.CrossRefGoogle Scholar
Zeder, M. A. (2011). The origins of agriculture in the Near East. Current Anthropology, 52, S221S235.Google Scholar
Zeuner, F. E. (1963). A History of Domesticated Animals. London: Hutchinson.Google Scholar

References

Barthel, S., Crumley, C. L., & Svedin, U. (2013). Bio-cultural refugia: combating the erosion of diversity in landscapes of food production. Ecology and Society, 18, 7188.Google Scholar
Beaufoy, G., Baldock, D., & Clark, J. (1994). The Nature of Farming: Low Intensity Farming Systems in Nine European Countries. London: Institute for European Environmental Policy.Google Scholar
Bentley, J. W. (1987). Economic and ecological approaches to land fragmentation: in defense of a much-maligned phenomenon. Annual Review of Anthropology, 16, 3167.Google Scholar
Biong Deng, L. (2010). Livelihood diversification and civil war: Dinka communities in Sudan’s civil war. Journal of Eastern African Studies, 4, 381–99.Google Scholar
Bloch, M. (1966). French Feudal History, translated by Janet Sondheimer. Originally published (1931) as Les caractères originaux de l’histoire rurale française. Berkley: University of California.Google Scholar
Bloch, M. (2015). The Historian’s Craft. Manchester: Manchester University Press. Originally published (1949) as Apologie pour l’Histoire ou Métier d’Historien. Paris: Armand Colin.Google Scholar
Block, S. & Webb, P. (2001). The dynamics of livelihood diversification in post-famine Ethiopia. Food Policy, 26, 333–50.Google Scholar
Blomkvist, P. & Larsson, J. (2013). An analytical framework for common-pool resource–large technical system (CPR-LTS) constellations. International Journal of the Commons, 7, 113–39.Google Scholar
Brännlund, I. & Axelsson, P. (2011). Reindeer management during the colonization of Sami lands: a long-term perspective of vulnerability and adaptation strategies. Global Environmental Change, 21, 10951105.Google Scholar
Bromley, D. W. (1991). Environment and Economy, Property Rights and Public Policy. Oxford: Basil Blackwell Ltd.Google Scholar
Chen, M. A. (1991). Coping with Seasonality and Drought. New Delhi: Sage Publications.Google Scholar
Cocks, M. (2006). Biocultural diversity: moving beyond the realm of ‘indigenous’ and ‘local’ people. Human Ecology, 34(2), 185200.Google Scholar
Couture, J. (2000). Native studies and the Academy. In Sefa Dei, G. J., Hall, B., & Goldin Rosenberg, D., eds., Indigenous Knowledges in Global Contexts: Multiple Readings of Our World. Toronto: University of Toronto Press, pp. 157–67.Google Scholar
Dahlström, A., Iuga, A., & Lennartsson, T. (2013). Managing biodiversity rich hay-meadows in the EU: a comparison of Swedish and Romanian grasslands. Journal of Environmental Conservation, 40(2), 194205.Google Scholar
Danell, Ö. (2000). Status, directions and priorities of reindeer husbandry research in Sweden. Polar Research, 19, 111–15.Google Scholar
Dasmann, R. F. (1991). The importance of cultural and biological diversity. In Oldfield, M. L. & Alcorn, J. B., eds., Biodiversity, Culture, Conservation, and Eco-development. Boulder, CO: Westview, pp. 715.Google Scholar
Davies, E. (1941). The patterns of transhumance in Europe. Geography, Journal of the Geographical Association, 117, 155–68.Google Scholar
Descola, P. & Pálsson, G. (1996). Introductory chapter. In Descola, P. & Pálsson, G., eds., Nature and Society. Anthropological Perspectives. London, New York: Routledge, pp. 121.Google Scholar
Dietz, T., Ostrom, E., & Stern, P. C.. (2003). The struggle to govern the commons. Science, 302, 1907–12.Google Scholar
Doak, D. F., Bigger, D., Harding, E. K., Marvier, M. A., O’Malley, R. E., & Thomson, D. (1998). The statistical inevitability of stability-diversity relationships in community ecology. The American Naturalist, 151, 264–76.Google Scholar
Durham, W. H. (1976). The adaptive significance of cultural behaviour. Human Ecology, 4, 89121.Google Scholar
Dynesius, M. & Jansson, R. (2000). Evolutionary consequences of changes in species’ geographical distributions driven by Milankovitch climate oscillations. Proceedings of the National Academy of Sciences of the United States of America, 97, 9115–20.Google Scholar
Ellis, E. C. (2011). Anthropogenic transformation of the terrestrial biosphere. Philosophical Transactions of the Royal Society A, 369, 1010–35.Google Scholar
Ellis, F. (1998). Household strategies and rural livelihood diversification. The Journal of Development Studies, 35, 138.Google Scholar
Ellis, F. (2000). Rural Livelihoods and Diversity in Developing Countries. Oxford: Oxford University Press.Google Scholar
Ellis, F. & Allison, E. (2004). Livelihood Diversification and Natural Resource Access. FAO Livelihood support programme Working paper 9.Google Scholar
Elton, C. S. (1958). The Ecology of Invasions by Animals and Plants. London: Methuens.Google Scholar
Eriksson, O. (2013). Species pools in cultural landscapes – niche construction, ecological opportunity and niche shifts. Ecography, 36, 403–13.Google Scholar
Feeny, D., Berkes, F., McCay, B. J., & Acheson, J. M. (1990). The tragedy of the commons: twenty-two years later. Human Ecology, 18(1), 119.Google Scholar
Fischer, A. G. (1960). Latitudinal variation in organic diversity. Evolution, 14, 6481.Google Scholar
Glacken, C. J. (1967). Traces on the Rhodian Shore. Nature and culture in Western thought from ancient times to the end of the eighteenth century. Berkeley, Los Angeles, London: University of California Press.Google Scholar
Goland, C. (1993). Field scattering as agricultural risk management: a case study from Cuyo Cuyo, Department of Puno, Peru. Mountain Research and Development, 13(4), 317–38.Google Scholar
Gonzalez, A. & Loreau, M. (2009). The causes and consequences of compensatory dynamics in ecological communities. Annual Review of Ecology, Evolution, and Systematics, 40, 393414.Google Scholar
Hardin, G. (1968). The tragedy of the commons. Science, 162, 1243–8.Google Scholar
Harper, J. L. (1969). The role of predation in vegetational diversity. Brookhaven Symposia in Biology, 22, 4861.Google Scholar
Hawkins, B. A., Field, R., Cornell, H. W., Currie, D. J., Guégan, J.-F., Kaufman, , Kerr, D. M., Mittelbach, J. T., Oberdorff, G. G., O’Brien, T., Porter, E. M., , E. E., & Turner, J. R. G. (2003). Energy, water, and broad-scale geographic patterns of species richness. Ecology, 84, 3105–17.Google Scholar
Hector, A., Hautier, Y., Saner, P., Wacker, L., Bagchi, R., Joshi, J., Scherer-Lorenzen, M., Spehn, E. M., Bazeley-White, E., Weilenmann, M., Caldeira, M. C., Dimitrakopoulos, P. G., Finn, J. A., Huss-Danell, K., Jumpponen, A., Mulder, C. P. H., Palmborg, C., Pereira, J. S., Siamantziouras, A. S. D., Terry, A. C., Troumbis, A. Y., Schmid, B., & Loreau, M. (2010). General stabilizing effects of plant diversity on grassland productivity through population asynchrony and overyielding. Ecology, 91, 2213–20.Google Scholar
Heggberget, T. M., Gaare, E., & Ball, J. P. (2002). Reindeer (Rangifer tarandus) and climate change: importance of winter forage. Rangifer, 22, 1332.Google Scholar
Helldin, J.-O. & Lennartsson, T. (2007). Agricultural landscapes in East Europe as reference areas for Swedish land management. In Surd, V. & Zotic, V., eds., Rural Space and Local Development. Cluj-Napoca: Cluj University Press, pp. 367–70.Google Scholar
Helle, T., Aspi, J., & Kilpelä, S. S. (1990). The effects of stand characteristics on reindeer lichens and range use by semi-domesticated reindeer. Rangifer Special Issue, 3, 107–14.Google Scholar
Hodder, I. (2012). Entangled: An Archaeology of the Relationships between Humans and Things. Oxford: Blackwell Wiley.Google Scholar
Holling, C. S. (1973). Resilience and stability of ecological systems. Annual Review of Ecology and Systematics, 4, 123.Google Scholar
Horstkotte, T. & Roturier, S. (2013). Does forest stand structure impact the dynamics of snow on winter grazing grounds of reindeer (Rangifer t. tarandus)? Forest Ecology and Management, 291, 162–71.Google Scholar
Horstkotte, T., Sandström, C., & Moen, J. (2014). Exploring the multiple use of boreal landscapes in northern Sweden: the importance of social-ecological diversity for mobility and flexibility. Human Ecology, 42, 671–82.Google Scholar
von Humboldt, A. (1808). Views of Nature, Translated by E. Otté. & H. Bohn. (1850). London: Bohn.Google Scholar
von Humboldt, A. & Bonpland, A. (1807). Essay on the Geography of Plants. Edited by Jackson, S. T.. (2009). Chicago, London: University of Chicago Press.Google Scholar
Ikonen, I. & Lammi, A., eds. (2000). Traditional Rural Biotopes in the Nordic Countries, the Baltic States and the Republic of Karelia: An International Seminar and Workshop in Turku May 2–May 4, 2000. Copenhagen: Nordic Council of Ministers.Google Scholar
IUCN. (2008). Report of the Director General on the Work of the Union since the IUCN World Conservation Congress, Bangkok, 2004. IUCN World Conservation Congress, Barcelona, Spain 2008. http://cmsdata.iucn.org/downloads/cgr_2008_8_dg_report.pdf.Google Scholar
Jochim, M. A. (1981). Strategies for Survival. Cultural Behaviour in an Ecological Context. New York: Academic Press.Google Scholar
Johnson, J. T. & Murton, B. (2007). Re/placing native science: indigenous voices in contemporary constructions of nature. Geographical Research, 45, 121–9.Google Scholar
Jörgensen, D. (2015). Rethinking rewilding. Geoforum, 65, 482–8.Google Scholar
Kivinen, S., Moen, J., Berg, A., & Eriksson, Å. (2010). Effects of modern forest management on winter grazing resources of reindeer in Sweden. Ambio, 39, 269–78.Google Scholar
Kricher, J. (2009). The Balance of Nature: Ecology’s Enduring Myth. Princeton, NJ: Princeton University Press.Google Scholar
Larsson, J. (2009). Fäbodväsendet 1550–1920: Ett centralt element i Nordsveriges jordbrukssystem [Summer Farms in Sweden 1550–1920: An Important Element in North Sweden’s Agricultural System]. Uppsala: Swedish University of Agricultural Sciences.Google Scholar
Larsson, J. (2012). The expansion and decline of a transhumance system in Sweden, 1550–1920. Historia Agraria, 56, 1139.Google Scholar
Larsson, J. (2014a). Boundaries and property rights: the transformation of a common-pool resource. Agricultural History Review, 62, 2040.Google Scholar
Larsson, J. (2014b). Labor division in an upland economy: workforce in a seventeenth-century transhumance system. The History of the Family, 19, 393410.Google Scholar
Leadley, P. W., Krug, C. B., Alkemade, R., Pereira, H. M., Sumaila, U. R., Walpole, M., Marques, A., Newbold, T., Teh, L. S. L, van Kolck, J., Bellard, C., Januchowski-Hartley, S. R., & Mumby, P. J. (2014). Progress towards the Aichi Biodiversity Targets: An Assessment of Biodiversity Trends, Policy Scenarios and Key Actions. Global biodiversity outlook 4, Technical Report, CBD Technical Series 78. Montreal, Canada: Secretariat of the Convention on Biological Diversity.Google Scholar
Lewontin, R. C. (1969). The meaning of stability. Brookhaven Symposia in Biology, 22, 1324.Google Scholar
Loh, J. & Harmon, D. (2005). A global index of biocultural diversity. Ecological Indicators, 5, 231–41.Google Scholar
Loreau, M., Naeem, S., Inchausti, P., Bengtsson, J., Grime, J. P., Hector, A., Hooper, D. U., Huston, M. A., Raffaelli, D., Schmid, B., Tilman, D., & Wardle, D. A. (2001). Biodiversity and ecosystem functioning: current knowledge and future challenges. Science, 294, 804–8.Google Scholar
Lundmark, H., Josefsson, T., & Östlund, L. (2013). The history of clear-cutting in northern Sweden. Driving forces and myths in boreal silviculture. Forest Ecology and Management, 307, 112–22.Google Scholar
Maffi, L. ed. (2001). On Biocultural Diversity: Linking Language, Knowledge, and the Environment. Washington, DC: Smithsonian Institution Press.Google Scholar
Maffi, L. & Woodley, E. (2010). Biocultural Diversity Conservation, a Global Sourcebook. New York: Earthscan.Google Scholar
Maurer, K., Weyand, A., Fischer, M., & Stöcklin, J. (2006). Old cultural traditions, in addition to land use and topography, are shaping plant diversity of grasslands in the Alps. Biological Conservation, 130, 438–46.Google Scholar
Mayr, E. (1991). One Long Argument – Charles Darwin and the Genesis of Modern Evolutionary Thought. Harvard University Press.Google Scholar
MacArthur, R. H. (1955). Fluctuations of animal populations, and a measure of community stability. Ecology, 36, 533–6.Google Scholar
MacArthur, R. H. & Wilson, E. O. (1967). The Theory of Island Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
McCloskey, D. N. (1975). The economics of enclosure: a market analysis. In Parker, N. William & Jones, E. L., eds., European Peasants and Their Markets: Essays in Agrarian Economic History, Princeton, NJ: Princeton University Press, pp. 123–60.Google Scholar
McCloskey, D. N. (1976). English open fields as behavior towards risk. Research in Economic History, 1, 124–70.Google Scholar
McCloskey, D. N. (1989). The open fields of England: rent, risk, and the rate of interest, 1300–1815. In Galenson, D. W., ed., Markets in History: Economic Studies of the Past. Cambridge: Cambridge University Press, pp. 551.Google Scholar
McCloskey, D. N. (1991). The prudent peasant: new findings on open fields. Journal of Economic History, 51, 343–55.Google Scholar
McKean, M. (1986). Management of traditional common lands (Iriaichi) in Japan. In Proceedings of the Conference on Common Property Resource Management. National Research Council, Washington, DC: National Academy of Sciences, pp. 533–89.Google Scholar
McNaughton, S. J. (1977). Diversity and the stability of ecological communities: a comment on the role of empiricism in ecology. The American Naturalist, 11, 515–25.Google Scholar
McNeely, J. A. (2000). Cultural factors in conserving biodiversity. In Wilkes, A., Tillman, H., Salas, M., Grinter, T., & Shaoting, Y., eds., Links between Cultures and Biodiversity. Proceedings of the Cultures and Biodiversity Congress. Yunnan Science and Technology Press, pp. 128–82.Google Scholar
McNeely, J. A. & Keeton, W. S. (1995). The interaction between biological and cultural diversity. In Droste, B. von, Plachter, H., & Rössler, M., eds., Cultural Landscapes of Universal Value: Components of a Global Strategy. Jena and New York: Fischer and UNESCO, pp. 2536.Google Scholar
Menge, B. A. & Sutherland, J. P. (1976). Species diversity gradients: synthesis of the roles of predation, competition, and temporal heterogeneity. The American Naturalist, 110, 351–69.CrossRefGoogle Scholar
Moen, J. (2008). Climate change: effects on the ecological basis for reindeer husbandry in Sweden. Ambio, 37, 304–11.Google Scholar
Moen, J., Andersen, R., & Illius, A. (2006). Living in a seasonal environment. In Danell, K., Bergström, R., Duncan, P., & Pastor, J., eds., Large Herbivore Ecology, Ecosystem Dynamics and Conservation. Cambridge: Cambridge University Press, pp. 5070.Google Scholar
Moen, J. & Danell, Ö. (2003). Reindeer in the Swedish mountains: an assessment of grazing impacts. Ambio, 32, 397402.Google Scholar
Moen, J. & Keskitalo, E. C. H. (2010). Interlocking panarchies in multi-use boreal forests in Sweden. Ecology and Society, 15(3), 17.Google Scholar
Monbiot, G. (2013). Feral: Searching for Enchantment on the Frontiers of Rewilding. London: Allen Lane.Google Scholar
de Moor, T. (2015). The Dilemma of the Commoners: Understanding the Use of Common Pool Resources in Long-Term Perspective. New York: Cambridge University Press.Google Scholar
Moore, J. L., Manne, L., Brooks, T., Burgess, N. D., Davies, R., Rahbek, C., Williams, P., & Balmford, A. (2002). The distribution of cultural and biological diversity in Africa. Proceedings of the Royal Society of London, 269, 1645–53.Google Scholar
Netting, R. M. C. (1981). Balancing on an Alp. Cambridge: Cambridge University Press.Google Scholar
Netting, R. M. C. (1993). Smallholders, Householders: Farm Families and the Ecology of Intensive, Sustainable Agriculture. Stanford, CA: Stanford University Press.Google Scholar
North, D. C. (1990). Institutions, Institutional Change and Economic Performance. Cambridge: Cambridge University Press.Google Scholar
North, D. C. (1991). Institutions. The Journal of Economic Perspectives, 5(1), 97112.Google Scholar
Oppermann, R., Beaufoy, G., & Jones, G., eds. (2012). High Nature Value Farming in Europe. Ubstadt-Weiher: Verlag Regionalkultur.Google Scholar
Orr, A., Blessings, M., & Saiti-Chitsonga, D. (2009). Exploring seasonal poverty traps: the ‘six-week window’ in southern Malawi. Journal of Development Studies, 45, 227–55.Google Scholar
Östlund, L., Zachrisson, , , O., & Axelsson, A.-L. (1997). The history and transformation of a Scandinavian boreal forest landscape since the 19th century. Canadian Journal of Forest Research, 27, 11981206.Google Scholar
Ostrom, E. (1990). Governing the Commons: The Evolution of Institutions for Collective Action. Cambridge: Cambridge University Press.Google Scholar
Ostrom, E. (1992). Crafting Institutions for Self-Governing Irrigation Systems. San Francisco, CA: Institute for Contemporary Studies Press.Google Scholar
Ostrom, E. (2005). Understanding Institutional Diversity. Princeton, NJ: Princeton University Press.Google Scholar
Ostrom, E., Gardner, R., & Walker, J. (1994). Rules, Games, and Common-Pool Resources, Ann Arbor: University of Michigan Press.Google Scholar
Ouma, C., Obando, J., & Koech, M. (2012). Post drought recovery strategies among the Turkana pastoralists in northern Kenya. Scholarly Journals of Biotechnology, 1, 90100.Google Scholar
Paine, R. T. (1966). Food web complexity and species diversity. The American Naturalist, 100, 6575.Google Scholar
Paine, R. T. (1969). A note on trophic complexity and community stability. The American Naturalist, 103, 91–3.Google Scholar
Panjek, A. & Larsson, J., eds. (2017). Integrated Peasant Economy in a Comparative Perspective. Alps, Scandinavia, and Beyond. Koper: University of Primorska Press.Google Scholar
Plieninger, T., Höchtl, F., & Spek, T. (2006). Traditional land-use and conservation in European rural landscapes. Environmental Science and Policy, 9, 317–21.Google Scholar
Plumwood, V. (2006). The concepts of a cultural landscape, nature, culture and agency in the land. Ethics and the Environment, 11, 115–50.Google Scholar
Posey, D. A., ed. (1999). Cultural and Spiritual Values of Biodiversity. Nairobi: UNEP and Intermediate Technology Publications.Google Scholar
Pretty, J., Adams, B., Berkes, F., Ferreira de Athayde, S., Dudley, N., Hunn, E., Maffi, L., Milton, K., Rapport, D., Robbins, P., Sterling, E., Stolton, S., Tsing, A., Vintinnerk, E., & Pilgrim, S. (2009). The intersections of biological diversity and cultural diversity: towards integration. Conservation and Society, 7, 100–12.Google Scholar
Prothero, R. (1961). English Farming: Past Present. 6th ed. Chicago: Quadrangle Books. First published in 1912 by Longmans, Green & Co. Ltd,Google Scholar
Quétier, F., Rivoal, , Marty, F., de Chazal, P., Thuiller, J., , W., & Lavorel, S. (2010). Social representations of an alpine grassland landscape and socio-political discourses on rural development. Regional Environmental Change, 10, 119–30.Google Scholar
Reardon, T., Taylor, J. E., Stamoulis, K., Lanjouw, P., & Balisacan, A. (2000). Effects of nonfarm employment on rural income inequality in developing countries: an investment perspective. Journal of Agricultural Economics, 51, 266–88.Google Scholar
Root, R. B. (1967). The niche exploitation pattern of the blue-grey gnatcatcher. Ecological Monographs, 37, 318–50.Google Scholar
Roturier, S. & Roué, M. (2009). Of forest, snow and lichen: Sami reindeer herders’ knowledge of winter pastures in northern Sweden. Forest Ecology and Management, 258, 1960–7.Google Scholar
Sanders, H. L. (1968). Marine benthic diversity: a comparative study. The American Naturalist, 102, 243–82.Google Scholar
Sandom, C., Donlan, J., Svenning, J.-C., & Hansen, D. (2013). Rewilding. In Macdonald, D. W. & Willis, K. J., eds., Key Topics in Conservation Biology 2. Wiley, pp. 430–51.Google Scholar
Sandström, C., Moen, , Widmark, J., , C., & Danell, Ö. (2006). Progressing toward co-management through collaborative learning: forestry and reindeer husbandry in dialogue. International Journal of Biodiversity Science and Management, 2, 326–33.Google Scholar
Scholl, M. D. (2008). The American Yeoman: An Historical Ecology of Production in Colonial Pennsylvania. Chapel Hill: University of North Carolina.Google Scholar
Scholl, M. D., Murray, D. S., & Crumley, C. L. (2010). Comparing trajectories of climate, class and production: an historical ecology of American yeomen. In Vaccaro, I., Smith, E. A., & Aswani, S., eds., Environmental Social Sciences: Methods and Research Design. Cambridge: Cambridge University Press, pp. 322–48.Google Scholar
Scott, G. H. (1963). Uniformitarianism, the uniformity of nature, and paleoecology. New Zealand Journal of Geology and Geophysics, 6, 510–27.Google Scholar
Simpson, G. G. (1964). Species density of North American recent mammals. Systematic Zoology, 13, 5773.Google Scholar
Siy, R. Y. (1982). Community Resource Management, Lessons from the Zinjera. Quezon City: University of the Philippines Press.Google Scholar
Skutnabb-Kangas, T., Maffi, L., & Harmon, D. (2003). Sharing a World of Difference: The Earth’s Linguistic, Cultural, and Biological Diversity. Paris: UNESCO, Terralingua, WWF.Google Scholar
Smith, D. R, Gordon, A., Meadows, K., & Zwick, K. (2001). Livelihood diversification in Uganda: patterns and determinants of change across two rural districts. Food Policy, 26, 421–35.Google Scholar
Smith, E. A. (2001). On the coevolution of cultural, linguistic, and biological diversity. In Maffi, L., ed., On Biocultural Diversity. Washington, DC: Smithsonian Institution Press, pp. 95117.Google Scholar
Snow, C. P. (1959). The Two Cultures and the Scientific Revolution. The Rede Lecture. London: Cambridge University Press.Google Scholar
Soulé, M. E. & Noss, R. F. (1998). Rewilding and biodiversity: complementary goals for continental conservation. Wild Earth, 8, 1928.Google Scholar
Stern, P. C. (2011). Design principles for global commons: natural resources and emerging technologies. International Journal of the Commons, 5, 213–32.Google Scholar
Swinnen, J. F. M. (1999). The political economy of land reform choices in Central and Eastern Europe. Economics of Transition, 7, 637–64.Google Scholar
Tylor, E. B. (1871). Primitive Culture: Researches into the Development of Mythology, Philosophy, Religion, Language, Art, and Custom Vols. 1–2. London: John Murray.Google Scholar
UNEP. (2007). Global Environment Outlook 4. Nairobi: UNEP.Google Scholar
UNESCO. (2001). Resolutions, vol. 1. Records of the General Conference, 31st session, Oct.–Nov.Google Scholar
Verdery, K. (2003). The Vanishing Hectare: Property and Value in Postsocialist Transylvania. Ithaca, NY: Cornell University Press.Google Scholar
Wallace, A. R. (1878). Tropical Nature and Other Essays. New York: Macmillan.Google Scholar
Whittaker, R. H. (1960). Vegetation on the Siskiyou Mountains, Oregon and California. Ecological Monographs, 30, 279–338.Google Scholar
Whittaker, R. H. (1972). Evolution and measurement of species diversity. Taxon, 21, 213–51.Google Scholar
Wilcox, B. A. & Duin, K. N. (1995). Indigenous cultural and biological diversity: overlapping values of Latin American ecoregions. Cultural Survival Quarterly, 18, 4953.Google Scholar
Williams, G. (1977). Differential risk strategies as cultural style among farmers in the Lower Chobut Valley, Patagonia. American Ethnologist, 4, 6583.Google Scholar
Winterhalder, B., Lu, F., & Tucker, B. (1999). Risk-sensitive adaptive tactics: models and evidence from subsistence studies in biology and anthropology. Journal of Anthropological Research, 7(4), 301–48.Google Scholar

References

Austin, Z., McVittie, A., McCracken, D., Moxey, A., Moran, D., & White, P. C. L. (2015). Integrating quantitative and qualitative data in assessing the cost-effectiveness of biodiversity conservation programmes. Biodiversity and Conservation, 24(6), 1359–75. doi: 10.1007/s10531-015-0861-4.Google Scholar
Baleé, W., ed. (1998). Advances in Historical Ecology. New York: Columbia University Press.Google Scholar
Barber, R. J. & Berdan, F. F. (1998). The Emperor’s Mirror: Understanding Cultures through Primary Sources. Tucson: University of Arizona Press.Google Scholar
Boix-Mansilla, V. (2010). Learning to synthesize: the development of interdisciplinary understanding. In Frodeman, R., Klein, J. T., & Mitcham, C., eds., The Oxford Handbook of Interdisciplinarity. Oxford: Oxford University Press, pp. 288306.Google Scholar
Burke, P. (2008). What Is Cultural History? 2nd ed. Cambridge: Polity Press.Google Scholar
Cadogan, D. (2014). Funding for research? Look to the crowd. College & Research Libraries Newsvol, 75(5), 268–71.Google Scholar
Chettiparamb, A. (2007). Interdisciplinarity: A Literature Review. Southampton, UK: University of Southampton. The Interdisciplinary Teaching and Learning Group, Subject Centre for Languages, Linguistics and Area Studies, School of Humanities.Google Scholar
Choi, B. C. & Pac, A. W. (2006). Multidisciplinarity, interdisciplinarity and transdisciplinarity in health research, services, education and policy: 1. Definitions, objectives, and evidence of effectiveness. Clinical and Investigative Medicine, 29(6), 351–64.Google Scholar
Creswell, J. W. (2015). A Concise Introduction to Mixed Methods Research. Los Angeles: Sage Publications.Google Scholar
Crumley, C. L., ed. (1994). Historical Ecology: Cultural Knowledge and Changing Landscapes. Santa Fe, NM: School of American Research Press.Google Scholar
Crumley, C. L. (1995). Heterarchy and the analysis of complex societies. Archaeological Papers of the American Anthropological Association, 6, 15. doi: 10.1525/ap3a.1995.6.1.1Google Scholar
Crumley, C. L. (2005). Remember how to organize: heterarchy across the disciplines. In Beekman, C. S. & Baden, W. S., eds., Nonlinear Models for Archaeology and Anthropology, Aldershot, UK: Ashgate Press, pp. 3550.Google Scholar
Crumley, C. L. (2007a). Historical ecology: integrated thinking at multiple temporal and spatial scales. In Hornborg, A. & Crumley, C. L.. eds., The World System and the Earth System: Global Socio-environmental Change and Sustainability since the Neolithic. Walnut Creek, CA: Left Coast Press, pp. 1528.Google Scholar
Crumley, C. L. (2007b). Heterarchy. In Darity, W. A., ed., International Encyclopedia of the Social Sciences, 2nd ed. Detroit, MI: Macmillan Reference USA, pp. 468–9.Google Scholar
Crumley, C. L. & Marquardt, W., eds. (1987). Regional Dynamics: Burgundian Landscapes in Historical Perspective. New York: Academic Press.Google Scholar
Dahlström, A., Lennartsson, T., & Iuga, A. M. (2013). Managing biodiversity rich hay meadows in the EU: a comparison of Swedish and Romanian grasslands. Environmental Conservation, 40(2), 194205.Google Scholar
Domaas, S. T. (2007). The reconstruction of past patterns of tilled fields from historical cadastral maps using GIS. Landscape Research, 32(1), 2343.Google Scholar
Donoghue, M. J. & Moore, B. R. (2003). Toward an integrative historical biogeography. Integrative and Comparative Biology, 43, 261–70.Google Scholar
Frodeman, R., Klein, J. T., & Mitcham, C. (2012). The Oxford Handbook of Interdisciplinarity, Reprint edition. Oxford: Oxford University Press.Google Scholar
Giles, J. (2012). Like it? Pay for it: with conventional sources of money drying up, some scientists are turning to crowd-funding. Nature, 481, 252–3.Google Scholar
Jaarsma, S. (2002). Handle with Care: Ownership and Control of Ethnographic Materials. Pittsburgh, PA: University of Pittsburgh Press.Google Scholar
Jones, E A. (2009). Multi-temporal landscape history in Burgundy: an innovative application of genealogy software. In Frischer, B., Crawford, J. W., & Koller, D., eds., Making History Interactive. Computer Applications and Quantitative Methods in Archaeology (CAA). Proceedings of the 37th Annual International Conference on Computer Applications to Archaeology. Oxford: British Archaeological Reports.Google Scholar
Kaplan, K. K. (2013). Crowd-funding: cash on demand. Nature, 497, 147–9. doi: 10.1038/nj7447-147aGoogle Scholar
Kesselman, M. A. & Watstein, S. B. (2009). Creating opportunities: embedded librarians. Journal of Library Administration, 49(4), 383400.Google Scholar
Kuhn, T. S. (1996). The Structure of Scientific Revolutions, 3rd ed. Chicago: University of Chicago Press.Google Scholar
Latour, B. (1999). Pandora’s Hope: Essays on the Reality of Science Studies. Cambridge, MA: Harvard University Press.Google Scholar
Latour, B. (2011). From multiculturalism to multinaturalism: what rules of method for the new socio-scientific experiments? Nature and Culture, 6(1), 117.Google Scholar
Lennartsson, T., Westin, A., Erikson, M., Flygare, I. A., Isacson, M., & Morell, M. (2015). Between nature and society: interpretation of an early 19th century farmer’s diary. Agricultural History Review, 63(2), 265–85.Google Scholar
Madison, D. S. (2012). Critical Ethnography: Method, Ethics, and Performance, 2nd ed. Thousand Oaks, CA: Sage Publications.Google Scholar
Madry, S. (2006). Hic sunt dracones (Here be dragons): The integration of historical cartographic data into geographic information systems. In Archer, S. & Bartoy, K., eds., Between Dirt and Discussion: Methods, Materials, and Interpretation in Historical Archaeology. New York: Springer, pp 3360.Google Scholar
Madry, S., Jones, E. A., Murray, S., & Tickner, A. (2011). Une Micro-Histoire De La Terre Et De L’utilisation Des Ressources: L’intégration des SIG-H (Systèmes d’information géographique historiques) et des données qui y sont liées en Bourgogne du sud. Le Monde des Cartes, Rapport Cartographique National revue de Comité Français de Cartographie, 208, 7594.Google Scholar
Madry, S., Jones, E. A., & Tickner, A. (2013). Interdisciplinary History of Rural Water and Land Use in Southern Burgundy, France. Technical report. www .doaks.org/research/garden-landscape/garden-and-landscape-project-grant-reports/madry-report.Google Scholar
Madry, S., Jones, E. A., Tickner, A. Murry, S., & Misner, T. (2015). Water and landscape dynamics in southern Burgundy: two and a half centuries of water management in an agricultural landscape. Water History, 7(3), 301–35. doi: 10.1007/s12685-015-0132-z.Google Scholar
Moen, J. (2015). On viewing your belly button or the world. In Mineur, E. & Myrman, B., eds., All of Science! 15 Researchers on Integrated Research. Stockholm: The Swedish Research Council, pp. 34–9.Google Scholar
Okely, J. (2012). Anthropological Practice: Fieldwork and the Ethnographic Method. London: Bloomsbury Publishing.Google Scholar
Olsson, R. (2008). Mångfaldsmarker: Naturbetesmarker en värdefull resurs. Uppsala: Centrum för biologisk mångfald.Google Scholar
O’Reilly, K. (2012). Ethnographic Methods, 2nd ed. London, New York: Routledge.Google Scholar
Piajet, J. (1972). The epistemology of interdisciplinary relationships. In Apostel, L., ed., Interdisciplinarity: The Problems of Teaching and Research in Universities. Paris: Organization for Economic Cooperation and Development, pp. 2739.Google Scholar
Project Management Institute. (2013). A Guide to the Project Management Body of Knowledge (PMBOK Guide), 5th ed. Project Management Institute.Google Scholar
Riggio, R. E., Chaleff, I., & Lipman-Blumen, J., eds. (2008). The Art of Followership: How Great Followers Create Great Leaders and Organizations. San Francisco: Jossey-Bass Press.Google Scholar
Sampson, R. J. & Laub, J. H. (1998). Integrating quantitative and qualitative data. In Giele, J. Z. & Elder, G. H. Jr., eds., Methods of Life Course Research: Qualitative and Quantitative Approaches. Thousand Oaks, CA: Sage Publications, pp. 213–30.Google Scholar
Schensul, J. J. & Lecompte, M. D. (2013). Essential Ethnographic Methods: A Mixed Methods Approach, 2nd ed. Lanham, MA: AltaMira Press.Google Scholar
Strang, V. (2009). Integrating the social and natural sciences in environmental research: a discussion paper. Environment, Development and Sustainability, 11, 118.Google Scholar
Svallfors, S. (2012). Kunskapens människa: Om kroppen, kollektivet och kunskapspolitiken. Stockholm: Santerus.Google Scholar
Swetnam, T. W., Allen, C. D., & Betancourt, J. L. (1999). Applied historical ecology: using the past to manage for the future. Ecological Applications, 9(4), 11891206.Google Scholar
Thompson-Klein, J. (2012). A taxonomy of interdisciplinarity. In Frodeman, R., Klein, J. T., & Mitcham, C. eds., The Oxford Handbook of Interdisciplinarity. Reprint edition. Oxford: Oxford University Press, pp. 1530.Google Scholar
Weigmann, K. (2013). Tapping the crowds for research funding. EMBO Reports, 14, 1043–6. doi: 10.1038/embor.2013.180.Google Scholar
Wheat, R. E., Wang, W., Byrnes, J. E., & Ranghanathan, J. (2013) Raising money for scientific research through crowdfunding. Trends in Ecology & Evolution, 28, 73–4.Google Scholar
Whyte, K. P. (2013). On the role of traditional ecological knowledge as a collaborative concept: a philosophical study. Ecological Processes, 2(7).Google Scholar
Yang, Y., Zhang, S., Yang, J., Chang, L., Bu, K., & Xing, X. (2014). A review of historical reconstruction methods of land use/land cover. Journal of Geographical Sciences, 24(4), 746–66.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
×